You are on page 1of 250

Copyright

by

Edward Louis Goldmann

2002
Committee Certification of Approved Version

The Dissertation Committee for Edward Louis Goldmann Certifies that this
is the approved version of the following dissertation:

Thermoreversible Gelation of Aromatic Hydrocarbons

Committee:

C. Grant Willson, Supervisor

Joel W. Barlow, co-Supervisor

Peter F. Green

Donald R. Paul

Stephen E. Webber
Thermoreversible Gelation of Aromatic Hydrocarbons

by

Edward Louis Goldmann, B.S., M.S.

Dissertation
Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin


December 2002
For Clyde
Acknowledgements

The author is indebted to many individuals for innumerable actions that

have made this work possible. First and formost I would like to thank my parents,

Louis and Elizabeth Goldmann, for their untiring support of this work and all of

my endeavors to date, and into the future. Their love, protection, and advice has

allowed me to grow in ways few sons could ever claim, and I am grateful for their

unwavering belief in whatever I have chosen even when the ‘going got tough!’

Next would be my entire family for their support, even as many have

grown tired of inquiring: “…you are STILL in school?!?!?!” With the successful

completion of this work that will be the end of my extended formal education but

of course there will always be learning to be had….just not in a formal classroom

or University laboratory. This is further backed up by the fact there are no more

degrees to get after a PhD.

I am indebted to my advisors, Dr. Joel W. Barlow and Dr. C. Grant

Willson, for their guidance during this work. Dr. Barlow’s experience as well as

his practical knowledge of what was under study allowed this work to come to

full fruition. Dr. Willson was gracious in serving as my Committee Chairman

after Dr. Barlow retired to sunny Florida. This indebtedness is also extended to

v
my full committee, whose valuable input and time allowed this work to be shaped

to its present form—an impossible task for a student all alone!

I also thank my friends, both close and distant, for their support, the good

times, and tension-breaking moments that only good friends can provide.

Thumping away in the laboratory during lunch to Def Leppard will sorely be

missed! I extend a special thanks to John Sanchez, my loyal labmate and research

partner for these five (and a half…) long years of work. We had some great times
together in good ol’ CPE 3.426. He should be finishing soon as well and he better

with a new baby! We might have goofed off a bit more than the average lab pair

might normally should have, but we kept things real yet productive enough when

it really mattered.

Edward Louis Goldmann

Austin, Texas

vi
Thermoreversible Gelation of Aromatic Hydrocarbons

Publication No._____________

Edward Louis Goldmann, Ph.D.

The University of Texas at Austin, 2002

Supervisors: Joel W. Barlow and C. Grant Willson

Fumed silica of nominal size 20 microns is used to fill many different

polymer and prepolymer mixtures creating filled composite materials. Of interest

in this work is silica combined with low-viscosity non-polar hydrocarbon liquids

(aryl-vinyl monomers). Unless they are continuously mixed, unmodified binary

mixtures that have flowable viscosity (<75 wt% solids) phase separate over time

scales of several days to pure liquid and a close-packed cake of silica. It is

desired that this phase separation not occur for at least several months with
minimal required physical or chemical modification.

This work investigates the thermodynamic and rheological (mainly elastic)

properties of binary solutions of ethylbenzene, styrene, toluene, xylene, and tert-

butyl styrene each with low concentrations of various molecular weight

syndiotactic polystyrenes (SPS). These systems exhibit a distinct thermally

reversible phase transformation when cooled, going from a low-viscosity polymer

vii
solution to a translucent waxy solid of varying physical properties. Various

methods have been used to characterize both the polymers used in generating this

behavior, as well as the resultant gels. Analyses include differential scanning

calorimetry, nuclear magnetic resonance spectroscopy, gel permeation

chromatography, and dynamic viscoelastic rheometry. For a given binary system

gel formation temperatures and physical properties were found to be primarily a

function of polymer loading and molecular weight. The molecular weight of


polymer studied was between 4,300 and 42,000 Mn; concentrations used typically

at or below 5 wt% polymer; observed modulii varied greatly and somewhat

erratically but were in the 2 kPa to 20 kPa range. Nearly all mechanical property

measurements were done in the elastic regime (0.1-0.2% strain; yield point 2-4%).

Elastic networks were observed at extraordinarily low concentrations.

viii
Table of Contents

List of Tables......................................................................................................... xv

List of Figures .....................................................................................................xvii

List of Illustrations ............................................................................................. xxvi

Chapter 1: Introduction .......................................................................................... 1


1.1 A Path For Research.............................................................................. 1
1.2 Commercial Product Development ....................................................... 1
1.3 Academic Work..................................................................................... 5
1.4 Conclusion........................................................................................... 11
References .................................................................................................... 13

Chapter 2: Background......................................................................................... 14
2.1 Introduction ......................................................................................... 14
2.2 Solution/Gel Thermodynamics ........................................................... 15
2.2.1 Definition of a Physical Gel ....................................................... 16
2.2.2 Theory of Gelation ..................................................................... 18
2.2.2.1 Crystallization from Solution ...................................... 19
2.3 Rheology of Gels................................................................................. 26
2.3.1 Cone and Plate Theory ............................................................... 28
2.4 Syndiotactic Polystyrene ..................................................................... 31
2.4.1 Syndiotactic Copolymer ............................................................. 33
2.5 Conclusion........................................................................................... 35
References .................................................................................................... 36

Chapter 3: Polymer Synthesis .............................................................................. 38


3.1 Introduction ......................................................................................... 38
3.2 Catalyst Selection................................................................................ 38
3.3 Catalyst Synthesis ............................................................................... 41

ix
3.4 Homopolymerizations ......................................................................... 45
3.4.1 Polymerizations with Tetrabenzyltitanium ................................ 45
3.4.2 Trichlorocyclopentadienyltitanium Polymerizations ................. 48
3.4.3 Substituted Styrene Polymerizations.......................................... 49
3.5 Syndiotactic Polymerization Mechanism............................................ 50
3.5.1 Catalyst Activation..................................................................... 50
3.5.2 Polymerization Initiation............................................................ 52
3.5.3 Propagation................................................................................. 52
3.5.3 Termination ................................................................................ 53
3.6 Copolymerizations .............................................................................. 55
3.7 Conclusion........................................................................................... 56
References .................................................................................................... 57

Chapter 4: Polymer Characterization ................................................................... 58


4.1 Introduction ......................................................................................... 58
4.2 Polymer Synthesis Purification ........................................................... 58
4.3 Differential Scanning Calorimetry ...................................................... 59
4.3.1 Syndiotactic Polystyrene DSC ................................................... 60
4.3.2 Questra MA405 Copolymer DSC .............................................. 64
4.3.3 Synthesized Syndiotactic Polystyrenes ...................................... 65
4.3.4 Heating/Cooling Rate Dependence of Phase Transitions .......... 68
4.3.4.1 Melting Point Time-Dependence ................................ 68
4.3.4.2 Non-Isothermal Crystallization ................................... 70
4.3.4.3 DSC Operating Envelope ............................................ 74
4.4 Stereoregularity Analysis .................................................................... 76
4.4.1 NMR Tube Charging.................................................................. 77
4.4.2 Tacticity Discernment ................................................................ 80
4.4.3 Catalyst Effect on Syndiotacticity Purity ................................... 82
4.4.4 NMR Tube Cleaning .................................................................. 84
4.5 Molecular Weight Analysis................................................................. 85

x
4.5.1 Validity of Non-Conventional GPC Technique ......................... 87
4.6 Conclusion........................................................................................... 88
References .................................................................................................... 89

Chapter 5: Gel Formation..................................................................................... 90


5.1 Introduction ......................................................................................... 90
5.2 Gel Formation Process ........................................................................ 90
5.2.1 Invention Disclosure .................................................................. 92
5.2.1.1 Invention Description.................................................. 92
5.2.1.2 Invention Uniqueness.................................................. 92
5.2.1.3 Invention Usefulness ................................................... 94
5.2.1.4 Invention Limitations .................................................. 95
5.2.1.5 Prior Art/Publicity ....................................................... 95
5.2.2 Invention Addendum.................................................................. 95
5.3 Polymer Solvation ............................................................................... 96
5.3.1 Volatility Issues.......................................................................... 96
5.3.2 Thermal Polymerization Minimization ...................................... 97
5.4 Gel Storage.......................................................................................... 98
5.5 Formation with Solids Present ............................................................ 99
5.6 Conclusion........................................................................................... 99
References .................................................................................................. 101

Chapter 6: Thermodynamic Gel Characterization ............................................. 102


6.1 Introduction ....................................................................................... 102
6.2 Gel-Sol Transitions ........................................................................... 103
6.2.1 Differential Scanning Calorimetry Methodology .................... 103
6.2.2 Composition (Solvent) Variation Thermograms...................... 105
6.2.2.1 Cooling Rate Sweep .................................................. 106
6.2.3 Concentration Variation Study................................................. 108
6.2.4 Molecular Weight Variation Study .......................................... 110

xi
6.3 Sample Polymerization ..................................................................... 112
6.4 Equilibrium Transitions..................................................................... 114
6.5 Long-Term Gel Ageing..................................................................... 115
6.6 Discussion ......................................................................................... 119
6.6.1 Melting Transitions .................................................................. 119
6.6.2 Gelling/Crystallization Transitions .......................................... 121
6.8 Conclusion......................................................................................... 125
References .................................................................................................. 127

Chapter 7: Mechanical Gel Characterization ..................................................... 128


7.1 Introduction ....................................................................................... 128
7.2 Bulk Polymer Modulus/Strength....................................................... 128
7.2.1 Two Macro-Phase Injection Molded Parts............................... 129
7.2.2 Annealed Injection Molded Parts............................................. 131
7.3 Dynamic Viscoelastic Rheometry ..................................................... 133
7.3.1 Cone and Plate Theory ............................................................. 134
7.3.2 Sample Loading Considerations............................................... 136
7.3.3 Thermal Response Study of Instrument ................................... 139
7.3.4 Testing Procedure..................................................................... 143
7.4 DVR Case Studies............................................................................. 144
7.4.1 Dow SPS in Styrene ................................................................. 144
7.4.2 Dow SPS in tert-Butyl Styrene................................................. 146
7.4.3 UT 0515 SPS in Ethylbenzene ................................................. 151
7.4.4 UT 0918 SPS in Ethylbenzene ................................................. 153
7.5 Discussion ......................................................................................... 155
7.5.1 Unforseen Multiple Phase Behavior ........................................ 155
7.5.1.1 Frequency Sweep of Aged Gel.................................. 158
7.5.1.2 Variation of Measurement Start-Time ...................... 159
7.5.2 Optical Observation of Sol-Gel................................................ 160
7.5.3 DVR Instrument/Measurement Variance................................. 162

xii
7.6 Future Work ...................................................................................... 164
7.7 Conclusion......................................................................................... 165
References .................................................................................................. 167

Chapter 8: Summary and Conclusions ............................................................... 168


8.1 Introduction ....................................................................................... 168
8.2 Gelling Polymers............................................................................... 168
8.2.1 Catalyst Synthesis .................................................................... 168
8.2.2 Polymer Synthesis .................................................................... 169
8.2.3 Polymer Characterization ......................................................... 169
8.3 Gel Characterization.......................................................................... 171
8.3.1 Thermal Studies........................................................................ 171
8.3.2 Rheological Studies.................................................................. 172
8.4 In Closing .......................................................................................... 174

Appendix A: Product Development ................................................................... 176


A.1 Introduction ....................................................................................... 176
A.2 Monomer Purification and Handling................................................. 177
A.3 Anionic Initiators............................................................................... 179
A.4 Filler Surface Treatment.................................................................... 181
A.4.1 Vinyl Siloxanes ........................................................................ 181
A.4.1.1 Application of A172 Siloxane .................................. 182
A.4.2 Surface Treatment Evaluation.................................................. 183
A.4.2.1 Accelerated Separation Test ..................................... 184
A.4.2.2 Equilibrium Water Uptake ....................................... 187
A.4.2.3 Surface Chemical Affinity........................................ 188
A.4.3 Siloxane Modified Polymer ..................................................... 190
A.4.3.1 Free-Radical Siloxane Copolymerization ................ 191
A.4.3.2 Anionic Siloxane Chain-Termination ...................... 191
A.4.4 Other Surface Treatments......................................................... 192

xiii
A.4.4.1 Styryl-Siloxane ......................................................... 192
A.4.4.2 Glycidyl-methacrylate siloxane................................ 194
A.5 High Loading Solids Suspension ...................................................... 194
A.5.1 Matrix Thickening.................................................................... 195
A.5.2 Early Matrix Gelation Methods................................................ 196
A.5.2.1 Aluminum Stearates ................................................. 196
A.5.3 Alternative Soap Chemistry ..................................................... 198
A.5.3.1 Polystyryl Soap......................................................... 199
A.5.3.2 Dicarboxylated Gelling Agent.................................. 202
A.5.4 Ultra-fine Particle-based Colloid ............................................. 203
A.6 Bulk Polymerizations ........................................................................ 203
A.6.1 Pouch Reactors ......................................................................... 203
A.6.2 Divinylbenzene Copolymerizations ......................................... 208
A.7 Paste Pigmentation ............................................................................ 209
A.8 Composite Packaging Foil Evaluation .............................................. 210
A.9 Conclusion......................................................................................... 213
References .................................................................................................. 214

Bibliography........................................................................................................ 215

Vita .................................................................................................................... 221

xiv
List of Tables

Table 2.1: Gel-sol transition temperatures of CPS copolymer in various

solvents. Calculated solubility parameters using the methods and

tables of Small (1953) ...................................................................... 34

Table 4.1: High-Temperature 13C NMR of Various Polystyrenes. All data

obtained on 500 MHz instrument in d-8 xylene at 120ºC for one

hour................................................................................................... 82

Table 4.2: Molecular Weights and Distributions of Three Syndiotactic

Polystyrenes Under Study. THF carrier solvent at 25ºC and

1cc/min flow rate. Injected from aliquot drawn from freshly

boiled polymer/THF sample............................................................. 86

Table 6.1: Different Solvent/SPS Combinations at 1% Polymer Loading

with Heating/Cooling Rates of 10ºC/min. It is useful to note at

this junction that the cooling rate of ~10ºC/min appears to be the

same transition temperature as the zero-rate extrapolated

transition temperature from the linear transition region. (a) Three

slightly detected transitions at 4, -10, and –18.7ºC. ....................... 108


Table 6.2: Comparison of Calculated vs. Observed Melting Points for Six

Solvents in 1 wt% Dow SPS, ºC. Three solvent/polymer binaries

agree pretty well, three not so well. ............................................... 121

Table 6.3: Extrapolated Sol-Gel Transition Temperatures and Pure Polymer

Comparison. Four different concentrations and pure polymer

shown. ............................................................................................ 125

xv
Table 7.1: Various Properties of Injection-Molded Syndiotactic Polystyrene

Under Varying Processing Conditions. Shear modulus

calculated assuming Poisson’s Ratio of 0.30 (Argon 1999). ......... 133

Table 7.2: Results of Ageing UT0515 in EB at Various Loadings and Age

Temperatures. Ramped from loading temperature to setpoint

over 15 minutes, held for additional 45 minutes. 2% sample

cooled to 5ºC detached both times from cone while observing. .... 153
Table A.1: Modern Electronics Packaging Material Minimum

Specifications. As given by Bajaj (1995)...................................... 176

Table A.2: Molecular Weight Results From Pouch Polymerizations in

Glovebox. Samples 1, 2, and 3 were of too high molecular

weight/crosslinked and were not THF soluble. Sample 5 failed

to inject properly (both vials). Duplicate samples were run for

each polymerization, each corresponding pair of molecular

weight data agrees within 3%......................................................... 206

xvi
List of Figures

Figure 2.1: Calculated melting point of SPS/xylene system. Using Equation

2.19, melting point depression calculation....................................... 24

Figure 2.2: Experimental Data with Melting Point Depression Calculation

for SPS/xylene system. Points are preliminary laboratory data

using falling ball indicator................................................................ 25

Figure 3.1: Formation of initiating species by TCCT/MAO catalytic system.... 51

Figure 3.2: Coordination of titanium complex with styrene molecule (h4 co-

ordination). ....................................................................................... 52

Figure 3.3: Coordination of additional styrene molecules with (a) nascent

polymer chain (one styrene unit), and (b) growing polymer chain

(more than one styrene unit) via h3 co-ordination............................ 53

Figure 3.4: Three different transfer reactions leading to MWD broadening.

(a) β-hydrido abstraction, (b) α-olefin metalation, and (c) and (d)

chain transfer to monomer (two results). ......................................... 54

Figure 4.1: DOW SPS Thermogram. First heating, 25ºC to 290ºC from

quenched state, 20ºC/min heating rate, 14.2 mg sample. (a) glass


transition then immediate recrystallization just above the glass

transition temperature, (b) melting endotherm, (c) crystallization

exotherm........................................................................................... 61

Figure 4.2: DOW SPS Thermogram. Cooling, 290ºC to 70ºC, 20ºC/min

cooling rate, 14.2 mg sample. .......................................................... 62

xvii
Figure 4.3: DOW SPS Thermogram. Second heating, 70ºC to 290ºC from

annealed state, 20ºC/min heating rate, 14.2 mg sample. .................. 63

Figure 4.4: DOW Syndiotactic Copolymer (CPS). Range 70ºC to 290ºC,

20ºC/min rates. (a) first (quenched) heating scan, (b) second

(annealed) heating scan, (c) cooling scan......................................... 64

Figure 4.5: UT 0515 SPS Homopolymer. Range 70ºC to 290ºC, 20ºC/min

rates. (a) first (quenched) heating scan, (b) second (annealed)


heating scan, (c) cooling scan. ......................................................... 66

Figure 4.6: UT 0918 SPS Homopolymer. Range 70ºC to 290ºC, 20ºC/min

rates. (a) first (quenched) heating scan, (b) second (annealed)

heating scan, (c) cooling scan. ......................................................... 67

Figure 4.7: DOW SPS Varying Heating Rates Thermogram. (a) 20ºC/min,

(b) 15ºC/min, (c) 10ºC/min, (d) 5ºC/min, (e) 2.5ºC/min. Cycled

at 10ºC/min down to 70ºC................................................................ 69

Figure 4.8: DOW SPS Varying Cooling Rates Thermogram. (a) 20ºC/min,

(b) 15ºC/min, (c) 10ºC/min, (d) 5ºC/min, (e) 2.5ºC/min.................. 70


Figure 4.9: Extrapolation of SPS Crystallization Point to Zero Cooling Rate.

(a) linear extrapolation to 0ºC/min cooling rate, (b) square

extrapolation to 0ºC/min cooling rate. ............................................. 72

Figure 4.10: Broad Cooling Rate Range Crystallization Points of SPS.

Region with excellent linear fit shown extrapolated to zero

cooling rate....................................................................................... 73

xviii
Figure 4.11: CPE 3.464 Perkin Elmer DSC 7 Operating Envelope. (a)

Intercooler OFF; (b) intercooler ON. Actual experimental data

using 22 mg (large) polymer sample and observing the controller

compliance LED............................................................................... 75
13
Figure 4.12: C NMR of SPS in d-8 Xylene. Scanned for one hour at 120ºC,

truncated to pertinent region of chemical shift (baseline noise

removed). Shown both polymer and solvent relative responses,


polymer peak normalized to 100...................................................... 79

Figure 4.13: Isotactic and Atactic Polystyrene 13C NMR Far-Downfield

Responses. Scans taken at 120ºC for 1 hour in d-8 xylene. (a)

atactic polystyrene and (b) isotactic polystyrene (correlates with

Kawamura ibid.)............................................................................... 81

Figure 4.14: Intra-Chain Syndiotactic Content from Two Catalysts. (a)

tetrabenzyltitanium/methylaluminoxane catalytic system and (b)

trichlorocyclopentadienyl titanium/methylaluminoxane catalytic

system............................................................................................... 83
Figure 5.1: Dow Questra MA405 Copolymer. DSC thermogram of a sample

taken from material that has been processed via this invention.

Inset: Closer look at the material’s slow crystallization as it is

slowly heated above its glass transition temperature (~100ºC)

from an amorphous state. ................................................................. 93

xix
Figure 6.1: DSC Thermograms of Dow SPS in EB (same sample pan).

Large range of cooling rates; rates are from top (a) to bottom (b):

5, 10, 15, 20, 25, 30, 35, 40, 45, and 50ºC/min. Sol-gel transition

peaks shown by open boxes. (The peaks are far more pronounced

when appropriate analysis scales are used; in the figure the peaks

are compressed by the large y-axis range.) .................................... 106

Figure 6.2: Sol-Gel Transition as Function of Cooling Rate. 1% Dow SPS in


EB, much akin to that shown in Figure 4.10 for pure SPS

polymer, but transition temperatures depressed about 200ºC.

Linear fit data from 20 to 50ºC/min. .............................................. 107

Figure 6.3: Various Dow SPS Loading Levels in EB Solvent, Sol-Gel

Transition as Function of Cooling Rate. Elevated polymer

loadings indicate more than one transition occurring (multiple

gel phases), dependent on cooling rate........................................... 109

Figure 6.4: ‘Dynamic’ Phase Diagram for Dow SPS/EB Binary System.

Broken lines indicate (simplified) phase boundary locations to


the extent data is available. (a) heating at 10ºC/min from gel, (b)

cooling at 10ºC/min from sol. ........................................................ 110

Figure 6.5: Various MW Polymers at 1 wt% in EB. Heavier polymers show

virtually direct translation as function of MW, very low MW

polymer (0918) indicating gelation process becoming less a

function of cooling rate. ................................................................. 112

xx
Figure 6.6: Extinction of Inhibitor in DSC Samples. Sample cycled at

8ºC/min from 25ºC to 140ºC, 7 passes. First cooling pass (a),

seventh cooling pass (b). Transition temperatures open circles..... 113

Figure 6.7: DSC of Two Year Old Toluene/CPS Gel. Lid of storage vessel

allowed solvent escape (via diffusion) of toluene solvent.

*Unknown actual composition. Suggesting heat-history and

ageing-time dependent phase behavior. ......................................... 116


Figure 6.8: Gel-Sol and Sol-Gel Transition for Highly Loaded (34%) Dow

SPS/EB system. 30ºC/min heating rate after being quenched

from a sol, (b) 10ºC/min cooling rate, (c) 10ºC/min heating rate.

Calculated melting point value is 132ºC via Flory......................... 117

Figure 6.9: Elevated Temperature Change Rate Cycling on 34wt% Dow

SPS/EB System. (a), (b), and (c) heating rates of 50, 70, and

90ºC/min......................................................................................... 118

Figure 6.10: Six Different Solvents in Dow SPS, Calculated Melting Phase

Boundaries Using Equation 2.19.................................................... 120


Figure 6.11: Extrapolation of Sol-Gel Transition Point to Zero Cooling Rate

for Various Concentration of Dow SPS in Styrene. Guenet

(1988) suggests linear extrapolation to zero rate provided for

‘equilibrium’ formation temperature, as well as representing a

maximum gelling rate..................................................................... 122

xxi
Figure 6.12: Sol-Gel Transitions Extrapolated for Four Concentrations of

Dow SPS in EB. Extrapolation range between 20 and

50ºC/minute cooling rate................................................................ 124

Figure 7.1: Stress-Strain of Two-Phase Injection Molded Polystyrene. Mold

temperature of 175ºC, 25ºC testing temperature, 0.2 in/min strain

rate.................................................................................................. 131

Figure 7.2: Stress-Strain of SPS Bars Aged at 125ºC for 24 Hours. Mold
temperature of 175ºC, tested at room temperature at 0.2 in/min

strain rate. ....................................................................................... 132

Figure 7.3: Elastic Regime Cycling of 2 wt% Dow SPS in EB. 70ºC load,

aged 30 minutes to 25ºC, each strain ramp lapsed 15 minutes, 10

minute pauses between ramps; filled shapes: storage modulus

(G’), open shapes: loss modulus (G”); (convention within this

work). ............................................................................................. 137

Figure 7.4: Step Change from 50 to 25ºC. 1% SPS in EB gel cast between

surfaces and the temperatures monitored over several minutes. .... 141
Figure 7.5: 2-minute Temperature Ramp-Down Program. Cone keeping

better pace with the plate now........................................................ 141

Figure 7.6: 4-minute Temperature Ramp-Down. Near-perfect agreement of

cone temperature with plate and program temperature this time. .. 142

xxii
Figure 7.7: Storage and Loss Response and Failure Point of Dow SPS in

Styrene. Loaded at 70ºC, immediate cool to 25ºC, aged 30

minutes, 1.5 Hz frequency, 30 seconds per point, 30 points, filled

symbols G’, open G”. Samples all failed in 1-3% strain range..... 145

Figure 7.8: Various Concentrations of Dow SPS in tBS Monomer. Loaded

at 60ºC, immediately cooled to 25ºC while measuring began,

0.1% strain, 1.5 Hz. ........................................................................ 147


Figure 7.9: Dow SPS 2.3 wt% in tBS aged at various temperatures per the

shown profile, 2 replicates. This time the modulus drops far

below the maximum reached upon initial formation. .................... 148

Figure 7.10: As in Previous Figure, 1.9 wt% Dow SPS in tBS Aged at Three

Different Temperatures. 2 replicates, 0.1% strain, 1.5 Hz ............ 149

Figure 7.11: 2.4 wt% Dow SPS in tBS Stepped Temperature Ageing. Loaded

at 60ºC, stepped down to 35ºC. ...................................................... 150

Figure 7.12: Ageing 1 wt% UT0515 in EB for 60 Minutes, Three Age

Temperatures of 5, 15, and 25ºC, 0.1% strain at 1.5 Hz. Notice


at gel incipient (all about 40ºC) the sharp rise in modulus, slight

drop, then larger rise to a peak then falling off. ............................. 152

Figure 7.13: UT 0918 Modulus vs. Time for Two Different Ageing

Temperatures. Loaded at 30ºC, ramped to ageing temperature

over 15 minutes, 0.1% strain, 30 seconds/point, 1.5 Hz. Erratic

modulus measured due to discontinuous sample medium. ............ 154

xxiii
Figure 7.14: Monitoring Formation of 0.52 wt% Dow SPS/EB Gel, 0.1%

Strain, 1.5 Hz. Loaded at 70ºC, measurements started

immediately while temperature approached noted setpoint, 30

sec/measuring point. (25ºC run measurement not begun for 3

minutes after load/temperature set.) ............................................... 156

Figure 7.15: Pulsed Oscillatory Measurements on Nascent 0.54% SPS/EB

Gel. Loaded at 70ºC, measurements started 3 minutes after


loading while cooling to 25ºC, 30 seconds/measuring points. 5

minute of measurement alternating with 15 minute pauses, 2

replicates......................................................................................... 157

Figure 7.16: Frequency Sweep 0.5 to 70 Hz at 0.1% Strain of 1% SPS/EB

System. Aged 30 minutes (diamonds); aged additional two

hours, then re-run (circles). ............................................................ 159

Figure 7.17: Varying Start Time of Oscillation After Cast, 90Hz. Three

different delays, 0.5, 1.5, and 2.5 minutes. These three start

times encompass the peak in modulus. .......................................... 160


Figure 7.18: Time-Resolved UV-Vis Observation of 1% SPS/EB System.

Two different phenomena in absorbance can be gleaned from this

first investigational run................................................................... 162

Figure 7.19: Nine Identical 0.1% Strain Storage Modulus Observations. 1%

SPS/EB gel, temperature slowly ramped down as indicated,

initial modulus signal occurred very repeatedly at 38ºC plate

temperature..................................................................................... 163

xxiv
Figure 7.20: 90% Confidence Interval of 0.1% Strain Experiment. Most

variance appears to be in when the peak rolls over and how broad

the downside of the modulus response is. ...................................... 164

Figure A.1: Equilibrium Water Uptake For Minco 40 micron Silica. Shown

as function of vapor pressure of water present............................... 188

Figure A.2: GPC of one of the bag polymers. Middle peak is main polymer

signal. Concentration of only 1 mg/cc in THF solvent still


caused signal to go off scale (over 650 mV for this particular

instrument). .................................................................................... 208

xxv
List of Illustrations

Illustration 3.1: (a) Trichlorocyclopentadienyltitanium (TCCT) molecule.

Dark yellow powder, 96% purity from Aldrich (b)

tetrabenzyltitanium (TBT), synthesized in-house (dark red),

(c) methylaluminoxane (in solution, n reported to vary from 3

to large, a distribution) ............................................................... 40

Illustration 3.2: TBT Catalyst Synthesis Reactor System. Bath maintains –

20ºC environment, BMC is charged to 3-neck flask, heptane

diluted titanium tetrachloride held in addition funnel, argon

cover gas dead-ended to system through top of addition

funnel. Rubber septa seal the flask and funnel necks................ 42

Illustration 4.1: Hot Solution Charging of 4 mm NMR Tube. Pressure source

at only a few psi, regulated via a fingerhole on hypodermic

needle Luer into test tube to ease in hot liquid with positive

pressure....................................................................................... 78

Illustration 4.2: Removing NMR Tube Charge and Cleaning. Performed in

boiling xylene solvent and pulling ~50 cc of solvent through


NMR tube................................................................................... 85

Illustration 5.1: Test-Tube Erlenmeyer Flask ‘Oven’. Propane flame used to

heat system to create small samples of polymer/solvent gel

quickly........................................................................................ 97

xxvi
Illustration 6.1: DSC Heat Flow Signal Interpretation. Small peak/response

temperatures taken at parallel-tangent peak maximum. (This

applies to melting and crystallization/gelation thermal

responses, not glass transitions.) .............................................. 105

Illustration 7.1: Cone and Plate Rheometer. Not to scale sectional schematic

of the instrument and measuring surfaces; nominal sample

meniscus shown (providing ~2% edge volume error); cone


apex truncated 50µ in axial length. .......................................... 135

Illustration A.1: A172 Vinyl Siloxane Silica Surface Modifying Chemical

(Whitco Chemical Company)................................................... 182

Illustration A.2a:Separation Profiles with No Liquid Modification and

Untreated Silica. ....................................................................... 185

Illustration A.2b:Separation Profiles with No Liquid Modification and 2%

A172 Surface-Treated Silica. ................................................... 185

Illustration A.3: UT Aromatic Oligomer Soaps. (a) polystyryl aluminum

hydroxide soap, (b) poly(tert-butyl styryl) aluminum


hydroxide soap; m was nominally attempted at 10 to 25

repeat units. .............................................................................. 200

Illustration A.4: Meta-Diisopropenylbenzene (DIPEB) Divinyl Molecule. It

is notable that the double bonds are not conjugated with the

ring, as is the case with the ortho- or para- isomers. ................ 202

Illustration A.5: Double dogbone loops for tensile testing of the composite

foil material (not to scale). ....................................................... 211

xxvii
Illustration A.6: Sealed joint ‘reverse’ dogbone for testing of linear bond

strength of the heat-sealed material (not to scale).................... 212

xxviii
Chapter 1: Introduction

1.1 A PATH FOR RESEARCH

This work primarily focuses on the phase and rheological behavior of

stereoregular polystyrene and stereoregular copolymers of styrene and ring-

substituted styrene in various low molecular weight aromatic solvents and

monomers. These systems have received minimal attention in the research

community until very recently (~15 years), primarily due to the relatively recent

successful synthesis of largely syndiotactic polystyrene (Ishihara 1986). Others

(Grassi 1987; Pellechia 1987) then proceeded to duplicate this work, citing

difficulty in maintaining the proper conditions, but nonetheless producing

polymer with significant purity of syndiotactic diads. Little work was done

initially on the polymer’s solution properties, presumably due to the absence of

any practical application; it even comes out of solution during polymerization!

Hence the polymer was initially recognized as being desirable for solvent resistant

applications.

This work desired to take advantage of this polymer’s gel-forming

behavior when solutions were formed (with difficulty) at elevated temperature

and subsequently cooled.

1.2 COMMERCIAL PRODUCT DEVELOPMENT

JemPac International of Round Rock, Texas, was founded to develop a

Reaction Injection Molding (RIM) process for encapsulation of microchips and

other electronic components. In this room-temperature process a pastelike slurry,

1
containing >75 wt% silica in tertiary-butyl styrene with divinyl benzene

crosslinking agent, is mixed with butyl lithium initiator and rapidly transferred to

a heated mold (70-80ºC) where it subsequently rapidly polymerizes. Early

development work indicated two major process/product deficiencies:

1. The anionic reaction chemistry is extremely fast and capable of

complete polymerization within 5-30 seconds at 80ºC. The chemistry

is, however, very sensitive to moisture. Great care must be taken


when preparing the slurry to minimize moisture content. Likewise

great care must be taken to prevent moisture vapor permeation into the

compound during storage/before use.

2. The simple slurry of silica filler and monomers showed severe phase-

separation during both processing and storage. Phase separation

during flow-processing caused blockage of lines and subsequent

process failure. Phase separation during storage caused the compound

to become not processable.

Moisture ingress was minimized through a series of steps. These included


vacuum-drying of the fumed silica, mixing of silica and dried monomers under

nitrogen gas blanketed conditions, and development of hermetically-sealed

packaging for the slurry.

