19.07.2013 Views

Collapse of polymer brushes grafted onto planar ... - Wageningen UR

Collapse of polymer brushes grafted onto planar ... - Wageningen UR

Collapse of polymer brushes grafted onto planar ... - Wageningen UR

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Frontis - <strong>Wageningen</strong> International Nucleus for Strategic<br />

Expertise<br />

Chain molecules at interfaces: SCF Theory and Experiments<br />

A symposium to the memory <strong>of</strong> Jan Scheutjens<br />

Conveners: Gerard J. Fleer and Frans A.M. Leermakers, Laboratory <strong>of</strong> Physical<br />

Chemistry and Colloid Science, <strong>Wageningen</strong> University


Contributors<br />

Dr. Victor Amoskov Institute <strong>of</strong> Macromolecular Compounds<br />

31 Bolshoy pr. V.O., St. Petersburg<br />

199004<br />

Russia<br />

Dr. ir. Peter Barneveld <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Dr. ir. Klaas Besseling <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Pr<strong>of</strong>. Kurt Binder Johannes-Gutenberg Universität, Inst. für<br />

Physik<br />

Postfach 3980, D-6500 Mainz 01<br />

Germany<br />

Dr. Edgar Blokhuis Leiden Institute <strong>of</strong> Chemistry<br />

P.O. Box 9502 , 2300 RA Leiden<br />

The Netherlands<br />

Dr. Oleg Borrisov <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Ir. Wouter Bosker <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Ir. Mireille Claessens <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

Pr<strong>of</strong>. dr. Martien Cohen<br />

Stuart<br />

Pr<strong>of</strong>. dr. Terence<br />

Cosgrove<br />

The Netherlands<br />

<strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

University <strong>of</strong> Bristol, School <strong>of</strong> Chemistry<br />

Cantock's Close , Bristol BS8 1TS<br />

United Kingdom<br />

Dr. Kostas Daoulas ICE/HT-FORTH<br />

Stadiou Street, Platani, GR 26500 Patras<br />

Greece<br />

Dr. Arie de Keizer <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Dr. Renko de Vries <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Pr<strong>of</strong>. dr. Gerard Fleer <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Telephone: +7-812-3288542<br />

Fax: +7-812-3286869<br />

Email: vic@avm.macro.pu.ru<br />

Telephone: +31-317-484962<br />

Fax: +31-317-483777<br />

Email: peter@fenk.wag-ur.nl<br />

Telephone: +31-317 484243<br />

Fax: +31-317-483777<br />

Email:<br />

klaas.besseling@fenk.wau.nl<br />

Telephone: -<br />

Fax: 49-6131-392991<br />

Email:<br />

binder@shiel.physik.uni-mainz.de<br />

Telephone: +31-71-5274542<br />

Fax: +31-71-5274397<br />

Email:<br />

e.blokhuis@chem.leidenuniv.nl<br />

Telephone: +31-317-482178<br />

Fax: +31-317-483777<br />

Email:<br />

Telephone: +31-317-482277<br />

Fax: +31-317-483777<br />

Email:<br />

wouter.bosker@fenk.wau.nl<br />

Telephone: +31-318-483710<br />

Fax: +31-317-483777<br />

Email: mireille@fenk.wau.nl<br />

Telephone: +31-317-482178<br />

Fax: +31-317-483777<br />

Email: martien@fenk.wag-ur.nl<br />

Telephone: +44-117-9287663<br />

Fax: +44-117-9250612<br />

Email:<br />

terence.cosgrove@bris.ac.uk<br />

Telephone: +30-610-965219<br />

Fax: +30-610-965223<br />

Email:<br />

daoulas@physics.upatras.gr<br />

Telephone: +31-317-416368<br />

Fax: +31-317-483777<br />

Email: arie@fenk.wau.nl<br />

Telephone: +31-317-484561<br />

Fax: +31-317-483777<br />

Email: devries@fenk.wau.nl<br />

Telephone: +31-317-482275<br />

Fax: +31-317-483777<br />

Email: fleer@fenk.wag-ur.nl


Pr<strong>of</strong>. dr. Hans Fraaije Leiden University<br />

P.O. Box 9502, 2300 RA Leiden<br />

The Netherlands<br />

Ana Belen Jodar-Reyes University <strong>of</strong> Granada<br />

Dpto. Fisica Aplicada, Facultad de<br />

Ciencias, Fuent, Granada 18071<br />

Spain<br />

Pr<strong>of</strong>. Toshihiro Kawakatsu Tohoku University<br />

Department <strong>of</strong> Physics, Tohoku<br />

University, Aza-Aoba, Sendai 980-8578<br />

Japan<br />

Dr. Leonid Klushin American University <strong>of</strong> Beirut, Department<br />

<strong>of</strong> Physics<br />

P.O.Box 11-0236, Beirut 1107 2020<br />

Lebanon<br />

Dr. Hiroya Kodama Mitsubishi Chemical Corporation<br />

Naruse 2-17-1-17-105, Machida 194-0044<br />

Japan<br />

Dr. Luuk Koopal <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Andriy Kyrylyuk Ph.D. Leiden University<br />

P.O..Box 9502, 2300 RA Leiden<br />

The Netherlands<br />

Yansen Lauw M.Sc. <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Dr. ir. Frans Leermakers <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Pr<strong>of</strong>. dr. Per Linse University <strong>of</strong> Lund, Dept. Physical<br />

Chemistry 1<br />

Chemical Centre, P.O. Box 124, S-22100<br />

Lund<br />

Sweden<br />

Drs. Kateryna Lyakhova Leiden University<br />

P.O. Box 9502, 2300 RA Leiden<br />

The Netherlands<br />

Pr<strong>of</strong>. dr. Hans Lyklema <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Dr. Vlasis Mavrantzas ICE/HT-FORTH<br />

Stadiou Street, Platani, GR 26500 Patras<br />

Greece<br />

Mihai Morariu University <strong>of</strong> Groningen<br />

Nijenborgh 4, 9742 AG Groningen<br />

The Netherlands<br />

Martin Olsson University <strong>of</strong> Lund Chemical Centre<br />

P.O. Box 124, S-22100 Lund<br />

Sweden<br />

Telephone: +31-71-5274243<br />

Fax: +31-71.5274243<br />

Email:<br />

j.fraaije@chem.leidenuniv.nl<br />

Telephone: +34-958-246175<br />

Fax: +34-958-243214<br />

Email: ajodar@ugr.es<br />

Telephone: +81-22-2176438<br />

Fax: +81-22-2176438<br />

Email:<br />

kawakatu@cmpt.phys.tohoku.ac.j<br />

p<br />

Telephone:<br />

Fax: +961-1-744461<br />

Email: leo@aub.edu.lb<br />

Telephone: +81-45-9633263<br />

Fax: +81-45-9633947<br />

Email: kodama@rc.m-<br />

kagaku.co.jp<br />

Telephone: +31-317-482629<br />

Fax: +31-317-483777<br />

Email: koopal@fenk.wau.nl<br />

Telephone: +31-71-527 4438<br />

Fax: +31-71 527 4537<br />

Email:<br />

a.kiriluk@chem.leidenuniv.nl<br />

Telephone: +31-317-485594<br />

Fax: -<br />

Email: yansen.lauw@fenk.wau.nl<br />

Telephone: +31-317 482268<br />

Fax: +31-317-483777<br />

Email: frans@fenk.wau.nl<br />

Telephone: +46-46-2228151<br />

Fax: +46-46-2224413<br />

Email: per.linse@fkem1.lu.se<br />

Telephone: +31-71-527 4438<br />

Fax: +31-71-5274397<br />

Email:<br />

e.lyakhova@chem.leidenuniv.nl<br />

Telephone: +31-317-482279<br />

Fax: +31-317-483777<br />

Email: hans.lyklema@fenk.wag-<br />

ur.nl<br />

Telephone: +30-610-965214<br />

Fax: +30-610-965223<br />

Email: vlasis@iceht.forth.gr<br />

Telephone: +31-50 363 4517<br />

Fax: +31-50-3634400<br />

Email: m.d.morariu@chem.rug.nl<br />

Telephone: +46-46-2221536<br />

Fax: +46 46-2224413<br />

Email:<br />

Martin.Olsson@fkem1.lu.se


Dr. Martijn Oversteegen Utrecht University<br />

Padualaan 8, 3584 CH Utrecht<br />

The Netherlands<br />

Dr. Johan Pluyter International Flavors and Fragrances<br />

1515 State Highway 36, Union Beach, NJ<br />

07735<br />

USA<br />

Pr<strong>of</strong>. Raj Rajagopalan University <strong>of</strong> Florida<br />

Dept <strong>of</strong> Chemical Engineering, Gainesville<br />

Florida 32611-6005<br />

USA<br />

Dr. Bart Reuvers Akzo Nobel Dep. CoRA<br />

Velperweg 76, P.O. Box 9300<br />

6800 SB Arnhem<br />

The Netherlands<br />

Pr<strong>of</strong>. Randal Richards IRC in Polymer Science<br />

University <strong>of</strong> Durham, Durham DH1 3LE<br />

United Kingdom<br />

Pr<strong>of</strong>. Thomas Russell University <strong>of</strong> Massachusetts<br />

Polymer Science & Engineering<br />

Amherst Mass. 01003<br />

USA<br />

Pr<strong>of</strong>. Sam Safran Weizmann Institute <strong>of</strong> Science<br />

Herzel Street, Rehovot 76100<br />

Israel<br />

Ir. Sonja Scheinhardt-<br />

Engels<br />

<strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Pr<strong>of</strong>. dr. Michael Schick University <strong>of</strong> Washington<br />

Dept. <strong>of</strong> Physics, Seattle WA 98195-1560<br />

USA<br />

Pr<strong>of</strong>. Alexander Semenov Moscow State University<br />

Physics Department, Moscow 119992<br />

Russia<br />

Dr. Agur Sevink Leiden University , Gorlaeus labs,<br />

P.O. Box 9502, 2300 RA Leiden<br />

The Netherlands<br />

Pr<strong>of</strong>. Yitzhak Shnidman Polytechnic University<br />

6 MetroTech Center, Brooklyn, NY 11201<br />

USA<br />

Dr. Nadezhda Shusharina University at Buffalo - SUNY<br />

303 Furnas Hall, Dept. <strong>of</strong> Chemical<br />

Engineering, Buffalo, NY 14260-4200<br />

USA<br />

Drs. Karl Isak Skau Leiden Institute <strong>of</strong> Chemistry<br />

P.O. Box 9502 , 2300 RA Leiden<br />

The Netherlands<br />

Pr<strong>of</strong>. dr. Alexander<br />

Skvortsov<br />

Chemical Pharmaceutical Academy,<br />

Pr<strong>of</strong>. Popova 14 , St. Petersburg 197022<br />

Russia<br />

Telephone: +31-30 253 2540<br />

Fax: +31-30-2533870<br />

Email:<br />

m.oversteegen@chem.uu.nl<br />

Telephone: +1-732 335-2496<br />

Fax: +1-32-3352591<br />

Email: johan.pluyter@iff.com<br />

Telephone: +1-352-3920868<br />

Fax: +1-352-3929513<br />

Email: Raj@ChE.UFL.edu<br />

Telephone: +31-26-3663415<br />

Fax: +31-26-3665827<br />

Email:<br />

Bart.Reuvers@AkzoNobel.com<br />

Telephone: +44-191-3743153<br />

Fax: +44-191-3744651<br />

Email:<br />

r.w.richards@durham.ac.uk<br />

Telephone: +1-413-5771516<br />

Fax: +1-413-5771510<br />

Email:<br />

russell@mail.pse.umass.edu<br />

Telephone: +972-8-9343362<br />

Fax: +972-8-9344138<br />

Email:<br />

sam.safran@weizmann.ac.il<br />

Telephone: +31-317-482277<br />

Fax: +31-317-483777<br />

Email: Sonja.Engels@fenk.wau.nl<br />

Telephone:<br />

Fax:<br />

Email:<br />

schick@mahler.phys.washington.<br />

edu<br />

Telephone:<br />

Fax:<br />

Email:<br />

semenov@polly.phys.msu.su<br />

Telephone: +31-71-5274233<br />

Fax: +31-71-5274397<br />

Email:<br />

a.sevink@chem.leidenuniv.nl<br />

Telephone: +1-718-260-3785<br />

Fax: +1-509-5617203<br />

Email: shnidman@poly.edu<br />

Telephone: +1-716-6452911<br />

Fax: +1-716-6452238<br />

Email: ns33@eng.buffalo.edu<br />

Telephone: +31-71-5274542<br />

Fax: +31-71-5274397<br />

Email:<br />

k.skau@chem.leidenuniv.nl<br />

Telephone:<br />

Fax:<br />

Email:


Pr<strong>of</strong>. Ullrich Steiner University <strong>of</strong> Groningen<br />

Nijenborgh 4, 9747 AG Groningen<br />

The Netherlands<br />

Pr<strong>of</strong>. Igal Szleifer Purdue University<br />

Dept. Of Chemistry, West Lafayette,<br />

IN 47907-1393<br />

USA<br />

Pr<strong>of</strong>. Doros Theodorou National Technical University <strong>of</strong> Athens<br />

School <strong>of</strong> Chemical Engineering,<br />

Department <strong>of</strong> Mate, GR 15780Athens<br />

Greece<br />

Pr<strong>of</strong>. dr. Theo van de Ven Paprican/McGill<br />

3420 University Street, Montreal QC H3A<br />

2A7<br />

Canada<br />

Ir. Jos van den Oever <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

The Netherlands<br />

Ir. Jasper van der Gucht <strong>Wageningen</strong> University, Laboratory <strong>of</strong><br />

Physical Chemistry and Colloid Science<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong><br />

Dr. ir. Katinka van der<br />

Linden<br />

The Netherlands<br />

E.I. Dupont de Nemours<br />

Antoon Spinoystraat 6, B-2800 Mechelen<br />

Belgium<br />

Dr. Marcel van Eijk CP Kelco<br />

Ved Banen 16, DK-4623 Lille Skensved<br />

Denmark<br />

Dr. Rene van Roij Utrecht University, Institute for Theoretical<br />

Physics<br />

Postbus 80.195, 3508 TD Utrecht<br />

The Netherlands<br />

Dr. Ad van Well IRI, Delft University <strong>of</strong> Technology<br />

Mekelweg 15, 3639 JB Delft<br />

María Eugenia Velázquez-<br />

Sánchez M. Sc.<br />

Dipl. Phys. Ludger<br />

Wenning<br />

The Netherlands<br />

Technical University <strong>of</strong> Eindhoven<br />

Helix STO 2.41, 5600 MB Eindhoven<br />

The Netherlands<br />

Johannes-Gutenberg Universität, Inst. für<br />

Physik<br />

Postfach 3980, D-6500Mainz 01<br />

Germany<br />

Pr<strong>of</strong>. Mark Whitmore Memorial University <strong>of</strong> Newfoundland<br />

Department <strong>of</strong> Physics, St. Johns NF A1B<br />

3X7<br />

Canada<br />

Pr<strong>of</strong>. Katya Zhulina 135 Sheffield Lane, McMurray, PA15317<br />

USA<br />

Dr. Pacelli Zitha Delft University <strong>of</strong> Technology<br />

Mijnbouwstraat 120, 2625 LK Delft<br />

The Netherlands<br />

Telephone: +31-50-3637888<br />

Fax: +31-50-3634400<br />

Email: u.steiner@chem.rug.nl<br />

Telephone: +1-765-494-5255<br />

Fax: +1-765-494-0239<br />

Email: igal@purdue.edu<br />

Telephone: +30-10-7723157<br />

Fax: +30-10-7723184<br />

Email: doros@central.ntua.gr<br />

Telephone: +1-514-3986177<br />

Fax: +1-514-398 6256<br />

Email: theo.vandeven@mcgill.ca<br />

Telephone: +31-317-484776<br />

Fax: +31-317-483777<br />

Email:<br />

Jos.vandenOever@fenk.wag-ur.nl<br />

Telephone: +31-317-482585<br />

Fax: +31-317-483777<br />

Email: jasper@fenk.wau.nl<br />

Telephone: +32-15-441615<br />

Fax: +32-15-441510<br />

Email: katinka.van-der-<br />

linden@bel.dupont.com<br />

Telephone: +45-56165895<br />

Fax: +45-56169446<br />

Email:<br />

marcel.vaneijk@cpkelco.com<br />

Telephone: +31-30-2537579<br />

Fax: +31-30-2535937<br />

Email: r.vanroij@phys.uu.nl<br />

Telephone: +31-15-2784738<br />

Fax: +31-15-2788303<br />

Email: vanwell@iri.tudelft.nl<br />

Telephone: +31-40-2473132<br />

Fax: +31-40-2445619<br />

Email: M.velazquez-<br />

sanchez@tue.nl<br />

Telephone: +49-6131-3925151<br />

Fax: +49-6131-3925441<br />

Email: wenning@uni-mainz.de<br />

Telephone: +1-709-7378832<br />

Fax: +1-709-7378739<br />

Email: markw@physics.mun.ca<br />

Telephone:<br />

Fax:<br />

Email: kzhulina@hotmail.com<br />

Telephone: +31-15-2788437<br />

Fax: +31-015-2781189<br />

Email: p.l.j.zitha@citg.tudelft.nl


Dr. Andre Zvelindovsky Leiden University<br />

P.O.Box 9502, 2300 RA Leiden<br />

The Netherlands<br />

Telephone: +31-71-5274234<br />

Fax: +31-71-5274537<br />

Email:<br />

a.zvelindovsky@chem.leidenuniv.<br />

nl


Preface<br />

Self-consistent field modelling has played a key role in the increase <strong>of</strong> our knowledge <strong>of</strong> chain<br />

molecules at interfaces. This is a topic where the disciplines <strong>of</strong> physics, chemistry, biology and<br />

technology meet. Typical keywords are:<br />

• Statics and dynamics <strong>of</strong> long and short <strong>polymer</strong>s in solution and at interfaces (coil, globule, helix,<br />

flower, living <strong>polymer</strong>s)<br />

• Self-organization <strong>of</strong> (bio-)amphiphiles into micelles, membranes, vesicles<br />

• Surface modification, co<strong>polymer</strong>, polyelectrolyte and polyampholyte adsorption, <strong>brushes</strong><br />

• Colloidal stabilization, wetting, capillary condensation, nucleation<br />

• (Micro-)emulsions, phase separation, curved interfaces<br />

• Polymer blends, structure (pattern) formation in <strong>polymer</strong> melts.<br />

Without exception all these topics have direct technological applications and consequences.<br />

The <strong>Wageningen</strong> School to model inhomogeneous systems was started by the seminal work <strong>of</strong> the<br />

late Jan Scheutjens. In August 2002 it was ten years ago that Jan Scheutjens died. This symposium<br />

was dedicated to his memory. Experts in the field <strong>of</strong> chain molecules at interfaces were invited to<br />

come together to evaluate the state <strong>of</strong> the art. Progress on numerical, analytical, scaling aspects as<br />

well as new insights resulting from "theory-guided" experiments was considered. The aim <strong>of</strong> the<br />

symposium was to assess the possibilities and challenges for the near and more distant future. The<br />

symposium was held from 25 to 28 August 2002 in the <strong>Wageningen</strong> International Conference Centre.


Chain Molecules at Interfaces: SCF Theory and Experiment: a summary<br />

The symposium "Chain Molecules at Interfaces: SCF Theory and Experiment" was organized<br />

by Frontis - <strong>Wageningen</strong> International Nucleus for Strategic Expertise, in collaboration with<br />

the Laboratory <strong>of</strong> Physical Chemistry and Colloid Science (LPCCS) <strong>of</strong> <strong>Wageningen</strong><br />

University and with financial support by the Royal Netherlands Academy <strong>of</strong> Sciences, to<br />

commemorate Jan Scheutjens, who died ten years ago in a car accident. The importance <strong>of</strong><br />

the self-consistent field (SCF) theory for mesoscopic physics can be compared with the<br />

importance <strong>of</strong> quantum mechanics for theoretical physics. Jan Scheutjens was one <strong>of</strong> the<br />

designers <strong>of</strong> the SCF theory, and the <strong>Wageningen</strong> LPCCS has built up a strong international<br />

position in this field. Ten years ago Gerard Fleer and his co-workers pledged to take care <strong>of</strong><br />

Jan Scheutjens' heritage. The attendance <strong>of</strong> so many <strong>of</strong> the world's most recognized<br />

scientists at this symposium proves that this field <strong>of</strong> research is still thriving at <strong>Wageningen</strong><br />

University.<br />

The participants were unanimous in their judgement that this symposium was extremely<br />

interesting; they were enthusiastic about the wealth <strong>of</strong> problems that can successfully be<br />

attacked by the SCF theory. The approach by Scheutjens (and Fleer) is still seen as a<br />

flexible tool to study complex problems involving chain molecules and concentration<br />

gradients. There was no doubt about the historic importance <strong>of</strong> Jan Scheutjens in the<br />

development <strong>of</strong> this theory.<br />

The symposium brought together various disciplines within physical chemistry. Theoretical<br />

modelling was the issue that united all those who attended the symposium, not only through<br />

typically theoretical approaches, but also in various experimental aspects. The wide range <strong>of</strong><br />

subjects that was covered may be deemed unique. New experimental data on <strong>polymer</strong>s at<br />

surface boundaries were discussed in relation with modern developments in modelling <strong>of</strong><br />

them. The emphasis was on equilibrium analyses, but also their dynamic modelling was<br />

discussed. Many <strong>of</strong> these systems have a direct technological relevance. In addition the<br />

modelling <strong>of</strong> self-assembling systems was thoroughly discussed. This topic has an obvious<br />

biological component. It is not surprising, therefore, that various speakers who usually find<br />

themselves in very different parts <strong>of</strong> science met each other for the first time during this<br />

symposium. These meetings can be characterized as "cross-pollinations" and must be<br />

considered one <strong>of</strong> the successes <strong>of</strong> the symposium. Without doubt <strong>Wageningen</strong> was put on<br />

the map once again as a centre for SCF modelling, not only in <strong>polymer</strong> physics, but also in<br />

areas <strong>of</strong> relevance to biophysics.<br />

The symposium was attended by ca 60 scientists. The 14 invited keynote speakers all had<br />

one hour for their presentations, including discussion. Five additional speakers, selected on<br />

the basis <strong>of</strong> their submitted poster abstracts, were each given 30 minutes for oral<br />

presentations. The limited number <strong>of</strong> participants has certainly contributed to the <strong>of</strong>ten very<br />

lively discussions that followed the thorough and generally excellent presentations. Also the<br />

LPCCS's PhD students were allowed and able to take part in the discussions, which, <strong>of</strong><br />

course, were continued after the formal sessions. The relatively many posters were the<br />

subject <strong>of</strong> lively discussions during breaks.<br />

The participants represented many countries: Canada (2), Denmark (1), Germany (3),<br />

Greece (3), Israel (1), Japan (2), Lebanon (1), Russia (3), Spain (1), Sweden (2), United<br />

Kingdom (2), United States (8) and The Netherlands (ca 30). The importance <strong>of</strong> the SCF<br />

method is recognized more strongly abroad than in The Netherlands; <strong>Wageningen</strong> University<br />

is better known in this field outside the Dutch borders than at home. A limited number <strong>of</strong><br />

participants (5) came from industry.<br />

Experimental aspects <strong>of</strong> <strong>polymer</strong>s at boundaries were discussed by T. Russel (how to place<br />

nano-sized rods perpendicular to a surface), T. Cosgrove (neutron scattering on surfaces


covered by <strong>polymer</strong>s), and R. Rajagopalan (surface force measurements <strong>of</strong> <strong>polymer</strong>-coated<br />

layers). Theoretical counterparts were F. Leermakers (accounting for intramolecular<br />

interactions), L. Klushin (adsorption transition <strong>of</strong> co<strong>polymer</strong>s), M. Whitmore (<strong>polymer</strong><br />

<strong>brushes</strong>), and A. Semenov (analytical theory for adsorption <strong>of</strong> semi-flexible macromolecules).<br />

Computer simulations for <strong>polymer</strong>s in confined spaces (wetting and capillary condensation)<br />

were presented by K. Binder. D. Theodorou discussed a method to use SCF results as input<br />

for simulations. Various aspects <strong>of</strong> dynamic modelling <strong>of</strong> concentrated <strong>polymer</strong> systems<br />

(melts) were treated by J. Fraaije, T. Kawakutsu and Y. Snidman. The emphasis was here on<br />

generating and understanding mesoscopic patterns that can only be made by a dynamical<br />

path. In this area theory and experiment come surprisingly close to each other. A contribution<br />

by V. Mavrantzas linked the classical adsorption theory with flow fields. P. Linse gave an<br />

overview <strong>of</strong> the modelling <strong>of</strong> self-assembling behaviour <strong>of</strong> non-ionogenic soap molecules. R.<br />

van Roij discussed the theory <strong>of</strong> rod-like particles (molecules) and their behaviour in confined<br />

spaces. Biologically oriented topics were brought forward by I. Szleifer (protein molecules at<br />

surfaces), S. Safran (self-assembling networks), and M. Schick (theory <strong>of</strong> vesicle fusions).<br />

The research done in the Laboratory <strong>of</strong> Physical Chemistry and Colloid Sciences at<br />

<strong>Wageningen</strong> has very close relations with the subjects discussed during the symposium.<br />

This was particularly clear through the many poster presentations <strong>of</strong> <strong>Wageningen</strong> graduate<br />

students. Some keywords describing the width <strong>of</strong> the research are: bending moduli <strong>of</strong><br />

membranes and stability <strong>of</strong> vesicles, modelling <strong>of</strong> pores in membranes and SCF modelling <strong>of</strong><br />

lipid bi-layers, stiffness <strong>of</strong> worm-like micelles, self-assembling <strong>polymer</strong>s at surfaces, wetting<br />

<strong>of</strong> a poly-electrolyte brush, allophobic and autophobic behaviour <strong>of</strong> a <strong>polymer</strong> melt on a<br />

brush, <strong>brushes</strong> made <strong>of</strong> polysaccharides, and the steady-state modelling <strong>of</strong> surfaces in a<br />

<strong>polymer</strong> solution.<br />

In conclusion the symposium has been very successful. Undoubtedly its results will have<br />

great impact on future research at <strong>Wageningen</strong>. New contacts were made and old ones<br />

renewed. <strong>Wageningen</strong> has confirmed or even strengthened its position in a large number <strong>of</strong><br />

disciplines in mesoscopic (bio)physics.


