Download as pdf or txt
Download as pdf or txt
You are on page 1of 364

Sebastian Seiffert

Physical Chemistry of Polymers


Also of Interest
Macromolecular Chemistry.
Natural and Synthetic Polymers
Elzagheid, 
ISBN ----, e-ISBN ----

Microfluidics.
Theory and Practice for Beginners
Seiffert, Thiele, 
ISBN ----, e-ISBN ----

Mechanochemistry.
A Practical Introduction from Soft to Hard Materials
Colacino, Ennas, Halasz, Porcheddu (Eds.), 
ISBN ----, e-ISBN ----

Polymer Synthesis.
Modern Methods and Technologies
Wang, Yuan, 
ISBN ----, e-ISBN ----

Smart Polymers.
Principles and Applications
García JM, García FC, Reglero Ruiz, Vallejos, Trigo-López, 
ISBN ----, e-ISBN ----
Sebastian Seiffert

Physical Chemistry
of Polymers

A Conceptual Introduction

2nd Edition
Author
Prof. Dr. Sebastian Seiffert
Johannes Gutenberg University Mainz
Department of Chemistry
Duesbergweg 10–14
D-55128 Mainz, Germany
sebastian.seiffert@uni-mainz.de

ISBN 978-3-11-071327-5
e-ISBN (PDF) 978-3-11-071326-8
e-ISBN (EPUB) 978-3-11-071339-8

Library of Congress Control Number: 2022944934

Bibliographic information published by the Deutsche Nationalbibliothek


The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie;
detailed bibliographic data are available on the Internet at http://dnb.dnb.de.

© 2023 Walter de Gruyter GmbH, Berlin/Boston


Cover image: Sebastian Seiffert
Typesetting: Integra Software Services Pvt. Ltd.
Printing and binding: CPI books GmbH, Leck

www.degruyter.com
Foreword to the 2nd edition
The original foreword to this book’s first edition said that its intention is to converge
polymer chemistry and physics, such to also converge the striking specificity and
ubiquitous universality of polymers. Based on that intent, the aim of this book was
sketched to be a bridge between polymeric structures & dynamics and their result-
ing properties, thereby arcing between macromolecular science and engineering.
This bridge has been identified to be the field of physical chemistry of polymers. It
was also said that a major challenge in that context is that the concepts in physical
chemistry of polymers commonly appear to be rather new and unknown to stu-
dents, and that it is therefore a mission of this book to converge deep conceptual
understanding (with many excursions into the underlaying basic physical chemical
concepts) with solid mathematical formalisms.
All the foresaid still holds true just as well for this present second edition. Yet,
there is a further aspect to be added: Just when the first edition appeared, the world
fell into a pandemic era, which shifted minds, societies, and universities. It also
shifted teaching formats, catapulting us into the digital age. Well, actually we should
have been in that era way before: In the 15th century, the invention of movable-type
printing by Gutenberg in Mainz ushered in the second media revolution in human
history. (The first before that was the transition from spoken to written language, and
the third after that was the advent of electronic mass media.) Today, the world is in
the fourth media revolution: digitization and networking. However, until recently,
universities (almost) exclusively taught in a way that dates back to pre-Gutenberg
times: frontal lecturing. The pandemic of the SARS-CoV2 pathogen in 2020, just when
the first edition of this book appeared, eventually propelled the academic world into
the 21st century and established digital teaching formats. This textbook provides a
basis for such a format in the subject of physical chemistry of polymers. It is a basis
for a blended learning course, i.e., a teaching format that consists of a self-study
phase (knowledge acquisition), a digital feedback unit (knowledge anchoring), and
an interactive classroom unit (knowledge application).
For this purpose, the book is divided into 23 thematically focused and modu-
larly applicable lesson units, which can be conceived as 90-min lectures each. This
number marks a plus of five new lesson units to the former eighteen ones of the
book’s first edition, adding to its general polymer-physical concepts some specific
material-centered contents. With that, the second book edition can be used in its
entirety for a 12-week course of 2 × 90 minutes per week (this corresponds to 4 SWS
in the German system). Alternatively, it can also be used for two separate, sequen-
tial courses of only 1 × 90 minutes per week each (corresponding to 2 SWS in the
German system), one covering lesson units 1–9 and one adding lesson units 10–23
(perhaps cutting unit 19 if time is too tight).

https://doi.org/10.1515/9783110713268-202
VI Foreword to the 2nd edition

The lesson-unit blocks each comprise about 10–20 pages, each to be worked
through in about 90 minutes of self-study. Each unit concludes with a set of con-
ceptual questions in a multiple-choice format. Lecturers can incorporate some of
those into an e-learning platform so that students can solve them there right after
reading the teaching unit. Many e-learning platforms even allow in-situ feedback
texts to be added to the respective answer options, which then immediately indi-
cate to the students whether and why their selected answer option is incorrect or
correct. The author of this book is happy to provide lecturers such answer texts
for the questions included here upon request. From the answer statistics, which
can also be easily generated in many e-learning platforms, lecturers can then see
which aspects of the topic are already well understood in the student group and
which are not, such to then tailor the subsequent classroom session accordingly.
In addition, further of these multiple-choice questions can be used in this atten-
dance unit to further deepen the material. This may best be realized with the peer
instruction method: In this approach, a conceptual multiple-choice question is
projected in the classroom, and the students first answer individually with the aid
of an audience response system (“clicker system”, e.g., smartphone-based). The
response statistics, which are then also projected by the teacher, give the students
direct and anonymous feedback on how their own selected response fits into the
overall cohort. Afterwards, the students are asked to exchange themselves with
their peers around them in groups of two or three, with the task to convince them
of the correctness of their first chosen answer.
A second round of voting after a few minutes will then almost always produce
the correct result with a clear majority, simply because those who have had the cor-
rect answer in the first place also have better arguments and can understand and
eliminate any gaps in their peers’ understanding — much better than any lecturer
ever could. That way, the students take an active role in deepening their knowl-
edge, are stimulated and motivated, and are interactively involved in the learning
process on several levels . . . whereas the lecturer merely takes the role of a modera-
tor. With that, the method fulfills one of the core claims of its inventor, Prof. Eric
Mazur (Harvard): “good teaching is to help students learn.”
As attractive as this teaching method may seem at first — it stands and falls
with the quality of the questions asked; and even more with the quality of the given
answer options. If it is immediately obvious which answer is the correct one, the
method is at best entertaining, but not particularly instructive. If, by contrast, one
of the answer options represents the “most common misconception“, i.e., the most
common and typical initial misunderstanding that students often have, then this
can be specifically addressed and eliminated. This is exactly where this book comes
in. It aims to provide well thought-out and prepared material for all three of the
above-mentioned teaching phases (self-study, e-learning feedback, and in-person
consolidation unit).
Foreword to the 2nd edition VII

A great acknowledgement goes to M.Ed. Julia Windhausen, who was a master


student in the author’s team at Mainz University in 2021 and co-developed most of
the foresaid multiple choice questions together with the author during that time.
Further assistance in that task was provided by PD Dr. Wolfgang Schärtl, who is a
university lecturer in the author’s team and who also co-authored other De-Gruyter
textbooks on Physical Chemistry with him.

Mainz, spring 2023


Foreword to the 1st edition
Polymer science is a field that requires good knowledge about both chemistry and
physics. This is because it is chemistry that sets the specificity of each polymer, but
behind that, it is physics that sets the universality of the properties of all polymers.
The convergence of both lies the ground for a huge variety of applications of poly-
mers. In such, though, a user (=a customer on the market) usually doesn’t care
about the material itself, but instead, only about the function that it provides. Thus,
material designers (be it in industry or academia) are demanded to develop materi-
als that provide functions of interest, whereby customers, however, do not actually
care about the beauty of the material that does that, but instead, only about its util-
ity. It is therefore obligatory for material designers to translate the desired functions
(=the customers’ wishes) into measurable physical material parameters, and then
to further translate these parameters into chemical structures. This translation can
only be done on the basis of a sound understanding of structure–property relations
of polymers. And that is what this book is targeted at. Physical chemistry of poly-
mers intends to bridge the structure of polymers, as provided by chemistry, to their
properties, as captured by physics. Once this bridge is erected, it can be passed in
either direction, thereby allowing us to understand from what structural character-
istics a certain beneficial property comes from, or vice versa, to predict what struc-
ture will provide a property (and therewith a function) of interest.
A challenge in this endeavor is that at a first view, the concepts and treatments
in physical chemistry of polymers appear to be rather new and unknown to stu-
dents. It is therefore another goal of this book to familiarize its readers with these
approaches, and to demonstrate that they are actually not at all new, but instead,
quite related to concepts known from elementary physical chemistry. Hence, in this
book, many relations are drawn to such classical physical chemistry contents,
and the whole focus of this book is generally more on conceptual universality
rather than on detailed specificity. For the purpose of the latter, all illustrations in
this book are styled as lecture-like schematics. The content is ordered into six
chapters that arc from fundamental polymer physical-chemical principles to ac-
tual structure–property relations, including a touch on experimental approaches
to characterize both of the latter. On top of this six-chapter structuring lies a sec-
ondary structure that portions the content into 18 lesson units, which could be
conceived as 90-min lectures each. The author teaches those units in two sequen-
tial classes, Physical Chemistry of Polymers 1 on bachelor level (Lessons 1–9) and
Physical Chemistry of Polymers 2 on master level (Lessons 10–18). With a semester
length of commonly 12 weeks, this portioning gives ample time to not only cover
the respective content, but also to leave some flexibility for additional lessons
to be filled with in-depthment, questions and answers, and a practicing exam.
Alternatively, in a long semester of 14 weeks (as it is the case in the winter term in
Germany), the full content can be covered if four lessons are left out; for example,

https://doi.org/10.1515/9783110713268-203
X Foreword to the 1st edition

these may be Lessons 9, 12, 17, and 18 if there shall be no such big emphasis on
analytical characterization of polymer systems, or these may be Lessons 12, 15,
and 16 (plus one more) if there shall be no such a deep focus on rheology.
The year of appearance of the first edition of this book is the inspiring
Staudinger year 2020, the year of polymers. Based on this big anniversary of our
field, this book aims to converge chemistry and physics of polymers and to make
it understandable and appetizing to students at levels as early as possible.

Mainz, spring 2020


Contents
Foreword to the 2nd edition V

Foreword to the 1st edition IX

Lessons XV

Literature basis XVII

1 Introduction to polymer physical chemistry 1


1.1 Targets of this book 1
1.2 Definition of terms 3
1.3 Irregularity of polymers 7
1.3.1 Nonuniformity of monomer connection in a chain 7
1.3.2 Polydispersity in an ensemble of chains 9
1.3.2.1 Derivation of the Schulz–Flory distribution 13
1.3.2.2 Derivation of the Schulz–Zimm distribution 16
1.3.2.3 Characteristic averages of the Schulz–Flory and Schulz–Zimm
distribution 17
1.4 Properties of polymers in contrast to low molar mass
materials 19
1.5 Polymers as soft matter 20
1.6 History of polymer science 22

2 Ideal polymer chains 31


2.1 Coil conformation 32
2.1.1 Micro- and macroconformations of a polymer chain 32
2.1.2 Measures of size of a polymer coil 35
2.2 Simple chain models 37
2.2.1 The random chain (phantom chain, freely jointed chain) 37
2.2.2 The freely rotating chain 39
2.2.3 The chain with hindered rotation 40
2.2.4 The Kuhn model 41
2.2.5 The rotational isomeric states model 43
2.2.6 Energetic versus entropic influence on the shape of a
polymer 45
2.2.7 The persistence length 46
2.2.8 Summary 48
2.3 The Gaussian coil 51
2.4 The distribution of end-to-end distances 53
2.4.1 Random-walk statistics 53
XII Contents

2.5 The free energy of ideal chains 58


2.6 Deformation of ideal chains 59
2.6.1 Entropy elasticity 59
2.6.2 A scaling argument for the deformation of ideal chains 62
2.7 Self-similarity and fractal nature of polymers 64

3 Real polymer chains 71


3.1 Interaction potentials and excluded volume 72
3.2 Classification of solvents 77
3.3 Omnipresence of the Θ-state in polymer melts 79
3.4 The conformation of real chains 83
3.4.1 Coil expansion 83
3.4.2 Flory theory of a polymer in a good solvent 85
3.5 Deformation of real chains 89
3.6 Chain dynamics 96
3.6.1 Diffusion 96
3.6.2 The Rouse model 98
3.6.3 The Zimm model 100
3.6.4 Relaxation modes 102
3.6.5 Subdiffusion 104
3.6.6 Validity of the models 106

4 Polymer thermodynamics 111


4.1 The Flory–Huggins mean-field theory 111
4.1.1 The entropy of mixing 115
4.1.2 The enthalpy of mixing 118
4.1.3 The Flory–Huggins parameter as a function 121
4.1.4 Microscopic demixing 124
4.1.5 Solubility parameters 125
4.2 Phase diagrams 131
4.2.1 Equilibrium and stability 131
4.2.2 Construction of the phase diagram 137
4.2.3 Mechanisms of phase separation 143
4.2.3.1 Spinodal decomposition 143
4.2.3.2 Nucleation and growth 143
4.3 Osmotic pressure 148
4.3.1 Connection of the second virial coefficient, A2, to the
Flory–Huggins parameter, χ, and the excluded volume, ve 151

5 Mechanics and rheology of polymer systems 159


5.1 Fundamentals of rheology 159
5.1.1 Elementary cases of mechanical response 159
Contents XIII

5.1.2 Different types of deformation in rheology 161


5.1.3 The tensor form of Hooke’s law 163
5.2 Viscoelasticity 167
5.2.1 Elementary types of rheological experiments 168
5.2.1.1 The relaxation experiment 168
5.2.1.2 The creep test 168
5.2.1.3 The dynamic experiment 169
5.3 Complex moduli 170
5.4 Viscous flow 174
5.5 Methodology in rheology 180
5.5.1 Oscillatory shear rheology 180
5.5.2 Microrheology 181
5.5.2.1 Active microrheology 181
5.5.2.2 Passive microrheology 182
5.6 Principles of viscoelasticity 184
5.6.1 Viscoelastic fluids: the Maxwell model 184
5.6.2 Viscoelastic solids: the Kelvin–Voigt model 191
5.6.3 More complex approaches 193
5.7 Superposition principles 198
5.7.1 The Boltzmann superposition principle 199
5.7.2 The thermal activation of relaxation processes 201
5.7.3 Time–temperature superposition 202
5.8 Viscoelastic states of a polymer system 209
5.8.1 Qualitative discussion of the mechanical spectra 209
5.8.2 Quantitative discussion of the mechanical spectra 211
5.9 Rubber elasticity 217
5.9.1 Chemical thermodynamics of rubber elasticity 218
5.9.2 Statistical thermodynamics of rubber elasticity 220
5.9.3 Swelling of rubber networks 227
5.10 Terminal flow and reptation 232
5.10.1 The tube concept 233
5.10.2 Rouse relaxation and reptation 234
5.10.3 Reptation and diffusion 237
5.10.4 Constraint release 239

6 Scattering analysis of polymer systems 243


6.1 Basics of scattering 243
6.2 Scattering regimes 246
6.3 Structure and form factor 247
6.4 Light scattering 255
6.4.1 Static light scattering 255
6.4.2 Dynamic light scattering 264
XIV Contents

6.4.3 Light scattering on polymer gels 267


6.4.3.1 Polymer networks and gels 267
6.4.3.2 Static light scattering on gels 268
6.4.3.3 Dynamic light scattering on gels 270

7 States of polymer systems 279


7.1 Polymer solutions 279
7.1.1 Formation 279
7.1.2 Concentration regimes 280
7.1.3 Dilute solutions 282
7.1.3.1 Structure 282
7.1.3.2 Dynamics 283
7.1.4 Semi-dilute solutions 292
7.1.4.1 Peculiarity 292
7.1.4.2 Structure 292
7.1.4.3 Dynamics 296
7.2 Polymer networks and gels 304
7.2.1 Fundamentals 304
7.2.2 Gelation 305
7.2.2.1 One-dimensional bond percolation 307
7.2.2.2 Two-dimensional bond percolation 307
7.2.2.3 Mean-field model of gelation 308
7.2.2.4 The gel point 312
7.2.3 Structure of polymer networks and gels 315
7.2.4 Dynamics of polymer networks and gels 317
7.3 Glassy and crystalline polymers 321
7.3.1 The glass transition 321
7.3.1.1 Fundamentals 321
7.3.1.2 Conceptual grounds of the glass transition 324
7.3.1.3 Structure–property relations for Tg 329
7.3.1.4 Experimental assessment of Tg 329
7.3.2 Polymer crystallization 330
7.3.2.1 Fundamentals 330
7.3.2.2 Partial crystalline microstructures 330
7.3.2.3 Experimental assessment of the degree of crystallinity 335
7.3.2.4 Structure–property relations for Tm 336
7.3.3 Comparison of the glass and the crystallization transition 337

8 Closing remarks 343

Index 345
Lessons
LESSON 1: INTRODUCTION 1

LESSON 2: IDEAL CHAINS 31

LESSON 3: GAUSSIAN COILS AND BOLTZMANN SPRINGS 51

LESSON 4: REAL CHAINS 71

LESSON 5: FLORY EXPONENT 83

LESSON 6: CHAIN DYNAMICS 96

LESSON 7: FLORY–HUGGINS THEORY 111

LESSON 8: PHASE DIAGRAMS 131

LESSON 9: OSMOTIC PRESSURE 148

LESSON 10: RHEOLOGY 159

LESSON 11: VISCOELASTICITY 167

LESSON 12: PRACTICE AND THEORY OF RHEOLOGY 180

LESSON 13: SUPERPOSITION PRINCIPLES 198

LESSON 14: MECHANICAL SPECTRA 209

LESSON 15: RUBBER ELASTICITY 217

LESSON 16: REPTATION 232

LESSON 17: SCATTERING METHODS IN POLYMER SCIENCE 243

LESSON 18: LIGHT SCATTERING ON POLYMERS 255

LESSON 19: LIGHT SCATTERING ON POLYMER GELS 267


XVI Lessons

LESSON 20: DILUTE POLYMER SOLUTIONS 279

LESSON 21: SEMI-DILUTE POLYMER SOLUTIONS 292

LESSON 22: POLYMER NETWORKS AND GELS 304

LESSON 23: POLYMERS IN BULK SOLID STATES 321


Literature basis
Isaac Newton once pointed out his work to be outstanding as he could “stand on
the shoulders of giants” (Letter from Sir Isaac Newton to Robert Hooke; Historical
Society of Pennsylvania). In the same sense, this textbook is based on the following
seminal existing ones:

M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, New York 2003
H. G. Elias: Makromoleküle, Wiley VCH, Weinheim 1999 and 2001
B. Tieke: Makromolekulare Chemie, Wiley VCH, Weinheim 1997
B. Vollmert: Grundriss der Makromolekularen Chemie, Springer, Berlin Heidelberg 1962
J. S. Higgins, H. C. Benoît: Polymers and Neutron Scattering, Clarendon Press, Oxford,
1994

Among those, the book by Rubinstein and Colby has given particular inspiration for
the present one. In the author’s view, these colleagues’ seminal book has a physics-
based approach (as its title also says) targeted at grad students (as its backcover-text
says), which would correspond to readers at a PhD-student level in the European sys-
tem. As a complement to that, the present textbook provides a physical-chemistry-
centered viewpoint addressing both grad and undergrad students, which corresponds
to students also on bachelor and master levels in the European system.
This book is based on a script to the author’s lecture series “Physical Chemistry
of Polymers” at JGU Mainz. The writing of this script has been assisted by Dr. Willi
Schmolke in the years 2018 and 2019, who was a PhD student in the author’s lab
at Mainz then; special thanks go to Willi for this assistance with the book’s fun-
dament, along with further thanks to PD Dr. Wolfgang Schärtl for proofreading it.
An even deeper fundament has been laid by another person named “Willi”: the
author’s own respected teacher in the field of polymer science at TU Clausthal in
the years 2003–2008, Prof. Dr. Wilhelm Oppermann.

https://doi.org/10.1515/9783110713268-206
1 Introduction to polymer physical chemistry

LESSON 1: INTRODUCTION
As a starting point of our endeavor, the first lesson of this book will introduce you to the
fundamental terms and basics in the field of polymer science, along with its historical devel-
opment. You will learn what a polymer actually is, why these large molecules are interest-
ing, and what the fundamental differences are in their treatment compared to classical
materials that are built up of small molecules.

1.1 Targets of this book

The prime goal of polymer physical chemistry is to understand the relations be-
tween structure and properties of polymers. Once such understanding is achieved,
molecular parameters can be rationally connected to macroscopic behavior of poly-
mer-based matter. This connection bridges the fields of polymer chemistry and poly-
mer engineering, as shown in Figure 1, thereby allowing polymer materials to be
tailored by rational design. Even though these fields may first seem far apart and
separated (by an impassable water body in Figure 1), the discipline of polymer
physical chemistry arcs and connects them. (And in the end, by looking underneath
the water, we realize that they are actually connected inherently.)
To achieve this goal, along our way, we need to learn some fundamental ap-
proaches in the physical chemistry of soft condensed matter; we will also learn why
polymers are a prime example of such matter. First, we need to consider the multi-
body nature of each polymer chain as well as the myriad of different local micro-
scopic conformations that it can adopt. This requires us to limit our focus on
averages over all these individual states, which we obtain by suitable statistical
treatment as well as mean-field modeling. Second, as a direct result of such consid-
eration of the conformational statistics of polymer chains, we will see that pretty
much of all their properties are coupled to each other in the form of scaling laws.
Hence, scaling discussions will be a frequent means in this book. With this treat-
ment, we will recognize that polymers exhibit both specificity and universality.
As an example, the relation between the size and the mass of a polymer chain fol-
lows a power-law proportionality, independent of what polymer exactly we look at;
this is universality set by physics. The proportionality factor, however, depends on
the specific polymer at hand; this is specificity set by chemistry. As a further exam-
ple, note that all polymers exhibit a glass-transition temperature, Tg, at which they

https://doi.org/10.1515/9783110713268-001
2 1 Introduction to polymer physical chemistry

Figure 1: Polymer physical chemistry builds a bridge between polymer chemistry, which focuses on
the primary molecular-scale structure of polymer chains, and polymer engineering, which focuses
on the macroscopic properties of polymer materials. This connection allows for a rational design of
polymer-based matter to obtain tailored materials. To erect the bridge, three pillars are required:
single chain statistics, multichain interactions, and chain dynamics. These pillars are in the focus
of the second, third, and fourth chapters of this book. The final bridge on top of this basis is in the
focus of the fifth chapter, which sheds a particular focus on polymers’ most important kind of
properties: their mechanical ones.

change from a hard, glassy state to a soft, leather-like state upon heating (and vice
versa). They do so because of universal physics shared by all polymers1. However,

 The glass transition is a phenomenon actually not only seen on polymers; basically all matter
can show it. When cooling a liquid, in classical chemistry we commonly see it to solidify by crystal-
lization. If the cooling is done rapidly, however, there is not enough time for the molecules to ar-
range themselves on well-defined lattice positions, but instead, we then observed freezing of the
molecular dynamics only. The result is a glass: an amourphous structure (like in a fluid) with no
translational dynamics (like in a solid). In polymers (and colloids as well), this kind of solidification
is seen often (we might actually even say “always”), because the large size of polymers and col-
loids, and even more the chain connectivity of polymers, requires quite long times to undergo cry-
strallization. In polymers, it is not the chains as a whole that arrange themselves on a crystal
lattice, but the monomeric segmental units within the chains, which requires high regularity of
monomer interconnection (even proper stereo-regularity) to actually be able to do so, and long time
to manage to do so. Both is often not given, and so glassy solidification is actually much more fre-
quent upon cooling than crystallization for polymers. The point where such glassy solidification
(also named “vitrification”) happens depends on the ease of monomer-segmental rotatability along
the chain, because this is the primary process by which the monomer units are to be arranged on a
lattice. This ease of rotatability, in turn, depends on the local primary monomer structure, such as
the bulkiness and interactions of the chemical side groups along the chain backbone. This is what
sets the chemical specificity of the individual glass-transition temperature of each polymer, whereas
the general occurrence of a glass-transition phenomenon in polymers is physical universality.
1.2 Definition of terms 3

the exact value of each polymer’s Tg is determined by each polymer’s specific chain
chemistry.

1.2 Definition of terms

The term macromolecule is derived from the Greek macro, meaning “large”, and the
word molecule in its usual meaning.2 Hence, this term refers to large and heavy mol-
ecules, with sizes, r, in the range of some ten to hundred nanometers, and molar
masses, M, in the range of 104–107 g·mol–1, as shown in Figure 2. Delimited from
that, the term polymer, which is derived from the Greek poly and meros meaning
“many” and “parts”, refers to chain molecules constituted of a sequence of many
repeating units. These units, named monomers, are usually covalently jointed.3
Therefore, whereas the term macromolecule generally refers to any sort of large mol-
ecules, the term polymer specifically refers to those built of a repeating sequence of
monomer units. Hence, each polymer is a macromolecule, but not each macro-
molecule is a polymer.

Figure 2: Molar mass, M, and size range, r, of various atomic and molecular structures.

Somewhat doubtful in this context are biopolymers such as proteins and DNA. On
the one hand, one may argue that they are composed of a repeating sequence of
units, which are peptides in the case of proteins and nucleotides in the case of
DNA, justifying them to be termed polymers. On the other hand, despite this repeti-
tion of peptides or nucleotides, it is not a monotonic sequence of just one out of the
20 natural amino acids or just one out of the four different nucleobases that are
available in nature, but a diverse variety of sequences of them along the chain. This

 Note: the term molecule actually derives from the Latin molecula, meaning “small mass”; as
such, the word macromolecule would mean “large small mass”, which is sort of an oxymoron.
 A rather new class of self-assembled matter is supramolecular polymers; it is based on chains
that are noncovalently jointed.
4 1 Introduction to polymer physical chemistry

argument may then question the adequateness of terming these polymers. This is,
however, a pea-picking discussion. Following this line of thought, even random co-
polymers, which are actually well accepted to be termed polymers, should not be
named such. We therefore refrain from this argument. Instead, we point out that by
virtue of their ability to store information in their primary monomer sequence,
which in turn is a basis for their highly specific function and functionality, proteins
and DNA are nothing less than the link between matter and life, that is, between
chemistry and biology.
The International Union of Pure and Applied Chemistry (IUPAC) defines a poly-
meric molecule as follows:

A molecule of high relative molecular mass, the structure of which essentially comprises the
multiple repetition of units derived, actually or conceptually, from molecules of low relative
molecular mass. In many cases, [. . .], the addition or removal of one or a few of the units has
a negligible effect on the molecular properties.

This definition does not only apply to molecular properties but also to macro-
scopic properties, as can be recognized when considering the example of simple
hydrocarbons in Figure 3. Starting from methane, a family of alkanes and poly-
ethylenes is obtained by adding –CH2– repeating units. For the first 17 additions,

Figure 3: Boiling and melting points, Tbp and Tmp, of alkanes and polyethylenes with N methylene
(–CH2–) units. Whereas the boiling point increases with N and eventually becomes inexistent from
N = 17 on (because beyond that, such high temperatures would be necessary for boiling that the
molecules would break rather than boil), the melting point first increases with N but then
eventually levels off in a plateau. In this plateau, addition or removal of a few –CH2– units does not
change TM considerably. Picture redrawn from H. G. Elias: Makromoleküle, Bd. 1 – Chemische
Struktur und Synthesen (6. Ed.), Wiley VCH, 1999.
1.2 Definition of terms 5

the boiling point, Tbp, is raised sharply, until then, further addition leads to solid
compounds that decompose instead of boiling at high temperatures. An upward
trend is also observed for the melting point, Tmp, up to even the first about hun-
dred additions. Then, however, the picture changes. Further addition does no lon-
ger change the melting point; instead, it has reached a plateau, which denotes the
polymeric regime. In this regime, addition or removal of single or few –CH2– re-
peating units has practically no impact on the material properties such as Tmp.
The materials’ appearance in this range of high N is that of hard solids rather than
that of oily or even volatile liquids at low N.
From the above notion, we may conclude that polymers are macromolecules
composed of a repeating sequence of units; these units are sometimes termed mono-
mer units and sometimes termed structural units. Note the difference between these
terms: the term monomer unit usually refers to the chemical species from which the
polymer has been made, or more precisely, to the chemical sequence along the
main chain into which the actual monomer has turned into upon polymerization,
whereas the term structural unit refers to the smallest possible repetitive entity
along the chain. Often this is the same, but sometimes it is not. For example, con-
sider polyethylene again. In this polymer, the monomer is ethylene, H2C=CH2,
which turns into a monomer unit of –H2C–CH2– upon polymerization, whereas the
smallest possible repetitive unit, that is, the structural unit, is just the methylene
group, –CH2–. Similarly, in polyamide, [–HN–(CH2)x–NH–CO–(CH2)y–CO–]n, the
structural unit is –HN–(CH2)x–NH–CO–(CH2)y–CO–, whereas the monomer units
are –HN–(CH2)x–NH– and –CO–(CH2)y–CO–, resulting from the respective diamines
and diacids upon polymerization.

With the latter examples, we have touched upon the process of polymer formation, which is
generally named polymerization. This process can occur in many different ways and is a matter
of own independent textbooks and lectures. For the sake of our consideration, let us just briefly
and rather conceptually summarize its bottom line. In general, polymers are synthesized by the
consecutive reaction of monomers. These are molecules that carry chemical groups which can
form connections to each other, most typically either double bonds that can get opened to be-
come reactive radicals that may then attack other double bonds on other monomer molecules,
or pairs of mutually connectable functionalities such as nucleophilic plus electrophilic moieties,
for example, alcohols or amines plus carboxylic acid groups. These different kinds of polymer-
izable groups can then lead to polymerization via two different pathways, as sketched illustra-
tively in the following conceptual schematic, in which people symbolize the monomers and
form chains in two different ways of growth.
6 1 Introduction to polymer physical chemistry

In the case of double-bond-containing monomers, a small fraction of them can get attacked by a
chemical initiator (symbolized by the “Go” flag in our schematic) to become a radical species.
This activated species then quickly attaches to a second monomer, which forms a bond with it
and transfers its active radical nature to that attached partner, which in turn does the same
with a third monomer, and so forth. As a result, the monomers quickly attach to each other to
form a chain one by one. This mechanism is referred to as chain-growth polymerization, not
because its result is a polymer chain, but because it occurs as a chain reaction, like a domino
cascade. Its predominant realization is by radical polymerization, as already indicated in the
process description just given. In our schematic sketch, this mechanism is illustrated as people
serially joining hands, whereby it is always the terminal person in the chain that reaches out to
the next.
In the case of nucleophilic plus electrophilic functionalized bifunctional monomers, by con-
trast, these species mutually add to one another in steps, to first form dimers, then tetramers,
then octamers and so forth, referred to as step-growth polymerization. Its predominant realiza-
tion is by polyaddition or, if the monomer connection occurs under spin-off of a small-molecule
by-product (most typically water), by polycondensation. In our schematic sketch, this mechanism
is illustrated as people sequentially joining hands in larger and larger groups.
Chemical modification of the resulting polymers, such as attaching further functional entities
to them (like sites to bind molecular cargo or to make the polymers attach to surfaces or biolog-
ical tissue) or labeling the polymers with dyes or certain isotopes (most importantly deuterium)
can be achieved either at the monomer level before the polymerization, or at the polymer level
after the polymerization. In our schematic illustration, such modification could mean that we
equip some of the persons in the chain (either before or after their incorporation in the chain)
with some kind of functional belt, vest, or backpack or with a headlight; it could also mean that
one or some of our persons in the chain can be somehow distinguished from the others, for
example, by wearing a different shirt. Such modification may also occur in the form of three-
dimensional interconnection of the chains in space, referred to as cross-linking. In our sche-
matic, this would mean that two people on two different chains would tie together somehow,
thereby joining the chains.
1.3 Irregularity of polymers 7

Because polymeric molecules consist of many monomeric units (hundreds to thousands),


they can be conceived in a coarse-grained view, without specific accounting for their local mo-
lecular-scale atomic structure, but instead, simply viewed and sketched globally in the form of
curved or curled lines, as also indicated in our abovementioned two schematics; this is the way
how we will actually draw and conceive polymers almost all over this textbook.

1.3 Irregularity of polymers

Polymers aren’t regular. This is because they are made by statistical processes dur-
ing most polymerization reactions, leading to inherent nonuniformity even if the
basic chain-propagation reaction step proceeds perfectly. On top of that, if there is
a nonperfect course of that step, each “side product” formed during it will be incor-
porated into the chains and create irregular local constitutions. As a result, poly-
mers display different types of nonuniformity and irregularity.

1.3.1 Nonuniformity of monomer connection in a chain

Even along each single chain, there is structural irregularity if multiple ways of
monomer–monomer interconnection are possible. One type of such irregularity is
the three different possibilities of head-to-head versus head-to-tail versus tail-to-tail
interconnection of asymmetric monomers such as the broad class of vinyl com-
pounds, as shown in Figure 4.

Figure 4: Head-to-head versus head-to-tail versus tail-to-tail interconnection of asymmetric


monomers such as vinyl compounds.

A second type of irregularity resulting from the connection of such asymmetric


monomers is different stereochemistries along the chain, referred to as tacticity, as
shown in Figure 5. This property is of high relevance when polymers are crystal-
lized, which does only work with isotactic or syndiotactic chains. This is because
crystallization requires atoms or molecules to be able to arrange themselves regu-
larly on a lattice. In polymers, though, this is tough, because the molecules (=the
8 1 Introduction to polymer physical chemistry

Figure 5: Different types of tacticities in vinyl polymers.

structural units) are connected to one another in the form of chains. Only if this
connection is very regular, the building blocks can form a regular lattice.
In the case of diene-monomers, even more ways of interconnectability exist, as
shown in Figure 6.

Figure 6: Different possibilities of diene-monomer interconnection in the example of isoprene.


1.3 Irregularity of polymers 9

1.3.2 Polydispersity in an ensemble of chains

On top of the local irregularity of the monomer connectivity along each individual
chain, another relevant and very frequent kind of irregularity in polymer systems is
non-uniformity of the chain lenghs and molecular weights (unit: Da) or molar mas-
sess (unit: g mol–1) in ensembles of multiple chains, referred to as polydispersity.4
As a result of the statistical nature of most polymerization processes, the distribu-
tion of these quantities can be quite marked. A typical shape of a distribution
curve, in two common ways of representation, is shown in Figure 7.

Figure 7: Schematic of a typical shape of a molar mass (unit: g mol–1) or molecular weight (unit: Da)
distribution in a polymer sample, represented as the frequency of occurrence (typically a relative
mass percentage), F, plotted on the ordinate, of different molar masses in the sample, M, plotted
on the abscissa. The upper-right inset shows a different representation of this distribution, which
is the integral of all molar masses, plotted on the ordinate, up to a specific one considered on the
abscissa.

For practical purposes, it is more convenient and often sufficient to not deal with
the full distribution but just with some characteristic averages of it. These are the
number-average molar mass, Mn, the weight-average molar mass, Mw, and the
centrifugal-average molar mass, Mz, as represented by dashed lines in Figure 7.
These averages may be calculated from the so-called moments of the distribution
function, which are generally defined as follows:

 In a strict view, the word “polydispersity” is a pleonasm. The word “disperse” already refers to
an ensemble with multiple different elements (here: multiple different chain lengths or chain
masses, alternatively expressed as chain molar masses or molecular weights), so there is actually
no need for the prefix “poly”.
10 1 Introduction to polymer physical chemistry

In statistics, the moments of a distribution quantify its characteristic properties.


The first moment generally corresponds to the arithmetic average, the second mo-
ment to the width of the distribution, and the third moment to its skewness. In the
formula above, the argument g denotes whether number fractions, weight fractions,
or z-fractions are used to weigh the contributions of different Mi in the right-hand
side calculation.
With this definition, we get the three characteristic average molar masses as
follows:
ð1Þ P X
μn ðMÞ i Ni Mi
– Number average: Mn = ð0Þ = P = xi Mi
μn ðMÞ i Ni i

where xi is the mole fraction of species i.


P P
Ni Mi2 μn ðMÞ X
ð1Þ ð2Þ
μw ð M Þ Wi Mi
– Weight average: Mw = ð0Þ = P i
= Pi = ð1Þ = wi Mi
μw ðMÞ i Wi i Ni Mi μn ðMÞ i

where wi is the weight fraction of species i.


ð1Þ P P ð2Þ P
μz ðMÞ Zi Mi Wi Mi2 μw ðMÞ Ni Mi3 μðn3Þ ðMÞ
– z-average: Mz = ð0Þ = P i
= P i
= ð1Þ = Pi 2
= ð2Þ
μz ðMÞ i Zi i Wi M i μw ðMÞ i Ni Mi μn ðMÞ
The number average puts the same emphasis on short and long chains in the sam-
ple and simply averages over their weight or molar mass by accounting for their
number fraction in the sample. An experimental means to determine this average
value is by methods that probe phenomena that specifically depend on the number
of molecules in a sample; a prime example of such methods is those that probe col-
ligative properties such as the osmotic pressure of a polymer solution. In contrast
to that, the weight average puts more emphasis on the long and therefore heavy
chains in the sample, as it averages over them not by accounting for their number
fraction in the sample but for their contribution to the total sample weight. As a
result, the heavy chains in a sample receive a greater pronunciation in the averag-
ing. The most typical experimental means to probe this average value is by static
light scattering. The z-average pronounces the heavy chains in the sample even
1.3 Irregularity of polymers 11

more severely. The standard technique to probe this average value is by ultracen-
trifugal analysis.
We may illustrate the difference of the three types of averages of the molar
mass in an example. Consider a sample of 100 chains, 10 × 100,000 g·mol–1,
50 × 200,000 g·mol–1, 30 × 500,000 g·mol–1, and 10 × 1,000,000 g·mol–1. Accord-
ing to the above formulae, Mn = 360,000 g·mol–1, Mw = 545,000 g·mol–1, and Mz =
722,000 g·mol–1. Note that quite different averages are obtained even though they
are all related to the absolute same sample. This is because Mn balances light and
heavy chains equally, whereas Mw (and Mz even more) puts more emphasis on the
heavy chains. Based on what we have said above,5 we may illustrate the values of
Mn and Mw as follows. If the polydisperse sample with composition as listed above
would be depolymerized down to its monomer units, and if these units were then
reconnected with the premise of now giving the same number of chains as previ-
ously (100) but with a monodisperse, that means, an infinitesimally narrow molar-
mass distribution,6 the result would be a sample with 100 chains of 360,000 g·mol–1
each. This new sample would display the same osmotic properties as the original
polydisperse one. We may also do the reconnection differently, with the premise of
not obtaining 100 but only 66 new monodisperse chains. These chains would then all
have a molar mass of 545,000 g·mol–1, and this new monodisperse sample would dis-
play the same light-scattering properties as the original polydisperse one.
The preceding example shows us that we receive quite different average values
from the same sample, depending on how much emphasis we put on the heavy
chains in it. This phenomenon may act as a direct means to express the broadness
of the chain-length and molar-mass distribution in just one quantity, which is the
polydispersity index, defined as PDI = Mw/Mn. A PDI of 1 denotes a monodisperse
sample, whereas a PDI greater than 1 denotes a polydisperse sample.7 In our example
above, we have a PDI of 1.51. Commonly, when uncontrolled free-radical chain-growth
polymerization or step-growth processes such as polycondensation or polyaddition
are employed to synthesize a polymer, a PDI not smaller than 2 can be achieved. This
can be calculated from the distribution function that is obtained from such polymeri-
zation processes, which is the Schulz–Zimm distribution

ð1 − αÞK + 1 K P
W ð PÞ = P α (1:1)
K!

 In the latter footnote, it has been pointed out already that the word “polydisperse” is actually a
doubling of terms, and strictly speaking it would be sufficient to just use the word “disperse”.
 The word “monodisperse” is actually an oxymoron. If a distribution is infinitesimally narrow,
then it is no distribution. So, any word of kind “disperse” is actually misleading and incorrect here
in this sense.
 A sample with PDI just slightly above 1, typically up to 1.1, is commonly referred to as “narrowly
disperse”.
12 1 Introduction to polymer physical chemistry

or its variant the Schulz–Flory distribution (with K = 1)

W ðPÞ = PαP ð1 − αÞ2 ≈ PαP ln2 α (1:2)

Günter Victor Schulz (Figure ) was born on


October , , in Łódź, which was then part of
the Russian Empire, and moved to Berlin in
. He did his undergraduate studies in
Freiburg and Munich, working together with
Heinrich Wieland and Gustaf Mie, before
returning to Berlin for his graduate studies on
the thermodynamics of solvation equilibria in
colloidal solutions of proteins at the Kaiser
Wilhelm Institute of Physical Chemistry and
Electrochemistry. He obtained his PhD in ,
being examined by no other than Fritz Haber on
the subject of physical chemistry. He acquired
his habilitation in Freiburg under Hermann
Staudinger in  and then held an associate
professorship at Rostock from , before he Figure 8: Portrait of Günter V. Schulz. Image
was appointed as full professor at the JGU Mainz reprinted with permission from Macromol.
in , where he established a new Institute of Chem. Phys. 2005, 206(19), 1913–1914.
Physical Chemistry out of World War II ruins, Copyright 2005 Wiley-VCH.
which he then headed until his retirement. He
died in Mainz on February , .

In eqs. (1.1) and (1.2), W(P) denotes the frequency of occurrence of a certain degree
of polymerization P in the sample, in the form of its weight fraction in the sample,
W(P) = mP/m. The parameter α denotes the conversion of functional groups (not
monomers!) in the case of step-growth processes or the probability for chain propa-
gation relative to the sum of all probabilities (propagation + termination + transfer)
for chain-growth processes. The parameter K denotes the degree of coupling, that
is, how many individually growing chains form one macromolecule in the end,
which is K = 1 in the case of termination by disproportionation or K = 2 in the case
of termination by recombination. In the Schulz–Flory case, with K = 1, we get a
number-average degree of polymerization of Pn = 1/(1 – α) and a weight-average de-
gree of polymerization of Pw = (1 + α) / (1 – α), and hence, PDI = Pw/Pn = 1 + α,
which approaches 2 if the reaction approaches full conversion (α ! 1) in the case of
step-growth processes. In chain-growth reactions, by contrast, we will get a PDI
greater than 2, as in this case, we get a Schulz–Flory distribution for each momen-
tary degree of conversion during the reaction. This is because the parameter α in
the Schulz–Flory equation denotes the probability for chain propagation relative to
the sum of all probabilities (propagation + termination + transfer) in the case of a
chain-growth reaction, and therefore it depends on the monomer concentration,
1.3 Irregularity of polymers 13

which changes continuously during the process. As a result, the final overall distri-
bution will be an overlay of all these momentary Schulz–Flory distributions, yield-
ing a broad sum distribution with PDI greater than 2. Typically, uncontrolled free-
radical polymerization gives PDI = 3. . .10, whereas with the aid of controlling
chain-transfer agents, the PDI can be narrowed down to about 2.5 or so. Even nar-
rower distributions with PDI smaller than 2, however, can only be obtained by truly
living or quasi-living polymerization processes. In the best controlled case of a liv-
ing anionic polymerization, the distribution function obtained will be the Poisson
distribution

νP − 1 P expð − νÞ


W ð PÞ = (1:3)
ðP − 1Þ! ðν + 1Þ

In this equation, ν is the “kinetic chain length”, which is the degree of polymeriza-
tion built from one active, growing molecule before termination; often ν = Pn.
With this function, we obtain a PDI of

Pw ν
=1+ (1:4)
Pn ðν + 1Þ2

that approaches 1 with increasing ν = Pn. This is because in a living anionic poly-
merization, a constant number of chains grow steadily. Any initial imbalance of
the lengths of these chains will then get less and less significant if they grow lon-
ger. As an analogue, consider two kids that are 4 and 2 years old. At that stage,
their “age-PDI” is 1.11, but 70 year later, at an age of 74 and 72, their age-PDI is
just 1.0002. Nevertheless, though, note that even such a narrow distribution in a
polymer sample still has a finite width. For example, consider a sample with num-
ber-average degree of polymerization Pn = 500 and Pw/Pn = 1.04. In this sample,
still about 32% of the chains have degrees of polymerization of less than 400 or
more than 600.
The PDI is related to the general measure of the width of a distribution, which
is its variance, σ2, by the relation σ2 = Mn2 (PDI – 1). As a result, the PDI yields infor-
mation on σ/Mn, but not on σ itself. This means that if two samples have different
Mn, the sample with higher PDI may not have the larger σ!

1.3.2.1 Derivation of the Schulz–Flory distribution


Due to its fundamental importance and universal applicability to both chain-growth
(i.e., free-radical) and step-growth (i.e., polycondensation and polyaddition) pro-
cesses of polymer synthesis, the Schulz–Flory and Schulz–Zimm distribution func-
tions, eqs. (1.1) and (1.2), shall be derived in the following. This analytical derivation
14 1 Introduction to polymer physical chemistry

was originally reported by Schulz for the case of (free-radical) chain-growth polymeri-
zation.8 On top of that, Schulz and Flory were both working on the sister problem of
step-growth polycondensation,9 and a further contribution was made by Zimm even
later,10 but Schulz’ work on that matter was first held back by his supervisor Stau-
dinger, so that Flory’s report appeared earlier. Both approaches are based on apprais-
ing the probability for chain extension steps to occur.
The probability for chain growth in a free-radical polymerization, wgrowth, which
is given by the ratio of the rate of chain growth relative to the sum of the rates of
chain growth and termination, is equal to the probability for a certain functional
group to have reacted during a step-growth polymerization, α:
νgrowth
wgrowth = =α (1:5)
νgrowth + νtermination

In turn, the probability for chain termination in free-radical polymerization,


wtermination, and the probability for a certain functional group not to have reacted
in step-growth polymerization are
νtermination
wtermination = =1−α (1:6)
νgrowth + νtermination

From these two equations, we can derive an expression for the probability of forma-
tion of a chain with degree of polymerization P, which we name as wP . To form
such a chain, we need P growth events, the likelihood of which is wgrowthP, and we
need one termination event, the likelihood of which is wtermination:

NP
wP = wgrowth P · wtermination = αP ð1 − αÞ = (1:7)
N

This probability directly translates to the number fraction of chains with degree of
polymerization P in the sample, NP =N, just like the frequency of occurrence of a cer-
tain number of points in a dice game with many rolls directly translates to the prob-
ability of occurrence of that number in each single dice-rolling event. Note that in
the polymer literature, αP – 1 is sometimes used instead of αP in eq. (1.7). This is be-
cause the start of a polymerization reaction can be seen as R* + M or as R–M* + M,
where R is the residue of an initiator, M is the monomer, and * denotes some kind
of active species, typically a radical or an ion. However, as P is usually much
greater than 1, we may say that P ≈ P – 1.

 G. V. Schulz, Z. Phys. Chem. 1935, 30B(1), 379–398.


 P. J. Flory, J. Am. Chem. Soc. 1936, 58(10), 1877–1885; G. V. Schulz, Z. Phys. Chem. 1938, 182A(1),
127–144.
 B. H. Zimm, J. Chem. Phys. 1948, 16(12), 1099–1116.
1.3 Irregularity of polymers 15

The total number of macromolecules in the sample, N, can also be expressed as


the total number of chain terminations during the reaction, n(1 – α), because each
such event forms one chain:

N = nð1 − αÞ (1:8)

where n denotes the total number of all reaction steps. We may limit our focus to
cases of α close to 1, as only in that limit, long chains will be present. In the limit of
α ≈ 1, n is equal to the number of addition reactions, nα:

n ≈ nα (1:9)

From this, it follows that n is practically identical with the number of monomer
units in the N macromolecules. The total weight of these N macromolecules, m, in
turn, can be expressed as the number of these monomer units times the mass of
each unit:

m = n · mmonomer (1:10)

Note that m is not the total weight of the sample, but only that of the N macromole-
cules in it, without including residual unreacted monomer or other components.
Analogously, the total weight of all macromolecules with the degree of poly-
merization P is given by

mP = NP · mmonomer · P (1:11)

With these two equations, along with eqs. (1.7) and (1.8), we can formulate an ex-
pression for the weight fraction of chains with degree of polymerization P

mP NP · mmonomer · P NP · P
= = = P · αP ð1 − αÞ2 (1:12a)
m n · mmonomer n

This is the Schulz–Flory distribution. In the polymer literature, a variant form is


often found as well:
mP
= P · αP ln2 α (1:12b)
m
P
This is because Schulz’ original derivation leads to ðmP =mÞ = P · αP ð1= PαP Þ. The
P P
sum in the denominator is a series of type Pα = α + 2α2 + 3α3 + 4α4 + · · · . As
2
α < 1, this series converges to 1=ð1 − αÞ , so that eq. (1.12a) is recovered. Alterna-
tively, if the derivation is conducted in a continuous fashion with the integral
Ð P P P 
Pα dP instead of the sum Pα , we obtain 1 ln2 α instead of 1=ð1 − αÞ2 ,11 thereby
leading to eq. (1.12b).

 Also note that ln α = ln (1 – (1 – α)), which may be expanded into a Taylor series with a first
term of –(1 – α), and so we again get ln2α ≈ (1 – α)2.
16 1 Introduction to polymer physical chemistry

1.3.2.2 Derivation of the Schulz–Zimm distribution


Let us now examine how such a distribution looks like for the case of free-radical
polymerization with chain termination by recombination. There are P/2 ways for two
macroradicals to recombine and form a chain with degree of polymerization P: 1 +
(P – 1); 2 + (P – 2); 3 + (P – 3);. . . . . . . . .; P/2 + P/2.12 The probability of two macro-
radicals with degrees of polymerization X and Y to be present at the same time,
which is a prerequisite to meet and combine with one another, is

wX + Y = 2 · wX wY (1:13)

The factor of 2 in eq. (1.13) accounts for the two possible combinations X + Y and
Y + X. As a result of each such combination, we can obtain a macromolecule with
degree of polymerization P = X + Y. Rearrangement of that term yields Y = P – X, so
that with eq. (1.7), we can formulate the likelihood of two macroradicals with de-
grees of polymerization X and Y = P – X to be present as follows:

wX = αX ð1 − αÞ (1:14a)

wP − X = αP − X ð1 − αÞ (1:14b)

This can be inserted into eq. (1.13) to yield

wX + ðP − XÞ = wP = 2 · αX ð1 − αÞ · αP − X ð1 − αÞ = 2 · αP ð1 − αÞ2 (1:15)

When we multiply this equation with P/2, we generate an expression that approxi-
mates the number fraction of macroradicals that can yield a chain with degree of
polymerization P upon combination:

NP
= P · αP ð1 − αÞ2 (1:16)
N

Note that in the literature, αP – 2 is sometimes used instead of αP.


From that, we directly get an expression for the number of chains with degree
of polymerization P that results from recombination of these macroradicals:
1
NP = · N · P · αP ð1 − αÞ2 (1:17)
2

The factor ½ in the latter equation is necessary because recombination of two mac-
roradicals yields only one polymer chain.
By inserting eq. (1.8) into the latter one, we generate

12 This is basically Gauss’ math class puzzle, in which he was asked to calculate the sum of all num-
bers from 1 to 100. Gauss did so by a clever approach: (100 + 1) + (99 + 2) + (98 + 3) +. . .+
P
(51 + 50) + (50 + 51), which is 50 × 101. In general: nk = 1 k = nðn2+ 1Þ.
1.3 Irregularity of polymers 17

1
NP = · n · P · αP ð1 − αÞ3 (1:18)
2

Analogous to eq. (1.12), we can derive an expression for the weight fraction of
chains with degree of polymerization P:
mP 1 2 P
= · P · α ð1 − αÞ3 (1:19)
m 2

In general:

ð1 − αÞK + 1 K P
W ð PÞ = P ·α (1:20)
K!
where K is the degree of coupling. Two boundary cases can be distinguished: If
K = 1, the reaction is only terminated by disproportionation, whereas K = 2 accounts
for termination mostly by recombination.

1.3.2.3 Characteristic averages of the Schulz–Flory and Schulz–Zimm distribution


From eq. (1.7), in its variant with αP – 1 rather than αP, we can derive NP = N · αP − 1
ð1 − αÞ, from which we may further derive the number-average degree of polymeriza-
tion, Pn :
P∞ P∞ P−1
P∞ P−1
P = 1 P · N ðPÞ P=1 P · N · α ð1 − αÞ P=1 P · α ð1 − α Þ
Pn = P ∞ = P ∞ −
= P∞ −
(1:21a)
P=1 N ð P Þ P=1 N · α P 1 ð1 − α Þ P=1 α P 1 ð1 − αÞ

where the numerator is the first moment of N(P) and the denominator is the zeroth
moment of N(P).
By factoring out (1 – α), we can write
1
1 − α 1α0 + 2α1 + 3α2 + 4α3 +    ð1 − αÞ2 1
Pn = · = = (1:21b)
1−α α0 + α1 + α2 + α3 +    1
1−α
1−α

Analogously, from the distribution of weight fractions, mP =m = P · αP − 1 ð1 − αÞ2 , eq.


(1.12a), we can derive the weight-average degree of polymerization, Pw :
P∞ 2 P∞ 2 P − 1
P · N ð PÞ P · α ð1 − α Þ
Pw = PP∞= 1 = PP∞= 1 P − 1 ð1 − αÞ
(1:22a)
P=1 P · N ð P Þ P=1 P · α

where the numerator is the second moment of N(P) and the denominator is the first
moment of N(P).
Again we can factor out (1 – α) to generate

1 − α 1α0 + 4α1 + 9α2 + 16α3 +    1+α


Pw = · = (1:22b)
1 − α 1α0 + 2α1 + 3α2 + 4α3 +    1−α
18 1 Introduction to polymer physical chemistry

The PDI, which is the ratio of weight and number-average degree of polymerization,
then turns out to be
Pw α!1
=1+α ) 2
Pn

The derivation for the case of chain termination via recombination, that is, K = 2, is
analogous to the latter one. It yields
2
Pn = (1:23)
1−α

and
2+α
Pw = (1:24)
1−α

From this, a PDI of 1.5 follows in the limit α ! 1

Pw 2 + α α!1
= ) 1.5 (1:25)
Pn 2

We can also calculate the maximum of the Schulz–Flory distribution. We do so for


the case of chain termination via disproportionation (K = 1):
h i
∂ ð1 − αÞ2 · P · αP !
= ð1 − αÞ2 · P · αP ln α + αP ð1 − αÞ2 = 0
∂P
, ð1 − αÞ2 · αPmax · ðPmax ln α + 1Þ = 0
, Pmax ln α + 1 = 0
−1 1
, Pmax = ≈ = Pn (1:26)
ln α 1 − α

Figure 9: Graphical representation of the number fraction, N(P), and the weight fraction, W(P), of
each degree of polymerization P in a polymer sample according to the Schulz–Flory distribution.
1.4 Properties of polymers in contrast to low molar mass materials 19

We see that the maximum of the distribution is equal to the number average Pn . The
same is true for the case of chain termination via recombination (K = 2) (Figure 9):
2
Pmax = = Pn (1:27)
1−α

1.4 Properties of polymers in contrast to low molar mass


materials
Due to their macromolecular, usually ill-defined structure, polymers exhibit proper-
ties that are fundamentally different from their low molar mass counterparts. A
comparison of a few such different properties is compiled in Table 1.

Table 1: Comparison of some properties of low molar mass versus polymeric materials.

Low molar mass materials Polymer materials

Composition Uniform Polydisperse and irregular


Solubility Finite solubility, no swelling “All or nothing”, often with swelling
Melt behavior Low viscous Highly viscous or even viscoelastic
Purification Distillation, crystallization Precipitation

First, in contrast to low molar mass materials that have a uniform and well-defined
molecular constitution (often even in view of their stereochemistry), polymers have
a polydisperse and irregular composition. As detailed in Section 1.3, this is because
they are commonly synthesized in statistical chain-growth or step-growth pro-
cesses, leading to nonuniform chain lengths in a polymer sample, and also leading
to irregularities along each individual chain, such as head-to-head versus head-to-
tail versus tail-to-tail monomer–monomer connections.
Second, low molar mass compounds can usually be dissolved up to a certain
solubility threshold, which is high or low depending on the chemical similarity of
the solvent and the compound. This is fundamentally different for polymers: their
solubility often practically exhibits an “all or nothing behavior”, meaning that
when a solvent is good it can dissolve virtually any amount of a polymer (at least in
the practically relevant concentration range), whereas if a solvent is not good it can
hardly dissolve any of the polymer. This is because usually, mixing a compound A
(the solute) with a compound B (the solvent) is energetically unfavorable (as A–A
and B–B contacts are trivially more similar and therefore more favorable than A–B
contacts)13 but entropically favorable (as entropy generally likes mixing, as that

 “Birds of a feather flock together.”


20 1 Introduction to polymer physical chemistry

creates disorder). If the compound A has a low molar mass, there’s many molecules
of A in a given mass of this substance, and hence, the mixing entropy is consider-
ably high, thereby compensating the commonly unfavorable mixing energy and
leading to mixing up to a certain threshold concentration at which the solution is
saturated. By contrast, if the compound A is a polymer, a given mass amount of it
contains just few molecules; hence, the entropy of mixing is so little that it cannot
compensate the usually unfavorable energy of mixing, thereby prohibiting dissolu-
tion more or less completely. Only if the polymer A and the solvent B are chemically
very similar, which is a situation referred to as “good solvent state” (or, even better,
“athermal solvent state”), the just little entropy of mixing is still powerful enough
to compensate the then also just minorly unfavorable energy of mixing. In this
case, the polymer will dissolve in almost any quantities, only limited by practical
impairment such as overly high viscosities at high concentrations.
Third, tying in to the preceding sentence, low molar mass molecules usually
show low viscosities in the melt and flow like Newtonian fluids (with exemptions
such as glycerol or related compounds that have strong mutual intermolecular as-
sociative forces causing high viscosities). This is because they are small and can
relax (i.e., change their positions relative to each other) comparably fast, without
need for input of extraordinarily large additional energies of activation. Polymers,
by contrast, are large molecules that need great energies of activation for relaxation
and flow, corresponding to high viscosities. Especially long-chain molecules can
also undergo mechanical entanglement with one another, which imparts further
hindrance on their motion. The result is a rich mechanical behavior that is very dif-
ferent from that of small molecules, ranging from elastic snap over viscoelastic
creep and relaxation to viscous flow. Whereas the macromolecular chain-like shape
of polymers causes this to hold true for all polymers in a universal fashion, the
chemical specificity of a given polymer determines what type of mechanical re-
sponse will be observed at what temperature and on which timescale exactly.
Fourth, polymers do not exist in the vapor phase, because they decompose be-
fore boiling (see Figure 3). Therefore, standard purification procedures such as distil-
lation cannot be applied for polymers. Fortunately, however, tying in to our second
point above, their “all or nothing” solubility allows polymers to be precipitated from
a solution by adding an excess of a nonsolvent, thereby keeping low molar mass im-
purities dissolved while the polymer separates from the solution and can be har-
vested just by filtration.

1.5 Polymers as soft matter

Polymers are a prime example of a material class named soft matter. This term
refers to materials that (i) consist of building blocks with sizes in the colloidal
domain, 10–1000 nm, and that (ii) have interaction energies in the range of
1.5 Polymers as soft matter 21

10–100 RT (as a benchmark to that, note that RT at room temperature is about


8 J mol–1 K–1 · 300 K ≈ 2.5 kJ mol–1).14 To put this into context, let us consider that a
covalent C–C bond has a dissociation energy of 350 kJ mol–1, which is more than 100
RT at room temperature; this means that it requires a huge amount of external energy
to be broken. A transient OH···H interaction, by contrast, has a dissociation energy of
just 18 kJ mol–1, and can therefore be broken with just little additional energy. To-
gether, the large size and weak interactions of the building blocks in soft matter make
it soft. We can appraise that by estimating the cohesive energy density, e = E/r³, where
E is the interaction energy and r is the distance between the building blocks. For hard
matter, we have E in the range of 10–18 J and r in the range of 0.1 nm, which results in
values of e around 1012 Nm–2; this is a typical elastic modulus for materials such as
diamond. For soft matter, in contrast, we have E in the range of 10–20 J and r in the
range of 10–100 nm, which results in values of e around 104–107 Nm–2. This is a typical
elastic modulus for a polymeric or colloidal melt or gel. The weak interaction energy
also gives rise to another characteristic property of soft matter: its ability to be assem-
bled, disassembled, and reassembled in a dynamic manner, thereby displaying a rich
dynamic and stimuli-sensitive phase behavior. Due to that and due to their colloidal-
scale size, these materials (iii) have relaxation times, which are the times needed for
their building blocks to move a distance of their own size, in the practically relevant
range of milliseconds to seconds. As a result, soft matter materials often show rich vis-
coelasticity, depending on the timescale of experimentation and temperature.
Following this definition, polymers can be classified as a prime example of soft
matter. First, the shape of a polymer chain is, in most cases, that of a random coil
with a couple of 10 nm in diameter. This puts polymers right in the colloidal domain.
Second, due to their colloidal-scale size, the polymeric coils exhibit relaxation times
in the range of a couple milliseconds to seconds.15 Third, multiple polymer chains
can interact with each other usually via (random-fluctuating, induced, or permanent)
dipolar interactions to form higher-order assemblies. These interactions have ener-
gies in the range of a couple 10 RT. Consequently, polymeric assemblies are stable at

 A little footnote in this context: the value of RT at room temperature should be in the mind of
each physical chemist. This is because it is the fundamental benchmark of the energy landscape,
which is a land that physical chemists often have to pass. As a comparison, consider the following
analogy: an information like “the building over there is 10 m tall” is only valuable for you if you
know what 1 m is. If you do not know that, then you have no clue whether 10 m is tall or short. The
same applies to energies, with which you need to argue over and over again in physical chemistry.
You can only appraise whether a 350 kJ mol–1 covalent C–C bond or a 20 kJ mol–1 hydrogen bond-
ing interaction is strong or weak if you know the baseline of RT at room temperature.
 The exact time depends on the architecture and the environment of the polymer chain. A short
chain will relax faster than a long one. It will also relax faster when it is surrounded only by solvent
molecules, as it would be the case in a dilute polymer solution, than when it is in direct contact
with other polymer chains that constrain its movement, as it would be the case in a polymer melt.
22 1 Introduction to polymer physical chemistry

room temperature, but can be restructured with relative ease, which is one out of two
reasons why nature often uses polymeric structures as building blocks for dynamic
complex architectures. (The second reason will be discussed in the context of the
Flory–Huggins theory in Chapter 4.) Aside from this fundamental role in nature, the
soft-matter characteristics of polymers have made tremendous impact on science.
The basic physics of ordering phenomena is the same for colloidal-scale soft matter
and for classical atomic- or small-molecule-scale hard matter, but soft matter, due to
the much larger size and slower motion of its building blocks compared to that of
atoms and small molecules, can be powerfully observed and studied. This notion led
Pierre-Gilles de Gennes to use polymers as fundamental studying objects in con-
densed-matter physics, which is much more elementary than “only” studying them
for their vast practical applicability. This groundshifting inspiration was honored
with the Physics Nobel Prize in 1991.

Pierre-Gilles de Gennes (Figure ) was born on


October , , in Paris. He was home-
schooled to the age of . By the age of , he
had adopted adult reading habits and was
visiting museums. Later on, he studied at the
École Normale Supérieure, from which he majored
in . From  to , he was a research
engineer at the Atomic Energy Center (Saclay), the
heart of French nuclear research at that time, and
obtained his PhD in  from the University of
Paris. After postdoctoral research in Berkeley in
 and a -month stint in the French Navy, he
Figure 10: Portrait of Pierre-Gilles de Gennes.
became an assistant professor in Orsay in .
Image reproduced with permission from
Later, in , he became full professor at the
Nature 2007, 448(7150), 149. Copyright 2007
Collège de France. From  to , he was the
Springer Nature.
director of the Ecole de Physique et Chimie. After
receiving the Physics Nobel Prize in , de
Gennes decided to give talks on science, innovation,
and common sense to high school students. He
died in Orsay on May , .

1.6 History of polymer science

For centuries, natural polymers have been employed without any knowledge about
their macromolecular nature. Table 2 shows a brief history of mankind’s use of
them.
1.6 History of polymer science 23

Table 2: Use of natural polymers in human history.

Time Material Discovery

 BC Cotton Mexico


 BC Silk China
 BC Bitumen Orient
 AD Rubber Mexico (Mayas)
 AD Guttapercha (trans-,-polyisoprene)
 AD Vulcanized rubber (cis-,-polyisoprene) Goodyear

Gutta “rubber” and percha “tree”.


1

In this list, Goodyear’s vulcanization of natural rubber, which denotes the


turnover of viscous natural caoutchouc into an elastic rubber by crosslinkage of its
polyisoprene chains through sulfur bridges, is a special milestone, because it is the
first industrial-scale modification of a natural polymeric material, and therefore sig-
nifies the advent of a chemical industry in this field.16 From thereon, chemists mod-
ified many biopolymers, but still had no knowledge about their structure or their
macromolecular nature. It was believed at that time that covalently jointed struc-
tures with molar masses of several thousand grams per mole simply could not exist.
Instead, a common belief was that polymers were colloidal structures, built up by
small molecular entities aggregated by noncovalent forces. The reason for this no-
tion might have been because organic chemistry was successful in assessing small
molecule structures to that time, whereas physical chemistry was successful in as-
sessing their intermolecular forces, so that scientists were prone to apply both poly-
mer-type matters as well. The high molar masses of these materials that were
measured in laboratories were therefore viewed to be that of the colloidal aggre-
gates, and hence, sort of artifacts. Table 3 displays the hypothetic and actual molec-
ular structures of caoutchouc and cellulose as they were perceived before 1920 and
as we know them today.
In 1920, Hermann Staudinger was able to prove that polymers, in fact, have ac-
tual rather than artifactual high molar masses. He did this by chemically modifying

 The turnover of natural caoutchouc to rubber has in fact been discovered much earlier, by Ama-
zon natives, who extrachted the sap of the rubber tree and smeared it onto their feet, which then
gave rubbery boots due to crosslinking of the polyisoprene chains by oxygen from air. A lot later,
though, Charles Goodyear achieved the same effect in a better and more powerful way through the
use of sulfur. This was in fact achieved by accident: in 1839, Goodyear (who was a self-trained
chemist) dropped a caoutchouc–sulfur mixture onto a hotplate, thereby obtaining a dry, durable,
elastic substance. Despite that luck and the tremendous impact it had on various industries (such
as the automotive business), Goodyear was never successful in commerce. He was even sentenced
to prison several times, as he could not pay his debt. There is a saying after which a newspaper
once wrote about him: “If you see a man in shoes, coat, and a hat of rubber, but without a cent in his
pocket, then you are looking at Charles Goodyear.”
24 1 Introduction to polymer physical chemistry

Table 3: “Colloidal” and actual structure of caotchouc and cellulose.

known polymers, as shown in Figure 11, with the premise that such chemical change
in their “colloidal” structures should severely alter their aggregation, just in the same
way that a colloidal fatty acid micelle breaks up upon esterification. What Staudinger
found instead was staggering: despite drastic changes in many other properties, the
high molar masses were retained in all the compounds upon such modification
(which has therefore been referred to as “polymer analogous modification” in the fol-
lowing). This finding strongly supports Staudinger’s hypothesis that polymers are not
associated by physical interactions only, but instead, covalently jointed chain mole-
cules with an inherently high molar mass. This mass doesn’t change upon polymer
analogous modification, whereas other properties do, as the modification alters the
side groups of the polymer, but not the length of its chain backbone.

Figure 11: Hermann Staudinger’s polymer analogous reactions that proved a high molar mass to be
an actual rather than just artificial polymeric property.

Staudinger expressed his findings rather carefully first:

If one is willing to imagine the formation and constitution of such high-molecular compounds,
one may assume that primarily, a combination of unsaturated molecules has taken place, sim-
ilar to the formation of four- or six-membered rings, but that for some reason, potentially a
1.6 History of polymer science 25

steric one, the four- or six-ring closure didn’t happen, and now many, perhaps hundreds of
molecules are jointed.17

Staudinger’s notion received immediate and harsh criticism by the scientific com-
munity of the early twentieth century, such as the following, which was expressed
by Heinrich Wieland, his predecessor at the University of Freiburg and Nobel Prize
laureate for the structural characterization of bile acids:

Dear colleague, refrain from the imagination of large molecules; organic molecules with a mo-
lecular weight above 5000 do not exist. Purify your products, such as the caoutchouc, and
then they will crystallize and turn out to be low-molecular compounds.18

But Staudinger proved persistent, and, as his hypothesis held true, grew increas-
ingly self-confident later, when he said:

Even though the formation of millions of compounds whose molecules are built of hundred or
several hundreds of atoms is possible, this realm of low-molecular organic chemistry is just a
pre-stage to actual organic chemistry, namely chemistry of compounds that are of fundamen-
tal importance for processes in life. This is because the latter compounds contain molecules
that are not built from some hundreds, but from thousands, ten-thousands, and perhaps even
millions of atoms.19

Staudinger eventually received the Nobel Prize for his work in 1953. By then, the
field of macromolecular chemistry that he founded had advanced impressively, and
also his above statement about polymers as the “true organic molecules” and back-
bones of life held true, as in the same year, Watson and Crick published the struc-
ture of DNA.

 The original German wording is: “Will man sich eine Vorstellung über die Bildung und Konstitu-
tion solcher hochmolekularen Stoffe machen, so kann man annehmen, daß primär eine Vereinigung
von ungesättigten Molekülen eingetreten ist, ähnlich einer Bildung von Vier- und Sechsringen, daß
aber aus irgendeinem, evtl. sterischen, Grunde der Vier- oder Sechsringschluß nicht stattfand, und
nun zahlreiche, evtl. hunderte von Molekülen sich zusammenlagern.”
 The original German wording is: “Lieber Herr Kollege, lassen Sie doch die Vorstellung mit den
großen Molekülen, organische Moleküle mit einem Molekulargewicht über 5000 gibt es nicht. Rei-
nigen Sie Ihre Produkte, wie z.B. den Kautschuk, dann werden diese kristallisieren und sich als
niedermolekulare Stoffe erweisen.”
 The original German wording is: “Wenn auch die Darstellung von Millionen von Stoffen möglich
ist, deren Moleküle aus Hundert oder einigen Hundert Atomen aufgebaut sind, so bliebe dieses Ge-
biet der niedermolekularen organischen Chemie trotzdem nur die Vorstufe der eigentlichen organi-
schen Chemie, nämlich der Chemie der Stoffe, die für die Lebensprozesse von grundlegender
Bedeutung sind. Denn in letzteren Verbindungen liegen Moleküle vor, die nicht aus einigen Hundert,
sondern aus Tausenden, Zehntausenden und vielleicht sogar Millionen von Atomen bestehen.”
26 1 Introduction to polymer physical chemistry

Hermann Staudinger (Figure ) was born on


March , , in Worms. His father, a leading
figure in the German union movement, wanted
his son to enter a hands-on profession, so Her-
mann became an apprentice in carpentry first.
After his training, he went to Halle to study
chemistry and obtained his PhD there. He then
held professorships in Karlsruhe, Zurich, and fi-
nally at Freiburg, where he conducted his Nobel
Prize research. He died in Freiburg on September
, .

Figure 12: Portrait of Hermann Staudinger.


Image reproduced with permission from
ETH-Bibliothek Zürich, Bildarchiv.

In the 40 years following Staudinger’s groundbreaking experiments (1920–1960),


the core concepts of polymer physics were developed:
– Kuhn model for polymer chain renormalization (Chapter 2)
– Flory theory of coil conformation in a good solvent (Chapter 3)
– Rouse and Zimm models of polymer dynamics (Chapter 3)
– Flory–Huggins mean-field theory of polymer thermodynamics (Chapter 4)
– Kuhn, Grün, Guth, Mark: statistical theory of rubber elasticity (Chapter 5)

Between 1960 and 1980, these concepts were further developed and refined to move
from the description of single chains to chain assemblies, culminating in the Nobel
Prize for Pierre-Gilles de Gennes in 1991 for discovering that methods developed for
studying order phenomena in simple systems can be generalized to more complex
forms of matter, in particular, to liquid crystals and polymers. We just name two
bullet points to summarize these milestones:
– Doi, Edwards, De Gennes: tube concept and reptation theory (Chapter 5)
– De Gennes: “soft matter physics”

Current topics of polymer research (with a special focus on the physical chemistry
branch of it) have introduced dynamics and adaptivity into polymer systems by
actively incorporating or making use of transient bonding. This leads to self-
assembly processes that create complex architectures potentially soon rivaling
those in nature; it also leads to sensitivity, responsiveness, and adaptivity of these
complex systems and materials. To freely quote Bruno Vollmert’s book Grundriss
der Makromolekularen Chemie: “Polymers are the highest level of complexity in
Questions to Lesson Unit 1 27

chemistry, and the lowest level of complexity in biology.” Current polymer research
focuses right on the bridge between these realms. Other modern topics include:
– Supramolecular polymers (J. M. Lehn, E. W. Meijer, T. Aida, M. A. Cohen-Stuart,
and many others)
– Conductive polymers and polymers for energy conversion and storage
– Biopolymer science and biomimetics
– Large-scale theoretical modeling and simulation

Aside and along with this academic development, industrial polymers have turned
mankind’s twentieth century into the “polymer age”.20 This vast and versatile use
of polymers throughout decades, however, has posed environmental challenges
that modern polymer research must address, in the form of reversing the tremen-
dous amounts of polymer-based waste that has been and still is being produced by
mankind. Modern research, both industrial and academic, therefore must have a
special focus on degradable polymeric materials. On top of that, it will be a socio-
logical challenge for us to learn about the great value of polymer-based everyday
life products, given their remarkable properties (mostly their strong mechanics
compared to their light weight), such as to not only use them once but multiple
times before discarding them.

Questions to Lesson Unit 1

(1) Which statement is correct?


a. Each polymer is a macromolecule.
b. Each macromolecule is a polymer.
c. Polymer and macromolecule are synonym terms.
d. Polymer and macromolecule are different terms that do not overlap in
meaning.

(2) Which of the following biological macromolecules is a polymer?


a. DNA
b. Proteins
c. Lignin
d. Polysaccharides

 Just as much longer before, mankind’s main materials of use led us to entitle former ages as the
stone age, the bronze age, and the iron age, and just as our world today is certainly justified to be
termed to be in the silicon age.
28 1 Introduction to polymer physical chemistry

(3) Given are two samples A and B with ten chains each. The molar mass distribution
is listed in this table:

Molar mass (g mol–) , , , , ,


Number of chains in sample A     
Number of chains in sample B –    –

Which statement regarding the number average Mn and weight average Mw is


correct?
a. Mw is identical for both samples, since both samples have identical Mn.
b. Mn is identical but Mw is different, because chains with higher molar mass
are weighted more.
c. Mn is identical and also Mw because sample A has a chain with higher
molar mass of 500,000 g mol−1 but also a chain with lowest molar mass
100,000 g mol−1, which practically cancels out.
d. Mn is identical and also Mw since both samples are symmetrically distributed.