Several different attempts were made to increase the monomer/silica

suspension’s rheological stability. The two things that could be varied in

attempting to accomplish this included increasing/changing the liquid

viscosity/properties and modifying the filler surface physical and chemical

2
properties. Quickly it became apparent that simply increasing the liquid viscosity

did little to improve the long-term stability of the suspension although it did

reduce phase-separation during flow-processing in the process. High molecular

weight commercial (atactic) polystyrene (APS) was used in upwards of 5 wt% of

total monomer, which greatly increased the bulk monomer viscosity to that of a

syrup, yet did little to reduce long-term separation of the silica from the monomer.

Although adding an even larger amount of polystyrene might have increased


settling time to at least several weeks, that would also have caused the bulk

system viscosity to be too high, effectively negating one of the major benefits

inherent to the RIM process (low energy requirements via easily flowable

materials).

Several different types of surface treatments/coatings were applied to the

filler; these were designed to change the filler’s affinity for the liquid material.

Again however, the suspension properties did not change significantly, and no

mixture exhibited long-term stability.

After many months of investigation, it was found that usage of an


aluminum stearate compound yielded excellent rheological control of the

monomer/silica mixture. Unfortunately, this additive interfered too much with the

anionic reaction chemistry. This approach was therefore abandoned. Although it

was abandoned, it demonstrated (Berezniak 1958) that a network-forming

additive was probably the best solution to the problem of filler settling. It is

unfortunate that the aluminum network-forming material operated via chemical

functionality that was incompatible with the rest of this particular process.

3
Shortly after abandoning the use of aluminum soaps for stabilizing the

reaction mixture it was found that syndiotactic polystyrene (SPS), manufactured

by Dow Chemical under the trade name Questra, provided a rather interesting

property: changing the viscosity of aromatic solvents by many orders of

magnitude with slight variations in temperature and at less than 1 wt% loading by

total weight of monomer, so around 0.25 wt% by total product weight. This

property is known as thermoreversible gelation and is the basis for this work.
The effect is physical in nature and due to a reversible association; there is

no chemical reaction as is normally the basis of forming gels/networks in polymer

systems. Other examples of physical gels arise from solutions of syndiotactic

poly(methyl-methacrylate) (SPMMA) in aromatic solvents (Berghams 1987) and

polyethylene in aliphatic solvents (Narh 1982). There do exist a few examples of

reversible chemical gels, which are quite extraordinary (Pezron 1988).

The syndiotactic polystyrene polymer product Dow manufactured was

normally filled with glass fibers (about 35 wt%) for use in elevated temperature

and solvent resistant applications. Dow however provided several pounds of pure
SPS polymer for use, as well as another stereoregular polystyrene copolymer that

contained 8 wt% paramethylstyrene (PMS) co-monomer. The copolymer

exhibited many of the same thermodynamic properties in solution as that of the

stereoregular homopolymer styrene, except that the sol/gel liquidus lines in

polymer/solvent systems are 20-30ºC depressed from those of pure solvent/SPS

systems, and the gels produced were more transparent at equivalent polymer

loadings.

4
1.3 ACADEMIC WORK

We found that a rather low concentration of gelling polymer (~1% by

weight or less) was sufficient to produce a gel with physical strength high enough

to achieve stable suspensions for at least six months. This finding was excellent

since one of the original goals was to minimize the amount of additive. However,

we also desired to quantify this response.

Once we determined that this type of polymer produced interesting results

and were set on studying its pure and solution properties for academic purposes,

we wished to synthesize stereoregular polymers in the laboratory. It was known

(Zucchini 1971) that rigorously inert conditions and exotic catalytic systems were

necessary to make high molecular weight polymers of this type. Two catalytic

systems were investigated for making polymers. One was selected for its reported

(Chien 1991) extraordinarily high efficiency in making syndiotactic polymer with

few intra-chain defects, and the other for its more “mainstream” properties used

by industry (Imabayashi 1991; Campbell 1993).

There are two oft-used citations in the literature that describe catalytic

systems for arriving at small, laboratory quantities of SPS and partially substituted

SPS: methylaluminoxane (MAO) co-catalyzing either tetrabenzyl titanium (TBT)

(Chien 1991), or titanium cyclopentadienyl trichloride (TCCT) (Grassi 1989);

both utilize toluene reaction diluent* and operating temperatures of about 50ºC.

* The term diluent is used in lieu of solvent as the polymerization reaction proceeds
heterogeneously under most conditions. Some authors use solvent and diluent seemingly
interchangeably, but this work’s own experimental observations warrant the use of diluent as all
polymerization reactions (successful, that is) did in fact precipitate and/or gel polymer very
quickly (order minutes) yet the reaction continued.

5
Inspection of the available information yielded different reasons for using each

catalytic system, of which will be discussed further in Chapter 3.

The MAO/TBT system was investigated first because it reportedly

produces SPS of very high syndiospecificity, i.e. low meso diad chain content

(short sequences of atactic polymer). Due to unforeseen difficulties in

synthesizing, purifying, and utilizing TBT catalyst however, commercially

available TCCT catalyst provided a much easier means for producing syndiotactic
polymer that was of still 99+% syndiospecific content. TCCT also offered higher

conversion of monomer to syndiotactic polymer under similar reaction conditions.

After polymer samples were either obtained or synthesized, several

analyses were performed on them. Differential scanning calorimetry (DSC) was

used to determine the polymer’s glass transition, melting, and subsequent

crystallization temperature(s), and the corresponding enthalpies of these

processes.

Nuclear magnetic resonance (NMR) spectroscopy, namely 13C-NMR, was

used to check the stereoregularity of each polymer sample. Nearly all of the
polymers tested generated very sharp specific peaks in the indicative range,

always occurring farthest downfield at around 147-148 ppm. A sharp, narrow

peak here for homopolymer is an excellent indicator of the sample being

stereoregular (Kawamura 1979).

Gel permeation chromatography (GPC) was used to determine the

molecular weights and distributions of the laboratory synthesized and

commercially obtained polymer samples. Since our facilities were not capable of

6
the high temperature di- or tri-chlorinated benzene carrier solvent analysis as is

typical with these polymers (Ishihara 1988), polymer was analyzed by standard

room temperature THF carrier solvent techniques. The only issue was getting the

polymer into solution and tested before a gel had formed, without any

fractionation occurring via molecular weight separation. The procedure used to

accomplish this unconventional GPC work. Supporting calculations showing it is

at least feasible to accomplish this are presented in Chapter 4.


Making a gel required dissolving the polymeric samples in a hydrocarbon

solution, then allowing this solution to cool. Since the polymer is highly

crystalline with a correspondingly high melting point, it proved impossible to

dissolve the polymer in any liquid at room temperature. For all systems, every

dissolving operation involved boiling (or very nearly so boiling) the liquid to get

the polymer to dissolve, always accompanied by rigorous mixing.

Syndiotactic polystyrene is soluble in aromatic liquids just as regular

atactic polystyrene is at high temperatures. Getting it into solution requires

virtually boiling the solvent and rigorous mixing due the polymer’s solvent
resistance via it being so crystalline (in the annealed state). A process was

developed to speed dissolution of polymer into solution, minimizing the time the

system is at elevated temperature. This was particularly useful when dissolving

polymer in monomer liquids, which are inherently prone to self-polymerization,

particularly at elevated temperatures. Additional amounts of appropriate

polymerization inhibitors were used (~10x above those normally present) to

minimize unwanted polymer from forming. The introduction of additional

7
inhibitor never exceeded 0.03 wt% (300 ppm) of the monomer, and the addition

of this component did not observably affect the measured thermal or mechanical

properties of the solutions or gels.

Boiling solvents such as toluene and xylene posed little problem to the

overall system, save maintaining a consistent mass balance for knowledge of the

actual system after boiling. Reactive monomers such as styrene and tertiary butyl

styrene can self-polymerize rapidly at elevated temperatures however, so slightly


different procedures were followed in order to get the desired polymer solution.

Thermal analysis with DSC was done on the newly formed gel samples in

order to determine the different transition temperatures each solvent/polymer

combination exhibits going from a gel to a sol and back again. The

ethylbenzene/SPS system was extensively studied using this technique in order to

obtain a preliminary phase diagram.

Knowledge of the gel’s thermal behavior is required to be able to use the

material effectively for solids suspension, as it was found in practice that a

mixture’s rheological and heat history were dependent on each other. Gel
formation must occur rapidly and produce good network formation while

minimizing any solid separation that might occur before the gel forms. In the

paste regime (50-80 wt% solids) mixing could be stopped and if the gel formed in

10 minutes or less negligible separation would occur, leading to a homogeneous

filled paste product. Less filled systems of course required gelation to take effect

faster in order to minimize solid settling out while the gel formed.

8
After the pure polymer samples were characterized, polymer-solvent

systems were made for further testing. Ranges of tests were performed to

determine the rheological response of mixtures with varying amounts and nature

of incorporated polymer additive. It was also desired to compare the minimum

agent concentration necessary to support the filler particles versus forming a gel

in an unfilled polymer/solvent system. Since the particles are of rather large size

the critical polymer concentration for gel formation was not expected to be
dependent on the presence of solid filler particles. Making gels is described in

Chapter 5, while characterization of the gels formed is detailed in Chapters 6 and

7.

Typically only low-loaded (<5 wt%) samples were analyzed for a few

reasons. One was due to our in-house equipment limitations and the fact that

other studies on low loaded systems (especially those below 1 wt%) are far fewer

than those on more concentrated systems. All samples were made by dissolving a

known amount of polymer in solvent in a vessel of known mass by boiling for

several seconds while rigorously shaking, then quickly cooling to room


temperature. Actual weight fractions were determined by weighing the vessel

with contents just prior to withdrawing sample, if that was 10 minutes or 10 days

after the sample was made. This was done to minimize the error associated with

solvent loss during solution preparation and/or storage. Samples were of

appropriate size to minimize sample waste and error, usually about 2 mL total

volume in a 12 mL test tube. Error associated with solvent-loss using this

technique were found to be quite low, less than a relative error of 0.5%.

9
Chapter 2 introduces a definition for the property and/or condition of

physical gelation, as well as corresponding relations to model the behavior.

Preliminary results forming a foundation for this research project are also

presented.

At the simplest level gels were characterized by simply noting whether a

gel formed or not at room temperature. If this elementary test indicated a

formulation produced any gelation behavior, it could then be subject to various


studies. Those most employed were DSC again to determine the system gel-sol

transition temperature and then sol-gel transition.

Gels were also extensively studied using a dynamic viscoelastic rheometer

with cone and plate measuring geometry. Very small oscillatory shear strains

were applied (0.1% to ~3%) over a range of low frequencies (0.5 to ~100 Hz) to

monitor the gel as it formed. This measurement was then continued as the gel

aged after the setpoint temperature of the sample and measuring system was

stabilized. This study produced unexpected and previously undocumented

results: a dilute gel of SPS polymer and “good” solvent, such as ethyl
benzene, exhibits a maximum in modulus at its incipient of formation (initial

formation, nascent state). The modulus then tapers off to a fraction (ranging from

5-80%) of its maximum level as the gel ages. It was not directly determined what

is causing this “self-weakening” of the gel as it ages, but it was found not to be

due to thermal gradients during measurement, response time issues in the

instrument, or to the continuous oscillatory mechanical disturbance gels were

subject to during formation and/or measurement of its modulus. These results are

10
documented in Chapter 7. The endotherm upon melting a formed gel does appear

to get larger with longer age times, which appears to extend to order days with

room-temperature aging. This is briefly reported in Chapter 6.

Another interesting feature is that physical networks have been realized

during this work at concentrations far below those that would normally be

associated with a dilute network formed by solute associations. This is subject to

the currently dubious definition that is still tossed back and forth by experts
(Guenet 1992) as to what really constitutes a true physical network in reference to

physical solvent induced gels. The working definition is based on the

requirement that a minimum physical strength be realized that is capable of

supporting its own weight in a nominally sized container (for instance, a 25 mm

test tube).

1.4 CONCLUSION

Many different aspects of stereoregular polymer preparation and

purification were explored and reported in this work. The most daunting task in

making polymers of these types is establishing the proper catalyst ratios as well as

maintaining the proper reaction environments, namely devoid of all traces of


water, oxygen, and any other proton-donating contaminants that will readily

deactivate the catalytic system.

Several ways to characterize physical gels were explored. The most

successful involved measuring the thermal properties of the materials via DSC.

DSC thermal scans of pure polymer demonstrated the large heat-history

dependence on the polymer’s degree of crystallinity, and the temperature at which

11
melting and crystallization occurred. DSC studies of gels provided preliminary

information that allowed the generation of partial phase diagrams to compare the

measured gel-sol transition temperatures to those predicted via theoretical means.

Agreement with the selected model (Flory) was not excellent for all systems

under study, but the model provided a good starting point to work from.

The viscoelastic studies done on cast gel sample between a cone and plate

were very interesting. Oscillatory measurements, depending on time and


temperature profiles applied to the materials, produced some surprising results. A

major problem in testing these gels with a cone and plate resulted from the

volume change that accompanied the sol-gel transition. The samples shrink by an

estimated 2%, which resulted in normal forces within the sample between the

cone and plate measuring surfaces that varied with radial position. Sometimes the

force was so great and/or uneven (compression in the middle, tension at the

edges) that the cone face would disengage from the gelled sample. This

occurrence was obvious in the instrument as the sudden reduction in torque

caused the cone to make a complete revolution as the controller struggled to


compensate. Hence, samples of 2 wt% or less were studied in order to minimize

this effect.

12
REFERENCES
Berezniak, A.; Zimoch, T., Biul. Wojsk. Akad. Tech. Prace. Chem., 7, 62, 1958

Berghams, H.; Donkers, A.; Frenay, L.; Stoks, W.; De Schryver, F.; Moldenaers,
P.; Mewis, J., Polymer, 28, 97, 1987

Campbell, R.; Hefner, J., US Pat. 5,196,490, 1993

Chien, J.; Salajka, Z., J. Poly. Sci., Part A: Poly. Chem., 29, 1243, 1991

Grassi, A.; Pellechia, C.; Longo, P.; Zambelli, A., Gazz. Chim. Ital., 19, 2465,
1987

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Imabayashi, H.; Ishikawa, K.; Yamamoto, K.; Izumi, T., US Pat. 5,037,907, 1991

Ishihara, N.; Seimiya, Y.; Kuramoto, M.; Uoi, M., Macromolecules, 19, 2464,
1986

Ishihara, N.; Kuramoto, M.; Uoi, M., Macromolecules, 21, 3356, 1988

Kawamura, T.; Uryu, T.; Matsuzaki, K., Makromolecular Chem., 180, 2001, 1979

Narh, K.; Barham, P.; Keller, A., Macromolecules, 15, 464, 1982

Pellechia, C.; Longo, P.; Grassi, A.; Ammendola, P.; Zambelli, A., Makromol.
Chem. Rapid Commun., 8, 277, 1987

Pezron, E.; Ricard, A.; Lafuma, F.; Audebert, R., Macromolecules, 21, 1121,
1988

Zucchini, U.; Albizzati, E.; Giannini, U., J. Organometal. Chem., 26, 357, 1971

13
Chapter 2: Background

2.1 INTRODUCTION

Traditionally the term gel implies the condition of a covalently bonded

chemical system: i.e. a specific reaction extent or coordinate in a polymerizing

system where an infinite molecular weight is achieved, and/or the system reaches

an inflowable and unmeltable state. However, in this work we studied gels that

are manifest without chemical bonds. Instead, it is the physical association of like

molecular species into an effectively infinite network that causes the gel to arise.

Evidence suggests (Guenet 1992) that this association is exclusively physical in

nature.

A thermoreversible gel is defined as a physically associative condition in

which the gel can be formed, melted, and reformed, any number of times. A

physical gel is produced by interactions between molecules or molecular groups

that are limited to Van der Waals (secondary bonds) or solvation effects (also

termed solvent-induced crystallization effects). Some even argue that some sort

of “compound formation” exists for certain polymer/solvent combinations

(Nijenhuis 1996). This is opposed to a chemical gel, which manifests through the

formation of irreversible covalent bonds, with few exceptions (Pezron 1988).

The properties of the polymers that may lead to gel formation are of

particular interest in this study. Initially this work focused on finding a

combination of materials that would exhibit the properties of a physical gel.

When such a composition was identified, the work then proceeded to the

14
characterization of the combinations of materials providing for that behavior.

This led to the study of highly syndiotactic polystyrene (SPS), copolymers of

SPS, and varying molecular weights of these polymers. Compared to ordinary

atactic polystyrene (APS), SPS has been available in quantities suitable for

research use for a rather short period (~12 years). Its cost of manufacture is now

such that commercial use has only recently been realized, mostly for glass fiber

filled heat and solvent resistant parts that take advantage of its extraordinarily
high melting temperature of 270ºC (Dow Questra).

In the present work, syndiotactic polystyrene was synthesized via two

different catalytic systems to provide a range of molecular weights and

polydispersities for study. The gel systems studied use low concentrations of

polymer (<5 wt%) compared to the work of others (Prasad 1990; Vittoria 1990;

Guenet 1995) that focus on these types of systems but at rather high polymer

loadings (>10 wt%).

2.2 SOLUTION/GEL THERMODYNAMICS

The gelling phenomena observed in this work arise from either very

localized polymer crystallization in solution, or through solvent-induced


crystallization. In solvent induced crystallization the solvent molecule is thought

to be very determining in how a specific system behaves, playing a role in the

polymer association (winding, folding, etc.) at the molecular level. Neutron

scattering experiments (both SANS and WAXS) verify that solvent molecules

intercalate between the pendent phenyl groups during polymer chain winding into

the T2G2 double helix configuration which is a specific assignment of the winding

15
chain pair (Guenet ibid.; Nijenhuis 1997). Each of these mechanisms has

previously been observed to manifest the gelled state quite differently, although

the end phenomenon exhibits the same macro-scale physical properties.

2.2.1 Definition of a Physical Gel

One investigator, Dr. Jean-Michel Guenet, has been prominent in studying

and reporting the behavior of SPS/solvent gelled systems. Debate has raged in the
literature as to how to define the physical gelled state and how to test that

proposed condition. In 1992 Guenet presented the most recent and generally

accepted definition that encompasses physical gel systems:

The simplest definition of a gel, which is supposed to behave as a solid, is

the existence of an elastic modulus at zero frequency (or infinite time). In

other words, over a long period of time, the system should show a constant

stress at constant deformation, or maintain a constant deformation at

constant stress.

He then immediately disclaimed this statement with previous observations

of his own (Guenet 1986) stating this rule may not always in fact apply to

reversible physical gels, given that sometimes the physical bonds are sufficiently
weak (of order kT) that they can actually break and reform, in equilibrium,

thereby reducing the actual strain on the stressed material, while still remaining a

“solid.” Due to this possibility, several criteria are therefore proposed to both

more extensively define and test for the condition of the existence of a physical

gel. Summarized here as given by Guenet (1994), four criteria are put forth:

16
1. A gel is first defined as a large system of lines, tubes, and wires, that cross

one another or are connected with one another (Flory 1947). Expanding

this to the macromolecular realm, the gel mesh size must be far larger than

the cross-section of the connecting objects. This therefore requires

actually determining the gel structure, which is not nearly as simple as

making a mechanical measurement. Determining the gel structure is

regularly done via infrared FTIR analysis or neutron scattering.


2. The second criterion specifies the phase transition that takes place when

either the gel melts (gel-sol) or forms (sol-gel). For a reversible physical

gel the transition of a gel-junction is first order as a function of

temperature. This implies that organized junctions are formed during the

sol-gel process.

3. Third, a gelled sample below the sol-gel transition temperature, when

immersed in excess preparation (matrix) solvent, will not disaggregate

under quiescent conditions. This is a simple test to perform, provided that

in the case of a very low physical strength gel, the immersion process does
not mechanically disturb the sample. The main issue being the

introduction of excess liquid may be disturbing enough to mechanically

mix the gel, thereby breaking it into smaller pieces and making a visual

determination of the gel sorbing solvent difficult.

4. Lastly, a gelled sample may be handled (the container moved or

accelerated) without irreversible damage being inflicted to the gelled

system, even at the molecular scale. This criteria is implicit for a purely

17
elastic network, yet obviously cannot be upheld as the limits of critical

concentration and/or transition temperature are approached. A minimum

concentration is historically and conveniently tested for via allowing a

sample to gel in a test tube, then simply tilting or inverting the test tube.

Invariably, this resulted in false-negatives due to weak systems (those

weak in crosslink concentration or very near transition temperature).

These may have very well been gels, but still exhibited slumping under
their own weight and flowing. Although not definitive it is a useful and

thrifty test for preliminary mechanical properties: i.e. is it a ‘strong’ or

‘weak’ gel? This test invariably includes materials of Bingham behavior

as well, so again one must use care in applying these results in an absolute

fashion.

2.2.2 Theory of Gelation

The systems in this work can be described as gels of the solvent-induced

type according to previous work (Guenet ibid 1992). Solvent-induced systems are

defined when the solvent plays a role in the formation of polymer-polymer

associations and remains as an integral part of the “compound” after network


formation. When the specific solvent plays a role in network formation, even a

constant chi interaction parameter can yield very different results for gel

formation, as then the specific solvent plays an integral part in interacting with the

polymer chain functionality. Meaning, solvent/polymer combinations, which

would regularly be treated as thermodynamically equivalent are in fact not, due to

other unaccounted for interactions of the polymer and solvents of like energy and

18
chemical potential. This activity is thought to be very dependent on the physical

geometry of the solvent, as it is in effect occupying or filling the cavities or holes

created by the pendent functional groups protruding from the polymer backbone

(Guenet ibid 1992). This response was tested for by using several different

aromatic solvents in our study.

2.2.2.1 Crystallization from Solution

Syndiotactic polystyrene is chemically identical to atactic polystyrene.

Both have essentially the same glass transition temperature, yet SPS melts at a far

higher temperature than they atactic version. Since all the solvents (and

monomers) used to make gels in this work could be termed excellent solvents for

atactic polystyrene (toluene, xylene, ethylbenzene), this reasonably eliminates

phase-separation as a possible mechanism leading to this type of gel formation.

Being that the polymers are chemically the same, crystallization must therefore

play some kind of role in gel formation. Therefore Flory’s theory of melting

point depression for solution-born crystals is employed as a means to model this

behavior.

To describe the melting point depression of a polymer by the presence of a

solvent, Flory (1942) started with an expression for the entropy of mixing of

polymer and solvent as a function of the component volume fractions

∆S M = −k ( n1 ln v1 + n2 ln v2 ) . (2.1)

The enthalpy of mixing is given by

19
∆H M = kTχn1v2 (2.2)

which introduces the dimensionless Chi parameter characterizing the

interaction energy per solvent molecule, divided by kT. This term is a sum of two

contributing components

χ = χH + χS (2.3)

where the entropic portion is empirically taken for nonpolar systems to be

0.34 (Scott 1949; Blanks 1964). The enthalpic component is related to the

Hildebrand (solubility, δ ) parameters as

V1
χH = (δ 1 − δ 2 ) 2 (2.4)
RT

to yield a new version of Equation 2.3

V1
χ= (δ 1 − δ 2 ) 2 + 0.34 (2.5)
RT

We apply more theory as derived by Flory (ibid.) in 1953. Using the

expression for free energy of mixing

∆GM = ∆H M − T∆S M (2.6)

which is then substituted with Equations 2.1 and 2.2

∆GM = kTχn1v2 − T ( − k ( n1 ln v1 + n2 ln v2 )) (2.7)

or simplified

∆GM = kT ( n1 ln v1 + n2 ln v2 + χn1v2 ) . (2.8)

20
Differentiating Equation 2.8 with respect to n1 and multiplied by

Avogadro’s number yields the chemical potential per mole of solvent

µ1 − µ10 = RT (ln(1 − v2 ) + (1 − 1 / x )v2 + χv22 ) . (2.9)

For polymers with distributed molecular weights, x is replaced with xn ,

the number average degree of polymerization

µ1 − µ10 = RT (ln(1 − v2 ) + (1 − 1 / xn )v2 + χv22 ) . (2.10)

Assuming large x (or xn ) for a high polymer (component 2) Equation

2.10 then becomes

µ1 − µ10 = RT (ln(1 − v2 ) + v2 + χv22 ) . (2.11)

Now differentiating Equation 2.8 with respect to n2 , times NA, and

performing like operations, the chemical potential of the polymer unit becomes

µ2 − µ20 = − RT ( v2 / v1 )((1 − v2 ) − χ (1 −v 2 ) 2 ) . (2.12)

Equilibrium between crystalline polymer and the polymer units in solution

is defined by making their chemical potentials equal

µ 2cry − µ 20 = µ 2 − µ 20 . (2.13)

Whereas the left hand side of Equation 2.13 represents the negative of the

free energy of fusion for pure polymer

µ 2cry − µ 20 = − ∆G2, fusion = −(∆H fusion − T∆S fusion ) (2.14)

which is preferably expressed as

21
µ 2cry − µ 20 = − ∆H fusion (1 − T / Tmelt
0
) (2.15)

assuming a constant ratio of ∆H 20 / ∆S 20 over the temperature range at

0 0
issue (T to Tmelt ) and that ratio is equal to Tmelt since

µ 2cry − µ 20 = 0 (2.16)

0
at the pure polymer’s melting temperature Tmelt .

Substituting for the left and right hand sides of Equation 2.13 with

Equations 2.12 and 2.15, and replacing T by Tm (now the dilute crystal’s melting

temperature), the equilibrium melting temperature for the mixture is now

1 1 RV2 (v1 − χv12 )


− = (2.17)
Tm , 2 Tm0, 2 ∆H 20V1

where: Tm, 2 is actual melting temperature of the crystals in solution;

Tm0, 2 is the pure polymer melting temperature;

∆H 20 is heat of fusion per mole of perfect polymer crystals; and

V1 , V2 are the respective molar [unit] volumes.

Solving for the gel’s melting temperature gives the desired melting point

of dilute crystals as

1
Tm, 2 = (2.18)
1 RV2 (v1 − χv12 )
+
Tm0, 2 ∆H 20V1

22
and recalling Equation 2.5 to make the Chi parameter a function of the

polymer and solvent via Hildebrand parameters, we obtain the predicted melting

point of dilute crystals (or gel) as a function of known variables

1
Tm = . (2.19)
1 V R  V 
+ 2 v − v12  1 (δ 1 − δ 2 )2 + 0.34  
0  1
Tm , 2 V1 ∆H 2 
0
 RT 

Equation 2.19 technically requires iteration to solve, as the Chi parameter

is a function of temperature. Since Chi is relatively insensitive to temperature in

this application due to the domination of the entropic contribution term, a

reasonable guess (within 30º) of the temperature on the RHS of the equation will

be very close to the actual value. This model was compared to preliminary

laboratory data and found to be fairly representative of the transition temperatures

observed for the system selected. Figure 2.1 illustrates the result when Equation

2.19 is invoked for the SPS/xylene system, with a corresponding infinite-dilution

melting temperature of about 110ºC. Parameters used in the calculation include:

δ 1 = 8.8 (cal/cc)½, δ 2 = 9.2 (cal/cc)½, V1 = 122 cc/mol, V2 = 99 cc/mol (repeat

units), and molar heat of fusion ( ∆H 20 ) for perfect polymer crystals of 1,322

cal/mol (repeat units).

23
280

260

240 Calculated Melting


Point
220
Melting Point (ºC)

200

180

160

140

120

100
0 0.2 0.4 0.6 0.8 1
Polymer Volume Fraction

Figure 2.1: Calculated melting point of SPS/xylene system. Using Equation


2.19, melting point depression calculation.

Taking a closer look at the dilute regime of this phase diagram, laboratory

data is shown compared to the predicted melting point curve. Figure 2.2 shows

transition in the very dilute polymer-loading regime of about 104ºC, aligning

fairly well with the predicted melting point depression calculation. The trend

follows essentially the same slope, yet is depressed several degrees. These initial

measurements were done with the simple yet not very exact falling-ball test. The

measuring instrument consisted of a 4 mm steel bearing resting on the gelled

24
sample’s surface inside a 100 mL test tube, and heated in a water bath. The ball

diameter to test tube diameter was such that the ball was free to move according

to local gel/sol conditions. This elementary measurement technique for

determining the gel-sol transition point was not exactly representative of the gel

point. Transition point determination was greatly improved upon using the DSC

to measure transitions via the endothermic response of the gel melting to sol.
118

116 Flory Predicted Melting


Falling Ball Data
114

112
Melting Point (ºC)

110

108

106

104 (linear fit)

102

100
0 0.02 0.04 0.06 0.08 0.1
Polymer Volume Fraction

Figure 2.2: Experimental Data with Melting Point Depression Calculation for
SPS/xylene system. Points are preliminary laboratory data using
falling ball indicator.

25
2.3 RHEOLOGY OF GELS

This work also was done to ascertain some important mechanical

properties of the gels formed. Although these gels are “solid” upon inspection,

the investigation of their properties is best approached by treatment as a

rheological study. Herein the elastic (spring) behavior dominates, compared to

the viscous (damper) behavior of the material, except at the sol-gel transition.

What enters the discussion again is the very definition of what a physical

thermally reversible gel is, and further, when does it actually go from a sol to a

gel, and back again. Heating what is already obviously a gel (in that it is solid,

rebounds under slight strain, etc) at nearly any temperature increase rate (up to

100ºC/min typ) will give a constant gel-sol transition temperature, accompanied

by a drastic reduction in system viscosity. This massive transformation from a

very elastic to a highly flowable state is easily observed via many means, some

examples include direct observation, falling ball, DSC, and light transmission.

When one desires to measure when/where exactly the transformation from

a low viscosity sol to a gel takes place, the experiment becomes much more

involved. Obviously if you subjected the sol to continuous mixing while cooling

it in hope of catching a drastic increase in viscosity, the test would fail as the

mixing obviously will interfere with (at least) the very initial formation of gel,

especially in dilute and/or weak systems, and at best you will be observing a

26
highly fractured granular material that had already gelled most of the way before

you could measure anything meaningful.

In order to avoid the problems of continuously disturbing the material

while observing its mechanical properties, a minute oscillatory disturbance is

introduced setting up the requirement of a measured oscillatory modulus

G * = G ′(ω ) + iG ′′(ω ) (2.20)

which is also equal to

G * = G (ω )(sin δ + i cos δ ) (2.21)

where δ is known as the loss angle.

G * is a complex number, referred to as the complex modulus. Equation

2.20 has two components, the elastic part and viscous part. The elastic part,

called the storage modulus, is represented by the G ′ term, and refers to the in-

phase resistance measured to an applied strain. The viscous part, called the loss

modulus, is represented by the G ′′ term, and refers to the out-of-phase resistance

measured to an applied strain.

As examples of what these terms represent by themselves, two different

cases are put forth. Materials that are obviously liquid, such as water, or a dilute

sol system above its melting temperature, the viscous term dominates, to the point

that the elastic term is non-existent. On the other extreme, tire rubber, being

chemically crosslinked, under strain should exhibit a purely elastic response, as

the viscous component will vanish, as the material does not flow at all. Gels are

27
somewhere in between these two examples, especially as the gel is forming from

the liquid sol state.

2.3.1 Cone and Plate Theory

In this work the sol-gel transition is measured, mechanically, with a cone

and plate viscometer. The theory of this measurement method and geometry is

explained here, with detailed experimental accounts appearing in Chapter 7.

A cone and plate geometry is particularly well-suited for studying the sol-

gel transition and resulting gel physical properties due to ease of which material

can be cast between the surfaces and the constant shear strain throughout the

sample. This as opposed to the varying shear stress resulting from parallel plate

geometry, another widely used geometry for rheological measurements. The only

“complication” with the cone and plate is that spherical coordinates must be used

to accurately and conveniently describe the motion.

Mooney and Ewart first proposed this geometry in 1934. After proper

consideration, the equations of motion for this measurement system reduce to (as

given by Macosko, 1994):

Radius, r -component

ρvφ2 1 ∂ 2 τ +τ
= ( r τ rr ) − θθ φφ (2.22)
r r ∂r
2
r

Horizon, θ -component

28
1 ∂ (τ θθ sin θ ) cot θτ θθ
0= − (2.23)
r sin θ ∂θ r

Azimuth, φ -component

1 ∂τ θφ 2
0= + cot θτ θφ . (2.24)
r ∂θ r

The shear stress is found by integrating Equation 2.24

C1
τ φθ = . (2.25)
sin 2 θ

Performing a torque balance on the plate, or the bottom surface (not the

cone, which “provides” the torque)

2π R 2πR 3
M =∫ ∫ r 2τ φθ drdφ = τ φθ . (2.26)
0 0 3

Substituting into Equation 2.26 from Equation 2.25 at the boundary

τ φθ |π / 2 = C1 = τ φθ (θ ) sin 2 θ (2.27)

to obtain

3M
τ φθ (θ ) = . (2.28)
2πR 3 sin 2 θ

Since the cone angle is so small, as in our instrument’s case (only 1º), and

shear stress is constant throughout the material it can be assumed

sin 2 (89º ) ≅ 1 (2.29)

so

3M
τ 12 = τ φθ = (2.30)
2πR 3

29
which is independent of radial position; shear stress is dependent only on

shear rate or strain.

Equation 2.30 applies for continuous rotation of the cone, used in

measuring the viscosity of flowable materials. In this work a very small

oscillatory strain is used to measure the gel forming and the gel after it has

formed. The applied strain is therefore

γ = γ 0 sin ωt (2.31)

and the shear stress

τ = τ 0 sin(ωt + δ ) (2.32)

reintroducing the loss angle previously shown in Equation 2.21.