INTERACTIONS BETWEEN BOVINE SERUM ALBUMIN AND BILE SALTS:<br />

A THERMODYNAMIC INVESTIGATION<br />

M.L. Antonelli, A. Palacios, C. La Mesa<br />

Dipartimento di Chimica, Università di Roma la Sapienza,<br />

P.le A. Moro 5, 00185, Rome, Italy<br />

email: marta.antonelli@uniroma1.it<br />

ABSTRACT<br />

Due to its ubiquitous nature, albumin is found in significant amounts in almost all human, or mammal, tissues.<br />

As such it is responsible for the complex phenomena associated to the formation <strong>of</strong> human bile, both hepatic<br />

and colecystic ones. In such tissues the composition and the overall amount <strong>of</strong> albumin and bile acid salts is<br />

dictated by certain rules.<br />

Most bile acid salts occurring in the humans interact with albumin by a combination <strong>of</strong> electrostatic and<br />

hydrophobic contributions. There is experimental evidence <strong>of</strong> a small selectivity between albumin and different<br />

bile acid salts (conjugated or not, mono, di- or tri-hydroxy derivatives, etc.) based on a relative hydrophobicity<br />

scale <strong>of</strong> the different bile acid salts.<br />

To check this hypothesis a thermodynamic investigation <strong>of</strong> the interactions between bovine serum<br />

albumin (BSA) and Sodium Taurodeoxycholate, (NaTDC), has been performed. The investigation deals with<br />

batch dilution calorimetric studies, activity coefficients findings (inferred by freezing point depression data)<br />

and surface tension behaviour. Studies have been performed at 25°C. The concentrations investigated cover<br />

a wide range <strong>of</strong> protein to surfactant ratios [1]. In this way information from the observed effects may be<br />

useful to quantify the energetics <strong>of</strong> binding phenomena in biologically relevant systems [2]. The data have<br />

been interpreted in terms <strong>of</strong> competition between NaTDC binding <strong>onto</strong> the protein and micelle formation.<br />

References<br />

1. Nielsen, A.D., Borch, K., Westh, P., 2000. Biochim. Biophys. Acta, 1479, 321-331.<br />

2. Samso, M, Daban, J:R., Hansen, S., Jones, G.R., 1995. Eur. J. Biochem., 212, 818-824.


FIRST-ORDER WETTING TRANSITION AT FINITE CONTACT ANGLE<br />

F.A.M. Leermakers, J. Maas and M.A. Cohen Stuart<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: frans@fenk.wau.nl<br />

ABSTRACT<br />

The wetting <strong>of</strong> a <strong>polymer</strong> brush by a melt <strong>of</strong> similar chains can have a window <strong>of</strong> complete wetting. In this<br />

case there is a classical allophobic wetting transition, at low grafting density σ , and an autophobic one at<br />

high σ . However, when the melt chains are much longer than the brush chains, the contact angle α goes<br />

through a non-zero minimum where ∂α ∂σ<br />

/ has a jump. An SCF analysis and experimental observations<br />

indicate a double-well disjoining pressure curve, consistent with a novel first-order wetting transition at finite<br />

α . The meta-stable contact angle can become zero. This is important information in order to understand the<br />

kinetics <strong>of</strong> dewetting. The results show that the classical wetting theory must be extended in order to account<br />

for incomplete wetting (i.e. wetting transitions at finite contact angle).


WETTING OF AN ANNEALED POLYELECTROLYTE BRUSH, A NUMERICAL SCF STUDY<br />

F.A.M. Leermakers 1) , A. Mercurieva 2) , E.B. Zhulina 2) , T.M. Birshtein 2)<br />

1) Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

2)<br />

Institute <strong>of</strong> Macromolecular Compounds <strong>of</strong> the Russian Academy <strong>of</strong> Sciences,<br />

199004, St. Petersburg, Russia<br />

email: frans@fenk.wau.nl<br />

ABSTRACT<br />

In this poster we will discuss the wetting properties <strong>of</strong> an annealed polyelectrolyte brush in an oil-salted water<br />

system. The ionic strength is chosen as the control parameter. The salted water interacts with the solid<br />

substrate as well as with the annealed brush. The first type <strong>of</strong> interaction has a short-range character; the<br />

other one is long range (at least proportional to the brush height). One can envision wetting transitions due to<br />

both interactions. The two wetting transitions appear to be <strong>of</strong> the first-order type and interfere with each other.<br />

The pre-wetting lines associated to these transitions cross and a triple point is found. At this triple point three<br />

(microscopic) film thicknesses coexist. There can only be one wetting transition and therefore one part <strong>of</strong> one<br />

<strong>of</strong> the pre-wetting lines is meta-stable in between a hidden wetting transition and the triple point. The other<br />

pre-wetting line is attached to the true wetting transition and has a kink at the triple point.


MOLECULAR SCF MODELLING OF PORES IN LIPID BILAYERS<br />

M.A.C. Roelen, F.A.M. Leermakers, M.M.A.E. Claessens, M.A. Cohen Stuart<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: frans@fenk.wau.nl<br />

ABSTRACT<br />

The stability <strong>of</strong> bilayer systems, e.g. lipid vesicles, under tension is determined by the ease with which rupture<br />

nuclei form. A molecular self-consistent-field theory with the discretisation scheme <strong>of</strong> Scheutjens and Fleer is<br />

applied to obtain structural and thermodynamical information on perforated (flat) bilayers. It is found that the<br />

critical tension on the bilayer for the long wavelength perturbation is much higher than the tension needed to<br />

rupture the bilayer if a small spherical hole is allowed for.


FIRST-ORDER ADSORPTION TRANSITION OF MINORITY CHAINS IN A BRUSH<br />

A.N. Skvortsov 1) , A.A. Gorbunov 2) , F.A.M. Leermakers 3) , G.J. Fleer 3)<br />

1) Chemical Pharmaceutical Academy,<br />

Pr<strong>of</strong>. Popova 14, 197022, St. Petersburg, Russia<br />

2) Institute for highly pure biopreparations<br />

Pudozhskaya 7, 197110 St Petersburg Russia<br />

3) Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: frans@fenk.wau.nl<br />

ABSTRACT<br />

The adsorption transition <strong>of</strong> homo<strong>polymer</strong>s <strong>onto</strong> a flat solid-liquid interface is known to be second order. The<br />

surface field is typically a short-range well near the surface. In this poster we consider <strong>polymer</strong> adsorption<br />

<strong>onto</strong> a surface that is decorated by a <strong>polymer</strong> brush. Now the surface field felt by adsorbing chains is more<br />

complex. Besides a short-range attraction, there is a long-range repulsive region. The range <strong>of</strong> the repulsive<br />

part <strong>of</strong> the surface potential is given by the brush height and the strength <strong>of</strong> the repulsion (which is <strong>of</strong> the<br />

excluded-volume type) is only a function <strong>of</strong> the grafting density. The adsorption transition becomes first-order<br />

in the thermodynamic limit. Simple scaling arguments are presented.


COLLAPSE OF POLYMER BRUSHES GRAFTED ONTO PLANAR OR CONVEX S<strong>UR</strong>FACE<br />

V.M. Amoskov, T.M. Birshtein, A.A. Mercurieva<br />

Institute <strong>of</strong> Macromolecular Compounds, Academy <strong>of</strong> Science <strong>of</strong> Russia,<br />

31, Bolshoy pr., V.O., St-Petersburg, 199004, Russia<br />

email: vic@avm.macro.pu.ru<br />

ABSTRACT<br />

The collapse <strong>of</strong> <strong>polymer</strong> <strong>brushes</strong> is investigated. Four types <strong>of</strong> the <strong>polymer</strong> systems are considered: (1)<br />

polyelectrolyte <strong>brushes</strong> in poor solvent, (2) non-ionizable and ionizable <strong>brushes</strong> in binary solvents, (3)<br />

<strong>brushes</strong> with n-cluster interactions and (4) thermotropic LC <strong>brushes</strong>. Both an asymptotic analytical analysis<br />

(for chain length N →∞)<br />

and exact numerical computations (for finite N) are applied for investigating these<br />

systems in a self-consistent field (SCF) framework. The SCF numerical calculations are carried out using the<br />

formalism developed by J.Scheutjens and G.Fleer in their classical works.<br />

For the <strong>brushes</strong> <strong>of</strong> <strong>planar</strong> geometry it has been stated that under suitable external and internal conditions,<br />

collapse in all <strong>of</strong> these systems can occur as a first order phase transition. The solvent quality (system 1), the<br />

fraction <strong>of</strong> better solvent (system 2), the energy <strong>of</strong> isotropic interactions (system 3) and the energy <strong>of</strong> anisotropic<br />

interaction (system 4) are the parameters that govern corresponding phase transitions. The <strong>brushes</strong><br />

<strong>grafted</strong> <strong>onto</strong> convex curved surfaces are also investigated. It is found that for the <strong>brushes</strong> <strong>grafted</strong> <strong>onto</strong><br />

external cylindrical or spherical surfaces (combs, stars, micelles and so on), collapse may also occur as a<br />

first order phase transition.<br />

Though investigated <strong>polymer</strong> systems differ in the interactions as well as in the grafting type, it is shown<br />

that a number <strong>of</strong> features are similar for the phase transitions observed in all <strong>of</strong> them.<br />

At small (or large) enough values <strong>of</strong> the governing parameter, anyone <strong>of</strong> the systems under consideration<br />

exists only in a single-phase state, namely, in extended (or collapsed) state. However, there is an intermediate<br />

range <strong>of</strong> the governing parameter values where the swollen and the collapsed microphases may<br />

coexist, such a state being usually referred as a micro-segregated brush (MSB). It is noteworthy that a more<br />

compact phase is typically located in the neighborhood <strong>of</strong> the grafting surface, while the swollen phase is at<br />

the periphery. If the brush is in the biphasic state, its chains are never partly collapsed and partly extended:<br />

on the contrary, the chains are strictly divided into two groups that are either fully collapsed or fully extended.<br />

The fraction <strong>of</strong> the chains responsible for one or another sublayer in the brush depends on the value <strong>of</strong> the<br />

corresponding governing parameter. <strong>Collapse</strong>d chains never participate in the formation <strong>of</strong> the swollen layer,<br />

thus decreasing the effective grafting density for the remaining chains and allowing their existence in the<br />

swollen state. The free end distribution for any brush in MSB state is bimodal. All these effects are<br />

irrespective <strong>of</strong> the grafting surface type.<br />

At relatively small values <strong>of</strong> the grafting densities and finite values <strong>of</strong> the number <strong>of</strong> segments N in each<br />

chain, the brush collapse takes place in a jump-like manner, that being typical <strong>of</strong> the first order phase<br />

transitions. The size <strong>of</strong> nuclei <strong>of</strong> the collapsed phase is defined as H ~ N δ and grows with the chain length N<br />

slower than the brush size. Hence the first order phase transitions cannot exist at the limit <strong>of</strong> N →∞.


MD AND SCF ON BILAYERS OF SOPC AND SDPC<br />

F.A.M. Leermakers 1) , N.K. Balabaev 2) , A.L. Rabinovich 3)<br />

1) Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

2) Institute <strong>of</strong> Mathematical Problems <strong>of</strong> Biology, Russian Academy <strong>of</strong> Sciences,<br />

Pushchino, Moscow Region, 142292, Russia<br />

3) Institute <strong>of</strong> Biology, Karelian Research Center, Russian Academy <strong>of</strong> Sciences,<br />

Pushkinskaja str 11, Petrozavodsk, 185610, Russia<br />

ABSTRACT<br />

Structural features <strong>of</strong> two lipid bilayer membranes composed <strong>of</strong> mono-unsaturated 1-stearoyl-2-oleoyl-snglycero-3-phosphatidylcholine<br />

SOPC and 1-stearoyl-2-docosahexaenoyl-sn-glycero-3-phosphatidylchol<br />

SDPC have been evaluated by an all-atom Molecular Dynamics simulation and compared to results<br />

obtained from a detailed numerical SCF method. SCF is computationally approximately 10.000 times<br />

more efficient than MD. The two methods appear complementary. The MD gives more details. In SCF it is<br />

possible to vary parameters. It is found that unsaturated lipid tails are effectively shorter (end-to-end<br />

distance. They reduce the interdigitation <strong>of</strong> tails into opposite monolayers (they modify the coupling<br />

between the two monolayers). Both MD and SCF predict equilibrium area per molecule close to the<br />

experimental data.


S<strong>UR</strong>FACE FORCES, SUPRAMOLECULAR POLYMERS, AND DIRECTIONALITY<br />

J. van der Gucht 1,2) , N.A.M. Besseling 1) , M.A. Cohen Stuart 1)<br />

1)<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

2)<br />

Dutch Polymer Institute,<br />

Postbus 902, 5600 AX Eindhoven, The Netherlands<br />

email: klaas@fenk.wau.nl<br />

ABSTRACT<br />

Interactions between surfaces, e.g. <strong>of</strong> dispersed colloidal particles are crucial in all heterogeneous s<strong>of</strong>t<br />

matter.(1) Supramolecular equilibrium <strong>polymer</strong>s (sometimes denoted as living <strong>polymer</strong>s) are <strong>polymer</strong>ic chains<br />

formed by reversible, non-covalent association <strong>of</strong> special small molecules.(2) The main purpose <strong>of</strong> this<br />

presentation is to point out the relevance <strong>of</strong> the symmetry <strong>of</strong> the monomers for their behaviour at surfaces.<br />

Monomers that are symmetric in the sense that their two binding sites are self-complementary form nondirectional<br />

chains, but monomers with mutually complementary binding sites, <strong>of</strong>ten denoted as donor and<br />

acceptors sites (or as host and guest, lock and key) form directional chains. In solution or in bulk this<br />

difference will not lead to any difference in behaviour. However, in the presence <strong>of</strong> surfaces, and especially<br />

when two surfaces approach each other, the directionality <strong>of</strong> the monomers and the ensuing chains may be<br />

crucial.<br />

Figure 1. Non-directional chains between two surfaces (a), and directional chains between two similar surfaces that bind<br />

preferentially to the acceptor binding sites <strong>of</strong> the monomers (b).<br />

Phenomenological Landau analysis indicates that non-directional equilibrium <strong>polymer</strong>s always yield attractive<br />

surface forces, but that directional equilibrium <strong>polymer</strong>s may yield repulsion between surfaces that binds<br />

preferentially either the donors or the acceptors. More detailed information is deduced from a lattice model<br />

analysed in a Bethe-Guggenheim approximation. The strength and the range <strong>of</strong> the surface force, and their<br />

concentration dependencies is predicted to be different for the directional and the non-directional cases.<br />

These principles may explain certain observations on heterogeneous systems, and might be used to<br />

manipulate the properties <strong>of</strong> such systems.<br />

References<br />

1. J. Israelachvili, Intermolecular and Surface Forces, Academic Press, 1991.<br />

2. A. Ciferri, Supramolecular Polymers, Marcel Dekker, 2000.<br />

3. J. van der Gucht, N.A.M. Besseling, M.A. Cohen Stuart, 2002. Surface forces, supramolecular <strong>polymer</strong>s,<br />

and inversion symmetry. J. Am. Chem. Soc. 124, 6202-6205.


PHASE TRANSITIONS OF BINARY POLYMER MIXT<strong>UR</strong>ES CONFINED BETWEEN COMPETING<br />

S<strong>UR</strong>FACES<br />

K. Binder, M. Müller<br />

Institut für Physik, Johannes-Gutenberg-Universität Mainz,<br />

Staudinger Weg 7, 55099 Mainz, Germany<br />

e-mail: kurt.binder@uni-mainz.de; marcus.mueller@uni-mainz.de<br />

ABSTRACT<br />

A symmetrical <strong>polymer</strong> blend (chain lengths NA = NB = N ) is considered in a thin film geometry, treating the<br />

case where one surface preferentially attracts component A, while the other surface attracts component B.<br />

This situation leads to an interesting competition between lateral versus perpendicular phase separation<br />

between an A-rich and a B-rich phase, which is studied both with the self-consistent field theory and with<br />

Monte Carlo simulations <strong>of</strong> the bond fluctuation model on the simple cubic lattice [1-5].<br />

For simplicity, the case <strong>of</strong> “antisymmetric walls” (same strength <strong>of</strong> the surface forces) is emphasized, while<br />

the general case <strong>of</strong> asymmetric surfaces is treated only briefly. It is argued that the increase <strong>of</strong> the Flory-<br />

Huggins χ parameter that drives phase separation in the bulk (if χ > χcb<br />

= 2/N in the mean field limit)<br />

leads to a rounded transition from a uniformly mixed state to a stratified structure in the thin film geometry (an<br />

A-rich phase is separated from the B-rich phase by a single interface, while in the case <strong>of</strong> “symmetric” walls<br />

which both attract the same component a stratified structure with two interfaces parallel to the walls results).<br />

Only for a distinctly larger value χc (D), controlled by the strength <strong>of</strong> the wall-monomer interactions and<br />

related to the wetting transition <strong>of</strong> a semi-infinite <strong>polymer</strong> blend, is a sharp phase transition encountered,<br />

related to lateral phase separation. For the case where in the semi-infinite geometry one has first-order<br />

wetting, the thin film shows a triple point, and two critical points appear as remnants <strong>of</strong> prewetting criticality.<br />

For asymmetric surfaces forces, the triple point can merge with one <strong>of</strong> the critical points, and a phase<br />

diagram with a single critical point remains. Also tricritical phenomena in thin films may occur. Ginzburg<br />

criteria are developed for a discussion <strong>of</strong> the validity <strong>of</strong> the mean field description. Finally, also some<br />

comments on related experiments [6] are made.<br />

References<br />

1. Müller, M. and Binder, K., 1998. Macromolecules 31, 8323.<br />

2. Müller, M, Binder, K. and Albano, E. V., 2000. Europhys. Lett 49, 724.<br />

3. Müller, M, Binder, K. and Albano, E. V., 2000. Physica A279, 188.<br />

4. Müller, M, Albano, E. V. and Binder, K., 2000. Phys. Rev. E62, 5281.<br />

5. Müller, M. and Binder, K., 2001. Phys. Rev. E63, 021602.<br />

6. Kerle, T, Klein, J. and Binder, K., 1996; 1999. Phys. Rev. Lett. 77, 1318; Europ. Phys. J. B7, 401.


SWEET BRUSHES: SYNTHESIS AND INTERFACIAL BEHAVIO<strong>UR</strong> OF POLYSTYRENE-<br />

POLYSACCHARIDE DI-BLOCK COPOLYMERS<br />

W.T.E. Bosker 1) , K. Ágoston 2) , M.A. Cohen Stuart 1) , W. Norde 1) , J.W. Timmermans 2) ,<br />

T.M. Slaghek 3)<br />

1) Department <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

2) Agrotechnological Research Institute ATO,<br />

<strong>Wageningen</strong> <strong>UR</strong>, 6700 AA <strong>Wageningen</strong>, The Netherlands<br />

3) TNO Nutrition and Food Research,<br />

3704 HE Zeist, The Netherlands<br />

email: bosker@fenk.wau.nl<br />

ABSTRACT<br />

Research has revealed that coating surfaces with hydrophilic <strong>polymer</strong> <strong>brushes</strong> can prevent or reduce protein<br />

adsorption and hence retard bi<strong>of</strong>ouling. In biological systems the use <strong>of</strong> natural <strong>polymer</strong>s may be preferred<br />

and polysaccharides seem to be the most plausible candidate. Linear block co<strong>polymer</strong>s <strong>of</strong> polystyrene and<br />

polysaccharide were synthesized using a block synthesis method with amino terminated polystyrene and<br />

sodium cyanoborohydride as reducing agent. Different types <strong>of</strong> polysaccharides, dextrans and maltodextrins,<br />

with various molecular weights were used.<br />

IR spectroscopy indicated a successful coupling. Yields are 75 to 95 wt%. Interfacial pressure measurements<br />

<strong>of</strong> monolayers <strong>of</strong> the co<strong>polymer</strong>s showed hysteresis between successive compression and expansion<br />

cycles. This is, to a large extent, due to the slow adsorption/desorption <strong>of</strong> the polysaccharide chains at the<br />

air-water interface and the formation <strong>of</strong> aggregates <strong>of</strong> co<strong>polymer</strong> at the interface, mainly by H-bonding<br />

between adjacent polysaccharide chains. When the time scale <strong>of</strong> compression and expansion is increased<br />

the hysteresis becomes smaller and reaches a time independent level, demonstrating a permanent change in<br />

conformation <strong>of</strong> the co<strong>polymer</strong>s. The relaxation time for hysteresis is 85 minutes. Interfacial pressure<br />

isotherms differ from theoretical <strong>polymer</strong> brush model predictions. The deviation is attributed to the relative<br />

short length <strong>of</strong> the polysaccharide chains.