(4) Complete the following sentence: Soft matter________


a. consists of building blocks smaller than a few nanometers, which can be
easily displaced against each other due to their small size.
b. has a cohesive energy density that is five to eight orders of magnitude
lower than that of conventional hard matter.
c. consists of building blocks of a few nanometers in size that are linked by
strong covalent bonds but can be displaced against each other due to the
small number of bonds.
d. consists of building blocks that are linked together via long-range interac-
tions, that is, based on a soft repulsive potential (as opposed to a hard-
sphere potential).

(5) On which experimental observation was Staudinger’s description of polymers


based?
a. Polymers are aggregates of small molecules in the colloidal domain; Stau-
dinger made that visible with the (at his time newly invented) method of
electron microscopy.
b. Polymers are aggregates of small molecules in the colloidal domain; upon
chemical modification, these aggregates change their size due to a change
in interactions.
c. Polymers consist of covalently bonded monomers; chemical modification
usually affects their side groups only, but not their connectivity. So, the chain
length is retained.
Questions to Lesson Unit 1 29

d. Polymers consist of covalently bonded monomers; chemical modification


usually changes their connectivity, either by chain breakage or by chain ex-
tension/linking/branching. Hence, the chain length is changed.

(6) Consider four samples with identical number-average molar mass, Mn, but dif-
ferent distributions:

From the following possibilities, select the polymerization-type assignment that


matches the molar mass distributions from left to right:
a. Biopolymer – living anionic polymerization – polyaddition – free radical
polymerization.
b. Living anionic polymerization – radical polymerization – polyaddition –
biopolymer.
c. Polyaddition – radical polymerization – living anionic polymerization –
biopolymer.
d. Radical polymerization – polyaddition – biopolymer – living anionic
polymerization.

(7) Which method is suitable for the purification of polymers and why?
a. Crystallization, because polymers based on covalent linkages can be puri-
fied by adjusting the polymer concentration in solution as high as possible.
b. Crystallization, because the long polymer chains can easily change into a
glassy arrangement upon melting and crystallize upon cooling.
c. Precipitation, because the mixing process of a polymer and a solvent is en-
tropy-driven, and the entropy in a polymer system is particularly high due
to the long chain length.
d. Precipitation, since the mixing process of a polymer and a solvent is en-
ergy-driven, whereas entropy hardly plays a role in polymer–solvent
mixtures.
30 1 Introduction to polymer physical chemistry

(8) Polymers are a prime representative of soft matter, because ______


a. they consist of small monomer units bound by weak interactions such as
van der Waals interactions.
b. they exhibit fast relaxation times of a few microseconds that are accessible
only to very sophisticated methods.
c. they consist of individual chains that have the shape of a random coil in the
colloidal size range and interact with each other via van der Waals or simi-
lar interactions.
d. they consist of covalently linked monomer units forming long chains that
interact with themselves.
2 Ideal polymer chains

LESSON 2: IDEAL CHAINS


In your elementary physical chemistry classes, the first and most simple type of matter that
you dealt with was the ideal gas. In polymer science, this has an analog. The first and most
simple type of polymeric matter that we will deal with is the ideal chain. In this lesson, you
will get to know the basics of the ideal chain model, see how this is conceptually very related
to the ideal gas analog, and you will see how this model allows us to simply capture and
quantify the structure of a polymer.

The goal of this book is to establish relations between the structure and properties
of polymers. To do so, we have to find a way to comprehensively describe the struc-
ture of the elementary building block of each polymeric material: the single poly-
mer chain. Polymers are made from a large number of repeating units; as such, a
polymer chain is a complex, multibodied entity that is hard to describe analytically.
We therefore have to rely on models that simplify the complexity while simulta-
neously retaining the physical realities that we observe in experiments. In the fol-
lowing, we will introduce a few of these models, starting with the simplest one: the
ideal polymer chain.
Let us recall a concept that you have encountered in the beginning of your
basic physical chemistry lectures, where you have considered the most simple state
of matter imaginable: the ideal gas. The description of an ideal gas is based on the
assumptions that (i) the gas molecules are point-like particles with no volume and
(ii) no interactions (other than simple elastic collisions). These assumptions are, of
course, not true: real gas molecules do have a finite volume, and they do interact
with each other in many ways, especially at short mutual distances, which is given
at high pressure. In many cases, however (for instance, at usual temperature21 and
pressure), gases in fact widely behave ideal, allowing us to use the above very sim-
plified model to treat them, which gives us (iii) a very simple equation of state:
pV = nRT, with p the pressure, V the volume, n the molar number of gas particles
(atoms or molecules) in our system, R the gas constant, and T the absolute tempera-
ture. The concept of the ideal polymer chain is based on the same simplistic princi-
ple: it imagines the chain to consist of (i) monomer segments with no volume and
(ii) no interactions with one another or the environment (other than the trivial inter-
action that they have due to their mutual connectivity, which, as we will see in
Chapter 3, impairs their independent statistical motion). The model of the ideal
polymer chain is the basis for any further development of refined polymer chain

 Strictly speaking: at a temperature well above the critical one.

https://doi.org/10.1515/9783110713268-002
32 2 Ideal polymer chains

models and can therefore rightfully be called a starting point of polymer physics.
We will see later in this chapter that this model will give us (iii) an equation of state
that is strikingly similar to the ideal gas law, both in view of its mathematical ap-
pearance and in view of its simplicity but yet powerful utility.

2.1 Coil conformation

2.1.1 Micro- and macroconformations of a polymer chain

Let us begin by considering how a polymer chain might look like. The shape of a
polymer is determined by three factors, as illustrated in Figure 13(A). The first factor
is the monomer segmental length, that is, the length of the elementary structural
unit plus the bond that connects it to the next. In many simple vinyl polymers,
such as polyethylene (PE), polypropylene (PP), polyvinylchloride, or polystyrene
(PS), this is a covalent carbon–carbon bond with a length of 1.54 Å. The second fac-
tor is the monomer bonding angle, φ. It has a value of 109.6° for an sp3-hybridized
carbon–carbon bond. Both of these parameters are determined by the specific
chemistry of the monomer, which means that they are fixed for a given polymer.
The third factor is the monomer bond torsion angle, θ. This factor is not fixed, be-
cause the bonds can rotate quite easily. As a consequence, the macroconformation
of a chain will result from the sequence of its monomer bond torsional microconfor-
mations. A polymer chain may therefore adopt many different shapes, two of which
are depicted in Figure 13(B) and (C). In Figure 13(B), you see a structure in which all
the bonds are trans-conformed. This all-trans conformation leads to an ordered,
rod-like macroconformation. Entropy, however, does not favor such ordered states.
More favorable, by contrast, is the structure depicted in Figure 13(C). Here, the mi-
croconformation is a random cis–trans mix that leads to a macroconformation of a
random coil, which does not show a high degree of order. The actual reason for the
entropic favor of this random-coil macroconformation is because it is realized by
many different microconformations of kind as the one shown in Figure 13(C), be-
cause there are many different possibilities for arrangement of a random cis–trans
mix along the chain contour, whereas the rod macroconformation is realized by just
one microconformation, namely the all-trans chain. With that, the core principle of
statistical thermodynamics applies: a macrostate that is realized by a high number
of microstates has a higher entropy and is more likely to be observed (both over
time and ensemble) than a macrostate that is realized by just a few microstates.
So far, we have discussed the monomer bond torsional angle θ and the resulting
structure of a polymer chain primarily from an entropic perspective. There is, how-
ever, also an energy side to the argument. This is because some monomer bond tor-
sional angles are energetically favorable, whereas others are not. The trans or gauche
conformations have lower free energies, ΔE, than the cis or the anti ones, as shown in
2.1 Coil conformation 33

Figure 13: (A) Characteristic quantities of the elementary chemical bonds in the backbone of a
polyethylene chain. In this simplest case of a polymer, the bond length between two methylene
units, here denoted as dark-shaded dots, corresponds to the segmental length, l. The monomer
bonding angle, φ, is the angle between two segment–segment (here: methylene–methylene)
bonds, and the monomer bond torsion angle, θ, defines the conformational positions of each unit.
Picture taken from H. G. Elias: Makromoleküle, Bd. 1 – Chemische Struktur und Synthesen (6. Ed.),
Wiley VCH, 1999. Depending on the angle θ, many different macroconformations of the polymer
chain can be realized, two of which are shown here. (B) If all microconformations of the chain are
trans, this leads to an ordered, rod-like macroconformation that is entropically unfavorable. (C) In
contrast, if a more favorable random microconformational mix of cis and trans is present, this
leads to a macroconformation of a random coil. This unordered structure is favored by entropy,
because it is realized by more microconformations than the all-trans structure in (B).

Figure 14 for the very typical case of a carbon main chain. The energy difference be-
tween those conformations, however, is just small: the energy of the trans or gauche
conformations differs by only 3 kJ mol–1, which corresponds to just about 1.2 RT at
room temperature. Furthermore, a conformational change between these states has
an energy of activation of only 13 kJ mol–1, that is, just about 5 RT. Even the biggest
energy gap, which is between the cis and the trans conformation, is only about
17 kJ mol–1, that is, 7 RT. This means that all the conformations available to the
34 2 Ideal polymer chains

Figure 14: Dependence of the molar energy, ΔE, on the bond torsion angle, θ, for basic
carbohydrates. Conformations are abbreviated as C = cis, G = gauche, A = anti, T = trans. The
energy has an absolute minimum at the trans conformation and two further local minima for the
gauche conformations G+ and G– that are 3 kJ mol–1 higher (ΔETG). The energy of activation needed
to change between T and G+ or G– is 13 kJ mol–1 (ΔE‡TG), and the energy gap between the most
favorable and the least favorable conformation, T and C, is 17 kJ mol–1 (ΔE‡). Picture redrawn from
H. G. Elias: Makromoleküle, Bd. 1 – Chemische Struktur und Synthesen (6. Ed.), Wiley VCH, 1999.

building blocks of a polymer chain are not very different in energy. Moreover, it is
easy to pass from one conformation to the other, because a conformational change
needs an activation energy in the range of only a couple RT at room temperature. As a
result, these changes happen fast, which is the prime reason why polymers are flexi-
ble above their glass transition temperature. This flexibility is lost if there is a more
severe energy barrier between the monomer segmental conformations, which is the
case for monomer units that carry bulky substituents. This is the reason why PS,
which has a bulky benzene side group, has a higher glass transition temperature than
PE, where no side groups are present, because it takes more thermal energy to activate
conformational change dynamics in the more obstructed PS backbone than in the flex-
ible PE backbone. (See: here we have our very first qualitative structure–property
relation!)
As another direct consequence of these low energy values, the entropy aspect is
mainly responsible for the macroconformation of a polymer chain and outweighs
the energy aspect. Therefore, the most probable structure of a polymer chain is
that of a flexible random coil. There are some cases where energy outweighs en-
tropy, but these only occur when an extremely high energy gain is associated with
the adoption of a certain specific chain structure. Proteins are a prime example:
they have specific, energetically favorable folded shapes, such as the α-helix or the
β-sheet. The energy they gain through secondary interactions when adopting these
shapes outweighs the entropy penalty that such an ordered structure imposes. This
ordered structure, in turn, is the basis for their biological function, often referred to
as key-and-lock principle. This notion has led Vollmert to formulate that biopoly-
mers such as proteins are the highest level of complexity in chemistry and the
2.1 Coil conformation 35

lowest level of complexity in biology. Most synthetic polymers, by contrast, do not


have such preferred structures, but instead shape up as random coils.
Now that we know that a polymer chain looks like a coil, how do we describe it
mathematically? Even if we restrict ourselves to just the energetically favored trans
and gauch microconformations, T, G+, and G–, there are 3N – 2 possible macroconfor-
mations that result in a chain with N segments. If we consider N to be 100, which is
still a comparably short polymer chain, that would mean that there are 6 × 1046 pos-
sible different macroconformations! On top of that, conformational change is very
fast due to the low energies involved. An appraisal based on an Eyring-type equation
with an energy barrier of 13 kJ mol–1, or 5 RT at room temperature, for the T ! G
transition shows that they happen on a timescale of just nanoseconds at room
temperature. All this means that the shape of a polymer chain cannot be described
analytically. It can, however, be described mathematically using suitable average
values.

2.1.2 Measures of size of a polymer coil

Two common averages are used to describe the size of a polymer coil. These are the
end-to-end distance, ~ r, and the radius of gyration, Rg, both depicted in Figure 15.
The end-to-end distance, ~ r, is calculated by simply summing up all individual bond
vectors, which creates a vector that points from one end of the polymer chain to the
other:
X
~ r1 +~
r =~ r2 +~
r3 +    = ~
ri (2:1)

It can be easily visualized and imagined, but it is hard to be determined experimen-


tally, which is why it is often preferred by polymer theorists. Note that the end-to-
end distance is only meaningful for linear chains, which have a defined start and
end of their polymer chain. Branched macromolecules, by contrast, have many
end-to-end distances, because they have multiple chain ends.
The radius of gyration, Rg, is calculated by the sum over all mass segments, m,
weighted by their distances to the center of mass:
ð
1 2  1X 1 2
Rg =
2 ~ ~ 2
~ 2
r 1 m1 + r 2 m2 + r 3 m3 +    = ~ 2
ri mi = r dm (2:2)
m m m

Rg corresponds to about the 1.3-fold of the radius of a sphere with the same moment
of inertia and the same density as our actual object of interest. It is not limited to
linear polymers, but can also be determined for branched or cross-linked polymers.
In fact, Rg can be calculated for any geometrical object. It is also much easier to
assess experimentally than the end-to-end distance, which is the reason why it is
often preferred by polymer experimentalists.
36 2 Ideal polymer chains

Figure 15: Schematic representation of two common average measures of the size of a polymer coil:
(A) The end-to-end distance ~r, simply calculated as the sum of all bond vectors, and (B) the radius of
gyration, Rg, calculated by a normalized sum over all mass segments weighted by their distances to
the center of mass of the polymer coil; Rg corresponds to the radius of a solid sphere (illustrated by a
dashed green circle) with the same moment of inertia and the same density as the fuzzy polymer coil.

Both quantities are connected to each other by simple relations that depend on
the geometry of the compound investigated. Table 4 compiles a selection of some of
these relations.

Table 4: Relation between the end-to-end distance and the radius of gyration for various types of
objects.

Type of molecule/ <Rg> – characteristic Characteristic length–mass


object length
 2  2  2
Random coil  2 ~ r ~
r ⁓ Rg ⁓ m
Rg =
6
 2  2
Coil in good  2 ~
r ~
r ⁓ Rg 2 ⁓ m2ν ν = Flory exponent (Chapter )
Rg =
solvent ð2ν + 1Þð2ν + 2Þ → ideal coils: .
→ with rep. interact: .

Rod with length L   L2 L⁓m


Rg 2 =
12
Disc with radius r  2  r2 r ⁓ m1=2
Rg =
2
Sphere with  2  3r 2 r ⁓ m1=3
radius r Rg =
5
2.2 Simple chain models 37

2.2 Simple chain models

2.2.1 The random chain (phantom chain, freely jointed chain)

The simplest model to describe a polymer chain is that of the random chain, also
referred to as freely jointed chain or phantom chain. It is based on the assumption
that all monomer bond conformations, assessed by the torsional angle, θ, and also
even all monomer–monomer bonding angles, φ, are possible. This is somewhat un-
intuitive, as the bonding angle is in fact set by the monomer’s specific chemistry.
Releasing this constraint and allowing this angle to be free would allow two mono-
mer segments to occupy the same spot in space, for example, if two adjacent mono-
mer segments would have a bonding angle of φ = 0°, which is, nevertheless,
allowed to be possible in the random chain model. This is why this model is also
referred to as the phantom chain model, as in this framework, the chain segments
can penetrate through each other like phantoms. With that simplifying assumption,
the random chain model relates the number of monomer segments, N, and their
length, l, to the end-to-end distance of the polymer chain, ~r, in a very simple fash-
ion, as we will show now by simply estimating the end-to-end distance through the
bond vector sum. For a reason that we will detail below, we use the square value of
the end-to-end vector, ~ r2 . With that, we calculate the square of the bond vector
sum. This must be done using a scalar product term, which gives us a sum of
squared lengths of the bonds vectors plus a cosine term of their angles:

X
~
r= ~li (2:3a)
 2
r2 = ~l1 +~l2 +~l3 +   
~
  X
= l1 2 + l2 2 + l3 2 +    + li lj cos φij (2:3b)

With this calculation, so far, we have only estimated the momentary-squared end-to-end
distance of one single polymer coil. This, however, is not representative, as the
coil conformation is subject to constant dynamic change, and as in a sample with
many chains, even in a momentary view, they won’t all exhibit the same end-to-
end distance, but a distribution of it. So, what we are actually interested in is the
mean of this distribution. If we would calculate it for ~ r directly, however, the
mean would always be zero. This is because ~ r is a vector quantity that has a direc-
tion. In a sample with many representants of this quantity, there will always be
pairs of same length that point into opposite directions, such that the mean of all
these would cancel out to zero. To get rid of this circumstance, we take the square of the
38 2 Ideal polymer chains

end-to-end vector first before averaging. That way, we eliminate the directional depen-
dence and obtain the mean-square end-to-end distance:
 2  2  X
~
r = l1 + l2 2 + l3 2 +    + li lj hcos φij i
 2 2 2

= l1 + l2 + l3 +   
= Nl2 (2:4)

In the latter form of the scalar product, the average of the angular dependence,
hcos φij i, is zero. This is because we assume that the bonding angle, φ, is free and
exhibits a random distribution, and so the cosine of that random angle in eq. (2.3b)
delivers random values in an interval of –1 and +1 around 0, which all occur with
same likelihood. So, each realization in the positive direction has an equally large
counterpart in the negative direction, thereby all cancelling to zero upon averaging
in eq. (2.4). The square of the end-to-end distance therefore only depends on the
number of monomer segments, N, and their length, l.
With this calculation, we have derived a simple power law for the chain mean-
square end-to-end distance as a function of the number of chain segments. To re-
linearize the physical dimension of the mean-square end-to-end distance, we can
take the square root and obtain the root-mean-square (rms) end-to-end distance,
 2 1=2
R= ~ r . Such rms values can be found all over the field of physical chemistry, as
they are a practical tool to handle vector quantities that are subject to a distribu-
tion. rms values eliminate the vectors’ directional dependencies and handle their
distributed nature by using the arithmetic mean of their square values. From this,
the square root is then formed to re-linearize the mean square quantity to its truly
meaningful physical dimensions. Following eq. (2.4), the rms end-to-end distance
of our phantom polymer chain scales with N1/2:

R = h~
r 2 i1=2 ⁓ N 1=2 (2:5)
Note that from eq. (2.5), it follows that a polymer that has twice as many segments
as another one is not twice as large as that other one in space, but only 21/2 ≈ 1.4
times as large. To get a polymer that is twice as large as another one in space, it
must have 22 = 4 times as many segments!
A very similar scaling law to eq. (2.5) can be found for freely diffusing particles
that move by a random walk. Here, the rms displacement, h~ x2 i1=2 , scales with the
square root of the number of diffusion steps, and therefore also with the square
root of time, t1=2 , if we assume that each elementary step takes a defined period.
This scaling is based on the Einstein–Smoluchowski equation:

h~
x2 i1=2 ⁓ t1=2 (2:6)

The similarity of eq. (2.5) and eq. (2.6) is striking. Upon closer scrutiny, however, it
is reasonable. The shape of a random phantom coil composed of bonds with fixed
2.2 Simple chain models 39

length l and free random bonding angles φ is the same as that of a path of diffusion
steps of fixed length l and free random directional change from step to step, as
shown in Figure 16.

Figure 16: Trajectory of a two-dimensional


random walk.

2.2.2 The freely rotating chain

So far, we have assumed no angular dependence in our calculations. However, we


actually know how such a dependence should be described, because we know the
bonding angle, φ, from the chemical specificity of each monomer. Thus, we can as-
sume that each bond vector~lj projects a component l · cos φ onto the next bond vec-
tor ~lj + 1 . Accounting for that, eq. (2.4) can be refined to
 2 1 − cos φ
~
r = Nl2 (2:7)
1 + cos φ

This equation is derived through a series expansion, which is explained in


detail in Rubinstein’s and Colby’s book Polymer Physics. When we calculate
ð1 − cos φÞ=ð1 + cos φÞ for an sp3-hybridized carbon bond with a bonding angle of
φ = 109.6°, we get a value of 2. Consequently, we see that we underestimate such a
polymer coil’s size by a factor of √2 if we disregard its fixed bonding angle, as we do
in the phantom chain model.
40 2 Ideal polymer chains

2.2.3 The chain with hindered rotation

We can further take into account the bond’s torsional angle, θ. We have seen in Sec-
tion 2.1 that the torsional angle is not fixed, but prefers specific conformations due
to their lower energies relative to other conformations. Accordingly, we can average
over all possible bond conformations per time and chain ensemble to determine the
average torsion angle hcos θi:
R +π  
ΔEðθÞ
−π exp − RT cos θ dθ
hcos θi = R +π   (2:8)
ΔEðθÞ
−π exp − RT dθ

Equation (2.8) is based on the integral of a Boltzmann term of the torsional-angle-


dependent potential energy, ΔE(θ), from Figure 14. With this value, we can append
a term for the torsional angle dependence to our chain model:
 2 1 − cos φ < 1 + cos θ >
~
r = Nl2 · (2:9)
1 + cos φ < 1 − cos θ >

We have now developed a model that takes into account both the universal physics of
a polymer coil, which is expressed by its universal scaling according to h~ r 2 i1=2 ⁓ N 1=2 ,
and that also contains terms that account for the chemical specificity of a given type of
monomer or structural unit, characterized by its bonding angle and its average tor-
sional angle. These specific values are constant for each kind of repeating unit,
which is why they are often summed up into polymer-specific parameters. The
torsional-angle-dependence h1 + cos θi=h1 − cos θi is often called the obstruction
parameter, σ2. It is determined by the bulkiness of the side groups of a repeating
unit: the more energy is needed for these units to change their conformation, the
higher this value. More often, the characteristic ratio, C∞ , is used; it includes
all chemistry-specific parameters in the form of

1 − cos φ h1 + cos θi
C∞ = · (2:10)
1 + cos φ h1 − cos θi

With that, we get the general scaling law for ideal polymer chains:
D 2E
!
r = C∞ Nl2

The characteristic ratio is large for stiff polymer chains with bulky or like-charged
substituents along the chain backbone, such as PS, poly(methylmethacrylate), or
poly(sodium acrylate), whereas it is low for flexible chains with uncharged and un-
bulky substituents, such as PE. Note that for short chains, the characteristic ratio
shows a dependence on the number of repeating units, N, as depicted in Figure 17;
it is then abbreviated by the symbol CN . From about 80–100 repeating units on,
2.2 Simple chain models 41

Figure 17: Characteristic ratio, CN, as a function of the number of repeating units, N. Initially, CN
rises with N, but after about 80–100 repeating units, it reaches a plateau for each polymer. That is
why the characteristic ratio is often referred to as C∞ . Note how its value rises from about 3 to
about 10 if we go from flexible chains with unbulky substituents such as polyethylene to chains
with bulky substituents such as poly(methylmethacrylate). Picture redrawn from H. G. Elias:
Makromoleküle, Bd. 2: Physikalische Strukturen und Eigenschaften (6. Ed.), Wiley VCH, 2001.

however, it reaches an N-independent plateau, as also shown in Figure 17; it is then


abbreviated by the symbol C∞ .

2.2.4 The Kuhn model

Figure 18: The Kuhn model sums up several actual polymer chain segments of length l to create a
new conceptual polymer chain made up of N=C∞ Kuhn segments of conceptual length lK = C∞ l, the
Kuhn length. The end-to-end distance of the original chain, ~
r, is retained. The bonding angles of
the Kuhn segments, however, are now random, whereas the bonding angles of the original chain
were fixed. As a consequence, the conceptual chain can be assessed as a random chain.
42 2 Ideal polymer chains

When we recapitulate what we have done in the preceding sections, we see that we
have taken up the most simplistic model and modified it with further terms to
achieve a closer resemblance to reality, thereby making it more complex. But
wouldn’t it be magnificent to have a model at hand that would retain the original
simplicity of the random chain and, at the same time, still capture chemical speci-
ficity? This can be done on the basis of an ingenious approach by Werner Kuhn,
which is the Kuhn model. In this model, a number (not necessarily a natural num-
ber) of several segments of a polymer chain with length l are grouped such to create
a new conceptual polymer chain made up of new conceptual segments, the so-
called Kuhn segments of new conceptual length lK, the so-called Kuhn length (see
Figure 18). The original chain end-to-end distance is retained for this new concep-
tual chain. The ingenious element of this approach is the following: the chemical
specificity of the original polymer is incorporated into the model by normalizing
the number N and length l of the repeating units to the characteristic ratio C∞. That
way, each given polymer “forgets its chemical specificity” and can be assessed by a
universal power law:
 2 N
~
r = C∞ Nl2 = ðC∞ lÞ2 = NK · lK 2 (2:11)
C∞

The new conceptual chain is made from NK = N =C∞ Kuhn segments of length lK = C∞ l.
The bonding angle φ between these segments is now random, whereas it was a fixed
constant in the original chain (Figure 18). As a result, the Kuhn chain can be assessed
like a random chain. In other words: on scales beyond the Kuhn-length, the chain
chemical specificity is no longer seen, and instead, each chain, independent of what
sort of polymer it actually is, displays universal, random-walk type statistics. On these
scales, hence, specific monomeric chemical primary structures do no longer matter,
which allows us to conceive and draw the chain not by its actual chemical composi-
tion, but as a universalized schematic of a wiggly line; this is what we actually do from
now on throughout the whole remainder of this book. With the Kuhn-approach, we
have traced our more and more complex models described in the previous sections
back to their simple origin. Note that with R2 = C∞ Nl2 as well as lK = C∞ l we get an in-
teresting formula: R2 = C∞ Nl2 = C∞ lNl = lK lcont , with lcont = Nl the contour length, that
is, the chain length in an hypothetic fully decoiled conformation.
2.2 Simple chain models 43

Werner Kuhn (Figure ) was born on February ,


 in Maur, Switzerland. He developed a lively
interest in natural phenomena in childhood days
already and was schooled and studied in Zürich,
where he also obtained his PhD degree with
research on the photochemical degradation of
ammonia in . After two years in Copenhagen,
where he co-worked with Niels Bohr and also
found his wife, Kuhn obtained his habilitation in
Zürich in . He then worked at Heidelberg
University from  till , where he developed
his famous theory for the optical activity of chiral
molecules. Kuhn then moved to Karlsruhe, where Figure 19: Portrait of Werner Kuhn. Image
he started his work on polymers, obtained a reprinted with permission from Chemie in
professorship at Kiel in , and happily accepted unserer Zeit 1985, 19(3), 86–94. Copyright
a position at Basel in . In the s, he 1985 Wiley VCH.
researched on membranes and gels, before he
became director of Basel University in .
Kuhn died unexpectedly on August ,  in
the middle of fruitful work. There is an anecdote
in the polymer community about Kuhn’s time in
Basel, which is very close to the place of Freiburg,
where Staudinger acted at the same time: Kuhn,
a physical chemist, and Staudinger, an organic
chemist, are said to having had kind of a “loyal
rivalry” on the viewpoint of how polymers are
properly conceived. For example, Kuhn’s work “Über
die Gestalt fadenformiger Moleküle in Lösung”
started with the statement that Staudinger’s view
on caoutchouc and cellulose to be elongated rods in
solution would in fact be outruled by Staudinger’s
very own experiments. The aftermath of that was
present up to Staudinger’s th birthday, where
it was expressed poetically as “die Kuhnschen
Knäuel sind uns hier ein Greuel”.

2.2.5 The rotational isomeric states model

We can also do the opposite of the Kuhn model’s simplicity and try to find a maxi-
mally realistic expression for C∞ for a given specific chain; this is the approach of
the rotational isomeric states (RIS) model.
It explicitly considers the energies of the most relevant conformational states T,
G+, and G–, denoted as ΔE(θ) in Figure 14, and then uses the Boltzmann distribution
to quantitatively appraise their populations:
44 2 Ideal polymer chains

ngauche − ΔEðθÞ
= 2 · exp (2:12)
ntrans RT

This equation determines the exact ratio of gauche and trans conformations in a
given polymer, as shown in Figure 20. The factor of 2 before the exponential term is
due to the existence of two gauche conformations, G+ and G–. In addition to the
gauche-to-trans conformational ratio, the RIS model also appraises how the confor-
mation of each bond i is depending on what the conformation of the preceding
bond i – 1 is; in other words, the model quantifies how many dyads of TT, TG, and
GG are present in the chain. This is done by quantifying the energies of these differ-
ent combinations and translating them into relative population ratios, as summa-
rized in Table 5.

Figure 20: The RIS model quantitatively appraises the populations of the low-energy
conformational states T, G+, and G–, and it determines the population of each of these states with
the Boltzmann distribution. The upper diagram is a replot of Figure 14 and shows the energies of
monomer–monomer bond torsional angles for butane; the lower diagram displays the relative
populations of these bond angles at a temperature of 300 K. Upper diagram redrawn from
H. G. Elias: Makromoleküle, Bd. 1 – Chemische Struktur und Synthesen (6. Ed.), Wiley VCH, 1999;
lower diagram redrawn from R. H. Boyd, P. J. Philips: The Science of Polymer Molecules, Cambridge
Solid State Series, 1993.
2.2 Simple chain models 45

Table 5: Relative population numbers of T and G conformations of a bond i in a polymer chain


depending on the conformation T or G of the preceding bond i – 1.

Bond i

0° (T) +120° (G+) –120° (G–)

1 0.54 0.54 0°(T)

Bond i – 1
1 0.54 0.05 +120° (G+)

1 0.05 0.54 –120° (G–)

2.2.6 Energetic versus entropic influence on the shape of a polymer

A key part of eq. (2.12) is the ratio of ΔE/RT. If this ratio is high, then the chain primar-
ily adopts the energetically most favorable shapes. In synthetic polymers, these are
either all-trans or alternating TGTGTG or TTGGTTGG conformational shapes, depend-
ing on the bulkiness of the substituents and their optimal low-energy arrangement in
space. The most well-ordered polymers in nature are proteins; they have very specific
energetically highly favorable folded shapes, and it is this specificity that gives rise to
their specific functions, because these specific outstanding ordered shapes make pos-
sible the lock-and-key principle that enables specific enzyme–substrate interactions.
By contrast, at a low ratio of ΔE/RT, a polymer chain adopts the entropically most
favorable conformation, which is that of a random coil.
A related ratio is that of the energy of activation for torsional change, ΔE ‡TG/RT.
This ratio corresponds to the height of the energy barrier between the T and G states
and thereby describes the dynamics of a polymer chain. A high ratio of ΔE ‡TG/RT is
encountered for stiff polymer chains that carry bulky side groups, such as PS. A low
ratio, on the contrary, is encountered for flexible polymer chains that carry unbulky
side groups, such as PE. This ratio is directly reflected in one of polymers’ most im-
portant property: the glass-transition temperature, which is the temperature needed
for activation of main-chain dynamics. This temperature is high for PS, whereas it
is low for PE due to the less easy activation of bond conformational changes in PS
than in PE. (See: again, here we have a structure–property relation!)
Both the latter relations are dependent on temperature, which is because

ΔG = ΔH − TΔS (2:13)

In general, the change of Gibb’s free energy, ΔG, needs to be negative for any pro-
cess to proceed spontaneously. We can see on the right-hand side of eq. (2.13) that
46 2 Ideal polymer chains

the change of enthalpy, ΔH, and that of entropy, ΔS, are connected to each other by
temperature, T. It directly follows that at low temperature, the entropy term has a
low significance, so that the enthalpy term dominates. As a result, a polymer coil
expands and adopts a more ordered macroconformation at low temperature. This
can even lead to higher-order chain assemblies such as polymer crystals. On the
opposite, at high temperature, the entropy term dominates; here, a polymer coils
up to a disordered macroconformation.

2.2.7 The persistence length

Due to the fixed bonding angles and somewhat preferred conformational arrange-
ments of the bonds in a polymer chain, as captured by the latter three models
above, each segment i imparts a directional preference onto the next following ones
i + 1, i + 2, i + 3, . . . along the chain. It therefore takes a certain number of following
segments until the directional “memory” of a given first segment is “forgotten”. This
“directional memory” or persistence is captured in the form of a quantity named per-
sistence length. Mathematically, it can be defined as the projection of all following
bond vectors i + j onto the direction of a given first one i, as shown in Figure 21.

Figure 21: Projection of all bonds i + j following a given one i onto its direction.

This projection is expressed as


1 X 
lp = li lj (2:14)
li j > i

Just as the characteristic ratio, C∞, and the Kuhn length, lK = C∞l, the persistence
length captures the stiffness of a polymer chain. The stiffer the chain is, the longer is
the length up to which the directional influence of a given segment is still recogniz-
able. It is not surprising that the two quantities lp and C∞ are related to one another:

lp
C∞ = 2 −1 (2:15)
l
2.2 Simple chain models 47

From that, we also obtain a relation between the persistence length, lp, and the
Kuhn length, lK:

lp
C∞ + 1 ≈ C∞ = 2
l
, C∞ l = lK = 2lp (2:16)

As an example, consider PS. In this polymer, we have C∞ = 10.2 and l = 0.154 nm;
hence, based on eq. (2.15), we get lP = 0.86 nm. This is still low! For comparison, con-
sider double-stranded DNA, which has a persistence length of 63 nm. This extremely
large value originates from the constrained primary constitution of the DNA double
strand and from the multiple repulsive electrostatic interactions between the charges
of its phosphate units. Chains with such high stiffness are no longer named flexible,
but instead semiflexible or wormlike. In these, a different geometrical view on the per-
sistence length (that reflects the same mathematics, of course) is commonly used,
given by the intersection angle of two tangents on the chain, as shown in Figure 22.

Figure 22: Schematic of a semiflexible, wormlike chain with two tangents constructed to it that
cross in a certain intersection angle.

In wormlike chains, the persistence length is the length that it takes to have the
average cosine of the angle shown in Figure 22 to decay down to 1/e (=0.368).
48 2 Ideal polymer chains

2.2.8 Summary

We have learned throughout this chapter that the most probable shape of polymer
chain is that of a random coil. We have developed the simplest model to describe this
coil, which is the random chain model. Based on that, we have continuously ex-
panded this model to incorporate the chemical specificity of the monomer segments
by adding contributions from the bonding angle, φ, in the freely rotating chain model,
and the bond torsion angle, θ, for the chain with hindered rotation (see Table 6). This
chemical specificity can be captured by the characteristic ratio C∞, which is taken in
the Kuhn model to resimplify and universalize the chain model while simultaneously
retaining the given specificity of each given polymer. Furthermore, we have seen that
the macroconformation of the chain is dependent on temperature and on the specific
chemical design of the monomeric units. Polymers with bulky side groups exhibit stiff
chains at ambient temperatures, whereas polymers with small side groups are flexible
at ambient temperatures. We have therefore created our first rational structure–
property relation.

Table 6: Summary of the characteristic simple chain models addressed in this chapter.

 2
R2 = ~
r = C∞ Nl2 Random chain Freely rotating chain Chain with RIS model
hindered rotation

Bond length (l) Fixed Fixed Fixed Fixed

Bond angle (φ) Free Fixed Fixed Fixed

Torsion angle (θ) Free Free Fixed average Discrete: T, G+, G–

Char. ratio (C∞)  1 − cos φ 1 − cos φ h1 + cos θi Specific


·
1 + cos φ 1 + cos φ h1 − cos θi

To get an impression on some numbers, consider PE with N = 20,000, l = 0.154 nm,


and C∞ = 6.87. Whereas the length of the maximally elongated chain, r = Nl, would
be 3080 nm, its rms end-to-end distance is R = <r2>1/2 = (C∞ N l2)1/2 = 57 nm. This esti-
mate clarifies two insights. First, the polymer is heavily coiled in its natural state.
Second, in that state, it has a size that falls right into the colloidal domain.
Questions to Lesson Unit 2 49

Questions to Lesson Unit 2

(1) The most likely conformation of a polymer chain is ______


a. that of a random coil, since the bond torsion angles can be changed by
small activation energies.
b. that of a random coil, since the bonding angles are statistically distributed
by the synthesis process of the polymer.
c. that of a defined coil, since bond length, angle, and torsion angle are prede-
termined and fixed by the chemical structure.
d. that of an elongated chain, since the bond torsion angles arrange them-
selves such that the segments are as far apart as possible.

(2) The characteristic ratio ______


a. reflects the ratio of the actual coil’s mean-square end-to-end distance to
that of a hypothetical ideal chain with the same segmental number and
length.
b. reflects the characteristic ratio of bonding angle and torsion angle for each
polymer.
c. allows the specific characteristics and chemical identities of polymers to be
clearly specified.
d. describes the characteristic behavior of a freely rotating chain.

(3) The Kuhn model ______


a. considers a random chain and is therefore independent of the characteristic
ratio.
b. is based on normalizing the segmental length and number to the character-
istic ratio.
c. does not reflect chemical specificity but is universal in nature.
d. divides the original chain into a number of arbitrarily small segments.

(4) The persistence length ______


a. is a measure of the durability of a chain.
b. agrees exactly with the Kuhn length.
c. is smaller for larger substituents along the chain.
d. is a measure of the stiffness of the chain’s backbone.
50 2 Ideal polymer chains

(5) Consider two equally sized ensembles of one-dimensional random walks: en-
semble “Walk A” with a number of four steps each and ensemble “Walk B”
with a number of ten steps each. Which statement regarding the random walks
is correct?
a. RMS and mean end-to-end distance reflect the same magnitude and there-
fore always have the same value, but they are greater for Walk B than for
Walk A.
b. For Walk A and B, the mean end-to-end distance is the same, but the RMS
end-to-end distance of Walk B is greater than that of Walk A.
c. For Walk A and Walk B, the RMS end-to-end distance is the same, but the
mean end-to-end distance of Walk B is greater than that of Walk A.
d. None of the previous statements is correct.

(6) Order the characteristic ratios of the following polymers from small to large:
PE, PP, PS.
a. PP – PE – PS
b. PE – PS – PP
c. PP – PS – PE
d. PE – PP –PS

(7) The scaling law R2 ¼ C∞ Nl2 ______


a. combines chemical specificity with the physical relationships that are uni-
versally valid for all polymers.
b. does not apply to ideal chains, since they have no specificity.
c. is a special case and cannot be applied to all polymers.
d. results from the assumption of a freely rotating chain.

(8) At high temperatures, a polymer takes the ______


a. enthalpically favored conformation of a rod.
b. entropically favored conformation of a rod.
c. enthalpically favored conformation of a coil.
d. entropically favored conformation of a coil.
2.3 The Gaussian coil 51

2.3 The Gaussian coil

LESSON 3: GAUSSIAN COILS AND BOLTZMANN SPRINGS


The last lesson has taught you that an ideal chain has the shape of a random coil. The fol-
lowing lesson will further refine this insight and make you understand how exactly the seg-
mental density is distributed inside the coil. It will also show how the most fundamental
thermodynamic property of such a coil, its entropy, reacts on deformation of the coil, thereby
introducing the most relevant property of flexible polymers: their entropy-based elasticity.

We now want to take a closer look at how exactly a polymer coil looks like. We start
by tying into the number example from the end of the preceding section and again
consider PE with N =20,000, l = 0.154 nm, and C∞ = 6.87, for which we have estimated
an rms end-to-end distance of R = <r2>1/2 = (C∞ N l2)1/2 = 57 nm. We now calculate the
volume of just all the segments together, that is, the volume that a dense globule
would have into which the coil might collapse: Vsegments = no. of segments × vol. per-
segment ≈ N l3. For comparison, we also consider the volume that the coil has in its
natural state: Vcoil = 4/3 π R3 = 4/3 π (C∞ N l2)3/2 ≈ 4 C∞3/2 N3/2 l3. When we compare
these two volumes by calculating their ratio, Vcoil/Vsegments = (4 C∞3/2 N3/2 l3)/(N l3) = 4
C∞3/2 N1/2, we realize that this ratio is more than 10,000! This means that most of the
volume of the polymer coil is actually empty and not occupied by the polymer chain
material itself. But how, then, is that chain material distributed within the coil?
The polymer chain segments in a coil are distributed according to a Gaussian
radial density profile:
!3=2 !
3 − 3r2
cseg = N · exp (2:17)
2πRg 2 2Rg 2

From the graphical representation of this equation, in Figure 23, we can see that
the highest polymer segmental density is at the coil’s center. For the blue curve in
Figure 23, denoting a polymer with N = 20,000 segments, we have a segmental den-
sity of about 11 segments per nm3 in the coil center; this corresponds to a molar con-
centration of about 20 mol L–1. The further we move away from the coil center, the
more does the density drop, which is especially true for short polymer chains with
only a few segments, whereas longer chains have a less steep radial segmental den-
sity profile. Nevertheless, note that the degree of coiling, Q, gets more pronounced at
high N, as shown in the following estimate:

Lcont Nl
Q= = = N 1=2 (2:18)
hr2 i1=2 N 1=2 l
52 2 Ideal polymer chains

Figure 23: Concentration of polymer segments in a coil as a function of the coil radial coordinate, r,
for two polyethylene chains with N = 20,000 and N = 80,000 repeating units. In both cases, the
majority of segments is located at the coil center, which is more pronounced for the shorter of the
two chains. In radial direction away from that center, the segmental density drops according to a
Gaussian profile. The dashed lines indicate the radii of gyration of both coils. The upper-right
sketch illustrates the radial segmental density profile in a schematic fashion.

Hence, longer chains are coiled more notably than shorter chains, but in terms of
their segmental density profile, long chains are more “dilute” at the coil center than
short chains but have a higher segmental density at the rim in turn.22
This last consideration is a good indication that we have now moved beyond
average values to learn about the shape of a polymer coil. In the following, we will
therefore look more closely at the coil statistics, with the aim to not only describe
its average size, but the full distribution of it.

 This circumstance is a quite astonishing consequence of the fractal nature of polymers, which
we will get to know in Section 2.7. For the moment, we can capture it only (but at least) from a
mathematical viewpoint: According to Eq. (2.17), we get a local segmental density proportional to
N–1/2 for r = 0 (where the exponential factor is exp(0) = 1) if we plug in Rg2 ⁓ N in the prefactor; we
may also get to this notion by simply estimating N/V ≈ N/R3 ⁓ N/(N1/2)3 ⁓ N–1/2. This means: the
higher the degree of polymerization, the lower is the segmental density in the core. The reason is
because random Gaussian coils have a fractal dimension of 2, as captured in the scaling R2 ⁓ N, but
when given 3 actual dimensions, we get terms with R3, which then leads to the above mathematical
result. By contrast, this effect vanishes if we match the geometrical dimension with the fractal one,
that means, if we consider a two-dimensional flat polymer on a surface. In this case, we get a segmen-
tal density that is independent of N, because then, our calculation reads N/A = N/R2 = N(N1/2l)2 = N0.
2.4 The distribution of end-to-end distances 53

2.4 The distribution of end-to-end distances

2.4.1 Random-walk statistics

We have learned in Section 2.2 about the fundamental scaling law between the mean-
square end-to-end distance and the number of monomer segments of an ideal polymer
 
chain, r2 ⁓ N. We have also noted the similarity to the Einstein–Smoluchowski law
 
of diffusion, x2 ⁓ t, and realized that this is due to very similar assumptions made in
the two models: the path of a diffusing particle has no memory and can cross itself,
and an ideal phantom polymer chain is assumed to be composed of volume-less seg-
ments without interactions, so it can cross itself, too. We now rely on this similarity to
appraise the distribution of the chain end-to-end distance, r, by adopting random-
walk statistics. We start by considering a one-dimensional random walk composed of
N steps on the x-axis:

N x Statistics

You may imagine this walk as follows: think that you place a marker at zero and
flip a coin. Every time the coin lands on heads, you move the marker up in the
x-direction, and every time the coin lands on tails, you move the marker down.
After one flip, the marker can be at x = ±1 with a probability of 50% each. After two
flips, the marker can either be back at x = 0, or it can be at x = ±2, with a probability
of 50% for x = 0 and 25% for each x = ±2. In general, for a total number of steps
N = N+ + N–, an end position of x = N+ – N– can be calculated. It is apparent from the
statistics column in the table above, which resembles Pascal’s triangle, that the most
54 2 Ideal polymer chains

probable outcome of any such random walk will be x = 0. This is because there’s
many walks that we can take for a given number of steps, but most of them lead us
back to zero, whereas just few of them lead us far apart from zero. Using the binomial
statistics behind the Pascal triangle in the table above, we can appraise the number
of walks reaching position x after N steps as

N!
W ðN, xÞ = N + x N − x (2:19)
2 ! 2 !

Equation (2.19) can be understood by illustrating the sequence of N steps as an or-


dered row of lots that we randomly draw out of a container. The first one that we
draw is one out of a total of N, the second is one out of a remaining total of (N – 1),
the third is one of a remaining total of (N – 2), and so forth. This gives us N · (N – 1) ·
(N – 2) · . . . = N! possible permutations to arrange that sequence. Each lot has two
possible values in a random walk: it can denote a step in the positive direction or a
step in the negative direction. As a result, we can divide this sequence into two sub-
ensembles. One comprises all those steps that go into the positive direction; we have
a total of N+ such steps. The other comprises all those steps that go into the negative
direction; we have a total of N– such steps. The order of these N+ and N– steps is irrel-
evant, because we always end up at the same final position of x = N+ – N–. Hence, we
have to divide our N! by the number of permutations of these N+ and N– steps, which
is N+! and N–! This gives us N!/(N+! · N–!) possible walks with a given number of posi-
tive and negative steps. We have seen already that the majority of these walks will
lead us close to the center at around x = 0, so we are particularly interested about
those walks with N+ = N/2 + (x/2) and N– = N/2 – (x/2). Replacing the N+ and N– in
the expression N!/(N+! · N–!) with these terms turns it into eq. (2.19).
The total number of walks that we can take at all is 2N. Together with eq. (2.19),
this gives us the probability of reaching position x after N steps:

W ðN, xÞ 1 N!
= N ·  N + x  N − x (2:20)
2 ! 2 !
2N 2

As W(N, x) is a big number, we take the logarithm of it and apply Stirling’s approxima-
tion ln(N!) ≈ N (ln(N) – 1) to eliminate the factorials. We may then expand it into a
Taylor series with x/N as the variable. This will give us a zeroth-order element that is
independent of x/N, whereas we will have no first-order element, as the distribu-
tion is symmetrical. The second-order element will be of kind (x/N)2. If we inter-
rupt the series after this element, it will be of type ln[W(N, x)] ≈ a – b·(x/N)2, so we have
W(N, x) ≈ exp [a – b·(x/N)2] = exp [a] · exp [–b·(x/N)2]. The exact calculation yields
rffiffiffiffiffiffiffi
W ðN, xÞ 2 − x2
ffi · exp (2:21a)
2 N πN 2N
2.4 The distribution of end-to-end distances 55

If we normalize our coordinate system from a step width of 1, as we had it in the


sketch above, to a step width of ½, we transform the integer distance on the x-axis
in the walks in the table above from 2 to 1. With that, we get the probability distri-
bution of the displacement x in a one-dimensional random walk:
rffiffiffiffiffiffiffiffiffi
1 − x2
p1d ðN, xÞ = · exp (2:21b)
2πN 2N

Equation (2.21b) differs from eq. (2.21a) just by a factor of 2, simply because the inte-
ger distance on the x-axis was 2 for the walks in the table above, whereas it has
now been normalized to 1.
So far in this book, we have considered the mean-square end-to-end distance,
which corresponds to the mean square displacement in the random walk. In gen-
eral, such mean values can be calculated from distribution functions by the use of
mathematical operators. Applying such an operator to eq. (2.21b) gives us the mean
square of the distribution:
+ð∞ rffiffiffiffiffiffiffiffiffi +ð∞
1 − x2
hx2 i = x2 p1d ðN, xÞdx = x2 exp dx = N (2:22)
2πN 2N
−∞ −∞

This simple result provides a useful identity, as it connects N to <x2>. With that, we can
replace one by the other in eq. (2.21b) and obtain a formula with just one variable (x):
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 − x2
p1d ðxÞ = · exp (2:23)
2π <x2 > 2 <x2 >

We now want to move into the three-dimensional realm, so we need to consider all
three spatial directions. Fortunately, a simple random walk does not have a pre-
ferred direction; in other words: we consider an isotropic case. In such a case, we
may construct a three-dimensional random walk simply by the superposition of
three independent components, one for each spatial dimension:
 
P3d ð~
rÞdrx dry drz = p1d ðrx Þdrx · p1d ry dry · p1d ðrz Þdrz (2:24)

In this simple treatment, the three-dimensional end-to-end distance is additive of


its three components:

 2  2  2 Nl2
r = Nl2 = hrx i2 + ry + hrz i2 · hrx i2 = ry = hrz i2 = (2:25)
3
With that, we can formulate an expression for a three-dimensional random walk as
follows:
56 2 Ideal polymer chains

nð~rÞ 3 3=2
− 3r2
p3d ð~
rÞ = P = · exp (2:26)
r n ð~r Þ 2π <r 2> 2 <r2 >

In this equation, nð~ rÞ is the number X of walks (or polymer chains) with a given dis-
placement (or chain length) ~ r, and r
nð~
rÞ is the total number of walks (or chains).
Let us now examine how a graphical representation of eq. (2.26) looks like. As an
example, a three-dimensional random-walk-type probability distribution of the end-to
-end distance for the example of a PE chain with N = 20,000 repeating units is shown
by the green curve in Figure 24. It visualizes the probability to have an end-to-end
vector ~r with a given length j~
rj and a given direction. This corresponds to the probabil-
ity of finding the second chain end in a certain specific volume element at a distance

Figure 24: Distribution of end-to-end distances of an ideal polyethylene coil composed of


N = 20,000 repeating units. The green curve is the outcome of a three-dimensional random-walk
statistical treatment (eq. (2.26)). Here, the most likely trajectory, or end-to-end vector, ~ r, is zero.
This corresponds to the likelihood of finding the second chain end exactly in a specific small
volume element in space if the first chain end is placed in the origin, as illustrated in the lower left
sketch. The blue curve filters off all directional dependences by multiplication with a sphere
surface area (eq. (2.27)), thereby reflecting the distribution of end-to-end vector lengths without
caring about into which direction they point. This corresponds to the likelihood of finding
the second chain end localized anywhere on a spherical orbit with radius j~ r j around the first end
that is again placed in the origin, as illustrated in the lower right sketch. (Note that the green p(r)
is actually of type pð~
r Þ, whereas the blue p(r) is actually of type pðj~
r jÞ. For common representation
on the same simple r-axis, they are both denoted in a simplified form as p(r) on both ordinates.
Their mathematical difference is apparent from the different physical units, though).
2.4 The distribution of end-to-end distances 57

and direction ~r from the origin if the first chain end is placed right there. This proba-
bility is maximal at an end-to-end vector of zero, as the random-walk statistics has
given us the highest likelihood of obtaining that very displacement,23 whereas the
probability distribution for other end-to-end vectors or displacements drops in the
form of a Gaussian bell curve.24
In elementary statistics, the likelihood of an event is proportional to the frequency
of its occurrence in a large ensemble. Hence, the Gaussian probability distribution
that we have just found is directly related to the Gaussian distribution of the monomer
segmental density from Section 2.3, which is also highest at the center of the polymer
coil. Due to that identity, ideal polymer chains are often named Gaussian chains.
We can filter off the directional dependence of the random walk by multiplying
it with the surface area of a sphere of radius r:
3=2
3 − 3r2
p3d ðj~
rjÞ = 4πr2 · exp (2:27)
2πhr2 i 2h r 2 i

That way, we generate a function that expresses the probability distribution of end-
to-end vector lengths, independent of their direction.25,26 The blue curve in Figure 24

 Here we make a simplifying approximation. Actually, only the one-dimensional random walk
has a most likely displacement of zero. Two-dimensional and three-dimensional random walks, by
contrast, have not. Of course, we may assume a two- or three-dimensional random walk to be com-
posed of a superposition of two or three independent one-dimensional walks, and of course, indi-
vidually, those indeed have a most likely displacement of zero, but to have an allover displacement
of zero also in the superposition, all of them must get to zero together. However, as the likelihood
for other displacements is also not too small, especially for small displacements, chances are high
that in two- or three-dimensions at least one of the two or three overlaid walks will actually not
show a zero displacement. So, two- and three-dimensional random walks actually do not have a
maximal likelihood of an overall displacement of zero. For the sake of keeping our treatment sim-
ple, though, we ignore that and simplify the three-dimensional displacement to be maximal at
zero, too, like the one-dimensional displacement.
24 Equation (2.26) is similar to an equation that you (hopefully) know from elementary physical
chemistry: the velocity distribution in the kinetic theory of gases, pð~ vÞ. That distribution has the
form of a Gaussian bell curve as well, as it essentially reflects the Boltzmann distribution of the
(kinetic) energies of the gas particles, which has its maximum at zero.
25 Again, something similar is known to you from the kinetic theory of gases, where pð~ vÞ turns into
pðj~
vjÞ, the Maxwell–Boltzmann distribution, by the same mathematical operation.
 Even more, you may have encountered a similar concept earlier in yet a different field of ele-
mentary physical chemistry already. In quantum mechanics, the square of the wave function of the
s-orbital is considered as a measure of the likelihood of finding the electron in a certain point in
space. This likelihood is maximal at a distance of zero from the nucleus and then drops when mov-
ing to larger distances. Just as with the ideal-polymer case treated here, in this problem, though,
the directional dependence is not of so much interest than the radial-distance-dependence. To ac-
count for that, the square of the wave function is multiplied with a spherical surface area of radius
r to obtain a function that gives the likelihood of finding the electron at a certain distance from the
nucleus, independent of the direction. Whereas the first function (the square of the wave function)
58 2 Ideal polymer chains

reflects this relation for the same PE compound considered above. It visualizes the
probability to find any end-to-end vector of length j~
rj independent of its direction,
that is, any end-to-end vector on a spherical circumference with radius r.

2.5 The free energy of ideal chains

We now have a formula at hand that reflects how many microconformations, charac-
terized as three-dimensional random walks, lead to a certain chain macroconforma-
tion, characterized by its mean-square end-to-end distance. This is information about
the structure, about the shape of the polymer chain. In this book, however, we have
set out to build structure–property relations. So, is it possible to use the structural in-
formation we have obtained and extract from them information about the polymer’s
properties? Would it be possible, for example, to calculate thermodynamic quantities
from them? As it turns out, this is feasible: we can use statistical thermodynamics,
which derives macroscopic thermodynamic quantities from the likelihood of micro-
scopic states of a system. In statistical thermodynamics, Boltzmann’s entropy formula
connects the entropy, S, to the number of possible microconfigurations, W, as S = kB
ln(W). When we insert the probability distribution of the end-to-end distances into
that formula, we can derive an expression for the entropy, S, of an ideal chain:

3kB r2
S = S0 + kB ln W ðN,~
rÞ ffi S0 − (2:28)
2h r 2 i

where S0 is the entropy at an end-to-end distance of zero. We see that the total en-
tropy, S, is maximal (with a value of S0) for an end-to-end distance of r = 0, because
the probability of obtaining that distance in a random walk is maximal (green curve
in Figure 24).
From this, we can further formulate the free energy, F, of an ideal polymer chain:

3kB Tr2
F = U − TS = F0 + (2:29)
2h r 2 i

where U is the internal energy and F0 is the free energy at an end-to-end distance of
zero. In the case of an ideal polymer chain, U is independent of the end-to-end dis-
tance, because we imagine the chain segments to have no interactions, so energy
doesn’t care about their spatial arrangement and microconformations. This means
that any change of the ideal-chain free energy stems from entropy alone.

drops with the radial coordinate, the second function (the sphere surface-area function) increases
as a power law (with a power of 2). The product of both functions first raises, then reaches a maxi-
mum, and then drops with the radial coordinate, with the intermediate maximum denoting the
atomic radius.
2.6 Deformation of ideal chains 59

2.6 Deformation of ideal chains

2.6.1 Entropy elasticity

When we slightly stretch an ideal polymer chain, the most feasible microscopic pro-
cess of accounting for that deformation is decoiling of the chain, that is, transforma-
tion of local gauche or cis conformations into trans conformations, as this does not
cost any significant amount of energy (see Figure 14). With that, though, we reduce
the number of possible microconformations that realize the coil’s macroconformation,
thereby forcing it into an entropically more unfavorable state. As a result, as soon as
the deformation cedes, the chain will relax back to the most probable coiled structure.
This phenomenon is based on entropy and is therefore called entropy elasticity.
When we differentiate the free energy with respect to rx, we can calculate the
force needed for a deformation in the x-direction:

~ ∂F ðN,~
rÞ 3kB T 3kB T
fx = = 2 ·~ rx = ·~
rx (2:30)
∂rx hr i Nl2

The latter formula tells us how much force is needed to stretch a chain by a distance
rx. Its form is that of Hooke’s law (f = κ·x), which connects the force necessary to
achieve a certain extent of deformation of a body, f ⁓ x, by its spring constant, κ, a
fundamental material property expressing how good mechanical deformation en-
ergy can be stored and released, and hence, how much the material deforms upon
application of a given force. In the case of an ideal polymer chain, we can consider

ð3kB T Þ Nl2 to be an entropic spring constant. As this quantity only contains struc-
tural information, namely, the number of segments and the segmental length, we
have established another structure–property relation.
Strictly speaking, eq. (2.30) and the concepts we have derived from it are only
valid for Gaussian chains in terms of the Kuhn model, that is, chains with a free seg-
mental bonding angle φ. How would this picture change when we would consider it
for the freely rotating chain that has a fixed segmental bonding angle φ? This ques-
tion can be answered on two levels. On a conceptual level, the freely rotating chain is
already more ordered than a Gaussian chain. It therefore already has a lower entropy,
such that a further loss of entropy would not have such a significant influence,
which means that these chains should be easier to stretch. On a mathematical level,
we would have to consider the characteristic ratio C∞ as a further factor in the de-
nominator of the entropic spring constant. This would result in a smaller entropic
spring constant, which would also mean that freely rotating chains are easier to
stretch.

The entropic spring constant 3kB T Nl2 has a temperature dependence in its nu-
merator. According to this dependence, it will rise with increasing temperature,
thereby making it harder to deform an ideal polymer chain at higher temperature.
This effect can be understood because entropy in thermodynamics always occurs
60 2 Ideal polymer chains

along with temperature in the form of TΔS. Hence, the reduction of entropy upon
chain stretching is even more pronounced at higher temperature, making it more
unfavorable and therefore harder to realize.
So far, with eq. (2.30), we operate at a single-chain level. When we consider an
ensemble of n chains, the essence of eq. (2.30) still holds for each single chain,
such that the total force needed to deform all n of them is just n times that of a sin-
gle chain, so a factor n will enter the right side of eq. (2.30). Furthermore, if we nor-
malize the force to an area, this translates into a pressure, p, on the left side of the
equation, whereas the length l in the denominator on the right side merges with
that newly introduced area to a volume. We then end up with an expression strik-
ingly close to the ideal gas law: p = nkBT/V. Once more, we have discovered astonish-
ing similarity between the physics of the ideal polymer chain and the ideal gas. For
our present discussion on entropy elasticity, this is straightforward to understand.
According to the ideal gas law, the pressure of an ideal gas rises if we compress it at
constant temperature. Why is that? Energetically, the point-like gas molecules do not
care about their mutual distance, and with that, about the volume we give them, be-
cause they do not have interactions, neither attractive nor repulsive. Entropically,
however, a reduction of the system’s volume reduces the number of possibilities for
the molecules to arrange themselves in space. In short: a reduction of the volume re-
duces the freedom of the gas molecules. This comes along with an entropic penalty,
which translates into an increase in free energy, and thus, to a backdriving force just
like the one according to eq. (2.30), which translates into a pressure if we normalize
it to area. The same is seen by a rise of the stress, σ = f/A, in a polymer sample that is
subject to stretching, which is also caused by the loss of conformational freedom,
and hence, the entropy penalty that comes along with that.
In contrast to the entropic origin of elasticity of ideal polymer chains and ideal
gases, the deformation of a classical solid such as a wire of metal is fundamentally
different. Here, upon deformation, the metal atoms are lifted out of their equilibrium
positions in the crystal lattice that correspond to a minimum in their (Lennard–Jones-
type) interaction potential, and as a result, an energetic-based restoring force arises.
Upon increase of temperature, such an energy-elastic body will expand, because the
additional energy enables the atoms to oscillate more extensively around their en-
ergy-minimum positions, and due to the skew shape of the (Lennard–Jones-type) in-
teraction potential well (which has a steeper incline at its left than at its right rim),
this stronger wiggling corresponds to a shift of the average atom positions in the lat-
tice to larger separations. This energetic excitation also allows the material to be de-
formed easier at higher temperatures. Ideal polymer chains, by contrast, will shrink at
elevated temperatures, as an energetically favorable but entropically unfavorable ex-
cess of local trans conformations in the chains will get less dominant in this case, due
to the fundamental coupling of temperature and entropy in the form of TΔS. As a re-
sult, the chain end-to-end distances in a rubber sample under slight load shrink upon
heating in order to achieve their entropically most favorable value of zero, and yet in
2.6 Deformation of ideal chains 61

Figure 25: Overview of the different possible modes of deformation for classical solids (metal),
polymers (rubber), and gases. Upon heating, a classical solid and a gas expand, whereas a polymer
such as rubber shrinks. In a classical solid, this is because the atoms in a crystal lattice oscillate
more heavily around their equilibrium positions at higher temperatures, and due to the skew shape
of their Lennard–Jones-type interaction potential well (which has a steeper incline at its left than at
its right rim), this stronger wiggling corresponds to a shift of the average atom positions in the
lattice to larger separations. In a gas, even more simply, the basic equation of state pV = nRT
explains expansion on rise of temperature. Ideal polymer chains, by contrast, shrink at elevated
temperatures, as an energetically favorable but entropically unfavorable excess of local trans
conformations in the chain will get less dominant in this case, due to the fundamental coupling of
temperature and entropy in the form of TΔS; as a result, the chain end-to-end distances in a rubber
sample under slight load shrink upon heating in order to achieve their entropically most favorable
value of zero, and yet in turn, the whole sample specimen shrinks. Upon mechanical deformation,
the atoms of a classical solid are moved out of their potential energy minima, thereby creating a
restoring force based on energy. Deformation of a polymer and a gas, by contrast, reduces the
number of conformations or freedom of arrangement of the molecules in space, respectively, thereby
creating a restoring force based on entropy. Picture inspired by J. E. Mark, B. Erman: Rubberlike
Elasticity: A Molecular Primer (2nd ed.), Cambridge University Press, 2007.
62 2 Ideal polymer chains

turn, the whole sample specimen shrinks. An ideal gas, yet in turn, does not contract
when temperature is increased, but instead, it shows thermal expansion, because its
volume is directly proportional to temperature (pV = nRT). All these various differen-
ces and similarities of the two fundamental examples of entropy-elastic materials,
which are the ideal gas and ideal polymer chains, as well as the classical example of
an energy-elastic material, which is a piece of metal wire, are illustrated in Figure 25.

2.6.2 A scaling argument for the deformation of ideal chains

As an alternative to the lengthy statistical treatment of the entropic elasticity of


ideal polymer chains that we have just discussed in the preceding sections, Rubin-
stein and Colby have introduced a clever treatment based on a blob concept (that
originates from De Gennes). In this concept, the chain is regarded as a sequence of
blobs of size ξ, each containing g segments. Up to the blob length scale, kBT is the
most relevant energy. As a consequence, any external energy is smaller than kBT
inside each blob, it is exactly kBT at the blob scale ξ, and it is larger than kBT above
the blob scale. When the chain is stretched by an external force fx, the segments
inside the blobs are unaffected by this deformation, because on scales up to the
blob scale ξ, the deformation energy is weaker than – or, in other words, screened
by – the ever-present thermal energy kBT. As a result, the deformation is effective
only on larger length scales, where enough deformation energy is cumulated to out-
weigh kBT. In this view, the chain segments inside the blobs always obey the ideal
scaling law

ξ 2 = gl2 (2:31)
 2
which is a blob-scale variant of the basic law R2 = ~ r = Nl2 that we have discussed
in terms of the ideal chain in Section 2.2.1 (eq. (2.4)). In contrast to these ideal statis-
tics on small scales, the stretched length of the entire chain can be viewed as a uni-
directional linear sequence of blobs, as shown in Figure 26.
Mathematically, a unidirectional linear sequence of blobs is expressed as the
number of blobs, (N/g), times the blob size ξ:

N
rx = ξ (2:32a)
g

Replacing g in the denominator by eq. (2.31) transforms this into

Nl2
rx = (2:32b)
ξ
2.6 Deformation of ideal chains 63

Figure 26: Schematic representation of the blob concept. Here, a chain with length rx is viewed to
be composed of conceptual units named blobs of a limiting length scale ξ, each containing g of the
actual monomer segments. At scales smaller than the blob size, any external energy is smaller
than kBT. This means that when the chain is stretched by a force fx, the deformation energy inside
each blob is screened by kBT, and deformation is only effective on length scales longer than ξ.
Picture redrawn from M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, 2003.

That can be rearranged to

Nl2
ξ= (2:33)
rx
Plugging that into eq. (2.32) gives

N 2 l2
g= (2:34)
rx 2
These latter formulae show that at stronger deformation (larger rx), the blobs get
smaller, meaning that the length scale up to which the deformation is screened by
the thermal energy gets smaller: thereby, the blobs of course also get more numer-
ous, because if they get smaller, then more of them are needed to build up the chain.
At maximal deformation (rx = Nl), the blob size is ξ = l (and g = 1), which means at
such extreme deformation, the blobs have shrunken down to the size of the actual
monomer; then, the deformation is notable on all length scales.
When the chain is conceived as such a sequence of blobs and then stretched in
x-direction, the blob sequence gets ordered from being random to being aligned in
that direction. This ordering comes along with a loss of one directional degree of free-
dom per blob, thereby raising the chain’s free energy by that very extent. Hence, the
free energy of stretching in the x-direction, Fx, is one kBT increment per blob:

N rx 2
Fx = kB T = kB T 2 (2:35)
g Nl
This energy is small at weak extent of stretching, where the chain is conceptually com-
posed of just few large blobs (meaning that the deformation energy is so small then
that up to quite large blob-lengthscales, it is screened by the thermal noise kBT),
whereas it is large at strong extent of stretching, where the chain is then conceptually
64 2 Ideal polymer chains

composed of many small blobs (meaning that the deformation energy is then so large
that only on very small blob-lengthscales, it is still screened by the thermal noise kBT).
Differentiation of that free energy to the stretching distance yields the force for defor-
mation, fx, in the limit of moderate stretching:

∂F rx
fx = ≈ kB T 2 (2:36a)
∂rx Nl

This result is qualitatively similar to that from the longer exact derivation in the pre-
ceding section, but with Rubinstein’s and Colby’s blob concept and scaling ap-
proach, it was obtained much quicker and easier. This is a great advantage of
scaling discussions as the one just led: they give results in good semiquantitative
agreement to those obtained from exact and often more complicated derivations,
but they do so in a much simpler and quicker fashion. Another great advantage is
that they make us see conceptual grounds. We may get such an insight by replacing

rx Nl2 in the last equation by eq. (2.33), yielding

∂F rx kB T
fx = ≈ kB T 2 ≈ (2:36b)
∂rx Nl ξ

Thus, we can conceive that the deformation energy is kBT per blob. As shown above,
the blobs get smaller (and then in turn more numerous) as the deformation gets
stronger (see eq. (2.33)). The greatest extent of deformation is if the chain is fully ex-
panded to a rod-like object of length Nl. In that extreme, according to eq. (2.33), the
blobs have shrunken down to the monomer segmental length of l (and accordingly,
following eq. (2.34), the number of monomers per blob is then just 1). In that extreme,
according to eq. (2.36b), the energy for deformation is kBT per monomer.
The reason why the change in scale that we have just used is valid and can still
be described by the same mathematical equations is the fact that polymers are fractal
and self-similar objects, a topic that we will discuss in the following paragraph.

2.7 Self-similarity and fractal nature of polymers

The relations between the mass and the characteristic size of any geometrical object
can be described by scaling laws. A three-dimensional sphere, for example, exhibits
scaling of m ⁓ r3 . A two-dimensional piece of paper exhibits scaling of m ⁓ r2 , and a
one-dimensional piece of wire exhibits scaling of m ⁓ r1 . Generally, any object ex-
hibits scaling according to

m ⁓ rd (2:37)

where d is its geometrical dimension.


2.7 Self-similarity and fractal nature of polymers 65

The same principle is valid for ideal polymers; these obey the basic scaling law
R ⁓ N 1=2 that we have developed in Section 2.2.1. N is proportional to their mass
(m = N · mmonomer); thus, comparison to the general relation m ⁓ rd denotes ideal
polymers to have a dimension of 2. This is a so-called fractal dimension, as it is
different from the geometrical dimension, which is 3 for a polymer in our three-
dimensional world. In the form introduced here, the fractal dimension is that of a
mass fractal, as it establishes a relation between the object’s mass and its size. This
fractality is not limited to ideal polymer chains. As we will see in the next main
chapter, the general scaling law of a real polymer chain is

R ⁓ Nν (2:38)

with ν = 1=dfractal , the Flory exponent.


Table 7 compiles a glimpse on fractal dimensions of various polymer types. In
general, a smaller fractal dimension denotes the object that it belongs to, to be less
dense. If we compare the fractal dimensions of ideal polymer chains, which is 2, to
that of real chains with short-range repulsion, which is 5/3, we see that the latter is
smaller. This means that real chains are less dense than ideal chains. The reason
for that is the short-range repulsion between the monomer segments in a real
chain, which pushes them apart from one another, thereby resulting in coil expan-
sion that comes along with a lower segmental density inside the coil.

Table 7: Overview of the fractality of ideal and real polymer chains with linear
or branched architecture.

Architecture Interactions Spatial dimension Fractal dimension

Linear None Any 


Linear Short-range repulsion  /
Linear Short-range repulsion  /
Branched None Any 
Branched Short-range repulsion  /
Branched Short-range repulsion  

Fractal objects are also self-similar. This concept can be illustrated by the example
shown in Figure 27(A). A two-dimensional square with side length L and area A has
a scaling law for its area of A = L2. A subsquare within that plane with a side length
l and area a has the same basic shape, and it also exhibits an area scaling law of
a = l2. Without information on the actual length scale that we look at, we can there-
fore not distinguish whether we look at the whole object or a subunit of it. This phe-
nomenon is called self-similarity. Many objects in nature are self-similar: clouds,
coastlines, the surface of broccoli, and more. For all these objects, when being shown
a picture of them, you cannot tell whether you see a small or large section of the ob-
ject, as it has the same appearance on different scales.
66 2 Ideal polymer chains

Figure 27: Illustration of the concept of self-similarity. (A) Consider a two-dimensional square with
side length L and area A, and within that, a subsquare with side length l and area a. Both squares
exhibit the same scaling law for their area, but on different length scales: A = L2 and a = l2. These
laws are self-similar. (B) The same principle holds true for polymers: The basic scaling law of an
ideal chain is R ⁓ N1=2 , which has a variant for subsegments of the chain of ξ ⁓ g1=2 . Again, these
laws are self-similar. Picture in (B) is redrawn from M. Rubinstein, R. H. Colby: Polymer Physics,
Oxford University Press, 2003.

The shape of an ideal polymer coil is also self-similar, as shown in Figure 27(B).
Although the shape of the whole chain shown in this figure is not exactly equal to
that of the magnified chain subsection, on average the same random-walk-type se-
quence is seen on these different length scales. With that, also the scaling law of
the ideal chain applies on both these (and further) different length scales. For the
whole chain of N monomers, we have a scaling of R ⁓ N 1=2 . For a subunit of the
chain with just g monomers, we have a similar scaling of ξ ⁓ g1=2 . Rearranging
these two scaling equations yields:

R = N 1=2 l , l = RN −1=2 (2:39a)

ξ = g1=2 l , l = ξg −1=2 (2:39b)

We can combine these latter two equations via their common variable l to obtain
1=2
N
RN − 1=2 = ξg− 1=2 , R = ξ (2:40)
g

This equation has again the same form as the one we started with: it relates the
polymer coil size, R, to the square root of a dimensionless number times an elemen-
tary unit size. In this notation, the chain can be viewed as a sequence of (N/g) seg-
ments of size ξ each, which was just the notation taken for the development of the
blob concept in Section 2.6.1, where ξ was set to be the blob size.
2.7 Self-similarity and fractal nature of polymers 67

Eq. (2.40) describes our chain, which is actually a random walk of N segments
of size l each, as a new conceptual random walk of (N/g) blobs of size ξ each. We
have done something similar before in Section 2.2.4: in the Kuhn concept, we have
also renormalized our actual chain of N segments of size l each as a new conceptual
chain of NK segments of size lK each. Now, the same was done with blobs as the
new conceptual segments. Due to the self-similarity of polymer chains, this kind of
renormalization works with any new scale, as long as we stay on scales larger than
the Kuhn length lK, because below that length, the polymer chain is no longer uni-
versal and self-similar but markedly exhibits its chemical specificity. In general,
such kind of scale transformation works with any object that is self-similar and
therefore scale invariant, meaning that it does not have a natural length scale that
sets its further properties (like its mass or surface area). Scale invariance is a funda-
mental phenomenon in nature, similar to symmetry.
As a summary, in both the Kuhn model and in the blob concept, we renormalize
a chain by a sequence of new conceptual rather than the actual repeating units and
thereby shift the segmental length and number to new scales. The only mathematical
function that allows such renormalization to be performed is power laws, which are
also named scaling laws for that reason. If another kind of function, for example, a
transcendental function27 like R = R0 ln(N/N0), would describe the N-dependence of
R, such rescaling would be mathematically impossible. Again, the reason for this is
the absence of a natural scale (as it would be captured by N0 and R0 in the latter
hypothetic equation) in objects that are scale invariant. Hence, the characteristics of
self-similar and therefore scale-invariant objects such as polymers are always power-
law type. This is the reason why power laws occur all over this textbook; they are
inherent to polymers, as polymers are self-similar and fractal. [One further note: it is
always single power-law terms that describe the relations between the characteristic
quantities of self-similar objects (such as the N-dependence of R of a polymer), but
not sums of them. Imagine, for example, that a sum of two power-law terms would
describe the N-dependence of R in a form like R = A·(Nl) + B·(Nl)2. Then, there would
be coefficients A and B with different physical dimensions; A would be dimension-
less, while B would have the unit m–1. Their ratio A/B would then have a unit of m
and therewith again reflect a natural length scale in the object under consideration.]

