29.11.2014 Views

UNIVERSITY OF MINNESOTA This is to certify that I have examined ...

UNIVERSITY OF MINNESOTA This is to certify that I have examined ...

UNIVERSITY OF MINNESOTA This is to certify that I have examined ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>UNIVERSITY</strong> <strong>OF</strong> <strong>MINNESOTA</strong><br />

<strong>Th<strong>is</strong></strong> <strong>is</strong> <strong>to</strong> <strong>certify</strong> <strong>that</strong> I <strong>have</strong> <strong>examined</strong> th<strong>is</strong> copy of a doc<strong>to</strong>ral thes<strong>is</strong> by<br />

Kwanho Chang<br />

and <strong>have</strong> found <strong>that</strong> it <strong>is</strong> complete and sat<strong>is</strong>fac<strong>to</strong>ry in all respects,<br />

and <strong>that</strong> any and all rev<strong>is</strong>ions required by the final<br />

examining committee <strong>have</strong> been made.<br />

David C. Morse, Chr<strong>is</strong><strong>to</strong>pher W. Macosko<br />

Name of Faculty Adv<strong>is</strong>ors<br />

Signature of Faculty Adv<strong>is</strong>ors<br />

Date<br />

GRADUATE SCHOOL


Block Copolymer Surfactants in Imm<strong>is</strong>cible<br />

Homopolymer Blends<br />

A THESIS<br />

SUBMITTED TO THE FACULTY <strong>OF</strong> THE GRADUATE SCHOOL<br />

<strong>OF</strong> THE <strong>UNIVERSITY</strong> <strong>OF</strong> <strong>MINNESOTA</strong><br />

BY<br />

Kwanho Chang<br />

IN PARTIAL FULFILLMENT <strong>OF</strong> THE REQUIREMENTS<br />

FOR THE DEGREE <strong>OF</strong><br />

DOCTOR <strong>OF</strong> PHILOSOPHY<br />

David C. Morse and Chr<strong>is</strong><strong>to</strong>pher W. Macosko, Adv<strong>is</strong>ors<br />

December 2005


c○ Kwanho Chang 2005


To my parents, Sonya, and Kwanjoo<br />

i


Acknowledgments<br />

<strong>Th<strong>is</strong></strong> thes<strong>is</strong> would <strong>have</strong> not been completed without collaboration and support of many<br />

people. First and foremost of them, I would like <strong>to</strong> thank my adv<strong>is</strong>ors, Professors David<br />

Morse and Chr<strong>is</strong> Macosko for their incessant guidance and encouragement throughout the<br />

entire period of my Ph.D. education. I <strong>have</strong> been incredibly lucky <strong>to</strong> <strong>have</strong> an opportunity<br />

<strong>to</strong> explore both theoretical and experimental research areas. Their advice and valuable<br />

d<strong>is</strong>cussion from outstanding knowledge and insight in polymer science has been critical <strong>to</strong><br />

expand my experience.<br />

With respect <strong>to</strong> polymer synthes<strong>is</strong>, I <strong>have</strong> greatly appreciated the help by Professor<br />

Timothy Lodge and Professor Frank Bates. All polymers used in th<strong>is</strong> thes<strong>is</strong> <strong>have</strong> been<br />

synthesized in their labs. I also would like <strong>to</strong> specially thank Professor Daniel Joseph in<br />

the Aerospace Engineering and Mechanics Department. He generously allowed me <strong>to</strong> use<br />

h<strong>is</strong> spinning drop tensiometer over two years.<br />

I <strong>have</strong> enjoyed being a part of both Morse and Macosko research groups and would like<br />

<strong>to</strong> thank all group members for helping my research in overall directions. D<strong>is</strong>cussion with<br />

Hyun Jeon was very helpful <strong>to</strong> initialize my research in polymer blends. I also thank her<br />

for introducing my future wife, Sonya, <strong>to</strong> me. For all the generous ass<strong>is</strong>tance and support<br />

of my research, I would like <strong>to</strong> thank Jongwhi Lee, Jonathan Schulze, Pieter Spietal,<br />

Jennifer Dean, Chr<strong>is</strong> Tyler, Drew Davidock, Kasiraman Kr<strong>is</strong>hnan, Jianbin Zhang and<br />

YongIl Kwon. Thanks go out <strong>to</strong> Dave Hultman and David Giles for modifying a spinning<br />

drop tensiometer, and my officemates L<strong>is</strong>a Lim, Lifeng Wu, Yoichiro Mori, Mickael Castro,<br />

ii


David Ackerman, Hyunwoo Kim for the most precious time in my life.<br />

Finally, I never thank enough my mom for her endless love and encouragement. Her<br />

letters sent me for six years are just two hundred. I do not find any proper words for<br />

appreciating my love, Sonya. Also, I thank my best friends, Byung Wook and Seungho,<br />

for their support.<br />

I would like <strong>to</strong> acknowledge IPRIME and the RTP company for financial support and<br />

the Minnesota Supercomputing Institute (MSI) for computation time.<br />

iii


(261words)<br />

Abstract<br />

Block copolymers <strong>have</strong> been often used as emulsifying agents in binary homopolymer<br />

blends. We <strong>have</strong> used both theory and experiment <strong>to</strong> study the competition between<br />

interfacial adsorption and micellization of copolymer in mixtures of copolymer and two<br />

imm<strong>is</strong>cible homopolymers.<br />

Self-cons<strong>is</strong>tent field theory (SCFT) has been used <strong>to</strong> characterize the elastic properties<br />

of a block copolymer monolayer dividing A- and B-rich homopolymer domains in the<br />

context of the Canham-Helfrich theory, and <strong>to</strong> study the thermodynamics of two phase<br />

systems in which one of the coex<strong>is</strong>ting phases contains micelles.<br />

Interfacial tension has been measured between poly<strong>is</strong>oprene and polydimethylsiloxane<br />

in the presence of poly(<strong>is</strong>oprene-b-dimethylsiloxane) copolymer using a spinning drop tensiometer.<br />

In the case of a symmetric block copolymer, we observed a decrease of interfacial<br />

tension by three orders of magnitude. Experimental measurements of interfacial tension<br />

were compared with the SCFT predictions for a system in which swollen spherical micelles<br />

are present in one phase. The predicted dependence of interfacial tension on the block<br />

composition f A agreed well with the experiment results for sufficiently asymmetric copolymers,<br />

with f A > 0.65. However, for nearly symmetric copolymers with 0.5 < f A < 0.6,<br />

interfacial tension was much lower than these predictions. For those nearly symmetric<br />

copolymers, the formation of a bicontinuous microemulsion phase has been observed.<br />

Interfacial tension and the cmc <strong>have</strong> also been measured in a system of polybutadiene<br />

and polystyrene with poly(styrene-b-butadiene) block copolymer. The cmc has been deiv


termined by both X-ray scattering and transm<strong>is</strong>sion electron microscopy. In th<strong>is</strong> system,<br />

measured interfacial tension was strongly affected by the slow diffusion of block copolymer.<br />

v


Contents<br />

L<strong>is</strong>t of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

xi<br />

L<strong>is</strong>t of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxiii<br />

1 Introduction and Background 1<br />

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 Overview of The Thes<strong>is</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.3 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.3.1 Phase Behaviors in Oil/Water/Nonionic Surfactant Systems . . . . . 7<br />

1.3.2 Interfacial Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.3.3 Wetting/Nonwetting Transition of Interface . . . . . . . . . . . . . . 12<br />

1.3.4 Helfrich Bending Elasticity . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.3.5 Stability of Surfactant Monolayer . . . . . . . . . . . . . . . . . . . . 17<br />

1.3.6 Thermal Fluctuation . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2 Self Cons<strong>is</strong>tent Field Theory 28<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.2 Formulation of SCFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.2.1 Field Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.2.2 Homogeneous Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

2.2.3 Interfacial Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

2.2.4 Nondimensionalization . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

vi


2.3 Numerical Implementation of SCFT . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.3.1 Finite Difference Method . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

2.3.2 Numerical Accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

2.4 Iteration Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.4.1 Relaxation Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.4.2 New<strong>to</strong>n-Raphson Algorithm . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3 Interfacial Bending Elasticity 44<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.2 Interfacial Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.2.1 Mechanical Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

3.2.2 Thermodynamic Variables . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.2.3 Helfrich Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.2.4 Lifshitz Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.3 SCFT Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.4 Symmetric Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.5 Asymmetric Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4 Swollen Micelles and Interfacial Tension 70<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2 Micellization - Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.2.1 Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.2.2 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.2.3 Critical Micelle Concentration . . . . . . . . . . . . . . . . . . . . . 79<br />

4.2.4 Interfacial Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.2.5 Helfrich Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

vii


4.3 SCFT Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

4.4 Results - Micelle Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

4.5 Results - Helfrich Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

4.6 Other Possible Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

4.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5 Transport of Block Copolymer Surfactant Between Two Homopolymer<br />

Phases 101<br />

5.1 No Micelles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

5.2 Micelles in Matrix, c I c < c II<br />

c K . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

5.2.1 Micellar Diffusion (No Exclusion Zone) . . . . . . . . . . . . . . . . 106<br />

5.2.2 Molecular Diffusion (Exclusion Zone) . . . . . . . . . . . . . . . . . . 107<br />

5.3 Micelles in Wrong Matrix, c II<br />

c < c I c/K . . . . . . . . . . . . . . . . . . . . . 113<br />

6 Interfacial Tension of PI/PDMS with PI-b-PDMS Copolymer 118<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

6.2 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

6.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

6.2.2 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

6.2.3 Interfacial Tension Measurement . . . . . . . . . . . . . . . . . . . . 125<br />

6.2.4 Small Angle X-Ray Scattering . . . . . . . . . . . . . . . . . . . . . 126<br />

6.2.5 Dynamic Mechanical Spectroscopy . . . . . . . . . . . . . . . . . . . 127<br />

6.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

6.3.1 Bare Interfacial Tension . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

6.3.2 Dependence on Copolymer Concentration . . . . . . . . . . . . . . . 128<br />

6.3.3 Dependence on Copolymer Composition . . . . . . . . . . . . . . . . 130<br />

6.3.4 Dependence on Homopolymer Molecular Weight . . . . . . . . . . . 134<br />

viii


6.3.5 A Larger Copolymer . . . . . . . . . . . . . . . . . . . . . . . . . . . 135<br />

6.3.6 Bicontinuous Microemulsion Phase . . . . . . . . . . . . . . . . . . . 136<br />

6.4 Analys<strong>is</strong> of Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

6.4.1 Estimate of Diffusion Time . . . . . . . . . . . . . . . . . . . . . . . 144<br />

6.4.2 Can Micelles Reach the Interface? . . . . . . . . . . . . . . . . . . . 145<br />

6.4.3 Experimental Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

6.5 D<strong>is</strong>cussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150<br />

6.5.1 Equilibrium vs. Nonequilibrium . . . . . . . . . . . . . . . . . . . . . 152<br />

7 Micellization and Interfacial Tension in V<strong>is</strong>cous Ternary PS/PB/P(S-b-<br />

B) Blends 158<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

7.2 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

7.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

7.2.2 Blending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163<br />

7.2.3 Preparation of SDT Samples . . . . . . . . . . . . . . . . . . . . . . 163<br />

7.2.4 Interfacial Tension Measurement . . . . . . . . . . . . . . . . . . . . 164<br />

7.2.5 Transm<strong>is</strong>sion Electron Microscopy (TEM) . . . . . . . . . . . . . . . 166<br />

7.2.6 Small Angle X-ray Scattering (SAXS) . . . . . . . . . . . . . . . . . 166<br />

7.2.7 Dynamic Mechanical Spectroscopy . . . . . . . . . . . . . . . . . . . 166<br />

7.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167<br />

7.3.1 CMC in PS4K/SB Binary Mixtures . . . . . . . . . . . . . . . . . . 167<br />

7.3.2 Bare Interfacial Tension . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

7.3.3 Interfacial Tension Reduction . . . . . . . . . . . . . . . . . . . . . . 174<br />

7.3.4 Blend Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179<br />

7.4 D<strong>is</strong>cussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179<br />

ix


7.4.1 Kinetics Below the CMC . . . . . . . . . . . . . . . . . . . . . . . . 182<br />

7.4.2 Kinetics Above the CMC . . . . . . . . . . . . . . . . . . . . . . . . 182<br />

7.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186<br />

8 Summary and Outlook 189<br />

8.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189<br />

8.2 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192<br />

A Flory-Huggins Interaction Parameter between PS and PB 196<br />

B Estimate of Diffusion Coefficient of P(S-b-B) Block Copolymer in the<br />

Homopolymer Phase. 202<br />

C The Effect of the Uniaxial Extensional Flow on Interfacial Tension 206<br />

C.1 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207<br />

C.1.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207<br />

C.1.2 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . 208<br />

C.1.3 Extensional Rheology . . . . . . . . . . . . . . . . . . . . . . . . . . 209<br />

C.1.4 Data analys<strong>is</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209<br />

C.1.5 Shear Rheology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210<br />

C.2 Results and D<strong>is</strong>cussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210<br />

C.2.1 Extensional V<strong>is</strong>cosity of the pure Homopolymers . . . . . . . . . . . 212<br />

C.2.2 Shrinkage of Multilayers . . . . . . . . . . . . . . . . . . . . . . . . . 214<br />

C.2.3 Extensional V<strong>is</strong>cosity of Multilayers . . . . . . . . . . . . . . . . . . 215<br />

C.2.4 Interfacial Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215<br />

x


L<strong>is</strong>t of Figures<br />

1.1 Schematic phase pr<strong>is</strong>m of a water-oil-nonionic surfactant (C i E j ) system in<br />

a vertical ax<strong>is</strong> of temperature. A water-rich two phase, a bicontinuous microemulsion<br />

middle phase, and the oil-rich two phase are denoted by 2¯, 3,<br />

and ¯2, respectively. The three test tubes on the right-hand side illustrate<br />

how the phases are observed in experiment. An upper critical solution temperature<br />

T β <strong>is</strong> not necessarily located at higher temperature than T u .[73].<br />

Redrawn from Strey.[76] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.2 A schematic phase diagram along an <strong>is</strong>opleth of the phase pr<strong>is</strong>m. T o = T m<br />

and φ ∗ s <strong>is</strong> the minimum surfactant concentration <strong>to</strong> reach a uniform middle<br />

phase. The lower <strong>is</strong> φ ∗ s , the more efficient <strong>is</strong> the surfactant. Reproduced<br />

from Wennerström et al. .[86] . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.3 Isothermal diagram of water/n-octane/C 10 E 5 mixtures at the mean temperature<br />

T = 44.6 ◦ C . Note the 1-phase region near the middle phase enclosed<br />

by the lamellar phase. Reproduced from Strey.[77] . . . . . . . . . . . . . . 11<br />

1.4 Schematic diagram of the individual interfacial tension curves of oil/water(γ ab ),<br />

water/emulsion(γ ac ), and oil/emulsion(γ bc ) on a log scale as a function of<br />

temperature. Redrawn from Sottmann and Strey.[80] . . . . . . . . . . . . 12<br />

1.5 Reduced oil-water interfacial tension γ ∗ as a function of reduced temperature<br />

τ ∗ . The full line <strong>is</strong> calculated from Eqn. 1.2. The definition of γ ∗ and τ ∗<br />

are given in ref. [80]. Redrawn from Sottmann and Strey.[80] . . . . . . . . 13<br />

xi


1.6 Lens of C 8 E 3 -rich middle phase floating on the interface between the lower<br />

water- and upper decane-rich phase at T = 21.8 ◦ C . Reproduced from<br />

Kahlweit et al. .[75] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.7 Phase stability of spherical, cylindrical, and lamellar micelles. The crosshatched<br />

region <strong>is</strong> a two-phase coex<strong>is</strong>tence of spheres and cylinders. Sphere<br />

<strong>is</strong> the most stable structure along the emulsification failure line. In th<strong>is</strong> plot,<br />

r = ρ ˜C o and x = −¯κ/(2κ + ¯κ). Reproduced from Safran.[62] . . . . . . . . 15<br />

1.8 The compar<strong>is</strong>on of the <strong>to</strong>pology among different structures. Only (C) has<br />

g = 1 where as others <strong>have</strong> g = 0, where g <strong>is</strong> the number of handles.<br />

Reproduced from Taddei.[65] . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1 Interfacial tension γ vs. the number of spatial nodes in a simulation box<br />

with the fixed size of 800 Å. The adaptive grids dramatically increase the<br />

accuracy of the calculation compared <strong>to</strong> the case with the same number of<br />

grid points. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

2.2 The interfacial tension difference ∆γ as a function of h 2 for a uniform grid. 37<br />

2.3 The chemical potential difference ∆µ A and ∆µ C as a function of (∆t) 2 . . . 38<br />

2.4 Interfacial tension γ with various time steps ∆t for a uniform grid. There<br />

<strong>is</strong> little effect of time step on γ. . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.5 The chemical potential difference ∆µ A and ∆µ C of the homopolymer A and<br />

block copolymer C determined at both boundaries of the simulation cells<br />

vs. the spatial step size h. . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

xii


3.1 Schematic view of the state of a two-phase system at fixed pressure as a<br />

function of T and of a copolymer activity Z C ∝ e µ C/kT , which <strong>is</strong> defined for<br />

th<strong>is</strong> purpose so <strong>that</strong> Z C <strong>is</strong> proportional <strong>to</strong> the concentration of copolymer<br />

d<strong>is</strong>solved in one of the two coex<strong>is</strong>ting phases. The Lifshitz point <strong>is</strong> the<br />

intersection of a line of critical points, for which the copolymer concentration<br />

tends <strong>to</strong> increase with decreasing temperature, with a saturation line e µ∗ C (T) ,<br />

for which copolymer concentration decreases with decreasing temperature<br />

as the copolymer becomes less soluble. . . . . . . . . . . . . . . . . . . . . 52<br />

3.2 SCFT results for interfacial tension γ vs. half curvature C/2 = 1/R for<br />

spherical interfaces with R ≥ 100 Å, for system with the copolymer of a<br />

fixed size of B block f B χN C = 10 and symmetric homopolymers of fixed<br />

lengths α A = α B = f B , with b A /b B = 1, for several values of f A . Solid<br />

lines are quadratic fits of the curvature dependence at small curvatures <strong>to</strong><br />

Eqn. 3.8, from which κ + and τ can be extracted. . . . . . . . . . . . . . . 56<br />

3.3 SCFT results for interfacial tension γ vs. curvature C = 1/R for cylindrical<br />

interfaces with R ≥ 100 Å, for the same mixtures as those studied in the<br />

previous figure. κ can be obtained. . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.4 The non-dimensional bending rigidity [κ + ] of a sphere vs. χN in symmetric<br />

mixtures with various values of α = α A = α B at fixed β = 1. . . . . . . . . 59<br />

3.5 The normalized bending rigidity [κ] vs. χN for symmetric systems with<br />

several values of α. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

3.6 The normalized Gaussian rigidity [¯κ] vs. χN for symmetric mixtures with<br />

several values of α. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

3.7 The ratio of the Gaussian rigidity <strong>to</strong> bending rigidity −[¯κ]/[κ] vs. χN for<br />

symmetric mixtures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

xiii


3.8 Bending parameter τ of block copolymer monolayer vs. a block ratio f A of<br />

the A block for systems with f B χN = α A χN = α B χN = 10. The system in<br />

th<strong>is</strong> plot <strong>is</strong> identical <strong>to</strong> <strong>that</strong> in Fig. 3.2 and 3.3. f bal<br />

A<br />

<strong>is</strong> always approximated<br />

as a limiting value linearly extrapolated <strong>to</strong> C = 0 as shown here. . . . . . . 63<br />

3.9 Bending rigidity κ + of sphere vs. a block ratio f A of the A block for systems<br />

with f B χN = α A χN = α B χN = 10. In th<strong>is</strong> article, κ + <strong>is</strong> always<br />

approximated as a limiting value linearly extrapolated <strong>to</strong> C = 0 as shown<br />

here. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.10 The variation of the balance point f bal<br />

A<br />

with fixed α = 1. The solid line<br />

indicates the balance point f L at an <strong>is</strong>otropic Lifshitz point. . . . . . . . . 64<br />

3.11 The variation of the balance point f bal<br />

A<br />

for asymmetric mixtures with a fixed<br />

minority block size f B χN C = 10. The solid line indicates the balance point<br />

f L at an <strong>is</strong>otropic Lifshitz point. . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.12 The variation of the balance point f bal<br />

A<br />

with asymmetric stat<strong>is</strong>tical segment<br />

length b. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

3.13 Rigidities κ + , κ, and ¯κ vs. β = α B /α A for systems with a copolymer<br />

minority block of size f B χN = 10, a B homopolymer of length α B χN = 5,<br />

and b A = b B , for varying size A homopolymers. Values for κ, κ + , and ¯κ are<br />

given in units with kT = 1 on the left scale, and corresponding values [κ],<br />

[κ + ], and [¯κ] on the right. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

xiv


4.1 Schematic phase diagram of a ternary system compr<strong>is</strong>ed of homopolymers A<br />

and B and diblock copolymer C, in which an A rich phase of swollen micelles<br />

(or, equivalently, microemulsion droplets) coex<strong>is</strong>ts with a phase of nearly<br />

pure B. The dotted line b-c <strong>is</strong> a line of the critical micelle concentration<br />

in a single A-rich phase. Micelles are drawn above th<strong>is</strong> line. A point b<br />

corresponds <strong>to</strong> the cmc in two phase ternary system. . . . . . . . . . . . . 73<br />

4.2 Dependence of concentrations of free copolymer and of copolymer within<br />

micelles upon the <strong>to</strong>tal concentration of copolymer within the micellar phase,<br />

along the coex<strong>is</strong>tence line. . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

4.3 Dependence of copolymer chemical potential upon <strong>to</strong>tal concentration of<br />

copolymer within the micellar phase, along the coex<strong>is</strong>tence line. The dotted<br />

line shows the continuation of ideal solution behavior beyond the cmc . . . 75<br />

4.4 Schematic of the ratio γ/γ 0 of macroscopic interfacial tension γ <strong>to</strong> its value<br />

γ 0 in the absence of copolymer, vs. the <strong>to</strong>tal volume fraction of copolymer<br />

within the A-rich micellar phase. As in Fig. 4.3, the solid line shows the<br />

behavior of a micellar solution, and the dotted line shows the continuation<br />

of ideal solution behavior, in the absence of micelles. . . . . . . . . . . . . 76<br />

4.5 Copolymer volume fraction φ cmc<br />

C<br />

at the critical micelle concentration of a<br />

two-phase ternary system (solid lines), and of a binary mixture of copolymer<br />

C in A (short dashed lines), vs. volume fraction f A of the corona<br />

block within the copolymer, for systems with homopolymer sizes α A χN =<br />

α B χN = 10, for three values of χN CB . Also shown <strong>is</strong> the copolymer volume<br />

fraction φ ∗ C<br />

(dot dashed line) at which the macroscopic interfacial tension<br />

of the two phase system extrapolates <strong>to</strong> zero. Note <strong>that</strong> φ ∗ C<br />

always exceeds<br />

the cmc φ cmc<br />

C<br />

of ternary two-phase system, but φ ∗ C<br />

drops below the cmc of<br />

the corresponding binary system for nearly symmetric copolymers. . . . . 87<br />

xv


4.6 Micelle core radius at the cmc for the ternary two-phase system and binary<br />

system considered in Fig. 4.5, vs. volume fraction f A of the corona block.<br />

The hollow and solid symbols represent the radii of unswollen and swollen<br />

micelles, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

4.7 Macroscopic interfacial tension γ cmc at the cmc, normalized by the value γ 0<br />

for a bare interface, vs. volume fraction f A of the copolymer corona block,<br />

for the same parameters as those used in Fig. 4.5. Leibler’s prediction based<br />

on the unswollen micelle <strong>is</strong> shown with a dotted line for compar<strong>is</strong>on. . . . 89<br />

4.8 Macroscopic interfacial tension γ cmc at the cmc, normalized by the bare tension<br />

γ 0 , vs. copolymer corona volume fraction f C for system with f B χN =<br />

10 and various values of the ratio β of copolymer molecular weights. . . . . 90<br />

4.9 Dependence of critical micelle concentration φ cmc<br />

C<br />

on χN BC = f B χN for<br />

systems with f A = 0.6 and α A χN = α B χN = 10. Symbols show predictions<br />

for the ternary two phase system by SCFT (filled triangles) and predictions<br />

for a binary system by SCFT (filled circles) and by the approximate strong<br />

LOW theory (open triangles). The dashed line <strong>is</strong> the function Ce −χN BC,<br />

with C = 376. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

4.10 Compar<strong>is</strong>on of SCFT predictions for copolymer cmc φ cmc<br />

C<br />

of a ternary twophase<br />

system (filled triangles) and a binary system (open triangles) <strong>to</strong> the<br />

strong stretching predictions for the binary system by Leibler (dot dashed<br />

line) and LOW (solid lines) for the systems considered in Fig. 4.5. . . . . . 91<br />

4.11 The prediction of γ/γ o from direct micelle simulation (hollow circle) and<br />

from Helfrich bending elasticity (solid circle) as a function of f A . They are<br />

cons<strong>is</strong>tent for f A < 0.55 but start <strong>to</strong> deviate for larger f A . Surpr<strong>is</strong>ingly, the<br />

quadratic function (solid line) of γ/γ o (Eqn. 4.23) with τ ′ and κ + evaluated<br />

at C = 0 accurately fits the entire results from the direct micelle simulation. 92<br />

xvi


4.12 The compar<strong>is</strong>on of micelle radii obtained from direct micelle simulation (hollow<br />

circle) and from Helfrich bending elasticity (solid circle) as a function of<br />

f A . They are cons<strong>is</strong>tent for f A < 0.55 but the latter starts <strong>to</strong> underestimate<br />

the radii for larger f A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

4.13 The normalized slope [τ ′ ] as a function of χN in a range of 0.25 ≤ α ≤ 2 . 94<br />

4.14 The prefac<strong>to</strong>r of γ flat ,<br />

τ ′2<br />

8κ + γ o<br />

, normalized by a bare interfacial tension γ o as<br />

a function of χN CB in a range of 0.25 ≤ α ≤ 2 . . . . . . . . . . . . . . . . 94<br />

5.1 A schematic diagram of concentration profile of surfactant perpendicular <strong>to</strong><br />

an interface z = 0 below the cmc. . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

5.2 A schematic diagram of concentration profile of surfactant and micelle perpendicular<br />

<strong>to</strong> an interface z = 0 without exclusion zone (Q < 1) above the<br />

cmc. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

5.3 A schematic diagram of concentration profile of surfactant perpendicular <strong>to</strong><br />

an interface z = 0 for Q > 1 above the cmc. A solid line <strong>is</strong> for the case of<br />

S ≫ 1 and a dotted line for the case of S ≪ 1. . . . . . . . . . . . . . . . . 108<br />

5.4 A schematic diagram of concentration profile of surfactant when surfactant<br />

<strong>is</strong> initially added <strong>to</strong> the wrong phase (I) (a) when Eqn. 5.21 and Eqn. 5.22<br />

are sat<strong>is</strong>fied (b) not sat<strong>is</strong>fied. . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

6.1 The transient diameter of the PI drop in the PDMS matrix as a function<br />

of time. (a) PI1 and PI6 in the absence of block copolymer (b) PI4 in<br />

PDMS1/IDMS6 and in PDMS2/IDMS12. The concentrations of IDMS6<br />

and IDMS12 in the corresponding PDMS matrices are 0.1 wt % and 0.2 wt<br />

%, respectively. The angular velocity <strong>is</strong> 6300 rpm. . . . . . . . . . . . . . . 128<br />

xvii


6.2 Bare interfacial tension between PI and PDMS1 in the absence of block<br />

copolymer as a function of molecular weights of PI. The solid line <strong>is</strong> the<br />

SCFT prediction using χ = 0.175. . . . . . . . . . . . . . . . . . . . . . . . 129<br />

6.3 The interfacial tension reduction of PI4/PDMS1 and PI4/PDMS2 as a function<br />

of the concentration of IDMS1 and IDMS12, respectively. The dashed<br />

lines are guides <strong>to</strong> the eye. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

6.4 The interfacial tension reduction between PI4 and PDMS1 as a function of<br />

IDMS1. The solid lines are the SCFT prediction with four different χ’s. . . 131<br />

6.5 Interfacial tension in PI4/PDMS1/IDMS system as a function of the block<br />

ratio f A of IDMS. The quadratic dependence of interfacial tension on f A<br />

agrees well with the SCFT prediction and the swollen micelle theory (hollow<br />

dots) for f A > 0.65. Leibler’s theory[23] (a dotted line) predicts <strong>that</strong> a wider<br />

range of f A will cause van<strong>is</strong>hing interfacial tension. . . . . . . . . . . . . . 133<br />

6.6 Data and calculations of Fig. 6.5 plotted on a logarithmic scale. Near a<br />

balance point, measured values are much lower than those predicted by<br />

SCFT. The concentrations of block copolymer are all greater than the cmc. 133<br />

6.7 The time-dependent diameter of a PI4 drop in a PDMS1 matrix at an<br />

angular velocity of 6300 rpm with two different concentrations of IDMS7<br />

(f A = 0.6). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

6.8 Interfacial tension in the system of PDMS1/PI/IDMS1 as a function of<br />

molecular weight of PI. The concentration of IDMS1 <strong>is</strong> 0.1 wt %. . . . . . 137<br />

6.9 Interfacial tension in the system of PDMS2/PI/IDMS13 as a function of<br />

molecular weight of PI. The concentration of IDMS13 <strong>is</strong> 0.2 wt %. The<br />

minimum of the interfacial tension <strong>is</strong> not ultra-low, but much greater than<br />

<strong>that</strong> in Fig. 6.8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

xviii


6.10 The variation of the diameter of the PI5 drop in the PDMS2 matrix at a<br />

concentration of 0.2 wt % of IDMS13. The rotational speed <strong>is</strong> 6300 rpm. . 138<br />

6.11 The ternary <strong>is</strong>opleth blends of the system PDMS1/PI4/IDMS3 for the various<br />

concentrations of block copolymer: 5, 25, 32, 45, 57, and 70 wt %<br />

from the leftmost vial. The samples are turbid below 30 wt % due <strong>to</strong> the<br />

macrophase separation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

6.12 The ternary <strong>is</strong>opleth blends of the system PDMS2/PI5/IDMS13 for the<br />

various concentrations of block copolymer: 10, 20, 30, 40, 50, and 70 wt %<br />

from the leftmost vial. The sample with 20 wt % of IDMS13 turns blu<strong>is</strong>h. 139<br />

6.13 The SAXS profiles for the ternary <strong>is</strong>opleth blends of PDMS1/PI4/IDMS3<br />

as a function of the concentration of IDMS3. All curves are presented on<br />

arbitrary linear scales for clarity. The primary peak continuously moves<br />

<strong>to</strong> a small q region as the mixtures are diluted with the homopolymers.<br />

Two peaks are d<strong>is</strong>tinct between 50 % and 55 % indicating a bicontinuous<br />

microemulsion phase coex<strong>is</strong>t with a lamellar phase. . . . . . . . . . . . . . 140<br />

6.14 The SAXS profiles for the ternary <strong>is</strong>opleth blends of PDMS2/PI5/IDMS13<br />

as a function of the concentration of IDMS13. The primary peaks of IDMS13<br />

appear at lower q than those of IDMS3 in Fig. 6.13 at the same concentration<br />

due <strong>to</strong> the twice longer block copolymer. . . . . . . . . . . . . . . . . . . . 140<br />

6.15 The variation of the domain spacing of a lamellar phase (ξ L ) and a bicontinuous<br />

microemulsion phase (ξ µE ) along the <strong>is</strong>opleth in the system of<br />

PDMS1/PI4/IDMS3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

6.16 The ratio of the pure lamellar spacing ((ξ Lo ) <strong>to</strong> the spacing at q = q ∗<br />

multiplied by a weight fac<strong>to</strong>r α as a function of the concentration of IDMS3.<br />

ξ µE /ξ L = 1.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

xix


6.17 The proposed phase diagram of the system PDMS1/PI4/IDMS3 in Fig. 6.13.<br />

Large three phase region ex<strong>is</strong>ts below 30 % of block copolymer. . . . . . . 143<br />

6.18 Time t <strong>to</strong> reach a center of spinning drop tensiometer drop by molecular<br />

diffusion as a function of interfacial tension γ. . . . . . . . . . . . . . . . . 146<br />

6.19 The time-independent interfacial concentration and the partition coefficient<br />

as a function of the block ratio f A in the block copolymer. The effect of the<br />

micelle diffusion on c int<br />

f<br />

<strong>is</strong> not considered. . . . . . . . . . . . . . . . . . . . 148<br />

6.20 The radii R of unswollen/swollen micelles and the dimensionless ratio Q<br />

with respect <strong>to</strong> f A . R <strong>is</strong> taken from the micelle simulation with f B χN c = 12<br />

in Fig. 4.6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148<br />

7.1 GPC curves of poly(styrene-b-butadiene) block copolymer before and after<br />

fractionation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

7.2 The sample preparation for a spinning drop tensiometer in a the heating<br />

oven. The heating tape <strong>is</strong> removed for clear demonstration. . . . . . . . . . 165<br />

7.3 The experimental setup <strong>to</strong> inject a PB drop in<strong>to</strong> a PS matrix. The region<br />

near a needle of a syringe <strong>is</strong> blown up at the right corner. . . . . . . . . . . 165<br />

7.4 TEM images of the binary blends of PS4K/SB1 at (a) 0.125 wt% (b) 0.25<br />

wt% (c) 0.5 wt% (d) 1 wt% (e) 2 wt% (f) 3 wt% (g) 4 wt% (h) 6 wt%. The<br />

scale bars are 50 nm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169<br />

7.5 The critical micelle concentration in the binary system of PS4K/SB1 determined<br />

by TEM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170<br />

7.6 TEM images of the binary blends of PS4K/SB2 at (a) 0.125 wt % (b) 0.25<br />

wt % (c) 0.5 wt % (d) 1 wt %. The scale bars are 100 nm. . . . . . . . . . 171<br />

7.7 TEM images of micelles in the PS4K matrix (a) 0.5 wt % SB3 and (b) 1 wt<br />

% SB4. The scale bars are 50 nm and 100 nm, respectively. . . . . . . . . 172<br />

xx


7.8 The normalized SAXS intensity in the binary system of PS4K/SB1. . . . . 173<br />

7.9 The critical micelle concentration in the binary system of PS4K/SB1 determined<br />

by SAXS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173<br />

7.10 Bare interfacial tension between PS4K and PB as a function of the molecular<br />

weights of PB. The experimental results are bound by the upper and lower<br />

limits in the SCFT prediction. The dotted line <strong>is</strong> given with χ = 0.054 and<br />

the solid line <strong>is</strong> with χ = 0.080, respectively. . . . . . . . . . . . . . . . . . 175<br />

7.11 Bare interfacial tension between PS12K and PB. . . . . . . . . . . . . . . . 175<br />

7.12 The interfacial tension reduction of PS4K and PB13K vs. the concentration<br />

of SB1 <strong>that</strong> <strong>is</strong> premixed with PS4K (circle) or PB13K (triangle). The solid<br />

line <strong>is</strong> the SCFT prediction with χ = 0.064 and the dashed line <strong>is</strong> with<br />

χ = 0.080. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177<br />

7.13 The change in a diameter of the PB drop as a function of time (min) in<br />

spin-up experiments. The data are plotted on a linear scale. . . . . . . . . . 178<br />

7.14 The change in a diameter of the PB drop as a function of time (min) in<br />

spin-up experiments. The data are plotted on a logarithmic scale. . . . . . . 178<br />

7.15 TEM images of the ternary blends of PS4K/PB13K/SB1 at various concentrations<br />

of SB1. (a) 1 wt% (b) 2 wt% (c) 4 wt% (d) 6 wt%. The scale bar<br />

corresponds <strong>to</strong> 400 nm in (a) and (b), and 200 nm in (c) and (d), respectively. 180<br />

7.16 The variation of Q as a function of the block ratio f A . The molecular weights<br />

of the homopolymers are fixed and c o m /c c = 5. f A of SB1 <strong>is</strong> indicated as an<br />

arrow. essentially identical <strong>to</strong> the CMC. . . . . . . . . . . . . . . . . . . . . 183<br />

7.17 The dependence of Q on the micelle concentration. . . . . . . . . . . . . . 183<br />

xxi


A.1 The Flory-Hugggins interaction parameter χ between PS and PB with various<br />

% 1,2 of PB in the literature. All values are based on a fixed v = 81.5<br />

cm 3 /mol. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199<br />

C.1 The extensional v<strong>is</strong>cosity of the PS homopolymer at different temperatures<br />

of 170, 180, 185 ◦ C . The extensional rate <strong>is</strong> 0.01 s −1 . Curves are averages<br />

of five different runs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211<br />

C.2 The extensional v<strong>is</strong>cosity of the PP homopolymer at 170, 180, and 185 ◦ C<br />

. The extensional rate <strong>is</strong> 0.01 s −1 . . . . . . . . . . . . . . . . . . . . . . . . 211<br />

C.3 The extensional v<strong>is</strong>cosity η ǫ of the pure PS homopolymer and the 2 wt %<br />

binary mixtures of PS(SEE) at extensional rates ˙ǫ of 0.01, 0.05, and 0.1 s −1<br />

at 180 ◦ C . Note <strong>that</strong> η ǫ of PS(SEE) <strong>is</strong> greatly reduced from <strong>that</strong> of PS due<br />

<strong>to</strong> SEE block copolymer. The extensional v<strong>is</strong>cosity of PS <strong>is</strong> cons<strong>is</strong>tent with<br />

3η + (t), which <strong>is</strong> given by a solid line. η ǫ of PS(SEE) <strong>is</strong> a weak function of<br />

˙ǫ approaching <strong>that</strong> of the pure PS. . . . . . . . . . . . . . . . . . . . . . . . 213<br />

C.4 The extensional v<strong>is</strong>cosity of the PP homopolymer at extensional rates of<br />

0.01, 0.05, and 0.1 s −1 . Experimental temperature <strong>is</strong> 180 ◦ C . The solid line<br />

represents 3η + (t) of PP in the LVE regime. . . . . . . . . . . . . . . . . . . 213<br />

C.5 Shrinkage in width of multilayer samples during annealing in the RME on<br />

a semi-logarithmic time scale. . . . . . . . . . . . . . . . . . . . . . . . . . 214<br />

C.6 The extensional v<strong>is</strong>cosity of the PS/PP 32-multilayer at extensional rates of<br />

0.01, 0.05, and 0.1 s −1 . Experimental temperature <strong>is</strong> 180 ◦ C . . . . . . . . 216<br />

C.7 Summary of the extensional v<strong>is</strong>cosities of all samples used in th<strong>is</strong> paper at<br />

˙ǫ = 0.01s −1 . Block copolymer reduces the extensional v<strong>is</strong>cosity by a fac<strong>to</strong>r<br />

of two. Three times zero shear v<strong>is</strong>cosities in the LVE region are given for<br />

the validity of the measurements. . . . . . . . . . . . . . . . . . . . . . . . 216<br />

xxii


C.8 Interfacial tension extracted from the multilayer samples as a function of<br />

time with ˙ǫ = 0.01, 0.05, and 0.1 s −1 . The solid line corresponds <strong>to</strong> equilibrium<br />

interfacial tension γ eq measured with a pendent drop method.[15] . . 218<br />

C.9 Maximum interfacial tension vs. the extensional rate. The linear dependence<br />

of γ on ˙ǫ can be easily observed. . . . . . . . . . . . . . . . . . . . . 218<br />

C.10 Interfacial v<strong>is</strong>cosity η ex = (γ − γ eq )/˙ǫ as a function of time for ˙ǫ = 0.01,<br />

0.05, and 0.1 s −1 . At t → 0, γ → γ eq . . . . . . . . . . . . . . . . . . . . . . 219<br />

C.11 Interfacial tension in the presence of block copolymer. It <strong>is</strong> interesting <strong>that</strong><br />

block copolymer increases interfacial tension. Only one experiment was<br />

conducted with ˙ǫ = 0.05s −1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 219<br />

xxiii


L<strong>is</strong>t of Tables<br />

6.1 Material Character<strong>is</strong>tics of Homopolymers . . . . . . . . . . . . . . . . . . . 123<br />

6.2 Material Character<strong>is</strong>tics of Block Copolymers . . . . . . . . . . . . . . . . . 124<br />

7.1 Material Character<strong>is</strong>tics of Homopolymers . . . . . . . . . . . . . . . . . . 160<br />

7.2 Material Character<strong>is</strong>tics of Block Copolymers . . . . . . . . . . . . . . . . 160<br />

7.3 The Reduction of Interfacial Tension . . . . . . . . . . . . . . . . . . . . . 179<br />

A.1 Summary of the Interaction Parameters Between PS and PB . . . . . . . . 198<br />

C.1 Extensional V<strong>is</strong>cosity at t = 10 s. . . . . . . . . . . . . . . . . . . . . . . . . 212<br />

xxiv


Chapter 1<br />

Introduction and Background<br />

1.1 Introduction<br />

Polymer blending has been of great interest in the polymer industry during the past three<br />

decades because it provides a convenient way <strong>to</strong> develop new materials. A major advantage<br />

<strong>is</strong> <strong>that</strong> it <strong>is</strong> a time and cost effective route <strong>to</strong> commercialization of products compared with<br />

the synthes<strong>is</strong> of an entirely new polymer. Polymer blends <strong>have</strong> achieved a sizable growth in<br />

au<strong>to</strong>motive and electric/electronic device markets, which <strong>to</strong>gether account for over 90 % of<br />

the <strong>to</strong>tal volume for polymer blends. Packaging, medical, household, and sports/recreation<br />

goods make up the remainder.[1, 2]<br />

With the exception of a few m<strong>is</strong>cible polymer systems <strong>that</strong> <strong>have</strong> been studied or<br />

commercialized,[3] most polymers are inherently imm<strong>is</strong>cible. Long polymer chains <strong>have</strong><br />

a low entropy of mixing, so <strong>that</strong> even a small positive enthalpic contribution <strong>to</strong> the <strong>to</strong>tal<br />

free energy can cause the polymer blends <strong>to</strong> macroscopically phase separate. Accordingly,<br />

a challenging task among researchers has been <strong>to</strong> attain a fine and uniform d<strong>is</strong>persion of<br />

components in the melt state which usually leads <strong>to</strong> enhancement of the interfacial strength<br />

in the solid state.<br />

The external flow field generated by mechanical mixing <strong>is</strong> a very powerful <strong>to</strong>ol in<br />

accompl<strong>is</strong>hing the above goals through the domain break-up process. For a New<strong>to</strong>nian<br />

drop under simple shear or extensional flow in the New<strong>to</strong>nian matrix, the droplet size <strong>is</strong> a<br />

1


function of the v<strong>is</strong>cosity ratio η r = η d /η m , where η d and η m <strong>is</strong> the v<strong>is</strong>cosity of the droplet<br />

and the matrix, respectively, and the capillary number Ca = dη m ˙ǫ/2γ where d <strong>is</strong> a drop<br />

diameter, ˙ǫ <strong>is</strong> a deformation rate, and γ <strong>is</strong> interfacial tension.[4]<br />

The minimum stable size of an <strong>is</strong>olated drop achieved by the balance between the shear<br />

force and interfacial tension force, known as the Taylor limit,[5, 6] <strong>is</strong><br />

d = γ<br />

˙ǫη m<br />

16η r + 16<br />

19η r + 16<br />

(1.1)<br />

for η r < 2.5. There <strong>have</strong> been many experiments <strong>to</strong> apply th<strong>is</strong> idea <strong>to</strong> polymer blends<br />

containing a small amount of the droplet phase by controlling η r and ˙ǫ. The diameter d<br />

in Eqn. 1.1 has often provided an asymp<strong>to</strong>tic limit of the drop size obtainable in dilute<br />

homopolymer blends. However, the morphology of highly v<strong>is</strong>cous polymer blends depends<br />

upon the complex flow h<strong>is</strong><strong>to</strong>ry during mixing.[7, 8, 9]<br />

An additional <strong>is</strong>sue in achieving fine d<strong>is</strong>persion <strong>is</strong> the problem of the stability of polymer<br />

blend morphology. Most imm<strong>is</strong>cible homopolymer blends with concentrated droplets<br />

are not stable against thermal annealing and suffer from drastic coalescence of droplets.<br />

One simple method <strong>to</strong> reduce interfacial tension and improve the thermal stability as well<br />

as decreasing the drop size (Eqn. 1.1) <strong>is</strong> <strong>to</strong> add a small amount of AB block copolymers<br />

<strong>to</strong> A and B homopolymer blends. [10, 11, 12, 13] Sometimes AB copolymer <strong>is</strong> replaced<br />

by AC or CD copolymer when specific interaction between monomers ex<strong>is</strong>ts.[14, 15] The<br />

role of block copolymer <strong>is</strong> analogous <strong>to</strong> <strong>that</strong> of a small molecule surfactant in emulsifying<br />

the mixtures of oil and water. Since each block has a favorable interaction with the corresponding<br />

homopolymer phase, block copolymers preferentially segregate at the interface<br />

and reduce the excess free energy of the interface. The traditional strategies for compatibilization<br />

(emulsification) of polymer blends with block copolymers. Copolymers can be<br />

either premade or created in situ from the reaction during mixing.[2, 7, 16] Although the<br />

latter <strong>is</strong> more often used in industry due <strong>to</strong> the kinetic advantage of directly creating block<br />

2


copolymer at the interface, the underlying principles <strong>that</strong> govern the thermodynamics of<br />

polymer blends should be identical in both cases.<br />

Block copolymers can modify interfacial properties in three ways:[2, 7, 9] 1) reduce<br />

interfacial tension, thus promoting drop break-up, 2) suppress particle coalescence and thus<br />

stabilize morphology, and 3) increase the interfacial adhesion in the solid state by bridging<br />

the d<strong>is</strong>tinct domains. In contrast <strong>to</strong> the linear dependence of the size of a New<strong>to</strong>nian droplet<br />

on the interfacial tension (Eqn. 1.1), recent theoretical [17] and experimental [8, 10, 11, 18]<br />

studies point out <strong>that</strong> the relatively small decrease in interfacial tension <strong>is</strong> not enough<br />

<strong>to</strong> account for the huge reduction of the drop size. Rather the suppression of particle<br />

coalescence has a dominant effect on the morphology development during melt blending.<br />

Lyu et al. showed <strong>that</strong> the minimum interfacial coverage of block copolymer required <strong>to</strong><br />

prevent drop coalescence <strong>is</strong> about 20 - 50 % of <strong>that</strong> required <strong>to</strong> form a lamellar phase<br />

of pure block copolymer [10] and nearly independent of shear rate. They attributed th<strong>is</strong><br />

result <strong>to</strong> the steric repulsion between block copolymer brushes.[8, 11, 19]<br />

When the copolymer concentration exceeds the critical micelle concentration in either<br />

homopolymer phase, block copolymer added in<strong>to</strong> the homopolymer blends starts <strong>to</strong> form<br />

micelles rather than segregating at an interface.[11, 20, 21, 39] Since micellization causes<br />

the chemical potential of block copolymer <strong>to</strong> become independent of copolymer concentration,<br />

the interfacial tension reduction will level off without further accumulation of block<br />

copolymer at an interface. Thus, from an economic point of view, excess block copolymers<br />

beyond the CMC are wasted as micelles.<br />

Since early 1980s, numerous theoretical studies <strong>have</strong> investigated the reduction of interfacial<br />

tension by addition of copolymer <strong>to</strong> binary homopolymer blends.[22, 23, 24, 28, 29, 30,<br />

39, 40, 41] or the formation of a spherical micelle in a selective solvent[39, 45, 46, 49, 50, 51]<br />

or a pure homopolymer melt.[39, 47, 48, 52, 53] Self cons<strong>is</strong>tent field theory (SCFT) has<br />

been used <strong>to</strong> predict the equilibrium density profiles of each component in addition <strong>to</strong><br />

3


earlier work based on analytical approximations.[39, 44, 45, 46, 48] Both numerical SCFT<br />

and analytical approximations <strong>have</strong> predicted <strong>that</strong> sufficiently high coverages of copolymer<br />

surfactants should be able <strong>to</strong> drive interfacial tension <strong>to</strong> zero.[22, 23, 24, 28, 29, 30, 40, 41]<br />

These results <strong>have</strong> inspired several researchers <strong>to</strong> directly measure interfacial tension using<br />

several techniques, including the spinning drop,[31] the pendant drop,[20, 33, 34, 35, 36]<br />

and the breaking thread method.[9, 37] These methods <strong>have</strong> been applied <strong>to</strong> various polymer<br />

pairs mixed with diblock,[20, 32, 33] triblock,[34] graft,[34] random,[35] three-arm[36],<br />

and gradient[40] copolymers. However, none of these experiments achieved the van<strong>is</strong>hing<br />

interfacial tension theoretically predicted for symmetric copolymers. Most measurements<br />

showed ∼ 50 % reduction of the original bare tensions [8, 20, 37, 38] up <strong>to</strong> a maximum<br />

reported reduction of ∼ 90 %.[34, 56]<br />

Most theoretical studies <strong>have</strong> only focused on either interfacial adsorption[1, 24], or<br />

the properties of micelles in binary solutions [4, 5, 9, 48, 49, 52] although both interfacial<br />

adsorption and micellization are always competitive in the ternary polymer blends.<br />

Leibler attempted <strong>to</strong> link both phenomena by equating the chemical potential of block<br />

copolymer at the CMC with <strong>that</strong> in a copolymer monolayer of copolymer adsorbed <strong>to</strong> a<br />

flat interface. He concluded <strong>that</strong> van<strong>is</strong>hing interfacial tension can precede micellization as<br />

long as block copolymer has the block ratio f A ranged from 0.31 <strong>to</strong> 0.69.[39] Unfortunately,<br />

h<strong>is</strong> theory neglected an important feature of micelles in ternary polymer systems, <strong>that</strong> <strong>is</strong>,<br />

a micelle can swell by emulsifying the minority homopolymer phase in its core. In Chapter<br />

4, we show <strong>that</strong> th<strong>is</strong> effect quantitatively changes the results.<br />

More recent studies <strong>have</strong> begun <strong>to</strong> recognize <strong>that</strong> hindered block copolymer diffusion<br />

could <strong>have</strong> a significant effect on interfacial tension measurement. [36, 55, 56] In experiments<br />

in which copolymers were premixed with either the drop or the matrix, they all<br />

observed <strong>that</strong> measured interfacial tensions were greatly dependent on the homopolymer<br />

phase <strong>to</strong> which copolymers were initially added. Some reported <strong>that</strong> the time <strong>to</strong> reach<br />

4


steady state was retarded by several orders of magnitude in highly v<strong>is</strong>cous system compared<br />

with the case in the absence of copolymer.[54, 57]<br />

Interpretation of the experimental data without an adequate understanding of thermodynamics<br />

and kinetic fac<strong>to</strong>rs often leads <strong>to</strong> very perplexing or apparently incons<strong>is</strong>tent<br />

conclusions. An example <strong>is</strong> the effect of the block copolymer length on reducing the domain<br />

size in imm<strong>is</strong>cible blends. Cigana et al. reported <strong>that</strong> small molecular weight poly(styreneb-ethylene-butylene-b-styrene)<br />

(SEBS) triblock copolymers were more efficient in producing<br />

PS/EPR blends with smaller domain size and higher impact strengths.[58] However,<br />

other researchers found <strong>that</strong> the drop morphology after mechanical mixing was nearly insensitive<br />

<strong>to</strong> the molecular weight of copolymer, but more efficiently stabilized with large<br />

copolymer.[9, 11, 25] Macosko and coworkers <strong>have</strong> suggested the ex<strong>is</strong>tence of an optimum<br />

molecular weight of block copolymer <strong>to</strong> achieve the maximum stability upon annealing.<br />

[8, 10, 12, 13]<br />

1.2 Overview of The Thes<strong>is</strong><br />

<strong>Th<strong>is</strong></strong> thes<strong>is</strong> pursues the ultimate goal of finding an optimum thermodynamic condition<br />

<strong>to</strong> emulsify imm<strong>is</strong>cible homopolymer blends by adding block copolymer surfactant, and<br />

largely cons<strong>is</strong>ts of two parts: the prediction of self cons<strong>is</strong>tent field theory (SCFT) in Chapter<br />

2, 3, and 4, and the experimental verification of the theoretical results in Chapter 5, 6<br />

and, 7.<br />

The rest of th<strong>is</strong> thes<strong>is</strong> <strong>is</strong> organized as follows. For better understanding of the thermodynamics<br />

of ternary polymer blends, we turn our attention <strong>to</strong> analogous systems of<br />

oil/water/nonionic surfactant in the subsequent sections of th<strong>is</strong> chapter. Based on wellknown<br />

phenomena in the small molecule surfactant system, we try <strong>to</strong> find a fundamental<br />

physics <strong>to</strong> govern the polymer system. Chapter 2 describes the underlying assumption and<br />

numerical implementation of SCFT in a grand canonical ensemble for ternary polymer<br />

5


systems containing A, B homopolymers and AB block copolymer.<br />

In Chapter 3, SCFT <strong>is</strong> used <strong>to</strong> calculate elastic parameters of the Canham-Helfrich<br />

theory. The simulations <strong>have</strong> been conducted for spherical and cylindrical interfaces of<br />

varying curvature in order <strong>to</strong> extract the bending rigidities and spontaneous curvature.<br />

The resulting quantities are used <strong>to</strong> search for a balance point where the spontaneous<br />

curvature of a monolayer van<strong>is</strong>hes.<br />

In Chapter 4, we study the thermodynamic competition between interfacial adsorption<br />

and micellization of block copolymers. Independent simulations are carried out at a<br />

flat interface and swollen micelles at equal chemical potentials in order <strong>to</strong> calculate the<br />

interfacial tension obtained at concentrations above the CMC.<br />

Chapter 5 <strong>is</strong> an analys<strong>is</strong> of the diffusion of block copolymer which <strong>is</strong> initially mixed with<br />

one of the homopolymer phases and the effect of transport limitations upon the measured<br />

interfacial tension.<br />

In Chapter 6, interfacial tension <strong>is</strong> measured between 1,4 poly<strong>is</strong>oprene (PI) and poly<br />

(dimethyl siloxane) (PDMS) in the presence of poly( 1,4 <strong>is</strong>oprene-b-dimethyl siloxane) (IDMS)<br />

block copolymer. The spontaneous curvature of a copolymer monolayer <strong>is</strong> controlled by<br />

changing the block ratio of IDMS block copolymer or the molecular weight of the PI<br />

homopolymers.<br />

In Chapter 7, we measure interfacial tension of polystyrene and 1,2 polybutadiene (PB)<br />

by adding poly(styrene-b- 1,2 butadiene) block copolymer in the PS matrix in Chapter 7.<br />

The critical micelle concentration <strong>is</strong> independently identified using transm<strong>is</strong>sion electron<br />

microscopy and small angle X-ray scattering. The highly v<strong>is</strong>cous PS matrix has a significant<br />

effect on the interfacial tension reduction.<br />

Finally, a summary of a d<strong>is</strong>cussion of thes<strong>is</strong> and the future directions are given in<br />

Chapter 8. In the Appendices, we present 1) the methods <strong>to</strong> determine the Flory-Huggins<br />

interaction parameter focusing on the system of polystyrene and polybutadiene, 2) the way<br />

6


<strong>to</strong> estimate the diffusion coefficient of block copolymer in the pure homopolymer phase, and<br />

3) the preliminary experiments of measuring interfacial tension in the uniaxial extensional<br />

flow with multilayer samples.<br />

1.3 Background<br />

<strong>Th<strong>is</strong></strong> subsection <strong>is</strong> devoted <strong>to</strong> the description of the phase diagram of oil/water/surfactant<br />

emulsions and a relevant theory based on the bending elasticity of an interface in a surfactant<br />

micelle.<br />

1.3.1 Phase Behaviors in Oil/Water/Nonionic Surfactant Systems<br />

The most carefully studied model oil/water/surfactant mixtures are mixtures of alkanes<br />

(C k H 2k+2 ), water, nonionic (C i E j ) alkyl polyether surfactants, which <strong>have</strong> been systematically<br />

studied by Strey and coworkers. [74, 76, 78] Here, C i and E j denote (CH 2 ) i H<br />

and [(CH 2 ) 2 O] j OH, respectively. We now briefly d<strong>is</strong>cuss the ternary phase behavior of<br />

these mixtures, focusing on the region in which the presence of a small concentration of<br />

surfactant causes the formation of a microemulsion phase. Specifically we mostly focus on<br />

the macroscopic two phase region between oil and water containing a sufficient amount of<br />

amphiphiles. A representative composition-temperature pr<strong>is</strong>m <strong>is</strong> reproduced in Fig. 1.1.<br />

The solubility pattern of surfactant (C) in water (A) or in oil (B) along the binary edge of<br />

A-C or B-C <strong>is</strong> characterized by an upper and lower critical solution temperature denoted<br />

T α and T β , respectively, within th<strong>is</strong> phase pr<strong>is</strong>m. Consequently, the relative d<strong>is</strong>tribution of<br />

nonionic surfactant in the oil-rich and water-rich phases in ternary mixtures <strong>is</strong> also sensitive<br />