STABILITY OF CHARGED LIPID VESICLES: EXPERIMENTS AND SCF-CALCULATIONS<br />

M.M.A.E. Claessens, B.F. van Oort, F.A.M. Leermakers<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: mireille@fenk.wau.nl<br />

ABSTRACT<br />

The unilamellar vesicle is a stable state for both anionic DOPG and zwitterionic DOPC bialyers in salt<br />

solutions. The equilibrium size <strong>of</strong> these vesicles strongly depends on the ionic strength and the type <strong>of</strong> salt<br />

solution. Molecularly realistic self consistent field calculations show that single component vesicles have no<br />

spontaneous curvature. The predictions for the mean bending moduli kc correlate well with the experimentally<br />

found trends in vesicle radii. This gives rise to the hypothesis that the vesicles are entropically stabilized.


EFFECT OF MIXING OF TWO CHARGED PHOSPHOLIPIDS ON VESICLE SIZE AND<br />

MEAN BENDING MODULI<br />

M.M.A.E. Claessens, B.F. van Oort, F.A.M. Leermakers<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: mireille@fenk.wau.nl<br />

ABSTRACT<br />

The mean bending modulus, kc, <strong>of</strong> a bilayer <strong>of</strong> a surfactant mixture is expected to depend on the<br />

composition. It is usually assumed that kc decreases towards the equimolar composition. As the the<br />

size <strong>of</strong> entropically stabilized vesicles is related to kc, a decrease in kc will result in smaller vesicles.<br />

Previously we showed that both the anionic DOPC and the zwitterionic DOPG form vesicles that are<br />

stabilized by undulation entropy. Mixtures <strong>of</strong> these bilayer forming phospholipids indeed have a<br />

decreased equilibrium size compared to the pure component vesicles. SCF calculations were used<br />

to obtain kc for mixtures <strong>of</strong> chared surfactants. The predictions for kc correlate well with the experimentally<br />

found trends in vesicle radii. The changes in kc are not only due to the mixing <strong>of</strong> lipids,<br />

variations in charge density also affect kc.


VOLUME FRACTION PROFILES OF ADSORBED POLYMER LAYERS<br />

T. Cosgrove 1) , J. Marshall 1) , J. Hone 1) , F. Leermakers 2) , T.M. Obey 1) , J. Marshall 1)<br />

1) University <strong>of</strong> Bristol, School <strong>of</strong> Chemistry,<br />

Cantock’s Close, Bristol BS8 1TS, United Kingdom<br />

2) Laboratory <strong>of</strong> Physical and Colloid Chemistry, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 BC <strong>Wageningen</strong>, The Netherlands<br />

email: terence.cosgrove@bris.ac.uk<br />

ABSTRACT<br />

Volume fraction pr<strong>of</strong>iles for adsorbed <strong>polymer</strong>s are central to the understanding and manipulation <strong>of</strong> colloidal<br />

dispersions. Small-angle neutron scattering has proven to be an elegant method for determining the shape <strong>of</strong><br />

the volume fraction pr<strong>of</strong>ile though there are some difficulties in dealing with concentration fluctuations. In this<br />

presentation we shall develop the SANS methods for both on and <strong>of</strong>f contrast scattering and use it together<br />

with scaling and mean field theories to provide the most detailed shapes yet for adsorbed <strong>polymer</strong> layers on<br />

colloidal particles. The improvement in these methods due to better resolution in particle size and momentum<br />

transfer (Q) has made it possible to expand the studies to look at complex formation at interfaces with<br />

surfactants and sugars.<br />

Results and Discussion<br />

-1<br />

I(Q) /cm<br />

1<br />

0.1<br />

0.01<br />

Q /Å -1<br />

0.01 0.1<br />

Figure 1 Figure 2<br />

(z) φ<br />

0.1<br />

0.01<br />

0.001<br />

112.1 K PEO Exponential Pr<strong>of</strong>iles<br />

112.1 K PEO Scaling Pr<strong>of</strong>iles<br />

112.1 K PEO Scheutjens-Fleer Pr<strong>of</strong>iles<br />

0 50 100 150 200 250<br />

Figure 1 shows the scattering obtained from a deuterated polystyrene latex with adsorbed polyethylene oxide<br />

(110K Mw) suspended in water that has the same scattering length density as the latex. Under these<br />

circumstances the latex is essentially invisible. However, unlike earlier data we observe strong oscillation in<br />

the scattering from the <strong>polymer</strong> layer. This is partly due to preparing a very monodisperse latex (radii<br />

625 ± 20 Ã and 443 ± 10 Ã) and by obtaining higher Q resolution on the D22 camera at Grenoble and NG3 at<br />

NIST. The data have been fitted using an exponential volume fraction pr<strong>of</strong>ile with fluctuations (solid line) and<br />

without fluctuations (dashed line) both smeared for Q resolution. It is clear that fluctuations do contribute to<br />

the scattering and without suitable safeguards fitting the data without fluctuations can be problematical.<br />

z /Å<br />

R g 112.1 K PEO


Figure 2 show a semi-log comparison <strong>of</strong> volume fraction pr<strong>of</strong>iles obtained from fits to the on-contrast layer<br />

scattering (black lines) and interference scattering (red lines) for 112 kD PEO at full surface coverage. The<br />

calculated radius <strong>of</strong> gyration (Rg) for 112 kD PEO is also shown. Three different functional forms have been<br />

chosen to fit the data, an exponential (as in Figure 1), a scaling pr<strong>of</strong>ile and a pr<strong>of</strong>ile generated by a modified<br />

Scheutjens Fleer (SF) theory. The pr<strong>of</strong>iles in the last case overlap completely. Further examples will be<br />

shown and the results indicate that up to ca 100k Mw the SF model can quantitatively predict the scattering<br />

data. Beyond this molecular weight as tails become progressively more extended and dilute the scattering<br />

becomes progressively less sensitive to the layer extent.<br />

Other systems that will be discussed include <strong>grafted</strong> and physically adsorbed PEO layers complexing with<br />

surfactants and cyclodextrins. Figure 3 below shows the volume fraction pr<strong>of</strong>iles for a <strong>grafted</strong> PEO layer and<br />

the effect <strong>of</strong> addition <strong>of</strong> α-cyclodextrin. Cyclodextrin forms spontaneous inclusion complexes with PEO in<br />

solution that are insoluble and precipitate. In the case <strong>of</strong> a <strong>grafted</strong> layer geometric constraints would prevent<br />

a complete complexation( the cyclic sugar threads on to the chain starting at one end). The data here show<br />

that these complexes can form at interfaces, causing the chains to stiffen and expand.<br />

Volume fraction<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

Increased CD<br />

concentration<br />

5.7Å αααα cyclodextrin<br />

0 100 200<br />

z/Å<br />

300 400<br />

References<br />

1. Fleer, G.J., Cohen Stuart, M.A., Scheutjens, J.M.H.M., Cosgrove, T., Vincent, B. Polymers at Interfaces,<br />

Chapman and Hall, London, 1993.<br />

2. Auvray, L., Auroy, P. Neutron, X-Ray and Light Scattering; Linder, P. and Zemb, T., Eds., Elsevier<br />

Science Publishers, B.V., Amsterdam, 1991.<br />

3. Hone, J.H.E., Cosgrove, T., Saphiannikova, M., Obey, T.M., Crowley, T.L., 2002. Langmuir 18, 855-864.


DETAILED ATOMISTIC MONTE CARLO SIMULATION OF GRAFTED POLYMER MELTS:<br />

THERMODYNAMIC, CONFORMATIONAL AND STRUCT<strong>UR</strong>AL PROPERTIES<br />

K.Ch. Daoulas, V.G. Mavrantzas<br />

Institute <strong>of</strong> Chemical Engineering and High-Temperature Chemical Processes,<br />

ICE/HT-FORTH, GR 26500, Patras, Greece<br />

email: daoulas@physics.upatras.gr<br />

ABSTRACT<br />

The thermodynamic and conformational properties <strong>of</strong> <strong>polymer</strong> melts <strong>grafted</strong> on a solid substrate as obtained<br />

from detailed, atomistic Monte Carlo simulations with the end-bridging algorithm are presented. The interface<br />

between a basal graphite plane (as well as a non-interacting hard surface) and a bulk polyethylene (PE) melt,<br />

a few or all chains <strong>of</strong> which are <strong>grafted</strong> on the plane, has been studied. Three different PE melts, <strong>of</strong> mean<br />

molecular length C78, C156 and C250, have been investigated, at grafting densities ranging from 0.54 nm –2 to<br />

2.62 nm –2 . For melts composed <strong>of</strong> <strong>grafted</strong> and free chains, it is observed that, at moderate to high surface<br />

densities (σ ≥ 1 nm –2 ), the region close to the substrate is fully occupied by segments belonging to <strong>grafted</strong><br />

chains, which are forced by their chemical grafting to have their first segment on the interface. As the grafting<br />

density increases, free chains are progressively expelled from the surface region, in agreement with scaling<br />

arguments and the predictions <strong>of</strong> a lattice-based self-consistent mean-field (SCF) theory. For melts <strong>grafted</strong><br />

on a graphite plane, it is also seen that the local melt density in the region closest to the interface is<br />

systematically higher than in the bulk, exhibiting distinct local maxima due to <strong>polymer</strong> adsorption. Results for<br />

the chain conformation tensor demonstrate that chains are significantly stretched in the direction<br />

perpendicular to the surface, even for moderate surface densities. For the C250 PE melt at a grafting density<br />

σ = 1.31 nm –2 , for example, the average chain dimension perpendicular to the interface is 1.9 times larger<br />

than its equilibrium value in the bulk. The pr<strong>of</strong>ile <strong>of</strong> the chain end density is also seen to exhibit universal<br />

behavior in agreement with the predictions <strong>of</strong> the SCF theory. Additional results concerning the mean chain<br />

conformational path, the structure <strong>of</strong> the interfacial area for both systems studied (fully <strong>grafted</strong> and mixtures<br />

<strong>of</strong> <strong>grafted</strong> and free chain systems) and their 2 H NMR spectra are also presented. Further, the dependence <strong>of</strong><br />

all these properties on chain polydispersity will be assessed.


MODELING OF STATIONARY POLYMER DIFFUSION: APPLICATIONS OF A NON-<br />

EQUILIBRIUM SCF-METHOD<br />

S.M. Scheinhardt-Engels, F.A.M. Leermakers and G.J. Fleer<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: sonja@fenk.wau.nl<br />

ABSTRACT<br />

We present a few applications <strong>of</strong> the Mean Field Stationary Diffusion (MFSD) method. This method allows<br />

the study <strong>of</strong> diffusion in <strong>polymer</strong> systems driven by (for example) a concentration gradient. We do not follow<br />

the evolution <strong>of</strong> the system towards the stationary state, but we obtain immediately the stationary state itself,<br />

which is the situation that there is no accumulation <strong>of</strong> material anywhere in the diffusion layer. The method<br />

has the Scheutjens-Fleer result for equilibrium inhomogeneous <strong>polymer</strong> solutions as its limit when gradients<br />

in chemical potential vanish. We show volume fraction pr<strong>of</strong>iles <strong>of</strong> such stationary <strong>polymer</strong> systems. In the<br />

case <strong>of</strong> ‘color’ diffusion, where there are no interactions between the diffusing substances, the strong influence<br />

<strong>of</strong> mobility-differences is immediately seen. If there are intermolecular interactions, an interface<br />

between diffusing substances may develop even for χ χ<br />

< critical . The flux-expressions <strong>of</strong> the MFSD method<br />

are used to obtain simple analytical approximations for the coexistence curves <strong>of</strong> <strong>polymer</strong> mixtures. We show<br />

that these approximations are more accurate than other analytical approximations. Finally, we present some<br />

results on diffusion through a barrier, where we employ the possibility to monitor the <strong>polymer</strong> conformations.


A SCF STUDY OF INHOMOGENEOUS ADSORPTION OF NON-IONIC S<strong>UR</strong>FACTANTS<br />

ON HYDROPHOBIC S<strong>UR</strong>FACES<br />

A.B. Jódar-Reyes, J.L. Ortega-Vinuesa, A. Martín-Rodríguez, F.A.M. Leermakers<br />

University <strong>of</strong> Granada, Dpto. Fisica Aplicada, Facultad de Ciencias, Fuent,<br />

Granada, 18071, Spain<br />

email: ajodar@ugr.es<br />

ABSTRACT<br />

The structure <strong>of</strong> non-ionic surfactant layer adsorbed on a hydrophobic surface is studied by means the<br />

SCF-A. The possibility <strong>of</strong> formation <strong>of</strong> surface aggregates instead <strong>of</strong> homogeneous layers is included.<br />

Applying the one-gradient version <strong>of</strong> this theory in combination with the two-gradient one, we could<br />

follow the evolution <strong>of</strong> the adsorption process for several lengths <strong>of</strong> the PEO head. A branched<br />

hydrocarbon tail model that has an important effect on micellisation and adsorption results is also<br />

used. Theoretical results are compared to the experimental adsorption isotherms <strong>of</strong> Triton X-100 and<br />

Triton X-405 <strong>onto</strong> a polystyrene Latex dispersion. The process including inhomogeneous adsorption is<br />

closer to the experimental results than the homogeneous one.


THEORY AND SIMULATION OF BILAYER MEMBRANE FUSION:<br />

FREE ENERGY AND STRUCT<strong>UR</strong>E OF INTERMEDIATES<br />

K. Katsov 1) , M. Schick 1) , M. Mueller 2)<br />

1) Department <strong>of</strong> Physics, University <strong>of</strong> Washington, U.S.A.<br />

2) Institut für Physik, Johannes-Gutenberg-Universität, Mainz, Germany<br />

email: schick@mahler.phys.washington.edu<br />

ABSTRACT<br />

We combine Monte Carlo simulations and SCF theory to study the microscopic mechanism <strong>of</strong> bilayer<br />

membrane fusion. Bilayers are composed <strong>of</strong> single chain amphiphiles in an excess homo<strong>polymer</strong> solvent.<br />

The fusion mechanism we observe in simulations is very different from a widely accepted "stalk" hypothesis.<br />

Instead <strong>of</strong> the radial stalk expansion into a hemifusion diaphragm with subsequent fusion pore formation, we<br />

observe very strong bilayer destabilization after the first local connections (stalks) are formed. This<br />

destabilization leads to a local rupture <strong>of</strong> the membranes in the vicinity <strong>of</strong> the stalk with a following sealing <strong>of</strong>f<br />

<strong>of</strong> the formed microscopic holes by propagation <strong>of</strong> the initial connection (stalk).<br />

The observed mechanism is supported by our SCFT calculations <strong>of</strong> the structure and free energy <strong>of</strong><br />

different intermediates that could be involved in the fusion process. The structures that we are able to study<br />

within the SCFT include bilayer edge, 3- and 4-junctions, point and linear stalks and some transition states.<br />

We calculate both relative stability <strong>of</strong> the intermediates and barriers encountered along the alternative fusion<br />

reaction paths. We also address the problem <strong>of</strong> mixed bilayers and find relatively strong deviations from the<br />

ideal mixing.<br />

The obtained results allow us to make very definite predictions that can be checked experimentally. This<br />

includes, e.g. the effect <strong>of</strong> tension and amphiphile architecture on the fusion rate, and the fact that the<br />

membrane fusion is a leaky process.


DYNAMIC EXTENSION OF THE SELF-CONSISTENT FIELD THEORY OF<br />

INHOMOGENEOUS POLYMER SYSTEMS<br />

T. Kawakatsu<br />

Department <strong>of</strong> Physics, Tohoku University,<br />

Aza-Aoba, Aramaki, Aoba-ku, Sendai 980-8578, Japan<br />

email: kawakatu@cmpt.phys.tohoku.ac.jp<br />

ABSTRACT<br />

1. Introduction<br />

Self-consistent field (SCF) theory is one <strong>of</strong> the useful techniques for studying mesoscopic structures <strong>of</strong><br />

inhomogeneous <strong>polymer</strong> systems, such as phase separating <strong>polymer</strong> blends, <strong>polymer</strong> films on solid<br />

surfaces, and <strong>polymer</strong> <strong>brushes</strong> (Fleer 1993, Matsen 1996). Combined with a numerical evaluation <strong>of</strong> the<br />

<strong>polymer</strong> path integral, i.e. canonical statistical weight <strong>of</strong> the chain conformation, this theory well reproduces<br />

the equilibrium domain structures and segment density pr<strong>of</strong>iles. Despite <strong>of</strong> such success in<br />

equilibrium systems, extensions to dynamic non-equilibrium systems have not completely understood yet.<br />

In the present study, we propose several techniques to introduce dynamic processes into the SCF theory,<br />

and evaluate their usefulness and efficiency.<br />

2. Dynamic SCF method for weakly non-equilibrium systems<br />

The dynamic extension <strong>of</strong> the SCF theory has been initiated by Fraaije (Fraaije 1993) by combining<br />

the SCF theory with the Fick’s law <strong>of</strong> segment diffusion driven by the gradient <strong>of</strong> the chemical potential.<br />

To obtain the chemical potential, the system is assumed to be in local equilibrium, which is essential for<br />

the use <strong>of</strong> the path integral formalism. Using this technique, one can calculate the ordering processes in<br />

phase separating <strong>polymer</strong> blends or block co<strong>polymer</strong> systems. For example, we can reproduce the phase<br />

diagram <strong>of</strong> a mixture <strong>of</strong> long and short block co<strong>polymer</strong>s that shows complex domain structures (Morita<br />

2002). Using the segment interaction parameters and the block ratio (the ratio between the lengths <strong>of</strong> the<br />

two blocks) as input parameters, the dynamic SCF calculation gives a quantitatively correct phase<br />

diagram and domain structures including metastable transient structures.<br />

The success <strong>of</strong> these dynamic SCF simulations owes to the weak non-equilibrium nature <strong>of</strong> the<br />

phenomena, which is crucial to the use <strong>of</strong> the path integral formalism <strong>of</strong> the <strong>polymer</strong> chain conformation.<br />

3. Highly non-equilibrium systems -- Rheological properties --<br />

As in the case <strong>of</strong> a sheared <strong>polymer</strong> melt, an external deformation force makes the chain conformation<br />

strongly stretched. In such a situation, the assumption <strong>of</strong> the local equilibrium becomes unreliable. To<br />

overcome this difficulty, we extend the path integral formalism by introducing the bond orientation tensor,<br />

and combined it with the reptation dynamics (Kawakatsu 2001, Shima 2002). In this formalism, the<br />

<strong>polymer</strong> path integral Q ( s, r; s′ , r ' ) , i.e. the statistical weight <strong>of</strong> a subchain between the s -th and s′ -th<br />

segments whose ends are found at r and r ′ , is given by<br />

∂<br />

1<br />

Q( s, r; s′ , r') = ∇∇: { A( s, r) Q( s, r; s′ , r') } −∇⋅{ u( s, r) Q( s, r; s′ , r') } −βV(<br />

r) Q( s, r; s′ , r ')<br />

,<br />

∂s<br />

2<br />

where u( s,<br />

r ) and A( s,<br />

r ) are the average and covariance <strong>of</strong> the distribution <strong>of</strong> the s -th bond vector,<br />

and V ( r ) is the self-consistent field. The temporal evolutions <strong>of</strong> u( s,<br />

r ) and A( s,<br />

r ) are determined by<br />

the reptation theory and the Navier-Stokes equation (Kawakatsu 2001, Shima 2002). A similar treatment<br />

has recently been used by Fredrickson (Fredrickson, 2002).<br />

We applied this method to a sheared two parallel plates on which melt <strong>brushes</strong> are <strong>grafted</strong>. Due to the<br />

disentanglement between the two <strong>brushes</strong>, the shear stress and the density pr<strong>of</strong>ile <strong>of</strong> the segments<br />

changes as is shown in Figure 1. Such a behavior is consistent with a microscopic Monte Carlo<br />

simulation using a many chain system.


Figure 1. Time evolution <strong>of</strong> the shear stress and the change in the segment density pr<strong>of</strong>ile.<br />

4. Conclusions<br />

We have proposed a new method to simulate rheological properties <strong>of</strong> the dense <strong>polymer</strong> systems using<br />

SCF technique. By applying it to sheared <strong>polymer</strong> <strong>brushes</strong>, we demonstrated its usefulness. This technique<br />

could also be applied to many non-equilibrium phenomena in dense <strong>polymer</strong> systems, such as<br />

adhesion, phase separation, and so on.<br />

References<br />

1. Fleer, G.J., Cohen Stuart, M.A., Scheutjens, J.M.H.M., Cosgrove, T., and Vincent, B. Polymers at<br />

Interfaces. Chapman & Hall, 1993.<br />

2. Fraaije, J.G.E.M., 1993. Dynamic density functional theory for microphase separation kinetics <strong>of</strong> block<br />

co<strong>polymer</strong> melts. J. Chem. Phys., 99 (11), 9202-9212.<br />

3. Fredrickson, G., Dynamics and rheology <strong>of</strong> inhomogeneous <strong>polymer</strong>ic fluids. A complex Langevin<br />

approach, preprint.<br />

4. Kawakatsu, T., 2001. Mesoscale modellings <strong>of</strong> multiphase <strong>polymer</strong> systems using density functional<br />

approaches, The Knowledge Foundation’s 2 nd International Conference on “Multiscale modeling –<br />

Bridging the gap between atomistic and mesoscales –“, Boston, 2001.<br />

5. Matsen, M. and Bates, F., 1996. Unifying weak- and strong-segregation block co<strong>polymer</strong> theories.<br />

Macromolecules, 29 (4), 1091-1098.<br />

6. Morita, H., Kawakatsu, T., Doi, M., Yamaguchi, D., Takenaka, M., and Hashimoto, T., 2002.<br />

Competition between micro- and macrophase separations in a binary mixture <strong>of</strong> block co<strong>polymer</strong>s. A<br />

dynamic density functional study. Macromolelues, in press.<br />

7. Shima, T., Kuni, H., Okabe, Y., Doi, M., Yuan, X.-F., and Kawakatsu, T., 2002. Self-consistent field<br />

theory <strong>of</strong> viscoelastic behavior <strong>of</strong> inhomogeneous dense <strong>polymer</strong> systems, Macromolecules,<br />

submitted.