 In a transcendental function, the argument is not allowed to have a physical unit; typical exam-
ples are exp, ln, sin, cos, and so on.
68 2 Ideal polymer chains

Questions to Lesson Unit 3

(1) Consider a coil A with 100 and a coil B with 300 chain segments. Which state-
ment is correct?
a. Coil B has a steeper drop in radial density than coil A, because the higher
number of chain segments in coil B results in a more distinct spherical ra-
dial density profile.
b. Coil A is less densely packed in the center than coil B, because shorter
chains are generally less dense.
c. Coil B is less densely packed in the center than coil A, because longer
chains generally expand over a larger space and “pull on themselves.”
d. Coil A shows a wider profile than coil B, because the longer a chain is, the
more coiled it is; hence, B is more coiled than A.

(2) Which statement about the random walk is wrong? For a one-dimensional ran-
dom walk ______
a. the probabilities of positions after a certain number of steps are binomially
distributed, that is, they can be predicted by Pascal’s triangle.
b. the probability to come out at a position far away from the origin is the
highest.
c. it is not possible to predict for a single walk at which position it will end.
d. the probability to come out at the origin position is the highest.

(3) What happens to the internal energy when stretching an ideal coil?
a. ΔU < 0
b. ΔU = 0
c. ΔU > 0
d. ΔU cannot be estimated as long as the type of interaction is not known.

(4) The free energy of an ideal polymer chain ______


a. does not depend on the chain’s end-to-end distance.
b. depends on the chain’s end-to-end distance via internal energy due to the
distance-dependent segmental interaction potential.
c. depends on the chain’s end-to-end distance by both a distance-dependent
energy and a distance-dependent entropy term.
d. depends on the end-to-end distance only by entropy due to the probability
distribution of the random walk.
Questions to Lesson Unit 3 69

(5) The entropic spring constant ______


a. is independent of temperature.
b. is dependent on temperature according to T.
c. depends on temperature according to T 1 .
d. depends on the temperature according to T 2 .

(6) An ideal coil ______


a. is also named Gaussian coil, because the probability of the end-to-end dis-
tance distribution according to random walk statistics is Gaussian and
therefore also the segmental density distribution of the coil has the shape of
a Gaussian bell.
b. is also named Gaussian coil, because the segmental density distribution of
the coil is Gaussian.
c. is also named Gaussian coil, because the probability of the end-to-end dis-
tribution is Gaussian due to random walk statistics.
d. is something else than a Gaussian coil.

(7) Which statement about the deformation behavior of ideal polymers is true?
a. Ideal polymers can be deformed better at low temperatures, because the re-
storing force to bring the coil back to its initial position is then lower.
b. Ideal polymers can be deformed less easily at low temperatures, because
the thermal energy is then too low to dislocate the chain segments.
c. Ideal polymers can be deformed better at high temperatures, because en-
tropy generally increases with increasing temperature, and thus the coil re-
quires more space.
d. Ideal polymers behave analogously to metallic solids in their deformation
behavior, and so does the temperature dependence.

(8) What is a result of the self-similarity of polymers?


a. The scaling factor of corresponding physical relations corresponds to the
fractal dimension.
b. Corresponding physical contexts always scale linearly in length.
c. Corresponding physical relations are equally valid on different length scales,
that is, they are scale invariant. Mathematically, this manifests itself in the
form of power-law dependencies.
d. Similar polymers can always be synthesized in similar ways.
3 Real polymer chains

LESSON 4: REAL CHAINS


In your elementary classes on physical chemistry, at some point, the ideal gas model
reached its limitations, and it was necessary to expand it such to account for the finite vol-
ume as well as interactions of the gas particles. A similar point is reached when interactions
and the own volume of polymer-chain segments can no longer be disregarded. This lesson
accounts for both and thereby introduces an elementary quantity named excluded volume.
This quantity, in turn, allows a spectrum of different types of solvents to be defined for
polymers.

Now that we have examined and understood the ideal chain model, we can take the
next step and model a polymer chain that resembles the real world more closely.
This leads us to the real chain model. We approach it by readjusting our premises.
So far, we have imagined a polymer chain to consist of monomer segments that
have no volume and show no interactions, neither with one another nor with a sol-
vent in the surrounding. Now, we overcome these simplifications and explicitly
consider that the chain segments have a finite covolume and that they have interac-
tions.28 Our first focus is on appraising how these two effects affect the shape of the
real polymer chain. We may get a first notion on that by pure intuition. If the mono-
mer segments of a real chain have a finite volume, then two of them cannot occupy
the same spot in space, as it was hypothetically possible for an ideal phantom
chain with a self-crossing random-walk-type coil shape, as shown in the left sketch
of Figure 28. A real chain, by contrast, has the shape of a self-avoiding random
walk, as shown in the right sketch of Figure 28. In such a chain, the blocking of
space by each monomer unit to be no longer occupiable by other monomer units
(i.e., the self-avoidance) causes part of the volume to be excluded volume. As a
consequence, the chain has less freedom of arrangement, and the coil must expand.
Indeed, the self-avoiding walk in Figure 28 has a larger end-to-end distance than
the self-crossing one. On top of that, if we also consider that the monomer segments
of a real chain display attractive and repulsive interactions with one another and
with their surrounding environment,29 then the ratio of these monomer–monomer
(M–M) and monomer–solvent (M–S) attractive and repulsive interactions will further

 This expansion of our simplistic model to a more realistic one is conceptually identical to the step
from the ideal gas to the real gas in elementary physical chemistry. Here, the gas molecules are also
considered to have a finite volume and interactions. This is done by inclusion of two parameters for
these two effects, thereby transforming the ideal gas law to the van der Waals equation.
 In that environment, we have either solvent molecules if we consider a polymer solution or seg-
ments of other chains if we consider a polymer melt.

https://doi.org/10.1515/9783110713268-003
72 3 Real polymer chains

Figure 28: Trajectories of a self-crossing random walk, which corresponds to the shape of an ideal
polymer coil, and a self-avoiding walk, which corresponds to the shape of a real polymer coil, in
which part of the volume around each monomer unit is excluded for occupation by other units.

determine the extent of coil expansion; it will also determine the polymer solubility
or miscibility.

3.1 Interaction potentials and excluded volume

As a start, let us examine the interaction potential between two monomer segments
in a chain that have no direct chemical bond to each other (i.e., segments that are
not direct neighbors along the chain). In fact, we do not even need to consider these
monomers to be connected in the form of a chain; instead, it is sufficient to just con-
sider them as molecular entities that have distance-dependent interactions through
space, both attractive and repulsive. A suitable functional form to describe these
interactions is the Lennard–Jones potential; it quantifies the distance-dependent
interaction energy, U(r), of two molecules, in a form similar to the illustration
in Figure 29. This function is generally called to be a 6–12 potential, because the
energetic contribution of the attractive interactions scales with the intermolecular
distance by r–6, whereas the energetic contribution of the repulsive interactions
scales with r–12.30 As a consequence, the repulsive interactions are more influential

 The repulsive part of the Lennard–Jones potential reflects the strong energy penalty that arises
if atoms or molecules are brought into contact so close that their occupied orbitals start to overlap;
this is basically what the Pauli principle expresses, after which two electrons cannot be identical in
all their quantum numbers (actually, this is expressed by a phenomenon named exchange interac-
tion in quantum mechanics). That strongly repulsive energy has a distance dependence that is actu-
ally not necessarily of type r–12; we could also use r–11 or r–13 for it. The r–12 exponent is chosen
3.1 Interaction potentials and excluded volume 73

Figure 29: Effective Lennard–Jones interaction potential, U(r), between two molecules with
attractive effective interaction. The optimal distance, at which we have the lowest U(r) value,
depends on the interactions between the two molecules themselves (shaded circles) and with the
surrounding medium (white circles). Picture inspired by M. Rubinstein, R. H. Colby: Polymer
Physics, Oxford University Press, 2003.

at short distances, where they contribute positive large U(r) values. At longer dis-
tances, by contrast, the attractive interactions dominate and contribute negative
U(r) values (that have a greater absolute magnitude than the small positive ones

somewhat arbitrarily, because it is an even number that makes a steep distance dependence. The
attractive part of the Lennard–Jones potential is more substantiated. It has its cause in electrostatic
interactions. You know a particularly strong representative of this type of interaction from your ele-
mentary chemistry education: the ionic bond, held together by Coulomb interaction. This type of
bond, though, is not meant here. But there are three kinds of related electrostatic interactions that
form the attractive part of the Lennard–Jones potential, all invloving electrically neutral molecules
subject to dipole moments. The first kind concernes molecules with a permanent dipole moment,
which is present if their bonds are covalent but polarized; an example is HCl. This dipole–dipole
interaction, also called Keesom interaction, scales with r–6. A second type of such interaction is
given if a dipolar species induces a complementary one in a non-polar neighboring molecule, so
that these dipole moments can then be paired. This interaction also scales with r–6; it is called
Debye interaction. A third type concerns completely neutral molecules: even those can have elec-
trostatically induced attractive interactions. This is due to so-called correlated quantum fluctua-
tions. Fluctuations exist for almost all quantities, including - for example - the dipole moment of an
argon atom. If these were purely random, the dipolar forces would cancel out on average. But this
is not the case, because if two or more dipole moments are aligned in pairs or groups, this lowers
the energy, and therefore such an alignment becomes favorable. This indirect mechanism leads to
attraction of the molecules; it is called London dispersion interaction. The strength of this kind of
interaction increases as the polarizability of the molecules increases, and the polarizability itself
increases with the number of electrons in the shell. This is why argon has a higher boiling point
than helium, although both are completely non-polar noble gases. The distance-dependence of the
London dispersion interaction is again an r–6 law. The Keesom, the Debye, and the London interac-
tion are summarized as van der Waals interactions.
74 3 Real polymer chains

contributed by the repulsive interactions there). The overlay31 of both causes a po-
tential well, in which a maximally negative U(r) value denotes the most favorable
equilibrium distance. At infinite distances, the potential levels off toward zero, be-
cause there the molecules are too far apart to “see each other”.
In a real polymer system, we need to consider the attractive and repulsive interac-
tions for two species: the monomer (M) and the surrounding medium, that is, the sol-
vent (S). We therefore have to take into account M–M and M–S interactions, both
having attractive and repulsive contributions each. To simplify the discussion, we dis-
cuss effective M–M interactions, in which we incorporate both the M–M and M–S con-
tributions. We do this by redefining repulsive M–S interactions to be just the same as
attractive M–M interactions, because both have the same effect: the monomers prefer
to stay in closer proximity to one another than to the solvent. The same works for the
opposite case: attractive M–S interactions just act like effectively repulsive M–M inter-
actions, as in both cases, the monomers prefer to be further apart from one another
than from the solvent. The resulting distance-dependent effective M–M interaction po-
tential has a Lennard–Jones-type appearance again, as depicted in Figure 29.
Based on this premise, we can delimit three boundary cases:
1. In most cases, M–M contact is favored over M–S contact, so we have attractive
effective M–M interactions and a minimum in U(r), as depicted in Figure 29.
This is due to the perfect structural match of two monomer units M to one an-
other, whereas that of M and S is not as perfect, but at most just similar. As a
result, whatever kind of interactions M can undergo (be it hydrogen bonding,
random-fluctuating, induced, or permanent, dipolar interaction, or whatever),
it will find a perfect partner to do so in another M, and a less perfect (at most
similarly good, but never as perfect) partner to do so in S, such that M–M is
more preferable than M–S.32 In most polymers, the M–M interactions are indeed
van-der-Waals-type dipolar forces, which comes along with effective M–M in-
teraction energies in the order of kBT; this magnitude quantifies the depth of
the well in the effective M–M interaction potential U(r).
2. In the second boundary case, the M–M interactions are equal to the M–S inter-
actions; this situation is encountered if M and S have practically the same
chemical structure, thereby allowing them to establish the same interactions
equally fine with one another or with each other. Consequently, the effective M–M

31 This is done by adding up the two parts as a sum, whereby the attractive one comes along with
a negative sign, as it lowers the total energy, whereas the repulsive one comes
 along with a positive
sign, as it increases the total energy. The exact formula is UðrÞ = U0 + ε ðrre Þ − 12 − 2ðrre Þ − 6 . In this
equation, re is the equilibrium distance at which the potential has its minimum, and ε is the depth
of the energy well at that separation. Note that the attractive term needs a numerical prefactor of 2,
because only then will the potential have a value of –ε at the equilibrium distance re along with a
potential minimum (i.e., a zero first derivative) at that distance.
 This general principle is reflected in the proverb “birds of a feather flock together”.
3.1 Interaction potentials and excluded volume 75

interaction is zero. The U(r)-potential then does not exhibit any potential well, but
only the hard sphere repulsive upturn in the limit r ! 0 (where U(r) ! ∞).
3. In the third boundary case, the M–M interactions are disfavored over the M–S in-
teractions. This is a rare special-case scenario that we may have in polyelectrolytes
in which each monomer unit carries a like charge, such that the monomer units
repel each other.33 Again, no potential well exists in U(r), but by contrast, an addi-
tional repulsive term is added on top of the inherent hard-sphere repulsion.

In the next step, we calculate the probability p to find two monomers at a certain
distance r. This is simply done by deriving a Boltzmann term from the interaction
energy potential U(r):34

− U ðr Þ
p ⁓ exp (3:1)
kB T

A graphical representation of this term is shown in Figure 30(A) for the case of at-
tractive effective M–M interactions. We can see that the probability to find two
monomers at very short distances is zero, which corresponds to the hard-sphere re-
pulsion term of the interaction energy potential, where U(r) ! ∞. The highest prob-
ability can be found at the most favorable distance r that corresponds to the
minimum in the potential well. In this region, U(r) has values of around –kBT, which
translates to values of the Boltzmann term in the range of about e (=2.718). At very
long distances, the probability is independent of r, because there are no operable in-
teractions between the two monomers at these large separations. Here, U(r) has a pla-
teau value of zero, such that the Boltzmann term has a plateau value of one.
When we normalize the Boltzmann expression such that the limit for r ! ∞
equals zero rather than one, we generate an auxiliary function called the Mayer
f-function, as shown in Figure 30(B):

− U ðr Þ
f ðrÞ = exp −1 (3:2)
kB T

By integrating over the Mayer f-function, we can calculate the excluded volume,
which corresponds to the area underneath the curve, shown in gray in Figure 30(B):

 We may also have such a case if there is specific attractive M–S interaction, which manifests
itself as if there were repulsive M–M interaction. This can be the case if the monomer and the sol-
vent are somehow mutually complementary, for example, if the one carries hydrogen-bonding
donor sites, whereas the other carries hydrogen-bonding acceptor sites, thereby being able to form
heterocomplementary bonding with one another that does not work among themselves alone.
34 In general, the likelihood (and therewith the population or frequency of occurrence) of a state
or situation
 of interest with a known energy U is estimated by the Boltzmann distribution:
nU −U
n = exp kB T ; the same principle holds for our situation of interest, which is the likelihood of hav-
ing a certain distance r between two monomers if that comes along with an energy of U(r).
76 3 Real polymer chains

Figure 30: (A) Probability to find two monomers at a specific distance, r, for the case of attractive
effective monomer–monomer (M–M) interactions. At short distances, the repulsive M–M
interactions are so strong that the probability to find two monomers at that separation is
practically zero. In contrast, the probability is highest at the U(r)-potential minimum, while it is
independent of r for r ! ∞. (B) Plot of the Mayer f-function; it normalizes the Boltzmann term
shown in Panel (A) to a value of zero at r ! ∞. From that functional form, the excluded volume can
be calculated by integrating the total area underneath the graph, shaded in gray. Picture redrawn
from M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, 2003.

ð ð
ve = − f ðrÞdr3 = − 4πr2 f ðrÞdr (3:3)

The excluded volume quantifies the space that each chain segment blocks in its sur-
rounding due to (i) its own volume and, on top of that, (ii) the M–M interactions; if
these interactions are repulsive, the blocking of (i) is further exacerbated, whereas if
these interactions are attractive, the blocking of (i) is attenuated. The repulsive term
in the potential U(r) at distances shorter than the equilibrium distance at which the
potential has its minimum, r < re, has a negative contribution to the integral over
the Mayer f-function, which translates into a positive contribution to the excluded
volume. By contrast, the attractive term at r > re has a positive contribution to the
integral over the Mayer f-function, and hence, a negative contribution to the excluded
volume. The example in Figure 30(B) shows a situation in which the attractive and
repulsive parts largely balance each other, leading to an excluded volume close to
zero. This is a very special state, named the Θ-state, in which the chain displays a
quasi-ideal conformation. That special state is very important in the field of polymer
physics, and we will treat it more deeply in the following.
3.2 Classification of solvents 77

3.2 Classification of solvents

Based on what we have learned about M–M and M–S interactions, we can compile
a list that classifies solvents by the extent of the resulting excluded volume ve that a
chain has in them.
When the M–M interactions are equal to the M–S interactions, the solvent is
called to be an athermal solvent. This terminology is because when the interac-
tions are the same, so is their temperature dependence, such that any change of
temperature will have no effect on the effective M–M interactions. Such a solvent is
structurally identical to the monomeric repeating unit of the polymer; a prime ex-
ample is ethylbenzene for polystyrene, which is practically equal to polystyrene’s
repeating unit. In an athermal solvent, there are no attractive effective M–M interac-
tions, leaving only the hard-sphere repulsion at short distances (r ! 0) in the effec-
tive M–M interaction potential U(r) (where U(r) ! ∞). As a result, there is no
positive contribution to the integral over the Mayer f-function, and hence, no nega-
tive contribution to the excluded volume. The outcome is a maximally positive ex-
cluded volume; it is equal to the covolume of the monomer segments: ve = l3.
In a good solvent, the M–M interactions are slightly more favorable than the
M–S interactions. This leads to a small well in the effective M–M interaction poten-
tial U(r), which somewhat compensates the inherent M–M hard-sphere repulsion.
The excluded volume, thus, will still be positive, but smaller than in the athermal
case: 0 < ve < l3. A typical example for a good solvent is toluene for polystyrene.
A very special case is the Θ-state, which is present in a Θ-solvent. In that state,
there are quite attractive effective M–M interactions,35 leading to a pronounced well
in the interaction potential U(r), which just exactly balances the hard-sphere M–M
repulsion at short distances in the potential, as indicated in Figure 29(B). At such bal-
ance, the negative and positive contributions to the integral over the Mayer f-function
cancel each other. For the coil, it then appears as if neither the hard-sphere repul-
sion, i.e., neither the monomer co-volume, nor the effective M–M interaction was
present—just like it is in an ideal chain made of pointlike and interaction-less seg-
ments. Hence, at that constellation, the excluded volume is zero, ve = 0, and the
chain is in a (pseudo-)ideal state36 and adopts the shape of a Gaussian coil with ran-
dom-walk conformation. The Θ-state is therefore beloved by theorists, as it allows
them to model the coil conformation by simple ideal Gaussian statistics. Experimen-
talists, by contrast, do not love the Θ-state as much, because it is highly temperature

 This means that there are quite repulsive M–S interactions, meaning that the solvent S is quite
dissimilar to the monomer repeating unit M.
 A truly ideal state is one without any interactions. A pseudo-ideal or quasi-ideal state is one
with attractive and repulsive interactions at balance.
78 3 Real polymer chains

dependent and can only be present at a special temperature, the Θ-temperature.37 This
temperature marks the borderline to the nonsolvent state, such that even a slight
change of temperature into the wrong direction will cause precipitation of the polymer,
requiring tedious redissolution and re-equilibration before experiments can be con-
ducted. A famous example of a Θ-state is polystyrene in cyclohexane at TΘ = 34.5 °C.
In a bad solvent, the M–M interactions are much more favorable than the M–S
interactions, leading to strong effective M–M attraction. This causes a strong mini-
mum in the effective M–M interaction potential U(r) and a negative excluded vol-
ume in the range of –l3 < ve < 0. An example is ethanol for polystyrene.
In an even more extreme case, the nonsolvent, the M–M interactions are so
much more favorable than the M–S interactions that all solvent is expelled from the
polymer coil and the excluded volume becomes maximally negative: ve = –l3. An
example is water for polystyrene.
The latter two cases cannot be realized in practice, because both actually lead
to nondissolution and can therefore only be studied by computer simulations. They
do, however, have practical applicability for preparatively working polymer chem-
ists: turning a good or a Θ-solvent into a bad or nonsolvent, for example, by suit-
able change of temperature or by addition of a bad or nonsolvent excess, can serve
to separate polymers from a mixture by precipitation, which is an easy means of
polymer purification.
To complete the listing above, note that another special case is the one where
the M–M interactions are disfavored over the M–S interactions. We may have that
situation in polyelectrolyte solutions in which each monomer unit carries a like
charge, such that the monomer units repel each other. We may also have such a
case if there is specific attractive M–S interaction, for example, mutual heterocom-
plementary hydrogen bonding, which manifests itself as if there were repulsive
M–M interactions. In that situation, just like in the athermal case, no well exists in
the interaction-energy potential U(r), and on top of that, there is an additional re-
pulsive term added to the inherent hard-sphere repulsion. As a result, the excluded
volume is even greater than in the athermal case: ve > l3.

 Again, we can find an analogy to gases: a real gas differs from an ideal one in a sense that there
are attractive and repulsive interactions between the gas particles and that the gas particles have a
finite covolume, which is in fact nothing else than the strongly repulsive low-distance branch of
the interaction-energy potential. Both these interactions are quantified by two parameters, a and b,
in the equation of state, through which the ideal gas law turns into the van der Waals equation. At
a special temperature, the Boyle temperature TBoyle, however, these interactions are at balance,
such that the van der Waals equation turns back into the ideal gas law. Hence, the Boyle tempera-
ture for a gas is analog to the Θ-temperature for a polymer–solvent system.
3.3 Omnipresence of the Θ-state in polymer melts 79

3.3 Omnipresence of the Θ-state in polymer melts

The Θ-state cannot only be realized in solution at a specific temperature, but it is


also always present in polymer melts, at any temperature. This can be understood
by the following line of thought: consider a polymer melt in which one chain is
somehow different in color, but structurally identical to the others. We can call this
a “blue” chain in a matrix of “black” chains, as shown in Figure 31. Due to the
chains’ structural identity, this is an athermal state: M–M interactions are the same
as M–S interactions, as the segments of the black “solvent” polymer chains are of
the same kind as those of our blue “dissolved” chain. The blue chain, thus, has a
strong tendency to expand. However, the black chains all have the same tendency.
Thus, all chains in the system want to expand at the same time, and as a result, no

Figure 31: In a polymer melt, the excluded volume interactions in a chain of interest, here shown in
blue color, are screened by overlapping segments of surrounding chains, here drawn in black
color. Picture redrawn from P. G. De Gennes: Scaling Concepts in Polymer Physics, Cornell
University Press, 1979.
80 3 Real polymer chains

chain can actually do so. The expansion tendencies all balance each other, and the
chains therefore all exhibit their unperturbed ideal conformation. In other words,
the excluded volume interactions of each coil are screened by overlapping seg-
ments of other coils.

Questions to Lesson Unit 4

(1) Which statement is not true for the Lennard–Jones potential?


a. The attractive interactions scale with r6 .
b. The attractive interactions are based on van der Waals forces.
c. The repulsive interactions scale with þr12 .
d. The repulsive interactions constitute a hard-sphere potential at short
distances.

(2) What is meant by “excluded volume”?


a. The volume that is inaccessible to the segments of a real polymer due to a self-
avoidance condition in the chain random-walk shape, which results from the
repulsive part of the effective monomer–monomer interaction potential.
b. The volume that is not occupied by the polymer due to the self-overlapping
random walk of the polymer chain segments.
c. The volume that is not occupied by the polymer due to the ideal random
walk of the polymer chain segments.
d. The volume that results from the sum of the volume of all monomers of the
polymer chain.

(3) Which statement is not true about effective monomer–monomer interactions?


a. Effective monomer–monomer interactions include both the correspond-
ing monomer–solvent interactions and the actual monomer–monomer
interactions.
b. Effective monomer–monomer interactions essentially include only the
corresponding monomer–solvent interactions and “translate” them into
something related only to one species, which is the monomer species.
c. Actual attractive monomer–solvent interactions are understood as effec-
tively repulsive monomer–monomer interactions.
d. Actual repulsive monomer–solvent interactions are understood to be effec-
tively attractive monomer–monomer interactions.

(4) Rank the different solvent classes in ascending order according to the excluded
volume of polymers dissolved in them.
a. θ-solvent, good solvent, athermal solvent, poor solvent, nonsolvent.
b. Athermal solvent, good solvent, θ-solvent, poor solvent, nonsolvent.
Questions to Lesson Unit 4 81

c. Nonsolvent, bad solvent, athermal solvent, good solvent, θ-solvent.


d. Nonsolvent, poor solvent, θ-solvent, good solvent, athermal solvent.

(5) The excluded volume ______


a. leads exclusively to compression of the polymer coil.
b. is larger the worse the solvent is.
c. can be determined by integration over the Mayer f-function.
d. is minimal in an athermal solvent.

(6) Match the given sequences of solvent types to the given properties (from top to
bottom).
1. The polymer shape is a Gaussian coil.
2. The integral over the Mayer f-function is positive but not maximally positive.
3. No effective attractive monomer–monomer interactions are present.
4. Monomer–monomer interactions are slightly more favorable than mono-
mer–solvent interactions.
5. The state in this solvent is opposite to the athermal state.
a. 1. θ-Solvent
2. Poor solvent
3. Athermal solvent
4. Good solvent
5. Nonsolvent
b. 1. Athermal solvent
2. Bad solvent
3. Nonsolvent
4. θ-Solvent
5. Good solvent
c. 1. θ-Solvent
2. Nonsolvent
3. Good solvent
4. Athermal solvent
5. Bad solvent
d. 1. Athermal solvent
2. Good solvent
3. θ-Solvent
4. Bad solvent
5. Nonsolvent

(7) Which statement regarding solvent and resulting state of a polymer solution is
correct?
a. The athermal solvent represents a quasi-ideal state, since here the mono-
mer–monomer interactions are equal to the monomer–solvent interactions.
82 3 Real polymer chains

b. The athermal solvent represents the ideal state since it is independent of


temperature.
c. The θ-solvent represents a quasi-ideal state, since repulsive and attractive
parts of the interaction potential balance each other here.
d. The θ-solvent represents an ideal state since there are no effective interactions.

(8) Which of the following is not a requirement for always having a θ-state rather
than an athermal state in polymer melts?
a. The polymer is surrounded by its own kind allover.
b. The polymer chains mutually block each other from expanding.
c. The monomer–monomer interactions are equal to the monomer–solvent
interactions.
d. The polymer chains all have the same structure.
3.4 The conformation of real chains 83

3.4 The conformation of real chains

LESSON 5: FLORY EXPONENT


The excluded volume introduced in the last lesson makes polymer coils expand in good sol-
vents. In this lesson, you will learn how severe this expansion is, and how the size of a coil
can be captured by a universal scaling law that introduces one of the most elementary pa-
rameters in polymer science: the Flory exponent.

3.4.1 Coil expansion

To quantify the size difference of a real polymer chain compared to an ideal one, an
expansion factor α is introduced:
1=2 1=2
hRg 2 ireal ≈ α hRg 2 iideal (3:4)

This factor has values above one, that is, α > 1, for athermal and good solvents, in
which the polymer coil expands. It is exactly one, which means, α = 1, for a Θ-
solvent, in which the coil has the same size as an ideal polymer chain. For bad or
nonsolvents, the expansion factor is smaller than one, that is, α < 1.
But how severe is this expansion? Let us consider polyethylene with NK = 50
Kuhn-segments as an example. In Chapter 2.2.8, the characteristic ratio of this
polymer was named to be C∞ = 6.87 (see also Figure 17), so we get a length of
lK = C∞ l = 6.87 · 0.154 nm = 1.06 nm for each Kuhn-segment.
At strong attractive effective M–M interactions, that is, at non-solvency condi-
tions, the polymer is collapsed to a dense globule with volume V ≈ NK lK 3 = 60 nm3 .
This corresponds to a linear dimension of R = V 1=3 = NK 1=3 lK = 3.9 nm, as shown
in Figure 32(A). Hence, in this collapsed state, the globule’s across diameter is just
about eight times the length of one of its Kuhn-segments (as also seen in Figure 32(A)).
This shows how densely packed the material is in this entity.
With balanced M–M interactions, or zero effective M–M interactions, we realize
the Θ-state. As discussed above, the polymer then has the shape of a random Gaussian
coil and obeys the ideal chain scaling law R = NK 1=2 lK = 7.5 nm, as shown in Figure 32
(B). Hence, in that state, compared to the nonsolvent situation discussed earlier, the
random coil is expanded to be about twice as large as the collapsed globule.
When we put the polymer chain into a good solvent, it has repulsive effective
M–M interactions, which manifests itself in a positive excluded volume ve. The
coil now has the shape of a self-avoiding walk and obeys scaling according to
R = NK 3=5 lK = 11 nm, as shown in Figure 32(C). (This scaling law will be derived in the
next section.) Now, the polymer has a size right in the typical colloidal domain.
84 3 Real polymer chains

Figure 32: Visualization of the different stages of expansion for a polymer with 50 Kuhn-segments.
(A) In its fully collapsed state, the polymer is a dense globule (R = 3.9 nm), whereas (B) at
Θ-conditions, it is a coil with the shape of a random walk (R = 7.5 nm). (C) Effective repulsive
M–M interactions cause the coil to expand to a self-avoiding walk shape (R = 11 nm). (D) At
maximally strong repulsive M–M interactions, the polymer fully expands to a stiff rod (R = 53
nm). Volume and size dependencies on the number of Kuhn segments, NK, and their length lK, for
each regime are shown underneath each image.

At maximally strong repulsive effective M–M interactions, meaning at a maxi-


mally large excluded volume ve, the chain is in its most expanded conformation,
which is a rodlike object, as shown in Figure 32(D). Its size can be calculated ac-
cording to R = NK lK = 53 nm, which is considerably larger than in the other states
discussed before.
We see from the above example that a better solvent quality leads to remark-
able expansion of the polymer coil, from a few single to several tens of nanometers.
This, however, comes at the cost of a lower segmental density in the coil interior.
Throughout our discussion above, we have treated the same polymer chain; we
have not added new segments to it, but only given it more space to arrange itself.
When we do this, we get a greater coil expansion, but, in turn, we naturally reduce
the number of monomer segments per volume. This is shown for the segmental den-
sity of polystyrene in either the θ-state or in a good-solvent state in Figure 33.
Generally, we can appraise the polymer coil’s size for all expansion regimes
by its root-mean-square end-to-end distance according to the following general
scaling law:

hr2 i1=2 ≈ N ν l (3:5)

In this law, ν is the Flory exponent that varies between 1/3 and 1 in our above line of
thought, depending on the solvent quality. It is this power-law exponent that causes
the difference in the size of a polymer at different states of solvency, because in the
3.4 The conformation of real chains 85

Figure 33: Segmental density of a polystyrene coil in a good and a Θ-solvent. The good solvent
expands the coil as compared to the Θ-solvent, leading to a greater occupancy of the volume far
from the coil center, but in turn, to a lower segmental density in its core.

power law (3.5), a large ν empowers a given number of (Kuhn-)segments N to a greater


extent than a small ν does. Due to the mathematical nature of power laws, this differ-
ent empowerment is even more pronounced when the number of segments in the
chain is large. For example, if we consider polyethylene again, but this time with
20,000 repeating units (just as we have done in Section 2.2.8), we have a polymer with
NK = CN∞ = 2899 Kuhn-segments of length lK = C∞ l = 1.06 nm each. In a fully collapsed
state, this polymer has a size of R = NK 1=3 lK = 15 nm, whereas at Θ-conditions, it has a
size of R = NK 1=2 lK = 57 nm (just as already calculated in Section 2.2.8). In a good sol-
vent, it is expanded even further, to a size of R = NK 3=5 lK = 127 nm, and when fully ex-
panded to a rod, the polymer exhibits a size of R = NK 1 lK = 3080 nm (as also calculated
already in Section 2.2.8). The size of this polymer differs even more markedly than the
one in the example of Figure 32 depending on its solvency state, spanning the whole
colloidal domain between single nanometers to a few micrometers.

3.4.2 Flory theory of a polymer in a good solvent

Let us consider an expanded polymer coil composed of N monomer units and a size
R bigger than that of the ideal Gaussian coil, R > R0 = N 1=2 l.
We can estimate the number of monomers that are located inside the excluded
volume of a given first one as follows:
86 3 Real polymer chains

N
ve · (3:6)
R3

Here, ve is the excluded volume of one monomer segment, which is the volume we
want to consider here, and RN3 is the number of segments in the coil per volume of
the coil, in other words, the general segmental density in the coil. In the optimal
case, the expression of eq. (3.6) should equal one. Then, the excluded volume of a
given monomer segment is only occupied by one such segment: by itself. Any addi-
tional monomers that intrude into the excluded volume impose an energy penalty
that can be appraised to be kBT per such unfavorable M–M contact:

N
Fexcl, per monomer = kB Tve (3:7)
R3

If there is only one monomer in the excluded volume (which is the optimal situation),
the resulting energy is kBT, which can be easily “paid” by the ever-present thermal
energy that is exactly kBT. Each additional monomer–monomer contact in the ex-
cluded volume, however, adds another kBT-increment to the penalty. Consequently,
additional energy is required. This is generally unfavorable, such that the system has
a tendency to avoid such M–M contacts, which can be achieved by coil expansion.
Upon such an expansion, R in eq. (3.7) will get larger, such that Fexcl,per monomer will
get lower, which means it is less unfavorable.
In a chain with N monomers, the excluded-volume-interaction energy appraised
by eq. (3.7) is present N times:

N2
Fexcl, per chain = kB Tve (3:8)
R3
Again, the only nonconstant parameter here that can be adjusted is the polymer
coil’s size R. Because R enters the equation in the denominator, the chain can mini-
mize this energy contribution by increasing its size, R ! ∞. The coil, thus, has an
expansion tendency.
On the other hand, coil expansion causes an entropy-based restoring force, cor-
responding to an energy Felast, due to the loss of microconformational freedom
upon coil expansion, as we have learned in Section 2.6. This can be appraised as

3 kB TR2
Felast = F0 + · (3:9)
2 Nl2
Again, the size R is the only adjustable parameter that, in this case, enters the equa-
tion in the numerator. From this, it follows that the chain can minimize this energy
contribution by decreasing its size, R ! 0. The coil, thus, has a contraction tendency.
To calculate the total energy of the chain subject to the two preceding influen-
ces, we have to sum up both the excluded-volume-interaction energy Fexcl and the
elastic energy Felast:
3.4 The conformation of real chains 87

N 2 3R2
F = Fexcl + Felast = kB T ve + (3:10)
R3 2Nl2

To find the coil size with the minimum total energy, we calculate the derivative and
set it zero:

∂F N 2 3R !
= kB T − 3ve 4 + 2 = 0 ) R ⁓ ve 1=5 l2=5 N 3=5 (3:11)
∂R R Nl

With that, we have shown that the N-dependent scaling of R of a coil in a good sol-
vent is

R ⁓ N 3=5 (3:12)

There is just one problem: so far, we always found a length-dependent scaling of


R ⁓ l1 . This is in contradiction to eq. (3.11), where we have found R ⁓ l2=5 instead.
The reason for this discrepancy is a mistake in the above derivation. In eq. (3.9), we
have used ideal-chain scaling of hr2 i = Nl2 in the denominator, even though all our
above line of argument in fact has the purpose to appraise nonideal, expanded coil
dimensions. This mistake causes the erroneous finding of R ⁓ l2=5 . Nevertheless,
our finding of R ⁓ N 3=5 is correct though. Why is that? It is because in view of the
N-dependence, our mistake is compensated by another one. In eq. (3.6), we have
appraised the segmental density in the coil to be uniform (in the form of N/R3),
whereas we know from Section 2.3 that it actually has a Gaussian radial profile.
This wrong estimate of the segmental density in eq. (3.6) compensates the incorrect
scaling in the denominator of eq. (3.9) in view of the N-dependence of R, whereas it
is not cancelled out in view of the l-dependence of R.
The general scaling of R as a function of N in the form of eq. (3.12) can be writ-
ten as

r2 i1=2 ⁓ N ν
R = h~ (3:13)

with ν the Flory exponent. For a polymer in a good solvent, we have just derived it
to be ν = 3/5. For an ideal chain, that is, a chain at θ-conditions, we have shown
earlier (in Section 2.2, eq. (2.5)) that ν = 1/2. For a coil that is fully collapsed to a
dense globule, that is, a chain in a nonsolvent, we have shown earlier (Section 3.4.1)
that ν = 1/3, whereas for the other extreme, a fully expanded rodlike chain, we
have shown that ν = 1.
If we conduct the preceding estimate for the general case of a d-dimensional
space, we have to use Rd in eqs. (3.6)–(3.8) and a numerical factor of d=2 in eq.
(3.9). With that, we get

r2 i1=2 ⁓ N 3=ðdgeometrical + 2Þ
R = h~ (3:14)
88 3 Real polymer chains

 
In this general form, the Flory exponent is ν = 3= dgeometrical + 2 , allowing us to illus-
trate the role of the geometrical dimension, dgeometrical, as follows. In earlier sections,
we have considered the three-dimensional case, d = 3, where the Flory exponent is
3/5. In a two-dimensional situation, d = 2, this is little different. Here, according to
eq. (3.14), the Flory exponent has a value of 3/4, which is larger than 3/5, indicating
that a coil with given N has a greater R in two dimension than in three dimension.
The reason is that there is less freedom for the polymer coil to arrange itself in a 2d-
space than in a 3d-space, because it has only two spatial directions to occupy. Hence,
the coil must expand more in 2d than in 3d to avoid unfavorable M–M contacts. This
trend is even more extreme in a one-dimensional situation, d = 1. Now, according to
eq. (3.14), the Flory exponent is 3/3 = 1. This is because in 1d, the coil has no other
way to avoid M–M contacts than to fully expanding itself into a rodlike object with
R = Nl. By extreme contrast, in a four-dimensional situation, d = 4, the coil has so
much freedom to arrange itself that M–M contacts are generally very unlikely. The coil
can therefore adopt its random Gaussian shape like an ideal chain or a real chain in
the Θ-state. Here, consequently, the Flory exponent is 3/6 = 1/2. (If we further follow
this line of thought, even higher geometrical dimensions such as d = 5, 6, . . . would
denote smaller Flory exponents such as ν = 3/7, 3/8, . . ., indicating that a coil in such
high dimensions would contract. This can be understood on the very same basis as the
discussion just led: the higher the geometrical dimensions, the smaller can be the coil
size without causing unfavorable M–M contacts. Higher dimensions, therefore, allow R
to be small, which minimizes the elastic energy according to eq. (3.9) (where R enters
in the form of R2, irrespective of d), without excessively increasing the excluded vol-
ume interaction energy according to eq. (3.8) (where R enters in the form of R–d, which
means that if d is high, there is lesser need for large R to make that term small).)
In addition to the preceding discussion of the geometrical dimension, we may
also do so for the fractal dimension, a concept that we have introduced in Sec-
tion 2.6.2. With this dimension, we get

R = h~
2
r i1=2 ⁓ N 1=d fractal (3:15)

Hence, the Flory exponent is nothing else than the inverse fractal dimension. When
looking on the compilation of Flory exponents in one to four geometrical dimen-
sions above, we see that the inverse of these exponents, that is, the fractal dimen-
sion, is getting lower and lower. In general, a smaller fractal dimension denotes the
object that it belongs to be less dense. If we compare the fractal dimension in four
geometrical dimensions, dfractal = 1/ν = 1/(1/2) = 2, to that in three geometrical di-
mensions, dfractal = 1/ν = 1/(3/5) = 5/3, and in two geometrical dimensions, dfractal
= 1/ν = 1/(3/4) = 4/3, we obtain a series of smaller and smaller values. This means
that chains in lower geometrical dimensions are less dense than in higher geometri-
cal dimensions. The reason for that is the excluded-volume repulsion between the
monomer segments, which pushes them apart from one another to avoid M–M
3.5 Deformation of real chains 89

contact; this does not need to be as pronounced in higher geometrical dimensions, as


such contact is generally less likely there. A special case is the one of a densely col-
lapsed globular polymer in a nonsolvent; here we have a Flory exponent of ν = 1/3,
corresponding to a fractal dimension of dfractal = 1/ν = 1/(1/3) = 3. This fractal di-
mension matches the geometrical one, thereby denoting a nonfuzzy, dense object,
which a collapsed polymer globule indeed is.

Paul John Flory (Figure ) was born on June ,


, in Sterling, Illinois. He studied at Manches-
ter College and obtained his Bachelor of Science
degree during the time of the great depression,
supporting himself with various jobs at the side.
During that period, his interest in science, partic-
ularly chemistry, was inspired by Professor Carl
W. Holl, who encouraged him to enter graduate
school at Ohio State University in . Flory fol-
lowed this advise, pursued his graduate studies
at Ohio State, and obtained his PhD degree in
. From  to , he worked in several
industrial research laboratories for companies
such as DuPont, Standard Oil, and Goodyear. He
was offered a faculty position at Cornell University
in , where he stayed until . Having trans- Figure 34: Portrait of Paul J. Flory. Image
ferred to and led the Mellon Institute in Pittsburgh reproduced with permission from Stanford
until , he became a full professor at Stanford University Libraries, Department of Special
University until he retired in . One year prior Collections and University Archives (SC0122,
to his retirement, he received the Nobel Prize in Stanford University News Service records,
Chemistry “for his fundamental achievements, Box 90, Folder 48, Paul Flory).
both theoretical and experimental, in the physical
chemistry of macromolecules”. He died on Septem-
ber , , at the age of  in Big Sur, California.

3.5 Deformation of real chains

Just like in Chapter 2, after having made ourselves a mind about the size and shape
of a real chain, we now want to translate that structural information into informa-
tion on properties, specifically, into the elastic properties upon deformation of the
chain. Other than for the simple case of an ideal chain, as treated in Section 2.6,
such a discussion for a real chain will be mathematically challenging. To simplify
it, Rubinstein’s and Colby’s clever scaling argument that has been used already in
Section 2.6.1 can be applied again to conceptually renormalize the polymer chain
within a blob concept. This approach is valid for both ideal (see Section 2.6.1) and
real chains, so both are treated simultaneously below.
90 3 Real polymer chains

The root-mean-square end-to-end distance of a polymer chain is calculated as


1=2
Ideal chain: hr2 i0 = R0 = N 1=2 l (3:16a)

Real chain: hr2 i1=2 = RF = N 3=5 l (3:16b)

Due to the self-similarity of polymer chains, the same scaling also applies to subsec-
tions of the chain that only encompass n monomers:

Ideal chain: r0 = n1=2 l (3:17a)

Real chain: rF = n3=5 l (3:17b)

We now regard a special subsection scale, named ξ. On scales smaller than ξ, the exter-
nal deformation energy is weaker that the ever-present thermal noise kBT; as a result,
on scales smaller than ξ, the chain segments show unperturbed random-walk-type (in
the case of an ideal chain) or excluded-volume-expanded (in the case of a real chain)
conformations, but they do not “feel” any external deformation. By contrast, on scales
larger than ξ, external deformation is effective, as its energy is stronger than kBT there.
Hence, although the chain segments on scales smaller than ξ are not affected by exter-
nal deformation, the whole chain is. It can therefore be conceived as an oriented se-
quence of deformation blobs of size ξ, inside of each being a chain subsegment of g
monomers that is not affected by the deformation, as shown in Figure 35. In this way,
the chain can maximize its entropy even under the external constraint of deformation.
By adaption of eq. (3.17) to the blob scale, the blob size is expressed as

Ideal chain: ξ = g1=2 l (3:18a)

Real chain: ξ = g3=5 l (3:18b)

On length scales larger than ξ, the chain is an oriented sequence of deformation


blobs; as a result, its end-to-end distance can be approximated as the blob size, ξ,
times the number of blobs, (N/g), which we may then rewrite further by using eq.
(3.18) (rearranged to g and then replacing g in N/g) and eq. (3.16) (replacing the then
occurring numerators of type Nl1/ν)

N Nl2 R0 2
Ideal chain: Rf ≈ ξ = = (3:19a)
g ξ ξ

N Nl5=3 RF 5=3
Real chain: Rf ≈ ξ = 2=3 = 2=3 (3:19b)
g ξ ξ

Rearranging these equations yields an expression for the blob size ξ


3.5 Deformation of real chains 91

Figure 35: Modeling of a polymer chain, ideal (upper sketch) and real (lower sketch), subject to an
external stretching force that leads to an end-to-end distance Rf as a conceptual object composed
of blob elements with size ξ. At scales below ξ, the polymer chain does not experience deformation
and displays ideal random-walk-type (ideal chain) or expanded (real chain) conformations. At
scales above ξ, by contrast, in both cases, the polymer is an oriented sequence of blobs, as on
these scales, the external deformation is effective. Picture redrawn from M. Rubinstein, R. H. Colby:
Polymer Physics, Oxford University Press, 2003.

R0 2
Ideal chain: ξ = (3:20a)
Rf

RF 5=2
Real chain: ξ = (3:20b)
Rf 3=2

As we have learned in Section 2.6.1, the free energy of deformation, F, is kBT per
blob. By using eqs. (3.19) and (3.20) to further rewrite, we get
2
N Rf Rf
Ideal chain: Fideal = kB T = kB T = kB T (3:21a)
g ξ R0
5=2
N Rf Rf
Real chain: Freal = kB T = kB T = kB T (3:21b)
g ξ RF

We have also learned that the force needed to deform the chain by a distance of Rf
corresponds to the ratio of the thermal energy kBT and the blob size ξ, which we
can express by eq. (3.20) to get

kB T kB T kB T Rf
Ideal chain: f = = 2 Rf = · (3:22a)
ξ R0 R0 R0
92 3 Real polymer chains

3=2
kB T kB T kB T Rf
Real chain: f = = 5=2 Rf 3=2 = · (3:22b)
ξ RF RF RF

From this expression we realize that the force needed to deform a real chain in-
creases stronger with Rf than the force needed to deform an ideal chain. The abso-
lute force values, however, are always smaller for a real chain due to their bigger
swollen, or “prestretched”, original shape. This fact is visualized in Figure 36. A
general equation for the deformation of a polymer chain can be expressed using the
Flory exponent ν:
1
Rf 1−ν
F = kB T (3:23)
Nνl

To summarize, the preceding scaling discussion has shown that both ideal and real
chains lose conformational freedom upon deformation, which is reflected by their
entropic spring constant kB T =ξ . However, they do so in different ways: an ideal
chain is deformed from its Gaussian coil dimensions, R0, whereas the real chain is
already prestretched due to excluded volume interactions to a bigger size, RF. From
this, it follows that the deformational force for real chains is smaller than that of
their ideal counterparts, albeit it increases more steeply upon deformation.

Figure 36: Force, f, needed to deform an ideal and a real polymer chain by a distance Rf. For an
ideal chain, the required force scales with the deformation distance Rf with a power-law exponent
of one, as expressed by eq. (3.22a), whereas for a real chain, the required force scales with the
deformation distance Rf with a power-law exponent of 3/2, as expressed by eq. (3.22b). Thus, the
force needed to deform a real chain increases more steeply than its counterpart for an ideal chain.
The absolute force values, however, are always smaller for a real chain due to their bigger swollen,
or “prestretched”, original shape. Picture redrawn from M. Rubinstein, R. H. Colby: Polymer
Physics, Oxford University Press, 2003.
Questions to Lesson Unit 5 93

Questions to Lesson Unit 5

(1) Match the following scaling laws with the appropriate polymer–solvent conditions
under which they apply (from left to right):

1 3 1
R ¼ ðNK Þ2 lK R ¼ ðNK Þ5 lK R ¼ ðNK Þ1 lK R ¼ ðNK Þ3 lK

a. No effective monomer–monomer interactions


Repulsive effective monomer–monomer interactions
Maximally repulsive effective monomer–monomer interactions
Maximally attractive effective monomer–monomer interactions
b. No effective monomer–monomer interactions
Maximally attractive effective monomer–monomer interactions
Maximally repulsive effective monomer–monomer interactions
Repulsive effective monomer–monomer interactions
c. Maximally attractive effective monomer–monomer interactions
Repulsive effective monomer–monomer interactions
Maximally repulsive effective monomer–monomer interactions
No effective monomer–monomer interactions
d. Maximally repulsive effective monomer–monomer interactions
Repulsive effective monomer–monomer interactions
Maximally attractive effective monomer–monomer interactions
No effective monomer–monomer interactions.

(2) By what factor is a polymer with a Kuhn segment number of 30,000 in a θ-solvent
larger than a polymer with a Kuhn segment number of 10,000 and the same
Kuhn length under the same conditions?
pffiffiffi
a. 3
b. 31
c. 3
d. 32

(3) By approximately what factor is a polymer with a Kuhn segment number of


30,000 in a good solvent larger than a polymer with the same Kuhn segment num-
ber and length in a θ-solvent?
3
a. 35
1
b. 30; 00010
1
c. 32
5
d. 33
94 3 Real polymer chains

(4) Which statement regarding the energies that are decisive for the resulting expan-
sion of a real coil is correct?
a. The energy for the occupation of the excluded volume and that resulting
from the entropy-elastic restoring force upon expansion are both positive,
which is why the polymer shows a tendency to contract to reach a minimum
total energy.
b. The energy for the occupation of the excluded volume as well as that result-
ing from the entropy-elastic restoring force upon expansion are both posi-
tive, which is why the polymer shows an expansion tendency to reach a
maximum total energy.
c. The energy for the occupation of the excluded volume results in a contraction
tendency of the polymer, whereas the energy resulting from the entropy-elas-
tic restoring force upon expansion results in an expansion tendency. The
overall energy minimum gives the optimal combination and therefore the op-
timal coil size.
d. The energy for the occupation of the excluded volume results in an expan-
sion tendency of the polymer, whereas the energy resulting from the entropy-
elastic restoring force upon expansion results in a contraction tendency. The
overall energy minimum gives the optimal combination and therefore the op-
timal coil size.

(5) The Flory theory provides a correct description of the coil size R of a real
chain ______
a. only with respect to the Kuhn segment number NK , since the theory as-
sumes an ideal chain entropy elasticity and a Gaussian segmental density
profile, and these assumptions provide a correct description with respect to
NK , but not with respect to the Kuhn segment length lK .
b. only with respect to the Kuhn segment number NK , because the theory as-
sumes an ideal chain entropy elasticity and a homogeneous segmental den-
sity, and these two simplifying approximations luckily compensate each
other with respect to NK , but not with respect to the Kuhn segment length lK .
c. with respect to the Kuhn segment number NK and the Kuhn segment length
lK , because it assumes an ideal chain entropy elasticity and a homogeneous
segmental density, and these two simplifying approximations luckily com-
pensate each other completely in the resulting exponents.
d. with respect to the Kuhn segment number NK and the Kuhn segment length
lK , since the dependence of NK compensates the simplifying approximations
from the assumption of an ideal chain entropy elasticity and a homogeneous
segmental density profile, while additionally the dependence of lK results
only from the assumption of an ideal chain and yields a correct exponent 52.
Questions to Lesson Unit 5 95

(6) The Flory theory is also applicable to polyelectrolytes, that is, charged poly-
mers. For this purpose, an additional energy term must be added to the Flory
theory approach. What is valid for this term?
a. The term corresponds to the Coulomb interaction, which contributes to the
contraction tendency of the coil.
b. The term corresponds to the Coulomb interaction, which contributes to the
expansion tendency of the coil.
c. The term corresponds to the van der Waals interaction, which contributes to
the expansion tendency of the coil.
d. The term corresponds to the van der Waals interaction, which contributes
to the contraction tendency of the coil.

(7) Which statement regarding the fractal dimension of polymers is not correct?
a. The Flory exponent is the reciprocal of the fractal dimension.
b. A higher fractal dimension results in a more densely packed coil.
c. The higher the geometric dimension, the lower the fractal dimension.
d. At maximum contraction, the fractal dimension equals the geometric one.

(8) Which statement regarding the deformation of a real chain is correct?


a. The force required for deformation of a real chain is always greater than
that required for the same deformation of an ideal chain.
b. The force required for deformation of a real chain is always smaller than
that required for the same deformation of an ideal chain.
c. A real chain with the same Kuhn length and number of segments is de-
3
formed by a factor N 5 more than the corresponding ideal chain if the same
force f is applied.
d. The force required to deform a real chain increases exponentially with the
Flory exponent.
96 3 Real polymer chains

3.6 Chain dynamics

LESSON 6: CHAIN DYNAMICS


So far, polymer chains were viewed to be static in this book. This chapter will go beyond
that view and introduce two fundamental frameworks to model and quantify chain dynamics:
the Rouse and the Zimm model. Both will show that it takes a certain time before a chain can
move as a whole, whereas below, only parts of it move. This time delimits the scale on which
a polymer is a viscous fluid from the scale where it is a viscoelastic body.

3.6.1 Diffusion

In the last chapters, we have looked at the shape of ideal and real polymer chains;
we have also discussed how their shape changes when they are deformed. In this
chapter, we will focus on the motion, the dynamics, of polymer chains. As a funda-
ment for that, we first recapitulate some elementary physical chemistry. The basic
description of the thermal diffusive translational motion of a molecular or colloidal
object is that of a random walk,38 a concept that we have already discussed in Sec-
tion 2.4.1. The trajectory of such a random walk is displayed in Figure 37.
A characteristic quantity for random walks is their mean-square displacement,
<~R2 > , which is expressed by the Einstein–Smoluchowki equation:39

 Often, the diffusive molecular motion is synonymized as Brownian motion. But note that this is
actually not the same. The Scottish botanist Robert Brown observed that plant pollen floating on a
surface of water show seemingly self-driven random motion. This Brownian motion, however, is
actually a result of the diffusive motion of the molecules in the surrounding medium, as is was cor-
rectly interpreted later on by Albert Einstein and Marian Smoluchowski. In Brown’s experiment,
the plant pollen are hit by the surrounding water molecules, but as the hitting is random, the num-
ber of hits from different directions are not equal at a time per particle. As a consequence, the over-
all momentum transfer at each timepoint pushes the pollen particles into randomly changing
directions, causing their seemingly self-driven random wiggling motion that Brown saw.
 In Figure 37, the displacement of the moving particle corresponds to the distance of the last
arrow of the sequence to the starting point of the walk. The mean-square of it is obtained by averag-
ing over many of such walks, whereby in that averaging, each displacement is first squared to get
rid of the directional dependence. Otherwise, if that was not done, the average would always be
zero, as each displacement would be cancelled by one showing exactly into the opposite direction
in a great ensemble of walks. As a result, the quantity of interest is the mean-square displacement –
the mean over many individual displacements in a squared form. Often, to relinearize the physical
dimension, people further take the square-root, thereby obtaining the root-mean-square displace-
ment. We also do so all over Chapters 2 and 3 when we talk about the root-mean-square end-to-end-
distance of polymer chains.
3.6 Chain dynamics 97

Figure 37: Trajectory of a (two-dimensional)


random walk.

h~
R2 i = hð~
rðtÞ −~
2
rð0ÞÞ i = 2 d D t (3:24)

Here, d denotes the geometrical dimension, and D is the translational diffusion co-
efficient, a quantity that expresses how mobile the moving molecule or particle is.
According to eq. (3.24), D has the unit m2·s–1, thereby quantifying what mean-
square distance the moving object passes per time. The diffusion coefficient can be
calculated by the Einstein equation:

kB T
D= (3:25)
f

This equation relates D to the ratio of the thermal energy that drives the diffusion,
kBT, and the friction that drags the diffusion; the latter is expressed by a friction
coefficient, f, that connects the frictional force to the velocity of the moving object,
~
f = f~
v. The friction coefficient of a spherical object is given by Stokes’ law as

f = 6πηrh (3:26)

Here, η expresses the viscosity of the surrounding medium (this is what actually ex-
erts the friction) and rh denotes the hydrodynamic radius, which is the radius of the
moving object itself plus its solvent shell (and potential swelling medium inside)
that is dragged with it during the motion.
Both the latter equations can be combined to give the Stokes–Einstein equation:

kB T
D= (3:27)
6πηrh

Equation (3.27) is fundamental in physical chemistry, as it relates the size of a mov-


ing molecule or particle, denoted by rh, to its mobility, denoted by D. Together with
the Einstein–Smoluchowki equation (3.24), this allows us to determine how far a
98 3 Real polymer chains

diffusing molecule or particle of given size rh can move in a prescribed time t, or


vice versa, how long it takes to move a given distance R2. This is of elementary rele-
vance in many scientific fields, for example, in the field of drug delivery, when the
time shall be appraised that a drug will need to reach a receptor in a cell or tissue
environment, or vice versa, when it shall be appraised how far at all the drug can
move in a given timeframe. The same question is also relevant in fields like chemi-
cal technology when it comes to appraising the efficiency of heterogeneous reac-
tions, where partners must find each other by diffusion across phase boundaries, or
in chemical engineering, when it comes to appraising how far diffusive smearing
will impair precision in micropatterning or 3d printing.
In the field of polymer and colloid science, a specified form of eq. (3.24) is rele-
vant; this is a form in which eq. (3.24) is rephrased such to express the timescale τ
for the displacement of a moving (macro)molecule or (colloid)particle by exactly its
own size, R:

R2 R2 f
τ= = (3:28)
2d · D 2dkB T
On timescales shorter than this characteristic time τ, the colloidal or polymeric build-
ing blocks of a material cannot move over distances at least corresponding to their
own size, meaning they are practically static. By contrast, on timescales longer than
τ, the material’s building blocks are macroscopically mobile. As a result, τ delimits
the time domain on which a material is a solid from that on which it is a liquid. For
polymeric and colloidal matter, this limiting time is often on experimentally relevant
orders of magnitude, which causes these materials to exhibit both solidlike and
liquidlike appearance, depending on the timescale of observation.
So far, all the above discussion accounts for simple molecules or particles.
When it comes to the dynamics of a flexible polymer coil, however, in addition to
its global motion, we also must consider that it has multiple kinds of coil-internal
dynamics, as it is a large multibody object. To account for this complexity, two dif-
ferent models have been developed: the Rouse model and the Zimm model.

3.6.2 The Rouse model

The Rouse model conceptually describes a polymer chain as a number of N spherical


beads, representing the monomeric units, connected through elastic springs of length
l, representing the bonds between the monomer segments, as depicted in Figure 38.
Each bead is assigned an individual segmental friction coefficient fsegment. Together,
the beads constitute a freely drained coil, which means that only the beads feel fric-
tion with the surrounding medium, but the springs do not. As a result, the solvent
can pass freely through the polymer coil and hit each bead, where it imparts a fric-
tional increment fsegment. The total friction coefficient of the coil is therefore simply
3.6 Chain dynamics 99

Figure 38: Spring-and-bead modeling of a chain in the Rouse model. Picture redrawn from
M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, 2003.

the sum of all these individual frictions: ftotal = N · fsegment. This generates a system
of coupled differential equations from the equations of motion of the beads subject to
drag by the springs and friction by the solvent. If we substitute ftotal in the Einstein
eq. (3.25) by N · fsegment, we get

kB T kB T
DRouse = = ⁓ N −1 (3:29)
ftotal N · fsegment

We now formulate an Einstein–Smoluchowski analog expression for the timescale


to achieve a displacement by exactly one coil size, the Rouse time:

R2 fsegment 2
τRouse = = NR (3:30)
2d · DRouse 2dkB T
This time denotes the upper extreme of a whole spectrum of relaxation times that
we will discuss in more detail below. On timescales longer than τRouse, the coil can
migrate over distances further than its own size; macroscopically, this means that
on these timescales the building blocks of a polymer material can effectively dis-
place against each other, such that the material exhibits flow.
The other extreme of the characteristic relaxation time spectrum is the time it
takes for at least each monomeric segment to be displaced by a distance equal to its
own size, l:

fsegment · l2
τ0 = (3:31)
2dkB T
On timescales shorter than τ0, no motion whatsoever is possible in the coil. On
these short timescales, the material therefore is an (energy-)elastic solid with glassy
appearance.
A polymer chain is a fractal object with size R = N ν l (cf. Section 2.6.2). When we
insert this into the above equation for the Rouse time, we can derive a generalized
expression that relates τRouse to τ0:
100 3 Real polymer chains

fsegment · l2 1 + 2ν
τRouse = N = τ0 N 1 + 2ν (3:32)
2dkB T

The latter equation combines the two characteristic ends of the relaxation time spec-
trum, and thereby delimits three different domains of dynamics. On timescales shorter
than τ0, that is, t < τ0, the polymer does not have sufficient time to move at all, neither
the entire chain nor any of its constituent segments. On this timescale, the polymer
forms an energy-elastic, glassy solid. On intermediate timescales, τ0 < t < τRouse, at
least monomeric segments and sequences of them (i.e., subchain segments) can move
over distances equal to their own size. The time is, however, not yet long enough for
the entire chain to effectively displace itself. On this timescale, the polymer is a visco-
elastic solid. Only on timescales longer than the Rouse time, t > τRouse, the whole poly-
mer coil can move over distances greater than its own size. Now, the material exhibits
flow and is a viscoelastic liquid.

3.6.3 The Zimm model

So far, we have considered the polymer coil to be freely drained by solvent. This
view, however, is not always true. M–S interactions can cause moving polymer
chain segments to drag solvent molecules with them. These, in turn, do the same to
the solvent molecules adjacent to them. This drag spreads from one solvent mole-
cule to another until, eventually, it reaches other segments of the polymer chain.
Hence, even distant segments are coupled to each other through space by hydrody-
namic interactions. As a result, each coil drags the solvent in its pervaded volume
with it on its motion; this trapped solvent is in diffusive exchange with the medium
in the surrounding, but does not drain through the coil. As a result, the coils appear
like solvent-filled nanogel particles, as visualized in Figure 39.
The Zimm model expands on the Rouse model to incorporate such hydrody-
namic interactions. It assumes that the polymer coil and the trapped solvent inside
act together as a united object of size R = Nνl. Substitution of the hydrodynamic
radius, rh, in the Stokes–Einstein equation with this expression yields

kB T kB T
DZimm = ≈ ⁓ N −ν (3:33)
6πηrh ηN ν l

Compared to the Rouse model, the power-law scaling is less steep here: it is just –ν
compared to –1.40 Hence, in the Zimm model, the diffusion coefficient has a less

 Note: for stiff, rodlike chains, with ν = 1, the Zimm-type scaling matches the Rouse-type scal-
ing, as for such a polymer, there is no trapped solvent in it, as it is fully decoiled. So, Zimm- and
Rouse-type dynamics become indistinguishable for this kind of chain.
3.6 Chain dynamics 101

Figure 39: Modeling of polymer coils as microgels entrapping the portion of solvent within their
pervaded volume due to hydrodynamic interactions and dragging it with them on their motion
through the free solvent in the surrounding medium. The trapped portion of solvent is in diffusive
exchange with that in the surrounding, but the coils are not drained by the medium. Picture
inspired by B. Vollmert: Grundriss der Makromolekularen Chemie, Springer, 1962.

pronounced dependence on N than in the Rouse model. The reason is the lack of
solvent draining in the Zimm model. In the Rouse model, viscous drag is exerted by
each bead that constitutes the coil, be it beads on the coil front or in the coil inte-
rior. In the Zimm model, this is done only by the beads (and the trapped solvent
between them) on the coil’s frontal face. Thus, in the Rouse model, any change of
the number of beads directly translates to the friction on the coil motion, and
hence, results in an inverse proportionality of its diffusion coefficient on N. By con-
trast, in the Zimm model, friction is exerted by the coil frontal face only, and the
size of that face scales with the size of the coil as a whole, that is, according to Nν,
so that the diffusion coefficient just inversely scales with N–ν.
As we have done in the Rouse model, we can formulate an Einstein–Smoluchowski
analog expression for the timescale of displacement by exactly one coil size, the
Zimm time:

R2 η 3 ηl3 3ν
τZimm = ≈ R ≈ N ≈ τ0 N 3ν (3:34)
2d · DZimm kB T kB T
Again, compared to the Rouse time, the power-law exponent is smaller here: 3ν
compared to (1 + 2ν).41 Again, this means that there is a weaker dependence of the
longest relaxation time τ on the number of monomer segments N in the Zimm

 Yet again, for stiff, rodlike chains, with ν = 1, the Zimm-type scaling matches the Rouse-type
scaling.
102 3 Real polymer chains

model than in the Rouse model. As a consequence, τZimm is shorter than τRouse. The
reason is the absence of solvent draining, which comes along with less viscous
drag, causing the time its takes for the coil to diffuse a given distance (such as that
of its own size) to be shorter than in the case of more pronounced viscous drag in
the Rouse scenario.

Bruno Hasbrouck Zimm (Figure ) was born


on October , , in Woodstock, New York.
He studied at Columbia University, where he ob-
tained his bachelor’s degree in , his mas-
ter’s degree in , and his PhD degree under
Joseph. E. Mayer in . He then moved across
town for postdoctoral work with Herman Mark
at the Polytechnic Institute of Brooklyn. In
, he transferred to the University of Califor-
nia in Berkeley, where he became an assistant
professor in –. After that, he was the
head of the General Electric research labs in
Schenectady. In , he became a full professor
at the University of California in San Diego. He re-
tired from research in . He is most famous Figure 40: Portrait of Bruno H. Zimm. Image
for his work on light scattering, where he devel- reprinted with permission from
oped the Zimm plot, his extension of the Rouse Macromolecules 1985, 18(11), 2095–2096.
model of polymer dynamics, and his groundbreak- Copyright 1985 American Chemical Society.
ing work on the structure of proteins and DNA.
He died on November , , at the age of 
in La Jolla, California.

3.6.4 Relaxation modes

We have learned in Section 2.6.2 that polymers are self-similar and fractal objects.
This self-similarity is also valid for chain dynamics: a subchain with g segments in
a whole of N segments relaxes like an individual chain composed of just g segments
in total, describable by the same Rouse and Zimm formalism as outlined earlier.
These subchain relaxations are appraised by so-called relaxation modes numer-
ated by an index p. The pth mode corresponds to the coherent motion of subchains
with N/p segments in our whole chain with N segments. At p = 1, coherent motion
of the entire chain is possible, meaning that the entire chain can relax and get dis-
placed by a distance equal to its own size, whereas at p = 2, only each half of it
can move coherently and relax and therefore get displaced by a distance equal to
its own size, respectively. At p = 3, just only each third of the chain can move co-
herently and relax and therefore get displaced by a distance equal to its own size,
and so forth. At p = N, only single monomeric units can relax and get displaced
3.6 Chain dynamics 103

against each other by their own size. Figure 41 visualizes this hierarchy of relaxa-
tion modes. At a time τp after abrupt deformation, all modes with index above p are
relaxed already, whereas all modes with index below p are still unrelaxed. In gen-
eral, the energy storage upon deformation is of order N · kBT. In a deformed visco-
elastic material, stress can be relaxed, whereby each mode relaxes a portion of kBT.
This means that the stored energy drops from a total of kBT per segment at τ0 to kBT
per chain at τRouse or τZimm. After an intermediate time of τp, only p · kBT remains
stored. From this notion, as well as from an expression for the time dependence of
the mode index – an equation that tells us up to which number p modes are already
relaxed at a time of interest τp – we may derive quantitative expressions for the
time-dependent energy storage and relaxation capabilities of polymer materials,
which we will do in Section 5.8.2.

Figure 41: Relaxation modes, indexed by a number p, of a schematic polymer chain. The first mode,
p = 1, relates to relaxation of the entire chain. In the second mode, p = 2, subchain segments
with length of just half of the chain can relax. The third mode, p = 3, corresponds to the relaxation
of subchain segments with length of only a third of the chain, and so on. In the last mode, p = N,
only single monomeric units can relax (not sketched here). Picture modified from H. G. Elias:
Makromoleküle, Bd. 2: Physikalische Strukturen und Eigenschaften (6. Ed.), Wiley VCH, 2001.

Table 8 summarizes the characteristic times that we have determined for the Rouse
and the Zimm model as well as the time dependence of the mode index p.

Table 8: Characteristic parameters of the Rouse and Zimm model. This table serves as a toolbox to
derive analytical expressions for the time-dependent mechanical spectra of polymer solutions and
melts in Chapter 5.

Rouse model Zimm model


1 + 2ν
Longest relaxation time τ 1 = τ Rouse = τ 0 N τ 1 = τ Zimm = τ 0 N3ν
 1 + 2ν  3ν
Relaxation time of the pth mode τ p = τ 0 Np τ p = τ 0 Np
Shortest relaxation time τN = τ0 τN = τ0
  −1  −1
τ 1 + 2ν τ 3ν
Time dependence of the mode index) p = τp ·N p = τp ·N
0 0

1)
Plugging in a value for τp allows us to calculate the mode down to which relaxation has already
proceeded after a time of interest τp. These equations are obtained from the ones in the second
row by simply rearranging those for p.
104 3 Real polymer chains

3.6.5 Subdiffusion

Let us get back to the Einstein–Smoluchowski equation we have introduced in Sec-


tion 3.6.1. Actually, it is not yet fully complete. To make it generally applicable, we
must extend it by an exponent α:

rð0ÞÞ i = 2 d D tα
2
hð~
r ðt Þ − ~ (3:35)

When α = 1, we get the regular Einstein–Smoluchowski equation that describes


normal Fickian diffusion. There are, however, many cases in which α ≠ 1. If α < 1,
the diffusion is constrained; this situation is called subdiffusion. A typical cause
for subdiffusion is if there is temporary trapping of the diffusing molecules, which
can be the case when they encounter binding sites on their way of motion.42 By con-
trast, if α > 1, the diffusion is promoted; this situation is called superdiffusion. A
typical cause for superdiffusion is if the diffusing molecules can ride on the back of
carriers for some time and thereby quickly span large distances on their way.43
Let us now consider a polymer chain with N segments, and within that, a sub-
 
chain with N/p segments. This subchain gets displaced by a distance R =~ rj τp −~
rj ð0Þ
of length l · (N/p)ν, which corresponds to its own size, during the time τp. Now imag-
ine that we somehow label one monomer segment j on the subchain and follow its
displacement. After τp it is

D 2 E 2 N 2v
τp
2v
1 + 2v
according to the Rouse model: ~rj ðτp Þ −~
rj ð0Þ =l =l2
(3:36a)
p τ0
D 2 E 2 N 2v
τp 2=3
according to the Zimm model: ~rj ðτp Þ −~
rj ð0Þ =l = l2 (3:36b)
p τ0

In the latter two equations, we have first written down a squared variant of the dis-
 
tance R =~ rj τp −~rj ð0Þ of length l · (N/p)ν that we consider here, and then we have
substituted the time dependence of the mode index p in the denominator by plug-
ging in the Rouse- or Zimm-model-related expression from Table 8. The resulting
exponents for the time dependence of the squared displacements that we obtain
with this approach are 1 +2ν2ν for the Rouse model (this is ½ in an ideal state with

 As an analogy, consider someone on a wine fest in Mainz who has enjoyed the wine too much
and therefore conducts a random walk home. The mean-square displacement of that person scales
linearly with time according to the Einstein–Smoluchowski equation. If, however, that person is
trapped on the way, for example, by meeting other people to chat with (if this is still possible in
that state) or by being attracted by more wine stands, the mean-square progress will be less than
proportional to time.
 As an analogy, consider our drunken friend at the wine fest again. This person may accelerate
the random-walk home by riding a bus on the way.
3.6 Chain dynamics 105

ν = 1/2) and 2/3 for the Zimm model, as shown in Figure 42. Both these exponents are
smaller than one. Thus, the segmental motion at times shorter than τRouse and
τZimm is subdiffusive. This is due to the hindrance imparted on the movement of each
monomer segment (such as our labeled one) by its neighbors, to whom it is bound by
chemical bonds. Our selected labeled monomer segment can move into a given direc-
tion only when the neighboring monomers do so as well. By contrast, when the
neighbors do not move into the same direction, our selected monomer is dragged on
its motion, thereby forcing its time-dependent displacement to be subdiffusive rather
than freely diffusive. As an analogy, remember our conceptual picture of a polymer
as a chain of people holding hands in Section 1.2. In such a chain, if one individual
wants to take some steps into a desired direction, this is possible only if the neigh-
bors join that motion. If they don’t, then that motion is constrained. Once more, the
motion is slower in the Rouse than in the Zimm scenario, as also seen in Figure 42,
where the line with slope 1/2 leads to smaller displacements in a given time than the
line with slope 2/3. Once more, this is due to the greater viscous drag imparted by the
freely draining solvent in the Rouse situation as compared to the solvent drag only
on the frontal surface of the solvent-filled coil in the Zimm situation.

Figure 42: Time dependence of the mean-square displacement of a “labeled” monomer unit
(named j) in the Rouse and Zimm scenario. At times smaller than the Rouse- or Zimm-time, the
power-law exponent in the two scaling laws is smaller than one, indicating subdiffusion of the
labeled segment. Picture redrawn from M. Rubinstein, R. H. Colby: Polymer Physics, Oxford
University Press, 2003.
106 3 Real polymer chains

3.6.6 Validity of the models

Now that we have treated two different models that describe the dynamics of a polymer
chain, a natural question is which of the two is able to make more accurate predictions.
As it turns out, both do, but their validity depends on the polymer’s surroundings. In
dilute solution, where the polymer concentration is low, hydrodynamic interactions be-
tween the segments in a coil are strong. In this case, the Zimm model is the more valid
one. We can see evidence for that in the left half of Figure 43, which shows the molar-
mass-dependent scaling of the diffusion coefficient of three types of polymers in dilute
solution. Each exhibits scaling to the power of the negative Flory exponent that ap-
plies to that specific polymer–solvent combination, in perfect agreement to eq. (3.33).
By contrast, in the semi-dulite concentration regime or in a melt, both characterized
by marked mutual interpenetration of the coils, the hydrodynamic interactions are
screened by overlapping segments of other polymer chains. This effect is similar to
the screening of excluded volume interactions in polymer melts that therefore always
show θ-type coil conformation. In these regimes, the Rouse model is better suited to
describe the polymer chain dynamics.44 We can see evidence for that in the right half

Figure 43: Scaling of (A) the polymer-chain translational diffusion coefficient, D, in dilute solutions
and of (B) the polymer-melt viscosity, η, both as a function of the polymer chain length, here
assessed by the polymer molar mass, M. Although dilute solutions show scaling of D ⁓ M–ν, as
predicted by the Zimm model, melts of chains with a length shorter than a certain critical one
exhibit η ⁓ M1, as predicted by the Rouse model (for a detailed, full explanation of this specific
scaling see Section 5.10.1).

 Actually, in the semidilute regime, polymers display both Rouse- and Zimm-type dynamics, de-
pending on the length scale under consideration.
Questions to Lesson Unit 6 107

of Figure 43, which shows the molar-mass-dependent viscosity of different polymer


melts. In the low molar-mass regime, this quantity rises linearly with the molar mass,
as will be proven to be Rouse-model based in Section 5.10.2. (In the high molar-mass
regime, the scaling is significantly steeper, indicating a different mechanism of chain
motion: the reptation mechanism, which we will also discuss in Section 5.10.2.)