Under oscillation the measured stress is decomposed into two waves of the

same frequency, but perfectly out of phase, conveniently done with trigonometry

τ = τ ′ + τ ′′ = τ 0′ sin ωt + τ 0′′ cos ωt (2.33)

Recalling Equation 2.21, it can be shown that

τ 0′′
tan δ = (2.34)
τ 0′

and the decomposition now suggest two (previously described) dynamic

moduli

τ 0′
G′ = , (2.35)
γ0

30
τ 0′′
G ′′ = , (2.36)
γ0

and the tangent of the loss angle

G′′
tan δ = . (2.37)
G′

Taking all this together, the final shear stress as a function of angular

frequency, peak strain, and the storage and loss modulii

τ = G′γ 0 sin ωt + G′′γ 0 cos ωt . (2.38)

2.4 SYNDIOTACTIC POLYSTYRENE

The literature (Berghams 1987; Daniel 1996) reported that stereoregular

polymers of styrene and methyl methacrylate were excellent physical gel formers

in many organic solvents, and that many of these gels are thermoreversible. It

was thought that this thermoreversible property could be harnessed for the filled

paste product and commercial process previously discussed, especially if a

transition temperature could be realized around 60-70ºC, as the final processing

temperature was ~80ºC.

Initially isotactic polystyrene (IPS) was tested as a possible gelling agent.

Itagaki (1997) reported temperature-sensitive gelling behavior at low polymer

loadings (<2 wt%) in the non-aromatic solvent decalin. To test this behavior in an

aromatic solvent, we dissolved 2 wt% of powdered iPS from Aldrich (MW=200k)

in TBS monomer by heating to near boiling (TBS b.p. 220ºC) with rigorous

agitation. This polymer was found to be slow to dissolve even at elevated

31
temperatures, most likely due to it being quite crystalline. After a few minutes

mixing at about 220ºC, the polymer was eventually dissolved (by inspection) so

the solution was then allowed to cool to room temperature. A solid material

obviously formed before the material reached room temperature, and it was

allowed to age at room temperature for several hours after that before disturbing.

The resultant material resembled a wax. It was nearly opaque (barely

translucent), yet easily fractured if strained. This material was then slowly heated
to 100ºC in a boiling water bath in attempt to melt it. It refused to melt, so

therefore was deemed unsuitable for our particular commercial process; we

required a transition of no greater than 70ºC.

Dow Chemical Company recently started producing syndiotactic

polystyrene (SPS). This was discovered about the same time information (Daniel

ibid. 1996) was found indicating that this stereoregular polymer, in addition to

being relatively “new” to the research world, would form thermoreversible gels in

benzene. This appeared far more attractive that the previous IPS/decalin

information, because benzene has a boiling point below that of the desired
transition temperature, and the fact that aromatic solvents were cited instead of

aliphatic ones (Prasad 1990) (since the monomers under study are aromatic).

Accompanied by low polymer loading requirements and lower gel-breaking

temperatures than IPS systems, this sounded perfect for our application. The

initial citations spoke of gels being formed and investigated from rather rich

systems (>50%) to very low levels (~1 wt%).

32
Dow provided us with a sample of pure SPS that they manufacture under

the trade name Questra (sample QA101, Mn ~45k, Mw ~250k). They also

supplied a crystalline copolymer of styrene (93%) and paramethylstyrene

(balance), hereafter referred to as CPS (Questra MA405, Mn ~40k, Mw ~200k).

The extent of randomness or short-range stereoregularity was not completely

resolved but laboratory analysis via DSC and NMR indicates that it is still highly

stereoregular even while containing the paramethylstyrene fraction.


Both the homopolymer material (SPS) and copolymer (CPS) produced

gels with excellent properties for the JemPac product, especially the CPS

syndiotactic copolymer. Less than one percent by weight (based on total

monomer content of the product) of the agent was required to make a gel strong

enough to suspend the filler material indefinitely. In addition, the gels broke at

much lower temperatures than those made from IPS. In TBS monomer the gel

melted at 85ºC for SPS and 62ºC for CPS. As the SPS/TBS gel aged at room

temperature in a closed container it exhibited slight syneresis, which was the

gelled network weeping or exuding a tiny amount of solvent, estimated to be


about 0.1 wt% (maximum) of the total solvent present.

2.4.1 Syndiotactic Copolymer

Since the CPS produced a gel that melted around 62ºC, which was ideal

for our application, further study was done with this polymer gelling agent

additive in several different aromatic solvents. During this round of preliminary

tests, it was found that CPS dissolved far more easily and quickly than SPS in like

solvents and conditions. This is probably due to the polymer being less

33
crystalline than pure SPS, verified via DSC; thermal analyses of these polymers

appear in Chapter 4. Table 2.1 summarizes these initial results of 1 wt% CPS in

various solvents. The test used for this information was the falling-ball

viscometer test, described earlier, wherein the gel supported the ball as it was

slowly heated at about 5ºC/min. When the ball fell at least its diameter, the bath

temperature was noted as the break temperature.

Break Temperature Solubility Parameter


Aromatic Solvent
ºC (cal/cc)1/2
Toluene 94 8.9
Xylene 85 8.8
tert-butyl Styrene 62 8.3 (calc)
divinyl Benzene 53 8.56 (calc)

Table 2.1: Gel-sol transition temperatures of CPS copolymer in various


solvents. Calculated solubility parameters using the methods and
tables of Small (1953)

Since they are provided in the industry-standard 1/8” cylindrical pellets,

dissolving some of the polymers as they were provided was problematic. Cryo-

grinding was used as one means to reduce the particle size (therefore increasing

surface area), although that was messy and is expensive. High temperatures are

required to dissolve these polymers in any event due to their high degree of

crystallinity; this can be detrimental to the properties of the monomer solvents

(causing them to partially polymerize prematurely). Consequently, we sought to

make a high surface area material that would provide for rapid dissolution. This

process was so successful in leading to a high-quality gel with minimum impact

on the final product that it was submitted for patent review through the University
34
of Texas at Austin Office of Technology Licensing and Intellectual Property

(UTA# A-024 BAR). This process is described in detail in Chapter 5.

2.5 CONCLUSION

Excellent solid-suspension properties were found using syndiotactic

homopolymers and copolymers of styrene and substituted styrene. This

phenomenon solved a major problem that was a roadblock in the development of


a usable prepolymer product for encapsulating microchips. In addition, this

gelling property provided an excellent place to start performing research in an

effort to better understand and describe its behavior.

This work aims at furthering the understanding of polystyrene-based gels.

Although our particular application remains in the incubation stage, the results of

this work can easily be extended for many other applications in industry.

35
REFERENCES
Berghams, H.; Donkers, A.;Frenay, L.; Stoks, W.; DeSchryver, F.; Moldenaers,
P.; Mewis, J., Polymer, 28, 97, 1987

Blanks, R.; Prausnitz, J., Ind. Eng. Chem. Fund., 3(1), 1, 1964

Daniel, Ch.; Deluca, M.; Guenet, J.; Brulet, A.; Menelle, A., Polymer, 37, 1273,
1996

Flory, P., J. Am. Chem. Soc., 63, 3096, 1941

Flory, P., J. Chem. Phys., 10, 51, 1942

Flory, P., J. Am. Chem. Soc., 69, 30, 1947

Flory, P., Principles of Polymer Chemistry, Cornell University Press: Ithaca, New
York, 1953, 1992 (15th Ed.)

Guenet, J.; McKenna, G., J. Polym. Sci., Polym. Phys. Edn., 24, 2499, 1986

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Guenet, J.; Daniel, C.; Dammer, C., Polymer, 35, 4243, 1994

Guenet, J.-M.; Daniel, Ch.; Deluca, M.; Brulet, A.; Menelle, A., Polymer, 37,
1273, 1996

Itagaki, H.; Nakatani, Y., Macromolecules, 30, 7793, 1997

Macosko, C., Rheology: Principles, Measurements, and Applications, VCH


Publishers: New York, 1994

Nijenhuis, K., Thermoreversible Networks, Springer-Verlag: Berlin Heidelberg,


Germany (In English) 1997

Pezron, E.; Ricard, A.; Lafuma, F.; Audebert, R., Macromolecules, 21, 1121,
1988

Prasad, A.; Mandelkern, L., Macromolecules, 23, 5041, 1990

Scott, R.; Magat, M., J. Poly. Sci., 4, 555, 1949

36
Small, P., J. App. Chem., 3, 71, 1953

Stockmayer, W., J. Chem. Phys., 11, 45, 1943

Tanaka, F.; Stockmayer, W., Macromolecules, 27, 3943, 1993

Tanaka, F.; Nishinari, K., Macromolecules, 29, 3625, 1995

Vittoria, V., Polymer Communications, 31, 263, 1990

37
Chapter 3: Polymer Synthesis

3.1 INTRODUCTION

Dow Chemical had provided us with two different stereoregular polymers,

syndiotactic homopolymer of styrene (SPS), and a copolymer of styrene and

paramethylstyrene (referred to herein as CPS). Both of these polymers produced

excellent gelled systems in various aromatic solvents such as toluene and

ethylbenzene (EB). The samples were labeled as having molecular weights of

250k and 200k, respectively. What was further required were polymers of lower

molecular weight to provide for a range of molecular weights to observe how the

solution behavior and mechanical properties of the gels depend on molecular

weight. Preferably, we desired several different molecular weight polymer

samples of low polydispersity.

3.2 CATALYST SELECTION

Making syndiotactic polystyrene in the laboratory is a relatively new

technology (Ishihara 1986; Pellechia 1987; Proto 1988). Isotactic polystyrene

(IPS) had been synthesized and the procedure published (Braun 1960) long before

the syndiotactic version, and attempts were made (Yamada 1961) to vary those

catalytic systems to produce SPS instead of IPS. These attempts were largely

unsuccessful however, usually small fractions of syndiotactic polymer were found

as a by-product from the synthesis of atactic or isotactic polymer via lithium-

based or other alkali-organo anionic polymerization catalytic systems.

38
To date the synthesis of SPS polymer requires very specific binary

catalytic systems consisting of an organometallic compound (metallocene)

combined with a large molar excess of methylaluminoxane (MAO) in toluene

diluent (Ishihara 1988). Many recent citations detail the various degrees of

success (or failure) testing individual combinations of organometallic and MAO

(anywhere from 50 to 2000 moles MAO/mole organometallic). A broadly

accepted mechanism for how the syndiotactic polymerization proceeds will also
be briefly discussed herein.

Dow is evidently using trichlorocyclopentadienyltitanium (TCCT) to

make their high molecular weight SPS (Campbell 1991; Siddall 1996), but they

must certainly have other proprietary chemistry and/or processing technologies in

order to achieve the unusually high molecular weights of 200k or more. Many

literature citations (Longo ibid 1988; Chien 1991) put the maximum molecular

weight at around 45,000 for this chemistry. These higher molecular weights in

the laboratory also meant smaller yields, sometimes only 2 or 3% of original

monomer polymerizing to stereoregular polymer, seemingly independent of the


catalytic system used for the polymerizations. Distributions were also noted in

the previous citations to be at least 2.5 and often exceeding three or more.

Two different catalyst systems were selected by this work for use in

polymerizing styrene to SPS. The first compound was

trichlorocyclopentadienyltitanium (TCCT) as shown in Illustration 3.1a,

commercially available and as such purchased from Aldrich Chemical Company.

This organometallic was noted to be rather efficient at producing syndiotactic

39
polymer under reasonable conditions and to produce higher molecular weights

than many other catalytic systems studied by one group (Grassi 1989). This

compound was used in conjunction with 100 to 1,000 times MAO cocatalyst (by

molar ratio).

(a)

(b)

Ti Cl
Cl Cl Ti

(c)

Al Al Al
O O O n

Illustration 3.1: (a) Trichlorocyclopentadienyltitanium (TCCT) molecule. Dark


yellow powder, 96% purity from Aldrich (b) tetrabenzyltitanium
(TBT), synthesized in-house (dark red), (c) methylaluminoxane
(in solution, n reported to vary from 3 to large, a distribution)

The other catalytic system selected utilized the standard MAO with a

metallocene: tetrabenzyltitanium (TBT) as shown in Illustration 3.1b. TBT is a

red-brown solid that could not be found commercially so was instead synthesized

in the laboratory per Zucchini (1971). The synthesis step required the moderately

low temperature of –20ºC and was simple to perform in a rigorously inert

atmosphere; the separation and purification of the TBT from the other fractions of

by-products and reactant ash proved extremely difficult and time consuming. It

40
took nearly nine months to develop a method for accomplishing this using

conventional laboratory equipment, as described below.

3.3 CATALYST SYNTHESIS

Tetrabenzyltitanium was synthesized according to the recipe and

procedures of Zucchini (ibid 1971) from benzyl magnesium chloride (BMC) and

titanium-IV tetrachloride (TTC). A three-neck flask was fitted with an addition


funnel and Teflon stir bar, and set in an acetone bath to achieve –20ºC. Cooling

was made by the controlled addition of small pieces of dry ice while constantly

monitoring the bath temperature. Illustration 3.2 depicts the synthesis apparatus.

The following equation (Equation 3.1) shows the simple Grignard reaction

leading to TBT product.

TiCl4 + 4 MgCl 4MgCl2 + Ti


4

(3.1)

41
Illustration 3.2: TBT Catalyst Synthesis Reactor System. Bath maintains –20ºC
environment, BMC is charged to 3-neck flask, heptane diluted
titanium tetrachloride held in addition funnel, argon cover gas
dead-ended to system through top of addition funnel. Rubber
septa seal the flask and funnel necks.

BMC was used as provided at 1.0 M in diethyl ether solvent and

introduced into the three-neck flask that had previously been evacuated, flamed,

and inerted with argon gas several times (cycled argon/vacuum). A typical batch

started with 100 mL of BMC solution. TTC (a dense, oily liquid, quarternary

functionality) from Aldrich at 98% was introduced volumetrically to 20:1 v/v of

freshly distilled heptane in the addition funnel, in an amount providing for a

42
stoichiometric reaction with the BMC (unity functionality). The diluted TTC was

slowly introduced dropwise over two hours to the rigorously mixed and cooled

BMC/ether charged reactor. A color change from pale yellow to dark red was

immediately observed. After all the TTC had been introduced the reaction was

stirred an additional two hours under same cold conditions to ensure maximum

conversion.

After the initial four hour step was complete, and while maintained very
cold at –20ºC, the reaction products were transferred via cannula through a fritted

filter into another inert flask using slight vacuum as the driving force. Two

equivalent volumes of freshly distilled ether was then introduced into the initial

reactor to further dissolve any remaining desirable material and transferred

through the filter into the receiving flask as before. The remaining solid material

in the reactor was dark brown to black in color, and cited to be non fully

substituted benzyl titaniums and product that had succumbed to degradation from

the solvent (Zucchini ibid).

The dark red newly filtered liquid in the receiving flask was then agitated
via a previously inserted Teflon stir bar and brought to cold-boil via vacuum.

Heat was provided during this operation by immersing the flask in a tap water

bath, although never being raised above room temperature. When the red residue

was observed to be virtually dry of solvent after several hours, the flask was

equalized with argon and freshly distilled heptane was introduced to redissolve

the desired TBT component.

43
A third flask was then used to receive the heptane rich in desired product

after passing through another fritted filter disc. This dark red liquid (although

lighter in color than the previous red filtered liquid) was then concentrated again

en vacuo, removing about 75% of the solvent, then slowly cooled to -20ºC

without disturbance for 12 hours in a freezer.

Sometimes (and I stress sometimes…) needle-like crystals formed in the

heptane solvent; most often the product crystallized completely from the solvent
if it crystallized at all as the solvent would go clear with red crystals on the

bottom. If crystallized product had indeed formed, while working as quickly yet

as quiescently as possible, the solvent was withdrawn via cannula as completely

as possible, then the solid crystals transferred to a vial from the flask in a helium

atmosphere glovebox with oxygen and water levels both indicated at sub-1.0 ppm

concentrations.

This procedure proved the most difficult work performed during this entire

research endeavor, with numerous encounters with less-than-inert reaction

conditions, equipment that required maintenance or modification, etc. The most


was learned from this phase of the work as well however, and a far greater

appreciation for the special care and techniques that must be employed when

working with highly reactive and pyrophoric organic syntheses.

Even with the solid catalyst material isolated and under inert atmosphere it

was still advisable to use it as quickly as possible due to its dubious chemical

stability. TBT is cited (Zucchini ibid. 1971) as having an infinite shelf life at

room temperature or below, although this author finds that very difficult to

44
believe. Perhaps a longer shelf life (greater than two weeks) could be realized if

the material’s purity was first verified, and if found less than desirably pure,

perform additional recrystallizations to further purify it. For this work purity was

not determined, but assumed to be high after completing the recovery steps. It

was then used in polymerizations within a few days to avoid degradation issues.

Certainly the material should be refrigerated (this capability was not available in

the dry box) and handled under darkness. Identification of specific impurities that
affect its shelf life would assist in providing a benchmark for being able to predict

its shelf life based on measured contaminant levels.

3.4 HOMOPOLYMERIZATIONS

Previous syntheses of SPS (Grassi ibid 1989; Chien ibid 1991) revealed

that low polydispersities (< 2.0) were not attainable. However, that would be

difficult to expect or achieve anyway given the nature of most polymerization

catalyst systems aside from simple anionic polymerizations. Reported

distributions ranged from 2.5 up to a little over three, sometimes more, so this

work would have to keep that in mind when making mechanical property

statements if a function molecular weight. Molecular weight analysis via GPC


would be used to determine the pertinent number average and weight average

molecular weight of each polymer produced in-house.

3.4.1 Polymerizations with Tetrabenzyltitanium

After tetrabenzyltitanium (TBT) had been synthesized and purified in a

usable amount, it was then used to polymerize styrene to syndiotactic polystyrene.

The first reactions were carried out in a three-neck flask held at 60ºC for three

45
hours, then the reaction quenched with 2% HCl in methanol (excess). A typical

TBT amount used in polymerization was 50 mg (110 µmol) in 100 mL of reaction

diluent, and 15 mL (0.13 mol) of freshly distilled styrene (so a 1500:1 monomer

to primary catalyst ratio).

The major hurdle in achieving good results with TBT catalyst was using

the catalyst when it was as fresh as possible. This meant keeping the catalyst out

of the toluene reaction solvent until the MAO had been introduced to scavenge
any remaining impurities from the freshly distilled toluene and not render the

rather small quantity of TBT added useless. Using the MAO as the initial

reaction “cleanser” of course depleted the amount of MAO present, but there was

such a large amount used (10-20 mL at 10 wt% active ingredient) that the small

amount consumed by reacting with trace contaminants was acceptable. Toluene

was purified via vacuum distillation over calcium hydride prior to injection into

the reactor.

Due to the pyrophoric nature of the catalyst, it had to be weighed and

added to a clean, dry, previously flamed flask inside a helium drybox under
rigorously inert conditions. Without very specialized laboratory scale intra-

reactor solids-handling equipment, the next best thing was done: thinly dispersing

the powder on the inside wall of the flask at room temperature to allow for later

introduction of the reaction diluent (toluene) to be done without contacting the

catalyst until after MAO introduction. Static electricity was relied upon for this.

After toluene and MAO introduction, the TBT was then dissolved by gently

46
swirling the flask, bringing the liquid into contact with the catalyst solid.* The

reactor was then brought to the reaction temperature and as the freshly distilled

styrene was introduced, that was noted as time zero for the polymerization

reaction. Styrene for polymerizations was prepared by vacuum distillation from a

larger amount of styrene over calcium hydride. The middle fraction of distillate

was collected and stored inerted, refrigerated, and under darkness until use, but

never used if more than seven days had passed since its distillation.
An inert nitrogen atmosphere was maintained over the reaction mixtures

during the polymerization procedure. The reaction color was initially reddish

brown to black, and after one or more hours of reacting at elevated temperature, it

turned to a black-green color. Co-catalyst ratios were varied from 200 to 5,000,

TBT catalyst to monomer ratios varied from 20 to 200, reaction times between 2

and 24 hours, and temperature from 40 to 80ºC, were varied to achieve (albeit ad

hoc) different molecular weights. Number average molecular weights were

achieved from 3,700 up to around 50,000, yet yields were rather low at about 5%

(which is in agreement with most literature citations).


Tetrabenzyl titanium, although very difficult to isolate and use for

polymerizations, was selected and used for this work due to its reported (Chien

ibid 1991) and evidently highly selective nature for initiating and maintaining

stereoregularity during chain propagation. In other words, it is better at producing

a stereoregular polymer chain with very few non-syndiotactic “defects” or diads.

* As an alternative, the solid catalyst material could have been introduced into a small vessel in the
glovebox and sealed with a septum. Then a small amount of toluene to dissolve it just prior to
injecting the resultant concentrated catalyst solution into the main reactor flask after introduction
of the MAO component.

47
Atactic polymer is produced as well during the polymerizations, presumably due

in part to the elevated temperatures leading to free-radical generation facilitating

regular addition polymerization. This atactic polymer is separated out when the

desired polymer is further processed after initial reaction solvent precipitation via

multiple methyl-ethyl ketone (MEK) Soxlet extractions.

3.4.2 Trichlorocyclopentadienyltitanium Polymerizations

TCCT was selected from among a list (Ishihara 1988) of many reported

successful SPS polymerization catalysts with less-than-negligible stereoregular

polymer yield. This particular species was selected due to its reportedly better

performance in terms of monomer conversion compared to other catalysts by the

same researchers, while maintaining a high degree of stereospecificity. It was

again specified to be used with a large ratio (>500:1) of MAO cocatalyst in order

to work properly. It was also later discovered that this is the catalyst system that

Dow is most likely using to make their commercial syndiotactic polymer product,

although their patents cite the typical huge range of “possible” catalysts that they

might be using (a standard and acceptable patent writing tactic).

Reactants were introduced in the same manner as before, using the TCCT
powder as delivered (98%) from Aldrich in a helium atmosphere glovebox.

Reactions using this catalyst exhibited a very broad spectrum of colors and color

transitions during the reaction, varying from lime green, dark yellow, orange,

brown, dark green, eventually going to nearly black after several hours.

48
3.4.3 Substituted Styrene Polymerizations

A report of substituted styrenes being polymerized to syndiotactic forms

was found, an example of which even noted (Grassi ibid 1989) the production of

syndiotactic poly(tert-butyl styrene) (SPTBS). Since this was the major monomer

component of our commercial product and PTBS has the desired high glass

transition temperature (almost, without crosslinker anyway) for electronics

packaging, perhaps SPTBS would function as an even higher-performance gelling


agent for the filled monomer mixture. A polymerization was performed under the

following conditions with TCCT main catalyst in an attempt to synthesize

SPTBS.

The reagents used for two attempts at making syntactic PTBS were 120

mL of toluene, then the following added to the inerted flask: 27 mg of TCCT (125

µmol), 18 mL of 10 wt% MAO solution in toluene, and finally the introduction of

13 mL (0.071 mol) of freshly distilled TBS monomer while briskly stirring the

reactor at 40ºC. The reaction was maintained at 40ºC and stirred for 24 hours per

the citation.
Purification and analysis of the recovered polymer from the limited

synthesis tests (two of them) indicated any polymer present was most likely

atactic in nature. The polymer showed no detectable crystallization exotherm

with DSC, and NMR of the quarternary ring-carbon attached to the backbone

chain was diffuse. The yield was also very low, so polymer probably only

resulted from thermal polymerization as the reaction was run for such a long time.

This polymer would not form a gel in an assortment of aromatic liquids, even at

49
high (10%) concentrations. Further attempts at performing this reaction might

lead to stereoregular polymer, but achieving this was not a primary goal of this

work so was not pursued. A high glass transition temperature hydrocarbon

gelling polymer would be very desirable however.

3.5 SYNDIOTACTIC POLYMERIZATION MECHANISM

A well-studied mechanism for the syndiotactic polymerization of styrene


to polystyrene will be summarized here. Four separate chemical sequences will

be briefly described and illustrated, as it has already been exhaustively studied

and reported (Chien 1991). The catalyst and mechanism presented will be the

TCCT/MAO combination as that is one of two used in this work, and appears to

be one of the most widely-used catalysts in research and industry for producing

syndiotactic polystyrene.

The four mechanistic steps of interest discussed herein are: 1) the active

species produced from the two catalytic components, 2) the initial attack of this

active species on a styrene molecule, 3) the propagation of this complex with

additional styrene molecules, and finally 4) methods that this process ceases by

termination. Much of the discussion here is a briefly summarized version of


Chien’s work (1991).

3.5.1 Catalyst Activation

The chemicals used for catalyzing syndiospecific styrene polymerization

do not lead to polymer with their initial form. First the co-catalysts react with

each other generating active charged species which then react with the styrene. It

is generally accepted for styrene polymerization that catalysts based on group

50
IVA metallocenes/MAO consist of cationic complexes resulting from the reaction

of the metallocene with the MAO.

The mechanism of TCCT and MAO reacting with each other are proposed

as follows (Bueschges 1989). Reduction of titanium from (IV) to (III) is made

possible by the free trimethylaluminum contained in MAO. The large molar

excess (100:1 or greater typ) of MAO to TCCT is necessary due to the known low

equilibrium constant of step B, as shown in Figure 3.1.

radical polymerization

TiCl3 + Al(CH3)3 TiCl2 + Al(CH3)2Cl + CH3

CH3
MAO (step B)
C2H6
TiClCH3 + [MAO Cl]

+ -
[ TiCH3] [MAO Cl2]

Figure 3.1: Formation of initiating species by TCCT/MAO catalytic system.

The methyl radicals formed due to Al(CH3)3 oxidation could initiate


radical polymerization to atactic polymer, but likely not consequential. The

lifetime of a primary radical is short-lived while accompanied by poor initiating

characteristics (Danusso 1958).

51
3.5.2 Polymerization Initiation

Initiation commences via coordination of a positively charged titanium

complex with a styrene molecule through the vinyl and one aryl double bond as

shown in Figure 3.2. This leads to a larger charged complex that then goes on to

react with further styrene molecules.


+

Ti
[ TiCH3]+ +
H3C

Figure 3.2: Coordination of titanium complex with styrene molecule (h4 co-
ordination).

3.5.3 Propagation

Propagation of the growing polymer chain from the initial species shown

in Figure 3.2 and subsequent reactions with additional styrene units proceeds as

follows in Figure 3.3. It has been found by using nuclear-tagged species that this

mechanism leads to cis-opening of the vinyl double bond very reliably; the

process being dependent on the last unit added only (Longo ibid. 1988). This

coordination of the nascent and growing chain is an h3 co-ordination species. It

has also been proposed that the toluene diluent/solvent may compete with styrene

in the co-ordination (Oliva 1989).

52
+
+

Ti
Ti

H3C polymer
(a) (b)

Figure 3.3: Coordination of additional styrene molecules with (a) nascent


polymer chain (one styrene unit), and (b) growing polymer chain
(more than one styrene unit) via h3 co-ordination.

3.5.3 Termination

Polymerizations of styrene using the TCCT/MAO catalytic system results

in polydispersities (MWD) of at least 2.5 and frequently exceed 3.0. Three

factors account for this and may contribute to MWD broadening (Po 1996):

1. existence of different catalytic centers;

2. existence of chain-transfer or termination reactions;

3. formation of some atactic PS due to side reactions of radical or ionic

initiation.

Case 3 is a specific case of Case 1. Termination reactions can be

considered absent due to the ionic nature of the growing/living chain under

rigorously inert conditions, and so must be intentionally killed with a proton

donor, typically done with acidified methanol. It has also been found that the

distribution actually decreases for a sample after extraction/removal of any atactic

polymer present (Ishihara ibid 1988). This leaves chain transfer as the “main

53
culprit” to molecular weight broadening, achieved through three different

proposed mechanisms.

The first is β-hydrido abstraction from the growing chain to the titanium

species, which could then go on to re-initiate a new polymer chain (Figure 3.4a).

Second, since MAO is involved and considering existing knowledge of α-olefin

polymerizations leading to aluminum-polymer bonds, the following can be

proposed (Figure 3.4b). Third, chain-transfer with monomer could follow two

different paths (Figure 3.4c and 3.4d), assigning an average rate constant to both

processes.


[Ti] CH CH2---- [Ti] H + Ph CH CH ----

(a) Ph
km
[Ti] CH CH2 ---- + Al O [Ti] CH3 + Al O

(b) Ph CH3 Ph CH CH2 ----

[Ti] CH CH2---- + (c) [Ti] CH CH3 Ph CH CH ----

(c,d) Ph ks Ph
Ph CH CH2 (d) [Ti] C CH2 Ph CH2 CH2----

Ph

Figure 3.4: Three different transfer reactions leading to MWD broadening. (a) β-
hydrido abstraction, (b) α-olefin metalation, and (c) and (d) chain
transfer to monomer (two results).

54
Chien then combined these proposed transfer reactions and assumed first-

order with respect to styrene/monomer concentration, to arrive at the following

equation for the number average degree of polymerization, DPn (Equation 3.2).*

1 k + k β + k s [ Ph − CH = CH 2 ]
= m (3.2)
DPn k p [ Ph − CH = CH 2 ]

which is only a function of styrene concentration; kp is the

propagation/polymerization rate constant (addition of styrene to active sites).

3.6 COPOLYMERIZATIONS

The copolymer provided by DOW was labeled as being a 7% crystalline

copolymer of styrene and paramethylstyrene of 200k molecular weight (per the

sample package labeling). This polymer still formed gels in various aromatic

solvents, with gel-sol transition temperatures about 25ºC lower that those formed

using pure homopolymer with everything else held constant. Although no

polymerizations were actually attempted in-house using styrene and a substituted

styrene, it is certainly may be advantageous to investigate the huge potential for

transition temperature control with a specific solvent and loading, by merely

changing the amount of co-monomer used and its substituted component. For our

specific application of TBS monomer, using a small fraction of purified TBS

along with styrene would be an excellent start. If more time was available this

would be the next polymerization that would be attempted in-house.

*This is how the equation was given in Chien’s work. This could simply be inverted to give the
actual degree of polymerization and not its inverse.

55
3.7 CONCLUSION

Acquisition and synthesis of suitable gelling polymers was the most

challenging portion of this work after this phenomenon was discovered.

Mastering the synthetic chemistry methods required for production of very pure

and active catalyst materials proved quite challenging, yet in the end produced

excellent results. Certainly a greater range of polymer molecular weights would

be desirable, but at the same time polymers produced via these methods result in

samples of such high polydispersity that further refining of a few polymer

samples might prove more useful.

Investigation into fractionating the syndiotactic polymers into low

distribution samples (1.2 or less) indicated this would be a rather expensive

operation in order to obtain enough material to perform experiments with.

Another’s work (Cohen 1989) includes results describing the production of

isotactic polystyrene via anionic synthesis at rather low distribution (~1.5),

interestingly the reaction reportedly requiring the participation of a minute

quantity of water (which is normally an ionic catalyst poison). Perhaps in the

future a similar method will lead to low-dispersity SPS via a multi-component

anionic complex synthesis.

56
REFERENCES
Braun, D.; Betz, W.; Kern, W., Makromolecular Chem., 42, 89, 1960

Bueschges, U.; Chien, J., J. Poly. Sci., Poly. Chem. Ed., 27, 1529, 1989

Campbell, R., US Pat. 5,066,741, 1991

Chien, J.; Salajka, Z., J. Poly. Sci., Part A: Poly. Chem., 29, 1243, 1991

Cohen, R.; Cazzaniga, L., Macromolecules, 22, 4125, 1989

Danusso, F.; Sianesi, D., Chim. Ind.. 40, 450, 1958

Grassi, A.; Longo, P.; Proto, A.; Zambelli, A., Macromomlecules, 22, 104, 1989

Ishihara, N.; Seimiya, Y.; Kuramoto, M.; Uoi, M., Macromolecules, 19, 2464,
1986

Ishihara, N.; Kuramoto, M.; Uoi, M., Macromolecules, 21, 3356, 1988

Longo, P.; Grassi, A.; Proto, A.; Ammendola, P., Macromolecules, 21, 24, 1988

Oliva, L.; Pellecchia, C.; Cinquina, P.; Zambelli, A., Macromolecules, 22, 1642,
1989

Pellechia, C.; Longo, P.; Grassi, A.; Ammendola, P.; Zambelli, A., Makromol.
Chem. Rapid Commun., 8, 277, 1987

Po, R.; Cardi, N., Prog. Polym. Sci., 21, 47, 1996

Proto, A.; Longo, P.; Grassi, A.; Ammendola, P., Macromolecules, 21, 24, 1988

Siddall, J., US Pat. 5,484,862, 1996

Yamada, A.; Kogyo Kagaku Zasshi, 64, 2204, 1961

Zambelli, A.; Longo, P.; Pellechia, C.; Grassi, A., Macromolecules, 20 2035,
1987

Zucchini, U.; Albizzati, E.; Giannini, U., J. Organometal. Chem., 26, 357, 1971

57
Chapter 4: Polymer Characterization

4.1 INTRODUCTION

After polymer samples were synthesized or obtained from a supplier,

many characteristics were obtained from them before they were used in making

gels. Several methods and instruments were used for gathering information on

the in-house synthesized polymers and samples obtained from Dow Chemical. To

perform this work the polymers were first subject to various processing methods

in order to ensure a high stereoregular chain fraction and minimize any other

impurities.