ADSORPTION OF PERIODIC AND RANDOM AB COPOLYMERS<br />

L. Klushin 1,2) , E. Maraachlian 1) , A. Skvortsov 3)<br />

1) American University <strong>of</strong> Beirut, Department <strong>of</strong> Physics, Beirut, Lebanon<br />

2)<br />

Institute for Macromolecular Compounds, Russian Academy <strong>of</strong> Sciences,<br />

St. Petersburg, Russia<br />

3)<br />

Chemical-Pharmaceutical Academy, St. Petersburg, Russia<br />

email: leo@aub.edu.lb<br />

ABSTRACT<br />

The goal <strong>of</strong> this work is to compare the effects <strong>of</strong> composition for random and periodic (multiblock) AB<br />

co<strong>polymer</strong>s on their localization at liquid-liquid and solid-liquid interfaces. Recently, it was extensively<br />

demonstrated that random monomer distribution promotes localization <strong>of</strong> co<strong>polymer</strong>s at liquid-liquid<br />

interface as compared to periodic co<strong>polymer</strong>s <strong>of</strong> the same average composition [1-3]. Here we<br />

investigate the effects <strong>of</strong> composition on adsorption at a <strong>planar</strong> solid surface. The energy <strong>of</strong> contact for<br />

a monomers A with the surface is ε , while the surface is assumed to be inert for monomers B<br />

(ε = 0 ). The energy <strong>of</strong> interaction per monomer averaged over composition is thus ε = f ε , where f<br />

B<br />

is the fraction <strong>of</strong> A monomers. An ideal lattice chain model is used to calculate numerically the thermodynamic<br />

average energy per segment and to determine the critical point <strong>of</strong> adsorption. We also<br />

calculate the average thickness <strong>of</strong> the adsorbed layer.<br />

For regular co<strong>polymer</strong>s (periodic composition) in the range 1≤f ≤0.3,<br />

the critical point <strong>of</strong><br />

adsorption is defined up to high accuracy by the composition average energy f ε = ε (homo) where<br />

c c<br />

ε (homo) is the critical energy for a homo<strong>polymer</strong>. This is in agreement with the earlier results <strong>of</strong> Di<br />

c<br />

Marzio et.al. [4]. We find that the dependence <strong>of</strong> the number <strong>of</strong> contacts and the thickness <strong>of</strong> the<br />

adsorbed layer on ε in the vicinity <strong>of</strong> the critical point is well described by the analytical expressions<br />

for a homo<strong>polymer</strong> with the effective energy per monomer ε .<br />

For adsorption <strong>of</strong> random co<strong>polymer</strong>s at a solid-liquid surface, a theory based on replica treatment<br />

[5] predicts the critical adsorption energy to be lower than expected from simple composition average<br />

arguments mentioned above, the deviation being especially prominent around f = 0.5. Our numerical<br />

calculations for chains <strong>of</strong> <strong>polymer</strong>ization index up to N = 10 4 show that the difference in the critical<br />

energy between regular and random co<strong>polymer</strong>s is quite negligible in the composition range<br />

1≤f ≤0.5.<br />

We develop an alternative analytical theory based on the trial potential approach<br />

introduced in [3] that supports this result.<br />

For lower values <strong>of</strong> f, (f < 0.2), the critical energy <strong>of</strong> adsorption for random co<strong>polymer</strong>s becomes<br />

noticeably lower than that for periodic chains. Both critical energies deviate significantly from the<br />

simple composition average prediction. For periodic co<strong>polymer</strong>s, the order parameter and the thickness<br />

<strong>of</strong> the adsorbed layer as functions <strong>of</strong> ( ε − εс) εcare<br />

still well described by “effective homo<strong>polymer</strong>”<br />

expressions. For random co<strong>polymer</strong>s, this description fails.<br />

References<br />

1. Sommer, J.-U., Daoud, M., 1995. Europhys. Let. 32, 407.<br />

2. Dai, C.-A., et.al., 1994; 1995. Phys. Rev. Lett. 73, 2472; 74, 2837.<br />

3. Chen, Z.Y., 1999; 2000. J. Chem. Phys. 111, 5603; 112, 8665.<br />

4. Di Marzio, E.A., Guttman, C.M., Ma, A., 1995. Macromolecules 28, 2930.<br />

5. Gutman, L., Chakraborty, A., 1995. J. Chem. Phys. 103, 10733.


SALT EFFECT TO RAYLEIGH DROPLETS OF HYDROPHOBIC POLYELECTROLYTES<br />

H. Kodama<br />

Mitsubishi Chemical Corporation,<br />

Naruse 2-17-1-17-105, Machida, 194-0044, Japan<br />

email: kodama@rc.m-kagaku.co.jp<br />

ABSTRACT<br />

A finite size droplets (aggregates) <strong>of</strong> weakly charged polyelectrolytes in poor solvents are analyzed in<br />

terms <strong>of</strong> mean-field Poisson-Boltzmann + Edwards approach. The standard chemical potential per<br />

chain is calculated as a function <strong>of</strong> the number <strong>of</strong> polyelectrolyte molecules in a droplet for a given salt<br />

concentration, and the droplet size distribution is computed. It will be shown that a classical Rayleightype<br />

argument [1] breaks down because <strong>of</strong> the counterion condensations. For low salt concentration, a<br />

finite size droplets can stably exist, while infinite aggregates emerge (macrophase separation) for high<br />

salt concentration.<br />

A simulation program "SUSHI" in the "OCTA" system (http://octa.jp) is used to perform the above<br />

calculations.<br />

Reference<br />

Joanny, J. F. and Leibler, L., 1990. J. Phys. France 51 545.


ELECTRIC FIELD VERSUS S<strong>UR</strong>FACE ALIGNMENT IN CONFINED FILMS OF A BLOCK<br />

COPOLYMER MELT<br />

A.V. Kyrylyuk, G.J.A. Sevink, A.V. Zvelindovsky, J.G.E.M. Fraaije<br />

Leiden University,<br />

P.O.Box 9502, 2300 RA Leiden, The Netherlands<br />

email: a.kiriluk@chem.leidenuniv.nl<br />

ABSTRACT<br />

The dynamics <strong>of</strong> alignment af a block co<strong>polymer</strong> microstructure in thin films in the presence <strong>of</strong> an<br />

external electric filed has been studied numerically. We considered in detail a symmetric diblock<br />

co<strong>polymer</strong> melt, exhibiting a lamellar morphology. The method used is a dynamic mean-field density<br />

functional method, derived from the generalized time-dependent Ginzburg-Landau theory. We allow<br />

for a slight compressibility by adding a Helfand penalty function to the free energy. We investigated<br />

the effect <strong>of</strong> an electric filed on block co<strong>polymer</strong>s under the assumption that the dipolar interaction<br />

induced due to the fluctuations <strong>of</strong> composition pattern is a dominant mechanism <strong>of</strong> electric field<br />

induced domain alignment. As a result, depending on the ratio between electric filed and interfacial<br />

interactions, perpendicular, parallel or intermediate (mixed) final lamellar structures were obtained.<br />

The intermediate mixed structures remains to be stable, though these structures do not correspond to<br />

the minimum <strong>of</strong> the free energy. While alignment process, the mesophases passes through<br />

fascinating states <strong>of</strong> coexistence <strong>of</strong> several lamellae clusters with different orientations.


MACROMOLECULES AT INTERFACES, A FLEXIBLE THEORY FOR HARD SYSTEMS<br />

F.A.M. Leermakers<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: frans@fenk.wau.nl<br />

ABSTRACT<br />

On January 11 1985 Jan Scheutjens defended his thesis entitled “Macromolecules at interfaces, a flexible<br />

theory for hard systems”. In this thesis the Scheutjens-Fleer self-consistent field SF-SCF theory is presented<br />

and applied mainly, but not exclusively, to homo<strong>polymer</strong>s at interfaces. It condensed the work done over a<br />

period <strong>of</strong> about 10 years in close harmony with Gerard Fleer and several MSc students <strong>of</strong> whom I was one,<br />

into a booklet <strong>of</strong> 168 pages. In retrospect one can say that Jan re-invented (almost single-handedly) the selfconsistent-field<br />

method. In the early days (before 1980) there was no knowledge in laboratory <strong>of</strong> Physical<br />

Chemistry and Colloid Science in <strong>Wageningen</strong> about comparable approaches in quantum mechanics. Also<br />

the close relation <strong>of</strong> his work to the Edwards diffusion equation was not known. The work has long been (and<br />

still is) associated to lattice models. The reason for this is clear, because the deep understanding <strong>of</strong> the<br />

theory was based on counting conformations and interactions <strong>of</strong> <strong>polymer</strong> chains on a lattice. The goal <strong>of</strong> this<br />

lecture is to prove that the discretisation scheme <strong>of</strong> Scheutjens (and Fleer) <strong>of</strong> the self-consistent field<br />

equations for <strong>polymer</strong> adsorption is even today a very powerful, flexible tool to obtain better insights in hard<br />

systems such as macromolecules at interfaces. I will do this in the context <strong>of</strong> the state <strong>of</strong> the art 10 years ago<br />

(when Jan passed away).<br />

I will discuss a new self-consistent-field method, that better than in the classical SF-SCF approach,<br />

accounts for the intra-molecular excluded-volume effects <strong>of</strong> the <strong>polymer</strong> molecules. The method that has<br />

been developed together with Ph.D’s Jan van Male and Jos van den Oever, features translationally constraint<br />

molecules in a two-gradient cylindrical coordinate system. The equilibrium between chains in the bulk and<br />

near the surface can only be established by hand. For this the accurate evaluation <strong>of</strong> appropriate thermodynamic<br />

quantities is necessary. A selection <strong>of</strong> applications will be discussed.<br />

Homo<strong>polymer</strong> adsorption<br />

The theory <strong>of</strong> homo<strong>polymer</strong>s at solid–liquid interfaces was the first problem considered by Jan<br />

Scheutjens. It also turns out to be (numerically) one <strong>of</strong> the hardest problems. The adsorption-desorption<br />

transition in <strong>polymer</strong> adsorption is smooth. It is shown that the <strong>polymer</strong> conformation near the adsorption<br />

transition transforms from the coil to the pancake. The radius scales as R ∝ α<br />

g N , whereα ≈ 0.6 for the coil<br />

and α = 3 / 4 for the pancake.<br />

Ten years ago, the self-similar scaling in the central region <strong>of</strong> the <strong>polymer</strong> adsorption pr<strong>of</strong>ile was basically<br />

understood (Van der Linden worked in close cooperation with the Strasbourg group on this issue) from a<br />

ground-state approximation. However, the mean-field power law ϕ ( ) ∝ −2<br />

z z that was found in the limit <strong>of</strong> long<br />

chains and strong adsorption, <strong>of</strong> course deviated from the prediction <strong>of</strong> De Gennes: ϕ ( ) ∝ −4/3<br />

z z . The failure<br />

was <strong>of</strong>ten attributed to the fact that ‘correlations’ were neglected. The new method (still being a mean-field<br />

theory) is in reasonable agreement with the result <strong>of</strong> De Gennes.


Polyelectrolyte adsorption<br />

Van der Steeg in our group considered polyelectrolyte adsorption in the limit <strong>of</strong> electrostatic interactions<br />

only. She showed that polyelectrolytes are attracted to an oppositely charged surface. The gain in entropy <strong>of</strong><br />

the small ions is the driving force for this. The strongest adsorption and the highest adsorbed amount, was<br />

found in the limit <strong>of</strong> low ionic strength. In the absence <strong>of</strong> non-electrostatic interactions no charge overcompensation<br />

was found. Accounting for intramolecular excluded-volume effects shows that this result<br />

should be reconsidered: I will report on the overcompensation <strong>of</strong> the surface charge that is found for<br />

polyelectrolyte stars electrostatically attracted to an oppositely charged surface (work together with Zhulina).<br />

Co<strong>polymer</strong> (protein analogues) near interfaces<br />

Jan Scheutjens was extremely interested in biological applications. The work on lipid bilayer membranes<br />

was just one example. I will show how the new approach can be used to obtain interesting results for the<br />

adsorption <strong>of</strong> charged partially collapsed co<strong>polymer</strong>s (protein analogues) adsorbing on charged surfaces.<br />

This work is done in collaboration with Jos van den Oever who carefully worked out the electrostatics in a<br />

two-gradient approach for systems in which the dielectric permittivity is not homogeneous.<br />

The effects <strong>of</strong> tails <strong>of</strong> adsorbed <strong>polymer</strong>s on the colloidal stability<br />

The effect <strong>of</strong> <strong>polymer</strong>s on the colloidal stability and the importance <strong>of</strong> tails were probably the two main<br />

insights <strong>of</strong>ten pushed by Jan Scheutjens. However, at full equilibrium, the interaction induced by the <strong>polymer</strong><br />

chains was found to be purely attractive. Now we know that this result shows that in fact the loops dominate<br />

the pair interaction. Only recently Semenov showed that there is a very small repulsive contribution <strong>of</strong> the<br />

tails that can be seen when the surfaces are a distance H ≈ R g apart. Together with Jan van Male I<br />

considered the adsorption <strong>of</strong> homo<strong>polymer</strong>s on a surface <strong>onto</strong> which short chains were <strong>grafted</strong>. This ‘fur’<br />

layer suppresses short loops. As a consequence tails dominate the adsorbed layer characteristics. In this<br />

case it is possible to obtain purely repulsive interaction potential for homo<strong>polymer</strong> adsorbing in full<br />

equilibrium. If time allows I will address this issue as well. It is already found for the classical SF-SCF theory.<br />

It is necessary to mention that our group still benefits from what Jan Scheutjens left behind. The best way<br />

to illustrate this is to mention that the SCF equations can usually only be solved numerically. Jan Scheutjens<br />

wrote a generic solver based on a Newton-like minimization algorithm optimized to tackle these ill-behaving<br />

functions. This solver is the hart <strong>of</strong> the computer program SFBox that has been used to generate results<br />

presented in this lecture. For good order, Jan van Male did most <strong>of</strong> the implementations <strong>of</strong> this unique<br />

program making use <strong>of</strong> many ‘old’ tricks.


APPLICATION OF HETEROGENEOUS LATTICE MEAN-FIELD THEORY APPLIED TO<br />

SYSTEMS WITH INTERNAL DEGREES OF FREEDOM, POLYELECTROLYTES, AND<br />

CAPILLARY INDUCED PHASE SEPARATION<br />

P. Linse<br />

Center for Chemistry and Chemical Engineering, Lund University,<br />

P.O. Box 124, S-221 00 Lund, Sweden<br />

email: per.linse@fkem1.lu.se<br />

ABSTRACT<br />

1. Introduction<br />

The impact <strong>of</strong> the ideas in the seminal papers by Jan Scheutjens and Gerard Fleer (SF) presented 1979<br />

and 1980 (Scheutjens 1979, Scheutjens 1980) can hardly be exaggerated. Today, about 200 papers have<br />

been published containing calculations based on their ideas, and ca. 10 doctoral theses have been devoted<br />

to developments and applications <strong>of</strong> the theory by SF.<br />

The start <strong>of</strong> my acquaintance with the SF-theory appeared in the second half <strong>of</strong> the 80:ies, when the<br />

original papers <strong>of</strong> SF came in my hand. At that time, I gave the PhD student Mikael Björling the task to merge<br />

the SF-theory with another idea <strong>of</strong> representing <strong>polymer</strong>s with internal states. Since then, the original SFtheory<br />

and its later numerous extensions have been one <strong>of</strong> the main theoretical tool in my research.<br />

Throughout, we have close contacts with the <strong>Wageningen</strong> group and heavily utilized their developments.<br />

The aim <strong>of</strong> my presentation is to (i) provide an overview <strong>of</strong> the developments and applications made in the<br />

Lund group during the last decade and to give a more in-dept account <strong>of</strong> three different areas, (ii.a) systems<br />

with internal states, (ii.b) polyelectrolytes, and (iii.c) capillary induced phase separation.<br />

2. Overview<br />

The SF-theory has been, and is still, used in our laboratory to provide a deepen understanding <strong>of</strong> <strong>polymer</strong><br />

in solution. The focus has been to apply the theory on systems currently under experimental investigation. In<br />

many occasions, it has been an extremely valuable tool for improving the analysis <strong>of</strong> experimental results.<br />

However, the results <strong>of</strong> the SF-theory has also been employed to examine scaling predictions and been<br />

confronted with results from Monte Carlo simulations.<br />

The overview will contain examples covering phase diagram <strong>of</strong> ethylene oxide (EO) and propylene oxide<br />

(PO) containing <strong>polymer</strong>s and co<strong>polymer</strong>s, micellization <strong>of</strong> corresponding co<strong>polymer</strong>s, adsorption <strong>of</strong> these<br />

co<strong>polymer</strong>s at interfaces/surfaces, partitioning <strong>of</strong> proteins in <strong>polymer</strong> two-phase systems, forces between<br />

solid surfaces mediated by block co<strong>polymer</strong>s. In addition, significant effort has been made to investigate the<br />

phase behavior in <strong>polymer</strong> mixtures containing polyelectrolytes, adsorption <strong>of</strong> polyelectrolytes to solid<br />

surfaces, and the self-association <strong>of</strong> charged block co<strong>polymer</strong>s. Moreover, examples <strong>of</strong> properties <strong>of</strong> novel<br />

charged block co<strong>polymer</strong>s <strong>grafted</strong> at a solid support will be provided.<br />

3. Systems with internal degrees <strong>of</strong> freedom<br />

An important class <strong>of</strong> <strong>polymer</strong>s as poly(propylene oxide) (PEO) display an decreasing solubility in water<br />

upon a temperature increase. This behavior is not captured by the standard Flory-Hugins theory for <strong>polymer</strong><br />

solutions. Instead <strong>of</strong> using an ad-hoc temperature dependent χ-parameter, Gunnar Karlström introduced on<br />

the basis <strong>of</strong> quantum mechanical calculations the concept <strong>of</strong> internal states (Karlström 1985). With this<br />

concept, a framework for describing the reverse temperature dependence was established. Despite that still a<br />

few parameters needs to be fitted, here to the phase diagram <strong>of</strong> PEO in aqueous solution, this approach<br />

provides a mechanism and understanding for the reverse temperature dependence.


An important step in our theoretical developments was the fusion <strong>of</strong> the SF-theory (at that time<br />

generalized by Olav Evers et al. (Evers 1990) to arbitrary number <strong>of</strong> components, co<strong>polymer</strong>s etc) with the<br />

concept <strong>of</strong> internal states (Linse 1991). This theoretical foundation was then extensively utilized to in great<br />

detail describe the self-association <strong>of</strong> PEO-PPO-PEO block co<strong>polymer</strong>s, the role <strong>of</strong> composition (Linse<br />

1993a), sphere-to-rod transition (Linse 1993b), impurities (Linse 1994a), and polydispersity (Linse 1994b) as<br />

well as the appearance <strong>of</strong> lyotropic liquid crystalline phases (Noolandi 1996, Svensson 1999) and the<br />

adsorption to surfaces and interfaces (Tiberg 1991, Malmsten 1992, Linse 1997). These predictions were<br />

made with essentially no fitted parameters <strong>of</strong> the actual systems. All parameters used (except those involving<br />

solid surfaces) were taken from previous studies, where they were determined by fitting to phase diagrams <strong>of</strong><br />

simple solutions <strong>of</strong> homo<strong>polymer</strong>s.<br />

4. Polyelectrolytes<br />

The properties <strong>of</strong> homogeneous systems containing polyelectrolytes can be successfully addressed by<br />

extending the Flory-Huggins theory and, similarly, properties <strong>of</strong> heterogeneous systems containing<br />

polyelectrolytes can be dealt within the SF-theory. In the former case, the simplest but yet powerful approach<br />

is to enforce charge neutrality <strong>of</strong> the individual phases being in equilibrium (Gottschalk 1998). In the latter<br />

case, the SF-theory was extended with the Poisson equation (Böhmer 1991), and typically the electrostatics<br />

are described on the same level as in the Poisson-Boltzmann equation, although subjected to an<br />

descretization implied by the layers.<br />

Examples <strong>of</strong> predictions <strong>of</strong> the phase behavior <strong>of</strong> solutions polyelectrolyte and mixtures <strong>of</strong> <strong>polymer</strong>s based<br />

on the extended Flory-Huggins theory will be given (Gottschalk 1998). Thereafter, the some selected<br />

properties <strong>of</strong> polyelectrolytes adsorbed or <strong>grafted</strong> to solid surfaces will be presented (Shubin 1995, Linse<br />

1996, Shusharina 2001).<br />

5. Capillary induced phase separation<br />

The capillary condensation may appear in many shapes and contexts. A few years ago, it displayed still<br />

another one. Experimentally, it was discovered (Freyssingeas 1998) that a mixture <strong>of</strong> two <strong>polymer</strong>s close to a<br />

segregative phase separation may give rise to a very long-range attraction between to surfaces inserted in<br />

the <strong>polymer</strong> mixture. The range <strong>of</strong> attraction exceeds far the <strong>polymer</strong> extension and has a thermodynamic<br />

origin and arises from a capillary induced phase separation (CIPS) (Wennerström 1998).<br />

Lattice mean-field calculations have been performed and provided an enhanced understanding <strong>of</strong> this<br />

phenomenon. An account <strong>of</strong> the long-range attraction associated with CIPS will be given (Joabsson 2002).<br />

Here, the lattice mean-field theory provided convicting argument supporting previous analytic explanation <strong>of</strong><br />

the long-range attraction. Presently, similar calculations are performed on polydisperse <strong>polymer</strong> solutions and<br />

solution <strong>of</strong> a single <strong>polymer</strong>.<br />

References<br />

1. Böhmer, M.R., Koopal, L.K. and Lyklema, J., 1991. Micellization <strong>of</strong> ionic surfactants. Calculations based<br />

on a self-consistent field lattice model. J. Phys. Chem., 1991, 9569.<br />

2. Evers, O.A., Scheutjens, J.M.H.M. and Fleer, G.J., 1990. Statistical thermodynamics <strong>of</strong> block co<strong>polymer</strong><br />

adsorption. 1. Formulation <strong>of</strong> the model and results for the adsorbed layer structure. Macromolecules,<br />

23, 5221.<br />

3. Freyssingeas, E., Thuresson, K., Nylander, T., Joabsson, F. and Lindman, B., 1998. A surface force,<br />

light scattering, and osmotic pressure study <strong>of</strong> semidilute aqueous solutions <strong>of</strong> ethyl(hydroxyethyl)<br />

cellulose - long-range attraction forces between two <strong>polymer</strong> coated surfaces. Langmuir, 14, 5877.<br />

4. Gottschalk, M., Linse, P. and Piculell, L., 1998. Phase stability <strong>of</strong> polyelectrolyte solutions as predicted<br />

from lattice mean-field theory. Macromolecules, 31, 8407.<br />

5. Joabsson, F. and Linse, P., 2002. Capillary-induced phase separation in mixed <strong>polymer</strong> solutions. A<br />

lattice mean-field calculation study. Langmuir, 106, 3827.<br />

6. Karlström, G., 1985. A new model for upper and lower critical solution temperature in poly(ethylene<br />

oxide) solutions. J. Phys. Chem., 89, 4962.<br />

7. Linse, P., 1993a. Micellization <strong>of</strong> poly(ethylene oxide)-poly(propylene oxide) block co<strong>polymer</strong>s in<br />

aqueous solution. Macromolecules, 26, 4437.