<strong>to</strong> the changes in temperature. At low temperatures, close the bot<strong>to</strong>m of the pr<strong>is</strong>m, most<br />

surfactants are d<strong>is</strong>solved in the (lower) water-rich phase as swollen micelles and constitute<br />

a micellar aqueous phase coex<strong>is</strong>ting with the (upper) excess phase of nearly pure oil. <strong>Th<strong>is</strong></strong><br />

macroscopic two phase system <strong>is</strong> denoted by or Winsor I.[71] As temperature increases,<br />

2¯<br />

7


surfactants become less soluble in the water-rich phase, and finally phase separate at a<br />

lower critical end temperature T l which <strong>is</strong> an end point of a critical line emanating from<br />

T β passing through the plait points P.<br />

Figure 1.1: Schematic phase pr<strong>is</strong>m of a water-oil-nonionic surfactant (C i E j ) system in a<br />

vertical ax<strong>is</strong> of temperature. A water-rich two phase, a bicontinuous microemulsion middle<br />

phase, and the oil-rich two phase are denoted by 2¯, 3, and ¯2, respectively. The three test<br />

tubes on the right-hand side illustrate how the phases are observed in experiment. An<br />

upper critical solution temperature T β <strong>is</strong> not necessarily located at higher temperature<br />

than T u .[73]. Redrawn from Strey.[76]<br />

With further r<strong>is</strong>ing temperature, the three phase coex<strong>is</strong>tence region (3 or Winsor III)<br />

opens as the water-rich phase moves <strong>to</strong>ward the pure A corner and the surfactant-rich phase<br />

moves <strong>to</strong>ward the pure B corner along the round trajec<strong>to</strong>ry. Such a middle phase remains<br />

stable until the temperature reaches an upper critical end temperature T u originating from<br />

8


T α passing through the plait points Q. The middle phase becomes completely soluble in<br />

the oil-rich phase in which surfactants form inverted swollen micelles. Above T = T u ,<br />

there <strong>is</strong> only the micellar oil-rich phase coex<strong>is</strong>ting with the excess water-rich phase (¯2 or<br />

Winsor II). The schematic pictures of the resulting samples are shown on the right hand<br />

side of the phase pr<strong>is</strong>m at the corresponding temperature. The volume of the middle<br />

phase depends on the amount of surfactants in the mixture. Near the mean temperature<br />

T m = (T u + T l )/2 between two end critical temperatures, the three phase body has a<br />

maximum size so <strong>that</strong> surfactant can <strong>have</strong> a maximum efficiency in emulsifying oil and<br />

water.[68, 70] As long as there ex<strong>is</strong>t two d<strong>is</strong>tinct (d<strong>is</strong>connected) critical end points, which<br />

<strong>is</strong> true of C i E j surfactants with j ≥ 1, the formation of the middle phase <strong>is</strong> generally<br />

observed in any surfactant system regardless of the strength of surfactant.<br />

Although temperature has been controlled <strong>to</strong> induce a 2¯-3-¯2 phase transition as an<br />

example in a nonionic surfactant system, a similar phase sequence can be also obtained at<br />

fixed temperature by changing the structure of surfactants,[74, 78] or by adding a fourth<br />

component such as a salt,[82] cosurfactant e.g. alcohols,[83, 85] or another polar solvent[77].<br />

The 2¯-3-¯2 phase sequence and the three phase body can also be illustrated in a phase<br />

diagram along an <strong>is</strong>opleth with an equal volume of oil and water. The schematic diagram<br />

given in Fig. 1.2 <strong>is</strong> often referred <strong>to</strong> as a ‘f<strong>is</strong>h cut’ and <strong>is</strong> roughly symmetric with respect<br />

<strong>to</strong> the mean temperature T m . It cons<strong>is</strong>ts of a ‘head’, which corresponds <strong>to</strong> a three phase<br />

coex<strong>is</strong>tence region, and a ‘tail’, which corresponds <strong>to</strong> a concentrated homogeneous middle<br />

phase. By increasing the surfactant concentration beyond φ ∗ s<br />

in the ‘tail’ portion, one can<br />

lead <strong>to</strong> a uniform middle phase (µE) followed by a lamellar phase (L α ). The ‘head’ portion<br />

of a f<strong>is</strong>h <strong>is</strong> fully described by two parameters: temperature interval, T u −T l , and the amount<br />

of surfactant, φ ∗ s , needed <strong>to</strong> achieve homogeneous mixture. The latter basically determines<br />

the height of a three phase triangle body (dark region) in Fig. 1.1. The location and size<br />

of a ‘head’ portion strongly depends on the type of a surfactant. In general, a head shrinks<br />

9


Figure 1.2: A schematic phase diagram along an <strong>is</strong>opleth of the phase pr<strong>is</strong>m. T o = T m and<br />

φ ∗ s <strong>is</strong> the minimum surfactant concentration <strong>to</strong> reach a uniform middle phase. The lower<br />

<strong>is</strong> φ ∗ s, the more efficient <strong>is</strong> the surfactant. Reproduced from Wennerström et al. .[86]<br />

and moves <strong>to</strong> a lower concentration of surfactant as a surfactant becomes strong. A f<strong>is</strong>h<br />

diagram <strong>is</strong> a character<strong>is</strong>tic feature in a small molecule surfactant system <strong>that</strong> <strong>is</strong> caused by<br />

complex interaction between water and surfactant.<br />

On the other hand, Fig. 1.3 shows an <strong>is</strong>othermal Gibbs phase triangle, which <strong>is</strong> a<br />

horizontal slice of the phase pr<strong>is</strong>m at a particular temperature (T = T m ).[77] One can<br />

easily notice <strong>that</strong> the homogeneous bicontinuous middle phase, denoted by 1, in a f<strong>is</strong>h<br />

‘tail’ has a narrow stable region around the <strong>is</strong>opleth whereas the lamellar phase <strong>is</strong> stable<br />

over a much wider water-<strong>to</strong>-oil ratio.[72]<br />

1.3.2 Interfacial Tension<br />

Since the mid 1970s people <strong>have</strong> noticed <strong>that</strong> the 2¯-3-¯2 phase transition <strong>is</strong> closely related<br />

<strong>to</strong> the variation of interfacial tension between T l and T u . [68, 69, 70, 74, 78, 80] As shown<br />

in a schematic diagram in Fig. 1.4, at T < T l the water and surfactant phases become<br />

identical and form a single micellar phase so <strong>that</strong> interfacial tension γ ac between the water<br />

10


Figure 1.3: Isothermal diagram of water/n-octane/C 10 E 5 mixtures at the mean temperature<br />

T = 44.6 ◦ C . Note the 1-phase region near the middle phase enclosed by the lamellar<br />

phase. Reproduced from Strey.[77]<br />

(a) and surfactant (c) phase van<strong>is</strong>hes in a region. As temperature r<strong>is</strong>es from T < T l , γ ac<br />

2¯<br />

continuously increases from zero at T = T l and a surfactant-rich phase separates out. Then<br />

γ ac coincides with the interfacial tension γ ab between the oil (b) and water phase at T =<br />

T u , when the oil and surfactant phases become identical. Conversely, interfacial tension<br />

γ bc between the oil and surfactant phase increases from zero at T = T u as temperature<br />

decreases, and becomes equal <strong>to</strong> γ ab at T = T l . When two tension curves, γ ac and γ bc ,<br />

intersect at T = T m , γ ab becomes a minimum. Empirical observation by Shinoda,[70] and<br />

Kahlweit and Strey[74] indicate <strong>that</strong> if two critical end points are close enough, T l ≃ T u ,<br />

such a minimum can be ultra-low because γ ac and γ bc meet at a very low value. Thus the<br />

minimum of γ ab can be observed in any ternary surfactant system in which the middle<br />

phase forms near T m regardless of the strength of surfactant.<br />

Inferring from the experimental observation <strong>that</strong> the phase behaviors of nonionic surfactant<br />

systems are associated with some character<strong>is</strong>tic length scale,[79] Strey et al. showed<br />

11


Figure 1.4: Schematic diagram of the individual interfacial tension curves of oil/water(γ ab ),<br />

water/emulsion(γ ac ), and oil/emulsion(γ bc ) on a log scale as a function of temperature.<br />

Redrawn from Sottmann and Strey.[80]<br />

<strong>that</strong> interfacial tensions in various systems can collapse in<strong>to</strong> a single master curve for a<br />

wide range of temperature when they are normalized by relevant length scales of internal<br />

domains.[80] All measured tensions were successfully fit <strong>to</strong> a semi-empirical expression<br />

derived by Leitao et al. [60]<br />

γ ∗ = τ ∗2 + 1 (1.2)<br />

as shown in Fig. 1.5.<br />

1.3.3 Wetting/Nonwetting Transition of Interface<br />

When the two phase micellar system or ¯2) traverses the three phase coex<strong>is</strong>tence region<br />

(2¯<br />

(3) through the end critical temperature T l or T u , a surfactant monolayer separating the<br />

oil-rich and water-rich phases undergoes one of two different types of bicontinuous middle<br />

phase transition. First, weak surfactants such as small alcohols (C i E 0 ) and C 4 E 1 form a<br />

12


Figure 1.5: Reduced oil-water interfacial tension γ ∗ as a function of reduced temperature<br />

τ ∗ . The full line <strong>is</strong> calculated from Eqn. 1.2. The definition of γ ∗ and τ ∗ are given in ref.<br />

[80]. Redrawn from Sottmann and Strey.[80]<br />

middle phase <strong>that</strong> always wets an interface in any temperature range between T u and T l .<br />

When the water/oil interface <strong>is</strong> wet by a macroscopically thick layer of the middle phase,<br />

the interfacial tension γ ab <strong>is</strong> simply given by An<strong>to</strong>noff’ rule[84]<br />

γ ab = γ ac + γ bc (1.3)<br />

as shown with a dashed line in Fig. 1.4. In th<strong>is</strong> case, interfacial tension <strong>is</strong> usually not so<br />

low (∼ 1 mN/m).<br />

On the other hand, strong surfactants C i E j with j ≥ 2 do not wet the interface except<br />

for a very small region of temperature near the critical end point. Instead, the middle<br />

phase <strong>is</strong> often observed <strong>to</strong> form small thin bl<strong>is</strong>ters along the oil/water interfaces as shown<br />

in Fig. 1.6. The presence of bl<strong>is</strong>ters <strong>is</strong> critical for the accurate measurement of ultra-low<br />

interfacial tension because they act as the surfactant reservoirs.<br />

In th<strong>is</strong> case, the measured interfacial tension γ corresponds <strong>to</strong> the actual interfacial<br />

13


Figure 1.6: Lens of C 8 E 3 -rich middle phase floating on the interface between the lower<br />

water- and upper decane-rich phase at T = 21.8 ◦ C . Reproduced from Kahlweit et al.<br />

.[75]<br />

tension γ ab between the water- and oil-rich phases, and γ ab becomes lower than the sum<br />

of γ ac and γ bc ,[80, 81]<br />

γ ab < γ ac + γ bc (1.4)<br />

as shown with a solid line in Fig. 1.4 at a surfactant chemical potential <strong>that</strong> <strong>is</strong> character<strong>is</strong>tic<br />

of the three phase coex<strong>is</strong>tence region.<br />

1.3.4 Helfrich Bending Elasticity<br />

When a small amount of surfactant <strong>is</strong> added <strong>to</strong> a mixture of oil and water, surfactant<br />

adsorbs <strong>to</strong> the oil/water interface, resulting in decrease of interfacial tension. The reduction<br />

of interfacial tension <strong>is</strong> limited at some critical concentration of surfactant by the formation<br />

of micelles in one of the bulk phases. As first noted by Schulman,[59] surfactants in a selfassembled<br />

monolayer essentially stay in the saturated state, in which interfacial tension <strong>is</strong><br />

zero because interfacial area per molecule should be adjusted <strong>to</strong> minimize the overall free<br />

energy. When interfacial tension becomes sufficiently low, the free energy of a surfactant<br />

monolayer (membrane) <strong>is</strong> governed primarily by the bending energy associated with the<br />

14


Figure 1.7: Phase stability of spherical, cylindrical, and lamellar micelles. The crosshatched<br />

region <strong>is</strong> a two-phase coex<strong>is</strong>tence of spheres and cylinders. Sphere <strong>is</strong> the most<br />

stable structure along the emulsification failure line. In th<strong>is</strong> plot, r = ρ ˜C o and x =<br />

−¯κ/(2κ + ¯κ). Reproduced from Safran.[62]<br />

curvature of the interface.[2, 61] For a weakly curved interface with the principal radii of<br />

curvature C 1 and C 2 , Helfrich bending elasticity <strong>is</strong> given by<br />

∫<br />

H =<br />

]<br />

1<br />

dA[<br />

2 κC2 + ¯κK − κC o C<br />

(1.5)<br />

in which C = C 1 + C 2 , K = C 1 C 2 , and C o are the mean, Gaussian, and spontaneous<br />

curvatures, respectively, and in which κ and ¯κ are the mean and Gaussian rigidities. Since<br />

the free energy of a surfactant monolayer varies depending upon the monolayer curvature,<br />

Eqn. 1.5 <strong>is</strong> very useful in predicting a phase transition between competing structures<br />

formed from monolayers. Safran et al. <strong>examined</strong> the phase transition between spherical,<br />

cylindrical, and flat surfaces in ternary oil/water/surfactant mixtures by comparing the<br />

15


corresponding bending energy per area (interfacial tension) γ = H/A, given by[62, 63]<br />

γ s =<br />

2κ + ¯κ<br />

R 2<br />

− 2κC o<br />

R<br />

for a sphere with C = 2/R and K = 1/R 2 , and<br />

γ c = κ<br />

2R 2 − κC o<br />

R<br />

for a cylinder with C = 1/R and K = 0, and<br />

γ l = 0<br />

for a lamellar phase with C = 0 and K = 0. From Eqn. 1.5, they found <strong>that</strong> the monolayer<br />

structure <strong>that</strong> minimizes the Helfrich bending energy along the emulsification failure line<br />

<strong>is</strong> a sphere with the curvature ˜C o = κC o /(κ + ¯κ/2). When a micellar phase coex<strong>is</strong>ts with<br />

an excess phase, the volume of material emulsified in the core of each micelle can be freely<br />

adjusted by transfer of material <strong>to</strong> or from the excess phase. In th<strong>is</strong> case, the optimal mean<br />

curvature can be obtained with either a spherical or cylindrical surface, but the Gaussian<br />

rigidity favors a spherical surface over a cylindrical surface of equal mean curvature when<br />

¯κ < 0. If the system enters a 1-phase micellar region from the emulsification failure line,<br />

however, the volume(V )-<strong>to</strong>-surface(A) ratio ρ of a micelle <strong>is</strong> determined by a s<strong>to</strong>ichiometry<br />

such <strong>that</strong> 3(V/A) s = R for a sphere and 2(V/A) c = 2R/3 for a cylinder where R <strong>is</strong> a radius<br />

of a sphere. By setting ρ = 3V/A, Fig. 1.7 shows <strong>that</strong> a dimensionless ratio r = ˜C o ρ can<br />

drive a phase transition between different structures in the absence of thermal fluctuation.<br />

If r > 1, excess bulk phase <strong>is</strong> simply rejected (emulsification failure) so as <strong>to</strong> maintain<br />

the saturation state in a micelle. As one decreases r from the unity by increasing the<br />

amount of surfactant or by decreasing <strong>to</strong>tal volume of the internal bulk phase, spherical<br />

micelles may be transformed in<strong>to</strong> either cylinders or lamellar phase, depending on the ratio<br />

x = −¯κ/(2κ + ¯κ). In th<strong>is</strong> thes<strong>is</strong>, we are primarily interested in the effect of copolymer<br />

16


upon macroscopic interfacial tension along an interface between a micellar phase and an<br />

excess homopolymer phase, and consider only spherical micelles.<br />

In fact, during the 2¯-3-¯2 transition, the spontaneous curvature C o of a surfactant monolayer<br />

dividing the oil-rich and water-rich phases continuously varies with temperature and<br />

changes sign at the mean temperature.[78] A surfactant monolayer with high spontaneous<br />

curvature forms a spherical (inverted) swollen micelle in the two phase (2¯, ¯2) region. If<br />

the curvature becomes sufficiently small, however, a phase transition in<strong>to</strong> a bicontinuous<br />

microemulsion phase <strong>is</strong> induced with the diverging length scale ζ ∼ 1/C o between T l and<br />

T u .<br />

1.3.5 Stability of Surfactant Monolayer<br />

Eqn. 1.5 can be rewritten in a more useful functional form <strong>to</strong> account for the stability of<br />

a membrane surface as<br />

H = 1 2<br />

∫<br />

dA<br />

[κ +<br />

(<br />

C + − κC o<br />

κ +<br />

) 2<br />

+ κ − C − 2 − κ +<br />

( κCo<br />

κ +<br />

) 2<br />

]<br />

(1.6)<br />

in which C ± ≡ C 1 ± C 2 , κ + ≡ κ + (1/2)¯κ, and κ − = −(1/2)¯κ. For a special case of a flat<br />

surface with C o = 0, Eqn. 1.6 <strong>is</strong> reduced <strong>to</strong><br />

H = 1 ∫<br />

dA [ κ + C 2 2 + + κ − C ] − . (1.7)<br />

2<br />

The above equations indicate <strong>that</strong> both κ + and κ − must be positive for a weakly curved<br />

surface <strong>to</strong> be stable. Otherw<strong>is</strong>e, a monolayer will break in<strong>to</strong> infinitesimally small spheres<br />

(C 1 = C 2 ) if κ + < 0, or will transform in<strong>to</strong> a multiply connected structure of an infinite<br />

minimal surface (C 1 = −C 2 ) via concentrated ’holes’ if κ − < 0.<br />

The density of ’holes’ <strong>is</strong> also related <strong>to</strong> the integral ∫ dAK of the Gaussian curvature<br />

over closed surface, which <strong>is</strong> a <strong>to</strong>pological invariant given by the Gauss-Bonnet theorem<br />

∫<br />

dAK = 4π(1 − g), where g <strong>is</strong> the number of handles on a surface.[64] For example, all<br />

shapes in Fig. 1.8 <strong>have</strong> the same <strong>to</strong>pology of g = 0 except a <strong>to</strong>rus (C) of g = 1. Therefore,<br />

17


if ¯κ − < 0 or ¯κ > 0, the Gaussian contribution <strong>to</strong> the free energy difference between a<br />

monolayer and a minimal surface increases linearly with g.<br />

Figure 1.8: The compar<strong>is</strong>on of the <strong>to</strong>pology among different structures. Only (C) has<br />

g = 1 where as others <strong>have</strong> g = 0, where g <strong>is</strong> the number of handles. Reproduced from<br />

Taddei.[65]<br />

Eqn. 1.6 also yields an important result for the interfacial tension γ flat of a macroscopic<br />

flat interface coex<strong>is</strong>ting with a phase containing swollen spherical micelle of curvature<br />

˜C o = (κC o )/κ + . Interfacial tension <strong>is</strong> the free energy per unit area for creating more<br />

macroscopic interface. Assuming the area per molecule <strong>is</strong> similar in the flat macroscopic<br />

interface and in the curved interface of a micelle, the creation of more macroscopic interface<br />

thus requires the destruction of an equal amount of micellar interfacial area. Therefore,<br />

interfacial tension can be the free energy penalty <strong>to</strong> transfer surfactant molecules from<br />

micelles <strong>to</strong> a flat interface, and obtained from Eqn. 1.6 [78]<br />

γ flat = 1 2 κ + ˜C o<br />

2<br />

. (1.8)<br />

Eqn. 1.8 stresses <strong>that</strong> γ flat has a quadratic dependence on the micelle curvature ˜C o and<br />

has a van<strong>is</strong>hing minimum at a balance point, where the spontaneous curvature van<strong>is</strong>hes.<br />

The balance point corresponds <strong>to</strong> the mean temperature in nonionic surfactant system.[80]<br />

The Helfrich theory has been extended <strong>to</strong> ternary polymer blends in which block<br />

copolymer coex<strong>is</strong>ts with two imm<strong>is</strong>cible solvents[89] or homopolymers.[39, 43, 90, 91, 92]<br />

Can<strong>to</strong>r[89] and Leibler[39] studied the elastic properties including C o and K + of block<br />

18


copolymer adsorbed at the spherically curved interface by assuming the constant density<br />

profile of copolymer brushes. Can<strong>to</strong>r was concerned with the scaling behaviors of K + and<br />

Leibler was particularly interested in the stability of the droplet with nonzero interfacial<br />

tension and prediction of the emulsification failure. Wang and Safran[90, 91] improved the<br />

results of Can<strong>to</strong>r and Leibler in homopolymer blends by considering the chain end d<strong>is</strong>tribution<br />

of copolymer at saturated state in the strong segregation limit. They analytically<br />

derived κ, ¯κ, and C o as a function of the block ratio in both states of the melt and swollen<br />

brush, and predicted the phase transition of copolymer in a plot of the surface-<strong>to</strong>-volume<br />

ratio vs. the block ratio, which <strong>is</strong> similar <strong>to</strong> Fig. 1.7.[90] Matsen used numerical SCFT <strong>to</strong><br />

calculate the elasticity of symmetric (mono or polyd<strong>is</strong>perse) block copolymer in homopolymer<br />

blends with equal lengths and concluded <strong>that</strong> block copolymer should be reasonably<br />

short so as <strong>to</strong> keep −¯κ ∼ kT, which <strong>is</strong> favorable <strong>to</strong> form a bicontinuous microemulsion<br />

phase.[43, 92]<br />

1.3.6 Thermal Fluctuation<br />

Beyond the mean field approximation of a rigid surface, deGennes and Taupin[93] considered<br />

the entropy effect on a flexible surface by introducing a concept of a ’pers<strong>is</strong>tence<br />

length’ ξ κ , which <strong>is</strong> given by<br />

( ) 4πκ<br />

ξ κ = a exp<br />

αkT<br />

(1.9)<br />

where a <strong>is</strong> a cut-off length of a molecular size, and α = 3.[64] Rearranging Eqn. 1.9 gives<br />

the renormalized bending rigidity κ(ξ κ ) as<br />

κ(ξ κ ) = κ o − αkT ( )<br />

4π ln ξκ<br />

a<br />

(1.10)<br />

, which implies <strong>that</strong> the rigidity of a monolayer van<strong>is</strong>hes beyond the pers<strong>is</strong>tence length.<br />

They argued <strong>that</strong> the membrane <strong>is</strong> effectively flat at length scales smaller than ξ κ , but<br />

becomes strongly wrinkled above ξ κ . Therefore, a lamellar phase should <strong>have</strong> a mean<br />

19


igidity of κ ∼ kT in order <strong>to</strong> melt in<strong>to</strong> a randomly fluctuating surface because ξ κ exponentially<br />

depends on κ. Also, they regarded a membrane surface as consecutive uncorrelated<br />

’patches’ with an area ξ 2 κ . In th<strong>is</strong> view, in order <strong>to</strong> obtain a one-phase mixture of<br />

oil/water/surfactant with a domain spacing ξ κ , the free energy cost γ ξ 2 κ <strong>to</strong> couple adjacent<br />

random patches should be comparable <strong>to</strong> kT or less, <strong>that</strong> <strong>is</strong>,<br />

γ ξ κ 2 ∼ kT (1.11)<br />

, which has been verified in many experiments. [78, 80] If γ ξ 2 κ kT, the system phase<br />

separates in<strong>to</strong> macroscopic domains.<br />

However, careful neutron scattering experiments by Porte et al. showed a lamellar phase<br />

could melt in<strong>to</strong> the sponge phase (L 3 ) in the AOT/brine system at a much smaller length<br />

scale than ξ κ .[67] It convinced researchers <strong>that</strong> the Gaussian rigidity ¯κ primarily governs the<br />

<strong>to</strong>pological transition of a lamellar phase <strong>to</strong> a L 3 phase. Partly for th<strong>is</strong> reason, Morse[94, 95]<br />

and Golubovic[96] argued <strong>that</strong> it was the renormalized Gaussian rigidity <strong>that</strong> controlled the<br />

competition between a lamellar and bicontinuous structure. The renormalized Gaussian<br />

rigidity <strong>is</strong> given by<br />

( )<br />

¯κ(ξ¯κ ) = ¯κ o + α′ kT<br />

4π ln ξ¯κ<br />

a<br />

(1.12)<br />

where α ′ = 10/3.[64] By comparing both length scales of ξ κ and ξ¯κ , they both observed<br />

<strong>that</strong> a phase transition of a lamellar phase in<strong>to</strong> a L 3 phase occurs when ¯κ+ 10 9<br />

κ > 0, which<br />

<strong>is</strong> the criterion <strong>that</strong> Matsen used for the formation of a bicontinuous microemulsion in<br />

ternary polymer systems.[92] The subject of the fluctuation effect on a membrane surface<br />

<strong>is</strong> actually beyond the scope of th<strong>is</strong> thes<strong>is</strong>. A more detailed review <strong>is</strong> given in ref [97].<br />

20


Bibliography<br />

[1] P-167 Polymer Alloys and Blends in Industry Research and Technical Market Analys<strong>is</strong>,<br />

Business Communications Co., Inc. 2005<br />

[2] Boning, C.; Van Duin., M.; Pagnoulle. C; Jerome. R. Prog. Ploy. Sci. 1998, 23,<br />

707<br />

[3] Ruzette. A.-V.; Leibler, L. Nature Materials 2005, 4, 19<br />

[4] Grace. H.-P. Chem. Eng. Commun. 1982, 14, 225<br />

[5] Taylor, G. I. Proc. Roy. Soc. (London), Ser. A 1932, 138, 41<br />

[6] Taylor, G. I. Proc. Roy. Soc. (London), Ser. A 1934, 146, 501<br />

[7] Ajji, A.; Utracki, L. A.; Polym. Eng. Sci. 1996, 36, 1574<br />

[8] Sundararaj, S. H.; Macosko, C. W. Macromolecules 1995, 22, 1449<br />

[9] Liang, H.; Fav<strong>is</strong>, B. D.; Yu, Y. S.; E<strong>is</strong>enberg, A. Macromolecules 1999, 32 1637<br />

[10] Macosko, C. W.; Guegan, P.; Khandpur, A. K.; Nakayama, A.; Marechal, P.; Inoue,<br />

T. Macromolecules 1996, 29, 5590;<br />

[11] Lyu, S.; Jones, T. D.; Bates, F. S.; Macosko, C. W. Macromolecules 2002, 35, 7845<br />

[12] Maric, M.; Macosko, C. W. Polym. Eng. Sci. 2001, 41, 118<br />

[13] Galloway, J. A.; Jeon, H. K.; Bell, J. R.; Macosko, C. W. Polymer 2005, 46, 183<br />

21


[14] Kim, J. R.; Jamieson, A. M.; Hudson. S. D.; Manas-Zloczower, I.; Ishida, H. Macromolecules<br />

1999, 32, 4582<br />

[15] Chun, S. B.; Han, C. D. Macromolecules 2000, 33, 3409<br />

[16] Paul, D. R. Polymer Blends; Paul, D. R., Newman, S., Eds.; Academic Press: New<br />

York, 1978; Vol. 2,<br />

[17] Milner, S. T.; Xi, H. W. J. Rheol 1996, 40, 663<br />

[18] Beck Tan, N. C.; Tai, S.-K.; Briber, R. M. Polymer 1996, 37, 3509<br />

[19] Lyu, S. Macromolecules 2003, 36, 10052<br />

[20] Anastasiad<strong>is</strong>, S. H.; Gancarz, I.; Koberstein, J. T. Macromolecules 1989, 22, 1449<br />

[21] Shull, K. R.; Kramer, E. J.; Hadziioannou, G.; Tang, W. Macromolecules 1990,<br />

23, 4780<br />

[22] Noolandi, J.; Hong, K. M. Macromolecules 1982, 15, 482<br />

[23] Noolandi, J.; Hong, K. M. Macromolecules 1984, 17, 1531<br />

[24] Shull, K. R.; Kramer. E. J. Macromolecules 1990, 23, 4769<br />

[25] Ma<strong>to</strong>s, M.; Fav<strong>is</strong>, B. D.; Lomellini, P. Polymer 1995, 36, 3899<br />

[26] Utracki, L. A. Polymer Alloys and Blends. Thermodynamics and Rheology; Hanser<br />

Publ<strong>is</strong>hers: Munich, Vienna, New York, 1989<br />

[27] Cre<strong>to</strong>n, C.; Kramer, E. J.; Hadziioannou, G. Macromolecules, 1991, 24, 1846<br />

[28] Lyatskaya, Y.; Gersappe, D.; Balazs, A. C. Macromolecules 1995, 6278<br />

[29] Lyatskaya, Y.; Gersappe, D.; Gross, N. A.; Balazs, A. C. J. Phys. Chem 1996,<br />

100, 1449<br />

22


[30] Lyatskaya, Y.; Balazs, A. C. Macromolecules 1996, 29, 7581<br />

[31] Verdier, C; Vinagre, H. T. M.; Piau, M.; Joseph, D. D. Polymer 2000, 41, 6683<br />

[32] Patterson, H. T.; Hu, K. H.; Grindstaff, T. H. J. Polym. Sci. Part C 1971, 34, 31<br />

[33] Retsos, H.; Margiolaki, I.; Messaritaki, A.; Anastasiad<strong>is</strong>, S. H. Macromolecules<br />

2001, 34, 5295<br />

[34] Wagner, M.; Wolf, B. A. Polymer 1993, 34, 1461; Jorzik, U.; Wolf, B. A. Macromolecules<br />

1997, 30, 4713<br />

[35] Stammer, A.; Wolf, B. A. Macromol. Rapid. Commun. 1998, 19, 123<br />

[36] Retsos, H.; Anastasiad<strong>is</strong>, S. H.; P<strong>is</strong>pas, S.; Mays, J. W.; Hadjichr<strong>is</strong>tid<strong>is</strong>, N.<br />

Macromolecules 2004, 37, 524<br />

[37] Elemans, P., H., M.; Janssen, J. M. H.; Meijer, H. E. H.; J. Rheo 1990, 34, 1311<br />

[38] Fle<strong>is</strong>cher, C., A.; Morales, A. R.; Koberstein, J. T.; Macromolecules 1994, 27, 379<br />

[39] Leibler, L. M akromol. Chem., Macromol. Symp. 1988, 16, 1<br />

[40] Shull, K. R. Macromolecules 2002, 35, 8631<br />

[41] Dadmun, M. Macromolecules 1996, 29, 3868<br />

[42] Vilg<strong>is</strong>, T. A.; Noolandi, J. Macromolecules 1990, 23, 2941<br />

[43] Thompson, R. B.; Matsen, M. W. Phys. Lett. Lett. 2000, 85, 670<br />

[44] Leibler, L; Orland, H.; Wheeler, J. C. J. Chem. Phys. 1983, 79, 3550<br />

[45] Whitmore, M. D.; Noolandi, J. Macromolecules 1985, 18, 657<br />

[46] Noolandi, J.; Hong, K. M. Macromolecules 1983, 16, 1443<br />

23


[47] Roe, R. -J. Macromolecules 1986, 19, 728<br />

[48] Mathur. D.; Hariharan, R.; Nauman, E. B. Polymer 1999, 40, 6077<br />

[49] van Lent, B.; Scheutjens, J. M. H. M. Macromolecules 1989, 22, 1931<br />

[50] Linse, P. Macromolecules 1993, 26, 4437<br />

[51] Leermakers, F. A. M.; Wijmans, C. M.; Fleer, G. J. Macromolecules 1995, 28, 3434<br />

[52] Shull, K. R. Macromolecules 1993, 26, 2346<br />

[53] Israels, R.; Jasnow, D.; Balazs, A. C.; Guo, L.; Krausch, G.; Sokolov, J.;<br />

Rafailovich, M. J. Chem. Phys. 1995, 102, 8149<br />

[54] Shi, T.; Ziegler, V. E.; Welge, I. C.; An, L.; Wolf, B. A. Macromolecules 2004,<br />

37, 1591<br />

[55] Welge, I.; Wolf, B. A. Polymer 2001, 42, 3467<br />

[56] Hu, W.; Koberstein, J. T.; Lingelser, J. P.; Gallot, Y. Macromolecules 1995, 28,<br />

5209<br />

[57] Cho, D.; Hu, W.; Koberstein, J. T.; Lingelser, J. P.; Gallot, Y. Macromolecules<br />

2000, 33, 5245<br />

[58] Cigana, P.; Fav<strong>is</strong>, B. D.; Jerome, R. J. Polym. Sci.: Polym. Phys. 1996, 34, 1691<br />

[59] Schulman, J. H.; Montagne, J. B. Ann. N.Y. Acad. Sci. 1961, 92, 366<br />

[60] Leitao, H.; Somoza, A. M.; Telo da Gama, M. M.; Sottmann, T.; Strey, R. J. Chem.<br />

Phys. 1996, 105, 2875<br />

[61] Szleifer, I.; Kramer, D.; Ben-Shaul, A.; Gelbart, W. M.; Safran, S. A. J. Chem.<br />

Phys. 1990, 92, 6800<br />

24


[62] Safran, S. A. Stat<strong>is</strong>tical Thermodynamics of Surfaces, Interfaces, and Membranes;<br />

Westview Press: pp243-259, 2003<br />

[63] Safran, S. A.; Turkevich, L. A.; Pincus, P. J. Phys. (Par<strong>is</strong>) 1984 45, L-69<br />

[64] David, F. Stat<strong>is</strong>tical Mechanics of Membranes and Surfaces, edited by Nelson, D.,<br />

Piran, T., and Weinberg, S.; World Scientific: 1989<br />

[65] Taddei, G. Coll. Ploym. Sci. 1994, 272, 1300<br />

[66] Porte, G.; Appell, J.; Bassereau,P.; Marignan, J. J. Phys. France, 1989, 50, 1335<br />

[67] Porte, G. J. Phys. II. France 1991, 1, 1121<br />

[68] Fletcher, P. D.; Horsup, D. I. J. Chem. Soc. Faraday. Trans. 1992, 88, 855<br />

[69] Pouchelon, A.; Meunier, J.; Langevin, D.; Chatenay, D.; Cazabat, A. M. Chem.<br />

Phys. Lett. 1980, 76, 277<br />

[70] Shinoda, K. Progr. Collo. Polym. Sci. 1983, 68, 1<br />

[71] Winsor, P. A. Trans. Faraday. Sco. 1948, 44, 376<br />

[72] Kunieda, H.; Shinoda, K. J. D<strong>is</strong>per. Sci. Tech. 1982, 3, 233<br />

[73] Kahlweit, M.; Lessner, E.; Strey, R. J. Phys. Chem. 1983, 87, 5032<br />

[74] Kahlweit, M.; Strey, R. Angew. Chem. Int. Ed. Engl. 1985, 24, 654<br />

[75] Kahlweit, M.; Strey, R.; Haase, D.; Firman, P. Langmuir 1988, 4, 785<br />

[76] Kahlweit, M.; Strey, R.; Busse, G. Phys. Rev. E 1993, 47, 4197<br />

[77] Strey, R. Ber. Bunsenges. Phys. Chem. 1993, 97, 742<br />

[78] Strey, R. Coll. Ploym. Sci. 1994, 272, 1005<br />

25


[79] Sottmann, T.; Strey, R. J. Phys.: Condens. Matter 1996, 8, A39<br />

[80] Sottmann, T.; Strey, R. J. Chem. Phys. 1997, 106, 8606<br />

[81] Widom, B. Lamgmuir 1987, 3, 12<br />

[82] Knickerbocker, B. M.; Pesheck, C. V.; Scriven, L. E.; Dav<strong>is</strong>, H. T. J. Phys. Chem.<br />

1979, 83, 1984<br />

[83] Bennett, K. E.; Dav<strong>is</strong>, H. T.; Scriven, L. E. J. Phys. Chem. 1982, 86, 3917<br />

[84] See<strong>to</strong>, Y.; Puig, J. E.; Scriven, L. E.; Dav<strong>is</strong>, H. T. J. Coll. Int. Sci. 1983, 96, 1983<br />

[85] Puig, J. E.; Hemker, D. L.; Gupta, A.; Dav<strong>is</strong>, H. T.; Scriven, L. E. J. Phys. Chem.<br />

1987, 91, 1137<br />

[86] Wennerstrom, H.; Söderman, O.; Olsson, U.; Lindman, B. Colloids Surf. 1997,<br />

123-124, 13<br />

[87] Schubert, K.-V.; Strey, R. J. Chem. Phys. 1991, 95, 8532<br />

[88] Schubert, K.-V.; Strey, R.; Kline, S. R.; Kaler, E. W. J. Chem. Phys. 1994, 101,<br />

5343<br />

[89] Can<strong>to</strong>r, R. Macromolecules 1981, 14, 1186<br />

[90] Wang, Z.-G.; Safran, S. A. J. Phys. France. 1990, 51, 185<br />

[91] Wang, Z.-G.; Safran, S. A. J. Chem. Phys. 1991, 94, 679<br />

[92] Matsen, M. W. J. Chem. Phys. 1999, 110, 4659<br />

[93] De Gennes, P. G.; Taupin, C. J. Phys. Chem. 1982 86, 2294<br />

[94] Morse, D. C. Phys. Rev. E 1994, 50, R2423<br />

26


[95] Palmer, K. M.; Morse, D. C. J. Chem. Phys. 1996, 105, 11147<br />

[96] Golubović, L. Phys. Rev. E 1994, 50, R2419<br />

[97] Morse, D., C. Curr. Opin. Colloid Interface Sci. 1997, 2, 365<br />

27


Chapter 2<br />

Self Cons<strong>is</strong>tent Field Theory<br />

2.1 Introduction<br />

Self Cons<strong>is</strong>tent Field Theory (SCFT) <strong>is</strong> a mean field theory of spatially inhomogeneous<br />

polymer melts. In SCFT, the stat<strong>is</strong>tical mechanics of interacting many polymer chains <strong>is</strong><br />

reduced <strong>to</strong> a single chain problem. Instead, the interaction between nearest neighboring<br />

chains <strong>is</strong> simplified as an unknown chemical potential field <strong>that</strong> bias the random walk of<br />

a Gaussian chain. The chemical potential field for each type of monomer at each point in<br />

space depends upon the local composition and the Flory-Huggins interaction parameter χ,<br />

and must be determined self-cons<strong>is</strong>tently.<br />

In th<strong>is</strong> chapter we review SCFT for the multicomponent polymer blends and d<strong>is</strong>cuss<br />

conventions and numerical methods used in Chapter 3 and 4. For generality, we consider a<br />

mixture of polymer species labeled by indices i and j <strong>that</strong> are constructed from monomers<br />

<strong>that</strong> are labeled by indices α or β. Polymers of type i contain the <strong>to</strong>tal number of N i<br />

monomers, whose position within the chain <strong>is</strong> parameterized by a variable 0 < s i < N i .<br />

2.2 Formulation of SCFT<br />

2.2.1 Field Equations<br />

SCFT requires the calculation of a stat<strong>is</strong>tical weight q i (r,s) for each species <strong>that</strong> <strong>is</strong> proportional<br />

<strong>to</strong> the constrained partition function of the subchain containing monomers 0 <strong>to</strong><br />

28


s of species i, when monomer s <strong>is</strong> constrained <strong>to</strong> point r, in a self-cons<strong>is</strong>tently determined<br />

chemical potential field. <strong>Th<strong>is</strong></strong> function sat<strong>is</strong>fies a modified diffusion equation<br />

[<br />

]<br />

∂q i (r,s)<br />

= − − b2 α(s)<br />

∂s 6 ∇2 + ω α(s) (r) q i (r,s) (2.1)<br />

with an initial condition q i (r,0) = 1 for all r. Here, b α <strong>is</strong> a stat<strong>is</strong>tical segment length and<br />

ω α (r) <strong>is</strong> a chemical potential field for α monomers, and α(s) <strong>is</strong> the type index for monomer<br />

s, which has different values within different blocks. An analogous function q † i<br />

(r,s) for<br />

the subchain containing monomers s <strong>to</strong> N i sat<strong>is</strong>fies a corresponding equation with the<br />

l.h.s. multiplied by −1, with a final condition q † i<br />

(r,N) = 1. In case of an imm<strong>is</strong>cible<br />

homopolymer blend we assume no flux at a unit cell boundary such as ∂q i (r,s)/∂r = 0.<br />

The chemical potential field ω α (r) <strong>is</strong> given<br />

ω α (r) = ∑ β<br />

χ αβ φ β (r) + ξ(r) , (2.2)<br />

where χ αβ <strong>is</strong> a binary Flory-Huggins interaction parameter, for which χ αβ = χ βα and<br />

χ αα = χ ββ = 0, φ α (r) <strong>is</strong> a local volume fraction of α monomers, and ξ(r) <strong>is</strong> a Lagrange<br />

multiplier field, which <strong>is</strong> chosen so as <strong>to</strong> sat<strong>is</strong>fy the incompressibility constraint,<br />

∑<br />

φ α (r) = 1 . (2.3)<br />

The local volume fraction φ β (r) <strong>is</strong> given by<br />

α<br />

φ β (r) = ∑ i<br />

¯φ i<br />

Q i<br />

∫ Ni<br />

0<br />

ds<br />

N i<br />

q i (r,s)q † i (r,s)δ β,α(s) , (2.4)<br />

where ¯φ i = M i N i v/V <strong>is</strong> the overall volume fraction of i molecules, M i <strong>is</strong> the <strong>to</strong>tal number<br />

of i molecules, v <strong>is</strong> a monomer reference volume, and<br />

Q i ≡ 1 ∫<br />

dr q i (r,N i ) . (2.5)<br />

V<br />

The s integral in Eqn. 2.4 <strong>is</strong> constrained <strong>to</strong> blocks of monomer type β by the the Kronecker<br />

δ function δ β,α(s) .<br />

29


SCFT may be implemented either in the canonical ensemble, in which one specifies a<br />

value of M i or ¯φ i for each species [1], or in grand-canonical ensemble, in which one specifies<br />

a chemical potential µ i for each species [2, 3]. The chemical potential and molecular volume<br />

fraction of a species i are related by the identity<br />

¯φ i = Q i e µ i/kT<br />

. (2.6)<br />

Eqn. 2.6 implicitly introduces a convention for the chemical potential in which µ i van<strong>is</strong>hes<br />

in a hypothetical standard state cons<strong>is</strong>ting of pure species i, with ¯φ i = 1, in a van<strong>is</strong>hing potential<br />

field ω α (r) = 0 for all α, for which q i (r,N i ) = 1 and Q i = 1. In the grand-canonical<br />

implementation of SCFT, local monomer densities are thus calculated by replacing the<br />

fac<strong>to</strong>r of ¯φ i /Q i in Eqn. 2.4 by the activity e µ i/kT . The <strong>to</strong>tal grand canonical free energy <strong>is</strong><br />

given by<br />

Φ<br />

kT<br />

= −V ∑ i<br />

+ 1 ∑<br />

2<br />

αβ<br />

¯φ i<br />

N i v − ∑ α<br />

∫ dr<br />

v χ αβ φ α (r)φ β (r) ,<br />

∫ dr<br />

v ω α(r)φ α (r) (2.7)<br />

where ¯φ i <strong>is</strong> replaced by Eqn. 2.6.<br />

In an incompressible liquid, SCFT predictions for the volume fraction fields φ i,α (r) and<br />

Helmholtz free energy F = Φ+ ∑ i µ iM i are invariant under a spatially homogeneous shift<br />

ξ(r) → ξ(r) + v δP (2.8)<br />

of the Lagrange multiplier field, which corresponds <strong>to</strong> a shift δP in hydrostatic pressure.<br />

Such a shift causes corresponding trivial shifts<br />

ω α (r) →<br />

ω α (r) + v δP<br />

µ i → µ i + N i v δP (2.9)<br />

in chemical potentials. The solution of the SCFT equations for a two-phase system <strong>is</strong> thus<br />

30


unique only <strong>to</strong> within such homogeneous shifts in ξ(r), unless a value <strong>is</strong> specified for either<br />

ξ or for the corresponding macroscopic pressure P in one of the two bulk phases.<br />

2.2.2 Homogeneous Mixtures<br />

Calculation of an interfacial excess free energy γ requires knowledge of the pressures in<br />

the surrounding homogeneous phases as functions of chemical potential. In homogeneous<br />

mixtures, SCFT reduces <strong>to</strong> a form of Flory-Huggins theory. The chemical potential in a<br />

homogeneous mixture, in the convention used here, <strong>is</strong> given by<br />

∑<br />

µ i = kT ln(¯φ i ) + N i f i,α ω α , (2.10)<br />

where f i,α <strong>is</strong> the fraction of monomers of type α on a chain of species i, and ¯φ i and ω α are<br />

homogeneous values of the fields. The macroscopic pressure P = −Φ/V for a homogeneous<br />

phase may be obtained from Eqn. 2.7, which yields<br />

Pv<br />

kT = ξ + ∑ i<br />

¯φ i<br />

N i<br />

+ 1 2<br />

α<br />

∑<br />

χ αβ ¯φα ¯φβ . (2.11)<br />

By using Eqn. 2.11 <strong>to</strong> express µ i as a function of P, rather than ξ, noting <strong>that</strong> M i =<br />

¯φ i V/N i v and ¯φ α = ∑ i ¯φ i f i,α , and evaluating the sum G = ∑ i M iµ i , we may confirm <strong>that</strong><br />

the corresponding Gibbs free energy per monomer <strong>is</strong> given by an expression<br />

Gv<br />

V kT = ∑ i<br />

which <strong>is</strong> cons<strong>is</strong>tent with Flory-Huggins theory.<br />

αβ<br />

( )<br />

¯φ i ¯φi<br />

ln + 1 ∑<br />

χ αβ ¯φα ¯φβ + Pv (2.12)<br />

N i e 2<br />

Using the Flory-Huggins theory, one can derive the equilibrium d<strong>is</strong>tribution of block<br />

copolymer molecularly d<strong>is</strong>solved in two coex<strong>is</strong>ting A- and B-rich homopolymer phases.<br />

αβ<br />

Setting the equal chemical potential of copolymer µ A C = µB C<br />

(Eqn. 2.10) in both phases at<br />

fixed bulk pressure P (Eqn. 2.11) gives the partition coefficient K ≡ φ A c /φ B c<br />

( 1<br />

ln K ≃ χN c (f A − f B ) + N C − 1 )<br />

N A N B<br />

(2.13)<br />

in the infinitely dilute copolymer solution, in which φ A A ≃ φB B ≃ 1.<br />

31


2.2.3 Interfacial Tension<br />

Interfacial tension γ <strong>is</strong> defined as the excess grand canonical free energy per unit area. For<br />

a system of volume V with a curved Gibbs dividing interface with a radius of curvature R,<br />

which separates an A-rich domain (phase I) of volume V I for r > R from a B-rich domain<br />

(phase II) of volume V II for r < R, interfacial tension <strong>is</strong> given by γA = Φ+P I V I +P II V II ,<br />

which yields<br />

γ<br />

kT<br />

= ∫ dr<br />

Av<br />

+<br />

∫ dr<br />

Av<br />

[<br />

e µ i/kT<br />

− ∑ q i (r,N i ) − ∑ ω α (r)φ α (r)<br />

N<br />

i i<br />

α<br />

⎡<br />

⎤<br />

(2.14)<br />

⎣ 1 ∑<br />

χ αβ φ α (r)φ β (r) + v<br />

2<br />

kT {P A Θ(r − R) + P B Θ(R − r)} ⎦ ,<br />

αβ<br />

where A <strong>is</strong> an area, and Θ(x) = 1 for x > 0 and Θ(x) = 0 for x < 0. For a cylindrically<br />

or spherically curved interface, the one dimensional integral over system volume per unit<br />

]<br />

interfacial area <strong>is</strong> given by<br />

∫ dr<br />

A = ∫<br />

(<br />

dz 1 +<br />

R) z d−1<br />

(2.15)<br />

where r = R + z, and the dimension d of the system <strong>is</strong> d = 1 for a flat surface, d = 2 for a<br />

cylinder, and d = 3 for a sphere, respectively.<br />

2.2.4 Nondimensionalization<br />

The modified diffusion equation can be non-dimensionalized by introducing reduced variables<br />

t ≡ s/N and ˜r = r/R o , with R 2 o = Nb 2 , where N <strong>is</strong> a reference degree of polymerization,<br />

for which we take N = N C , and b <strong>is</strong> a reference stat<strong>is</strong>tical segment length, for which<br />

we take b = b A in th<strong>is</strong> work. <strong>Th<strong>is</strong></strong> yields the non-dimensionalized diffusion equation<br />

with ˜∇ 2 ≡ R 2 o∇ 2 and<br />

∂q i (˜r,t)<br />

∂t<br />

[<br />

= − − 1 ]<br />

6 (b α<br />

b )2 ˜∇2 + ˜ω α (˜r) q i (˜r,t) , (2.16)<br />

˜ω α (˜r) ≡ Nω α (˜r) = ∑ β<br />

˜χ αβ φ β (˜r) + ˜ξ(˜r) (2.17)<br />

32


where ˜χ αβ ≡ Nχ αβ and ˜ξ(˜r) ≡ Nξ(˜r). By non-dimensionalizing the self-cons<strong>is</strong>tency<br />

conditions in terms of the same variables, it may be shown <strong>that</strong> the self-cons<strong>is</strong>tent solution<br />

for q i near a curved interface <strong>is</strong> a function of t, ˜r, the chemical potentials µ A , µ B and µ C ,<br />

and the dimensionless parameters α A , α B , f A , b B /b A , ˜χ αβ , and C √ Nb alone.<br />

The grand-canonical free energy per interfacial area may be expressed in non-dimensional<br />

form as an integral<br />

∫<br />

Φ<br />

AkT = b<br />

N 1/2 v<br />

∫<br />

b<br />

+<br />

N 1/2 v<br />

{<br />

dẑ − ∑ i<br />

e µ i/kT<br />

α i<br />

q i (˜r,α i ) − ∑ }<br />

˜ω α (˜r)φ α (˜r)<br />

α<br />

dẑ ∑ ˜χ αβ φ α (˜r)φ β (˜r) , (2.18)<br />

αβ<br />

where the integrand <strong>is</strong> a function of dimensionless d<strong>is</strong>tance ẑ from the interface given by<br />

dẑ =<br />

(<br />

1 + z ) d−1 dz<br />

R N 1/2 b<br />

. (2.19)<br />

A and R <strong>is</strong> the area and radius of the Gibbs dividing surface, and α i = N i /N, so <strong>that</strong><br />

α C = 1 with our convention N = N C . To obtain the excess free energy density γ, we must<br />

subtract from bulk pressure in the neighboring phases from the integrand, subtracting<br />

the bulk pressure of a given phase throughout the domain on the corresponding side of<br />

the Gibbs dividing surface, and thus cancel contributions <strong>to</strong> Φ from points far from the<br />

interface. <strong>Th<strong>is</strong></strong> yields an interfacial tension of the form<br />

γ<br />

kT =<br />

b<br />

N 1/2 v [γ](α A,α B , b A<br />

b B<br />

,f A ,χN,[C],µ C ) (2.20)<br />

in which [γ] <strong>is</strong> a dimensionless interfacial tension, and [C] ≡ CN 1/2 b <strong>is</strong> a dimensionless<br />

curvature.<br />

2.3 Numerical Implementation of SCFT<br />

In th<strong>is</strong> work, the modified diffusion equation for q i (r,s) <strong>is</strong> numerically implemented using<br />

a finite difference method in one dimensional real space. Since the chemical potential<br />

33


field ω α (r) <strong>is</strong> initially unknown, it <strong>is</strong> determined in a self-cons<strong>is</strong>tent manner using proper<br />

iteration schemes.<br />

2.3.1 Finite Difference Method<br />

The modified diffusion equation (Eqn. 2.1) <strong>is</strong> spatially d<strong>is</strong>cretized in a unit cell with a flat,<br />

cylindrical, or spherical geometry and time integrated with the Crank-Nicholson technique<br />

as previous studies of the ternary polymer blends.[1, 4, 5] The Crank-Nicholson method <strong>is</strong><br />

unconditionally stable and has a truncation error of order O((∆t) 2 +h 2 ) in time and space<br />

where ∆t and h are the time and spatial step, respectively.[6]<br />

The time and spatial grid points are labeled as (s j ,r k ) where s j = j(∆t) for j =<br />

0,... , N i<br />

∆t of a molecule of type i, and where r k = kh for k = 0,... ,n for uniform grid<br />

points. In order <strong>to</strong> increase the numerical accuracy and save the computational time,<br />

the adaptive grid points are also used by adjusting the d<strong>is</strong>tribution of n + 1 node with<br />

a stretching function S(k), which <strong>is</strong> a mixture of a hypertangent and a linear function,<br />

sat<strong>is</strong>fying S(0) = 0 and S( n 2 ) = 1,<br />

⎧<br />

⎨ tanh[A( 2k −1)]+B( 2k −1)<br />

n n<br />

tanh(A)+B<br />

+ 1 : r k ≥ R, k = 0,... , n 2<br />

S(k) =<br />

⎩ tanh(A 2k n )+B 2k n<br />

tanh(A)+B<br />

: r k < R, k = 0,... , n 2<br />

where R <strong>is</strong> a radial position of a Gibbs dividing surface, and A and B are the parameters<br />

<strong>to</strong> control the node concentration near R. Such adaptive grids were usually adopted for<br />

the simulation of a weakly curved membrane in Chapter 3 by fixing R at the center of a<br />

simulation cell with A = 3 and B = 0 whereas the fixed grids were used for the micelle<br />

simulation in Chapter 4 because a micelle radius <strong>is</strong> spontaneously determined given the<br />

chemical potential of block copolymer in a grand canonical ensemble.<br />

Approximating the first and second derivative of q i (r,s) with respect <strong>to</strong> r as<br />

∂q i<br />

∂r<br />

= 1 (q i,k+1 − q i,k−1 ) (2.21)<br />

L 1<br />

( ) ]<br />

h1<br />

+ 1 q i,k + q i,k−1 (2.22)<br />

h 2<br />

∂ 2 q i<br />

∂r 2 = 1 L 2<br />

[<br />

h1<br />

h 2<br />

q i,k+1 −<br />

34


in the spatial steps h 1 and h 2 with arbitrary sizes defined by<br />

h 1 = r k − r k−1 h 2 = r k+1 − r k (2.23)<br />

L 1 = h 1 + h 2 L 2 = 1 2 h 1(h 1 + h 2 ) (2.24)<br />

transforms Eqn. 2.1 <strong>to</strong> a tridiagonal band matrix of<br />

C j+1<br />

k−1 qj+1 i,k−1 + Cj+1 k<br />

q j+1<br />

i,k<br />

+ Cj+1 k+1 qj+1 i,k+1<br />

(2.25)<br />

= C j k−1 qj i,k−1 + Cj+1 k<br />

q j i,k + Cj+1 k+1 qj i,k+1<br />

where each coefficient C j k<br />

<strong>is</strong> given by<br />

(<br />

C j+1<br />

k−1<br />

= b 2 α h 2 − (d − 1)L )<br />

2<br />

(2.26)<br />

r k L 1<br />

(<br />

C j+1<br />

k<br />

= −<br />

[b 2 αL 1 + 6L 2 ω α,k + 2 )]<br />

(2.27)<br />

∆t<br />

(<br />

C j+1<br />

k+1<br />

= b 2 α h 1 − (d − 1)L )<br />

2<br />

(2.28)<br />

r k L 1<br />

(<br />

C j k−1<br />

= −b 2 α h 2 − (d − 1)L )<br />

2<br />

(2.29)<br />

r k L 1<br />

(<br />

C j k<br />

=<br />

[b 2 α L 1 + 6L 2 ω α,k − 2 )]<br />

(2.30)<br />

∆t<br />

(<br />

C j k+1<br />

= −b 2 α h 1 − (d − 1)L )<br />

2<br />

. (2.31)<br />

r k L 1<br />

<strong>Th<strong>is</strong></strong> finite difference equation <strong>is</strong> solved under no flux boundary condition at the unit cell<br />

boundary. If a system <strong>is</strong> defined in terms of the mean curvature C = (d − 1)/R instead<br />

of r, the inverse radial d<strong>is</strong>tance r −1 = (R + z) −1 can be replaced by r −1 =<br />

−∆ ≤ z ≤ ∆.<br />

2.3.2 Numerical Accuracy<br />

C<br />

d−1+Cz where<br />

The numerical accuracy of a SCFT code with a finite difference method greatly depends on<br />

the way <strong>to</strong> d<strong>is</strong>cretize the nodes in the simulation box and the way <strong>to</strong> integrate the system<br />

in time. In th<strong>is</strong> subsection, therefore, we examine the sensitivity of interfacial tension γ<br />