4.2 POLYMER SYNTHESIS PURIFICATION

Polymers produced in the laboratory were initially a mixture of reaction

solvent, unconverted monomer, methanol, spent catalysts, and polymer. To

obtain clean purified polymer a few different separation processes were

employed. To remove reaction solvent and monomer, the reaction mixture was

precipitated into a vessel containing at least three times of the poured volume of

chilled methanol while rigorously mixing. The solid precipitate was then vacuum

filtered with a Buchner Funnel and twice washed with chilled methanol.

This solid material was dried in an air oven and then a vacuum oven to

constant weight to prepare it for extraction. A standard method (Grassi 1989) was

used for separating syndiotactic from atactic polymer by extracting the polymer

mixture with butanone (MEK) in a Soxlet extraction apparatus for several hours.

This process extracted each polymer sample for at least four hours, the unit

58
cycling every five minutes or so. After this procedure it could normally be

assumed that only syndiotactic polymer remained in the thimble (if any, and

resultant polymer was confirmed as stereoregular via NMR analysis), although it

quickly became evident that a significant quantity of spent catalyst and other

insoluble reaction products were also still present in/with the desired polymer.

The procedure to typically remove such impurities was not located in any text or

journal, so a double Soxlet extraction was used with two separate solvents to
separate the ashes and to isolate the stereoregular fraction from atactic polymer.

To remove the polymer from the spent/deactivated catalysts, the same

thimble was oven dried, then extracted for at least six hours with xylene at about

130ºC with a six minute cycle time. This process doubled to remove the polymer

from the ash as well as putting the polymer back into solution for final

precipitation into a quenched non-crystalline (or at least way less crystalline than

annealed) state. This hot xylene solution was then immediately poured into a 4x

volume of chilled methanol to precipitate the polymer. This polymer was vacuum

filtered, rewashed with methanol again, air dried, oven dried, then vacuum-oven
dried to constant weight. The resultant material was usually quite powdered to

flaky, and of very high surface area. Samples were stored in capped glass vials.

4.3 DIFFERENTIAL SCANNING CALORIMETRY

All differential scanning calorimetry (DSC) work was done on a Perkin

Elmer DSC7 with aluminum sample pans. Of great interest in the case of

syndiotactic polystyrene is the crystallization temperature and enthalpy of this

transition. Other properties that were also found include the glass transition

59
temperature, melting temperature, heat of fusion, and for a quenched polymer the

lowest temperature at which the polymer would begin to crystallize from heating,

not cooling. This can be measured when the polymer has been trapped in an

amorphous state by near-instant cooling during purification processing.

A differential scanning calorimeter operates via controlling the

temperature of a sample under measure very precisely. A thermal response is

observed by changing the temperature of the sample via a preprogrammed routine


and controlling the amount of heating (or cooling) required in order to maintain

the setpoint/programmed temperature. DSC thermograms are presented for each

polymer used in this study. Thermograms are presented as the rate of change of

heat capacity as a function of temperature, versus sample temperature.

Endothermic transitions create peaks and exothermic transitions depressions using

this convention.

A total of four polymers presented here: the DOW homo- and copolymers,

as well as two different in-house synthesized homopolymers of far lower

molecular weights than the DOW materials. Each thermogram is then analyzed
for their pertinent properties and compared to one another.

4.3.1 Syndiotactic Polystyrene DSC

A thermogram of DOW SPS homopolymer is shown below in Figure 4.1,

after quenching preparation described in Chapter 5 under the invention disclosure.

60
(b)
Heat Flow, Endo Up

(c)
(a)

70 120 170 220 270


Sample Temperature (°C)

Figure 4.1: DOW SPS Thermogram. First heating, 25ºC to 290ºC from
quenched state, 20ºC/min heating rate, 14.2 mg sample. (a) glass
transition then immediate recrystallization just above the glass
transition temperature, (b) melting endotherm, (c) crystallization
exotherm.

This initial temperature rise, being from a nearly amorphous state due to

the rapid cooling processing, shows crystallization first via an endotherm

occurring at around 105ºC followed by an exotherm peaking at 112ºC, this

occurring just at and above the expected glass transition temperature. After

another small unexplained exotherm at 195ºC, a broad melting endotherm,

peaking at 271ºC, heat of fusion of 30.8 J/gram, or 3.2 kJ/mol repeat units. This

61
melting point agrees well with reported (Chien 1991) values of 271ºC for SPS.

Figure 4.2 shows the cooling response of this now molten sample.
Heat Flow, Endo Up_

70 120 170 220 270


Sample Temperature (ºC)

Figure 4.2: DOW SPS Thermogram. Cooling, 290ºC to 70ºC, 20ºC/min cooling
rate, 14.2 mg sample.

The crystallization temperature is significantly lower than the melting

point at about 228ºC. The heat of crystallization integrated to be 2.9 kJ/mol

repeat units. This number is about 10% less than the 3.2 kJ/mol reported on

melting, but statistically they are not significantly different due to how dependent

the peak areas are on the baseline selection for integrating and instrumental

62
inaccuracies. The height of the peak is dependent on the cooling rate used as

well.

The second temperature upscan now shows only the glass transition

temperature at about 105ºC, with no other activity until the melting point due to

near-equilibrium slow-cooling. Figure 4.3 illustrates the slight endotherm step of

the glass transition at 105ºC and melting point, this time peaking at a slightly

lower number of 269ºC.


Heat Flow, Endo Up

Tg

70 120 170 220 270


Sample Temperature (°C)

Figure 4.3: DOW SPS Thermogram. Second heating, 70ºC to 290ºC from
annealed state, 20ºC/min heating rate, 14.2 mg sample.

63
The dual peak is proposed (Chien ibid 1991) to be a phenomenon of the

polymer chains melting, first into a less-ordered thermoplastic gum, then full-

blown melting to a liquid. This dual peak nature is not observed when heated

from a quenched state, nor in lower molecular weight samples.

4.3.2 Questra MA405 Copolymer DSC

The copolymer was scanned using the same heating program as that for
DOW SPS homopolymer. Figure 4.4 shows both temperature rise and the cooling

scans.

(a)
Heat Flow, Endo Up_

(b)

(c)

70 120 170 220 270


Sample Temperature (ºC)

Figure 4.4: DOW Syndiotactic Copolymer (CPS). Range 70ºC to 290ºC,


20ºC/min rates. (a) first (quenched) heating scan, (b) second
(annealed) heating scan, (c) cooling scan.

64
Upon initial temperature rise the copolymer shows a striking endotherm at

105ºC then an immediate exotherm, indicating chain mobilization since it was

trapped due to quenching from solution. This is also sometimes referred to as

‘pan wiggles’ as the first heating of a prepared DSC sample often shows a slightly

erratic thermal response; this response can also can occur from slow annealing of

the original sample below its Tg before further heating. Another exotherm at

175ºC indicates further crystallization, up to a very broad single melting peak


tangent-centered at 252ºC. Since the paramethylstyrene content of this polymer is

not known, the heat of fusion is reported as 24.0 J/gram of polymer. The molar

heat of fusion ranges linearly from 2.5 kJ/mol repeat-units for pure styrene to 2.8

kJ/mol for pure paramethyl styrene, a difference of less than 15%. If the content

is assumed to be 7% as this fraction is given on the label, the heat of fusion for

this CPS then becomes 2.53 kJ/mol repeat-units (therefore a per-unit average).

Crystallization upon cooling from the melt occurred at 196ºC, ∆Hf=19.5 J/g, or

2.17 kJ/mol.

The polymer was later ascertained to contain 7% paramethylstyrene and


the balance styrene (via private communication with DOW). Second heating of

the annealed sample yields a glass transition at about 103ºC and a very sharp

melting endotherm at 246ºC, 19.6 J/gram. The peak is much narrower than before

but of the same unit enthalpy.

4.3.3 Synthesized Syndiotactic Polystyrenes

Thermograms were also obtained for polymer produced in-house using

two different catalysts and analyzed as prepared from the separation process. The

65
summary of each trace is shown in Figs. 4.5 and 4.6. Both polymers were of

lower molecular weight than the Dow SPS.

(a)
Heat Flow, Endo Up_

(b)

(c)

70 120 170 220 270


Sample Temperature (ºC)

Figure 4.5: UT 0515 SPS Homopolymer. Range 70ºC to 290ºC, 20ºC/min rates.
(a) first (quenched) heating scan, (b) second (annealed) heating scan,
(c) cooling scan.

66
(a)
Heat Flow, Endo Up_

(b)

(c)

70 120 170 220 270


Sample Temperature (ºC)

Figure 4.6: UT 0918 SPS Homopolymer. Range 70ºC to 290ºC, 20ºC/min rates.
(a) first (quenched) heating scan, (b) second (annealed) heating scan,
(c) cooling scan.

The 0515 polymer sample exhibits many of the same responses and

artifacts as does the DOW homopolymer, aside from slightly lower melting and

largely lower crystallization peak temperatures, implying a lower molecular

weight. The 0918 sample shows a much lower melting temperature that for the

two previous homopolymers, yet a higher crystallization temperature (for a like

heating/cooling schedule). It was later found that this polymer was of such a low

molecular weight (4,300, ~42 repeat units) that it was acting “between” that of a

67
high-polymer and low molecular weight species, in that the spread between

melting and crystallization temperature responses was getting smaller for a given

heating/cooling rate.

4.3.4 Heating/Cooling Rate Dependence of Phase Transitions

Syndiotactic polystyrene crystallizes rather quickly, far more quickly than

its isotactic counterpart, but is by no means instantaneous. For purposes of


brevity and a matter of convention crystallization and melting temperatures are

reported as observed under heating or cooling rates of 10 or 20ºC/minute unless

otherwise noted. Due to the kinetics of these transitions it is more accurate to

report the equilibrium transition temperature. Since a sample in the DSC would

give a negligible detectable thermal response at a vanishingly small cooling or

heating rate, several corresponding rates were employed, to later be extrapolated

to an equilibrium ‘zero-rate’ transition condition.

4.3.4.1 Melting Point Time-Dependence

Figure 4.7 demonstrates the thermal response under several different

heating rates for Dow SPS homopolymer. These heating passes are all second

heatings of a sample that was cooled at 10ºC/min from 290ºC to 70ºC (or lower).

68
(a)
Heat Flow, Endo Up_

(b)

(c)

(d)

(e)

250 260 270 280 290


Sample Temperature (ºC)

Figure 4.7: DOW SPS Varying Heating Rates Thermogram. (a) 20ºC/min, (b)
15ºC/min, (c) 10ºC/min, (d) 5ºC/min, (e) 2.5ºC/min. Cycled at
10ºC/min down to 70ºC.

The ultimate melting temperature is neither time nor heating rate

dependent for this polymer, as would be expected. What is seen though is for

slower heating rates several minor transitions occur as the polymer goes from a

flexible thermoplastic to the melt. This is cited to be (Hoffman 1965) minor

changes in the crystal structures and pitch changes in chain winding. Melting is

demonstrated to be virtually the same temperature regardless of how fast it is

heated.

69
4.3.4.2 Non-Isothermal Crystallization

Figure 4.8 shows several different cooling rates, yielding varying

temperatures at which the exothermic crystallization process takes place.

(a)

(b)
Heat Flow, Endo Up_

(c)

(d)

(e)

215 225 235 245 255


Sample Temperature (ºC)

Figure 4.8: DOW SPS Varying Cooling Rates Thermogram. (a) 20ºC/min, (b)
15ºC/min, (c) 10ºC/min, (d) 5ºC/min, (e) 2.5ºC/min.

Crystallization measurements performed isothermally affords better

agreement with available crystallization models (Chen 2002; Sudduth 2002).

Since syndiotactic polystyrene crystallizes so rapidly compared to other

crystalline polymers, the available equipment and software was not capable of

performing reliable isothermal experiments at more than a few degrees of

70
subsooling. Non-isothermal means of measuring crystallization would have to be

pursued. Models do exist however for modeling non-isothermal crystallization,

but they are inherently more complex, and do not agree with experimental

observations as well as their isothermal counterparts (Park 2000; Yuan 2000).

Since large cooling rates are used with our instrument data acquired from elevated

cooling rates (>20ºC/min.) is also verified experimentally as to the instruments

maximum capabilities.
Cooling the polymer at varying rates does result is large differences in the

temperature at which crystallization occurs. Compiling the crystallization

transition peak temperatures and plotting linearly versus rate of cooling yields a

dataset that can then be extrapolated to a zero-rate or equilibrium transition

temperature. Although cited to be a linear extrapolation (Shinozaki 1997), the

author herein shows both linear and square extrapolations and their corresponding

zero-rate intersections shown in Figure 4.9 since neither is a very good fit.

71
255

250
Crystallization Temp (ºC)_

245
(a)

240

(b)
235

230

225
0 5 10 15 20
Cooling Rate (ºC/min)

Figure 4.9: Extrapolation of SPS Crystallization Point to Zero Cooling Rate. (a)
linear extrapolation to 0ºC/min cooling rate, (b) square extrapolation
to 0ºC/min cooling rate.

Both of these fits are fairly well representative of what has been

previously reported for the crystallization of SPS and for the formation of SPS

gels. A broader cooling rate spectrum was then selected, to determine if indeed a

regime of cooling rates would indeed produce a linear fit. What results is Figure

4.10, which is in fact linear at cooling rates of 20ºC/min and above.

72
260

255

250
Crystallization Temp (ºC)_

245

240

235

230

225

220

215

210
0 20 40 60 80 100
Cooling Rate (ºC/min)

Figure 4.10: Broad Cooling Rate Range Crystallization Points of SPS. Region
with excellent linear fit shown extrapolated to zero cooling rate.

Crystallization is expected to occur anywhere between the glass transition

temperature and melting temperature of the polymer (Wesson 1994). Aside from

instantly quenching it below its glass transition temperature at probably

10,000+ºC/min cooling rate, not allowing for any ordering to take place, as long

as the polymer is resident for some period of time between 100 and 270ºC it will

crystallize. The extrapolated temperature of 233ºC is proposed to be useful for

predicting the maximum temperature at which the material could be cooled to

73
before heat-transfer or chain-mobility limitations apply. At higher crystallization

temperatures (slow cooling rates) the thermodynamics of the material contribute

to its crystallizing, and larger molecular weight species crystallize first; lower

temperatures (faster cooling rates) on the other hand result in a huge

thermodynamic driving force to crystallize, but the process becomes diffusion

limited (via higher viscosity), hence smaller molecular weight chains order first.

For slightly higher molecular weight SPS samples (300k Mw) the
maximum in crystallization kinetics occurs at about 200ºC (Wesson ibid.),

although it is not noted if this is directly correlated from kinetic cooling data or

extrapolated to an isothermal cooling model. In the next chapter this same

analysis will be performed on gels with very little SPS component, to determine if

the same behavior is observed for gel formation.

4.3.4.3 DSC Operating Envelope

High rates of cooling were employed for the non-isothermal crystallization

studies. Depending on how fast of cooling is desired, the sample’s temperature

(or temperature where observable activity is expected), and the instruments heat

history, the DSC temperature controller cannot always maintain the actual
program temperature. Several runs were therefore performed with the specific

intent of only generating an operating envelope for this particular DSC

instrument.

On the exterior of the DSC instrument is an LED indicating whether the

internal temperature controller is “keeping up” with the program or not. If

program temperature cannot be maintained in the sample head, the red light goes

74
off. Figure 4.11 illustrates the operating envelope of the DSC instrument used for

this work, using a large (22 mg) sample in order to conservatively obtain the

operating curves. Two separate investigations were done, one with the instrument

sinking heat at room temperature, and another while operating the –40ºC capable

intercooler.
220

180
Floor Controlled Temperature (ºC)

140
(a)
100

60

(b)
20

-20

-60
0 20 40 60 80 100 120 140
Cooling Rate (ºC/min)

Figure 4.11: CPE 3.464 Perkin Elmer DSC 7 Operating Envelope. (a) Intercooler
OFF; (b) intercooler ON. Actual experimental data using 22 mg
(large) polymer sample and observing the controller compliance
LED.

75
For pure polymer studies the intercooler was off since all of the cooling

transitions of interest took place at sufficiently high temperatures to not

necessitate the need for it. DSC Studies of gels, as described in detail in Chapter

6, were performed with the intercooler on to accurately detect gel transition

temperatures as low as 10ºC.

4.4 STEREOREGULARITY ANALYSIS

In order to determine how stereoregular our syndiotactic polymers were,


13
C nuclear magnetic resonance (13C-NMR, or just NMR) spectra were collected

for each polymer sample. Since liquid/solution samples were required for this

analysis and these polymers like to form gels in solution, elevated temperature

analysis was required. The solvent must be deuterated for the sample signal to be

discerned.

This analysis was selected to perform conformational analysis of the

polymer as it has been cited previously (Kawamura 1979; Grassi ibid 1989) as a

simple yet effective test for stereoregularity. The peculiar thing is they did not

cite the deuterated solvent used for the analysis. Deuterated dimethylsulfoxide

(DMSO) was tried as the measurement solvent since that is normally employed
for high temperature NMR analysis. Sufficient polymer could not be dissolved in

this solvent and upon cooling what solution had been made phase separated.

Another solvent was necessary that would completely dissolve the required

amount of polymer (about 5 wt%) and become/remain a liquid at an elevated

temperature while neither boiling nor exceeding 120ºC. The only other candidate

solvents for this analysis were heavier deuterated aromatic solvents. Benzene

76
boils too low, so deuterated toluene was next in line. Sufficient polymer

dissolved in toluene but the material would not melt to a liquid without some

boiling, which was not possible with a sample within the NMR probe. So,

deuterated xylene was used, and it provided excellent results.

4.4.1 NMR Tube Charging

Dissolving the polymer samples in deuterated solvent was a fairly simple


procedure. Getting the solution into a 3 mm bore NMR tube to the proper level

while not allowing it to gel (keeping it hot) was another matter requiring some

creative plumbing. To accomplish this, a 12” length of heavy polyethylene

cannula was made into a U shape with beveled ends and one end run through a

sealed cap into the hot solution and the other end placed into the NMR tube to the

desired final level. Positive pressure was used to force hot solution from the test

tube into the NMR tube to the proper level or above. After filling, a long

hypodermic needle was used to withdraw any excess solution till the tube was

filled to precisely 4 cm in height. Normally precise NMR tube filling is not

necessary but since these samples were run at such elevated temperature, precise

filling was necessary to avoid thermal gradients within the sample while within
the NMR probe. Illustration 4.1 shows the NMR tube filling apparatus devised.

77
Illustration 4.1: Hot Solution Charging of 4 mm NMR Tube. Pressure source at
only a few psi, regulated via a fingerhole on hypodermic needle
Luer into test tube to ease in hot liquid with positive pressure.

After preparation, charged tubes were submitted for analysis on a Bruker


13
AMX-500 (500 MHz coil) instrument. Figure 4.12 shows the downfield C

response of a sample of Dow SPS in d-8 xylene.

78
138.7 ppm
(4º solvent ring)

Relative Intensity
148.4 ppm
(4º polymer ring)

152 150 148 146 144 142 140 138 136


Chemical Shift (ppm)

13
Figure 4.12: C NMR of SPS in d-8 Xylene. Scanned for one hour at 120ºC,
truncated to pertinent region of chemical shift (baseline noise
removed). Shown both polymer and solvent relative responses,
polymer peak normalized to 100.

The analysis focuses on the farthest downfield signal, this being the

quarternary carbon on the pendent group, around 148 ppm downfield. A single

peak here that is very sharp can indicate the polymer is quite stereoregular, which

was observed for all of our polymers. In the NMR operator’s opinion the

polymers were very stereoregular and of rather low polydispersity, although these

polymers were all later found to have polydispersities of 2.5 or more, determined

79
via other analysis such as GPC. A sharp peak could not exclusively be due to

only syndiotactic diads though, as it could also indicate only isotactic diads.

4.4.2 Tacticity Discernment

In order to compare the response of these polymers relative to other

polystyrenes, high temperature 13C-NMR was also done with isotactic polystyrene

(iPS) and regular atactic polystyrene under identical conditions. The chemical
shifts of the quarternary carbon and the peak shapes are quite different, hence

confirming the previous results showing how syndiotactic the three studied

polymers were. Figure 4.13 shows both isotactic and atactic downfield scans.

SPS has a singlet at 148.4 ppm while IPS produces a peak at 148.9 ppm.

80
(a), broad spectrum
around 148.7 ppm
aPS 4º ring carbons

Relative Intensity
(b), 148.9 ppm
(4º iPS ring carbon

152 150 148 146 144 142 140 138 136


Chemical Shift (ppm)

Figure 4.13: Isotactic and Atactic Polystyrene 13C NMR Far-Downfield


Responses. Scans taken at 120ºC for 1 hour in d-8 xylene. (a) atactic
polystyrene and (b) isotactic polystyrene (correlates with Kawamura
ibid.)

The peaks for quarternary carbons are indeed different in these cases, that

for Isotactic being narrow yet a different shift from those of Figure 4.12, and that

for atactic polymer being shorter and quite broad, indicating the totally random

81
dispersal of repeat unit diads shielding the quarternary ring carbon in several

different degrees.

Other analysis was performed on in-house generated polymers to verify

their tacticity as well. Table 4.1 summarizes the results of all the NMR work

performed on polymers for this research. The same analysis was also done on the

DOW copolymer to determine if it was stereoregular despite having a significant

portion of substituted styrene in the polymer.

Material Main 4º Peak Quality Side Peaks?


Dow SPS 148.4 very narrow no
UT0515 148.4 very narrow no
UT0918 148.4 narrow Yes, 148.8
Dow CPS 148.4 narrow Yes, 148.1
aPS 148.7 centered very broad broad, continuous
iPS 148.9 narrow no

Table 4.1: High-Temperature 13C NMR of Various Polystyrenes. All data


obtained on 500 MHz instrument in d-8 xylene at 120ºC for one
hour.

4.4.3 Catalyst Effect on Syndiotacticity Purity

Tetrabenzyltitanium (TBT) was selected due to its previously described

highly selective nature for producing syndiotactic polystyrene without ‘forgetting

then remembering’ syndiotactic addition during chain propagation. Since another

catalyst, trichlorocyclopentadienyl titanium (TCCT), was also used to make

polymer, NMR results were compared to determine if any difference in tacticity

efficiency could be observed in our prepared samples. This is no published

information as to how efficient TCCT is during chain propagation. Figure 4.14

82
shows the relative quarternary carbon chemical shifts for the two catalyst systems

used during polymerizations.

148.4 ppm
(a)

Relative Intensity
148.4 ppm
(b)

152 150 148 146 144 142 140 138 136


Chemical Shift (ppm)

Figure 4.14: Intra-Chain Syndiotactic Content from Two Catalysts. (a)


tetrabenzyltitanium/methylaluminoxane catalytic system and (b)
trichlorocyclopentadienyl titanium/methylaluminoxane catalytic
system.

Comparing these results qualitatively indicates that the TBT synthesized

polymer is virtually 100% syndiotactic diads, while the TCCT produced polymer

does have quarternary-carbon peak characteristics slightly alike those of atactic


83
polymer, although it is still mostly syndiotactic. A further extraction was

performed on this polymer to make sure the side peak was not due to a non-

extracted fraction of atactic chains, but the results were the same.

4.4.4 NMR Tube Cleaning

After analysis, the NMR tube was left with a solid cylinder of gelled

solvent, of which could only be removed by remelting it and extracting it hot.


Another apparatus was used to completely submerge the NMR tube in xylene

solvent, it boiled until the charge had melted, then removed via a vacuum

capillary tube, pulling excess solvent through the tube to scavenge all remaining

polymer while boiling the entire apparatus. Illustration 4.2 shows how this was

done.

84
Illustration 4.2: Removing NMR Tube Charge and Cleaning. Performed in
boiling xylene solvent and pulling ~50 cc of solvent through
NMR tube.

4.5 MOLECULAR WEIGHT ANALYSIS

A molecular weight and polydispersity analysis was performed on all of

our polymers under study by Gel Permeation Chromatography (GPC) on a Waters

Model 150C with a Viscotek Model 250 refractive index detector and

divinylbenzene/styrene columns, 1 cc/min flow rate carrier solvent. This

provided the information required to correlate the observed macroscopic

mechanical properties of the formed gels to molecular weight. Standard analyses

85
described in the literature (Ishihara 1988) prescribe that a di- or tri-chlorinated

benzene carrier solvent be employed at high temperature (135ºC) in the

chromatography column. Since this capability was not at our disposal, other

methods were investigated. The only available equipment utilized standard

crosslinked polystyrene/divinylbenzene columns running THF carrier solvent at

room temperature.

It was found that by rigorously boiling a quench-precipitated sample of


syndiotactic polymer in THF for several hours, enough polymer had dissolved in

order to perform the GPC analysis at room temperature, although only about 0.5

mg/cc of THF carrier solvent. Using a polystyrene standard to compare elution

times and signals, the molecular weights and distributions of the polymers under

study are presented in Table 4.2.

Sample Mn Mw Distribution
Dow SPS 41,500 206,700 5.0
UT0515 34,100 92,000 2.7
UT0918 4,300 9,800 2.2
Dow CPS 40,300 174,000 4.3

Table 4.2: Molecular Weights and Distributions of Three Syndiotactic


Polystyrenes Under Study. THF carrier solvent at 25ºC and 1cc/min
flow rate. Injected from aliquot drawn from freshly boiled
polymer/THF sample.

Sometimes polymer remained undissolved after this boiling operation, so

the remaining undissolved polymer was recovered and reboiled in a fresh sample

of THF. Serial boiling and subsequent GPC testing was done to ensure that since

the method deviated from what is standard procedure, it was in fact not skewing

86
the results in favor of lower molecular weight components that were preferentially

dissolving in the solvent. Tests show this was not the case however; the polymer

was not fully dissolving due to its lack of solubility, independent of molecular

weight and molecular weight distribution. If the boiling temperature of the THF

could be raised by putting the whole thing in an autoclave at elevated pressure

that would certainly assist in getting the SPS into THF solution.

4.5.1 Validity of Non-Conventional GPC Technique

Since room temperature GPC was done in THF solvent instead of high

temperature (120+ºC) in chlorinated solvents, this raises questions as to whether

the results are in fact valid. The main issue is whether during the long period of

boiling the polymer in THF did the polymer fractionate, i.e. lower molecular

weight polymer dissolving while higher MW’s staying out of solution entirely.

Since THF boils at only 66ºC at atmospheric pressure, this is far below the glass

transition temperature of the polymer so polymer that had been quench-processed

was used to maximize its solubility, essentially in a 100% amorphous state.

Using very small fractions of polymer it was found, by inspection, that

polymer would entirely dissolve in THF by boiling for several hours at a polymer
loading of 1 mg/cc of solvent. Not a large amount but sufficient for GPC

analysis. Upon cooling this solution to room temperature it remained a clear

homogenous liquid, for at least several hours. After at least several hours the

solution would begin to visibly gel as the polymer typically does far more quickly

in heavier aromatic liquids. For this reason GPC samples were prepared/boiled

immediately before analysis to avoid this delayed-onset gelation.

87
To provide additional evidence of the validity of this method of molecular

weight characterization, a solubility parameter calculation was done to

demonstrate, via a method not dependant on physical observation, that it is indeed

possible to get polymer to dissolve into THF at atmospheric pressure. Using the

known solubility parameter of THF of 9.1 cal/cc and Equation 2.19, the calculated

melting point of THF/SPS crystals at infinite dilution is about 53ºC, which is well

below the boiling point of THF. This of course does not take into account the
difficulty in getting it to dissolve, just shows it is possible.

4.6 CONCLUSION

The various polymers acquired and synthesized for study have been

characterized, so that any solution behavior can be properly associated with

polymer properties and solvent properties. Molecular weight in particular is

interesting as it contributes to the gel’s physical strength and transition

temperatures. This is the main goal of this study—attempt to correlate polymer

molecular weight to the gel modulus and sol-gel transition point.

Non-conventional means were employed to characterize polymer

molecular weight but those means were shown to be nearly as effective as those
normally used. Molecular weights ranging from about 4,000 to 42,000 were

ascertained and are the polymers used for further work in this study.

Syndiotacticity and stereospecificity were verified to be very high for all

polymers to be examined via NMR analysis. Thermal studies showed the

polymer’s crystallization behavior to be very time and temperature dependent.

88
REFERENCES
Chen, Q.; Yu, Y.; Na, T.; Zhang, H.; Mo, Z., J. App. Poly. Sci., 83, 2528, 2002

Chien, J.; Salajka, Z., J. Poly. Sci., Part A: Poly. Chem., 29, 1243, 1991

Grassi, A.; Longo, P.; Proto, A.; Zambelli, A., Macromolecules, 22, 104, 1989

Hoffman, J.; Weeks, J., J. Chem. Phys., 42, 4301, 1965

Ishihara, N.; Kuramoto, M.; Uoi, M., Macromolecules, 21, 3356, 1988

Kawamura, T.; Uryu, T.; Matsuzaki, K., Makromolecular Chem., 180, 2001, 1979

Park, J.; Kwon, M.; Park, O., J. Polym. Sci.: Part B: Polym. Phys., 38, 3001, 2000

Shinozaki, D.; Lawrence, S., Polymer Engineering and Science, 37, 1825, 1997

Sudduth, R.; Yarala, P., Poly. Engr. and Sci., 42, 694, 2002

Wesson, R., Poly. Eng. and Sci., 34, 1157, 1994

Yuan, Z.; Song, R.; Shen, D., Polym. Int., 49, 1377, 2000

89
Chapter 5: Gel Formation

5.1 INTRODUCTION

After the properties of the polymers had been obtained, they were used to

make gels. Successful gels were then studied as quickly as possible to minimize

errors brought about by solvent loss, especially those experiments done that

required open-air handling of the gels at elevated temperatures. Some even

required handling of the gel heated to where it was a sol for casting into the

appropriate measuring geometry, as in the dynamic viscoelastic rheometry (DVR)

measurements. For volatility reasons very high vapor pressure solvents such as

benzene and toluene were excluded from some experimental procedures due to

the possibility of high solvent losses and losing knowledge of the composition.

5.2 GEL FORMATION PROCESS

Our process for making gels from organic liquid and solid polymer

dramatically improved the quality of gels that were formed both in the laboratory

and during the earlier product development phase of this work. The major issue

was long exposures at high temperature led to solvent loss through boiling off

and/or monomer polymerization if the liquid bore the necessary functionality.

Commercially produced pellets of syndiotactic polystyrene were far too difficult

to weigh in small quantities and even more difficult still to dissolve in any solvent

even with rigorous boiling. Only by boiling for ten or more minutes could pellets

be fully dissolved, and this was usually at temperatures of over 200ºC. Boiling

time would not normally be an issue but when dealing with reactive monomers it

90
is very desirable if not necessary to minimize the time the monomer is exposed to

high temperatures.

A method had to be devised to get the gelling agent polymer in solution

quickly enough to minimize if not exclude altogether this thermal polymerization

while trying to dissolve the polymer. Boiling an aryl-vinyl monomer for several

minutes usually leads to too much thermal polymerization of the liquid monomer

to permit its use in reaction injection molding operations, even by introducing


additional inhibitor to consume the large quantities of free-radicals generated

from the elevated temperatures. Too much polymerization has occurred when the

formed polymer makes the liquid too thick (viscous) for the desired application,

or if crosslinker is present the matrix has reached the critical degree of

polymerization that it has indeed crosslinked, or chemically gelled.

We decided to precipitate the polymer as if we had synthesized it

ourselves, greatly increasing its surface area, as well as resulting in a polymer that

was more amorphous and presumably easier to dissolve in solvents. The entire

process of going from commercial pellets to a monomer liquid based gel was
submitted for patent review with the University of Texas at Austin Office of

Technology Licensing under the title: “Thermoreversible Gelation of Reactive

Aromatic Hydrocarbon Monomers” (UTA# A – 024 BAR).

Since the process for potentially patenting this invention has not been

pursued beyond the disclosure phase over the past two years, it will be presented

here as it was to the University of Texas Office of Technology Licensing and

Intellectual Property (UTOTLIP):

91
5.2.1 Invention Disclosure

Invention Titles (information as provided to UTOTLIP):

1) Thermoreversible Gelation of Reactive Aromatic Monomers, and

2) Process of Making and Maintaining Reactive Aromatic Monomer Gel

5.2.1.1 Invention Description

This invention consists of two parts. The first is a process for combining
several existing materials in a way to minimize material degradation. The second

is a new composition of matter, the incorporation of an existing material to

change the rheological stability of the overall composition. This new material

consists of reactive monomer(s), solid filler particles, and the rheological control

agent, also referred to as the gelling agent, in our particular case Dow Chemical

Company Questra MA405 crystalline copolymer, 7% PMS.

5.2.1.2 Invention Uniqueness

First is in the preparation of the gelling agent polymer. The polymer must

be dissolved in a non-reactive solvent (such as xylene or decalin) at the solvent’s

boiling point accompanied by rigorous agitation. Given the geometry of the

commercial gelling agent polymer pellets as provided to us, this operation usually

takes about 20-30 minutes. This resulting solution is then slowly poured into a

agitated chilled vessel of methanol. This process greatly enhances the surface

area of the gelling agent polymer as well as trapping it in a thermodynamically

amorphous state, causing it to readily dissolve in other solvents without long

periods of high temperature. Figure 1 shows the amorphous condition the

polymer is in after processing in this manner

92
Heat Flow_
Heat Flow, Endo Up_

90 110 130 150 170 190


Sample Temp (ºC)

70 120 170 220 270


Sample Temperature (ºC)

Figure 5.1: Dow Questra MA405 Copolymer. DSC thermogram of a sample


taken from material that has been processed via this invention. Inset:
Closer look at the material’s slow crystallization as it is slowly
heated above its glass transition temperature (~100ºC) from an
amorphous state.