8. Linse, P., 1993b. Phase behavior <strong>of</strong> poly(ethylene oxide)-poly(propylene oxide) block co<strong>polymer</strong>s in<br />

aqueous solution. J. Phys. Chem., 97, 13896.<br />

9. Linse, P., 1994a. Micellization <strong>of</strong> poly(ethylene oxide)-poly(propylene oxide) block co<strong>polymer</strong> in aqueous<br />

solution: Effect <strong>of</strong> <strong>polymer</strong> impurities. Macromolecules, 27, 2685.<br />

10. Linse, P., 1994b. Micellization <strong>of</strong> poly(ethylene oxide)-poly(propylene oxide) block co<strong>polymer</strong>s in<br />

aqueous solution: Effect <strong>of</strong> <strong>polymer</strong> polydispersity. Macromolecules, 27, 6404.<br />

11. Linse, P., 1996. Adsorption <strong>of</strong> weakly charged polyelectrolytes at oppositely charged surfaces. Macromolecules,<br />

29, 326.<br />

12. Linse, P. and Björling, M., 1991. Lattice theory for multicomponent mixtures <strong>of</strong> co<strong>polymer</strong>s with internal<br />

degrees <strong>of</strong> freedom in heterogeneous systems. Macromolecules, 24, 6700.<br />

13. Linse, P. and Hatton, T.A., 1997. Mean-field lattice calculations <strong>of</strong> ethylene oxide and propylene oxide<br />

containing homo<strong>polymer</strong>s and triblock co<strong>polymer</strong>s at the air/water interface. Langmuir, 13, 4066.<br />

14. Malmsten, M., Linse, P. and Cosgrove, T., 1992. Adsorption <strong>of</strong> peo-ppo-peo block co<strong>polymer</strong>s at silica.<br />

Macromolecules, 25, 2474.<br />

15. Noolandi, J., Shi, A.-C. and Linse, P., 1996. Theory <strong>of</strong> phase behavior <strong>of</strong> poly(oxyethylene)-poly-<br />

(oxypropylene) - poly(oxyethylene) triblock co<strong>polymer</strong>s in aqueous solutions. Macromolecules, 29, 5907.<br />

16. Scheutjens, J.M.H.M. and Fleer, G.J., 1979. Statistical theory <strong>of</strong> the adsorption <strong>of</strong> interacting chain<br />

molecules. 1. Partition function, segment density distribution and adsorption isotherms. J. Phys. Chem.,<br />

83, 1619.<br />

17. Scheutjens, J.M.H.M. and Fleer, G.J., 1980. Statistical theory <strong>of</strong> the adsorption <strong>of</strong> interacting chain<br />

molecules. 2. Train, loop, and tail size distributions. J. Phys. Chem., 84, 178.<br />

18. Shubin, V. and Linse, P., 1995. Effect <strong>of</strong> electrolytes on adsorption <strong>of</strong> cationic polyacrylamide on silica:<br />

Ellipsometric study and theoretical modeling. J. Phys. Chem., 99, 1285.<br />

19. Shusharina, N.P. and Linse, P., 2001. Oppositely charged polyelectrolytes <strong>grafted</strong> <strong>onto</strong> <strong>planar</strong> surfaces:<br />

Mean-field lattice theory. Eur. Phys. J. E6, 147.<br />

20. Svensson, M., Alexandridis, P. and Linse, P., 1999. Phase behaviour and microstructure in binary block<br />

co<strong>polymer</strong>/selective solvent systems: Experiment and theory. Macromolecules, 32, 637.<br />

21. Tiberg, F., Malmsten, M., Linse, P. and Lindman, B., 1991. Kinetic and equilibrium aspects <strong>of</strong> block<br />

co<strong>polymer</strong> adsorption. Langmuir, 7, 2723.<br />

22. Wennerström, H., Thuresson, K., Linse, P. and Freyssingeas, E., 1998. Long range attractive surface<br />

forces due to capillary-induced <strong>polymer</strong> incompatibility. Langmuir, 14, 5664.


MORPHOLOGIES IN THIN TRIBLOCK COPOLYMER FILMS<br />

K.S. Lyakhova 1) , A. Horvat 2) , G.J.A. Sevink 2) , A.V. Zvelindovsky 1) , R. Magerle 2) ,<br />

J.G.E.M. Fraaije 1)<br />

1)<br />

S<strong>of</strong>t Condensed Matter Group, Leiden Institute <strong>of</strong> Chemistry,<br />

P.O. Box 9502, 2300 RA Leiden, The Netherlands<br />

2)<br />

Physicalishe Chemie II, University Bayreuth,<br />

D-95440 Bayreuth, Germany<br />

email: e.lyakhova@chem.leidenuniv.nl<br />

ABSTRACT<br />

The physics behind morphology formation in thin block co<strong>polymer</strong> films is not well understood. Recent<br />

experiments with free polystyrene- polybutadiene-polystyrene (SBS) films (in collaboration with Bayreuth<br />

University) result in complex morphologies, with terrace formation at integer domain distance film heights and<br />

connecting structures that are not present in bulk. In an article recently published [1], we compared these<br />

experiments with DDFT simulations for a geometry confined between two hard walls and symmetric <strong>polymer</strong>wall<br />

interaction. We identified two important parameters, the width <strong>of</strong> the slit and the surface interaction, and<br />

two relevant mechanisms: frustration and surface reconstruction. The experiments and the simulations match<br />

perfectly. In the present study, we extend our simulations to the more general case <strong>of</strong> asymmetric <strong>polymer</strong>wall<br />

interactions. The resulting asymmetry in surface reconstruction, combined with frustration, leads to a<br />

much richer phase diagram with many so-called hybrid structures. We compare the results to the symmetric<br />

phase diagram, and identify, similar to the symmetric case, the relevant mechanisms.<br />

Reference<br />

Knoll, A., Horvat, A., Lyakhova, K.S., Kraush, G., Sevink, G.J.A., Zvelindovsky, A.V., Magerle, R., 2002.<br />

Physical Review Letters 89 (3), 035501.


MODELING INTERFACIAL EFFECTS ON THE CONFORMATION AND RHEOLOGY OF<br />

POLYMER SOLUTIONS THROUGH A HIERARCHICAL, CONTINUUM MODEL<br />

V.G. Mavrantzas 1) , A.N. Beris 1,2)<br />

1)<br />

Institute <strong>of</strong> Chemical Engineering and High-Temperature Chemical Processes,<br />

ICE/HT-FORTH, GR 26500, Patras, Greece<br />

2)<br />

University <strong>of</strong> Delaware, Department <strong>of</strong> Chemical Engineering,<br />

Newark, DE 19716, U.S.A.<br />

email: vlasis@iceht.forth.gr<br />

ABSTRACT<br />

The Hamiltonian formulation <strong>of</strong> transport phenomena is employed to study the behavior <strong>of</strong> <strong>polymer</strong> solutions<br />

near a solid surface under both equilibrium (static) and nonequilibrium (flowing) conditions. The methodology<br />

is hierarchical and derives macroscopic flow equations which, through the use <strong>of</strong> selected internal variables,<br />

couple with the system microstructure. The bridging <strong>of</strong> the different length scales in the problem is achieved<br />

through the specification <strong>of</strong> the system Hamiltonian for which a microscopic model (at the level <strong>of</strong> Kuhn<br />

segments) is invoked. The macroscopic equations are derived through a two-fluid Hamiltonian model and<br />

consist <strong>of</strong>: (a) the concentration equation for the <strong>polymer</strong> component, (b) the momentum equation for the<br />

average fluid velocity, and (c) the constitutive equation for the stress-strain rate relationship which is a<br />

generalized Oldroyd-B model accounting for flow inhomogeneities. Under equilibrium conditions, the<br />

governing equations reduce to a minimization problem for the free energy <strong>of</strong> the system; this results into the<br />

well known equilibrium condition that the chemical potentials <strong>of</strong> all chain conformations in the interfacial area<br />

should be equal. However, the proposed methodology is more general and allows extending equilibrium<br />

considerations under flowing conditions, as well. In both cases, chain conformational characteristics near the<br />

solid boundary is taken into account through the solution <strong>of</strong> a diffusion equation for the chain propagator.<br />

Flow field effects are accommodated in the model by allowing the propagator to further depend on the<br />

Cauchy-Green strain tensor. Results are presented in two Parts for two cases: (a) the homo<strong>polymer</strong><br />

adsorption problem and (b) the flow <strong>of</strong> a <strong>polymer</strong> solution past a non-interacting solid surface. In the first case<br />

(Part I), our model reduces to the continuum analogue <strong>of</strong> the original Scheutjens-Fleer lattice model, with<br />

which a thorough comparison is presented. In the second case (Part II), our model calculates selfconsistently<br />

the velocity pr<strong>of</strong>ile in the interfacial area, and predicts the development <strong>of</strong> an apparent slip<br />

velocity near the wall <strong>of</strong> length scale commensurate with the size <strong>of</strong> the flowing <strong>polymer</strong> molecules.


HIERARCHICAL STRUCT<strong>UR</strong>E FORMATION AND PATTERN TRANSFER INDUCED BY AN<br />

ELECTRIC FIELD<br />

M.D. Morariu 1) , N.E. Voicu 1) , E. Schaeffer 2) , U. Steiner 1)<br />

1) Polymer Physics, Rijksuniversiteit Groningen,<br />

Nijenborgh 4, 9747 AG, Groningen, The Netherlands<br />

2) Max Planck Institute <strong>of</strong> Molecular Cell Biology and Genetics,<br />

Pfotenhauerstrasse 108, 01307 Dresden, Germany<br />

email: u.steiner@chem.rug.nl<br />

ABSTRACT<br />

Fabrication <strong>of</strong> smaller and smaller surface structures is a current trend in materials science. Many different<br />

techniques were developed during the past years such as photolithography, microcontact printing, imprint<br />

lithography and electrically induced structure formation. We present here a new concept <strong>of</strong> pattern formation<br />

that make use <strong>of</strong> a sequence <strong>of</strong> instabilities. The technique is based on the process <strong>of</strong> destabilizing the<br />

surface <strong>of</strong> <strong>polymer</strong> films. By imposing a lateral variation <strong>of</strong> the force field, we are able to direct the instability<br />

to replicate a master pattern with high precision and 100% reproducibility.<br />

Electrohydrodynamic (EHD) instabilities have received considerable attention [1] and surface instabilities<br />

were studied using a variety <strong>of</strong> liquids. The triggering factor was the experimental observations <strong>of</strong> the rupture<br />

<strong>of</strong> a liquid surface by an external applied electric field. [2,3] Two different cases can be distinguished: [4-6]<br />

(i) Accumulation <strong>of</strong> charges at the interface due to a difference in the conductivities between liquid-air or<br />

liquid-liquid interface; and (ii) induction <strong>of</strong> polarization charges due to the external electric field. In the case <strong>of</strong><br />

two <strong>polymer</strong>ic liquids, the electrostatic pressure is smaller at the interface due to a lower difference in the<br />

dielectric constants and the dynamics is slowed down because <strong>of</strong> the viscous coupling and the dissipation in<br />

the second liquid layer. While these effects weaken the instability, the decrease in the interfacial tension<br />

enhances it.<br />

The control <strong>of</strong> patterns on sub-micrometer lateral length scales is <strong>of</strong> considerable technological interest.<br />

Since in our experiment the wavelength <strong>of</strong> the instability is on the order <strong>of</strong> microns, we were able to control<br />

the resulting pattern by using topographically structured electrodes. Any conventional lithographic technique<br />

[7] is sufficient to produce a master pattern that can be multiply used. A replicated structure is shown in fig. 1.<br />

Figure 1. The replicated topography and the subsequent structure that surrounds the main structure.


Using this patterned electrode, we induce a lateral anisotropy in the electric field. The instability is focused<br />

towards regions <strong>of</strong> highest fields and the final structure is a replica <strong>of</strong> the master pattern. In our experiments,<br />

due to the use <strong>of</strong> a <strong>polymer</strong> bilayer, a hierarchical pattern is formed. The first instability replicated the topography<br />

<strong>of</strong> the upper electrode. A subsequent instability creates a <strong>polymer</strong> structure that surrounds the main<br />

structure. In conclusion we have developed here a new non–optical lithographic technique, that can produce<br />

structures on sub-100 nm lateral scales. The method is very simple and does not need special equipment or<br />

materials. By tuning the system parameters, such as the applied voltage (U), distance between capacitor<br />

plates (d) or <strong>polymer</strong> film thickness and imposing a lateral non-uniformity in the electric field by using a<br />

patterned master electrode, we can produce hierarchical structures with high aspect ratio.<br />

References<br />

1. Tonks, L., 1935. Phys.Rev. 48, 562.<br />

2. Melcher, J.R. and Smith Jr., C.V., 1969 Phys. Fluids 12, 778.<br />

3. Melcher, J.R., 1961. Phys. Fluids 4, 1348.<br />

4. Melcher, J.R., 1963. MIT Press, Cambridge, Mass.<br />

5. Reynolds, M., 1965. Phys. Fluids 8, 161.<br />

6. Swan, J.W. 1987. Proc. R. Soc. (London) 62, 38.<br />

7. Xia, Y., Rogers, J.A., Paul, K.E. and Whitesides, G.M., 1999. Chem.Rev. 99, 1823.


CAPILLARY-INDUCED FORCES BETWEEN PARTICLES IN POLYMER SOLUTIONS NEAR<br />

PHASE SEPARATION<br />

M. Olsson, F. Joabsson, P. Linse, L. Piculell<br />

Physical Chemistry 1, Center for Chemistry and Chemical Engineering,<br />

Lund University, P.O. Box 124, S-221 00 Lund<br />

email: Martin.Olsson@fkem1.lu.se<br />

ABSTRACT<br />

Capillary-induced phase separation (CIPS) leads to a long-ranged attractive force between two surfaces<br />

immersed in a bulk solution. The formation <strong>of</strong> a new phase in the gap between the surfaces is driven by a<br />

lower surface energy for the new (capillary) phase compared to the bulk phase. The decrease in free energy<br />

is larger at shorter separations between the particles than for longer ones. An important condition for CIPS to<br />

arise is that the particle-free solution has to be close to phase separation.<br />

CIPS forces are known to affect particle aggregation in mixed solvents and in <strong>polymer</strong> blends as well as<br />

protein aggregation in biological membranes. We have instead looked at <strong>polymer</strong> solutions with added<br />

colloidal dispersed particles both experimentally and theoretically. Previous surface force measurements<br />

have showed that CIPS occurs in mixed <strong>polymer</strong> solutions as well as in quasi-binary <strong>polymer</strong> solutions [1,2].<br />

A complementary experimental study [3] for the mixed <strong>polymer</strong> system displays phase separation at<br />

compositions far into the one-phase area for the original phase diagram, when colloidal particles are added.<br />

To investigate more generally the parameters affecting CIPS for the mixed <strong>polymer</strong> system, a theoretical<br />

mean-field lattice study has been done [4]. This study shows that CIPS occurs close to phase separation for<br />

the system and is affected by parameters like the molecular weight <strong>of</strong> the <strong>polymer</strong>, the affinity <strong>of</strong> the <strong>polymer</strong><br />

to the surface, the composition <strong>of</strong> the sample and the distance between the particles. A recent study [5]<br />

shows that CIPS can also appear in quasi-binary solutions. This has also been shown experimentally.<br />

References<br />

1. Freyssingeas, E., Thuresson, K., Nylander, T., Joabsson, F., Lindman, B., 1998. Langmuir 14, 5877-5889.<br />

2. Wennerström, H., Thuresson, K., Linse, P., Freyssingeas, E., 1998. Langmuir 14, 5664-5666.<br />

3. Joabsson, F., Calatayu, A., Thuresson, K., Piculell, L., submitted. Journal <strong>of</strong> Physical Chemistry B.<br />

4. Joabsson, F. and Linse, P. “Capillary-Induced Phase Separation in Mixed Polymer Solutions. A Lattice<br />

Mean-Field Calculation Study”, manuscript.<br />

5. Olsson, M., Linse, P., Piculell, L. “Capillary-Induced Phase Separation in Binary and Quasi-binary Solutions.<br />

A Lattice Mean-Field Calculation Study”, manuscript.


span the gap between the electrodes. Throughout the course <strong>of</strong> growth, the characteristic wavelength<br />

remains invariant. The characteristic length scale coupled with the sharpening <strong>of</strong> the peaks <strong>of</strong> the fluctuations<br />

gives rise to a hexagonal packing <strong>of</strong> the cylinders laterally.<br />

By replacing the air gap with a second fluid or <strong>polymer</strong> layer between the electrodes reduces the energetic<br />

cost <strong>of</strong> generating interfacial area by replacing the surface energy with an interfacial energy. In general, the<br />

interfacial energy is approximately one order <strong>of</strong> magnitude smaller than the surface energy. This translates<br />

into a reduction in the characteristic fluctuation wavelength. The experimentally measured wavelengths are in<br />

excellent agreement with the predictions. Theoretically, significant changes are unexpected, but a significant<br />

decrease is observed. While unexpected, this represents a tremendous advantage for end-use <strong>of</strong> this<br />

phenomenon.


POLYMER-MEDIATED FORCES: OPTICAL TENSIOMETRY AND OPTICAL FORCE<br />

MICROSCOPY<br />

R. Rajagopalan<br />

Department <strong>of</strong> Chemical Engineering, University <strong>of</strong> Florida,<br />

Gainesville, Florida 32611-6005, USA<br />

email: Raj@ChE.UFL.edu<br />

ABSTRACT<br />

Introduction<br />

Intensive research over the last two decades has established that <strong>polymer</strong> adsorption on surfaces is far<br />

more subtle than previously recognized (Scheutjens 1985; Fleer 1993). Whether interaction between <strong>polymer</strong>-laden<br />

surfaces is attractive or repulsive depends on the amount <strong>of</strong> <strong>polymer</strong> between the two surfaces<br />

and on the precise conformational details <strong>of</strong> the adsorbed chains, a fact that is all the more important in the<br />

case <strong>of</strong> physisorbed layers, especially under “starved” (undersaturated) conditions. The latter situation is <strong>of</strong><br />

particular interest in a number <strong>of</strong> practical contexts, as physisorbed <strong>polymer</strong> layers under less-than-saturation<br />

conditions are <strong>of</strong>ten the rule in colloid stabilization or in <strong>polymer</strong> interfaces <strong>of</strong> biomedical and biological<br />

interest. The structure <strong>of</strong> adsorbed <strong>polymer</strong> chains (especially the role <strong>of</strong> tails) has been debated since the<br />

early work <strong>of</strong> Scheutjens and Fleer and has been clarified to some extent only recently by Semenov and<br />

coworkers (e.g., Semenov 1996), who revisited this problem in the last few years using an extension <strong>of</strong> the<br />

ground state dominance approximation within a mean-field framework, and by Fleer and coworkers<br />

(1999a,b), who developed an analytical approximation for the numerical mean-field theory and related it to<br />

the approach <strong>of</strong> Semenov and coworkers. Simultaneously, a parallel effort to study interaction forces<br />

between <strong>polymer</strong> layers using simulations has also been initiated (Jimenez 1998, 2000). Although these<br />

simulations are based on self-avoiding random walk chains and therefore cannot be compared directly (or,<br />

used to “test”) mean-field theories, such simulations do provide a way to assess how mean-field approximations<br />

perform relative to realistic chains. Despite the approximations or simplifications involved, the<br />

Scheutjens-Fleer (SF) lattice mean-field theory provides an important framework for studying <strong>polymer</strong>/<br />

interface systems that are not amenable to scaling theories and to the Edwards-equation-based analytical<br />

mean-filed formulations. In particular, the SF theory allows one to examine low-molecular-weight <strong>polymer</strong>s<br />

and chains <strong>of</strong> complicated molecular architecture.<br />

The extensive work available in the literature based on mean-filed theories and simulations have also set<br />

the stage for systematic experimental measurements <strong>of</strong> <strong>polymer</strong>-mediated forces down to the level <strong>of</strong> single<br />

chains. The extension to single chains implies that one should be able to measure forces <strong>of</strong> the order <strong>of</strong> tens<br />

<strong>of</strong> femtoNewtons – a level <strong>of</strong> sensitivity not readily accessible through atomic force microscopy (AFM).<br />

However, the required level <strong>of</strong> sensitivity and resolution (in force) can be achieved if one uses optical<br />

tweezers to hold and manipulated <strong>polymer</strong> chains and as force transducers. In this talk, we describe an<br />

optical force microscope (analogous to an AFM), in which a probe particle held by a diffraction-limited optical<br />

trap is used in place <strong>of</strong> the cantilever <strong>of</strong> an atomic force microscope. Examples <strong>of</strong> static and dynamic<br />

measurements and related possibilities will be described.<br />

Optical Force Microscopy<br />

As well-known, members <strong>of</strong> the family <strong>of</strong> scanning probe microscopes differ from each other with respect<br />

to the type <strong>of</strong> probe used and the type <strong>of</strong> substrate/probe interaction that is used as the basis <strong>of</strong> the measurement.<br />

The sensitivity <strong>of</strong> the force measurement and the resolution <strong>of</strong> the distances depend on the<br />

characteristics <strong>of</strong> the probe and the probe/substrate interactions, among others. The stiffness <strong>of</strong> a typical<br />

AFM cantilever is typically <strong>of</strong> the order <strong>of</strong> 1000 pN/nm, but cantilevers with stiffnesses in the neighborhood <strong>of</strong><br />

1 pN/nm can be fabricated. Nevertheless, when dealing with s<strong>of</strong>t interfaces such as a <strong>polymer</strong> layer, it is


desirable to use a probe with a much lower spring constant. The use <strong>of</strong> a spherical particle held in place by<br />

an optical trap was first suggested as a probe about a decade back by Ghislain and Webb (Ghislain 1993) to<br />

meet the need for probes <strong>of</strong> very low stiffness for measuring single-molecule force measurements at the pico-<br />

Newton level in biology and biophysics. The focused laser beam in this case acts as an optical “tweezer” to<br />

hold and spatially manipulate the probe. Such an optical trap, known as a tweezer trap, behaves like a linear<br />

spring, and the particle and the trap replace the AFM cantilever in the measurement. The above arrangement<br />

is also more compatible with optical microscopy, and a combination <strong>of</strong> the two allows one to perform simultaneous<br />

videomicroscopy on the specimen under examination. A scanning probe microscope based on the<br />

above principle is known as an optical force microscope (OFM). More recent developments and applications<br />

<strong>of</strong> optical manipulation <strong>of</strong> colloids and <strong>polymer</strong>s are available in a forthcoming book (Rajagopalan 2003).<br />

The tweezer trap functions as a harmonic spring in the axial (z) as well as in the lateral (x and y)<br />

directions. For example, in the z-direction (normal to the interacting surface), which is the only one we shall<br />

consider here, one may write the trap potential φtr as φtr = (ktr/2)(z−ztr) 2 , where z is the distance along the axis<br />

<strong>of</strong> the beam, ztr is the location <strong>of</strong> the trap center and ktr is the “stiffness” <strong>of</strong> the trap potential. Similar<br />

approximations can be written for the lateral directions. As in the case <strong>of</strong> an AFM cantilever, the above<br />

spring-like function <strong>of</strong> the trap allows one to use the trap as a force transducer to determine any external<br />

force exerted on the probe particle, as has been demonstrated by a number <strong>of</strong> investigators (see, for<br />

example, Dogariu 2000 and Rajagopalan 2003 and references therein). Moreover, as the trap stiffness can<br />

be adjusted by changing the power <strong>of</strong> the trap laser to magnitudes <strong>of</strong> the order <strong>of</strong> 0.001 pN/nm (about four<br />

orders <strong>of</strong> magnitude smaller than the typical stiffness <strong>of</strong> the cantilevers <strong>of</strong> AFMs), the use <strong>of</strong> tweezer traps<br />

allows one to measure very small forces directly and to image “s<strong>of</strong>t” <strong>polymer</strong> interfaces.<br />

Measurement <strong>of</strong> Polymer-Mediated Forces<br />

Lack <strong>of</strong> space does not permit a detailed discussion <strong>of</strong> the method and measurements here and details<br />

may be found elsewhere (Dogariu 2000; Rajagopalan 2003). In this talk, we shall present illustration <strong>of</strong> the<br />

use <strong>of</strong> an OFM for measuring electrical double-layer forces, force between a <strong>planar</strong> surface and a colloidal<br />

particle with physisorbed <strong>polymer</strong> chains (poly(ethylene oxide) and polyacrylic acid) on both surfaces, and<br />

single-chain elasticity. Possibilities for examining the dynamic response <strong>of</strong> the <strong>polymer</strong> layer to an oscillating<br />

probe particles and “microrheology” <strong>of</strong> <strong>polymer</strong> layers will also be illustrated.<br />

References<br />

1. Dogariu, A.C. and Rajagopalan, R., 2000. Optical Tweezers as Force Transducers: The Effects <strong>of</strong><br />

Focusing the Trapping Beam through a Dielectric Interface. Langmuir 16, 2770-2778.<br />

2. Fleer, G., Cohen Stuart, M., Scheutjens, J., Cosgrove, T., and Vincent, B., Polymers at Interfaces, London:<br />

Chapman and Hall, 1993.<br />

3. Fleer, G.J., van Male, J. and Johner, A., 1999a. Analytical Approximation to the Scheutjens-Fleer Theory<br />

for Polymer Adsorption from Dilute solution. 1. Trains, Loops, and Tails in terms <strong>of</strong> Two Parameters: The<br />

Proximal and Distal Lengths. Macromolecules, 32, 825-844.<br />

4. Fleer, G.J., van Male, J. and Johner, A., 1999b. Analytical Approximation to the Scheutjens-Fleer Theory<br />

for Polymer Adsorption from Dilute solution. 2. Adsorbed Amounts and Structure <strong>of</strong> the Adsorbed Layer,<br />

Macromolecules, 32, 845-862.<br />

5. Ghislain, L.P. and Webb, W.W., 1993. Scanning Force Microscope Based on an Optical Trap, Opt. Lett.,<br />

18, 1678-1680.<br />

6. Jimenez, J. and Rajagopalan, R., 1998. A New Simulation Method for the Determination <strong>of</strong> Forces in<br />

Polymer/Colloid Systems. Euro. Phys. J. B, 5, 237-243.<br />

7. Jimenez, J., deJoannis, J., Bitsanis, I. and Rajagopalan, R., 2000. Interaction between Undersaturated<br />

Polymer Layers: Computer Simulations and Numerical Mean-Field Calculations. Macromolecules, 33,<br />

8512-8519.<br />

8. Rajagopalan, R. and Dogariu, A.C., Eds., Optical Manipulation <strong>of</strong> Particles and Macromolecules, Cambridge:<br />

Cambridge Univ. Pr., 2003.<br />

9. Scheutjens, J. and Fleer, G.J., 1985, Interaction between Two Adsorbed Polymer Layers. Macromolecules<br />

18, 1882-1900.<br />

10.Semenov, A.N., Bonet-Avalos, J., Johner, A., and Joanny, J.-F., 1996. Adsorption <strong>of</strong> Polymer Solutions<br />

<strong>onto</strong> a Flat Surface. Macromolecules 29, 2179-2196.