35


7<br />

∆ t = 0.5<br />

6.5<br />

γ o<br />

6<br />

10 6 γ /kT (Å −2 )<br />

5.5<br />

5<br />

4.5<br />

adaptive grids<br />

fixed grids<br />

4<br />

0 500 1000 1500 2000 2500 3000 3500<br />

the number of grids (#)<br />

Figure 2.1: Interfacial tension γ vs. the number of spatial nodes in a simulation box<br />

with the fixed size of 800 Å. The adaptive grids dramatically increase the accuracy of the<br />

calculation compared <strong>to</strong> the case with the same number of grid points.<br />

and the chemical potentials µ i <strong>to</strong> the variation of the spatial and time steps in the case<br />

of χ = 0.1, N A = N B = f A N C = 100, and b α = 6Å at an arbitrary chemical potential of<br />

block copolymer.<br />

Fig. 2.1 compares interfacial tension γ calculated on fixed grids with <strong>that</strong> on adaptive<br />

grids. As <strong>to</strong>tal number of nodes increase in the system, or decreasing fixed grid step, γ<br />

increases by about 50 % and approaches a plateau. On the other hand, the adaptive grids<br />

dramatically improve the numerical accuracy giving the better estimate of γ with fewer<br />

nodes.<br />

Fig. 2.2 and Fig. 2.3 show the dependence of the interfacial tension difference ∆γ<br />

relative <strong>to</strong> the plateau value γ o marked in Fig. 2.1 and the chemical potential difference<br />

of phase I and II, ∆µ A (= µ I A − µII A ) and ∆µ C(= µ I C − µII C<br />

), at both boundaries of the<br />

simulation box on the uniform spatial grid step h for a uniform grid and the time step ∆t,<br />

respectively. As expected for a Crank-Nicholson method, these quantities <strong>have</strong> the second<br />

36


3<br />

2.5<br />

∆ t = 0.5<br />

10 6 ∆ γ / kT ( Å −2 )<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 1 2 3 4 5<br />

h 2 [ Å 2 ]<br />

Figure 2.2: The interfacial tension difference ∆γ as a function of h 2 for a uniform grid.<br />

order accuracy in space and time.<br />

On the contrary, interfacial tension and the chemical potential difference are nearly<br />

independent of the time and spatial steps, respectively, as shown in Fig. 2.4 and Fig. 2.5.<br />

For cons<strong>is</strong>tency, we applied the same time step ∆t = 0.25 for all SCFT calculations in th<strong>is</strong><br />

thes<strong>is</strong>, but adaptive grids were used for the membrane simulation, whereas h = 1Å for the<br />

micelle simulation.<br />

2.4 Iteration Scheme<br />

In th<strong>is</strong> thes<strong>is</strong>, two types of the one dimensional SCFT calculations are carried out, in both<br />

grand canonical ensemble. One <strong>is</strong> the calculation of an entire (spherical) micelle, in which<br />

the size of a whole micelle <strong>is</strong> allowed <strong>to</strong> adjust so as <strong>to</strong> minimize the grand canonical free<br />

energy. <strong>Th<strong>is</strong></strong> <strong>is</strong> presented in Chapter 4. The other <strong>is</strong> the calculation of an interface with a<br />

flat, cylinder, or spherical geometry, in which we impose an extra constraint on the position<br />

of the interface within a simulation cell. Two iteration schemes <strong>have</strong> been employed <strong>to</strong><br />

obtain the solutions, depending on the type of calculation: relaxation method <strong>is</strong> used only<br />

37


10 −1 10 −2 10 −1 10 0 10 1<br />

10 −2<br />

∆ µ A / kT, ∆ µ C / kT<br />

10 −3<br />

10 −4<br />

10 −5<br />

10 −6<br />

∆ µ A<br />

∆ µ C<br />

(∆ t) 2<br />

Figure 2.3: The chemical potential difference ∆µ A and ∆µ C as a function of (∆t) 2 .<br />

7<br />

6.5<br />

10 6 γ /kT ( Å −2 )<br />

6<br />

5.5<br />

5<br />

4.5<br />

h = 0.5 Å<br />

h = 1 Å<br />

4<br />

0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

∆ t<br />

Figure 2.4: Interfacial tension γ with various time steps ∆t for a uniform grid. There <strong>is</strong><br />

little effect of time step on γ.<br />

38


∆ t = 0.5<br />

10 −2<br />

∆ µ A / kT , ∆ µ C / kT<br />

10 −3<br />

10 −4<br />

10 −5<br />

∆ µ A<br />

∆ µ C<br />

10 −1 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

10 −6<br />

h [ Å ]<br />

Figure 2.5: The chemical potential difference ∆µ A and ∆µ C of the homopolymer A and<br />

block copolymer C determined at both boundaries of the simulation cells vs. the spatial<br />

step size h.<br />

for a swollen micelle simulation and a New<strong>to</strong>n-Raphson method has been used both for<br />

unswollen micelle simulation and for all interface simulations.<br />

2.4.1 Relaxation Method<br />

Recently Drolet and Fredrickson developed a powerful Picard-type relaxation algorithm,<br />

which was originally used <strong>to</strong> screen multiblock copolymer morphology in real space.[7, 8]<br />

In th<strong>is</strong> thes<strong>is</strong> th<strong>is</strong> method <strong>is</strong> extended <strong>to</strong> ternary polymer blends in a grand canonical<br />

ensemble.<br />

Initial guess of ω α (r) and ξ(r) are obtained from the analytical solutions for the binary<br />

homopolymer blends with infinite molecular weights given by Helfand and Tagami.[9] Let<br />

ω j α(r) and ξ j (r) be the approximate fields at the beginning of the jth iteration.<br />

1. Given ω j α(r), new density profiles φ j+1<br />

α (r) are calculated using Eqn. 2.1 and Eqn. 2.4.<br />

39


2. The potential fields ωα<br />

j+1 (r) are locally updated with a proper linearized mixing rule<br />

ω j+1<br />

α (r) = ω j α(r) + λ δω j α(r) (2.32)<br />

where the field difference δω j α <strong>is</strong> given by<br />

δω j α (r) = ∑ β<br />

χ αβ φ j+1<br />

α (r) + ξj (r) − ωα j (r) . (2.33)<br />

The weight fac<strong>to</strong>r λ, which takes the value between 0.1 and 0.2, decides the speed <strong>to</strong><br />

approach a local minimum of the <strong>to</strong>tal free energy along the local gradient direction.<br />

3. Finally a local pressure field ξ j+1 (r) <strong>is</strong> updated using Eqn. 2.2 and Eqn. 2.3. For<br />

instance, in the ternary system of A/B/AB<br />

ξ j+1 (r) = 1 2 (ωj+1 (r) + ωj+1(r) − χ) (2.34)<br />

A<br />

B<br />

where χ = χ AB . At th<strong>is</strong> step, the chemical potential fields ω α (r) and a Lagrange multiplier<br />

ξ(r) are not forced <strong>to</strong> <strong>have</strong> van<strong>is</strong>hing spatial average as suggested by Rasmussen and<br />

Kalosakas.[10]<br />

4. The whole procedure <strong>is</strong> repeated until max(|δωα|) j drops below a set <strong>to</strong>lerance (10 −8 ).<br />

Through th<strong>is</strong> thes<strong>is</strong>, most SCFT results were obtained using a New<strong>to</strong>n-Raphson method<br />

which <strong>is</strong> described in detail in the next subsection because the relaxation method usually<br />

showed the limited stability, especially in the flat geometry, only when χN of the system<br />

was small enough. However, we had <strong>to</strong> rely on only th<strong>is</strong> iteration scheme for the purpose<br />

of simulating highly swollen micelles of nearly symmetric block copolymer emulsifying the<br />

minor homopolymer phase in the core (Chapter 4) since in th<strong>is</strong> case the New<strong>to</strong>n-Raphson<br />

method did not converge due <strong>to</strong> the extremely shallow minima of the free energy. In th<strong>is</strong><br />

case, the number of iteration steps usually exceed several millions as the block ratio f A<br />

approaches 1/2.<br />

40


2.4.2 New<strong>to</strong>n-Raphson Algorithm<br />

A New<strong>to</strong>n-Raphson iteration scheme <strong>is</strong> based on the first order Taylor expansion of the<br />

residual equations by assuming the variation of the equations <strong>is</strong> locally linear near the<br />

solutions. Different variants of New<strong>to</strong>n-Raphson iteration were used for micelle simulations,<br />

in which we solve only the basic SCF equations, and for simulations of interfaces, in which<br />

we solve SCF equations subject <strong>to</strong> a constraint on the interface position.<br />

We first describe the algorithm used in micelle simulations. In SCFT, the constraints of<br />

the chemical potential fields ω α (r) in Eqn. 2.2 and the incompressibility in Eqn. 2.3 can be<br />

used <strong>to</strong> set up the residual vec<strong>to</strong>r R (α,k) with a dimension of C(n+1) where α = A,B,... , C<br />

and k = 0,... ,n in case of the canonical system. For example, in the ternary system of<br />

A/B/AB with C = 2, R (α,k) can be defined by<br />

R (A,k) = χ(φ A,k − φ B,k ) − (ω B,k − ω A,k ) (2.35)<br />

R (B,k) = φ A,k + φ B,k − 1 . (2.36)<br />

from which the Jacobian matrix J (α,k)(β, l) can be establ<strong>is</strong>hed as<br />

J (α,k)(β, l) = ∂R (α,k)<br />

∂ω β,l<br />

(2.37)<br />

with a dimension of 2(n + 1) × 2(n + 1) and further reduced <strong>to</strong><br />

J (A,k)(β, l)<br />

J (B,k)(β, l)<br />

(<br />

∂φA,k<br />

= χ − ∂φ )<br />

B,k<br />

− δ k, l (δ β,B − δ β,A ) (2.38)<br />

∂ω β,l ∂ω β,l<br />

(<br />

∂φA,k<br />

= δ k, l + ∂φ )<br />

B,k<br />

(2.39)<br />

∂ω β,l ∂ω β,l<br />

where β = A,B,... , C and l = 0,... ,n. The Jacobian matrix was numerically obtained in<br />

th<strong>is</strong> thes<strong>is</strong> by perturbing ω β,l by an amount of δω β . Finally, the chemical potential fields<br />

are updated as<br />

ω j+1<br />

α,k = ωj α,k − J −1<br />

(α,k)(β, l) R (β,l) (2.40)<br />

41


for the next iteration step j + 1 until the solutions converge.<br />

We next d<strong>is</strong>cuss the iteration scheme for simulations of interfaces. For a flat interface,<br />

the position of an interface <strong>is</strong> arbitrary because the pressure <strong>is</strong> identical through an interface.<br />

Thus we fix the location of an interface at the center of a simulation box imposing<br />

an extra constraint of equal amount of A and B homopolymers in the system,<br />

R (B,n+1) = ¯φ A − ¯φ B (2.41)<br />

by slightly adjusting the chemical potential of the B homopolymer, µ B .<br />

For a weakly curved phase, the radius of curvature <strong>is</strong> a very sensitive function of the<br />

pressure (or chemical potential) difference. The position of a curved interface can also<br />

be determined in a similar manner by matching the prescribed curvature with the Gibbs<br />

dividing surface of no excess A monomers such as<br />

I<br />

R (B, n+1) = V I ¯φ A + V II<br />

II ¯φ A − V ¯φ A (2.42)<br />

where V I and V II are the volume of the I and II phases with the constraint of V = V I +V II ,<br />

and ¯φ I<br />

A<br />

and ¯φ<br />

II<br />

A<br />

are the bulk volume fractions of A monomer in the corresponding phases.<br />

To sat<strong>is</strong>fy the extra constraint, µ B <strong>is</strong> updated using the expanded (2n + 3) × (2n + 3)<br />

Jacobian matrix as<br />

µ j+1<br />

B<br />

= µj B − J −1<br />

(B, n+1)(β,l) R (β,l) (2.43)<br />

where l = 0,... ,n + 1.<br />

42


Bibliography<br />

[1] Nalani, J.; Hong, K. M. Macromolecules 1982, 15, 482;<br />

[2] Matsen, M. W. Phys. Rev. Lett. 1995, 74, 4225; Matsen, M. W. Macromolecules 1995,<br />

28, 5765<br />

[3] Matsen, M. W. J. Chem. Phys. 1999, 110, 4659<br />

[4] Noolandi, J.; Hong, K. M. Macromolecules 1984, 17, 1531<br />

[5] Shull, K. R.; Kramer, E. J. Macromolecules 1990, 23, 4769<br />

[6] Burden, R. L.; Faires, J. D. Numerical Analys<strong>is</strong> (Brooks/Cole Publ<strong>is</strong>hing Company)<br />

1997<br />

[7] Drolet, F.; Fredrickson, G. H. Phys. Rev. Lett. 1999, 83, 4317<br />

[8] Drolet, F.; Fredrickson, G. H. Macromolecules 2001, 34, 5317<br />

[9] Helfand, E.; Tagami, Y. J. Chem. Phys. 1972, 56, 3592<br />

[10] Rasmussen, K, ∅; Kalosakas, G. J. Polym. Sci.: B 2002, 40, 1777<br />

43


Chapter 3<br />

Interfacial Bending Elasticity<br />

Abstract<br />

Self-cons<strong>is</strong>tent field theory (SCFT) has been used <strong>to</strong> calculate the Helfrich bending elastic<br />

constants for monolayers of A/B diblock copolymers at interfaces between imm<strong>is</strong>cible A<br />

and B homopolymer liquids. We focus here on the properties of saturated monolayers,<br />

with van<strong>is</strong>hing interfacial tension. We also determine values of the volume fraction f A<br />

of the A-block within the copolymer for which the monolayer has van<strong>is</strong>hing spontaneous<br />

curvature in systems with asymmetric homopolymers. The procedure used here <strong>to</strong> define<br />

and calculate the curvature-dependent interfacial tension <strong>is</strong> somewhat different than <strong>that</strong><br />

used previously by Matsen.<br />

3.1 Introduction<br />

Mixtures of two imm<strong>is</strong>cible homopolymers and a compatibilizing diblock copolymer surfactant<br />

exhibit phase behavior <strong>that</strong> <strong>is</strong> often closely analogous <strong>to</strong> <strong>that</strong> of mixtures of oil, water,<br />

and a small molecule surfactant. Some such polymer mixtures form equilibrium structures<br />

cons<strong>is</strong>ting of domains of nearly pure A or B homopolymer separated by copolymer monolayers,<br />

including both d<strong>is</strong>ordered microemulsion phases and swollen ordered phases. When<br />

the character<strong>is</strong>tic interfacial radii of curvature are sufficiently large (i.e., much larger than<br />

the monolayer thickness), the thermodynamic competition between phases of similar com-<br />

44


position <strong>that</strong> differ in the geometrical arrangement of interfaces can be described by the<br />

Canham-Helfrich theory of interfacial bending elasticity.[1, 2] The chemical potential of<br />

surfactant in such structures tends <strong>to</strong> remain very close <strong>to</strong> the value at which the microscopic<br />

interfacial tension van<strong>is</strong>hes [3, 4, 5], creating so-called saturated monolayers, for<br />

reasons <strong>that</strong> were first articulated by Schulman.[6]<br />

Of particular interest for some purposes <strong>is</strong> the identification of surfactants <strong>that</strong> form<br />

saturated monolayers with van<strong>is</strong>hing spontaneous curvature, which are referred <strong>to</strong> here as<br />

balanced monolayers. Both theory and experiments with small molecule surfactants [7, 8]<br />

suggest <strong>that</strong> minimum possible macroscopic interfacial tension between oil- and waterrich<br />

(or A- and B-rich) phases (which can be several orders of magnitude below <strong>that</strong><br />

obtained with no surfactant) <strong>is</strong> obtained when balanced monolayers are formed. The creation<br />

of balanced monolayers <strong>is</strong> also believed <strong>to</strong> be a necessary condition for the ex<strong>is</strong>tence<br />

of three-phase coex<strong>is</strong>tence of excess A- and B-rich phases with a balanced bicontinuous microemulsion,<br />

within which surfactant monolayers divide space in<strong>to</strong> A- and B-rich domains<br />

with nearly equal volume fractions. In mixtures of oil, water, and some small nonionic<br />

C i E j surfactants, for which the conformation of the water-soluble block of the surfactant<br />

<strong>is</strong> very sensitive <strong>to</strong> changes in temperature, balanced monolayers can be created by adjusting<br />

the temperature, leading <strong>to</strong> the formation of a balanced bicontinuous microemulsion at<br />

a specific temperature. Balanced bicontinuous microemulsion phases <strong>have</strong> been observed<br />

for ternary polymer systems of A and B homopolymers mixed with both symmetric AB<br />

copolymers [9, 10, 11] and carefully chosen AC [12, 13] block copolymers. The design<br />

of balanced bicontinuous microemulsions for systems with asymmetric homopolymers <strong>is</strong><br />

of interest as a strategy for the design of polymer alloys with interpenetrating network<br />

structures.<br />

The bending elasticity of a copolymer monolayer has been considered in several previous<br />

theoretical studies [3, 4, 14]. The Helfrich elastic constants were calculated by Wang<br />

45


and Safran [3, 4] using strong stretching theory, and by Matsen [14] using numerical SCFT,<br />

as also done here. Matsen proposed a general method for calculating the Helfrich elastic<br />

constants, but gave results only for symmetric systems containing homopolymers of equal<br />

molecular weights and a symmetric block copolymer, for which the monolayer has a van<strong>is</strong>hing<br />

spontaneous curvature by symmetry. Matsen devoted as least as much attention <strong>to</strong><br />

systems with nonzero interfacial tensions as <strong>to</strong> saturated monolayers. Here, we give results<br />

for asymmetric as well as symmetric mixtures, and focus on more fully characterizing the<br />

case of saturated monolayers <strong>that</strong> <strong>is</strong> relevant <strong>to</strong> self-assembly. In mixtures with asymmetric<br />

homopolymers, we emphasize the identification of values of the copolymer composition<br />

required <strong>to</strong> form a balanced monolayer. The SCFT calculations presented here are also<br />

based upon a definition of the excess free energy of a curved monolayer somewhat different<br />

from <strong>that</strong> used by Matsen, as d<strong>is</strong>cussed in more detail below.<br />

3.2 Interfacial Thermodynamics<br />

We consider an interface in an incompressible ternary mixture of two imm<strong>is</strong>cible homopolymers,<br />

A and B, and an AB diblock copolymer, under conditions for which system phase<br />

separates in<strong>to</strong> two phases rich in A and B, respectively. Monomers of types A and B<br />

occupy an equal volume v, and <strong>have</strong> stat<strong>is</strong>tical segment lengths b A and b B . Let N A , N B ,<br />

and N C be the degrees of polymerization of the two homopolymers and the copolymer<br />

(C), respectively. Let f A be the volume fraction of the A block within the copolymer, and<br />

f B = 1 − f A . Let α A ≡ N A /N C , α B ≡ N A /N C , and β ≡ N A /N B .<br />

It <strong>is</strong> convenient <strong>to</strong> formulate both the classical thermodynamics and the SCFT calculations<br />

for a monolayer in the grand canonical stat<strong>is</strong>tical ensemble. The grand canonical<br />

free energy Φ of any system <strong>is</strong><br />

Φ = F − ∑ i<br />

µ i M i (3.1)<br />

where F <strong>is</strong> the Helmholtz free energy, and µ i <strong>is</strong> the chemical potential and M i the <strong>to</strong>tal<br />

46


number of molecules of species i in the system. For any macroscopic phase of volume V<br />

and pressure P, Φ = −PV .<br />

Interfacial tension γ of a flat macroscopic interface <strong>is</strong> the interfacial excess grandcanonical<br />

free energy per unit area of interface, so <strong>that</strong> Φ = −PV + Aγ for a system of<br />

volume V with an interface of area A. Here, we consider a curved interface, and define γ<br />

in the same manner, by defining<br />

Φ ≡ −P I V I − P II V II + Aγ (3.2)<br />

for some choice of Gibbs dividing surface <strong>that</strong> separates an A-rich domain (phase I) from<br />

a B-rich domain (phase II), where V I and V II are the volumes of the domains, P I and P II<br />

are the pressures of bulk phases I and II at the temperature and chemical potentials of<br />

interest, and A <strong>is</strong> the area of the dividing surface.<br />

In the case of a flat interface, the value of γ <strong>is</strong> independent of our choice of a convention<br />

for position of the dividing surface. Mathematically, th<strong>is</strong> <strong>is</strong> a result of the fact <strong>that</strong> in th<strong>is</strong><br />

case, for which P I = P II = P, both the sum P I V I + P II V II = PV and the area A are<br />

independent of the position of the interface. For a curved interface, however, P I and P II<br />

are generally unequal (for reasons d<strong>is</strong>cussed below), and A and γ both depend upon the<br />

choice of the dividing surface. The relations among interfacial excess quantities <strong>that</strong> are<br />

obtained below are, however, valid for any choice of dividing surface.<br />

3.2.1 Mechanical Equilibrium<br />

For a curved surface <strong>to</strong> be in mechanical equilibrium, the <strong>to</strong>tal free energy Φ must be an<br />

extrema with respect <strong>to</strong> normal d<strong>is</strong>placements of the interface at fixed chemical potentials.<br />

For a flat surface, th<strong>is</strong> mechanical equilibrium condition reduces <strong>to</strong> the statement <strong>that</strong><br />

pressures must be equal on either side of the interface.<br />

Consider the mechanical equilibrium condition for a curved interface. For simplicity, we<br />

restrict ourselves <strong>to</strong> situations in which the inner B-rich domain (II), <strong>is</strong> either a spherical<br />

47


or cylindrical domain of radius R. Differentiating Eqn. 3.2 with respect <strong>to</strong> R while holding<br />

temperature and all chemical potentials fixed yields the condition<br />

0 = ∂Φ<br />

∂R = (P I − P II ) ∂V II<br />

∂R + ∂(γA)<br />

∂R<br />

, (3.3)<br />

The required derivative of volume V II with respect <strong>to</strong> radius <strong>is</strong> area: ∂V II /∂R = −∂V I /∂R =<br />

A. The derivative of the area <strong>is</strong> given by ∂A/∂R = AC, where<br />

C = d − 1<br />

R<br />

(3.4)<br />

<strong>is</strong> the mean curvature, in which d = 3 for a sphere of radius R and d = 2 for a cylinder.<br />

Substitution in Eqn. 3.3 then yields a pressure difference<br />

P II − P I = Cγ + ∂γ<br />

∂R<br />

(3.5)<br />

If we parameterize γ at fixed temperature and chemical potential as a function of C, rather<br />

than R, th<strong>is</strong> becomes<br />

P II − P I = Cγ −<br />

C2 ∂γ<br />

d − 1 ∂C . (3.6)<br />

<strong>Th<strong>is</strong></strong> mechanical equilibrium condition reduces <strong>to</strong> the usual Young-Laplace condition for<br />

the pressure drop across a curved interface in the limit γ ≫ |C ∂γ<br />

∂C<br />

| in which the curvature<br />

dependence of γ <strong>is</strong> negligible. It further reduces <strong>to</strong> the condition P I = P II in the limit<br />

C = 0, thus recovering the usual condition for macroscopic phase coex<strong>is</strong>tence only in the<br />

case of a flat interface.<br />

3.2.2 Thermodynamic Variables<br />

A straightforward extension of the Gibbs phase rule may be used <strong>to</strong> show <strong>that</strong> interfacial<br />

tension γ of a mechanically stable spherical or cylindrical interface in an incompressible<br />

ternary mixture <strong>is</strong> a function of three parameters, which we may take <strong>to</strong> be temperature<br />

T, block copolymer chemical potential µ C , and interfacial curvature C. To show th<strong>is</strong>, we<br />

first specify the state of such a system in terms of five degrees of freedom, which may<br />

48


e taken <strong>to</strong> be temperature, three chemical potentials (which must be the same in both<br />

phases), and interfacial curvature. The requirement <strong>that</strong> the interface be in mechanical<br />

equilibrium removes one degree of freedom, leaving four degrees of freedom, one of which<br />

<strong>is</strong> the curvature. The usual Gibbs phase rule, which yields three degrees of freedom for a<br />

ternary two-phase system, corresponds <strong>to</strong> the case C = 0 of a flat interface.<br />

The restriction <strong>to</strong> incompressible fluids reduces the number of physically relevant degrees<br />

of freedom by one more: Changes in the pressure of an incompressible mixture <strong>have</strong><br />

no effect upon the state of the system other than <strong>to</strong> cause trivial shifts in the values of the<br />

chemical potentials, without changing γ or any other interfacial excess properties. <strong>Th<strong>is</strong></strong><br />

trivial dependence on hydrostatic pressure <strong>is</strong> d<strong>is</strong>cussed in the context of SCFT in Chapter<br />

2.2.1. The state of a mechanically stable curved interface in an incompressible ternary<br />

mixture may thus be specified as a function of three variables, e.g., T, µ C , and C.<br />

In order <strong>to</strong> unambiguously define a functional relationship γ(T,µ C ,C) for an incompressible<br />

system, however, one must introduce a standard state for chemical potential (as<br />

always) and a convention <strong>to</strong> fix a value of the pressure in one of the two bulk phases. In numerical<br />

work, for flat interfaces, we evaluate γ(T,µ C ) in the macroscopic state in which the<br />

bulk pressure van<strong>is</strong>hes in both phases. For curved interfaces, we <strong>have</strong> chosen <strong>to</strong> calculate<br />

γ(T,µ C ,C) in a state in which µ A <strong>is</strong> kept fixed at the value obtained in the macroscopic<br />

two-phase with P I = P II = 0 at the prescribed value of µ C , in which µ B must be adjusted<br />

so as <strong>to</strong> create a stable interface with the chosen curvature. The results obtained with<br />

th<strong>is</strong> convention are expected <strong>to</strong> be very similar <strong>to</strong> those <strong>that</strong> would be obtained by instead<br />

requiring <strong>that</strong> P = 0 in the A-phase for all values of C: Changing µ B at fixed µ A changes<br />

the pressure in the A-rich phase only as a result of changes in the osmotic pressure ar<strong>is</strong>ing<br />

from changes in the concentration of d<strong>is</strong>solved B homopolymer, which <strong>is</strong> generally very<br />

small in the situations of interest.<br />

An interface in which γ = 0 will hereafter be referred <strong>to</strong> as a saturated interface.<br />

49


Let µ ∗ C (T) denote the value of µ C at which γ = 0 for a flat interface. In phases <strong>that</strong><br />

contain an extensive area of surfactant monolayers between A- and B-rich domains, such<br />

as a microemulsion or swollen lamellar phase, the interfaces are generally very nearly<br />

saturated. [6, 5]<br />

3.2.3 Helfrich Expansion<br />

We consider a Helfrich expansion of γ about the flat saturated state, at fixed temperature<br />

T, as a function of the principal curvatures of the interface and of the deviation<br />

δµ ≡ µ C − µ ∗ C . (3.7)<br />

Expanding <strong>to</strong> quadratic order in curvature and linear order in δµ yields an expansion of<br />

the form<br />

γ = −τC + 1 2 κC2 + ¯κK + Γ ∗ δµ + ΛδµC (3.8)<br />

where C = R −1<br />

1 +R −2<br />

2 <strong>is</strong> the mean curvature and K = 1/(R 1 R 2 ) <strong>is</strong> the Gaussian curvature<br />

of a surface with principal radii of curvature R 1 and R 2 . For a sphere, C = 2/R and<br />

K = 1/R 2 , while for a cylinder C = 1/R and K = 0. The coefficients κ and ¯κ are bending<br />

and Gaussian rigidities, respectively, while<br />

τ ≡ ∂γ<br />

∣<br />

∂C<br />

∣<br />

µC =µ ∗ C<br />

(3.9)<br />

<strong>is</strong> a parameter <strong>that</strong> controls the asymmetry of the monolayer. The coefficient of the term<br />

linear in δµ <strong>is</strong> given by the Gibbs adsorption equation<br />

Γ ∗ = − ∂γ<br />

∂µ C<br />

∣<br />

∣∣∣C=0<br />

(3.10)<br />

in which Γ ∗ ≡ Γ ∗ C<br />

<strong>is</strong> the excess number density (copolymers per area) of copolymers in a<br />

saturated flat interface. The coefficient Λ <strong>is</strong> the second derivative Λ ≡<br />

∂2 γ<br />

∂µ C ∂C .<br />

50


The interfacial tension of a cylindrical or spherical surface at the constant saturation<br />

chemical potential µ C = µ ∗ C<br />

may be expressed as a harmonic function<br />

γ = −τC + 1 2 κ′ C 2 (3.11)<br />

where κ ′ = κ for a cylindrical surface or κ ′ = κ + ≡ κ + ¯κ/2 for a spherical surface. The<br />

minimum of interfacial tension with respect <strong>to</strong> C at µ C = µ ∗ C<br />

<strong>is</strong> given for a cylindrical<br />

surfaces by the ratio τ/κ, which <strong>is</strong> often referred <strong>to</strong> as the spontaneous curvature (though<br />

“spontaneous cylindrical curvature” would arguably be more appropriate). The corresponding<br />

spontaneous curvature for a spherical surface <strong>is</strong> given by τ/κ + . The spontaneous<br />

curvatures τ/κ and τ/κ + correspond <strong>to</strong> the thermodynamically preferred radii of swollen<br />

cylindrical or spherical micelles in an A matrix <strong>that</strong> coex<strong>is</strong>ts with an excess B-rich phase.<br />

In the absence of a coex<strong>is</strong>ting excess phase, micelles generally adopt smaller radii as the<br />

result of a s<strong>to</strong>ichiometric constraint on the amount of available B homopolymer.<br />

A saturated monolayer for which τ = 0 will be referred <strong>to</strong> here as a balanced monolayer.<br />

Symmetric mixtures, with f A = 1/2, N A = N B , and b A = b B , must form balanced<br />

monolayers, by symmetry. In systems with asymmetric homopolymers, with N A ≠ N B<br />

and/or b A ≠ b B , let f bal<br />

A<br />

denote the value of f A necessary <strong>to</strong> create a balanced monolayer.<br />

3.2.4 Lifshitz Point<br />

Consider an incompressible ternary system containing two coex<strong>is</strong>ting phases rich in A<br />

and B. For a specified set of three molecules, the behavior of such a system depends<br />

on only surfactant chemical potential µ C and temperature. For simplicity, we assume in<br />

the following d<strong>is</strong>cussion <strong>that</strong> χ mono<strong>to</strong>nically decreases with increasing T, and ignore the<br />

possibility of three-phase coex<strong>is</strong>tence.<br />

In the absence of copolymer, A and B are completely m<strong>is</strong>cible above a binary critical<br />

temperature T c . As T <strong>is</strong> decreased below T c , a two-phase region of Gibbs phase triangle<br />

emerges from the binary critical point (a), as shown in Fig. 3.1. At temperatures slightly<br />

51


Figure 3.1: Schematic view of the state of a two-phase system at fixed pressure as a function<br />

of T and of a copolymer activity Z C ∝ e µ C/kT , which <strong>is</strong> defined for th<strong>is</strong> purpose so <strong>that</strong><br />

Z C <strong>is</strong> proportional <strong>to</strong> the concentration of copolymer d<strong>is</strong>solved in one of the two coex<strong>is</strong>ting<br />

phases. The Lifshitz point <strong>is</strong> the intersection of a line of critical points, for which the<br />

copolymer concentration tends <strong>to</strong> increase with decreasing temperature, with a saturation<br />

line e µ∗ C (T) , for which copolymer concentration decreases with decreasing temperature as<br />

the copolymer becomes less soluble.<br />

52


elow T c , the two phase region generally terminates at a critical point (b) <strong>that</strong> moves <strong>to</strong><br />

higher surfactant concentration with decreasing T. Within the two phase region, each<br />

tie-line may be characterized by a value of µ C and a corresponding value of the interfacial<br />

tension γ(T,µ C ,C = 0) of a flat macroscopic interface. <strong>Th<strong>is</strong></strong> interfacial tension generally<br />

decreases with increasing µ C and, at temperatures only slightly below T c , van<strong>is</strong>hes only at<br />

the critical point, where the the compositions of the coex<strong>is</strong>ting phase merge. At temperatures<br />

below a Lifshitz temperature T L , however, the macroscopic interfacial tension may<br />

be driven <strong>to</strong> zero at chemical potential µ ∗ C<br />

below the value at the critical point (c), producing<br />

a saturated monolayer between d<strong>is</strong>tinct A- and B-phases. At temperatures T < T L<br />

for which such a saturated monolayer can ex<strong>is</strong>t, increasing temperature causes the critical<br />

point c <strong>to</strong> move <strong>to</strong> lower copolymer concentrations and the saturation tie line <strong>to</strong> move<br />

<strong>to</strong>ward higher copolymer concentrations, until they merge at the Lifshitz point, denoted<br />

by L in Fig. 3.1. By examining instabilities of the d<strong>is</strong>ordered phase, the Lifshitz point<br />

can also be identified as the point along the line of critical points (a − L) at which the<br />

divergence of the structure fac<strong>to</strong>r S(q) at q = 0, which signals an instability <strong>to</strong> macroscopic<br />

phase separation, <strong>is</strong> first preempted by divergence of S(q) at q ≠ 0, signalling an instability<br />

<strong>to</strong>wards formation of a mesophase. As the Lifshitz point <strong>is</strong> approached along the<br />

saturation line µ ∗ C<br />

(T), the compositions of the two coex<strong>is</strong>ting phases must merge, and so<br />

the Helfrich elastic constants of a saturated monolayer must approach zero at the Lifshitz<br />

temperature.<br />

Broseta and Fredrickson [15] <strong>have</strong> considered thermodynamics of symmetric mixtures,<br />

with f = 1/2, α A = α B = α, and b A = b B . For such systems, the line of critical points<br />

must stay within an <strong>is</strong>opleth φ A = φ B . In th<strong>is</strong> case, a stable Lifshitz point occurs at<br />

(χN C ) L = 2(1 + 2α2 )<br />

α<br />

(φ C ) L = 2α2<br />

1 + 2α 2 (3.12)<br />

for all α < 1. [15, 16]. For α > 1, they found <strong>that</strong> the appearance of the Lifshitz point<br />

53


<strong>is</strong> preempted by the appearance of a region of coex<strong>is</strong>tence between three homogeneous<br />

phases.<br />

A Lifshitz point will generally also ex<strong>is</strong>t in systems with slightly asymmetric homopolymers<br />

and/or an asymmetric copolymer. For each choice of a set of values N A , N B , and N C ,<br />

however, there ex<strong>is</strong>ts a unique value of f A , given by the limit lim T →TL fA bal (T), for which<br />

a balanced saturated monolayer will be formed at temperatures infinitesimally below T L .<br />

Fredrickson and Bates [16] <strong>have</strong> proposed the identification of such balanced Lifshitz points<br />

as a strategy for the formulation of balanced bicontinuous microemulsions in mixtures with<br />

asymmetric polymers. They also proposed an approximate criterion for identifying such<br />

points: They required <strong>that</strong> the critical fluctuation mode at the Lifshitz point not involve<br />

any fluctuation in surfactant concentration. <strong>Th<strong>is</strong></strong> <strong>is</strong> equivalent <strong>to</strong> requiring <strong>that</strong> the slope<br />

of the tie-lines infinitesimally close <strong>to</strong> the critical point at T = T L should be parallel <strong>to</strong><br />

the binary A-B homopolymer edge of the Gibbs phase triangle. <strong>Th<strong>is</strong></strong> criterion was found<br />

<strong>to</strong> be sat<strong>is</strong>fied for any critical point with<br />

f A = √ β/(1 + √ β). (3.13)<br />

<strong>Th<strong>is</strong></strong> criterion does not appear <strong>to</strong> us <strong>to</strong> be equivalent <strong>to</strong> the requirement τ = 0 near the<br />

Lifshitz point, but the two criteria are shown <strong>to</strong> yield similar results in section 3.4.<br />

3.3 SCFT Methodology<br />

SCFT calculations of γ for spherically and cylindrically curved interfaces with a radius<br />

R <strong>have</strong> been carried out using the grand-canonical formulation, which <strong>is</strong> d<strong>is</strong>cussed in<br />

Chapter 2.2.2. Calculation of γ for a weakly curved interface requires the construction of<br />

a spherically or cylindrically symmetric solution <strong>to</strong> the SCF equations within an annular<br />

domain R −<br />

< r < R + , where r <strong>is</strong> a radial coordinate, in which R − and R + are far<br />

enough from the dividing surface so <strong>that</strong> the solution <strong>is</strong> insensitive <strong>to</strong> changes in these cell<br />

54


oundaries. To calculate the elastic parameters for an expansion about a saturated flat<br />

interface, for a specified set of molecules at a specified value of χ, µ ∗ C<br />

<strong>is</strong> first determined<br />

by calculating interfacial tension γ(µ C ,C = 0) of a flat interface for several values of µ C .<br />

In these simulations, the chemical potentials µ A and µ B are chosen for each value of µ C so<br />

as <strong>to</strong> sat<strong>is</strong>fy the condition P I = P II = 0. The required values of chemical potential can be<br />

calculated from Flory-Huggins theory. Once µ ∗ C<br />

<strong>is</strong> known for a specific system, calculations<br />

of γ(µ ∗ C<br />

,C) are conducted for both cylindrical and spherically curved interfaces at several<br />

values of R or C. For each such calculation, the values of µ A and µ C are kept fixed at the<br />

values obtained for a saturated flat monolayer at zero bulk pressure, and µ B <strong>is</strong> adjusted<br />

in each calculation as <strong>to</strong> obtain a prescribed radius of curvature R for the Gibbs dividing<br />

surface. We take the Gibbs dividing surface <strong>to</strong> be the equimolar surface for either type of<br />

monomer, defined so <strong>that</strong> there <strong>is</strong> a van<strong>is</strong>hing interfacial excess of either A or B monomers,<br />

when local A or B monomer concentrations are taken <strong>to</strong> include contributions from both<br />

one homopolymer and one block of the copolymer. Values of κ and κ + are obtained for<br />

each system of interest by fitting SCFT results of γ(µ ∗ C<br />

,C) for cylindrical and spherical<br />

surfaces, separately, <strong>to</strong> Eqn. 3.8.<br />

For both flat and curved interfaces, a New<strong>to</strong>n-Raphson iteration scheme <strong>is</strong> used <strong>to</strong><br />

solve the grand canonical SCF equations constrained simultaneously with a constraint<br />

requiring <strong>that</strong> the dividing surface <strong>have</strong> a radial position R = (R + + R − )/2 in the middle<br />

of the unit cell. The macroscopic chemical potential µ B <strong>is</strong> adjusted so as <strong>to</strong> sat<strong>is</strong>fy th<strong>is</strong><br />

additional constraint, and each iteration thus involves a simultaneous adjustment of the<br />

SCFT chemical potential fields ω α at all spatial grid points and the macroscopic chemical<br />

potential µ B . In grand-canonical ensemble, it <strong>is</strong> necessary <strong>to</strong> force the dividing surface <strong>to</strong><br />

take a specified position within the unit cell in order <strong>to</strong> obtain a unique, stable solution<br />

even for a flat interface, because the position of a flat interface within the unit cell <strong>is</strong><br />

otherw<strong>is</strong>e indeterminate.<br />

55


0.1<br />

0.05<br />

f A =0.5098<br />

0<br />

f A =0.5455<br />

−0.05<br />

f A =0.5833<br />

γ / γ ο<br />

−0.1<br />

−0.15<br />

−0.2<br />

0.003<br />

γ / γ ο<br />

f A =0.5098<br />

f A =0.6667<br />

−0.25<br />

−0.001<br />

0 0.0025<br />

R −1 [Å −1 ]<br />

−0.3<br />

0 0.002 0.004 0.006 0.008 0.01<br />

R −1 [Å −1 ]<br />

Figure 3.2: SCFT results for interfacial tension γ vs. half curvature C/2 = 1/R for<br />

spherical interfaces with R ≥ 100 Å, for system with the copolymer of a fixed size of<br />

B block f B χN C = 10 and symmetric homopolymers of fixed lengths α A = α B = f B ,<br />

with b A /b B = 1, for several values of f A . Solid lines are quadratic fits of the curvature<br />

dependence at small curvatures <strong>to</strong> Eqn. 3.8, from which κ + and τ can be extracted.<br />

0.04<br />

0.02<br />

0<br />

f A =0.5098<br />

f A =0.5455<br />

γ / γ ο<br />

−0.02<br />

−0.04<br />

f A =0.5833<br />

−0.06<br />

0.003<br />

f A =0.5098<br />

−0.08<br />

γ / γ ο<br />

f A =0.6667<br />

−0.1<br />

−0.001<br />

0 0.0025<br />

R −1 [Å −1 ]<br />

−0.12<br />

0 0.002 0.004 0.006 0.008 0.01<br />

R −1 [Å −1 ]<br />

Figure 3.3: SCFT results for interfacial tension γ vs. curvature C = 1/R for cylindrical<br />

interfaces with R ≥ 100 Å, for the same mixtures as those studied in the previous figure.<br />

κ can be obtained.<br />

56


Fig. 3.2 and Fig. 3.3 show examples of results for γ vs. 1/R at µ C = µ ∗ C<br />

, for spherical<br />

and cylindrical geometries respectively, for a series of systems in which N A = N B = f B N C<br />

and f B χN C = 10, and in which the size of the A block of the copolymer <strong>is</strong> varied <strong>to</strong> change<br />

f A . For both spherical and cylindrical deformations, the Helfrich expansion <strong>is</strong> found <strong>to</strong> be<br />

accurate for th<strong>is</strong> choice of parameters for surfaces with radii R 150 Å, below which γ<br />

starts <strong>to</strong> deviate significantly from Eqn. 3.8. The radius at which th<strong>is</strong> deviation becomes<br />

significant <strong>is</strong> comparable <strong>to</strong> <strong>that</strong> of a ”dry” binary micelle of the copolymer used here, in<br />

which no B homopolymer <strong>is</strong> contained in the core. The minimum in γ vs. 1/R for nearly<br />

symmetric copolymers corresponds <strong>to</strong> the thermodynamically preferred (i.e., spontaneous)<br />

curvature for a swollen cylindrical micelle or spherical micelle. Surpr<strong>is</strong>ingly, as well as<br />

1/R, f A strongly affects th<strong>is</strong> preferred curvature so <strong>that</strong> it exceeds the range of values for<br />

which the Helfrich theory remains valid when f A ≥ 0.58. The Helfrich theory <strong>is</strong> thus found<br />

<strong>to</strong> be valid over a rather wide range of curvatures in the case of symmetric copolymer,<br />

as noted previously by Matsen, [14] but <strong>is</strong> useful for asymmetric monolayers only over a<br />

rather limited range of f A , because of the rapid variation of spontaneous curvature with<br />

f A .<br />

We <strong>have</strong> shown in Chapter 2.2.4 <strong>that</strong>, as noted previously by Matsen [14], the SCF<br />

expression for interfacial tension of a spherical or cylindrical surface of radius R may be<br />

expressed in the dimensionless form<br />

γ =<br />

b<br />

N 1/2 [γ] (3.14)<br />

v<br />

where we <strong>have</strong> chosen N = N C as a reference degree of polymerization and b = b B as<br />

reference stat<strong>is</strong>tical segment length, and where [γ] <strong>is</strong> dimensionless interfacial tension <strong>that</strong><br />

can depend only upon the dimensionless arguments χN, f A , α A , α B , b B /b A , µ C /kT, and<br />

the dimensionless radius R/(N 1/2 b). Following Matsen, we express the Helfrich expansion<br />

57


for a membrane with µ C = µ ∗ C<br />

in nondimensional form as<br />

[γ] = 1 2 [κ][C]2 + [¯κ][K] 2 − [τ][C] (3.15)<br />

where [C] ≡ C N 1/2 b and [K] ≡ KNb 2 are dimensionless mean and Gaussian curvatures,<br />

and<br />

κ =<br />

¯N 1/2 [κ]<br />

¯κ = ¯N 1/2 [¯κ] (3.16)<br />

τ =<br />

¯N<br />

1/2<br />

N 1/2 b [τ] ,<br />

where ¯N ≡ Nb 6 /v 2 , and where [κ], [¯κ], and [τ] are dimensionless elastic parameters <strong>that</strong><br />

depend only on χN, f A , α A , α B , and b B /b A ,<br />

As noted in Introduction, the definition of the excess free energy density γ for a curved<br />

interface used here <strong>is</strong> slightly different from <strong>that</strong> used by Matsen.[14] Solutions <strong>to</strong> the SCF<br />

equations are always extrema of a corresponding SCF free energy functional, and can thus<br />

only describe mechanical equilibrium states. In order <strong>to</strong> simulate curved interfaces with<br />

a range of curvatures, Matsen anchored the chemical potentials <strong>to</strong> the values at the coex<strong>is</strong>tence<br />

condition for a flat interface in all simulations, but instead stabilized each curved<br />

interface with desired curvature by adding <strong>to</strong> the SCF free energy functional an additional<br />

term whose sole purpose <strong>is</strong> <strong>to</strong> exert a stabilizing force on the interface. The prefac<strong>to</strong>r of th<strong>is</strong><br />

added term <strong>is</strong> treated as a Lagrange multiplier, which can be adjusted so as <strong>to</strong> provide any<br />

required force and thereby allow the interfacial curvature <strong>to</strong> be controlled independently of<br />

the pressure difference across the interface. Matsen required <strong>that</strong> the macroscopic chemical<br />

potentials µ A , µ B , and µ C always be chosen so as <strong>to</strong> sat<strong>is</strong>fy the condition P I = P II for<br />

macroscopic phase coex<strong>is</strong>tence. As a result, the magnitude of the Lagrange multiplier <strong>that</strong><br />

he introduced <strong>to</strong> res<strong>to</strong>re mechanical stability must van<strong>is</strong>h in the limit of a flat interface,<br />

but must become nonzero for any other curvature. The form chosen by Matsen for the<br />

58


0.35<br />

0.3<br />

α = 1<br />

0.5<br />

0.25<br />

0.25<br />

0.2<br />

[κ + ]<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 5 10 15 20 25 30 35 40 45 50<br />

χ N C<br />

Figure 3.4: The non-dimensional bending rigidity [κ + ] of a sphere vs. χN in symmetric<br />

mixtures with various values of α = α A = α B at fixed β = 1.<br />

term added <strong>to</strong> the free energy introduced δ-function spikes in the SCF chemical potential<br />

fields ω A (r) and ω B (r) along the dividing surface along which φ A (r) = φ B (r) in h<strong>is</strong> definition.<br />

The addition of a such a δ-function contribution <strong>to</strong> the SCF chemical potentials must<br />

perturb the structure calculated for any curved interface. Specifically, it must produce a<br />

d<strong>is</strong>continuity in the gradients of the monomer volume fraction fields at the dividing surface.<br />

<strong>Th<strong>is</strong></strong> construction seems conceptually awkward <strong>to</strong> us, since no such ”magic finger” ex<strong>is</strong>ts<br />

<strong>to</strong> stabilize the interface in any physical situation. The procedure used here, in which a<br />

curved interface <strong>is</strong> stabilized by a pressure difference between the neighboring phases, <strong>is</strong><br />

instead more natural for the physical situation in any self-assembled system.<br />

3.4 Symmetric Mixtures<br />

In th<strong>is</strong> section, we d<strong>is</strong>cuss the bending elasticity of saturated monolayers in symmetric<br />

mixture with f A = 1/2 and symmetric homopolymers, with α A = α B = α and b A = b B = b.<br />

Fig. 3.4-3.6 show variation of the dimensionless elastic constants [κ], [κ + ], and [¯κ] as<br />

59


0.35<br />

0.3<br />

0.25<br />

α = 1<br />

0.5<br />

0.25<br />

0.2<br />

[κ]<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 5 10 15 20 25 30 35 40 45 50<br />

χ N C<br />

Figure 3.5: The normalized bending rigidity [κ] vs. χN for symmetric systems with several<br />

values of α.<br />

0<br />

−0.05<br />

−0.1<br />

[ − κ ]<br />

−0.15<br />

−0.2<br />

α = 1<br />

0.5<br />

0.25<br />

−0.25<br />

0 5 10 15 20 25 30 35 40 45 50<br />

χ N C<br />

Figure 3.6: The normalized Gaussian rigidity [¯κ] vs. χN for symmetric mixtures with<br />

several values of α.<br />

60


− [ − κ ] / [κ]<br />

1<br />

α = 1<br />

0.5<br />

0.8<br />

0.25<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 5 10 15 20 25 30 35 40 45 50<br />

χ N C<br />

Figure 3.7: The ratio of the Gaussian rigidity <strong>to</strong> bending rigidity −[¯κ]/[κ] vs. χN for<br />

symmetric mixtures.<br />

functions of χN, for several values of α. Here, we show data only for values of α ≤ 1,<br />

which corresponds <strong>to</strong> the range in which Broseta and Fredrickson find a stable Lifshitz<br />

point. Within th<strong>is</strong> range, [κ], [κ + ] and [¯κ], all extrapolate <strong>to</strong> zero at the value of (χN) L<br />

given in Eqn. 3.12, and appear <strong>to</strong> vary linearly with χN near the Lifshitz point. Note <strong>that</strong><br />

the Gaussian rigidity ¯κ <strong>is</strong> negative for all χN > (χN) L , as required for a flat membrane,<br />

or a lamellar phase, <strong>to</strong> be stable against the formation of a bicontinuous minimal surface<br />

structure.<br />

Fig. 3.7 shows the ratio −[¯κ]/[κ] for saturated monolayers with α ≤ 1. In the strong<br />

stretching limit, Wang and Safran predict a value of −[¯κ]/[κ] = 4<br />

15<br />

= 0.266 for the bending<br />

rigidity of a monolayer in the dry brush (melt state) limit [3, 4], and −[¯κ]/[κ] = 17<br />

30 = 0.5667<br />

in the limit of a strongly stretched but fully solvated brush with a parabolic concentration<br />

profile. Our numerical results for moderately segregated monolayers appear <strong>to</strong> be qualitatively<br />

cons<strong>is</strong>tent with these strong stretching predictions: For χN C > 20, we obtain values<br />

in the range 0.2 < −[¯κ]/[κ] < 0.6, with higher values of −[¯κ]/[κ] in systems with lower<br />

61


values of α, for which the monolayer <strong>is</strong> more swollen. For all α < 1 and χN C > 20, the<br />

ratio −[¯κ]/[κ] also sat<strong>is</strong>fies the condition −[¯κ]/[κ] < 10/9, which has been proposed [17, 18]<br />

as a criterion for systems in which, upon swelling, a lamellar phase of balanced membranes<br />

will ultimately melt in<strong>to</strong> a bicontinuous microemulsion rather than a d<strong>is</strong>persion of spheres.<br />

3.5 Asymmetric Mixtures<br />

In th<strong>is</strong> subsection we switch our attention <strong>to</strong> asymmetric blends, in which α A ≠ α B ,<br />

f A ≠ 1/2, and/or a A ≠ b B . Note <strong>that</strong> we redefine α B ≡ N B /(2f B N C ) because we fix only<br />

the B block length of the copolymer rather than <strong>to</strong>tal copolymer length.<br />

Figs. 3.8 and 3.9 show how the elastic parameters τ and κ vary with f A by changing<br />

only the A block length of the copolymer in a series of systems with N A = N B and<br />

f B N C = N B , as in Figs. 3.2 and 3.3. A linear dependence between τ and f A holds over<br />

approximately the same range of f A 0.58 as <strong>that</strong> for which the spontaneous curvature<br />

was found <strong>to</strong> lie within the range of validity of the Helfrich expansion.<br />

Figs. 3.10 and 3.11 show the the balance point f bal<br />

A<br />

at which τ = 0 a function of the<br />

ratio β = α A /α B for several different sets of systems with b A = b B . Fig. 3.10 shows the<br />

results for a set of systems for which the size of the B homopolymer and the B block of<br />

the copolymer are equal, with several values of χN B = f B N C χ, while we <strong>have</strong> varied N A<br />

<strong>to</strong> vary β and the size of the A block of the copolymer <strong>to</strong> vary f A . Fig. 3.11 presents<br />

f bal<br />

A<br />

vs. β for several values of the ratio f BN C /N B for systems in which the B block of<br />

the copolymer has a fixed size f B χN C = 10. In both cases, increasing β by increasing N A<br />

leads <strong>to</strong> an increase in f bal<br />

A . <strong>Th<strong>is</strong></strong> <strong>is</strong> because increasing N A decreases the tendency of the A<br />

homopolymer <strong>to</strong> swell the A brush, which must be compensated by an increase in the size<br />

of the A block. The sensitivity of f bal<br />

A<br />

<strong>to</strong> changes in β decreases as the brush <strong>is</strong> made drier<br />

by increasing either χN C or α B , In the limit of a completely dry brush, we would expect<br />

fA bal <strong>to</strong> approach 1 2 , independent of β. 62


0.1<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

τ<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.5 0.6 0.7 0.8<br />

f A<br />

Figure 3.8: Bending parameter τ of block copolymer monolayer vs. a block ratio f A of<br />

the A block for systems with f B χN = α A χN = α B χN = 10. The system in th<strong>is</strong> plot<br />

<strong>is</strong> identical <strong>to</strong> <strong>that</strong> in Fig. 3.2 and 3.3. fA<br />

bal <strong>is</strong> always approximated as a limiting value<br />

linearly extrapolated <strong>to</strong> C = 0 as shown here.<br />

2.4<br />

2.2<br />

2<br />

κ +<br />

1.8<br />

1.6<br />

1.4<br />

0.5 0.6 0.7 0.8<br />

f A<br />

Figure 3.9: Bending rigidity κ + of sphere vs. a block ratio f A of the A block for systems<br />

with f B χN = α A χN = α B χN = 10. In th<strong>is</strong> article, κ + <strong>is</strong> always approximated as a<br />

limiting value linearly extrapolated <strong>to</strong> C = 0 as shown here.<br />

63


0.6<br />

α B = 0.5<br />

0.55<br />

f A<br />

bal<br />

0.5<br />

0.45<br />

0.4<br />

f L<br />

f B χ N C = 5<br />

10<br />

20<br />

0.35<br />

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

β<br />

Figure 3.10: The variation of the balance point fA<br />

bal with fixed α = 1. The solid line<br />

indicates the balance point f L at an <strong>is</strong>otropic Lifshitz point.<br />

0.6<br />

f B χ N C = 10<br />

0.55<br />

0.5<br />

f A<br />

bal<br />

0.45<br />

0.4<br />

f L<br />

α B = 0.25<br />

0.5<br />

1<br />

0.35<br />

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

β<br />

Figure 3.11: The variation of the balance point fA<br />

bal for asymmetric mixtures with a fixed<br />

minority block size f B χN C = 10. The solid line indicates the balance point f L at an<br />

<strong>is</strong>otropic Lifshitz point.<br />

64


0.6<br />

0.55<br />

0.5<br />

f A<br />

bal<br />

0.45<br />

0.4<br />

f L<br />

b A /b B = 1<br />

2<br />

0.35<br />

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

Figure 3.12: The variation of the balance point fA<br />

bal<br />

length b.<br />

β<br />

with asymmetric stat<strong>is</strong>tical segment<br />

Eqn. 3.13 for the value of f A at which the balanced Lifshitz point <strong>is</strong> obtained <strong>is</strong> shown<br />

as a solid line in Figs. 3.10 and 3.11. It appears <strong>to</strong> provide a reasonably accurate approximation<br />

<strong>to</strong> the balance point <strong>that</strong> <strong>is</strong> obtained here by setting τ = 0, which remains useful<br />

over the the entire range of values of χN C and α B explored here.<br />

Fig. 3.12 explores the effect of changes in stat<strong>is</strong>tical segment length upon the balance<br />

point f bal<br />

A for system containing a B block of fixed size f BN C χ = 10 and α B = f B .<br />

Increasing b A /b B by a fac<strong>to</strong>r of 2 <strong>is</strong> found <strong>to</strong> cause a rather slight increase in f bal<br />

A . The<br />

sign of the change may be unders<strong>to</strong>od by noting <strong>that</strong> an increase in b A decreases the free<br />

energy cost of stretching the A block of the copolymer, by decreasing the spring constant<br />

kT/b 2 A<br />

in the Gaussian stretching energy. Thus it causes a tendency for the monolayer <strong>to</strong><br />

curve around the A block <strong>that</strong> may be compensated by increasing f A .<br />

At last, results for κ, ¯κ, and κ + are shown in Fig. 3.13 as a function of β for a series of<br />

asymmetric systems with a fixed size of the B block f B χN = 10, α B = 0.5, and b A = b B .<br />

Note <strong>that</strong> all three rigidities seem <strong>to</strong> extrapolate <strong>to</strong> zero at a value of β slightly below the<br />

65


3<br />

2<br />

κ, κ + , − κ<br />

1<br />

0<br />

κ<br />

κ +<br />

− κ<br />

−1<br />

−2<br />

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

β<br />

Figure 3.13: Rigidities κ + , κ, and ¯κ vs. β = α B /α A for systems with a copolymer minority<br />

block of size f B χN = 10, a B homopolymer of length α B χN = 5, and b A = b B , for varying<br />

size A homopolymers. Values for κ, κ + , and ¯κ are given in units with kT = 1 on the left<br />

scale, and corresponding values [κ], [κ + ], and [¯κ] on the right.<br />

lower limit of the range shown in th<strong>is</strong> plot (beyond which the simulations become difficult<br />