Dissolving the agent in reactive solvent, i.e. monomer (tert-butyl styrene)

requires the monomer be heated as rapidly as possible to just below its boiling

point after the proper amount (0.2 to 4 wt%) of quench-processed polymer

powder has been added to it. This operation should cause the powder to dissolve

in less than a minute, therefore minimizing the exposure of the monomer to high

temperatures, which can lead to excessive thermal polymerization of the

93
monomer. A small amount of thermal polymerization is ok, but too much will

render the material useless for reaction injection molding operations.

After the monomer is treated with gelling agent polymer, it can be mixed

with any other desired components, including solid filler particles. Since the

particles are typically too large to provide liquid-solid rheological stability on

their own via Van der Waals interactions, the gelling agent in the monomer matrix

provides all the rheological stability. As all the materials are combined in an
appropriate mixing device, the resultant mixture must be dispensed above the sol-

gel transition temperature, in one particular case that is between 60 and 70ºC.

After dispensing into an appropriate container is must not be disturbed for several

hours until it has cooled to below the gelation temperature throughout, below

50ºC even at the center. After cooling the mixture should have fully gelled and

now be of infinite bulk viscosity yet of low physical strength, so it will still flow

when external pressure force is applied. This material should then not ever phase-

separate into distinct liquid and solid-rich components.

5.2.1.3 Invention Usefulness

Suspending solid particles in a low viscosity liquid for long-term


stability/storage is the major usefulness of this invention. There exist a large

assortment of fillers in liquids, in our particular case fumed silica of “large”

particle size (~20 microns) or glass beads in a low molecular weight aromatic

monomer liquid.

94
5.2.1.4 Invention Limitations

The length of time the reactive monomer is held at elevated temperature

has a direct effect on the success of this process and invention, the longer it is hot

the more detrimental thermal polymerization can occur. Our process is currently

quite acceptable in terms of how much material reacts, but it can be made even

better with more specialized processing equipment. Getting it hot, mixing it hot,

then cooling it back to the prescribed dispensing temperature as quickly as


possible is the best means to deliver an acceptable product.

5.2.1.5 Prior Art/Publicity

No public disclosure has ever been made of this invention up to this point.

Jempac International provided the corporate support to make this work possible.

One references which assisted to lead to the initial development of the ideas and

processes for this invention were primarily gleaned from a work of Guenet

(1996).

5.2.2 Invention Addendum

Concentrated samples of monomer and gelling agent polymer were also

made and then diluted to the desired polymer concentration with additional

monomer after dissolution to further reduce the amount of net thermally-

polymerized polymer in the resultant monomer solution. This is a more desirable

way to process the polymer, requiring smaller high-temperature equipment to get

the polymer into solution, and allows for greater control over the final gelling

polymer concentration in the resultant product.

95
5.3 POLYMER SOLVATION

To dissolve the polymer on a small scale, oven-dried test tubes were used

to create samples usually of about one to three mL for immediate use. The test

tube was tared, polymer powder weighed into it, and re-tared. Solvent was then

introduced to create the desired weight fraction, then a little bit more for the

ensuing boiling/dissolving operation, which usually resulted in some solvent loss.

Actual mass fractions were determined on the spot by weighing the entire vessel

just before removing a sample, either a solid piece or remelting the system for

liquid handling if so necessary.

5.3.1 Volatility Issues

Questions arose as to the error this methodology introduced when

transferring gel samples, particularly those of elevated temperature of high vapor

pressure solvents. Several tests concluded the repeatability of the employed

transfer techniques to exceed 99% accuracy for moderate vapor pressure solvents.

This method was not suitable for high vapor pressure solvents.

In order to quickly produce laboratory samples of gel, small disposable

test tubes were used and sample of about two mL made in each. The charged test

tube was capped with a septum, a hypodermic needle inserted to vent it, then it

placed into an empty Erlenmeyer flask suspended on a ring stand. A propane

torch was used to gently heat the flask for a minute or two, such that the

polymer/solvent in the test tube was allowed to gently boil until all the polymer

was dissolved. Illustration 5.1 shows how this was done. A slight fire hazard

exists when doing this but as long as the solvent is carefully boiled, and not boiled

96
over, problems were never encountered. The vessel used for this should be

laboratory glassware (Pyrex) and not regular glass jars however.

Illustration 5.1: Test-Tube Erlenmeyer Flask ‘Oven’. Propane flame used to heat
system to create small samples of polymer/solvent gel quickly.

5.3.2 Thermal Polymerization Minimization

In the case of monomer being used as the solvent, about 5 mg of tertiary-

butyl catechol (tBC) free radical polymerization inhibitor was added per cc of

monomer before and after this boiling operation to minimize thermal

polymerizations. This amount of tBC increased the nominal concentration of

97
inhibitor in the monomer by over 10-fold, but was found through subsequent

testing to not significantly interfere with any thermodynamic or rheological

measurements performed on the treated systems. Of course too much inhibitor

could adversely interfere with the RIM polymerization chemistry. To

commercially apply this technology the exact amount of inhibitor should be

determined to protect the monomer while subject to dissolving the gelling

polymer. In addition, this process should consume nearly all the inhibitor, leaving
only a trace amount to provide for room temperature/long term storage chemical

stability (as the monomer is normally treated in the pure state).

5.4 GEL STORAGE

Since the gels used in this study were all of quite low polymer loading

(<10wt%) they still possessed significant vapor pressure and were therefore

subject to solvent loss dependent on the container they were stored in, namely the

container’s closure. Most samples were used immediately or within a few days if

made and stored in a test tube with silicone rubber stopper. Longer term storage

was done in jars with tight fitting lids.

Some samples exhibited syneresis over extended time periods (several


weeks). Syneresis is a small amount of solvent being squeezed out of the gelled

slug. This process usually did not observably continue. Though. Only a small

amount of pure solvent would collect on the top surface of samples, then they

ceased exuding solvent.

98
5.5 FORMATION WITH SOLIDS PRESENT

Making a successful gel proved rather easy but making a successful gel

while maintaining a homogeneous system that is filled with silica introduced an

additional complication. The hot silica-filled sol system was still susceptible to

partial silica settling before and while the gel is forming. The gel forming is a

time and temperature dependent process, so the filled material was mixed and

brought to a temperature just at the gel-sol transition temperature, then mixing

ceased and the material cooled as quickly as practically possible. Of course in

large samples limited thermal conductivity through the system resulted in an

exterior shell of gel with sol in the middle for some period of time.

Highly loaded systems (> 70 wt%) were found to possess enough inherent

viscosity to minimize silica settling before the gel formed, counter to those of low

loading. In this respect a more homogeneous sample results from high loaded

systems than those containing low amounts of silica. Low loaded samples

required constant agitation even while the gel was forming. This interfered with

the gel’s formation. To achieve results comparable to those in a highly loaded

system, a higher gelling polymer concentration was necessary in order to speed

initial gel formation and therefore immobilization of the filler particles.

5.6 CONCLUSION

Forming stable gels did not result in much quantifiable information, yet

provided one of the most important developments in this work. Solving the

gelation problem for aromatic monomers allowed the silica filled monomer paste

system to be tested and used in pilot-scale product delivery tests. The resultant

99
gelled silica-filled paste as a solid but still rather flowable when subject to normal

pumping forces in the process equipment.

This research focuses on liquid/polymer systems with no silica filler.

Further study into the formation of a gelled matrix with varying levels of filler

would be beneficial to ensure a sufficiently strong gel while keeping the gelling

polymer concentration low. The next two chapters delve into the experimental

results from studying non-filled gels of SPS and various aromatic solvents, mainly
focusing on the SPS/ethylbenzene binary system.

100
REFERENCES
Guenet, J.; Daniel, Ch.; Deluca, M.; Brulet, A.; Menelle, A., Polymer, 37, 1273
1996

101
Chapter 6: Thermodynamic Gel Characterization

6.1 INTRODUCTION

The study of physical gels has garnered great interest over the past decade

(Guenet 1996; Nijenhuis 1996; Tanaka 1994). The mechanism of gel formation is

quite well known in dilute solution (Guenet ibid. 1996). The form of the gel

usually takes on a double-helix winding for the crystallizing polymer chains over

some range (several repeat units) to create short range order, this alternating with

the chain extending to another region in the solvent matrix to engage in further

winding with another. Neutron scattering and FTIR experiments have already

extensively characterized these phenomena (Guenet 1992; Nijenhuis 1997). This

effectively creates a physically crosslinked system much akin to what happens in

styrene-butadiene-styrene (SBS) block copolymer when cooled appropriately,

physical linkages of rubber connecting glassy regions of styrene-rich polymer.

One goal of this work is to produce physically durable gel capable of

suspending solids in a composite mixture with the smallest amount of polymer

possible, ideally below 1 wt%. So, the mechanical properties of the dilute

physical gel are of great interest. Very little study has been done on the physical

strength and/or modulus of dilute gels; this chapter will delve into the modulus of

several different gels, focusing on the SPS/EB system. The modulii of several

physical gels have been measured and are reported in Chapter 7.

Differential scanning calorimetry was used to determine the solutions sol-

gel points, enthalpies for these transitions, and time/temperature dependence for

102
these transitions. Molecular weight may also play a role in the transitions

temperatures. Very low molecular weight polymer forms more of a paste than a

gel, but still exhibits a measurable sol-gel transition exotherm. Three different

molecular weights were studied in this work.

6.2 GEL-SOL TRANSITIONS

The qualitative test for the sol-gel transition/condition is accomplished by


simply dissolving polymer in solvent, then allowing it to cool to room

temperature. If it turns into a waxlike substance over several minutes to hours, a

gel has formed. More quantitative methods require carefully controlling the

temperature while the solution is cooling, and detecting the transition via an

enthalpy change, for gels that involve some sort of crystallization that is. Gel

formation is also detected by a marked increase in the material’s modulus, going

from essentially zero Pascals to a modulus of at least several thousand Pascals.

6.2.1 Differential Scanning Calorimetry Methodology

The Perkin-Elmer DSC-7 (DSC) was used to characterize gel thermal

properties, via controlled heating and cooling at varying rates. The studies are

similar to those carried out on the pure polymer, however all the transition

temperatures in question are far depressed below those observed with pure

polymer, in line with Flory ibid. (herein Chapter 2).

To prepare suitable samples for DSC, liquid sample pans were required to

contain the material, as many tests cause the sample’s vapor pressure to get rather

high, sometimes exceeding atmospheric pressure. This limitation also required

studies to be made at lower concentrations as concentrations of polymer over

103
about 5 wt% would frequently cause the hermetically crimped aluminum liquid

pans to rupture. Pans were capable of handling systems at 5-10º above their

solvent’s boiling point (for dilute systems) but the vapor pressure became too high

if heated too far above the normal boiling point.

Addition of the polymer to the solvent does of course lower the vapor

pressure, but sometimes to achieve complete transition the temperature became

too high and the vapor pressure exceeded the positive pressure capabilities of the
aluminum DSC sample pans.

The loading of sample pans was performed by placing about 10 mg of

gelled material of known composition into a previously weighed sample pan

bottom seated in the crimping die. The previously weighed top was immediately

set in place and the two halves crimped together in a matter of several seconds,

minimizing solvent loss and therefore maintaining knowledge of the composition.

Of samples that were successfully tested, thermograms were analyzed

based on the parallel-tangent peak maximums, since many signals were quite faint

and/or diffuse. Peak onset or area calculation was often ineffective. Illustration
6.1 shows how peaks were interpreted; these peaks are quite broad and defined for

purposes of demonstrating how all were interpreted.

104
Illustration 6.1: DSC Heat Flow Signal Interpretation. Small peak/response
temperatures taken at parallel-tangent peak maximum. (This
applies to melting and crystallization/gelation thermal responses,
not glass transitions.)

6.2.2 Composition (Solvent) Variation Thermograms

Several different solvents were tested to determine their corresponding

gel-sol transition temperatures, as well as their sol-gel transition temperatures. As

demonstrated in Chapter 4, the sol-gel transition is highly dependent on the

cooling rate of the sample under study. The higher the cooling rate, the lower

temperature the transition occurs.

105
6.2.2.1 Cooling Rate Sweep

Figure 6.1 shows a typical DSC study of a single gelled sample to provide

transition temperatures over a large range of cooling rates. This particular sample

is 1.0 wt% Dow SPS in ethyl benzene (EB).

(a)
Heat Flow (Endo Up)

(b)

0 10 20 30 40 50
Sample Temperature (ºC)

Figure 6.1: DSC Thermograms of Dow SPS in EB (same sample pan). Large
range of cooling rates; rates are from top (a) to bottom (b): 5, 10, 15,
20, 25, 30, 35, 40, 45, and 50ºC/min. Sol-gel transition peaks shown
by open boxes. (The peaks are far more pronounced when
appropriate analysis scales are used; in the figure the peaks are
compressed by the large y-axis range.)

The transition temperatures are then picked and plotted versus cooling rate

to obtain the “dynamic” phase diagram shown in Figure 6.2.

106
45

40

35
Sol-Gel Transition (ºC)

30

25

20

15

10
0 10 20 30 40 50
Cooling Rate (ºC/min)

Figure 6.2: Sol-Gel Transition as Function of Cooling Rate. 1% Dow SPS in


EB, much akin to that shown in Figure 4.10 for pure SPS polymer,
but transition temperatures depressed about 200ºC. Linear fit data
from 20 to 50ºC/min.

This analysis was carried out for several different solvents including

toluene, xylene, styrene, and tertiary-butyl styrene. Narrower cooling rate ranges

were used when studying other solvents as suggested in previous information

(Guenet 1988) for linear extrapolation, although it has now been demonstrated

that the larger range of cooling rates provides much more useful information.

107
Table 6.1 summarizes the responses of different solvents at ~1% Dow SPS

polymer loading.

Solvent Molar Vol. T(melt) (ºC) T(gel) (ºC) Tm-Tgel


Toluene 106.5 111.0 28.0 83.0
o-Xylene 122.0 112.0 31.0 81.0
Ethyl Benzene 122.5 104.4 35.0 69.4
Styrene 114.6 115.0 46.5 68.5
t-Butyl Styrene 183.2 92.0 various(a) N/A
p-Methyl Styrene 131.8 109.0 21.5 87.5

Table 6.1: Different Solvent/SPS Combinations at 1% Polymer Loading with


Heating/Cooling Rates of 10ºC/min. It is useful to note at this
junction that the cooling rate of ~10ºC/min appears to be the same
transition temperature as the zero-rate extrapolated transition
temperature from the linear transition region. (a) Three slightly
detected transitions at 4, -10, and –18.7ºC.

6.2.3 Concentration Variation Study

To determine the effect of varying amounts of polymer loading on

gelation dynamics, different loadings of polymer were used with the same

solvent. The system selected for this study was Dow SPS in EB, with loadings of
1, 3, 5, 10, and 34 wt%. Depending on the cooling rate for the 34 wt% sample,

there is evidence that more than a single sol-gel phase transition may be taking

place (Figure 6.3). This split in transitions may indicate more than one phase is

involved, the sol becoming one phase of gel then with further ageing/cooling,

proceeding to another gel phase. This may happen with dilute systems as well but

essentially as a continuum, with only one detectable exothermic transition.

108
Richer systems (> 10 wt%) may also, in fact, split into two phases before the sol-

gel transition, into a polymer rich and polymer lean phase, where each of these

phases may go through a respective sol-gel transition, which exhibit different

transition temperatures because they have different compositions.


55

45
Transition Temperature (ºC)

35

25

15 1 wt%
3 wt%
5 wt%
5 10 wt%
34 wt%
-5
0 10 20 30 40 50
Cooling Rate (ºC/min)

Figure 6.3: Various Dow SPS Loading Levels in EB Solvent, Sol-Gel Transition
as Function of Cooling Rate. Elevated polymer loadings indicate
more than one transition occurring (multiple gel phases), dependent
on cooling rate.

Some sort of maximum in the phase diagram is also evidenced at loadings

between 5 and 10 wt% polymer, where a local maximum occurs in the transition

temperature. This may be the region where the split in transitions is slowly
109
separating into two different distinct temperatures. From these various studies a

‘dynamic’ phase diagram is constructed in Figure 6.4 for all sol-gel transition

information at heating and cooling rates of 10ºC/min.

250
Transition Temperature (ºC)

200

150
(a)

100

(b)
50

0
0 0.2 0.4 0.6 0.8 1
Polymer Weight Fraction

Figure 6.4: ‘Dynamic’ Phase Diagram for Dow SPS/EB Binary System. Broken
lines indicate (simplified) phase boundary locations to the extent
data is available. (a) heating at 10ºC/min from gel, (b) cooling at
10ºC/min from sol.

6.2.4 Molecular Weight Variation Study

Different molecular weights of syndiotactic polystyrene were studied in

ethylbenzene (EB) to investigate how molecular weight affects the gel formation

110
dynamics. Three different molecular weight polymers were used, those

previously described in Chapter 4 for this work. Melting points of these three

gels were all measured at heating rates of 20ºC/min with DSC. The Dow and

0515 polymers had melting points of 105.4ºC and 105.0ºC, respectively. This

compares well with a predicted melting temperature of 109.5ºC. The very low

molecular weight polymer melted at 96.5ºC, about 10 degrees lower. Given that

its molecular weight is only 4,300, it is only about 40 times larger than the solvent
molecules, so perhaps not quite ‘polymer-like’ enough for Equation 2.19 to apply.

Figure 6.5 shows the formation responses of each of these polymers in EB solvent

at 1% loading. The lowest molecular weight polymer’s transition temperature is

less dependent on cooling rate that the two higher molecular weight species,

possibly due to crystallization from solution and not forming a “pro forma” gel.

111
45

40
Dow sPS, 41.5kMW
35 UT 0515, 34.1kMW
Transition Temperature (ºC)

UT 0918, 4.3kMW
30

25

20

15

10

0
0 10 20 30 40 50
Cooling Rate (ºC/min)

Figure 6.5: Various MW Polymers at 1 wt% in EB. Heavier polymers show


virtually direct translation as function of MW, very low MW
polymer (0918) indicating gelation process becoming less a function
of cooling rate.

6.3 SAMPLE POLYMERIZATION

Sample polymerization occurred when reactive solvents were used such as

styrene and tertiary butylstyrene (tBS). To prevent this from happening when

samples were subject to continuous periods of elevated temperatures, extra

tertiary-butyl catechol (tBC) inhibitor was added to the system when the gelling

polymer was being dissolved. Even with this extra inhibitor the sample

112
polymerized after repeated runs (thermal scans in the DSC). Figure 6.6 shows

what happens when the inhibitor has been ‘run out’ of the system, by the

systematic translation of the sol-gel transition peaks to higher temperatures.


Heat Flow (Endo Up)

(a)

(b)

45 55 65 75 85
Sample Temperature (ºC)

Figure 6.6: Extinction of Inhibitor in DSC Samples. Sample cycled at 8ºC/min


from 25ºC to 140ºC, 7 passes. First cooling pass (a), seventh cooling
pass (b). Transition temperatures open circles.

If a specific sample was suspected of undergoing polymerization, several

different samples were made and compared to gather the necessary data for each

weight fraction of polymer. Replicate experimental runs were done using

identical conditions on samples to test this theory with no significant hysteresis,

113
so the creation of multiple samples from the same initial solution yielded identical

results (i.e. no error introduced by making several different pans.)

6.4 EQUILIBRIUM TRANSITIONS

Measuring exothermic and endothermic transitions in the DSC was a

simple procedure when performed at constant heating or cooling rate, but this

methodology does not give (direct) measurement of what is happening on an


equilibrium basis. The melting of gel has been found to be virtually independent

of heating rate, whereas the formation of gel is quite dependent on the cooling

rate employed. Ideally isothermal analysis would be used to measure time to

transition at varying temperature setpoints, but was found to not be applicable for

the system under study and available equipment limitations.

To measure the sol-gel transition isothermally, the sample would have to

be melted, then rapidly cooled to the desired temperature to bring about gelation.

A wait would then ensue, and the transition would present itself at some

measurable time from the moment its temperature was brought below the melting

temperature. The isothermal temperature could be plotted versus inverse time to

determine an equilibrium temperature.


Unfortunately, it proved difficult to measure isothermal crystallization

transitions on the available DSC equipment due the rapid crystallization (and gel-

forming) kinetics of the systems under study. Several attempts were made at

isothermal observation of the sol-gel transition but the DSC controller lag

consumed the transition signal at subcooling setpoint temperatures of ten degrees

or more. In subsequent sections other methods are discussed and explored for

114
extrapolating to a zero cooling rate (Guenet ibid. 1988) as has been shown by

linear fits in Figures 4.10 and 6.2, previously.

6.5 LONG-TERM GEL AGEING

Gels formed by dissolving polymer into solvent then cooling to room

temperature exhibited excellent long-term stability, provided the container or

closure was not permeable to the solvent used. Although the gel is a physical
network it still possess a solvent vapor pressure very near that of pure solvent for

low concentration gels. Over several months to several years the gels slowly lose

solvent, shrinking in size to in some cases leave just pure polymer. A sample of

what was originally a 5 wt% CPS solution/gel in toluene was tested in the DSC

and found to be very rich in polymer, demonstrated in Figure 6.7 by its glass

transition and melting peaks comparable to those found in Figure 4.4.

It is still not entirely known if these types of gels will eventually lose even

more solvent at room temperature if given the opportunity. Presumably the gelled

network will lose solvent until the glass transition temperature reaches the storage

temperature for the sample given enough time. Given the behavior of most gels,

that would be a temperature about 75 degrees lower than that for pure polymer, so
a system with polymer content of about 75 vol% would have a glass transition

temperature of room temperature (per Figure 2.1).

115
Heat Flow, Endo Up

80 100 120 140 160


Sample Temperature, ºC

Figure 6.7: DSC of Two Year Old Toluene/CPS Gel. Lid of storage vessel
allowed solvent escape (via diffusion) of toluene solvent. *Unknown
actual composition. Suggesting heat-history and ageing-time
dependent phase behavior.

A 34% SPS polymer containing gel was made by dissolving 0.15 g of SPS

in about 10 mL of EB solvent, followed by solvent removal via vacuum until

nearly all solvent had been removed, giving the resultant polymer-rich material.

This hard waxy material was quickly ground to equalize any concentration

gradients or segregation that had taken place and reweighed to make the final

composition calculation. This sample was run in the DSC to see how its melting

temperature compared to that predicted by Equation 2.19 for SPS in EB. The

116
DSC melting and formation response is given in Figure 6.8 for this highly loaded

material at 10ºC/min heating and cooling rates. The melting point agrees

somewhat with the calculated value for this materials melting point of 132ºC, but

that depends on which transition peak you choose. Here two peaks exist, one at

about 109ºC and another at 139ºC.


Heat Flow (Endo Up)

(a)

(c)

(b)

0 50 100 150
Sample Temperature (ºC)

Figure 6.8: Gel-Sol and Sol-Gel Transition for Highly Loaded (34%) Dow
SPS/EB system. 30ºC/min heating rate after being quenched from a
sol, (b) 10ºC/min cooling rate, (c) 10ºC/min heating rate. Calculated
melting point value is 132ºC via Flory.

117
Of course there exist two signals depending on the sample’s heat history.

Apparently a slow cooled system segregates into different phases, which may be

in following behavior described by (Berghmans 1994). Highly loaded (>10%)

samples however appear to exhibit only one phase transformation in both

directions of heating and cooling if the temperature change rates are fast enough.

The same sample as that observed in the previous Figure was then subject to

further elevated rate cycles of 50, 70, and 90ºC/min heating rates as shown in
Figure 6.9, and the two peaks were observed merging into one peak.

(c)
Heat Flow (Endo Up)

(b)

(a)

60 80 100 120 140 160


Sample Temperature (ºC)

Figure 6.9: Elevated Temperature Change Rate Cycling on 34wt% Dow SPS/EB
System. (a), (b), and (c) heating rates of 50, 70, and 90ºC/min.

118
This phenomenon may have implications for an unusual phenomenon in

the forming gel’s modulus, observations that appeared when performing

viscoelastic studies on dilute polymer gelled systems, which are discussed in the

next chapter.

6.6 DISCUSSION

Since so much information was gathered studying physical gelation for


this work the preceding did not address all interesting issues regarding these

materials. More rigorous analysis and discussion will be presented in this section

regarding the observed thermal behavior of the gels and comparing that to

theoretical calculations for the same materials.

Studying gelled material in the DSC was a fairly straightforward task yet

the results were not always what might be expected. At the beginning of this

work it was not thought that the gelation phenomenon was so cooling rate

dependent. At this point however it is apparently so dependent that the cooling

rate may in fact affect what phase of gel the material ends up taking, introducing a

few complications for further work. Melting point information was not nearly as

heating-rate dependent but did not exactly occur as predicted via the theory set up
in Chapter 2.

6.6.1 Melting Transitions

To begin, Equation 2.19 is used to show via Figure 6.10 the predicted

(equilibrium) melting temperatures for six different solvents for which data was

gathered during this study.

119
270

250

230
Predicted Melting Point (ºC)

210

190

170
Toluene
150 o-Xylene
EB
130 Styrene
tBS
110
pMS
90
0 0.2 0.4 0.6 0.8 1
Polymer Weight Fraction

Figure 6.10: Six Different Solvents in Dow SPS, Calculated Melting Phase
Boundaries Using Equation 2.19.

This trend suggests that melting phenomenon is more a function of solvent

molar volume than anything else, as the Chi parameter (a function of solubility

parameter differences) caused little variation in liquidus line prediction. Table 6.2

compares what has calculated and what has been observed for these six solvents

at 1% polymer loading.

120
Solvent Tmelt observed Tmelt calculated Tobs-Tcalc
Toluene 111.0 91.5 19.5
o-Xylene 112.0 109.1 2.9
EB 104.4 109.6 -5.2
Styrene 115.0 98.4 16.6
tBS 92.0 172.9 (high!) -80.9
pMS 109.0 109.2 -0.2

Table 6.2: Comparison of Calculated vs. Observed Melting Points for Six
Solvents in 1 wt% Dow SPS, ºC. Three solvent/polymer binaries
agree pretty well, three not so well.

So much for the simplicity of the Flory equations for melting point

depression over a broad range of different solvents! It is probably coincidental,

yet fortunate, that the solvent most studied with SPS in this work, EB, agrees well

during this comparison. At first glance it appears that the behavior is generally

better predicted as the solvent becomes “more like” the polymer. More is

involved as there is most likely a dependence on the previous

gelation/crystallization conditions and heating history while cooling that affect the

temperature of the melting.

6.6.2 Gelling/Crystallization Transitions

The sol-gel transition at this point appears to depend on cooling rate

regardless of the solvent type or concentration. Earlier work (Guenet ibid. 1988)

suggested that several transition temperatures could be recorded and plotted as a

function of cooling rate, to enable [linear] extrapolation to an equilibrium or zero

cooling rate condition. This method and cooling rate regime as reported is shown

121
with observed data for styrene in Dow SPS in Figure 6.11, although the fits are

not very linear with R values as low as they are.


70
1%
2%
Sol-Gel Transition Temperature (ºC)

65 5%
10%

60

55

50

45
0 2 4 6 8 10
Cooling Rate (ºC/min)

Figure 6.11: Extrapolation of Sol-Gel Transition Point to Zero Cooling Rate for
Various Concentration of Dow SPS in Styrene. Guenet (1988)
suggests linear extrapolation to zero rate provided for ‘equilibrium’
formation temperature, as well as representing a maximum gelling
rate.

As demonstrated by the work on the pure polymer summarized in Figure

4.10, the condition of a discontinuity between melting temperature and

gelation/crystallization temperature does not in fact exist. As cooling rate is

lowered the sol-gel transition temperature “asymptotically” approaches the

122
equilibrium melting temperature. This test was not confirmed with dilute gel due

to the difficulty of detecting the exothermic gel formation peak when subject to

extremely slow cooling rates. Dilute systems must be cooled at least one-half

degree per minute in order to show a detectable transition, whereas pure polymer

cooled at 0.25ºC/min still crystallized about 15 degrees below its melting point of

271ºC.

As an alternative to extrapolating points in the relatively low cooling rates


regime, larger cooling rates can be employed and a linear region of formation

dynamics be identified from that. This is demonstrated in Figures 4.10 and 6.2.

Evaluation of many solvent/polymer combinations suggests the linear region for

this method is applicable for cooling rates of 20 to 50ºC/minute. This method is

only useful when a single phase transformation, either dilute polymer/solvent

systems or pure polymer, is occurring. The effect of concentration is shown in

Figure 6.12.

123
55

50
1 wt%
Transition Temperature (ºC)

45 3 wt%
5 wt%
40 10 wt%

35

30

25

20

15
0 10 20 30 40 50
Cooling Rate (ºC/min)

Figure 6.12: Sol-Gel Transitions Extrapolated for Four Concentrations of Dow


SPS in EB. Extrapolation range between 20 and 50ºC/minute
cooling rate.

The slopes of all of the extrapolations in Figure 6.12 are essentially the

same. These extrapolations are then compared for the extrapolation made for

pure SPS polymer over the same 20 to 50ºC/min cooling rate range (Figure 4.10)

in Table 6.3.

124
Material Tgel @ 10ºC/min Textrap % diff (abs) Slope, (min)
1 wt% 34.3 32.7 -0.521% -0.346
3 wt% 38.3 40.6 0.739% -0.418
5 wt% 46.2 47.2 0.313% -0.417
10 wt% 42.5 42.3 -0.063% -0.381
Pure 234.8 233.1 -0.335% -0.219

Table 6.3: Extrapolated Sol-Gel Transition Temperatures and Pure Polymer


Comparison. Four different concentrations and pure polymer shown.

This analysis provides a new way to extrapolate sol-gel transformation

information and have it agree within 1% of that for transitions measured at

cooling rates of 10ºC/min. The linear fit slopes for the 10 wt% polymer and

below all agree within 10% of each other, while that for pure polymer is a little

flatter at a little more than half the more dilute dynamic transition fit slopes.

One thing that should be investigated to test the validity of this proposal is

if this phenomenon is not just a nuance associated with this particular DSC’s

temperature controller, or the DSC’s ability to transfer heat in a rapid fashion.

Since all samples tested here are of more or less the same mass (10-15 mg) larger

and smaller masses would probably show if this were the case.

6.8 CONCLUSION

Various methods were employed to characterize physical gels of

syndiotactic polystyrene and aromatic solvents, with most work focusing on the

thermal behavior of SPS/ethylbenzene systems. An extrapolation was proposed

which is different from protocols found in the literature regarding extrapolating

rate-dependant thermal response to a theoretical “zero-rate” conditon.

125
Acquiring data of this type raised issues with whether the instrument was

capable of making such measurements as such high cooling rates and low

temperatures. Using the DSC intercooler along with experimentally verifying the

operating envelope of the instrument quelled these issues. A linear extrapolation

of rate-dependent sol-gel transition data results in a “zero-rate” very much in

agreement with the transition temperature indicated for 10ºC/min cooling

schemes. This information may be useful for designing processing equipment if


gel were to be produced on a large scale for controlling its temperature profile

during processing.

Chapter 7 delves into the mechanics of measuring and evaluating the

physical properties of dilute gels, both in the aged “solid” state, but as well as

real-time during the sol-gel transition with a nearly non-invasive technique.

These experiments produced very interesting and previously unreported

phenomenon, namely, the undocumented maximum in a gel’s modulus as it

initially forms.

126
REFERENCES
Berghmans, H.; Roels, T.; Deberdt, F., Macromolecules, 27, 6216, 1994

Birkinshaw, C.; Duff, S.; Tsuyama, S.; Iwamoto, T.; Fujibayashi, F., Polymer, 42,
991, 2001

Guenet, J.-M.; McKenna, G., Macromolecules, 21, 1752, 1988

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Nijenhuis, K., Thermoreversible Networks, Springer-Verlag: Berlin Heidelberg,


Germany (In English) 1997

Prasad, A.; Mandelkern, L., Macromolecules, 23, 5041, 1990

Tanaka, F.; Stockmayer, W., Macromolecules, 27, 3943, 1993

127
Chapter 7: Mechanical Gel Characterization

7.1 INTRODUCTION

The previous chapter dealt with understanding the thermodynamics of the

gel forming process; with that knowledge the mechanical properties of the gel

could then be investigated, both at long ageing times, but more importantly, the

properties measured as the gel is in its infancy. Since virtually no results of this

kind have been reported in the literature, this chapter represents some of the most

interesting findings in this entire work. The experiments and some of the tests

have been roughly modeled after the works of others (Aoki and Watanabe et al

1997-98), but their work focused on the behavior of PVC/DOP solution/gel

behavior. In addition, phenomenon previously uncharacterized in the literature in

any fashion has been observed, documented, and described in Section 5.