DENT720CONTROLLING POLYMERS AT INTERFACES ON THE NANOSCOPIC AND<br />

MACROSCOPIC LEVELS<br />

T.P. Russell<br />

Polymer Science and Engineering Department, University <strong>of</strong> Massachusetts,<br />

Amherst, MA 01003 USA<br />

email: russell@mail.pse.umass.edu<br />

ABSTRACT<br />

Developing routes to manipulate the behavior <strong>of</strong> <strong>polymer</strong>s at interfaces is key in designing and developing<br />

devices base on thin <strong>polymer</strong> films. Block co<strong>polymer</strong>s <strong>of</strong>fer a wealth <strong>of</strong> self-assembling, ordered, nanoscopic<br />

morphologies that have tremendous potential as lithographic templates, high-density porous media or as<br />

scaffolds for nanoscopic devices. Realizing the full potential <strong>of</strong> these materials in bulk or thin film applications,<br />

however, mandates control over the spatial orientation <strong>of</strong> the structures. External fields can be used to this<br />

end. These may be passive, as with interfacial interactions, or active, as with mechanical shear. In liquid<br />

crystal displays, for example, electric fields are used to control molecular alignment and, hence, the optical<br />

properties <strong>of</strong> the device. Much less attention has been devoted to amorphous systems where the<br />

anisotropies on the molecular level are small. Block co<strong>polymer</strong>s are an exception where mechanical shear<br />

and electric fields have been used to produce well-ordered arrays <strong>of</strong> nanoscopic domains. However, in thin<br />

co<strong>polymer</strong> films, the orientation <strong>of</strong> the structures normal to the surface introduces an interesting twist. In<br />

particular, there is a competition between the applied electric field and interfacial interactions that gives rise to<br />

a threshold field strength that must be exceeded to fully align the morphology. This threshold field strength<br />

has revealed a novel means <strong>of</strong> determining differences in the interfacial energies <strong>of</strong> the components<br />

comprising the co<strong>polymer</strong>.<br />

Producing well-ordered thin films can take hours to achieve or there are restrictions on the film thickness.<br />

Both present barriers to their end-use. Here, a very rapid route is described by which the cylindrical<br />

microdomains <strong>of</strong> an asymmetric diblock co<strong>polymer</strong> <strong>of</strong> polystyrene, PS and polyethyleneoxide, PEO, denoted<br />

P(S-b-EO) can be oriented normal to the surface <strong>of</strong> a film over very large areas. Results are shown where,<br />

within seconds, arrays <strong>of</strong> nanoscopic cylindrical domains <strong>of</strong> PEO are produced in a glassy PS matrix in films<br />

with thickness several times the period <strong>of</strong> the co<strong>polymer</strong>. This was found over a wide range in molecular<br />

weight, allowing control over the size <strong>of</strong> the cylindrical domains and the areal density <strong>of</strong> the arrays. In<br />

addition, since PEO is water soluble, the oriented arrays <strong>of</strong> nanoscopic cylinders opening a novel route<br />

towards water permeable membranes with well-defined channel sizes.<br />

An electric field applied normal to an interface, due to the gradient in the dielectric constant across the<br />

interface, produces a field gradient that translates into an electrostatic pressure that amplifies thermal<br />

fluctuations. Surface and interfacial energies, on the other hand, act to minimize the interfacial area and,<br />

therefore, tend to suppress the growth <strong>of</strong> the fluctuations. A linear stability analysis taking these opposing<br />

forces into account yields a preferential wavelength <strong>of</strong> fluctuations that is amplified with a characteristic<br />

relaxation time. Optical microscopy studies on homo<strong>polymer</strong> films between two electrodes with and air gap<br />

between the <strong>polymer</strong> surface and the upper electrode show the growth <strong>of</strong> theses fluctuations which,<br />

ultimately lead to the formation <strong>of</strong> solid cylinders <strong>of</strong> the <strong>polymer</strong> spanning the gap between the two<br />

electrodes. The average separation distance between adjacent cylinders and the characteristic growth rate <strong>of</strong><br />

the cylinders agree surprisingly well with linear theoretical arguments. Studies from the early stages <strong>of</strong> the<br />

growth <strong>of</strong> the fluctuations show features that a reminiscent <strong>of</strong> spinodal phase separation in <strong>polymer</strong> mixtures.<br />

An interconnected fluctuation pattern is observed on the surface with a characteristic length scale. With time<br />

the fluctuations sharpen significantly, grow in height until, finally, they form cylinders <strong>of</strong> solid <strong>polymer</strong> that


span the gap between the electrodes. Throughout the course <strong>of</strong> growth, the characteristic wavelength<br />

remains invariant. The characteristic length scale coupled with the sharpening <strong>of</strong> the peaks <strong>of</strong> the fluctuations<br />

gives rise to a hexagonal packing <strong>of</strong> the cylinders laterally.<br />

By replacing the air gap with a second fluid or <strong>polymer</strong> layer between the electrodes reduces the energetic<br />

cost <strong>of</strong> generating interfacial area by replacing the surface energy with an interfacial energy. In general, the<br />

interfacial energy is approximately one order <strong>of</strong> magnitude smaller than the surface energy. This translates<br />

into a reduction in the characteristic fluctuation wavelength. The experimentally measured wavelengths are in<br />

excellent agreement with the predictions. Theoretically, significant changes are unexpected, but a significant<br />

decrease is observed. While unexpected, this represents a tremendous advantage for end-use <strong>of</strong> this<br />

phenomenon.


ENTROPIC NETWORKS IN POLYMERS AND MICROEMULSIONS<br />

A. Zilman, S. A. Safran<br />

Weizmann Institute <strong>of</strong> Science,<br />

Rehovot, Israel 76100<br />

email: sam.safran@weizmann.ac.il<br />

ABSTRACT<br />

Self-assembling networks and branched structures are common in both natural and synthetic materials and<br />

form under a variety <strong>of</strong> equilibrium and non-equilibrium conditions. Here, we review the theory <strong>of</strong> the structure<br />

and thermodynamic properties <strong>of</strong> equilibrium networks <strong>of</strong> cylindrical micelles [1], microemulsions [2] and selfassembling<br />

<strong>polymer</strong>ic gels [3]. In all these systems, the networks consist <strong>of</strong> self-assembling, locally, onedimensional<br />

objects (e.g., <strong>polymer</strong> chains in the case <strong>of</strong> a gel, cylindrical micelles or microemulsions or<br />

chains <strong>of</strong> dipolar colloids [4]). The study <strong>of</strong> network phases is <strong>of</strong> both theoretical [5] and practical interest.<br />

From the practical point <strong>of</strong> view, network forming amphiphiles and sol-gel systems are at the core <strong>of</strong> many<br />

industrial, biological and bio-medical applications. The existence <strong>of</strong> junctions and their dynamics is important<br />

in understanding the rheology <strong>of</strong> worm-like micelles. Network formation may also be responsible for the<br />

unusual "closed loop" phase diagrams <strong>of</strong> microemulsions and dipolar liquids. Cryo-electron microscopy [1]<br />

shows clear evidence <strong>of</strong> coexisting network phases in dilute microemulsions. In many <strong>of</strong> the systems<br />

mentioned above, the chains themselves are self-assembled in an equilibrium manner from a large number<br />

<strong>of</strong> monomers.<br />

We begin with a simple estimate <strong>of</strong> the density and free energy contribution <strong>of</strong> ends and junctions in a<br />

system <strong>of</strong> self-assembling chains or tubes and show that the junction entropy introduces an effective<br />

attraction (negative second-virial coefficient) that can lead to phase separation. The relationship <strong>of</strong> this<br />

thermodynamic transition to the topological, connectivity transition (percolation or gelation) can be predicted<br />

in a unified manner using a modification <strong>of</strong> the n = 0 spin-model for <strong>polymer</strong>s .The extension <strong>of</strong> these ideas<br />

to systems with crosslinks (e.g. actin gels) is also discussed and the theoretical phase diagrams are<br />

compared with recent experiments [6]. Finally, we show how entropic phase separation can also explain<br />

recent observations [7] in a combined system <strong>of</strong> telechelic <strong>polymer</strong>s and microemulsions that form network<br />

structures.<br />

References<br />

1. Berheim-Groswasser, A., Tlusty, T., Safran, S.A., Talmon, Y., 1999. Langmuir 15, 5448; Bernheim-Grosswasser,<br />

A., Wachtel, E., Talmon, Y., 2000. Langmuir 16, 4131.<br />

2. Tlusty, T, Safran, S.A., Strey, R., 2000. Phys. Rev. Lett. 84, 1244.<br />

3. Zilman, A. and Safran, S.A., submitted.<br />

4. Tlusty, T, Safran, S.A., 2000. Science 290, 1328.<br />

5. Drye, T., Cates, M.E., 1992. J. Chem. Phys. 96, 1367; Zilman, A., Safran, S.A., Phys. Rev. E, in press.<br />

6. Tempel, M., Isenberg, G., Sackmann, E., 1996. Phys. Rev. E 54, 1802.<br />

7. Filali, M., Ouazzani, M.J., Michel, E., Aznar, R., Porte, G. Appell, J., 2001. J. Phys. Chem B, 105, 10528.


MEMBRANE FUSION: RESULTS FROM SIMULATION AND SELF-CONSISTENT FIELD<br />

THEORY<br />

Michael Schick<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Washington,<br />

Box 351560, Seattle, WA 98195-1560, U.S.A.<br />

email: schick@mahler.phys.washington.edu<br />

ABSTRACT<br />

Our group has been studying self assembly in systems <strong>of</strong> block co<strong>polymer</strong> and <strong>of</strong> biological lipids, as there is<br />

much similarity in the phase behavior <strong>of</strong> the two. Our studies have focussed on defects in these systems. In<br />

the first part <strong>of</strong> my talk, I will discuss various grain boundaries in lamellar phases <strong>of</strong> block co<strong>polymer</strong>s, work<br />

which was accomplished utilizing self-consistent field theory only, and in the Fourier representation. In the<br />

second part, I will discuss our work on the mechanism <strong>of</strong> membrane fusion, which was carried out utilizing<br />

both Monte Carlo simulations and self-consistent field theory in real-space representation.<br />

Grain Boundaries<br />

Grain boundaries in smectics, such as the lamellar phases <strong>of</strong> block co<strong>polymer</strong>s, are easier to study than<br />

those in crystalline solids due to the lack <strong>of</strong> in-plane order in the smectics. Matsen first showed how scft in the<br />

Fourier representation could be applied to symmetric-tilt grain boundaries. A notable feature was the<br />

occurrence <strong>of</strong> a symmetry-breaking transition as a function <strong>of</strong> tilt angle. We have extended these studies to<br />

include twist grain boundaries and, most recently, T-junctions. Symmetric tilt grain boundaries and Tjunctions<br />

both can connect lamellar grains <strong>of</strong> different orientations. Presumably the frequency with which one<br />

observes one or the other depends upon their relative free energy. Our motivation was a series <strong>of</strong><br />

experiments which showed that T-junctions did not occur very <strong>of</strong>ten in a system consisting <strong>of</strong> block<br />

co<strong>polymer</strong> only. However the addition <strong>of</strong> homo<strong>polymer</strong> significantly increased their occurrence. Our studies<br />

showed that in the system <strong>of</strong> co<strong>polymer</strong> only, there was only a small range <strong>of</strong> angles over which the free<br />

energy <strong>of</strong> the T-junction was less than that <strong>of</strong> the symmetric tilt grain boundary. However this range <strong>of</strong> angles<br />

increases appreciably with the addition <strong>of</strong> homo<strong>polymer</strong>. Of additional interest is the appearance <strong>of</strong> a<br />

morphological change as the angle between grains approaches 180 degrees. This change is rather similar to<br />

that which occurs in the symmetric tilt boundaries.Membrane FusionAlthough the fusion <strong>of</strong> membranes is an<br />

enormously important biological process, it is poorly understood. One knows that specialized proteins are<br />

needed to bring two fluctuating membranes close to one another, a process which puts the membranes<br />

under tension. Beyond this, little is known. Almost all theories begin with an initial state consisting <strong>of</strong> a<br />

cylindrically symmetric stalk, a localized region in which the lipids <strong>of</strong> each <strong>of</strong> the proximate lipid layers<br />

rearrange to form a continuous hydrophobic region joing the hydrophobic cores <strong>of</strong> the two bilayers. The final<br />

fusion pore, whose hydrophilic region traverses both bilayers, is also cylindrically symmetric. It has been<br />

assumed by all theories that all intermediate states in the fusion process were equally symmetric. We<br />

investigated this process by Monte Carlo simulation and found the process took a completely different<br />

symmetry-broken route. Knowing the form <strong>of</strong> the intermediate, we have used self consistent field theory to<br />

calculate the free energy barriers along previously assumed paths and the one we found. Our calculations<br />

help us understand why the path we did see is the one we should have seen.


ADSORBED POLYMER LAYERS: LOOPS, TAILS AND STIFFNESS.<br />

A.N. Semenov<br />

Physics Department, Moscow State University,<br />

Moscow 119992, Russia<br />

email: semenov@polly.phys.msu.su<br />

ABSTRACT<br />

Polymer adsorption is a remarkable phenomenon that is widely used to modify properties <strong>of</strong> surfaces and<br />

interfaces. I will review theoretical results on the structure <strong>of</strong> adsorbed <strong>polymer</strong> layers focusing on the role <strong>of</strong><br />

loops and tails on the one hand, and the chain stiffness on the other hand.<br />

1. Adsorbed layers formed by flexible <strong>polymer</strong>s<br />

Self-similar structure <strong>of</strong> saturated <strong>polymer</strong> layers adsorbed from a good solvent <strong>onto</strong> an attracting surface<br />

was elucidated by De Gennes long ago [1]: in the central region (at the distances z from the surface longer<br />

than the molecular extrapolation length b but shorter than the coil size R) the monomer concentration decay<br />

as φ = z −a<br />

, where a = 4/3( a = 2 in the marginal solvent regime). The layer can be viewed as a system <strong>of</strong><br />

loops <strong>of</strong> different lengths (chains sections starting and terminating at the adsorbing surface).<br />

More recently it was shown [2,3,4] that loops do not dominate everywhere in the central region: the<br />

contribution <strong>of</strong> tails (chain sections starting at the surface and terminating in the bulk) is important at<br />

distances z > z * , where z * is a length-scale that is much longer than the monomer size, but much shorter<br />

than the coil size R ( u z* N , where u = 1/ 2 in good solvent and u = 1/ 3 in marginal solvent). Hence two<br />

subregions: the inner sublayer z < z * formed by loops, and the outer sublayer z > z * dominated by tails (the<br />

tail subregion resembles a layer <strong>of</strong> broadly polydisperse end-<strong>grafted</strong> chains). Hence a significant portion <strong>of</strong><br />

the adsorbed layer contains mostly tail monomers. It is remarkable that this conclusion was first obtained in<br />

the numerical studies by J. Scheutjens, G. Fleer and their co-workers [5,6] and then stayed unexplained for a<br />

long while.<br />

The tail sublayer is considerably thicker than the loop layer (at a given distance z from the surface). This<br />

results in a surprising behavior: monomer concentration at the given point z ( z >> b) in a saturated layer <strong>of</strong><br />

shorter chains (N1) may be higher (by a large factor ~ 10) than concentration in a layer <strong>of</strong> longer chains (N2)<br />

provided that z * (N1) < z < z * (N2). The effect <strong>of</strong> <strong>polymer</strong> tails is responsible for a number <strong>of</strong> important<br />

effects. The `thicker' tail sublayer is relatively less permeable: it is rather difficult for an unadsorbed chain to<br />

penetrate through this layer, the main mechanism <strong>of</strong> penetration being <strong>of</strong>ten end-entry (rather than hairpin<br />

entry) [7]. Two saturated adsorbed layers repel each other when their tail sublayers are overlapping (at the<br />

distances z* < h< R). This is in contrast with the well-known normal behavior: attraction due to bridging by<br />

<strong>polymer</strong> fragments which is predicted at shorter distances (in the regime <strong>of</strong> loops) [8].<br />

2. Adsorption <strong>of</strong> a stiff worm-like macromolecule<br />

Polymers with stiff backbone (including many biological <strong>polymer</strong>s) show a rather different adsorption<br />

behavior as compared to that characteristic <strong>of</strong> flexible chains [9]. I present a theoretical description <strong>of</strong><br />

adsorption <strong>of</strong> a single semiflexible chain <strong>onto</strong> a flat surface. Both scaling and quantitative approaches are<br />

discussed. A self-similar monomer concentration pr<strong>of</strong>ile cx ( )~ z −4/3 is predicted near the surface (when the<br />

distance to the surface z is much smaller than the chain persistence length / 2 ). The typical conformation <strong>of</strong><br />

a weakly adsorbed chain can be viewed as a sequence <strong>of</strong> alternating 2-dimensional (flat) and 3-dimensional<br />

blobs forming a double-layer structure: a thin contact layer (<strong>of</strong> molecular thickness d, d


the transition temperature T * when τ = 1 −T<br />

*/ T is smaller than ( d / ) −4/3;<br />

otherwise (at larger τ ) the lion's<br />

share <strong>of</strong> the chain is staying in the d-layer very close to the surface. The adsorption transition is continuous<br />

although it looks like a discontinuous (first-order) transition outside a narrow region around the adsorption<br />

temperature, τ > ( d / ) −4/3.<br />

References<br />

1 De Gennes, P.G., 1981. Macromolecules 14, 1637.<br />

2 Semenov, A.N., Joanny, J.F., 1995. Europhys.Lett. 29, 279.<br />

3 Semenov, A.N., Joanny, J.F., Johner, A. Polymer Adsorption: mean-field theory and ground state dominance<br />

approximation. In: Theoretical and Mathematical Models in Polymer Science, Chapter 2, pp.37-81<br />

(Ed. A. Grosberg, Academic Press, Boston, 1998).<br />

4 Johner, A., Bonet-Avalos, J., Van der Linden, C.C., Semenov, A.N., Joanny, J.F., 1996. Macromolecules<br />

29, 3629.<br />

5 Scheutjens, J., Fleer, G., 1980. J. Phys.Chem. 84, 178.<br />

6 Scheutjens, J., Fleer, G. In: The Effect <strong>of</strong> Polymers on Dispersion Properties, Th.F. Tadros Ed., Academic<br />

Press, New York, 1982.<br />

7 Semenov, A.N., Joanny, J.-F., 1995. J. de Physique 5, 859.<br />

8 Semenov, A.N., Joanny, J.F., Johner, A., Bonet-Avalos, J., 1997. Macromolecules 30, 1479.<br />

9 Van der Linden, C.C., Leermakers, F.A.M., Fleer, G.J., 1996. Macromolecules 29, 1172.


MEAN-FIELD THEORY PREDICTION OF PHASE BEHAVIOR OF ABC TRIBLOCK<br />

COPOLYMER IN SELECTIVE SOLVENT<br />

N.P. Shusharina 1) , P. Alexandridis 1) , S. Balijepalli 2) , H.J.M. Gruenbauer 3)<br />

1) Department <strong>of</strong> Chemical Engineering, University at Buffalo,<br />

The State University <strong>of</strong> New York, Buffalo, New York 14260-4200, U.S.A.<br />

2) New Product & Mathematical Modeling Group, Corporate R&D,<br />

The Dow Chemical Co., Midland, MI 48674, U.S.A.<br />

3) New Business Development, Dow Benelux N.V.,<br />

4530 AA Terneuzen, The Netherlands<br />

email: ns33@eng.buffalo.edu<br />

ABSTRACT<br />

A mean-field lattice theory is applied to predict the self-assembly into ordered structures <strong>of</strong> ABC triblock<br />

co<strong>polymer</strong>s in selective solvent. More specifically, the composition-temperature phase diagram has been<br />

constructed for the system (C)14(PO)12(EO)17/water, where C stands for methylene, PO for propylene oxide<br />

and EO for ethylene oxide. The model predicts thermotropic phase transitions between the ordered<br />

hexagonal, lamellar, reverse hexagonal, and reverse cubic phases, as well as the disordered phase. The<br />

thermotropic behavior is a result <strong>of</strong> the temperature dependence <strong>of</strong> water interaction with EO- and POsegments.<br />

The lyotropic effect (caused by changing the solvent concentration) on the formation <strong>of</strong> different<br />

structures has been found weak. The structure in the ordered phases is described by analyzing the species<br />

volume fraction pr<strong>of</strong>iles and the end segment and junction distributions. A "triple-layer" structure has been<br />

found for each <strong>of</strong> the ordered phases, with each layer rich in C-, PO-, and EO-segments, respectively. The<br />

electrolyte effects on the macroscopic phase behavior are also considered. The addition <strong>of</strong> NaCl is found to<br />

sharpen the thermotropic phase transition and shift them to lower temperatures compared to the salt-free<br />

case. The parameters describing interactions between different species are obtained from simple experiments,<br />

and thus this model can be useful in capturing the effects <strong>of</strong> various electrolytes on non-ionic<br />

surfactants phase behavior.