<strong>to</strong> converge), indicating the ex<strong>is</strong>tence of the ’balanced’ Lifshitz point at th<strong>is</strong> value of χN<br />

and its dependence on β.<br />

3.6 Conclusions<br />

We <strong>have</strong> used numerical SCFT <strong>to</strong> investigate the bending elasticity of block copolymer<br />

monolayers in melts of imm<strong>is</strong>cible homopolymers. The method used <strong>to</strong> calculate the free<br />

energy of an interface as a function of an imposed curvature <strong>is</strong> somewhat different than<br />

<strong>that</strong> employed previously by Matsen. In both methods, a force must be exerted upon an<br />

interface in order <strong>to</strong> impose an arbitrary curvature. Here, the required force <strong>is</strong> provided<br />

by a pressure difference.<br />

The Helfrich expression for the free energy of a curved interface <strong>is</strong> a Taylor expansion<br />

about a flat reference state. <strong>Th<strong>is</strong></strong> expansion <strong>is</strong> useful for describing equilibrium structures<br />

66


only if it remains accurate at curvature comparable <strong>to</strong> the spontaneous curvature. The<br />

Helfrich expansion for a symmetric monolayer <strong>is</strong> found <strong>to</strong> be accurate over a surpr<strong>is</strong>ingly<br />

wide range of imposed curvatures, extending almost <strong>to</strong> the curvature of a ”dry” micelle,<br />

with no homopolymer in its core. However, th<strong>is</strong> expansion appears <strong>to</strong> be useful only over a<br />

rather limited range of values of the copolymer block ratio f A because of a rapid variation<br />

of spontaneous curvature with f A . For the moderately segregated systems considered here,<br />

with χN C ≃ 20, the spontaneous radius of curvature of a spherical micelle <strong>is</strong> significantly<br />

greater than the core radius of a dry micelle only for 0.5 f A 0.6 (for micelles in an A<br />

matrix), which limits the potential usefulness of the expansion <strong>to</strong> roughly the same range<br />

of values.<br />

The three elastic constants κ, ¯κ, and τ were shown <strong>to</strong> all continuously approach zero at<br />

the analytically predicted Lifshitz point for symmetric systems with α ≤ 1. The Gaussian<br />

rigidity for moderately segregated symmetric systems was found <strong>to</strong> always lie with the<br />

range 0 < −¯κ < (10/9)κ (with a κ > 0) in which the lamellar phase <strong>is</strong> mechanically<br />

stable but <strong>is</strong> expected <strong>to</strong> be susceptible <strong>to</strong> melting in<strong>to</strong> a bicontinuous microemulsion upon<br />

swelling, as the result of effects of interfacial fluctuation <strong>that</strong> are not captured by SCFT.<br />

The value f bal<br />

A<br />

of the copolymer composition at which the spontaneous curvature van<strong>is</strong>hes<br />

has been determined for a variety of systems with asymmetric homopolymers. Knowledge<br />

of th<strong>is</strong> ’balance point’ <strong>is</strong> potentially useful for the design of optimal polymeric surfactants.<br />

The sensitivity of f bal<br />

A<br />

<strong>to</strong> changes in the ratio β decreases as the brush becomes drier<br />

either because of an increase in χN or an increase in α B . The analytical approximation<br />

for f bal<br />

A<br />

proposed by Fredrickson and Bates, which was obtained from an analys<strong>is</strong> of the<br />

critical mode at the Lifshitz point, <strong>is</strong> found <strong>to</strong> provide a reasonably accurate estimate even<br />

far from the Lifshitz point.<br />

67


Bibliography<br />

[1] Canham, P. B. J. Theoret. Biol. 1970, 26, 61<br />

[2] Helfrich, W. Z. Naturforsch. C 1973, 28, 693<br />

[3] Wang, Z.-G.; Safran, S. A. J. Phys. France. 1990, 51, 185<br />

[4] Wang, Z.-G.; Safran, S. A. J. Chem. Phys. 1991, 94, 679<br />

[5] De Gennes, P. G.; Taupin, C. J. Phys. Chem. 1982 86, 2294<br />

[6] Schulman, J. H.; Montagne, J. B. Ann. N.Y. Acad. Sci. 1961, 92, 366<br />

[7] Strey, R. Coll. Ploym. Sci. 1994, 272, 1005<br />

[8] Sottmann, T.; Strey, R. J. Chem. Phys. 1997, 106, 8606<br />

[9] Bates. F. S.; Maurer, W.; Lipic, P. M.; Hillmyer, M. A.; Almdal, K.; Mortensen, K.;<br />

Fredrickson, G. H.; Lodge, T. P. Phys. Rev. Lett 1997, 79, 849<br />

[10] Washburn, N. R.; Lodge, T. P.; Bates. F. S. J. Phys. Chem. B 2000, 104, 6987<br />

[11] Hillmyer, M. A.; Maurer, W.; Lodge, T. P.; Bates. F. S.; Almdal, K. J. Phys. Chem.<br />

B 1999, 103, 4814<br />

[12] Jeon, H. S.; Lee, J. H.; Newstein, M. C. Macromolecules 1998, 31, 3340; 34, 6557<br />

[13] Lee, J. H.; Ruegg, M. L.; Balsara, N. P.; Zhu. Y.; Gido, S. P.; Kr<strong>is</strong>hnamoorti, R.;<br />

Kim, M.-H. Macromolecules 2003, 36, 6537<br />

68


[14] Matsen, M. W. J. Chem. Phys. 1999, 110, 4659<br />

[15] Broseta, D.; Fredrickson, G. H. J. Chem. Phys. 1990, 93, 2927<br />

[16] Fredrickson, G. H.; Bates, F. S. J. Polym. Sci.: Part B: Polym. Phys. 1997, 35, 2775<br />

[17] Morse, D. C. Phys. Rev. E 1994, 50, R2423<br />

[18] Golubovic, L. Phys. Rev. E 1994, 50, R2419<br />

69


Chapter 4<br />

Swollen Micelles and Interfacial<br />

Tension<br />

Abstract<br />

We combined self-cons<strong>is</strong>tent field theory (SCFT) and the Helfrich theory of interfacial<br />

bending elasticity <strong>to</strong> analyze the thermodynamics of a ternary two-phase systems of two<br />

imm<strong>is</strong>cible homopolymers A and B and a small amount of diblock copolymer AB. We<br />

focus on the effect of micellization upon the interfacial tension in a two-phase state in<br />

which copolymer micelles form in the A rich phase. SCFT has been utilized <strong>to</strong> calculate the<br />

critical micelle concentrations (cmc’s) and macroscopic interfacial tensions for the interface<br />

between A- and B-rich phases. The behavior of systems of nearly balanced copolymers,<br />

which tend <strong>to</strong> form highly swollen micelles, <strong>is</strong> also d<strong>is</strong>cussed within the context of the<br />

Helfrich theory, using elastic constants obtained from SCFT simulations of weakly curved<br />

monolayers.<br />

4.1 Introduction<br />

There <strong>have</strong> been numerous efforts <strong>to</strong> emulsify or compatibilize imm<strong>is</strong>cible homopolymer<br />

blends by adding small amounts of block copolymer. Added block copolymers adsorb at<br />

interface between homopolymer phases, and stabilize the d<strong>is</strong>persion of minor phases by<br />

70


educing interfacial tension, which promotes droplet break-up, and by suppressing droplet<br />

coalescence.[1, 2] Once the concentration of copolymer d<strong>is</strong>solved in either phase exceeds<br />

the critical micelle concentration (cmc), additional block copolymer instead form micelles,<br />

preventing further increase in the interfacial coverage of copolymer at the interfaces and<br />

any further decrease in the interfacial tension.<br />

Although interfacial adsorption and micellization are always competitive in the ternary<br />

polymer blends, most theoretical studies of these phenomena <strong>have</strong> focused either on interfacial<br />

adsorption, while ignoring the possibility of micellization [3, 4], or on predictions<br />

of the cmc in binary mixtures of homopolymer and copolymer.[5, 6, 7, 8, 9, 10] Leibler<br />

has studied the saturation of interfacial tension γ above the cmc in a two-phase ternary<br />

mixture [11], by equating the chemical potential at the cmc with <strong>that</strong> of the copolymer adsorbed<br />

<strong>to</strong> a macroscopic interface, while using a strong stretching theory <strong>to</strong> describe both<br />

the micelle and the adsorbed monolayer. Leibler concluded <strong>that</strong> a van<strong>is</strong>hing macroscopic<br />

interfacial tension, γ = 0, should be obtained below the cmc for systems with a rather<br />

wide range of values of block copolymer compositions f A . However, Leibler calculated<br />

the free energy of formation of a micelle using a model appropriate <strong>to</strong> a binary system<br />

of copolymer in a homopolymer matrix, and thus neglected the fact <strong>that</strong> a micelle in an<br />

A-rich phase <strong>that</strong> coex<strong>is</strong>ts with an excess phase of the nearly pure B homopolymer may<br />

swell by emulsifying the B homopolymer within its core. Allowing for the formation of<br />

swollen micelles (which are sometimes referred <strong>to</strong> as microemulsion droplets [12, 13]) will<br />

generally reduce the calculated free energy of formation of a micelle, thereby reducing the<br />

predicted value of the copolymer chemical potential at the cmc, and ra<strong>is</strong>ing the predicted<br />

saturation value of γ above the cmc.<br />

In th<strong>is</strong> Chapter we numerically solve SCFT <strong>to</strong> examine the effect of the formation of<br />

spherical micelles on macroscopic interfacial tension within a ternary two-phase mixture,<br />

while allowing the formation of swollen micelles. In contrast <strong>to</strong> Leibler, we find <strong>that</strong> micelles<br />

71


always form at the concentration of a copolymer below <strong>that</strong> needed <strong>to</strong> drive the interfacial<br />

tension <strong>to</strong> zero, yielding a nonzero interfacial tension above the cmc, except in the limit of<br />

a perfectly balanced surfactant monolayer with van<strong>is</strong>hing spontaneous curvature.<br />

4.2 Micellization - Theory<br />

We consider a ternary system of two imm<strong>is</strong>cible homopolymers, A and B, and an AB<br />

diblock copolymer (denoted C), under condition for which system phase separates in<strong>to</strong><br />

two phases rich in A and B, respectively, with a small concentration of copolymer C.<br />

Let N A , N B , and N C denote the degrees of polymerization of the two homopolymers<br />

and copolymer, respectively, and f A denote the volume fraction of A monomers within the<br />

copolymer. Let α A ≡ N A /N C , α B ≡ N B /N C , and β ≡ N A /N B . For specificity, we consider<br />

the case in which, above the critical micelle concentration, spherical micelles preferentially<br />

form in the A-rich phase. In the case of symmetric homopolymers with β = 1 and equal<br />

stat<strong>is</strong>tical segment lengths, th<strong>is</strong> corresponds <strong>to</strong> the case f A > 1/2.<br />

4.2.1 Phenomena<br />

The dilute region of the ternary phase diagram for such a system <strong>is</strong> shown schematically<br />

in Fig. 4.1. The copolymer <strong>is</strong> assumed <strong>to</strong> form micelles in the A-rich phase above cmc line,<br />

shown by a dotted line c-b, which extends from the cmc of the binary A-C system (point c)<br />

<strong>to</strong> the cmc of an A-rich phase <strong>that</strong> coex<strong>is</strong>ts with a B-rich phase (point b). The dash-dotted<br />

tie line <strong>that</strong> intersects the phase boundary at point b divides the two-phase region in<strong>to</strong> a<br />

lower region of very low copolymer concentration in which neither of the two coex<strong>is</strong>ting<br />

phases contains micelles, and an upper region of higher copolymer concentration in which<br />

a micellar A-rich phase coex<strong>is</strong>ts with a micelle-free B-rich phase. Below th<strong>is</strong> cmc tie-line,<br />

the boundaries of the A and B-rich phases are reasonably well described by Flory-Huggins<br />

theory. Above the cmc, the limit of solubility of B within the micellar phase becomes<br />

72


Figure 4.1: Schematic phase diagram of a ternary system compr<strong>is</strong>ed of homopolymers<br />

A and B and diblock copolymer C, in which an A rich phase of swollen micelles (or,<br />

equivalently, microemulsion droplets) coex<strong>is</strong>ts with a phase of nearly pure B. The dotted<br />

line b-c <strong>is</strong> a line of the critical micelle concentration in a single A-rich phase. Micelles are<br />

drawn above th<strong>is</strong> line. A point b corresponds <strong>to</strong> the cmc in two phase ternary system.<br />

73


higher than <strong>that</strong> predicted by Flory-Huggins theory as a result of the solubilization of the<br />

B homopolymer within the cores of swollen micelles. Also, the upper long-dashed two<br />

phase boundary (the emulsification failure line) <strong>is</strong> almost straight because a micelle radius,<br />

or the ratio of a core volume <strong>to</strong> an interfacial area, <strong>is</strong> nearly constant due <strong>to</strong> the fixed<br />

chemical potentials until the emulsification failure <strong>to</strong> exceed the limit of solubility of the<br />

B homopolymer in the core occurs.<br />

The chemical potential µ C of copolymer within the A-rich phase <strong>is</strong> related by an ideal<br />

solution law,<br />

µ C = µ 0 C + kT ln(ρf C /ρ0 C ) , (4.1)<br />

<strong>to</strong> the number concentration ρ f C<br />

of ”free” copolymers d<strong>is</strong>solved as single molecules (outside<br />

of any micelles) within the A-rich phase. Here, ρ 0 C<br />

<strong>is</strong> a standard state concentration and<br />

µ 0 C<br />

<strong>is</strong> a corresponding standard state chemical potential for copolymer d<strong>is</strong>solved in A. A<br />

similar relation ex<strong>is</strong>ts between the chemical potential µ B of the B homopolymer and the<br />

concentration of B d<strong>is</strong>solved within the A matrix. Above the cmc line, the concentrations<br />

and the corresponding chemical potentials of free B and C molecules within the matrix of<br />

a two-phase system always remain very close <strong>to</strong> those obtained at the cmc.<br />

The resulting saturation of ρ f C (or the volume fraction φf C ) and µ C <strong>to</strong> nearly constant<br />

values at concentrations above the cmc are shown schematically in Fig. 4.2 and Fig. 4.3.<br />

In a two-phase system, micelles generally form in only one phase (here, the A-rich phase)<br />

because the saturation of µ C caused by micellization in the A-rich phase prevents µ C from<br />

ever reaching the higher value required for micellization in the B-rich phase, unless the<br />

values of µ cmc<br />

C<br />

in the two phases are extremely similar.<br />

Interfacial tension γ of a macroscopic interface between coex<strong>is</strong>ting A- and B-rich phases<br />

mono<strong>to</strong>nically decreases with increasing µ C while the interfacial coverage of adsorbed<br />

copolymer increases. As shown in Fig. 4.4, interfacial tension extrapolates <strong>to</strong> zero at a<br />

74


φ c<br />

m<br />

φ c<br />

(f,m)<br />

φ c<br />

f<br />

φ c<br />

Figure 4.2: Dependence of concentrations of free copolymer and of copolymer within micelles<br />

upon the <strong>to</strong>tal concentration of copolymer within the micellar phase, along the<br />

coex<strong>is</strong>tence line.<br />

µ c - µ c<br />

o<br />

φ c<br />

Figure 4.3: Dependence of copolymer chemical potential upon <strong>to</strong>tal concentration of<br />

copolymer within the micellar phase, along the coex<strong>is</strong>tence line. The dotted line shows the<br />

continuation of ideal solution behavior beyond the cmc<br />

75


1<br />

γ / γ o<br />

0<br />

φ c<br />

cmc<br />

φ<br />

*<br />

c<br />

φ c<br />

Figure 4.4: Schematic of the ratio γ/γ 0 of macroscopic interfacial tension γ <strong>to</strong> its value γ 0<br />

in the absence of copolymer, vs. the <strong>to</strong>tal volume fraction of copolymer within the A-rich<br />

micellar phase. As in Fig. 4.3, the solid line shows the behavior of a micellar solution,<br />

and the dotted line shows the continuation of ideal solution behavior, in the absence of<br />

micelles.<br />

saturation chemical potential µ ∗ C or a corresponding concentration ρ∗ C<br />

of free d<strong>is</strong>solved<br />

copolymer. If µ cmc<br />

C<br />

<strong>is</strong> less than the saturation chemical potential µ ∗ C<br />

, then γ will decrease<br />

only until the cmc <strong>is</strong> reached, and thereafter remain very close <strong>to</strong> the nonzero value<br />

γ cmc obtained at the cmc. If µ cmc<br />

C<br />

were somehow <strong>to</strong> exceed µ ∗ C<br />

, however, then it would<br />

be possible <strong>to</strong> drive the interfacial tension <strong>to</strong> zero at a copolymer concentration below the<br />

cmc.<br />

Leibler estimated µ ∗ C<br />

and µcmc<br />

C<br />

from the free energies of flat interfaces and spherical<br />

micelles using a simple strong-stretching calculation. The most important limitation of<br />

h<strong>is</strong> calculation, as noted in the introduction, <strong>is</strong> <strong>that</strong> he did not allow for the possibility<br />

of swelling of the micelle cores. Leibler found µ cmc<br />

C<br />

< µ ∗ C<br />

for sufficiently asymmetric<br />

copolymers, with |f A − 0.5| > 0.19, but µ ∗ C < µcmc C<br />

for more symmetric copolymers, with<br />

0.31 < f A < 0.69. These calculation suggested <strong>that</strong> it might be possible <strong>to</strong> produce<br />

76


macroscopic interfaces with γ = 0 using block copolymers with a rather wide range of<br />

values of f A . The nearly symmetric copolymers for which Leibler found µ ∗ C < µcmc C<br />

also,<br />

however, prefer <strong>to</strong> form monolayers with large radii of curvature, and thus form highly<br />

swollen micelles. Any theory <strong>that</strong> prohibits swelling of such micelles will overestimate the<br />

free energy of formation of a micelle in a ternary system, and thereby overestimate µ cmc<br />

C .<br />

We show in what follows <strong>that</strong>, if the micelles of nearly symmetric copolymers are allowed<br />

<strong>to</strong> swell <strong>to</strong> their preferred size by emulsifying the B homopolymer, µ cmc<br />

C<br />

always remains<br />

less than or equal <strong>to</strong> µ ∗ C<br />

, causing the macroscopic interfacial tension <strong>to</strong> always saturate <strong>to</strong><br />

a non-negative value.<br />

4.2.2 Thermodynamics<br />

It <strong>is</strong> convenient <strong>to</strong> formulate both the classical thermodynamics of th<strong>is</strong> system and the<br />

corresponding SCFT calculations in grand canonical ensemble. The grand canonical free<br />

energy Φ of any system <strong>is</strong> given by the difference<br />

Φ = F − ∑ i<br />

µ i M i (4.2)<br />

where F <strong>is</strong> the Helmholtz canonical free energy, and µ i <strong>is</strong> the chemical potential and M i<br />

the <strong>to</strong>tal number of molecules of species i in the system of interest, for i = A,B,C. For<br />

any macroscopic system of volume V and pressure P, Φ = −PV .<br />

The grand-canonical free energy of a dilute solution of micelles at a specified set of<br />

chemical potentials and a specified number density ρ m of nearly monod<strong>is</strong>perse micelles<br />

may be expressed as a function<br />

Φ<br />

V = −P h (T,µ) + ρ m<br />

[<br />

Φ m (T,µ) + kT ln(ρ m /ρ 0 m e)] (4.3)<br />

where P h <strong>is</strong> the pressure of a hypothetical homogeneous (i.e., micelle-free) phase at the<br />

specified set of chemical potentials and Φ m (T,µ) <strong>is</strong> the extrapolated excess grand-canonical<br />

free energy per micelle of a hypothetical micellar solution with a standard state micelle<br />

77


concentration ρ 0 m . Here and hereafter, we use µ, with no subscripts, <strong>to</strong> refer <strong>to</strong> a set of<br />

chemical potentials µ ≡ {µ A ,µ B ,µ C }. The true equilibrium grand-canonical free energy<br />

at a specified T and µ <strong>is</strong> obtained by minimizing Eqn. 4.3 with respect <strong>to</strong> variations in the<br />

number density ρ m of micelles yielding<br />

ρ m = ρ 0 m exp [−Φ m (µ,T)/kT] . (4.4)<br />

Substituting th<strong>is</strong> back in<strong>to</strong> Eqn. 4.3 produces a macroscopic pressure P ≡ −Φ/V = P h +<br />

ρ m kT in which the micelles simply contribute an ideal-solution osmotic pressure of ρ m kT.<br />

In Eqn. 4.3, we <strong>have</strong> not attempted <strong>to</strong> account separately for different ”species” of<br />

micelles containing different numbers of molecules of C copolymer or solubilized B homopolymer.<br />

The free energy Φ m (T,µ) should thus be unders<strong>to</strong>od <strong>to</strong> be an excess grandcanonical<br />

free energy per a micelle of fluctuating size and composition. The numbers of<br />

copolymer and homopolymer molecules within each micelle can fluctuate by exchanging<br />

molecules between the micelle and the surrounding matrix. Correspondingly, ρ m <strong>is</strong> the<br />

number density of micelles of any size or composition. <strong>Th<strong>is</strong></strong> formulation assumes only <strong>that</strong><br />

it would be possible <strong>to</strong> accurately count the <strong>to</strong>tal number of micelles in any microscopic<br />

state of a micellar solution, but does not require us <strong>to</strong> exactly define the d<strong>is</strong>tribution of<br />

molecules in a system by deciding how many molecules of each type reside ”within” each<br />

micelle, and how many are part of the surrounding matrix.<br />

Let the excess aggregation number of species i within a micelle, denoted M ex<br />

i , be<br />

defined <strong>to</strong> be difference between the average number of molecules of species i in a system<br />

(or simulation) <strong>that</strong> <strong>is</strong> constrained <strong>to</strong> contain one micelle and <strong>that</strong> in a system of equal<br />

volume and equal chemical potentials with no micelles. These excess aggregation numbers<br />

are related <strong>to</strong> the corresponding excess free energy Φ m by the thermodynamic identity<br />

M i = − ∂Φm (µ,T)<br />

∂µ i<br />

(4.5)<br />

by dropping the superscript ex out for brevity. In an incompressible fluid, the excess<br />

78


aggregation numbers for the three species are related by the fact <strong>that</strong> the <strong>to</strong>tal number<br />

of monomers are conserved regardless of the introduction of a micelle so <strong>that</strong> they must<br />

sat<strong>is</strong>fy 0 = ∑ i N iM i for i = A,B,C. <strong>Th<strong>is</strong></strong> definition thus yields positive values for the<br />

copolymer excess M C , and (in an A-rich phase) for the excess M B of solubilized B, but<br />

negative values for the ”excess” M A of matrix homopolymer.<br />

Fluctuations in the excess number of molecules of each type within a polyd<strong>is</strong>perse<br />

ensemble of micelles can be related <strong>to</strong> the susceptibility of the average aggregation numbers<br />

<strong>to</strong> changes in the chemical potential by the stat<strong>is</strong>tical mechanical identity<br />

〈δM i δM j 〉 = kT ∂M i<br />

∂µ j<br />

= −kT ∂2 Φ m (µ)<br />

∂µ i ∂µ j<br />

(4.6)<br />

in which δM i <strong>is</strong> the deviation of the ”actual” excess number of molecules of type i in a<br />

micelle from its average value M i . <strong>Th<strong>is</strong></strong> <strong>is</strong> a special case of the equilibrium fluctuationresponse<br />

theorem.<br />

The interfacial tension γ of a flat macroscopic interface between A- and B-rich phases <strong>is</strong><br />

also most conveniently expressed in grand-canonical ensemble, as an excess grand canonical<br />

free energy per unit area<br />

γ = f int − ∑ i<br />

µ i Γ int<br />

i (4.7)<br />

where f int <strong>is</strong> excess interfacial Helmholtz free energy per unit area, and Γ int<br />

i<br />

<strong>is</strong> the interfacial<br />

excess number density of molecules of type i. Interfacial excess quantities, such as f int<br />

and Γ i must be defined using some choice of a Gibbs dividing surface. For a flat interface,<br />

however, the interfacial tension γ <strong>is</strong> independent of one’s choice of convention for the<br />

position of th<strong>is</strong> surface, because the bulk grand-canonical free energy per unit volume −P<br />

must be the same for two bulk phases separated by a flat interface.<br />

4.2.3 Critical Micelle Concentration<br />

A useful approximation for the critical micelle concentration may be obtained by simply<br />

neglecting the translational entropy of the micelles, and thus removing the logarithmic term<br />

79


in Eqn. 4.3. In th<strong>is</strong> simplified expression, the formation of micelles <strong>is</strong> favorable whenever<br />

Φ m (T,µ) ≤ 0. The cmc line (b − c) in Fig. 4.1 thus corresponds in th<strong>is</strong> approximation <strong>to</strong><br />

the set of chemical potentials for which<br />

0 = Φ m (T,µ) . (4.8)<br />

In th<strong>is</strong> approximation, the thermodynamic analys<strong>is</strong> of micellization becomes mathematically<br />

identical <strong>to</strong> the analys<strong>is</strong> of the limit of solubility of a second phase (e.g., a precipitate)<br />

in a dilute solution.<br />

When the translational entropy of the micelles <strong>is</strong> taken in<strong>to</strong> account, the critical micelle<br />

concentration ceases <strong>to</strong> be a sharp phase transition. Instead, in the limit of sufficiently<br />

large micelles, it marks the position of a rapid but continuous crossover, which approaches<br />

the behavior of a true phase transition. For finite micelles, the definition of the cmc<br />

<strong>is</strong> necessarily somewhat arbitrary. We choose <strong>to</strong> define the cmc of copolymer <strong>to</strong> be the<br />

concentration of free copolymers at the chemical potential µ cmc<br />

C<br />

of unimers <strong>is</strong> equal <strong>to</strong> <strong>that</strong> of copolymers within micelles, i.e., for which<br />

for which the concentration<br />

ρ cmc<br />

C = ρ f C = ρ mM C . (4.9)<br />

By substituting th<strong>is</strong> definition in<strong>to</strong> Eqn. 4.4, we find <strong>that</strong><br />

0 = Φ m + kT ln(ρ cmc<br />

C /ρ0 m M C) (4.10)<br />

at the cmc, where Φ m (T,µ) <strong>is</strong> evaluated at the chemical potential for which ρ f C = ρcmc C .<br />

By using Eqn. 4.5 for the derivative of Φ m with respect <strong>to</strong> µ C , one may easily show<br />

<strong>that</strong> the value of µ C at which Eqn. 4.10 <strong>is</strong> sat<strong>is</strong>fied differs from <strong>that</strong> at which the simplified<br />

condition of Eqn. 4.8 <strong>is</strong> sat<strong>is</strong>fied by an amount δµ C ≃ kT ln(ρ cmc<br />

C<br />

/ρ0 m M C)/M C . The<br />

corresponding shift in ρ cmc<br />

C<br />

, which ar<strong>is</strong>es from our neglect of the translational entropy of a<br />

micelle in Eqn. 4.3, <strong>is</strong> small in the limit M C ≫ 1 of a large micelle, and corresponds <strong>to</strong> a<br />

very small fractional change in ρ cmc<br />

C<br />

under all circumstances of interest here.<br />

80


4.2.4 Interfacial Model<br />

We <strong>have</strong> emphasized above <strong>that</strong> micelles in a ternary system generally swell by emulsifying<br />

the B homopolymer within the core of the micelle. In a two-phase system coex<strong>is</strong>ting with<br />

a reservoir of bulk B, nearly symmetric copolymers can form highly swollen micelles (or,<br />

equivalently, microemulsion droplets). A core of each micelle <strong>is</strong> rich in the B homopolymer<br />

and has a composition similar <strong>to</strong> <strong>that</strong> of the coex<strong>is</strong>ting B-rich phase surrounded by a<br />

weakly-curved monolayer with a structure also similar <strong>to</strong> <strong>that</strong> of a macroscopic saturated<br />

interface. <strong>Th<strong>is</strong></strong> suggests a view of the free energy of highly swollen micelles as a sum of<br />

bulk and interfacial contribution, in which the latter <strong>is</strong> a function of curvature.<br />

Now let the phase I and II be the region of an A-rich matrix and a B-rich core,<br />

respectively, separated by a Gibbs dividing surface of area A. The excess free energy<br />

Φ m ≡ Φ+P I V of a spherical or cylindrical micelle with a core region of volume V II = V −V I<br />

may be expressed as a sum<br />

Φ m = (P I − P II )V II + Aγ (4.11)<br />

of the excess interfacial free energy Aγ plus an excess bulk free energy of the core, which<br />

ar<strong>is</strong>es from the fact <strong>that</strong> the grand-canonical free energy density −P <strong>is</strong> generally different in<br />

the core and in the matrix. The simplified condition for the cmc, which we use throughout<br />

th<strong>is</strong> section, requires <strong>that</strong> Φ m = 0 at the cmc.<br />

We focus in th<strong>is</strong> chapter on the case of a micellar phase <strong>that</strong> coex<strong>is</strong>ts with a second<br />

excess phase. The condition for mechanical equilibrium between the two macroscopic<br />

phases separated by a curved monolayer with a radius of curvature R <strong>is</strong> given by<br />

P II − P I = Cγ −<br />

C2 ∂γ<br />

d − 1 ∂C . (4.12)<br />

where the mean curvature C = (d−1)/R with d = 3 for a sphere or d = 2 for a cylinder. If<br />

an interface <strong>is</strong> flat, the pressure in the micellar A-rich phase be equal <strong>to</strong> <strong>that</strong> in the B rich<br />

phase. Note <strong>that</strong> the local pressure within the matrix of the micellar phase slightly differs<br />

81


from the macroscopic pressure of the whole system as a result of an osmotic pressure ρ m kT<br />

ar<strong>is</strong>ing from the translational entropy of the micelles, which we ignore here. Because the<br />

chemical potentials in B-rich core of a highly swollen micelle must be the same as those in a<br />

coex<strong>is</strong>ting B-rich macroscopic phase in equilibrium, the local pressure within the core must<br />

equal <strong>to</strong> <strong>that</strong> in the B-rich phase which also equals <strong>to</strong> <strong>that</strong> in matrix, implying P I = P II .<br />

Combining th<strong>is</strong> requirement with Φ m = 0 at the cmc yields a requirement <strong>that</strong><br />

γ = 0 (4.13)<br />

at the cmc of a two-phase system implying interfacial tension within a micelle <strong>is</strong> essentially<br />

zero. Combining the conditions P I = P II and γ = 0 with the mechanical equilibrium<br />

condition yields the additional requirement <strong>that</strong><br />

∂γ<br />

∂C ∣ = 0 , (4.14)<br />

µ<br />

where C <strong>is</strong> the mean curvature and the derivative <strong>is</strong> evaluated with all chemical potentials<br />

held constant. Eqn. 4.14 <strong>is</strong> a statement of the fact <strong>that</strong>, when the micellar phase coex<strong>is</strong>ts<br />

with an excess B-rich phase, the preferred micelle radius <strong>is</strong> <strong>that</strong> which minimizes the excess<br />

free energy density of the interfacial monolayer.<br />

Though we focus in th<strong>is</strong> thes<strong>is</strong> primarily upon two-phase systems, it <strong>is</strong> interesting <strong>to</strong><br />

briefly consider the case of a ternary micellar solution <strong>that</strong> does not coex<strong>is</strong>t with a second<br />

phase (1φ region in Fig. 4.1). Combining Eqn. 4.11 for Φ m with the mechanical equilibrium<br />

condition, which requires <strong>that</strong> the micelle radius minimize Φ m , yields<br />

Φ m = A d<br />

[<br />

γ + C ∂γ ]<br />

∂C<br />

(4.15)<br />

where d = 3 for a sphere or d = 2 for a cylinder. Combining th<strong>is</strong> with the simplified cmc<br />

condition, Φ m = 0, yields the requirement <strong>that</strong><br />

γ = −C ∂γ<br />

∂C<br />

82<br />

(4.16)


at the cmc. Substituting th<strong>is</strong> back in<strong>to</strong> the mechanical equilibrium condition (Eqn. 4.12)<br />

yields a pressure difference<br />

P II − P I = −<br />

d<br />

d − 1<br />

C2<br />

∂γ<br />

∂C =<br />

d Cγ (4.17)<br />

d − 1<br />

everywhere along the cmc line. In a single-phase solution, the preferred micelle radius <strong>is</strong><br />

restricted by a s<strong>to</strong>ichiometric constraint on the amount of available the B homopolymer,<br />

which causes the micelle interface <strong>to</strong> adopt a smaller radius, or higher curvature, than the<br />

spontaneous curvature adopted along the phase-coex<strong>is</strong>tence line. <strong>Th<strong>is</strong></strong> leads <strong>to</strong> a situation<br />

in which ∂γ/∂C > 0, and in which a negative relative pressure P II < P I develops in the<br />

core of the micelle, so as <strong>to</strong> prevent the micelle from expanding <strong>to</strong> the radius corresponding<br />

<strong>to</strong> the spontaneous curvature. Along the cmc line, the micelle radius <strong>is</strong> thus expected <strong>to</strong><br />

steadily decrease, and magnitude of the pressure difference P I −P II <strong>to</strong> steadily increase, as<br />

one moves away from the emulsification failure line and <strong>to</strong>wards the point c of the binary<br />

A-C system as shown in Fig. 4.1.<br />

4.2.5 Helfrich Theory<br />

Combining the CMC and mechanical equilibrium conditions with the Canham-Helfrich<br />

theory,[15] which <strong>is</strong> expanded about zero curvature and δµ = µ C − µ ∗ C<br />

while assuming a<br />

spherical interface, yields the conditions<br />

0 = γ = −τC + 1 2 κ +C 2 − Γ ∗ δµ − Λδµ C C (4.18)<br />

0 = ∂γ<br />

∂C = −τ + κ +C − Λδµ (4.19)<br />

1<br />

in which κ + ≡ κ + 2¯κ <strong>is</strong> the bending rigidity for a spherically curved interface. Eqn. 4.19<br />

yields an equilibrium curvature<br />

C ≃ τ<br />

κ +<br />

(4.20)<br />

83


with corrections of O(τ 2 ). In the same limit, Eqn. 4.18 and Eqn. 4.19 yield a chemical<br />

potential shift<br />

−Γ ∗ δµ = 1 2 κ +C 2 (4.21)<br />

<strong>that</strong> van<strong>is</strong>hes in the limit τ → 0, or C → 0.<br />

The interfacial tension γ flat of a macroscopic flat interface between a micellar A-rich<br />

solution and an excess B-rich phase can be obtained by evaluating Eqn. 4.18 with C = 0<br />

and replacing δµ by Eqn. 4.21. <strong>Th<strong>is</strong></strong> yields a predicted interfacial tension<br />

γ flat = −Γ ∗ δµ = 1 2 κ +C 2 (4.22)<br />

<strong>that</strong> varies quadratically with the spontaneous spherical curvature C ≃ τ/κ + , and <strong>that</strong><br />

van<strong>is</strong>hes in the limit of a balanced monolayer.<br />

It <strong>is</strong> reasonable <strong>to</strong> assume <strong>that</strong> τ linearly depends on any variable <strong>to</strong> affect the spontaneous<br />

curvature near at a balance point. For instance, the nonionic surfactant system has<br />

utilized temperature <strong>to</strong> control the spontaneous curvature of a micelle which linearly varies<br />

close a mean temperature where c o van<strong>is</strong>hes.[14] Analogously, for the polymeric system the<br />

polymer chain length can be used as a control parameter instead of temperature since a<br />

Flory-Huggins interaction parameter χ <strong>is</strong> usually a mono<strong>to</strong>nic function of temperature unlike<br />

the complex interaction between water and surfactant. Assuming a constant bending<br />

rigidity κ + ≃ κ + | C=0 and τ ∝ f A − fA bal , Eqn. 4.22 <strong>is</strong> rewritten as<br />

where a balance point f bal<br />

A<br />

γ flat = (τ ′ ) 2<br />

2κ +<br />

(f A − f bal<br />

A )2 (4.23)<br />

<strong>is</strong> a volume fraction of A block within block copolymer giving<br />

the zero spontaneous curvature and τ ′ = dτ/df A | C=0 . Eqn. 4.23 completely characterizes<br />

γ flat with two parameters: the ”sensitivity” (τ ′ ) 2<br />

κ +<br />

of a parabolic curve <strong>to</strong> a curvature and<br />

a position of a zero minimum which <strong>is</strong> a balance point f bal<br />

A .<br />

84


4.3 SCFT Methodology<br />

We carry out two separate SCFT simulations in grand-canonical ensemble in order <strong>to</strong><br />

calculate the excess free energy Φ m of an <strong>is</strong>olated micelle within an A-rich matrix, and the<br />

corresponding interfacial tension γ of a flat interface at common chemical potentials for<br />

sequentially increasing µ C . To obtain the coex<strong>is</strong>ting conditions appropriate <strong>to</strong> a two-phase<br />

system the values of µ A and µ B are taken, for each value of µ C , <strong>to</strong> follow a line a − b of a<br />

phase boundary in Fig. 4.1 adjusting a pressure P = 0 in both phases. The required values<br />

of µ A and µ B are calculated using Flory-Huggins theory, which <strong>is</strong> the homogeneous limit<br />

of SCFT.<br />

SCFT simulations of a spherical micelle and a macroscopic flat interface are carried<br />

out in a one dimensional unit cell of the corresponding geometry with a finite-difference<br />

d<strong>is</strong>cretization. In the micelle simulations, use of grand-canonical ensemble <strong>is</strong> convenient<br />

because it allows the micelle <strong>to</strong> au<strong>to</strong>matically adjust its size, composition, and structure<br />

so as <strong>to</strong> minimize the relevant free energy: Solutions of the SCF equations for the grandcanonical<br />

ensemble may be shown <strong>to</strong> be extrema of the excess grand free energy Φ m (µ,T)<br />

at fixed chemical potential and temperature.<br />

In the simulation of a flat interface, a mechanically stable interface may be formed<br />

only at a coex<strong>is</strong>tence condition which requires equal pressures in both phases. In grandcanonical<br />

ensemble, however, the position of the interface within the simulation cell <strong>is</strong><br />

indeterminate so <strong>that</strong> small pressure difference can easily cause the instability. To obtain<br />

a unique, stable solution, we impose on the SCFT equations a requirement <strong>to</strong> specify Gibbs<br />

dividing surface for the interface within the unit cell. The combined set of equations are<br />

solved by New<strong>to</strong>n-Raphson iteration (as d<strong>is</strong>cussed in Chapter 2).<br />

The number density ρ m of micelles at each value of µ C may be calculated from Eqn. 4.4,<br />

in which we approximate Φ m by the excess grand-canonical free energy calculated for<br />

85


an <strong>is</strong>olated stationary micelle by SCFT. Values for the critical micelle concentration are<br />

obtained from Eqn. 4.9. The standard-state micelle concentration ρ 0 m <strong>is</strong> taken <strong>to</strong> be the<br />

hypothetical state in which the excess volume fraction of copolymer contained within<br />

the whole micelles extrapolates <strong>to</strong> unity, i.e., a dense unswollen micelle such as ρ 0 m =<br />

1/(V C M C ), where V C <strong>is</strong> the volume occupied by a copolymer chain. Fig. 4.2-Fig. 4.4,<br />

previously illustrated as schematic plots, are actually the results of SCFT calculations for<br />

a system with f A = 0.635, f B χN C = 10, and N A = N B = 100, for which φ cmc<br />

C<br />

= 0.0051.<br />

The crossover at the cmc in the derivatives of µ C and φ f C<br />

with respect <strong>to</strong> <strong>to</strong>tal copolymer<br />

volume fraction at the cmc <strong>is</strong> continuous, but appears quite sharp, which gives a welldefined<br />

apparent cmc at a concentration <strong>that</strong> agrees well with <strong>that</strong> obtained from Eqn. 4.9.<br />

4.4 Results - Micelle Simulations<br />

Fig. 4.5 shows the effect of micelle swelling upon the cmc, by comparing SCFT predictions<br />

for the copolymer volume fraction φ cmc<br />

C<br />

at the cmc of a ternary two-phase system (point<br />

b in Fig. 4.1) and at the cmc of a corresponding binary system (point c in Fig. 4.1) in<br />

which the micelles are necessarily unswollen. Data <strong>is</strong> shown for three different values of<br />

χN CB = f B χN C , where N CB <strong>is</strong> the size of the core block, in systems with homopolymers<br />

of N A = N B = 100. Each line in th<strong>is</strong> figure thus corresponds <strong>to</strong> the systems with a series<br />

of copolymers with the same core block χN CB but the varying corona block length. In<br />

either binary or ternary system, the cmc exhibits a roughly exponential dependence on<br />

χN CB , and a weaker dependence on the corona block length.<br />

It <strong>is</strong> important <strong>to</strong> note <strong>that</strong> the cmc of the ternary two-phase system <strong>is</strong> always lower<br />

than <strong>that</strong> of the corresponding binary system as a result of the freedom of the ternary<br />

system <strong>to</strong> minimize the micelle free energy by emulsifying B so as <strong>to</strong> relax the curvature of<br />

the copolymer layer. The difference between the cmc of the swollen micelles in a ternary<br />

mixture and <strong>that</strong> of unswollen micelles in a binary mixture <strong>is</strong> greatest for symmetric<br />

86


φ c<br />

cmc<br />

10 -1 χ N<br />

10 -2<br />

CB = 10<br />

χ N CB = 12<br />

10 -3<br />

cmc<br />

φ c (binary)<br />

cmc<br />

φ c (ternary)<br />

*<br />

φ c<br />

10 -4<br />

0.5 0.6 0.7 0.8 0.9<br />

f A<br />

χ N CB = 8<br />

Figure 4.5: Copolymer volume fraction φ cmc<br />

C<br />

at the critical micelle concentration of a twophase<br />

ternary system (solid lines), and of a binary mixture of copolymer C in A (short<br />

dashed lines), vs. volume fraction f A of the corona block within the copolymer, for systems<br />

with homopolymer sizes α A χN = α B χN = 10, for three values of χN CB . Also shown <strong>is</strong><br />

the copolymer volume fraction φ ∗ C<br />

(dot dashed line) at which the macroscopic interfacial<br />

tension of the two phase system extrapolates <strong>to</strong> zero. Note <strong>that</strong> φ ∗ C always exceeds the<br />

cmc φ cmc<br />

C<br />

of ternary two-phase system, but φ ∗ C<br />

drops below the cmc of the corresponding<br />

binary system for nearly symmetric copolymers.<br />

copolymers due <strong>to</strong> the highest degree of swelling, but <strong>is</strong> always modest by less than a<br />

fac<strong>to</strong>r of two in the present cases.<br />

<strong>Th<strong>is</strong></strong> modest difference between the cmc’s of swollen and unswollen micelles <strong>have</strong> a<br />

profound effect on the predictions of interfacial tensions. The dotted-dashed lines in Fig. 4.5<br />

show the saturation concentration φ ∗ C<br />

of free copolymer within the A-rich phase at which<br />

macroscopic interfacial tension extrapolates <strong>to</strong> zero as shown in Fig. 4.4. The cmc in<br />

the ternary two-phase system thus always remains less than or equal <strong>to</strong> the saturation<br />

concentration φ ∗ C<br />

, implying <strong>that</strong> the interfacial tension will saturate <strong>to</strong> a nonzero (or at least<br />

non-negative) value above the cmc. The predicted cmc of a binary system does, however,<br />

drop below the saturation concentration as f A approaches 0.5, in a manner cons<strong>is</strong>tent with<br />

87


R (Å)<br />

500<br />

450<br />

χ N CB = 8<br />

400<br />

10<br />

350<br />

12<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0.5 0.6 0.7 0.8<br />

f A<br />

Figure 4.6: Micelle core radius at the cmc for the ternary two-phase system and binary<br />

system considered in Fig. 4.5, vs. volume fraction f A of the corona block. The hollow and<br />

solid symbols represent the radii of unswollen and swollen micelles, respectively.<br />

what Leibler found in a more approximate calculation. The values of f A for which the<br />

saturation concentration and the binary cmc become equal are surpr<strong>is</strong>ingly close <strong>to</strong> the<br />

value of 0.69 obtained by Leibler. However, Leibler’s prediction of a wide range of values of<br />

f A for which φ cmc<br />

C<br />

> φ ∗ C<br />

<strong>is</strong> clearly an artifact of h<strong>is</strong> two-component model for the micelles.<br />

The convergence of φ cmc<br />

C<br />

and φ ∗ C at f A = 1/2 implies interfacial tension for a flat<br />

interface at and above cmc extrapolates <strong>to</strong> zero in the limit of the symmetric system.<br />

Fig. 4.6 shows the core radii of micelles in the same set of ternary two-phase systems<br />

and binary systems as those considered in Fig. 4.5. These radii are calculated using a<br />

stat<strong>is</strong>tical segment length of 6Å for both monomers. In the ternary two-phase system, the<br />

micelle radius swells dramatically and diverges as f A approaches 1/2.<br />

In Fig. 4.7 we shows the SCFT predictions of macroscopic interfacial tension saturated<br />

at the cmc for the same set of systems as those considered in two previous figures. As<br />

suggested in Eqn. 4.23, the interfacial tension converges <strong>to</strong> zero at a balance point f bal<br />

A =<br />

88


0.25<br />

γ cmc / γ o<br />

0.2<br />

0.15<br />

0.1<br />

χN CB = 8<br />

10<br />

12<br />

Leibler<br />

0.05<br />

0<br />

0.5 0.55 0.6 0.65 0.7 0.75 0.8<br />

f A<br />

Figure 4.7: Macroscopic interfacial tension γ cmc at the cmc, normalized by the value γ 0<br />

for a bare interface, vs. volume fraction f A of the copolymer corona block, for the same<br />

parameters as those used in Fig. 4.5. Leibler’s prediction based on the unswollen micelle<br />

<strong>is</strong> shown with a dotted line for compar<strong>is</strong>on.<br />

1/2 in the symmetric limit, in which the micelle radius diverges, and exhibits a parabolic<br />

dependence on f A . For compar<strong>is</strong>on Leibler’s prediction <strong>is</strong> given with a dotted line showing<br />

a much wider region of f A for which interfacial tension <strong>is</strong> saturated earlier than the critical<br />

micelle concentration.<br />

The position of a balance point <strong>is</strong> also affected by the asymmetry of the system. Fig. 4.8<br />

shows the same plot as the previous one, but for the systems with asymmetric homopolymers<br />

(β ≠ 1). In each case, there <strong>is</strong> a one value of the corona volume fraction f A for which<br />

the interfacial tension extrapolates <strong>to</strong> zero.<br />

Fig. 4.9 and Fig. 4.10 compare the cmc’s in both binary and ternary systems calculated<br />

by SCFT with those in the binary mixtures by simplified analytic models. The theory of<br />

Leibler, Orland, and Wheeler (LOW) [5] <strong>is</strong> based on a simple strong stretching calculation<br />

of the free energy of a copolymer micelles in a single A homopolymer matrix. The<br />

89


0.6<br />

γ sat /γ o , γ flat /γ o<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

β = 0.5<br />

1<br />

2<br />

0.1<br />

0<br />

0.3 0.4 0.5 0.6 0.7 0.8<br />

f A<br />

Figure 4.8: Macroscopic interfacial tension γ cmc at the cmc, normalized by the bare tension<br />

γ 0 , vs. copolymer corona volume fraction f C for system with f B χN = 10 and various values<br />

of the ratio β of copolymer molecular weights.<br />

10 0 8 9 10 11 12 13 14 15 16<br />

10 -1<br />

f A = 0.6<br />

SCFT (b)<br />

SCFT (t)<br />

LOW<br />

10 -2<br />

φ c<br />

cmc<br />

10 -3<br />

10 -4<br />

10 -5<br />

χ N CB<br />

Figure 4.9: Dependence of critical micelle concentration φ cmc<br />

C<br />

on χN BC = f B χN for systems<br />

with f A = 0.6 and α A χN = α B χN = 10. Symbols show predictions for the ternary<br />

two phase system by SCFT (filled triangles) and predictions for a binary system by SCFT<br />

(filled circles) and by the approximate strong LOW theory (open triangles). The dashed<br />

line <strong>is</strong> the function Ce −χN BC, with C = 376.<br />

90


0.014<br />

0.012<br />

0.01<br />

SCFT (b)<br />

SCFT (t)<br />

Leibler<br />

LOW<br />

0.008<br />

φ cmc<br />

0.006<br />

0.004<br />

0.002<br />

0<br />

0.5 0.6 0.7 0.8<br />

f A<br />

Figure 4.10: Compar<strong>is</strong>on of SCFT predictions for copolymer cmc φ cmc<br />

C<br />

of a ternary twophase<br />

system (filled triangles) and a binary system (open triangles) <strong>to</strong> the strong stretching<br />

predictions for the binary system by Leibler (dot dashed line) and LOW (solid lines) for<br />

the systems considered in Fig. 4.5.<br />

density profiles are assumed uniform in the core and corona regions, but some penetration<br />

of homopolymer in<strong>to</strong> the core and corona <strong>is</strong> allowed. The theory denoted ’Leibler’ <strong>is</strong> a<br />

more simplified version of the LOW theory <strong>that</strong> ignores the effect of the homopolymer<br />

penetration and the translational entropy of a micelle. [11] These analytic theories capture<br />

the dominant exponential dependence of the cmc on χN CB at the value of f A = 0.6 in<br />

Fig. 4.9, but exhibit much <strong>to</strong>o weak a dependence on f A in Fig. 4.10.<br />

4.5 Results - Helfrich Theory<br />

The interfacial tension reduction <strong>is</strong> thermodynamically inhibited by the onset of micellization.<br />

Therefore, interfacial tension γ cmc saturated at the CMC (Fig. 4.4) should be<br />

identical <strong>to</strong> γ flat predicted by the Helfrich membrane theory of the bending elasticity of<br />

a micelle coex<strong>is</strong>ting with two bulk phases. Fig. 4.11 tests the validity of Eqn. 4.23 in two<br />

ways. First, a parabolic curve of Eqn. 4.23 constructed with two parameters τ ′ and κ +<br />

91


0.7<br />

γ / γ o<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

γ flat<br />

γ cmc<br />

τ 2 /2κ +<br />

0.2<br />

0.1<br />

0<br />

0.5 0.55 0.6 0.65 0.7 0.75 0.8<br />

f A<br />

Figure 4.11: The prediction of γ/γ o from direct micelle simulation (hollow circle) and<br />

from Helfrich bending elasticity (solid circle) as a function of f A . They are cons<strong>is</strong>tent for<br />

f A < 0.55 but start <strong>to</strong> deviate for larger f A . Surpr<strong>is</strong>ingly, the quadratic function (solid<br />

line) of γ/γ o (Eqn. 4.23) with τ ′ and κ + evaluated at C = 0 accurately fits the entire<br />

results from the direct micelle simulation.<br />

evaluated at a balance point f bal<br />

A = 1/2 <strong>is</strong> compared with τ2 /2κ + in combined Eqn. 4.20<br />

and Eqn. 4.22 directly obtained in each membrane simulation. The two approaches are<br />

reasonably cons<strong>is</strong>tent for f A < 0.55 but start <strong>to</strong> deviate thereafter. Second, interestingly<br />

Eqn. 4.23 agrees fairly well with γ cmc from the direct micelle simulation, <strong>that</strong> <strong>is</strong>, the<br />

quadratic relationship between γ and f A can be extended <strong>to</strong> the system with highly asymmetric<br />

block copolymer. The Helfrich membrane theory, however, fails even for moderately<br />

asymmetric systems. Both Fig. 4.7 with the symmetric homopolymers and Fig. 4.8 with<br />

the asymmetric homopolymers clearly show <strong>that</strong> γ cmc remarkably follows the quadratic<br />

curvature dependence of the Helfrich theory.<br />

Fig. 4.12 compares the radii of swollen micelles calculated by both micelle and Helfrich<br />

theories as a function of f A . The radius of a micelle <strong>is</strong> defined as R = ( 3<br />

4π (M BV B + M C V C )) 1 3<br />

in a micelle simulation, or as the spontaneous spherical radius of curvature of R = 2κ + /τ in<br />

92


R (Å)<br />

1000<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

unswollen micelle<br />

swollen micelle<br />

2 κ + / τ<br />

0.5 0.6 0.7 0.8<br />

f A<br />

Figure 4.12: The compar<strong>is</strong>on of micelle radii obtained from direct micelle simulation (hollow<br />

circle) and from Helfrich bending elasticity (solid circle) as a function of f A . They are<br />

cons<strong>is</strong>tent for f A < 0.55 but the latter starts <strong>to</strong> underestimate the radii for larger f A .<br />

the Helfrich membrane theory, respectively. The systems are identical <strong>to</strong> those in Fig. 4.11.<br />

One can easily notice <strong>that</strong> the radii determined by the Helfrich membrane theory are bigger<br />

than those of swollen micelles by the micelle simulation for f A < 0.55 because the simplified<br />

cmc condition Φ m = 0 used in the Helfrich theory causes micellization in the ternary<br />

systems <strong>to</strong> occur at a few kT higher energy states than those at the cmc in the micelle<br />

simulation at which the translational entropy of micelles are considered. However, the radii<br />

become smaller than even those of unswollen micelles for highly asymmetric copolymers.<br />

<strong>Th<strong>is</strong></strong> <strong>is</strong> cons<strong>is</strong>tent with the observation <strong>that</strong>, in Fig. 4.11 and in Chapter 2, the Helfrich<br />

membrane theory <strong>is</strong> accurate only for f A < 0.6.<br />

We can observe more interesting features of a block copolymer monolayer in the plots<br />

of [τ ′ ] in Fig. 4.13 and τ ′2<br />

8κ + γ o<br />

in Fig. 4.14 as a function of χN CB . Here γ o <strong>is</strong> bare interfacial<br />

tension of homopolymers in the absence of block copolymer and [τ ′ ] <strong>is</strong> evaluated<br />

by changing only the corona A block while fixing the core block so as <strong>to</strong> keep the cmc<br />

93


2.5<br />

2<br />

1.5<br />

α = 2<br />

1<br />

0.7<br />

0.5<br />

0.25<br />

[τ’]<br />

1<br />

0.5<br />

0<br />

0 5 10 15 20 25 30<br />

χ N CB<br />

Figure 4.13: The normalized slope [τ ′ ] as a function of χN in a range of 0.25 ≤ α ≤ 2<br />

5<br />

4.5<br />

(τ’) 2 /(8κ + γ o )<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

α = 2<br />

1<br />

0.7<br />

0.5<br />

0.25<br />

0<br />

0 5 10 15 20 25 30<br />

χ (N CB )<br />

Figure 4.14: The prefac<strong>to</strong>r of γ flat ,<br />

τ ′2<br />

8κ + γ o<br />

, normalized by a bare interfacial tension γ o as a<br />

function of χN CB in a range of 0.25 ≤ α ≤ 2<br />

94


nearly constant. Both quantities directly indicate how sensitively γ flat and c o respond <strong>to</strong> a<br />

small curvature induced by f A ≠ f bal<br />

A . If one assumes γflat <strong>is</strong> an exact harmonic function<br />

in a range of 0 ≤ f A ≤ 1 which sat<strong>is</strong>fies γ flat = γ o at f A = 0 and f A = 1,<br />

τ ′2<br />

8κ + γ o<br />

= 1<br />

becomes a criterion <strong>to</strong> decide the relative ’flatness’ of a parabolic γ flat curve. When the<br />

length of the homopolymer <strong>is</strong> equal <strong>to</strong> or less than <strong>that</strong> of the copolymer (α ≤ 1), both<br />

τ ′2<br />

8κ + γ o<br />

and [τ ′ ] always continuously decrease <strong>to</strong> zero as the system approaches the Lifshitz<br />

point (χN CB ) L = (1 + 2α 2 )/α. However, for α > 1 these terms sharply diverge at low<br />

χN CB implying the interface becomes unstable because small deviation from a balance<br />

point can lead <strong>to</strong> a dramatic increase of the interfacial tension, which makes it harder <strong>to</strong><br />

compatibilize the large homopolymer with short copolymer. With small block copolymers,<br />

say χN CB < 7, the parabolic γ flat curve <strong>is</strong> greatly affected by the relative length of homopolymer<br />

<strong>to</strong> block copolymer but it becomes insensitive <strong>to</strong> α with larger copolymers due<br />

<strong>to</strong> the increasing rigidities of interfaces as shown in Chapter 2.<br />

4.6 Other Possible Phenomena<br />

In th<strong>is</strong> chapter, we focus primarily upon the formation of swollen spherical micelles and its<br />

relation <strong>to</strong> interfacial tension in ternary two-phase systems. There are, however, several<br />

other possible phenomena <strong>that</strong> should be considered.<br />

Cylindrical Micelles It <strong>is</strong> possible for dilute solutions of block copolymers <strong>to</strong> form<br />

cylindrical, rather than spherical, micelles. In binary systems of A homopolymer and<br />

AB copolymer, analytic strong stretching calculations by Mayes and de la Cruz [16] and<br />

numerical SCFT calculations by Matsen [17], and Janert and Schick [18] <strong>have</strong> predicted the<br />

formation of either cylindrical micelles [16] or highly swollen periodic hexagonal mesophases<br />

[17, 18] in some systems of symmetric and slightly asymmetric copolymers.<br />

In an analys<strong>is</strong> of the behavior of highly swollen micelles in ternary systems, which was<br />

based upon the Helfrich theory of weakly curved saturated interfaces, Safran and Turkevich<br />