Questions to Lesson Unit 6

(1) How does the root-mean-square displacement of a diffusing molecule or particle


differ if we give it twice as much time?
a. It is larger by a factor of 2.
pffiffiffi
b. It is larger by a factor of 2.
c. It is larger by a factor of 22 .
d. This requires a more extensive calculation.

(2) On which basic assumption is the Rouse model not based?


a. The coil is freely drained by the surrounding medium.
b. The chain is modeled as a series of beads connected by springs.
c. The total friction coefficient is the sum of the individual segmental friction
coefficients fSegment .
d. Both beads and springs experience friction together as they flow through
the surrounding medium.

(3) Match the appropriate timescale to the material behavior.


1. Individual segments or small chain sections can move; the material has the
properties of a viscoelastic solid.
2. Not even single segments can move; the material has the property of a glass.
3. Time is sufficient for the entire polymer chain to move, such that displace-
ment beyond the chain size occurs; the material has the property of a visco-
elastic fluid.
a. 1. τ0 < t < τRouse
2. t < τ0
3. t > τRouse
b. 1. τ0 < t < τRouse
2. t > τRouse
3. t < τ0
c. 1. t < τ0
2. τ0 < t < τRouse
3. t > τRouse
108 3 Real polymer chains

d. 1. t > τRouse
2. τ0 < t < τRouse
3. t < τ0

(4) On which basic assumption is the Zimm model not based?


a. The chains can be considered as nanogel particles filled with solvent.
b. The moving chain segments exert viscous drag on neighboring solvent
molecules.
c. The entrapped solvent remains in the coil.
d. Because of hydrodynamic interactions, even distant chain segments can af-
fect each other.

(5) Which statement concerning relaxation modes is correct?


a. At a time τp ; all relaxation modes with index smaller than p are already re-
laxed, whereas all modes with index higher than p are still unrelaxed.
b. The zeroth relaxation mode (index p ¼ 0) corresponds to the relaxation of
the entire chain.
c. The concept of relaxation modes is applicable only within the Rouse-model
framework but not within the Zimm model, as it does not account for hydro-
dynamic coupling.
d. The relaxation of the chain can be conceived as a sequential relaxation of
chain subsegments, which become larger with increasing time index.

(6) Why does the relaxation timescale have a higher exponent on N in the Rouse
model than in the Zimm model?
a. In the Zimm model, due to the non-draining condition, only the segments on
the frontal surface of the coil experience friction, whereas in the Rouse
model, all segments are affected because the solvent drains through the coil.
b. This is a pure result of the derivation of the corresponding formula and has
no trivial physical meaning.
c. In the Zimm model, the relaxation time is shortened by hydrodynamic inter-
actions, because the segments can interact with each other even at a higher
distance.
d. The hydrodynamic interactions in the Zimm model compensate the friction
due to the solvent, which makes the friction on the individual segments
smaller compared to the Rouse model.
Questions to Lesson Unit 6 109

(7) Assign the mentioned properties to the phenomena of subdiffusion or superdif-


fusion; choose the correct assignment order.
1. The exponent in the generalized Einstein–Smoluchowski equation is greater
than 1.
2. Diffusing particles are temporarily slowed down when they encounter bind-
ing sites.
3. Diffusing particles are temporarily transported in a directed manner, for ex-
ample, by docking to transport systems in a living cell.
4. The exponent in the generalized Einstein–Smoluchowski equation is less
than 1.
a. 1. Subdiffusion
2. Subdiffusion
3. Superdiffusion
4. Superdiffusion
b. 1. Superdiffusion
2. Subdiffusion
3. Superdiffusion
4. Subdiffusion
c. 1. Superdiffusion
2. Superdiffusion
3. Subdiffusion
4. Subdiffusion
d. 1. Subdiffusion
2. Superdiffusion
3. Subdiffusion
4. Superdiffusion

(8) The motion of individual segments within a polymer chain is subdiffusive ______
a. since this motion is inhibited by attractive interactions with the solvent.
b. for timescales smaller than τ0 (time for relaxation of a monomer).
c. since individual segments restrict each other’s free mobility due to their
binding to each other.
d. if good-solvent conditions are fulfilled.
4 Polymer thermodynamics

LESSON 7: FLORY–HUGGINS THEORY


All the previous discussion on polymers in this book was limited to single chains. The follow-
ing lesson moves beyond and introduces a concept to model and quantify the thermodynam-
ics of mixing of multiple chains with either a low-molecular solvent or with another polymer
species. You will see from this modeling that the mixing entropy, which is majorly responsi-
ble for the ease of mixing of many substances in classical chemistry, plays a minor role only
in polymer science, such that polymers only mix when there is very high enthalpic compati-
bility. This is captured in a fundamental quantity: the Flory–Huggins interaction parameter.

Throughout this textbook, so far, we have investigated how a polymer chain is


shaped, we have learned how it interacts with itself and with a solvent, and we
have obtained a picture on how it moves. The common factor in all of this is that,
so far, we have focused only on single chains. In the following, we want to expand
our view to multi-chain systems. We now ask ourselves: how do many chains inter-
act with one another and with a solvent? Answering that question will be a complex
endeavor, because it will require the appraisal of a huge number of interactions,
and is, thus, mathematically hard to describe. In addition, the entire multichain
system is subject to constant dynamic change, making it even more difficult for us
to describe it adequately. So, how do we solve this?
We approach this challenge by just considering average monomer–monomer
and monomer–solvent interactions. In short, rather than trying to appraise how
many M–M and how many M–S interactions we have at a time in the system, and
rather than trying to appraise where exactly in the system they occur, we are satis-
fied with the knowledge of how many of such interactions we have on average (over
space and time). A very convenient aspect of that simplification is that this number
simply scales in proportion to the respective volume fractions of M and S in the sys-
tem. This approach is called a mean-field treatment.

4.1 The Flory–Huggins mean-field theory

Using a mean-field approach, we are able to conceptualize the thermodynamics of


polymer–solvent and polymer–polymer mixtures. Said mean-field approach was in-
dependently developed by Maurice L. Huggins and Paul J. Flory and is called the

https://doi.org/10.1515/9783110713268-004
112 4 Polymer thermodynamics

Flory–Huggins mean-field theory.45 It appraises the change of the Gibbs free en-
ergy of mixing, ΔGmix, which must be negative for two components to readily mix
with each other:

ΔGmix = GAB − ðGA + GB Þ = ΔHmix − TΔSmix (4:1)

Here, GAB is the free energy of the mixture, whereas GA and GB are the free energies
of the demixed components. ΔGmix can also be expressed in terms of the enthalpy
and entropy of mixing, ΔHmix and ΔSmix, respectively. In general, entropy always fa-
vors mixing, as that decreases the order of the system, whereas enthalpy usually dis-
favors mixing, because that would force the components to interact with each other
rather than with themselves, which is a less perfect mutual match in most cases.

Maurice Loyal Huggins (Figure ) was born on


September , , in Berkeley, California. He
studied chemistry at the University of California
in Berkeley, where he obtained his Master’s
degree in . He obtained his PhD degree two
years later, in , under Charles M. Porter for
work on the structure of benzene. He then worked
at different institutes, including Stanford Research
Institute, Johns Hopkins University (Baltimore,
MD), before he joined Eastman Kodak (Rochester,
NY) in . From , he returned to Stanford
Research Institute and eventually retired from
research in . During his active time, he
independently conceived the idea of hydrogen
bonding in , was an early advocate for its
role in stabilizing protein secondary structures, Figure 44: Portrait of Maurice L. Huggins. Image
and produced a model of the α-helix in , reproduced with permission from Oregon State
roughly eight years ahead of the modern model University Libraries Special Collections &
of Linus Pauling, Robert Corey, and Herman Archives Research Center.
Branson. He also developed the mean-field theory
for polymer solutions, the Flory–Huggins theory.
He died on December , , in Woodside,
California.

 Flory himself suggested naming Huggins first (see J. Chem. Phys. 1942, 10(1), 51–61), as Huggins
was in fact the erstwhile of the two scientists to publish the mean-field theory (Huggins’ paper was
printed in the May issue of J. Chem. Phys. 1941, whereas Flory’s paper appeared in the August issue
of that journal). By the time of Flory’s suggestion, though, the former name was already established
in the scientific community, such that it is still used in the order of Flory–Huggins today.
4.1 The Flory–Huggins mean-field theory 113

In classical mixtures of low-molar-mass compounds, the inherently favorable


entropy influence in eq. (4.1) is powerful enough to outweigh the usually unfavor-
able enthalpy influence, such that mixing occurs. (And if that does not occur at
room temperature, it often helps to heat up, as this will give the entropy term
in eq. (4.1) more emphasis.) However, the favorable entropy term is much less
powerful for polymer systems. This is because polymers are long chain-like mol-
ecules that bring a large number of monomer units into quite some preorder,
simply by tying them together in the form of chains. In that state, these mono-
mers cannot arrange themselves in a solvent with that much freedom as they
could if they were not polymerized, and as a result, there is much less entropy
gain in the polymeric state than it could be in a nonpolymeric state upon mixing
with a solvent. Consequently, dissolution and intermixing of polymers is mostly
dependent on the enthalpic term, as we have learned already in Section 3.2.
The Flory–Huggins mean-field theory now aims at independently estimating the
entropic and the enthalpic contributors to eq. (4.1). It is based on a lattice model, as
shown in Figure 45. Component A is represented by black beads, whereas component
B is represented by white beads. These beads can be placed on lattices, either sepa-
rate ones for each species (left-hand side of Figure 45), which represents the demixed
state, or a common lattice for both species (right-hand side of Figure 45), which rep-
resents the mixed state. We assume that all molecules involved, be it the solvent
(white beads) or the dissolved component (black beads; note: if that compound is a
polymer, the beads are its monomeric units!), have the same volume.

Figure 45: Mixing of black and white beads with the same volume on a lattice. Picture redrawn from
M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, 2003.

Based on this premise, the volumes of the individual demixed lattices, VA and VB, are
additive upon mixing and sum up to the volume of the combined lattice, Vmix:

Vmix = VA + VB (4:2)

When the mixing is thermodynamically favored (which we aim to appraise here),


all beads will randomly occupy new positions within the new combined volume
114 4 Polymer thermodynamics

Vmix. The resulting volume fractions46 occupied by component A and B, ϕA and ϕB,
are given as follows:
VA
ϕA = (4:3a)
VA + VB
VB
ϕB = = 1 − ϕA (4:3b)
VA + VB

We define the unit volume of each single lattice site to be ν0. Then, the total number
of lattice sites, n, is

VA + VB
n= (4:4)
ν0

Component A occupies nϕA of the lattice sites, whereas component B occupies nϕB
of the lattice sites.
We can also use ν0 to calculate the chain volumes, νA and νB, of both components:

νA = NA ν0 (4:5a)

νB = NB ν0 (4:5b)

Here, NA and NB are the numbers of adjacent lattice sites occupied by each molecule
of component A or B, respectively. For a polymeric sample, NA and NB correspond
to the degrees of polymerization of the polymer chains (which scales to their sizes
as RA ⁓ NAν and RB ⁓ NBν).
We can delimit three different cases, depending on the relative size of NA and
NB, all three of which are shown in Figure 46. In the first scenario, depicted in
Panel (A), both NA and NB are one. In this case, we talk about a regular solution of
low-molar-mass compounds. Such a mixing process is highly entropy driven, as
also seen by the quite messy appearance of Panel (A). This is very different for sce-
nario two, shown in Panel (B). Here, NA >> 1 (in the figure, it is NA = 10), correspond-
ing to long polymer chains, and NB = 1, corresponding to solvent molecules. This is
the scenario of a polymer solution. Due to their connectivity to chains, the mono-
mer segments now have much less possibilities to arrange themselves than in the
first scenario, as seen by the more tidy distribution of the black beads in Panel (B)
as compared to Panel (A); only the solvent molecules still can arrange themselves
quite freely in that kind of system. In scenario three, shown in Panel (C), both NA
and NB are >> 1 (in the figure, NA = NB = 10). This is the case for a mixture of two
different polymers, a polymer blend. In that case, the freedom of arrangement of

46 The volume fraction is a common measure of concentration in polymer science. It denotes what
fraction of the volume of a system under consideration is occupied by a species of interest. You
know a related quantity from elementary physical chemistry: the mole fraction. If there’s two com-
n n
ponents A and B, it calculates as xA = n +A n and xB = n +Bn = 1 − xA .
A B A B
4.1 The Flory–Huggins mean-field theory 115

Figure 46: Arrangement of 50 black and 50 white beads on a 10 × 10 lattice with elementary-unit
volume v0 to model the binary mixing of two components in (A) a regular solution of two low-molar-mass
components A and B, both consisting of molecules with size equal to the lattice-site size,
(B) a polymer solution of a low-molar-mass component A that is built of monomer units (here: 10 per
chain) that each occupy one lattice site, such that each macromolecule of the whole species occupies NA
adjacent sites, dissolved in a low-molar-mass solvent B that consists of molecules equal to the lattice-
site size, and (C) a polymer blend of one high-molar-mass species A mixed with another high-molar-
mass species B, both occupying multiple adjacent lattice sites, NA and NB (here: both 10 per chain).
Pictures redrawn from M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, 2003.

both the black and the white beads is quite limited, as both are constrained by their
connectivity to chains.

4.1.1 The entropy of mixing

The entropy of mixing, ΔSmix, can be estimated by the number of microstates that
realize the two macrostates mixed and demixed. This is a statistical approach,
through which we can calculate the entropy by Boltzmann’s formula, S = kB ln(W).
In a homogeneous A–B mixture, each bead has n possible positions where it can be
placed on the common lattice. In other words, each bead can sit anywhere on the
lattice. In the demixed phases, component A only has nϕA possible positions in the
A phase; the rest is occupied by component B in the B phase. From this, we get the
entropy of mixing for a single A-bead, ΔSA, as47
1
ΔSA = kB ln n − kB ln nϕA = kB ln = − kB ln ϕA (4:6a)
ϕA

Here, kB ln n relates to the mixed, and kB ln nϕA to the demixed state.

 Do not confuse the index B at kB, where it denotes Boltzmann, to the index B at nB, NB, and ϕ B,
where it denotes species B.
116 4 Polymer thermodynamics

Analogously, the entropy of mixing for a single B-bead is

ΔSB = − kB ln ϕB (4:6b)

With both these expression, we can calculate the total entropy of mixing, ΔSmix

ΔSmix = nA ΔSA + nB ΔSB = − kB ðnA ln ϕA + nB ln ϕB Þ (4:7)

If we express the number of A-molecules, which are single beads in the case of a
low molar-mass substance or chains built up of multiple beads in the case of a
polymer, nA, as n · ϕA =NA as well as the number of B-molecules, nB, as n · ϕB =NB
in eq. (4.7) and then divide by n, we get the entropy of mixing per lattice site:

ΔSmix ϕ ϕ
ΔSmix = = − kB A ln ϕA + B ln ϕB (4:8)
n NA NB

In this equation, note again that ϕA and ϕB are the volume fractions of the components
A and B, and NA and NB are their degrees of polymerization, meaning how many adja-
cent lattice sites are occupied per A and B (macro)molecule. If we apply this formula to
a regular mixture, with NA = NB = 1, and with equal molar volumes of the A- and B-
beads (which we assume allover the present discussion anyways), we may replace the
volume fractions in eq. (4.8) by mole fractions, thereby yielding a form of the equation
that you know from elementary physical chemistry: ΔSmix = − kB ðxA ln xA + xB ln xB Þ.
When inspecting eq. (4.8), we can get some conceptual insights. As the numeri-
cal values for ϕA and ϕB are both 0. . .1, the logarithms in eq. (4.8) will always be
negative (or zero at most). Together with the minus sign in front of the right-hand
side of the equation, this will always lead to a positive mixing entropy. That makes
sense, as entropy always favors mixing. The extent of positivity, however, depends
on the denominators NA and NB. The larger they are, the less positive will be the
mixing entropy. This means that, even though entropy always favors mixing, it
does so with less power at high chain length. This insight matches the qualitative
discussion that we have just led on the preceding pages.
Let us deepen our conceptual insight by some quantitative calculation and esti-
mate the entropy of mixing for each of the three cases displayed in Figure 46. Here,
the volume fractions of component A and B are equal, ϕA = ϕB = 0.5. In our first sce-
nario, Panel (A), we have 50 white and 50 black beads that we can arrange freely
on the lattice. This means that there are 100! / (50! 50!) = 1029 different possible mi-
crostates to realize the macrostate “mixed”. The entropy of mixing per lattice site
normalized by Boltzmann’s constant, ΔSmix =kB , is 0.69 in this case. In Panel (B),
our second scenario, we retain the 50 white solvent beads, but change the structure
of component A, represented by the black beads, such that it is now a 10-segment
polymer, thereby now represented by five 10-segment chains of black beads. Be-
cause the black beads are connected to each other in this fashion, the normalized
entropy of mixing is markedly reduced, ΔSmix =kB = 0.38. In the last scenario, Panel
4.1 The Flory–Huggins mean-field theory 117

(C), this trend continues. Now, the white beads are also connected together in five
polymer chains with 10 segments each. As a result, the number of possible arrange-
ments on the lattice for both components drops to only about 103. Consequently, the
normalized entropy of mixing is further reduced, ΔSmix =kB = 0.069.
To visualize these findings, the normalized entropy of mixing is plotted as a func-
tion of the volume fraction of component A in Figure 47. Again, we realize from this
plot that the entropic gain from mixing a polymer with a solvent is drastically lower
than that of a regular mixture of low-molar-mass compounds. Furthermore, we see
that the entropic gain of the polymer blend is almost negligible. What we also realize
from this plot is that the extent of the entropic gain also depends on the volume frac-
tion; it is greatest at a 50:50 mixture (ϕA = 0.5) if the mixing components have equal
sizes, which is the case for our scenario A with NA = NB = 1 and for our scenario C
with NA = NB = 10, whereas the entropy maximum is at a different ϕA (of about 0.65)
if the mixture consists of compounds with different sizes, as it is the case in scenario
B, where we have NA = 10 but NB = 1.

Figure 47: Entropy of mixing for the three different cases shown in Figure 46.

The just minor gain of entropy upon mixing polymers with solvents or other poly-
mers is one of two reasons for nature to rely the build-up of complex structures on
polymeric building blocks: since their mess-up would bring just minor entropic
gain, their tiding to ordered structures does not cost much either.48

 The second reason is that, due to their interaction energies of just a few couple RT, polymeric
building blocks can serve to build entities that are stable enough to persist, but also labile enough
to be de- and rearrangeable with not too much energy input.
118 4 Polymer thermodynamics

4.1.2 The enthalpy of mixing

Let us now focus on the energetic contributions to the mixing of two components A
and B. To do this, we have to quantify all possible interactions, which are A–A, B–B,
and A–B. This, however, cannot be done analytically. Remember that a regular 1:1
mixture of even just 50 A-beads and 50 B-beads gives 1029 possible combinations. To
address this challenge, we use a mean-field approach. In that approach, we conceptu-
ally “hydrolyze” our polymer to unconnected segments that have the same size as
their surrounding molecules, and distribute them randomly on the lattice. We then
only discuss interactions between two direct neighbors on this lattice. This gives us
three different interaction energies, u, that we must discuss: uAA, uBB, and uAB.
The average pairwise interaction energy of a monomer A with one of its
neighboring lattice sites is given as

UA = uAA ϕA + uAB ϕB (4:9a)

Here, uAA is the interaction energy in case of an A–A contact and ϕA the probability
for the neighboring site of the A-monomer under consideration to be indeed occu-
pied by another A. This value is equal to the volume fraction of A in the system.
The second term denotes the same for the case of an A–B contact.
Analogously, the average pairwise interaction energy of component B with one
of its lattice sites is

UB = uAB ϕA + uBB ϕB (4:9b)

Arguing this way exposes the simplicity and advantage of a mean-field approach.
Normally, the interaction of one molecule (or bead) of component A is either exactly
uAA or exactly uAB, depending on what type of molecule (or bead) sits next to it.
However, we do not know which of the two interactions is present at a given time
and spot on the lattice, so we assume an average mean interaction of both based on
their frequency of occurrence in the system, ϕA and ϕB.
We know that each of the n lattice sites has z nearest neighbors to interact with.
From the single component average pairwise interaction energies, we can calculate
the total average interaction energy in the mixed state:
z·n
U= ðUA ϕA + UB ϕB Þ (4:10a)
2

with UA and UB the average pairwise interaction energies given by eq. (4.9a) and
(4.9b). The factor of ½ rows back double-counting of the pairwise interactions.
The average interaction energy in the demixed states can also be determined. In
this case, we use the individual interaction energies, uAA and uBB, in an expression
just like eq. (4.10a):
4.1 The Flory–Huggins mean-field theory 119

z·n
U0 = ðuAA ϕA + uBB ϕB Þ (4:10b)
2

With the interaction energies in the mixed and demixed states, we can calculate the
energy change upon mixing. We do this by subtracting the energies and normalize
the resulting difference by dividing it by the number of lattice sites, n:

 mix = U − U0 z
ΔU = ð2uAB − uAA − uBB Þ ϕA ϕB = kB TχϕA ϕB (4:11)
n 2

This equation introduces a very important quantity for the discussion of polymer
thermodynamics: the Flory–Huggins interaction parameter:
z 2uAB − uAA − uBB
χ= · (4:12)
2 kB T

This parameter is a dimensionless measure of the difference of the pairwise interac-


tion energies before mixing (uAA and uBB) and after mixing (2uAB), normalized to the
lattice geometry (z/2) and the elementary thermal-energy increment (kBT). During
mixing of A and B, we break A–A and B–B contacts but in turn establish two new
A–B contacts per such breaking. This means that we “loose” uAA and uBB, but we
gain 2uAB. Hence, according to eq. (4.12), when χ < 0, this means that there is more
energy gained than lost upon this exchange. In that case, 2uAB is smaller, that
means either less positive (=thermodynamically less unfavorable) or (even better)
more negative (= thermodynamically more favorable), than uAA and uBB are to-
gether; this corresponds to an exothermic mixing process. When χ > 0, by contrast,
there is more energy lost than gained upon mixing. In that case, 2uAB is larger, that
means either more positive (=thermodynamically more unfavorable) or less nega-
tive (= thermodynamically less favorable), than uAA and uBB together; this corre-
sponds to an endothermic mixing process. A Flory–Huggins interaction parameter
of χ = 0 indicates no energy change upon mixing. The mixing process is then driven
by its entropic contribution only, a situation commonly referred to as ideal mix-
ture.49 Most usually, mixing of two components A and B is disfavorable, as two

 Do not confuse such an ideal mixture with what we have referred to as an ideal state (or
pseudo-ideal state to be precise) in Section 3.2, when introducing the Θ-state. In an ideal mixture
(sometimes synonymized as “ideal solution”), interactions of kind A–A, B–B, and A–B are all the
same; we refer to that as an “ideal mixture”, as mixing is driven only by entropy in that case. In
polymer systems, such a case corresponds to a system where M–M, S–S, and M–S interactions are
all the same; we have referred to this as an “athermal solution” in Section 3.2. An ideal state, by
contrast, is one without any interactions, which is merely theoretical. A truly existing variant is a
pseudo- or quasi-ideal state which shows rather strong attractive effective M–M interactions, so
strong that they just balance the M–M hard-sphere repulsion; we name this state a pseudo- or
quasi-ideal one, as it is not free of interactions, but displays a delicate balance of effective M–M
attraction and M–M hard-sphere repulsion (just like a real gas at the Boyle temperature shows
120 4 Polymer thermodynamics

A–B contacts just cannot be as perfect matches as an A–A and a B–B contact.50 This
corresponds to an endothermic mixing process, characterized by a positive χ.
If we assume, for the sake of simplicity, that our mixing process is isochoric,
which means that the total volume does not change upon mixing, ΔVmix = 0, then we
can substitute the enthalpic term, ΔHmix, in eq. (4.1) by ΔUmix, that is, by eq. (4.11).
This yields the change of the Gibbs free energy upon mixing:

 mix = kB T ϕA ϕ
ΔG ln ϕA + B ln ϕB + χϕA ϕB (4:13)
NA NB

We have already stated that ΔGmix must be negative for mixing to occur, so let us
look a little bit closer at the latter equation. The combinatorial entropy term always
contributes something negative, as the numerical values for ϕA and ϕB are both 0
. . . 1, so the logarithms in eq. (4.13) will always be negative (or zero at most). This
makes sense, as entropy always favors mixing. For the enthalpic term, however, it
depends on the difference of the pairwise interaction energies before and after mix-
ing, indicated by the Flory–Huggins parameter χ. As just said above, in most cases,
χ > 0, because the components rather want to interact with themselves than with
one another. The enthalpic term, then, counteracts the entropic contribution. It
therefore depends on how strong the entropic contribution is, and how large the
counteracting enthalpic term is, which is most dominantly regulated by the magni-
tude of χ· As we have discussed earlier in this chapter, the entropy of mixing is very
small for polymer solutions and blends (see Figure 47). As a result, χ must be lower
than a critical value (which is 0.5 in the case of polymer solutions and close to zero
in the case of polymer blends, as will be demonstrated in Section 4.2) for mixing
still to happen. Just in very rare cases, χ can be smaller than 0. For that to be, there
must be a mutually complementary favorable interaction between A and B, such
that there will be an additional energy gain when both components interact with
each other rather than with themselves. A clever realization of that premise was pre-
sented in Freiburg and Mainz in the 1980s by Reimund Stadler. He equipped polymer
chains with complementary hydrogen bonding motifs, as shown in Figure 48. The
energy gained from these transient interactions very much fuels the mixing of
these polymers.

balanced effective A–A attraction and hard-sphere A–A repulsion). Be aware of the risk of confu-
sion of the terms ideal state (which is a polymer in a Θ-solvent at Θ-temperature) and ideal mixture
(which is a polymer in an athermal solvent).
 “Birds of a feather flock together.”
4.1 The Flory–Huggins mean-field theory 121

Figure 48: Reimund Stadler’s approach of equipping polymers with self-complementary hydrogen
bonding motifs to fuel a favorable enthalpic contribution to the free energy of mixing. Pictures
reprinted with permission from R. Stadler, L. L. de Lucca Freitas, Coll. Polym. Sci. 1986, 264(9),
773–778, copyright 1986 Steinkopff Verlag (now Springer Nature), and R. Stadler, L. L. de Lucca
Freitas, Macromolecules 1987, 20(10), 2478–2485, copyright 1987 American Chemical Society.

Reimund Stadler (Figure ) was born on October ,


, in Stühlingen, Germany. He studied chemistry
at the Albert Ludwigs University Freiburg, where
he received his diploma and PhD degree, with a
thesis on “Viscoelasticity and Crystal Melting of
Thermoplastic Elastomers”. He then became a
postdoc at Porto Alegre in Brazil before his
habilitation in  under Hans-Joachim Cantow
back in Freiburg. Stadler was appointed as full
professor at the Johannes Gutenberg University
Mainz, and then switched to the University of
Bayreuth, where he died suddenly just one year Figure 49: Portrait of Reimund Stadler. Image
later, on June , . To his honor, the German reproduced with permission from Designed
Chemical Society gives out an award named after Monomers and Polymers 1999, 2(2), 109–110.
him to rising-star researchers in the field of polymer Copyright 1999 Taylor & Francis.
science every other year.

4.1.3 The Flory–Huggins parameter as a function

The Flory–Huggins theory is a simplified approach to polymer thermodynamics: it


appraises the entropy of mixing based on a combinatorial argument, and the energy
of mixing based on mutual average interactions in a mean-field treatment. So far,
so good. There are, however, several aspects that are not at all or at least not ade-
quately captured by that theory, for example, additional entropic effects such as
the need for favorable mutual orientation of the molecules to each other to interact,
or mistakes in the enthalpic part of the theory that come from the mean-field treat-
ment, ignoring the rather inhomogeneous distribution of the two species in the sys-
tem, where we have high densities of the black species inside the polymer coils, but
none of it between the coils. All these deviations from reality are lumped into the
122 4 Polymer thermodynamics

Flory–Huggins parameter χ. It is therefore not a given and fix number, but a func-
tion that itself is composed of an entropic part, χS, and a temperature-weighted en-
thalpic part, χH.
χH
χ = χS + (4:14a)
T

It is therefore the interplay of both contributions, χS and χH, and temperature that
determines the miscibility or immiscibility of a polymer and a solvent. The role of
the entropy part and the enthalpy part of χ can be understood based on the follow-
ing thought: Let us consider all contributions to the mixing free enthalpy, that
means, not only the simple configurational entropy term and the simple pairwise-
interaction enthalpy term in the Flory–Huggins eq. (4.13), but also all further spe-
cific entropic and enthalpic effects that the mean-field approach doesn’t account
for; we name this full set of contributions ΔGmix, excess . We may then write the follow-
ing equation to connect this to the (then phenomenological) parameter χ:
 χ 
ΔGmix, excess = ΔHmix, excess − T · ΔSmix, excess = kB T · χS + H (4:14b)
T

Based on that, we identify − kB · χS as an excess mixing entropy ΔSmix, excess that is not
based on simple statistical-combinatorial considerations, but lumps all the specificity
of the molecules acting as the solvent and the solute polymer. kB · χH = ΔHmix, excess still
reflects the mixing enthalpy based on mean-field average pairwise interactions in the
form of χH = ðz=2kB Þð2uAB − uAA − uBB Þ, so that ΔHmix, excess = ðz=2Þð2uAB − uAA − uBB Þ.
(Note that in this notion, χH has a physical unit [K], whereas χS has no unit.)
Let us examine the enthalpic contribution to the Flory–Huggins parameter, χH, a
little more closely. In unpolar systems, it is commonly positive, χH > 0. This is due to
mostly attractive intramolecular interactions such as van der Waals interactions in
those systems, which oppose mixing as they are better built between each compo-
nent itself but not so perfectly between the components each other.51 These interac-
tions, however, can be overcome at high temperatures, such that mixing is promoted
at high temperatures. Mathematically, this situation is reflected by the temperature
dependence in the denominator of eq. (4.14a). If χH > 0, then we add something posi-
tive to χs, such that our total χ gets larger; this is bad for mixing, as χ should be small
for that, smaller than a certain critical limit (which we will discuss in the next sec-
tion). As T reduces the χH contribution by its position in the denominator, however,
this unfavorable part of χ is attenuated if T is high. In the opposite case, usually en-
countered in polar, protic systems, χH can be negative, that is, χH < 0. This situation is
encountered if we have specific intercomponent interactions, such as hydrogen

 A quantitative justification for χH > 0 in systems that prefer intra- over intercomponent contact
is given in Footnote No. 58 in Section 4.2.1.
4.1 The Flory–Huggins mean-field theory 123

bonds or dipole–dipole interactions. These transient bonds favor the mixing, but
they are broken at high temperature. Mathematically, we can see this again from the
temperature dependence in the denominator of eq. (4.14a). If χH < 0, then we add
something negative to χS, such that our total χ gets smaller; this is good for mixing.
As T reduces the χH contribution by its position in the denominator, however, this
favorable part of χ is attenuated at high T.

Figure 50: Poly(N-isopropylacrylamide) is a polymer


with a largely unpolar backbone and an unpolar side
group that both dislike water, whereas a polar amide
group in between can promote mixing with water due
to hydrogen bonding. The hydrogen bonds, however,
are broken at high temperature, and the polymer then
precipitates from the solution.

A prime example of the latter case is an aqueous solution of poly(N-isopropylacryla-


mide), pNIPAAm, whose structure is depicted in Figure 50. This polymer has a
largely unpolar backbone and an unpolar side group that both disfavor interactions
with water, but it also has a polar amide group that can form hydrogen bonds
with water molecules. This promotes mixing overall, but only up to a temperature
of 32 °C. At higher temperatures, the hydrogen bonds break and the polymer pre-
cipitates, because then, the so-called hydrophobic effect comes into play: Water
molecules that find themselves in close proximity to hydrophobic domains, such
as the isopropyl moieties in pNIPAAm, form clusters and therefore assume a more
ordered intermolecular arrangement than they would have in their natural state,
because in that ordered arrangement they avoid unfavorable hydrophobic–hydro-
philic contact. This effect causes an additional unfavorable excess mixing entropy
ΔSmix, excess < 0. With increasing temperature, we find that the contribution of this
excess entropy to ΔGmix, excess , which is − T · ΔSmix, excess , increases linearly, and there-
fore the excess free enthalpy of mixing ΔGmix, excess can compensate the usual configu-
rational Flory–Huggins free enthalpy of mixing, eventually leading to ΔGmix, total > 0,
thereby inducing phase separation.
Such a solubility-temperature barrier is called Lower Critical Solution Temper-
ature, LCST. There is a whole class of LCST-type polymers such as pNIPAAm, all of
which exhibiting LCSTs reasonably close to the human body temperature.52 This
makes these polymers interesting for biomedical applications, possibly in targeted

 This is because hydrogen-bonding interactions have a strength of some few singles to tens of
RT, which is easily activated to open in the temperature window of 30–50 °C. Nature uses the same
principle for immune reactions (e.g., for fever, which denaturates hostile proteins by breaking their
hydrogen-bonding interactions) or to bind and unbind functional entities on demand, such as the
double strands of DNA.
124 4 Polymer thermodynamics

drug-release systems. For example, consider that inflammatory tissue has a little
higher temperature than healthy tissue. If an LCST polymer is designed such to dis-
play a coil-to-globule (i.e., swollen-to-deswollen) transition right in the range of that
temperature difference, a nanocapsule system may be made from that polymer such
that it can collapse and thereby release an anti-inflammatory drug only in the inflam-
matory tissue regions, but not in the healthy ones. Furthermore, we may expect that
protonation or deprotonation of these polymers may further drastically affect their
LCST, and with that, their regions of solubility and nonsolubility in water. As a result,
we may also use these polymers for pH-dependent active systems, such as nanocap-
sules that release drugs only in cancer tissue but not in healthy tissue, making use of
the circumstance that there is a pH difference between these two kinds of tissues.
As a closing remark, note that in addition to the temperature dependence, the
Flory–Huggins parameter also often exhibits a concentration dependence that may
be captured in a virial-series form:
χH
χ = χS + + χ 1 ϕ + χ 2 ϕ2 +    (4:15)
T

This does even more underline our above statement that χ is not a unique number,
but a phenomenological parameter that itself depends on multiple influences, the
most relevant being temperature (variable T in eq. (4.15)), composition of the sys-
tem (variable ϕ in eq. (4.15)), and chain length (not included in eq. (4.15); there
could be further terms in it to account for that variable as well).

4.1.4 Microscopic demixing

We have seen that the entropic gain of a polymer–polymer blend is negligible. As a


consequence, it is almost impossible to dissolve one polymer in another. Even when
the polymers are almost chemically identical, the enthalpic penalty is still high
enough to cause demixing.53 This immiscibility has severe consequences for systems
where two polymers are chemically connected to one another, as it is the case in
block copolymers. In these systems, the two blocks usually want to demix, but they
can do so only on nanoscopic scales, because they are tied together. As a result, the
block-copolymer system will show microphase separation, with a morphology that
depends on the block lengths, as shown in Figure 51. At equal block length, shown
in the center part of Figure 51, both blocks will arrange themselves in lamellar
phases, whereas if one of the blocks is shorter, it will undergo coiling while the

 Polystyrene and deuterated polystyrene are a prime example of this. Both are chemically almost
the same (they just differ by exchange of hydrogen for deuterium), but they still have a positive,
albeit very small, Flory–Huggins parameter, χ = 10–4. This causes both polymers to be immiscible if
their molar mass is greater than about M = 3 × 106 g·mol–1.
4.1 The Flory–Huggins mean-field theory 125

segments of the longer block will be less coiled, thereby changing the lamellae thick-
ness. If the block-length mismatch gets too severe, different other morphologies will
result, as shown in the left and right outer parts of Figure 51. These have astonishing
order on both microscopic and mesoscopic scales, which is determined by the mu-
tual pinning of locally microseparated phases by chain-blocks that partition in each
of them. The overall phase microstructure therefore results from the length, degree
of coiling, and extent of partition of either chain block. The astonishing order of the
different morphologies shown in Figure 51 is therefore not due to spontaneous self-
assembly of the polymers to such states, but instead, due to the hindered dis-
assembly of the immiscible polymer blocks that results from their connectivity.

Figure 51: Various block copolymer phases that can form depending on the fraction of the
constituent blocks A and B.

Note that a stark contrast to block copolymers is random or alternating copolymers.


These copolymers manage to incorporate two chemically different building-block spe-
cies along their chains. As such, they can serve to incorporate both these components
in one polymeric system, without chain–chain immiscibility issues. This circumstance
makes random or alternating copolymerization so attractive, as it is often the only
way to realize such a material combination, due to the foresaid extreme difficulty of
realizing polymer blends. So to speak, in a random or alternating copolymer, the mix-
ing of the different chemical species is done “along the chain backbone”.

4.1.5 Solubility parameters

The Flory–Huggins parameter quantifies polymer–solvent interactions; it does so,


however, for pairs of these only, such that it is somewhat unhandy to tabulate it.
More practical would be pairs of parameters, one for the polymer and one for the
126 4 Polymer thermodynamics

solvent, that can be tabulated independently and then taken together to appraise
the miscibility of both compounds. Such parameters have been introduced by Hil-
debrandt and Scott. They are based on quantifying the attractive interactions of the
solvent molecules or the monomer units of a polymer chain amongst themselves
via their cohesive energy, ΔEA, which can be determined by measurement of the
heat of combustion. With that, Hildebrandt and Scott defined
rffiffiffiffiffiffiffiffi
ΔEA
δA = (4:16)
vA

Here, δA is the solubility parameter for compound A, ΔEA is its cohesive energy, and
vA its molecular volume.
In the lattice model that we have discussed above, the interaction energy for a
z ΔEA
given lattice site in a plain-A system is uAA , which is equal to − v0 and thus,
2 vA
according to eq. (4.16), also equal to = − v0 δA 2 :

z ΔEA
uAA = − v0 = − v0 δA 2 (4:17a)
2 vA

The interaction energy of the second component in a plain-B system is calculated


analogously

z ΔEB
uBB = − v0 = − v0 δB 2 (4:17b)
2 vB

By the geometric mean of both parameters, δA and δB, we can estimate the mutual
interaction energy, uAB
z
uAB = − v0 δA δB (4:17c)
2

Strictly speaking, the solubility parameter as defined above exists only for liquid
substances, that is, only for solvents, but not for polymers, as the latter do not ex-
hibit combustion. It can be presumed, though, that this is only due to the covalent
connectivity of the monomers, such that if this connection would be lost, they
would combust just as an own independent substance composed of molecules with
a structure that matches the one of the repeating units in the polymer. Thus, the
solubility parameter of the polymer matches the one of such a hypothetic fluid.54 In
other words, although the solubility parameter of a solvent can be measured di-
rectly from its heat of combustion, the one for a polymer can be appraised by find-
ing a fluid with a best identical structure to the monomer units in the polymer and

 Based on our classification in Section 3.2, that fluid would be an athermal solvent for the
polymer.
4.1 The Flory–Huggins mean-field theory 127

then estimate the solubility parameter of that fluid. As a practical alternative, we


may also just seek a fluid that best dissolves the polymer, as assessed by maximum
coil expansion in that solvent, which we may determine by light scattering or vis-
cometry, and then presume this polymer–solvent mixture to be an athermal one
with interaction energies of kind uAA = uBB = uAB, such that we can set the solubility
parameter of the polymer practically equal to that of the solvent.
The Flory–Huggins parameter, χ, is related to the solubility parameters by
v0
χ= ðδA − δB Þ2 (4:18)
kB T

With that equation, χ can be directly calculated from the δ-parameters of a polymer
and solvent of interest. Solubility parameters therefore have the advantage that
they can be used singularly for these two components and do not have to be speci-
fied for each possible pair, as it is the case with the Flory–Huggins parameter. In
this way, lists such as Table 9 can be compiled that collect many different solubility
parameters, and these collections can then be used to select a good solvent for a
given polymer. For this purpose, we just have to choose a solvent with a solubility
parameter that is similar to that of the polymer. This even works for solvent mix-
tures, because the net-δ value is the average of its constituents. In that way, solvent
mixtures tailored to the polymer’s solubility parameter can easily be prepared.55

Table 9: Solubility parameters according to Hildebrandt and Scott for some typical solvents
and polymers.

Solvent δ (cal·cm–)/ Polymer δ (cal·cm–)/

Cyclohexane . Polyethylene .


Benzene . Poly(vinyl chloride) .
Chloroform . Polystyrene .
Acetone . Poly(methyl methacrylate) .
Methanol . Polyamine  .
Water . Polyacrylonitrile .

The unit for the solubility parameter, (cal·cm–3)1/2, is commonly named 1 Hildebrandt.

We see that χ can only be positive according to eq. (4.18). That implies perfect solu-
bility of the two components when δA = δB, and thus χ = 0, according to the proverb
“similia similibus solvuntur”, meaning “like dissolves like”. Note that δ-values are
large for polar solvents, because these have high cohesive energies ΔEA. Such polar

 Most interestingly, even mixtures of two nonsolvents can constitute a solvent for a polymer
with a δ in the middle range of Table 9, namely if the mixture brings together one nonsolvent with
a too large δ and another nonsolvent with a too small δ, whose average will then lie in the middle
and thereby match the δ of the polymer.
128 4 Polymer thermodynamics

solvents actually require a three-dimensional solubility parameter that takes into


account possible secondary interactions:
 1=2
δ3d = δvdW 2 + δDipole − Dipole 2 + δH − Bonds 2 (4:19)

Questions to Lesson Unit 7

(1) Which one of the following is true for an ideal mixture?


a. The mixing process is entropically controlled, since the mixing of two com-
ponents A and B increases entropy, whereas in view of energy, the interac-
tions of type A–B are inferior to those of types A–A and B–B.
b. The mixing process is purely entropically controlled, since the mixing of
two components A and B increases entropy, whereas in view of energy, the
interactions of type A–B are equal to those of types A–A and B–B.
c. The mixing process is entropically controlled, since the mixing of two com-
ponents A and B increases entropy, whereas in view of energy, the interac-
tions of type A–B, despite being favorable, contribute less than the entropy.
d. The mixing process is enthalpically controlled, since the interaction ener-
gies of two miscible species are significantly larger in magnitude than the
extent of entropy increase that results from their mixing.

(2) What is meant by a “regular solution”?


a. A solution in which particles of component A and particles of component B
are not randomly distributed and the interaction energies between A and B
differ.
b. A solution for which H E ≠ 0 and SE ¼ 0 holds.
c. A solution for which V E ≠ 0 holds, that is, there is an excess volume when
mixing.
d. A solution under thermodynamically regular conditions (standard pressure
and standard temperature).

(3) What are the differences between the mixing process of polymers and the mix-
ing process of low-molecular-weight compounds?
a. There are no differences; both mixing processes are exclusively entropically
driven.
b. There are no differences; both mixing processes are mainly enthalpically
driven.
Questions to Lesson Unit 7 129

c. The mixing process in polymers is more entropically driven than that in


low-molecular-weight compounds, because the larger molar mass means
that entropy is more strongly involved in the mixing process.
d. The mixing process in polymers is less entropically driven than that in low-
molecular-weight compounds, because the given arrangement of the mono-
mers in chains limits their freedom of arrangement in a mixture.

(4) Which of the following is not true for the entropy of mixing?
a. It can be calculated with the help of the Boltzmann formula.
b. Its values do not differ for a polymer solution on the one hand and a regular
solution on the other hand; the decisive factor is the presence of a solution.
c. Its value is always positive, that is, the entropy is always in favor of a
mixture.
d. For polymers, it is very low, which is one way to explain why many ordered
structures in nature are based on polymers.

(5) What does the Flory–Huggins interaction parameter χ indicate?


a. The difference of the pairwise interaction energies before and after mixing,
normalized to kB T and the lattice geometry.
b. The ratio of the pairwise interaction energies before and after mixing, nor-
malized to kB T and the lattice geometry.
c. The difference of the pairwise interaction energies before and after mixing,
normalized to RT and the number of all pairs.
d. The ratio of the pairwise interaction energies before and after mixing, nor-
malized to RT and the number of all pairs.

(6) Which of the following is not true for the Flory–Huggins interaction parameter χ?
a. χ describes the energy balance of a mixing process, comparable to the en-
ergy balance of a chemical reaction.
b. χ can take values below, above, and of exactly zero
c. χ usually has a value of zero.
d. A value of χ ¼ 0 is achieved only in an ideal mixture.

(7) The Flory–Huggins theory does not pick up all entropic and enthalpic aspects. If
one includes additional entropic and enthalpic aspects, the Flory–Huggins pa-
rameter results as a function, which can be split into an entropic part χS and an
enthalpic part χH : χ ¼ χS þ ðχH =T Þ. What are the consequences with respect to
χH for a nonpolar system and a polar-protic system, respectively?
a. In the nonpolar system, intramolecular attractive interactions dominate,
which can be formed more strongly at higher temperatures because of the
better mobility of the molecules, and thus mixing is inhibited.
130 4 Polymer thermodynamics

b. In the nonpolar system, intramolecular attractive interactions dominate;


these are temperature-independent, in contrast to the entropic part, so that
miscibility depends only on the entropic part.
c. In the polar-protic system, intercomponent interactions dominate, favoring
or inhibiting mixing depending on the interaction type.
d. In the polar-protic system, intercomponent interactions dominate; the re-
sulting bonds are broken at high temperatures and thus mixing is favored
only at low temperatures.

(8) The effect outlined in the previous question for nonpolar polymers in polar-
protic solvents such as water is also called the hydrophobic effect. This can also
be viewed from the entropic side. Which statement correctly describes this?
a. Since there is no interaction partner for the water molecules in case of dis-
solution of a hydrophobic species, they distribute themselves as widely as
possible to maximize their entropy. This makes the dissolution process of
the existing polymer more difficult – especially at higher temperatures.
b. The water molecules near the contact surfaces of the polymer are more or-
dered than “free” water molecules. To minimize these contacts, entropy
causes the formation of polymer globules, which becomes more pronounced
at increasing temperature.
c. At high temperatures, the system strongly tends to maximize its entropy. In
principle, this is only possible in the state where polymer and solvent are
separate.
d. To maximize entropy, existing hydrate shells are preferentially broken at
high temperatures to obtain more possibilities in the arrangement of water
molecules and thus higher entropy.
4.2 Phase diagrams 131

4.2 Phase diagrams

LESSON 8: PHASE DIAGRAMS


Based upon the Flory–Huggins theory introduced in the last lesson, the following will show
at what combinations of relevant variables, mostly temperature and composition, a polymer
will or will not mix with a solvent or with another polymer. A plot of these parameters gives a
phase diagram, of which the following lesson will introduce two fundamentally different
types for polymer systems.

The Flory–Huggins theory has given us conceptual and quantitative insight about
the miscibility of polymers with solvents or with other polymers. In the following,
we want to expand this insight to make even more detailed quantitative predictions
at what conditions polymer–polymer and polymer–solvent systems are miscible or
immiscible. In other words, we want to construct phase diagrams.
A phase diagram is a plot of two practically relevant variables against each
other such to create a map with regions that show at what combination of these
variables the system is in what state. You may remember the most simple case of a
phase diagram from your elementary physical chemistry classes: a one-component
phase diagram, which is a map of pressure versus temperature that delimits at
which p–T pair the substance is a solid, a liquid, or a gas. This map also delimits in
which regions we have coexistence of two or even three of these phases. In general,
the construction of a phase diagram is based on seeking the minimum of the free
energy for each region in the map. In the following, we want to derive such a map
for two-component polymer mixtures; thus, one of our two practically relevant vari-
ables will be the system’s composition, that is, the volume fraction of the polymer
of interest in the mixture, ϕ. The other variable will be a measure of the energetic
interactions between the components. Naturally, this variable is χ but as we will see
later, that can actually be translated easily to an even more practical variable, T.

4.2.1 Equilibrium and stability

The mixing and demixing of a system with a component A at the initial composition
ϕ0 is determined by the free energy of its mixed phase, Fmix(ϕ0), and the free energy
of the separated phases α and β, Fαβ(ϕ0). When Fmix(ϕ0) < Fαβ(ϕ0), then the mixture
is stable; in this case, when brought together, both components will mix spontane-
ously and stay mixed. By contrast, when Fmix(ϕ0) > Fαβ(ϕ0), then the mixture is
unstable and will phase-separate, or the components would not even mix in the
132 4 Polymer thermodynamics

first place. To determine this ratio, we need to determine functions that describe
Fmix(ϕ0) and Fαβ(ϕ0).
In the phase-separated state, an initial system composition ϕ0 is constituted by

ϕ0 = fα ϕα + fβ ϕβ (4:20)

Here, fα is the relative fraction of phase α, and ϕα is the volume fraction of compo-
nent A in that phase α. Analogously, fβ is the relative fraction of phase β, and ϕβ is
the volume fraction of component A in that phase β.
The energy of the phase-separated state, Fαβ(ϕ0), is a simple sum of similar
kind: Fαβ = fα Fα + fβ Fβ . By combination with eq. (4.20), we get
   
ϕ β − ϕ 0 F α + ϕ 0 − ϕ α Fβ
Fαβ = fα Fα + fβ Fβ = (4:21)
ϕβ − ϕα

This expression has the form of a linear equation and gives a straight line when plot-
ted with ϕ0 not as a single composition point but as variable in the interval [ϕα; ϕβ].
The energy of the mixed state, Fmix(ϕ0), can be estimated based on the energy
change upon mixing, as expressed by the Flory–Huggins formula from the last
chapter (eq. 4.13). We substitute ϕA = ϕ and ϕB = 1 – ϕ and get

mix = kB T ϕ ð1 − ϕÞ
ΔF ln ϕ + lnð1 − ϕÞ + χϕð1 − ϕÞ (4:22)
NA NB

This expression has the form of a curve when plotted.56


Both functions, Fmix(ϕ)57 and Fαβ(ϕ), are shown in Figure 52 for the case of a
system where demixing is more favorable (Panel A) and one where mixing is more
favorable (Panel B). We can see for the phase-separated system in Panel (A) that the
curve of the mixed state at composition ϕ0, where we have a value of Fmix, lies
above the straight line of the demixed state, where we have a value of Fαβ. Thus,
when mixed initially, the system can spontaneously lower its energy by phase sepa-
ration and thereby drop from Fmix to Fαβ at the composition ϕ0. We will then have
two separate phases, α and β, with separate free energies Fα and Fβ, which have an

 The energy change expressed by eq. (4.22) adds upon the initial state, which is the phase-
separated one at an energy of Fαβ(ϕ0), when mixing occurs, with either positive or negative sign
depending on whether the mixing is disfavorable (positive sign) or favorable (negative sign). After
the mixing, we therefore have a free energy of Fmix(ϕ) = Fαβ(ϕ) + ΔFmix(ϕ).
 This is the function just “derived” in the latter footnote: Fmix(ϕ) = Fαβ(ϕ) + ΔFmix(ϕ). It consists
of a summand ΔFmix(ϕ), which is the Flory–Huggins formula (4.22), which is added to (and there-
fore “sits upon”) the straight line of the phase-separated state, Fαβ(ϕ) (4.21). At conditions for favor-
able mixing, the curve lies below the straight line as the ΔFmix(ϕ)-summand has negative values,
and vice versa.
4.2 Phase diagrams 133

Figure 52: Free energy, F, as a function of the volume fraction of a component of interest, ϕ, in
mixed or phase-separated polymer systems. In both plots, the straight line corresponds to the
phase-separated state, whereas the curve corresponds to the mixed state (it represents the
Flory–Huggins equation). In (A), the concave curve lies above the straight line in a region between
compositions ϕα and ϕβ, meaning that the mixed state is less favorable than the phase-separated
state there. As a result, a mixed system that has a composition ϕ0 will spontaneously decompose
into separated phases α and β, in which the component of interest is present in volume fractions
ϕα and ϕβ, respectively. In (B), the convex curve lies below the straight line in a region between
compositions ϕα and ϕβ, meaning that the mixed state is more favorable that the phase-separated
state. Pictures redrawn from M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press,
2003.

average value of Fαβ. The straight line that denotes the demixed state in Figure 52(A)
connects these points, whose abscissa values denote the volume fractions of our
component of interest A in the two phases, which are ϕα in the α-phase and ϕβ in
the β-phase. The relative fractions of the two phases are given by the length of the
two parts of the straight line left and right of the point at (ϕ0;Fαβ), commonly re-
ferred to as the lever rule. In full contrast to all that, we can see for the mixed sys-
tem in Panel (B) that the curve of the mixed state at composition ϕ0, where we
have an energy of Fmix, lies below the straight line of the demixed state, where we
have an average energy of Fαβ at the composition ϕ0, which is actually a mean of
two separated phases α and β with separate free energies Fα and Fβ and volume
fractions of our component A of ϕα in the α-phase and ϕβ in the β-phase. In that
situation, the system can spontaneously lower its energy by mixing and drop from
the Fαβ(ϕ0) line to the Fmix(ϕ0) curve.
We can delimit the two different scenarios in Figure 52 by analyzing the curva-
ture of the ΔFmix-curve. This can be calculated by the second derivative of ΔFmix. In
case of the phase-separated system, the curve has a concave curvature. This means

that the second derivative of ΔFmix is negative, ∂2 ΔFmix ∂ϕ2 < 0. The mixed state, by
contrast, exhibits a convex curvature, and the second derivative of ΔFmix is positive,

∂2 ΔFmix ∂ϕ2 > 0.
134 4 Polymer thermodynamics

Let us look at some boundary-case examples to understand this better. For an


ideal mixture, the enthalpic contribution is zero, that is, ΔHmix = 0 or ΔUmix = 0. In that
case, only the temperature-dependent entropic contribution remains (see eq. 4.1).
The second derivative of ΔFmix then is
 mix
∂2 ΔF ∂2 ΔSmix 1 1
= −T = kB T + >0 (4:23)
∂ϕ 2
∂ϕ 2 N A ϕ NB − ϕÞ
ð 1

The outcome is a positive value at any ϕ, which means that ΔFmix(ϕ) has a convex
curvature. Hence, mixing is always favorable, which is reasonable: entropy always
favors mixing independent of the mixture’s composition.
In an opposite example, let us ignore the entropy term and only look at the en-
ergetic contributions. The second derivative of ΔFmix then is
 mix
∂2 ΔF  mix
∂2 ΔU
2
= = − 2χkB T (4:24)
∂ϕ ∂ϕ2

This represents the findings that we have derived in the last chapter. When χ < 0,
which according to our discussion in Section 4.1.2 denotes that mixing is favored,
then the curvature is convex, which also denotes mixing to be favored according to
what we have just said above. When χ > 0, which according to our discussion in Sec-
tion 4.1.2 denotes that mixing is disfavored, then the curvature is concave, which
also denotes mixing to be disfavored according to what we have just said above.58
In an actual mixture, both the energetic and the entropic term are of course
relevant:
 mix
∂2 ΔF  mix
∂2 ΔU ∂2 ΔSmix 1 1
= −T = kB T + − 2χkB T (4:25)
∂ϕ 2
∂ϕ 2
∂ϕ 2 NA ϕ NB ð1 − ϕÞ

Let us look at how the curves in an F(ϕ) diagram look like in that situation.

58 We may get a further insight when we write out the temperature-dependent form of χ, namely
χ = χS + χH/T, in eq. (4.24):
 mix
∂2 ΔF  mix
∂2 ΔU  χ 
= = − 2χkB T = − 2 χS + H kB T = − 2kB TχS − 2kB χH
∂ϕ2 ∂ϕ2 T
 mix ∂2 ΔU
∂2 ΔF  mix
For T → 0, this leads to 2
= = − 2kB χH def
=
− z · ð2uAB − uAA − uBB Þ
∂ϕ ∂ϕ2
From this, we see that
uAA + uBB
– if uAB > , which corresponds to an unstable mixture, then we have χH > 0
2
uAA + uBB
– if uAB < , which corresponds to a stable mixture, then we have χH < 0
2
4.2 Phase diagrams 135

Figure 53 displays an example of an unsymmetric polymer blend (NA ≠ NB).


Shown is the change of the free energy of mixing, ΔFmix, as a function of the volume
fraction of component A, ϕ, at different temperatures. At high temperature, the en-
tropy term dominates and creates a global minimum. The curve is convex allover,
which means that mixing is favored at all compositions. At low temperature, by
contrast, the energy term that usually disfavors mixing creates a miscibility gap. In
this region, the curve is locally concave, and the system can decrease F by phase
separation down to a value represented by the straight line that connects the energy
minima. This line is the common tangent, because that kind of line is the deepest
possible one touching the curve such that ∂F/∂ϕ is the same for the curve and the
straight line in the two phase-coexistence points, respectively. This must be the
case, as ∂F/∂ϕ corresponds to the chemical potential, which must be the same in
the coexistence points ϕ′ and ϕ″ to have equilibrium there.

Figure 53: Free energy of mixing, ΔFmix, as a function of the volume fraction of a component of
interest, ϕ, in a polymer system at different temperatures. At high temperature, dominance of the
entropy term creates a distinct minimum, thereby favoring mixing at all compositions. At low
temperature, the energy term that usually disfavors mixing creates a miscibility gap at
intermediate compositions.

According to this latter thought, the common tangent delimits the stable “rim-ϕ
regions” from the unstable “mid-ϕ region”. However, a “bad” concave curvature is
present actually only in the innermost-ϕ region, between the inflection points.
Hence, we can delimit three different regimes. First, we have the rim-ϕ regime be-
yond the common-tangent touching points. This corresponds to two regions, one
at the left ϕ-rim and the other at the right ϕ-rim, that both enable stable mixtures.
In contrast, the second, innermost-ϕ regime between the inflection points delimits
the region in which the mixture is truly unstable. The third domain lies between
these two extremes, and is called the metastable domain. Here, we have a convex
curvature, indicating favorable mixing, but we are already above the deepest pos-
sible straight line, the common tangent, which indicates unfavorable mixing.
136 4 Polymer thermodynamics

Figure 54: The miscibility gap is further subdivided into unstable and metastable domains. (A) At
the edges of the miscibility-gap region, the curvature of the F(ϕ)-curve is convex, which would
correspond to a stable mixture, but we still have Fmix > Fphase-sep. there. As a result, this is a
metastable domain. (B) The separation between the single phase and the metastable regime is
called the binodal line, whereas the separation between the metastable and the unstable regime is
called the spinodal line. Both these lines coincide at the critical point (χc;ϕc).

We can take a closer look at the miscibility gap by considering a symmetric poly-
mer blend with two components that have the same N, as shown in Figure 54(A).
Between the inflection points, that is, for the volume fractions ϕsp1–ϕsp2, the curva-
  
ture is concave ∂2 ΔFmix ∂ϕ2 < 0 . The mixture is unstable, and even smallest fluctua-
tions induce phase separation, which is called spinodal decomposition. Between the
inflection points and the common-tangent touching points,59 for the volume frac-
  
tions ϕ′–ϕsp1 and ϕsp2–ϕ″, the curvature is convex ∂2 ΔFmix ∂ϕ2 > 0 , but still Fmix
(ϕ0) > Fαβ(ϕ0). In this regime, the mixture is metastable, which means that it is lo-
cally stable against small fluctuations, whereas phase separation will set in at larger
fluctuations. In other words, phase separation needs nucleation and growth here.
Figure 54(B) shows the same system in a plot of the Flory–Huggins parameter,
χ, as a function of the polymer volume fraction, ϕ. This is a phase diagram. At low
χ, the components like or at least tolerate each other, and mixing is possible at all
compositions. To the contrary, at high χ, the components do not like each other,

 In the case of a symmetric blend, the common tangent is a horizontal line with slope zero, so its
touching points with the ΔFmix(ϕ)-curve are the minima of that curve.
4.2 Phase diagrams 137

such that demixing will occur at certain compositions, delimited by the same
boundaries as just discussed for the F(ϕ)-plot above. The point where this can first
happen is called the critical point at χc;ϕc.60 The regions of nonstability are delim-
ited by two lines. The binodal line separates the stable from the nonstable domain.
The nonstable domain itself is composed of the unstable and the metastable re-
gime, which are delimited from each other by the spinodal line.

4.2.2 Construction of the phase diagram

So far, we have derived a diagram of F versus ϕ with a set of T-dependent curves to


estimate at what composition, at a given temperature, the system is mixed or dem-
ixed. More intuitive, though, would be a diagram with two practical variables, T
and ϕ, in which we can directly map domains of miscibility and immiscibility.
To construct this, we take two steps. First, we derive a χ–ϕ diagram as plotted
in Figure 54(B). Second, we derive a T–ϕ diagram from the χ–ϕ diagram by tak-
ing into account the temperature dependence of the Flory–Huggins parameter χ.
We begin by calculating the binodal and spinodal lines of the first diagram.
As a toolkit, we need the energy of mixing given by

mix = kB T ϕ ð1 − ϕÞ
ΔF ln ϕ + lnð1 − ϕÞ + χϕð1 − ϕÞ (4:26a)
NA NB

its first derivative given by

∂ΔF mix ln ϕ 1 lnð1 − ϕÞ 1


= kB T + − − + χð1 − 2ϕÞ (4:26b)
∂ϕ NA NA NB NB

and its second derivative given by


 mix
∂2 ΔF 1 1
= kB T + − 2χ (4:26c)
∂ϕ 2 NA ϕ NB ð1 − ϕÞ

The phase boundary, or the binodal line, is given by the common tangent. For sim-
plicity, we consider a symmetrical blend with NA = NB = N. In this case, the common
tangent is a horizontal line with slope zero. This can be calculated by setting the
first derivative of Fmix to zero:

 This point is conceptually identical to the critical point in a one-component phase diagram and
the related van der Waals equation at (pc;Tc;Vc), at which liquefaction of a gas can first occur.
138 4 Polymer thermodynamics

∂ΔF mix ln ϕ lnð1 − ϕÞ !


ϕ = ϕ′
= kB T − + χð1 − 2ϕÞ = 0 (4:27)
∂ϕ ϕ = ϕ′′
N N

Rearrangement for χ yields the binodal χ(ϕ) curve:


 
ϕ
1 ln ϕ lnð1 − ϕÞ ln 1−ϕ
χBinodal = − = (4:28)
2ϕ − 1 N N ð2ϕ − 1ÞN

The inflection points delimit the metastable from the unstable regime. They can be
calculated by setting the second derivative of Fmix to zero:
 mix
∂2 ΔF 1 1 !
= kB T + − 2χ = 0 (4:29)
∂ϕ 2 Nϕ N ð 1 − ϕ Þ

Rearrangement for χ yields the spinodal χ (ϕ) curve:

1 1 1
χSpinodal = + (4:30)
2 Nϕ N ð1 − ϕÞ

The minimum of both the latter curves denotes the critical point, (χc;ϕc). This point
marks the very first possibility for demixing. Below χc, mixing is possible for all vol-
ume fractions. We can calculate the critical point by setting the first derivative of
the latter equation for the spinodal line to zero to estimate its minimum:
!
∂χSpinodal 1 1 1 !
= − + =0 (4:31a)
∂ϕ 2 Nϕ2 N ð1 − ϕÞ2

At this point of our discussion, we can actually drop our above simplification and
consider the general case of a nonsymmetric blend again. This is mathematically
imprecise, but rather based on a logical argument: the N in the first summand be-
longs to the volume fraction ϕ of compound A (note that in this section, we write ϕ
for what we have earlier denoted as ϕA), whereas the N in the second summand
belongs to the volume fraction 1 – ϕ (i.e., 1 – ϕ), which is that of compound B (be-
cause 1 – ϕA = ϕB). So, we get
!
∂χSpinodal 1 1 1 !
= − + =0 (4:31b)
∂ϕ 2 NA ϕ2 NB ð1 − ϕÞ2

Solving this equation for ϕ yields the critical composition, ϕc:


pffiffiffiffiffiffi
NB
ϕc = pffiffiffiffiffiffi pffiffiffiffiffiffi (4:32)
NA + NB
4.2 Phase diagrams 139

According to that latter formula, a symmetric blend with NA = NB has a critical com-
position of ϕc = 0.5. This value is understandable, because at such a 50:50 composi-
tion, we have the greatest extent of disfavorable hetero contacts in our system, and
so it should be that composition at which demixing is most likely to happen first.
Precisely, this happens if χ is too unfavorable, meaning if the components are not
compatible enough; and this is the case if χ lays below χc. That critical threshold χc
can in turn be quantified by inserting ϕc for ϕ in eq. (4.30):
2
1 1 1
χc = pffiffiffiffiffiffi + pffiffiffiffiffiffi (4:33)
2 NA NB

Let us examine the latter equation in some detail.


In a regular small-molecule solution, where NA = NB = 1, the critical interaction
parameter is χc = 2. As a result, everything with χ < χc = 2 will mix, whereas every-
thing with χ > χc = 2 may demix if ϕ is around ϕc.
In a polymer solution, with NA >> 1 and NB = 1, the picture is different. Here,
χc = ½, which is much lower than in the small-molecule scenario. Here, every-
thing with χ < χc = ½ will mix, whereas everything with χ > χc = ½ may demix if ϕ is
around ϕc. A polymer solution in which χ is exactly ½ is in the Θ-state, which means
at the borderline between miscibility and immiscibility.
In a polymer blend, where both NA >> 1 and NB >> 1, the critical Flory–Huggins
parameter is basically zero, χc ≈ 0. As a result, mixing is hardly possible at all in
this case, as for mixing to occur, χ must be smaller than χc, which is hard to achieve
when χc ≈ 0 already. Mixing two polymers therefore practically only works for χ = 0,
which is very rare! We therefore have to resort to other strategies than mixing if the
properties of different polymers shall be combined. A common way to achieve that
is by copolymerization of their different monomers in one copolymer species (“mix-
ing within the chain”), or by incorporation of attractive side groups in the to-be-
mixed polymers that aid to overcome the enthalpy penalty of mixing.

As a numerical example, we calculate the volume fractions at the phase boundaries of the spino-
dal curve of a polymer solution at a degree of polymerization of NA = 1000 with a Flory–
Huggins interaction parameter of χ = 1.5; hence, we consider a system beyond χc that already
exhibits a decent two-phase regime. The spinodal curve separates the two-phase regime
from the metastable coexisting regime. Mathematically, it is determined by the set of inflec-
tion points of the function ΔFmix(ϕA); these points can be calculated by setting the second
derivative of this function to zero. The resulting polymer volume fractions then correspond to
the polymer concentrations of the two separate phases that arise from the phase separa-
tion.61 For our example, the calculation goes as follows:

 Note that these volume fractions are temperature dependent due to the temperature depen-
dence of the Flory–Huggins parameter χ(T), as will be discussed in the next paragraph.
140 4 Polymer thermodynamics

ΔFmix ϕA ϕ
= · ln ϕA + ϕB · ln ϕB + χϕA ϕB = A · ln ϕA + ð1 − ϕA Þ · lnð1 − ϕA Þ + χϕA ð1 − ϕA Þ
kB T NA NA
The first derivate of that function is:
 
ΔFmix
∂ kB T 1 ϕA 1 −1
= · + · ln ϕA + ð1 − ϕA Þ · − lnð1 − ϕA Þ + χ − 2χϕA
∂ϕA NA ϕA NA ð 1 − ϕA Þ
1 1
= + · ln ϕA − 1 − lnð1 − ϕA Þ + χ − 2χϕA
NA NA
We further calculate the second derivative, plug in the numbers from the example and set the
entire equation equal to zero:
 
ΔFmix
∂2 kB T 1 1 −1 1 1 1 !
= · − − 2χ = · + − 2 · 1.5 = 0
∂ϕA 2 NA ϕA ð1 − ϕA Þ 1000 ϕA ð1 − ϕA Þ

Further rearrangement yields:

1 1 1
· ð1 − ϕA Þ + ϕA − 3ð1 − ϕA ÞϕA = 0 , 3ϕA 2 + 1 − − 3 ϕA + =0
1000 1000 1000

, ϕA 2 − 0.667ϕA + 0.0003333 = 0
We have now generated a quadratic equation in its reduced form and can solve for its two solu-
tions using the p–q formula:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0.667 0.667 2 0.667
ϕA = ± − 0.0003333 = ± 0.333 = 0.6665 and 0.0005
2 2 2

As we can see from the resulting volume fractions, spinodal decomposition does in fact not produce
pure solvent and pure polymer phases, but instead, a polymer-rich phase (here: 67% polymer) and
a polymer-lean phase (here: 0.5‰ polymer) are obtained. This means that the phase containing
most of the polymer still contains a lot of solvent (here: 33%), and thus needs to be properly
dried to obtain the pure polymer.

We have learned in Section 4.1.3 that the Flory–Huggins parameter is not a con-
stant, but rather a temperature-dependent function according to eq. (4.14a). This
allows us to translate the χ–ϕ representations we have discussed so far (Figure 54
(B)) into more practical T–ϕ diagrams, as shown in Figure 55. These can have two
very different appearances, depending on the sign of the χH-contributor, corre-
sponding to two types of miscibility. In the first case, depicted on the left-hand side
of Figure 55, χH > 0, which means that the overall χ parameter decreases as the tem-
perature rises. These systems display a volume-fraction-dependent upper critical
solution temperature (UCST), above which mixing is possible. In the second case,
χH < 0, and the overall χ parameter increases with rising T. Such mixtures exhibit a
volume-fraction-dependent lower critical solution temperature (LCST), above
which mixing is impossible!
4.2 Phase diagrams 141

Figure 55: The sign of the enthalpy-component of the T-dependent Flory–Huggins parameter
determines whether mixing is favorable or disfavorable at high or low T, which are two opposite
scenarios assessed by either an upper critical solution temperature (UCST), as shown on the left,
or a lower critical solution temperature (LCST), as shown on the right. In both schematics, we see
that there is also an impact of the molar mass of the chains.

A UCST is the usual case for many polymer–solvent mixtures without specific
intercomponent interactions such as mutual hydrogen bonding, as we have dis-
cussed in Section 4.1.3. The exact value of the UCST is also molar-mass dependent,
as can also be seen in Figure 55. In the UCST graph, we see that upon decrease of
temperature, an accompanying decrease of the solvent quality such that χ exceeds ½,
and with that, an onset of precipitation is encountered for long chains sooner than for
short chains. Analogously, in the LCST graph, we see that upon increase of tempera-
ture, an accompanying decrease of the solvent quality such that χ exceeds ½, and
with that, an onset of precipitation is encountered for long chains sooner than for
short chains. In both cases, this circumstance can be used for fractionation of poly-
disperse samples, simply by dissolving such a sample in a solvent and then gradu-
ally worsening the solvent quality, either by suitable change of temperature or by
successive addition of a nonsolvent. Upon that, first the longest chains precipitate,
then the next shorter fraction, and so forth. Isolating these fractions before precipitat-
ing the next thereby allows the broadly polydisperse sample to be fractionated into
several less polydisperse ones. This may either be used preparatively or at least analyt-
ically to determine a coarse-grained molar-mass distribution by analyzing the average
142 4 Polymer thermodynamics

molar mass and the amount of each fraction through means like light scattering, os-
mometry, and gravimetry.
More complex phase diagrams may arise in special cases due to the sometimes
complex temperature dependence of the Flory–Huggins interaction parameter that
might change its sign multiple times. Figure 56 shows some of these special cases.
Panel (A) is for a mixture with both a UCST and an LCST, Panel (B) is for a mixture
with a closed miscibility gap, and Panel (C) displays an hourglass-shaped phase dia-
gram. The first case has a rather simple explanation, shown in a χ–T representation in
Panel (D). Initially, we see a regular Flory–Huggins-type UCST upon increase of tem-
perature. On top of that, however, comes another effect: an increase of temperature
often leads to thermal expansion of both the solvent and the polymer. This expansion
is usually more pronounced for the solvent than for the polymer. Hence, in our lattice
picture, we have to increase the lattice-site volume for the solvent more than for the
polymer’s repeating units upon increase of temperature, meaning that we need to dis-
tort our lattice to fit them both on, which entails an entropy penalty. If this penalty is
too marked, the system will phase-separate again, creating an LCST in the process.
Hence, we actually always have both an LCST and a UCST, and depending on where
they and their two binodal curves lie, we may have phase diagrams of kind (A), (B),
or (C). The reason why we often nevertheless (seemingly) observe simple phase dia-
grams such as those shown in Figure 55, with just one critical temperature (LCST or
UCST, but not both of them) is that many polymer–solvent systems actually exhibit
both, but one of them is usually “hidden”, either in the form of a UCST that lies

Figure 56: Complex shapes of phase diagrams that result from the interplay of the T-dependent χ-
parameter in the enthalpy term and the T-dependent TΔS term in the entropy term of the
Flory–Huggins equation. Case (A) has an illustrative explanation, as visualized in (D): the usual
UCST-type T-dependence of χ leads to a decrease of it upon raise of T (making mixing more
favorable, or better saying “less disfavorable” at high T), whereas the different thermal expansion
of the solvent and the polymer that comes along with a temperature increase causes energy and
entropy penalties that oppose this; together, this creates a minimum of χ (T), corresponding to a
closed miscibility gap in the phase diagram. The sketch in Panel (D) is inspired by J.M.G. Cowie:
Chemie und Physik der synthetischen Polymeren, Vieweg, 1991.
4.2 Phase diagrams 143

below the melting point of the solvent, or in the form of an LCST that is located
above the boiling point of the solvent.

4.2.3 Mechanisms of phase separation

4.2.3.1 Spinodal decomposition


In the unstable two-phase regime (see Figure 54(B)), any random concentration
fluctuation self-amplifies and eventually leads to macroscopic phase separation.
This is because the gradient of the chemical potential μ is opposite to the gradient
of concentration in this regime. In very general, concentration fluctuations in a sys-
tem are equilibrated by a diffusive flux of matter along the gradient of the chemical
potential μ, which is defined as μ = ð∂G=∂nÞ = ð∂G=∂ϕÞ and exhibits a monotonic
concentration dependence in mixed phases: μmix = μ0 + RT ln ϕ. The volume frac-
tion ϕ in the second equation necessitates a high chemical potential at high con-
centrations, and conversely, a low chemical potential at low concentrations. If
   
there is a gradient in the chemical potential, ∂μ=∂ϕÞ = ∂2 G ∂ϕ2 > 0, diffusion
leads to an overall flux of matter from the higher to the lower potential until the
gradient is equilibrated. Normally, such a chemical-potential gradient is in line
with the concentration gradient, so the diffusive flux goes from high to low concen-
trations, referred to as downhill diffusion. In the two-phase spinodal decomposi-
tion regime, by contrast, the gradient of the chemical potential is reverse to that in
   
concentration, ∂μ=∂ϕÞ = ∂2 G ∂ϕ2 < 0. Diffusion still takes place along the chemi-
cal-potential gradient, but now this means it occurs against the concentration gradi-
ent. The resulting self-amplification of concentration differences is called uphill
diffusion and eventually leads to a macroscopic phase separation, the so-called
spinodal decomposition. This phenomenon occurs on the characteristic length-
scale of the concentration fluctuations in the system. If that lengthscale is large,
diffusion has to cover large distances, which is unlikely; if the lengthscale is short,
many new polymer–solvent interfaces must be created, which is disfavored, too.
Therefore, spinodal decomposition occurs on an intermediate lengthscale that con-
stitutes the best compromise between the two effects. This lengthscale remains con-
stant at first and is amplified only in its amplitude. During later stages of the
process, the lengthscale diverges, and the phase separation actually becomes mac-
roscopic. This last phase can be observed, for example, by optical microscopy, if
one of the components is colored. Earlier stages can be observed by complementary
methods such as neutron or X-ray scattering.