PROTEIN ADSORPTION ON S<strong>UR</strong>FACES WITH GRAFTED POLYMERS<br />

Igal Szleifer<br />

Department <strong>of</strong> Chemistry, Purdue University,<br />

West Lafayette, IN 47907, USA<br />

email: igal @ purdue.edu<br />

ABSTRACT<br />

Introduction<br />

Protein adsorption plays a key role in a variety <strong>of</strong> biological processes and it is very important in many<br />

biotechnological processes. For example, when a foreign body is put in contact with the blood stream, large<br />

blood proteins e.g. fibrinogen, adsorb to the surface <strong>of</strong> the material. This protein adsorption is necessary for<br />

the following adhesion <strong>of</strong> platelet and the possibility <strong>of</strong> surface induced thrombosis. It has been shown that<br />

when the surface adsorption <strong>of</strong> fibrinogen is below a certain threshold, platelet adhesion does not occur.<br />

Therefore, one <strong>of</strong> the important steps in increasing biocompatibility in a material is by preventing the<br />

adsorption <strong>of</strong> blood proteins into the surface <strong>of</strong> the material. Another important example in which protein<br />

adsorption plays a key role is in the design <strong>of</strong> biosensors. In some <strong>of</strong> these cases it is necessary to obtain an<br />

ordered array <strong>of</strong> enzymes. Further, the proteins should adsorb in their biological active conformation for<br />

proper functioning <strong>of</strong> the device. These two examples show the importance <strong>of</strong> protein adsorption. However, in<br />

one case it is necessary to induce protein adsorption while in the other it is important to prevent the<br />

adsorption <strong>of</strong> the macromolecules.<br />

Protein adsorption is a very complex process since it involves very large energy scales. Typical van der<br />

Waals interactions between surfaces and proteins are in the order <strong>of</strong> tens to hundred times the thermal<br />

energy. Further, proteins are highly heterogeneous molecules that have the ability to change their conformation<br />

upon adsorption on the surface. This change <strong>of</strong> conformation is induced by the presence <strong>of</strong> an<br />

inhomogeneous interaction field due to surface. Another complexity arises from the fact that in general<br />

proteins have charged amino acids and therefore electrostatic interactions play a key role on the adsorption<br />

behavior.<br />

In this talk we will discuss the ability <strong>of</strong> <strong>grafted</strong> <strong>polymer</strong> layers to control protein adsorption. This is aimed<br />

to use <strong>grafted</strong> <strong>polymer</strong>s as surface modifiers that can be tuned to prevent protein adsorption or in other cases<br />

to control the kind <strong>of</strong> protein conformation that adsorbs on the surface. We will show our theoretical studies <strong>of</strong><br />

how different types <strong>of</strong> <strong>grafted</strong> <strong>polymer</strong>s can be used to tune the adsorption process. Our theoretical approach<br />

is based on a generalization <strong>of</strong> a molecular mean-field theory that has been very successful to describe the<br />

properties <strong>of</strong> tethered <strong>polymer</strong>s. For example, the theory is capable to quantitatively describe the pressurearea<br />

isotherms <strong>of</strong> polystyrene-poly ethylene oxide (PS-PEO) diblock co<strong>polymer</strong>s spread at the water-air<br />

interface. We have also shown that the predictions <strong>of</strong> the theory are in excellent quantitative agreement with<br />

experimental observations <strong>of</strong> the structure <strong>of</strong> PEO molecules <strong>grafted</strong> on surfaces with rigid collagen<br />

molecules. Furthermore, we will also show that the predictions <strong>of</strong> the theory are in very good agreement with<br />

experimental observations for the adsorption isotherms <strong>of</strong> lysozyme and fibrinogen on surfaces with <strong>grafted</strong><br />

PEO. These comparisons have been done over a very large range <strong>of</strong> <strong>polymer</strong> chain length, from short<br />

oligomers to long PEO molecules.<br />

Thermodynamic vs. Kinetic Control <strong>of</strong> Protein Adsorption<br />

Due to the very large energy scales involved on the adsorption <strong>of</strong> proteins it is important to understand the<br />

equilibrium and kinetic process. When a surface is modified by grafting <strong>polymer</strong> molecules the modified<br />

surface-protein interactions give rise to changes in both the kinetics <strong>of</strong> adsorption as well as the equilibrium<br />

amount <strong>of</strong> proteins adsorbed. The question is then what type <strong>of</strong> control is necessary depending on the type <strong>of</strong>


application. For example, in the case <strong>of</strong> biocompatible materials it is important to obtain thermodynamic<br />

prevention <strong>of</strong> protein adsorption, since the biomaterial is expected to be in contact with the blood stream for<br />

long periods <strong>of</strong> time. However, in the case <strong>of</strong> drug carriers, control <strong>of</strong> the kinetics <strong>of</strong> adsorption is enough,<br />

since in this case one is interested in preventing protein adsorption on the time scale that the drug is<br />

delivered to the target cell.<br />

We will show our theoretical studies on both the kinetic and thermodynamic control <strong>of</strong> protein adsorption.<br />

Most <strong>of</strong> our studies are on model PEO molecules since these are the <strong>polymer</strong>s that have been most<br />

extensively used experimentally. PEO is a water-soluble <strong>polymer</strong> that also has affinity for hydrophobic<br />

surfaces. Therefore, the structure <strong>of</strong> the <strong>polymer</strong> layer and thus, its ability to control protein adsorption<br />

depends upon the <strong>polymer</strong>-surface interactions.<br />

Overall we find that when the <strong>polymer</strong>s are attracted to the surfaces the <strong>polymer</strong> layer is more effective for<br />

the thermodynamic control <strong>of</strong> protein adsorption than surfaces that do not attract the <strong>polymer</strong>. For the kinetic<br />

control the opposite is true. Namely, surfaces with <strong>polymer</strong>s that are not attracted to them exert a stronger<br />

steric repulsion to approaching proteins than surfaces with attracting <strong>polymer</strong>s. In all cases <strong>polymer</strong> surface<br />

coverage is the most important parameter to prevent protein adsorption. As the surface coverage increases,<br />

the ability <strong>of</strong> the <strong>polymer</strong> to prevent protein adsorption increases. At fixed <strong>polymer</strong> surface coverage, the<br />

equilibrium adsorption <strong>of</strong> proteins is independent <strong>of</strong> <strong>polymer</strong> molecular weight once the thickness <strong>of</strong> the brush<br />

is larger than the size <strong>of</strong> the protein. However, increasing chain length at fixed surface coverage results in an<br />

exponential increase <strong>of</strong> the time scale for adsorption. This is the result <strong>of</strong> the increased range and strength <strong>of</strong><br />

the effective surface-protein interactions as <strong>polymer</strong> molecular weight increases.<br />

The kinetic process <strong>of</strong> protein adsorption on surfaces with <strong>grafted</strong> <strong>polymer</strong>s will be shown to be rather<br />

complex. This is due to the fact that as proteins adsorb the <strong>polymer</strong> layer is deformed and therefore, the<br />

interactions that the proteins arriving from solution feel changes with time.<br />

The changes in the adsorption process due to the ability <strong>of</strong> the proteins to change their conformation upon<br />

adsorption will be described in detail as well as the effect <strong>of</strong> charges on the proteins, <strong>polymer</strong>s and surfaces.<br />

References<br />

Most <strong>of</strong> the work presented in this talk appears in the following publications.<br />

1. Szleifer, I., 1997. "Protein Adsorption on Surfaces with Grafted Polymers: A Theoretical Approach", Biophysical<br />

J. 72, 595-612.<br />

2. Szleifer, I., 1997. “Protein Adsorption on Tethered Polymer Layers: Effect <strong>of</strong> Polymer Chain Architecture<br />

and Composition”, Physica A 244, 370-388.<br />

3. Szleifer, I., 1997. “Polymers and Proteins: Interactions at Interfaces: Current Opinion in Solid State and<br />

Materials Science 2, 337-344.<br />

4. McPherson, T., Kidane, A., Szleifer, I. and Park, K., 1998. “Prevention <strong>of</strong> Protein Adsorption by Tethered<br />

PEO Layers: Experiments and Single Chain Mean Field Analysis”. Langmuir 14, 176-186.<br />

5. Satulovsky, J., Carignano, M.A. and Szleifer, I., 2000. “Kinetic and Thermodynamic Control <strong>of</strong> Protein<br />

Adsorption”. Proc. Nat. Acad. Sci. 97, 9037-9041.<br />

6. Carignano M.A. and Szleifer, I., 2000. “Prevention <strong>of</strong> Protein Adsorption by Flexible and Rigid Chain<br />

Molecules”, Colloids and Surfaces B: Biointerfaces 18, 169-182.<br />

7. Szleifer I. and Carignano, M.A., 2000. “Tethered Polymer Layers: Phase Transitions and Reduction <strong>of</strong><br />

Protein Adsorption”. Feature Article in: Macromolecular Rapid Communications 21, 423-448.<br />

8. Carignano M.A. and Szleifer, I., 2002. “Adsorption <strong>of</strong> Model Charged Proteins on Charged Surfaces with<br />

Grafted Polymers”. Mol. Phys. (in press).<br />

9. Fang Fang and Szleifer, I., 2002. “Effect <strong>of</strong> Molecular Structure on the Adsorption <strong>of</strong> Protein on Surfaces<br />

with Grafted Polymers”. Langmuir 18, 5497-5510.


MODELLING STRUCT<strong>UR</strong>E AND ADHESION AT POLYMER-POLYMER INTERFACES<br />

D.N. Theodorou<br />

School <strong>of</strong> Chemical Engineering, National Technical University <strong>of</strong> Athens,<br />

GR-15780 Athens, Greece and ICE/HT-FORTH, GR-26500 Patras, Greece<br />

email: doros@central.ntua.gr<br />

ABSTRACT<br />

Co<strong>polymer</strong>s are remarkably effective in improving adhesion between immiscible <strong>polymer</strong>s, or between<br />

<strong>polymer</strong>s and solid substrates on which the <strong>polymer</strong> chains do not adsorb strongly. Predicting the molecular<br />

organisation and mechanical response upon deformation <strong>of</strong> <strong>polymer</strong>/<strong>polymer</strong> and <strong>polymer</strong>/solid interfaces<br />

strengthened with co<strong>polymer</strong>s can serve as a basis for the “molecular engineering” design <strong>of</strong> materials with<br />

optimal interfacial properties.<br />

Using the work <strong>of</strong> Scheutjens and Fleer (Scheutjens 1979) as a starting point, we have developed a<br />

lattice-based self-consistent field (SCF) theory for capturing the molecular-level structure <strong>of</strong> interfaces formed<br />

by <strong>polymer</strong> melts in contact with other <strong>polymer</strong> melts or solid substrates, in the absence or presence <strong>of</strong><br />

co<strong>polymer</strong>s. Real <strong>polymer</strong> and co<strong>polymer</strong> chains are mapped <strong>onto</strong> the model representation invoked in the<br />

theory in a manner that respects their mass density in the homogeneous bulk, their c<strong>onto</strong>ur length, and their<br />

conformational stiffness (characteristic ratio). The propagation <strong>of</strong> conformations is envisioned as a secondorder<br />

Markov process, involving an energetic penalty for chain bending. Interaction (χ) parameters between<br />

dissimilar segments, when needed, are obtained from fitting binary interfacial widths between the<br />

corresponding <strong>polymer</strong> melts.<br />

Application <strong>of</strong> the theory to interfaces between polystyrene (PS) and poly(methyl methacrylate) (PMMA)<br />

(Fischel 1995) in the presence <strong>of</strong> a symmetric PS-PMMA diblock co<strong>polymer</strong> gave segment density pr<strong>of</strong>iles in<br />

very favourable agreement with neutron reflectivity measurements (Russell 1991). The end-segment, mean<br />

square junction-to-end distance normal to the interface, shape parameter, and bond-order parameter pr<strong>of</strong>iles<br />

indicate that co<strong>polymer</strong> chains are significantly stretched under experimental conditions. For fixed surface<br />

density <strong>of</strong> the co<strong>polymer</strong>, with increasing block length, a broad transition from reflected random coil to “brush”<br />

behaviour occurs over the region where the block radii <strong>of</strong> gyration are 1.2 to 1.7 times the mean interchain<br />

spacing. At very high surface densities, the co<strong>polymer</strong> can no longer reside as a monolayer at the surface.<br />

Surface-phase equilibrium calculations predict that patches <strong>of</strong> trilayer will appear when the volume <strong>of</strong><br />

co<strong>polymer</strong> per unit surface is roughly three times the block unperturbed root mean square radius <strong>of</strong> gyration.<br />

This predicted appearance <strong>of</strong> structures more complex than a monolayer at the surface is consistent with the<br />

emergence <strong>of</strong> <strong>of</strong>f-specular scattering in neutron reflectivity measurements. It suggests an upper limit to the<br />

degree <strong>of</strong> mechanical strengthening <strong>of</strong> the interface by addition <strong>of</strong> larger amounts <strong>of</strong> block co<strong>polymer</strong>, as<br />

seen experimentally.<br />

Our SCF theory, in parallel with neutron reflectivity measurements, has been applied to probe interfaces<br />

between PS <strong>of</strong> molar mass 186 kg/mol and polished silicon wafer substrates (SiO2), in the presence <strong>of</strong><br />

diblock co<strong>polymer</strong>s <strong>of</strong> poly(2-vinyl pyridine) (PV2P) and PS as adhesion promoters (Retsos 2002). In these<br />

systems, prepared from the melt, the PV2P (“anchor”) block <strong>of</strong> the co<strong>polymer</strong> adsorbs strongly on the<br />

surface, while the “dangling” PS block mixes with the PS homo<strong>polymer</strong>. The molar mass <strong>of</strong> the dangling<br />

block, which was deuterated and therefore visible in the experiments, was kept constant at 75 kg/mol, while<br />

that <strong>of</strong> the anchor block was varied systematically between 3.4 and 102 kg/mol. The dangling block pr<strong>of</strong>ile<br />

was found to exhibit a maximum, which decreases in height and moves away from the surface as the length<br />

<strong>of</strong> the anchor block is increased. Results from the SCF calculation are in very good agreement with the<br />

experiment. The anchoring block is not flat (pancakelike) as <strong>of</strong>ten assumed, but forms a layer <strong>of</strong> thickness


10-50 Å, which increases with the block length. Both experiments and theory reveal evidence for the<br />

existence <strong>of</strong> three regimes regarding the configuration <strong>of</strong> the dangling blocks: a “wet brush” regime, a<br />

“mushroom” regime, and a broad transition regime in between.<br />

SCF calculations can serve as a starting point for predicting mechanical failure at interfaces subjected to<br />

large-scale deformations. A hierarchical simulation approach has been designed for this purpose. Using as<br />

input the chemical constitutions and relative amounts <strong>of</strong> chain species present at the interface, the<br />

composition pr<strong>of</strong>iles and conformational characteristics for all these species are first calculated by SCF<br />

theory (Terzis 2000a). The statistical weights derived from the model are then used within a Monte Carlo<br />

procedure to generate large ( ≈ 0.1 µm-sized) three-dimensional computer “specimens” <strong>of</strong> the interfacial<br />

region. In these specimens the material is represented, in a coarse-grained sense, as a network <strong>of</strong><br />

entanglement points. Each entanglement point is shared by two chains. Each chain is defined by the<br />

positions <strong>of</strong> its ends and <strong>of</strong> all entanglement points it traverses, as well as by its c<strong>onto</strong>ur lengths between<br />

these points (Terzis 2000b). A coarse-grained free energy function incorporating excluded volume, dispersion<br />

attraction, and conformational entropy contributions is written for the network. An efficient numerical<br />

procedure is invoked for finding local minima <strong>of</strong> this free energy function with respect to the positions <strong>of</strong> all<br />

chain ends and entanglement points, and thereby imposing the condition <strong>of</strong> mechanical equilibrium.<br />

Deformation <strong>of</strong> the network to fracture (Figure 1) at prescribed temperature and strain rate is simulated<br />

through a kinetic Monte Carlo procedure, which tracks elementary events <strong>of</strong> chain slippage across<br />

entanglements, chain disentanglement, chain reentanglement, and chain rupture (Terzis 2002). This<br />

hierarchical procedure has been applied to study interfaces between polypropylene (PP) and polyamide 6<br />

(PA6) compatibilised with the reaction product between maleic anhydride-functionalised PP (PP-g-MA) and<br />

PA6. Such interfaces can be viewed as consisting <strong>of</strong> a homo<strong>polymer</strong> (free PP) next to an impenetrable solid<br />

surface (PA6), <strong>onto</strong> which are <strong>grafted</strong> chains <strong>of</strong> the same chemical constitution as the homo<strong>polymer</strong> (end<strong>grafted</strong><br />

PP). The effects <strong>of</strong> the surface density <strong>of</strong> <strong>grafted</strong> PP and <strong>of</strong> the molecular weight distribution <strong>of</strong> free<br />

PP and <strong>grafted</strong> PP have been explored. It is clearly seen that increasing the surface grafting density does not<br />

necessarily enhance adhesion. For high surface grafting densities, a “brush” <strong>of</strong> <strong>grafted</strong> PP builds up next to<br />

the PA6 surface; as a consequence, the region over which <strong>grafted</strong> and free PP chains interentangle is <strong>of</strong><br />

limited width. For monodisperse <strong>grafted</strong> PP <strong>of</strong> molar mass 40 kg/mol in a free PP matrix <strong>of</strong> molar mass 60<br />

kg/mol, optimal adhesion (a maximum in the work required to destroy the interface) is seen at 0.1 <strong>grafted</strong><br />

chains/nm 2 .<br />

TRUE STRESS (MPa)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

1% per sec<br />

10% per sec<br />

0<br />

1 2 3 4<br />

DRAW RATIO<br />

Figure 1: Shapshots and stress-draw ratio curves from tensile deformation computer experiments carried out on<br />

network specimens modelling the PP/PP-g-MA/PA6 interface. Grafted and free PP chains are shown in red and<br />

green, respectively. The stress-strain curves have been obtained at two strain rates for a surface grafting density <strong>of</strong><br />

0.10 chains/nm 2 . The specimen, <strong>of</strong> initial dimensions 45 nm × 45 nm × 70 nm, contained 1300 chains.


References<br />

1. Fischel, L.B. and Theodorou, D.N., 1995. Self-consistent field model <strong>of</strong> the <strong>polymer</strong>/diblock co<strong>polymer</strong>/<br />

<strong>polymer</strong> interface. J. Chem. Soc. Faraday Trans., 91 (16), 2381-2402.<br />

2. Retsos, H., Terzis, A.F., Anastasiadis, S.H., Anastassopoulos, D.L., Toprakcioglu, C., Theodorou, D.N.,<br />

Smith, G.S., Menelle, A., Gill, G.E., Hadziioannou, G., Gallot, Y., 2002. Mushrooms and <strong>brushes</strong> in thin<br />

films <strong>of</strong> diblock co<strong>polymer</strong>/homo<strong>polymer</strong> mixtures. Macromolecules, 35(3), 1116-1132.<br />

3. Russell, T.P., Anastasiadis, S.H., Menelle, A., Felcher, F.P., Satija, S.K., 1991. Segment density<br />

distribution <strong>of</strong> symmetric diblock co<strong>polymer</strong>s at the interface between two homo<strong>polymer</strong>s as revealed by<br />

neutron reflectivity. Macromolecules, 24(7), 1575-1582.<br />

4. Scheutjens, J.M.H.M. and Fleer, G.J., 1979. Statistical theory <strong>of</strong> the adsorption <strong>of</strong> interacting chain<br />

molecules 1. Partition function, segment density distribution, and adsorption isotherms. J. Phys. Chem.,<br />

83(12), 1619-1635.<br />

5. Terzis, A.F., Theodorou, D.N., Stroeks, A., 2000. Entanglement network <strong>of</strong> the polypropylene/polyamide<br />

interface. 1. Self-consistent field model. Macromolecules, 33(4), 1385-1396.<br />

6. Terzis, A.F., Theodorou, D.N., Stroeks, A., 2000. Entanglement network <strong>of</strong> the polypropylene/polyamide<br />

interface. 2. Network generation. Macromolecules, 33(4), 1397-1410.<br />

7. Terzis, A.F., Theodorou, D.N., Stroeks, A., 2002. Entanglement network <strong>of</strong> the polypropylene/polyamide<br />

interface. 3. Deformation to fracture. Macromolecules, 35(2), 508-521.


NEUTRON REFLECTIVITY STUDY OF POLY(STYRENE MALEIC ANHYDRIDE) ON THE<br />

WATER-AIR INTERFACE<br />

C. Malardier-Jugroot 1) , T.G.M. v.d. Ven 1) , M.A. Whitehead ) , R. Richardson 2) , T. Cosgrove 2)<br />

1) Department <strong>of</strong> Chemistry, McGill University, Montreal, Canada<br />

2) School <strong>of</strong> Chemistry, University <strong>of</strong> Bristol, United Kingdom<br />

email: theo.vandeven@mcgill.ca<br />

ABSTRACT<br />

Poly(styrene maleic anhydride) (SMA) is an interesting alternating co<strong>polymer</strong>, used in paper treatment<br />

applications. Previous studies showed interesting behavior <strong>of</strong> SMA in aqueous solutions, characterized by<br />

self-association, depending on pH and ionic strength. The neutron reflectivity study was performed to see if<br />

surface properties show similar interesting behavior.<br />

The results for the ethyl esters in DO 2 were indistinguishable from the reflectivity <strong>of</strong> pure DO, 2 indicating<br />

that the adsorption layer is very thin. The data can be accurately fitted to the theoretical reflectivity <strong>of</strong> DO, 2<br />

using a surface roughness <strong>of</strong> 2-3 Å.<br />

The neutron reflectivity data on the ethyl ester <strong>of</strong> SMA in NRW showed surprisingly little effect on pH. The<br />

2<br />

data were analyzed by plotting<br />

RQ vs. Q. A linear relation was obtained from which the adsorbed amount<br />

Γ and the adsorption layer thickness δ can be obtained from the intercept and the slope respectively. We<br />

found an interesting relaxation effect for the reflectivity <strong>of</strong> the ethyl ester SMA at pH 7, absent at other pH’s.<br />

The results obtained after 4 hours were significantly different from the initial data. With time, Γ decreased<br />

from 1.34 to 1.11 mg/m 2 and δ decreased from 12.6 to 9.6 Å (for the ethyl deuterated ethyl ester <strong>of</strong> SMA).<br />

The adsorption at pH 3 was identical to the relaxed adsorption at pH 7. The styrene deuterated compound<br />

always showed a slightly higher adsorbed amount compared to the ethyl deuterated one: Γ = 1.2 and 1.1<br />

mg/m 2 resp.<br />

The results at pH 13 showed that the ethyl ester <strong>of</strong> SMA was partially hydrolyzed, resulting in a reflectivity<br />

pr<strong>of</strong>ile close to that <strong>of</strong> SMA at the same pH.<br />

The adsorbed amounts for SMA were found to be extremely low, both for pH 7 and pH 13, with adsorbed<br />

amounts <strong>of</strong> 0.38 and 0.24 mg/m 2 resp. The thickness <strong>of</strong> the adsorption layer was in the range 13 – 16 Å.


MULTIDOMAIN CONFORMATIONS OF RANDOM HETEROPOLYMERS<br />

J.M.P. van den Oever<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: oever@fenk.wau.nl<br />

ABSTRACT<br />

Polymer molecules can attain various random-coil and compact structures in solution and at interfaces.<br />

The <strong>polymer</strong>'s primary structure and its environment determine which conformation is prevalent. The<br />

influence <strong>of</strong> primary structure on stability and higher structure is an area <strong>of</strong> intensive research, especially<br />

in protein science. A novel two gradient SCF model is used to study various types <strong>of</strong> hetero<strong>polymer</strong>s. In<br />

this method the complete partition function <strong>of</strong> the <strong>polymer</strong> system is calculated. Intramolecular excluded<br />

volume is taken into account partially. The average conformations <strong>of</strong> molecules under consideration are<br />

obtained in the form <strong>of</strong> volume fraction pr<strong>of</strong>iles. Random co<strong>polymer</strong> sequence space has been sampled<br />

by investigating molecules <strong>of</strong> varying length and monomer ratio. The number <strong>of</strong> domains attained in<br />

equilibrium was observed and summarized in a sequence space diagram. Random hetero<strong>polymer</strong>s with<br />

pH-dependant characteristics mimicing those <strong>of</strong> proteins have been studied in a wide pH-range. We see<br />

that some molecules form multiple domains when going to extreme pH values. Domain formation can also<br />

be observed in hetero<strong>polymer</strong>s approaching an oppositely charged surface.