95


[19] found <strong>that</strong> spherical micelles are always preferred over cylinders along the emulsification<br />

failure line. In th<strong>is</strong> theory, a transition from spherical <strong>to</strong> cylindrical micelles can,<br />

however, be induced in a homogeneous micellar phase by decreasing the volume fraction<br />

of B. Within the context of the Helfrich theory the preference for spherical micelles <strong>is</strong> a<br />

result of a contribution of a negative Gaussian rigidity ¯κ, which favors the formation of<br />

spherical surface over a cylindrical curvature with the same mean curvature.<br />

The Helfrich theory <strong>is</strong> valid only for systems containing monolayers with low spontaneous<br />

curvatures, which tend <strong>to</strong> form highly swollen micelles near the emulsification<br />

failure line. In the opposite limit of highly asymmetric copolymers, however, we also expect<br />

spherical micelles, even in a binary system, as a result of packing constraints. These<br />

considerations suggest <strong>that</strong> spherical micelles will always be preferred over cylindrical micelles<br />

in a two-phase system.<br />

The above considerations also suggest <strong>that</strong> addition of B homopolymer <strong>to</strong> a binary<br />

system <strong>that</strong> forms cylindrical micelles must generally cause a transition from cylindrical <strong>to</strong><br />

spherical micelles somewhere along the line of critical micelle concentrations <strong>that</strong> extends<br />

from the binary system <strong>to</strong> the emulsification failure line. We <strong>have</strong> not systematically<br />

pursued th<strong>is</strong> here. As an example, however, we <strong>have</strong> compared the cmc’s for cylindrical and<br />

spherical micelles in a single binary system with f A = 1/2, χN C = 20 and α A = α B = 1/2,<br />

for which we find <strong>that</strong> spherical micelles are preferred over cylinders even in absence of<br />

B homopolymer. Cylindrical micelles would presumably be more favorable for symmetric<br />

copolymers with larger values of χN CB or α, for which the corona block would be less<br />

strongly swollen by homopolymer. [16]<br />

Lamellar Phases: The model of a non-interacting spherical micelle d<strong>is</strong>cussed above predicts<br />

an infinite micelle radius in a system with symmetric copolymer and homopolymers.<br />

<strong>Th<strong>is</strong></strong> prediction <strong>is</strong> a result of our implicit assumption <strong>that</strong> excluded volume interactions<br />

between swollen micelles remain irrelevant even as their radii diverge. Systems with bal-<br />

96


anced surfactants are expected <strong>to</strong> instead form either a lamellar phase or a bicontinuous<br />

microemulsion.<br />

Monolayers of symmetric block copolymers with molecular weights less than <strong>that</strong> of either<br />

homopolymer are known <strong>to</strong> exhibit a weak entropic attraction when separated by<br />

d<strong>is</strong>tances for which opposing copolymer brushes slightly overlap. <strong>Th<strong>is</strong></strong> attraction can<br />

prevent swelling of the lamellar spacing, and instead possibly lead <strong>to</strong> a broad region of<br />

three phase coex<strong>is</strong>tence of a collapsed lamellar phase with two phases of essentially pure<br />

homopolymers.[20] Thompson and Matsen [21] <strong>have</strong> used numerical SCFT <strong>to</strong> study the<br />

effective interaction between symmetric monolayers within a ternary lamellar phase, and<br />

identified a line of values of α as a function of χN at which the interaction between monolayers<br />

changes from effectively repulsive (at lower α) <strong>to</strong> attractive (at higher α). <strong>Th<strong>is</strong></strong><br />

critical value of α decreases very slowly with increasing χN, varying from α ≃ 1.3 near<br />

the critical point <strong>to</strong> α = 0.8 at χN > 100. Symmetric systems with α < 0.8 are thus<br />

not expected <strong>to</strong> form a collapsed lamellar phase. The attraction between brushes <strong>that</strong><br />

can cause the collapse of a lamellar phase may analogously cause ordered dense crystals<br />

of spherical or cylindrical micelles under similar conditions, but the interactions between<br />

spheres and cylinders <strong>have</strong> not been so carefully studied.<br />

Bicontinuous Microemulsion: When the interactions between monolayers in a lamellar<br />

phase are repulsive, the lamellar spacing may increase with decreasing the surfactant<br />

concentration, eventually leading, within the context of SCFT, <strong>to</strong> an unbinding transition<br />

at which the lamellar spacing diverges at a nonzero surfactant volume fraction. Systems<br />

for which SCFT predicts such an unbinding transition are are, however, likely <strong>to</strong> form a<br />

bicontinuous microemulsion phase over a narrow range of surfactant concentrations near<br />

the predicted unbinding concentration. The bicontinuous microemulsion phase <strong>is</strong> believed<br />

<strong>to</strong> be stabilized by interfacial fluctuations <strong>that</strong> cannot be correctly described by SCFT.<br />

[22] <strong>Th<strong>is</strong></strong>, however, lies far beyond the scope of the present thes<strong>is</strong>.<br />

97


4.7 Conclusion<br />

In th<strong>is</strong> chapter, we formulated micellization of block copolymer in a grand canonical ensemble<br />

and connect it <strong>to</strong> Helfrich bending elasticity of a block copolymer monolayer in a<br />

micelle for the two-phase ternary polymer system. While a micelle radius <strong>is</strong> simply determined<br />

by a s<strong>to</strong>ichiometry of available B and C molecules in a single phase system, a micelle<br />

coex<strong>is</strong>ting with the excess B phase should swell until a spontaneous spherical curvature<br />

<strong>is</strong> accompl<strong>is</strong>hed at equal pressure across an interface along the emulsification failure line.<br />

Also, due <strong>to</strong> the slightly reduced excess free energy of a micelle, the interfacial saturation<br />

<strong>is</strong> always preempted by micellization producing non-negative interfacial tension.<br />

In the limit of highly swollen micelles, a weakly curved interface within a micelle<br />

<strong>is</strong> successfully described by Helfrich bending elasticity of a block copolymer monolayer.<br />

Combining with the simplified cmc condition in which the translational entropy of a micelle<br />

<strong>is</strong> ignored, the macroscopic interfacial tension saturated at the cmc was found <strong>to</strong> <strong>have</strong> a<br />

quadratic dependence on the spontaneous spherical curvature and van<strong>is</strong>h at a balance<br />

point, which was verified with both micelle simulation and the Helfrich membrane theory.<br />

These parabolic curves of saturated interfacial tension were significantly affected by the<br />

relative length of homopolymer <strong>to</strong> block copolymer if block copolymers are small but start<br />

<strong>to</strong> become nearly constant with increasing copolymer lengths in the context of [τ ′ ] and<br />

τ ′2<br />

8κ + γ o<br />

. As long as α < 1, both terms continuously decrease <strong>to</strong> zero at the Lifshitz point<br />

but sharply diverge for α > 1.<br />

Although here we only focus on the formation of spherical micelle, th<strong>is</strong> approach may<br />

be extended <strong>to</strong> other structures of micelles and <strong>to</strong> phase transitions between them.<br />

98


Bibliography<br />

[1] Linc, H.; Fav<strong>is</strong>, B. D.; Yu, Y. S.; E<strong>is</strong>enberg, A. Macromolecules 1999, 32, 1637 .<br />

[2] Macosko, C. W.; Guegan, P.; Khandpur, A. K.; Nakayama, A.; Marechal, P.; Inoue, T.<br />

Macromolecules 1996, 29, 5590<br />

[3] Noolandi, J.; Hong, K. M. Macromolecules 1982, 15, 482; Noolandi, J.; Hong, K. M.<br />

Macromolecules 1984, 17, 1531<br />

[4] Shull, K. R.; Kramer, E. J. Macromolecules 1990, 23, 4769<br />

[5] Leibler, L.; Orland, H.; Wheeler J. C. J. Chem. Phys. 1983, 79, 3550<br />

[6] Whitmore, M. D.; Noolandi, J. Macromolecules 1985, 18, 657<br />

[7] Mathur. D.; Hariharan, R.; Nauman, E. B. Polymer 1999, 40, 6077<br />

[8] Lent, B. V.; Scheutjens, J. M. H. M. Macromolecules 1989, 22, 1931; Evers, O. A.;<br />

Scheutjens, J. M. H. M.; Fleer, G. J. Macromolecules 1990, 23, 5221;<br />

[9] Linse, P.; Malmsten, M. Macromolecules 1992, 25, 5434<br />

[10] Shull, K. R. Macromolecules 1993, 26, 2346<br />

[11] Leibler, L. Makromol. Chem., Macromol. Symp. 1988, 16, 1<br />

[12] Palmer, K. M.; Morse, D. C. J. Chem. Phys. 1996, 105, 11147<br />

99


[13] Shull, K. R.; Kellock, A. J.; Deline, V. R.; MacDonald, S. A. J. Chem. Phys 1992,<br />

97, 2095<br />

[14] Strey, R. Coll. Ploym. Sci. 1994, 272, 1005<br />

[15] Helfrich, W. Z. Naturforsch. C 1973, 28, 693<br />

[16] Mayes, A. M; de la Cruz, M. O. macromolecules 1988, 21, 2543<br />

[17] Matsen, M. W. Phys. Rev. Lett. 1995, 74, 4225; Matsen, M. W. Macromolecules 1995,<br />

28, 5765<br />

[18] Janert, P. K.; Schick, M. Macromolecules 1998, 31, 1109<br />

[19] Safran, S. A.; Turkevich, L. A.; Pincus, P. J. Phys. (Par<strong>is</strong>) 1984 45, L-69<br />

[20] Broseta, D.; Fredrickson, G. H. J. Chem. Phys. 1990, 93, 2927<br />

[21] Thompson, R. B.; Matsen, M. W. J. Chem. Phys. 2000, 112, 6863<br />

[22] Fredrickson, G. H.; Bates, F. S. J. Polym. Sci.: Part B: Polym. Phys. 1997, 35, 2775<br />

100


Chapter 5<br />

Transport of Block Copolymer<br />

Surfactant Between Two<br />

Homopolymer Phases<br />

The measurements of interfacial tension between v<strong>is</strong>cous homopolymers are often seriously<br />

affected by slow kinetics in the system when block copolymer <strong>is</strong> selectively added <strong>to</strong> either<br />

homopolymer phase. Although equilibrium interfacial tension <strong>is</strong> uniquely determined by<br />

thermodynamics, measured interfacial tension can depend on the actual interfacial concentration<br />

of block copolymer at steady state during experiments. Thus the purpose of th<strong>is</strong><br />

chapter <strong>is</strong> <strong>to</strong> provide the comprehensive understanding about the effect of limited transport<br />

of block copolymer on interfacial tension depending on the various initial conditions.<br />

It should be stated <strong>that</strong> the primary authorship of th<strong>is</strong> chapter belongs <strong>to</strong> my adv<strong>is</strong>or,<br />

Prof. Morse, but I <strong>have</strong> included it here because the analys<strong>is</strong> of kinetics associated with<br />

transport of block copolymer in Chapter 6 and 7 has used the results given in th<strong>is</strong> chapter.<br />

Here we consider the transport of AB block copolymer surfactant <strong>that</strong> <strong>is</strong> initially<br />

homogeneously mixed throughout the matrix of the nearly pure homopolymer A (phase<br />

I) in<strong>to</strong> the drop of the initially pure homopolymer B (phase II). <strong>Th<strong>is</strong></strong> configuration <strong>is</strong><br />

closely related <strong>to</strong> the initial state of the interfacial tension measurement in a spinning drop<br />

tensiometer, which <strong>is</strong> described in the following two chapters.<br />

101


1<br />

0.8<br />

c f (z) / c o<br />

0.6<br />

0.4<br />

II<br />

c f<br />

I,int<br />

I<br />

0.2<br />

c f<br />

II,int<br />

0<br />

z<br />

0<br />

Figure 5.1: A schematic diagram of concentration profile of surfactant perpendicular <strong>to</strong> an<br />

interface z = 0 below the cmc.<br />

The problem may involve dynamics over several time regimes. Initially, some relatively<br />

short time τ int <strong>is</strong> required <strong>to</strong> build up a block copolymer monolayer at the interface. Once a<br />

monolayer <strong>is</strong> establ<strong>is</strong>hed, a steady state <strong>is</strong> achieved at all times τ int ≪ t ≪ τ eq if we assume<br />

<strong>that</strong> the rate of accumulation of copolymer at the interface <strong>is</strong> negligible compared <strong>to</strong> the<br />

rate of flux <strong>to</strong> the interface from phase I or from the interface in<strong>to</strong> phase II. At much<br />

longer time t > τ eq copolymer finally diffuses <strong>to</strong> the center of the drop and an equilibrium<br />

concentration of copolymer <strong>is</strong> obtained throughout the drop. We focus primarily in what<br />

follows upon the copolymer diffusion in the intermediate time regime τ int ≪ t ≪ τ eq with<br />

a negligible rate of accumulation at the interface, and negligible res<strong>is</strong>tance <strong>to</strong> transport<br />

through the interface. In th<strong>is</strong> period, the problem becomes equivalent <strong>to</strong> <strong>that</strong> of diffusion of<br />

a substance from one semi-infinite medium from one in<strong>to</strong> another across an essentially flat<br />

boundary. [1] To d<strong>is</strong>cuss th<strong>is</strong> time regime, we thus consider the one-dimensional problem<br />

of copolymer transport across a flat interface at z = 0 and consider the evolution of the<br />

copolymer concentration c(z) along a direction perpendicular <strong>to</strong> the interface. Phase I <strong>is</strong><br />

102


taken <strong>to</strong> occupy the half space z > 0 with an initial copolymer concentration c o , while<br />

phase II occupies the half space z < 0 with a van<strong>is</strong>hing initial concentration, as shown in<br />

Fig. 5.1.<br />

To d<strong>is</strong>cuss transport under conditions in which micelles may be present, we must d<strong>is</strong>tingu<strong>is</strong>h<br />

at each point a local concentration c f (z) of molecularly d<strong>is</strong>solved ”free” copolymer<br />

from a local concentration c m (z) of copolymer in micelles, such <strong>that</strong> c(z) = c f (z) + c m (z).<br />

We adopt a slightly simplified view of micellization in which the concentration of free<br />

molecules <strong>is</strong> equal <strong>to</strong> the critical micelle concentration (cmc), c f (z) = c c , where c c <strong>is</strong> the<br />

cmc in the phase of interest, at any point with c(z) > c c . <strong>Th<strong>is</strong></strong> description assumes both<br />

<strong>that</strong> the equilibrium between free copolymer and micelles can be adequately described as<br />

a simple limit of solubility (which can be shown <strong>to</strong> be equivalent <strong>to</strong> neglecting the comparatively<br />

small translational entropy of the micelles), and <strong>that</strong> the formation and d<strong>is</strong>solution<br />

of micelles occur essentially instantaneously, so <strong>that</strong> local equilibrium <strong>is</strong> always maintained.<br />

The assumed constancy of c f (z) in regions in which micelles are present has an important<br />

implication for transport: There can be no diffusive flux of free molecules in regions with<br />

micelles since there <strong>is</strong> no gradient in the concentration of free copolymer. All transport<br />

of copolymer must thus occur by micellar diffusion alone in regions with c(z) > c c and by<br />

diffusion of free molecules alone in regions with c(z) < c c .<br />

Let D I f<br />

and DII<br />

f<br />

be the diffusion coefficients of free copolymer in phases I and II,<br />

respectively, and D I m<br />

and DII m<br />

be the diffusivities of micelles in the corresponding phases.<br />

Let c I c<br />

and cII c<br />

be the critical micelle concentrations in phases I and II, and c I f,int<br />

and cII<br />

f,int<br />

be the concentrations of free copolymer infinitesimally close <strong>to</strong> the interface in phases I<br />

and II, respectively. Let K be the partition coefficient for free copolymers in either phase,<br />

defined by c I f,int = KcII f,int<br />

, so <strong>that</strong> the copolymer <strong>is</strong> more soluble in phase I if K > 1.<br />

In what follows we consider three cases: First, we review as background the relatively<br />

simple case in which there are no micelles in either phase. Next, we d<strong>is</strong>cuss the case in<br />

103


which c o > c I c , so <strong>that</strong> micelles are initially present in phase I, and in which the value µI c<br />

of the copolymer chemical potential required <strong>to</strong> form micelles in phase I (i.e., the value<br />

of µ at the cmc in phase I) <strong>is</strong> lower than the corresponding value µ II<br />

c<br />

in phase II. The<br />

condition µ I c < µII c<br />

<strong>is</strong> equivalent <strong>to</strong> the condition c I c<br />

/K < cII c . In th<strong>is</strong> case, micelles do not<br />

appear in phase II, and the mechan<strong>is</strong>m by which copolymer <strong>is</strong> transported from phase<br />

I in<strong>to</strong> phase II <strong>is</strong> shown <strong>to</strong> depend <strong>to</strong> depend upon the magnitude of the diffusivity of<br />

micelles in phase I relative <strong>to</strong> the diffusivity of free copolymer in phase II. Finally we<br />

consider the case in which c o > c I c<br />

but µII c<br />

< µ I c , or cII c<br />

< c I c /K, in which micelles are<br />

initially present only in phase I, but in which micelles would appears only in phase II in<br />

thermodynamic equilibrium.<br />

5.1 No Micelles<br />

When c o ≪ c c in both phases, so <strong>that</strong> no micelles are present in either phase, transport<br />

occurs by Fickian diffusion of d<strong>is</strong>solved copolymer. In th<strong>is</strong> case we thus consider the<br />

evolution of the copolymer concentration c(z) for a one-dimensional diffusion across an<br />

interface, in which the concentrations on either side of the interface sat<strong>is</strong>fy a boundary<br />

condition c I f,int = KcII f,int , where cI f,int = c f(z = 0 + ) and c II<br />

f,int = c f(z = 0 − ).<br />

It <strong>is</strong> known[1] <strong>that</strong> there ex<strong>is</strong>ts an analytical solution <strong>to</strong> th<strong>is</strong> problem in which the<br />

concentration c(z) changes with time everywhere except at the interface, where the c I f,int<br />

and c II<br />

f,int<br />

are actually independent of time. To confirm th<strong>is</strong>, we assume a time-independent<br />

concentration on either side of the interface, and consider the well known error-function<br />

solution for free diffusion from or in<strong>to</strong> a semi-infinite slab from an interface held at fixed<br />

concentration. For the problem at hand, in which the copolymer concentration approaches<br />

c o deep in phase I and van<strong>is</strong>hes deep in phase II, th<strong>is</strong> proposed solution yields a <strong>to</strong>tal flux<br />

104


J I (t) <strong>to</strong> the interface and J II (t) from the interface <strong>that</strong> both decrease as 1/ √ t, given by<br />

√<br />

Df<br />

I J I (t) =<br />

πt (co − c I f,int )<br />

√<br />

Df<br />

II<br />

J II (t) =<br />

πt cII f,int . (5.1)<br />

Requiring <strong>that</strong> J I (t) = J II (t) yields a constant interfacial concentration of copolymer in<br />

phase I<br />

c I f,int =<br />

co<br />

1 + P<br />

(5.2)<br />

where<br />

√<br />

P =<br />

√ DII f<br />

Df<br />

I<br />

1<br />

K . (5.3)<br />

For P ≫ 1, the interfacial concentration thus stays at the much lower concentration than<br />

<strong>that</strong> far from an interface, c I f,int ≪ co . If the diffusion occurs in<strong>to</strong> a finite drop (rather than<br />

a semi-infinite slab) the concentration remains at th<strong>is</strong> value until the entire drop reaches its<br />

equilibrium concentration at t ≃ τ eq . Conversely, for P ≪ 1, the interfacial concentration<br />

c I f,int reaches a value very close <strong>to</strong> co , even at times much less than τ eq . Therefore, a high<br />

interfacial concentration may be quickly achieved for systems in which the solubility <strong>is</strong><br />

much higher in the matrix than in the drop, so <strong>that</strong> K ≫ 1, or in which the diffusivity in<br />

the matrix <strong>is</strong> much higher than <strong>that</strong> in the drop, so <strong>that</strong> D I f ≫ DII f .<br />

5.2 Micelles in Matrix, c I c < c II<br />

c K<br />

We next consider the case c o > c I c , in which micelles are initially present in phase I, but<br />

in which µ I c < µII c , or cI c /K < cII c , so <strong>that</strong> the formation of micelles in phase I <strong>is</strong> preferred<br />

over the formation of micelles in phase II. In th<strong>is</strong> case, micelles never ex<strong>is</strong>t in phase<br />

II. Two qualitatively different situations can ar<strong>is</strong>e in phase I: One possibility <strong>is</strong> <strong>that</strong><br />

c(z) <strong>is</strong> greater than c I c throughout phase I, so <strong>that</strong> c f(z) = c I c<br />

for all z > 0 at all times<br />

of interest. In th<strong>is</strong> case, transport in phase I <strong>is</strong> governed by micellar diffusion alone, as<br />

105


3<br />

c m<br />

o<br />

c f (z) / c c<br />

I , cm (z) / c c<br />

I<br />

2<br />

1<br />

II<br />

I<br />

c c<br />

I<br />

0<br />

z<br />

0<br />

Figure 5.2: A schematic diagram of concentration profile of surfactant and micelle perpendicular<br />

<strong>to</strong> an interface z = 0 without exclusion zone (Q < 1) above the cmc.<br />

a result of the absence of a gradient in the concentration of free molecules. The other<br />

possibility <strong>is</strong> <strong>that</strong> there ex<strong>is</strong>ts a region near the interface in which c(z) < c I c, from which<br />

micelles are completely excluded. <strong>Th<strong>is</strong></strong> will be referred <strong>to</strong> in what follows as an ”exclusion<br />

zone”. Whether an exclusion zone <strong>is</strong> created depends upon a combination of kinetic and<br />

thermodynamic properties of both phases, as d<strong>is</strong>cussed below.<br />

5.2.1 Micellar Diffusion (No Exclusion Zone)<br />

We first consider the case in which there <strong>is</strong> no exclusion zone, so <strong>that</strong> c(z) > c I c for all z > 0.<br />

In th<strong>is</strong> case, surfactant in phase I <strong>is</strong> carried all the way <strong>to</strong> the interface by micellar diffusion<br />

alone. Since the interfacial concentration c I f,int <strong>is</strong> equal <strong>to</strong> cI c , the interfacial concentration<br />

in phase II must be c II<br />

f,int = cI c/K as shown in Fig. 5.2. The flux in<strong>to</strong> phase II <strong>is</strong> thus<br />

given again by the solution for the previous case c(z) < c I c, which gives<br />

√<br />

Df<br />

II c I c<br />

J II (t) =<br />

πt K<br />

(5.4)<br />

106


for a system with an interfacial concentration c II<br />

f,int /cI c /K. Postulating an analogous concentration<br />

profile for micelles in the micellar phase, controlled by a diffusivity D I m of a<br />

micelle with a time-independent interfacial concentration c m (0 + ) in phase I yields the flux<br />

J m (t) =<br />

√<br />

D I m<br />

πt [co m − c m(0 + )] (5.5)<br />

<strong>that</strong> also decreases as 1/ √ t. By equating the prefac<strong>to</strong>rs of 1/ √ t in Eqns. (5.4) and (5.5),<br />

we find<br />

c m (0 + )<br />

c o m<br />

= 1 − Q (5.6)<br />

where<br />

Q ≡<br />

√<br />

D II<br />

f<br />

D I m<br />

c c<br />

c o m<br />

1<br />

K . (5.7)<br />

A non-negative interfacial concentration, c m (0 + ) > 0, <strong>is</strong> obtained only for Q < 1. <strong>Th<strong>is</strong></strong><br />

requirement Q < 1 <strong>is</strong> equivalent <strong>to</strong> the requirement <strong>that</strong> the flux <strong>that</strong> would be obtained<br />

by micelle diffusion <strong>to</strong> an absorbing boundary, with c m (0 + ) ∼ 0, must exceed the rate<br />

of leakage of copolymer in<strong>to</strong> phase II, as given by Eq. (5.4). Note <strong>that</strong> the condition<br />

involves both the micellar diffusivity D I m<br />

in phase I and the molecular diffusivity DII<br />

f<br />

in<br />

phase II, but not molecular diffusivity D I f<br />

in phase I, since no molecular diffusion occurs<br />

in phase I. If the condition Q < 1 <strong>is</strong> sat<strong>is</strong>fied, the concentration of free copolymer near<br />

the interface can thus reach the cmc over times much faster than the equilibrium time τ eq<br />

needed <strong>to</strong> fill the drop, thus rapidly establ<strong>is</strong>hing the highest interfacial coverage and the<br />

lowest interfacial tension <strong>that</strong> thermodynamics can allow.<br />

5.2.2 Molecular Diffusion (Exclusion Zone)<br />

If Q > 1, micelle diffusivity alone cannot provide the flux of copolymer in<strong>to</strong> domain II<br />

needed <strong>to</strong> maintain a concentration c I f,int = cI c<br />

at the interface. As a result, the concentration<br />

c f (z) of copolymer near the interface must drop below the cmc, and cause the<br />

formation of a region near the interface from which micelles are completely excluded. We<br />

107


3<br />

h (t)<br />

d m (t)<br />

c m<br />

o<br />

c f (z) / c c<br />

I , cm (z) / c c<br />

I<br />

2<br />

1<br />

II<br />

I<br />

c c<br />

I<br />

0<br />

z<br />

0<br />

Figure 5.3: A schematic diagram of concentration profile of surfactant perpendicular <strong>to</strong> an<br />

interface z = 0 for Q > 1 above the cmc. A solid line <strong>is</strong> for the case of S ≫ 1 and a dotted<br />

line for the case of S ≪ 1.<br />

may thus divide phase I at any time t in<strong>to</strong> an ”exclusion zone” 0 < z < h(t) in which<br />

c f (z) < c c and c m (z) = 0, and a micellar zone z > h(t) in which c f (z) = c c and c m (z) > 0.<br />

A schematic of th<strong>is</strong> behavior <strong>is</strong> shown Fig. 5.3. As shown there, the micelle concentration<br />

c m (z) must van<strong>is</strong>h along the moving edge z = h(t) of the exclusion zone. Within the<br />

micellar zone z > h(t), there will generally be a ”depletion region” of some width d m <strong>that</strong><br />

precedes the edge of the exclusion zone, in which the concentration c m (z) <strong>is</strong> significantly<br />

less than c o m, within which micellar diffusion <strong>is</strong> significant. For copolymer <strong>that</strong> starts in<br />

a micelle <strong>to</strong> reach the interface between phases I and II, it must first be transported by<br />

micellar diffusion <strong>to</strong> the moving edge of the diffusion zone, where the micelle d<strong>is</strong>solves, and<br />

then undergo molecular diffusion across the exclusion zone.<br />

When an exclusion zone ex<strong>is</strong>ts, we are no longer able <strong>to</strong> find an exact analytic solution<br />

<strong>to</strong> the problem. Significant insight can be gained, however, by analyzing two possible<br />

limiting cases:<br />

108


One limit <strong>is</strong> the situation in which the width d m of the depletion region <strong>is</strong> very narrow<br />

compared <strong>to</strong> the width h(t) of the exclusion zone. <strong>Th<strong>is</strong></strong> corresponds <strong>to</strong> the case of negligible<br />

micelle diffusivity. In th<strong>is</strong> limit, the presence of the narrow depletion region <strong>is</strong> irrelevant,<br />

and the magnitude of the flux of copolymer <strong>to</strong> the interface at z = 0 <strong>is</strong> controlled by<br />

the rate at which essentially immobile micelles <strong>that</strong> d<strong>is</strong>solve at the edge z = h(t) of the<br />

exclusion zone can diffuse across the exclusion zone.<br />

The opposite limit occurs when the width h(t) <strong>is</strong> very narrow compared <strong>to</strong> the width of<br />

the depletion region. The edge of the exclusion zone thus acts in th<strong>is</strong> case as an essentially<br />

stationary absorbing boundary, along which c m (z) = 0, along a surface <strong>that</strong> remains very<br />

near z = 0. In th<strong>is</strong> case, the micelle concentration c m (z) <strong>is</strong> thus well approximated by the<br />

standard solution for diffusion of an initially homogeneous concentration of micelles <strong>to</strong> an<br />

absorbing boundary, in which the width of the depletion region increases as √ Dm I t. In<br />

th<strong>is</strong> case, the overall flux of copolymer <strong>to</strong> the interface <strong>is</strong> controlled by micelle diffusivity,<br />

and the presence of a very narrow exclusion zone <strong>is</strong> essentially irrelevant. We will show<br />

in what follows <strong>that</strong> th<strong>is</strong> limit <strong>is</strong> obtained in the limit of a van<strong>is</strong>hing cmc, in which the<br />

diffusion of free copolymer <strong>is</strong> limited by the limited solubility of free copolymer.<br />

Limit i) Negligible Micelle Diffusivity<br />

We first consider the case of negligible micelle diffusivity D I m<br />

→ 0, in which the width of<br />

the depletion zone goes <strong>to</strong> zero. In th<strong>is</strong> limit the flux of copolymer <strong>to</strong> the interface, which<br />

occurs by diffusion of free molecules across the exclusion zone, <strong>is</strong> provided by d<strong>is</strong>solution of<br />

micelles at the boundary z = h(t) between the depletion and micellar zones. The micellar<br />

zone <strong>is</strong> thereby eaten away, in a process analogous <strong>to</strong> <strong>that</strong> which occurs (in the absence of<br />

convection) when a soluble solid <strong>is</strong> immersed in solvent.<br />

In th<strong>is</strong> limit, the flux of free copolymers at z = h(t) must equal the rate<br />

J I (t) = c o dh(t)<br />

m<br />

dt<br />

109<br />

(5.8)


at which micelles are d<strong>is</strong>solved along the moving boundary h(t). If we make a quasi-steady<br />

state approximation in which we neglect accumulation of copolymer within the depletion<br />

zone, and thus approximate c f (z) within the depletion zone by a linear function of z, the<br />

diffusive flux throughout th<strong>is</strong> region <strong>is</strong><br />

J I (t) = Df<br />

I c I c − c I f,int (t)<br />

h(t)<br />

. (5.9)<br />

By considering a candidate solution in which c I f,int<br />

(t) <strong>is</strong> postulated <strong>to</strong> be independent of<br />

time, and equating the above two expressions for the flux, we obtain an ordinary differential<br />

equation for h(t), which has a solution<br />

h(t) =<br />

√<br />

2D I f t (cI c − cI f,int )/co m (5.10)<br />

Combining th<strong>is</strong> with Eq. (5.8) yields a flux<br />

√<br />

Df<br />

I J I (t) =<br />

2t (cI c − c I f,int )co m . (5.11)<br />

Note <strong>that</strong> in th<strong>is</strong> case the flux again decreases with time as 1/ √ t, but <strong>that</strong> the prefac<strong>to</strong>r<br />

<strong>is</strong> a nonlinear function of c c , c I f,int , and co m.<br />

Matching the prefac<strong>to</strong>r of 1/ √ t in the r.h.s. of Eq. (5.11) <strong>to</strong> the corresponding prefac<strong>to</strong>r<br />

on the r.h.s. of Eq. (5.1) for the flux in<strong>to</strong> domain II again yields a solution with a time<br />

independent value for c I f,int<br />

. In th<strong>is</strong> case, we find <strong>that</strong> the ratio<br />

x ≡ c I f,int /cI c , (5.12)<br />

sat<strong>is</strong>fies a quadratic equation<br />

0 = Rx 2 + x − 1 (5.13)<br />

in which<br />

R ≡ 2 π<br />

D II<br />

f<br />

D I f<br />

c I c<br />

c o m<br />

1<br />

K 2 . (5.14)<br />

110


The solution x = (−1 + √ 1 + 4R)/2R has the limits<br />

x =<br />

{ 1 − R for R ≪ 1<br />

R −1/2 for R ≫ 1<br />

. (5.15)<br />

At times t ≪ τ eq , the systems with essentially immobile micelles and R ≫ 1 thus exhibit an<br />

interfacial concentration c I f,int ≪ cI c much less than the cmc as the result of the relatively<br />

high diffusivity in the domain II. In the extreme limit of D II<br />

f<br />

→ ∞, c I f,int<br />

= 0 at an<br />

absorbing boundary. On the other hand, the systems with R ≪ 1 exhibit c I f,int ≃ cI c .<br />

Both cases lead <strong>to</strong> apparently time-independent values for the interfacial concentration,<br />

and thus a time-independent interfacial tension, at times t ≪ τ eq .<br />

When micelle diffusivity <strong>is</strong> small but nonzero, the width d m (t) of the depletion region<br />

may be estimated as follows: The gradient dc m /dz of the micelle concentration at<br />

z infinitesimally greater than h(t) may be calculated by matching the flux Dmdc I m /dz of<br />

micelles <strong>to</strong> edge of the exclusion zone <strong>to</strong> the rate at which micelles d<strong>is</strong>solve along th<strong>is</strong> edge,<br />

which <strong>is</strong> equal <strong>to</strong> the <strong>to</strong>tal flux J I (t). By approximating dc m /dz ≃ c o m/d m (t) and matching<br />

fluxes we obtain a width<br />

d m (t) ∼ Sh(t) ≃ √ √<br />

S Dmt I (5.16)<br />

where S <strong>is</strong> a dimensionless ratio<br />

S ≡<br />

D I mc o m<br />

D I f (cI c − c I f,int ) . (5.17)<br />

The approximation considered above, in which we neglected micelle diffusivity, <strong>is</strong> value<br />

when S ≪ 1. In th<strong>is</strong> case, d m (t) remains much less than both the width h(t) of the<br />

exclusion zone and the width √ Dm I t of the region <strong>that</strong> would be depleted in the limit of<br />

an infinitesimal exclusion zone. In the opposite case S ≫ 1, we obtain an exclusion zone<br />

<strong>that</strong> <strong>is</strong> much narrower than the depleted region, which <strong>is</strong> d<strong>is</strong>cussed below.<br />

111


Limit ii) Negligible Solubility<br />

We now consider the situation h(t) ≪ d m (t). <strong>Th<strong>is</strong></strong> <strong>is</strong> found <strong>to</strong> occur when S ≪ 1. Since<br />

the micelle diffusivity D I m <strong>is</strong> always much less than the free molecule diffusion D I f , the<br />

limit S ≫ 1 may be obtained only in systems in which the cmc c I c, and thus the difference<br />

c I c − cI f,int<br />

, <strong>is</strong> much less than the <strong>to</strong>tal concentration. <strong>Th<strong>is</strong></strong> limit <strong>is</strong> thus best unders<strong>to</strong>od in<br />

practice as a limit of very low molecular solubility, in which the exclusion zone in which<br />

c f (z) < c I c must become very narrow in order <strong>to</strong> produce a concentration gradient large<br />

enough <strong>to</strong> support the mass flux provided by micellar diffusion.<br />

In th<strong>is</strong> case, the edge of the exclusion zone acts as a nearly stationary absorbing<br />

boundary along a surface h(t) <strong>that</strong> stays very near z = 0. In the limit h(t) → 0 of interest,<br />

c m (z) may thus be approximated by the error-function solution for diffusion of an initially<br />

homogeneous d<strong>is</strong>tribution <strong>to</strong> an absorbing boundary at z = 0. <strong>Th<strong>is</strong></strong> yields a depletion<br />

region of width<br />

d m (t) ∼<br />

√<br />

D I m t (5.18)<br />

and a <strong>to</strong>tal flux<br />

J I (t) =<br />

√<br />

D I m<br />

πt co m . (5.19)<br />

The concentration c I f,int<br />

of free molecules at the interface may be obtained by matching<br />

Eq. (5.19) <strong>to</strong> Eq. (5.4) for the flux in<strong>to</strong> phase II. <strong>Th<strong>is</strong></strong> yields a value<br />

c I f,int<br />

c I c<br />

= 1 Q , (5.20)<br />

for the dimensionless ratio x = c I f,int /cI c . A reasonable approximation for x for Q > 1 and<br />

any value of S might be obtained by using the larger of the values for x obtained from<br />

Eq. (5.20) and from the solution of Eq. (5.13), or by an appropriate interpolation between<br />

these values.<br />

112


5.3 Micelles in Wrong Matrix, c II<br />

c < c I c /K<br />

Finally, we consider the situation in which we put micelles in what <strong>is</strong>, from a thermodynamic<br />

point of view, the wrong phase: We consider the situation in which copolymer<br />

<strong>is</strong> initially mixed in<strong>to</strong> phase I with a concentration c o > c I c, but in which µ II<br />

c<br />

< µ I c, or<br />

c II<br />

c<br />

< c I c /K so <strong>that</strong> the global equilibrium state <strong>is</strong> one in which appear only in phase II.<br />

<strong>Th<strong>is</strong></strong> might occur, for example, if one mixed a diblock copolymer with f A < 1/2 in<strong>to</strong> a<br />

phase of nearly pure A, forcing the formation of inverted micelles in which the B core block<br />

<strong>is</strong> longer than the A corona, and tracked the diffusion of the copolymer in<strong>to</strong> phase B, in<br />

which the longer B block becomes the corona.<br />

In th<strong>is</strong> case, the concentration c II<br />

f,int<br />

in phase II near the interface can never exceed<br />

the critical micelle concentration c II<br />

c . As a result, the concentration of free molecules near<br />

the interface in phase I can never exceed c II<br />

c K. Since, by assumption, c II<br />

c K < c I c, the<br />

interfacial concentration of free molecules in phase I must remain below c I c , forcing the<br />

formation of an exclusion zone. In th<strong>is</strong> case there must thus always an exclusion zone near<br />

the interface in phase I.<br />

Two possibilities remain: One possible outcome <strong>is</strong> an interfacial concentration c II<br />

f,int <<br />

c II<br />

c . In th<strong>is</strong> case no micelles appear in phase II. (More prec<strong>is</strong>ely, in th<strong>is</strong> case no micelles<br />

can appear in phase II until times t > τ eq required <strong>to</strong> for copolymer <strong>to</strong> diffuse over the<br />

entire volume of phase II). The other possibility <strong>is</strong> one in which c II<br />

f,int = cII c , and in which<br />

micelles ex<strong>is</strong>t in phase II near the interface.<br />

When no micelles are present in phase II, the mathematical model needed <strong>to</strong> describe<br />

transport <strong>is</strong> identical <strong>to</strong> <strong>that</strong> developed above <strong>to</strong> describe the situation in which there <strong>is</strong> an<br />

exclusion zone in phase I (Q > 1) and no micelles in phase II, but in which c II<br />

c<br />

> c I c/K:<br />

The only d<strong>is</strong>tinction between the two situations <strong>is</strong> the value of c II<br />

c , which <strong>is</strong> irrelevant until<br />

micelles actually appear in phase II.<br />

113


4<br />

h (t)<br />

d m (t)<br />

c m<br />

o<br />

c f (z) / c c<br />

I , cm (z) / c c<br />

I<br />

3<br />

2<br />

1<br />

c c<br />

II<br />

II<br />

I<br />

c c<br />

I<br />

c c<br />

II K<br />

0<br />

z<br />

0<br />

(a)<br />

4<br />

h (t)<br />

d m (t)<br />

c m<br />

o<br />

c f (z) / c c<br />

I , cm (z) / c c<br />

I<br />

3<br />

2<br />

1<br />

c c<br />

II<br />

II<br />

I<br />

c c<br />

I<br />

c c<br />

II K<br />

0<br />

z<br />

0<br />

(b)<br />

Figure 5.4: A schematic diagram of concentration profile of surfactant when surfactant <strong>is</strong><br />

initially added <strong>to</strong> the wrong phase (I) (a) when Eqn. 5.21 and Eqn. 5.22 are sat<strong>is</strong>fied (b)<br />

not sat<strong>is</strong>fied.<br />

114


Micelles appear in phase II near the interface if the model <strong>that</strong> neglects th<strong>is</strong> possibility<br />

predicts an interfacial concentration c II<br />

f,int ≥ cII c , or c I f,int ≥ cII c K. Since we do not <strong>have</strong><br />

an exact expression for c II<br />

f,int<br />

for arbitrary values of S, we can d<strong>is</strong>cuss only the limiting<br />

behaviors. In the limit S ≪ 1 of van<strong>is</strong>hing micelle diffusivity, we find by setting c II<br />

f,int ≥ cII c<br />

in Eqn. 5.13 <strong>that</strong> micelles appear when<br />

2 Df<br />

II<br />

π Df<br />

I<br />

c II<br />

c<br />

c o m K + 1 <<br />

cI c<br />

Kc II c<br />

(5.21)<br />

Note <strong>that</strong> c I c /KcII c<br />

> 1 by assumption, so the rhs always exceeds one. In the opposite limit<br />

S ≫ 1, we find from Eqn. 5.20 <strong>that</strong> micelles form when Q < c I c/(c II<br />

c K), or<br />

√<br />

D II<br />

f<br />

D I m<br />

c I c<br />

c o mK <<br />

cI c<br />

c II c K<br />

(5.22)<br />

For both large and small S, the formation of micelles within phase II near the interface <strong>is</strong><br />

favored by a low diffusivity in phase II, since a high diffusivity in phase II tends <strong>to</strong> suppress<br />

the concentration at the interface by sweeping copolymer rapidly in<strong>to</strong> the copolymerpoor<br />

interior of phase II. When micelles do appear in phase II, they are formed by a<br />

creation reaction <strong>that</strong> occurs at the interface. When th<strong>is</strong> happens, the chemical potential<br />

of copolymer at the interface reaches the maximum value allowed by thermodynamics,<br />

which in th<strong>is</strong> case corresponds <strong>to</strong> a concentration c II<br />

f,int = cII c , during a time regime in<br />

which most of phase II contain no copolymer.<br />

If micelles appear near the interface in phase II, the structure of micellar zone and<br />

exclusion zone <strong>that</strong> ex<strong>is</strong>ts in phase I must be repeated in phase II. There must a micellar<br />

zone near the interface in phase II, within which c m (z) > 0 and c f (z) = c II<br />

c , and an<br />

semi-infinite micelle-free zone further from the interface, within which c m (z) = 0 and<br />

c f (z) < c II<br />

c . In th<strong>is</strong> case, transport of copolymer between the two phases must thus<br />

involve: micellar diffusion <strong>to</strong> the edge z = h(t) of the exclusion zone in phase I, d<strong>is</strong>solution<br />

of the micelles at th<strong>is</strong> edge, molecular diffusion across the exclusion zone, reformation of<br />

115


micelles on the phase II side of the interface, micellar diffusion across the micellar zone<br />

in phase II, d<strong>is</strong>solution of the micelles at the edge of the micellar region in phase II, and<br />

molecular diffusion in<strong>to</strong> the micelle-free interior of phase II.<br />

The evolution of such a system over very long times can follow different scenarios,<br />

depending upon the relative volumes of phases I and II, and the initial concentration<br />

of copolymer in phase I. The simplest scenario <strong>is</strong> for the micellar region of phase II <strong>to</strong><br />

expand until it fills all of phase II, the micelle-free region region <strong>to</strong> expand until it fills<br />

phase I, and transport of copolymer from phase I <strong>to</strong> phase I <strong>to</strong> then continue until the<br />

concentration of free molecules in phase I reaches c f = c II<br />

c K throughout. <strong>Th<strong>is</strong></strong> leaves<br />

the system in a final state in which there are micelles throughout phase II and none in<br />

phase I. A complicated scenario must occur when the excess of the initial copolymer<br />

concentration above the final concentration c I c<br />

K in phase I <strong>is</strong> insufficient <strong>to</strong> fill phase II<br />

at a homogeneous concentration c I c (i.e., when difference co − c II<br />

c K times the volume of<br />

phase I <strong>is</strong> less than c II<br />

c<br />

times the volume of phase II). In th<strong>is</strong> case, the final state will<br />

contain no micelles in either phase, and so any micelles <strong>that</strong> form in phase II near the<br />

interface during the exchange of copolymer between phases must ultimately re-d<strong>is</strong>solve.<br />

116


Bibliography<br />

[1] Crank, J. The Mathematics of Diffusion (Oxford:Clarendon Press) 1975<br />

117


Chapter 6<br />

Interfacial Tension of PI/PDMS<br />

with PI-b-PDMS Copolymer<br />

Abstract<br />

Equilibrium interfacial tension γ has been measured with a spinning drop tensiometer<br />

for 1,4 poly<strong>is</strong>oprene (PI) and polydimethylsiloxane (PDMS) homopolymers mixed with<br />

poly(<strong>is</strong>oprene-b-dimethylsiloxane) (IDMS) block copolymers. Measurements <strong>have</strong> been<br />

conducted for 12 copolymers in which the volume fraction f A of the PDMS block of the<br />

copolymer <strong>is</strong> varied from 0.49 <strong>to</strong> 0.73. The interfacial tension <strong>is</strong> found <strong>to</strong> be independent of<br />

copolymer concentration above a critical concentration of order 0.1 wt % copolymer, which<br />

we associate with the formation of micelles or some other structure of aggregated copolymer.<br />

In the case of a symmetric copolymer, the measured interfacial tension <strong>is</strong> three orders<br />

of magnitude below <strong>that</strong> obtained in the absence of copolymer. Experimental results <strong>have</strong><br />

been compared <strong>to</strong> the predictions of numerical self-cons<strong>is</strong>tent field theory for an interface<br />

between a PDMS phase containing swollen spherical micelles and a phase of nearly pure<br />

PI. Agreement between th<strong>is</strong> theory and experiments <strong>is</strong> excellent for sufficiently asymmetric<br />

copolymers, with f A > 0.65. Measured values of γ are much lower than predictions for<br />

0.5 < f A ≤ 0.6, however, and exhibit an apparently d<strong>is</strong>continuous change with increasing<br />

f A over a narrow range 0.60 < f A < 0.65. The ex<strong>is</strong>tence of a bicontinuous microemulsion<br />

118


phase has been identified by results of small angle X-ray scattering experiments on ternary<br />

blends containing symmetric copolymer.<br />

6.1 Introduction<br />

When small amounts of surfactant are added <strong>to</strong> a mixture of oil and water, the oil-water<br />

interfacial tension initially decreases, and the amount of adsorbed copolymer increases,<br />

with increasing concentration of d<strong>is</strong>solved surfactant. Both the interfacial tension and the<br />

interfacial coverage generally saturate, however, when the surfactant concentration exceeds<br />

a critical value, above which additional surfactant begins <strong>to</strong> form aggregates in one of the<br />

bulk phases. [1, 2]<br />

<strong>Th<strong>is</strong></strong> saturation of macroscopic interfacial tension <strong>is</strong> observed independent of the form<br />

taken by surfactant aggregates: It occurs whether surfactant aggregates in<strong>to</strong> swollen spherical<br />

or cylindrical micelles, or in<strong>to</strong> monolayers within a bicontinuous microemulsion. The<br />

value of the saturated macroscopic interfacial tension depends, however, upon the structure<br />

and free energy per molecule of the aggregates. Extremely low interfacial tensions can<br />

be obtained in systems for which the thermodynamically preferred form of aggregation <strong>is</strong><br />

a surfactant monolayer with a van<strong>is</strong>hing spontaneous curvature (i.e., a van<strong>is</strong>hing thermodynamically<br />

preferred value of interfacial mean curvature). Systems of such ”balanced”<br />

surfactants generally form either a bicontinuous microemulsion or a lamellar phase, either<br />

of which may exhibit regions of two and three phase coex<strong>is</strong>tence with phases of excess<br />

oil and/or water. In some mixtures of oil, water, and nonionic alkyl-polyethylene oxide<br />

surfactants [3, 4, 5], in which the monolayer spontaneous curvature <strong>is</strong> a sensitive function<br />

of temperature, the macroscopic interfacial tension between oil and water phases reaches<br />

a minimum at a ”mean” temperature T m at which the monolayer spontaneous curvature<br />

passes through zero, with a minimum value <strong>that</strong> can be a few orders of magnitude below<br />

the bare oil-water interfacial tension.<br />

119


In mixtures of imm<strong>is</strong>cible A and B homopolymers with small amounts of AB diblock<br />

copolymer, the copolymer plays a role analogous <strong>to</strong> <strong>that</strong> of small molecule surfactants in<br />

oil-water mixtures.[6] Some symmetric ternary mixtures <strong>have</strong> been shown <strong>to</strong> form balanced<br />

bicontinuous microemulsions, [7, 8, 9]. Direct measurements of interfacial tension <strong>have</strong> not<br />

previously been conducted for these systems. The spontaneous curvature of monolayers<br />

of AB block copolymer in mixtures of A and B homopolymers cannot, however, be easily<br />

controlled by varying the temperature. Here, we instead vary spontaneous curvature by<br />

conducting experiments at fixed temperature with a sequence of mixtures in which we vary<br />

both the molecular weight of one block of the copolymer and the molecular weight of one<br />

of the two homopolymers.<br />

One potential advantage of the study of block copolymers as model surfactants <strong>is</strong><br />

the ability of numerical self-cons<strong>is</strong>tent field theory (SCFT) <strong>to</strong> provide rather accurate<br />

predictions of free energies of polymeric monolayers and micelles. We recently used <strong>to</strong><br />

SCFT <strong>to</strong> study the interfacial tension between a phase of nearly pure B homopolymer and<br />

a coex<strong>is</strong>ting A-rich micellar phase containing spherical micelles of AB diblock copolymer,<br />

in which the micelles may be swollen with emulsified B. For mixtures of homopolymers<br />

with equal degrees of polymerization, and monomers with equal stat<strong>is</strong>tical segment lengths,<br />

th<strong>is</strong> model predicts <strong>that</strong> the interfacial tension <strong>that</strong> van<strong>is</strong>hes for a symmetric copolymer,<br />

with f A = 1/2, and varies quadratically around with the volume fraction f A of the A block.<br />

One goal of the present study <strong>is</strong> <strong>to</strong> quantitatively test the predictions of th<strong>is</strong> theory.<br />

From an experimental point of view, the most important difference between smallmolecule<br />

and macromolecular surfactants <strong>is</strong> the difference in equilibration time scales,<br />

rather than any difference in equilibrium behavior. The low diffusivity of any macromolecular<br />

surfactant in a polymeric matrices make it difficult <strong>to</strong> establ<strong>is</strong>h a state of true phase<br />

equilibrium in a macroscopic apparatus for measuring interfacial tension. In both the pendant<br />

drop method [10] and the spinning drop method used here [11, 12], the tension of<br />

120


an interface between a macroscopic (e.g., millimeter) drop and a surrounding matrix <strong>is</strong><br />

inferred from the deformation of the droplet caused by gravitational or centrifugal forces.<br />

To guarantee <strong>that</strong> such a system reaches a state of global thermodynamic equilibrium,<br />

it would be necessary <strong>to</strong> allow time for copolymer <strong>to</strong> diffuse across the entire drop and<br />

matrix, and thereby establ<strong>is</strong>h homogeneous copolymer chemical potential throughout the<br />

sample.<br />

The ideal samples for interfacial tension measurement are thus macroscopically phase<br />

separated mixtures <strong>that</strong> <strong>have</strong> been allowed time <strong>to</strong> fully equilibrate. Such samples <strong>have</strong><br />

been used in the small molecule surfactant system.[4] For polymeric surfactants, however,<br />

it can be difficult or impossible <strong>to</strong> obtain a state of global equilibrium by diffusion of the<br />

copolymer. In samples in which the copolymer <strong>is</strong> initially mixed in<strong>to</strong> one of two imm<strong>is</strong>cible<br />

homopolymers, it can be difficult <strong>to</strong> assure <strong>that</strong> copolymer diffusion will <strong>have</strong> sufficient<br />

time <strong>to</strong> establ<strong>is</strong>h an equilibrium concentration of copolymer at the interface, or <strong>to</strong> confirm<br />

<strong>that</strong> th<strong>is</strong> has occurred. Thus, it has been challenging <strong>to</strong> measure equilibrium interfacial<br />

tension in the presence of block copolymer. Over the last decade, several researchers<br />

<strong>have</strong> noted <strong>that</strong> the slow diffusion of block copolymer can affect the results of interfacial<br />

tension measurements.[10, 13, 14] The results of pendant drop tensiometer experiments<br />

differ significantly depending upon whether the copolymer <strong>is</strong> initially premixed with the<br />

drop or the matrix material.<br />

The goals of th<strong>is</strong> study are <strong>to</strong> measure equilibrium interfacial tensions in a system<br />

whose character<strong>is</strong>tics <strong>have</strong> been chosen <strong>to</strong> minimize the effect of slow copolymer diffusion<br />

upon the measured interfacial tension, and <strong>to</strong> compare the results <strong>to</strong> the predictions of the<br />

SCFT presented in Chapter 4.<br />

121


6.2 Experiment<br />

All of the experiments reported in th<strong>is</strong> chapter <strong>have</strong> been carried out on spinning drop tensiometer<br />

samples containing a drop surrounding a drop of poly(dimethylsiloxane) (PDMS)<br />

surrounded by a matrix of 1,4 poly<strong>is</strong>oprene (PI) mixed with poly( 1,4 <strong>is</strong>oprene-b-dimethylsiloxane)<br />

(IDMS). The copolymer was always initially mixed with the PDMS matrix.<br />

6.2.1 Materials<br />

All polymers used in th<strong>is</strong> study were synthesized by anionic polymerization.[15] Isoprene<br />

(Aldrich) monomers were stirred with calcium hydride (CaH 2 ) overnight and purified with<br />

n-butyl lithium twice under vacuum for 2 hours each. Polymerization of poly<strong>is</strong>oprene (PI)<br />

homopolymer/block was initiated with sec-butyl lithium in 10 vol % cyclohexane solution<br />

and carried out at 40 ◦ C for 6 hours. The solution of PI homopolymer was terminated<br />

in degassed methanol, washed with deionized water several times and finally dried in a<br />

vacuum oven until the weight remained constant.<br />

For the synthes<strong>is</strong> of poly(dimethylsiloxane) (PDMS) homopolymer, hexamethylcyclotr<strong>is</strong>iloxane<br />

(D 3 , Aldrich) was stirred over calcium hydride at 80 - 90 ◦ C for 4 hours<br />

after which it was degassed 3 times and further purified with dibutylmagnesium under<br />

vacuum for 1.5 hours. After initiation with sec-butyl lithium, at least 18 hours was required<br />

<strong>to</strong> assure the complete activation of D 3 in 10 vol % cyclohexane at 25 ◦ C and<br />

afterwards actual polymerization started upon adding an equal volume of tetrahydrofuran<br />

(THF). After 2 - 3 hours, the reaction was terminated with 10-fold molar excess of<br />

chlorotrimethylsilane (Aldrich). The conversion was approximately 40 <strong>to</strong> 45 %. The final<br />

solution was red<strong>is</strong>solved in cyclohexane after the majority of THF was removed. It was<br />

then immediately washed with 5 <strong>to</strong> 10 wt % sodium bicarbonate (NaHCO 3 ) aqueous solution,<br />

and then washed several times in deionized water. The products filtered with 0.2<br />

µm pore filter d<strong>is</strong>ks were further dried in a vacuum oven for two <strong>to</strong> three weeks in order <strong>to</strong><br />

122


Table 6.1: Material Character<strong>is</strong>tics of Homopolymers<br />

sample code M n (g/mol) a M w /M n N b η o (Pa·s) c γ o (mN/m) d<br />

PI1 2900 1.08 42 1.3 2.6<br />

PI2 4600 1.07 68 3.0 3.1<br />

PI3 6000 1.09 88 4.1<br />

PI4 6600 1.06 97 7.2 3.4<br />

PI5 7700 1.08 113 8.9 3.4<br />

PI6 9600 1.06 141 11.7 3.5<br />

PI7 15800 1.06 231 47.0 3.6<br />

PDMS1 6200 1.08 84 0.1<br />

PDMS2 8000 1.10 107<br />

a The number average molecular weights determined by GPC and converted using<br />

polystyrene standards. b Based on a referece volume v = 126.1 Å. c Zero shear v<strong>is</strong>cosity<br />

measured with parallel plate rheometry at 298K. d Bare interfacial tension with PDMS1.<br />

remove the residual materials such as solvent and unreacted D 3 monomer. The molecular<br />

character<strong>is</strong>tics of homopolymers are l<strong>is</strong>ted in Table 6.1.<br />

Poly(<strong>is</strong>oprene-b-dimethylsiloxane) (IDMS) block copolymer was polymerized by the<br />

sequential addition of <strong>is</strong>oprene and D 3 monomers in the same way described above for<br />

homopolymer, except <strong>that</strong> the IDMS solution was precipitated in a 3:1 vol mixture of<br />

methanol and 2-propanol. Before adding D 3 monomers for the second block, an aliquot<br />

of PI block was removed and quenched in methanol for analys<strong>is</strong>. The molecular character<strong>is</strong>tics<br />

of <strong>that</strong> block are summarized in Table 6.2. In order <strong>to</strong> synthesize pairs of block<br />

copolymers with the same length of PI, but PDMS blocks of different length, a small<br />

amount of the IDMS solution was cannulated out during the synthes<strong>is</strong> of the DMS block<br />

and terminated in a separate pre-flamed flask. The IDMS pairs made in th<strong>is</strong> way are<br />