4.2.3.2 Nucleation and growth


In the metastable regime (again, see Figure 54(B)), the system is able to tolerate small
concentration fluctuations, and thus, spinodal decomposition is prohibited. Here, a
144 4 Polymer thermodynamics

different phase-separation mechanism is in action: nucleation and growth. Instead


of a macroscopic phase separation, small polymer-rich phase clusters form locally.
As can be seen in Figure 57, these clusters need to grow to a specific minimal size
with a radius r* for the nucleation process to be energetically favorable. This is be-
cause the unfavorable interfacial energy of the nuclei first is stronger than their favor-
able volume energy, but this ratio reverses from a cluster minimal size r* on. If the
nuclei do not make it to that size, nucleation does not surpass the minimum size for
macroscopic phase separation, so the system can remain two-phased. By contrast,
once the minimal size r* is reached, macroscopic phase separation sets in. Therefore,
impurities with a radius bigger than r*, such as dust particles, often function as a
seed for nucleation. Alternatively, artificial nucleation seeds can be added on pur-
pose to facilitate the process. This is called heteronucleation. Once the critical nucle-
ation size is exceeded and a nucleus is formed, more material adds to it easily, and
the new phase grows.

Figure 57: Graphical representation of the free energy of nucleation. Two factors contribute to the
free energy: the interface and the volume of the nucleus. Nucleation is energetically favorable only
after the cluster has grown to a minimal size r*.

Figure 58 summarizes the two polymer phase-separation mechanisms in view of the


system’s state of stability or instability. In the unstable regime, spinodal decompo-
sition is the mode of phase separation. In the metastable regimes, phase separation
happens through nucleation and growth.
Questions to Lesson Unit 8 145

Figure 58: Schematic representation of phase separation regimes of polymer–solvent or


polymer–polymer mixtures. Shown is the mixing free energy as a function of the polymer volume
fraction, like in Figure 53. In the unstable regime, spinodal decomposition takes place. In the
metastable regimes, phase separation occurs by nucleation and growth.

Questions to Lesson Unit 8

(1) What applies to phase diagrams of polymer mixtures or solutions? (In the fol-
lowing, both are referred to as a mixture)
a. Like the phase diagram of a single component, in a two-component system, a
phase diagram represents the state of aggregation, that is, whether it is a solid,
liquid, or gas phase, in the form of a pressure versus temperature diagram.
b. Like the phase diagram of a single component, in a two-component system,
a phase diagram represents the mixing state, that is, whether it is a mixed
or demixed system, in the form of a pressure versus temperature diagram.
c. Unlike the phase diagram of a single component, in the case of mixtures,
instead of pressure versus temperature, we plot the interaction energy in
terms of χ, which can also be reformulated as a temperature dependence
versus the composition of the mixture in terms of mole fractions xi.
d. Unlike the phase diagram of a single component, in the case of mixtures,
instead of pressure versus temperature, we plot the interaction energy in
terms of χ, which can also be reformulated as a temperature dependence
versus the composition of the mixture in terms of volume fractions ϕi .
146 4 Polymer thermodynamics

(2) Which statement applies to the state of a mixture in view of the free energy of
mixing Δ Fmix ?
a. A concave curvature of the mixing energy curve always implies miscibility
of the components.
b. A concave curvature of the mixing energy curve always means nonmiscibil-
ity of the components.
c. A convex curvature of the mixing energy curve always means miscibility of
the components.
d. A convex curvature of the mixing energy curve always means nonmiscibil-
ity of the components.

(3) A phase diagram of a two-component system is classically represented in the


form of a χ–ϕ diagram. This can be derived from the F–ϕ diagram. The χ–ϕ
diagram then consists of ______
a. a curve connecting all minima of the curve in the F–ϕ diagram and a curve
connecting all inflection points of the curve in the F–ϕ diagram.
b. a curve connecting all maxima of the curve in the F–ϕ diagram and a curve
connecting all inflection points of the curve in the F–ϕ diagram.
c. a curve connecting all minima of the curve in the F–ϕ diagram and a curve
connecting all maxima of the curve in the F–ϕ diagram.
d. a curve connecting all extrema of the curve in the F–ϕ diagram and a curve
connecting all inflection points of the curve in the F–ϕ diagram.

(4) What is the critical point?


a. The point in the phase diagram with values χc and ϕc at which segregation
is first possible.
b. The point in the phase diagram with values χc and ϕc at which the tempera-
ture is so high that mixing is no longer possible.
c. The point in the phase diagram with values χc and ϕc at which it is no lon-
ger possible to distinguish whether it is a mixture or two demixed phases.
d. The point in the phase diagram with the values χc and ϕc at which the
metastable state transitions into the unstable state; accordingly, there are
two critical points.
Questions to Lesson Unit 8 147

(5) What do the terms LCST and UCST denote?


a. The LCST is the temperature below which mixing is impossible; UCST is the
temperature above which mixing is possible.
b. The LCST is the temperature above which mixing is impossible; the UCST is
the temperature above which mixing is possible.
c. The LCST is the temperature below which mixing is possible; the UCST is
the temperature above which mixing is impossible.
d. The LCST is the temperature below which mixing is impossible; the UCST is
the temperature below which mixing is possible.

(6) Which statement regarding the metastability of a mixture is correct?


a. A mixture that shows a greater tendency to phase separation at high tem-
peratures is called metastable.
b. A mixture is metastable if it remains mixed for only a limited time after mix-
ing and then segregates.
c. If the composition of a mixture is above a critical value ϕc , spontaneous
segregation can always occur, and the mixture is metastable.
d. If the free energy of mixing exhibits a convex curvature in its course but is
energetically above the total free energy of the system, this region is called
metastable, since phase separation can only occur through nucleation here.

(7) What is the critical value of the Flory–Huggins parameter χc of a polymer


solution?
a. It is always 2.
b. It is always 21.
c. It is always 0.
d. It depends on the system at hand.

(8) For a polymer mixture, mixing is possible only when χ ≈ 0. To achieve this,
what is true for the individual components?
a. They must be as different as possible.
b. They must be as similar as possible.
c. Their χ-values must be as large as possible.
d. Their χ-values must be as small as possible.
148 4 Polymer thermodynamics

4.3 Osmotic pressure

LESSON 9: OSMOTIC PRESSURE


You have got to know the Flory–Huggins parameter as an elementary tool to appraise
whether or not polymers are miscible with solvents or other polymers. What you are lacking,
though, is an experimental technique to determine this parameter. Such a technique will be
introduced in the following lesson. You will see that in addition to quantifying interactions
between a polymer and its surrounding, and therewith determining the Flory–Huggins pa-
rameter, this method also allows you to determine the molar mass of a polymer. Further-
more, you will again see the striking similarity of the physical-chemical concepts behind real
polymer chains and real gases.

So far we have looked at polymer thermodynamics from a theoretical viewpoint and


introduced the Flory–Huggins parameter χ as a simple means to quantify the
polymer–solvent interaction and to predict the state of solvency of one in the other
(meaning if it is a good-solution state, a θ-state, or a bad- or nonsolvent state). Ear-
lier in this book, in Chapter 3, we have introduced a quantity that expresses pretty
much the same: the excluded volume, ve. Now, we will take a look on how to deter-
mine χ and ve experimentally; this discussion will also show us how these two pa-
rameters are related to each other.
Consider the experimental setup that is shown in Figure 59. This setup features
a vessel containing a polymer solution immersed in a bath with excess solvent.
Both are separated by a semipermeable membrane that allows for solvent interdif-
fusion but blocks polymer interdiffusion, simply because its pores are too tight for
the large polymer coils to diffuse through, whereas the small solvent molecules
can. As a result, since the polymer cannot diffuse out of its vessel to spread over the
whole system (and thereby equilibrate its chemical-potential gradient), in turn, part
of the solvent will diffuse into the polymer vessel such to dilute the polymer solu-
tion in there. This influx of solvent into the polymer chamber creates a mechanical
pressure, in the present case by raise of a fluid column in a vertical tube. Once this
mechanical pressure, p, is equal to the osmotic pressure of the solution, ∏, equilib-
rium is reached. At this state, the elevation of the fluid column in the vertical tube,
Δh, has come to a final value. We may then write:

m · g Vρ · g πr2 Δhρ · g
Π=p= = = = Δhρ · g (4:34)
Area Area πr2
4.3 Osmotic pressure 149

Figure 59: Experimental setup to determine the osmotic pressure of a polymer solution. The
polymer cannot pass through the pores of a membrane separating it from excess surrounding
solvent, whereas the solvent molecules can. As a result, influx of solvent into the solution-
containing compartment will cause a mechanical pressure, for example, by rise of a fluid column in
a tube. Equilibrium is reached when this mechanical pressure matches and balances the osmotic
pressure of the solution. Picture inspired by B. Tieke: Makromolekulare Chemie, Wiley VCH, 1997.

With eq. (4.34), we have a simple means to measure ∏, simply by measuring


Δh. ∏, in turn, is a central parameter in the equation of state of an osmotic system,
which is Van’t Hoff’s law for dilute solutions:

Π · V = n · RT (4:35)

Rearrangement yields a relation between the osmotic pressure, ∏, and the polymer
mass-per-volume concentration, c:

n · RT m RT c · RT RT
Π= = · = ) Mn = (4:36)
V Mn V Mn Π=c

With that, we can determine the number-average molar mass of our polymer simply
by a concentration-dependent osmometric experimental series.62 Note that in the

62 In general, the number-average molar mass of a dissolved species can be determined by solution-
based effects whose magnitude depends on the number of dissolved molecules. Such effects are known
as colligative properties (derived from Latin collegere = to collect), the prime examples of which being
the osmotic pressure, the boiling-point elevation, and the freezing-point depression of a solution. All
these effects have a magnitude that is proportional to the number of dissolved molecules, and therefore
they allow the molar amount of the dissolved compound in the solution to be determined. If the solu-
tion has been made from a known mass of that dissolved compound, then dividing the one by the
other gives the weight per mole, and therewith, the molar mass. We may express this general principle
150 4 Polymer thermodynamics

upper notation, the concentration c = m/V has the unit g·L–1, as it is common in
polymer science. In classical physical chemistry, by contrast, concentration is com-
monly defined as n/V with the unit mol·L–1. In this form, Van’t Hoff’s law would
read ∏ = c RT instead of the polymer-science variant noted in eq. (4.36).
Van’t Hoff’s law is valid for dilute solutions. Nondilute solutions deviate from
that simple relation and are best described by a virial series expansion of it

1
Π=c = RT + A 2 c + A3 c 2 + . . . (4:37)
Mn

In the notation of eq. (4.37), the osmotic pressure is given in a reduced form, ∏/c.63
A key parameter in the above equation is the second virial coefficient, A2. It ac-
counts for deviations from the ideal Van’t Hoff law due to interactions at higher
concentration. Thus, A2 is an important parameter to quantify such interactions.
Often, experimental data can be well fitted to eq. (4.37) in a linear form, where the
series is expanded only up to the linear A2 term. Figure 60 sketches such a dataset
and linear fit. From the intercept, RT/Mn, we get the polymer’s number-average
molar mass, and from the slope, RT·A2, we get the second virial coefficient A2.
The second virial coefficient quantifies the interaction between a polymer and a
solvent. With that, it is closely connected to the Flory–Huggins parameter χ, as we
will discuss in more detail below. Generally, a positive A2-value means good solvation

quantitatively as follows. Let Y be a colligative property (i.e., one of the above three effects), whose
extent is proportional to the number of dissolved molecules of species i per volume, Ni/V, with a con-
N
stant of proportionality K: Y = K Vi . If there are several different species in the solution, for example,
P
N
several different molar masses of a polymer with a polydisperse distribution, then we get Y = K Vi i . In
P P
w NM
polymer science, a typical measure of concentration is mass per volume, that is, c = Vi i = Ni Vi i . If
A
we normalize our colligative property Y by the concentration c (see the next footnote for some back-
P
N
ground on this way of normalization), we obtain: Yc = K Vi i · P AN M . Together with the definition of
N V
i i i
P
NM
the number-average molar mass, Mn = P i i i , this yields: Y = KNA . We see from that identity that all
N c Mn
i
colligative properties are connected to the number-average molar mass Mn, and therefore they all
allow it to be determined experimentally. The extent of how strongly a colligative property Y depends
on N is set by the constant of proportionality K. For the osmotic pressure, this constant is particularly
high, so the effect is very strong for this kind of colligative property. This is the reason why it is fa-
vored as a method in polymer science: in a given mass-amount of polymer sample we only have a
few molecules (but each one very large in turn), and so we need a colligative property where even
such few (large) molecules have a strong effect that is well measurable. Osmotic pressure is such a
favorable property. On top of that, there is another reason why this method is beloved by polymer
scientists: in addition to Mn, it also gives quantitative information on the quality of polymer–solvent
interactions. This is what the following discussion in this section is about.
 A reduced quantity is one that is normalized by concentration. Other such normalized quanti-
ties are known to you from elementary physical chemistry already, for example, specific quantities
(normalized by mass) or molar quantities (normalized by the molar number).
4.3 Osmotic pressure 151

Figure 60: Plot of the series expansion of the concentration-dependent osmotic pressure in its
typical form.

of the polymer, whereas a negative A2-value means nonsolubility; in the Θ-state, A2


is zero.64 Figure 61 schematizes typical data for poly(methyl methacrylate) in various
solvents as an example. The quality of these solvents can be determined by the abso-
lute values of A2, and with that, by the slopes in the plot. m-Xylene turns out to be a
Θ-solvent with a slope of zero. Dioxane has a positive linear slope and therefore posi-
tive A2, denoting it to be a good solvent. For the best solvent, chloroform, the concen-
tration dependence of the osmotic pressure is even stronger, and it even deviates from
the linear form, which makes a third nonlinear virial term necessary to fit the exper-
imental data. Despite all these different slopes, however, the intercept is the same
for all samples, as this is only dependent on the molar mass average, Mn, but not
on the solvent quality.

4.3.1 Connection of the second virial coefficient, A2, to the Flory–Huggins


parameter, χ, and the excluded volume, ve

We have already hinted at the close interrelation between the second virial coeffi-
cient A2 to the Flory–Huggins interaction parameter χ, as both are a measure of
polymer–solvent interactions. The following paragraph will examine this relation-
ship and determine it quantitatively.

64 Again, we can make an analogy to a real gas, whose equation of state can be written by a virial
 ð1 + V
series as p = RT
V  + . . .Þ with the second virial coefficient B = b − RT , wherein a and b are the van
B a

der Waals coefficients, with a accounting for attractive intermolecular interactions, and b account-
ing for the covolume and other repulsive interactions of the gas molecules. At the Boyle tempera-
ture, attractive and repulsive interactions are at balance, causing B to be zero and the virial series
thereby to simplify to the ideal gas law. The real gas then displays pseudo-ideal behavior, analo-
gous to polymers in a solvent at the Θ-temperature.
152 4 Polymer thermodynamics

Figure 61: Schematic of the concentration-dependent reduced osmotic pressure for poly(methyl
methacrylate) in different solvents. It exhibits strongly differing slopes, corresponding to strongly
differing second virial coefficients, which is due to the different solvent qualities for the polymer.
Picture inspired by B. Vollmert: Grundriss der Makromolekularen Chemie, Springer, 1962.

As a starting point, we take the thermodynamic definition of the osmotic pressure:

∂ΔFmix n · ∂ΔF mix


Π≡ − = − (4:38)
∂V nA ∂V nA

We have seen in the last section that the second virial coefficient A2 is related to the
reduced osmotic pressure, ∏/c. Hence, we must transform the above equation into
a practical expression for the osmotic pressure as a function of concentration. To
achieve that, we use the expression for ΔFmix that we have appraised with the lat-
tice model of the Flory–Huggins mean-field theory in Section 4.1. This expression
uses the volume fraction, ϕ, as a measure of concentration, which is connected to
the total volume of the system, V, by

nA NA v0
ϕA = (4:39a)
V

Here, nA is the number of A-molecules on the lattice, NA the number of lattice sites
per A-molecule, which is equal to the degree of polymerization of species A, and v0
the volume of a single lattice site.
Rearranging eq. (4.39a) and expressing it in a differential form yields

1 nA NA v0
∂V ¼ nA NA v0 ∂ ¼ ∂ϕ (4:39b)
ϕ ϕ2

whereby the identity d(1/x) = –x–2 dx was used. We can insert this relation into the
above expression for the osmotic pressure and further use the identity nϕA = nA NA =>
n = nA NA / ϕ to change the variable V to the variable ϕ:
4.3 Osmotic pressure 153

0  1 0  1

ΔF 
 mix 2 ∂ n N · mix =ϕ 2 ∂ ΔF mix =ϕ
n · ∂ΔF ϕ A A ϕ
Π≡ − =@ · A =@ · A
∂V nA nA NA v0 ∂ϕ v0 ∂ϕ
nA nA

(4:40)

The Flory–Huggins mean-field solution for the Free Energy of mixing is

mix = kB T ϕ ð1 − ϕÞ
ΔF ln ϕ + lnð1 − ϕÞ + χϕð1 − ϕÞ (4:22)
NA NB

We now have to differentiate this term. Unfortunately, this is difficult for the B com-
ponent part of the equation, as our variable ϕ enters in a complicated form here. To
simplify this point of our calculation, we use a series expansion according to
!
ð 1 − ϕÞ 1 ϕ2 ϕ3
lnð1 − ϕÞ = −ϕ+ + + ... (4:41)
NB NB 2 6

By plugging this into the term for ΔFmix, we generate


!
2
mix = kB T ϕ ln ϕ + ϕ χ − 1 + ϕ 1 ϕ3
ΔF − 2χ + + ... (4:42)
NA NB 2 NB 6NB

This expression, in turn, can be inserted into our original formula for the osmotic
pressure, eq. (4.40)
!
k B T ϕ ϕ2 1 ϕ3
Π= + − 2χ + + ... (4:43)
v0 NA 2 NB 3NB

In the next step, we transform further by inserting the concentration as a variable


according to c = ϕ/v0. This variable has the unit L–1; it expresses the number of
chain segments per volume, that is, per liter:

c c2 1
Π = kB T + v0 − 2χ + . . . (4:44a)
NA 2 NB

Division of this formula by c generates the needed expression for the reduced os-
motic pressure, ∏/c

1 c 1
Π=c = kB T + v0 − 2χ + . . . (4:44b)
NA 2 NB

When comparing the general form of the above formula to the one of eq. (4.37), we
realize that they both are of kind
154 4 Polymer thermodynamics

1
Π=c = Thermal Energy + Interaction Parameter · c (4:45)
Chain Length

Equation (4.44b) has a dimensionless first summand in the parentheses; this means
that the second summand must be dimensionless, too. As a result, since c has the
unit L–1, the interaction parameter in front of it must have a unit of liters, L. Hence,
this parameter must be some kind of volume that quantifies the polymer–solvent
interactions. A quantity doing this is the excluded volume, ve, which therefore
serves as the second virial coefficient in eq. (4.44b):

1
ve = v0 − 2χ (4:46a)
NB

Here, NB is the degree of polymerization of the second component, which is either a


solvent or a second polymer. In a polymer solution, we have NB = 1. When we now
plug in the Flory–Huggins parameter for the Θ-state, χ = ½, then we end up with an
excluded volume of ve = 0, exactly as the definition of the Θ-state demands. For the
case of an athermal solvent, by contrast, χ = 0, and as a result, ve = v0. The excluded
volume is then maximally positive and coincides with the monomer-segmental vol-
ume, as we have already seen in Section 3.2. In a mixture of two different polymers,
a polymer blend, we have NB >> 1. A Flory–Huggins parameter close to zero, χ ≈ 0,
which is needed to successfully blend two polymers, now leads to an excluded vol-
ume also close to zero, ve = v0/NB ≈ 0. This shows us that, even though the A–B in-
teraction in a polymer blend is very good (it must be so to mix!), the chains adopt
their ideal conformation, as they do in the Θ-state.
We can further express the above equation as

v0 l3
ve = − 2v0 χ = − 2l3 χ (4:46b)
NB NB
Here, l3/NB is the volume of the polymer segments (normalized to the degree of poly-
merization of the solvent), which corresponds to the hard-sphere branch of the Mayer
f-function. 2l3χ is a measure of the effective segment–segment attraction, which corre-
sponds to the potential-well part of the Mayer f-function. The concert of both deter-
mines the excluded volume (remember that this volume is in fact calculated from the
negative integral over the Mayer f-function). We now see once more that for the Θ-
state (ve = 0), coil expansion due to the co-volume of the segments just exactly balan-
ces the coil contraction due to the net-attractive segment–segment interaction.
So far, we have expressed the reduced osmotic pressure on the molecular level,
meaning we use kBT for the thermal energy and a concentration c = ϕ/v0 with the
unit L–1 to account for the number of molecules per volume. To make this more
practical, we can transform eq. (4.44b) to molar values by using RT for the thermal
energy and c = ϕ/vspecific,polymer with the unit g·L–1 as the concentration. For a poly-
mer solution (NB = 1), this yields
Questions to Lesson Unit 9 155

!
1 v2specific, polymer 1
Π=c = RT +  Solvent −χ c+ ... (4:47)
Mn V 2

When we compare this to the Van’t Hoff equation

1
Π=c = RT + A2 c + . . . (4:37)
Mn

we immediately recognize the connection between the second virial coefficient, A2,
and the Flory–Huggins parameter, χ:

v2specific, polymer 1
A2 =  Solvent −χ (4:48)
V 2

It is thanks to this relation that we can determine χ through an osmotic pressure


experiment as described at the beginning of this chapter.

Questions to Lesson Unit 9

(1) Which molar-mass average can be determined by colligative phenomena?


a. The number-average molar mass Mn , since colligative properties depend on
the number of (macro)molecules in a solution.
b. The viscosity-average molar mass Mη, since colligative properties depend
on the viscosity of a (macro)molecular solution.
c. The weight-average molar mass Mw , since colligative properties depend on
the mass of the (macro)molecules in a solution.
d. The centrifugal-average molar mass Mz , since colligative properties depend
on the amount of (macro)molecules in a solution.

(2) Which colligative property is preferred for studying polymer solutions?


a. The depression of the freezing point, since this is most pronounced for poly-
mer solutions.
b. The elevation of the boiling point, since this is most pronounced for poly-
mer solutions.
c. The osmotic pressure, since this is most pronounced for polymer solutions.
d. All three of the previously mentioned colligative properties can be used
equally since all three depend on the number average of the polymer mass.
156 4 Polymer thermodynamics

(3) An osmometer consists of two chambers, one inside the other; the inner cham-
ber contains a polymer solution at the beginning and the outer chamber con-
tains only solvent. The chambers are separated by a semipermeable membrane.
What does that semipermeable membrane do?
a. It allows only the polymer to pass, whereas the solvent remains in the inner
chamber. Pressure builds up from outside on the inner chamber.
b. It allows only the solvent to pass, so the solvent flows out of the inner
chamber in this case. Pressure builds up from the outside on the inner
chamber.
c. It only allows the solvent to pass, so the solvent flows out of the outer
chamber into the inner chamber. The incoming solvent increases the pres-
sure in the inner chamber.
d. It allows the solvent and polymer to pass, but they can only escape from
the inner chamber into the outer chamber and not vice versa. The solution
therefore flows into the outer chamber and thereby exerts pressure onto the
inner chamber.

(4) The excluded volume ve , the second virial coefficient A2 of the modified Van’t
Hoff equation, and the Flory–Huggins parameter χ are related because ______
a. all are linearly dependent on the concentration of the polymer.
b. all are dimensionless quantities.
c. all are a measure of polymer–solvent interactions.
d. all are a measure of the degree of polymerization of the polymer.

(5) What is the rising height in an osmosis cell as the one shown in Figure 59 for a
(dilute) aqueous polymer solution at a concentration of 0.1 mol L−1 at normal
conditions?
a. 2.5 cm
b. 25 cm
c. 2.5 m
d. 25 m

(6) The Van’t Hoff equation can be modified for the undiluted, nonideal case by a vi-
rial series expansion – similar to the way the ideal gas equation can be trans-
formed into a more general form by a related virial series expansion. At a special
temperature, the Boyle temperature, the real gas equation simplifies to the ideal
gas equation. Which case is analog to this for a polymer solution?
a. Athermal solvent, A2 ¼ 0
b. θ -Solvent, A2 ¼ 0
c. Good solvent, A2 > 0
d. Nonsolvent, A2 < 0
Questions to Lesson Unit 9 157

(7) Particularly good solvents show a deviation of the commonly linear concentration
dependence of the reduced osmotic pressure. To what extent can this course be
described?
a. The course cannot be described with the given theory; for this scenario, a
new equation based on a more complex theory is needed.
b. The course can only be described section by section; small sections corre-
spond to a locally linear course.
c. The course can be modeled by including another virial coefficient A3 , which
represents a quadratic dependence of the reduced osmotic pressure on the
concentration.
d. The progression can be modeled by adding an appropriate constant C to the
existing equation.

(8) How can the θ-temperature for a dilute solution be determined via osmometry?
a. The temperature at which the concentration-dependent profile of the re-
duced osmotic pressure passes through the origin is determined; this tem-
perature is the θ-temperature.
b. The temperature is determined at which the concentration-dependent course
of the reduced osmotic pressure has a slope of zero; this temperature is the θ-
temperature.
c. The temperature is determined at which the concentration-dependent course
of the reduced osmotic pressure has a minimum; this temperature is the θ-
temperature.
d. The temperature at which the concentration-dependent course of the re-
duced osmotic pressure deviates from the linear course is determined; this
temperature is the θ-temperature.
5 Mechanics and rheology of polymer systems

LESSON 10: RHEOLOGY


The most relevant properties of polymers, on which ground they have taken dominance in
our materials world, are their mechanical ones. In this lesson, you will be introduced to the
fundamentals of a methodology to assess these mechanics: rheology. You will become ac-
quainted with some elementary physical and working principles in this field and get to know
the most relevant material property of polymers: the elastic modulus.

5.1 Fundamentals of rheology

At the very beginning of this book, it has been pointed out that the prime goal of
polymer physical chemistry is the understanding of the relations between the struc-
ture and properties of polymer-based materials, thereby bridging the fields of
polymer chemistry and polymer engineering, as shown in Figure 1. To achieve
this goal, so far, we have studied the structure and dynamics of single chains and
the thermodynamics of multichain systems in the preceding chapters; with that, we
have erected three fundamental pillars of that foresaid bridge. Now, the construc-
tion of the actual bridge will be the goal of the forthcoming chapter. This will help
us to rationally and quantitatively understand why many of the polymeric materials
that we encounter in our everyday lives behave the way they do. Our prime focus
will be on those properties of polymers that have made greatest impact on our lives
in the past decades, which is their mechanical properties. Hence, we will first deal
with a fundamental introduction to the field of rheology.
The term rheology derives from the Greek words rheos, meaning “flow”, and
logos, meaning “understanding”. The name reminds of a famous aphorism voiced
by the Greek philosopher Simplicus, who had stated that panta rhei, “everything
flows”. As such, rheology is the study of the flow of matter, and therefore the
branch of physics that deals with the deformation and flow of materials, both solids
and liquids.

5.1.1 Elementary cases of mechanical response

Two elementary extreme cases can be distinguished from one another in view of a
material’s mechanical properties: it can behave either as an elastic solid, like a
rubber band, or as a viscous fluid, like water or syrup.
In the first case of an ideal elastic solid, the energy introduced into the material
through exertion of an external force, f, which can also be expressed in a normalized

https://doi.org/10.1515/9783110713268-005
160 5 Mechanics and rheology of polymer systems

form as a stress, σ = f/A (with A being the cross-sectional area of the material speci-
men onto which the force acts), is stored elastically. In this case, the material is im-
mediately deformed to an extent proportional to the applied stress, or vice versa, a
counteracting stress builds up in it upon and proportional to deformation by a given
extent. This proportionality of the extent of deformation, or strain, ε, to the stress, σ,
is captured by Hooke’s law:

σ=E·ε (5:1)

In this law, the strain is the relative extent of deformation, ε = ΔL/L0, that is, the
change of the material specimen length, ΔL, relative to its length before deformation,
L0. The constant of proportionality is the elastic modulus, also referred to as Young’s
modulus, E. In a mathematical view, from eq. (5.1), we realize that the elastic or
Young’s modulus has a unit of N·m–2 = Pa, because the stress is described by the unit
N·m–2 = Pa as well, whereas the strain, ε = ΔL/L0, is a dimensionless quantity. In a
conceptual view, the modulus is an elementary material parameter that quantifies
how much energy can be stored in the material upon deformation by a given extent.
In this view, it is illustrative to note that the unit N·m–2 may also be written as J·m–3,
which denotes the energy density in the material that we need to work against when
we want to deform it; it originates from the energies and sizes of the material’s build-
ing blocks (e.g., polymer-network strands of energy kBT and size ξ in a rubber). As we
already see from the latter note, the capability for energy storage is directly related to
the internal microscopic structure and dynamics of the material. Hence, the modulus
is nothing less than a quantitative expression of a structure–property relation. With
knowledge of the modulus, we can calculate from eq. (5.1) how hard it is to achieve a
certain extent of deformation upon application of a given stress, which is of direct rel-
evance for applications. As soon as the stress cedes, the stored energy is released, and
the material snaps back to its original shape. This means that elastic deformation is
reversible. This type of mechanical response can be visualized illustratively in the
form of an elastic spring, as shown in Figure 62(A).
In the second case of an ideal viscous fluid, the exertion of an external stress
also causes a deformation. Here, however, the stress, σ, is proportional not to the
deformation itself but to its rate, dε/dt. It follows that the material will reside in its
deformed state as soon as the stress cedes. There is no energy storage in this case;
instead, the energy of deformation is dissipated to the environment as heat, which
means that viscous deformation is irreversible. The proportionality between the
rate of deformation, dε/dt, and the stress, σ, is captured by Newton’s law:


σ=η· = η · ε_ (5:2)
dt
5.1 Fundamentals of rheology 161

Figure 62: Elementary cases of mechanical response in rheology. (A) Hooke’s law, relating the
stress, σ, imposed on an ideal elastic solid, to its strain, ε, via the Young’s modulus, E. This model
can be visualized in the form of an elastic spring with a spring constant of E. (B) Newton’s law,
relating the stress, σ, imposed on an ideal viscous fluid to its time-dependent rate of deformation,
dε/dt, via the viscosity, η. This model can be visualized in the form of a dashpot filled with a fluid
of viscosity η.

In this law, the constant of proportionality is the viscosity, η. Just like the modu-
lus in the elastic case above, the viscosity is an elementary material parameter that
quantifies how much resistance a material exerts on flow.65 As we will see later in
this chapter, the capability for exerting resistance on flow can be directly related to
the internal microscopic structure and dynamics of a material. Hence, just like the
elastic modulus, the viscosity is a quantitative expression of a structure–property re-
lation. The viscous type of mechanical response can be visualized illustratively in the
form of a dashpot filled with a fluid, as illustrated in Figure 2(B). In view of eq. (5.2),
we realize that the viscosity has the unit Pa·s, because the stress is described by the
unit N·m–2 = Pa, and the rate of deformation has the unit s–1.

5.1.2 Different types of deformation in rheology

Strain can be applied to a material in many directions, but we may simplify our view
to two particularly relevant cases. To visualize these, consider a cube with edges A,
B, and C. In this cube, A is the edge in and out-of-plane, B is the vertical edge, and C
the horizontal one (see Figure 63).
In the first case, a uniform stress exerted onto the A–B face causes uniaxial strain.
This strain, ε, can be calculated from the ratio between the deformation, ΔC, and the
original length of the edge C. For the two elementary cases of an elastic solid and a

 Later on, in Sections 5.6.1 and 5.7.1, we will see that the elastic modulus and the viscosity are
connected to one another by one of the most central equations in the field of rheology: η = E τ.
162 5 Mechanics and rheology of polymer systems

Figure 63: Two ways to deform a material: (A) a uniform strain exerted onto the A–B face is called
uniaxial strain. (B) A strain exerted onto the A–C face is called shear. In both cases, the stress, σ,
can be calculated with Hooke’s law for an ideal elastic material or with Newton’s law for an ideal
viscous fluid.

viscous fluid, Hooke’s and Newton’s law can be applied as stated in Section 5.1.1 to
calculate the stress, σ, related to such a uniaxial strain (see Figure 63(A)). In the second
case, the stress is exerted onto the A–C face. This situation is called shear. The result-
ing shear strain, γ, is calculated from the ratio between the deformation, Δc, and the
original length of edge B, which corresponds to the tangent of the shearing angle α
(see Figure 63(B)). For an ideal viscous fluid, the stress, σ, is still calculated using New-
ton’s law as stated in Section 5.1.1. Hooke’s law, however, needs to be modified to in-
corporate a new proportionality constant: the shear modulus, G.
Both the Young’s modulus, E, and the shear modulus, G, are connected to one
another in the form of

E = 2Gð1 + μÞ (5:3)

In eq. (5.3.), µ is named Poisson’s ratio; it is a measure of the Poisson effect, a phe-
nomenon according to which a material tends to expand in directions perpendicu-
lar to the direction of compression or, vice versa, tends to laterally contract upon
extension; both is a consequence of the retention of the material body’s volume
upon compression or extension. µ quantifies this effect by relating the relative lat-
eral contraction of a material specimen, expressed by the change of its thickness
relative to the one before deformation, Δd/d0, to its axial strain, ΔL/L0:

Δd=d0
μ= (5:4)
ΔL=L0

Incompressible objects such as fluids have a Poisson’s ratio of µ = 0.5. This value
also commonly applies to polymers, because linear extension of a polymer sample
5.1 Fundamentals of rheology 163

is usually accompanied by a concurrent lateral contraction, meaning that it does


not undergo a volume change upon stretching. In this case, E = 3G.

5.1.3 The tensor form of Hooke’s law

The equations for Hooke’s and Newton’s laws named above are actually simplified
forms for a stress acting normal to the sample and in one direction only. In their gen-
eralized forms, they are tensor equations that account for all possible spatial direc-
tions. In this generalized form, Hooke’s law reads

τ=c·γ (5:5)

with τ the stress tensor, c the elasticity tensor, and γ the strain tensor. With this
nomenclature, we follow a common convention, in which – unfortunately – the ten-
sor form of the stress is abbreviated with the symbol τ. This is inconvenient, be-
cause in polymer physics, τ is very commonly used to abbreviate the relaxation
time as well.66 To avoid this confusion, we use the symbol σ for the special-case
variant of normal stress in this textbook; this is the stress with both its directional
indices being the same, that is σ11 = τ11, σ22 = τ22, and σ33 = τ33.
To illustrate the relation of eq. (5.5), again consider a cube representing a volume
section of an elastic body. If we impose a stress τik in the direction k, the cube will
deform, but now, the resulting strain will not be limited to just one spatial direction,
but instead, it will have contributions for all spatial directions (see Figure 64).
This relation is reflected by the following equation:
X
3 X
3
τik = ciklm · γlm = cik11 · ε11 + cik12 · γ12 + cik13 · γ13 +
l=1 m=1

cik21 · γ21 + cik22 · ε22 + cik23 · γ23 + cik31 · γ31 + cik32 · γ32 + cik33 · ε33 (5:6)

The strength of each contribution is given by the elastic constant ciklm. Spatial direc-
tions for each value are indicated by the indices l and m.
For a comprehensive consideration, we would have to evaluate all nine stress
components (σ11 , τ12 , τ13 , τ21 , σ22 , τ23 , τ31 , τ32 , σ33 ), which are connected by nine elastic
constants ciklm to nine strain components (ε11 , γ12 , γ13 , γ21 , ε22 , γ23 , γ31 , γ32 , ε33 ). This
adds up to 81(!) single components to fully describe the deformation of a cubic vol-
ume. Fortunately, if we consider the symmetry of the body, this expression can be
severely simplified. If the body is isotropic, meaning it is uniform in all orientations,

 If you are unsure whether a symbol τ in a textbook or a paper represents stress or (relaxation)
time, check its physical unit: if it is Pascal, then τ represents a stress, and if it is seconds, then τ
represents a time.
164 5 Mechanics and rheology of polymer systems

Figure 64: Deformation of an elastic body taking into account all spatial directions. (A) A stress τik
imposed in the direction k results in a (B) corresponding strain γik in the same direction. Picture
redrawn from C. Wrana: Polymerphysik, Springer Verlag Berlin, Heidelberg, 2014.

the number of independent constants is reduced to two. Polymers exhibit homoge-


neous and amorphous structures and can therefore be considered to be isotropic.
We can now apply this to the two cases of uniaxial strain and shear deformation,
which leads to the following equations for Hooke’s law considering all three spatial
direction x, y, and z:

Uniaxial strain: Shear deformation:


σ xx = E · εxx ; σ yy = E · εyy ; σ zz = E · εzz τ xy = G · γxy ; τ xz = G · γxz ; τ yz = G · γyz

These are just the equations we introduced in Section 5.1.1. We can see that ε and γ
are, in fact, special-case variants of γ, σ and τ are special-case variants of τ, and E
and G are special-case variants of c.

Questions to Lesson Unit 10

(1) Which property of polymers cannot be probed by rheological experiments?


a. Their compressibility
b. Their elasticity
c. Their flow rate
d. Their viscosity

(2) Which are the two extreme cases of a polymer’s ideal mechanical response?
a. The elastic solid and the viscous fluid
b. The elastic solid and the viscous solid
c. The glassy solid and the elastic fluid
d. The glassy solid and the viscous fluid
Questions to Lesson Unit 10 165

(3) Which statement does not hold for the case of an elastic solid?
a. The strain of the material scales linearly with the stress applied to it.
b. The deformation of the material is irreversible.
c. An elastic solid can be simply modeled as a spring.
d. The mechanical response is described by Hooke’s law with the elastic mod-
ulus as a proportionality factor.

(4) What is the relation between the elastic modulus and the microscopic structure
of an elastic solid?
a. There is no trivial relation between them.
b. As the elastic modulus is a parameter that quantifies the elasticity, it is di-
rectly related to the lifetime of the transient agglomerates formed due to
deformation.
c. As the elastic modulus is a parameter that quantifies the extent of deforma-
tion, it is directly related to the number of bonds between the building
blocks that are broken during deformation.
d. As the elastic modulus is a parameter that quantifies the stored energy, it is
directly related to the coupling and sizes of the building blocks that are
moved due to deformation.

(5) What is stress (in mechanics)?


a. Stress is a different denomination for pressure. They both have the same di-
mension with the unit “Pascal.”
b. Stress is a measure of the internal forces in a continuous medium that
builds up because of an external load.
c. Stress is the deformation of a continuous medium caused by an applied ex-
ternal force.
d. Stress is the internal tension in a continuous medium caused by an external
force per volume segment of the medium.

(6) Which statement about strain is true?


a. Strain is the relative deformation of a continuous medium caused by an ap-
plied force or stress.
b. Strain is the angular deformation of a continuous medium caused by an ap-
plied force or stress.
c. A body always retains its volume when being strained.
d. A body always retains its shape when being strained.
166 5 Mechanics and rheology of polymer systems

(7) Which statement does not hold for the case of a viscous fluid?
a. The strain of the material scales linearly with the stress applied to it.
b. The deformation is irreversible.
c. A viscous fluid can be simply modeled as a dashpot.
d. The behavior is described by Newton’s law with the viscosity as a propor-
tionality factor.

(8) Which model describes the actual mechanical response of a polymer?


a. That of an ideal elastic solid since the bonds between the building blocks
work as springs which can snap back into shape after deformation.
b. That of an ideal viscous fluid since the movement of the building blocks
due to deformation is irreversible.
c. It depends on the experimental parameters. Based on those, polymers dis-
play either fully elastic or fully viscous behavior.
d. It depends on the experimental parameters. Based on those, polymers may
display elastic or viscous or both characteristics. This is described as the
term “viscoelasticity.”
5.2 Viscoelasticity 167

5.2 Viscoelasticity

LESSON 11: VISCOELASTICITY


The rich mechanical properties of polymers emerge because they are neither a true fluid nor
a true solid; instead, they display characteristics of both, referred to as viscoelasticity. This
lesson will show how this duality manifests itself in a polymer’s time- or frequency-
dependent stress relaxation or strain creep. It will also show how this is captured by com-
plex-number material constants, such as the complex modulus.

Often, materials exhibit both elastic as well as viscous properties. These materials
are named viscoelastic. Strictly speaking, all matter exhibits this dualism. For ex-
ample, imagine watching a glacier for one day. You will not be able to observe any
perceivable movement of the ice and could therefore say that it is a solid. Now
imagine visiting the same glacier once a day over the course of an entire year. Every
time you visit, you take a photograph, and in the end you edit all pictures together
to create a time-lapse movie. The glacier will appear to be flowing like a fluid. An-
other example is the big, ornamental glass windows that can be found in old cathe-
drals: often, they are thicker at their bottom than at the top. This difference is
mainly related to the process with which they are produced: in historic times,
church windows were manufactured upright, which meant that the cooling glass
melt would flow a little from top to bottom inside the manufacturing frame. In addi-
tion to this manufacture-related thickness gradient, however, the glass of these
windows still flows today, and over very long timescales (longer than centuries in
fact), it causes further thickening of the window bottoms compared to their tops.
Viscoelasticity is a phenomenon most strikingly encountered in polymer
systems. In contrast to glaciers and glass windows, polymer systems usually transi-
tion from a regime where their elastic properties dominate to one where the viscous
properties are dominant on timescales that are right in the practically and experimen-
tally relevant window of milliseconds to seconds, depending on the specific system
at hand. This makes polymers highly relevant for practical applications, because
their mechanical properties are versatile, dynamic, and, with the proper understand-
ing of the underlying concepts that govern their mechanics, can be rationally de-
signed to serve a specific purpose.
Polymer’s mechanical properties are usually assessed and quantified in three
types of rheological experiments, as will be discussed in the following subsection.
168 5 Mechanics and rheology of polymer systems

5.2.1 Elementary types of rheological experiments

5.2.1.1 The relaxation experiment


In a relaxation experiment, uniaxial or shear strain, ε or γ, is applied to a sample
and the resulting stress in the material, σ, is recorded as a function of time, as shown
in Figure 65. An elastic sample shows direct proportionality of both quantities accord-
ing to Hooke’s law. The proportionality factor is the time-dependent relaxation mod-
ulus E(t). As long as strain is applied, an equivalent stress is recorded. An ideal
viscous sample, instead, displays two spikes of the stress, one at the moment where
the stress is abruptly increased and one at the moment where the stress is abruptly
decreased. This is because according to Newton’s law, only the time-dependent rate
of deformation is considered, and hence, at times where the strain is constant, no
stress is recorded, whereas an infinite spike of stress occurs during an abrupt change
of the strain. In the case of a viscoelastic sample, the initial response is similar to the
elastic one: the stress increases proportionally to the strain. Once the strain is kept at
a constant level, though, some of the stored energy of deformation is dissipated as
heat, leading to stress relaxation over time. If the sample is a viscoelastic solid, the
stress will reach a plateau value with a plateau modulus Eeq = E(t ! ∞).

Figure 65: In a relaxation experiment, strain, ε or γ, is applied to a sample, and the time-dependent
stress, σ, that builds up in the sample is recorded. An ideal elastic sample shows direct
proportionality of both quantities according to Hooke’s law. In an ideal viscous sample, the stress
displays spikes at times of abrupt change of the strain. In a viscoelastic sample, the stress in the
material relaxes to a plateau value for t ! ∞, because some of the energy of deformation is
dissipated to heat over time.

5.2.1.2 The creep test


In a creep test, stress, σ, is applied to a sample, and the resulting uniaxial or shear
strain of the specimen, ε or γ, is recorded as a function of time (see Figure 66). An
elastic sample shows direct proportionality of both quantities according to Hooke’s
5.2 Viscoelasticity 169

Figure 66: In a creep test, stress, σ, is applied to a sample, and the resulting time-dependent strain
of it, ε or γ, is recorded. An ideal elastic sample shows direct proportionality of both quantities
according to Hooke’s law. In an ideal viscous sample, the strain increases linearly with time as
long as constant stress is applied; the material exhibits creep. In a viscoelastic sample, the
response is a mix of both: the recorded strain value rises instantly, but not to same extent as in the
elastic sample, and with time it rises further due to creep of the material. A viscoelastic solid
reaches a plateau for t ! ∞, whereas a viscoelastic liquid proceeds creeping steadily. Once the
application of stress is released, the sample creeps back with time.

law. The proportionality factor is called the time-dependent creep compliance, J(t).
As long as stress is applied, an equivalent strain is recorded. An ideal viscous sample
exhibits a strain that increases linearly as long as the stress is kept constant; the ma-
terial exhibits creep. Once the test is finished by release of the stress, the strain does
not return its original level, thereby showcasing the irreversibility of viscous flow. In
case of a viscoelastic sample, the response is a mix of both: once a constant amount
of stress is applied, the recorded strain value rises instantly, but not to the same ex-
tent as in the elastic sample. With time, however, it rises further due to creep of the
material. If the sample has more elastic than viscous character, it would reach a pla-
teau for t ! ∞, whereas it would proceed creeping steadily if the sample has more
viscous than elastic character. Once the test is finished and the application of stress
is released, the sample exhibits creep in the direction back.

5.2.1.3 The dynamic experiment


Another method to probe a material’s mechanical properties is the dynamic exper-
iment. Here, a sinusoidally modulated stress, σ = σ0 exp(iωt), is applied to the sam-
ple, and the time-dependent strain, ε or γ, is recorded, as illustrated in Figure 67.
This can also be done vice versa, that is, a sinusoidally modulated strain is applied
and the emerging time-dependent stress in the material is recorded. In the case of
an ideal elastic sample, both stress and strain are in phase, as both quantities are
directly proportional to each other at each time according to Hooke’s law; as a
170 5 Mechanics and rheology of polymer systems

Figure 67: In a dynamic experiment, a sinusoidal stress, σ, is applied to a sample, and the time-
dependent strain, ε or γ, is recorded. Both the stress and strain curves are proportional and in
phase for an ideal elastic sample. By contrast, the time-dependent strain curve of an ideal viscous
sample exhibits a phase shift to the stress curve with a phase angle, δ, of 90° or π/2. In the
intermediate case of a viscoelastic sample, the stress and strain curves display a phase angle
between 0 and π/2. In this case, δ reflects the individual contribution of the elastic and viscous
parts of the sample.

result, their phase angle, δ, is zero. In the case of an ideal viscous sample, accord-
ing to Newton’s law, it is not the strain itself but its derivative that is proportional
to the applied stress; hence, an original sinewave condition of strain causes a co-
sine answer of stress, or vice versa, an original sinewave condition of stress causes
a minus-cosine answer of strain. In either case, as a result, stress and strain have a
phase shift of 90°; this means that the phase angle, δ, is π/2. A viscoelastic sample
will be intermediate between these two extreme cases and display a phase shift
with phase angle, δ, in the range 0 < δ < π/2. From δ, the individual contribution of
the elastic and viscous parts of the viscoelastic sample can be calculated. The closer
δ is to zero, the more does the elastic contribution dominate, whereas the closer δ is
to π/2, the more does the viscous contribution dominate.

5.3 Complex moduli

In Hooke’s law, eq. (5.1), the elastic modulus reflects the ratio of stress, σ, to strain,
ε or γ. In the case of a dynamic experiment as just discussed, this modulus is a com-
plex quantity, E* or G*, connecting the ratio of the sinusoidally modulated stress
and strain and the phase angle between them, δ, as follows:

σ σ0 expðiωtÞ σ0
E* = = = expðiδÞ (5:7a)
ε ε0 expðiðωt − δÞÞ ε0
5.3 Complex moduli 171

σ σ0 expðiωtÞ σ0
G* = = = expðiδÞ (5:7b)
γ γ0 expðiðωt − δÞÞ γ0

With Euler’s formula, this can be rewritten as a trigonometric function with a real
and an imaginary part; these two parts can be viewed to be two different parts of
the complex modulus:
σ0
E* = ðcos δ + i sin δÞ
ε0

= E′ + iE′′ (5:8a)
σ0
G* = ðcos δ + i sin δÞ
γ0

= G′ + iG′′ (5:8b)

The first part is the storage modulus, E′ or G′; it captures the sample’s elastic prop-
erties. The second part is the loss modulus, E″ or G″; it captures the sample’s vis-
cous properties.
Storage modulus:
σ0
E′ = cos δ (5:9a)
ε0
σ0
G′ = cos δ (5:9b)
γ0

Loss modulus:
σ0
E′′ = sin δ (5:10a)
ε0
σ0
G′′ = sin δ (5:10b)
γ0

Depending on the phase angle δ, the first or the second contribution will dominate
the overall modulus E* or G*. If δ = 0, then E″ and G″ are zero, such that only the
elastic parts E′ or G′ contribute to E* or G*. By contrast, if δ = π/2, then E′ and G′ are
zero, such that only the viscous parts E″ and G″ contribute to E* or G*. In between
these two extremes, both parts contribute to E* or G*, and δ determines with what
relative magnitude they do.
The extent of the storage and the loss modulus contributions to E* and G* can
be judged from their ratio E″/E′ and G″/G′. Following the simple trigonometric eval-
uation shown in Figure 68, or alternatively, the simple trigonometric identities of
eqs. (5.11), these ratios can be described as the tangent of δ. It is called the loss
tangent, and its value describes the contributions of the elastic and the viscous
parts to a sample’s viscoelastic properties in one simple number. tan δ > 1 means
172 5 Mechanics and rheology of polymer systems

that the value of the loss modulus is larger than that of the storage modulus. The sam-
ple’s mechanical properties are then dominated by its viscous part, and it is therefore
called a viscoelastic liquid. When tan δ < 1, by contrast, the sample’s mechanical prop-
erties are dominated by its elastic contribution. The value of its storage modulus is
then larger than that of its loss modulus, and it is therefore called a viscoelastic solid.
When tan δ = 1, both the elastic and the viscous parts contribute equally.

σ0
E″ ε sin δ sin δ
= 0 = = tan δ (5.11a)
E′ σ 0 cos δ cos δ
ε0

σ0
G″ γ0 sin δ sin δ
= = = tan δ (5.11b)
G′ σ 0 cos δ cos δ
γ0

Figure 68: Visual representation of the complex modulus E* in an Argand diagram. The abscissa
represents the real part of the complex modulus, E′, whereas the ordinate represents the imaginary
part, E″. A vector connecting the value of E* and the origin exhibits a slope of tan δ, the loss
tangent.

The contributions of the storage modulus, E′ or G′, and the loss modulus, E″ or G″,
to the complex modulus E* or G* can be visualized in an experiment as shown in
Figure 69. When a person drops a rubber ball from the hand to the floor, it bounces
off the floor and shoots back up, but it does not reach its original height, because
some of the energy is dissipated to heat during the bounce. This amount of energy
can be directly related to the loss in height, and this is the contribution of the loss
modulus, E″ or G″, to E* or G*. The height that the ball is able to gain back when
bouncing is directly related to the energy stored elastically within the material. This
is the contribution of the storage modulus, E′ or G′ to E* or G*.
The latter example is a direct illustration of a structure–property relation. The
extent of the rubber ball bounce is a direct resemblance of the loss tangent of its
rubbery material, which is a ratio of the moduli E″ (or G″) and E′ (or G′). These mod-
uli, in turn, have to do with the material’s structure. As we will see later, in Sec-
tion 5.9, the ability of a rubbery polymer network to store mechanical energy is
proportional to its polymer-chain crosslinking density, as this determines the tight-
ness of the polymer-network meshes. The ability for energy dissipation, by contrast,
has to do with structural motifs that are linked to the network with only one extrem-
ity, such as loops or dangling chains, as these may dissipate external deformation
energy. All of the latter has to do with the network structure, and based on the
aforesaid, this structure directly translates into properties. Although the example
with a rubber ball may appear silly, the same holds for other rubbery products such
5.3 Complex moduli 173

Figure 69: Visualization of the contributions of the loss modulus, E″, and the storage modulus, E′,
to the complex modulus E*. In an experiment, a rubber ball is dropped, but does not bounce back
to its original height. The amount of energy dissipated to heat corresponds to the loss in height
after the bounce; this is the loss modulus, E″. The residual height of the bounce corresponds to the
energy stored elastically within the material; this is the storage modulus, E′. An ideal elastic ball
would regain its original position, whereas an ideal viscous ball would not bounce at all.

as shoe soles. Tailoring their properties to either good damping (as requested by
customers for leisure activities) or good bounce (as requested by customers for pro-
fessional marathon running) can be achieved on the very same basis.67

 For this structure–property relationship to actually apply, however, the polymer must be in an
entropy-elastic rubbery state, which means that (network-strand) chains must be capable of ac-
counting for chain stretching by local monomer-segmental bond rotation, that is, conversion of
local narrow cis or gauche into wide trans conformations. This is only possible at temperatures
above the glass-transition temperature, T > Tg. At temperatures below the glass-transition tempera-
ture, T < Tg, by contrast, a completely different material state is present, namely an energy-elastic
glassy state. Here the modulus is much higher, and the material is not elastic and ductile but hard
and brittle. A tragic example of a material application where this was not taken into account is the
Challenger disaster. On the morning of January 28, 1986, the Challenger space shuttle with seven
astronauts on board was to be launched from the Kennedy Space Center in Florida on a 6-day
space mission. On that morning, the temperature was low, and so rubber sealing rings in the fuel
system were glassy and therefore not tight. This led to fuel leakage, causing a catastrophe that took
all seven lives on board. We see from this tragic example: the ratio of the designated application
temperature T to the glass-transition temperature Tg has a decisive, here literally life-decisive influ-
ence on the state and thus the mechanical properties of a polymer material. Tg, in turn, is correlated
to how easy or difficult it is to rotate individual monomer bonds, which has to do with the local
chemical substitution pattern, meaning the bulkiness and interactions of the chemical side groups
on the polymer backbone. This is assessed by parameters such as the characteristic ratio, the per-
sistence length, or the Kuhn length, that we have got to know in our second chapter. We see from
all this, that the microscopic chemical and polymer topological structure is coupled to macroscopic
properties on many micro- and even nanoscopic levels, and that we have to take all this into ac-
count in applications (and not lose sight of the operation window such as the temperature-range
intended for an application). Expressed more positively, we also see that we actually can do this,
and we see how certain desired material properties at foreseen defined application conditions can
in fact be rationally translated into specific polymer chain structures.
174 5 Mechanics and rheology of polymer systems

5.4 Viscous flow

In addition to the elastic Young’s or shear modulus E* or G*, other quantities that
we have introduced in the previous paragraphs, such as the creep compliance, J,
and the viscosity, η, can also be determined in a dynamic experiment and therefore
be expressed as complex quantities.
The complex creep compliance, J*, which is the reciprocal of the complex
modulus, 1/E* or 1/G*, reflects the ratio of strain, ε or γ, to stress, σ. For the case of
oscillatory shear deformation, it calculates as follows:
γ γ0 1 1
J* = = expð − iδÞ = = (5:12a)
σ σ0 G* G′ + iG′′
′ ′′
We can expand this expression by G′ − iG′′ to gain
G − iG

G′ − iG′′ G′ − iG′′ G′ G′′


J* =   = 2 = 2 −i 2 = J ′ − iJ ′′ (5:12b)
′ ′′ ′
G + iG G − iG ′′ G′ + G′′
2
G′ + G′′
2
G′ + G′′
2

The complex viscosity, η*, is calculated analogously. It reflects the ratio of stress, σ,
to the strain rate, dε/dt or dγ/dt. For an oscillatory shear deformation experiment, it
calculates as follows:

σ σ0 expðiωtÞ σ0 expðiωtÞ σ0 1
η* =   = = = expðiδÞ (5:13a)

dt 0 expðiðωt − δÞÞÞ
d
ð γ iωγ0 ðexpðiðωt − δÞÞÞ γ0 iω
dt

σ0
Insertion of G* = γ0 expðiδÞ transforms the equation to

G* G′ + iG′′ G′ iG′′
η* = = = + (5:13b)
iω iω iω iω
By expanding the first summand with i=i and cancelling out the i in the second
summand, we can write

iG′ G′′ iG′ G′′ G′′ G′


+ = + = − i = η′ − iη′′ (5:13c)
iiω ω − 1ω ω ω ω
We can see that the complex viscosity, η*, is connected to the complex moduli, E*
and G*, by the frequency, ω. It should be noted here that η′ is connected to E″ and
G″, whereas η″ is connected to E′ and G′. This makes sense: the real part of the vis-
cosity accounts for the dissipation of energy by a material, just as the imaginary
part of E* and G* does.
This viscosity of a polymer solution is often quite different than that of a solu-
tion of small molecules that exhibits Newtonian flow according to eq. (5.2). The
latter’s viscosity is independent of the strain rate or shear rate dε/dt or dγ/dt. If we
5.4 Viscous flow 175

plot the stress, σ, as a function of that rate (such a plot is named flow curve), the
viscosity is the slope of a straight line in this case, as this plot is a simple graphical
representation of Newton’s law. In polymer solutions and melts, however, the rela-
tion of stress to the strain rate or shear rate is often nonlinear. Instead, polymer sys-
tems often exhibit shear-thinning, that is, a decrease of the viscosity at increasing
shear rate. The reason for this is because shear forces destroy possible associate
structures and orient the chains along the direction of flow, thereby enabling the
polymer solution or melt to flow more freely by reducing the friction and mutual
impairment between their chains. Due to this structural change, shear-thinning has
been termed structural viscosity by Wolfgang Ostwald junior in 1825.
Because their viscosity does not relate linearly to the shear rate, polymer sys-
tems are often discussed in terms of the zero shear viscosity, η0, which is obtained
by extrapolating the shear-thinning flow curve in Figure 70(A) to a shear rate of
zero. In this limit, the shear-thinning flow curve coincides with the ideal Newtonian
one. Shear-thinning properties, nevertheless, are actually wanted for many applica-
tions. For example, wall paint is specifically engineered such to be shear-thinning,
because in this way, it can be applied with ease onto a wall through the exertion of
shear by brushing, whereas once this is done, it will stay on the wall and dry in-
stead of drip. Shear-thinning polymers are also used in everyday-life products such

Figure 70: Plot of a flow curve, which is a graphical representation of stress, σ, as a function of
shear rate, dγ/dt, of a flowing liquid. (A) In case of a Newtonian fluid, we obtain a straight line
passing through the origin in which the slope is the viscosity, η, which is independent of the shear
rate. This is different for the case of non-Newtonian flow, which can be observed in different
manifestations. In the case of shear-thickening, the viscosity increases with the of shear rate,
whereas in the case of shear-thinning, the viscosity decreases with the shear rate. The latter is
often found in polymer systems, as shear breaks associates and orients polymer chains, thereby
enabling them slide against each other more easily. (B) Further deviations from Newtonian flow are
cases in which the flow curve exhibits an intercept different from zero. A Bingham fluid is solid at
low stress, but starts to flow beyond a specific stress threshold. A Casson fluid is a special case of
such a fluid that exhibits additional shear-thinning once flow sets in.
176 5 Mechanics and rheology of polymer systems

as shrink foils. These foils are produced by extrusion of a shear-thinning polymer


melt in which the chains orient themselves when squeezed through the extrusion
nozzle. After quick cooling and solidification, this chain orientation is frozen in the
foil material. When wrapping an object with this foil and heating it, the oriented
chains will relax to their original random conformation, and thus, the foil shrinks.
Resolidification then locks this state. This is especially useful for materials that re-
quire an airtight packaging such as consumables.
A special case is the Bingham fluid: This is a material that has a flow threshold at
a specific yield stress. It acts as a solid at zero and very low shear below the threshold,
but flows like a Newtonian fluid when the shear rate is increased above it. This me-
chanical behavior is often attributed to secondary interactions such as Van der Waals
forces or hydrogen bonds in the material, which build up a three-dimensional network
structure that gives the material solid-like properties. These assemblies, however, can
be broken by shear due to the low binding energies of the transient interactions in-
volved. Bingham fluids known from everyday life include tooth paste, mayonnaise, or
ketchup. Similar to that is a Casson fluid, which acts like a Bingham fluid but then
furthermore exhibits shear-thinning. Melted chocolate is an example of a Casson fluid.
The opposite type of shear-altering behavior is called shear-thickening, which
is a common feature of particulate suspensions, whereas it is less common for poly-
mer solutions and melts. Here, transient hydrodynamic clusters of particles can
form and break spontaneously in the material. At high shear rates, the shear oscilla-
tion period gets shorter than the transient lifetime of these clusters, such that they
become permanent on the timescale of the experiment, thereby imparting hin-
drance on flow. A prime example is a water–sand suspension, that means wet
sand. When carefully burying your feet into it on the beach, you can easily remove
them from the sand when imposing only low shear by moving your feet slowly.
However, if you impose high shear and try to abruptly pull your feet out of the wet
sand, they will be stuck as if they were cemented in concrete.
Two further variants of complex rheological behavior are thixotropy and rheo-
pexy. In a thixotropic case, the viscosity is observed to decrease over time during
the experiment. An illustrative explanation for such a behavior is a time-dependent
breakdown of a house-of-cards structure in the sample. In the opposite case, named
rheopexy, the viscosity increases over time. There is no illustrative picture for that.
Questions to Lesson Unit 11 177

Questions to Lesson Unit 11

(1) Which is not a typical rheological experiment?


a. Creep test
b. Relaxation experiment
c. Dynamic oscillatory experiment
d. Plateau test

(2) Which statement describes the stress in a relaxation experiment when a box-
profiled strain is applied to a viscoelastic solid?
a. The applied constant strain is followed by an instant response of the stress
that slowly decreases with time until it reaches a plateau value; it does not
return to zero.
b. The applied constant strain is followed by an instant response of the stress
that slowly decreases with time until the point where the applied strain is
gone and it immediately returns to zero.
c. The applied constant strain is followed by a delayed response of the stress
that slowly decreases with time until it reaches an equilibrium value.
d. The applied constant strain is followed by a delayed response with a sharp
rise of stress which drops sharply back to zero till the applied strain ends,
followed by another spike in the stress signal.

(3) Choose the answer with the correctly assigned experimental setups that re-
sulted in the following graphs:
178 5 Mechanics and rheology of polymer systems

a.
① Dynamic experiment with an ideal viscous liquid
② Dynamic experiment with a viscoelastic material.
③ Relaxation experiment with a viscoelastic solid.
④ Creep test with a viscoelastic solid.
b.
① Relaxation experiment with a viscoelastic solid.
② Creep test with a viscoelastic solid.
③ Dynamic experiment with an ideal viscous liquid.
④ Dynamic experiment with a viscoelastic material.
c.
① Creep test with a viscoelastic solid.
② Relaxation experiment with a viscoelastic solid.
③ Dynamic experiment with a viscoelastic material.
④ Dynamic experiment with an ideal viscous liquid.
d.
① Dynamic experiment with an ideal viscous liquid.
② Dynamic experiment with a viscoelastic material.
③ Creep test with a viscoelastic solid.
④ Relaxation experiment with a viscoelastic solid.

(4) What is the benefit of expressing the complex moduli as a sum of sine and
cosine?
a. They are easier to handle mathematically compared to the exponential
expression.
b. The signal of a dynamic experiment is sinusoidal.
c. The different summands capture the elastic or the viscous properties.
d. One can identify the phase angle δ more easily.

(5) What is the benefit of expressing the complex moduli as complex numbers?
a. They are easier to handle mathematically compared to the trigonometrical
expression.
b. The expression eiωt has the form of an exponential build up.
c. The ratio between the real and the imaginary parts is given by a simple sine.
d. One can identify the elastic or the viscous properties more easily.
Questions to Lesson Unit 11 179

(6) When plotted in a complex number plane, the ratio of E′′ to E′ is given by the
tangent of the phase angle δ. Which is the correct value assignment to the fol-
lowing three characteristics (left to right):
Viscoelastic solid; viscous and elastic contribute equally; viscoelastic liquid
a. tan δ > 1; tan δ < 1; tan δ = 1
b. tan δ < 1; tan δ = 1; tan δ > 1
c. tan δ = 1; tan δ < 1; tan δ > 1
d. tan δ < 1; tan δ > 1; tan δ = 1

(7) Which statement about the complex viscosity is true?


a. The complex viscosity is related to the imaginary part of the complex modu-
lus because it represents the viscous properties.
b. The complex viscosity is related to the reciprocal of the real part of the com-
plex modulus which represents their elastic properties.
c. The real part of the complex viscosity is related to the real part of the com-
plex modulus and is the same for the imaginary parts; in short: real to real,
imaginary to imaginary.
d. The real part of the complex viscosity is related to the imaginary part of the
complex modulus and vice versa since both quantities are quantifying con-
trary properties.

(8) What is not a good and frequent reason for researchers’ interests in polymers as
viscoelastic materials?
a. The timescale on which their rather elastic behavior changes into a rather
viscous behavior is in between ms and s, which are practically relevant
timescales.
b. The comparably long timescales of motion combined with the comparably
big size of the polymer building blocks (compared to low molar mass mole-
cules) make them easily accessible for experimental methods.
c. The viscoelastic properties result in a very robust material, that is quite re-
sistant to fluctuations in temperature.
d. The viscoelastic properties are very dynamic, which results in a broad
range of possible applications.
180 5 Mechanics and rheology of polymer systems

5.5 Methodology in rheology

LESSON 12: PRACTICE AND THEORY OF RHEOLOGY


So far, you got a notion about the viscoelastic nature of polymers, how it manifests itself in
relaxation and creep experiments, and how it is reflected in complex material parameters. The
following lesson adds information on how these parameters are actually probed in experi-
ments, and how the phenomenology of viscoelasticity can be treated in simplifying mechanical
models: the Maxwell model and the Kelvin–Voigt model. We will see how these models are
conceptually very simple but yet powerful in quantitatively capturing a polymer’s time- or fre-
quency-dependent mechanics.

5.5.1 Oscillatory shear rheology

The most common and experimentally easiest way to assess the viscoelasticity of a
sample is the method of oscillatory shear rheology. Here, a sample is placed on a
lower plate that is usually an even plane, and an upper plate is lowered onto it, as
shown in Figure 71. A motor then turns the upper part in an oscillatory manner,
which results in a torque that is exerted by the sample-medium’s resistance in be-
tween the two plates. From this torque, the values for the moduli are determined,
and if that kind of estimate is performed at different oscillation frequencies, they are
even obtained in a frequency-dependent fashion. Sometimes, a very shallow cone is
used instead of a plate as the upper geometry for the following reason: the turning
motion creates a shear-force gradient facing outward from the sample’s center. The
cone’s angled geometry eliminates this gradient and ensures a constant shear force
throughout the sample. This, however, only works at a very defined gap size, which is
the distance between the upper and lower geometry. Polymer solutions are therefore

Figure 71: Schematic representation of a setup for a shear rheology experiment. The sample is
applied onto the lower part of the measuring geometry, which is usually an even plane. Then, the
upper part of the geometry is lowered onto the sample and sheared against the lower plate, either
in a continuous or in an oscillating manner. This upper part is either also an even plate, thereby
enabling the use of a user-defined gap size (A), or it is a shallow cone, thereby eliminating
possible disruptive shear-force gradients throughout the sample, which, however, only works at a
predetermined gap size (B).
5.5 Methodology in rheology 181

often measured with a cone–plate geometry, whereas polymer gels and melts are
mostly measured using plate–plate geometries.

5.5.2 Microrheology

Another way to capture a sample’s viscoelastic properties is the method of microrheol-


ogy. Here, a population of nano- or micrometer-sized probes is immersed within the
sample, which then explores it, as sketched in Figure 72. By quantifying the relative
ease of these probes’ movement through the sample medium, the sample’s viscoelas-
tic characteristics can be quantified. As the probes are small, this information can be
obtained on a local microscopic scale. Furthermore, if methods of imaging microscopy
are used to measure the probe motion, the information on the sample’s viscoelastic
mechanics is even obtained in a spatially resolved manner. With that, potential het-
erogeneity of the sample composition or structure can be determined.

Figure 72: Schematic of a polymer sample loaded with nanoparticles that explore it by either
random-thermal or externally directed motion. In the method of microrheology, observation of this
probe motion is used to draw quantitative conclusions about the viscoelastic properties of the
surrounding sample medium.

There are two variants of microrheology. In an active microrheology experiment, the


probe particles are displaced by external forces. In a passive microrheology experi-
ment, by contrast, only the ever-present thermal energy moves the probes.

5.5.2.1 Active microrheology


A prime example of an active microrheological method is magnetic-bead rheome-
try. Here, the sample is loaded with spherical magnetic microparticles that can be
moved by application of an external magnetic field. The magnetic force, ~
fmag , is re-
lated to the product of the magnetic moment induced in the bead, M ~ ðtÞ, and the
external field gradient, ∂~BðtÞ=∂x, which can also be expressed as the magnetic
182 5 Mechanics and rheology of polymer systems

susceptibility of the bead, χ, multiplied by its volume, V, and the field strength
~
times its gradient, ~
BðtÞ ∂B∂xðtÞ:

~ ∂~
~ ~ ðtÞ ∂BðtÞ = χV~
fmag, x ðtÞ = M BðtÞ
BðtÞ
(5:14)
∂x ∂x
The resulting displacement of the bead can be monitored by microscopy and then
translated into the viscoelastic properties in a spatially resolved manner. In the fol-
lowing, we first detail on how this works for the elementary case of a purely viscous
medium. In such a medium, the application of a constant magnetic force, ~ fmag , ac-
celerates the probe particles, accompanied by emergence of a counteracting fric-
tional force, ~ffrict . This force is related to the particle’s velocity, ~
ν, and the viscosity
of the medium, η, by Stoke’s law. In a steady state, these two forces are at balance,
and the particles reach a constant velocity. At this state, knowledge of the external
magnetic force on the basis of eq. (5.14) along with knowledge of the probe particle
size, r, and measurement of the particle velocity in the sample, ~ ν, allows the viscos-
ity, η, to be calculated:

fmag =~
~ ffrict = 6πηr~
ν (5:15)

When the sample is viscoelastic, its properties are best captured in a dynamic experi-
ment, in which the magnetic field is sinusoidally modulated. The resulting time-
dependent displacement of the bead, x(t), is given by the following expression:

xðtÞ = x0 expðiðωt − ’ÞÞ (5:16)

As we have seen in Section 5.3, a complex exponential can be rewritten as a trigo-


nometric function using Euler’s formula. Analogous to the dynamic macrorheologi-
cal experiment, we can find the frequency-dependent storage and loss moduli, G′(ω)
and G″(ω), as follows:

f0
G′ðωÞ = cos ’ (5:17a)
6πrjx0 ωj
f0
G′′ðωÞ = sin ’ (5:17b)
6πrjx0 ωj

5.5.2.2 Passive microrheology


Passive microrheology experiments are based on the same principles as active
ones, but in contrast to those, the passive variant solely relies on thermal energy,
kBT, to drive the probe-particle motion through the sample. Based on that premise,
the probes’ mean-square displacement, Δx2 ðtÞ, is observed as a function of time to
characterize the sample’s viscoelastic properties.
5.5 Methodology in rheology 183

In a purely viscous medium, only Fickian diffusion of the probes takes place. In
this case, the probes’ mean-square displacement obeys the Einstein–Smulochowski
equation:

hΔx2 ðtÞi = 6Dt (5:18)

Together with the Stokes–Einstein equation for the diffusion of spherical particles
in a viscous medium, we can express the sample’s viscosity as follows:
  kB T kB Tt
η =^ G′′ðtÞ = = (5:19)
6πDr πrhΔx2 ðtÞi

In a purely elastic medium, the diffusion of the probes is constrained. They may move
a certain short distance during which they do not yet realize the constraint exerted by
the solid medium, but as soon as they feel the elastic trapping of their surrounding,
they cannot move further. This means that their time-dependent mean-square displace-
ment, hΔx2 ðtÞi, first exhibits free diffusion with Einstein–Smulochowski-type scaling,
but then eventually reaches a plateau value, hΔx2 ip (Figure 73). Balancing the driving
thermal energy to the dragging elastic energy, with the latter expressed in a Hooke-
type form with the material’s spring constant, κ, yields:
1  
kB T = κ Δx2 p (5:20)
2

Figure 73: Mean-square displacement of probe particles, hΔx 2 ðt Þi, in an elastic medium as a
function of time, t. At short times, the particles travel so short distances that they do not yet
realize the elastic trapping in their surrounding. At long times, by contrast, the trapping is fully
operative and prevents motion of the probes further than a distance reflected by the balance of
their thermal driving energy and the counteracting elastic energy of the medium.