7.2 BULK POLYMER MODULUS/STRENGTH

Understanding the physical strength of the gels was very important for the

product development portion of this work. A material was desired that provided

excellent solids-suspension properties at extremely low polymer gelling agent

loadings, yet at the same time was easily sheared under applied force and

effectively “crumbled” into a highly flowable slurry or paste. Preliminary

observations of this phenomenon during proofing for product indicated it was

functioning quite well, but more detailed work has been performed in an attempt

to directly relate the effect of polymer concentration and molecular weight on gel

modulus and/or strength.

128
To possibly correlate bulk syndiotactic polymer strength with dilute gel

physical strength, injection molded “dog bones” were made from pure Dow SPS

for tensile testing. The tensile modulus was then converted to shear modulus (via

Poisson’s Ratio) for comparison to the shear modulii observed for various gels

measured via dynamic viscoelastic rheometry.

Bars were made on an Arburg 305 single-screw injection molding

machine with a standard ASTM six-inch “dog bone” mold with effective test
section of 0.125” x 0.500”. Since the polymer melts over a range centered at

270ºC, the final injection molder extruder barrel temperature was set to 300ºC.

Occasional plugging occurred, but machine operation was fairly smooth at these

conditions. Two different mold temperatures were used, 52ºC (125ºF) and 79ºC

(175ºF). Due to the crystalline nature of the polymer, it was expected that two-

phase parts (on a macro-scale) would be produced because the glass transition

temperature is greater than the mold temperature. Some of the produced parts

were annealed producing homogeneous parts; after this treatment the samples

appeared as homogeneous opaque samples, as the original polymer pellets do.

7.2.1 Two Macro-Phase Injection Molded Parts

Two-zone parts did result, quite amorphous transparent quickly-cooled

polymer at the specimen surface, and the slower-cooled center of partially

crystallized polymer. Simple inspection indicated the resultant parts contained 25

to 75% opaque volume in the centermost zones of the specimen (which is most

likely annealed fully crystallized polymer, 50-60% crystalline assumed, two

phases in itself, Birkinshaw 2001), irrespective of mold operating temperature,

129
determined by direct observation. This was evident by the cloudy opaque

polymer center surrounded by transparent amorphous syndiotactic polystyrene,

which was cooled fast enough by the mold surfaces to not allow for any

crystallization to take place. The entire specimen could be classified as still

consisting of only two phases, but with non-homogeneous distribution of

amorphous and amorphous/crystalline mixed zones. Tensile tests were performed

on an Instron machine with a 1,000 pound load cell and 1” linear strain gauge at
the standard strain rate of 0.200”/min. (for glassy polymers) moving crosshead at

room temperature

Several samples of two-phase parts exhibited brittle failure at about 1.9%

strain. Modulus was taken at 0.2% per standard for solid rigid samples, averaged

to 3.50 GPa. A stress-strain curve of this syndiotactic polystyrene as ejected from

the mold is shown in Figure 7.1.

130
70

60

50
Stress, MPa

40

30

20

10

0
0 0.5 1 1.5 2
Strain, %

Figure 7.1: Stress-Strain of Two-Phase Injection Molded Polystyrene. Mold


temperature of 175ºC, 25ºC testing temperature, 0.2 in/min strain
rate.

7.2.2 Annealed Injection Molded Parts

Although parts formed via injection molding are solid testable samples,

they are not necessarily representative or comparative to the gels under study due

to their multi-phasic condition after mold ejection as the gels under study are. As

far as can be ascertained, dilute gels are a single phase, at the macro-scale level.

To see what happens to the solid bars when they are at their maximum polymer

131
crystallinity, samples were aged in an over at 125ºC for over 24 hours and air-

cooled to room temperature.

Two different mold operating temperatures were used for parts that were

annealed, herein is shown a stress-strain for a parts molded at 175ºC and annealed

at 125ºC for 24 hours.


70

60

50
Stress, MPa

40

30

20

10

0
0 0.5 1 1.5 2
Strain, %

Figure 7.2: Stress-Strain of SPS Bars Aged at 125ºC for 24 Hours. Mold
temperature of 175ºC, tested at room temperature at 0.2 in/min strain
rate.

Polymer samples that were annealed to a more crystalline state as in

Figure 7.2 possessed an average modulus of 3.7 GPa, about 5% higher than for

unannealed parts. Ultimate strengths and maximum strain-to-break was higher

132
for annealed parts as well. Results of this bulk polymer property study are

summarized in Table 7.1. Shear modulus is found using the equation relating

tensile and shear modulus via Poisson’s ratio

E = 2G (1 + ν ) (7.1)

which is rearranged to find the shear modulus G

E
G= . (7.2)
2(1 + ν )

Sample 2-phase Annealed Annealed


Mold Temp, ºC 175 125 175
Anneal Temp, ºC X 125 125
Modulus, GPa 3.5 3.7 3.7
Ultimate Stress, MPa 50.5 57.9 59.0
Ultimate Strain, % 2.01 1.83 1.92
Shear Modulus, GPa 1.30 1.37 1.37

Table 7.1: Various Properties of Injection-Molded Syndiotactic Polystyrene


Under Varying Processing Conditions. Shear modulus calculated
assuming Poisson’s Ratio of 0.30 (Argon 1999).

7.3 DYNAMIC VISCOELASTIC RHEOMETRY

To perform a stress/strain analysis on a formed gel, samples were subject

to oscillatory strain on a Paar Physica MCR 300 Compact Rheometer at constant

frequency, increasing the strain until the sample failed. This was typically done

after hot-casting the sample between the measuring faces and then the measuring

133
temperature set accordingly. This is traditionally performed in a parallel-plate

geometry configuration at varying angular oscillatory velocities and frequencies.

Since the nascent (incipient, first-forming) gel is of rather low physical strength,

especially at low polymer loadings, a constant low frequency of 1.5 Hz was used

in all experiments. In addition since we are most concerned with an intact gel’s

shear modulus, cone and plate experimental geometry was used in order to obtain

a constant strain field throughout the testing medium regardless of radial position.

Material modulus was reported over shear strength, since measuring ultimate

strength resulted in wildly varying values. Normally measurements were done at

strain levels far below the yield point. On average, 1 wt% polymer gels failed at

2-3% shear strain, but sometimes as high as 7%.

7.3.1 Cone and Plate Theory

The theory for the cone and plate rheometer is given by Macosko (1994),

for a moving cone and stationary plate originally proposed by Mooney and Ewart

in 1934. The equations governing the cone and plate shear rheometry have been

previously given and described in Chapter 2, Section 3, Equations 2.20 through

2.38. This geometry is ideal for this analysis due to the constant shear strain

(before failure, elastic regime) and rate of shear (after failure, pseudo-viscous

region).

134
Illustration 7.1: Cone and Plate Rheometer. Not to scale sectional schematic of
the instrument and measuring surfaces; nominal sample meniscus
shown (providing ~2% edge volume error); cone apex truncated
50µ in axial length.

As a recap, for this system it is acceptable to assume:

1. Steady, laminar flow when liquid, constant shear strain when solid;

2. Cone angle ( β ) less than 6º (ours is only 1º), more important for

liquid samples;

3. Negligible body forces, and;

4. Spherical meniscus at the edge.

135
7.3.2 Sample Loading Considerations

The cone is carefully turned stainless steel with a 16 rms surface that

provides smooth intimate continuous contact with the sample. One drawback in

the case for studying gels (non-liquid materials) is the fact that sometimes the gel

would detach from the cone surface, resulting in the cone slipping along not

engaging the sample anymore. This only occurred in about 4% of tested samples.

It was suggested that modifying the measuring face surface(s) via sand or bead

blasting could improve adhesion even under normal stresses, but this was not

done as others were using the instrument at the time—their experiments requiring

the stock smooth measuring surfaces.

Cone detachment was more prevalent in richer gelled polymer

preparations. This was most likely due to the sample volume change (decrease of

1-3%) upon the sol-gel transition occurring, setting up varying normal forces

throughout the sample radius; in the center sample under compression while

farther out at thicker regions sample under tension. Cone disengagement hardly

ever occurred for samples containing less than 2 wt% polymer, so virtually all

studies on this instrument were kept in the low concentration regime (<2 wt%).

A test was performed to ensure the strains selected for elastic studies were

in fact in the elastic region for these materials. A 2 wt% solution of Dow SPS in

EB was cast at 70ºC, aged to 25ºC for 30 minutes, then cycled from 0.01 to 1.0%

strain three times with 10 minute pauses in between strain ramps. Figure 7.3

136
shows how the storage and loss modulus respond during the three 15 minute

passes.

1E+05

Storage, G’
Modulus (Pa)

1E+04
1st Pass
Loss, G”
2nd Pass
3rd Pass

1E+03
0.01% 0.10% 1.00%
Strain (%)

Figure 7.3: Elastic Regime Cycling of 2 wt% Dow SPS in EB. 70ºC load, aged
30 minutes to 25ºC, each strain ramp lapsed 15 minutes, 10 minute
pauses between ramps; filled shapes: storage modulus (G’), open
shapes: loss modulus (G”); (convention within this work).

The fact that the third pass appears slightly degraded may have something

to do with the samples phase behavior upon continued ageing, not due to

mechanical disturbance effects (discussed later). Aside from that the elastic

nature of the gel is rather evident due to the traces overlapping one another after

137
being highly strained (to ~50% of known yield strain), although this preliminary

test is not definitive as the gel could possibly be “healing” itself of any micro-

fractures that took place due to the gel bonds being of near-room-temperature

energy (Chapter 2).

Solvent loss was another consideration in the loading and testing of gels

using the cone and plate viscometer geometry. The greatest uncertainty occurred

when sol samples were loaded from hot solution into a hot instrument. It was

necessary to either observe the sample dynamically as it progressed through the

sol-gel transition, or to statically cast the sample between the cone and plate and

then begin observation after an equilibrium temperature and/or gel had been

obtained. Measurements were made via both methods; both were successful if

temperatures above room temperature were minimized and/or lower vapor

pressure solvents were employed in the samples. Solvent loss with sample

loading was determined to introduce less than 2% composition variance. This

was determined by estimating the change in volume of the meniscus of the sample

between the cone and plate. Elevated observation temperatures (>35ºC) or

experiments longer than one hour usually caused excessive solvent loss, to the

point of the sample virtually vanishing from between the measurement faces. The

stock solvent-cup and cover were used to provide a solvent-rich atmosphere to

minimize this effect. A three-inch diameter O-ring was placed on the plate

surface and 15-20 drops of solvent introduced, which the O-ring held in place via

138
surface-tension. Since this solvent was in direct contact with the temperature

controlled plate yet not in contact with the sample, this provided a better saturated

atmosphere and further reduced solvent loss from the sample.

Thermal effects after sample loading were another issue that was daunting

for some time. The carbide-coated bottom planar measuring face also doubled as

a –30/+70ºC Peltier device that could achieve and control with great accuracy

temperature swings of 100ºC in only a couple of minutes. The issue of overall

temperature control is still as issue however, since the Peltier plate only provides

for temperature control at the very bottom of the sample. Some lag or response

time was obviously present, since both the sample and the measuring cone are

comparatively poor thermal conductors. It was initially estimated that in the

temperature ranges of interest (0 to 80ºC) the response time of this system was

about 5 minutes.

7.3.3 Thermal Response Study of Instrument

The Paar rheometer uses a Peltier temperature control system on the plate

(bottom) surface of the measuring apparatus. The computer program allows one

to program a single temperature and/or temperature profile while measurements

are being taken. At issue is the response of the instrument to temperature

changes, namely what is the response time of the entire measuring system to

program changes in temperature?

139
To measure our instrument’s response to temperature changes, an external

thermocouple was secured to the top surface of the measuring cone normally used

for this work. The machine was prepared and preheated as if a normal gel

rheological measurement was to be taken, only with a 0.300 mm gap instead of

the 0.050 mm final measuring gap (theoretical cone point touching plate). The

plate was heated to 50ºC for 10 minutes, then different temperature profiles

programmed to determine the lag the cone’s temperature experienced in response


to changes in the plate temperature with 1% EB gel cast between the surfaces.

The thermocouple taped to the top surface of the cone represents a “worst case” of

measuring temperature, meaning the actual differential cannot be larger than what

is measured by this method. The thermal conductivity of the stainless steel cone

is at least 30 times greater than that for either the 1% SPS/EB sol or gel, though,

so it is also representative of the cone’s bottom temperature (sample’s top

temperature). Figure 7.4 shows the response of a step change in plate temperature

from 50 to 25ºC.

140
55
50
45 Plate Temp
Temp (ºC)

40 Cone Temp
35 Program Temp
30
25
20
0 50 100 150 200 250 300
Time (sec)

Figure 7.4: Step Change from 50 to 25ºC. 1% SPS in EB gel cast between
surfaces and the temperatures monitored over several minutes.

Then another test was done, the same way, this time assigning a 2 minute

ramp to bring the temperature down to 25ºC (via stepping to 45, then ramp to 25,

Figure 7.5).
55
50
45 Plate Temp
Temp (ºC)

40 Cone Temp
35 Program Temp
30
25
20
0 50 100 150 200 250 300
Time (sec)

Figure 7.5: 2-minute Temperature Ramp-Down Program. Cone keeping better


pace with the plate now.

141
Finally in Figure 7.6, a 4-minute ramp on the temperature program, this

time the measured cone temperature tracking the plate temperature almost

exactly.
55
50
45 Plate Temp
Temp (ºC)

40 Cone Temp
35 Program Temp
30
25
20
0 50 100 150 200 250 300
Time (sec)

Figure 7.6: 4-minute Temperature Ramp-Down. Near-perfect agreement of


cone temperature with plate and program temperature this time.

With this knowledge of the temperature response, all samples were cast by

bringing the temperature down at 10ºC/minute cooling rate unless otherwise

noted. It should be noted that since all of the liquid loading, casting, and

measurement operations on the DVR took place at temperatures that were below

the melting point for the gels. All transitions occurred due to kinetic processes,

not equilibrium processes. The initial loading temperatures (of ~50-60ºC) were

selected for these studies to provide enough time to prepare the machine to cast a

good sample while not too high of a temperature for solvent loss through

evaporation to be a problem over time scales of the experiments. The response

time is actually far less than the estimated 5 minutes; it is now evidenced that the

142
response time is only about a minute when a step-change in program temperature

is made.

7.3.4 Testing Procedure

To prepare the rheometer for testing of a gel, the measuring faces were

carefully cleaned with acetone, assembled, and the faces brought to within 0.300

mm of their eventual testing position (so the cone slightly raised). Liquid samples
were loaded at elevated temperatures; a close gap allowed the cone to be heated

radiatively and convectively from the Peltier flat bottom plate while leaving

enough space for liquid sample to be drawn in without entraining air. The plate

temperature was selected such that the sample could be injected between the

plates, the cone lowered, and testing program initiated before any gel formed.

This was performed to ensure that a homogeneous, continuous sample with no

fractures in it was produced between the measuring surfaces. This was typically

taken to be 10 or 15 degrees above the expected formation temperature, usually

between 45 and 60ºC, as previously ascertained via DSC at moderate cooling

rates (~20ºC/min). The temperature was set and held for 10 minutes, whilst the

testing program was entered into the controller and the sample prepared.
When it was time to begin a test, first the gel sample was made liquid by

rapid heating with a propane torch. Then a small amount of pure solvent was

placed in the reservoir above the moving cone (intended for this purpose) to

maintain a solvent-rich atmosphere around the testing locality. The gel’s

resistance to imbibing more solvent was relied upon to assume that no additional

solvent was adsorbed into it, that along with the rather small relative exposed area

143
at the cone/plate meniscus. The sample weight may have also been verified if it

was the first draw for that particular sample, then 0.62 mL immediately

withdrawn from the sample vessel and pipetted between the plates. Pipet

dispensing took about two seconds to accomplish. It was done as quickly as

possible to minimize solvent loss yet slow enough to allow the liquid to be drawn

between the measuring faces via surface tension and not pooling away from the

cone/plate gap.
After injection the cone was immediately lowered to the testing position of

0.050 mm, the solvent cover installed, and the program initiated. The amount of

time it took from sample injection to program start was always less than 15

seconds.

7.4 DVR CASE STUDIES

Several different dilute syndiotactic polymer/solvent combinations were

tested in the cone and plate instrument. The most interesting results have been

documented and described here, with additional results further discussed and

presented in Section 5 of this chapter.

7.4.1 Dow SPS in Styrene

A small amount of Dow SPS polymer was diluted in styrene monomer

solvent and tested at 1.5 Hz under the cone and plate geometry. This test was an

early attempt to correlate system concentration with bulk modulus, as well as

gradually increasing the applied strain until the material failed. The data for four

different concentrations of polymer in styrene are given in Figure 7.7.

144
1.E+04

0.38 wt%
0.56 wt%
0.86 wt%
1.E+03 1.05 wt%
Modulus (Pa)

1.E+02

1.E+01
0.001 0.01 0.1 1
Strain (%)

Figure 7.7: Storage and Loss Response and Failure Point of Dow SPS in
Styrene. Loaded at 70ºC, immediate cool to 25ºC, aged 30 minutes,
1.5 Hz frequency, 30 seconds per point, 30 points, filled symbols G’,
open G”. Samples all failed in 1-3% strain range.

This test does not directly correlate gel strength with polymer

concentration; the strongest material is actually the second lowest concentration

material at 0.56 wt% polymer. This of course is not expected. All these materials

exhibited shear failure at between 1% and 7% strain, with the three richest gels

failing at an average of 2% strain.

145
7.4.2 Dow SPS in tert-Butyl Styrene

Several DVR studies were performed with Dow SPS in tBS as this work

has always had a keen interest in the properties of tBS, since it was the mainstay

of our product development. It also has an excellent potential future an

engineering thermoplastic polymer. Although our product incorporated Dow

copolymer for its depressed transition temperatures, syndiotactic polymer was

employed here as that is the focus of this academic study.


First we are interested in the strength of resulting tBS gels at room

temperature. Several gel concentrations were made and aged then subjected to a

constant strain of 0.1% at 1.5 Hz for about an hour, as in Figure 7.8.

146
14000

12000
2.3 wt%
1.2 wt%
10000 0.62 wt%
Storage Modulus (Pa)

8000

6000

4000

2000

0
0 10 20 30 40 50 60
Time (min)

Figure 7.8: Various Concentrations of Dow SPS in tBS Monomer. Loaded at


60ºC, immediately cooled to 25ºC while measuring began, 0.1%
strain, 1.5 Hz.

It is notable that for the 0.62 wt% SPS in tBS system gel forming modulus

response in the previous Figure, it is the only observed case where the forming gel

did not exhibit a maximum in its modulus while ageing (and was confirmed via

several more like tests). The two other higher concentration gels are interesting in

that their equilibrium modulus is a greater percentage than that for EB solvent,

147
lending more information to the fact that each and every polymer/solvent binary

in this study is quite different in its phase behavior.

Starting with a 2.3 wt% SPS in tBS solution, gel was cast and the

temperature immediately dropped to 15ºC, then systematically raised per the

temperature profile shown in Figure 7.9.


80000 60

70000
50
60000

Sample Temperature (ºC)


40
50000
Modulus (Pa)

40000 30

30000
20
20000
10
10000

0 0
0 20 40 60 80
Time (min)

Figure 7.9: Dow SPS 2.3 wt% in tBS aged at various temperatures per the
shown profile, 2 replicates. This time the modulus drops far below
the maximum reached upon initial formation.

This is not anything like what was observed in the previous test using tBs

as solvent; 10ºC lower initial ageing temperature made the gel get far weaker,

148
suggesting some sort of activation temperature may have been crossed. Another

test was done this time reversing the ageing temperature profile order (Figure

7.10).
40000 60

35000
50
30000
40

Temperature (ºC)
25000
Modulus (Pa)

20000 30

15000
20
10000
10
5000

0 0
0 20 40 60 80
Time (min)

Figure 7.10: As in Previous Figure, 1.9 wt% Dow SPS in tBS Aged at Three
Different Temperatures. 2 replicates, 0.1% strain, 1.5 Hz

And finally a down-up (35/15/35) temperature profile of this system

(Figure 7.11).

149
70000 60

60000
50

50000
40

Temperature (ºC)
Modulus (Pa)

40000
30
30000

20
20000

10
10000

0 0
0 20 40 60 80
Time (min)

Figure 7.11: 2.4 wt% Dow SPS in tBS Stepped Temperature Ageing. Loaded at
60ºC, stepped down to 35ºC.

The past several Figures demonstrate that there is quite a lot involved in

obtaining repeatable observations, and obviously there are instrumental/software

control issues given the discontinuities in measured modulus with simple changes

in program temperature (which is not instantaneous). At this point this binary

system is left as providing quite interesting phenomenon and not delved into

further. At the very least the technique for measuring properties such as these

probably has some problems. This most evident by the step changes in

150
temperature resulting in immediate artifacts in the measured modulii, most likely

an instrument control and/or software issue.

7.4.3 UT 0515 SPS in Ethylbenzene

Higher molecular weight UT produced syndiotactic polymer was tested in

EB at two concentrations, 1 and 2 wt%. This study investigated the potential

affect that ageing temperature has on the maximum and equilibrium modulus for
the gel. Three different temperature ageing levels were selected from the loading

temperature, done at either 55 or 60ºC. Temperature was slowly dropped from

load to age over 15 minutes, then the sample aged for a further 45 minutes.

Longer ageing would be desirable for this exercise but solvent loss from the

sample became an issue at that point. Figure 7.12 shows the response of various

ageing temperatures at 1% concentration.

151
120000
60
100000
50

80000

Sample Temperature (ºC)


40
Modulus (Pa)

60000
30

40000
20

20000 10

0 0
0 10 20 30 40 50 60
Ageing Time (min)

Figure 7.12: Ageing 1 wt% UT0515 in EB for 60 Minutes, Three Age


Temperatures of 5, 15, and 25ºC, 0.1% strain at 1.5 Hz. Notice at
gel incipient (all about 40ºC) the sharp rise in modulus, slight drop,
then larger rise to a peak then falling off.

The trend to note here is how broad the peaks are for the maximum

modulus rolling over. Although the maximum values vary nearly 20%, the higher

the ageing temperature the faster the modulus departs from the maximum value.

Table 7.2 summarizes these results for both the 1 and 2 wt% concentrations of

UT0515 in EB. Since these tests were only ran for 60 minutes (due to solvent

loss problems) the %-drop values in this Table are not equilibrium values. The

152
lower %-drop numbers for 2% loaded material may be a result of the slower

equilibration of the richer system due to its denser chain structure.

Conc. Tage, ºC Modulus (max), Pa Modulus (60min) % drop


1% 25 85,000 63,000 25.9%
1% 15 100,000 80,000 20.0%
1% 5 89,000 88,000 1.1%
2% 25 210,000 197,000 6.2%
2% 15 190,000 183,000 3.7%
2% 5 X (detach) X X

Table 7.2: Results of Ageing UT0515 in EB at Various Loadings and Age


Temperatures. Ramped from loading temperature to setpoint over
15 minutes, held for additional 45 minutes. 2% sample cooled to
5ºC detached both times from cone while observing.

7.4.4 UT 0918 SPS in Ethylbenzene

The lowest molecular weight polymer used in this study was tested using

ethylbenzene at 2.0 wt% UT0918 polymer. Preliminary benchtop gel-forming

propensity tests indicated this polymer required far longer to form a gel (at room
temperature), so far lower temperatures than normal were employed in the DVR

tests performed. Two runs were done using the same 2% composition pot at two

different ageing temperatures, as shown in Figure 7.13.

153
16000 35

14000 25

Sample Temperature (ºC)


Storage Modulus (Pa)

12000 15
to -20ºC
to 0ºC
10000 5

8000 -5

6000 -15

4000 -25
0 10 20 30 40 50 60
Time (min)

Figure 7.13: UT 0918 Modulus vs. Time for Two Different Ageing Temperatures.
Loaded at 30ºC, ramped to ageing temperature over 15 minutes,
0.1% strain, 30 seconds/point, 1.5 Hz. Erratic modulus measured
due to discontinuous sample medium.

What is immediately apparent in this figure is the noise in the measured

modulus as a function of time. The material is still exhibiting solid-like

characteristics due to its high modulus, but is more like a concentrated slurry of

hard particles grinding around in a lower viscosity matrix of solvent and some

polymer. Due to the erratic response of the material’s modulus, per the original

definitions this would not necessarily be a gel, so there is probably some

154
minimum molecular weight that some observable difference between a true gel

and this behavior represents. Since the gap is rather large between the polymers

studied in this work (4,300, 34,000, and 42,000), more polymer samples of

molecular weights between these numbers, especially lower ones would be useful

for determining if that is the case.

7.5 DISCUSSION

Over 260 individual samples were tested in the cone-and-plate instrument

under various conditions. What quickly became evident consistent measurement

of properties such as elastic shear modulus and shear strength were rather

difficult. Samples that were too rich in polymer (>2 wt%) exhibited so much

volume change upon gelation (shrinking) that high normal forces on the

measuring faces frequently caused dewetting/detachment of the material from the

measuring face. In addition solvent loss was an issue for tests run for over an

hour or two. An improved saturated-solvent environment was provided for late in

this work by trapping solvent on the measuring face with a thin O-ring ½” larger

in diameter than the measuring cone.

7.5.1 Unforseen Multiple Phase Behavior

Oscillatory testing at low strains did make known another rather

interesting phenomenon. It is already known from previous thermal studies that

highly-loaded SPS/solvent systems exhibit more than one phase transformation as

a sample is cooled, this being dependent on the cooling rate among other things.

This author proposes that this is also happening at low polymer concentrations as

well, even if not observed directly via thermal means. This is thought to be

155
evidenced primarily by when measuring a gelling sample directly after injection

before a gel has formed at very low concentrations and low (0.1%) strains, the gel

is apparently stiffest upon its initial formation, then its strength plateaus to some

fraction of the maximum strength (Figure 7.14).


35000

30000
15ºC
25ºC
25000
35ºC
Modulus (Pa)

20000

15000

10000

5000

0
0 10 20 30 40 50 60
Age Time (min)

Figure 7.14: Monitoring Formation of 0.52 wt% Dow SPS/EB Gel, 0.1% Strain,
1.5 Hz. Loaded at 70ºC, measurements started immediately while
temperature approached noted setpoint, 30 sec/measuring point.
(25ºC run measurement not begun for 3 minutes after
load/temperature set.)

When this “gel-weakening” was first observed it was attributed to the

oscillating cone (albeit very little) interfering with the formation of gel via a non-

156
quiescent state. Another test was done this time pulsing the measuring activity

with periods of when no oscillation was taking place. However, the phenomenon

was still observed as shown in Figure 7.15.


20000

16000
Modulus (Pa)

12000

8000

4000

0
0 20 40 60 80 100 120
Time (min)

Figure 7.15: Pulsed Oscillatory Measurements on Nascent 0.54% SPS/EB Gel.


Loaded at 70ºC, measurements started 3 minutes after loading while
cooling to 25ºC, 30 seconds/measuring points. 5 minute of
measurement alternating with 15 minute pauses, 2 replicates.

It was never independently confirmed whether this gel weakening is due

to the gel going through more than one phase of not, but a little independent

evidence may have been found to support this theory. Prasad (1990) speaks of

157
observations in which ageing temperature affects formation time and gel quality.

For his SPS/trans-decalin systems at a given polymer content, solid gels resulted

from higher cooling temperatures, whereas solution crystals formed from lower

cooling temperatures. Since the rapidity of temperature change is not given near

step-change is assumed during his tests.

It is reasonable to assume that the bulk strength and/or modulus of a

discontinuous medium would be less than a continuous one of the same material,
hence why lower temperatures and/or longer age times result in a gel that forms,

then “breaks itself” into a bunch of little pieces. Since the physical gel bond

strengths have been reported (Guenet ibid. 1992) to be rather weak, of order kT,

then it may be capable of doing this even at low temperatures. Another

possibility is that gel-forming polymer (intercrystallite/interconnecting chains) is

removed from the gel to further form crystals (intracrystallite/crystalline chains)

as the gel ages. Since the associative bonds are demonstrably weak, this is

certainly possible that the gel is “tearing itself apart”, reducing chain

entanglements as well weakening the bulk gel material.

7.5.1.1 Frequency Sweep of Aged Gel

One test to ensure the machine is capable of making the measurements it

is reporting is to perform a traditional frequency-sweep on an aged gel material.

To accomplish this the same 1% SPS/EB system was employed, cast at 50ºC,

cooled to 25ºC, then aged for 30 minutes with no measurements. Measurement

was then commenced over a range of frequency from 0.5 Hz to about 70 Hz,

taking the instrument to its limits. Figure 7.16 shows the aged gel’s response is

158
virtually 100% Hookean, and a further test on the same sample two hours later

overlays the first trace.


1E+05 1E+04

Complex Viscosity (Pa*s)


1E+04 1E+03
Modulus (Pa)

1E+03 1E+02

1E+02 1E+01
0.1 1 10 100
Frequency (Hz)

Figure 7.16: Frequency Sweep 0.5 to 70 Hz at 0.1% Strain of 1% SPS/EB


System. Aged 30 minutes (diamonds); aged additional two hours,
then re-run (circles).

7.5.1.2 Variation of Measurement Start-Time

To further test if the peak in modulus is an artifact or induced by the

slightly-moving/vibrating cone, sample was loaded, cast, then motion started at

varying delays. So in this instance, no sample disturbance at all until the

159
programmed start time. Figure 7.17 is one such experiment, done three times at

three separate delays (0.5 min, 1.5min, 2.5min) at 90 Hz.


1E+05 50

8E+04
Storage Modulus (Pa)

→ 40

Temperature (ºC)
6E+04

4E+04
30

2E+04

0E+00 20
-120 0 120 240 360 480 600 720
time (sec)

Figure 7.17: Varying Start Time of Oscillation After Cast, 90Hz. Three different
delays, 0.5, 1.5, and 2.5 minutes. These three start times encompass
the peak in modulus.

7.5.2 Optical Observation of Sol-Gel

In order to possibly discern what was happening while the gel aged from a

sol to a room temperature gel, a dilute gel of 2wt% SPS in EB was cast between

two glass slides and sealed with metal tape. This slide was then placed in a

160
controlled hot-stage microscope set to 60x magnification. The temperature was

cycled to 120ºC at 10ºC/min, held for one minute, then cooled back to room

temperature at 10ºC/min.

Upon heating the gel was observed going to a transparent sol, which also

allowed for precise setting of the focal length onto the sample. A gel was

observed to form at about 45ºC, as expected via previous DSC measurements

(although by peak-interpretation convention DSC gelation is reported at 35ºC).


Small square domains of presumably gelled polymer nucleated and grew to fill

the field of view, but as the temperature progressed down no significant change

could be observed after initial gel growth. Potential multiple-phase behavior was

not directly observed (or observable anyway) via this contrivance.

A further test was done optically, this time using a wide-band UV-vis

absorbance meter. A 1 cm pathlength cuvette was charged with 1% SPS/EB sol

and observations immediately started. Since the sample holder was an aluminum

block and at room temperature, the sample was quickly cooled, so observations

were made every 5 seconds (no temperature control and/or knowledge). Only one
run was performed, and the time-resolved results of this are summarized in Figure

7.18. An initial inspection of these plots show two different processes may be

going on during the sol-gel transition and immediately thereafter. After several

minutes the material had aged into its equilibrium translucent waxy-gelled state.

161
t = long

t = 45 sec
Absorbance (normalized)

t = 35 sec

t = 30 sec

t = 25 sec

t = zero

100 200 300 400 500 600 700 800 900


Wavelength (nm)

Figure 7.18: Time-Resolved UV-Vis Observation of 1% SPS/EB System. Two


different phenomena in absorbance can be gleaned from this first
investigational run.

7.5.3 DVR Instrument/Measurement Variance

It was noticed that this particular method of measuring the gel’s modulus

via cone/plate geometry resulted in traces that did not always overlay for repeated

runs under the same conditions. This could be related to crystal nucleation and

particle growth issues within the gelling sample. Even with the cone/plate

stationary there are still other external factors contributing to things which may

162
lead to variance, although those which could reasonable be assumed (sample fill

volume, saturation environment, [bulk] temperature issues) were found to not be

of sufficient magnitude to lead to such non-repeatability of certain measurements.

For this reason the associated error of making these measurements as reported

here is given, along with a calculated confidence interval of a typical

measurement. Figure 7.19 shows the results of nine identical runs using 1% SPS

in EB.
1.6E+05 65

1.4E+05
55
1.2E+05
Storage Modulus (Pa)

1.0E+05

Temperature (ºC)
45

8.0E+04

35
6.0E+04

4.0E+04
25
2.0E+04

0.0E+00 15
0 300 600 900 1200
time (sec)

Figure 7.19: Nine Identical 0.1% Strain Storage Modulus Observations. 1%


SPS/EB gel, temperature slowly ramped down as indicated, initial
modulus signal occurred very repeatedly at 38ºC plate temperature.

163
These nine runs are combined to generate a 90% confidence interval for a

measurement of this type. Figure 7.20 shows the average, upper bound, and

lower bound, assuming a 90% confidence interval.


1.6E+05

1.4E+05
Average
Upper 90%
1.2E+05
Lower 90%
Storage Modulus (Pa)

1.0E+05

8.0E+04

6.0E+04

4.0E+04

2.0E+04

0.0E+00
0 300 600 900 1200
time (sec)

Figure 7.20: 90% Confidence Interval of 0.1% Strain Experiment. Most variance
appears to be in when the peak rolls over and how broad the
downside of the modulus response is.