IDMS2/IDMS9, IDMS4/IDMS5, IDMS6/IDMS7, and IDMS11/IDMS12. A small portion<br />

of PDMS homopolymer can be produced during the polymerization of IDMS, [16] but we<br />

did not identify a d<strong>is</strong>tinct peak associated with PDMS homopolymer in gel permeation<br />

chroma<strong>to</strong>graphy (GPC) curves. The sensitivity of th<strong>is</strong> GPC measurement <strong>to</strong> homopolymer<br />

<strong>is</strong> limited, however, by the fact <strong>that</strong> a small homopolymer peak can be buried in the tail<br />

123


Table 6.2: Material Character<strong>is</strong>tics of Block Copolymers<br />

sample code M PI (g/mol) a M w /M n fA b γ (0.1 %) c γ (0.2 %) γ (0.3 %) γ (0.4 %)<br />

IDMS1 4800 1.06 0.49 0.0011 0.0014 0.0013<br />

IDMS2 4400 1.15 0.50 0.0031<br />

IDMS3 4000 1.08 0.50 0.0007<br />

IDMS4 4800 1.09 0.53 0.0060<br />

IDMS5 4800 1.11 0.57 0.0094<br />

IDMS6 4900 1.07 0.58 0.0084<br />

IDMS7 4900 1.09 0.60 0.050 0.016 0.020<br />

IDMS8 5200 1.09 0.61 0.12 0.064<br />

IDMS9 4400 1.13 0.63 0.17 0.064<br />

IDMS10 4400 1.13 0.66 0.26 0.25 0.22<br />

IDMS11 4700 1.05 0.71 0.435 0.43<br />

IDMS12 4700 1.04 0.73 0.50 0.49<br />

IDMS13 9300 1.07 0.50<br />

a The number average molecular weights of PI blocks determined by GPC and converted<br />

using polystyrene standards. b The volume fraction of PDMS within IDMS. c Interfacial<br />

tension (mN/m) in the system of PI4/PDMS1/IDMS.<br />

of the IDMS peak due <strong>to</strong> the similar hydrodynamic volumes. No further fractionation was<br />

performed.<br />

The number (M n ) and weight (M w ) averaged molecular weights were determined with<br />

GPC in THF for PI based on the PS standards. Those of PDMS and IDMS were determined<br />

with GPC in <strong>to</strong>luene based on the refractive index increment dn/dc by way of both a<br />

refractive index detec<strong>to</strong>r and a light scattering detec<strong>to</strong>r. NMR analyses were used <strong>to</strong><br />

determine a block ratio f A = N CA /N C of IDMS, where N CA <strong>is</strong> a block length of PDMS<br />

block and N C <strong>is</strong> a <strong>to</strong>tal length of IDMS. Also, NMR was used <strong>to</strong> calculate 1,4 fractions of<br />

PI homopolymer and blocks which were at least 94 % and double check molecular weights<br />

of PDMS homopolymer. Hereafter a symbol A cons<strong>is</strong>tently stands for PDMS and B for<br />

PI.<br />

124


6.2.2 Sample Preparation<br />

IDMS block copolymers were premixed with pure PDMS homopolymer at room temperature<br />

(298K) at a concentration of 0.1 - 0.4 wt %, which <strong>is</strong> above the cmc (Fig. 6.3). A<br />

<strong>to</strong>tal volume of a prepared sample was 1 mL. The samples were kept in a vacuum oven<br />

until block copolymer d<strong>is</strong>appeared v<strong>is</strong>ually. <strong>Th<strong>is</strong></strong> <strong>to</strong>ok 4 <strong>to</strong> 5 hours at room temperature<br />

without mechanical stirring.<br />

6.2.3 Interfacial Tension Measurement<br />

The interfacial tension was measured with a spinning drop tensiometer (SDT). An electric<br />

mo<strong>to</strong>r rotates a horizontal glass tube containing a liquid drop in a denser matrix at a<br />

constant angular velocity. The PDMS matrix (density ρ A = 0.970 g/mol)[17] was poured<br />

in<strong>to</strong> a glass tube, a 6 mm inner diameter, 15 cm long. One end of the glass tube <strong>is</strong> sealed<br />

by a plug equipped with a set of stepped calibration posts. We used either of two different<br />

plugs, which contained terraced cylindrical posts <strong>that</strong> range from 0.3 mm <strong>to</strong> 0.5 mm and<br />

0.5 mm and 2 mm, respectively, depending on the expected final diameter of the drop.<br />

Vacuum was pulled through the glass tube for a few hours until no air bubbles came out<br />

of the gap between the O-rings around a plug and a glass wall. Then a pure PI drop (ρ B<br />

= 0.90 g/mol)[17] with an appropriate diameter was injected in the middle of the PDMS<br />

matrix using a glass pipette, and another plug with a threaded hole at the center was<br />

quickly pushed in<strong>to</strong> the glass tube until the first drop of the PDMS matrix was pushed out<br />

through the hole in the plug. The hole was then tightly sealed with a bolt and an O-ring<br />

before the PI drop could r<strong>is</strong>e significantly and the glass tube was loaded in<strong>to</strong> the SDT.<br />

In an attempt <strong>to</strong> reduce the effects of the slow diffusion of block copolymer in<strong>to</strong> the<br />

drop, interfacial tension measurements were carried out on samples <strong>that</strong> were s<strong>to</strong>red for<br />

one <strong>to</strong> four weeks after injection of the pure PI drop in<strong>to</strong> a PDMS/copolymer matrix.<br />

Assuming a cylindrical drop of length l and diameter d with l/d > 4 capped with<br />

125


hem<strong>is</strong>pherical ends, the interfacial tension γ can be calculated as [18]<br />

γ = 1<br />

32 ∆ρω2 d 3 (6.1)<br />

where ω <strong>is</strong> an angular velocity. The density difference <strong>is</strong> ∆ρ = ρ A − ρ B = 0.07 g/mol.<br />

The initial drop size and angular velocity was chosen for each sample so as <strong>to</strong> give l/d ∼ 5<br />

at a speed of 3000 <strong>to</strong> 6000 rpm. The rotation of the sample was synchronized with the<br />

frequency of a stroboscope in order <strong>to</strong> obtain still images of the rotating drop with a video<br />

camera. Since γ <strong>is</strong> proportional <strong>to</strong> d 3 , it was critical <strong>to</strong> get images of sharp drop edges.<br />

The stroboscope minimized the effect of off-centered movements of both the drop and<br />

calibration posts.<br />

During the measurements, both spin-up and spin-down experiments were carried out<br />

<strong>to</strong> confirm <strong>that</strong> the system had reached hydrodynamic equilibrium.[11, 12] Also, a few<br />

samples were run at two different angular velocities <strong>to</strong> confirm <strong>that</strong> the same value of γ<br />

was obtained at different rotation rates. Individual measurements were reproducible within<br />

ca. ± 5%. <strong>Th<strong>is</strong></strong> error <strong>is</strong> primarily due <strong>to</strong> the uncertainty in diameter measurement.<br />

The observation of an apparently time-independent drop diameter <strong>is</strong> not necessarily<br />

sufficient <strong>to</strong> prove <strong>that</strong> a tensiometer sample has reached thermodynamic equilibrium. The<br />

analys<strong>is</strong> presented in Chapter 5 shows <strong>that</strong>, under some conditions, it <strong>is</strong> possible for the<br />

concentration of copolymer at an interface, and the corresponding interfacial tension, <strong>to</strong><br />

remain at apparently time-independent values at times much less than <strong>that</strong> required <strong>to</strong><br />

establ<strong>is</strong>h a global thermodynamic equilibrium. An analys<strong>is</strong> of the kinetics of copolymer<br />

transport in th<strong>is</strong> experiment <strong>is</strong> given in Sec. 6.4.<br />

6.2.4 Small Angle X-Ray Scattering<br />

Small Angle X-ray scattering (SAXS) experiments were conducted using a 6m beamline<br />

at University of Minnesota.[19] Ternary blend samples of PDMS1/PI4/IDMS3 were prepared<br />

by adding various concentrations of block copolymers <strong>to</strong> 50:50 vol mixture of the<br />

126


homopolymers. Dilute block copolymer samples were mechanically stirred with a magnetic<br />

bar overnight. The concentrated samples were solution blended with benzene at a concentration<br />

of 20 wt % of polymers and vacuum dried for a few days at room temperature. All<br />

<strong>is</strong>otropic 2D scattering patterns were azimuthally integrated <strong>to</strong> produce 1D plots of the<br />

intensity I vs. the magnitude of the scattering wavenumber q = 4π/λ sin(θ/2) where θ <strong>is</strong><br />

the scattering angle and λ <strong>is</strong> the wavelength (1.54 Å) of the incident radiation.<br />

6.2.5 Dynamic Mechanical Spectroscopy<br />

The zero shear v<strong>is</strong>cosity η o of each homopolymer was measured with a TA instruments<br />

ARES rheometer using a steady rate sweep mode at room temperature with 25 mm parallel<br />

plates, a 1 mm gap, and 10 % strain.<br />

6.3 Results<br />

6.3.1 Bare Interfacial Tension<br />

Bare interfacial tension γ o between PI and PDMS1 in the absence of block copolymer was<br />

measured as a function of molecular weight of the PI homopolymer, denoted M PI . In the<br />

absence of copolymer, the drop diameter typically became independent of time within 20<br />

<strong>to</strong> 30 seconds in both spin-up and spin-down experiments as shown in Fig. 6.1 (a). Fig. 6.2<br />

shows γ o as a function of PI molecular weight. The interfacial tension reaches an apparent<br />

plateau value of 3.6 mN/m for the highest PI molecular weights. <strong>Th<strong>is</strong></strong> plateau value for γ o<br />

<strong>is</strong> in a reasonable agreement with the literature value (3.2 mN/m),[20] which was measured<br />

using commercial grade homopolymers with a density difference ∆ρ = 0.064 g/cm 3 , rather<br />

than 0.07 g/cm 3 obtained for our sample.<br />

The solid line in Fig. 6.2 <strong>is</strong> a SCFT prediction for th<strong>is</strong> series of samples. The SCFT<br />

calculations were conducted using the reported molecular weights and a Flory-Huggins<br />

parameter χ = 0.175 based on a reference volume v = 126.1 Å3 . <strong>Th<strong>is</strong></strong> value of χ was<br />

127


3<br />

1<br />

2.5<br />

0.8<br />

f A = 0.73 (IDMS12)<br />

D (mm)<br />

2<br />

1.5<br />

1<br />

PI6 (9.6K)<br />

PI1 (2.9K)<br />

D (mm)<br />

0.6<br />

0.4<br />

f A = 0.58 (IDMS6)<br />

0.5<br />

0.2<br />

0<br />

0 1 2 3 4 5<br />

t (min)<br />

(a)<br />

0<br />

0 100 200 300 400 500 600 700<br />

t (min)<br />

(b)<br />

Figure 6.1: The transient diameter of the PI drop in the PDMS matrix as a function of<br />

time. (a) PI1 and PI6 in the absence of block copolymer (b) PI4 in PDMS1/IDMS6 and in<br />

PDMS2/IDMS12. The concentrations of IDMS6 and IDMS12 in the corresponding PDMS<br />

matrices are 0.1 wt % and 0.2 wt %, respectively. The angular velocity <strong>is</strong> 6300 rpm.<br />

chosen <strong>to</strong> fit measurements of the initial decrease in interfacial tension with small amounts<br />

of added copolymer and the cmc observed in experiments <strong>that</strong> are reported in the following<br />

subsection. Our chosen value of χ = 0.175 <strong>is</strong> <strong>is</strong> slightly larger than the value χ = 0.146<br />

at T = 298K reported by Cochran et. al [19], which was obtained by fitting T ODT of two<br />

different symmetric PI-b-PDMS block copolymers with the equation χ = A/T + B and<br />

assuming (χN C ) ODT = 10.495.[21] SCFT predictions of the bare interfacial tension using<br />

χ = 0.175 underestimate the interfacial tension at high PI molecular weights by about 20<br />

%. Use of the lower literature value would make the d<strong>is</strong>crepancy slightly worse.<br />

6.3.2 Dependence on Copolymer Concentration<br />

Fig. 6.3 shows the reduction of interfacial tension γ with different block copolymers on a<br />

logarithmic scale. The dashed lines are just guides for eyes. When the highly asymmetric<br />

block copolymer IDMS12, with f A = 0.73, was added <strong>to</strong> PI4/PDMS2, γ became saturated<br />

128


5<br />

4<br />

γ (mN/m)<br />

3<br />

2<br />

1<br />

0<br />

2 4 6 8 10 12 14 16<br />

M PI (kg / mol)<br />

Figure 6.2: Bare interfacial tension between PI and PDMS1 in the absence of block copolymer<br />

as a function of molecular weights of PI. The solid line <strong>is</strong> the SCFT prediction using<br />

χ = 0.175.<br />

at 0.44 mN/m. <strong>Th<strong>is</strong></strong> <strong>is</strong> an 87 % reduction from the bare interfacial tension, which <strong>is</strong><br />

comparable <strong>to</strong> the maximum reduction found in previous studies. [10, 22]<br />

However, when the symmetric block copolymer IDMS1 was added <strong>to</strong> PI4/PDMS1,<br />

we obtained a strikingly low interfacial tension of approximately 0.0015 mN/m, at all<br />

concentrations above 0.05 wt % of block copolymer. In th<strong>is</strong> case, the interfacial tension <strong>is</strong><br />

reduced by more than three orders of magnitude. <strong>Th<strong>is</strong></strong> dramatic decrease has commonly<br />

been observed for balanced small molecule surfactants, and predicted theoretically for<br />

balanced surfactants of either large or small molecule, but had not been observed previously<br />

in a polymeric system.<br />

The value of copolymer concentration beyond which γ becomes independent of concentration<br />

<strong>is</strong> close <strong>to</strong> 0.5% for these two systems. <strong>Th<strong>is</strong></strong> apparent cmc <strong>is</strong> associated with the<br />

aggregation of copolymer in<strong>to</strong> self-assembled structures, though th<strong>is</strong> saturation of interfacial<br />

tension <strong>is</strong> expected whether the aggregates are spherical micelles, as appears likely for<br />

129


f A = 0.73, or a bicontinuous microemulsions, as appears <strong>to</strong> be the case for the symmetric<br />

copolymer.<br />

The predictions of SCFT for the cmc and the slope dγ/dφ c of interfacial tension vs<br />

concentration below the cmc are extremely sensitive <strong>to</strong> the choice of χ parameter. The cmc<br />

<strong>is</strong> known <strong>to</strong> vary as e −χN Cf B<br />

. Fig. 6.4 compares our results for IDMS1 <strong>to</strong> SCFT predictions<br />

for several slightly different χ values which are all defined using a reference volume v =<br />

126.1 Å3 . All of the other SCFT results presented in th<strong>is</strong> chapter use the value χ = 0.175<br />

chosen <strong>to</strong> fit th<strong>is</strong> data.<br />

In the presence of block copolymer, the transient behavior in the spinning drop tensiometer<br />

<strong>is</strong> much slower than in the absence of copolymer. Fig. 6.1 (b) shows representative<br />

transient behaviors of the diameter of the PI4 drop during each spin-up and spin-down<br />

experiment with 0.1 wt % of IDMS6 in PDMS1 and 0.2 wt % of IDMS12 in PDMS2,<br />

respectively. The time required for the drop <strong>to</strong> reach its final diameter <strong>is</strong> several hours<br />

at concentrations above the cmc, while less than a minute <strong>is</strong> required in the absence of<br />

copolymer. Allowing the apparatus <strong>to</strong> spin for 24 hours was found <strong>to</strong> be sufficient <strong>to</strong> obtain<br />

a time-independent value in all but a few samples <strong>that</strong> contain block copolymer. (The<br />

exceptions are are d<strong>is</strong>cussed below).<br />

6.3.3 Dependence on Copolymer Composition<br />

We <strong>have</strong> measured the interfacial tension at concentrations above the apparent cmc for each<br />

of the 12 PI-b-PDMS copolymers IDMS1-IDMS12 l<strong>is</strong>ted in Table 6.2. These copolymers<br />

all <strong>have</strong> PI blocks of similar size, corresponding <strong>to</strong> χN CB ≃ 12, but varying PDMS corona<br />

block. The volume fraction of PDMS in th<strong>is</strong> series varies from 0.49 <strong>to</strong> 0.73. All of these<br />

experiments were carried out with homopolymers PI4 and PDMS1. Measurements were<br />

taken with copolymer concentrations varying from 0.1 wt % <strong>to</strong> 0.4 wt % in order <strong>to</strong> confirm<br />

<strong>that</strong>, within th<strong>is</strong> range, the interfacial tension <strong>is</strong> generally independent of concentration.<br />

130


10 1 0 0.05 0.1 0.15 0.2 0.25 0.3<br />

10 0<br />

f A = 0.73 (IDMS12)<br />

γ (mN/m)<br />

10 -1<br />

10 -2<br />

10 -3<br />

f A = 0.49 (IDMS1)<br />

10 -4<br />

φ c (%)<br />

Figure 6.3: The interfacial tension reduction of PI4/PDMS1 and PI4/PDMS2 as a function<br />

of the concentration of IDMS1 and IDMS12, respectively. The dashed lines are guides <strong>to</strong><br />

the eye.<br />

10 1 0 0.05 0.1 0.15 0.2 0.25 0.3<br />

γ (mN/m)<br />

10 0<br />

10 -1<br />

10 -2<br />

χ=0.15<br />

0.17<br />

0.175<br />

0.18<br />

10 -3<br />

10 -4<br />

φ c (%)<br />

Figure 6.4: The interfacial tension reduction between PI4 and PDMS1 as a function of<br />

IDMS1. The solid lines are the SCFT prediction with four different χ’s.<br />

131


Fig. 6.5 show the saturated interfacial tension as a function of f A . Fig. 6.6 shows the<br />

same data using a logarithmic ax<strong>is</strong> for interfacial tension. For most copolymers, values<br />

are shown for more than one concentration. It should be apparent <strong>that</strong> γ <strong>is</strong> not a smooth<br />

function of f A . Starting from an ultra-low interfacial tension of order 10 −3 mN/m near<br />

the symmetric copolymer f A = 0.49 (IDMS1), γ increases relatively slowly <strong>to</strong> about 10 −2<br />

mN/m near f A = 0.6, and then changes by roughly an order of magnitude over a narrow<br />

range 0.6 < f A < 0.65. For f A > 0.65, γ again r<strong>is</strong>es smoothly with f A . <strong>Th<strong>is</strong></strong> seemingly<br />

d<strong>is</strong>continuous dependence on f A <strong>is</strong> more obvious in Fig. 6.6.<br />

The open circles in Fig. 6.5 are the results of SCFT predictions of γ for two systems.<br />

The required calculations were similar <strong>to</strong> the micelle simulations presented in chapter 4,<br />

in which we <strong>have</strong> determined the cmc from simulations of a swollen spherical micelle and<br />

used a separate SCFT simulation of a flat interface <strong>to</strong> determine the interfacial tension at<br />

the cmc. The solid line <strong>is</strong> the result of an expansion of the Helfrich theory for γ around a<br />

balance point f bal<br />

A , of the form γ = τ ′2<br />

2κ +<br />

(f A − f bal<br />

A )2 (6.2)<br />

in which we <strong>have</strong> used SCFT predictions for the balance point f bal<br />

A<br />

= 0.48, the bending<br />

rigidity κ + = 1.92kT, and for τ ′<br />

= ∂τ/∂f A = 0.270 kT Å−1 . <strong>Th<strong>is</strong></strong> simple parabolic<br />

expansion agrees extremely well with the results of the two micelle calculations shown by<br />

open circles. Leibler’s prediction[23] of interfacial tension, in which the micelle core was<br />

not allowed <strong>to</strong> swell, <strong>is</strong> also shown as a dotted line in the same plot for compar<strong>is</strong>on.<br />

For sufficiently asymmetric copolymers, with f A > 0.65, the SCFT predictions are<br />

remarkably cons<strong>is</strong>tent with the measured interfacial tensions. However, SCFT predictions<br />

of γ for f A < 0.6 are significantly higher than the measurements, particularly near the<br />

apparent d<strong>is</strong>continuity at 0.6 < f A < 0.65.<br />

In the transition region between 0.6 < f A < 0.65, it was very difficult <strong>to</strong> obtain<br />

132


γ (mN/m)<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.1 %<br />

0.2 %<br />

0.3 %<br />

0.4 %<br />

swollen micelle<br />

Leibler<br />

0<br />

0.3 0.4 0.5 0.6 0.7 0.8<br />

f A<br />

Figure 6.5: Interfacial tension in PI4/PDMS1/IDMS system as a function of the block<br />

ratio f A of IDMS. The quadratic dependence of interfacial tension on f A agrees well with<br />

the SCFT prediction and the swollen micelle theory (hollow dots) for f A > 0.65. Leibler’s<br />

theory[23] (a dotted line) predicts <strong>that</strong> a wider range of f A will cause van<strong>is</strong>hing interfacial<br />

tension.<br />

10 0 0.3 0.4 0.5 0.6 0.7 0.8<br />

10 -1<br />

γ (mN/m)<br />

10 -2<br />

10 -3<br />

10 -4<br />

0.1 %<br />

0.2 %<br />

0.3 %<br />

0.4 %<br />

swollen micelle<br />

Leibler<br />

f A<br />

Figure 6.6: Data and calculations of Fig. 6.5 plotted on a logarithmic scale. Near a balance<br />

point, measured values are much lower than those predicted by SCFT. The concentrations<br />

of block copolymer are all greater than the cmc.<br />

133


eproducible results, because the relaxation of the drop diameter became dramatically<br />

slower, and because the results depend on concentration. Fig. 6.7 shows the relaxation of<br />

the drop diameter for IDMS7 (f A = 0.6). In th<strong>is</strong> case, the drop diameter continued <strong>to</strong><br />

decrease throughout 3 days of observation, and the results depended significantly upon the<br />

concentration of copolymer. <strong>Th<strong>is</strong></strong> transient behavior <strong>is</strong> very different from <strong>that</strong> observed<br />

for either f A > 0.65 or f A < 0.6, as shown in Fig. 6.1 (b).<br />

0.6<br />

0.55<br />

0.5<br />

D (mm)<br />

0.45<br />

0.4<br />

0.1 wt %<br />

0.35<br />

0.3<br />

0.2 wt %<br />

0.25<br />

0 1000 2000 3000 4000<br />

t (min)<br />

Figure 6.7: The time-dependent diameter of a PI4 drop in a PDMS1 matrix at an angular<br />

velocity of 6300 rpm with two different concentrations of IDMS7 (f A = 0.6).<br />

6.3.4 Dependence on Homopolymer Molecular Weight<br />

An alternative way <strong>to</strong> control the spontaneous curvature of a monolayer <strong>is</strong> <strong>to</strong> change<br />

the relative lengths of the homopolymers. In th<strong>is</strong> subsection we examine the effect of<br />

varying molecular weight M PI of poly<strong>is</strong>oprene on interfacial tension. We <strong>have</strong> measured<br />

the interfacial tension between PDMS1 matrix <strong>that</strong> initially contains 0.1 wt % of the<br />

symmetric block copolymer IDMS1 and a series of PI homopolymers.<br />

The results are shown in Fig. 6.8. The results show an essentially quadratic dependence<br />

134


on the M PI , with a very low minimum in γ at a balance point between 6 K and 6.6 K. The<br />

solid lines are an approximate SCFT prediction, of the form γ = τ ′2<br />

2κ +<br />

(β − β bal ) 2 analogous<br />

<strong>to</strong> Eqn. 6.2, in which β = N PI /N PDMS and τ ′ ≡ ∂τ/∂β. The d<strong>is</strong>played prediction uses<br />

parameters M PI,bal = 6.2 K, κ + = 2.00 kT and τ ′ = ∂τ/∂β = 0.0237 kT Å−1 <strong>that</strong> were<br />

obtained from SCFT calculations similar <strong>to</strong> those presented in Chapter 3.<br />

6.3.5 A Larger Copolymer<br />

Another series of measurements was carried out with copolymer IDMS13, PDMS2 matrix,<br />

and PI homopolymers of varying molecular weight. IDMS13 <strong>is</strong> a symmetric copolymer<br />

with a molecular weight nearly twice <strong>that</strong> of IDMS1, and a PI block size roughly twice<br />

<strong>that</strong> of any of the other copolymers used in our experiments. IDMS13 was initially added<br />

the PDMS2 matrix at a concentration of 0.2 wt %.<br />

Results for the interfacial tension inferred from the final observed drop radii in experiments<br />

<strong>that</strong> were each run for several days are shown in Fig. 6.9 for systems with several<br />

different PI homopolymers. The minimum value of γ <strong>is</strong> much higher than observed with<br />

IDMS1, and overall agreement with SCFT predictions for the equilibrium interfacial tension<br />

<strong>is</strong> d<strong>is</strong>appointing. These results are not, however, reliable measurements of equilibrium<br />

interfacial tension, because drop diameter may not <strong>have</strong> reached its true steady state.<br />

In spinning drop experiments with IDMS13, equilibration of the drop size was dramatically<br />

slower than in experiments with the smaller symmetric copolymer. Fig. 6.10<br />

shows the time-dependence of the drop diameter in a system with PDMS2, IDMS13, and<br />

PI5. In th<strong>is</strong> example, the drop diameter appears <strong>to</strong> still be noticeably decreasing after the<br />

tensiometer has been left running for one week. The relaxation time for the drop radius<br />

in th<strong>is</strong> experiment appears <strong>to</strong> be at least 50 times greater than the relaxation time of a<br />

few hours observed in IDMS1. We find it difficult <strong>to</strong> understand th<strong>is</strong> dramatic difference<br />

in time scales on the bas<strong>is</strong> of the difference in diffusivities alone for two molecules whose<br />

135


molecular weights differ by only a fac<strong>to</strong>r of two. We do not understand the reason for th<strong>is</strong><br />

dramatic difference in equilibration times.<br />

6.3.6 Bicontinuous Microemulsion Phase<br />

Some systems containing symmetric diblock copolymers, like some systems of balanced<br />

small molecule surfactants, <strong>have</strong> been found <strong>to</strong> form a bicontinuous microemulsion phase.<br />

In order <strong>to</strong> determine whether th<strong>is</strong> also occurs in some of the systems studied here, we<br />

<strong>have</strong> studied the phase behavior of mixtures of the symmetric homopolymer IDMS1 with<br />

equal volume fractions of PI4 and IDMS3, as well as mixtures of the larger symmetric<br />

copolymer, IDMS13 with equal volume of PI5 and PDMS2. These systems correspond <strong>to</strong><br />

the minima of the parabolic tension curves in Fig. 6.8 and Fig. 6.9.<br />

Pho<strong>to</strong>graphs of the samples are shown in Fig. 6.11 and Fig. 6.12, respectively. Mixtures<br />

of PDMS1/PI4/IDMS3 were turbid at block copolymer concentrations up <strong>to</strong> 25 wt %,<br />

indicating phase separation in<strong>to</strong> two or three coex<strong>is</strong>ting phases. The sample with 32<br />

wt % IDMS1 was slightly blu<strong>is</strong>h, which <strong>is</strong> a character<strong>is</strong>tic signal of the formation of a<br />

uniform bicontinuous microemulsion phase µE with internal structures large enough <strong>to</strong><br />

induce Rayleigh scattering. At higher concentrations, samples of th<strong>is</strong> mixture become<br />

<strong>to</strong>tally transparent. A similar sequence <strong>is</strong> observed for PDMS2/PI5/IDMS13. In th<strong>is</strong> case,<br />

however, the 10 wt % <strong>is</strong> turbid, and 20 wt % sample <strong>is</strong> blu<strong>is</strong>h. These observations are<br />

cons<strong>is</strong>tent with the ex<strong>is</strong>tence of two or three phase coex<strong>is</strong>tence of a microemulsion with<br />

PDMS and/and PI rich phases at low copolymer concentrations ( below 25 % for IDMS3<br />

and 10 % for IDMS13), the formation of a microemulsion ( near 32 wt % for IDMS3), and<br />

the formation of an optically clear lamellar phase at higher concentrations. <strong>Th<strong>is</strong></strong> <strong>is</strong> the<br />

normal sequence for a system <strong>that</strong> forms a balanced bicontinuous microemulsion.<br />

We <strong>have</strong> also carried out small angle X-ray experiments on these samples. One dimensional<br />

SAXS intensities are shown as a function of scattering wavenumber q in Fig. 6.13<br />

136


10 1<br />

10 0<br />

2 4 6 8 10<br />

γ (mN/m)<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

M PI (kg/mol)<br />

Figure 6.8: Interfacial tension in the system of PDMS1/PI/IDMS1 as a function of molecular<br />

weight of PI. The concentration of IDMS1 <strong>is</strong> 0.1 wt %.<br />

10 1<br />

10 0<br />

2 4 6 8 10<br />

γ (mN/m)<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

M PI (kg/mol)<br />

Figure 6.9: Interfacial tension in the system of PDMS2/PI/IDMS13 as a function of molecular<br />

weight of PI. The concentration of IDMS13 <strong>is</strong> 0.2 wt %. The minimum of the interfacial<br />

tension <strong>is</strong> not ultra-low, but much greater than <strong>that</strong> in Fig. 6.8.<br />

137


1<br />

0.8<br />

D (mm)<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 2000 4000 6000 8000 10000<br />

t (min)<br />

Figure 6.10: The variation of the diameter of the PI5 drop in the PDMS2 matrix at a<br />

concentration of 0.2 wt % of IDMS13. The rotational speed <strong>is</strong> 6300 rpm.<br />

and Fig. 6.14 for both sets of mixtures. In both systems, at high concentrations, we see<br />

the scattering peak character<strong>is</strong>tic of a lamellar phase. In Fig. 6.13, as the concentration of<br />

IDMS3 decreases, the peak wavenumber q ∗ decreases until φ c = 0.59. The corresponding<br />

variation of the domain spacings ξ = 2π/q ∗ with copolymer concentration <strong>is</strong> shown given<br />

in Fig. 6.15. Then another peak suddenly appears in front of a primary lamellar peak,<br />

cons<strong>is</strong>tent with the formation of a bicontinuous microemulsion phase coex<strong>is</strong>ting with a<br />

lamellar phase. Within a range of between 0.49 < φ c < 0.58, it appears <strong>that</strong> neither peak<br />

position changes with concentration, which we interpret as two-phase coex<strong>is</strong>tence between<br />

a lamellar and microemulsion phase. The ratio of the domain spacing of a bicontinuous<br />

microemulsion phase <strong>to</strong> <strong>that</strong> of a lamellar phase (ξ µE /ξ L ) <strong>is</strong> 1.1 as compared in Fig. 6.16.<br />

Ideally th<strong>is</strong> ratio should be 1.5 in case of the structure with the perfect minimal surface.[24]<br />

In fact, these double peaks also appear with an ideal ratio of ξ µE /ξ L = 1.5 in Fig. 6.14<br />

at the lower concentration of IDMS13 (φ c ∼ 0.45). Since the primary peaks for φ c <<br />

138


Figure 6.11: The ternary <strong>is</strong>opleth blends of the system PDMS1/PI4/IDMS3 for the various<br />

concentrations of block copolymer: 5, 25, 32, 45, 57, and 70 wt % from the leftmost vial.<br />

The samples are turbid below 30 wt % due <strong>to</strong> the macrophase separation.<br />

Figure 6.12: The ternary <strong>is</strong>opleth blends of the system PDMS2/PI5/IDMS13 for the various<br />

concentrations of block copolymer: 10, 20, 30, 40, 50, and 70 wt % from the leftmost<br />

vial. The sample with 20 wt % of IDMS13 turns blu<strong>is</strong>h.<br />

139


10 −4 I (a.u.)<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

φ c<br />

0%<br />

5%<br />

10%<br />

20%<br />

30%<br />

32%<br />

35%<br />

40%<br />

48%<br />

50%<br />

55%<br />

57%<br />

60%<br />

70%<br />

80%<br />

90%<br />

100%<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

q (Å −1 )<br />

Figure 6.13: The SAXS profiles for the ternary <strong>is</strong>opleth blends of PDMS1/PI4/IDMS3 as a<br />

function of the concentration of IDMS3. All curves are presented on arbitrary linear scales<br />

for clarity. The primary peak continuously moves <strong>to</strong> a small q region as the mixtures are<br />

diluted with the homopolymers. Two peaks are d<strong>is</strong>tinct between 50 % and 55 % indicating<br />

a bicontinuous microemulsion phase coex<strong>is</strong>t with a lamellar phase.<br />

14<br />

φ c<br />

10 −4 I (a.u.)<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

10%<br />

20%<br />

30%<br />

35%<br />

40%<br />

45%<br />

48%<br />

50%<br />

60%<br />

70%<br />

80%<br />

90%<br />

100%<br />

0 0.01 0.02 0.03 0.04 0.05 0.06<br />

q (Å −1 )<br />

Figure 6.14: The SAXS profiles for the ternary <strong>is</strong>opleth blends of PDMS2/PI5/IDMS13<br />

as a function of the concentration of IDMS13. The primary peaks of IDMS13 appear at<br />

lower q than those of IDMS3 in Fig. 6.13 at the same concentration due <strong>to</strong> the twice longer<br />

block copolymer.<br />

140


0.4 are hidden inside the low q intensity, further analys<strong>is</strong> of the morphology transition<br />

<strong>is</strong> not possible. However, a bicontinuous microemulsion phase unambiguously forms with<br />

IDMS13 even though interfacial tension <strong>is</strong> not ultra-low. Returning <strong>to</strong> Fig. 6.13, if the<br />

concentration of the homopolymers increases in the blends, a uniform bicontinuous microemulsion<br />

phase <strong>is</strong> encountered between 0.31 < φ c < 0.49, and finally a wide region of<br />

three phase coex<strong>is</strong>tence, denoted 3 µE , opens for φ c < 0.3. A 3 µE phase can be identified<br />

by both the turbid blends and a constant peak position at q ∗ = 0.17 Å −1 as clearly shown<br />

in Fig. 6.11 and Fig. 6.12.<br />

A macroscopic two phase region could not be verified here, but it should be stable<br />

below the cmc. Surpr<strong>is</strong>ingly, the proposed diagram of the system PDMS1/PI4/IDMS3<br />

in Fig. 6.17 agrees very well with the theoretical diagram for a small molecule surfactant<br />

system with κ = kT by Daicic et al. [25] In addition, the semiquantitative relationship,<br />

γ ξ 2 ≃ kT, between the domain spacing and interfacial tension in the small molecule<br />

surfactant system,[3, 4] <strong>is</strong> also valid in th<strong>is</strong> system using ξ µE = 52.5 nm in a 3 µE region<br />

and γ ≃ 0.0007 mN/m.<br />

In contrast <strong>to</strong> the previous literature of the ternary polymer blends [7, 8, 9, 11, 26]<br />

in which a uniform bicontinuous microemulsion phase was observed at low concentration<br />

of φ c ≃ 0.1 in the limit of unbinding transition of a lamellar phase,[7] the present system<br />

has a denser microemulsion phase at a much higher concentration of φ c = 0.30 without<br />

unbinding transition. Such a uniform bicontinuous microemulsion phase <strong>is</strong> often found at<br />

the similar concentration in the nonionic surfactant system depending on the strength of the<br />

surfactants.[27] We conjecture the different location of the Lifshitz point given by (φ c ) L =<br />

2α 2 /(1+2α 2 ) and (χN c ) L = 2(1+2α 2 )/α <strong>is</strong> implicitly related <strong>to</strong> those results. Within the<br />

mean field approximation[28] the Lifshitz point are expected <strong>to</strong> be (φ c ) L = 0.08 for α = 0.2<br />

in the most literature, but (φ c ) L = 0.47 for α = 0.66 in our system. Although our system <strong>is</strong><br />

far from the Lifshitz point (χN c ) L = 5.7, these predictions are reasonably cons<strong>is</strong>tent with<br />

141


60<br />

50<br />

40<br />

ζ (nm)<br />

30<br />

20<br />

10<br />

ζ µE<br />

ζ L<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

φ c<br />

Figure 6.15: The variation of the domain spacing of a lamellar phase (ξ L ) and a bicontinuous<br />

microemulsion phase (ξ µE ) along the <strong>is</strong>opleth in the system of PDMS1/PI4/IDMS3.<br />

1<br />

0.8<br />

ζ L<br />

ζ Lo<br />

/ ζ ∗<br />

0.6<br />

0.4<br />

ζ µE<br />

0.2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

φ c<br />

Figure 6.16: The ratio of the pure lamellar spacing ((ξ Lo ) <strong>to</strong> the spacing at q = q ∗ multiplied<br />

by a weight fac<strong>to</strong>r α as a function of the concentration of IDMS3. ξ µE /ξ L = 1.1.<br />

142


2 φ 3 µE µE µE<br />

+<br />

L<br />

L<br />

0 10 20 30 40 50 60 70 80 90 100<br />

φ c (%)<br />

Figure 6.17: The proposed phase diagram of the system PDMS1/PI4/IDMS3 in Fig. 6.13.<br />

Large three phase region ex<strong>is</strong>ts below 30 % of block copolymer.<br />

the experimental data. Moreover, in case of the system PDMS2/PI5/IDMS13, (φ c ) L = 0.26<br />

for α = 0.42 so <strong>that</strong> a bicontinuous microemulsion phase could be located at more dilute<br />

regime.<br />

6.4 Analys<strong>is</strong> of Kinetics<br />

In th<strong>is</strong> section, we analyze the kinetics of copolymer diffusion in the spinning drop tensiometer,<br />

in order <strong>to</strong> estimate the time necessary for the copolymer chemical potential <strong>to</strong><br />

reach its equilibrium value at the PI/PDMS interface. The analys<strong>is</strong> <strong>is</strong> based on <strong>that</strong> of<br />

Chapter 5. That analys<strong>is</strong> suggests <strong>that</strong> the copolymer chemical potential at the interface<br />

of the drop should reach its equilibrium value of µ I c , corresponding <strong>to</strong> the cmc in PDMS,<br />

either if the copolymer has time <strong>to</strong> diffuse <strong>to</strong> the center of the drop and fill the drop <strong>to</strong><br />

its equilibrium concentration, or if there <strong>is</strong> no exclusion zone in the PDMS phase near<br />

the interface even at early times, in which case the chemical potential at the interface <strong>is</strong><br />

expected <strong>to</strong> reach µ cmc in a time comparable <strong>to</strong> <strong>that</strong> required <strong>to</strong> establ<strong>is</strong>h an interfacial<br />

143


monolayer.<br />

6.4.1 Estimate of Diffusion Time<br />

We first estimate the time for free copolymer <strong>to</strong> diffuse <strong>to</strong> the center of the PI drop. <strong>Th<strong>is</strong></strong><br />

requires an estimate of the diffusion coefficient D f of block copolymer in the PI matrix.<br />

At the low molecular weights used in our experiments, we may use the Rouse model <strong>to</strong><br />

predict the dependence of diffusivity on molecular weight. We thus require estimates of<br />

the monomer friction coefficient for the PI and PDMS blocks of IDMS copolymer in the<br />

PI matrix. We are not aware of any direct measurement of friction coefficient for a PDMS<br />

tracer in PI. The reasoning used <strong>to</strong> estimate th<strong>is</strong> friction coefficient, which <strong>is</strong> based on<br />

a review of the literature on friction coefficient in mixtures, <strong>is</strong> d<strong>is</strong>cussed in Appendix B.<br />

Briefly, we assume <strong>that</strong> the segmental dynamics of IDMS block copolymer <strong>is</strong> governed by<br />

the solvent (homopolymer) v<strong>is</strong>cosity rather than the friction coefficient ζ i of each block,<br />

because th<strong>is</strong> assumption has been found <strong>to</strong> provide a reasonable estimate in systems <strong>that</strong><br />

are chemically similar <strong>to</strong> PI and PDMS. The monomer friction coefficient PI homopolymer<br />

at 298 K has been reported <strong>to</strong> be [29] ζ PI = 1.5×10 −6 dyn·s/cm for a reference volume of 99<br />

cm 3 /mol. Because the friction coefficient for pure PI (i.e., a PI tracer in the PI matrix) <strong>is</strong> of<br />

order 10 2 times larger than of pure PDMS, we assume <strong>that</strong> the friction coefficient for PDMS<br />

in PI <strong>is</strong> less than or equal <strong>to</strong> <strong>that</strong> of PI in PI. With th<strong>is</strong> assumption, the tracer diffusion<br />

coefficient of block copolymer in the PI homopolymer must lie between <strong>that</strong> calculated by<br />

ignoring the friction of the PDMS block, giving D o = D PI = kT/Nζ PI ≃ 5 × 10 −10 cm 2 /s,<br />

where N corresponds <strong>to</strong> the length of PI block (ca. 4800 g/mol) and <strong>that</strong> obtained taking<br />

the monomer friction coefficient of PDMS <strong>to</strong> equal <strong>that</strong> of PI, which yields a diffusivity<br />

roughly half as large.<br />

The diffusion coefficient of PDMS can be directly obtained in ref. [30] or [31] as<br />

D PDMS = 1 × 10 −7 cm 2 /s for M PDMS = 5200 g/mol.<br />

144


The time for copolymer <strong>to</strong> diffuse <strong>to</strong> the center of an essentially cylindrical droplet of<br />

diameter d <strong>is</strong> given by<br />

t = d 2 /4kD PI<br />

f (6.3)<br />

A numerical fac<strong>to</strong>r k = 5.8 <strong>is</strong> obtained by identifying t with the relaxation time of the<br />

slowest decaying mode in an eigenmode expansion of the concentration field in a cylindrical<br />

geometry. For th<strong>is</strong> diffusion problem the concentration at the edge of the drop <strong>is</strong> held<br />

constant and then the eigenmodes are zeroth order Bessel functions. The final radius<br />

of a drop in a spinning drop tensiometer depends upon its interfacial tension, and <strong>is</strong><br />

substantially different for samples with widely differing interfacial tensions. Assuming the<br />

diffusion coefficient D PI<br />

f<br />

of block copolymer in the PI phase as D PI<br />

f<br />

= D o (1 − f A ) with<br />

increasing PDMS blocks, Fig. 6.18 shows the character<strong>is</strong>tic time t as a function of γ. <strong>Th<strong>is</strong></strong><br />

curve <strong>is</strong> obtained by combining Eqn. 6.3 with Eqn. 6.1 for the relationship between γ and<br />

drop diameter d = 2R for the current system. Fig. 6.18 demonstrates <strong>that</strong> only nearly<br />

symmetric block copolymers, which produce γ < 0.01 mN/m, corresponding <strong>to</strong> f A < 0.6,<br />

can diffuse <strong>to</strong> the center of the drop in less than about 10 hours.<br />

6.4.2 Can Micelles Reach the Interface?<br />

We now consider the question of whether the concentration of free copolymer in PDMS near<br />

the PI/PDMS interface reaches the cmc during the early stages of diffusion, at times much<br />

less than time required <strong>to</strong> fill the drop with copolymer, or whether it remains depressed<br />

below the cmc at early times, leading <strong>to</strong> the formation of an exclusion zone with no micelles<br />

near the interface. According <strong>to</strong> the analys<strong>is</strong> given in Chapter 5, the choice between these<br />

scenarios <strong>is</strong> determined by the value of a parameter<br />

√<br />

Df<br />

PI c PDMS<br />

c<br />

Q =<br />

Dm<br />

PDMS c 0 mK<br />

defined in Eqn. 5.6, in which D PI<br />

f<br />

D PDMS<br />

m<br />

(6.4)<br />

<strong>is</strong> the diffusivity of free copolymer in the PI drop,<br />

<strong>is</strong> the diffusivity of a micelle in the PDMS matrix, c PDMS<br />

c<br />

145<br />

<strong>is</strong> the critical micelle


10 3 0 0.1 0.2 0.3 0.4 0.5 0.6<br />

10 2<br />

t (hr)<br />

10 1<br />

10 0<br />

γ (mN/m)<br />

Figure 6.18: Time t <strong>to</strong> reach a center of spinning drop tensiometer drop by molecular<br />

diffusion as a function of interfacial tension γ.<br />

concentration in PDMS, and c o m <strong>is</strong> the initial concentration of copolymers in micelles within<br />

PDMS, given by c o m = co −c PDMS<br />

c , where c o <strong>is</strong> the <strong>to</strong>tal initial concentration of copolymers.<br />

<strong>Th<strong>is</strong></strong> dimensionless number <strong>is</strong> the ratio of the flux of copolymer <strong>that</strong> leaks in<strong>to</strong> the PI drop,<br />

calculated by assuming <strong>that</strong> the concentration of free copolymer in PDMS <strong>is</strong> held at the<br />

cmc, <strong>to</strong> the maximum flux of copolymer <strong>that</strong> can be provided by copolymer diffusion,<br />

which <strong>is</strong> the flux of micelles <strong>to</strong> an absorbing boundary. If Q < 1, micellar diffusion alone<br />

<strong>is</strong> sufficient <strong>to</strong> supply the required flux in<strong>to</strong> the drop, and micelles ex<strong>is</strong>t throughout the<br />

PDMS matrix. If Q > 1, the concentration of free copolymer at the interface must be<br />

depressed below the cmc <strong>to</strong> match the flux in<strong>to</strong> the drop, creating a micelle-free region<br />

near the interface at early times.<br />

The partition coefficient K ≡ φ A c /φB c<br />

for the equilibrium d<strong>is</strong>tribution of d<strong>is</strong>solved<br />

copolymer between A and B homopolymer phases can be derived from the Flory-Huggins<br />

146


theory, which yields<br />

ln K ≃ χN c (f A − f B ) + N C<br />

( 1<br />

N A<br />

− 1<br />

N B<br />

)<br />

. (6.5)<br />

The partition coefficient <strong>is</strong> shown in 6.19 for a series of systems, similar <strong>to</strong> those studied<br />

here, in which χN C f A = 12 and the symmetric homopolymers <strong>have</strong> the length N A = N B =<br />

Nf A equal <strong>to</strong> <strong>that</strong> of the core block. The partition function <strong>is</strong> mainly an exponential<br />

function of incompatibility difference between two blocks so <strong>that</strong> K becomes huge for<br />

asymmetric copolymer.<br />

The ratio c int<br />

f /co of interfacial concentration <strong>to</strong> bulk concentration shown in Fig. 6.19<br />

<strong>is</strong> calculated using Eqn. 5.2, and <strong>is</strong> <strong>that</strong> which would be obtained in the absence of<br />

micelles. <strong>Th<strong>is</strong></strong> was calculated using an estimated ratio of copolymer tracer diffusivities<br />

Df<br />

PDMS /Df<br />

PI<br />

= 200. For f 0.55, th<strong>is</strong> ratio very closely approaches the unity because<br />

the partition coefficient K becomes so large, indicating <strong>that</strong> the copolymer <strong>is</strong> nearly insoluble<br />

in the PI drop. For f = 0.5, where K = 1, the predicted ratio c int<br />

f /co ≃ 0.93 remains<br />

high because the diffusivity <strong>is</strong> much higher in the matrix than in the drop. <strong>Th<strong>is</strong></strong> plot <strong>is</strong><br />

directly relevant, however, only at bulk concentrations below the cmc.<br />

The diffusion coefficient of a micelle has been estimated from the S<strong>to</strong>kes-Einstein equation<br />

D m = kT/6πη I oR, using estimated micelle hydrodynamic radius R from SCFT calculations.<br />

Results for the radii of swollen micelles in two-phase ternary systems and of<br />

unswollen micelles in binary mixtures are shown on the left ax<strong>is</strong> in Fig. 6.20. The estimates<br />

are taken from the series of systems shown in Fig. 4.6, for systems in which<br />

f B χN = 12 with the homopolymers of equal length N A = N B = fN C , which <strong>is</strong> similar<br />

<strong>to</strong> the PI/PDMS/IDMS1 system. The predicted radius of a swollen micelle diverges<br />

as f A approaches 1/2, causing the predicted micelle diffusivity <strong>to</strong> van<strong>is</strong>h. The assumption<br />

<strong>that</strong> the copolymer forms spherical micelles <strong>is</strong>, however, dubious in th<strong>is</strong> limit, where the<br />

equilibrium structure appears <strong>to</strong> be a bicontinuous microemulsion phase.<br />

147


1<br />

10 12<br />

0.98<br />

10 10<br />

10 8<br />

/ c o<br />

c f<br />

I,int<br />

0.96<br />

10 6<br />

K<br />

0.94<br />

10 4<br />

0.92<br />

10 2<br />

0.9<br />

0.5 0.55 0.6 0.65 0.7 0.75 100<br />

f A<br />

Figure 6.19: The time-independent interfacial concentration and the partition coefficient as<br />

a function of the block ratio f A in the block copolymer. The effect of the micelle diffusion<br />

on c int<br />

f<br />

<strong>is</strong> not considered.<br />

1000<br />

900<br />

unswollen micelle<br />

swollen micelle<br />

10 1<br />

10 0<br />

R (Å)<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

10 −1<br />

10 −2<br />

10 −3<br />

10 −4<br />

10 −5<br />

Q<br />

200<br />

10 −6<br />

100<br />

0.5 0.55 0.6 0.65 0.7 0.75 10−7<br />

f A<br />

Figure 6.20: The radii R of unswollen/swollen micelles and the dimensionless ratio Q with<br />

respect <strong>to</strong> f A . R <strong>is</strong> taken from the micelle simulation with f B χN c = 12 in Fig. 4.6.<br />

148


Our results for Q as a function of f A are shown in Fig. 6.20, for systems with χNf A = 12<br />

and N A = N B = f A N C . Results for swollen and unswollen micelles were obtained using<br />

micelle diffusivities calculated based on the corresponding micelle radii shown in the same<br />

figure. For unswollen micelles, Q ≪ 1 for all values of f A . <strong>Th<strong>is</strong></strong> <strong>is</strong> in part a reflection of the<br />

fact <strong>that</strong> diffusivity D PDMS<br />

m<br />

of unswollen micelles in PDMS <strong>is</strong> actually somewhat higher<br />

than the free copolymer tracer diffusivity D PI<br />

f<br />

in PI. For swollen micelles, the diverging<br />

micelle radius causes the micelle diffusivity <strong>to</strong> approach zero, and thus causes Q <strong>to</strong> diverge<br />

as f A → 1/2. Despite th<strong>is</strong> divergence, however, we find Q > 1 only for a very narrow<br />

range 0.5 < f A < 0.501, and Q < 1 everywhere else. For f A ≃ 0.6, where we observe an<br />

apparent d<strong>is</strong>continuity in γ as a function of f A , K ∼ 10 3 , indicating <strong>that</strong> the copolymer<br />

has become essentially insoluble in PI, and Q < 10 −3 as a result. For f A ≃ 0.5, where th<strong>is</strong><br />

model yields Q < 1 over a very narrow range of values, it appears <strong>that</strong> the hydrodynamic<br />

equilibrium diameter in the spinning drop tensiometer becomes small enough <strong>that</strong> there<br />

actually should be time for copolymer <strong>to</strong> diffuse <strong>to</strong> the center of the drop within a few hours.<br />

In th<strong>is</strong> case, however, the whole analys<strong>is</strong> become dubious, because we do not know what<br />

kind of aggregate the copolymer forms within the PDMS phase near the interface. The<br />

above analys<strong>is</strong> suggests <strong>that</strong> all measured interfacial tensions in th<strong>is</strong> experiment represent<br />

the equilibrium values.<br />

6.4.3 Experimental Strategy<br />

From the above results we can propose the general experimental strategy <strong>to</strong> reach, or at<br />

least mimic, a global equilibrium during experiments.<br />

First, block copolymer should be always added <strong>to</strong> the matrix for two reasons: <strong>to</strong> keep<br />

the initial concentration of block copolymer constant due <strong>to</strong> the much larger volume of<br />

the matrix as addressed by Hu et al. [10] and <strong>to</strong> reduce the diffusional d<strong>is</strong>tance of block<br />

copolymer than the opposite case. As described in Experiment section, we premixed IDMS<br />

149


lock copolymer with the PDMS matrix.<br />

Second, block copolymer should favor the initial phase (i.e., matrix) where micelles<br />

preferentially form. We could simply sat<strong>is</strong>fy th<strong>is</strong> condition either by using block copolymer<br />

with the longer PDMS block or by varying the molecular weight of the PI homopolymer.<br />

Third, it <strong>is</strong> recommended <strong>that</strong> both homopolymer phases, especially the matrix, <strong>have</strong><br />

low v<strong>is</strong>cosity in order <strong>to</strong> allow the fast diffusion of block copolymer, and block copolymer<br />

can be added as much as possible so as <strong>to</strong> reduce the value of Q via c c /c o m.<br />

Fourth, in order <strong>to</strong> systematically control the spontaneous curvature of a block copolymer<br />

monolayer, the length of the core (PI) block can be fixed <strong>to</strong> keep the cmc constant<br />

while sequentially increasing the corona (PDMS) block.<br />

6.5 D<strong>is</strong>cussion and Conclusion<br />

We <strong>have</strong> used a spinning drop tensiometer <strong>to</strong> measure interfacial tensions of PI/PDMS<br />

interfaces with adsorbed IDMS diblock copolymers. We <strong>have</strong> characterized three different<br />

types of systems: one with copolymers of varying compositions and nearly symmetric<br />

homopolymers, and two others with PI homopolymers of varying molecular weight and<br />

a nearly symmetric copolymer. For all of the systems studied, the IDMS copolymer <strong>is</strong><br />

expected <strong>to</strong> be more soluble in the PDMS matrix, and <strong>to</strong> thus form micelles in PDMS at a<br />

lower chemical potential of copolymer than <strong>that</strong> required <strong>to</strong> form micelles in the PI drop.<br />

The copolymer was always added <strong>to</strong> the PDMS matrix.<br />

In these systems, the tracer diffusivity of the copolymer <strong>is</strong> estimated <strong>to</strong> be of 200 times<br />

higher in the PDMS matrix than in the PI drop.<br />

Measured interfacial tensions at copolymer concentration above the apparent cmc <strong>have</strong><br />

been compared <strong>to</strong> SCFT predictions <strong>that</strong> assume swollen spherical micelles in the PDMS<br />

matrix coex<strong>is</strong>t with an excess PI phase. These predictions agreed extremely well with<br />

our measurements for more asymmetric copolymers, with 0.65 < f A < 0.73. In a system<br />

150


with a symmetric copolymer, with f A ≃ 0.5, we observe an ultra-low interfacial tension,<br />

γ ≃ 10 −3 mN/m. <strong>Th<strong>is</strong></strong> behavior <strong>is</strong> cons<strong>is</strong>tent with theoretical predictions of van<strong>is</strong>hing<br />

or very low interfacial tensions for symmetric systems, and with many observations of<br />

ultra-low tensions in balanced small molecule systems. However, it had not previously<br />

been directly observed in a polymeric system. For slightly asymmetric copolymers, with<br />

0.5 < f A < 0.6, the observed dependence of γ on increasing f A <strong>is</strong> much less than predicted<br />

by the SCFT described above. Notably, the measured interfacial tension seems <strong>to</strong> change<br />

d<strong>is</strong>continuously with f A at a value near f A = 0.6, where measured values change by roughly<br />

an order of magnitude. Within the narrow range of 0.6 < f A < 0.65, we were unable <strong>to</strong><br />

obtain a reliable value of equilibrium interfacial tension, because the drop diameter was<br />

found slowly decrease over very long times, in a manner <strong>that</strong> depends upon copolymer<br />

concentration.<br />

Experiments in which we varied the molecular weight of PI homopolymer in mixtures<br />

containing a symmetric copolymer yielded very different results for two symmetric copolymers<br />

with significantly different <strong>to</strong>tal molecular weights. In experiments <strong>that</strong> used the<br />

lower molecular weight copolymer (IDMS1) a very low tension was observed at the balance<br />

point and a dependence of γ on PI molecular weight <strong>that</strong> agreed well with the SCFT<br />

predictions.<br />

Experiments with a higher molecular weight copolymer (IDMS13) showed a dramatically<br />

slower relaxation of the drop size than observed with IDMS1, making it impossible<br />

for us <strong>to</strong> obtain reliable measurements of equilibrium interfacial tension. It <strong>is</strong> not obvious<br />

at present why the relaxation of drop size <strong>is</strong> so much slower for the larger copolymer. The<br />

difference seems <strong>to</strong> be <strong>to</strong>o large <strong>to</strong> be accounted for by a difference in a tracer diffusivity<br />

alone.<br />

To determine structure and phase behavior of symmetric mixtures, we also <strong>examined</strong><br />

two sets of bulk mixtures with equal volumes of PI and PDMS homopolymers mixed with<br />

151


a symmetric copolymer, over a wide range of copolymer concentrations. The results of<br />

SAXS experiments and optical observations both suggest the ex<strong>is</strong>tence of a bicontinuous<br />

microemulsion <strong>that</strong> (for the mixtures with IDMS1) coex<strong>is</strong>ts with phases nearly pure PI<br />

and/or PDMS homopolymer at copolymer concentration up <strong>to</strong> about 30 % copolymer.<br />

6.5.1 Equilibrium vs. Nonequilibrium<br />

In th<strong>is</strong> subsection we d<strong>is</strong>cuss the possible equilibrium and nonequilibrium fac<strong>to</strong>rs <strong>that</strong><br />

might cause the seemingly d<strong>is</strong>continuous change of γ in the spinning drop tensiometer.<br />