The loss modulus, G″, can be neglected in a purely elastic environment; as a conse-
quence, the storage modulus, G′, can be determined as
κ kB T
) G′ðtÞ ⁓ ≈ (5:21)
r rhΔx2 ip
184 5 Mechanics and rheology of polymer systems

The intermediate case of a viscoelastic medium relates to the intermediate part of


the curve in Figure 73, where the probes partially but not yet fully experience elastic
constraint on their diffusive motion. In this range, the Einstein–Smulochowski
equation (5.18) is extended by an exponent α(t):

hΔx2 ðtÞi ⁓ tα (5:22)

In a purely viscous medium, α(t) = 1 and the probes can diffuse freely. In a purely
elastic medium, by contrast, α(t) = 0, and we find a plateau value for the probes’
mean-square displacement, as described above. When 0 < α(t) < 1, the probe dis-
placement is called subdiffusive. This is the case in a medium that has both viscous
and elastic properties. The curve in Figure 73 shows that α(t) is time dependent. At
short times, the potential elastic constraint of the surrounding medium is not yet
felt by the moving particles, such that α(t) ≈ 1. At longer times, the particles more
and more experience such constraint, such that α(t) < 1 (more and more as time pro-
ceeds), and eventually, in the long-time limit, the elastic constraint is fully felt and
prevents particle motion further than the plateau value of hΔx2 ip , such that α(t) = 0.
From α(t), the complex modulus, G*, with its two parts G′ and G″, can be deter-
mined based on a calculation by Mason and Weitz:
h πi
G′ðωÞ = GðωÞ cos αðωÞ (5:23a)
2
h πi
G′′ðωÞ = GðωÞ sin αðωÞ (5:23b)
2

with

k T
GðωÞ =  1 B (5:23c)
πrhΔx2 ω iΓ½1 + αðωÞ

5.6 Principles of viscoelasticity

5.6.1 Viscoelastic fluids: the Maxwell model

Now that we have understood that many materials, especially the soft-matter type, dis-
play mechanical properties with both elastic and viscous contributions, we have to
find a way to mathematically describe this viscoelastic mechanics. A suitable approach
is to start from simple mechanical models that combine both the stereotype mechani-
cal element of an elastic body, which is a spring, and the stereotype element of a vis-
cous fluid, which is a dashpot. The easiest way to combine these two elements is to
connect them in series, as shown in Figure 74. This is done in the Maxwell model.
When stress is applied to this model by an external deforming force, the elastic spring
5.6 Principles of viscoelasticity 185

Figure 74: A Maxwell element is composed of a


purely viscous damper (a dashpot) and a purely
elastic spring connected in series. When stress is
applied, the elastic spring deforms instantly,
followed by delayed irreversible deformation of
the viscous damper. Once the stress is released,
only the spring snaps back to its original position,
whereas the dashpot remains deformed.

element deforms instantly, followed by a delayed irreversible deformation of the vis-


cous damper element that relaxes the stress in the deformed spring as far as possible.
If that relaxation is not allowed to come to completion, there is residual stress in the
system, but once this stress is released externally, only the spring element will snap
back to its original position, whereas the dashpot element will remain deformed
irreversibly.
The applied stress is equal in both parts, σ = σ1 = σ2 , whereas the total strain is
the sum of the strain of the two separate parts, ε = ε1 + ε2 . The same holds true for its
time derivative

dε dε1 dε2
= + (5:24)
dt dt dt

The elastic contribution by the spring is given by Hooke’s law; in differential form it
reads

dσ dε1
=E (5:25)
dt dt

The viscous contribution from the dashpot is described by Newton’s law as

dε2
σ=η (5:26)
dt

We can now use the latter two equations and insert them in eq. (5.24) to generate

dε 1 dσ σ
= + (5:27)
dt E dt η

Solving this differential equation with the boundary condition of constant deforma-
tion, dε=dt = 0, yields

Et t
σðtÞ = σ0 exp − = σ0 exp − (5:28)
η τ

The latter equation quantifies the time-dependent stress relaxation of the Maxwell el-
ement; it contains a parameter that is essential for that: the relaxation time, τ = η=E.
This parameter connects two macroscopic quantities, the elastic modulus, E, and the
186 5 Mechanics and rheology of polymer systems

viscosity, η, by a microscopic time constant, τ.68 In a viscoelastic medium, τ is the


characteristic time of molecular rearrangement. On timescales longer than τ, the
molecules (or macromolecules in case of a polymer sample) can fully rearrange them-
selves, such that viscous flow dominates the viscoelastic properties. By contrast, on
timescales shorter than τ, the (macro)molecules cannot yet rearrange themselves,
such that elastic solidity dominates the viscoelastic properties. At the timescale of τ,
the one behavior transitions into the other, and we observe marked viscoelastic prop-
erties. For polymers, τ is in the range of microseconds to seconds. With that, it falls
right into the domain that many experiments and practical applications cover. As a
result, polymers display rich viscoelastic mechanics in practice. The exact value of τ,
and with that, the extent of viscous or elastic dominance on their mechanics, de-
pends on parameters such as the polymer size, shape, and interactions.
From the preceding mathematical consideration, we see that the Maxwell ele-
ment is good to model an experimental situation of constant strain, which leads to
time-dependent stress relaxation, as shown on the left-hand side of Figure 75; this is

Figure 75: Relaxation experiment (left) and creep test (right) applied to a Maxwell element. During
the relaxation experiment, the spring is first stretched, but then the damper follows and relieves
the spring, such that stress can relax over time. In the creep experiment, the spring first shows an
initial response that is then followed by the dashpot, which starts to exhibit normal Newtonian
flow, such that the deformation steadily proceeds. Note that the latter behavior is not creep in a
sense as we have defined it in Section 5.2.1.2, but instead, just flow. Hence, the Maxwell element
is good at modeling stress relaxation in a viscoelastic fluid, but not creep of a viscoelastic solid.

 At this point, the identity η = E τ was introduced as a pure mathematical tool to replace η/E in
eq. (5.28) by a time constant τ to turn the exponential decay function exp (–Et/η) into the mathemati-
cally general form exp (–t/τ). With that, we have obtained the identity η = E τ as a sort of “byprod-
uct”. It is, however, one of the most central equations in the field of rheology, with very fundamental
physical meaning. Later on, in Section 5.7.1, we will derive it again “properly”. Anyhow, we may al-
ready understand it conceptually here: viscosity (η) is composed (= a product) of the ability of a me-
dium to initially store energy upon deformation (E) times how long it takes to dissipate that (τ).
5.6 Principles of viscoelasticity 187

typical for a viscoelastic fluid. By contrast, if we consider the opposite case of a con-
stant stress, dσ=dt = 0, then eq. (5.27) just turns into Newton’s law. As a result, in
this case, the Maxwell element just responds with some initial strain of the spring
followed by normal Newtonian flow of the dashpot, as shown on the right-hand side
of Figure 75. This is not realistic; we know from Section 5.2.1.2 that a viscoelastic
solid exhibits time-dependent creep under the condition of constant stress instead.
As an alternative to the purely mathematical discussion above, we can also in-
terpret the mechanical responses of the Maxwell element shown in Figure 75 in a
conceptual way. If we apply constant strain to the Maxwell element, as shown in
the upper-left schematic in Figure 75, the resulting time-dependent stress initially
rises sharply, but then drops back to its original value over time, as sketched in the
lower left in Figure 75. This is because the spring is stretched instantly, but then the
dashpot follows and relieves the strained spring. In this way, stress can relax over
time, following an exponential function as in eq. (5.28):

t t
σðtÞ = σ0 exp − = ε0 E0 exp − (5:29a)
τ τ

From that, we can calculate a time-dependent elastic modulus, E, as follows:

t
EðtÞ = E0 exp − (5:29b)
τ

By contrast, if we apply constant stress to the Maxwell element, as sketched in the


upper right of Figure 75, the resulting time-dependent strain initially rises to some
extent, and then grows linearly as long as stress is applied, as shown in the lower
right of Figure 75. Once the stress ceases, the strain drops by the exact same amount
as it rose at the start, but retains the increase it has gained over time. Conceptually,
the initial response can be attributed to the instant deformation of the spring. The
dashpot follows and starts to flow, such that the deformation proceeds to rise. After
the experiment, the spring returns to its original unstrained form, thus creating the
drop in the strain signal. Note that this kind of behavior, however, is not creep in a
sense as we have defined it in Section 5.2.1.2, but instead, just flow. Hence, the Max-
well element is good at modeling stress relaxation in a viscoelastic fluid, but not
creep of a viscoelastic solid.
For the latter experimental situation, as a complement to the time-dependent
elastic modulus in the preceding case, a time-dependent creep compliance can be
calculated:

σ0 1 t
εðtÞ = ε0 + t = σ0 + (5:30a)
η0 E0 η0
t
J ðt Þ = J0 + (5:30b)
η0
188 5 Mechanics and rheology of polymer systems

In a dynamic experiment, we apply a sinusoidally modulated strain to the Maxwell


element. We may imagine easily how the response should be in the limit of high
and low frequencies. At high frequencies, the spring element can respond instantly
even to a quickly modulad sinusoidal load, whereas the dashpot cannot respond
tue to its inertia. As such, in that regime, the response of the Maxwell element is
predominatly elastic. By contrast, at low frequencies, the dashpot can fully follow
the motion of the then just slowly agitated spring to relieve it from the external
strain. As such, in that regime, the response of the Maxwell element is predomi-
nantly viscous. In addition to that qualitative line of though, we may also treat the
Maxwell element quantitatively. As we have learned from Section 5.3, the involved
complex strain, ε*, and the complex stress resulting from it, σ*, can be expressed as
exponential functions, of which we also need the derivatives in the following:

dε*
ε* = ε0 expðiωtÞ ) = iωε0 expðiωtÞ (5:31a)
dt
dσ*
σ* = σ0 expðiðωt + δÞÞ ) = iωσ0 expðiðωt + δÞÞ (5:31b)
dt

The starting point for mathematical modeling of the Maxwell element under the
condition of a dynamic experiment is again eq. (5.27), which we expand by a factor
of 1 = η=η on the right side, as this allows us to rearrange the equation as follows:

dε 1 η dσ σ 1 η dσ dε η dσ
= + = +σ , η = +σ (5:32)
dt E η dt η η E dt dt E dt

When we apply the fundamental equation η = E0 τ0 to the left-hand side of eq. (5.32)
along with applying a rearranged form of it, τ0 = η=E0 , to the right-hand side
of eq. (5.32), we get

dε dσ
E0 τ0 = τ0 +σ (5:33)
dt dt

Plugging in the exponential functions given in eq. (5.31a) and (5.31b) yields

E0 τ0 iωε0 expðiωtÞ = τ0 iωσ0 expðiðωt + δÞÞ + σ0 expðiðωt + δÞÞ

= σ0 expðiðωt + δÞÞðτ0 iω + 1Þ (5:34)

That can be rearranged as follows:

E0 τ0 iω σ0 expðiðωt + δÞÞ σ
= = = E* (5:35)
τ0 iω + 1 ε0 expðiωtÞ ε

Expansion of the left-hand side of eq. (5.35) with the conjugate complex number
allows us to write out the real and imaginary parts of the right-side, E* = E′ + iE′′:
5.6 Principles of viscoelasticity 189

E0 τ0 2 ω2
E′ = (5:36a)
τ0 2 ω2 + 1
E0 τ0 ω
E′′ = (5:36b)
τ 0 2 ω2 + 1

Analogous calculations can be made for the complex oscillatory shear modulus
G* = G′ + iG″.
Let us examine the latter two equations more closely by first focusing again on
two extreme frequency regimes, a high and a low one.
At low frequencies, the denominators of both eq. (5.36a) and (5.36b) have a value
close to 1, because the τ0 2 ω2 term in them becomes very small then, and therefore
negligible compared to the 1 in them. Then, the power-law form of the remaining nu-
merators of eq. (5.36a) and (5.36b) gives straight-line graphs when plotted in a dou-
ble-logarithmic representation, with a slope determined by the exponent of the
variable ω. Figure 76 shows such a log–log plot, in which we find a slope of 2 for E′,
and a slope of 1 for E″.

Figure 76: Frequency-dependent elastic modulus of a Maxwell element, E(ω), in a double-


logarithmic representation. The graphs can be separated into two frequency regimes, separated by
the crossover point of E′ and E″. At low ω, the material has enough time to relax, and its
viscoelastic properties are dominated by its viscous contributions, so E″ > E′. At high ω, the
material cannot fully relax anymore, and its viscoelastic properties are dominated by its elastic
contribution, so E″ < E′.

At high frequencies, by contrast, the τ0 2 ω2 term in the denominators of both eq. (5.36a)
and (5.36b) is much larger than the 1 in them, such that the latter can be neglected.
Many of the variables in the numerators and in the denominators cancel out then.
For E′, no dependence of ω remains due to that cancelling, such that its graph is a
190 5 Mechanics and rheology of polymer systems

flat line with a value of E0 at high frequencies. For E″, we obtain a frequency depen-
dence of ω − 1 due to that cancelling, such that its graph is a straight line with slope –1
in a double-logarithmic representation such as the one shown in Figure 76.
Both frequency regimes are separated by the terminal relaxation time, τ0, at a
frequency of ω0 = 1/τ0. At that frequency, according to eq. (5.36a) and (5.36b), both
E′ and E″ have values of ½E0, as also seen in Figure 76; hence, both the viscous
and elastic part contribute equally here.
Now let us do what we have done several times already in the course of this
book and use this mathematical insight to obtain conceptual insight into the time-
dependent mechanical properties of a viscoelastic Maxwell-type fluid. Again, we
discuss this in terms of a low- and a high-frequency regime.
At low frequencies, there is much time for relaxation of the system. A deformed
material then has enough time to rearrange its building blocks on a microscopic scale;
for example, if it is a polymer system, the chains have enough time to relax back to
their equilibrium conformations after deformation. Reconsidering the Maxwell element
from Figure 74, the time is long enough for the damper to relax the energy stored in
the spring. As a result, at such long timescales, the material’s mechanical properties
are dominated by its viscous contribution. This becomes more and more dominant the
longer the timescale is or, in turn, the lower the frequency is. We see that clearly in
Figure 76: the lower the frequency, the more does the E″-curve outweigh the E′-curve.
At high frequencies, the picture is to the contrary: here, there is not enough time for
relaxation of the system. Visually speaking, the damper does not have time to be ac-
tive. On these short timescales, the material therefore behaves as an elastic solid. This
becomes more and more dominant the shorter the timescale is or, in turn, the higher
the frequency is. Again, we see that clearly in Figure 76: the higher the frequency, the
more does the E′-curve outweigh the E″-curve. The transition from the low-frequency
viscous regime to the high-frequency elastic regime happens at the terminal relaxation
time, τ0. It denotes the time it takes for the building blocks of any material, which in
our case are polymer chains, to rearrange themselves over a distance equal to their
own size. On timescales longer than τ0, such effective displacement of the microscopic
building blocks against one another is possible, causing flow on the macroscale. On
timescales shorter than τ0, such displacement is not possible, and the material re-
sponds elastically. When the time is exactly τ0, then both the viscous and the elastic
parts of the sample’s mechanical spectrum contribute equally, and E′ and E″ have
identical values of E0/2. We can observe this as the crossover point of the graphs in
Figure 76.
Polymers and colloids, which are composed of large building blocks with high
molar masses, exhibit τ0 in the milliseconds to seconds range. This is advantageous,
because it makes these relaxation processes observable. The same basic processes
are also present in common low molar-mass materials, such as water, but in those
materials, τ0 is in the nanosecond range. These timescales are much harder or even
impossible to assess experimentally. Fortunately, however, the knowledge gained
5.6 Principles of viscoelasticity 191

from observation of the slower polymeric or colloidal systems can be adapted to the
hard-to-observe classical materials due to their common physical grounds. This insight
of the French physicist Pierre-Gilles de Gennes was awarded with the Nobel Prize in
Physics in 1991. In addition to leading to that scientific breakthrough, the millisecond-
to-second timescale of relaxation of polymeric and colloidal matter makes these rich
in mechanical behavior on practically relevant timescales, opening a plethora of me-
chanical application possibilities. It is for these that polymers are in fact most famous.

5.6.2 Viscoelastic solids: the Kelvin–Voigt model

The Maxwell model is not the only possible combination of a Hookian spring and a
Newtonian dashpot. When the two are connected in parallel, we generate a system
that is the basis of the Kelvin–Voigt model, as depicted in Figure 77.69 Upon appli-
cation of a constant stress, the Kelvin–Voigt element exhibits creep and deforms at
a decreasing rate, asymptotically approaching the steady-state strain σ0/E. When
the stress is released, the material gradually relaxes to its undeformed state. A
graphical representation of this experiment is shown in Figure 78. Since the defor-
mation is reversible (though not suddenly), the Kelvin–Voigt model describes an
elastic solid that also exhibits some viscous contribution.

Figure 77: Kelvin–Voight element composed of a purely viscous damper


and a purely elastic spring connected in series.

By simple geometrical need, the total strain of the Kelvin–Voigt element is the same
as the individual strain of each element: ε = ε1 = ε2 , whereas the total stress is the
sum of the individual stresses: σ = σ1 + σ2 .
We can express the elastic contribution to the Kelvin–Voigt body by Hooke’s
law as

 The model was originally proposed by Oskar Emil Meyer in 1874 in his essay “Zur Theorie der
inneren Reibung”, published in the Journal for Pure and Applied Mathematics. It was later indepen-
dently “rediscovered” by the British physicist William Thomson, who later became the first Baron
Kelvin and by the German physicist Woldemar Voigt in 1892.
192 5 Mechanics and rheology of polymer systems

Figure 78: Creep test of a Kelvin–Voigt element. The response of the spring is delayed by the
simultaneous but slow response of the damper. The final deformation will eventually approach that
for the pure elastic part, σ0/E. Once the stress is released, the entire Kelvin–Voigt element slowly
relaxes to its original position, which means that the deformation is reversible.

σ1 = Eε (5:37)

and the viscous contribution by Newton’s law as


σ2 = η (5:38)
dt

We can now plug the latter two equations into σ = σ1 + σ2 to generate

dε σ Eε
= − (5:39)
dt η η

Solving this differential equation with the boundary condition of a constant applied
stress of σ = σ0 yields

σ0 Et σ0 t
εðtÞ = 1 − exp − = 1 − exp − (5:40)
E η E τ

Here, τ is the retardation time: it quantifies the delayed response to an applied


stress and is a material-specific parameter.
The Kelvin–Voigt model is good at predicting creep, because in the infinite
σ
time limit the strain approaches a constant value of limt!∞ ε = E0 , whereas the Max-
well model predicts an infinite linear relationship between strain and time. Thus,
the Kelvin–Voigt model element is good to model a viscoelastic solid. By contrast,
as we have seen above, the Maxwell model is good to model a viscoelastic liquid.
5.6 Principles of viscoelasticity 193

5.6.3 More complex approaches

By addition of further Hookian spring and Newtonian damper elements, a plethora


of more complex models have been created to describe the viscoelastic properties of
many substances. These models include but are not limited to the standard linear
solid or Zehner model, the generalized Maxwell model, the Lethersich, the Jeffreys,
and the Burgers model. The latter one, which is depicted in Figure 79, combines a
Kelvin–Voigt with a Maxwell element by connecting them in series. It was devel-
oped by the Dutch physicist Johannes Martinus Burgers in 1935 to model the behav-
ior of bitumen. It is sometimes also applied to polymer-based materials.

Figure 79: The Burgers model of viscoelastic media by a combination of a Kelvin–Voigt and a
Maxwell element. It is tailored to accurately predict the viscoelastic properties of bitumen, but is
sometimes applied to polymers as well.

In this model, the strain is the sum of the individual strains of the mechanical ele-
ments, just like in the Maxwell model:

εð t Þ = ε 1 + ε 2 + ε 3 = ε 1 + ε 2 ð t Þ + ε 3 ð t Þ

The stress is equal in the three parts, as also in the Maxwell model:

σ = σ1 = σ2 = σ3 = σ0

According to the Kelvin–Voigt model, the total stress in the second part is the sum
of those in the elastic and the viscous mechanical elements:

σ0 = σ2 = σ2e + σ2v

Plugging both the latter equations into the first one gives the time-dependent strain
εðtÞ for the Burgers model:
194 5 Mechanics and rheology of polymer systems

σ0 σ0 E2 t σ0 t
εðtÞ = ε1 + ε2 ðtÞ + ε3 ðtÞ = + · 1 − exp − +
E1 E2 η2 η3

Questions to Lesson Unit 12

(1) What is a major disadvantage of classical oscillatory shear rheology?


a. It is only feasible for polymer gels or melts and does not work for low-vis-
cous polymer solutions.
b. The temperature cannot be kept at a constant level since the sample gets
warmed up with time due to dissipation.
c. The amount of sample needed is large compared to the amount one obtains
in a typical laboratory-scale synthesis.
d. After the measurement the sample cannot be used for further analytics, be-
cause the structure is damaged during the experiment.

(2) What is the relaxation time?


a. The time needed for the microscopic building blocks of a material to move
distances of their own size.
b. The time necessary for a building block to move distances of the size of a
polymer chain.
c. The time interval it takes for stress relaxation.
d. The time that a sample needs to re-equilibrate after a rheological experiment.
Questions to Lesson Unit 12 195

(3) Remember: The Maxwell model consists of a dashpot and a spring connected
in series. Choose the correct stress response σðtÞ to the given applied strain sig-
nal εðtÞ for a viscoelastic fluid.

a. ①
b. ②
c. ③
d. ④
196 5 Mechanics and rheology of polymer systems

(4) Remember: The Kelvin–Voigt model consists of a dashpot and a spring con-
nected in parallel. Choose the correct strain response εðtÞ to the given applied
stress signal σðtÞ for a viscoelastic solid.

a. ①
b. ②
c. ③
d. ④

(5) What is a major advantage of the small probe volume in microrheology?


a. The measurement is a lot faster.
b. The obtained data applies locally; even spatial resolution is possible.
c. The embedded probe particles can be easily removed. The sample is avail-
able for further measurements after that.
d. The experimental realization is easy.
Questions to Lesson Unit 12 197

(6) What is the meaning of the exponent α in the adapted Einstein–Smoluchowski-


like relation for the mean-square displacement of a particle hΔ x2 ðtÞi ⁓ tα ?
a. α captures the elastic properties only; the higher the exponent, the more
dominant is the elasticity of the medium.
b. α captures the viscous properties only; the higher the exponent, the more
dominant is the viscosity of the medium.
c. α captures the elastic properties as well as the viscous ones; for α ¼ 1, the
medium shows purely viscous characteristics, whereas for α ¼ 0, the me-
dium is purely elastic with a plateau value Δ xp2 .
d. α captures the elastic properties as well as the viscous ones; for α ¼ 1, the
medium is purely viscous, whereas for α ¼ 2, it is purely elastic with a pla-
teau value Δ x2p .

(7) What happens to a particle in a not purely viscous medium?


a. The particle is partly dragged, which leads to superdiffusion.
b. The particle is partly trapped, which leads to superdiffusion.
c. The particle is partly dragged, which leads to subdiffusion.
d. The particle is partly trapped, which leads to subdiffusion.

(8) A probe sample consists of a polymer with network structure and embedded
probe particles. Which statements about the time-dependent displacement of
the particles isare false?
a. At long timescales, the particles are constrained by the polymer network;
their diffusion is restrained.
b. At long timescales, the particles are constrained more and more up to a
value at which the elastic outnumbers the driving thermal energy, and the
diffusion is stopped.
c. At long timescales, the particles are constrained more and more until the
energies of kB T and of the elastic constraint are at balance and the plateau
value is reached.
d. At short timescales, the particles do not “feel” any constraint, and they can
diffuse freely.
198 5 Mechanics and rheology of polymer systems

5.7 Superposition principles

LESSON 13: SUPERPOSITION PRINCIPLES


Παντα ρηει – everything flows. This fundamental aphorism of rheology holds for polymers in
particular, because even if their momentary appearance may be that of an elastic solid, in
many cases, waiting long enough will let them flow. In this lesson, you will see which funda-
mental physical principle lies behind this, and how it allows us to superimpose data collected
in different experiments to generate mechanical spectra spanning multiple decades in time or
frequency.

Up to this point, we have looked upon some simplified models to describe the me-
chanical response of a viscoelastic material. But what do we observe in a real experi-
ment? Let us look at a typical amorphous polymer and consider its creep compliance,
J, as a function of time, t, for a range of temperatures, T, in Figure 80. (Note that J
and t are the inverse quantities of E and ω, respectively.) At high temperatures, J has
high values and increases proportionally with time t, which means that the material
exhibits purely viscous flow. We can understand that by making ourselves clear that
in this limit of viscous dominance, J is dominated by its viscous part J”, and since it
is the inverse of the elastic modulus, this part is connected to 1/E”, for which we
have seen from the quantitative treatment of the Maxwell body in Figure 76 that in

Figure 80: Creep compliance, J, of an amorphous polymer probed as a function of time, t, at various
temperatures. At high temperature, J is proportional to t, meaning that the material exhibits purely
viscous flow. At low temperature, by contrast, J is fully independent of t, which means that the
material shows a purely elastic response. At intermediate temperatures, J is first independent of t
and then becomes proportional to it.
5.7 Superposition principles 199

the limit for which it dominates the mechanics (which is in the limit of low frequen-
cies, corresponding to long times), it has a frequency-dependent scaling of E” ⁓ ω1,
and so its inverse has a frequency-dependent scaling of J” ⁓ t1. At low temperatures,
by contrast, J has low values and is independent of t, which means that the material
shows a purely elastic response. Again, we can understand that by making ourselves
clear that in the limit of elastic dominance, J is dominated by its elastic part J’, and,
since it is the inverse of the elastic modulus, this part is connected to 1/E’, for which
we have seen from the quantitative treatment of the Maxwell body in Figure 76 that
in the limit for which it dominates the mechanics (which is in the limit of high fre-
quencies, corresponding to short times), it is frequency- (and therefore also time-)in-
dependent and has high values (approaching E0), so as a result, its inverse, 1/E’ = J’,
is also frequency- and time-independent and then has low values. At intermediate
temperatures, J has intermediate values, and it is first independent of t but then be-
comes proportional to it. To sum up, we see that the polymer displays a rich mechan-
ical behavior depending on both time, t, and temperature, T.
In this section, we will see how the quantities t and T are related to each other
in a dataset such as the one shown in Figure 80, and, while doing so, we will also
learn some fundamental principles of the overlay, or superposition, of rheological
data.

5.7.1 The Boltzmann superposition principle

The Boltzmann superposition principle formulates that the current state of stress
or deformation of a viscoelastic material is the result of its history.70 In mathemati-
cal terms, this means that the total stress or deformation applied to a material is the
sum of the time-dependent evolution of every single stress or deformation incre-
ment that it has experienced. In other words, the system remembers any former
stresses or deformations and continues to relax or creep from them even when new
ones are applied on top, as depicted in Figure 81.
Let us consider an example to illustrate the Boltzmann Superposition principle.
The sum of all shear deformations applied to a material can be expressed as
X
γ= γi (5:41)
i

According to Hooke’s law, the resulting sum of stresses can be expressed by the
sum of the shear modulus G multiplied by the individual deformation γi:

 In a philosophical view, this is just like with us as individuals: our histories have made us, or
“shaped us”, into the persons we are today.
200 5 Mechanics and rheology of polymer systems

Figure 81: When multiple relaxation experiments (left) or creep tests (right) are performed in a
sequence, the system remembers the stresses or deformations of each single one and continues to
creep or relax from these even when new ones are applied on top.

X X
σ= σi = Gðt − ti Þγi (5:42)
i i

Here, t is the time of measurement, whereas ti is the time of application of the strain
increment γi. We can derive an equation for the case of continuous change by
integration:
ð ðt ðt
    dγ
σ ðt Þ = d σ = ′
G t − t dγ = G t − t′ dt′ (5:43a)
−∞ −∞ dt

To simplify this expression, we use the method of substitution and write t − t′ ≡ u to


generate
ð0 ð∞
σ ðt Þ = − _ =+
GðuÞγdu GðuÞγ_ du (5:43b)
−∞ 0

This can be solved if we have an expression of the time-dependent modulus at


hand. As a simple example, we take the relaxation function GðuÞ = G0 expð− uτÞ from
the Maxwell model and insert it:
ð∞  u
σðtÞ = G0 exp − γ_ du (5:44)
0 τ

In a continuous-shear scenario, the shear rate γ_ is constant. For a monodisperse


polymer sample, this yields
ð∞  u h  ui∞
_
σ ð t Þ = G0 γ exp − du = G0 γ_ − τ exp − _
= G0 γτ (5:45)
0 τ τ 0
5.7 Superposition principles 201

We can rearrange that equation to derive an expression for the stress divided by the
shear rate, which, according to Newton’s law, is the viscosity η:
σ
≡ η = G0 τ (5:46)
γ_

Equation (5.46) is a very fundamental relation in the field of rheology, because it


connects the macroscopic properties viscosity, η, and shear modulus, G, to the mi-
croscopic property of the relaxation time, τ, that contains information about the
structure of the system’s building blocks (here: polymer coils).71 With that, we have
a structure–property relation. This relation defines flow as a relaxation process, in
which the viscosity is composed of the relaxation modulus and the relaxation time.
We can understand that quite illustratively: viscosity (η) quantifies the ability of a
fluid medium to initially store energy upon shear deformation (G) times how long it
takes to dissipate that by subsequent relaxation (τ), that means, by positional
change of the fluid’s building blocks (molecules or particles). The higher both these
contributors are, the harder it is to stir or pump the fluid.
In the more complex case of a polydisperse relaxation time spectrum, the vis-
cosity is expressed as the sum of multiple relaxation contributions:
X u σ X
GðuÞ = G exp −
i i
) ≡η= Gτ
i i i
(5:47)
τi γ_

This is the case when either the system’s building blocks are polydisperse in size,
or if monodisperse building blocks exhibit multiple, hierarchical relaxations, such
as coil-internal Rouse modes.
In summary, we see that relaxation occurs on the timescale τ, or on a time spec-
trum of multiple overlaying τi, and that these determine the viscosity by the ele-
mentary formula η = G0 τ. In the next section, we will take a closer look at the
temperature dependence of that.

5.7.2 The thermal activation of relaxation processes

Flow requires the molecules or particles that constitute a material to change their
positions. This necessitates them to slip by each other, for which they have to over-
come an energetic activation; it also necessitates some free volume around them to
move into. Both these prerequisites are temperature dependent, which means that,
conceptually, they can be treated together as an effective activation energy barrier,

 Here at this point, we have derived the fundamental identity η = G τ “properly”, whereas be-
fore, in Section 5.6.1, it was obtained more as a “mathematical by-product”.
202 5 Mechanics and rheology of polymer systems

Ea,eff, that determines the temperature-dependent duration of positional changes τ,


in the form of an Arrhenius-type expression:72

Ea V* Ea;eff
τ ⁓ η ⁓ exp · exp ≈ exp (5:48)
kB T Vf kB T

In eq. (5.48), the first term features the activation energy, Ea, for each molecular or
particulate positional change past other molecules or particles, whereas in the second
term, V*=Vf reflects the probability to find enough free space in the vicinity during
such an attempt of positional change, assessed by the free volume in the system, Vf,
and the volume required for each molecular or particulate rearrangement, V*.73
The effective activation energy Ea,eff is more often overcome at high temperatures,
which means that high temperatures facilitate positional change of the molecules.
Thus, high temperatures lead to small (=short) relaxation times and therefore also
small (=low) viscosities, whereas low temperatures lead to high (=long) relaxation
times and therefore also high viscosities. The lowest possible temperature at which
movement is still possible at all is the glass-transition temperature, Tg. Below this spe-
cific temperature, thermal activation is not possible anymore, all movement is frozen,
and the relaxation time is infinite, τ ! ∞.

5.7.3 Time–temperature superposition

The fundamental connection of temperature to the rate of chain relaxation, which –


in turn – is connected to the viscosity according to eq. (5.46), allows us to shift and
superimpose viscoelastic data that were recorded at different temperatures so as all
superimpose on a common time axis; this is done by shifting these data along the
time axis by application of a shift factor, aT:

t
Gðt, T Þ = bT G , T0 (5:49)
aT

The left-hand side of eq. (5.49) features G(t,T), which is the shear modulus measured
at a time t and temperature T. The right-hand side of the equation tells us that the

72 This is analogous to the Arrhenius relation for the rate constant k of a chemical reaction. This rate
constant increases with temperature, because the activation energy is overcome more frequently at
higher temperature. The rate constant, in turn, is inversely related to the reaction half-life time, τ, in
the form of k ⁓ τ1, whereby the constant of proportionality depends on the order of the reaction. In
that inverse form, the exponential term on the right side of the equation has a negative argument,
which is indeed the common way ofnotation of the Arrhenius law in reaction kinetics.
 Both these contributions have first been introduced independent of one another according to
the free volume theory by Doolittle (J. Appl. Phys. 1951, 22(12), 1471–1475) and the theory of rate
processes by Eyring; later on, however, it was recognized that they must both be accounted to-
gether by Macedo and Litovitz (J. Chem. Phys. 1965, 42(1), 245–256).
5.7 Superposition principles 203

same modulus will also be measured at a different temperature, T0, if the timescale is
shifted by a factor aT. Often, a suitable standard reference temperature is taken for
T0, for example, room temperature or the polymer’s glass-transition temperature, Tg.
To also account for potential density differences of the material at the temperatures T

and T0 (due to thermal expansion), a further shift factor bT = ρT ρ0 T0 is often applied
as well, as also seen in eq. (5.49).
The shift factors are not just loose arbitrary quantities, but they can be concep-
tualized and understood to be part of further interrelations. One such relation is the
Williams–Landel–Ferry (WLF) equation. It was originally discovered empirically,
but it can also be derived from eq. (5.48) as follows: if two viscosities η1 and η2 are
measured at two temperatures T1 and T2, their ratio according to eq. (5.48) is

η1 V* V*
= exp − (5:50)
η2 Vf1 Vf2

if we assume that the temperatures T1 and T2 are high enough such that the influ-
ence of the temperature on the fundamental positional change process can be ne-
glected ðkEaT − k EaT ≈ 0Þ. Furthermore, the relationship between the free volumes at
B 1 B 2
T1 and T2 is as follows:

Vf2 = Vf 1 + ΔαVm ðT2 − T1 Þ (5:51)

Here, Δα is the difference between the material’s thermal expansion coefficients at


both temperatures, and Vm is the eigenvolume of the chain material.
Logarithmization of eq. (5.50) as well as rearrangement and insertion of eq.
(5.51) yields

η1 V* T2 − T1
ln = ·  (5:52)
η2 Vf1 Vf
ΔαVm + ðT2 − T1 Þ
2

If T1 is replaced by the glass-transition temperature of the material under consider-


ation, Tg, the equation for any second temperature T = T2 is
   
Vf, g · T − Tg
V*
ηg
ln = V    (5:53)
ηT f, T
+ T − Tg
ΔαVm

Based on temperature-dependent viscosity measurements, Williams, Landel, and Ferry


estimated the volume V* required for a microscopic positional change as V* ≈ 40 · Vf, g
and the free volume at the glass-transition temperature, Vf,g, as Vf, g ≈ 52 · ΔαVm . Fur-
thermore, in a polymer melt, V* can be approximately equated with the eigenvolume
of the chain material Vm. If we do that, then the stepwise change in the thermal expan-
sion coefficient Δα is determined to be Δα = 4.8·10–4 K–1.
204 5 Mechanics and rheology of polymer systems

By plugging in these values into eq. (5.53), we obtain the WLF equation that
relates the shift factor aT at temperature T to the glass-transition reference state:
   
τ ηðT Þ − c1 T − Tg − 17.44 T − Tg
log aT = log ffi log   =  =   (5:54)
τg η Tg c2 + T − Tg 51.6K + T − Tg

In the latter form, we have further transferred to decimal logarithms ðlog x = ln x · log eÞ.
The numbers c1 = 17.44 and c2 = 51.6 K in eq. (5.54) are universal for many dif-
ferent polymers if Tg is taken as the reference temperature.
The utility of the shift factors introduced in eq. (5.49), which we may either de-
termine empirically or calculate with the WLF eq. (5.54), is that they allow us to
create a master curve of many individual datasets of G(T, t) referenced to new varia-
bles G(T0, t/aT), as shown in Figure 82 (note that this plot has a frequency axis in-
stead of a time axis, which is just the inverse). Such a master curve may then
display the full rheological spectrum of a polymer, covering many orders of magni-
tude in G (from single pascals to gigapascals!) and t (or ω) (from nano- or millisec-
onds to decades!), which would be inaccessible by actual experimentation.

Figure 82: Construction of a rheological master curve according to eq. (5.49). Full lines denote the
storage modulus, G′, whereas dashed lines denote the loss modulus, G″. The reference temperature
in this example is the glass-transition temperature. By assembling the single curves obtained at
various temperatures over a frequency span of about three decades, as collected in the right part of
the figure, and shifting them along the frequency axis through application of suitable shift factors aT,
along with potential additional shifting along the modulus axis to account for density differences at
different temperatures through the use of shift factors bT, we can create a master curve spanning 12
(!) orders of magnitude on the time axis, as shown in the left part of the figure.

In such a full rheological spectrum, we can recognize several distinctly different re-
gimes, depending on the timescale and on the temperature of observation. Fig-
ures 83 and 84 display these different regimes in a schematic fashion. Note that
Figure 83 is flipped compared to both Figures 82 and 84, because it shows the mod-
uli E and G as a function of time t rather than frequency ω. However, this is just
Questions to Lesson Unit 13 205

Figure 83: Schematic representation of a polymer’s full mechanical spectrum. At low temperatures
or on very short times, the polymer is a glassy, energy elastic solid. When the temperature is
raised or the time extended, the polymer enters a leather-like regime in which it is a viscoelastic
solid described by the Kelvin–Voigt model. For samples that are composed of long chains, we then
observe an intermediate rubbery plateau regime. When the temperature is sufficiently high or the
time sufficiently long, the material will eventually flow like a viscoelastic liquid described by the
Maxwell model. The Pascal values on the ordinate are typical for polymers in the melt state.

transformation of the type of displaying, but the characteristic domains remain


identical in either representation.

Questions to Lesson Unit 13

(1) Which statement about the creep compliance J is true?


a. The creep compliance is the inverse property of the Young’s modulus,
meaning that if the storage modulus E′ is described by a cosine, the corre-
sponding part of the creep compliance J ′ has the form of a sine, and the
sames the case for the imaginary parts.
b. The creep compliance is the inverse property of the Young’s modulus,
meaning that the real and imaginary parts of J, J ′, and J ′′ are given by the
reciprocal value of the respective part of the elastic moduli E′ and E′′;
−1 −1
J ′ = E′ , J ′′ = E′′ .
c. The creep compliance is a contrary property to the Young’s modulus, mean-
ing that the real and imaginary parts of J, J ′, and J ′′ are given by the oppo-
site part of the elastic moduli E′′ and E′.
d. The creep compliance is a contrary property to the Young’s modulus, mean-
ing that the real and imaginary parts of J, J ′, and J ′′ are given by 1 minus
the respective part of the elastic moduli E′ and E′′; J ′ = 1 − E′, J ′′ = 1 − E′′.
206 5 Mechanics and rheology of polymer systems

(2) What is the meaning of the Boltzmann superposition principle?


a. Stress is additive up to the maximal value of all applied single stresses, the
“superposition threshold.”
b. Under sequential deformations of equal magnitude, stress always increases
linearly.
c. Stress is additive; upon application of deformation to a viscoelastic sample, the
resulting stress adds upon preexisting ones. The same holds for deformations.
d. Stress is amplified by the already given stresses in the material. The same in-
crement in the applied deformation leads to a comparably higher increment
in stress if the material is pre-stressed. The same goes for deformations.

(3) What statement about the relation η ¼ G0 τ is true?


a. It is a structure–property relation because the macroscopic concept of flow
is given by the viscosity, which is connected to the shear modulus that
stands for the stored energy upon deformation and the microscopic prop-
erty of the relaxation time of the building blocks.
b. It is a structure–property relation that is simply based on the mathematical
connection of macroscopic properties (η; G0 ) and microscopic ones (τ).
c. It is not a structure–property relation, since all properties are macroscopic.
There is no easy way to describe viscosity, the shear modulus, or the relaxa-
tion time on a microscopic scale.
d. It is not a structure–property relation, because none of the properties can
be derived through analyzing the sample’s structure.

(4) The relaxation time of a material that exhibits flow can be estimated with the
following expression:

Ea V Ea; eff
τ ⁓ η ⁓ exp · exp ≈ exp
kB T Vf kB T

Which answer describes the different contributions to the final rightmost ex-
pression correctly?
a. The relaxation time depends on the ratio of the already relaxed building
blocks to the total building blocks, which is reflected in a Boltzmann distri-
bution-like expression times a volumetric term.
b. The relaxation time is given by a ratio of still unrelaxed to relaxed states in
a Boltzmann distribution-like expression. The temperature dependence
comes from the thermal energy that is released during relaxation.
c. Relaxation requires an activation energy, which is represented by a temper-
ature-dependent Arrhenius-like term times a term that represents the ratio
of the unrelaxed volume to the total volume after relaxation.
Questions to Lesson Unit 13 207

d. Relaxation requires an activation energy as well as volume available to oc-


cupy. Both contributions can be compiled in a final Arrhenius-like tempera-
ture-dependent term.

(5) The expression for the relaxation time of a flow-exhibiting material has an Ar-
rhenius-like form with

Ea; eff
τ ⁓ exp
kB T

Hint: The Arrhenius equation from reaction kinetics has the following form:
Ea; eff
k ⁓ exp −
kB T

What statement describes the relation between both correctly?


a. The Arrhenius equation gives an expression for the temperature depen-
dence of the rate constant k of a chemical reaction. Since this rate constant
is the reciprocal of the reaction’s half-life time it has basically the same
form as the expression for the relaxation time.
b. The Arrhenius equation gives an expression for the temperature depen-
dence of the rate constant of a chemical reaction. Since this rate constant
decreases during a reaction, there needs to be a negative sign in the expo-
nent. Both expressions are not the same but very alike.
c. The Arrhenius equation gives an expression for the temperature depen-
dence of the rate constant k of a chemical reaction. Since both expressions
describe a time-based quantity with the unit of seconds, they can be consid-
ered as similar expressions.
d. There is no trivial relation between both expressions.

(6) How can we describe the fundamental principle on which the time–tempera-
ture superposition is based?
a. The same relaxation processes can happen at different relaxation times at
different temperatures.
b. Higher temperatures lead to a greater expansion, which means the relaxa-
tion is accelerated.
c. Different temperatures cause a change in the way of relaxation. The differ-
ent relaxation processes can be superimposed.
d. Through different frequencies, relaxation processes can be recorded that
need different activation energies, which results in a wide range of temper-
atures that can be measured.
208 5 Mechanics and rheology of polymer systems

(7) Why is time–temperature superposition essential for measuring mechanical


spectra in rheology?
a. There is only a small range of timescales at which measuring is practical.
Increasing the range would be at the expense of the measurement’s preci-
sion. By measuring at different temperatures and shifting the data one can
obtain data for a wider range of timescales.
b. Technically rheometers can only cover a time range of three decades. By
measuring at different temperatures and shifting the obtained data it is pos-
sible to cover a time range of up to 12 decades.
c. Probed materials are often very thermosensitive. The time–temperature su-
perposition allows measuring at different timescales at the same tempera-
ture to generate data for different temperatures by shifting.
d. Since the heating supplies of a rheometer are very limited in their tempera-
ture range, time–temperature superposition allows measuring at different
timescales and shifting the obtained data to get a wider temperature range.

(8) A common PhD thesis takes a time of 3 years. Which frequency regime would
you be able to cover in that time in a regular oscillatory rheological measure-
ment?
a. 10–1 rad s –1
b. 10–3 rad s –1
c. 10–5 rad s –1
d. 10–7 rad s –1
5.8 Viscoelastic states of a polymer system 209

5.8 Viscoelastic states of a polymer system

LESSON 14: MECHANICAL SPECTRA


The principle of time–temperature superposition that you got to know in the last lesson allows
master curves of the time- or frequency-dependent moduli of polymer samples to be composed.
Such rheological spectra feature multiple different domains, depending on in which state of
mobility or trapping the chains in the sample are. The following lesson will give an overview of
these states and summarize the rich variety of viscoelastic states of polymers.

5.8.1 Qualitative discussion of the mechanical spectra

Let us discuss the schematic of a polymer viscoelastic spectrum in Figure 83 from left
to right along the time axis (that can also be viewed as a temperature axis due to the
principle of time–temperature superposition), or alternatively, the variant schematic
in Figure 84 from the right to the left on the frequency axis (which is nothing else
than a flipped time axis on log scales). At very short timescales, corresponding to
very high frequencies and to low temperatures, the elastic modulus has high values in
the range of gigapascals, and its storage part, G′ or E′, exceeds its loss part, G″ or E″.
This means that the material is a tough elastic solid in this regime. A molecular-scale
interpretation is that on these very short timescales, or at that low temperatures, no
motion is possible at all in the sample, neither of the polymer chains as a whole nor
of just parts of them. This is because the temperature is either so low that any motion
is frozen, or because we consider timescales so short that these motions can just not
be achieved yet. As a result, we have a sample with amorphous structure and without
any dynamics: a glass. This glassy state features energy-elastic mechanics with a high
modulus. At a first characteristic time, τ0,74 the shortest possible relaxation time
(which is typically around 1 ns at room temperature for a polymer melt), at least single
monomeric units have enough time to relax in a sense that they can diffuse over dis-
tances equal to their own size. At times longer than that, more and more additional
relaxation modes are activated, in a sense that now also sequences of pairs, triples,
quadruples, and so on of monomers can be displaced over distances equal to their
own size. This sequential activation of more and more relaxation modes more and
more drops the modulus if we go from left to right on the time axis in Figure 83 or
from right to left on the frequency axis in Figure 84. The polymer is then in a markedly

 We hereby name the shortest possible relaxation time of a polymer τ0; this is commonly done
so in many textbooks, referring to the “elementary” relaxation time of the units in the chain, which
somehow appears to be well represented by the index zero. In the framework of relaxation modes
treated in Chapter 3.6.4, we had named that time τN instead (see Table 8 in that chapter). Note here
that τ0 in the context of the present nomenclature equals τN in the context of the nomenclature
related to the relaxation-mode concept.
210 5 Mechanics and rheology of polymer systems

viscoelastic leather-like regime, as we see in Figures 83 and 84, where both the stor-
age and loss moduli, G′ and G″, have very similar values and very similar time- or
frequency-dependent scaling. The mechanical properties in that state can be treated
analytically by the Rouse model (for modeling the polymer subchain relaxation
modes and the dropping of the time- or frequency-dependent modulus as more and
more of these modes are activated) or phenomenologically by the Kelvin–Voigt model
(to phenomenologically treat the viscoelastic solid-like appearance of the specimen).

Figure 84: Full rheological spectrum of a polymer melt at room temperature. At the shortest
possible relaxation time, τ0, which lays on the right extreme of the spectrum shown in the present
frequency-dependent form, only subchains of single monomeric units can relax; following upon
that, at frequencies left of that extreme, comes a viscoelastic regime in which sequential activation
of relaxation modes of pairs, triples, quadruples, and so on of monomer units leads to dropping of
the moduli. At the entanglement time, τe, chain segments realize mutual entanglement, leading to
an intermediate elastic plateau. From the Rouse time on, τR, there is sufficient time for the entire
chain to be displaced, such that it could actually fully and freely relax and flow already if it would
not be trapped by entanglements. Getting loose of that constraint is eventually possible on
timescales longer than the reptation time, τrep. From that time on, on the very left end of the
spectrum shown here, the material is a viscoelastic liquid phenomenologically described by the
Maxwell model, with G′ ⁓ ω2 and G″ ⁓ ω.

At the end of the viscoelastic leather-like regime that features just subchain motion,
intuitively, we would expect to observe a transition to a free-flowing regime in
which relaxation and displacement of the whole chains can occur. From the Rouse
(and also from the Zimm) model of polymer dynamics, we know that this should
happen on times longer than the Rouse (or the Zimm) time, τR (or τZ), which is de-
noted to be about 1 ms in Figure 84. Such a transition will indeed be observed in
the case of samples with short chains. In samples with long chains, by contrast,
such as those shown in Figure 82–84, before that free-flow regime, we observe
something else: an intermediate rubbery elastic plateau. This plateau starts at
a second characteristic time: the entanglement time τe, which is around 1 μs in Fig-
ure 84. At that time, segments of long chains realize that they are trapped by mu-
tual entanglement. This constraint hinders them from relaxing fully and freely,
5.8 Viscoelastic states of a polymer system 211

thereby leading to a regime in which deformation energy cannot be relaxed but is


stored, with a modulus in the range of about a megapascal. As in that regime, mo-
tion of the chains is already activated but topological entanglement still hinders
them to relax, deformation of the sample is accommodated by decoiling of the
chains; as a result, the elasticity in this regime is of entropic origin. When the time
is eventually long enough for the polymer chains to disentangle and move by a
mechanism named reptation on timescales longer than the reptation time τrep,
which is around 1 s in Figure 84, the polymer shows terminal viscous flow. In this
regime, the material is a viscoelastic liquid phenomenologically described by the
Maxwell model, with G′ scaling with ω2 and G″ scaling with ω. From then on, the
more time we give the material, the more will G″ exceed G′, meaning that the more
will the material mechanics be dominated by its viscous properties.

5.8.2 Quantitative discussion of the mechanical spectra

Now that we are able to visualize the full mechanical spectrum of a polymer and to
describe it phenomenologically, let us discuss these spectra quantitatively. For this
purpose, we first focus on unentangled polymer melts or solutions that both do not
display any intermediate rubbery plateau. We know from Section 5.6.1 that the me-
chanics of viscoelastic liquids is described by the Maxwell model, and we know
that the dynamics of polymer chains is quantified by the Rouse and the Zimm
model. We have also just learned that, in addition to the chain as a whole on time-
scales longer than the Rouse or Zimm time, smaller chain segments of various
lengths can relax on shorter timescales. These subchain relaxations are appraised
by so-called relaxation modes numerated by an index p. The pth mode corresponds
to the coherent motion of a subchain with N/p segments if N is the total number of
monomer segments in the whole chain. This means that at p = 1, the entire chain
relaxes and can get displaced by a distance equal to its own size, whereas at p = 2,
only each half of it relaxes and can get displaced by a distance equal to the half-
chain size, respectively. At p = 3, just only each third of the chain relaxes and can
get displaced by a distance equal to one-third of the whole chain size, and so forth.
At p = N, only single monomeric units can relax and get displaced against each
other by their own size. Figure 85 visualizes this hierarchy of relaxation modes. At a
time τp after abrupt deformation, all modes with index above p are relaxed already,
whereas all modes with index below p are still unrelaxed. Each of the p subchains
that belong to the mode with index p relax like own independent chains of length
N/p, with a relaxation time that can be appraised by the Rouse or Zimm formalism
if this were applied to a chain with not N segments, as usual, but only N/p seg-
ments; this relaxation time may also be plugged into a Maxwell-type formalism to
phenomenologically model the mechanical spectra of these subchains. Superposi-
tion of all these possible relaxation spectra yields
212 5 Mechanics and rheology of polymer systems

XN
τp 2 ω2
G′ðωÞ = νkB T (5:55a)
τ 2 ω2 + 1
p=1 p

X
N
τp ω
G′′ðωÞ = νkB T (5:55b)
p=1
τ p
2 ω2 + 1

Here, ν is the number-per-volume concentration of chains in the sample, whereas N


is the number of monomer units per chain, and with that, also the total number of
possible relaxation modes per chain.

Figure 85: Relaxation modes, indexed by a number p, of a schematic polymer chain. The first mode,
p = 1, relates to relaxation of the entire chain. In the second mode, p = 2, subchain segments with
a length of just half of the chain can relax. The third mode, p = 3, corresponds to the relaxation of
subchain segments with a length of only a third of the chain, and so on. In the last mode, p = N,
only single monomeric units can relax (not sketched here). Picture modified from H. G. Elias:
Makromoleküle, Bd. 2: Physikalische Strukturen und Eigenschaften (6. Ed.), Wiley VCH, 2001.

We can see from this expression that at frequencies ω < 1/τ1, all coil-internal modes
can relax. G′ then scales with ω2 and G″ scales with ω, because the denominators in
eq. (5.55) can be assumed to be one for such small ω-values, as the τp 2 ω2 parts in
them become negligible then. At frequencies ω > 1/τ1, by contrast, this changes:
now, each unrelaxed mode contributes an energy-storage increment of kBT to the
moduli. The moduli therefore increase from kBT per chain in the sample volume at
τ1 (=τRouse or τZimm) to kBT per monomer in the sample volume at τN (=τ0). Note: by
the normalization of these energies to the sample volume, which is done by the fac-
tor ν (number-per-volume concentration of chains in the sample) in eqs. (5.55), we
obtain a quantity with unit Nm/m3 = N/m2 = Pa. In that regime, the time depen-
dence of G′ and G″ (or inversely, the frequency dependence of these moduli) is
therefore given by the time- (or frequency-) dependence of the mode index. This
can be obtained from the Rouse model for polymer melts or the Zimm model for
polymer solutions, as summarized in Table 8 in Section 3.6.4. According to eq.
(5.55a) and (5.55b), we can calculate the frequency-dependent storage and loss
moduli for the individual modes p, Gp′ and Gp″, and visualize how they contribute
to the overall storage and loss moduli, G′ and G″. This is shown in Figure 86 for an
ideal polymer with N Kuhn monomer units. We see that at times longer than τ1 (=τRouse
or τZimm), there is enough time for the entire chain to relax by thermal motion, such that
the energy-storage capability is just kBT per chain. At a shorter time, τ2, subsegments
5.8 Viscoelastic states of a polymer system 213

with a length of up to only half of the entire chain can relax, whereas the mode τ1 is
still unrelaxed. The energy-storage contribution to the modulus is now kBT for
each unrelaxed mode. This trend continues: the shorter the time, the more modes
are still unrelaxed, with each unrelaxed mode contributing one kBT energy-storage
increment to the modulus. The overlay of still-unrelaxed and already-relaxed modes
leads to a scaling of G′ and G″ with ω½. At times shorter than τN (=τ0), not even
single-monomeric segmental relaxation is possible anymore, and the polymer
chain motion is practically frozen. All the unrelaxed modes now each contribute
an energy-storage increment of kBT to the moduli, such that G′ has achieved its
maximum value of G∞ = νN kBT.

Figure 86: Frequency-dependent storage (left) and loss (right) moduli for the individual modes p,
Gp′(ω) and Gp″(ω), as well as the entire chain, G′(ω) and G″(ω), of a polymer with N Kuhn segments.
Redrawn from C. Wrana: Polymerphysik, Springer, 2014.

Indeed, such mechanical spectra can be observed for unentangled polymer melts
and solutions. The first type of sample (i.e., polymer melts) is better described by the
Rouse model, whereas the second (i.e., polymer solutions) is better described by the
Zimm model. Figure 87 shows schematics of the mechanical spectra of these types of
samples. Again, at times below t = τRouse and t = τZimm, the polymer’s viscoelastic
properties are those of a viscoelastic liquid, with scaling according to G′ ⁓ ω2 and
G′′ ⁓ ω. Here, the chains have enough time to fully relax. At intermediate times, τRouse
or τZimm < t < τN, only relaxation of chain segments is possible. The material is a visco-
elastic solid and both moduli, G′ and G″, have similar values and both scale with ω
according to power laws with exponents corresponding to the inverse time dependence
(i.e., the frequency dependence) of the mode index from Table 8 in Section 3.6.4, giving
values in the range of 0.5–0.6 if ν is set to be 0.5 (Θ-state) to 0.6 (good-solvent state).
At times shorter than τN (=τ0), no relaxation is possible anymore, and the material is a
glassy solid.
214 5 Mechanics and rheology of polymer systems

Figure 87: Schematic viscoelastic spectra of unentangled polymers. Note that the names τ0 and τR
correspond to the names τN and τ1 in the concept of relaxation modes in the Rouse and Zimm
model, as discussed in Chapter 3.6.4 (see Table 8 there).

Questions to Lesson Unit 14

(1) Which answer assigns the right regimes in a full rheological spectrum to their
following characteristics?
① G′ and G′′ have similar values and courses.
② G′ drops rapidly and G′′ outnumbers G′.
③ G′ and G′′ have both high values in the range of GPa.
④ G′ shows an intermediate plateau value.
a. ① → Glassy, ② → viscous, ③ → leather-like, ④ → rubbery
b. ① → Leather-like, ② → viscous, ③ → rubbery, ④ → glassy
c. ① → Leather-like, ② → viscous, ③ → glassy, ④ → rubbery
d. ① → Viscous, ② → rubbery, ③ → leather-like, ④ → glassy

(2) What is the microscopic explanation for the glassy state?


a. At very short timescales, no motion of the building blocks is possible in the
sample, especially when it has a crystal structure.
b. At very short timescales, no motion of the building blocks is possible in the
sample, even if it has an amorphous structure.
c. At very short timescales, only single building blocks can move distances of
their own size, so the system flows like a glass.
d. At very short timescales, the sample is fully transparent due to the arranged
building blocks, looking like glass.
Questions to Lesson Unit 14 215

(3) Which is the shortest possible relaxation time?


a. τ0 , because at this time single monomeric units can move distances equal
to their own size.
b. τRouse , because at this time Rouse subsegments can move distances equal
to their own size.
c. τZimm , because at this time Zimm subsegments can move distances equal to
their own size.
d. τrep , because at this time the disentanglement of entangled chains is possible.

(4) What properties does a material in the leather-like regime display?


a. The material is viscoelastic since G′ and G′′ show very similar scaling as
well as almost identical overall values of the modulus.
b. The material is viscoelastic since G′ and G′′ have the same values despite
different scaling with frequency.
c. The material is a viscoelastic solid since the values of G′ are sufficiently
higher than those of G′′.
d. The material is a viscoelastic liquid since the values of G′′ are sufficiently
higher than those of G′.

(5) Which statement about the relaxation modes is true?


a. At τp the relaxation modes with index < p are already relaxed, whereas the
ones with index > p are still unrelaxed.
b. The zeroth relaxation mode (p ¼ 0) means the relaxation of the whole chain.
c. For a given chain, there is any number of relaxation modes with index p
possible.
d. The relaxation of the chain can be understood as sequential relaxation of
the chain’s subsegments. Their sizes increase with longer timescales.

(6) Why do G′ and G′′ scale with ω1=2 for intermediate timescales?
a. The overlay of the still-unrelaxed modes and the relaxed modes leads to a
scaling of ω1=2 .
b. Considering the Zimm model and a θ-state, the scaling equals 1 + 2 1ð0:5Þ = 0:5.
c. Since the time scaling is given by t2 ; the frequency scaling needs to be the
reciprocal value.
d. At intermediate timescales, the movement of the building blocks is subdif-
fusive. Therefore, the scaling is < 1.
216 5 Mechanics and rheology of polymer systems

(7) Why do the moduli drop for intermediate timescales in the leather-like regime?
a. The longer the time, the more chain entanglements can be disentangled,
which leads to a drop of kB T per chain.
b. The longer the time, the more modes can relax, and each mode leads to a
drop of kB T in the modulus.
c. The longer the time, the more chains can show, viscous flow leads to a sig-
nificant drop, especially in the storage modulus.
d. The longer the time, the more the chain dynamics can be described by the
Maxwell model, which scales with a steeper slope in ω than the Kelvin–
Voigt model.

(8) Which statement about the terminal viscous flow is true?


a. Terminal viscous flow is always possible when the time exceeds the Rouse
time, respectively, the Zimm time τZimm .
b. The reptation time marks the onset of terminal viscous flow, visible by an
increase of the loss modulus with time behind that point.
c. The mechanics of terminal viscoelastic flow can be described by the Max-
well model, scaling with ω2 for the storage modulus and with ω1=2 for the
loss modulus.
d. When the loss modulus exceeds the storage modulus, terminal viscous flow
sets in. Behind that point, the dominance of the loss modulus further in-
creases with time.
5.9 Rubber elasticity 217

5.9 Rubber elasticity

LESSON 15: RUBBER ELASTICITY


In Lesson 3, you have got to know polymers as entropic springs. This principle causes poly-
mer samples with entangled or crosslinked chains to show a characteristic plateau in the
time- or frequency-dependent modulus. In this lesson, you will get to know how the height
of this plateau can be conceptually and mathematically understood, and how this allows you
to fundamentally connect the mesh structure of a polymer network (be it a permanent cross-
linked or just a transiently entangled one) to its elastic modulus.

In the last lesson, we have examined the viscoelastic properties of short polymer
chains. They transition directly from the viscoelastic leather-like regime at τ < τRouse
to the viscous terminal-flow regime at t > τRouse, as shown in Figure 87. The picture,
however, is different if the polymer chains are long enough such that they can form
entanglements with one another and thereby mutually impair each other’s relaxa-
tion. As we will see later, this is possible once the chains are longer than a certain
minimal length, which is connected to a minimal degree of polymerization, Ne, or an-
alog, a minimal molar mass, Me, from which on chain entanglement can first occur.
If chains are longer than Nel (and therewith heavier than Me), mutual entanglement

Figure 88: Mechanical spectra of chains with molar masses below (blue), at (green), and above
(black) the entanglement molar mass, Me. The latter shows a distinct rubbery plateau between the
1
viscoelastic (G′ ⁓ ω2 Þ and the viscous (G′ ⁓ ω2 ) regimes. If an even higher molar mass were shown
in addition, it would display an even more extended plateau. At the high-frequency end of the
spectra, all data coincide, as here, subchain relaxation modes and transition to the glassy regime
is covered, which is the same for a given type of polymer at given experimental conditions,
independent of the total chain length of that material.
218 5 Mechanics and rheology of polymer systems

impairs their relaxation, and only on long timescales, these entangled chains can es-
cape this mutual constraint. On short timescales, by contrast, the entanglements act
like crosslinking points between the chains, thereby allowing for elastic energy storage
that manifests itself in the appearance of an intermediate plateau regime in the visco-
elastic spectrum, the so-called rubbery plateau depicted in Figure 88. The longer the
chains are, the longer does it take them to disentangle from one another, such that the
extent of the rubbery plateau regime on the time or frequency axis is greater the higher
the molar mass is. To quantitatively assess the impact of the entanglements on the me-
chanical properties of these long polymer chains, we regard them as permanent cross-
links and spend a focus on crosslinked polymer networks in this section.

5.9.1 Chemical thermodynamics of rubber elasticity

In chemical thermodynamics, free energy, F, is given as the relation between inter-


nal energy, U, and temperature, T, multiplied by entropy, S:

F = U − TS (5:56)

The force of deformation of a material specimen f is the length derivative of this


function:

∂F ∂U ∂S
f= = −T (5:57)
∂l ∂l ∂l

The entropy derivative in eq. (5.57) can be rewritten based on the identity S = − ð∂F =∂T Þ
and the symmetry of second derivatives (Schwarz’ theorem) for total differentials:

∂S ∂ ∂F ∂ ∂F ∂f
=− =− =− (5:58)
∂l ∂l ∂T ∂T ∂l ∂T

Reinserting this into eq. (5.57) generates

∂U ∂f
f= +T (5:59)
∂l ∂T

This expression can be plotted as a linear equation with an intercept of ∂U =∂l and a
slope of ∂f =∂T . Such a plot is shown for a typical rubber in Figure 89, in which the
force has been normalized to the area, thereby yielding the more general quantity
stress; the illustrated data can be imagined to be from an experiment in which the
force to achieve a certain deformation is measured at various temperatures.
At temperatures below Tg, where all chain dynamics is frozen, the material is
energy elastic. In that state, deformation pulls the motionless monomer segments
apart from each other against their attractive interaction potentials. This is easier to
achieve at higher temperatures, leading to less stress in the material at higher
5.9 Rubber elasticity 219

Figure 89: Relation of stress to temperature for a typical rubbery material. Below the glass-transition
temperature, Tg, the material is energy elastic, making it harder to deform the lower the temperature
is. By contrast, above Tg, the material is entropy elastic, making it harder to deform the higher the
temperature is. Picture inspired by B. Tieke: Makromolekulare Chemie, Wiley VCH, 1997.

temperature upon deformation, which causes a negative slope in the left branch of
the dataset in Figure 89. By contrast, above Tg, where chain dynamics is activated,
the material is entropy elastic. In that state, deformation leads to uncoiling of the
coiled chains by transition of local cis into local trans segmental conformations.
This chain uncoiling reduces the chain entropy, thereby creating an entropy-based
backdriving force. Due to the fundamental coupling of temperature and entropy,
that kind of deformation gets harder at higher temperature, leading to a positive
slope in the right branch of the dataset in Figure 89. Conversely, the material heats
up upon stretching in that state. We can see this from the following expression:

∂T ∂T ∂S ∂T ∂H ∂S 1 ∂S
= = =− T (5:60)
∂l ∂S ∂l ∂H ∂S ∂l cl ∂l

When the entropy term ∂S=∂l is negative, which it is upon deformation, the tempera-
ture term ∂T =∂l is positive, meaning that the temperature increases upon deformation.
We also see in Figure 89 that the extrapolated intercept of the positive-slope
part of the dataset is small; this underlines once more that the energetic contribu-
tion to stress in a deformed rubbery sample is small.
220 5 Mechanics and rheology of polymer systems

5.9.2 Statistical thermodynamics of rubber elasticity

In addition to the chemical-thermodynamic argument above, a statistical-thermody-


namic approach can give us particularly valuable quantitative information about the
rubbery elastic state. In such an approach, we generally appraise the most central
quantity in thermodynamics, entropy, as the degree to which a specific macrostate of
interest, in our case the shape of the polymer coil, is realized by different possible mi-
crostates, in our case the local segmental cis or trans conformations. We have learned
in Chapter 2 that the ideal shape of a single polymer chain is that of a Gaussian random
coil, with an entropy given by random-walk statistics:

3kB r2 3kB r2
S = S0 − = S0 − (5:61)
2hr2 i0 2Nl2

In eq. (5.61), hr2 i0 is the equilibrium mean-square end-to-end distance of the poly-
mer chain, whereas r is the end-to-end distance in a given situation of interest.
When we stretch the chain, we extend this latter distance, such that r2 > hr2 i0 . Be-
cause the variable r2 is in the numerator, whereas the constant hr2 i0 is in the de-
nominator of a term with negative sign in front of it, we reduce the entropy S upon
increase of r, that is, upon stretching. Note that the latter equation was actually de-
rived for single uncrosslinked polymer chains; we now presume that it also applies
to each network chain in a crosslinked rubber.
Figure 90 displays the relevant parameters that are needed to treat the deforma-
tion of a polymer network sample, both on the macro- and the microscale. Due to
the self-similar nature of polymers, we can express the entropy change ΔS of a sin-
gle ideal network chain, which is a subchain ranging from one crosslink to another
in the network, upon deformation as follows:75
       
~ ~ 3kB rx 2 + ry 2 + rz 2 3kB rx, 0 2 + ry, 0 2 + rz, 0 2
ΔS = S R − S R0 = − + =
2hr2 i0 2hr2 i0

3kB  2  2  2  2  2  2 
= λx − 1 rx, 0 + λy − 1 ry, 0 + λz − 1 rz, 0 (5:62)
2hr2 i0

The entropy change of the entire network is obtained simply by summation over all
its n network chains.

3kB  2  Xn 2  2  X n  2  2  X n
ΔS = − λx − 1 ð rx, 0 Þ + λy − 1 r y, 0 + λz − 1 ðrz, 0 Þ2i
2hr2 i0 i=1 i i=1 i
i=1
(5:63)

 Note again that in this context, deformation means the decoiling of the polymer chain, but not
stretching of the bonds between its monomeric units.
5.9 Rubber elasticity 221

Figure 90: When a polymer-network specimen is deformed, its new dimensions Lj can be described
as the original ones Lj,0 multiplied by strain ratios λj (with j denoting the geometrical dimension x,
y, and z). In an affine scenario, we presume the microscopic deformation of the chains that
constitute the sample to be equal to the macroscopic deformation of the entire body. As a result,
the strain ratios λj are identical on both microscopic and macroscopic scales. Picture redrawn from
M. Rubinstein, R. H. Colby: Polymer Physics, Oxford University Press, 2003.

In an isotropic sample, the average end-to-end distances are equal in all directions:

1 Xn 2 hr2 iu
i=1
ð rx, 0 Þ i = hr x, 0
2
i = hr y, 0
2
i = hrz, 0
2
i = (5:64)
n 3
Here, hr2 iu is the mean-square end-to-end distance of the network chains in the un-
deformed state.
We can simply rearrange this to
Xn Xn  2 Xn nhr2 iu
ðr Þ2 =
i = 1 x, 0 i i=1
ry, 0 i
= ðr Þ2 =
i = 1 z, 0 i
(5:65)
3
in which n is the number of network chains.
We can insert this expression into eq. (5.63) to generate

3kB  2  n 2   n 2   n 2
ΔS = − λx − 1 hr iu + λy 2 − 1 hr iu + λz 2 − 1 hr iu =
2hr2 i0 3 3 3

nkB hr2 iu  2 
=− λx + λy 2 + λz 2 − 3 (5:66)
2 hr2 i0

From that equation, it becomes clear that the entropy reduction upon stretching actu-
ally stems from two contributions. First, the entropy is reduced more if we stretch to
a greater extent, which is expressed in the form of higher strain ratios λj in eq. (5.66).
222 5 Mechanics and rheology of polymer systems

Second, the number of chains n that are stretched by that extent proportions the en-
tropy reduction. Remember again in this context that we treat network chains, mean-
ing chain segments ranging from one crosslinking junction to another. Consequently,
a high crosslinking degree also amplifies the change of entropy, because it corresponds
to a high number of (short) network chains. The proportionality factor hr2 iu =hr2 i0 in
eq. (5.66) describes the ratio of the mean-square end-to-end distance of the network
chains in the undeformed state to the mean-square end-to-end distance of an analog
uncrosslinked ideal chain. This factor depends on the state of the network chains dur-
ing the network formation. Often, the network is made in a state where the chains that
are crosslinked (or formed in-situ during crosslinking) are in the same state (e.g., an
ideal or θ-state) as they are during the network deformation experiment; in this case,
the factor hr2 iu =hr2 i0 in eq. (5.66) is one. It may deviate from one when the states of
network preparation and network probing are different, for example, when the net-
work was formed in a swollen state but is measured in a deswollen state, or vice versa.
For the rest of our argumentation, we assume the usual case of hr2 iu = hr2 i0 .
Now that we have derived an expression for the change of entropy, ΔS, we can
insert it into the fundamental equation for the free energy, eq. (5.56), and obtain

nkB T  2 
ΔF = ΔU − TΔS = + λx + λy 2 + λz 2 − 3 (5:67)
2

The internal energy, U, is independent of the network-chain end-to-end distance, as


we presume these chains to be ideal, such that ΔU = 0. Any change of the network
chains’ free energy, ΔF, therefore stems from entropy alone.
Let us consider a typical experimental situation: isochoric uniaxial strain. For
that situation, the latter expression can be simplified further. Isochoric means that
the total volume change upon deformation is zero, ΔV = 0, which means that the
product of the strain ratios has to be one, λx λy λz = 1. Uniaxial means that the strain
is only applied along one spatial dimension, x, leading to one defined strain ratio
λx = λ, whereas the two other strain ratios must follow according to the isochoric
pffiffiffi
condition: λy = λz = 1= λ.
Both considerations lead to a simplified expression:

nkB T 2 2
ΔF = λ + −3 (5:68)
2 λ

The deformative force fx can again be calculated by the length derivate along the
axis of deformation:

∂ΔF ∂ΔF nkB T 1


fx = = = λ− 2 (5:69)
∂Lx ∂ðλLx, 0 Þ Lx, 0 λ
By normalizing it to the plane of deformation, we can formulate an expression for
the stress σxx
5.9 Rubber elasticity 223

fx nkB T 1 nkB T 1
σxx = = λ− 2 = λ− 2 (5:70a)
Ly Lz Lx, 0 Ly Lz λ L λ L λ L
x, 0 y y, 0 z z, 0 λ
pffiffiffi
Due to the isochoric condition, λy = λz = 1= λ, we can eliminate the direction-specific
strain ratios and simplify the expression to

nkB T 1 nkB T 2 1
σxx = λ λ− 2 = λ − (5:70b)
Lx, 0 Ly, 0 Lz, 0 λ V λ

In the very beginning of our discussion on rheology, in eq. (5.1), we have defined
the extensional strain as ε = ΔL/L0, that is, the change of the material specimen
length, ΔL, relative to its length before deformation, L0; for uniaxial strain, this is
equal to ε = λ – 1, so λ2 − λ1 ≈ ε2 + 2ε + 1 − ð1 − εÞ = ε2 + 3ε,76 which for small ε is nearly
3ε. With that, we get

nkB T
σxx = 3ε (5:70c)
V

This equation is strikingly similar to Hooke’s law, eq. (5.1), that connects stress to
nk T
strain by a proportionality factor. In eq. (5.70c), this factor is 3 VB ; it therefore cor-
responds to the elastic modulus:

nkB T
E=3 (5:71a)
V

From that, along with eq. (5.3), we also get the shear modulus:

E nkB T
G= = (5:71b)
3 V

The latter formula is quite an extraordinary expression, as it connects the micro-


scopic structural information of the network chain concentration, n/V, to the macro-
scopic property of the shear modulus, G, thereby creating a quantitative structure–
property relation. We now realize that upon deformation, each network chain
stores an energy increment of kBT. The modulus increases with temperature due to
the entropy elastic nature of rubber elasticity. We also realize that a higher degree of
crosslinking causes a higher modulus, as the network chains are then more numer-
ous (and shorter), leading to a higher n in eq. (5.71a).
We can formulate the latter identity on a molar scale as well:
ρ RT
G= (5:71c)
Mx

3 3
76 The calculation for this is as follows: λ2 − λ1 = ðε + 1Þ2 − ε +1 1 = ðεε++11Þ − ε +1 1 = ðε +ε1+Þ 1 − 1 = ε +ε3ε+ 1+ 3ε.
3 2

Polynomial long division turns that into ε2 + 2ε + 1 − ε +1 1. Taylor series approximation of ε +1 1 ≈ 1 − ε (as
first described by Isaac Newton) then gives ε2 + 2ε + 1 − ð1 − εÞ.
224 5 Mechanics and rheology of polymer systems

This equation relates the network density or network concentration, ρ, in the unit
g·L–1, to the molar mass of the network chains, Mx, in the unit g·mol–1. From this, we
again realize that a higher degree of crosslinking causes a higher modulus, as the
network chains are then shorter (and more numerous), leading to a lower Mx in the
denominator of eq. (5.71b).

Equation (5.71b) is strikingly similar to the ideal gas law in elementary physical chemistry. This
is because rubber elasticity at all is very similar to the concept of pressure for an ideal gas.
Both are based on the reduction of freedom of arrangement for the system’s building blocks
(gas particles in space, or monomer units in a polymer chain) upon application of deformation
such as compression of the ideal gas, which limits the freedom of spatial arrangement of the
gas particles, or stretching of the polymer chain, which limits the freedom of arrangement of
the monomers along the chain from a mix of cis and trans local conformation to more trans. We
can actually use the same statistical thermodynamic approach that we have used above to ap-
praise the pressure of an ideal gas: we appraise the probability, Ω, for a gas molecule to sit in a
subvolume V of a container with total volume V0 to be Ω = V/V0. From this follows that the prob-
ability for n gas molecules all to sit in that subvolume at a time is Ωn = (V/V0)n. We estimate the
system’s entropy, S, by inserting this into the Boltzmann formula: S = kB lnΩn = kBn ln(V/V0).
From physical chemistry lectures, we know that the pressure, p, in thermodynamics is calcu-
lated as p = –(∂G/∂V)T = –(∂H/∂V)T + T(∂S/∂V)T. In our case, we are looking at an ideal gas,
whose particles do not interact. This means that there is no energetic change upon change of
volume, that is, (∂H/∂V) = 0. This simplifies the latter expression to p = T(∂S/∂V)T = (kBn·T)/V.
This is exactly equal to eq. (5.71a).77

An alternative calculation can be made for an isotropic shear experiment. In this


case, there is no deformation in the z-direction; hence λz = 1. The two other strain ra-
tios are then defined to be λx = λ and λy = 1=λ. For the free energy, this yields

nkB T  2  nkB T  2 nkB T 2


ΔF = λ + λ−2 − 2 = λ − λ−1 = γ (5:72)
2 2 2

Subsequent calculation of the stress gives

nkB T
σ= γ (5:73)
V

Again, we end up with an expression that has the form of Hooke’s law. In this case,
we identify the proportionality factor nkB T =V to be the shear modulus G:

nkB T
G= (5:74a)
V

 Note from this identity that pressure and the elastic modulus also have the same unit: Pascal.
5.9 Rubber elasticity 225

Once again, we can formulate the latter identity on a molar scale as well:

ρ RT
G= (5:74b)
Mx

So far, we have discussed rubber elasticity on the basis of permanent crosslinks. An-
other form of hindering polymer chains from relaxing, at least temporarily, is mutual
entanglement, as shown in Figure 91(A). Hence, on intermediate timescales, such
entanglements act like permanent crosslinks and create a rubbery plateau in the
time- or frequency-dependent elastic modulus, as shown in Figure 91(B).