COIL-SIZE OSCILLATORY PACKING IN POLYMER SOLUTIONS NEAR A S<strong>UR</strong>FACE<br />

J. van der Gucht, N.A.M. Besseling, J. van Male, M.A. Cohen Stuart<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: jasper@fenk.wau.nl<br />

ABSTRACT<br />

Depletion <strong>of</strong> <strong>polymer</strong>s at a surface has been studied extensively before. Theories predict that the segment<br />

concentration pr<strong>of</strong>ile decreases monotonously with decreasing distance to the surface 1,2) . Furthermore it is<br />

generally accepted that the only relevant lengthscale above the overlap concentration is the “blob size”. The<br />

coil size is believed to be irrelevant in this regime 1,2) . We have performed detailed calculations <strong>of</strong> the density<br />

pr<strong>of</strong>ile <strong>of</strong> nonadsorbing <strong>polymer</strong>s near a surface, using the theory developed by Scheutjens and Fleer 1) . Our<br />

results indicate that there is a damped oscillatory contribution to the density pr<strong>of</strong>ile. Both the period <strong>of</strong> the<br />

oscillations and the decay length are proportional to the size <strong>of</strong> the individual coils and are independent <strong>of</strong> the<br />

<strong>polymer</strong> concentration, also above the overlap concentration. This indicates that the size <strong>of</strong> the individual<br />

coils is still a relevant lengthscale above the overlap concentration. The oscillations are associated with a<br />

liquid-like layering <strong>of</strong> <strong>polymer</strong> coils near the surface. In dilute solutions no oscillations are observed, because<br />

the decay length <strong>of</strong> the oscillations is smaller than the depletion correlation length. This is similar to the<br />

behaviour <strong>of</strong> simple fluids, where the Fisher-Widom line marks the transition from monotonic to oscillatory<br />

decay <strong>of</strong> the density correlation function 3) . The Fisher-Widom line in <strong>polymer</strong> solutions is proportional to the<br />

overlap concentration. On the oscillatory side <strong>of</strong> the Fisher-Widom line the interaction energy between two<br />

plates immersed in a solution <strong>of</strong> nonadsorbing <strong>polymer</strong>s is an oscillatory function <strong>of</strong> the separation distance.<br />

The size <strong>of</strong> the oscillations is too small to be detected experimentally, but the effect is expected to be<br />

stronger in for example branched <strong>polymer</strong> solutions. As indicated by Evans et al. the oscillations might be <strong>of</strong><br />

importance in, for example, wetting phenomena 4) .<br />

References<br />

1. Fleer, G.J., Cohen Stuart, M.A., Scheutjens, J.M.H.M.; Cosgrove, T., Vincent, B. Polymers at Interfaces.<br />

Chapman and Hall: London, 1993.<br />

2. De Gennes, P.G., Scaling Concepts in Polymer Physics; Cornell University Press: Ithaca, London.<br />

3. Fisher, M.E., and Widom, B., 1969. J. Chem. Phys. 50, 3756.<br />

4. Evans, R., Leote de Carvalho, R.J.F., Henderson, J.R. and Hoyle, D.C., 1993. J. Chem. Phys., 100, 591.


SUPRAMOLECULAR COORDINATION POLYMERS: CHAINS AND RINGS<br />

J. van der Gucht<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6708 HB <strong>Wageningen</strong>, The Netherlands<br />

email: jasper@fenk.wau.nl<br />

ABSTRACT<br />

Bifunctional monomers were synthesized, consisting <strong>of</strong> two ligand groups bridged by a short spacer. With<br />

metal ions, these molecules can form coordination <strong>polymer</strong>s. This can be monitored by measuring the<br />

viscosity. For linear chains one expects a maximum in the visocity at a 1:1 ratio <strong>of</strong> metal to monomers. At low<br />

concentrations however, a dip is measured there. This dip can be explained using a simple model for<br />

equilibrium <strong>polymer</strong>ization into rings and chains, which predicts that short rings dominate at low<br />

concentrations at a 1:1 ratio, causing a minimum in the average length, and, hence, in the viscosity.


ASSOCIATION BEHAVIO<strong>UR</strong> OF GLUCITOL AMINE GEMINI S<strong>UR</strong>FACTANTS<br />

SELF-CONSISTENT-FIELD THEORY AND MOLECULAR DYNAMICS SIMULATIONS<br />

M. van Eijk, M. Bergsma, S.-J. Marrink<br />

CP Kelco,<br />

Ved Banen 16, Lille Skensved, 4623, Denmark<br />

email: marcel.vaneijk@cpkelco.com<br />

ABSTRACT<br />

The association behaviour <strong>of</strong> a number <strong>of</strong> glucitol amine gemini surfactants has been investigated by means<br />

<strong>of</strong> molecular dynamics and self-consistent-field calculations. We have shown that the titratable head group <strong>of</strong><br />

the surfactant is responsible for a micelle-to-membrane transition when changing the pH. Furthermore, the<br />

association structure <strong>of</strong> this group <strong>of</strong> surfactants is shown to be very sensitive to ionic strength. The<br />

combination <strong>of</strong> a charged head group, a spacer, and the hydrophilic glucitol side chains is responsible for the<br />

possible structural transitions in the associates as a function <strong>of</strong> ionic strength and pH.


INFLUENCE OF POLYMERS ON THE BENDING MODULI OF BILAYERS<br />

J. van Male, F.A.M. Leermakers<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: vanmale@fenk.wau.nl<br />

ABSTRACT<br />

Surfactant bilayers have a low surface tension. The rigidity <strong>of</strong> bilayers determines the phase behavior <strong>of</strong> the<br />

bilayer membranes. We investigated the effect <strong>of</strong> homo<strong>polymer</strong>s on the bending moduli <strong>of</strong> bilayer membranes<br />

by numerical self-consistent field calculations. Adsorbing <strong>polymer</strong>s increase the mean bending<br />

modulus k c , resulting in stiffer bilayers and decrease the Gaussian bending modulus k , disfavoring handles<br />

between the membranes. Overall, adsorbing <strong>polymer</strong>s stabilize the bilayers. Calculations that approximate<br />

the membrane as an inert solid wall give opposite results. This shows that the membrane should be explicitly<br />

taken into account.


INTERFACES, WETTING, AND CAPILLARY NEMATISATION IN COLLOIDAL ROD<br />

SUSPENSIONS<br />

R. van Roij<br />

Institute for Theoretical Physics, Utrecht University,<br />

Postbus 80.195, 3508 TD, Utrecht, The Netherlands<br />

email: r.vanroij@phys.uu.nl<br />

ABSTRACT<br />

We study the thermodynamics and structure <strong>of</strong> interfaces and films in colloidal hard-rod fluids within an<br />

Onsager-like density functional theory. For a one-component system, we find complete wetting near a hard<br />

wall and capillary nematisation (with a capillary critical point) in a slit confined by two hard walls. For binary<br />

mixtures <strong>of</strong> hard rods we study the <strong>planar</strong> interfaces <strong>of</strong> coexisting isotropic and nematic phases. We calculate<br />

the surface tension, the interface-induced biaxiality, and the density and order parameter pr<strong>of</strong>iles through the<br />

interface. In the case that the bulk phase diagram exhibits an isotropic-nematic-nematic (I-N1-N2) triple point,<br />

we find complete wetting <strong>of</strong> the I-N2 interface by N1 upon approach <strong>of</strong> the triple point. This triple-point wetting<br />

is an entropy-driven analogue <strong>of</strong> e.g. surface melting <strong>of</strong> a metal upon approach <strong>of</strong> the melting temperature.<br />

References<br />

1. Van Roij, R., Dijkstra, M. and Evans, R., 2000. Europhys. Lett. 49, 350.<br />

2. Shundyak, K. and van Roij, R., 2002. Phys. Rev. Lett. 88, 205501.


SPINODAL DECOMPOSITION IN BINARY POLYMERIC BLENDS<br />

M.E. Velázquez-Sánchez, P.D. Anderson, P.G.T. van der Varst, G. de With<br />

Technische Universiteit Eindhoven,<br />

Postbus 513, 5600 MB Eindhoven, The Netherlands<br />

email: M.velazquez-sanchez@tue.nl<br />

ABSTRACT<br />

This work aims to predict the morphology development in films, on a substrate, made from binary <strong>polymer</strong>ic<br />

blends that undergo phase separation via a spinodal decomposition mechanism. From thermodynamics it is<br />

possible to predict the spinodal and binodal regions <strong>of</strong> a system, once the Flory-Huggins interaction<br />

parameter is known; however to know how the morphology develops as a function <strong>of</strong> time, the diffusion<br />

equation must be solved. A diffuse-interface model that couples the thermodynamics and the hydrodynamics<br />

<strong>of</strong> a binary system is used. As an input, this model uses a chemical potential where the wall interaction and<br />

the non-local effects in the enthalpy and the entropy are taken into account. Preliminary results show that the<br />

introduction <strong>of</strong> the wall potential plays an important role in the development <strong>of</strong> the spinodal decomposition,<br />

resulting in a more complex morphology, where formation <strong>of</strong> macroscopic layers close to the substrate is<br />

obtained. Additionally, the growth rate <strong>of</strong> concentration fluctuations is significantly influenced by the wall.


EXPERIMENTAL AND THEORETICAL BEHAVIO<strong>UR</strong> OF END-TETHERED POLYMERS FROM<br />

THE MUSHROOM TO THE BRUSH LIMITS<br />

M.D. Whitmore 1) , M. Pépin 1) , M.S. Kent 2) , G. Grest 2) , J. Douglas 3)<br />

1)<br />

Dept. <strong>of</strong> Physics and Physical Oceanography, Memorial University <strong>of</strong> Newfoundland,<br />

St. John's, NF, Canada, A1B 3X7<br />

2) Sandia National Laboratory,<br />

Albuquerque, NM 87185, U.S.A.<br />

3)<br />

Polymers Division, National Institute <strong>of</strong> Science and Technology,<br />

Gaithersburg, MD 20899, U.S.A.<br />

email: markw@physics.mun.ca<br />

ABSTRACT<br />

We present systematic experimental results on the thickness <strong>of</strong> end-tethered <strong>polymer</strong> layers in good and Θ<br />

solvents. These results clearly show that most experimental systems fall between the mushroom and brush<br />

regimes. We then show that numerical self-consistent field theory, which uses appropriate values for<br />

monomer and solvent volumes and statistical segment lengths, provides quantitative agreement with<br />

experiment in these ranges. We go on to introduce a new parameter for characterizing the tethered-layer<br />

coverage, and use it to propose a unified description <strong>of</strong> these systems that applies from the mushroom<br />

regime, through the intermediate regime where most experiments occur, to deep into the brush limit, for<br />

layers in both good and Θ solvents. The analysis is based on a generalization <strong>of</strong> the asymptotic SCF theory,<br />

numerical SCF theory, Monte Carlo and molecular dynamics simulations, and the experimental data.


PERSISTENCE LENGTH OF WORM-LIKE MICELLES OF NON-IONIC S<strong>UR</strong>FACTANTS<br />

Y. Lauw, F.A.M. Leermakers<br />

Laboratory <strong>of</strong> Physical Chemistry and Colloid Science, <strong>Wageningen</strong> University,<br />

Dreijenplein 6, 6703 HB <strong>Wageningen</strong>, The Netherlands<br />

email: yansen@fenk.wau.nl<br />

ABSTRACT<br />

The persistence length <strong>of</strong> worm-like micelles is calculated by using two-gradient self-consistent field (SCF)<br />

theory within a cylindrical coordinates system. The CnEm non-ionic surfactants are used and forced into a<br />

toroidal-shaped micelles. This leads to a persistent length l p which scales with the tail length n as l x<br />

p ∞ n in<br />

which x is about 2.5. Furthermore the calculated bending energy at high total curvatures J, shows a deviation<br />

that can be attributed to a higher order term proportional to J 3 .


COMPLEXATION OF POLYELECTROLYTES AND POLYAMPHOLYTES WITH CHARGED<br />

OBJECTS<br />

E.B. Zhulina<br />

Institute <strong>of</strong> Macromolecular Compounds <strong>of</strong> the Russian Academy <strong>of</strong> Sciences,<br />

199004 St.Petersburg, Russia<br />

email: kzhulina@hotmail.com<br />

ABSTRACT<br />

1. Introduction<br />

Charged <strong>polymer</strong>s are capable <strong>of</strong> aggregating and forming complexes <strong>of</strong> various morphologies in solutions.<br />

The equilibrium structure and the thermodynamic stability <strong>of</strong> the aggregates are determined by a<br />

number <strong>of</strong> factors: the distribution <strong>of</strong> charges on macromolecules, the <strong>polymer</strong> concentration, the pH and the<br />

ionic strength <strong>of</strong> the solution, etc. In this lecture we focus on two different systems: (1) the complexes <strong>of</strong><br />

random polyampholytes with charged spherical particles and (2) ionized <strong>polymer</strong> micelles interacting with<br />

oppositely charged polyions. We analyze the equilibrium properties <strong>of</strong> the complexes by using the scaling<br />

model and the numerical Scheutjens and Fleer self-consistent-field model (Fleer 1993). We demonstrate that<br />

in spite <strong>of</strong> evident differences between the two systems, the soluble complexes exhibit similar features in a<br />

range <strong>of</strong> conditions. Namely, in the salt-free dilute solutions, they can be envisioned as effectively neutral<br />

star-like <strong>polymer</strong>s.<br />

2. Interaction <strong>of</strong> polyampholytes with charged spherical particles<br />

We consider a water solution <strong>of</strong> long polyampholyte molecules, each comprising N >> 1 monomers <strong>of</strong> size<br />

a. The chains are flexible, and the Bjerrum length lB is on the order <strong>of</strong> the monomer size, IB≅a.The positive<br />

and negative charges are distributed in a random fashion along the chain. The respective fractions f + and f−<br />

<strong>of</strong> charged monomers are assumed to be almost equal ( f+ ≈ f− = f ) and small ( fN 2


For small particles with radius R < R ≅ 1/2<br />

o aN , the two effects give rise to the novel “pseudo-star”<br />

conformation <strong>of</strong> polyampholyte chain (Zhulina 2001). Here, the equilibrium structure <strong>of</strong> the complex is<br />

determined by the interplay <strong>of</strong> three factors: the polarization <strong>of</strong> the loops, the stretching <strong>of</strong> the loops, and the<br />

ternary contacts between monomers (theta-solvent conditions). The analysis (Zhulina 2001; Dobrynin 2001)<br />

indicates that in the pseudo-star conformation, the loops have a bimodal distribution. The smaller loops <strong>of</strong><br />

size ~ R decorate the particle (the core <strong>of</strong> the pseudo-star). The larger loops <strong>of</strong> size H ~ aN1/2( Q2f ) −1/<br />

6 form<br />

the “branches” <strong>of</strong> the pseudo-star. The number <strong>of</strong> branches P ~( Q2f ) 2/3 does not depend on chain length N<br />

and is determined solely by particle charge Q and degree <strong>of</strong> chain ionization f.<br />

In excess <strong>of</strong> polyampholyte molecules in the bulk solution, each complex is comprised <strong>of</strong> many chains.<br />

One can envision such a situation as formation <strong>of</strong> a semi-dilute adsorbed layer near the charged particle. The<br />

structure <strong>of</strong> the adsorbed layer (the <strong>polymer</strong> density pr<strong>of</strong>ile, the distribution <strong>of</strong> loops, etc.) depends on the<br />

particle size and the surface charge density. For small particles with R < R ≅ 1/2<br />

o aN , the internal part <strong>of</strong> the<br />

adsorbed layer is found in the pseudo-star conformation. (It contains P chains each forming a single branch<br />

<strong>of</strong> the pseudo-star). The <strong>polymer</strong> density pr<strong>of</strong>ile cr ()~ P1/2r −1<br />

decays inversely proportional to the distance r<br />

from the particle. The exterior part <strong>of</strong> the layer is formed by the less polarized chains (“poles” stretched in the<br />

electric field <strong>of</strong> spherical symmetry). For very large particles, adsorption occurs as on a <strong>planar</strong> surface<br />

(Dobrynin 1999). We calculate the adsorption isotherms for particles <strong>of</strong> different sizes and obtain the<br />

thickness <strong>of</strong> the adsorbed layer as a function <strong>of</strong> bulk <strong>polymer</strong> concentration. The results are compared with<br />

the experimental data on adsorption <strong>of</strong> gelatin.<br />

3. Interaction <strong>of</strong> ionized micelles with oppositely charged polyions<br />

We consider the association <strong>of</strong> an ionized quenched <strong>polymer</strong> micelle with an oppositely charged polyion.<br />

The star-like micelle consists <strong>of</strong> a neutral core and a charged corona. The respective degrees <strong>of</strong> ionization <strong>of</strong><br />

the corona and <strong>of</strong> the polyion, α and β, are assumed to be relatively small to ensure the stability (solubility) <strong>of</strong><br />

the complex. The number <strong>of</strong> <strong>polymer</strong> chains in the micelle is fixed, and the length <strong>of</strong> the polyion varies. An<br />

analytical self-consistent-field model indicates that when the polyion is long enough, the complex is virtually<br />

electroneutral as a whole (Simmons 2001). The compensating amount <strong>of</strong> oppositely charged polyion is<br />

dragged inside the micelle to release the small counterions from the corona. The structure <strong>of</strong> the equilibrium<br />

complex is rationalized in terms <strong>of</strong> the effective second ( v eff ) and third ( w eff ) virial coefficients <strong>of</strong> monomermonomer<br />

interactions, v = + α β 2<br />

eff v (1 / ) + χα/ β and w = + α β 3<br />

eff w (1 / ) . Here, v and w are the actual virial<br />

coefficients <strong>of</strong> monomer-monomer interactions <strong>of</strong> the corona chains, and χ is the Flory interaction parameter.<br />

We then use the numerical Scheutjens and Fleer model to explore the structure <strong>of</strong> the micelle/polyion<br />

complex in more detail. By performing calculations at various values <strong>of</strong> v, α, β, χ, and different amounts <strong>of</strong><br />

added polyion, we demonstrate that the equilibrium properties <strong>of</strong> the complex can be rationalized in terms <strong>of</strong><br />

the theory <strong>of</strong> neutral star-like <strong>polymer</strong>s. We also focus on the effect <strong>of</strong> charge distribution on the polyion and<br />

consider three different cases: (1) a homo<strong>polymer</strong> with the smeared distribution <strong>of</strong> charge, (2) a diblock<br />

co<strong>polymer</strong> with a charged and a neutral block, and (3) a multi-block co<strong>polymer</strong>. Although the fine details <strong>of</strong><br />

polyion distribution in the micelle/polyion complex are different in the three cases, the Scheutjens and Fleer<br />

model confirms the essential electroneutrality <strong>of</strong> the complex. The numerical findings are in reasonable<br />

agreement with predictions <strong>of</strong> the analytical self-consistent-field theory.<br />

Acknowledgements<br />

The results presented in this lecture were obtained in collaboration with Dr. Andrey Dobrynin (UNC), Pr<strong>of</strong>.<br />

Michael Rubinstein (UNC), Chris Simmons (UT), Pr<strong>of</strong>. Steve Webber (UT), Dr. Jan van Male (WU) and Dr.<br />

Frans Leermakers (WU). The financial support from the National Science Foundation USA (grants DMR-<br />

9973300 and DMR-9730777) and NWO Project 047.009.016 “Self-Organization and Structure <strong>of</strong> Bionanocomposites”<br />

is greatly acknowledged.


References<br />

1. Dobrynin, A.V., Obukhov, S.P. and Rubinstein, M., 1999. Long-Range Mutichain Adsorption <strong>of</strong> Polyampholytes<br />

on a Charged Surface. Macromolecules, 32(17), 5689-5700.<br />

2. Dobrynin, A.V., Zhulina, E.B. and Rubinstein, M., 2001. Structure <strong>of</strong> Adsorbed Polyampholyte Layers at<br />

Charged Objects. Macromolecules, 34(3), 627-639.<br />

3. Fleer, G.J., Cohen Stuart, M.A., Scheutjens, J.M.H.M., Cosgrove, T. and Vincent, B. Polymers at Interfaces,<br />

Chapman & Hall, 1993.<br />

4. Simmons, C., Webber, S.E. and Zhulina, E.B., 2001. Association <strong>of</strong> Ionized <strong>polymer</strong> Micelles with<br />

Oppositely Charged Polyelectrolytes. Macromolecules 34 (14), 5053- 5066.<br />

5. Zhulina, E.B., Dobrynin, A.V. and Rubinstein, M., 2001. Adsorption <strong>of</strong> a Polyampholyte on a Charged<br />

Spherical Particle. Eur. Phys. J. E , 5(1), 41-49.<br />

6. Zhulina, E.B., Dobrynin, A.V. and Rubinstein, M., 2001. Adsorption Isotherms <strong>of</strong> Polyampholytes at<br />

Charged Spherical Particles .J. Phys. Chem. B, 105(37), 8917-8930.


MEAN-FIELD THEORY FOR THE ADSORPTION OF FLEXIBLE POLYMERS D<strong>UR</strong>ING FLOW<br />

THROUGH GRANULAR POROUS MEDIA<br />

P. L.J. Zitha<br />

Delft University <strong>of</strong> Technology, Department <strong>of</strong> Applied Earth Sciences,<br />

Mijnbouwstraat 120, 2628 RX Delft, The Netherlands<br />

email: p.l.j.zitha@citg.tudelft.nl<br />

ABSTRACT<br />

We have developed a mean-field theory for the adsorption <strong>of</strong> flexible <strong>polymer</strong>s during flow through granular<br />

porous media. There exist two situations (Fig. 1): (a) At low velocity gradients (weak flows) the chains retain<br />

the coil conformation and adsorb according to the classic layer mode. The adsorption density w increases<br />

with time θ, first according to a linear kinetic equation leading to an exponential saturation law<br />

w( θ) = w e[1−exp(<br />

θ / τ)]<br />

and then following a non-linear kinetic process. The transition from the linear to the<br />

non-linear regimes is associated with the development <strong>of</strong> a strong reptation barrier scaling like w 9/4 , when<br />

the adsorbed chains begin to overlap. This occurs at a well-defined critical adsorption density w*<br />

corresponding to the critical overlap concentration for the bulk <strong>polymer</strong> solution. (b) At high velocity gradient<br />

(strong flow), the coiled chains coexist with strongly stretched ones giving rise to a bi-modal adsorption: the<br />

coiled chains adsorb according to the layer mode while the long (stretched) ones bridge over the smallest<br />

pores (bridging adsorption). These two adsorption modes are strongly correlated and the corresponding<br />

adsorption densities w and z obey a system <strong>of</strong> coupled non-linear kinetic equations. The transition from linear<br />

to non-linear regimes occurs now at a concentration v* = ( w + κ z ) , where κ is the fraction <strong>of</strong> segments <strong>of</strong> the<br />

bridging chains that are in the adsorbed layer. Remarkably, the mechanical trapping <strong>of</strong> the stretched chains<br />

overcomes the reptation barrier when bridging adsorption is consecutive to layer adsorption.<br />

Figure 1. Cross-section <strong>of</strong> a pore showing the adsorbed layer (hydrodynamic thickness εH) and one bridging chain. Two<br />

portions <strong>of</strong> the bridging chain (degree <strong>of</strong> <strong>polymer</strong>isation Nb,1 and Nb,2) penetrate the adsorbed layer while the central<br />

portion (degree <strong>of</strong> <strong>polymer</strong>ization Nb,c) is exposed to flow.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!