7.6 FUTURE WORK

Future study of physical gels stemming from this work most certainly

should entail first learning more about the time-temperature-concentration

164
behavior of syndiotactic polystyrene in various low molecular weight aromatic

solvents. Many of the studies for this work used SPS in ethylbenzene (EB). EB

was chosen because it has virtually the same shape as the repeat units on the

polymer, simplifying the thermodynamics, yet not reactive like its

dehydrogenated version styrene (high temperatures requires lots of extra inhibitor

additives). It was also heavy enough that its vapor pressure did not significantly

contribute to mass-balance problems via evaporation except over several hour


experiments in open systems.

More samples of different molecular weights would be very useful in

determining properties that depend on a minimum molecular weight. Of note is

the condition of a gel existing or not, per the definition given in Chapter 2. Of

even greater interest would be syndiotactic polystyrenes of many different

molecular weights from 4,000 to 60,000, each usable sample of say a

polydispersity of 1.5 or less. Separating a polymer sample into enough usable

fractions (using chromatography) would be extremely expensive however.

7.7 CONCLUSION

This final formal chapter describes the most novel work done by this
author, most notably the observation of a forming aromatic/SPS gel exhibiting a

maximum in modulus when it initially forms. Hundreds of runs were performed

to ensure these observations were not in fact artifacts of some sort, as well as

other tests in an attempt to verify this unusual response in modulus. It was never

directly determined whether this is a direct result of some suttle phase change in

the gel from the time it initially networks/forms to a longer ageing time.

165
Inspecting the material in bulk shows it does form a clear solid from liquid, then

the solid proceeds to become cloudy.

As this unexpected response in measured modulus was interesting, other

tests had to be performed to ensure it was in fact real. An investigation into the

instruments temperature response revealed it controlled and maintained

temperature with response times far shorter than the time scales involved in the

evolution of the modulus maximum. A brief UV-vis analysis indicated a few


different processes going on but due to time limitations was not investigated

further. Finally, traditional frequency-sweep experiments were done on cast aged

gel to ensure the instrument was producing reliable numbers. Since the gel gave a

virtually Hookean response (as expected) the results can be deemed reliable. Of

course further 100% non-invasive investigation into the phase behavior of these

gels in the future will shed more light on what is really going on.

Chapter 8 will summarize the research as it has been presented in this

work. The topic will be the synthesis and characterization of both pure polymers

and polymer gels, in addition to some overall conclusions and recommendations.

166
REFERENCES
Aoki, Y.; Li, L., Macromolecules, 30, 7835, 1997

Aoki, Y.; Uchida, H.; Li, L.; Yao, M., Macromolecules, 30, 7842, 1997

Aoki, Y.; Li, L., Macromolecules, 31, 740, 1998

Aoki, Y.; Li, L.; Uchida, H.; Kakiuchi, M.; Watanabe, H., Macromolecules, 31,
7472, 1998

Aoki, Y.; Li, L.; Kakiuchi, M., Macromolecules, 31, 8117, 1998

Argon, A.; Cohen, R.; Patel, A., Polymer, 40, 6991, 1999

Berghmans, H.; Roels, T.; Deberdt, F., Macromolecules, 27, 6216, 1994

Guenet, J.-M.; McKenna, G., Macromolecules, 21, 1752, 1988

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Macosko, C., Rheology: Principles, Measurements, and Applications, VCH


Publishers: New York, 1994

Nijenhuis, K., Thermoreversible Networks, Springer-Verlag: Berlin Heidelberg,


Germany (In English) 1997

Watanabe, H.; Sato, T.; Osaki, K.; Aoki, Y.; Li, L.; Kakiuchi, M.; Yao, M.,
Macromolecules, 31, 4198, 1998

167
Chapter 8: Summary and Conclusions

8.1 INTRODUCTION

This chapter will serve as summarizing everything done and observed in

this work. In addition to summarization, additional conclusions will be given, as

well as any recommendations where appropriate, both for future work, as well as

how work herein may have been performed differently.

8.2 GELLING POLYMERS

Two different sources were utilized for polymers in this work. Dow

Chemical Company provided the “seed” polymer, which piqued our interest with

its unusual solution phenomena. Additional polymer specimens were synthesized

in-house in order to provide different molecular weights to obtain data as a

function of molecular weight. Although more often than not the higher molecular

weight Dow polymer was used in forming and characterizing gels, this was due to

the extremely small quantities of in-house polymers that were successfully

synthesized (after proper purification).

8.2.1 Catalyst Synthesis

Initial inspection of the literature showed many groups successfully

synthesizing syndiotactic polystyrene in the laboratory. What became quickly

evident though was that many of the catalysts necessary to make polymer were

not commercially available, so these in turn had to be synthesized from other

chemicals as well. For this work catalyst synthesis was moderately successful,

168
yet was not ever an end goal of this work. For this reason extensive catalyst

characterization was not performed, although it would be desirable to characterize

the catalysts in the future to ensure that the isolated product is indeed what it was

herein presumed to be.

8.2.2 Polymer Synthesis

Several successful polymerizations were carried out using two different


primary polymerization catalysts that were known to lead to a syndiotactic

homopolymer of polystyrene. One catalyst being commercially available, the

other synthesized according to previously described procedures. Several different

molecular weight samples were produced, spanning a range of about 2,000 to

35,000. Due to polymerization yields that were rather low (as expected per

literature); only a few samples of sufficient mass were isolated to carry out further

experiments on.

8.2.3 Polymer Characterization

Polymers that were to be used were characterized with differential

scanning calorimetry (DSC), gel permeation chromatography (GPC), and nuclear

magnetic resonance (NMR). Obtaining this characterization information allowed

gels that were subsequently made with the polymer to be then be properly

characterized.

Thermal studies performed on the various polymers under study via DSC

resulted in thermal responses which agreed very well with previous literature for

these materials. The same high melting point was observed (270ºC) for in-house

synthesized polymers as well as for the provided Dow polymer product. Glass

169
transition temperatures were all measured to be around 100-105ºC. Erratic

thermal response for a polymer that has been instantly quenched from the melt or

solution to below its glass transition temperature has also been observed,

indicating that this polymer, although having a high propensity to crystallize, can

in fact be “trapped” in a virtually pure amorphous state. Annealed polymer

samples (or those that have been slow cooled from the melt) exhibit sharply

observable phenomenon for both the glass transition and melting endotherm.
GPC was used to determine polymer number average and weight average

molecular weights, and polymer polydispersity. Polydispersities ranged around

3.0-3.5, with number-average molecular weights ranging from 3,000 to about

50,000 for all polymers under study. Of interest to this work is that a room

temperature method was used on ordinary equipment using THF solvent instead

of multi-chlorinated benzenes at 100ºC above ambient as is typical in the

literature. Although questions were raised as to the viability of this method,

calculations and experiments verify that the results obtained here are reliable.
13
NMR, namely C-NMR, was successfully employed in order to verify
tacticity and stereoregular purity of the polymers under study. This method of

tacticity discernment and verification was rather common in the literature,

although no citation actually mentioned the specific deuterated solvent used for

the reported NMR analysis. Syndiotactic polystyrene solutions are always, as far

as is known, a solid/gel at ambient temperature, so elevated temperature in the

probe was obviously necessary. Deuterated toluene was tested, but the gel would

not melt under the quiescent conditions in the NMR probe; deuterated xylene did

170
however produce a liquid polymer solution at 120ºC which resulted in definitive

experimental results. All polymers suspected to being highly syndiotactic were in

fact verified to have greater than 99% purity in syndiotactic diads.

8.3 GEL CHARACTERIZATION

Making gels was a straightforward task: combine the desired amounts of

polymer and solvent and boil until dissolved. This required high temperatures so
when monomer solvents were used the high-temperature exposure time was

minimized. After gels had been successfully made, they were characterized via

several available methods, namely DSC and dynamic viscoelastic rheometry

(DVR).

8.3.1 Thermal Studies

DSC studies of gels provided information regarding the thermal behavior

of gels, both going from a gel to a sol and vice-versa. This resulted in the

generation of a preliminary phase diagram for the SPS/EB system, as well as low-

loaded transition information for several other solvent/polymer binary systems.

By varying the cooling rates while observing binaries going from the sol

to a gel, transition-temperature vs. cooling rate figures were prepared. Contrary

to existing literature in which some authors speak of extrapolating low cooling

rate data (5-20ºC/min) to a zero cooling rate, this work found that performing a

linear fit over that temperature range was inappropriate for extrapolating

transition information to slower cooling rates. In fact, this work proposes that for

infinitesimally small cooling rates, the sol-gel transition temperature will

approach and equal that for the gel’s melting temperature.

171
An alternative extrapolation was proposed however. By inspection of the

cooling transition data, a linear region is found for cooling rates of 20ºC/min and

greater. This data was then extrapolated to a zero cooling rate, which in nearly all

cases resulted in the zero-rate interception being very near the actual transition

temperature for a 10ºC/min cooling rate. One important issue that arose during

this work was the high cooling rate data had to be verified since heat-transfer was

sometimes an issue. For this reason an instrument operating envelope was


generated and any data obtained that lied outside that envelope was discarded

since the recorded/program temperature was not in agreement with the actual

sample temperature.

8.3.2 Rheological Studies

DVR studies of gels provided some of the most interesting data during this

work. When studies were initially commenced, it was desired to only measure a

gel’s physical strength, as one of the original premises for this work required a gel

of a minimum physical strength. It quickly became apparent though that, by the

method selected of an oscillatory cone and plate, this was a flawed method of

analysis for several reasons. One is initial experimental design neglected to take
into account the volume decrease the material exhibits when going from a sol to a

gel. Since in order to test between the cone and plate the material had to be cast

from a sol, the transition resulted in large discontinuous normal forces between

the measuring surfaces, which led to surface detachment. Also, adhesion between

the gel and measuring surface (ceramic and stainless steel) was poor, resulting in

extremely erratic results for material ultimate strength/yield points. More often

172
than not the gel was not fracturing, the gel was disengaging from the stainless

cone surface. It was suggested to increase adhesion that the cone surface be

conditioned via light sandblasting or a similar operation, but that would render the

cone useless for other experiments that the instrument was simultaneously being

employed for.

For this reason experimental procedures focused on keeping sample

concentrations low in order to minimize the volume change effect. Polymer


loadings of less than 2% very rarely exhibited detachment. In addition

measurement focus went from ramping deflection to fail after a gel was cast and

aged, to constant deflection at a strain 10 times lower than the yield point while

observing the sol-gel transition real-time. Both storage and loss modulus are now

being measured, as the sample is still being a liquid and cooled to temperatures

near ambient. As the material gelled the storage modulus then dominated the loss

modulus by at least an order of magnitude, but a very unexpected unusual

response in the storage modulus as the material was gelling was observed.

For a typical experimental observation, a 1% SPS/EB solution was cast


between the plates at ~50ºC and the plate/cone gap set. The oscillation was

started at ~0.1% strain and 1.5 Hz, and the temperature ramped to ambient over a

several minute period. At a particular temperature the modulus would spike

nearly discontinuously from zero to several thousand Pascals, slowly roll over for

a minute or two, then decrease. Many times this phenomenon was also observed

under isothermal conditions (i.e. the sample had already approached the setpoint

temperature). This unusual response in modulus was pored over for several

173
months, and the literature assaulted in an attempt to determine what may be

occurring. No source was specific enough in explaining this phenomenon, so a

hypothesis was proposed in this work.

This work proposed that this unusual apparent maximum in gel modulus is

a result of the system at the instant of the theoretical gel point is an infinite

network of a large number of nascent crystallites, so the bulk of the polymer

chains are connecting microjunctions to one another, resulting in a “rigid”


material. At the low polymer concentrations under study, the gelled material is

still above the materials glass transition temperature, so crystals continue to grow,

resulting in “denetworking”; as more polymer crystallizes, less polymer connects

those growing crystals together.

8.4 IN CLOSING

Much new and interesting phenomenon have been studied and reported in

this work. Most notable of all experimental observations is the still-poorly

understood behavior of a suitable low-loaded polymer/solvent binary system as it

is going through the sol-gel transition as a result of cooling. A maximum in

modulus is observed just as the gel forms, which decreases significantly to an


equilibrium value as the gel ages. If any further work is done on these types of

materials in the context of this study, this phenomenon should be investigated

more fully using appropriate real-time capable methods.

To finalize this work an Appendix is provided to recap the work that was

performed by this researcher for the first few years of graduate school. This

174
Appendix details the various research and product development projects

undertaken during the product-developments days of 1997 and 1998.

175
Appendix A: Product Development

A.1 INTRODUCTION

Starting in the early 1990’s research was accelerated at the University of

Texas investigating the feasibility of revamping the method by which the

electronics industry finished various products. This is commonly referred to as

the electronics packaging industry. Not referring to how it is packaged so

someone will see it and buy it, but the packaging method and material that will

protect the electronic part or array from detrimental environmental conditions.

Any material that is a candidate for electronics packaging is subject to numerous

preexisting conditions as shown in Table A.1.

Glass Transition Temperature > 160ºC


Cure Temperature < 175ºC
Linear Cure Shrinkage < 0.5%
Thermal Expansion Coefficient < 20 ppm
Moisture Absorption, 24hr immersion < 0.5%
Dielectric Constant 3.0-5.0
Flammability UL94V-0 (a standard)
Extractable Halogens < 20 ppm
Hydrolyzable Halogens < 80 ppm
Flexural Strength > 14 ksi
Copper Peel Strength > 6 lbs/inch
Mold Cycle Time < 60 secs
Material Cost (Bulk) < $7.00/lb

Table A.1: Modern Electronics Packaging Material Minimum Specifications.


As given by Bajaj (1995).

176
Work progressed at UT until a suitable material had been isolated that

utilized a brand new reaction chemistry, specifically tailored to be a Reaction

Injection Molding (RIM) process. This process has demonstrated (Your 1989) to

be far cheaper and quicker than more traditional and developed methods for

delivering polymeric materials. With this in hand, Jempac International was

chartered to deliver this material to the marketplace. The first two years of

research done by this author were directly related to product development for this
company.

The aspects of work performed were: purification of reactant chemicals,

development of various anionic initiating chemicals, evaluation of various

candidate materials for hermetically storing our prepolymer product, pigmentation

of the prepolymer product, rheological control of the liquid monomer/fumed silica

paste mixtures, and providing for adhesion of our product to existing substrates

using the least invasive modifications and treatments necessary.

A.2 MONOMER PURIFICATION AND HANDLING

Butyl lithium initiators react strongly with water and carbon dioxide; trace

amounts of these contaminants can lead to error in reaction kinetics and too much
partially or un-reacted material after processing. The vinyl reactants we are using

for our process (styrene, paratertiarybutyl styrene (tBS), and divinylbenzene

(dVB)) are sold with varying small amounts of inhibitor and normally contain

small amounts of other undesired species such as water and oxygen. Therefore,

methods were required to purify these chemicals, in a way to minimize their

177
degradation, while maintaining reasonable costs and shelf lives of the resultant

materials.

One of the most common aryl-vinyl monomer polymerization inhibitors

(preservatives) is paratertiarybutyl catechol (tBC), present in commercially

delivered products at 10 to 500 ppm to scavenge free radicals and prevent radical

polymerization while sitting in the container. In the laboratory tBC is normally

removed by passing the liquid through a short bed of activated basic alumina
(Allock 1990). Alumina is activated for this procedure by heating it in an air oven

for at least an hour at 350ºC. Another traditional method for tBC removal is via

agitation in a concentrated aqueous sodium hydroxide solution, but this invariably

introduces more water as a result. A 6” deep x 5/8” diameter alumina bed is

capable of processing about a gallon of monomer at 100 ppm original inhibitor

content at or less than about 5% breakthrough before the bed material should then

be replaced. It is also reported that small amounts of dissolved water are removed

by the alumina if it has been freshly baked and cooled inertly.

For further purification the liquid monomer was degassed en vacuo then
mixed rigorously while being exposed to a circulating inert gas such as nitrogen

to strip out additional dissolved gases and water. Of course this is rather

expensive so was only done when highly quantitative polymerization reactions

were carried out. Otherwise alumina-filtered monomer was deemed acceptable

for production material. After purification the monomers were refrigerated as

without inhibitor they would self-polymerize in a matter of weeks at room

temperature.

178
When the most rigorous purification is required for anionic chemistry

work, solid sodium metal ingots were used to scavenge remaining contaminant

species. Handling of the sodium was done in a inert atmosphere glovebox to

avoid catching the organic liquids on fire while exposing the sodium to air. This

is rather impractical for everyday use though so was discontinued.

Testing for contaminant removal was done by taking an aliquot of

monomer and injecting a very small amount of butyllithium while looking for any
color change (deviation from colorless to orange or red). Since our reaction

system is based on butyllithium, this property was well known and the chemical

was at hand. Since the color is quite rich this method is very sensitive, at the

same time this also scavenges any remaining impurities before the polymerization

ensues, a small amount of which is acceptable.

A.3 ANIONIC INITIATORS

Butyllithium was the initiator of choice for our rapid reacting RIM

process. Being quite reactive, it is delivered in pentane or hexane solution, both

the 1º and 2º substituted butane varieties. The species that was most widely used

in this work was normal or 1-butyllithium (nBL). Shipped with a labeled


concentration of 10.0 molarity in hexanes, the vials always exhibited some degree

of degradation illustrated by its cloudy nature or presence of ash at the bottom,

ever increasing with time even if refrigerated. Since the material is indeed so

reactive, a method was necessary to determine the initiator’s actual concentration

on the spot when it was dispensed to achieve accurate results in modeling its

reaction behavior.

179
A relatively quick procedure was therefore used to double-titrate an nBL

aliquot, first described by Gilman in 1955. The procedure is as follows: 1 mL of

nBL is withdrawn from its storage vial and introduced into a sealed vessel

containing 30 mL of diethylether (ether) under an argon atmosphere while mixing

briskly. This mixture was then slowly quenched with 10 mL of pure water, a

quite violent reaction. The resultant precipitate/water/ether mixture was then

titrated to visible phenolphthalein endpoint using dilute HCl solution to yield total
base present in the initiator sample. This gives total base in the initiator solution,

but does not take into account the already spent reactive components that also

contribute to total base present.

To account for already spent reactive species, a second titration was then

performed on another aliquot of the same initiator solution in ether to which has

been added a slightly greater than stoichiometric (estimation) amount of

benzylchloride (BzCl). BzCl consumes the active nBL species into non-basic

carbonaceous compounds.

After this reaction, subsequent water quench and endpoint titration yielded
total adventious basic material, presumed to be those of spent initiator from self-

degradation. Taking the difference from the first and second titrations yields the

actual nBL concentration, at that particular time of course since the solution is

always slowly degrading. Others (Cartledge 1964) have questioned the use of the

BzCl double titration method as the BzCl reaction may not be exactly

stoichiometric, but we found it to be suitable for the compounds we were dealing

with in this work.

180
A.4 FILLER SURFACE TREATMENT

Fumed silica as it is normally provided is more or less very pure fine sand,

the average grain size for our applications being about 20 to 40µ. Due to it’s

“large” particle size, this material exhibits virtually no self-association when in a

liquid matrix, as can be seen when the particle size drop below about 1µ through

van der Waals interactions. Coating the silica particles with an additive was done

for several reasons, the two most important being to enhance its affinity for the

liquid monomer matrix material is was suspended in before reaction, and to

increase its interfacial interaction with the resultant matrix polymer after reaction,

increasing the bulk physical strength and decrease interfacial mass transport, in

effect “waterproofing” the silica.

A.4.1 Vinyl Siloxanes

Much of the surface treatment work focused on the method of applying,

and then characterizing the success of, coating the silica particles with a siloxane

based chemical. The compound utilized most widely for this was vinyl-tris(2-

methoxyethoxy)siloxane (herein referred to as A172, its trade name). The

siloxane works by chemically bonding to the silica particle surface through

hydrolyzable methoxy groups on the A172, condensating methanol and/or ethanol

upon its reaction. Ideally the result is a coating that covers the entire particle,

making it more “organic” in nature with respect to its thermodynamic association

with its nearest neighbors. A172 (a liquid in pure form) is shown in Illustration

A.1.

181
O Si O
O
O
O

Illustration A.1: A172 Vinyl Siloxane Silica Surface Modifying Chemical


(Whitco Chemical Company).

In this particular case the left over pendant vinyl group would remain

protruding from the silica surface, providing an attachment point for a

polymerizing system. Also bonding the organic compound to the silica surface in

effect increases the particle size and decreases its density, providing for increased

form drag and buoyant force. This also created a hydrophobic coating on the

particle helping to exclude water from the system, an important feature given the
RIM’s process sensitivity to contaminants such as water. Several processes for

coating silica powder with treatments are given below.

A.4.1.1 Application of A172 Siloxane

Initially silica was treated by making a slurry of filler in methanol (50/50

w/w), stirring in the siloxane liquid, then cooked without mixing overnight in a

baking pan at 80ºC to dryness. This ‘pan treated’ silica resulted in a hard cake of

material which was then reduced to powder again in a Wiley Mill. This silica

182
exhibited longer settling times and a greater ultimate packing density than

untreated silica, meaning if it did eventually settle out, it was far more difficult to

remix the silica with its matrix liquid.

The method arrived at that resulted in a very uniformly coated material

while maintaining it as a powder was realized when the powder was continuously

agitated in a Brabender heated sigma-bladed mixing chamber. A concentrated

siloxane solution in methanol was slowly sprayed into the chamber while the
powder was turned and sheared, simultaneously reacting and drying the material.

The resultant silica material appeared as the starting material: dry, no clumps,

requiring no further processing. The major drawback was the small batch size

(~300 grams, three hours per batch) and the apparent abrading of a small amount

of stainless steel into the powder, which would not affect laboratory tests but

would certainly not be acceptable in a non-conductive finished product. Silica

treated via the continuously stirred process showed a marked improvement in

settling rates (slower) and less agglomeration.

Spraying on the siloxane surface treatment is the preferred and best way to
date to introduce our own surface modifications to silica particles. A higher

shearing mixhead along with hardened surfaces on mixer parts would provide for

an ideal surface treatment apparatus. Barring those things silica produced in this

method provided good surface modification properties as detailed below.

A.4.2 Surface Treatment Evaluation

Surface treatments were evaluated via several methods, but the three most

used tests here were: 1) accelerated separation tests in tBS monomer, 2) water

183
uptake by a bulk filler sample over time, and 3) surface chemical activity/affinity

for an anionic polymerizing system. Each test is briefly described and results

discussed below.

A.4.2.1 Accelerated Separation Test

This test uses dilute solutions/slurries of a 10 wt% filler sample with the

balance inhibited tBS monomer. As a standard 25 mL of monomer was used in a


100 mL test tube, resulting in a liquid depth of about 6 cm. Fillers were weighed

in, tube capped, sample dispersed (shaken), then allowed to stand vertically while

observed for settling. Depending on the type of filler and surface treatment used

for a test, two distinct settling behaviors were observed, one possessing a single

apparent interface while the particles settled, and another dual-interface settling

mode separating the slurry into three distinct concentration zones. These two

different modes of filler separation are shown in Illustrations 2a and 2b.

184
Illustration A.2a:Separation Profiles with No Liquid Modification and Untreated
Silica.

Illustration A.2b:Separation Profiles with No Liquid Modification and 2% A172


Surface-Treated Silica.

185
These observations are obviously qualitative yet gave excellent indications

as to how well different surface treatment chemicals and application methods

affected the filler’s interaction with a standard matrix liquid and like processing

conditions. It was assumed the longer filler took to settle out, the better its

affinity for the liquid was and the stronger interfacial interactions were, leading to

enhanced physical properties in a finished product.

Settling time is taken to be the surprisingly distinct interface formed by the


descending silica particles that originated at the uppermost surface of liquid in the

test tube column. As this moved downward its level could be recorded while

timing the fall of this line. Untreated filler settled so quickly in pure monomer

liquid that 80% of equilibrium packing would be realized in just about three

minutes. In these studies 40µ silica particles were used. The region of silica-rich

liquid packed rather slowly, hence the overall assignment of two zones during

these observations.

Other treated silica particles would pack together far quicker than they

appeared to settle out, so three distinct zones during settling observations are
described. The upper interface demarking virtually pure liquid from that

containing any silica was more diffuse than that for untreated silicas, possibly

indicating some particles had far greater affinity for the liquid than most of the

silica settling out. The region at the bottom of the column also showed the silica

to be very tightly close-packed and even agglomerated, so much so that

vigorously shaking a settled vial would not redistribute the particles—they had to

be sheared with a sharp instrument and remixed to do this. Times for all treated

186
silica to settle out were far greater than those for untreated silica, most exceeding

20 minutes. The fact it ultimately formed a hard cake was rather undesirable

though. This phenomenon did eventually lead to research modifying the matrix

liquid in order to prevent silica particles from settling out.

A.4.2.2 Equilibrium Water Uptake

Fumed silica will readily adsorb water onto its surface via hydrogen
bonding as the surface is a sea of pendant Si-OH bonds. Ideally the surface

treatment would completely exclude any water sorption but this is not necessarily

the case due to coating efficiency. Water sorption has been greatly reduced

however with our current siloxane treatment technology. Figure A.1 shows that

equilibrium water adsorption is greatly reduced using varying amounts of surface

modification.

187
0.7

0.6
Untreated
0.5 2.5% A172
3.0% A172
% Mass Increase

0.4

0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1
Relative Humidity @ 21ºC

Figure A.1: Equilibrium Water Uptake For Minco 40 micron Silica. Shown as
function of vapor pressure of water present.

A.4.2.3 Surface Chemical Affinity

A siloxane compound with vinyl functionality was specifically selected in

order to provide some degree of chemical bonding between silica particles and

matrix polymer. After the siloxane had bonded to the silica particle surface the

unreacted vinyl group would be protruding and act like a co-monomer fraction

during the anionic polymerization process, since pendent vinyl groups from a

silicon atom are in fact anionically active (Hirao 1987).

188
In order to check for this activity, samples of surface treated filler were

carefully dried, weighed, and introduced into vials which were then sealed,

evacuated, and charged with about 50% by weight tBS monomer. At room

temperature, a small amount of nBL was introduced and the vials shaken; an

immediate orange color indicated the reaction was ‘living’ so was allowed to

proceed without further disturbance. Depending on how much initiator was used,

in 3 to 10 minutes the reactions had accelerated and completed, resulting in a


silica filled slug of black-red polymer.

After cooling for an hour or two, the vials were carefully broken and the

reacted slug reweighed to double-check the mass balance, then the slug placed in

an excess of toluene. This dissolved the slug into a polymer solution and silica

powder, this powder then carefully recovered by filtration and dried under

vacuum at low temperature. Reweighing the resultant dry silica found no increase

in mass, not a good sign as some mass increase was expected from polymer

chains chemically bonding to the silica surface during reaction. The error

associated with handling the silica, filtering it, filter hold-up, etc would not
exceed the mass gain expected from a reasonable fraction (>10%) of silica surface

pendant vinyl groups reacting with the growing PtBS chains.

As another test for surface reactivity, a dilute pot of living anionically

synthesized tBS oligomers was made in toluene at about 5% non-solvent

component, then a measured dry amount of A172 treated silica filler introduced to

this environment of living short chains. This living slurry was mixed for over an

hour with persistent orange color, terminated with methanol, then filtered to

189
isolate the filler component. The filler was washed and filtered through toluene

several times, then dried and weighed. Again, no significant weight gain was

realized when the results of this experiment were compared to the same one done

with untreated silica.

This result had profound implications for developing a hermetically

capable silica filled system. Earlier work at the University assumed, without

proof, that the reaction between living aryl-vinyl monomer chains and silica
surface-bound vinyl groups was indeed occurring, although the efficiency of this

chemical bonding was never ascertained. These early composition also fared

poorly in autoclave temperature cycling tests, which could be attributed to the

non-chemical association between the filler and polymer matrix. The most recent

work indicates that there is no, measurable anyway, chemical coupling between

polymerizing chains and surface treated filler. This is most likely due to the

pendent vinyl groups bonded to the silica surface either being sterically hindered

to reacting with the propagating anionically forming polymer matrix, or the vinyl

group is somehow deactivated electronically from such a reaction. Additional


filler surface treatments were explored for their post surface application bulk

reactivity although far more expensive than the existing A172 treatment liquid.

A.4.3 Siloxane Modified Polymer

While siloxane treatment gives good water adsorption control, it has not

sufficiently increased slurry settling performance. It was decided to synthesize

PtBS/siloxane copolymers before their application to silica filler, as it was well-

documented (Jeong 1996, Hirao ibid.) that this reaction would indeed work,

190
without interfering with the siloxane’s ability to later be bound to the silica

surface. Both free-radical and anionic based polymerization methods were

investigated for this purpose. The resultant material was then applied to untreated

silica material to determine its reactivity and solution properties.

A.4.3.1 Free-Radical Siloxane Copolymerization

A172 and tBS were combined in a reactor at 50ºC in toluene solvent in


nearly stoichiometric amounts. The siloxane component was selected to achieve

5% available surface treatment to 500 g of 40µ fumed silica. 3 wt% by total

monomer of azobisisobutyronitrile (AIBN) was used to initiate the five hour

reaction. After several hours an increase in the reactor viscosity indicated that

some reaction had taken place. Since the resultant polymer was of such high

molecular weight, spray-applying a solution of it would require far too much

solvent, so instead a thicker solution of the resultant polymer was just poured into

the Brabender mixer to apply the treatment and sheared as described previously.

Filler treated in this manner settled in slurry-solution at virtually the same

rate as regular A172 treated filler. One advantage is that filler treated with the

copolymer easily flows after settling to the bottom of the vessel (no hard cake
formation) while other low molecular weight treated species form a hard cake of

inflowable material.

A.4.3.2 Anionic Siloxane Chain-Termination

Living anionically synthesized oligomers of tBS (10-15 repeat units) were

made in toluene solution using nBL then were subject to an equivalent of A172

vinylsiloxane. The resultant siloxane terminated polymer was precipitated,

191
washed, dried, then redissolved in toluene for surface treatment application. The

polymerization reaction was performed at room temperature then lowered to 0ºC

before addition of the siloxane to minimize the side reactions the methoxyethoxy

functionality might engage in whilst in the anionic environment.

When filler treated with this terminated polymer was evaluated significant

mass gain was observed, indicating that the filler had indeed attached to the silica

and since the mass gain was nearly 10% while total A172 involved was about 1%
of the filler mass, there must have been anionic coupling between the living tBS

and A172, then subsequent reaction of the surviving methoxyethoxy groups with

silica, resulting in an apparent glassy hydrocarbon coating on the silica particles.

Testing this via the regular 10% slurry method showed the silica stayed in

solution far longer that previously treated fillers, and some particles staying in

dilute solution above the settled bed for over a week.

A.4.4 Other Surface Treatments

Since the chemical activity of the pendant vinyl group on surface-bound

siloxane was brought in question, other functionalities were selected to test in

replacement of this apparently ineffective chemistry. Two different siloxanes


attempted for this work are glycidyl methacrylate substituted methoxy silane, and

parastyryl substituted siloxane.

A.4.4.1 Styryl-Siloxane

If a vinyl group directly bonded to a siloxane is anionically active, it was

thought then certainly a styryl group bonded to a like siloxane certainly would be.

In addition this para vinyl-silyl substituted ring would perhaps be chemically

192
active even after surface bonding unlike the regular vinylsiloxane. A small study

was performed using styrylsiloxane (SS) to determine if it was surface-reactive

but regardless of outcome this compounds prohibitive cost most certainly kept it

out of contention from viable large-scale implementation.

300 g of 40µ silica was to be treated with a binary surface treating solution

of A172 and SS (10 mL and 0.2 mL, respectively). The regular Brabender mixing

application was used, the surface treatment diluted in 40 mL of methanol for


misting application. Lower temperatures were used in this run to limit the

possible degradation of the styrene functionality of the siloxane, so the liquid was

sprayed on at 40ºC and mixed for only ten minutes. An additional five minutes of

mixing was done at 70ºC to remove residual methanol. The resultant powder was

removed and stored in a vacuum oven at 55ºC.

Subsequent testing of this material showed no chemical uptake of

monomer in an anionic polymerization environment. In retrospect this test was

insufficient to prove this, as so little of the SS component was present compared

to the regular A172 agent used. These amounts were selected as only 1 cc of the
styryl siloxane was available, yet to effectively process material 300 grams of

powder was required to run the big mixing machine. A better test would be to

attempt to treat a smaller amount of silica using pure SS surface treatment.

A small batch was therefore attempted using pure SS in a methanol slurry

of silica, 0.5 cc (0.5 g) of SS to 25 grams of silica, a 2% surface coating yield. 0.3

grams of surface treatment was taken up based on before and after massing of the

silica, yet again as before the left-over styryl group showed no chemical affinity

193
for anionically polymerizing tBS based on exposing this treated silica to an

anionically polymerizing environment as previously discussed.

A.4.4.2 Glycidyl-methacrylate siloxane

The bulk material purchased for use in the product was delivered with a

glycidal-methacrylate surface preparation already on it. This material could be

easily identified by its peculiar strong smell in its container, probably leftover
surface treatment that was not washed from the material in the factory. Tests

performed as previously described demonstrated this surface chemistry was not

reactive toward the anionic polymerization environment as well, so no chemical

bonding was being realized in the test parts that were produced.