Figure 91: (A) Schematic representation of a polymer network composed of permanent crosslinks
and chain entanglements. Both hinder the chains from relaxing and therefore contribute to the
ability of elastic energy storage, thereby both contributing to the plateau modulus in the form of
Gp ffi Gx + Gent . For permanently crosslinked networks, the plateau in the mechanical spectra,
shown in (B), is infinite and reaches out to t ! ∞, whereas if no permanent crosslinks are present,
the polymer chains can untie their mutual entanglement at long times. The extent of the rubbery
plateau is defined by the polymer molar mass, which must exceed a specific minimum molar mass,
Me, to show mutual entanglement at all.

A crosslinked network in fact usually has both modes of polymer connection: cross-
linking and (trapped) entanglements, as shown in Figure 91(A). We can therefore
split the modulus G into two contributions, one from actual permanent crosslinks,
Gx, and one from (trapped) entanglements, Gent:

1 1
G ffi Gx + Gent = ρRT + (5:75)
Mx Ment

This means that even uncrosslinked polymer systems, in which Gx = 0, are elastic
on short timescales solely due their mutual entanglements, whereby Gent typically
is in the range of a megapascal:
226 5 Mechanics and rheology of polymer systems

ρRT
Gent = ≈ 106 Pa (5:76)
Ment

There is just one prerequisite for this: to entangle with one another, chains must
have a certain minimum length, which translates to a certain minimum molar mass,
named the entanglement molar mass, Me. This quantity is related to the one found
in the denominator of the last equations, named Ment there. Typically, Me ≈ 2 Ment,
because Me denotes the minimum molar mass (and therewith also the minimum de-
gree of polymerization Ne, and yet in turn, the minimum contour length Nel) that
chains must have to form entanglements, whereas Ment denotes the length of a strand
segment between two entanglement points in a well-entangled network. A simple
geometrical thought leads us to the notion that Me must be twice as large as Ment,
because to form entanglements at all, the chain must be minimally twice as long as
one segment between two entanglement nodes in an entangled network.
The reason for the existence of a minimum molar mass for entanglement is be-
cause the chains have a certain stiffness or persistence, which must be overcome by
sufficient chain length to form entanglements. There is an analogy to everyday life.
In large canteens, for example, in the Mensa, when spaghetti are cooked and
served, they are short. This is because the kitchen personnel wants the noodles to
be shorter than their entanglement length, as entangled noodles are much harder
to process than unentangled ones. The stiffer the main chain is (which, on the
nanoscale, translates to the ease or non-ease of monomer-bond rotatability), the
longer must it be to be able to form entanglements. For polystyrene, an entangle-
ment molar mass of Me ≈ 20 kg mol–1 is practically relevant.
As a closing note, it shall be mentioned that all the above appraisal is based on
the entropy of an ideal chain, which we have obtained in Section 2.6.1 by applying
Boltzmann’s formula to the distribution of end-to-end distances that we have de-
rived by random-walk statistics before. Hence, all the above applies to chains only
with these statistics. This is valid only in the limit of low deformation, whereas at
large deformation, we actually disturb the random-walk-type distribution of end-to-
end distances considerably, such that it is no longer valid. In this limit, other statis-
tics have to be used, for example, the Langevin model.78 In addition, in polymers
with high main-chain regularity and tacticity, crystallization is possible. Stretching
of the network chains may then bring them in so close distance and good order that
crystallization may occur, leading to crystalline nodes that act as further crosslinks.
This phenomenon is named strain-induced hardening.79

 Again, there is an analogy to ideal gases here: at large pressure, they deviate from ideality and
must be treated by a different model: the van der Waals equation.
 And once more, there is an analogy to gases, which may liquefy at high pressure.
5.9 Rubber elasticity 227

5.9.3 Swelling of rubber networks

In addition to stretching or shearing, there is another way to deform a polymer net-


work: swelling. In the process of swelling, a fluid medium enters the network and
expands it from the interior. The basis of this process is that upon contact of solvent
molecules with a dry network, its chains want to dissolve. Their mutual connectiv-
ity, however, hinders them to do so freely. Hence, rather than separating the chains
from one another completely, as it would be the case in an uncrosslinked sample,
solvent can only penetrate into the network and stretch the polymer network chains
in it. The degree of stretching, however, is limited by the entropy-elastic backdriv-
ing force that comes along with that, as we have just quantified above. At equilib-
rium, the polymer–solvent mixing energy is at balance with the entropy-elastic
energy. The point where this equilibrium lies, and with that, the swelling capacity
of the network, depends on the ratio of the network chain length (or in other words,
the density of crosslinking points in the network) and the thermodynamic quality of
the solvent (athermal, good, or Θ).
To quantify the swelling equilibrium, we introduce the degree of swelling q as

V Vuptaken solvent
q= =1+ (5:77)
V0 Vunswollen network

To estimate this value, we need to know both contributors: the free energy of mix-
ing and the entropy-elastic backdriving energy. This is best gauged via the chemical
potentials related to these energies.
The contribution of the free energy of mixing is given by the Flory–Huggins
formula:
   
Δμmix = RT ln 1 − q − 1 + q − 1 + χq − 2 (5:78)

The potential of the entropy-elastic backdriving force is calculated by

∂ΔFelast ∂ΔFelast ∂q
Δμelast = = · (5:79)
∂n ∂q ∂n

The elastic energy is given by the rubber elasticity formula:


3  
ΔFelast = NkB T q2=3 − 1 (5:80)
2

Here, N denotes the total number of network chains.


Assuming that both the volumes of the dry network and the solvent are addi-
tive, we can formulate the degree of swelling as follows:
228 5 Mechanics and rheology of polymer systems

 ∂q  solvent
 solvent + V0 ) q = V = nsolvent V solvent + 1 )
V = nsolvent V =
V
V0 V0 ∂n V0

∂ΔFelast ∂q 3 2  solvent
V
) Δμelast = · = NkB T · q − 1=3 ·  solvent (5:81)
= vRTq − 1=3 · V
∂q ∂n 2 3 V0

Here, v denotes the concentration of network chains, which is inversely related to


the network chain length. This is because in a network with a given amount of poly-
mer material in it, if the network chains are long, then there is just few of them and
vice versa. Hence, v can be considered to be both the molar number of network
chains n per volume unit V [mol·L– 1], or the network density ρ [g·L–1] divided by
the network chain molar mass Mx [g·mol–1]:
n ρ
v= = (5:82)
V Mx

At the swelling equilibrium, both the free energy of polymer–solvent mixing as well
as the entropy-elastic backdriving energy are identical:

Δμmix = Δμelast (5:83)

This allows us to calculate the molar concentration of network chains as

ρ lnð1 − q − 1 Þ + q − 1 + χq − 2
v= =−  solvent q − 1=3 (5:84)
Mx V

We can now understand the value of doing a swelling experiment. It allows us to


determine the degree of crosslinking at a known Flory–Huggins parameter χ, or,
vice versa, the determination of the Flory–Huggins parameter at a known degree of
crosslinking.
The swelling ability and often soft appearance of polymer gels give rise to vari-
ous areas of application. Particularly relevant in this context are hydrogels, which
are polymer networks swollen with up to 99% (w/w) of water. This is, in principle,
the same ratio of water and solid material as in biological matter, which is why hy-
drogels are often used as matrixes in the field of life science for procedures such as
electrophoresis or chromatography. Dried hydrogels can also be found in hygiene
products as super absorbers due to their superior ability to absorb moisture – such
as urin in diapers.
In more sophisticated applications, the polymer network can be designed to
react to external stimuli that trigger either swelling or shrinking. Such a stimulus-
sensitive reversible swelling is sketched in Figure 92. Stimuli-responsive gels have
potential uses as artificial muscles, chemo-mechanical actuators, or as “intelligent”
capsules for drug release. As an example for the latter, consider the following: can-
cer tissue has a marginally lower pH than the surrounding healthy tissue. In a
Questions to Lesson Unit 15 229

Figure 92: Schematical representation of a stimulus-sensitive polymer network that can undergo
swelling or deswelling upon external triggers. This stimuli-responsiveness renders such materials
useful for potential applications such as chemo-mechanical switches, artificial muscles, or “smart”
drug-release carriers.

specifically tailored polymer network, the pH change would trigger a deswelling or


a swelling process, thus initiating controlled release of an cytostatic drug (either by
squeezing it out of the gel in the case of deswelling or by letting it diffuse out of the
gel upon opening diffusive paths through the widened network meshes in the case
of swelling). The same principle might be used for the controlled release of anti-
inflammatory agents. Inflamed tissue has a higher local temperature than its envi-
ronment; hence, in that case, the temperature difference could provide a stimulus
for the hydrogel to shrink or swell.

Questions to Lesson Unit 15

(1) Which relation to the chain length holds true for the elastic rubbery plateau?
a. The elastic rubbery plateau has higher values of the storage modulus the
longer the chains are.
b. The elastic rubbery plateau expands over a larger frequency/time/tempera-
ture range the longer the chains are.
c. The elastic rubbery plateau is shifted toward shorter timescales/higher fre-
quencies the longer the chains are.
d. The elastic rubbery plateau shows a more significant drop in the loss modu-
lus the longer the chains are.
230 5 Mechanics and rheology of polymer systems

(2) Chain entanglements in the elastic rubbery regime can be treated as _____
a. defect structures in the polymer chain, analog to dangling ends.
b. cross-links between the polymer chains.
c. sequential relaxation modes.
d. interactions between the polymer chains in the magnitude of van der Waals
interactions.

(3) What does the glass-transition temperature for a rubbery material not denote?
a. The change of a crystalline to an amorphous structure.
b. The change of energy-based to entropy-based elasticity
c. The change of easier to harder deformation with increasing temperature.
d. The change of a negative slope of the stress-to-temperature relation to a
positive one.

(4) The affine network model ______


a. describes the shrinking of a rubbery material that has been stretched to a
size that is smaller than its initial one.
b. is only applicable for cross-linked polymer networks.
c. is based upon the assumption that when deformed, the strands between
two cross-link points of the network are stretched to the same extend as the
network in total.
d. states that entanglements in the chains attract each other, so that the
chains get even more entangled.

(5) What holds when comparing the shear modulus G for the case of applying iso-
choric uniaxial strain with the one when an isotropic shear deformation is
applied?
a. They have completely different form, because they are based on different
conditions.
b. They both are connected to the number of chains n, temperature, and vol-
ume but have a different scaling.
c. They are proportional to the number of chains n, temperature, and volume.
Just the values of the proportionality factors are different due to the differ-
ent geometrics in shear vs. uniaxial deformation.
d. They both are the same.
Questions to Lesson Unit 15 231

(6) Why do chains need to have a minimum molar mass to show a rubbery elastic
plateau of the modulus?
a. The chains need a minimum molar mass to be long enough to form van der
Waals interactions, which get stronger with increasing molar mass.
b. The chains need a certain length to overcome their own stiffness such as to
be flexible enough to form entanglements.
c. Only chains that are sufficiently long can form ties to others.
d. The height of the plateau depends on the molar mass. If the molar mass is
too low, the plateau is not visible in the mechanical spectrum.

(7) What does not affect the swelling equilibrium of a polymer network?
a. Its cross-linking density
b. Its molar mass
c. The pH value of the solution
d. The quality of the solvent

(8) We can also use the statistical approach that we used to access the concept of
rubber elasticity, to derive the ideal gas law. Which are the basic thoughts that
lead to the final formulation of the law?
a. The Boltzmann formula and the thermodynamic expression for pressure.
b. The Boltzmann formula and the thermodynamic expression for entropy.
c. The thermodynamic expression for entropy and the statistical formula of
pressure.
d. The thermodynamic expression for entropy and the statistical formula of oc-
cupation probabilities.
232 5 Mechanics and rheology of polymer systems

5.10 Terminal flow and reptation

LESSON 16: REPTATION


When a noncrosslinked polymer sample is given enough time, it will display viscous flow,
even if the chains are long and entangled. On a microscopic scale, such flow requires the
chains to change their positions against each other. Describing this microscopic motion is
an essence to describe macroscopic material properties such as the polymer fluid viscosity.
In this lesson, you will get to know a simple but powerful approach to do so: the reptation
model. This model simplifies the direct environment of a chain in an entangled surrounding
to be a rigid tube, out of which the chain can creep only by curvilinear motion.

So far, our knowledge on the viscoelastic spectra of polymers has two sides: a
phenomenological macroscopic one and a conceptual microscopic one. On the one
side, phenomenologically, we can describe the macroscopic viscoelastic solid-like
mechanics of the short-time or high-frequency regime with the Kelvin–Voigt model,
which gives us equations to approximate the leather-like creep properties of a sam-
ple in that domain. As a complement to that, we can use the Maxwell model to de-
scribe the viscoelastic liquid-like stress relaxation of the sample in the long-time or
low-frequency domain. On the other side, conceptually and microscopically, we
can understand how the viscoelastic solid-like appearance of the sample in the first
domain originates from viscoelastic relaxation modes, that is, motion of only parts
of the chains but not yet the chains as a whole, which we can quantify based on the
Rouse or the Zimm model. We also understand how chain entanglement may intro-
duce a further domain into the mechanical spectra, namely, that of a rubbery pla-
teau. What we are lacking, though, is a conceptual microscopic understanding of
the relaxation and flow in the terminal regime. Delivering that is the intent of the
following section. We have already seen that the transition between the rubbery
plateau and the terminal flow regime depends on something called the entangle-
ment molar mass, Me (see Figure 91(B)), without further discussing this quantity;
we will do so in this subchapter.
Describing the long-term relaxation of an entangled multichain system is a
challenging proposition. Many chains move simultaneously through the sample
volume, part in concert with each other, part independently. They may also interact
with each other or, if present, with a solvent. This is a very complex multibody phe-
nomenon that can hardly be described analytically. The solution to this challenge
is as easy as it is clever: we focus only on one test chain, and then simplify its envi-
ronment drastically.
5.10 Terminal flow and reptation 233

5.10.1 The tube concept

Consider a test chain in an ensemble of overlapping and entangled chains that we


have somehow modified so we can identify it from the background and be able to
follow its path of motion, as sketched in Figure 93(A). To describe its surroundings,
we neglect every other chain that does not have contact with our test chain. Even the
ones that have contact only do so with small portions of their own length, such that
we must only regard these sections of them, as illustrated in Figure 93(B). These con-
tacts form an array of obstacles in the direct environment of our test chain. Now, a
smart approach by Samuel Frederick Edwards comes up. Edwards modeled the direct
constraining environment of the test chain by a solid tube, as shown in Figure 93(C).
The test chain can move only along this tube in a curvilinear manner, whereas per-
pendicular movement to the tube is prohibited. With this kind of motion, the chain
can creep out of its confining tube, as illustrated in Figure 93(D).

Figure 93: Tube concept to simplify the complex multibody surrounding of a polymer chain in a
system of many other mutually entangled chains. (A) We consider a test chain, here labeled in blue
color, which is somehow distinguishable from its background into which it is embedded and with
which it is entangled. (B) To simplify our view, we only consider the direct environment of the test
chain. (C) The directly surrounding matrix chain segments around the test chain can be modeled as
a confining solid tube that constrains the motion of the test chain to be possible only along the
tube contour. (D) Over time, the test chain can creep out of this confining tube by curvilinear
motion, which is named reptation. Note: as the surrounding matrix is still there, the test chain then
finds itself entrapped in a new tube (not shown here). The overall diffusion of the test chain
through the surrounding matrix can be viewed as a sequence of steps from one tube to another,
each of which occurring by curvilinear creep along the tube contour. Pictures inspired by
W. W. Graessley: Entangled linear, branched and network polymer systems – Molecular theories,
Adv. Polym. Sci. 1982, 47 (Synthesis and Degradation Rheology and Extrusion), 67–117.
234 5 Mechanics and rheology of polymer systems

This approach is called the tube concept. Mathematically, Edwards assumed


that each neighboring segment imparts a parabolic confining potential on the test
chain. The energetically most favorable way for the test chain to arrange itself in
the tube and also to move along it is through the path of the potential minima
through space, which is called the primitive path. The test chain fluctuates around
this primitive path, but only up to an extent that the ever-present thermal energy
kBT can activate, because anything further would require additional energy input.
The tube diameter, a, is therefore given by the transverse distance at which the
confining potential is exactly kBT. In other words, segments of the test chain fluctu-
ate with an average displacement of a.

5.10.2 Rouse relaxation and reptation

In the tube concept, the motion of a chain is largely restricted only to curvilinear
creep along the tube contour, whereas migration in directions perpendicular to that
is impossible. Pierre-Gilles de Gennes used this premise to mathematically describe
the test chain motion along the tube, which, as it resembles the creep of a reptile
through the woods, is referred to as reptation. De Gennes based his approach upon
the Rouse model for polymer dynamics in the melt, in which the diffusion along the
tube would be described by a one-dimensional Rouse-type diffusion coefficient:

kB T
DTube = (5:85)
Nfseg

The diffusion of a polymer chain is related to its mean-square displacement by the


Einstein–Smoluchowski equation that we have encountered numerous times in Sec-
tion 3.6: hx2 i = 2dDt, with d the geometrical dimension of the motion. We can now
adopt this for the reptating polymer: the tube has a contour length equal to the
total length of the polymer test chain, L = Nl, which will leave the tube after a spe-
cific time τrep, the reptation time:

hLTube 2 i N 2 l2
τrep = = ⁓ N3 (5:86)
2 · DTube 2kB T=Nfseg

We see that the degree of polymerization, and with that, the molar mass of the test
chain, has a paramount influence on the reptation time, as it scales with a power-
law exponent of 3(!).
Furthermore, we have learned about the fundamental connection between the
shear modulus, G, and the viscosity, η, by the relaxation time, τ: η = G0τ (eq. 5.46).
To estimate the viscosity, we only need to plug in the relevant values:
5.10 Terminal flow and reptation 235

In nonentangled melts, the relaxation is delimited by the Rouse time, τRouse.


Under melt conditions, the chain conformation is ideal, and in that case, the Flory
exponent is ν = ½, such that we get from the Rouse model (cf. Section 3.6.2):

τRouse = τ0 N 1 + 2ν = τ0 N 2 (5:87)

The latter equation features two characteristic timescales: τ0 , which denotes the
lower limit for any chain motion at all (below that timescale, not even single mono-
mer segments can get displaced by a distance at least equal to their own size), and
τRouse , which denotes the upper time limit for complete chain motion (above that
timescale, the whole coil can get displaced by distances equal to or even greater
than its own size). In the time domain between these two characteristic limits, the
modulus G relaxes from an energy-storage capacity of kBT per monomer segment at
τ0 to an energy-storage capacity of kBT per chain at τRouse (see Section 5.8.2). The
shear modulus at the Rouse time is therefore given by

ϕ
GðτRouse Þ = kB T (5:88)
Nl3

Here, ϕ is the polymer volume fraction, and Nl3 is the volume per chain in a fully

collapsed state, such that ϕ Nl3 is an expression for the number of chains per vol-
ume in the sample, denoted ν in Section 5.8.2. With that, the modulus at τRouse is
kBT per chain in the sample volume.
According to eq. (5.46), the viscosity at the Rouse time can then be calculated as

η = GðτRouse Þ · τRouse ⁓ N − 1 N 2 ⁓ N 1 (5:89)

We realize that it scales linearly (power-law scaling exponent of 1) with the degree
of polymerization, and with that, with the molar mass.
The picture is very different for entangled melts. Here, the relaxation is delim-
ited by the reptation time, τrep

τrep ⁓ N 3 (5:90)

At τrep, we find ourselves just at the end of the rubbery plateau of the shear mod-
ulus G. Here, the modulus just depends on the entanglement molar mass, Me,
(see eq. (5.76)), but not on he overall molar mass, Mtotal. This leads to a different
dependence of the viscosity on the molar mass, and with that, on the degree of
polymerization N, than that of eq. (5.89):
 
η = G τrep · τrep ⁓ const · N 3 (5:91)

In eqs. (5.89) and (5.91), we have again connected the macroscopic properties viscos-
ity, η, and shear modulus, G, to the microscopic property relaxation time, τ, that con-
tains information about the molecular-scale characteristics of the system’s building
236 5 Mechanics and rheology of polymer systems

blocks. In particular, we were able to delimit two very different viscosity–molar-mass


scaling laws, depending on whether a specific molar mass, the entanglement molar
mass, is surpassed. These dependencies are visualized in Figure 94.

Figure 94: Plot of a polymer melt sample’s viscosity as a function of the chains’ molar mass. Up to
certain molar mass, Me, the dependence is linear. Once this molar mass is exceeded, however, the
chains entangle with each other. This imparts severe topological constraint on the polymer chains’
motion, which is now to be described by P.G. de Gennes’ reptation model. Here, the viscosity has a
severe molar-mass dependence with a power-law exponent of 3.

But how come that we need a minimal Me for chain entanglement? The reason is the
polymer chain stiffness, as quantified by parameters such as the characteristic ratio,
the Kuhn length, or the persistence length, as all introduced in Chapter 2. The stiff-
ness of the chains requires them to have at least a certain length to bend enough to
form entanglements, which is longer if the chains are stiffer. An every-day life anal-
ogy for the phenomenon of a minimum length for entanglement is spaghetti: when
being served in a proper Italian restaurant, they are long and entangle with one an-
other quite markedly. By contrast, when being served in a large canteen, for example,
a university Mensa, they are short such that they can not entangle. The kitchen per-
sonnel does that to avoid processing problems of entangled noodles. So, even spa-
ghetti have a minimum length (or “molar mass”) for entanglement.
Getting back to Figure 94: note that experimental data actually suggest an η(M)
scaling-law exponent of 3.4 rather than 3 for the initial reptation regime. This is due
to fluctuations of the length of the confining tube. The chain ends fluctuate on
5.10 Terminal flow and reptation 237

timescales t < τrep. Because of this fluctuation, the effective tube length is actually
shorter, which decreases τrep. This effect, however, gets less significant when the
molar mass M increases. Hence, in addition to the primary effect of an M-increase on
η, as captured by eq. (5.90), comes a secondary one by the no-longer-effectiveness of
tube-length fluctuation. Together, this leads to the observed stronger molar-mass-
dependence of a power of 3.4. In the very high molar-mass regime, the effect of such
tube length fluctuation looses significance, such that the original reptation power-
law exponent of 3 is recovered there.

5.10.3 Reptation and diffusion

In addition to discussing the macroscopic viscosity of our sample, we may also quan-
tify a microscopic characteristic parameter of it: the translational diffusion coefficient
of the chains. On timescales shorter than τrep, the chains exhibit a one-dimensional
tube diffusion coefficient, DTube, as captured by eq. (5.85). On timescales longer
than τrep, by contrast, each chain diffuses three-dimensionally through space with a
Fickian diffusion coefficient Drep. We can imagine this diffusion to be a hopping
process from tube to tube, each of these hops taking a time increment of τrep. Based
on that picture, the macroscopic chain displacement in a three-dimensional space
can again be quantified by an Einstein–Smoluchowski equation:

R2 Nl2
Drep ≈ ⁓ 3 ⁓ N −2 (5:92)
τrep N

This power-law scaling has indeed been well confirmed by multiple sorts of
experiments.
When we focus on the microscopic diffusion of polymer chains in a reptation
scenario, as we have just started to do above by quantifying the two relevant chain
diffusion coefficients, Dtube and Drep, we can also take an even more detailed view.
In that detailed view, we consider one monomer segment on a reptating chain to be
somehow labeled, for example, by radioactive marking or fluorescent tagging.
When we follow the mean-square displacement, hΔr2 i, over time, t, of that labeled
segment, we can distinguish four different regimes, as shown in Figure 95(B).
1) On timescales between the segmental relaxation time, τ0, below which no seg-
mental motion at all is possible, up to the entanglement time, τe, from which
on the segments first realize that they are trapped in a tube environment, the
segmental displacement is impaired by the connectivity to neighboring seg-
ments. This constraint gives us a time dependence of the mean-square displace-
ment according to the Rouse model:
hΔr2 i ⁓ t1=2 (5:93)
238 5 Mechanics and rheology of polymer systems

Figure 95: (A) Visualization of a chain entrapped in a confining tube with end-to-end distance r and
contour length s. (B) Subdiffusive displacement of a labeled chain segment as a function of time on
different timescales during a reptation process. The mean-square displacement, <Δr2>, of the
labeled segment is proportional to time, t, only on scales longer than τrep, because only on these
long timescales, the labeled segment follows the three-dimensional Fickian diffusion of the chain
through space (which we may imagine as a hopping process from one tube to another with an
overall three-dimensional diffusion coefficient DRep). On shorter scales, by contrast, the motion of
the chain segment under consideration is sub-diffusive. First, on very short timescales, we have
subdiffusive spreading of the labeled segment due to the constraint imparted by its connectivity to
neighbor segments, which gives a time dependence of <Δr2> ⁓ t½ according to the Rouse model
(which applies in polymer melts). Second, on longer timescales, from the entanglement time τe on,
additional constraint comes into play, as our labeled segment now realizes that it is also
entrapped in a tube and can only move along the tube contour, which in 3D space is a random walk
with scaling of r ⁓ N½ such that the overall scaling exponent is ½ · ½ = ¼. Third, on even longer
timescales, from the Rouse time τR on, the first constraint is lost, but the second constraint is still
active, such that we still have subdiffusive scaling, now again according to <Δr2> ⁓ t½. Fourth, only
on timescales longer than the reptation time, τRep, also the latter constraint is lost and the chain
moves freely in space, and so does the labeled segment on it.

2) At the entanglement time, τe, the chain segment also takes notice of the addi-
tional confining tube. Up to the Rouse time, τRouse, the segmental displacement
along the tube is still Rouse type, but in three dimensions it is
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffi
hΔr2 i ⁓ hs2 i ⁓ t1=2 ⁓ t1=4 (5:94)

This is because the tube itself has a random walk shape with r ⁓ s1=2 . In other
words, in the time window τe–τRouse, our labeled segment is constrained by
both its connectivity to other segments, giving us one scaling-exponent contri-
bution of ½ according to eq. (5.29), and on top of that, it now also realizes that
it is entrapped in a tube and can only move along the tube contour, which in
pffiffiffiffiffiffiffiffi
three-dimensional space is a random walk with scaling of hr2 i ⁓ hs2 i, giving
us another scaling-exponent contribution of ½, such that the overall scaling ex-
ponent is ½ · ½ = ¼.
5.10 Terminal flow and reptation 239

3) At times longer than the Rouse time, τRouse, we now discuss the motion of the
entire coil along the tube. Its mechanism of motion is still a one-dimensional
Rouse-type diffusion, but in three dimensions it is
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi
hΔr2 i ⁓ hs2 i ⁓ DTube t ⁓ t1=2 (5:95)

4) On timescales longer than the reptation time, t > τrep, we have macroscopic
three-dimensional diffusion of the entire coil according to the Einstein–
Smoluchowski equation:

hΔr2 i ⁓ 6Drep t (5:96)

The polymer chain now has enough time to “hop” from one confining tube to an-
other and thereby move freely through space.

5.10.4 Constraint release

In the years after development of the reptation model, several refinements have been
added to it to better match to experimental data. An impactful refinement was the dis-
covery of the so-called constraint release mechanism. It takes into account that, in
addition to the test chain, all other chains of the system are in motion as well. As a
consequence, a topological constraint resulting from an entanglement point might re-
solve itself by diffusion of the constraining chain, as depicted in Figure 96. The effec-
tive diffusion is, thus, determined by reptation and the finite lifetime of the momentary
tube:

Figure 96: Constraint release as an additional mechanism for chain relaxation in entangled
systems. (A) Schematic of two entangled chains entrapped in their respective tubes. (B) Reptation
of one chain opens a degree of freedom for the other chain to rearrange itself not only by
curvilinear motion along its tube contour, but also lateral to it. Picture redrawn from M. Rubinstein,
R. H. Colby: Polymer Physics, Oxford University Press, 2003.
240 5 Mechanics and rheology of polymer systems

R2 R2
Deff ≈ + (5:97)
τrep τTube

Depending on which process is faster, the overall diffusion is dominated by repta-


tional motion or the constraint release mechanism. As a consequence, both the
tracer and matrix molar mass are important.

Questions to Lesson Unit 16

(1) Why can an entangled chain be described by the tube concept?


a. Because the entangled chain tries to achieve a preferably high end-to-end
distance, whereby a maximal value is achieved in the shape of a tube.
b. Because the surrounding chain segments overlap and form a conceptual
tube that wraps around the entangled chain.
c. Because the only shape that the entangled chain can take is that of a tube.
d. Because the tube shape shows the entropically most favorable conformation.

(2) Which potential describes the constraint of the test chain by its neighboring
segments?
a. A Coulomb potential, because of the electrostatic repulsion of the surround-
ing segments and the test chain.
b. A Lennard–Jones potential, because of the repulsive and attractive interac-
tions between the surrounding segments and the test chain.
c. A hard-sphere potential, because of the hard-sphere repulsion of the test
chain and the surrounding segments when colliding during movement.
d. A parabolic potential, which is just a simple and pragmatic concept without
explicitly referring to any type of interactions.

(3) How does the total molar mass of a polymer affect the rubbery elastic plateau
in its mechanical spectrum?
a. A higher molar mass leads to an overall increase of the shear modulus,
meaning a higher plateau value.
b. A higher molar mass leads to a greater extent of the plateau over a wider
range of time/frequency.
c. A higher molar mass leads to both a higher plateau value of the modulus
and a more extended plateau.
d. A higher molar mass does not affect the plateau in any manner.
Questions to Lesson Unit 16 241

(4) What holds for the scaling of the viscosity η with the degree of polymerization
N for the Rouse and the reptation time?
a. At the Rouse time, η scales with N, whereas at the reptation time it does
with N 3 .
b. At the Rouse time, η scales with N 3 , whereas at the reptation time it does
with N.
c. At the Rouse time and the reptation time, η scales with N 3 .
d. At the Rouse time and the reptation time, η scales with N.

(5) How do you derive the macroscopic property viscosity η from the reptation
model which describes a microscopic process?
a. The viscosity is directly implemented in the mathematical description of the
reptation model.
b. The viscosity can be calculated through the reptation time, which is derived
from the reptation model, by simply multiplying with the storage modulus G′.
c. The viscosity can be calculated through the reptation time, which is derived
from the reptation model, by simply multiplying with the loss modulus G′′.
d. The calculation of the viscosity from the reptation time must be done nu-
merically by computer simulation.

(6) Why do experimental data show a higher scaling law exponent of 3:4 for the
scaling of η with M (for molar masses above the minimal entanglement molar
mass Me ) compared to the theoretical value of 3?
a. With higher molar mass, the length of the polymer chains increases, which
leads to an overall increase in the relaxation time or viscosity, respectively.
b. An increase in the molar mass results in a higher pronounced constraint of
the test chain and therefore a smaller diameter of the tube according to the
tube concept, which leads to an increase in viscosity since the motion of
the test chain is more confined.
c. The term for the frictional coefficient only scales linearly in N if you consider
comparably shorter chains. With increasing molar mass, meaning with in-
creasing degreepffiffi of polymerization, the friction is described by a model that
scales with N 2 ≈ N 1:4 leading to an overall power-law exponent of 3:4.
d. The reptation time is effectively lower at the tube ends due to fluctuations
of the tube length, which are especially pronounced at the tube ends. This
effect decreases in magnitude with increasing molar mass.
242 5 Mechanics and rheology of polymer systems

(7) The diffusion coefficient can be derived from the reptation time by using an
Einstein–Smoluchowski approach. How does the diffusion coefficient scale in N?
a. N 2
b. N 2
c. N 1=2
d. N 3

(8) Which times mark the range in which the constraint by the tube is effective?
a. τ0 and τRouse
b. τe and τRouse
c. τRouse and τrep
d. τe and τrep
6 Scattering analysis of polymer systems

LESSON 17: SCATTERING METHODS IN POLYMER SCIENCE


The overall goal of polymer physical chemistry is to bridge between polymer structures and
properties. This requires both to be assessed. In this lesson, you will get to know scattering
methods as a prime experimental approach to assess polymer structures on different scales.
You will learn about the scattering vector as the “magnifying glass” in scattering, and you
will see how the scattering intensity as a function of this vector is connected to microstruc-
tural aspects.

6.1 Basics of scattering

The basic target of this book – and of the field of physical chemistry of polymers in
general – is to unravel relations between structure and properties of polymer sys-
tems. In Chapter 5, we have learned that the most relevant class of properties of poly-
mers are their mechanical characteristics, and we have learned about a class of
methods to quantify these: rheology. What we are lacking still is an equivalent class
of methods to characterize polymer structures. We know that the size of polymers
and their superstructures is in the colloidal domain of 10. . .1000 nm, so the method-
ology for structural characterization that we seek for must be able to resolve that.
Most polymer systems are amorphous, while only a few can show regular crystalline
structures. Hence, we also need a methodology that is generally able to assess both
in the colloidal size range. A class of methods that serves all that is based on scatter-
ing. The physics behind scattering has its ground in the interaction between quan-
tum objects and matter. The most prominent quantum objects in our focus are
neutrons, x-rays, and light; they all obey the particle–wave dualism, which is con-
nected by the De Broglie wavelength

h
λ= (6:1)

If a quantum object hits a particle or scattering center in a sample, it is re-irradiated


into space, and the overlay of multiple such scattered quantum objects that are re-
irradiated by multiple adjacent scattering centers in a sample causes a scattering in-
tensity pattern. Analysis of this pattern allows us to draw conclusions on the spatial
distribution of the scattering centers in the sample, and with that, about the structure
of the sample that we investigate. Depending on the wavelength of the scattered quan-
tum objects, we can do so on different length scales. The three mentioned kinds of
quantum objects can therefore all be used as “magnifying glasses” to provide structural
information about the sample, with different extent of magnification depending on their

https://doi.org/10.1515/9783110713268-006
244 6 Scattering analysis of polymer systems

wavelength. However, they do so from slightly different perspectives. This is caused by


the different ways in which contrast between the primary beam and the sample is gen-
erated during the scattering experiment. When light is used, the scattering contrast
stems from spatial variation of the refractive index in the sample, which is correlated to
spatial variation of the polarizability. For x-rays, contrast is caused by spatial distribu-
tion of electron densities, while for neutrons, contrast is caused from differences in the
so-called scattering lengths of certain atoms in the sample, which is most pronounced
for the isotope pair of hydrogen and deuterium. All these different sources of scattering
eventually trace down to the spatial distribution of matter in the sample, or in other
words, to the sample nanostructure, which can therefore be explored by scattering. To
properly account for the different origins of contrasting in light, neutron, and x-ray scat-
tering, special sample preparation is necessary in some of these methods. In the case of
neutron scattering, we need to tag interesting parts of the samples by exchange of hy-
drogen atoms by deuterium atoms.80 For x-ray scattering, we must have atoms with a
high electron density in the sample. While this might seem tedious at first glance, it of-
fers the chance to selectively label certain areas of a polymer, for example, its side
chains or chain ends, and then to selectively investigate them. Often, combinations of
the three scattering methods are used to generate a complimentary dataset.
A simple scattering experiment is the Bragg diffraction on a crystal lattice; it is
schematically shown in Figure 97. An incoming beam causes each scattering center

Figure 97: Bragg diffraction of a neutron or x-ray beam on crystal layers with a lattice constant of d.
The part of the wave front scattered at the lower lattice plane has to travel further to reach the
detector than the part scattered at the upper lattice plane. The extra distance traveled is
highlighted in blue color. Following the geometry, it is 2d sin θ. Note that due to the reflection at
an interface of an optically thin to an optically thick medium, the phase of the wave is shifted by
180° at the reflection point.

 The most simple way to realize that is to exchange the normal solvent by a deuterated one, as
also used in NMR spectroscopy.
6.1 Basics of scattering 245

(here: atoms in the sample) to re-irradiate incident x-rays as an isotropic spherical


wave. When looking from a specific scattering angle, θ, we detect an overlay of all
the radiation that is deflected into this specific direction. The sketch in Figure 97
shows that the part of the wave front scattered at the lower lattice plane has to travel
further than the part scattered at the upper lattice plane. Following the geometry, the
extra distance traveled is 2d sin θ. This causes an interference that is positive and
creates scattering peaks at the detector if that extra distance matches an integer mul-
tiple of the wavelength. This criterion is expressed in the Bragg equation

2d sin θ = nλ, with n = 1, 2, 3, . . . (6:2)

An alternative way to sketch this scattering experiment is with the concept of a


wave vector, as shown in Figure 98. Here, ~ ke is the wave vector of the entering
beam and ~ ks the wave vector of the scattered beam. The difference of both vectors is
called the scattering vector ~
q. The absolute values of both vectors are equal to

~ 2π
ke = ~
ks = (6:3)
λ

Figure 98: Vector representation of a scattering process. The entering wave vector is marked ~ ke ,
whereas ~ ks denotes the scattered vector. The difference of both generates the scattering vector ~q.
Picture redrawn from J. S. Higgins, H. C. Benoît: Polymers and Neutron Scattering, Clarendon Press,
Oxford, 1994.

From geometry (see Figure 98), we obtain the scattering vector ~


q as follows:

j~
qj = q = sinðθ=2Þ (6:4)
λ

With that, the Bragg condition for positive interference reads

2π d
= (6:5)
q n

The equation shows how the structural feature d relates to the scattering vector q:
the smaller the structures, the larger must q be to probe them. In other words, a low
q-value means a low magnification, and vice versa.
246 6 Scattering analysis of polymer systems

6.2 Scattering regimes

The scattering vector, our “magnifying glass”, can be varied by the wavelength of
the scattered quantum objects. Neutrons and x-rays have short wavelengths
(λ = 0.01–10 nm) and thereby realize high scattering vectors, whereas light has long
wavelengths (green laser λ = 543 nm, red laser λ = 633 nm), and its scattering therefore
covers small q-values. Further fine adjustments can be made by the scattering angle θ.
This allows the investigation of various structural characteristics of polymeric systems,
ranging from the entire polymer coil down to its local chemical structure. The accessi-
ble structural motifs of a polymer system are schematically shown in Figure 99. Light
scattering probes the first two domains. In the first, our zoom level is so low that the
polymer coils appear as pointlike masses, which allows for determination of their
molar mass (“zero-angle domain”). In the second regime, zoomed in a little more, the
polymer coil dimensions become visible (Guinier domain). To reach even more magni-
fication, we have to switch from light to smaller wavelength quantum objects. By
using either x-rays or neutrons, we zoom in a lot further and reach the third regime,
in which parts of the chain become visible (if we observe at small scattering angles).
Here, no more information on the chain size is gathered, and hence, the degree of

Figure 99: Scheme of the accessible scattering regimes. Light scattering can reach the first two
domains, in which the polymer chains appear either pointlike (A) or, in the Guinier domain, as objects
whose dimensions, shape, and mass can be determined (provided their radius of gyration is larger
than about 50 nm) (B). Small-angle neutron or x-ray scattering allows for further magnification. Here,
we lose information about degree of polymerization and polydispersity, but we gain knowledge about
substructural characteristics such as the persistence length in the case of linear chains in a dilute
solution (C‘) or the correlation length in the case of cross-linked or overlapping chains in a network at
semi-dilute concentration (C”). At further magnification, we start to observe single chain segments only,
until we finally resolve the local chemical structure of the polymer segmental units (D). Picture redrawn
from J. S. Higgins, H. C. Benoît: Polymers and Neutron Scattering, Clarendon Press, Oxford, 1994.
6.3 Structure and form factor 247

polymerization and the polydispersity of it becomes irrelevant. Instead, substructural


characteristics such as the persistence length or, in solution, the solution-network corre-
lation length are revealed. Even further magnification, as achieved at higher scattering
angles, gives us a glimpse on single chain segments that may appear rodlike already if
their persistence length is larger than the zoom level. Having reached the final regime,
at even higher scattering angles, we can resolve the local chemical structure and gain
information about chain tacticity and other (side-)chain orientations.

6.3 Structure and form factor

If we take Figures 97 and 99 together, we can guess already that there must be a
general connection between the q-dependence of the scattering intensity, I(q), and
the structure of the sample. This is indeed the case, and that connection is given by
the so-called structure factor, S(q). It is defined as

I ðqÞ = Δb2 SðqÞ (6:6)

The factor Δb is a measure of the difference in contrast to the species of interest and
its surroundings. For light scattering, this is connected to relative polarizability dif-
ferences. In small-angle x-ray scattering (SAXS), this is connected to relative elec-
tron-density differences that are especially strong if parts of the specimen are
labeled with heavy atoms. For small-angle neutron scattering (SANS), contrast is
generated by the so-called scattering-length density difference that is particularly
pronounced between deuterium and hydrogen atoms. Hydrogen substitution for
deuterium therefore allows for custom contrasting in SANS, but yet does not alter
the polymer too much in its other properties.
The reduced form of the above equation describes the differential scattering
cross section per unit volume:

1 ∂σ 1 lð~
qÞr2
· = · (6:7)
V ∂Ω V I0
where σ denotes the scattering cross section, or number of scattered quantum objects
per detector area, and Ω the solid angle under which the sample is observed. The ex-
ðlð~qÞr2 Þ
pression on the right side of the equation, V1 · I , is also called the Rayleigh ratio.
0
The structure factor, S(q), can itself be split into two contributors. For neutron
scattering, we have

SðqÞ = Nz2 PðqÞ + N 2 z2 QðqÞ (6:8a)

where N corresponds to the number of molecules of the species of interest in the


scattering volume (volume cross section of beam and detector view), where each
molecule contains z scattering centers. P(q) is the form factor that accounts for
248 6 Scattering analysis of polymer systems

intramolecular interferences happening when scattering waves originate from dif-


ferent scattering centers of the same (macro)molecule. It is directly connected to
the shape of that molecule, and is therefore called form factor. Q(q) accounts for
intermolecular interferences.
In light scattering, by contrast, we have

SðqÞ = Nz2 PðqÞ · Q′ðqÞ (6:8b)

where Qʹ(q) accounts for interferences of light scattered from the centers of mass of
different molecules. In the limit of a dilute solution, Q(q) → 0 and Qʹ(q) → 1, which
simplifies both expressions to

SðqÞ = Nz2 PðqÞ (6:8c)

The crux now lies in determining an analytical expression for P(q) that matches the
recorded dataset from the experiment. Note that at a concentration of c ! 0, when
no intermolecular interference is possible, P(q)! 1 for q ! 0. Hence, in the zero-
angle limit, the scattering intensity becomes independent of the scattering vector q
and reflects the number of molecules in the scattering volume as well as the num-
ber of scattering centers per molecule. This can then be translated to a molar mass,
and this quantity can therefore be determined from the scattering intensity in the
low-q limit. The mathematical definition of the form factor, P(q), is given by
1X z X z  
PðqÞ = 2
hexp − i~
q~
rij i (6:9)
z i=1 j=1

This equation defines a pair correlation function. We can understand what this is
and why it is the basis to estimate a form factor (and therewith a scattering intensity
pattern in the end) as follows: to calculate a spatially-dependent (that means, a
q-dependent) scattering intensity, we would actually need to have precise informa-
tion about the specific spatial distribution of the scattering centers in the detection
volume (for example, the monomer units in a polymer coil if our detection volume
shall be such that it just fits this one coil). What we only have, though, is average
information on their density in space (such as the Gaussian segmental density profile
in a polymer coil). A correlation function is some kind of “compromise” beween the
needed rich and the given poor information. The most simple type of correlation
function is the pair correlation: it estimates the probability to find two scattering cen-
ters at a directed (that means, vectorial) distance ~rij from each other. So, in this defi-
~
nition rij is a vector that joins two scattering centers i and j in the (macro)molecule
under consideration. The operator h. . .i indicates averaging over all orientations and
conformations. The orientational average of a function f(θ,ϕ) is
6.3 Structure and form factor 249

2ðπ ðπ
1
hf ðθ, ϕÞi = f ðθ, ϕÞ sin θdϕdθ (6:10)

ϕ=0 θ=0

When we apply this to eq. (6.9), we generate



1Xz X z
sin qrij
PðqÞ = 2 (6:11)
z i=1 j=1 qrij

For continuous bodies, the following alternative expression is also valid:


ð
1 sin qrij
PðqÞ = Ð nð r Þ dr (6:12)
nðrÞ dr qrij

where n(r) relates to the radial density function of the object of interest.

Figure 100: Sketch of the mean-square-average scattering intensity, <|I(qR)|2>, as a function of the
reduced scattering vector, qR for spherical particles. In the low Guinier region (qR << 1), the
scattering intensity does not depend on the scattering vector q, but only on the number of spheres
in the scattering volume. In the region at higher q, by contrast, characteristic peaks appear that
allow the size of the spherical particles to be determined.
250 6 Scattering analysis of polymer systems

The simplest geometry that a scatterer can have is that of a regular sphere.
Here, the scattering intensity does not depend on its orientation due to symmetry.
This yields

ðR !2
3 sin qr 2 9
Psphere ðqÞ = r dr = 6
ðsinðqRÞ − ðqRÞ cosðqRÞÞ2 (6:13)
R3 qr ðqRÞ
0

In that latter equation, the scattering vector appears as a dimensionless reduced


variable qR, which means that the scattering vector q is normalized to the length
scale of observation R. A typical result for a spherical object is shown in Figure 100.
Note that there is no dependence on the scattering vector q at the very low-q end of
the Guinier region. The scattering intensity there depends on the mass of spheres
in the scattering volume, thereby allowing their molar mass to be estimated.
A spherical object is easy to treat theoretically, because it is fully symmetric. For
more complicated objects such as polymer coils, we have to simplify the derivation
by looking onto the low-q and high-q extremes separately.
To capture the realities at low-q values corresponding to light scattering meas-
urements, we can expand the basic formula given by eq. (6.11) into a Taylor series
utilizing the expansion

sin x x2 x4
=1− + −  (6:14)
x 3! 5!
This yields

q2 X z X z
q2 X z X z
PðqÞ = 1 − hrij
2
i + hrij 4 i −    (6:15)
3!z2 i = 1 j = 1 5!z2 i = 1 j = 1

In this series expansion, the second term gives the radius of gyration, Rg, which is
defined as
1X z
Rg 2 = hri 2 i (6:16a)
z i=1

where ri is the vector joining the scattering point i to the center of mass of the object.
An alternative expression in an ij two-point notation is
1 X z X z
Rg 2 = hrij 2 i (6:16b)
2z2 i = 1 j = 1

With that, we generate

q2 2
PðqÞ = 1 − Rg +    ðZimmÞ (6:17)
3
6.3 Structure and form factor 251

The above equation allows us to construct a Zimm plot, which is the conventional
means of deriving the radius of gyration Rg, the weight-average molar mass Mw,
and the second virial coefficient A2.
The form factor may also be approximated by an exponential to generate the
so-called Guinier function

Rg 2 q2
pðqÞ ffi exp − for qRg < 1 ðGuinierÞ (6:18)
3

At high q-values, that is, in the regime of SANS and SAXS experiments, we resort to
a scaling discussion. We retain the normalized scattering vector, qR, as a variable
and use the following identity:

SðqÞ = Nz2 PðqRÞ = Vϕz PðqRÞ (6:19)

where V is the sample volume, N the number of scattering particles therein, and ϕ
the number of scattering units of degree of polymerization z per unit volume. Be-
cause we are only looking at a single, or possibly even just part of a single polymer
chain, N = 1. Nz now directly corresponds to the number of scattering centers per
chain. We can detect every Kuhn segment as a scattering center, and thus z corre-
sponds to the degree of polymerization.
We assume that at large scattering vectors, P(qR) has an asymptotic tail of type
(qR)–α. Furthermore, we know several relations between the mass and size of poly-
mers. For a Gaussian coil, it is given as R ⁓ z1=2 ; for a chain in a good solvent, it is
R ⁓ z3=5 ; and for a rodlike chain, it is R ⁓ z1 . In general, we can say that the mass
scales with the Flory exponent, R ⁓ zν . Replacing that for R in the above equation
yields

SðqÞ = Vϕz Pðqzν Þ = Vϕðqzν Þ − α = Vϕq − α zð1 − ναÞ (6:20)

At high q-values, the scattering intensity becomes independent of the degree of po-
lymerization z, as we look into each coil but do not see it in its totality. This, how-
ever, is only fulfilled if α = ν1, because only then will the z-factor be z0. This yields
q-dependencies for the aforementioned polymer geometries: SðqÞ ⁓ q − 2 for Gauss-
ian chains, SðqÞ ⁓ q − 5=3 for polymer coils in a good solvent, and SðqÞ ⁓ q − 1 for max-
imally elongated polymeric rods. Generally, we can say that SðqÞ ⁓ q − 1=ν .
Let us again come back to the most simple example of a regular sphere. For
that geometry, in the higher q regime,81 scattering can only come from the sphere’s
surface, because if we look completely inside or outside of it, there will be no scat-
tering as there is no spatial contrast difference. As a result, the scattering intensity

 To be precise: for q > 1/R, denoted as the Mie regime.


252 6 Scattering analysis of polymer systems

is proportional to the extent of the surface in the scattering volume. For the surface
S we can express R ⁓ S0.5 and Nz2 = NV 2 = NS3 . Plugging this into eq. (6.8b), again
under assumption of an asymptotic power-law tail, yields
  −α
SðqÞ = Nz2 PðqRÞ = NS3 qS0.5 (6:21)

Thus, the scattering intensity must be proportional to the total surface of the sam-
ple in the scattering volume, which is proportional to NS. That can only be fulfilled
at α = 4.

SðqÞ ⁓ q − 4 (6:22)

This is called Porod’s law and visualized in Figure 100 by the envelope slope of –4.

Questions to Lesson Unit 17

(1) Depending on the quantum objects used, three relevant types of scattering ex-
periments for polymer analysis can be differentiated: light, x-ray, and neutron
scattering. What holds for the suitability of these three variants?
a. Different variants are suitable for different types of polymer samples.
Whereas fully swollen polymers can be investigated using light, fully col-
lapsed polymers require the use of neutrons.
b. Whether or not a quantum object is suitable to probe a given polymer sample
depends on the size of the polymers, which should be in the range of the De
Broglie wavelength of the quantum object.
c. Each variant can be used on the same type of polymer sample. The only
difference lies in the different structural features that are imaged in the dif-
ferent variants, whereas light scattering images single coils, neutron scat-
tering images the structure along the polymer chain.
d. Each variant can be used on the same type of polymer sample, resulting in
different insights into the features of polymers, such as their molar masses
or persistence lengths.

(2) What is the criterion for a suitable wavelength to perform scattering analysis of
a given material?
a. The wavelength should match the distances of the scattering centers in the
sample.
b. The wavelength should be as different as possible from the distances of the
scattering centers in the sample.
c. The wavelength should be smaller than the distances of the scattering cen-
ters in the sample.
Questions to Lesson Unit 17 253

d. The wavelength should be bigger than the distances of the scattering cen-
ters in the sample.

(3) What is the role of data obtained from scattering analysis?


a. The data can be used to predict a microscopic structure.
b. The data can be used to determine the microscopic structure.
c. The data can be used to support a predicted image of the microscopic structure.
d. The data can be used for reaction monitoring.

(4) What holds for the relation of the interlattice distance d and the scattering vec-
tor q?
a. The smaller the distance d the lower must q be to achieve a smaller magnifica-
tion.
b. The smaller the distance d the lower must q be to achieve a higher magnifi-
cation.
c. The smaller the distance d the higher must q be to achieve a smaller magni-
fication.
d. The smaller the distance d the higher must q be to achieve a higher magnifi-
cation.

(5) Besides the wavelength of the scattered quantum objects, which is a parameter
that can be varied to achieve the desired magnification?
a. The temperature
b. The detection angle
c. The sample thickness
d. The intensity of the incident beam

(6) What does the form factor PðqÞ account for?


a. The form factor summarizes the contributions of intermolecular and intra-
molecular interferences.
b. The form factor accounts for intermolecular interferences.
c. The form factor accounts for intramolecular interferences.
d. The form factor accounts for the number of intermolecular interferences,
namely the number of scattering centers.

(7) Why does the mean-square-average scattering intensity of solid regular spheres
show characteristic peaks in the higher q Porod region but levels off to a con-
stant value at the lower q Guinier region?
a. At higher q, the magnification is high enough to detect single spheres, whereas
at lower q, this differentiation is gone. The boundary line for the two cases is
determined by the sphere’s radius.
254 6 Scattering analysis of polymer systems

b. At higher q, only parts of the sphere have an impact on the scattering inten-
sity, whereas at lower q, the whole sphere creates a scattering intensity that
levels off in a plateau.
c. At lower q, scattering is not pronounced, leading to a plateau of the inten-
sity. Only at higher q values, the scattering is pronounced enough to show
characteristic decreases in the intensity.
d. At lower q, the scattering vector does not suit the size of the spheres, lead-
ing to no scattering response. Only at higher q values, the scattering vector
and the sphere size are sufficiently similar, resulting in a noticeable scat-
tering signal.

(8) What is the common procedure when planning and analyzing scattering meas-
urements on samples? Choose the answer that sorts the following steps in the
right order.
① Use pair correlation function.
② Assume structure of the sample.
③ Compare experimental and theoretical data.
④ Conduct the type of scattering experiment.
⑤ Predict angular dependence of the sample.
a. ③ → ① → ⑤ → ④ → ②
b. ③ → ⑤ → ① → ④ → ②
c. ② → ⑤ → ① → ④ → ③
d. ② → ① → ⑤ → ④ → ③
6.4 Light scattering 255

6.4 Light scattering

LESSON 18: LIGHT SCATTERING ON POLYMERS


The structure of many polymer systems can be assessed by different scattering methods,
each with its specific pros and cons. The most widespread technique is light scattering. The
following lesson sheds a special view to this method and shows how light scattering can
serve to assess a polymer’s molar mass, shape, and interactions with the surrounding.

Based on the above general thoughts about scattering, we now look on how this is
reflected in actual experiments. We focus on light scattering, as this method is
much easier to realize than neutron scattering or x-ray scattering, and as it is less
demanding in view of the sample composition, in a sense that no heavy element or
isotope labeling is necessary. According to the discussion of the scattering vector q
in the last lesson, light scattering covers the zero-angle and Guinier regimes. This
means that information about the molar mass and the size of the polymer coils is
accessible.

6.4.1 Static light scattering

As we have learned earlier, light scattering generates its contrast from the relative
differences of the refractive indices of the sample and its environment. But how
does the light scattering process itself work? It starts with an incident, polarized
light beam that hits a polarizable object, that is, a scattering center. There, the light
induces a dipole moment μ according to

μind = α · Eincident = α · E0 cosðωtÞ (6:23)

where α is the polarizability of the object under investigation. Due to that induced
dipole moment, the scatterer starts to vibrate, and as a result, it emits radiation into
all directions. The scattered electrical field can be estimated according to

1 ∂2 μind 1
Es ⁓ sin ’ · ⁓ sin ’ · E0 αω2 · −cosðωtÞ (6:24)
r ∂t2 r

where 1r sin ’ · E0 αω2 is the amplitude of the scattered light that depends on the
distance, angle, and oscillation frequency of the source; it corresponds to E0,s. The
expression − cosðωtÞ denotes the periodicity of the scattered light that is also the
periodicity of the incident light. The intensity of the scattered light scales with the
square of the electrical field, I ⁓ E2 . This yields
256 6 Scattering analysis of polymer systems

2
E0, s Is
= ⁓ α2 ω4 (6:25)
E0 I0

We realize that the scattering intensity has a strong dependence on the frequency.
Blue light is scattered much stronger than other colors. This is the reason why we
see that part of the sunlight majorly when looking at the sky, which therefore ap-
pears blue to us82 (while the sun appears yellow, because we see the remaining
spectrum without the blue when looking at it).83 The exact expression is given by

Is π2 · α2 · sin2 ’
= (6:26)
I0 ε0 2 · r2 · λ4

where ε0 is the vacuum dielectric constant and λ the wavelength of the utilized
light.
Now that we know how to treat a single particle, we can expand our view to
multiple-particle systems. The easiest such system is that of a dilute ideal gas, be-
cause it contains N particles of the same nature as discussed above (spheres). Con-
sequently, we only have to extend the last equation by the number of particles, N:

Is, total π2 · α2 · sin2 ’


=N · (6:27)
I0 ε0 2 · r 2 · λ 4

The opposite to a dilute ideal gas is an ideal crystal. Such a material, however, can-
not be investigated by light scattering. This is because the distance of the lattice
planes is in the range of some picometers, whereas the wavelength λ of the utilized
visible light is a couple of hundreds of nanometers. This causes each beam scat-
tered in any direction to always find a counterpart point source that causes destruc-
tive interference, and therefore extinction of the scattered light. This is the reason
why crystals are transparent. They might at most be colored, but this is caused by
light absorption rather than scattering.
At a first glance, the latter conceptual picture is also similar for liquids: al-
though there is no perfect order like in a crystal, the molecules in a liquid are still
so densely packed that their average distances from one another are still much
smaller than the light’s wavelength. Again, as a result, the light is extinguished and
the liquid has a transparent appearance. In contrast to a crystal, however, density
fluctuations can occur due to the thermal motion of the molecules. As a result, not
all of the scattered rays interfere with an extincting counterpart at a given time, and

 The oceans appear blue, too, because they reflect the blue color of the sky above them.
 To be precise, the sky actually appears blue to us only when the light passes a short distance
and is therefore scattered not too often, which is the case at the height of the day. In the morning
and evening, by contrast, the light path is longer, so the blue components are scattered away
completely, and we see the remaining red-orange parts at dawn and dusk.
6.4 Light scattering 257

so we do observe some scattering. As an example, consider liquid water: it scatters


200 times stronger than the gaseous water. It should be kept in mind, however, that
the particle-number density in liquid water is 1200 times higher than in gaseous
water. The “normalized” scattering therefore is only 1/6 times higher, so there is in
fact still notable extinction due to destructive interference.
The polarizability of a molecule is usually anisotropic. If the molecule rotates
and vibrates, this manifests itself such as if the polarizability α would oscillate:

α = α0 + αk cosðωk tÞ (6:28)

where ωk is the eigenfrequency of the molecule that is closely connected to its rota-
tional and vibrational eigenfrequencies. We can insert the above expression into
eq. (6.23) to generate

μind = α · Eincident = α0 · Eincident + αk · cosðωk tÞ · Eincident j Eincident = E0 · cosðωtÞ

= α0 · E0 · cosðωtÞ + αk · E0 · cosðωtÞ · cosðωk tÞ


1
= α0 · E0 · cosðωtÞ + αk · E0 · ½cosðω − ωk Þt + cosðω + ωk Þt (6:29)
2

where the first term α0 · E0 · cosðωtÞ accounts for the Rayleigh scattering that is a
predominantly elastic scattering by particles much smaller than the wavelength of
the radiation. Elastic scattering means that the frequency of the scattered light is
the same as that of the incident light, which means that Rayleigh scattering is not
accompanied by any energy exchange between the photon and the scatterer. The
other two terms account for Raman scattering, which encompasses the fraction of
the light that is scattered inelastically, whereby the scattered photons have a fre-
quency and energy different from those of the incident photons. When the scatterer
absorbs energy and the scattered photon has less energy than the incident one, we
have Stokes Raman scattering, whereas an energy transfer to the scattered photon
so that its energy is higher than that of the incident one is called anti-Stokes
Raman scattering. More important for static light scattering of liquids or solutions
is the Rayleigh scattering part of the process, which is why we limit our further dis-
cussion to this type.
Let us first examine the case of a dilute polymer solution. Its refractive index,
nsolution, is given by
2
dn square dn dn
nsolution = nsolvent + c ) n2solution = n2solvent + 2nsolvent c+ c (6:30)
dc dc dc

Due to its negligible influence, we will disregard the final term of this expression,
2
dn
c .
dc
258 6 Scattering analysis of polymer systems

The refractive indices n are connected to the polarizability α via the following
expression:

ðN=V Þ
n2 − 1 = α (6:31)
ε

Applied to the refractive index of the solution, nsolution, we generate


     
Nsolvent Nsolute Nsolute
V − V V
n2solution − 1 = αsolvent + αsolute (6:32)
ε0 ε0
The corresponding expression for the refractive index of the solvent, nsolvent, is
 
Nsolvent
V
n2solvent −1= αsolvent (6:33)
ε0
By subtracting the former two expressions from one another and taking eq. (6.30)
into account, we can derive a formula that connects the refractive index differences
to the incremental refractive index change with concentration, dn dc :
 
Nsolute
V dn
n2solution − n2solvent = ðαsolute − αsolvent Þ = 2nsolvent · c (6:34)
ε0 dc

Solving this for an equation describing the polarizability difference, Δα = αsolute − αsolvent ,
yields

dn c dn M N c · NA
Δα = 2nsolvent · ε0 = 2nsolvent · ε0 , with = (6:35)
dc N=V dc NA V M

We can now insert this expression into the Rayleigh formula from above (eq. 6.27):84
 2 2
Is, total π2 · 4n2solvent · ε0 2 dn
dc M cNA V
=N · · sin2 ’ jN=
I0 NA · ε0 · r · λ4
2 2 2 M
 2 2
cNA V π2 · 4n2solvent · ε0 2 dn
dc M
= · · sin2 ’
M NA · ε0 · r · λ4
2 2 2

2
dn 4π2 · n2solvent · sin2 ’
=c·M·V · · (6:36)
dc NA · r2 · λ4

 Equation (6.27) has α in it, because it holds for gases, where refractive index differences occur
because we either have a gas molecule (with polarizability α) or nothing in a spot; in the case of
solutions, we need to work with Δα instead, because then, refractive index differences occur be-
cause we either have a solvent molecule (with polarizability αsolvent) or a solute molecule (with po-
larizability αsolute) in a spot.
6.4 Light scattering 259

While this equation might seem complicated at first glance, all parameters except
for the concentration c and the molar mass M are constant for a given system and
setup. This allows us to determine the molar mass M of a compound by performing
concentration-dependent static light scattering measurements.
As light is scattered into all directions, there is a trivial sample–detector dis-
tance dependence of r–2. Furthermore, the scattering volume V has a trivial influ-
ence on the scattering intensity detected: the larger the volume, the larger is the
I
relative intensity, s, Itotal . To get experimental setup-independent results, we normal-
0
ize to these two factors and introduce the Rayleigh ratio, Rθ, that has the unit m–1:

Iθ r2 4π2 n2 dn 2 2
Rθ = = sin ’ · cM (6:37a)
I0 V NA λ4 dc

We can simplify this expression greatly by combining all the constant values into a
constant K. This yields

Iθ r 2
Rθ = =K·c·M (6:37b)
I0 V
In a polydisperse sample, all components with molar mass Mi and the correspond-
ing concentration ci scatter independently, yielding:
X
Rθ = K cM
i i i
(6:38)

We know from Section 1.3.2 that the weight-average molar mass of a polymer is de-
fined as
P P
Ni Mi 2 i ci Mi
Mw = Pi = P (6:39)
i N i Mi i ci

Inserting this into the expression for the Rayleigh ratio, Rθ, yields

Kc 1
= (6:40)
Rθ Mw

With this equation, we can directly calculate the weight-average molar mass from
the Rayleigh ratio measured during a scattering experiment.
An alternative viewpoint can be gained from fluctuation theory, which com-
bines thermodynamics with light scattering. To account for and quantify concentra-
tion fluctuations, which are the actual cause of scattering from solutions, we regard
our system to be composed of many elementary cells δV, that are large compared to
λ 3
molecular scales but smaller than ð20Þ . Concentration fluctuations occur due to mo-
lecular exchange between these cells, and are described by

 + δα
c = c + δc ) α = α (6:41)
260 6 Scattering analysis of polymer systems

Note in this context that strictly speaking, there are fluctuations of the polarizability
in each elementary cell, α, not only due to concentration fluctuations, but also due to
other types of fluctuations in the system, such as temperature and pressure:

∂α ∂α ∂α
δα = ∂c + ∂p + ∂T (6:42)
∂c p, T ∂p c, T ∂T c, p

However, these fluctuations affect both the solvent and the solute equally, so we
may disregard them.
Going back to the original scattering equation (6.27), we need an expression for
α2. It is calculated from eq. (6.41) as

 + δαÞ2 = α
α2 = ðα  2 + 2α
δα + ðδαÞ2 (6:43)

2 , is
Let us examine the three terms of this expression a little closer. The first one, α
the same for all volume elements very much like in an ideal crystal, so there is no
contribution due to interference. The second term, 2αδα, can either have a positive
or a negative value, but it will be zero upon averaging. This only leaves the third
term ðδαÞ2 to actually cause the observed scattering intensity. By inserting this last
term into the Rayleigh equation, we generate

Is, total π2 · ðδαÞ2 · sin2 ’


=N · (6:44)
I0 ε0 2 · r2 · λ4

Note that the sin2 ’ term is due to light polarization in this case.
The mean-square fluctuation of the polarizability α is connected to the refrac-
tive indices n as follows:

dn 2 2
hδα2 i ⁓ n2 hδc i (6:45)
dc

The mean-square fluctuation of the concentration, hδc2 i, is given by statistical ther-


modynamics as
RTc
hδc2 i = ∂π (6:46)
∂c T

Plugging this into our above formula and using the virial series expansion of the con-
centration-dependent osmotic pressure, we get:

Kc 1
= + 2A2 c +    (6:47)
Rθ Mw

where A2 is the second virial coefficient of the osmotic pressure, obtained as a


z-average.
6.4 Light scattering 261

λ
For molecules that are larger than a twentieth of the used wavelength, 20 , or
roughly 20 nm, there will be additional intramolecular interference due to optical
path differences as sketched in Figure 101. This leads to an additional angular depen-
dence of the scattered light, and necessitates to account for an angular-dependent
particle form factor, P(θ):

Kc 1
= + 2A2 c +    (6:48)
Rθ Mw PðθÞ

Figure 101: Schematic of the optical path difference, Δλ, of two interfering beams scattered at
different points of a polymer coil with different angles of observation. The lower left pair of beams
has a bigger scattering angle, resulting in a bigger optical path difference, whereas the lower right
pair of beams has a smaller scattering angle and therefore a smaller optical path difference.

We can use the formula for the form factor that we derived in the last section
(eq. 6.17) and insert into it the expression for the scattering vector q (eq. 6.4) to
generate
1 16π2 n2 θ
PðθÞ = 1 − q2 hRg 2 i = 1 − hRg 2 isin2 (6:49)
3 3λ2 2

Note that sin2 θ here is not due to light polarization, but comes from the definition
of the scattering vector q.
Inserting the above expression into eq. (6.48) yields

Kc 1 16π2 n2 θ
= 1+ hRg 2 isin2 + 2A2 c (6:50)
Rθ M w 3λ2 2

Measuring the Rayleigh ratio Rθ as a function of the scattering angle θ and the con-
centration c allows three parameters to be determined simultaneously: the weight-
average molar mass Mw, the radius of gyration Rg, and the second virial coefficient
A2. This determination is often done graphically using a Zimm plot. For this purpose,
262 6 Scattering analysis of polymer systems

experiments are performed at several angles that all satisfy the condition qRg < 1, and
at least at four concentrations. Instead of creating two graphs for the angle- and the
concentration-dependent series, both can be plotted in a single graph, as shown in
Figure 102. From the linear extrapolation of the angle-dependent series as well as the
concentration-dependent series to θ ! 0 and c ! 0, respectively, the weight-average
molar mass can be determined. The radius of gyration can be calculated from the
slope of the angle series at the zero concentration, and the second virial coefficient
can be calculated from the slope of the concentration series at the zero angle.85
In statistical thermodynamics, the mean-square concentration fluctuation may
also be given as

RTc ∂G
hδc2 i =   with g = ðGibbs free energy per volume elementÞ (6:51)
∂2 g ∂V
∂c2 p, T

From this expression, we see that

Is RTc
⁓2  (6:52)
I0 ∂ g
∂c2 p, T

At a certain critical demixing point, we have a horizontal slope of G(c), meaning


2
that ð∂∂cG2 Þ = 0, as shown in Figure 103. At this point we observe very strong scatter-
ing, the so-called critical opalescence.

85 Note, in this context, that the quantities Rg and A2 are obtained as z-averages from the Zimm
plot, whereas the molar mass is obtained as a w-average, Mw. This is due to the following reason-
ing. In general, the electrical field amplitude, E, of a single light-scattering particle (or of a polymer
coil in the limit of θ ! 0, where intramolecular interferences are irrelevant) is proportional to its
mass: E ⁓ m (or, if you wish, to its molar mass: E ⁓ M). The intensity of the scattered light scales
with the square of that electrical field amplitude: I ⁓ E2 . This means that I ⁓ m2 (or I ⁓ M2 if you
like that better). For an ensemble of N particles, we get I ⁓ N·m2 (or, if we express it in molar num-
bers, I ⁓ (n/NA)·M2, with NA the Avogadro number), which already resembles the basic definition
of a quantity’s z-average (whereas that of a w-average is related to n·M1 and that of an n-average
relates to n·M0). As a result, any quantity obtained from light scattering is generally received as a
z-average, including the hydrodynamic radius RH (that is determined from dynamic light scatter-
ing, as detailed in the following section), the radius of gyration Rg, and the second virial coeffi-
cient A2. The molar mass, however, is determined from the Zimm equation, RKc = M1w (6.40) (which
θ
reflectsPthe Zimm plot in the limit of θ ! 0 and c ! 0). The concentration in the numerator equals
ni Mi P
to c = i , whereas the Rayleigh ratio in the denominator equals to Rθ = K i ci Mi according to
V P P
nMM nM2
eq. (6.38). Together, this gives Rθ = K i i i i = K i i i . Plugging both into the Zimm formula
P V V
nM
gives P in iM i2 = M1w , which directly reflects the definition of a w-average for the molar mass accord-
i i i
ing to eq. (6.39).
6.4 Light scattering 263


Figure 102: Plot of RKc as a function of sin2 θ2 + kc, the so-called Zimm plot. The constant k can be
θ
chosen arbitrarily to make the plot look clear in a sense that the data points are spread equally in
the x- and y-directions. From the common y-intercept of the extrapolations for θ ! 0 and c ! 0,
the weight-average molar mass, Mw, can be determined. On top of that, the slope of the θ ! 0
extrapolated line gives the second virial coefficient, A2, whereas the slope of the c ! 0
extrapolated line gives the radius of gyration, Rg. Picture redrawn from B. Tieke: Makromolekulare
Chemie, Wiley VCH, 1997.

Figure 103: At the critical demixing point, the slope of G(c) is horizontal, such that very intense
scattering, called critical opalescence, occurs.
264 6 Scattering analysis of polymer systems

6.4.2 Dynamic light scattering

Figure 104: Data processing in dynamic light scattering. (A) Time-dependent intensity fluctuations
as recorded by dynamic light scattering. (B) Intensity autocorrelation as a function of τ, shown both
in a linear and in a (C) semilogarithmic plot. The characteristic time τ0 corresponds to the decay of
the autocorrelation function to 1/e of its original value and can be easily spotted in the
semilogarithmic representation as the inflection point of the function.

When light scattering measurements are carried out in a time-dependent fashion, we


talk about dynamic light scattering. With this method, as its name implies, we can
obtain information about the dynamics of a sample. This is because scattering centers
in the probe volume are subject to thermal motion: they constantly change their mu-
tual distances, thereby causing temporal fluctuation of the scattering intensity due to
temporal fluctuation of the intermolecular interference. A time-dependent graph of
these fluctuations is shown in Figure 104(A). The fluctuating scattering intensity can
be mathematically treated by autocorrelation:

hI ðtÞI ðt + τÞi
g2 ðτÞ = (6:53)
hI ðtÞi2

The index 2 indicates that this is an intensity autocorrelation function. An index


of 1 denotes the electrical field autocorrelation function, with E instead of I. In such
an autocorrelation analysis, we shift a duplicate of the fluctuation intensity signal
along a conceptual time axis by a lag time τ. The further we shift, the less similar is
the duplicated signal to the original. As a result, the autocorrelation function g2(τ)
decays, as shown in Figure 104(B). A characteristic point is the decay to 1/e of the
original at the characteristic time τ0, as also indicated in Figure 104(B). It is easily de-
termined as the inflection point in a semilogarithmic plot, as shown in Figure 104(C).
At this time, the scattering centers have displaced themselves by a distance of 1/q. This
results in the following Einstein–Smoluchowski analog equation:
1 1
≈ Dτ0 , = Γ ≈ Dq2 (6:54)
q2 τ0

Measuring τ0 at different scattering vectors q, that is, at different scattering angles,


shows whether the motion in the sample is diffusive or not. For diffusive motion, we
Questions to Lesson Unit 18 265

need to find a linear interdependence of 1/τ0 and q2. If this is the case, then we can
calculate the prime resulting parameter obtained by DLS: the diffusion coefficient
D. Using the Stokes–Einstein equation, we can translate this diffusion coefficient into
a hydrodynamic radius, Rh. This is not, however, the precise hydrodynamic radius of
the sample, but that of an imaginary, perfectly spherical particle that diffuses ideally
and has the same diffusion coefficient as the one obtained from the experiment. For
polymers, this assumption is valid most of the time, because the shape of a Gaussian
coil can be assumed to be spherical on timescales longer than the Rouse or Zimm
time.

Questions to Lesson Unit 18

(1) Why does the sky appear blue?


a. Because the other parts of the spectrum are absorbed by molecules in the
atmosphere, only the blue parts remain.
b. Pollution in the atmosphere scatters mainly the blue parts of the spectrum.
c. At certain altitudes, molecules in the air appear blue.
d. Due to its higher frequency, blue light is scattered stronger than the other
parts of the visible spectrum.

(2) Why can an ideal crystal not be investigated by light scattering?


a. The lattice sites are so close that light with a shorter wavelength (UV at
least) would be necessary.
b. The lattice planes are so close that their distance is way smaller than the
used wavelength. Every scattered beam always finds a counterpart and in-
terferes destructively.
c. The distances of the lattice planes are all the same. A scattered beam imme-
diately gets absorbed by the next scattering center that it meets, which then
scatters the beam again. The crystal becomes impermeable for light.
d. The distances in the crystal are well defined. That is why monochromatic
beams are necessary, which cannot be achieved in the visible light spectrum.