First, we consider possible explanations based on equilibrium phase behavior. As previously<br />

mentioned, controlling the block ratio of copolymer <strong>is</strong> analogous <strong>to</strong> varying temperature<br />

in the nonionic surfactant system. Thus if we gradually decrease the block ratio<br />

from 0.73 <strong>to</strong> 0.5 at a copolymer concentration slightly greater than the cmc, it corresponds<br />

<strong>to</strong> increasing temperature from T < T l where T l <strong>is</strong> a lower critical end temperature. At<br />

a certain point corresponding <strong>to</strong> T = T l , we expect the three phase body <strong>to</strong> continuously<br />

open as the spontaneous curvature of copolymer becomes sufficiently small. In fact, the<br />

opening of a third phase does not cause d<strong>is</strong>continuous change in interfacial tension at T l<br />

or T u in nonionic surfactant systems. There are several reasons for th<strong>is</strong>. The chemical<br />

potentials of all three components in a three phase body vary continuously with T or f A in<br />

th<strong>is</strong> experiment, even through a critical end point. The interfacial tension of an oil-water<br />

interface with adsorbed surfactant thus also changes continuously. Only a d<strong>is</strong>continuous<br />

change in the structure of the macroscopic interface could cause a d<strong>is</strong>continuous change in<br />

γ. <strong>Th<strong>is</strong></strong> could, in principle, be caused by the appearance of a third phase <strong>that</strong> wets the<br />

interface. However, it has been observed <strong>that</strong> th<strong>is</strong> does not occur in mixtures containing<br />

strong nonionic surfactants, and the middle phase does not wet oil-water interface. In<br />

addition, the behavior near a critical end point T = T l <strong>is</strong> continuous because it <strong>is</strong> a critical<br />

point, at which the compositions and (presumably) the structures of the water-rich phase<br />

152


and the surfactant-rich phase must merge causing γ ow and γ o,µE <strong>to</strong> merge and γ w,µE <strong>to</strong><br />

van<strong>is</strong>h. We conclude, on the bas<strong>is</strong> of th<strong>is</strong> compar<strong>is</strong>on <strong>to</strong> the behavior observed in systems<br />

of small molecule surfactants, <strong>that</strong> it <strong>is</strong> unlike <strong>that</strong> the seemingly d<strong>is</strong>continuous change in<br />

γ observed in th<strong>is</strong> experiment can be explained by equilibrium thermodynamics.<br />

We now consider what non-equilibrium processes might be relevant <strong>to</strong> the observed<br />

phenomena. The analys<strong>is</strong> given in Chapter 5 was an attempt <strong>to</strong> account for the effects of<br />

limitations on the rate of copolymer transport between phases on the measured interfacial<br />

tension. That analys<strong>is</strong> ignored any limitations on the rate at which micelles can swell by<br />

absorbing homopolymer as they approach the interface, on transfer of copolymer between<br />

micelles and the surrounding solution, and on the rate at which entire micelles can be<br />

destroyed at or near the interface. Swelling of the micelles, which must be unswollen far<br />

from the interface, <strong>is</strong> necessary <strong>to</strong> maintain the equilibrium structure predicted in our SCFT<br />

simulations. Transfer of copolymer between micelles and the solution and/or destruction of<br />

entire micelles <strong>is</strong> necessary <strong>to</strong> maintain local equilibrium between the interface and nearby<br />

micelles, i.e., <strong>to</strong> maintain equal chemical in the interface equal <strong>to</strong> <strong>that</strong> in nearby micelles.<br />

Destruction of micelles <strong>is</strong> unavoidable if there <strong>is</strong> a non-negligible flux of copolymer in<strong>to</strong> the<br />

second phase.<br />

If micelles had insufficient time <strong>to</strong> swell as they approached the interface, the chemical<br />

potential of copolymers within the micelles would be higher than <strong>that</strong> expected in a swollen<br />

micelle. If the chemical potential in the interface nonetheless stayed equal <strong>to</strong> <strong>that</strong> in<br />

the resulting unswollen micelles, th<strong>is</strong> would result in a higher interfacial concentration of<br />

adsorbed copolymer, and thus a lower interfacial tension than obtained in equilibrium. <strong>Th<strong>is</strong></strong><br />

<strong>is</strong> essentially the scenario described by Leibler, who assumed equilibrium between micelles<br />

and interface, but did not allow for swelling of the micelles, and found a van<strong>is</strong>hing interfacial<br />

tension for f A < 0.69 as a result. Our experimental results resemble Leibler’s predictions,<br />

insofar as we obtain an interfacial tension much lower than predicted for swollen micelles<br />

153


for nearly symmetric copolymers. It thus seems possible <strong>that</strong> the interfacial tension <strong>is</strong><br />

reduced by a limitation on the rate at which the micelles can swell.<br />

154


Bibliography<br />

[1] Safran, S. A. Stat<strong>is</strong>tical Thermodynamics of Surfaces, Interfaces, and Membranes 2003<br />

Westview<br />

[2] Strey, R. Coll. Polym. Sci. 1994, 272, 1005<br />

[3] Kahlweit, M.; Strey, R.; Busse, G. Phys. Rev. E 1993, 47, 4197<br />

[4] Sottmann, T.; Strey, R. J. Chem. Phys. 1997, 106, 8606<br />

[5] Kahlweit, M.; Strey, R. Angew. Chem. Int. Engl. 1985, 24, 654<br />

[6] Koning, C.; Van Duin., M.; Pagnoulle. C; Jerome. R. Prog. Polym. Sci. 1998, 23, 707<br />

[7] Bates. F. S.; Maurer, W.; Lipic, P. M.; Hillmyer, M. A.; Almdal, K.; Mortensen,<br />

K.; Fredrickson, G. H.; Lodge, T. P. Phys. Rev. Lett 1997, 79, 849<br />

[8] Jeon, H. S.; Lee, J. H.; Newstein, M. C. Macromolecules 1998, 31, 3340<br />

[9] Lee, J. H.; Ruegg, M. L.; Balsara, N. P.; Zhu. Y.; Gido, S.P.; Kr<strong>is</strong>hnamoorti, R.;<br />

Kim, M.-H. Macromolecules 2003, 36, 6537<br />

[10] Hu, W.; Koberstein, J. T.; Lingelser, J. P.; Gallot, Y. Macromolecules 1995, 28,<br />

5209<br />

[11] Joseph, D. D.; Arney, M. S.; Gillberg, G.; Hu., H.; Hultman, D.; Verdier, C.;<br />

Vinagre, T. M. J. Rheol. 1992, 36, 621<br />

155


[12] Joseph, D. D.; Arney, M.; Ma, G. J. Coll. Int. Sci. 1992, 148, 291<br />

[13] Welge, I.; Wolf, B. A. Polymer 2001, 42, 3467<br />

[14] Retsos, H.; Anastasiad<strong>is</strong> S. H.; P<strong>is</strong>pas, S.; Mays, J. W.; Hadjichr<strong>is</strong>tid<strong>is</strong>, N. Macromolecules<br />

2004, 37, 524<br />

[15] Ndoni, S.; Papadak<strong>is</strong>, C. M.; Bates, F. S.; Almdal, K Rev. Sci. Instrum. 1995, 66,<br />

1090<br />

[16] Cavicchi. K. A.; Lodge, T. P. Macromolecules 2003, 36, 7158<br />

[17] Fetters, L. J.; Lohse, D. J.; Richter, D.; Witten, T. A.; Zerkel, A. Macromolecules<br />

1994, 27, 4639<br />

[18] Vonnegut, B. Rev. Sci. Instrum. 1942, 13, 6<br />

[19] Cochran, E.; Morse, D. C.; Bates, F. S.; Macromolecules 2003, 36, 782<br />

[20] Kitade, S.; Ichikawa. A.; Imura, N.; Takahashi, Y.; Noda, I. J. Rheol. 1997, 41,<br />

1039<br />

[21] personal communication with K. Almdal, 2005<br />

[22] Jorzik, U.; Wolf, B. A. Macromolecules 1997, 30, 4713<br />

[23] Leibler, L. Makromol. Chem., Macromol. Symp. 1988, 16, 1<br />

[24] Skouri, M.; Marignan, J.; Appell, J.; Porte, G. J. Phys. II. (France) 1991, 1, 1121<br />

[25] Daicic, J.; Olsson, U; Wennerstrom, H. Lamgmuir, 1995, 11, 2451<br />

[26] Washburn, N. R.; Lodge, T. P.; Bates. F. S. J. Phys. Chem. B 2000, 104, 6987<br />

[27] Kahlweit, M.; Strey, R; Haase, D.; Firman, P. Lamgmuir, 1988, 4, 785<br />

156


[28] Broseta, D.; Fredrickson, G. H. J. Chem. Phys. 1990, 93, 2927<br />

[29] Chapman, B. R.; Hamersky, M. W.; Milhaupt, J. M.; Kostelecky, C. Lodge, T. P.;<br />

von Meerwall. E. D. Smith, S. D. Macromolecules 1998, 31, 4562<br />

[30] Fle<strong>is</strong>cher, G.; Appel, M. Macromolecules 1995, 28, 7281<br />

[31] Roberts, C.; Cosgrove, T.; Schmidt, R. G.; Gordon, G. V. Macromolecules 2001,<br />

34, 538<br />

157


Chapter 7<br />

Micellization and Interfacial<br />

Tension in V<strong>is</strong>cous Ternary<br />

PS/PB/P(S-b-B) Blends<br />

Abstract<br />

We <strong>have</strong> measured the interfacial tension (γ) between polystyrene (PS) and 1,2 polybutadiene<br />

(PB) in the presence of poly(styrene-b-butadiene) (SB) block copolymer concentration using<br />

a spinning drop tensiometer (SDT). The critical micelle concentration (CMC) of a<br />

binary system of PS and SB copolymer was independently determined from small angle<br />

X-ray scattering (SAXS) and transm<strong>is</strong>sion electron microscopy (TEM). The value of γ<br />

obtained at high copolymer concentrations agreed very well with the equilibrium value<br />

predicted by Self Cons<strong>is</strong>tent Field Theory (SCFT). The measured tension continues <strong>to</strong> fall<br />

<strong>to</strong> th<strong>is</strong> predicted value and becomes saturated until the copolymer concentration reaches<br />

a value of 6 %, substantially above the measured CMC, < 0.5%. We account for th<strong>is</strong><br />

d<strong>is</strong>crepancy by considering the kinetics of copolymer diffusion <strong>to</strong> the interface, and how it<br />

<strong>is</strong> affected by the bulk concentration of copolymer.<br />

158


7.1 Introduction<br />

<strong>Th<strong>is</strong></strong> chapter reports the results of experiments on systems containing polystyrene (PS)<br />

and polybutadiene (PB) homopolymers mixed with poly(styrene-b-butadiene) (SB) block<br />

copolymers. Most of the experiments reported here were carried out before the experiments<br />

on the system of PI/PDMS/IDMS in the previous chapter. The choice of a polymer pair<br />

including PS, which <strong>is</strong> a highly v<strong>is</strong>cous polymer with a high glass transition temperature,<br />

was motivated by a desire <strong>to</strong> use transm<strong>is</strong>sion electron microscopy (TEM) <strong>to</strong> v<strong>is</strong>ualize<br />

micelles at room temperature. For th<strong>is</strong> system we thus used both TEM and small angle<br />

X-ray scattering (SAXS) <strong>to</strong> independently determine the critical micelle concentration in<br />

binary mixtures of PS and SB copolymers, in addition <strong>to</strong> measurements of interfacial<br />

tension in ternary samples.<br />

The kinetics of copolymer diffusion in our spinning drop tensiometer are, however,<br />

vastly different in th<strong>is</strong> system than in the PI-PDMS system. In spinning drop tensiometer<br />

experiments, a drop of PB (the less dense material) <strong>is</strong> surrounded by a matrix of PS. The<br />

diffusivity of copolymer in the PS matrix <strong>is</strong> several orders of magnitude slower than <strong>that</strong><br />

in the PB drop, and also much lower than the diffusivity in either PI or PDMS. The slow<br />

diffusion of copolymer in the PS matrix introduces transport limitations in th<strong>is</strong> system<br />

<strong>that</strong> complicate the behavior observed in tensiometer experiments, as d<strong>is</strong>cussed in more<br />

detail below.<br />

7.2 Experiment<br />

7.2.1 Materials<br />

We <strong>have</strong> studied several mixtures of polystyrene(PS), 1,2 polybutadiene (PB) and poly(styreneb-<br />

1,2 butadiene) (SB) block copolymer. Character<strong>is</strong>tics of these materials are l<strong>is</strong>ted in Table<br />

7.1 and Table 7.2.<br />

159


Table 7.1: Material Character<strong>is</strong>tics of Homopolymers<br />

sample code M n (g/mol) a M w /M n N b %1,2 η o (Pa·s) c γo d (PS4K) γo d (PS12K)<br />

PS4K 3900 1.08 48 60<br />

PS12K 12700 1.05 159 3500<br />

PB4K 4000 1.09 57 94 0.1 0.82 1.5<br />

PB5K 5300 1.13 76 94 0.2 0.95 1.67<br />

PB13K 13200 1.06 188 93 1 1.23 1.79<br />

PB26K 26000 1.05 374 95 10 1.35 2.01<br />

a The number average molecular weights determined by GPC and converted using<br />

polystyrene standards. b Based on a reference volume v = 81.5 cm 3 /mol at 140 ◦ C .<br />

[2] c Zero shear v<strong>is</strong>cosity measured at 140 ◦ C with ARES d Interfacial tension measured<br />

with PS4K and PS12K, respectively.<br />

Table 7.2: Material Character<strong>is</strong>tics of Block Copolymers<br />

sample code M PS (g/mol) a M PB (g/mol) a M w /M n fPS b %1,2<br />

SB1 c 9900 9700 1.08 0.47 88<br />

SB2 15000 8700 1.03 0.59 84<br />

SB3 22000 12600 1.05 0.62 95<br />

SB4 16300 9200 1.04 0.61 5<br />

a The number average molecular weights of PI blocks determined by GPC and converted<br />

using polystyrene standards. b The volume fraction of PS in SB determined by NMR. c<br />

Purchased from Polymer Source.<br />

160


Polystyrene (PS) homopolymer was anionically synthesized in the experiment. Styrene<br />

(Aldrich) monomers were stirred with calcium hydride (CaH 2 ) for a day and purified<br />

with dibuthylmagnesium under vacuum for 4 hours at 298K. Polymerization was initiated<br />

with sec-butyl lithium and carried out in 10 vol % cyclohexane at 40 ◦ C for 4 hours.<br />

After terminated with degassed methanol, the polymer solution was washed with deionized<br />

water ten times followed by filtering it with 0.2-µm filter d<strong>is</strong>k and then quenched in a 3:1<br />

vol mixture of methanol and 2-propanol. The precipitated product was purged under<br />

nitrogen gas for a day and further dried in a vacuum oven for a month. Polymerization<br />

of polybutadiene (PB) homopolymer was similar <strong>to</strong> the above procedure except for the<br />

following details. Butadiene monomers were successively stirred with dibuthylmagnesium<br />

and n-butyl lithium at 0 ◦ C for four hours each. A two-fold molar excess of purified DIPIP<br />

(1,2-dipiperidinoethane) was initially added <strong>to</strong> cyclohexane with sec-butyl lithium in order<br />

<strong>to</strong> promote 1,2-addition. Finally, an antioxidant Irganox was added <strong>to</strong> PB homopolymer<br />

at a concentration of 0.1 wt %.<br />

Poly(styrene-b-butadiene) (SB) block copolymer was synthesized by sequentially adding<br />

styrene then butadiene monomer. Before adding butadiene monomer an aliquot of PS<br />

block was cannulated out and terminated in methanol <strong>to</strong> determine the molecular weight.<br />

A small amount of low molecular weight PS homopolymer was observed in the gel permeation<br />

chroma<strong>to</strong>graphy (GPC) curve in Fig. 7.1. <strong>Th<strong>is</strong></strong> was removed by d<strong>is</strong>solving the final<br />

block copolymer in ace<strong>to</strong>ne, which <strong>is</strong> a good solvent for PS but a poor solvent for SB, at 5<br />

wt % solution for 3 hours three times. Afterwards the phase separated SB was quenched<br />

in methanol and dried in a vacuum oven.<br />

Experiments <strong>have</strong> been carried out on several different types of samples. TEM experiments<br />

<strong>have</strong> been used <strong>to</strong> determine the critical micelle concentration (CMC) in binary<br />

mixtures of PS4K and all of the copolymers. SAXS experiments <strong>have</strong> been used <strong>to</strong> determine<br />

the CMC in a mixture of PS4K and SB1. The CMC in such a binary mixtures<br />

161


0.3<br />

0.25<br />

0.2<br />

before<br />

1st<br />

2nd<br />

3rd<br />

I (a.u.)<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

19 19.5 20 20.5 21 21.5 22<br />

V e (ml)<br />

Figure 7.1: GPC curves of poly(styrene-b-butadiene) block copolymer before and after<br />

fractionation<br />

<strong>is</strong> expected <strong>to</strong> be somewhat above the equilibrium CMC in the PS-rich phase of corresponding<br />

equilibrated two-phase ternary mixture (the situation relevant <strong>to</strong> an interfacial<br />

tension measurement), in which the freedom <strong>to</strong> emulsify PB within the core of the micelle<br />

can slightly reduce the free energy of formation of a micelle. Measurements of bare homopolymer<br />

interfacial tension <strong>have</strong> been carried out between PS4K and PS12K and all of<br />

the PB homopolymers. The measurements of interfacial tension in the presence of block<br />

copolymer were all conducted on samples containing PS4K, PB13K, and SB1. In all but<br />

one measurement, the copolymer was initially mixed in<strong>to</strong> the PS matrix. D<strong>is</strong>ordered emulsions<br />

of PS4K and PB13K compatibilized by SB1 (i.e., blends, in the usual sense) <strong>have</strong><br />

also been prepared by mechanical mixing and studied by transm<strong>is</strong>sion election microscopy.<br />

Though copolymer SB1 <strong>is</strong> nearly symmetric (f PS = 0.47), it <strong>is</strong> expected <strong>to</strong> be somewhat<br />

more soluble in PS4K than in PB13K, and <strong>to</strong> preferentially forms micelles in PS4K, as<br />

a result of the lower molecular weight of the PS homopolymer ”solvent”. Flory-Huggins<br />

theory with χ = 0.08 predicts a partition coefficient K = 10 for th<strong>is</strong> molecule between<br />

162


PS4K and PB13K. The other copolymers (which <strong>have</strong> been less heavily studied) <strong>have</strong> a<br />

stronger tendency <strong>to</strong> segregate in<strong>to</strong> PS.<br />

Experimental results for both critical micelle concentrations and interfacial tensions<br />

<strong>have</strong> been compared with SCFT predictions. Appendix A reviews literature values for the<br />

Flory-Huggins interaction parameter χ(T) for PS-PB mixtures, which are needed as an<br />

input <strong>to</strong> SCFT. Appendix B reviews the way <strong>to</strong> estimate the diffusion coefficient of block<br />

copolymer in the homopolymer matrix using the friction fac<strong>to</strong>rs of the pure homopolymers.<br />

7.2.2 Blending<br />

The PS matrix used in the measurement of interfacial tension was prepared by d<strong>is</strong>solving<br />

the PS homopolymer and block copolymer in 20 wt % benzene and by subsequently evaporating<br />

the solvent with the freeze-drying method. The homogeneous polymer solution was<br />

rapidly quenched in liquid nitrogen and vacuum dried for three days. The resulting porous<br />

powder of the binary PS/SB mixture was vacuum annealed at 140 ◦ C for a hour before the<br />

spinning drop experiments so as <strong>to</strong> remove most air bubbles trapped in the freeze-dried<br />

sample and form micelles in the PS matrix.<br />

D<strong>is</strong>ordered emulsions of PS/PB/SB were obtained by mixing an equal volume of the<br />

pure PB homopolymer with the a PS copolymer mixture taken from the matrix of completed<br />

spinning drop tensiometer samples after completion of the interfacial tension measurement.<br />

Mechanical mixing was conducted in a 13 mm diameter, cup-and-ro<strong>to</strong>r miniature<br />

mixer (Mini Max CS-183MMX, Cus<strong>to</strong>m Scientific Instruments, Inc.). The samples were<br />

blended at a speed of 320 rpm with three steel balls for 20 minutes at 140 ◦ C under a<br />

nitrogen gas blanket and rapidly quenched in liquid nitrogen <strong>to</strong> reserve the morphology.<br />

7.2.3 Preparation of SDT Samples<br />

A cylindrical glass tube was vertically installed in a preheated oven shown in Fig. 7.2 after<br />

its one end <strong>is</strong> sealed with a sliding plug. The plug was equipped with a set of the calibration<br />

163


posts from 0.5 mm <strong>to</strong> 2.0 mm. Temperature was set <strong>to</strong> 160 - 180 ◦ C , which <strong>is</strong> high enough<br />

for the PS matrix <strong>to</strong> easily flow on the glass wall under small pressure difference. Once a<br />

matrix chunk was loaded, vacuum was pulled through the glass tube for an hour until all<br />

air bubbles were completely removed. <strong>Th<strong>is</strong></strong> procedure was repeated several times until the<br />

tube was filled with the required volume of PS matrix.<br />

For each sample, a 1 ml gas-tight syringe (Hamil<strong>to</strong>n) containing the PB homopolymer<br />

was prepared separately. The syringe and the glass tube in the heating oven were carefully<br />

aligned using the translational stage shown in Fig. 7.3. After the syringe was plunged in<strong>to</strong><br />

the middle of the matrix, it was important <strong>to</strong> wait for a while until the PB drop grew<br />

<strong>to</strong> a proper size at the tip of a needle due <strong>to</strong> the volume expansion instead of pushing a<br />

syringe p<strong>is</strong><strong>to</strong>n. When the needle was withdrawn, a trace of the PB homopolymer always<br />

remained in the matrix and eventually broke in<strong>to</strong> tiny droplets due <strong>to</strong> Rayleigh instability.<br />

Thus there were always several small drops coex<strong>is</strong>ting with a main drop of interest during<br />

experiments. Finally, a conjugate sliding plug <strong>that</strong> has a threaded hole at a center was<br />

inserted in<strong>to</strong> the glass tube until the first drop of the molten matrix was squeezed through<br />

the hole, and then sealed with a metal gasket and a bolt.<br />

7.2.4 Interfacial Tension Measurement<br />

The instrumental description and the detailed operating procedure of a SDT are summarized<br />

in Chapter 6. All interfacial tension measurements were carried out at 140 C. Samples<br />

were loaded inside a heating chamber preheated <strong>to</strong> 140 ± 1 ◦ C and covered with a glass<br />

enclosure <strong>to</strong> minimize heat loss by convection of air around the rotating tube. Temperature<br />

was measured right next <strong>to</strong> the rotating tube using a wire-type thermometer.<br />

When the drop diameter d becomes time independent at a constant angular velocity<br />

ω, the interfacial tension γ <strong>is</strong> calculated from the operating equation[1]<br />

γ = 1<br />

32 (ρ PS − ρ PB )ω 2 d 3 . (7.1)<br />

164


Figure 7.2: The sample preparation for a spinning drop tensiometer in a the heating oven.<br />

The heating tape <strong>is</strong> removed for clear demonstration.<br />

Figure 7.3: The experimental setup <strong>to</strong> inject a PB drop in<strong>to</strong> a PS matrix. The region near<br />

a needle of a syringe <strong>is</strong> blown up at the right corner.<br />

165


where are ρ PS = 0.983 g/cm 3 and ρ PB = 0.860 g/cm 3 [2] are the densities of PS and PB,<br />

respectively. In each experiment on a system containing block copolymer, the apparatus<br />

was allowed <strong>to</strong> spin at constant angular velocity for 24 hours.<br />

A small amount of the PS matrix was taken with the PB drop after every experiment<br />

for GPC analys<strong>is</strong>. The cross-linking and thermal degradation of the homopolymers and<br />

block copolymer were negligible.<br />

7.2.5 Transm<strong>is</strong>sion Electron Microscopy (TEM)<br />

Samples were micro<strong>to</strong>med (Ultracut Micro<strong>to</strong>me, Reichert) <strong>to</strong> thin slices with a thickness of<br />

50 nm using a diamond knife either at room temperature for the binary mixtures, or at -150<br />

◦ C for the ternary mixtures. The resulting sections collected on a 400 mesh copper grid<br />

were stained by vapor of osmium tetraoxide (OsO 4 ) aqueous solution for 30 minutes. The<br />

sample morphologies were observed using JEOL 1210 transm<strong>is</strong>sion electron microscope<br />

(TEM) at an accelerating voltage of 120 kV. The PB domains are preferentially stained,<br />

and so PB appears dark in the TEM images.<br />

7.2.6 Small Angle X-ray Scattering (SAXS)<br />

The SAXS measurement were performed for the binary blends of PS/SB in a 6m small<br />

angle beam line at the University of Minnesota. Samples were placed in a chamber under<br />

a helium atmosphere at 140 ◦ C . The integration of <strong>is</strong>otropic two dimensional scattering<br />

data in an azimuthal direction produced the 1D plots of the scattering intensity I vs. the<br />

magnitude of the scattering vec<strong>to</strong>r q = 4π/λ sin(θ/2) where θ and λ are the scattering<br />

angle and the wavelength (1.54 Å), respectively.<br />

7.2.7 Dynamic Mechanical Spectroscopy<br />

The zero shear v<strong>is</strong>cosity η o of each homopolymers was measured through dynamic mechanical<br />

spectrometry with a TA instruments ARES strain-controlled rheometer using 25 mm<br />

166


parallel plates. The PS powder was molded in<strong>to</strong> 1 mm thick d<strong>is</strong>ks using a hot press while<br />

liquid PB was directly used. Samples were protected under nitrogen gas.<br />

7.3 Results<br />

7.3.1 CMC in PS4K/SB Binary Mixtures<br />

The CMC of block copolymer in the PS4K matrix was determined using the TEM and<br />

SAXS analys<strong>is</strong>. To verify the micelle morphology had reached thermal equilibrium, we<br />

prepared binary blend samples by four different methods. First, binary mixtures of PS4K<br />

and SB were flame-sealed in ampules under high vacuum and annealed in an oil bath at<br />

140 ◦ C for a day and quenched in liquid nitrogen or ice water. Second, samples placed<br />

on an aluminum foil in a vacuum oven were annealed for a day and directly quenched in<br />

liquid nitrogen in order <strong>to</strong> increase the cooling rate of the bulk sample by removing a glass<br />

wall of the ampule. Most samples were prepared in th<strong>is</strong> way. Third, we annealed samples<br />

inside the ARES chamber under nitrogen atmosphere for an hour and quenched it in liquid<br />

nitrogen so as <strong>to</strong> simulate the actual experimental condition in a sense <strong>that</strong> the spinning<br />

drop experiments were carried out at atmospheric pressure. Fourth, samples d<strong>is</strong>solved in<br />

5 wt% <strong>to</strong>luene solution were cast in<strong>to</strong> a 1 mm thick film on a glass plate, left in the hood<br />

for a week, and vacuum dried for three weeks until there was no change in weight. These<br />

samples were then annealed for 2 days.<br />

Fig. 7.4 shows a series of TEM images of micelles in PS4K/SB1 mixture with increasing<br />

concentrations of block copolymer. By counting the number of micelles ρ m per unit area,<br />

we estimated the volume fraction of micelles φ m as<br />

φ m = 4πR3 ρ m<br />

3h<br />

1<br />

f PB<br />

(7.2)<br />

where a radius of the micelle core R was approximated <strong>to</strong> be 20 nm and a thickness of the<br />

sample h <strong>to</strong> be 50 nm + 2R.<br />

167


(a)<br />

(b)<br />

(c)<br />

(d)<br />

168


(e)<br />

(f)<br />

(g)<br />

(h)<br />

Figure 7.4: TEM images of the binary blends of PS4K/SB1 at (a) 0.125 wt% (b) 0.25 wt%<br />

(c) 0.5 wt% (d) 1 wt% (e) 2 wt% (f) 3 wt% (g) 4 wt% (h) 6 wt%. The scale bars are 50<br />

nm.<br />

169


8<br />

7<br />

6<br />

5<br />

φ m<br />

4<br />

3<br />

1<br />

2<br />

1<br />

0<br />

0 1 2 3 4 5 6 7 8<br />

φ c (%)<br />

Figure 7.5: The critical micelle concentration in the binary system of PS4K/SB1 determined<br />

by TEM.<br />

From the plot of φ m vs. φ c in Fig. 7.5, the CMC (φ cmc<br />

c ) was determined <strong>to</strong> be approximately<br />

< 0.5 wt % by extrapolating φ m <strong>to</strong> zero. The CMC we obtained for PS4K/SB1 <strong>is</strong><br />

much smaller than <strong>that</strong> reported by Kinning et al. [3] for a very similar system of PS 3.9K<br />

and SB 12K-10K, in which the PB block <strong>is</strong> predominantly 1,4 addition, for which they<br />

reported the CMC of 2.7 ± 0.2 wt % at 115 ◦ C . Since χ <strong>is</strong> inversely related <strong>to</strong> temperature,<br />

we initially expected <strong>to</strong> get the significantly higher CMC at 140 ◦ C than reported<br />

by these authors. Since the CMC <strong>is</strong> scaled as e −χf PBN C<br />

, and thus very sensitive <strong>to</strong> small<br />

changes in χ, we considered the possibility <strong>that</strong> the difference could be due differences in<br />

the fractions of 1,2 and 1,4 <strong>is</strong>omers between our copolymer SB1 (which <strong>is</strong> predominantly<br />

1,2) and the copolymer used by Kinning et al., which contained predominantly 1,4 PB. To<br />

explore th<strong>is</strong> possibility, we also measured the CMC by TEM for mixtures of PS4K with<br />

block copolymer SB4, which contains a predominantly 1,4 PB block, as well as with the<br />

other two copolymers SB2 and SB3 described in Table 7.2.<br />

170


(a)<br />

(b)<br />

(c)<br />

(d)<br />

Figure 7.6: TEM images of the binary blends of PS4K/SB2 at (a) 0.125 wt % (b) 0.25 wt<br />

% (c) 0.5 wt % (d) 1 wt %. The scale bars are 100 nm.<br />

171


(a)<br />

(b)<br />

Figure 7.7: TEM images of micelles in the PS4K matrix (a) 0.5 wt % SB3 and (b) 1 wt %<br />

SB4. The scale bars are 50 nm and 100 nm, respectively.<br />

Nevertheless, as shown in Fig. 7.6 and Fig. 7.7, we found no substantial differences in the<br />

number of micelles <strong>that</strong> still ex<strong>is</strong>t at concentrations less than 0.5 wt % regardless of 1,2<br />

<strong>is</strong>omers in PB or the length of the PS block. Though it would be difficult for us <strong>to</strong> rule out<br />

the CMC far below 0.5 wt %, our results do seem <strong>to</strong> rule out the CMC as high as <strong>that</strong> found<br />

by Kinning et al. The CMC in the PS4/SB1 mixture has been independently determined<br />

by SAXS. The one dimensional normalized intensity I(q,φ) for various concentrations φ<br />

of SB1 in the PS4K matrix <strong>is</strong> shown in Fig. 7.8. The difference between I(q,φ) and the<br />

intensity I(q,0) of of the pure PS homopolymer was integrated <strong>to</strong> calculate the integrated<br />

excess scattering intensity [4]<br />

∫<br />

Q(φ) = 4π<br />

dq q 2 [I(q,φ) − I(q,0%)] (7.3)<br />

If th<strong>is</strong> excess scattering intensity <strong>is</strong> proportional <strong>to</strong> the concentration of micelles, the CMC<br />

can be taken <strong>to</strong> be the concentration at which Q(φ) <strong>is</strong> linearly extrapolated <strong>to</strong> zero, as<br />

shown in Fig. 7.9. The CMC obtained by th<strong>is</strong> method was obviously less than 0.5 wt %,<br />

172


log I (a.u.)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0 %<br />

0.125 %<br />

0.25 %<br />

0.5 %<br />

1.0 %<br />

2.0 %<br />

3.0 %<br />

4.0 %<br />

6.0 %<br />

0.2<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05<br />

q (Å −1 )<br />

Figure 7.8: The normalized SAXS intensity in the binary system of PS4K/SB1.<br />

180<br />

160<br />

140<br />

120<br />

10 -3 Q (nm -3 )<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-20<br />

0 1 2 3 4 5 6 7 8 9<br />

φ c (%)<br />

Figure 7.9: The critical micelle concentration in the binary system of PS4K/SB1 determined<br />

by SAXS.<br />

173


ut could not be prec<strong>is</strong>ely determined due <strong>to</strong> inaccuracy of the baseline correction for small<br />

concentrations of SB1.<br />

From the above results with TEM and SAXS, the highest χ = 0.080 in Fig. A.1 in<br />

Appendix A, which corresponds <strong>to</strong> the value for 93 % of 1,2 PB, was chosen at 140 ◦ C in<br />

th<strong>is</strong> chapter for the SCFT calculation. It gives φ cmc<br />

c<br />

7.3.2 Bare Interfacial Tension<br />

∼ 0.5% in the binary system.<br />

We <strong>have</strong> measured bare interfacial tension γ o in the binary homopolymer system of PS/PB<br />

with various molecular weights l<strong>is</strong>ted in Table 7.1 in order <strong>to</strong> test the accuracy of SCFT<br />

predictions for th<strong>is</strong> quantity. With increasing molecular weights of the PB homopolymers,<br />

denoted M PB , γ o approaches the respective plateau value in Fig. 7.10 and Fig. 7.11, in<br />

which the PS matrix was fixed <strong>to</strong> PS4K and PS12K. Although the dependence of interfacial<br />

tension on χ, which <strong>is</strong> scaled as ∼ √ χ, <strong>is</strong> much less than the exponential dependence<br />

of the CMC, SCFT still produces a rather wide range of prediction for γ <strong>to</strong> cover the<br />

whole experimental results. The upper and lower bounds are obtained with χ = 0.080<br />

and χ = 0.054 <strong>that</strong> correspond <strong>to</strong> the value for the highest (93 %) and lowest (7 %) 1,2-<br />

fraction in the PB homopolymer. Both figures commonly show slower increase of measured<br />

interfacial tension relative <strong>to</strong> the theoretical prediction. Note <strong>that</strong> the sign of the error in<br />

the theoretical prediction of γ o <strong>is</strong> opposite <strong>to</strong> <strong>that</strong> found in the PI/PDMS system.<br />

7.3.3 Interfacial Tension Reduction<br />

Fig. 7.12 shows the interfacial tension reduction between PS4K and PB13K at 140 ◦ C with<br />

increasing volume fraction of SB1. In all of the measurements but one, the copolymer was<br />

initially mixed in the PS phase at the concentration show. The data point shown by a filled<br />

triangle represents an experiment in which the copolymer was initially mixed with the PB<br />

drop. The numerical values of interfacial tension in th<strong>is</strong> plot are also given in Table 7.3.<br />

The dotted and solid lines are SCFT predictions for two slightly different values of<br />

174


3<br />

2.5<br />

2<br />

γ (mN/m)<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 5 10 15 20 25 30<br />

M PB (kg/mol)<br />

Figure 7.10: Bare interfacial tension between PS4K and PB as a function of the molecular<br />

weights of PB. The experimental results are bound by the upper and lower limits in the<br />

SCFT prediction. The dotted line <strong>is</strong> given with χ = 0.054 and the solid line <strong>is</strong> with<br />

χ = 0.080, respectively.<br />

3<br />

2.5<br />

2<br />

γ (mN/m)<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 5 10 15 20 25 30<br />

M PB (kg/mol)<br />

Figure 7.11: Bare interfacial tension between PS12K and PB.<br />

175


χ = 0.08 and χ = 0.064. In these calculations, the CMC was calculated for an equilibrated<br />

two phase ternary system, in which the micelle core swells with PB homopolymer. The<br />

CMC predicted with χ = 0.064 <strong>is</strong> 4.6 % for th<strong>is</strong> ternary system, and 5.9 % for the binary<br />

system of PS4/SB1 whereas the CMC with χ = 0.08 <strong>is</strong> 0.83 % for the ternary system<br />

and 1.0 % for the binary system. Note <strong>that</strong> the predicted CMC <strong>is</strong> very sensitive <strong>to</strong> small<br />

changes in χ, but the predicted value of γ above the CMC <strong>is</strong> nearly insensitive <strong>to</strong> changes<br />

in our estimate of χ, as noted in Chapter 4.<br />

The interfacial tension measured in experiments in which copolymer <strong>is</strong> premixed with<br />

PS4K initially decreases with increasing copolymer concentration, and appears <strong>to</strong> level<br />

off above roughly 6% copolymers. The value of γ obtained above th<strong>is</strong> concentration <strong>is</strong><br />

very close <strong>to</strong> <strong>that</strong> predicted by SCFT, for either choice of χ. The concentration above<br />

which the tension saturates <strong>is</strong>, however, far above the CMC of 0.5 % or less <strong>that</strong> was<br />

obtained for a binary PS4K/SB1 system by both TEM and SAXS. The CMC in a ternary<br />

two-phase system <strong>is</strong> always somewhat lower than <strong>that</strong> of the corresponding binary system,<br />

because the freedom of a micelle <strong>to</strong> swell by absorbing PB homopolymer generally lowers<br />

the free energy of formation of a micelle. The predicted difference between CMC in binary<br />

and ternary systems thus increases the d<strong>is</strong>crepancy between TEM observations and these<br />

interfacial tension measurements. We believe <strong>that</strong> the d<strong>is</strong>crepancy <strong>is</strong> the result of transport<br />

limitations, as we d<strong>is</strong>cuss below.<br />

We conducted one interfacial tension measurement on a sample in which 4 wt % SB1<br />

block copolymer was added <strong>to</strong> the PB drop, rather than the PS matrix. The result <strong>is</strong><br />

indicated as a solid triangle in Fig. 7.12. In th<strong>is</strong> case we found an interfacial tension was<br />

very similar <strong>to</strong> the plateau value (0.13 mN/m) found in the other experiments at somewhat<br />

higher concentrations of copolymer, and very similar <strong>to</strong> the SCFT predictions for the true<br />

equilibrium. In th<strong>is</strong> experiment, we also noted <strong>that</strong> the drop radius reached its final value<br />

much more rapidly than in experiments in which copolymer was added <strong>to</strong> the matrix. It<br />

176


2<br />

1.8<br />

γ (mN/m)<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

χ=0.064<br />

0.080<br />

SB1/PS<br />

SB1/PB<br />

0<br />

0 1 2 3 4 5 6 7 8 9<br />

φ c (%)<br />

Figure 7.12: The interfacial tension reduction of PS4K and PB13K vs. the concentration<br />

of SB1 <strong>that</strong> <strong>is</strong> premixed with PS4K (circle) or PB13K (triangle). The solid line <strong>is</strong> the<br />

SCFT prediction with χ = 0.064 and the dashed line <strong>is</strong> with χ = 0.080.<br />

177


1.4<br />

d (mm)<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

1 %<br />

2 %<br />

4 %<br />

6 %<br />

8 %<br />

4%(PB)<br />

0.4<br />

0.2<br />

0<br />

0 200 400 600 800 1000 1200<br />

t (min)<br />

Figure 7.13: The change in a diameter of the PB drop as a function of time (min) in<br />

spin-up experiments. The data are plotted on a linear scale.<br />

d (mm)<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1 %<br />

2 %<br />

4 %<br />

6 %<br />

8 %<br />

4%(PB)<br />

0<br />

10 0 10 1 10 2 10 3<br />

t (min)<br />

Figure 7.14: The change in a diameter of the PB drop as a function of time (min) in<br />

spin-up experiments. The data are plotted on a logarithmic scale.<br />

178


Table 7.3: The Reduction of Interfacial Tension<br />

γ b (mN/m)<br />

0 1.23<br />

0.125 1.03 1.05<br />

0.25 0.935<br />

0.5 0.84 0.93<br />

1 0.58 0.60 0.72<br />

2 0.35 0.45<br />

4 0.21 0.23 0.12 c<br />

6 0.13<br />

8 0.14<br />

a wt % of SB1 b interfacial tension of PS4K/PB13K c SB1 was premixed with the PB13K.<br />

φ c<br />

a<br />

<strong>is</strong> easily noticed in Fig. 7.13 on a linear scale and in Fig. 7.14 on a logarithmic scale.<br />

7.3.4 Blend Morphology<br />

Fig. 7.15 shows TEM images of the morphology of ternary blends of PS4K/PB13K/SB1<br />

with 50:50 vol mixture of the homopolymers at various concentration of block copolymer<br />

premixed in<strong>to</strong> the PS.<br />

However, the average drop size dramatically decreases as φ c increases up <strong>to</strong> 6 wt %.<br />

With 1 - 2 wt % of SB1, the PB domain size <strong>is</strong> actually out of the scope of TEM. These<br />

samples with low concentrations of SB1 also were sticky, suggesting macroscopic phase<br />

separation, while those containing more block copolymer were not. We also <strong>examined</strong> the<br />

morphology of a blend made from PB homopolymer containing 4 wt % copolymer mixed<br />

with pure PS4K. The morphology for th<strong>is</strong> sample <strong>is</strong> very similar <strong>to</strong> <strong>that</strong> found for 1 - 2 wt<br />

% copolymer in PS, with large droplets, despite the low interfacial tension found for th<strong>is</strong><br />

sample in the spinning drop tensiometer.<br />

7.4 D<strong>is</strong>cussion<br />

In th<strong>is</strong> chapter we <strong>have</strong> observed <strong>that</strong> interfacial tension shown in Fig. 7.12 becomes saturated<br />

only above an apparent CMC of roughly 6 % copolymer <strong>that</strong> <strong>is</strong> significantly higher<br />

than the true CMC of under 0.5 % obtained by the TEM and SAXS. However, the sat-<br />

179


(a)<br />

(b)<br />

(c)<br />

(d)<br />

Figure 7.15: TEM images of the ternary blends of PS4K/PB13K/SB1 at various concentrations<br />

of SB1. (a) 1 wt% (b) 2 wt% (c) 4 wt% (d) 6 wt%. The scale bar corresponds <strong>to</strong><br />

400 nm in (a) and (b), and 200 nm in (c) and (d), respectively.<br />

180


urated value of interfacial tension was in a good agreement with SCFT predictions. In<br />

th<strong>is</strong> section, we consider whether the unexpectedly small reduction in γ at low copolymer<br />

concentrations could be a result of transport limitations. Specifically, we consider whether<br />

the concentration of copolymer at the interface might be suppressed <strong>to</strong> a value below the<br />

equilibrium value for <strong>to</strong>tal copolymer concentrations above the true CMC, and remain suppressed<br />

throughout the time scale of the experiment due <strong>to</strong> a slow transport of copolymer<br />

from the matrix <strong>to</strong> the PB drop<br />

Our d<strong>is</strong>cussion <strong>is</strong> based on the analys<strong>is</strong> presented in Chapter 5. There, we considered<br />

the diffusion of copolymer from a micellar phase (phase I, the PS matrix) in<strong>to</strong> a phase of<br />

initially pure homopolymer (phase II, the PB drop). We found <strong>that</strong> the time required for<br />

the concentration of free molecules at the interface <strong>to</strong> first reach the CMC c I c in phase I<br />

depends upon whether there ex<strong>is</strong>ts an exclusion zone with no micelles near the interface in<br />

phase I during the early stages of the transport process. When no exclusion zone ex<strong>is</strong>ts,<br />

the concentration of free molecules <strong>is</strong> expected <strong>to</strong> reach the CMC quite quickly, in a time<br />

comparable <strong>to</strong> <strong>that</strong> required <strong>to</strong> establ<strong>is</strong>h a nearly saturated monolayer. When an exclusion<br />

zone <strong>is</strong> present, however, the concentration of free molecules will reach the CMC only<br />

when the drop <strong>is</strong> filled with a homogeneous concentration c I c /K, where K <strong>is</strong> a partition<br />

coefficient of copolymer between a phase I and II.<br />

In order <strong>to</strong> provide a plausible explanation for our observation, <strong>that</strong> <strong>is</strong>, interfacial tension<br />

remains above the saturation value over a range of copolymer concentrations beyond<br />

c I c , it would be necessary <strong>to</strong> verify both conditions <strong>that</strong> there could ex<strong>is</strong>t an exclusion zone<br />

under the conditions of interest and <strong>that</strong> there might be insufficient time <strong>to</strong> fill the drop <strong>to</strong><br />

its final concentration within the 24 hour experiments. In addition, a complete explanation<br />

would also need <strong>to</strong> show <strong>that</strong> either condition might not be sat<strong>is</strong>fied at higher copolymer<br />

concentrations, where we seem <strong>to</strong> observe the equilibrium interfacial tension. In addition,<br />

we provide a brief analys<strong>is</strong> on kinetics in the case <strong>that</strong> block copolymer was initially added<br />

181


<strong>to</strong> the PB drop which thermodynamically res<strong>is</strong>t <strong>to</strong> micellization in equilibrium.<br />

7.4.1 Kinetics Below the CMC<br />

Below the CMC, the diffusion of block copolymer <strong>is</strong> well defined by the standard free<br />

diffusion in the semi-infinite composite media.[5] Given the diffusion coefficients of block<br />

copolymer SB1 estimated in Appendix B, which are D I f ≃ 1 × 10−11 cm 2 /s in the PS<br />

matrix and D II<br />

f<br />

≃ 3 × 10 −8 cm 2 /s in the PB drop, and the partition coefficient K =<br />

c I f /cII f<br />

≈ 13 from the Flory-Huggins theory (Eqn. 6.5), the time-independent interfacial<br />

concentration c I f,int of block copolymer relative <strong>to</strong> the bulk concentration co <strong>is</strong> uniquely<br />

fixed <strong>to</strong> c I f,int /co ≃ 0.19 from Eqn. 5.2. It implies <strong>that</strong> the actual reduction of interfacial<br />

tension obtained during steady state will be about five times smaller than the reduction in<br />

equilibrium, which <strong>is</strong> reasonably cons<strong>is</strong>tent with the result at c o = 0.125 wt % in Fig. 7.12.<br />

7.4.2 Kinetics Above the CMC<br />

Next we consider the micellar system above the CMC and determine whether an exclusion<br />

zone might ex<strong>is</strong>t in the PS matrix before a homogeneous solution of copolymer <strong>is</strong> created<br />

in the drop. According <strong>to</strong> the analys<strong>is</strong> in Chapter 5, an exclusion zone should form when<br />

the dimensionless parameter Q defined in Eqn. 5.7 <strong>is</strong> greater than 1. To calculate Q, we<br />

may estimate the diffusivity of a micelle in PS using the S<strong>to</strong>kes-Einstein equation D I m =<br />

kT/(6πη o R m ),[6] where η o <strong>is</strong> the measured v<strong>is</strong>cosity of PS 4K and R m <strong>is</strong> the hydrodynamic<br />

radius of a micelle. R m was obtained as a function of f A in Fig. 6.20, but the balance<br />

point was shifted <strong>to</strong> f bal<br />

A<br />

= 0.4 assuming K = 1 at the balance point. As a result, we used<br />

R m = 380 Å, which <strong>is</strong> somewhat swollen compared with an unswollen micelle (R m = 170<br />

Å).<br />

Fig. 7.16 shows our estimate of Q as a function of block ratio f A of block copolymer<br />

with the fixed homopolymer lengths and the concentration of micelles c o m/c c = 5. For the<br />

present system marked with an arrow, Q <strong>is</strong> slightly greater than the unity, regardless of<br />

182


10 4 0.4 0.45 0.5 0.55 0.6 0.65<br />

10 2<br />

unswollen micelle<br />

swollen micelle<br />

10 0<br />

Q<br />

10 −2<br />

10 −4<br />

f A<br />

Figure 7.16: The variation of Q as a function of the block ratio f A . The molecular weights of<br />

the homopolymers are fixed and c o m/c c = 5. f A of SB1 <strong>is</strong> indicated as an arrow. essentially<br />

identical <strong>to</strong> the CMC.<br />

10 2 0 5 10 15 20 25 30<br />

10 1<br />

Q<br />

10 0<br />

10 −1<br />

c m /c c<br />

Ι<br />

Figure 7.17: The dependence of Q on the micelle concentration.<br />

183


swelling of a micelle core, so <strong>that</strong> the exclusion zone would open in the PS matrix due <strong>to</strong> an<br />

insufficient micelle flux. Also, we plot Q as a function of with varying initial concentration<br />

of copolymer c I c/c o m in Fig. 7.17 for the current PS/PB/SB1 system with K = 13. The<br />

resulting estimate for Q passes through unity at c I c /co m = 12, which corresponds <strong>to</strong> 6.5 %<br />

assuming c c ≈ 0.5%. In light of uncertainties in the determination of the diffusivity D II<br />

f<br />

of copolymer in the PB drop and of the CMC c I c in PB, the exact value obtained for the<br />

critical concentration at which Q = 1 should not be taken overly seriously. The results do<br />

indicate, however, <strong>that</strong> it <strong>is</strong> plausible <strong>that</strong> an exclusion zone may ex<strong>is</strong>t at concentrations<br />

for which we observe an interfacial tension greater than the equilibrium value, and may<br />

close at slightly higher concentrations.<br />

We next consider the time <strong>to</strong> fill the drop. In th<strong>is</strong> system, it <strong>is</strong> important <strong>to</strong> d<strong>is</strong>tingu<strong>is</strong>h<br />

between the time required for diffusion <strong>to</strong> establ<strong>is</strong>h a uniform concentration of copolymer<br />

throughout the drop, which we will call τ d , and the time required for <strong>that</strong> concentration <strong>to</strong><br />

r<strong>is</strong>e <strong>to</strong> its equilibrium level c I c/K, which we will call τ eq . Because diffusivities of copolymer<br />

are much lower in the PS matrix than in the PB drop, we find <strong>that</strong> the dominant res<strong>is</strong>tance<br />

<strong>to</strong> copolymer transport ar<strong>is</strong>es from the matrix, and <strong>that</strong> as a result, τ d ≪ τ eq . We estimate<br />

τ d < d 2 /kD II<br />

f<br />

≈ 1 hour with a drop diameter d = 0.5 mm and k = 5.8. The drop <strong>is</strong> thus<br />

expected <strong>to</strong> contain an essentially homogeneous concentration of copolymer by the end of<br />

a 24 hour experiment.<br />

To roughly calculate the time τ eq required <strong>to</strong> ”fill” the drop when an exclusion zone<br />

<strong>is</strong> present (Q > 1), we <strong>have</strong> used the early-time solution of the transport problem given<br />

in section 5.2.2 of Chapter 5 <strong>to</strong> calculate the <strong>to</strong>tal amount N(t) of copolymer transported<br />

in<strong>to</strong> a cylindrical drop of a specified diameter d per unit length as a function of time. Then<br />

we calculated the time at which the <strong>to</strong>tal amount of material transferred corresponds <strong>to</strong> an<br />

average concentration per length 4N(t)/πd 2 equal <strong>to</strong> the equilibrium concentration c I c /K.<br />

To describe situations in which Q <strong>is</strong> very close <strong>to</strong> 1, for which we expect the exclusion zone<br />

184


<strong>to</strong> remain narrow, we may use the limit d<strong>is</strong>cussed in subsection 5.2.2. <strong>Th<strong>is</strong></strong> approximation<br />

yields a time<br />

τ eq = π 64<br />

d 2<br />

D I m<br />

( c<br />

I<br />

c<br />

c o m K ) 2<br />

. (7.4)<br />

For d = 0.5 mm, K = 13, and Dm I = 1×10−12 cm 2 /s, an estimated time τ eq spans from ∼ 2<br />

hour for c I c /co m = 1/10 up <strong>to</strong> 8.4 days for cI c /co m = 1. About cI c /co m = 3, which corresponds<br />

<strong>to</strong> 2 % as a <strong>to</strong>tal, gives the estimate of a day. If we consider Eqn. 7.4 provides the lower<br />

limit of the time <strong>to</strong> fill the drop due <strong>to</strong> the fact <strong>that</strong> the flux in<strong>to</strong> the drop should decrease as<br />

the concentration increases, we can conjecture the drop will be saturated at slightly higher<br />

concentration. It thus appears <strong>to</strong> be plausible <strong>that</strong> there ex<strong>is</strong>ts a range of concentrations<br />

slightly above the CMC in which there ex<strong>is</strong>ts an exclusion zone and in which there was<br />

insufficient time <strong>to</strong> fill the drop over the course of the experiment. Because Eqn. 7.4 for<br />

τ eq decreases as 1/(c o m )2 for a drop of fixed diameter, it also appears <strong>to</strong> be possible <strong>that</strong><br />

the time <strong>to</strong> fill the drop could become accessible at higher concentrations studied in our<br />

experiments. These estimates thus leave it unclear whether the convergence of γ <strong>to</strong> its<br />

equilibrium value above a copolymer concentration of roughly 6 % <strong>is</strong> due <strong>to</strong> the closing of<br />

an exclusion zone as shown in Fig. 7.16 under conditions in which τ eq lies beyond the time<br />

scale of the experiment, or a reduction in τ eq under conditions for which an exclusion zone<br />

<strong>is</strong> still present. Either hypothes<strong>is</strong> <strong>is</strong>, however, sufficient <strong>to</strong> explain the observation.<br />

Finally, we consider the experiment in which we mixed 4 % copolymer in<strong>to</strong> the PB<br />

drop. As shown in Fig. 7.12, we found interfacial tension equal <strong>to</strong> the saturation value<br />

obtained at concentrations higher than 6 % when copolymer was added <strong>to</strong> the PS matrix.<br />

<strong>Th<strong>is</strong></strong> <strong>is</strong> a case in which we <strong>have</strong> added copolymers <strong>to</strong> the wrong phase, since the solubility<br />

<strong>is</strong> higher and the critical chemical potential for micellization <strong>is</strong> lower in the PS matrix than<br />

in the PB drop. By redefining phase I <strong>to</strong> be the PB drop and phase II the PS matrix,<br />

the analys<strong>is</strong> of th<strong>is</strong> situation given in section 5.3 indicates <strong>that</strong> there must be an exclusion<br />

185


zone in the PB drop near the interface due <strong>to</strong> the constraint of c I,int<br />

f<br />

≤ c II<br />

c K < cI c , but<br />

<strong>that</strong> micelles may form at the interface in the PS matrix. If micelles form in PS near the<br />

interface during the early stages of transport, interfacial concentration in the PS matrix<br />

<strong>is</strong> pinned <strong>to</strong> the value of c II<br />

c , so <strong>that</strong> interfacial tension <strong>is</strong> expected <strong>to</strong> rapidly drop <strong>to</strong> the<br />

equilibrium value. The criteria for the formation of micelles in PS near the interface are<br />

given by Eqn. 5.21<br />

2 Df<br />

II<br />

π Df<br />

I<br />

c II<br />

c<br />

c o m K + 1 <<br />

cI c<br />

Kc II c<br />

(7.5)<br />

for S ≪ 1 and by Eqn. 5.22<br />

√<br />

D II<br />

f<br />

D I m<br />

c I c<br />

c o mK <<br />

cI c<br />

c II c K<br />

(7.6)<br />

for S ≫ 1. If either condition <strong>is</strong> sat<strong>is</strong>fied, then micelles will form near the interface because<br />

<strong>to</strong>tal flux should be dominated by the one which gives the larger flux. Although the CMC<br />

c I c<br />

in the PB drop <strong>is</strong> undetermined in th<strong>is</strong> experiment, the first condition can be safely<br />

fulfilled due <strong>to</strong> very small diffusivity ratio of D II<br />

f /DI f = 1/3000 unless the CMC cI c<br />

in the<br />

PB drop <strong>is</strong> <strong>to</strong>o close an interfacial concentration Kc II<br />

c<br />

defined in the true equilibrium.<br />

Note <strong>that</strong> (2/π)(c II<br />

c /c o mK) ∼ 1. Also, one can easily show the second condition <strong>is</strong> always<br />

sat<strong>is</strong>fied from the fact <strong>that</strong> Df II/DI<br />

m < 1 and cII c /co m < 1. <strong>Th<strong>is</strong></strong> result implies <strong>that</strong> micelles<br />

in the PB phase supply enough copolymer <strong>to</strong> produce inverted micelles in PS near the<br />

interface thus achieving an equilibrium interfacial concentration of copolymer. <strong>Th<strong>is</strong></strong> <strong>is</strong><br />

cons<strong>is</strong>tent with our observation of an equilibrium interfacial tension.<br />

7.5 Conclusion<br />

In th<strong>is</strong> chapter the interfacial tension in a PS/PB/SB system was measured with a SDT,<br />

and the CMC was independently determined by TEM and SAXS. The value obtained for<br />

interfacial tension at high copolymer concentrations agreed well with SCFT predictions,<br />

but a copolymer concentration well above the observed CMC was required <strong>to</strong> drive the<br />