A.5 HIGH LOADING SOLIDS SUSPENSION

Maintaining a homogeneous paste of filler particles and monomer liquid

was found very early on to be critical for a useful product to come to fruition. A

binary mixture of low viscosity monomer (1 cP) and large silica particles

separates over several days, whereas it is desired that this not take place, if ever,

for several months to provide for an acceptable shelf life. Since the filler particles

are so ‘large’ at 20 or 40µ, their interparticle interactions (van der Waals forces)

are insufficient to overcome gravity. Changing the liquid (matrix) viscosity

would affect separation rate, in addition to modifying the surface character of the

filler, previously studied but not leading to sufficient solids suspension. The

focus now changes to modifying the matrix component.

194
A.5.1 Matrix Thickening

Dissolving commercial high molecular weight (250k) atactic polystyrene

(APS) in the tBS monomer before introducing the filler reduced filler settling.

The amount of polymer used to test this was rather high, about 10% of the total

monomer weight. While a significant increase in viscosity was realized, ~100 cP

from 1 cP, this did not significantly increase the slurry’s rheological stability, it

only resulted in a thicker paste that would still fully separate as before. In
addition this mode of modificaton was contrary to one of the main product goals:

easy handling and injection of prepolymer material via low viscosity.

A further viscosity modification test was done with atactic PtBS made

anionically (50k MW) at about 5% w/w monomer with like results as with regular

PS. Higher viscosity liquid only delayed the inevitable: dense solid particles

falling under the influence of gravity in a less dense matrix. Something had be

done to essentially ‘glue’ the particles in place to keep them immobilized virtually

indefinitely. Polymer may still be dissolved in the matrix to tailor the final fluid

viscosity once the solids loading is locked in place. This is due to earlier work
indicating that high friction between fluidized filler particles combined with a low

liquid viscosity can lead to flow instabilities in which phase separation occurs

while material is being pumped, plugging the delivery lines. Work then focused

to creating something of virtually infinite viscosity yet very low physical strength.

This was a characteristic of a gel.

195
A.5.2 Early Matrix Gelation Methods

To achieve true filler suspension in monomer work then shifted to

searching for materials that would gel tBS monomer (and later monomer also

containing a significant dVB fraction) without being detrimental to any other

process requirement previously specified. Namely, the weight fraction would be

maintained as low as possible as to not adversely lower the glass transition

temperature of the reacted material, as well as not adversely affect the anionic
polymerization chemistry. For these reason a product containing a gelled matrix

was pursued because it can exhibit the properties of a very viscous mixture, that is

being purely elastic below a yield point (static system), yet require very little

energy input to break and flow easily under higher stresses, at very low gel-

causing additive levels.

A.5.2.1 Aluminum Stearates

The first materials identified as gelling agents having potential for our

application were aluminum stearate soaps or various chemistries. These were

provided to us by Whitco Company under the their product #18 (dicarboxylated-

monobasic) and #22 (tricarboxylated) aluminum soaps. The majority of these


products contained the soap component (83+%) yet are also listed to contain

significant quantities of ash (~8 wt%) and free stearic (fatty) acid (~9 wt%).

Initially the acid was removed for our use but it was later discovered (Hill 1956)

that it assists in forming the gel network by serving as a peptizing agent. The ash

is also listed as an anti-agglomerant during final soap comminution.

196
Experiments with the Whitco products produced excellent gelling

behavior in tBS monomer when used at the 2 wt% level or greater. The procedure

for making a gelled system required the desired amount of soap to be added to the

monomer then the mixture heated to over 90ºC which served to dissolve and

activate the soap’s network formation capability. What was interesting is that in

the temperature ranges studied for these soaps, the systems became more viscous

as the temperature was raised from room temperature up to about 130ºC.


To test this new system’s solids suspension capability, gelled

monomer/filler solutions (50/50 w/w) were made at ~200 mL volumes in jars,

well blended at room temperature (after gel activation), and let stand for over a

week. Mixing occurred after heating the gel to activation due to this gel’s

activation temperature being too high for use in the commercial process. The gel

had to deliver on suspension stability at below-activation temperatures or it would

be unusable. Systems that contained 2 wt% or greater aluminum soap

(dicarboxylated, monomer basis weight fraction) showed no observable particle

settling or stratification in the vessel. This was very good news, as even after
shearing up the formed gel it still provided static rheological stability. The

viscosity probably increased by the higher temperatures allowing the soap to melt,

thereby greatly increasing its mobility, which led to a far greater number of

associations of ionic centers.

Systems with less than 2% aluminum soap separated as per how much

gelling agent had been used quite linearly. Samples with 1 wt% settled about half

way that samples with no agent would. This phenomenon indicated by whatever

197
mechanism the soap was providing particle support there exists some minimum

weight fraction or concentration of soap that provides for sticking solid particles

together even when the gel’s ionic network had been sheared up into little pieces.

An exciting property of this gelled system was that it would flow under its

own weight. This is of course studied at 50 wt% solids loading instead of 75 or

80 wt%, but still illustrates that it is possible to suspend dense particles while

exhibiting low dynamic viscosity. Another important observation is that the gel is
further broken when nBL is injected into it. Most likely the nBL is reacting with

the protons in the free fatty acid present, breaking the gel, but also consuming

initiator; at the time this was deemed acceptable if too much initiator was not

consumed. This would lead to an even lower delivery viscosity aiding in this

materials injection into intricate mold runners and gates. Other soap chemistry

was investigated in an attempt to tailor the organic non-associative component of

a soap molecule to the matrix monomers that were associated with this research.

A.5.3 Alternative Soap Chemistry

Due to the success of controlling the viscosity and filler suspension

properties of monomer with aliphatic aluminum soaps, the work now turned to
modifying the organic portion of the soap molecule while maintaining the

dicarboxylated monohydroxyl aluminum component. If the aliphatic chain

(currently C18 in length) could be substituted with phenyl groups or even better

yet tertiarybutyl phenyl groups, it would certainly be more soluble in the matrix

monomer and perhaps require less material to make a strong enough gel.

198
A.5.3.1 Polystyryl Soap

Water-insoluble aliphatic soaps gel aromatic solvents, but ones with aryl-

substituted chains might be even better. This could be realized by making short

polymer chains of the desired organic branch component, then adding the

carboxylate and aluminum components to make a di(poly-styryl)ated monobasic

or di(poly-t-butylstyryl)ated monobasic aluminum soap, the structures of which

are shown in Illustration A.3.

199
O

OH
C4H9 O Al
3-n
m

n=1 or 2
n
(a)
O

OH
C4H9 O Al
3-n
m

n=1 or 2
n
(b)

Illustration A.3: UT Aromatic Oligomer Soaps. (a) polystyryl aluminum


hydroxide soap, (b) poly(tert-butyl styryl) aluminum hydroxide
soap; m was nominally attempted at 10 to 25 repeat units.

These materials would also certainly be of higher glass transition

temperature than an aliphatic-based soap, affecting the final product properties

less so. Changing to a substituted soap organic backbone has raised many

discussions as to how long it should be made to have the same effective length

and function of the regular C-18 (18 carbon) saturated straight chain soap.

200
Several different lengths were attempted from 20 to 50 carbon units, so 10 to 25

monomer repeat units.

To synthesize our own agent, chains were grown anionically to the desired

length, then terminated by bubbling pure CO2 then treated with aqueous HCl to

result in an acid-terminated oligomer of tBS. It had been touted (Porter 1967) as a

simple procedure to recover the PtBS-carboxylic acid and then further react this

to the desired soap. Apparently (Hsieh 1996) making long aryl-substituted chain
acid is not that simple, as several other termination mechanisms compete with the

anionic carbon carboxylation reaction. Yield can be improved however (Quirk

1989) by using extremely low temperatures (-78ºC) and adding a touch of THF to

facilitate dimmer disassociation. Even if acid was obtained in high yield however

it is subsequently subject to decarboxylation due to the substituted nature of the

dangling chain (J. March 1992) even at room temperature. We have observed

what we believe to be decarboxylation of our synthesized samples at about 160ºC

via DSC.

After several failed attempts at recovery and purification of the acidified


oligomer from a hot aqueous phase, the entire saponification reaction was instead

carried to completion in the same reactor. The biggest problem with this was the

saponification step is typically done in aqueous phase, while our oligomeric acid

was not water soluble, and would not melt even at 100ºC (although it got quite

viscous). To accomplish the saponification anhydrous aluminum chloride was

used in proper amount to yield tricarboxylated aluminum PtBS soap (IUPAC

nomenclature still up in the air…) Lots of runs were tried with a few yielding

201
recoverable material, or which gave varying properties. Not enough was known

about the complex organic chemistry taking place, and since the regular

aluminum soap was performing as desired, tailored organic soap chemistry was

not pursued much beyond that described herein.

A.5.3.2 Dicarboxylated Gelling Agent

One other path attempted to lead to an ionic gelling agent was emulating
the acid-laden properties of CarbopolTM, typically used to gel solid-rocket fuel.

An experiment started by initiating meta-diisopropenylbenzene (DIPEB,

Illustration A.4) with nBL at low temperature, adding tBS monomer to 20 repeat

units per side, then acid terminated via CO2/HCl treatment. This did not lead to a

recoverable product however, as DIPEB apparently (Fetters 1979) is unreliable as

far as obtaining a dually initiated compound with no (or very little) crosslinking

during initiation.

Illustration A.4: Meta-Diisopropenylbenzene (DIPEB) Divinyl Molecule. It is


notable that the double bonds are not conjugated with the ring, as
is the case with the ortho- or para- isomers.

202
A.5.4 Ultra-fine Particle-based Colloid

A very fine hydrophobic silica of 100nm size range was tested for gelling

monomer liquid under the trade name CabosilTM. Being of such small particle

size, this phenomenon is most likely due to particle-particle interactions (van der

Waals forces) causing a liquid containing this material to behave as inflowable

goo. This behavior was found to not be significantly affected by temperature over

our range of use (20-100ºC). Loading levels of nearly 10% were required to
achieve sufficient solids suspension (just much larger silica particles) yet so much

of this material present caused the upper level of solids that could be added to the

complete system to fall below 70 wt%. When the gel was broken it maintained a

very weak Bingham behavior for several hours before reforming into a material

requiring much greater stress to break it again (very time dependent association at

room temperature).

A.6 BULK POLYMERIZATIONS

High molecular weight polymer was useful for controlling the final

dynamic viscosity of the product paste, so PtBS was synthesized anionically for

this purpose of various molecular weights. Since large amounts of polymer were
necessary once production began, a new easy to use disposable reactor system

was developed which allowed for quick production of a desired amount and

molecular weight of polymer.

A.6.1 Pouch Reactors

In the past, single neck vacuum flasks were sacrificed for carrying out

non-solvent bulk polymerizations, but a new disposable pouch reactor was

203
developed that was far more economical and safer to use (no broken glass). The

pouches were made from ScotchPak heat sealable PE/PET/PE ternary laminate

film bags fitted with a small silicone rubber injection port and further sealed with

a few drops of silicone red-RTV and a staple. Prototypes were subject to several

simples tests and demonstrated they were airtight and could handle the heat of

even a truly adiabatic reaction, which would be accompanied by a temperature

rise of about 200ºC.


Pouches were prepared for reaction by repeatedly filling with inert gas and

evacuating. The polymerizations were performed by injecting a nominal amount

of 100 mL of previously purified monomer liquid using a large barrel syringe.

Then a carefully measured amount of nBL (25 to 200 µL) was then injected with

a microsyringe and dispersed by massaging the pouch a little. The now orange

(initiated) monomer was left on the bench at room temperature, given time to self-

heat and eventually therm and fully react.

Induction times varied from about 20 seconds up to 10 minutes, or never

reacting at all. From low concentration pouches that did finally react, a non-
reacted skin of monomer was present on the slug, indicating the pouches were

permeable to air at a non-negligible rate. Of course this interfered with the

kinetics, molecular weight, and distribution of the final product, and most likely

led to non-reaction when very low initiator concentrations were used—it was

consumed by permeating air before bulk polymerization could occur.

Given this slight drawback to the pouch system, the pouches were then

constructed and purged as before, then put in an inert glovebox. In here

204
chemicals could be injected without contamination through the end of the

reaction. Like experiments inside the glovebox resulted in induction times 50-

80% lower than those performed outside the glovebox, indicating all bench-top

polymerizations were subject to air diffusion contamination. After

polymerizations were allowed to cool overnight in the glovebox they were

removed, the bags peeled off, and the still living red slug of pure polymer crushed

by first cooling with liquid nitrogen. Higher molecular weight samples (>50k)
were found to be VERY strong and tough for an unmodified glassy polymer, not

brittle as expected. Low molecular weight samples crushed very easily, even

between the fingers at room temperature.

After samples were reduced to small pieces and powder, molecular weight

analysis was performed by making 5 mg/cc solutions of polymer in THF per the

GPC instrument operator’s recommendations. For calculated molecular weights

over about 100k the polymer was insoluble in THF however, only swelling up to

50 times its original volume. Several other solvents were used to attempt to

dissolve the polymer with no success, only swelling. It was learned several years
after this work was done from the manufacturer of the tBS under use that after

commercial processing a small amount of divinyl monomer resulted in the final

product (dehydrogenated tertiary butyl group presumably). This most likely was

lightly crosslinking the PtBS polymer when high molecular weights were

attempted via using a low enough concentration of initiator to allow the

polymerization to reach the critical chemical gellation point. Another possibility

proposed was that since the living polymerized samples were exposed to air,

205
adventious water could have led to termination mechanisms forming dimeric

ketones and trimeric alcohols (Quirk 1982), thus increasing the apparent

molecular weight.

GPC analysis was performed on the same instrument as previously cited,

using a crosslinked polystyrene column at room temperature and 1 mL per minute

flow rate. Molecular weight results supported the theory of molecular weight

enhancement by water-involved termination steps, but at the same time several


number average molecular weight results were below those calculated from

reaction conditions. Table 2 summarizes the results for polymer made in pouch

reactors in the glovebox using various initiator/monomer molar ratios.

Sample 10M nBL tBS Predicted Actual Actual


Distribution
Number µL mL Mn, Mw Mn Mw
41 75 100 117350 87700 156800 1.79
42 75 100 117350 86600 157000 1.81
61 100 100 88000 84600 158600 1.87
62 100 100 88000 81200 144800 1.78
71 100 100 88000 77800 138300 1.78
72 100 100 88000 76800 138000 1.80
81 200 70 30800 27400 36700 1.34
82 200 70 30800 27700 36700 1.32

Table A.2: Molecular Weight Results From Pouch Polymerizations in


Glovebox. Samples 1, 2, and 3 were of too high molecular
weight/crosslinked and were not THF soluble. Sample 5 failed to
inject properly (both vials). Duplicate samples were run for each
polymerization, each corresponding pair of molecular weight data
agrees within 3%.

206
Number average molecular weights were below that of calculated, while

weight averages were much high, yielding distributions of 1.3 to 1.9, while this

type of polymerization is reasonably expected to given distributions of 1.0 to 1.1.

Inspection of the GPC detector traces shows the maximum detector count

was exceeded, so the top of the peak was chopped off in software introducing

some error during peak integration. For monodisperse polymers (or very nearly

so) it is most likely advisable to use a much lower sample concentration, say one
or two mg/cc of carrier solvent. Figure 2 illustrates a characteristic GPC signal

for these polymers, hence why lower injection concentrations would lower the

peak to a more detector-friendly range.

207
600

500
Detector Response (mV)

400

300

200

100

0
5 10 15 20 25 30
Retention Volume (mL)

Figure A.2: GPC of one of the bag polymers. Middle peak is main polymer
signal. Concentration of only 1 mg/cc in THF solvent still caused
signal to go off scale (over 650 mV for this particular instrument).

A.6.2 Divinylbenzene Copolymerizations

To obtain a rough understanding of how divinylbenzene (dVB)

copolymerizes with tBS with anionic initiator a few bulk polymerizations were

performed. Some information indicated at high concentrations dVB would

polymerize with itself, resulting in a microgel, and leaving a large fraction of the

tBS unreacted even if the polymerization system remained active (living). Since

208
to obtain the required glass transition temperature performance nearly 30% by

weight dVB was required, this was certainly a high concentration system.

Tests were made using 3:1 tBS/dVB by weight, mixing well, and injecting

various amounts of nBL initiator giving yields of 40 to 300 vinyl groups per

initiator molecule. Low initiator concentrations formed microgels at room

temperature with very low ultimate reaction extent, whereas at elevated starting

temperatures the reaction would go to completion but a rather brittle material


resulted. High initiator concentrations would react to completion at all starting

temperatures but a very brittle, to the point of self-crumbling, polymer resulted.

Divinylbenzene did not appear to be a very good crosslinking material for this

application but is all that was commercially viable at the time. Sanchez (2002) is

pursuing various forms of economical molecules of dual aryl-vinyl functionality

that are not conjugated.

A.7 PASTE PIGMENTATION

Electronics encapsulation materials are typically black with white

identifying lettering on them. Jemstone is light to dark gray in color depending

on the filler type and loading used. Some kind of pigment was therefore required
to make the final appearance black. Of course as little as possible of this material

is desired, at the same time the material cannot adversely interfere with the RIM

system chemistry or rheological delivery behavior. Cabot Monarch 120 carbon

black was found to be suitable for this purpose, and yielded excellent finished

appearance at very small loading levels, the maximum necessary being 0.1%

based on preliminary tests to pigment the reacted paste.

209
A.8 COMPOSITE PACKAGING FOIL EVALUATION

When the issue of packaging the prepolymer paste (Jemstone) in a

hermetic manner was raised, several specifications for a suitable package were

made: 1) it must exclude entirely any environmental species that would

contaminate the anionic polymerization chemistry, 2) must be easy to use in an

automated process, 3) be durable enough to allow normal handling, and 4) possess

enough physical strength to deform while Jemstone is loaded and dispensed from

it, and 5) be economical.

Given the environmental requirement a barrier containing an impermeable

metal film was then decided; a multi-layer composite laminate film of PE,

aluminum, and PET at 1 mil per layer, 3 mils total thickness. This material

provided both a suitable barrier and melt-sealing capability with a simple process.

To test for mechanical properties samples of this film were subject to both film

and sealed-joint tensile tests.

To find the tensile strength of the film, a small loop was made in the form

of a “double dogbone” following the geometry by which rigid materials are

subject to tensile stress. Illustration A.5 shows how the film was cut, then joined

to form a loop at the wider ends, then loaded into the small Instron for stretching.

210
Illustration A.5: Double dogbone loops for tensile testing of the composite foil
material (not to scale).

Dynamic tests were done at a strain rate of two inches/minute, while the

effective strain length of the test subject was four inches. The material was found

to be elastic up to ~2.5% elongation and applied stress of 5.3 ksi, and ultimately

failed at 25% elongation and ultimate stress of 7.6 ksi. Given what we knew of

the process at the time this strength was deemed sufficient for use in the process.

Peel strength of the heat-sealed joints were then tested via a ‘reverse’ double-

dogbone test configuration, the heat sealed length being narrower than the width

of the rest of the loop as shown in Illustration A.6.

211
Illustration A.6: Sealed joint ‘reverse’ dogbone for testing of linear bond strength
of the heat-sealed material (not to scale).

Joints formed by just exposing two PE surfaces showed very low bond

strengths, barely exceeding one pound per inch (linear strength). After cleaning

the PE surface with acetone then sealing did a much stronger joint result with a

peel-failure strength of 20 pounds/inch. So to effectively heat seal this material,

the sealing surfaces had to be simply wiped of any bulk contaminants or dust.

212
A.9 CONCLUSION

Many different ideas and processes were explored in the pursuit of the

ultimate electronics packaging material. In the end the final product was not

commercially viable and the company disbanded, but the idea still remains:

superior performance and economics could very well be realized in the future

using this type of polymer if the technology were to be further developed.

213
REFERENCES
Allock, H.; Lampe, F., Contemporary Polymer Chemistry, Prentice-Hall: New
York, 1990, p 69

Bajaj, R., Doctoral Dissertation, The University of Texas at Austin, 1995

Cartledge, F.; Gilman, H., J. Organometallic Chem., 2, 447, 1964

Fetters, L.; Kamiensky, C.; Morrison, R.; Young, R., Macromolecules, 12, 344,
1979

Gilman, H.; Haubein, A., J. Am. Chem. Soc., 1515, 1944

Hill, P.; Van Strien, R.; Towle, P., U.S. Patent 2,751,359, 1956

Hirao, A.; Hatayama, T.; Nagawa, T.; Yamaguchi, M.; Yamaguchi, K,


Nakahama, S., Macromolecules, 20, 242, 1987

Hsieh, H.; Quirk, R., Anionic Polymerization, Dekker: New York, 1996, p 263

Jeong, J.; Jang, J.; Lee, J., U.S. Patent 5,525,677, 1996

March, J., Advanced Organic Chemistry, Wiley Interscience: New York, 1992, p
627

Porter, C.; Titus, D., Proc. Soc. Exp. Biol. Med., 124, 500, 1967

Quirk, R.; Chen, W., Makromolecular Chem., 183, 2071, 1982

Quirk, R.; Fetters, L.; Yin, J., Macromolecules, 22, 85, 1989

Sanchez, J., Doctoral Dissertation, The University of Texas at Austin, Currently


under Composition

Your, J.-J.; Karles, G.; Ekerdt, J.; Trachtenberg, I.; Barlow, J., Ind. Eng. Chem.
Res. 28, 1456, 1989

214
Bibliography

Berezniak, A.; Zimoch, T., Biul. Wojsk. Akad. Tech. Prace. Chem., 7, 62, 1958

Berghams, H.; Donkers, A.; Frenay, L.; Stoks, W.; De Schryver, F.; Moldenaers,
P.; Mewis, J., Polymer, 28, 97, 1987

Campbell, R.; Hefner, J., US Pat. 5,196,490, 1993

Chien, J.; Salajka, Z., J. Poly. Sci., Part A: Poly. Chem., 29, 1243, 1991

Grassi, A.; Pellechia, C.; Longo, P.; Zambelli, A., Gazz. Chim. Ital., 19, 2465,
1987

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Imabayashi, H.; Ishikawa, K.; Yamamoto, K.; Izumi, T., US Pat. 5,037,907, 1991

Ishihara, N.; Seimiya, Y.; Kuramoto, M.; Uoi, M., Macromolecules, 19, 2464,
1986

Ishihara, N.; Kuramoto, M.; Uoi, M., Macromolecules, 21, 3356, 1988

Kawamura, T.; Uryu, T.; Matsuzaki, K., Makromolecular Chem., 180, 2001, 1979

Narh, K.; Barham, P.; Keller, A., Macromolecules, 15, 464, 1982

Pellechia, C.; Longo, P.; Grassi, A.; Ammendola, P.; Zambelli, A., Makromol.
Chem. Rapid Commun., 8, 277, 1987

Pezron, E.; Ricard, A.; Lafuma, F.; Audebert, R., Macromolecules, 21, 1121,
1988

Zucchini, U.; Albizzati, E.; Giannini, U., J. Organometal. Chem., 26, 357, 1971

Berghams, H.; Donkers, A.;Frenay, L.; Stoks, W.; DeSchryver, F.; Moldenaers,
P.; Mewis, J., Polymer, 28, 97, 1987

Blanks, R.; Prausnitz, J., Ind. Eng. Chem. Fund., 3(1), 1, 1964

215
Daniel, Ch.; Deluca, M.; Guenet, J.; Brulet, A.; Menelle, A., Polymer, 37, 1273,
1996

Flory, P., J. Am. Chem. Soc., 63, 3096, 1941

Flory, P., J. Chem. Phys., 10, 51, 1942

Flory, P., J. Am. Chem. Soc., 69, 30, 1947

Flory, P., Principles of Polymer Chemistry, Cornell University Press: Ithaca, New
York, 1953, 1992 (15th Ed.)

Guenet, J.; McKenna, G., J. Polym. Sci., Polym. Phys. Edn., 24, 2499, 1986

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Guenet, J.; Daniel, C.; Dammer, C., Polymer, 35, 4243, 1994

Guenet, J.-M.; Daniel, Ch.; Deluca, M.; Brulet, A.; Menelle, A., Polymer, 37,
1273, 1996

Itagaki, H.; Nakatani, Y., Macromolecules, 30, 7793, 1997

Macosko, C., Rheology: Principles, Measurements, and Applications, VCH


Publishers: New York, 1994

Nijenhuis, K., Thermoreversible Networks, Springer-Verlag: Berlin Heidelberg,


Germany (In English) 1997

Pezron, E.; Ricard, A.; Lafuma, F.; Audebert, R., Macromolecules, 21, 1121,
1988

Prasad, A.; Mandelkern, L., Macromolecules, 23, 5041, 1990

Scott, R.; Magat, M., J. Poly. Sci., 4, 555, 1949

Small, P., J. App. Chem., 3, 71, 1953

Stockmayer, W., J. Chem. Phys., 11, 45, 1943

Tanaka, F.; Stockmayer, W., Macromolecules, 27, 3943, 1993

Tanaka, F.; Nishinari, K., Macromolecules, 29, 3625, 1995

216
Vittoria, V., Polymer Communications, 31, 263, 1990

Braun, D.; Betz, W.; Kern, W., Makromolecular Chem., 42, 89, 1960

Bueschges, U.; Chien, J., J. Poly. Sci., Poly. Chem. Ed., 27, 1529, 1989

Campbell, R., US Pat. 5,066,741, 1991

Chien, J.; Salajka, Z., J. Poly. Sci., Part A: Poly. Chem., 29, 1243, 1991

Cohen, R.; Cazzaniga, L., Macromolecules, 22, 4125, 1989

Danusso, F.; Sianesi, D., Chim. Ind.. 40, 450, 1958

Grassi, A.; Longo, P.; Proto, A.; Zambelli, A., Macromomlecules, 22, 104, 1989

Ishihara, N.; Seimiya, Y.; Kuramoto, M.; Uoi, M., Macromolecules, 19, 2464,
1986

Ishihara, N.; Kuramoto, M.; Uoi, M., Macromolecules, 21, 3356, 1988

Longo, P.; Grassi, A.; Proto, A.; Ammendola, P., Macromolecules, 21, 24, 1988

Oliva, L.; Pellecchia, C.; Cinquina, P.; Zambelli, A., Macromolecules, 22, 1642,
1989

Pellechia, C.; Longo, P.; Grassi, A.; Ammendola, P.; Zambelli, A., Makromol.
Chem. Rapid Commun., 8, 277, 1987

Po, R.; Cardi, N., Prog. Polym. Sci., 21, 47, 1996

Proto, A.; Longo, P.; Grassi, A.; Ammendola, P., Macromolecules, 21, 24, 1988

Siddall, J., US Pat. 5,484,862, 1996

Yamada, A.; Kogyo Kagaku Zasshi, 64, 2204, 1961

Zambelli, A.; Longo, P.; Pellechia, C.; Grassi, A., Macromolecules, 20 2035,
1987

Zucchini, U.; Albizzati, E.; Giannini, U., J. Organometal. Chem., 26, 357, 1971

Chen, Q.; Yu, Y.; Na, T.; Zhang, H.; Mo, Z., J. App. Poly. Sci., 83, 2528, 2002

217
Chien, J.; Salajka, Z., J. Poly. Sci., Part A: Poly. Chem., 29, 1243, 1991

Grassi, A.; Longo, P.; Proto, A.; Zambelli, A., Macromolecules, 22, 104, 1989

Hoffman, J.; Weeks, J., J. Chem. Phys., 42, 4301, 1965

Ishihara, N.; Kuramoto, M.; Uoi, M., Macromolecules, 21, 3356, 1988

Kawamura, T.; Uryu, T.; Matsuzaki, K., Makromolecular Chem., 180, 2001, 1979

Park, J.; Kwon, M.; Park, O., J. Polym. Sci.: Part B: Polym. Phys., 38, 3001, 2000

Shinozaki, D.; Lawrence, S., Polymer Engineering and Science, 37, 1825, 1997

Sudduth, R.; Yarala, P., Poly. Engr. and Sci., 42, 694, 2002

Wesson, R., Poly. Eng. and Sci., 34, 1157, 1994

Yuan, Z.; Song, R.; Shen, D., Polym. Int., 49, 1377, 2000

Guenet, J.; Daniel, Ch.; Deluca, M.; Brulet, A.; Menelle, A., Polymer, 37, 1273
1996

Berghmans, H.; Roels, T.; Deberdt, F., Macromolecules, 27, 6216, 1994

Birkinshaw, C.; Duff, S.; Tsuyama, S.; Iwamoto, T.; Fujibayashi, F., Polymer, 42,
991, 2001

Guenet, J.-M.; McKenna, G., Macromolecules, 21, 1752, 1988

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Nijenhuis, K., Thermoreversible Networks, Springer-Verlag: Berlin Heidelberg,


Germany (In English) 1997

Prasad, A.; Mandelkern, L., Macromolecules, 23, 5041, 1990

Tanaka, F.; Stockmayer, W., Macromolecules, 27, 3943, 1993

Aoki, Y.; Li, L., Macromolecules, 30, 7835, 1997

Aoki, Y.; Uchida, H.; Li, L.; Yao, M., Macromolecules, 30, 7842, 1997

218
Aoki, Y.; Li, L., Macromolecules, 31, 740, 1998

Aoki, Y.; Li, L.; Uchida, H.; Kakiuchi, M.; Watanabe, H., Macromolecules, 31,
7472, 1998

Aoki, Y.; Li, L.; Kakiuchi, M., Macromolecules, 31, 8117, 1998

Argon, A.; Cohen, R.; Patel, A., Polymer, 40, 6991, 1999

Berghmans, H.; Roels, T.; Deberdt, F., Macromolecules, 27, 6216, 1994

Guenet, J.-M.; McKenna, G., Macromolecules, 21, 1752, 1988

Guenet, J.-M., Thermoreversible Gelation of Polymers and Biopolymers,


Academic Press Inc.: San Diego, 1992

Macosko, C., Rheology: Principles, Measurements, and Applications, VCH


Publishers: New York, 1994

Nijenhuis, K., Thermoreversible Networks, Springer-Verlag: Berlin Heidelberg,


Germany (In English) 1997

Watanabe, H.; Sato, T.; Osaki, K.; Aoki, Y.; Li, L.; Kakiuchi, M.; Yao, M.,
Macromolecules, 31, 4198, 1998

Allock, H.; Lampe, F., Contemporary Polymer Chemistry, Prentice-Hall: New


York, 1990, p 69

Bajaj, R., Doctoral Dissertation, The University of Texas at Austin, 1995

Cartledge, F.; Gilman, H., J. Organometallic Chem., 2, 447, 1964

Fetters, L.; Kamiensky, C.; Morrison, R.; Young, R., Macromolecules, 12, 344,
1979

Gilman, H.; Haubein, A., J. Am. Chem. Soc., 1515, 1944

Hill, P.; Van Strien, R.; Towle, P., U.S. Patent 2,751,359, 1956

Hirao, A.; Hatayama, T.; Nagawa, T.; Yamaguchi, M.; Yamaguchi, K,


Nakahama, S., Macromolecules, 20, 242, 1987

Hsieh, H.; Quirk, R., Anionic Polymerization, Dekker: New York, 1996, p 263

219
Jeong, J.; Jang, J.; Lee, J., U.S. Patent 5,525,677, 1996

March, J., Advanced Organic Chemistry, Wiley Interscience: New York, 1992, p
627

Porter, C.; Titus, D., Proc. Soc. Exp. Biol. Med., 124, 500, 1967

Quirk, R.; Chen, W., Makromolecular Chem., 183, 2071, 1982

Quirk, R.; Fetters, L.; Yin, J., Macromolecules, 22, 85, 1989

Sanchez, J., Doctoral Dissertation, The University of Texas at Austin, Currently


under Composition

Your, J.-J.; Karles, G.; Ekerdt, J.; Trachtenberg, I.; Barlow, J., Ind. Eng. Chem.
Res. 28, 1456, 1989

220
Vita

Edward Louis Goldmann was born on 12 September 1974 in Jackson,

California, USA, the first of four children by parents Louis and Elizabeth

Goldmann. The family moved to Benton City, Washington State in 1979 where

Edward grew up attending elementary, middle, and high school in the Kiona-

Benton City School District of Benton County. After completing high school he

began attending the University of Washington (Seattle, Washington) in the fall of

1993. While being initiated into and remaining active in the Phi Kappa Sigma

International Fraternity for four years he received his Bachelor of Science in

Chemical Engineering in June 1997 with two departmental minors in Nuclear

Engineering and Materials Science. Whilst in the final year of attending the

University of Washington, an offer was accepted to pursue a Doctor of

Philosophy in Chemical Engineering at the University of Texas at Austin. Along

the way to the completion of this degree a Master of Science in Chemical

Engineering was also awarded for the extensive graduate coursework undertaken.

The successful binding/publishing of this work must mean the author has also

been granted the before-mentioned PhD in Chemical Engineering, which most

likely is the end of his formal education. Edward has accepted an offer of

employment from Intel Corporation in Hillsboro, Oregon, USA, which he should

commence in January 2003.

(continued next page)

221
Permanent address for Edward Goldmann:

50207 N Whan Rd, Benton City, WA 99320. The author can also be directly

found at anytime at the email address: edward@goldmann.com or the website:

http://www.goldmann.com

This dissertation was typed by the author.

222

You might also like