186


interfacial tension down <strong>to</strong> th<strong>is</strong> value. An analys<strong>is</strong> of the kinetics of copolymer transport<br />

suggests <strong>that</strong> th<strong>is</strong> <strong>is</strong> because, at intermediate copolymer concentrations, the concentration<br />

at the interface <strong>is</strong> suppressed below its value at the CMC in the PS matrix during the<br />

diffusion of copolymer from the matrix in<strong>to</strong> the drop.<br />

187


Bibliography<br />

[1] Vonnegut, B. Rev. Sci. Instrum. 1942, 13, 6<br />

[2] Fetters, L. J.; Lohse, D. J.; Richter, D.; Witten, T. A.; Zerkel, A. Macromolecules<br />

1994, 27, 4639<br />

[3] Kinning, D. J.; Thomas, E. L.; Fetters, L. J. J. Chem. Phys. 1989, 90, 5806<br />

[4] Rigby, D.; Roe, R.-J. Macromolecules 1984, 17, 1778<br />

[5] Crank, J. The Mathematics of Diffusion (Oxford:Clarendon Press) 1975<br />

[6] Einstein, A. Ann. Phys. 1906, 19, 289<br />

188


Chapter 8<br />

Summary and Outlook<br />

Throughout th<strong>is</strong> thes<strong>is</strong>, we <strong>have</strong> investigated the role of AB block copolymer as an efficient<br />

surfactant in the imm<strong>is</strong>cible A/B homopolymer blends using both theoretical and<br />

experimental <strong>to</strong>ols. We <strong>have</strong> analyzed the competition between interfacial adsorption and<br />

micellization of copolymer using SCFT and the possible effects of limited kinetics on the<br />

thermodynamic prediction. In th<strong>is</strong> chapter we summarize our critical results in the thes<strong>is</strong>,<br />

and propose future research directions.<br />

8.1 Summary<br />

Self cons<strong>is</strong>tent field theory (SCFT) has been formulated in Chapter 2 for multi-component<br />

polymer systems in a grand canonical ensemble and numerically implemented with a finite<br />

difference method on a one dimensional adaptive grid. The chemical potentials and<br />

macroscopic pressure in homogeneous mixtures are derived in the limit of Flory-Huggins<br />

theory so as <strong>to</strong> obtain an expression for interfacial tension.<br />

In Chapter 3, the elastic properties of a saturated AB block copolymer monolayer<br />

separating two coex<strong>is</strong>ting A- and B- rich homopolymer phases in A/B/AB ternary systems<br />

<strong>have</strong> been characterized using SCFT calculation combined with the Canham-Helfrich<br />

theory <strong>to</strong> calculate the bending energy of weakly curved surfactant membranes. A curved<br />

interface was stabilized in a natural way by adjusting the pressure difference between<br />

189


coex<strong>is</strong>ting phases in mechanical equilibrium. For block copolymers equal or larger than<br />

homopolymers, all elastic constants extrapolated <strong>to</strong> zero with decreasing χN at the Lifshitz<br />

point at which no d<strong>is</strong>tinct boundary can ex<strong>is</strong>t between coex<strong>is</strong>ting phases. For several<br />

sets of asymmetric systems in the length of the homopolymers and/or in the stat<strong>is</strong>tical<br />

segment lengths, we could determine the balance points by adjusting the volume fraction<br />

f A of the A block within block copolymer. At the balance point, the copolymer monolayer<br />

has the van<strong>is</strong>hing spontaneous curvature. As the copolymer brush becomes drier,<br />

the effect of such asymmetry in systems on the balance point decreases. In the limit of a<br />

completely dry brush, block copolymer would be always symmetric at the balance point.<br />

The analytical expression for the Lifshitz points (Eqn. 3.13) predicted the balance points<br />

in the asymmetric systems reasonably well.<br />

In Chapter 4, the thermodynamic competition between interfacial adsorption and micellization<br />

of block copolymer has been <strong>examined</strong> under the condition in which swollen<br />

spherical micelles ex<strong>is</strong>t in one homopolymer phase. Unlike an unswollen micelle in a A/AB<br />

binary system, a micelle coex<strong>is</strong>ting with the excess B phase would swell until the spontaneous<br />

curvature <strong>is</strong> obtained in a micelle. It leads <strong>to</strong> the slight but nontrivial reduction<br />

in the excess free energy of the micelle formation, which causes the interfacial saturation<br />

<strong>to</strong> be always preempted by micellization except at a balance point. As a result, interfacial<br />

tension saturated at the cmc remains at a non-negative value. The thermodynamic<br />

behavior of the systems containing a nearly balanced copolymer monolayer can also be<br />

explained within the context of the Helfrich theory. Combining with the simplified cmc<br />

condition <strong>that</strong> ignores the translational entropy of a micelle, the Helfrich theory predicted<br />

interfacial tension would quadratically vary with the spontaneous curvature of a copolymer<br />

monolayer and van<strong>is</strong>h at a balance point. <strong>Th<strong>is</strong></strong> prediction could be independently verified<br />

with the micelle simulation for various symmetric and asymmetric systems.<br />

Transport of block copolymer from the initial phase I <strong>to</strong> the other phase II has been<br />

190


analyzed in Chapter 5 depending on the ex<strong>is</strong>tence of micelles and the choice of the initial<br />

phase. Below the CMC, a copolymer concentration at the interface <strong>is</strong> independent of time<br />

and explicitly determined by the ratio of the free chain diffusivity in phase I and II, and<br />

the partition coefficient K. Above the CMC, micelle diffusion becomes very important<br />

<strong>to</strong> increase the interfacial concentration up <strong>to</strong> the CMC. When block copolymer <strong>is</strong> added<br />

<strong>to</strong> the wrong phase, the interfacial concentration can be fixed <strong>to</strong> the value in the true<br />

equilibrium and micelles form in both phases as long as either condition of Eqn. 5.21 or<br />

Eqn. 5.22 <strong>is</strong> fulfilled.<br />

In Chapter 6, equilibrium interfacial tension between polydimethylsiloxane (PDMS)<br />

and 1,4 poly<strong>is</strong>oprene (PI) has been measured using the spinning drop tensiometer (SDT) by<br />

initially adding poly(<strong>is</strong>oprene-b-dimethylsiloxane) (IDMS) block copolymer <strong>to</strong> the PDMS<br />

matrix. The spontaneous curvature of the block copolymer monolayer was controlled by<br />

varying the PDMS block or molecular weight of the PI homopolymer. Near the balance<br />

point, symmetric block copolymers were able <strong>to</strong> decrease interfacial tension by three orders<br />

of magnitude <strong>to</strong> 10 −3 mN/m. The predicted quadratic dependence of interfacial tension<br />

on the block ratio of the PDMS block f PDMS agreed very well with the experimental<br />

results for f PDMS > 0.65. However, for 0.5 < f PDMS < 0.6, interfacial tension was<br />

much lower than the theoretical prediction with a d<strong>is</strong>continuous change within a narrow<br />

range of 0.6 < f PDMS < 0.65. Using small angle X-ray scattering (SAXS) for the ternary<br />

blends of PDMS/PI/IDMS along the <strong>is</strong>opleth, we observed the formation of a bicontinuous<br />

microemulsion phase upon achieving ultra-low interfacial tension.<br />

Interfacial tension in the system of 1,2 polybutadiene (PB) and polystyrene (PS) premixed<br />

with poly(styrene-b-butadiene) (SB) block copolymer has also been measured using<br />

the SDT in Chapter 7. The cmc in th<strong>is</strong> system could be independently identified using<br />

SAXS and TEM. While the PI/PDMS/IDMS system <strong>is</strong> ideal <strong>to</strong> access equilibrium interfacial<br />

tension for the reasons d<strong>is</strong>cussed in Chapter 5 and 6, slow diffusion of copolymer in<br />

191


the highly v<strong>is</strong>cous PS matrix greatly affects our measurements in th<strong>is</strong> system. As a result,<br />

we <strong>have</strong> observed <strong>that</strong> interfacial tension becomes saturated <strong>to</strong> the equilibrium value only<br />

above an apparent CMC <strong>that</strong> <strong>is</strong> significantly higher than the true CMC.<br />

8.2 Future Directions<br />

In Chapter 6, the ratio of the length of PI and PDMS homopolymer <strong>to</strong> <strong>that</strong> of IDMS<br />

block copolymer maintained cons<strong>is</strong>tently less than the unity following several experimental<br />

systems in the literature on a polymeric bicontinuous microemulsion phase. Since the<br />

formation of a bicontinuous microemulsion phase has been observed upon achieving ultralow<br />

interfacial tension in th<strong>is</strong> thes<strong>is</strong>, the remaining question <strong>is</strong> whether interfacial tension<br />

can still be ultra-low even in the condition in which the length of homopolymer <strong>is</strong> longer<br />

than <strong>that</strong> of block copolymer. We <strong>have</strong> not found any reference <strong>to</strong> clearly answer th<strong>is</strong><br />

question. However, we speculate <strong>that</strong> it might not occur because the unbinding transition<br />

in the binary A/AB system <strong>is</strong> unstable if the ratio of the length of homopolymer <strong>to</strong><br />

<strong>that</strong> of block copolymer <strong>is</strong> greater than 1.1,[1] which guarantees the appearance of the<br />

collapsed lamellar phase in the ternary system. In addition, it <strong>is</strong> not obvious why a uniform<br />

bicontinuous microemulsion phase appears at much higher copolymer concentration (φ c =<br />

0.3) compared with φ c ∼ 0.1 in the literature. It <strong>is</strong> probably either due <strong>to</strong> the difference in<br />

the molecular weight ratio of components, or the difference in χN c of copolymer. Therefore,<br />

it will be of particular interest <strong>to</strong> measure interfacial tension in the PI/PDMS/IDMS<br />

system by increasing molecular weight of the PDMS homopolymer. It only requires the<br />

minimal syntheses of the PDMS homopolymers since several PI homopolymers were already<br />

prepared in Table 6.1.<br />

Next it <strong>is</strong> worth further testing the analys<strong>is</strong> of the effect of copolymer transport on<br />

the interfacial concentration, given in Chapter 5. For systems with low v<strong>is</strong>cosity such<br />

as PI/PDMS/IDMS, a spinning drop experiment <strong>is</strong> relevant <strong>to</strong> quantify the interfacial<br />

192


concentration by measuring interfacial tension. For example, by synthesizing a few more<br />

symmetric IDMS copolymers with <strong>to</strong>tal molecular weights between 10 kg/mol (IDMS3)<br />

and 20 kg/mol (IDMS13), we can find an optimum molecular weight <strong>that</strong> can reach an equilibrium<br />

during the experiment because both block copolymer gave significantly different<br />

results in interfacial tension measurements. Also, we suggest <strong>that</strong> adding block copolymer<br />

<strong>to</strong> the wrong phase, in which micellization <strong>is</strong> unfavorable, beyond the cmc would be a better<br />

choice <strong>to</strong> measure equilibrium interfacial tension. A preliminary experiment has been<br />

reported in Chapter 7, in which SB1 copolymer was added <strong>to</strong> the less v<strong>is</strong>cous PB drop<br />

rather than the PS matrix. If we apply the same idea <strong>to</strong> the system of PI/PDMS/IDMS by<br />

adding copolymer <strong>to</strong> the more v<strong>is</strong>cous PI drop, we would be unable <strong>to</strong> get ultra-low interfacial<br />

tension because copolymers should be swept <strong>to</strong> the PDMS matrix. On the other hand,<br />

a thin film experiments could be ideal for highly v<strong>is</strong>cous systems such as PS/PDMS/PSb-PDMS.<br />

Compared with a spinning drop experiment, it has an advantage <strong>to</strong> carry out<br />

multiple parallel experiments with several thin film specimens. By solvent casting polymer<br />

solutions on silicon wafers using a spin coater, we can construct a bilayer of PS and PDMS<br />

where block copolymer <strong>is</strong> initially added <strong>to</strong> either homopolymer phase with concentrations<br />

below and above the CMC. The thickness of each layer can <strong>have</strong> a range of 100 - 500 nm<br />

depending on the diffusion coefficient of copolymer in the corresponding layer. Forward<br />

recoil spectrometry (FRES) and transm<strong>is</strong>sion electron microscopy (TEM) would be good<br />

experimental techniques <strong>to</strong> identify the interfacial concentration of copolymer and the population<br />

of micelles along the direction perpendicular an interface. <strong>Th<strong>is</strong></strong> type of experiment<br />

has been performed by Schulze.[3]<br />

Testing the various situations of copolymer transport using a well defined geometry<br />

either in the SDT or in thin film experiments will be greatly helpful <strong>to</strong> understand nonequilibrium<br />

behaviors of block copolymer in the ternary polymeric emulsions. Furthermore,<br />

the results might be applied <strong>to</strong> the reactive blends because block copolymer created by the<br />

193


interfacial reaction will follow the kinetic pattern of premade copolymer.[2, 3]<br />

194


Bibliography<br />

[1] Janert, P. K.; Schick, M. Macromolecules 1997, 30, 3916<br />

[2] Jones, T. D. Ph.D Thes<strong>is</strong>, University of Minnesota, 2000<br />

[3] Schulze, J. S. Ph.D Thes<strong>is</strong>, University of Minnesota, 2001<br />

195


Appendix A<br />

Flory-Huggins Interaction<br />

Parameter between PS and PB<br />

The Flory-Huggins interaction parameter χ directly affects the thermodynamic properties<br />

of the polymer blends such as interfacial tension and the CMC, which are scaled as γ ∼ √ χ<br />

and φ cmc<br />

c<br />

∼ e −f BχN c<br />

, respectively. Here we review seven different methods <strong>to</strong> measure χ in<br />

the literature and present all χ equations we <strong>have</strong> collected for the system of polystyrene<br />

(PS) and polybutadiene (PB) which <strong>is</strong> used in Chapter 7.<br />

There <strong>have</strong> been two main approaches <strong>to</strong> experimentally determine χ: One <strong>is</strong> <strong>to</strong> use the<br />

binary homopolymer blends and the other <strong>is</strong> <strong>to</strong> use block copolymer melts (or solutions).<br />

The former fits the cloud points of the binary homopolymer blends using the molar Flory-<br />

Huggins free energy of mixing per unit volume (method #1), [1, 2, 3]<br />

∆G M<br />

vkT = [<br />

φA<br />

V A<br />

ln φ A + φ B<br />

V B<br />

ln φ B<br />

]<br />

+ α φ A φ B (A.1)<br />

where ∆G M <strong>is</strong> a molar free energy of mixing per unit volume, a reference volume v <strong>is</strong><br />

usually taken as v = (v A v B ) 1 2 giving α = χ/v the units of cm 3 /mol, and φ i and V i are the<br />

overall volume fraction and the molar volume of a molecule of type i, respectively. The<br />

other method <strong>to</strong> use the binary homopolymer blends <strong>is</strong> <strong>to</strong> fit the inverse structure fac<strong>to</strong>r<br />

S −1 (q = 0)(∝ I −1 (q = 0)) at q = 0 obtained from small angle neutron scattering (SANS)<br />

196


using the random phase approximation (RPA) by deGennes [5] (method #2)[11] given as<br />

S −1 (0) = N −1<br />

A φ−1 A + N −1<br />

B φ−1 B − 2χ<br />

(A.2)<br />

where N A and N B denote the length of the A and B homopolymers. If thermal composition<br />

fluctuation <strong>is</strong> taken in<strong>to</strong> account above the critical temperature, the susceptibility S(q = 0)<br />

<strong>is</strong> fitted using a crossover function introduced by Belyakov and Kieselev (method #3).[4, 5]<br />

On the other hand, the latter can be categorized in<strong>to</strong> two types: The first type fits<br />

a plot of I(q ∗ ) at q = q ∗ of d<strong>is</strong>ordered block copolymer vs. 1/T[7, 8, 11] (or the I(q)<br />

profile in the entire range of q at each temperature)[9] using the RPA of block copolymer<br />

given by Leibler (method #4)[12] or using the theory with fluctuation correction given<br />

by Fredrickson-Helfand [10] near the order-d<strong>is</strong>order transition temperature T ODT (method<br />

#5). [5, 13, 11] The second type simply relates T ODT <strong>to</strong> a critical point χN = 10.495 in<br />

the mean field (method #6) or χN = 10.495 + 41.0 ¯N −1/3 with thermal fluctuation where<br />

¯N = Na 6 v −2 (method #7) for a few symmetric block copolymers with different lengths of<br />

N.[13, 11, 14]<br />

Usually χ <strong>is</strong> assumed <strong>to</strong> <strong>have</strong> a functional form of A/T + B. In the case of the homopolymer<br />

blends, it often includes additional compositional dependence of φ A /T,[2, 3]<br />

which normally exerts a minor contribution <strong>to</strong> χ. Among the above methods, we believe<br />

<strong>that</strong> χ extracted from the binary homopolymer blends gives the most fundamental<br />

measurement of the monomer-monomer interaction as d<strong>is</strong>cussed by Maurer et al. [11] The<br />

equations of χ (or α) for PS and PB reported in the literatures are summarized in Table A.1<br />

with the value at 140 ◦ C based on v = 81.5 cm 3 /mol. Fig. A.1 compares the temperature<br />

dependence of each χ equation for various fractions of 1,2 <strong>is</strong>omers in PB. Although χ looks<br />

like increasing with the more 1,2 addition in PB, χ <strong>is</strong> also substantially affected by the<br />

types of the measurement methods.<br />

Interestingly, given the temperature range, all χs from the Flory-Huggins equation<br />

197


Table A.1: Summary of the Interaction Parameters Between PS and PB<br />

Ref System Interaction Parameter χ h (140 ◦ C ) % 1,2<br />

[1] PS/PB α = −0.0009 + 0.750/T 0.075 20 - 25<br />

[2] PS/PB α = −0.001308 + 0.8756/T − 0.02516φ A /T b 0.064 - 0.080 6, 34<br />

[3] PS/PB α = −0.001436 + 0.9401/T − 0.0329φ A /T c 0.054 - 0.080 7 - 93<br />

[4] PS/PB α = −0.0017 + 0.99/T d 0.038 - 0.064 7, 54, 91<br />

[5] SB α = 0.001 + 0.68/T 0.053 7<br />

[7] SB χ = −0.027 + 28/T e 0.041 e n/a<br />

[8] SB χ = −0.021 + 25/T f 0.042 95<br />

[9] SBS/DOP a χ = 0.00659 + 13.6/T g 0.041 7.4, 9.8<br />

a P(S-b-B-b-S) (SBS) triblock copolymer in dioctyl phthalate (DOP) solvent. b For PS<br />

(M w = 2400 g/mol) and 6% 1,2-PB (M w = 2660 g/mol) among 5 blends. c For PS (M w<br />

= 1500 g/mol) and 43% 1,2-PB (M w = 3960 g/mol) among 16 blends. d For PS (M w =<br />

2030 g/mol) and 54% 1,2-PB (M w = 1870 g/mol) among 3 blends. e v <strong>is</strong> not available. f<br />

Based on v = 77.7 cm 3 /mol. g Based on v = 79.0 cm 3 /mol. h Based on v = 81.5 cm 3 /mol<br />

and φ A =0.5.<br />

(method #1) [1, 2, 3] cons<strong>is</strong>tently <strong>have</strong> the larger slopes A and magnitudes than those from<br />

the RPA of block copolymers (method #4).[7, 8, 9] Frielinghaus et al. pointed out <strong>that</strong> the<br />

enthalpic contribution of the homopolymer blend <strong>is</strong> larger than <strong>that</strong> of block copolymer<br />

in most systems except the case of polyethylene(PE)/poly(ethylenepropylene)(PEP). [6]<br />

Also, Fig. A.1 shows <strong>that</strong> the thermal fluctuation correction [4, 5] gives χ values <strong>that</strong><br />

deviate considerably from the others in both the binary homopolymer blends and block<br />

copolymer, i.e., χ (#1) > χ (#3) and χ (#5) > χ (#4) . It <strong>is</strong> cons<strong>is</strong>tent with Maurer et al. ’s<br />

observation,[11] <strong>that</strong> <strong>is</strong>, the Fredrickson-Helfand theory of block copolymer (#5 and #7)<br />

produced χ with much higher slopes relative <strong>to</strong> the mean field theory. Moreover, they<br />

observed <strong>that</strong> only χ determined from the measurement of T ODT in the mean field (#6)<br />

was quantitatively close <strong>that</strong> of the homopolymer blend, but any other theories with the<br />

d<strong>is</strong>ordered block copolymer (#4 and #5) caused significant errors in both coefficients in<br />

χ, which may explain why all χ’s of SB block copolymer are much lower than those of the<br />

PS/PB blends. In Chapter 7, we used the χ from ref [3] for compar<strong>is</strong>on.<br />

198


0.18<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

ref [2] 6%<br />

ref [3] 93%<br />

84%<br />

61%<br />

43%<br />

7%<br />

ref [4] 7%<br />

ref [5] 7%<br />

ref [7] n/a<br />

ref [8] 95%<br />

ref [9] 7.4%<br />

blend [2],[3]<br />

χ<br />

0.08<br />

0.06<br />

0.04<br />

} flucutuation [4],[5]<br />

} copolymer [7],[8],[9]<br />

0.02<br />

0<br />

2.4 2.5 2.6 2.7 2.8<br />

1000/T (K -1 )<br />

Figure A.1: The Flory-Hugggins interaction parameter χ between PS and PB with various<br />

% 1,2 of PB in the literature. All values are based on a fixed v = 81.5 cm 3 /mol.<br />

199


Bibliography<br />

[1] Rounds, N. A. Ph.D. Thes<strong>is</strong>, University of Akron, 1971<br />

[2] Roe, R.-J.; Zin, W.-C. Macromolecules 1980, 13, 1221<br />

[3] Han, C. D.; Chun, S. B.; Hahn, S. F.; Harper, S. Q.; Savickas, P. J.; Meunier, D.<br />

M.; Li, L.; Yalcin, T. Macromolecules 1998, 31, 394<br />

[4] Frielinghaus, H.; Schwahn, D.; Willner, L. Macromolecules 2001, 34, 1751<br />

[5] Frielinghaus, H.; Abbas, B.; Schwahn, D.; Willner, L. Europhys. Lett. 1998, 44, 606<br />

[6] Frielinghaus, H.; Mortensen, K.; Almdal, K. Macromol. Symp. 2000, 149, 63<br />

[7] Hewel, M.; Ruland, W. Makromol. Chem. Macromol. Symp. 1986, 4, 197<br />

[8] Owens, J. N.; Gancarz, I. S.; Koberstein, J. T.; Russell, T. P. Macromolecules 1989,<br />

22, 3380<br />

[9] Sakurai, S.; Mori, K.; Okawara, A.; Kim<strong>is</strong>hima, K.; Hashimo<strong>to</strong>, T. Macromolecules<br />

1992, 25, 2679<br />

[10] Fredrickson, G. H.; Helfand, E. J. Chem. Phys. 1987, 87, 697<br />

[11] Maurer, W. W.; Bates, F. S.; Lodge, T. P.; Alamdal, K.; Mortensen, K.; Fredrickson,<br />

G. H. J. Chem. Phys. 1998, 108, 2989<br />

[12] Leibler, L. Macromolecules 1980, 13, 1602<br />

200


[13] Bates, F. S.; Rosedale, J. H.; Fredrickson, G. H. J. Chem. Phys. 1990, 92, 6255;<br />

Rosedale, J. H.; Bates, F. S.; Almdal, K.; Mortensen, K.; Wignall, G. D. Macromolecules<br />

1995, 28, 1429<br />

[14] Almdal, K.; Hillmyer, M. A.; Bates, F. S. Macromolecules 2002, 35, 7685<br />

201


Appendix B<br />

Estimate of Diffusion Coefficient of<br />

P(S-b-B) Block Copolymer in the<br />

Homopolymer Phase.<br />

In Appendix B, we provide a simple way <strong>to</strong> estimate the diffusion coefficient of AB block<br />

copolymer in the corresponding homopolymer matrix using the zero shear v<strong>is</strong>cosity of<br />

the pure A and B homopolymers for the system of PS/PB/SB in Chapter 7. The same<br />

argument was applied <strong>to</strong> other system PI/PDMS/IDMS in Chapter 6.<br />

We first estimate the diffusion coefficient D SB of block copolymer using homopolymer<br />

data. According <strong>to</strong> recent studies on composition and temperature dependence of the<br />

monomeric friction fac<strong>to</strong>r ζ(φ,T) in imm<strong>is</strong>cible polymer pairs such as PS/PI (poly<strong>is</strong>oprene)<br />

[1, 2, 3] and PS/PMMA (poly(methyl methacrylate)),[4] segmental dynamics are governed<br />

by two independent mechan<strong>is</strong>ms: the inter- and intramolecular interactions. The former<br />

<strong>is</strong> dominant in the PS/PI system and the latter in the PS/PMMA system depending on<br />

the bulkiness of the monomer segments, <strong>that</strong> <strong>is</strong>, how much the backbone rearrangements<br />

of the chain are hindered by the surrounding medium. It was inferred from the fact<br />

<strong>that</strong> the effective friction fac<strong>to</strong>r ζ eff of d<strong>is</strong>ordered PS-PI (multi)block copolymer with the<br />

composition φ exhibits a similar temperature dependence and collapses in<strong>to</strong> a single master<br />

curve relative <strong>to</strong> the glass transition temperature (T −T g ) whereas it does not happen in the<br />

202


PS/PMMA system due <strong>to</strong> the different temperature dependence of monomer segments.[3, 4]<br />

Since 1,2 PB has a relatively small side group and a similar chemical structure <strong>to</strong> PI, it<br />

would be reasonable <strong>to</strong> anticipate <strong>that</strong> segmental dynamics in the PS/PB system <strong>is</strong> also<br />

dominated by the intermolecular interaction. With th<strong>is</strong> assumption, the friction fac<strong>to</strong>r<br />

ζ i (φ,T) of any molecule of type i in the common effective matrix becomes all identical,<br />

i.e.,<br />

ζ SB (φ,T) ≈ ζ PS (φ,T) ≈ ζ PB (φ,T)<br />

(B.1)<br />

while the friction fac<strong>to</strong>r ζi o (T) of the pure homopolymer differs by a few orders of magnitude.<br />

For compar<strong>is</strong>on ζ PS (φ,T) ≪ ζ PMMA (φ,T) in the PS/PMMA system.[4] Therefore,<br />

block copolymer diffusing in the pure homopolymer matrix with φ = 0 or 1 be<strong>have</strong>s<br />

as if it were imaginary random copolymer of <strong>to</strong>tal length N with the friction fac<strong>to</strong>r<br />

ζ SB . Furthermore, applying the Rouse mixing rule for the effective friction fac<strong>to</strong>r,<br />

ζ eff = f A ζ A (φ,T)+f B ζ B (φ,T) within the mean field approximation,[5, 4] gives ζ SB ≈ ζ eff<br />

using Eqn. B.1.<br />

The Arrhenius mixing rule <strong>is</strong> often adopted <strong>to</strong> predict ζ eff (φ,T) from ζi o (T) of each<br />

component,<br />

log ζ eff (φ,T) = φ A log ζ o A (T) + φ B log ζ o B (T).<br />

(B.2)<br />

In fact, Eqn. B.2 usually overestimates ζ eff (φ,T) of the PS/PI system because overall<br />

segmental dynamics <strong>is</strong> more facilitated by the effect of PI on plasticizing PS than <strong>that</strong> of<br />

PS on retarding the motion of PI.[2, 3, 4] Nonetheless, it should be at least accurate for<br />

the pure homopolymer matrix, <strong>that</strong> <strong>is</strong>, ζ eff = ζi o. The zero-shear v<strong>is</strong>cosity η o in Table 7.1<br />

<strong>is</strong> directly related <strong>to</strong> ζi o(T). Since M PS4K < M e,PS (= 13000 g/mol),[6] ζ and D of the<br />

PS4K matrix <strong>is</strong> given by<br />

η =<br />

ρN av<br />

36m o<br />

Nb 2 ζ<br />

203<br />

D = kT<br />

Nζ<br />

(B.3)


in the Rouse regime, but since M PB13K ≫ M e,1,4−PB (= 1800 g/mol), those of PB13K are<br />

by<br />

η =<br />

5ρN avm o<br />

48Me<br />

2 N 3 b 2 ζ<br />

D = kT 4N e<br />

Nζ 15N<br />

(B.4)<br />

in the reptation regime where N av <strong>is</strong> Avogadro’s number, m o and b are the molar mass and<br />

the stat<strong>is</strong>tical segment length of a monomer segment.[7] As a result, we obtain ζPS o (140◦ C) ≃<br />

2 × 10 −5 dyn-s/cm and ζPB o (140◦ C) ≃ 3 × 10 −10 dyn-s/cm. The corresponding diffusion<br />

coefficient of block copolymer with N = 260 <strong>is</strong> D SB ≃ 1 × 10 −11 cm 2 /s in the PS matrix<br />

and D SB ≃ 3 ×10 −8 cm 2 /s in the PB drop, respectively. These results are directly used in<br />

Chapter 7. Therefore, the character<strong>is</strong>tic time τ for block copolymer <strong>to</strong> diffuse by a d<strong>is</strong>tance<br />

of a typical drop radius x = d/2 ≃ 0.25 mm <strong>is</strong> τ ≃ x 2 /2D PB ≃ 10 4 sec in PB13K and<br />

τ ≃ 10 7 sec in PS4K suggesting <strong>that</strong> it takes a couple of hours <strong>to</strong> reach the center of the<br />

PB drop whereas it takes ∼ 10 2 days in the PS matrix.<br />

204


Bibliography<br />

[1] Hamersky, M. W.; Tirrell, M.; Lodge, T. P. J. Polym. Sci.:Part B 1996, 34, 2899<br />

[2] Milhaupt, J. M.; Chapman, B. R.; Lodge, T. P.; Smith, S. D. J. Polym. Sci.:Part<br />

B 1998, 36, 3079<br />

[3] Chapman, B. R.; Hamersky, M. W.; Milhaupt, J. M.; Kostelecky, C. Lodge, T. P.;<br />

von Meerwall. E. D. Smith, S. D. Macromolecules 1998, 31, 4562<br />

[4] Milhaupt, J. M.; Lodge, T. P.; Smith, S. D.; Hamersky, M. W. Macromolecules<br />

2001, 34, 5561<br />

[5] Green, P. F. Macromolecules 1995, 28, 2155<br />

[6] Fetters, L. J.; Lohse, D. J.; Richter, D.; Witten, T. A.; Zerkel, A. Macromolecules<br />

1994, 27, 4639 ‘<br />

[7] Doi, M.; Edwards, S. F. The Theory of Polymer Dynamics (Oxford:Clarendon Press)<br />

1986<br />

205


Appendix C<br />

The Effect of the Uniaxial<br />

Extensional Flow on Interfacial<br />

Tension<br />

Introduction<br />

The interfacial phenomena has been rather widely studied in simple shear flow. Given<br />

homopolymer blends with droplet/matrix morphology, interfacial tension can be obtained<br />

by fitting the elastic modulus of the blend using the model described by Palierne.[1, 2, 3]<br />

Also, interfacial slip, which <strong>is</strong> one of the interesting phenomena in the shear flow, often<br />

causes anomalously low shear v<strong>is</strong>cosity in the polymer blends.[4] Such an interfacial slip<br />

can be prevented by adding a small amount of block copolymer.[5] Recently, Moldenaers<br />

showed <strong>that</strong> interfacial tension responds linearly <strong>to</strong> the shear strain,[2] <strong>that</strong> <strong>is</strong>,<br />

γ = γ eq + β ′ ǫ<br />

(C.1)<br />

where γ eq <strong>is</strong> an <strong>is</strong>otropic interfacial tension, β ′ <strong>is</strong> a complex interfacial dilation modulus,<br />

and ǫ <strong>is</strong> a shear strain.<br />

On the other hand, there are fewer studies measuring interfacial tension in an extensional<br />

flow. Gramespacher and Me<strong>is</strong>sner developed a method <strong>to</strong> obtain interfacial tension<br />

using the elastic recovery of the blends after extension.[6] Using multilayer samples Levitt<br />

et al. found <strong>that</strong> <strong>to</strong>tal extensional force <strong>is</strong> always greater than the bulk average of each<br />

206


component due <strong>to</strong> the ex<strong>is</strong>tence of interfacial tension and increases with the number of<br />

layers[7] or the interfacial reaction.[7, 8]<br />

In addition, it would be of great interest <strong>to</strong> examine the effect of block copolymer on<br />

the interfacial property. It <strong>is</strong> widely known <strong>that</strong> block copolymer can effectively prevent<br />

interfacial slip in a shear flow.[5] However, the effect of block copolymer on the interfacial<br />

property of polymer blends has not been studied in an extensional flow in which no relative<br />

motion between adjacent phases ex<strong>is</strong>ts.<br />

In th<strong>is</strong> Appendix, an extensional rheometer (RME) <strong>is</strong> used <strong>to</strong> measure interfacial tension<br />

between polystyrene (PS) and i polypropylene (PP) with multilayer samples, which<br />

cons<strong>is</strong>t of 32 PS and PP alternating layers. For some multilayer samples, poly(styrene-bethyl<br />

ethylene) symmetric diblock copolymer (SEE) was initially added <strong>to</strong> the PS layers.<br />

For the rectangular samples used during extension, the macroscopic dimensions will follow<br />

L(t) = L o exp(˙ǫt)<br />

W(t) = W o exp(− ˙ǫ 2 t)<br />

H(t) = H o exp(− ˙ǫ 2 t)<br />

(C.2)<br />

(C.3)<br />

(C.4)<br />

in which L, W, and H are the length, width, and height of the sample, respectively.<br />

Therefore, it <strong>is</strong> much simpler <strong>to</strong> analyze the dimensional change for each phase compared<br />

with the blends.<br />

C.1 Experiment<br />

C.1.1 Materials<br />

Molecular character<strong>is</strong>tics of all materials used in th<strong>is</strong> paper are described in detail in<br />

ref [5]. For homopolymers, polystyrene (Styron 666D, Dow Chemical), denoted PS, and<br />

linear i polypropylene (Exceed 4062, ExxonMobil), denoted PP, <strong>have</strong> M w = 200 kg/mol<br />

with M w /M n = 2.0 and M w = 140 kg/mol with M w /M n = 2.4, respectively. Both<br />

207


homopolymers were supplied as pellets. Poly(styrene-b-ethyl ethylene) symmetric block<br />

copolymer, denoted SEE, as anionically synthesized in our lab by Jones [12], has M w = 200<br />

kg/mol with M w /M n = 1.05. Block copolymer was initially melt blended with polystyrene<br />

at a concentration of 2.0 wt % using a twin-screw extruder and then pelletized. For the<br />

multilayer samples, PP was coextruded either with the pure PS or with a mixture of PS<br />

and SEE, denoted PS(SEE), in<strong>to</strong> the stacks of 32 layers using a multilayer coextruder<br />

through layer multiplication dies by Zhao.[5, 11]<br />

C.1.2 Sample Preparation<br />

The pellets were directly molded in<strong>to</strong> 60 × 1 × 8 mm rectangles at 200 ◦ C in a hot press<br />

after being vacuum dried overnight. Sometimes final samples had the nonuniform cross<br />

section, which were usually thinner by ∼ 0.1 mm in the middle part, due <strong>to</strong> the different<br />

thermal expansion coefficient between the polymers and the stainless steel mold during a<br />

water-cooling process. Thus the average values were taken from at least three points in<br />

both width and height.<br />

In case of the multilayer samples, it was crucial <strong>to</strong> completely remove the residual stress<br />

<strong>that</strong> resulted from a sudden quench by a double chill roller at the last stage of the multilayer<br />

preparation.[5] Otherw<strong>is</strong>e, the multilayers immediately shrank in<strong>to</strong> irregular shapes upon<br />

melting. For th<strong>is</strong> purpose, the multilayer samples were annealed at 185 ◦ C for 15 minutes<br />

under slight pressure between two aluminum plates separated with shims a bit thicker than<br />

the initial sample heights and then slowly cooled in air. Also, they were trimmed using<br />

a hot wire cutter <strong>to</strong> prevent a significant delamination. All irregular molten edges were<br />

carefully trimmed with a razor blade in order <strong>to</strong> get rid of any potential extra stress from<br />

the irregular surfaces.<br />

208


C.1.3 Extensional Rheology<br />

The extensional rheology of the materials were measured with the Rheometric Scientific<br />

(TA instruments) RME which <strong>is</strong> an extensional rheometer based on Me<strong>is</strong>sner’s original<br />

design.[9] Samples are suspended using heated nitrogen (N 2 ) gas and elongated at a constant<br />

strain rate, ˙ǫ, up <strong>to</strong> maximum Hencky strain, ǫ = ˙ǫt = 7, using counter rotating belts.<br />

To ensure the reproducibility, all measurements were repeated at least five times unless<br />

otherw<strong>is</strong>e stated. The images of all elongated samples were recorded with VCR <strong>to</strong> determine<br />

the true strain rate from a plot of log(W(t)/W o ) vs. t where the slope corresponds<br />

<strong>to</strong> −˙ǫ/2, and where W(t) and W o are the sample width averaged at three different places<br />

of the sample at time t and t = 0, respectively.[13] The force transducer was regularly<br />

calibrated with known weights and its damping system was always cooled with circulating<br />

water during the experiments. Measured <strong>to</strong>tal force F(t) was corrected following guidelines<br />

given in ref [10] . Since the maximum force did not exceed 5 cN up <strong>to</strong> ˙ǫ= 0.1 s −1<br />

for any sample, in most experiments we employed only two lower belts with double side<br />

tapes <strong>to</strong> minimize the slip between the belt and polymer. The results with two belts were<br />

cons<strong>is</strong>tent with those obtained using all four belts.<br />

C.1.4 Data analys<strong>is</strong><br />

The extensional v<strong>is</strong>cosity <strong>is</strong> directly calculated by measuring <strong>to</strong>tal force F(t) and by assuming<br />

the affine deformation of samples,<br />

η ǫ (t) = F(t)<br />

A o ˙ǫ exp(˙ǫt) .<br />

(C.5)<br />

The initial cross sectional area A o at temperature T should be corrected based on the<br />

information at T = 298K as<br />

A o = A 298K<br />

o<br />

∑<br />

i<br />

209<br />

( ) ρ<br />

298K 2/3<br />

φ i (C.6)<br />

iρ i


where φ i and ρ i are the volume fraction and the density of the homopolymer of type i. φ i<br />

<strong>is</strong> determined from the average layer thickness measured using an optical microscopy.[5]<br />

Interfacial contribution ∆F(t) <strong>to</strong> <strong>to</strong>tal force <strong>is</strong> obtained by<br />

∆F(t) = F(t) − ∑ i<br />

φ i F i (t)<br />

(C.7)<br />

where F i (t) <strong>is</strong> the extensional force of the homopolymer of type i.[7] Since each sample<br />

has a slightly different dimension, instead it <strong>is</strong> more convenient <strong>to</strong> use the stress, denoted<br />

σ(t) = F(t)/(WH). Defining the interfacial stress ∆σ(t) by<br />

∆σ(t) = σ(t) − ∑ i<br />

φ i σ i (t) ,<br />

(C.8)<br />

interfacial tension γ(t) can be determined using the force balance as<br />

(N − 1)γ(t)W(t) = ∆F(t) = ∆σ(t)W(t)H(t) (C.9)<br />

, which leads <strong>to</strong><br />

where H T o = H298K o<br />

∆F(t)<br />

γ(t) =<br />

(N − 1)W(t) = HT o ∆σ(t)<br />

(N − 1)<br />

∑ ( )<br />

i φ ρ 298K 1/3.<br />

i<br />

i ρi<br />

exp(− ˙ǫ 2 t)<br />

(C.10)<br />

C.1.5 Shear Rheology<br />

The transient shear v<strong>is</strong>cosity η + (t) of the homopolymers was measured by start-up experiment<br />

using a strain-controlled TA instruments ARES rheometer with 25 mm parallel<br />

plates of 1 mm gap.<br />

C.2 Results and D<strong>is</strong>cussion<br />

In th<strong>is</strong> section, we measure the extensional v<strong>is</strong>cosity, or equivalently the extensional stress<br />

σ(t) = η ǫ (t) ˙ǫ, of the homopolymer and the multilayer so as <strong>to</strong> extract interfacial tension<br />

using Eqn. C.10.<br />

210


10 6 T = 170 o C<br />

180 o C<br />

185 o C<br />

10 -1 10 0 10 1 10 2 10 3<br />

10 5<br />

η ε (Pa.s)<br />

10 4<br />

10 3<br />

Time (sec)<br />

Figure C.1: The extensional v<strong>is</strong>cosity of the PS homopolymer at different temperatures of<br />

170, 180, 185 ◦ C . The extensional rate <strong>is</strong> 0.01 s −1 . Curves are averages of five different<br />

runs.<br />

10 6 T = 170 o C<br />

180 o C<br />

185 o C<br />

10 -1 10 0 10 1 10 2 10 3<br />

10 5<br />

η ε (Pa.s)<br />

10 4<br />

10 3<br />

Time (sec)<br />

Figure C.2: The extensional v<strong>is</strong>cosity of the PP homopolymer at 170, 180, and 185 ◦ C .<br />

The extensional rate <strong>is</strong> 0.01 s −1 .<br />

211


C.2.1 Extensional V<strong>is</strong>cosity of the pure Homopolymers<br />

Here we present the bulk extensional v<strong>is</strong>cosities of the components in the multilayers at the<br />

various experimental conditions. Fig. C.1 and Fig. C.2 show the extensional v<strong>is</strong>cosity of the<br />

PS and PP homopolymer, respectively, with ˙ǫ = 0.01 s −1 above melting temperature of PP,<br />

which <strong>is</strong> 165 - 170 ◦ C .[16] Although the v<strong>is</strong>cosity of PS <strong>is</strong> greatly affected by temperature,<br />

<strong>that</strong> of PP <strong>is</strong> nearly temperature independent, because the test temperatures of 170 <strong>to</strong> 185<br />

◦ C are much higher than the glass transition temperature of PP, -17 ◦ C ,[16] than <strong>that</strong> of<br />

PS, 100 ◦ C . Hereafter, all material properties are given at T = 180 ◦ C so as <strong>to</strong> remove<br />

the effect of the crystallinity in PP.<br />

Fig. C.3 and Fig. C.4 show <strong>that</strong> there <strong>is</strong> little extensional thickening for the PS and<br />

PP homopolymers at three extensional rates ˙ǫ= 0.01, 0.05, and 0.1 s −1 at 180 ◦ C . The<br />

extensional v<strong>is</strong>cosity agrees with three times the low shear rate transient v<strong>is</strong>cosity 3η + (t)<br />

(solid line). The deviations of η ǫ from 3η + (t) for t ≤ 1 s in both figures due <strong>to</strong> initial sag<br />

of the samples.[9, 7, 13]<br />

η ǫ of the binary PS(SEE) mixture <strong>is</strong> given in Fig. C.3. Interestingly, it <strong>is</strong> observed <strong>that</strong><br />

the absolute magnitude of η ǫ <strong>is</strong> reduced <strong>to</strong> about half of <strong>that</strong> of the pure PS homopolymer<br />

with ˙ǫ= 0.01 s −1 and then increases with increasing strain rate. The values at t = 10 s are<br />

summarized in Table C.1 for compar<strong>is</strong>on. The effect of block copolymer on the v<strong>is</strong>cosity<br />

has not been thoroughly investigated in the literature. Decrease in complex shear v<strong>is</strong>cosity<br />

of the homopolymer in the presence of block copolymer has also been found by Jones[12],<br />

and Yurekli and Kr<strong>is</strong>hnamoorti.[17] Such reduction in shear v<strong>is</strong>cosity <strong>is</strong> often found in<br />

Table C.1: Extensional V<strong>is</strong>cosity at t = 10 s.<br />

˙ǫ(s −1 ) η ǫ (PS) (Pa·s) η ǫ (PS(SEE)) (Pa·s)<br />

0.01 86000 51000<br />

0.05 74000 56000<br />

0.1 75000 63000<br />

212


10 6 10 -1 10 0 10 1 10 2 10 3<br />

10 5<br />

•<br />

PS ε = 0.01<br />

0.05<br />

0.1<br />

PS(PEE) 0.01<br />

0.05<br />

0.1<br />

3η +<br />

η ε (Pa.s)<br />

10 4<br />

10 3<br />

t (sec)<br />

Figure C.3: The extensional v<strong>is</strong>cosity η ǫ of the pure PS homopolymer and the 2 wt %<br />

binary mixtures of PS(SEE) at extensional rates ˙ǫ of 0.01, 0.05, and 0.1 s −1 at 180 ◦ C .<br />

Note <strong>that</strong> η ǫ of PS(SEE) <strong>is</strong> greatly reduced from <strong>that</strong> of PS due <strong>to</strong> SEE block copolymer.<br />

The extensional v<strong>is</strong>cosity of PS <strong>is</strong> cons<strong>is</strong>tent with 3η + (t), which <strong>is</strong> given by a solid line. η ǫ<br />

of PS(SEE) <strong>is</strong> a weak function of ˙ǫ approaching <strong>that</strong> of the pure PS.<br />

10 6 •<br />

ε = 0.01<br />

0.05<br />

0.1<br />

3η +<br />

10 5<br />

10 -1 10 0 10 1 10 2 10 3<br />

η ε (Pa.s)<br />

10 4<br />

10 3<br />

t (sec)<br />

Figure C.4: The extensional v<strong>is</strong>cosity of the PP homopolymer at extensional rates of 0.01,<br />

0.05, and 0.1 s −1 . Experimental temperature <strong>is</strong> 180 ◦ C . The solid line represents 3η + (t)<br />

of PP in the LVE regime.<br />

213


1<br />

0.9<br />

W / W o<br />

0.8<br />

0.7<br />

PS/PP<br />

0.6<br />

PS/PP/P(S-EE)<br />

0.5<br />

0.4<br />

10 0 10 1 10 2 10 3<br />

t (sec)<br />

Figure C.5: Shrinkage in width of multilayer samples during annealing in the RME on a<br />

semi-logarithmic time scale.<br />

hyperbranched polymer blends.[19]<br />

C.2.2 Shrinkage of Multilayers<br />

After the residual stress <strong>is</strong> removed, as mentioned in section C.1.2, there <strong>is</strong> another subtle<br />

problem <strong>to</strong> be considered for the multilayer samples. Homopolymer samples retain their<br />

initial configuration in the melt state for a relatively long time due <strong>to</strong> their high v<strong>is</strong>cosity<br />

and the surface tension force. However, the multilayer samples continuously shrink in<br />

width and increase in height, because interfacial tension between the many layers tries<br />

<strong>to</strong> minimize interfacial area within multilayers until the inward interfacial tension force <strong>is</strong><br />

balanced by the outward pressure force caused by surface tension (Young-Laplace equation)<br />

on the edge of each layer.<br />

The multilayers were annealed inside a RME chamber before extension for about 10<br />

minutes. Fig. C.5 shows the shrinkage in the width of the multilayer samples with or<br />

without block copolymer as a function of time on a semi-logarithmic plot. Although the<br />

214


multilayers with block copolymer shrink slightly faster, there <strong>is</strong> no substantial difference<br />

between the curves. The true strain rate was obtained by subtracting the contribution of<br />

the shrinkage from the apparent change in width in the image analys<strong>is</strong>. Interfacial tension<br />

<strong>is</strong> calculated by replacing the height H o in Eqn C.10 by<br />

H ′ o(t)| t=<strong>to</strong> =<br />

H o<br />

(W(t)/W o ) t=<strong>to</strong><br />

(C.11)<br />

assuming the length of the multilayer <strong>is</strong> invariant.<br />

C.2.3 Extensional V<strong>is</strong>cosity of Multilayers<br />

The extensional v<strong>is</strong>cosity of the PS/PP multilayers are given in Fig. C.6 for ˙ǫ = 0.01, 0.05,<br />

0.1, 0.5, and 1 s −1 . Error bars are not indicated for the data with the last two strain<br />

rates because the measurements were carried out only once. Up <strong>to</strong> ˙ǫ = 0.1 s −1 , there<br />

<strong>is</strong> no significant extensional thickening behavior. Fig. C.7 plots all extensional v<strong>is</strong>cosities<br />

measured at ˙ǫ= 0.01 s −1 , including the data of the PS(SEE)/PP multilayers. One can<br />

easily notice <strong>that</strong> the v<strong>is</strong>cosity of the PS homopolymer <strong>is</strong> almost ind<strong>is</strong>tingu<strong>is</strong>hable from<br />

those of the PS/PP and PS(SEE)/PP multilayer even though the multilayers cons<strong>is</strong>t of 50<br />

vol % PP homopolymer, which has much lower v<strong>is</strong>cosity. It unambiguously demonstrates<br />

the nontrivial effect of interfacial tension on <strong>to</strong>tal extensional force.<br />

C.2.4 Interfacial Tension<br />

Interfacial tension γ(t, ˙ǫ) under the external flow can be obtained using Eqn. C.10. Fig.<br />

C.8 shows the variation of interfacial tension of PS/PP with ˙ǫ= 0.01, 0.05, and 0.1 s −1<br />

as a function of time. In th<strong>is</strong> figure, one can observe a similar pattern for each interfacial<br />

tension curve. That <strong>is</strong>, it increases from the equal starting value, reaches a maximum,<br />

and quickly drops <strong>to</strong> zero. It <strong>is</strong> cons<strong>is</strong>tent with the previous observations. [7, 3]. The<br />

initial plateau, indicated as a solid line, <strong>is</strong> in good agreement with equilibrium interfacial<br />

tension γ eq = 6.48 mN/m, which was measured using a pendent drop method.[15] We do<br />

215


10 6 10 -1 10 0 10 1 10 2 10 3<br />

10 5<br />

•<br />

ε = 0.01<br />

0.05<br />

0.1<br />

0.5<br />

1<br />

η ε (Pa.s)<br />

10 4<br />

10 3<br />

Time (sec)<br />

Figure C.6: The extensional v<strong>is</strong>cosity of the PS/PP 32-multilayer at extensional rates of<br />

0.01, 0.05, and 0.1 s −1 . Experimental temperature <strong>is</strong> 180 ◦ C .<br />

10 6 10 -1 10 0 10 1 10 2 10 3<br />

10 5<br />

PS<br />

PP<br />

PS(SEE)<br />

PS/PP<br />

PS(SEE)/PP<br />

3η + (PS)<br />

3η + (PP)<br />

η ε (Pa.s)<br />

10 4<br />

10 3<br />

Time (sec)<br />

Figure C.7: Summary of the extensional v<strong>is</strong>cosities of all samples used in th<strong>is</strong> paper at<br />

˙ǫ = 0.01s −1 . Block copolymer reduces the extensional v<strong>is</strong>cosity by a fac<strong>to</strong>r of two. Three<br />

times zero shear v<strong>is</strong>cosities in the LVE region are given for the validity of the measurements.<br />

216


not fully understand at present why γ has a maximum at nearly equal time. Vinckier et al.<br />

argued <strong>that</strong> the layer structures might be destroyed at high strain.[3] However, Eqn. C.10<br />

suggests <strong>that</strong> it may also be possible for the decrease in interfacial tension <strong>to</strong> occur when the<br />

interfacial stress ∆σ(t) reaches a plateau when the system reaches steady state. Moreover,<br />

Fig. C.9 shows <strong>that</strong> the maximum of interfacial tension γ max has a linear dependence on<br />

˙ǫ implying <strong>that</strong> the interface can be regarded as a two dimensional Bingham plastic in an<br />

extensional flow[18] as given by<br />

γ(t, ˙ǫ) = γ eq + η ex (t)˙ǫ<br />

(C.12)<br />

where an interfacial v<strong>is</strong>cosity η ex (t) <strong>is</strong> assumed <strong>to</strong> be only a function of time. <strong>Th<strong>is</strong></strong> relationship<br />

between γ and ˙ǫ <strong>is</strong> cons<strong>is</strong>tent with the previous study.[18] Laun and Münstedt<br />

reported <strong>that</strong> interfacial stress between linear low density polyethylene (LDPE) and the<br />

surrounding silicon oil increased linearly with the extensional rate ˙ǫ. Eqn. C.12 also indicates<br />

<strong>that</strong> equilibrium interfacial tension can be obtained only when ˙ǫ → 0. Fig. C.10<br />

plots η ex (t) = (γ − γ eq )/˙ǫ vs. time. It <strong>is</strong> not obvious whether or not η ex (t) <strong>is</strong> independent<br />

of ˙ǫ with the current experimental results. Finally, Fig. C.11 gives the interfacial tension<br />

in the presence of SEE block copolymer. Surpr<strong>is</strong>ingly, the maximum interfacial tension <strong>is</strong><br />

larger, not smaller, than <strong>that</strong> obtained without block copolymer. These results are plotted<br />

<strong>to</strong>gether in Fig. C.9 showing <strong>that</strong> interfacial contribution actually increase due <strong>to</strong> the<br />

presence of block copolymer in the extensional flow.<br />

217


120<br />

•<br />

ε = 0.01<br />

0.05<br />

0.1<br />

100<br />

80<br />

γ (mN/m)<br />

60<br />

40<br />

20<br />

γ eq<br />

0<br />

10 -1 10 0 10 1 10 2 10 3<br />

t (sec)<br />

Figure C.8: Interfacial tension extracted from the multilayer samples as a function of time<br />

with ˙ǫ = 0.01, 0.05, and 0.1 s −1 . The solid line corresponds <strong>to</strong> equilibrium interfacial<br />

tension γ eq measured with a pendent drop method.[15]<br />

10 3 PS/PP<br />

PS(SEE)/PP<br />

10 -4 10 -3 10 -2 10 -1 10 0<br />

10 2<br />

γ max (mN/m)<br />

10 1<br />

10 0<br />

•<br />

ε<br />

Figure C.9: Maximum interfacial tension vs. the extensional rate. The linear dependence<br />

of γ on ˙ǫ can be easily observed.<br />

218


1200<br />

η ex = ( γ − γ eq<br />

) / • ε (mN.s / m)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

•<br />

ε = 0.01<br />

0.05<br />

0.1<br />

0<br />

10 -1 10 0 10 1 10 2 10 3<br />

t (sec)<br />

Figure C.10: Interfacial v<strong>is</strong>cosity η ex = (γ − γ eq )/˙ǫ as a function of time for ˙ǫ = 0.01, 0.05,<br />

and 0.1 s −1 . At t → 0, γ → γ eq .<br />

120<br />

100<br />

•<br />

ε = 0.01<br />

0.05<br />

80<br />

γ (mN/m)<br />

60<br />

40<br />

20<br />

0<br />

10 -1 10 0 10 1 10 2 10 3<br />

t (sec)<br />

Figure C.11: Interfacial tension in the presence of block copolymer. It <strong>is</strong> interesting <strong>that</strong><br />

block copolymer increases interfacial tension. Only one experiment was conducted with<br />

˙ǫ = 0.05s −1 .<br />

219


Bibliography<br />

[1] Palierne, J. F. Rheol Acta 1990, 29, 204<br />

[2] Van Hemelrijck, E.; Van Puyvelde, P.; Velankar, S.; Macosko, C. W.; Moldenaers, P.<br />

J. Rheol. 2004, 48, 145<br />

[3] Vinckier, I.; Schweizer, T.; Laun, H. M. J. Polym. Sci.:Part B. 2002, 40, 679<br />

[4] Utracki, L. A. J. Rheol. 1991, 35, 1615<br />

[5] Zhao, R.; Macosko, C. W. J. Rheol. 2002, 46, 145<br />

[6] Gramespacher, H.; Me<strong>is</strong>sner, J. J. Rheol. 1997, 41, 27<br />

[7] Levitt, L.; Macosko, C. W.; Schweizer, T.; Me<strong>is</strong>sner, J. J. Rheol. 1997, 41, 671<br />

[8] Sai<strong>to</strong>, T.; Macosko, C. W. Polym. Eng. Sci. 2002, 42, 1<br />

[9] Me<strong>is</strong>sner, J.; Raible, T.; Stephenson, S. E. J. Rheol. 1981, 25, 1<br />

[10] Schweizer, T. Rheol. Acta 2000, 39, 428<br />

[11] Zhao, R. Ph.D. Thes<strong>is</strong>, University of Minnesota, 2002<br />

[12] Jones, T. D. Ph.D. Thes<strong>is</strong>, University of Minnesota, 2000<br />

[13] Schulze, J. S.; Lodge, T. P.; Macosko, C. W.; Hepperle, J.; Munstedt, H.; Bastian, H.;<br />

Ferri, D.; Groves, D. J.; Kim, Y. H.; Lyon, M.; Schweizer, T.; Virkler, T.; Wassner, E.;<br />

Zoetelief, W. Rheol Acta, 2001, 40, 457 ‘<br />

220


[14] Schulze, J. S. Ph.D. Thes<strong>is</strong>, University of Minnesota, 2001<br />

[15] Kamal, M. R.; Lai-Fook, R.; Demarquette, N. R. Polym Eng. Sci. 1994, 34, 1823<br />

[16] Brandrup, J.; Immergut, E. H. Polymer Handbook, 3rd ed. John and Wiley & Sons,<br />

1989<br />

[17] Yureki, K.; Kr<strong>is</strong>hnamoorti, R. Macromolecules 2002, 35, 4075<br />

[18] Laun, H. M.; Munstedt, H. Rheol. Acta 1978, 17, 444<br />

[19] Kim, Y. H.; Webster, O. W. Macromolecules 1992, 25, 5561<br />

221

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!