(3) What do Rayleigh and Raman scattering correspond to?


a. Inelastic scattering with energy loss and energy gain.
b. Elastic and inelastic scattering.
c. Elastic scattering based on visible and UV light.
d. Inelastic scattering of particles smaller and bigger than the used wavelength.
266 6 Scattering analysis of polymer systems

(4) What is the benefit of introducing the Rayleigh ratio?


a. By normalization of the relative intensity with V and r2 , we get a setup-
independent variable.
b. The Rayleigh ratio is an intensity normalized to the concentration, meaning
that it is independent of the sample’s concentration.
c. The Rayleigh ratio is directly related to the molar mass of a polydisperse
sample, leading to an easy experimental determination of this parameter.
d. Since the unit of the Rayleigh ratio is m1 , it is directly related to the wave-
number of the scattered light.

(5) Which parameter has no influence on the optical constant K?


a. The wavelength
b. The polarizability
c. The refractive index
d. The vacuum permittivity

(6) Why is a virial-series expansion of the Zimm equation necessary?


a. It is not possible to determine the radius of gyration with the Zimm equa-
tion. This parameter can be introduced by a virial-series expansion.
b. For real polymer solutions that are not infinitely dilute, additional interac-
tions must be considered. This is pragmatically accounted for by a virial-se-
ries expansion.
c. By a virial-series expansion, more complicated special cases like multiple
scattering can be included.
d. With increasing polymer size, the form factor cannot be neglected anymore;
it can be identified as one coefficient of the virial-series expansion.

(7) Which parameters can be derived from the Zimm plot?


a. Radius of gyration, second virial coefficient, weight-average molar mass
b. Hydrodynamic radius, second virial coefficient, weight-average molar mass
c. Radius of gyration, second virial coefficient, Z-average molar mass
d. Weight-average molar mass, form factor, radius of gyration

(8) What is the outcome of a DLS experiment?


a. Dref
b. RH
c. τcorrel
d. η
6.4 Light scattering 267

LESSON 19: LIGHT SCATTERING ON POLYMER GELS


Polymer gels consist of three-dimensional networks of interlinked chains swollen by solvent;
they are a prime representative of polymer-based soft matter, with diverse applications relying
on their viscoelastic soft mechanics and their ability to take up and be penetrated by solvents
and solutes. These applications have a strong dependence on the gel microscopic structures,
given by the polymer network mesh sizes in a range of 1–10 nm and by spatial inhomogenei-
ties in the gel cross-linking density in a range of 10–100 nm. To quantitatively assess these
structural features, scattering techniques are the method of choice. The following excursion
presents the principle and prospect of static and dynamic light scattering to characterize the
structure of polymer gels.

6.4.3 Light scattering on polymer gels

6.4.3.1 Polymer networks and gels


Polymer gels are three-dimensional networks loaded with solvent, whose chains are
cross-linked physically (that means, through the action of coordinative or hydrogen
bonds, ion pairs, or other weak interactions) or chemically (that means, by covalent
bonds). In principle, they can be formed by any of the following mechanisms:
(1) Copolymerization of mono- and multifunctional monomers
(2) Polymer analogous reaction of pre-polymerized linear chains with cross-link-
able groups
(3) End-linking of linear polymer chains with multifunctional cross-linkers
(4) Cross-linking of star-shaped macromonomers as precursors with suitable end-groups
(5) Cross-linking of linear chains by a (typically harsh) transfer reaction

In an idealized view, a polymer gel consists of a single network with homogeneous


distribution of cross-linking points on scales of some few nanometers. In a realistic
view, by contrast, the structure of polymer gels is usually more complex due to
micro- and/or nanostructural inhomogeneities and irregularities, whose magni-
tude depends on which of the above ways of gel formation has been used.86 Shi-
bayama and Norisuye suggested the following classification:
(1) Spatial inhomogeneities due to a spatially inhomogeneous cross-linking
density
(2) Topological inhomogeneities due to local defects such as loops, cross-linker–
cross-linker shortcuts, or loose dangling chain ends in the mesh structure of
the network
(3) Connectivity defects during the process of gel percolation

 S. Seiffert: “Origin of Nanostructural Inhomogeneity in Polymer-Network Gels.” Polymer Chem.


2017, 8, 4472–4487.
268 6 Scattering analysis of polymer systems

The last two points can be summarized as local defects and in turn cause the spa-
tial inhomogeneities on larger scales that are listed in the first point. On top of
that classification, Okay further delimited the term “inhomogeneity,” referring to
“fluctuation of the cross-link density in space” from the term “heterogeneity,” refer-
ring to “the existence of phase-separated domains in polymer gels.”87 Many of the
most important gel properties, including the viscoelastic soft mechanics, the optical
clarity, and the ability to take up and be penetrated by solvents and solutes, are
decisively influenced by the (inhomogeneous) structure of the polymer network. As
a result, the quantitative assessment of that structure is a key to rational materials
design of polymer gels.

6.4.3.2 Static light scattering on gels


If the scattering intensity of a polymer gel is compared with that of an equally pre-
pared non-cross-linked polymer solution of the same concentration (in the semi-di-
lute regime, that means, in a range where chains overlap already), it turns out that
the scattering intensity of the gel is higher. This is because light scattering in a solu-
tion is caused only by temporal local random concentration fluctuations, whereas
in a gel, there are two types of concentration fluctuations:
(1) dynamic thermal concentration fluctuations caused by molecular motion,
δcF ð~

(2) immobile “frozen” concentration fluctuations by an inhomogeneous distri-
bution of cross-linking points across the gel, δcC ð~

The total concentration fluctuation δcð~


rÞ at a position ~
r is composed of both these
contributions:

rÞ þ δcC ð~
rÞ ¼ δcF ð~
δcð~ rÞ (6:55)

Figure 105 schematizes this overlay. Analogous to that, the scattering intensity, and
with that also the Rayleigh ratio of a gel, RGel , is a superposition of two independent
scattering contributions: the fluid scattering intensity RF by thermal fluctuations,
which is largely equal to the scattering intensity of an alike uncross-linked polymer
solution, and the excess scattering REx caused by the frozen spatial inhomogenei-
ties of the gel:

RGel ¼ RF þ REx (6:56)

Just like the position-dependent “frozen” concentration fluctuation, δcC ð~


rÞ, the scat-
tering intensity of a gel is also position-dependent, as shown in Figure 106. Because

 This view is in line with an earlier perspective by Dušek and Prins, who denoted “heteroge-
neous” gels to be formed by microsyneresis during their gelation.
6.4 Light scattering 269

Figure 105: Scattering intensity of a polymer gel as a function of the local nanometer-scale spatial
position r due to thermal fluctuations, “frozen” spatial inhomogeneities, and both.

of that permanent position dependence of these gel properties, a polymer gel is a


non-ergodic system. Therefore, light-scattering measurements must be conducted
at several different sample positions, and a proper statistical treatment of the result-
ing data is necessary to characterize the structure in detail, as discussed further. In
practice, this can be achieved through the use of a rotating cuvette unit in combina-
tion with computer-based repetitive measurement protocols.

Figure 106: Speckle pattern of the time-averaged scattering intensity of a gel, hIit;p , probed at
different positions p, along with a denotation of the ensemble average hIiE and the ensemble
average of a hypothetic fluid-only reference, IFE .

Figure 106 displays a typical speckle pattern of the position-dependent scattering


intensity in a gel, obtained by measurement of the time-averaged scattering inten-
sity hIit;p at different positions p. We see that the scattering intensity differs from
position to position. By averaging over several positions, the ensemble average hIiE
can be determined. With the assumption that the minimum of the scattering intensity
270 6 Scattering analysis of polymer systems

hIF iE corresponds to a plain fluid reference, the Rayleigh ratio of the excess scattering
hREx i can be calculated:

hREx i ¼ hRiE  hRFE i (6:57)

Analogous to the scattering intensity of large particles in solution (which is subject


to intraparticulate interferences), the excess scattering hREx ðqÞi is angular depen-
dent. In a simple approach, the angular-dependent excess scattering hREx ðqÞi is
conceived as a product of the zero-angle scattering intensity REx ð0Þ and a q-depen-
dent structure factor, SEx ðqÞ, which can be treated analogous to the form factor of
particle solutions:

REx ðqÞ = REx ð0Þ · SEx ðqÞ = 4π · K · Ξ 3 n2 · SEx ðqÞ (6:58)

In these equations, Ξ is the static correlation length, which is a measure of the aver-
age distance over which spatial concentration fluctuations occur in the gel; it typi-
 
cally has values of some tens to hundreds of nanometers. n2 is the mean-square
variation of the refractive index, which is a measure of how strong the concentra-
tion varies over that distance; it can be further translated to concentration fluctua-
tions by knowledge of the refractive index increment, dn/dc, which usually turns
out to be several single to tens of percent of that of the overall polymer concentra-
  
tion in the gel. K ¼ 8π2 n2solvent λ4 is an optical parameter.
Several empirical theories deliver expressions for the structure factor of gels, as
summarized in Table 10. With the linearized equations in the lowermost row of
Table 10, the static correlation length Ξ and the mean-square variation of the re-
 
fractive index n2 are accessible by plotting and linear fitting (to obtain the slope
m and intercept b) of the dataset, thereby delivering information on the magnitude
and lengthscale of spatial inhomogeneities in the gel.

6.4.3.3 Dynamic light scattering on gels


Analogous to the previous consideration for static light scattering, the temporally fluc-
tuating intensity of a gel in dynamic light scattering is composed of two contributions:

IGel ðtÞ ¼ IF ðtÞ þ IC (6:59)

(1) Time-dependent fluctuating intensity IF ðtÞ caused by diffusive motion of


polymer chain segments
(2) Constant scattering contribution IC caused by immobilized static inhomoge-
neities in the gel

Analytically, a temporal intensity autocorrelation is performed by multiplying the


intensity I ðq; tÞ at a time t with the intensity I ðq; t þ τÞ shifted by a lag time τ. A
Table 10: Prominent theoretical approaches to model the structure factor of polymer gels.

Theory Debye–Bueche Ornstein–Zernike Guinier

Explanation

Two-phase system with sharp boundaries One species distributed in another Randomly distributed domains
with different densities
1
 
Structure factor, SEx ðqÞ 1 SOZ ðqÞ = SGU ðqÞ = exp q2 Ξ GU 2
SDB ðqÞ = 1þq2 Ξ OZ 2
ð1þq2 Ξ DB 2 Þ2
 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1 qffiffiffiffiffiffiffiffi 1 1
 
Linearized form 1=2 Ξ DB 2 REx ðqÞ = þ 4πKΞ1 2 q2 lnðREx ðqÞÞ = ln 4π · K · Ξ GU 3 n2  Ξ 2Gu q2
REx ðqÞ = 2 πKΞ DB 3 n2 + 21 πKn 2q
4πKΞ OZ 3 n2 OZ n

m1=2 m1=2
Ξ b b jmj1=2
 1  1  1
Æn2 æ 4π · K · m3=2 b1=2 b1=2 4π · K · m3=2 expðbÞ · 4π · K · jmj3=2
6.4 Light scattering
271
272 6 Scattering analysis of polymer systems

normalized form of the intensity autocorrelation function is obtained from the in-
tensity measurement signal by hardware correlators:

hI ðq; tÞ · I ðq; t þ τÞi


gð2Þ ðq; τÞ ¼ (6:60)
hI ðq; tÞÞi 2

For dilute solutions, the field correlation gð1Þ ðq; τÞ is accessible from the intensity
correlation gð2Þ ðq; τÞ by the Siegert relation:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  
gð1Þ ðq; τÞ ¼ gð2Þ ðq; τÞ  1 ¼ exp Dq2 τ (6:61)

If, as in solutions, the light is only scattered by the (cooperative) diffusive motion of
polymer network-mesh chain segments, then this is called homodyne scattering.
In polymer networks, though, there are also constant scattering contributions caused
by static inhomogeneities, causing partial heterodyning. Both are illustrated in Fig-
ure 107.

Figure 107: Homodyne and heterodyne scattering in gels. Picture redrawn from M. Shibayama, Bull.
Chem. Soc. Jpn 2006 79(12), 1799–1819.

The field correlation function gð1Þ for one segment subject to thermal motion de-
creases exponentially according to expðDq2 τÞ, while the field, due to contribu-
tions of static inhomogeneities, is constant over time. Since the measured intensity
corresponds to the product of two scattering fields, two cases can occur for the in-
tensity correlation function:
     
gð2Þ ðq; τÞ  1 ⁓ exp Dq2 τ · exp Dq2 τ ¼ exp 2Dq2 τ ðhomodyneÞ (6:62)
   
gð2Þ ðq; τÞ  1 ⁓ 1 · exp Dq2 τ ¼ exp Dq2 τ ðheterodyneÞ (6:63)

This means that in the case of plain heterodyne (HD) scattering, the field correlation
is measured directly (cf. eqs. (6.61) and (6.63)).
6.4 Light scattering 273

As shown in Figure 105, the constant scattering contribution varies for each
measurement position; the system is inhomogeneous and non-ergodic. This means
that measurements at many different positions p are required to fully characterize a
gel by dynamic light scattering.
In 1973, Tanaka, Hocker, and Benedek (THB) were the first providing the physi-
cal principles for the interpretation of experimental results obtained from a poly-
acrylamide hydrogel probed by dynamic light scattering. The THB theory treats a
gel as a viscoelastic polymer network of mesh sizes from 1 to 100 nm dispersed in a
continuum of solvent. The hydrodynamic coupling of density fluctuations from sol-
vent and polymer segments located between two junctions allows to distinguish
longitudinal from transversal modes of fluctuations. With the assumption of a ran-
dom distribution of cross-linking points, the scattering is approximately homodyne.
According to the viscoelastic properties of a gel, the following formalism can be
found for the collective diffusion coefficient:
4
D ¼ Kos þ G=ζ (6:64)
3

where Kos is the osmotic bulk modulus, G is the shear modulus, and ζ is the poly-
mer–solvent friction coefficient. D is further related to the density correlation func-
tion according to
ð

1 kT
D¼ gðrÞ dr (6:65)
3N 6πξ h
0

with N being the degree of polymerization of the network strands. With the spatial
correlation function for semi-dilute polymer solutions given by

ξh r
g ðr Þ ¼ exp  (6:66)
r ξh

we obtain a Stokes–Einstein analog equation for the hydrodynamic correlation


length

kT
ξh ¼ (6:67)
6πηsolvent D

A different approach by Joosten considers the time-averaged intensity fluctuation at


a certain measuring point p and takes into account the partial HD scattering by use
of the following equation for the intensity correlation function:
 2 
ð2Þ ð1Þ   h ð1Þ i
gHD;p ðq; τÞ ¼ 1 þ Xp2 gF ðq; τÞ þ 2Xp 1  Xp gF ðq; τÞ (6:68)

homodyne heterodyne
274 6 Scattering analysis of polymer systems

The homodyne scattering ratio Xp as a scaling factor between both terms is defined
as the ratio between the time-averaged fluid scattering intensity hIF it and the time-
averaged total intensity hIit;p at the position p:

h IF i t
Xp ¼ (6:69)
hI it;p

In the case of plain homodyne scattering, we obtain Xp ¼ 1. With increasing hetero-


dyne (HD) scattering contribution, the total intensity hIit;p increases and Xp gets
smaller. Also, the amplitude of the intensity correlation σ2 gets smaller with de-
creasing homodyne scattering ratio:
ð2Þ  
σ2 ¼ gHD;p ð0Þ  1 ¼ Xp 2  Xp (6:70)

While the intensity correlation function varies from position to position, the fluid cor-
ð1Þ
relation function gF ðq; τÞ can be assumed as a simple position-independent expo-
nential function, since it now only captures the dynamic properties of the network:
ð1Þ  
gF ðq; τÞ ¼ exp DHD q2 τ (6:71)

Shibayama took up the approach of Joosten by assuming only homodyne scattering


ð2Þ
for the intensity correlation gHD;p ðq; τÞ and taking into account the Siegert relation,
eq. (6.61), which is strictly not valid, to determine an apparent diffusion coefficient
for different measurement positions Dapp;p . By comparison to the approach of Joos-
ten, a relation between the apparent diffusion coefficient Dapp;p and the diffusion
coefficient DHD concerning partial heterodyning can be established:
 
DHD ¼ 2  Xp Dapp;p (6:72)

With eq. (6.69), we obtain

hIit;p 2 hIF it;p


¼ · hIit;p  (6:73)
Dapp;p DHD DHD

Linearized plotting and fitting of a dataset according to eq. (eq. 6.73), which is then
based on simply experimentally accessible quantities only, enables one to deter-
mine the diffusion coefficient DHD and the fluid scattering intensity hI iFt;p . With the
help of eq. (6.67), the dynamic correlation length ξ h can be determined from the
diffusion coefficient DHD , which in turn serves as a measure of the mesh size, that
is, the spatial distance between two cross-linking points in a polymer network gel.
Questions to Lesson Unit 19 275

Toyoichi Tanaka (Figure ) was born in Nagaoka


city, Niigata Prefecture, Japan, in . He received
his BSc (), MSc (), and doctorate degree
() in physics from the University of Tokyo. Ta-
naka joined the physics faculty at M.I.T., Cam-
bridge, MA, in , where he applied the method
of dynamic light scattering to study biomedical
phenomena. At the same time, he also discovered
and detailed upon the volume phase transition of
environmentally responsive polymer gels. In the
s, Tanaka further evolved that concept to cre-
ate “smart gels” that mimic functions of proteins.
His idea was to create gels from a mixture of dif-
ferent monomers (playing the role of amino acids)
with sequences designed through a process of Figure 108: Portrait of Toyoichi Tanaka. Image
“imprinting.” In the middle of that work, Tanaka by Len Irish.
unexpectedly died of heart failure on May ,
, while playing tennis at age .

Questions to Lesson Unit 19

(1) What holds true when comparing the scattering intensity of a polymer gel and
a polymer solution of the same polymer and same concentration?
a. The scattering intensity of a polymer solution is higher than that of a poly-
mer gel, because the light can transmit easier between the polymer coils,
allowing multiple scattering to occur.
b. The scattering intensity of a polymer solution is higher than that of a poly-
mer gel, because the polymer gel absorbs parts of the light that can no lon-
ger be scattered, leading to a diminished scattering intensity.
c. The scattering intensity of a polymer gel is higher than that of a polymer solu-
tion, because on top of the temporal concentration fluctuations there are perma-
nent fluctuations due to spatial inhomogeneities in the cross-linking density.
d. The scattering intensity of a polymer gel is higher than that of a polymer
solution, because the density of the scattering centers in general is higher,
leading to a higher probability for an incident photon to be scattered.

(2) What are the consequences of the different contributions to the scattering in-
tensity for the Rayleigh ratio in a polymer gel?
a. The Rayleigh ratio of the gel is the sum of the Rayleigh ratios of the different
scattering contributions.
b. The Rayleigh ratio of the gel is the difference of the Rayleigh ratios of the
different scattering contributions.
276 6 Scattering analysis of polymer systems

c. The Rayleigh ratio of the gel is the mean of the Rayleigh ratios of the differ-
ent scattering contributions.
d. The Rayleigh ratio of the gel is the product of the Rayleigh ratios of the dif-
ferent scattering contributions.

(3) What is ergodicity?


a. When a process is aperiodic, it is referred to as ergodic in statistics.
b. Ergodicity is fulfilled when a system shows the same averages over time
and ensemble.
c. A system with correlation decay, that is, loss of its memory after a specific
amount of time, is called an ergodic system.
d. Ergodicity is the characteristic of a system in which every parameter at a
given time t is determined by the parameters that were present a timestep
earlier, at t – τ.

(4) Which of the following is an example for a non-ergodic system?


a. A microparticle in a liquid
b. An ideal polymer coil
c. An ideal gas
d. A polymer gel

(5) What holds for the Rayleigh ratio of excess scattering?


a. It depends on the time and the position, which is why averaging over time
and positions is necessary.
b. It depends on the time and the position, which is why averaging over time
and positions as well as normalizing to a minimum value is necessary.
c. It depends on the time and the position, which is why averaging over time
and positions as well as correcting by a minimum value is necessary.
d. It depends on the time and the position, which is why averaging over either
time or positions is necessary, but only possible if the system is ergodic.

(6) Because it is so complex, the structure factor of a polymer gel cannot be solved
analytically. Therefore, there are multiple theoretical approaches which have
been derived from empirical data.
Which of the following approaches is not suitable to theoretically model a poly-
mer gel?
a. A system of two phases with clear boundaries
b. A given ensemble, whose molecular dynamics is simulated by the given
potentials
c. Variable densities, which are present in randomly distributed domains
d. Considering an overall main species in which another species is distributed
Questions to Lesson Unit 19 277

(7) How can you derive the intensity correlation from the field correlation function?
a. gð2Þ ðq; τÞ ¼ gð1Þ ðq; τÞ2 þ 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
b. gð2Þ ðq; τÞ ¼ gð1Þ ðq; τÞ þ 1
c. gð2Þ ðq; τÞ ¼ gð1Þ ðq; τÞ2  1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
d. gð2Þ ðq; τÞ ¼ gð1Þ ðq; τÞ  1

(8) What causes a heterodyne scattering contribution in polymer gels?


a. Static light scattering
b. Dynamic light scattering
c. Static inhomogeneities
d. Dynamic inhomogeneities
7 States of polymer systems

LESSON 20: DILUTE POLYMER SOLUTIONS


Polymer solutions are a prime state of appearance during many polymer syntheses and in
almost all polymer-analytical investigations; also, many biopolymers find themselves in a
solution state in natural environments. This lesson unit treats the most simple concentration
regime of such solutions: the dilute regime. In that state, the polymer volume fraction is so
low that the chains do not touch or overlap. Hence, the solution is alike a colloidal suspen-
sion of solvent-filled nanogel particles. This consideration allows the viscosity of such dilute
solutions to be appraised conceptually, and as it turns out, that delivers a basis to deter-
mine the size and shape, and with that the molar mass and solution state of polymers by the
simple yet precise experimental method of dilute-solution viscometry.

7.1 Polymer solutions

So far, in the course of this textbook, we have learned about the fundamentals of
polymer structures, dynamics, and properties, the latter particularly focused on me-
chanical features. In the following, we go ahead and deepen our knowledge by
shedding a specific view to the characteristic and application-relevant states of ap-
pearance of polymer systems, discussing their structure, dynamics, and (mechani-
cal) properties separately and specifically. We start by taking a closer look at
polymer solutions, before the following subchapters will do the same for networks
and gels and for polymers in glassy and crystalline solid states. The melt state, by
contrast, is not considered specifically, as we have in fact said quite everything that
is relevant for melts already in our preceding chapters, in a sense that in this state,
the coils display ideal conformations, have Rouse-type dynamics, and exhibit visco-
elastic properties as treated in our large Chapter 5.

7.1.1 Formation

In Chapter 4, we have learned that the dissolution of a polymer in a solvent has just a
little entropy contribution, and hence, the process of dissolution is mostly governed
by the change in enthalpy, ΔH, that comes along with the establishment of mono-
mer–solvent (M–S) contacts in a polymer solution replacing former monomer–mono-
mer (M–M) contacts in the solid polymer body and solvent–solvent (S–S) contacts in
the former plain liquid medium. From classical physical chemistry, we also know that
a dissolution process is generally an equilibrium with a back-and-forth “reaction,” in
a sense that each molecule in a solute species is subject to a solution–dissolution

https://doi.org/10.1515/9783110713268-007
280 7 States of polymer systems

equilibrium. However, in a polymer, these molecules are connected to each other as


segments in a chain, and so this dynamic equilibrium has to be on the dissolved side
for all these segments until a whole chain can get dissolved. On top of that, the chains
may be entangled with one another in the solid polymer specimen; in this case, it is
not even enough to bring just one chain fully into the dissolved state for it to separate
from the solid body, but other chains that are entangled with it must get dissolved,
too. Therefore, dissolution of a solid polymer specimen usually takes a long time,
often hours or days, sometimes even weeks! During such a long dissolution process,
the small and therefore highly mobile solvent molecules will first penetrate into the
(usually largely amorphous) polymer solid and fill the free volume between the indi-
vidual chains, causing the material to swell. Hence, we usually observe swelling of
the solid polymer body before it eventually dissolves. If the solvent is not a good one,
the dissolution equilibrium of the monomer segments is less favorable, such that it
will never be on the dissolved side for all monomers at a time. If that is the case,
then only swelling occurs, but no free dissolution of the polymer chains is achieved.

Practical note: In contrast to small-molecule solutions, sonication of a polymer solution is


not a recommended method to speed up dissolution, as it will break the chemical bonds of
the polymer chains! In fact, if you sonicate a mixture of a solid polymer specimen and a sol-
vent, dissolution is indeed apparently fast, but this is because the sonication breaks the
chains, and so their fragments dissolve quickly. The fragmentation of the chains is easily rec-
ognized by the observation that the final solution has a low viscosity, whereas that of a solu-
tion made by patient waiting rather than sonication is usually high. This is because η ⁓ M3, so
a low M of fragmented short chains causes a much lower η than a high M of intact long chains.

7.1.2 Concentration regimes

A polymer solution can be prepared in different characteristic concentration re-


gimes, as schematically shown in Figure 109. In a dilute solution, each coil is dis-
solved and surrounded by solvent only but has no contact to other coils. In other
words: the solution is so dilute that all the coils are far apart from each other.88 If
we increase the polymer concentration, we add more coils to the given fixed volume
of our system, such that the spacing between the individual coils shrinks. At the
overlap concentration, c*, the polymer coils first start to have physical contact
with one another. From this point on, they stop acting as single entities, but in-
stead, they interpenetrate. This is possible because each coil is a loose fuzzy object

 Nevertheless, remember what we have learned about the Gaussian coil: the segmental density
inside each single coil is comparably high, corresponding to a local molar concentration of some
tens of moles per liter (depending on the degree of coil expansion, and with that, on the solvent
quality). In a dilute solution, however, the segmental density between the coils is zero, as they are
well separated by solvent.
7.1 Polymer solutions 281

with a lot of free volume inside89 that can be occupied by either solvent (as it is in a
dilute solution) or by segments of other coils (as it starts to be the case beyond the
overlap concentration). In such a state, we speak about a semi-dilute solution.
This name refers to the duality of this concentration regime: on the one hand, the
concentration is still comparably low (only a few wt%), so that the solution is still
kind of “dilute,” but on the other hand, we have a situation qualitatively different
from the true dilute one (where coils are isolated from one another). If we further
increase the concentration by adding more polymer, this must somehow be accom-
modated in the given volume of the system. The only way how this is achievable is
by more interpenetration of the coils. By that, the excluded-volume interactions in-
side each coil become progressively screened by overlapping segments of other
coils, such that effectively, the good-solvent situation that each coil has found itself
in at the dilute concentration turns into a less good one. As a result, all the coils
shrink. At a special point, they have shrunken down to their unperturbed size as
they would have in a state without any excluded-volume interactions, that is, in the
θ-state. From this point on, our solution has turned into a melt-like fluid. We then
speak about a concentrated solution.
The overlap concentration, c* or ϕ*, is reached when the volume fraction of the
chain segments in solution is equal to the volume fraction of chain segments in
each single coil, which we appraise as the volume per segment, l3, times the num-
ber of segments, N, over the coil volume, R3:

Nl3
ϕ* = ⁓ N 1 − 3ν (7:1)
R3
We can readily see that the overlap concentration decreases with increasing degree of
polymerization N. In other words, long chains (i.e., larger coils according to R ⁓ Nν)
overlap at lower concentrations already.
Often, an alternate expression is used for the overlap concentration:90
  3M
c* g · L − 1 = (7:2)
4πNA Rg 3

It calculates c* in the very common unit of g·L−1.


Let us quantify c* for an exemplary polymer compound: polystyrene with a molar
mass of M = 1,000,000 g·mol−1. In a good solvent, its c* is only 3.6 g·L−1, or 0.36 wt%,
whereas in a Θ-solvent, its c* is 14.7 g·L−1, or 1.47 wt%. The first value is quite low,

 Again, remember that the ratio of the coil volume to the volume of the actual polymer segmen-
tal mass in it is about 10,000:1!
 There is actually a whole lot of different expressions for the overlap concentration, as summa-
rized by Ying and Chu in Macromolecules 1987, 20(2), 362–366.
282 7 States of polymer systems

Figure 109: Schematic representation of the concentration regimes of a polymer solution. A key
parameter is the overlap concentration, c*, at which the polymer chains start to have physical
contact with one another. The concentration regime below c* is called the dilute regime, whereas
the one above c* is called the semi-dilute regime. Provided that the polymer chains are sufficiently
long, another parameter, the entanglement concentration, ce*, becomes relevant. At ce*, the
polymer chains are so overlapped that they start to form mutual entanglements. The semi-dilute
regime is therefore subdivided into a non-entangled and an entangled one. Commonly, ce* is the
twofold to tenfold of c*. At very high concentrations, the solution resembles a polymer melt, albeit
residual solvent is still present. This concentration regime is called the concentrated regime.

illustrating how much the polymer coil expands in a good solution. Thus, even a low
amount of polymer already creates an overlapping polymer solution.

7.1.3 Dilute solutions

7.1.3.1 Structure
The dilute solution regime is the easiest concentration regime to be captured con-
ceptually, as the individual polymer coils are not in contact with one another in
that regime. The overall coil conformation depends on the solvent quality according
to the general scaling law R ⁓ N ν . Note, however, that on scales below the thermal
blob size, a concept introduced in Section 2.6.2, there might be a potential change
to ideal statistics. Up to this specific lengthscale ξT, the most relevant energy is the
thermal energy kBT, outweighing any other energy such as the excluded-volume in-
teraction. As a result, we can separate two lengthscales: at scales smaller than the
thermal blob size, r < ξT, the excluded-volume interaction energy is smaller than
kBT, and the polymer chain segments on these small scales have ideal conformation
according to a three-dimensional random walk, whereas at scales larger than the
thermal blob size, r > ξT, the excluded-volume interaction energy is larger than kBT,
and the polymer chain segments on these scales have real coil conformation ac-
cording to a three-dimensional self-avoiding walk. The structure, thus, is dependent
on the lengthscale of observation.
7.1 Polymer solutions 283

7.1.3.2 Dynamics
In the dilute regime, the Zimm model that we have encountered in Section 3.6.3 best
describes the polymer chain dynamics. Here, M–S interactions cause moving polymer
chain segments to drag solvent molecules with them. These, in turn, do the same to
the solvent molecules adjacent to them. This drag spreads from one solvent molecule
to another until, eventually, it reaches other segments of the same polymer chain.
Hence, even distant segments couple to each other by these hydrodynamic interac-
tions through space. As a result, each coil drags the solvent in its pervaded volume
with it. The trapped solvent is in diffusive exchange with the surrounding solvent
molecules, but it does not drain through the coil. As a result, the coils appear like
solvent-filled nanogel particles, as visualized in Figure 110. Given that shape, it is the
polymer coil size and degree of expansion in the solvent medium that makes the
main effect of the solution viscosity. Hence, in turn, experimental estimation of di-
lute-solution viscosities can actually serve as a means to gain information about the
size and shape of the coils, that is, on their average molar mass and on the extent of
swelling of the coils, which further corresponds to the solvent quality. Based on this
premise, the method of solution viscometry is a precise and easy means to character-
ize polymers, with outcome information similar to those obtained from osmometry,
namely, a molar-mass information and a solvent-quality information.

Figure 110: Polymer coils entrapping solvent within


their pervaded volume and dragging it with them
on their motion due to hydrodynamic interactions.
This portion of solvent is in diffusive exchange
with the surrounding solvent molecules, but the
coils are not drained by the medium on their
motion through it. Picture inspired by B. Vollmert:
Grundriss der Makromolekularen Chemie,
Springer, 1962.

The viscosity of a dilute polymer solution can be assessed in multiple variant forms,
all tracing back to the relative viscosity:

ηsolution
ηrel = (7:3)
ηsolvent

It quantifies the relative extent of viscosity elevation of a solvent due to the pres-
ence of polymer coils in it.
284 7 States of polymer systems

A variant form is the specific viscosity:


ηsolution η − ηsolvent ηpolymer
ηsp = ηrel − 1 = − 1 = solution = (7:4)
ηsolvent ηsolvent ηsolvent

It “isolates out” only the effect of the polymer to the relative viscosity elevation.
We may guess already that the extent of viscosity elevation of a fluid due to the
presence of a polymer in it will certainly be concentration-dependent. Whenever we
have to do with such dependencies, it is helpful to introduce so-called reduced
quantities, in which we normalize to concentration; we know such a quantity al-
ready from Section 4.3, where we introduced and discussed the reduced osmotic
pressure. Following that line of thought, we introduce the reduced viscosity:
ηsp
ηred = (7:5)
c

As even small concentrations of polymer will already have a drastic effect on the
viscosity, which is due to the large size of the polymer coils, a universal measure is
introduced by extrapolation of the latter quantity to zero concentration; this limit is
named limiting viscosity number91 or Staudinger index:
1 ηsolution − ηsolvent
½η = lim ηred = lim (7:6)
c!0 c!0 c ηsolvent

Let us get back to the concentration dependence we have touched upon already.
For ηred, this is given by a formula introduced by Huggins:

ηred = ½η + kHuggins ½η2 c (7:7)

We may rearrange that to:


 
ηsolution = ηsolvent 1 + ½ηc + kHuggins ½η2 c2 (7:8)

This is a form similar to the virial-series expansion of the osmotic pressure, in


which the first virial coefficient has to do with the polymer molar mass, while the

 Actually, we should speak about a number only when a quantity has no physical unit. All the
viscosity values we have introduced above have no unit, as they are all based on the relative viscos-
ity introduced in eq. (7.3), in which the same unit appears in the numerator and the denominator
and therefore cancels out. This means that the concentration in the denominator of eq. (7.5) must
have no unit as well to give a dimensionless quantity [η]; this is only fulfilled if we use a dimension-
less measure of concentration, such as the volume fraction. If we use a measure for the concentra-
tion that has a unit, though, such as the mass-per-volume concentration that is quite common in
polymer science, then [η] has a unit that is the inverse of that (i.e., L g–1 or so). This is fine, but
strictly speaking, we should not name it limiting viscosity number then, as it is not a number in
that case.
7.1 Polymer solutions 285

second virial coefficient accounts for polymer–solvent interactions. The first role is
taken by [η] in the latter equation, whereas kHuggins plays the second role.92
To understand the relation between [η] and the polymer molar mass, we have
to look into the field of colloid science.93 In that field, the relative viscosity of a sus-
pension of hard colloidal spheres (i.e., noninteracting spherical particles with sizes
in the range of some tens to hundreds of nanometers, large enough to no longer
show molecular characteristics, but small enough to still show no gravitational sed-
imentation) at a volume fraction ϕ is given by Einstein’s law:94

ηrel = 1 + 5=2ϕ (7:9)

Application of that relation to calculate the limiting viscosity in the form of


 .    
½η = lim ηred = lim ηsp ϕ = lim ð1=ϕÞ ηrel − 1 yields a number value of 5/2 for
ϕ!0 ϕ!0 ϕ!0
spherical colloids.95 Other geometries like ellipsoids, rods, platelets, or whatever
will lead to different numbers. With that, we see that [η] is a quantity that has to
do with the shape of the solute colloids. We may transfer this to polymers by the
following line of thought:
First, even with the simple Einstein law, we might actually be able to distinguish
between monomeric and dimeric species in the solution. If we assume the monomeric
species to be spherically shaped, it will deliver a value of 5/2 for [η], whereas if we
assume the dimeric species to be dumbbell-shaped with an aspect ratio of 2:1, it will
deliver a value of 3 for [η], and thereby raise the viscosity more than spheres at same
volume fractions would do. This trend persists: trimeric dumbbells with an aspect

 Strictly speaking, the parameter [η] already includes information on polymer–solvent interac-
tions, as it is actually a factor that reflects the size and shape but not strictly the molar mass of the
solute polymer, which of course depends on the solvent quality and therefore on polymer–solvent
interactions. This actually explains why that factor is also present in the second virial-series
term of eq. (7.8). Here, however, that parameter alone would be “a bit too much,” such that it
gets “attenuated” by the additional factor kHuggins in the second virial-series term. This factor
has numerical values of typically kHuggins = 0.38 in a good solvent, kHuggins = 0.5–0.8 in a theta-
solvent, and kHuggins = 1–1.3 in a bad solvent.
 The field of colloid science is actually a good neighboring discipline to polymer science, as pol-
ymers are in fact a (very versatile and interesting) type of colloidal soft matter. Before Staudinger’s
notion about their macromolecular chain-like architecture, they were in fact believed to be colloidal
clusters of small molecules. With Staudinger’s insight, the fields of colloid and polymer science
separated, but with the merger of supramolecular and polymer chemistry in the most recent up to
present times, also the fields of polymer and colloid science reapproached each other.
 Note: that law is only valid for small volume fractions. This gets clear to us if we think about
the limiting case of a dense-packed suspension; this has a viscosity of infinity, that is, much higher
than what the Einstein law predicts.
 Here we get a number, as we use the dimensionless volume fraction as the measure of concen-
tration; we may then rightfully speak about the limiting viscosity number in that case.
286 7 States of polymer systems

ratio of 3:1 would lead to a value of 3.6 for [η]. From that we see clearly that the limit-
ing viscosity is a shape parameter.
Second, we now assume a solution of polymer coils to be a simple limiting case
by presuming that the polymer coils are spheres. In that case, the volume fraction
of the coils in the solution can be appraised as

ϕH ⁓ n · rH 3 (7:10)

where n is the number of coils and rH is the hydrodynamic radius of each coil,96
which makes rH3 to be a kind of measure of the coil volume, and hence, n·rH3 to be
a measure of the volume of all coils in the sample. Division of that volume by the
whole sample volume gives the polymer coil volume fraction in the sample, as ex-
pressed in a simplified proportionality form in eq. (7.10). This volume fraction is de-
noted with an index H in eq. (7.10) to strike out that it is a hydrodynamic volume
fraction in a sense that it is not equal to just the volume fraction of chain segments
inside each coil, but instead, also accounts for potentially additional hydrodynami-
cally trapped solvent in the coil.97 This is indeed very relevant in dilute solutions,
such that ϕH in eq. (7.10) does not simply denote the polymer volume fraction but
the polymer coil volume fraction, which is composed of the chain material itself
plus the entrapped solvent.
In a hypothetic case where the polymer coils are fully collapsed to solid spheres
without any entrapped solvent, we have a poor-solvent scaling of

rH ⁓ Rg ⁓ hr2 i1=2 ⁓ N 1=3 (7:11)

At a given polymer volume fraction and therefore also a given polymer mass-per-
volume concentration, the number of coils n is inversely proportional to the size
and therewith also the weight per coil, which both scales inversely to its degree of
polymerization N. With that, we get:
 3
ηrel ⁓ ϕ ⁓ n · rH 3 ⁓ N − 1 · N 1=3 ⁓ N 0 (7:12)

This is exactly what the Einstein law expresses: the relative viscosity of a suspen-
sion of spheres is in fact proportional to the sphere volume fraction, but not explic-
itly to the size of the spheres!
At the more common condition of Gaussian coils with some solvent inside, we
have a theta-state scaling of

 We know that quantity from Section 3.6.1: it quantifies the radius of a hypothetic ideal sphere
with the same diffusive properties as our object of interest, here: as our polymer coils.
 This is the scenario treated in the Zimm model of polymer dynamics.
7.1 Polymer solutions 287

rH ⁓ Rg ⁓ hr2 i1=2 ⁓ N 1=2 (7:13)

and hence we get


 3
ηrel ⁓ ϕ ⁓ n · rH 3 ⁓ N − 1 · N 1=2 ⁓ N 1=2 (7:14)

In the case of a good solvent, the coils are expanded and carry a lot of solvent with
them on their motion. Such coils exhibit good-solvent scaling according to

rH ⁓ Rg ⁓ hr2 i1=2 ⁓ N 3=5 (7:15)

such that we get


 3
ηrel ⁓ ϕ ⁓ n · rH 3 ⁓ N − 1 · N 3=5 ⁓ N 4=5 (7:16)

A similar general form of a relation has also been found experimentally by Mark,
Houwink, and Sakurada:

½η = K · Mη a (7:17)

in which Mη is the so-called viscosity-average molar mass


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P ffi
a Ni Mi a + 1
Mη = P i
(7:18)
i Ni Mi

It commonly lies between the number average and the weight average.98
K is a constant of proportionality, and a is a scaling exponent that has to do
with the extent of swelling of the polymer coils in the solvent of interest. Compari-
son of eq. (7.17) to eqs. (7.6) and (7.4), along with eqs. (7.12), (7.14), (7.16) indicates

a = 3ν − 1 (7:19)

whereby ν is the Flory exponent from Section 3.4.


The Mark–Houwink–Sakurada equation99 is therefore an important basis for
experimental determination of the polymer molar mass, hereby obtained as the vis-
cosity average. Such kind of estimation requires knowledge of the values of K and a

 In a good solvent, Mη is closer to Mw than to Mn, which you see by plugging in the good-solvent
value of a = 0.8 (see eq. (7.16)), which turns eq. (7.18) into a form close to the definition of the
weight-average molar mass Mw in Chapter 1.
 This equation is encountered with numerous names in the literature by featuring all the three
we name here or only two of them (most often Mark–Houwink only), or by additionally or alterna-
tively featuring the names Staudinger and/or Kuhn.
288 7 States of polymer systems

for the given polymer–solvent pair, which are tabulated for many different poly-
mer–solvent mixtures.

Herman Franz (Francis) Mark (Figure ) was


born on May , , in Vienna. Already at
the age of , he frequently visited a scientific
laboratory at the University of Vienna. Mark
started to study chemistry during World War I
and obtained his PhD in  and then became
an assistant of Wilhelm Schlenk at the University
of Berlin in the same year. Just one year later,
Mark was invited by Fritz Haber to work at the
newly established Kaiser Wilhelm Institute of
fiber chemistry in Berlin. In , he accepted an
offer from I.G. Farben, which later became BASF
at Ludwigshafen, to become a vice director of
research; his scientific focus there was on x-ray
analytics of fiber-forming polymers. In those
times, Mark still opposed Staudinger’s theory of Figure 111: Portrait of Herman F. Mark. Image
chain-like macromolecules, and not before  reprinted with permission from Chemical
did Mark start to accept that insight. Apart from Heritage Foundation/Science Photo Library.
fundamental studies on rubber elasticity together
with Guth, Mark explored the commercialization
of polystyrene, polyvinyl chloride, and polyvinyl
alcohol. After fascism took power in Germany,
Mark moved to Vienna and later to New York,
where he established the Institute of Polymer
Research at Polytechnic Institute, Brooklyn, in
, which he led until his retirement in .
Mark died on April , , in Austin, TX.

The experimental determination of η is based on a capillary viscometer, the most


famous one being the Ubbelohde viscometer, as sketched in Figure 112. In this ap-
paratus, we determine the time Δt that a given definite volume ΔV of a polymer so-
lution with density ρ needs to flow through a thin capillary of radius rc and length
lc, measured at different polymer concentrations including concentration zero, that
is, including the pure solvent. We may then convert these efflux times into the vis-
cosity using the Hagen–Poiseuille equation:

ρghπrc 4 Δ t
η= = const · ρΔ t (7:20)
8lc Δ V
The risk in this kind of evaluation is the high power of 4 that the capillary radius rc
has in eq. (7.20). Any imprecision in measuring that quantity (and any variation
that it may have over the length lc of the capillary) will therefore cause a strong
error. However, fortunately, our whole set of equations that leads to the limiting
viscosity [η] is based on a relative viscosity, which is simply determined as
7.1 Polymer solutions 289

ρsolution Δ tsolution Δ tsolution


ηrel = ≈ (7:21)
ρsolvent Δ tsolvent Δ tsolvent

So, any error in rc appears in both the estimates of the solutions and the estimates
of the pure solvent, and therefore cancels out.
From ηrel, we can calculate the reduced quantity ηred and then plot it as a func-
tion of concentration, which gives a Huggins plot according to eq. (7.8), as sketched
in Figure 113. From the slope, we can determine the Huggins constant, which as-
sesses polymer–solvent interactions, and from the intercept, we can determine the
limiting viscosity [η], which will then lead to the viscosity-average molar mass ac-
cording to the Mark–Houwink–Sakurada equation. In addition, [η] can also serve to
appraise the chain overlap concentration in the solvent considered, which is roughly
given as c* = [η]−1.

Figure 112: Schematic of an Ubbelohde viscometer to determine the quantities treated in this
section. A sample solution (typically about 20 mL is needed) is filled into the right tube. Then, the
left thin venting tube is closed (as can be done manually by placing a finger onto it), and part of the
fluid is pumped up into the upper reservoir by application of pressure to the right filling tube (as can
be exerted manually by hand bellows). Removing the pressure and reopening the venting tube
(by lifting off the finger from it, for example) then lets the fluid flow down the middle tube, through
the thin capillary. Measurement of the time it takes for the fluid meniscus to pass from the upper to
the lower light barrier gives the characteristic efflux time featured in eqs. (7.20) and (7.21).
290 7 States of polymer systems

Figure 113: Schematic of a Huggins plot to determine the limiting viscosity [η] and the Huggins
constant from a concentration-dependent series of estimates of the reduced viscosity ηred.

Questions to Lesson Unit 20

(1) Why is sonication not suitable to speed up the dissolution process of a polymer?
a. Sonication reinforces the movement of the chains, leading to more entan-
glement and therefore to a hindered dissolution process.
b. Sonication may break the bonds in the polymer chains, resulting in quickly
dissolving but shorter polymer chains.
c. Sonication leads to a forced, nonrandom distribution of the polymer chains
with an entropy loss, which disfavors the dissolution process.
d. Sonication speeds up the drainage of the solvent, which is trapped inside
the polymer coils, leading to deswelling/shrinking and therefore to a hin-
dered dissolution.

(2) What does the term “overlap” in the overlap concentration c* refer to?
a. The overlap of the dilute and semi-dilute regime
b. The overlap of chain segments of the same polymer chain
c. The overlap of chain segments of different polymer chains
d. The overlap of entire polymer chains

(3) The boundary between which concentration regimes does the overlap concen-
tration denote?
a. The dilute and the semi-dilute solution regime
b. The semi-dilute and the concentrated solution regime
c. The dilute and the concentrated solution regime
d. The concentrated and the saturated solution regime
Questions to Lesson Unit 20 291

(4) Which energy determines the size of a blob according to the blob concept?
a. The potential energy of the chain-internal interactions
b. The energy of the chemical bonds connecting the chain segments
c. The thermal energy kB T
d. The internal energy U

(5) The thermal blob size ______


a. marks the lengthscale that draws the line between an ideal coil conforma-
tion and a real coil conformation.
b. gives the size of the swollen polymer coils in thermal equilibrium.
c. is the upper critical size of a blob for which the Zimm model approach
needs to be changed to a Rouse model one.
d. describes the size of a conceptual blob in which the polymer chain moves
due to thermal excitation.

(6) Which assumption is not part of the Zimm model?


a. The coils can be considered as nanogel particles filled with solvent.
b. The moving chain segments drag solvent along with them.
c. The trapped solvent remains within the coil.
d. The hydrodynamic interactions allow interactions of chain segments that
are far apart from each other.

(7) Which properties can be determined via viscosity measurements?


a. The molar mass and the polymer size
b. The molar mass and the solvent quality
c. The polymer size and the excluded volume
d. The polymer size and the solvent quality

(8) Because of the very thin capillary, a measurement in a capillary viscometer is


very sensitive to errors in the capillary’s radius. Why do these errors neverthe-
less not affect the viscosity?
a. The derived viscosity is a relative quantity, meaning errors in the radii can-
cel out.
b. Since we perform a measurement series at different concentrations and de-
termine the viscosity by the slope of the linear fit of the data, the error,
which only leads to an offset of the data points, does not affect the slope.
c. The errors in radii are from a similar magnitude than the common measure-
ment errors in time. These errors compensate each other in the calculation
of the viscosity.
d. Deviations in the radius of the capillary are as much common as deviations
in the capillary’s length for manufactory reasons. In the viscosity calcula-
tion, these errors cancel out.
292 7 States of polymer systems

LESSON 21: SEMI-DILUTE POLYMER SOLUTIONS


Other than in the dilute regime, at concentrations beyond the onset of coil overlap, polymer
chains interpenetrate and perhaps even entangle with one another in a solution. This over-
lap drastically alters the structure and especially the dynamics of the coils in the solution,
heavily impacting the solution viscosity and chain diffusivity. A simple yet powerful ap-
proach to treat that with the aim of deriving equations for the chain length and concentration
dependence of these quantities is a scaling concept introduced by De Gennes.

7.1.4 Semi-dilute solutions

7.1.4.1 Peculiarity
Polymer solutions with concentrations above c* are somewhat peculiar, as their
structure and dynamics are dependent on the lengthscale under consideration. In
such a semi-dilute solution, the system can be viewed to be a space-filling array of
correlation blobs, which are conceptual spherical entities with a size of ξ, the cor-
relation length. An illustrative representation of ξ is shown in Figure 114. At dis-
tances smaller than ξ, a randomly chosen monomer unit on a chain is surrounded
only by solvent molecules or by other monomers from the same chain, whereas the
presence of monomer units on other chains is not noticed. On scales larger than ξ,
by contrast, overlapping segments of other chains can be “seen” by our chosen
monomer, and these other segments screen its interactions such as excluded-vol-
ume or hydrodynamic interactions.

7.1.4.2 Structure
Up to the correlation length, r < ξ, a chosen monomer can only ”see” solvent and
monomers on its same chain – precisely a number of g of them that sit on a chain
segment which occupies and defines the blob around the chosen monomer. The
conformation of that chain segment within the blob is that of a real chain with a
scaling of

ξ = gν · l (7:22)

Beyond the correlation length, r > ξ, the chain conformation is an ideal random
walk of correlation blobs, each of which with a size ξ and of which we have a num-
ber of N/g. That gives a scaling of
1=2
N
R=ξ · (7:23)
g

These correlation blobs are space filling, as is indicated in Figure 114. Hence, the
polymer volume fraction of a single blob, given by the ratio of the segmental
7.1 Polymer solutions 293

Figure 114: Illustrative representation of the correlation length in a semi-dilute solution of


overlapping polymer chains. In such a solution, the system can be viewed to be composed of a
space-filling array of correlation blobs with size ξ each. Inside each blob sits on segments of one
and the same chain only, whereas overlapping segments of other chains are present only on scales
larger than the correlation-blob size. As a result, at small distances, a selected monomer unit is
surrounded only by solvent or by other monomers from the same chain, whereas the presence of
monomer units from other chains is noticed not until scales larger than ξ are viewed. Picture redrawn
from to P. G. De Gennes: Scaling Concepts in Polymer Physics, Cornell University Press, 1979.

volume inside the blob, gl3, and the blob volume, ξ3, is equal to the polymer volume
fraction of the entire system:

gl3
ϕtotal system = ϕper blob ≈ (7:24)
ξ3

This is basically a blob-scale variant of the overlap concentration formulated in


eq. (7.1).
From the latter equation, along with some further scaling arguments detailed
in Rubinstein’s and Colby’s textbook on polymer physics (Section 5.3 in there), one
can derive a scaling-law dependence for the correlation length ξ as a function of the
volume fraction ϕ:
ν
ξ ⁓ ϕ− 3ν1 (7:25)

and for the dependence of the polymer coil size R on the volume fraction ϕ
ν0:5
R ⁓ ϕ− 3ν1 (7:26)

In a good solvent, with ν = 3/5, the latter scaling laws are ξ ⁓ ϕ − 3=4 and R ⁓ ϕ − 1=8 ,
whereas in a Θ-solvent, with ν = 1/2, they are ξ ⁓ ϕ − 1 and R ⁓ ϕ0 , as shown in Figure 115.
We see that both the correlation length ξ and the coil size R decrease upon in-
crease of the polymer volume fraction ϕ or concentration c beyond the overlap
294 7 States of polymer systems

threshold. This is reasonable for both parameters: as we add more polymer to a


semi-dilute solution, there is more coil interpenetration, such that we force a ran-
domly chosen monomer on a coil segment to see monomers from other segments at
smaller distances already, or in other words, we decrease the correlation length ξ. A
similar argument is valid for the polymer size R. The polymer coil is swollen in a
dilute solution of a good solvent. Above the point of overlap, however, the coil ex-
pansion is mutually hindered because the space for each coil to expand into is re-
stricted by the presence and overlap of other coils, and because coil-expanding
excluded-volume interactions are screened by overlapping segments of other coils.
Consequently, each coil shrinks in size. This shrinking goes on upon further in-
crease of the concentration, until it has proceeded down to the coil’s ideal Gaussian

Figure 115: Dependence of the correlation length ξ and polymer coil size R on the polymer volume
fraction ϕ, plotted in a log–log representation. Below the overlap concentration, ϕ < ϕ*, the chains
are separate entities with Flory radius R = RF that do not exhibit any concentration dependence, as
in this region, they do not yet realize each other’s presence. As soon as they start to overlap, at
ϕ = ϕ*, a randomly chosen monomer on a chain can for the first time see monomers from other
chains at distances greater than the correlation length, ξ, which is exactly the Flory radius RF at
that point, as coils just first touch each other here. At higher concentrations, ϕ > ϕ*, screening of
coil-expanding excluded-volume interactions sets in due to the overlap of interpenetrating
segments from other coils, such that the polymer coil is reduced in size according to eq. (7.26),
which gives a scaling of R ⁓ ϕ − 1=8 if ν = 3/5 (good solvent value); along with that, the more and
more pronounced coil interpenetration causes the length until which a monomer can see only
solvent or monomers from its same chain to decrease according to eq. (7.25), which gives a scaling
of ξ ⁓ ϕ − 3=4 if ν = 3/5 (good solvent value). In the concentrated regime, at ϕ > ϕ**, the coil has
shrunken down to its ideal dimensions with the ideal Gaussian radius of R0, and R does then no
longer display any further concentration dependence. The screening of excluded-volume interactions
inside each coil is then maximally operative, such that the solvent quality has turned from that of a
good solvent (with strong excluded-volume interactions) to that of a θ-solvent (with no excluded-
volume interactions); hence, the scaling of eq. (7.25) then turns out to be ξ ⁓ ϕ − 1 , with ν = 1/2
(θ-solvent value). The correlation length itself is then as small as the thermal blob size, ξT, below which
residual excluded-volume interactions are weaker than the ever-present thermal energy anyways.
7.1 Polymer solutions 295

dimensions. This point is called ϕ** or c**, and the regime above, ϕ > ϕ** or
c > c**, is called the concentrated regime. When we reach it, the coil size stays
constant at its ideal Gaussian value of R0 , and the correlation length has decreased
down to the thermal blob size, ξ = ξT, and then persists decreasing down to its abso-
lute lowermost limit, which is the monomer size (named l in Figure 115); at that
limit, we have reached the state of a polymer melt with a volume fraction of ϕ = 1.

Note: the first characteristic threshold ϕ* is usually just a few wt% in many polymer–solvent
systems, whereas the second threshold ϕ** is commonly very high. Hence, the largest portion
of the ϕ-scale in many systems actually corresponds to the semi-dilute regime. This is why we
spend so much emphasis on it here.

The scaling laws for R(ϕ) and ξ(ϕ) given earlier were alternatively and very cleverly
derived by Pierre-Gilles de Gennes as well, who realized that the power-law-type
nature of almost all relations in polymer science allows the implementation of the
following strategy: De Gennes first a priori assumed some kind of power-law depen-
dence to be operative, then included reliable boundary conditions for the upper
and lower limits of applicability of that power law, and finally solved for the numer-
ical value of the unknown power-law exponent. As it turned out, his method is as
accurate as simple.
Based on this premise, the derivation of R(ϕ) relies on the following argument:
we know that up to the overlap concentration, ϕ*, a polymer coil has a conforma-
tion as in a dilute solution according to RF ⁓ N ν . We now assume a power-law de-
pendence for concentrations higher than ϕ* according to R ≈ RF ðϕ=ϕ Þ , which we
x

can formulate because we know by definition that, at the onset of the overlap,
ϕ = ϕ*, the polymer coil has its expanded size, R = RF. Taking into account both
RF ⁓ N ν and ϕ ⁓ N 1 − 3ν ,
x
ϕ
R ≈ RF ⁓ N ν + 3νx − x ϕx (7:27)
ϕ

With the knowledge that the polymer coil in a semi-dilute solution is a random
walk of correlation blobs according to R ⁓ N 1=2 , we must have
1 0:5 − ν
ν + 3νx − x = ) x= (7:28)
2 3ν − 1

The result is the same scaling law that we have previously derived in eq. (7.26).
A similar derivation can be made for ξ(ϕ). Again, we assume a power-law de-
pendence to be operative, using the same boundary condition as we did above
(RF ⁓ N ν and ϕ ⁓ N 1 − 3ν )
ν
ξ ≈ ϕ− 3ν1 ⁓ N ν + 3νy − y ϕy (7:29)
296 7 States of polymer systems

We know that the correlation length is independent of the total chain length be-
yond the chain overlap threshold, so the numerical value of the power-law expo-
nent at the N must be zero. With that we get
−ν
ν + 3νy − y = 0 ) y = (7:30)
3ν − 1

Again, we arrive at the same expression we have derived previously in eq. (7.25).

7.1.4.3 Dynamics
To discuss the chain dynamics in semi-dilute solutions, we need to further subdi-
vide the concentration range into the unentangled (ϕ* < ϕ < ϕe*) and the en-
tangled (ϕ > ϕe*) semi-dilute domain.

Unentangled semi-dilute dynamics (ϕ* ≤ ϕ ≤ ϕe*)


Let us start our discussion with the less complex scenario of an unentangled semi-
dilute polymer solution; in that regime, we can use the same model(s) we have
used for dilute polymer solutions before. Remember that the Zimm model describes
a dilute solution with hydrodynamic interactions (see Section 3.6.3), whereas with-
out these interactions, the Rouse model is applicable. So well then, which of these
models shall we use now? This depends on the lengthscale of observation. On
lengthscales below the correlation length ξ, hydrodynamic interactions are opera-
tive, so the Zimm model is the appropriate choice. On lengthscales above ξ, by con-
trast, hydrodynamic interactions are screened by overlapping segments of other
polymer coils, so the Rouse model is correct here. For a quantitative treatment, we
again imagine a polymer chain in the semi-dilute concentration regime to be a se-
quence of correlation blobs with size ξ. Within each blob, excluded-volume interac-
tions lead to a scaling of ξ ⁓ gν , and hydrodynamic interactions are active as well.
This causes a Zimm-type relaxation of the chain segments of length inside ξ each
blob, described by

ηξ 3
τξ ≈ (7:31a)
kB T

As we know the relation between the correlation length ξ and the volume fraction ϕ
(eq. 6.25), we can reformulate eq. (7.31a) to
η − 3ν
τξ ⁓ ϕ 3ν1 (7:31b)
kB T

For the entire chain that consists of N=g blobs, by contrast, excluded-volume and
hydrodynamic interactions are screened. This leads to a Rouse-type relaxation of
7.1 Polymer solutions 297

the whole polymer chain, which we conceptualize to be a random-walk sequence of


correlation blobs that relaxes according to
1 + 2ν
N
τRouse = τξ (7:32a)
g

On scales above ξ we observe ideal statistics, and the Flory exponent is ν = 1/2 (as
just said: the whole chain can be seen as a random walk of blobs). This simplifies
the last expression to
2
N
τRouse = τξ (7:32b)
g

Just like in Section 5.8.2 (in the context of Figure 87), we can now plot the fre-
quency-dependent relaxation spectrum of a semi-dilute unentangled polymer solu-
tion, which consists of four regimes: (i) a Maxwell-type full relaxation regime on
timescales longer than τRouse , (ii) a Rouse-type leather-like domain with relaxation
of chain segments longer than ξ on times between τRouse and τξ , (iii) a Zimm-type
leather-like domain with relaxation of chain segments shorter than ξ on times
shorter than τξ , and (iv) a glassy domain with no relaxation at all on times shorter
than τ0 , as sketched in Figure 116.

Figure 116: Relaxation spectrum of an unentangled semi-dilute polymer solution. On times longer
than the Rouse time, τRouse, the chains can diffuse freely, such that the viscoelastic response is
described by the Maxwell model. The leather-like regime on timescales shorter than that (that is,
on frequencies higher than the inverse of τRouse) is subdivided into two parts. In the Rouse domain,
chain segmental motion takes place on scales larger than the correlation length ξ, such that
hydrodynamic as well as excluded-volume interactions are screened. This is followed by the Zimm
domain: here, chain segmental motion takes place on scales smaller than ξ, where hydrodynamic
and excluded-volume interactions are present.
298 7 States of polymer systems

We have already seen that the correlation blobs are space filling, as is indicated in
Figure 114. From this, we have concluded:

gl3
ϕtotal system = ϕper blob ≈ (7:24)
ξ3
ν
Together with eq. (7.25), specified to a form of ξ ≈ l ϕ − 3ν − 1 , we get a scaling-law de-
pendence of
1
g ⁓ ϕ− 3ν1 (7:33)

Taking eq. (7.32b) and replacing τξ by eq. (7.31b) as well as g by the latter eq. (7.33)
then yields
2 − 3ν
τRouse ⁓ N 2 ϕ 3ν1 (7:34)

We can specify this for different solvent qualities as follows: in a good solvent, with
ν = 3/5, the relaxation time τRouse scales with τRouse ⁓ ϕ1=4 , whereas in a Θ-solvent,
with ν = 1/2, it scales with τRouse ϕ~1 .
Another illustrative and practically measurable quantity that captures chain re-
laxation is the translational chain diffusion coefficient D. We estimate it based on a
simplified Einstein–Smoluchowski equation

R2
D≈ (7:35)
τRouse
Here we can insert what we have learned about the relations of the coil size and
relaxation time to its degree of polymerization and volume fraction, basically ex-
pressed by eqs. (7.26) and (7.34), to obtain
ν0:5 2ν1
R ⁓ N 1=2 ϕ − 3ν1 ) R2 ⁓ N 1 ϕ − 3ν1 (7:26)
2 − 3ν
τRouse ⁓ N 2 ϕ 3ν1 (7:34)

Plugging that into eq. (7.35) generates


1−ν
D ⁓ N −1 ϕ− 3ν1 (7:36)

We can specify this for different solvent qualities as follows: in a good solvent, with
ν = 3/5, we get D ⁓ N − 1 ϕ − 0:5 , whereas in a Θ-solvent, with ν = 1/2, we get D ⁓ N − 1 ϕ − 1 .
As a further practically relevant quantity, we estimate the viscosity of our unen-
tangled polymer solution. For this purpose, we can again use the scaling approach
according to De Gennes that we have already encountered when discussing the con-
centration dependence of the coil and correlation blob size in the preceding section:
7.1 Polymer solutions 299

x
ϕ
η = ηϕ = ϕ (7:37)
ϕ

Using eq. (7.1), we can specify to

η ⁓ ηϕ = ϕ N ð3ν1Þx ϕx (7:38)

Long-term relaxation modes are Rouse type with η ⁓ N 1 . Thus, we can solve for the
power-law exponent:
1
ð3ν − 1Þx = 1 ) x = (7:39)
3ν − 1

Again, we can determine scaling laws for different solvent qualities: in a good sol-
vent, the viscosity scales with η ⁓ N 1 ϕ2 , whereas in a Θ-solvent it scales with
η ⁓ N 1 ϕ5=4 .

Entangled semi-dilute dynamics (ϕ ≥ ϕe*)


The dynamics of an entangled semi-dilute polymer solution is similar to that of an
unentangled one on both long and short timescales, but on intermediate scales, a
further domain is added to the mechanical spectrum; this domain is dominated by
chain entanglements. By assuming a reptation-like chain motion in this domain, we
can estimate the concentration dependence of the reptation time, τRep, by a scaling
argument, again according to De Gennes. For this purpose, two boundary condi-
tions are introduced. First, for ϕ > ϕe*, we expect a scaling of τRep ⁓ N 3 . Second, for
ϕ < ϕ*, there is no reptation, and the longest chain relaxation time is the Zimm time
τZimm. We assume a power law for the range between these two domains and spec-
ify it through the use of eqs. (3.34) and (7.1):
m
ϕ
τRep = τZimm ⁓ τ0 N 3ν ϕm N − ð1 − 3νÞm ⁓ ϕm N 3νð1 − 3νÞm (7:40)
ϕ

Satisfying the first boundary condition lets us solve for the power-law exponent:
3 − 3ν
3ν − ð1 − 3νÞm = 3 ) m = (7:41)
3ν − 1

Thus, we can formulate a scaling law for the reptation time according to
3 − 3ν
τRep ⁓ N 3 ϕ 3ν1 (7:42)

Once again, just like above, we can plot the frequency-dependent relaxation spectrum
of a semi-dilute entangled polymer solution, which now consists of five regimes: (i) a
Maxwell-type full relaxation regime on timescales longer than τRep ; (ii) a plateau re-
gion on timescales between the reptation time τRep and the entanglement time τe , on
300 7 States of polymer systems

Figure 117: Relaxation time spectrum of an entangled semi-dilute polymer solution. As an addition
to the Rouse and Zimm domain that splits the leather-like regime in the case of semi-dilute
unentangled solutions, a plateau region appears when entanglements are present. These
entanglements dominate the mechanics of the system and lead to an entropy-elastic plateau that is
delimited by the reptation time, τRep, after which free relaxation and Maxwell-type flow can set in.

which entanglements impair the chain relaxation, thereby leading to a sample with
the ability for entropy-elastic energy storage. Beyond that, just like in the unentangled
solution case, we have (iii) a Rouse-type leather-like domain with relaxation of chain
segments longer than ξ on times between τe and τξ ; (iv) a Zimm-type leather-like do-
main with relaxation of chain segments shorter than ξ on times shorter than τξ , and
(v) a glassy domain with no relaxation at all on times shorter than τ0 , as sketched in
Figure 117.
Yet again, as an illustrative and meaningful microscopic quantity, we calculate
the translational chain diffusion coefficient as follows:

R2
D≈ (7:43)
τRep

We again take what we have learned about the size and time relations:
Size:
ν0:5 2ν1
R ⁓ N 1=2 ϕ − 3ν1 ) R2 ⁓ N 1 ϕ − 3ν1 (7:26)

Time:
3 − 3ν
τRep ⁓ N 3 ϕ 3ν1 (7:42)

and insert that into the expression for the diffusion coefficient to generate the fol-
lowing scaling law:
7.1 Polymer solutions 301

Figure 118: Set of scaling laws for the translational chain diffusion coefficient, D, as a function of
the chain degree of polymerization, N, and the polymer volume fraction in the system, ϕ.

ν2
D ⁓ N − 2 ϕ3ν1 (7:44)

When we specify that for different solvent qualities, we get D ⁓ N − 2 ϕ − 1:75 for a
good solvent and D ⁓ N − 2 ϕ − 3 for a θ-solvent.100
To sum up, the dynamics of a polymer solution depends on many factors, such
as the concentration regime and the lengthscale observed. We have derived a vari-
ety of scaling laws that all use the Flory exponent ν to take into account different
solvent qualities. Figure 118 lists all these scaling laws for the chain diffusion coeffi-
cient and gives the numerical values of the power-law exponents for each concen-
tration regime. We see that the scaling generally steepens for both parameters N
and ϕ as we go from one concentration regime to the next on the ϕ-axis.

 Note: the correct literature reference to the derivation of these scaling laws by De Gennes (that
we have recapitulated here) is Macromolecules 1976, 9(4), 587–593. Many experimental studies, in-
stead, refer to De Gennes’ book Scaling Concepts in Polymer Physics (Cornell University Press, 1979)
in that context. This is wrong, as this book actually only features a (rather carefully expressed) deri-
vation of a formula for the reptation time in athermal solvents, whereas expressions for the concen-
tration dependence of the diffusion coefficient are in fact not given in that book; by quite contrast,
the book rather states that “Reptation studies on solutions at variable c have been started recently
[. . .] – but precise exponents are not yet known.”
302 7 States of polymer systems

Questions to Lesson Unit 21

(1) What are correlation blobs?


a. Conceptual blobs that are in the size of the correlation length.
b. Conceptual blobs that are present for the timespan of the correlation time.
c. Conceptual blobs that move distances of their own size within the correla-
tion time.
d. Conceptual blobs for which the correlation function has a value of 1.

(2) What holds for the chain conformation on lengthscales bigger than the correla-
tion length ξ?
a. The conformation is that of a real chain of chain segments.
b. The conformation is that of a real chain of correlation blobs.
c. The conformation is that of an ideal chain of chain segments.
d. The conformation is that of an ideal chain of correlation blobs.

(3) The polymer volume fraction of a total sample is _____


a. the same as that of a single correlation blob, because within the correlation
blob the conformation of the chain is the same as for lengths bigger than
the correlation length.
b. the same as that of a single correlation blob, because the correlation blobs
are space filling.
c. the same as that of a single correlation blob, because the correlation blobs
are subvolumes of arbitrary sizes of the whole polymer solution.
d. not the same as that of a single correlation blob.

(4) What happens to the polymer coil size when the polymer volume fraction
increases?
a. The coil expands due to more screening of excluded-volume interactions.
b. The coil expands due to an increase in the correlation length upon increase
of the overall number of polymer chains in the solution
c. The coil shrinks due to mutual hindrance in the overlap with other coils.
d. The coil shrinks due to coil internal interactions that are enforced with in-
creasing polymer concentration.

(5) Which coil size is present at the concentrated regime?


a. Real coil
b. Random coil
c. Fully collapsed coil
d. Random coil that decreases to a fully collapsed state with increasing con-
centration.
Questions to Lesson Unit 21 303

(6) Choose the answer that assigns the relevant frequencies in the following relax-
ation spectrum of an unentangled semi-dilute polymer solution correctly.

−1
a. ① = τRouse , ② = τξ− 1 , ③ = τ0− 1
−1 −1
b. ① = τRouse , ② = τZimm , ③ = τ0− 1
c. ① = τ0 , ② = τRep , ③ = τξ− 1
−1 −1

d. ① = τ0− 1 , ② = τξ− 1 , ③ = τZimm


−1

(7) The frequency-dependent moduli of an unentangled and an entangled semi-di-


lute solution are ______
a. completely the same, just the values of the relevant timescales differ.
b. similar for shorter and longer timescales but not for intermediate ones.
c. similar for shorter timescales but different for longer ones.
d. completely different and therefore need different theoretical approaches.

(8) Comparing the overall chain diffusion coefficients, what holds for the scaling
of N and ϕ with increasing concentration and solvent quality?
a. The scaling of N and ϕ flattens.
b. The scaling of N steepens, whereas the scaling of ϕ flattens.
c. The scaling of N flattens, whereas the scaling of ϕ steepens.
d. The scaling of N and ϕ steepens.
304 7 States of polymer systems

LESSON 22: POLYMER NETWORKS AND GELS


A system of cross-linked polymer chains is named a network, and when immersed in (and
thereby swollen by) a solvent, it is named a gel. Networks and gels are used in a myriad of
applications, be it in vehicle tires, in hygiene products, or in medicine. This lesson unit
sheds a spotlight on how we can conceive the formation of networks from a statistical per-
spective in the framework of the percolation theory, and what the most relevant structural
and dynamic properties of networks and gels are, with a specific view to nanostructural inho-
mogeneities in them.

7.2 Polymer networks and gels

7.2.1 Fundamentals

This chapter investigates the ability of polymers to form network structures that
when being formed or immersed in a solvent medium turn out to be gels. These
materials have valuable properties in our everyday life, be it as hygiene products,
in food, or in the life sciences and medicine. In fact, life-forms themselves can be
viewed to be gels: whereas animals are protein-based hydrogels, plants are polysu-
gar-based hydrogels.101 It is therefore surprising to realize that this class of materi-
als is lacking a basic common definition. This dilemma was first recognized by the
British scientist Dorothy Jordan Lloyd in 1926, who stated that “the colloidal state,
the gel, is one which is easier to recognize than to define.” In 1946, Dutch scientist P.
H. Hermans defined a gel as “a coherent system of at least two components, which
exhibits mechanical properties characteristic of a solid, where both the dispersed
component and the dispersion medium extend themselves continuously throughout
the whole system.”102 It was famous American polymer scientist and Nobel laureate
Paul John Flory who eventually formulated a comprehensive definition of a gel in
1974,103 one that also comprises the above definitions. He delimited four classes of
gels:
1. Well-ordered lamellar structures (e.g., soap gels and clays)
2. Covalently cross-linked polymer networks filled with a swelling medium (i.e.,
polymer gels)
3. Fibrillar networks formed by dynamic arrest of stiff (bio)polymers (e.g., actin
fibers)
4. Particulate disordered structures (i.e., colloidal gels)

 Freely quoted from Vollmert: Grundriss der Makromolekularen Chemie, Springer, 1962.
 Colloid Science, Volume II: Reversible Systems, ed. H. R. Kruyt, Elsevier Publishing Company,
New York, 1949, pp. 483–494.
 Faraday Discuss. Chem. Soc., 1974, 57, 7–18.
7.2 Polymer networks and gels 305

A further scheme of classification introduced by Paul S. Russo distinguishes be-


tween “fishnet gels,” that is, cross-linked network structures, and “lattice gels,”
that is, all gel-type systems with a space-filling structure somehow different than a
network mesh-type one.104 In the following paragraphs, we take a closer look at the
fishnet type, corresponding to Flory’s second class of gels, made from polymeric
chains. It should be noted here that samples of such fishnet gels comprised of a
polymer network and a solvent almost always also incorporate unbound linear or
branched polymer chains. These are part of the sample, but by definition not part
of the gel.

7.2.2 Gelation

Gelation describes the process of forming a gel. During this process, monomers or
polymer chains bond to each other, thereby building up and branching larger and
larger molecules up to a specific point, the gel point. At this point, a first “infinite
molecule,” meaning a molecule that spans the entire system, appears, and we can
speak of a gel for the first time. Further gelation increases this infinite molecule
into an infinite cluster, the so-called gel fraction of the sample. Polymer chains
that are not (yet) connected to the network structure, by contrast, belong to the sol
fraction. These might also form non-system-spanning clusters, which might later
be connected to the gel and thereby eventually become incorporated into the gel
fraction. Gelation happens due to a multitude of reasons, but they all share one un-
derlying principle, which is the creation of cross-linking points in a polymer net-
work. The various ways of creating such points are listed in Figure 119. Based on
the nature of the junction, the broad area of polymer gels (of the fishnet type) can
be subdivided into the two classes of physical and chemical gels. In physical gels,
the cross-linking points are based on physical (often transient) interactions such as
ionic or van der Waals interactions between the chains or between certain func-
tional groups along their backbones, or local helical or crystalline domains in
which multiple chains (at least two) participate. This class can be further subdi-
vided into weak and strong physical gels, depending on the strength of that inter-
action, determining on what timescales the gel is stable. In chemical gels, by
contrast, the cross-linking process occurs by formation of chemical bonds be-
tween the chains. Chemical polymer gels can be produced via two different routes.
They can either be made from cross-linking reactions of monomers during polyad-
dition, polycondensation, or radial polymerization, or they can be built from al-
ready preformed polymer precursors that are end- or side-group functionalized

 P. S. Russo in ACS Symposium Series, Vol. 350, ed. P. S. Russo, ACS, Washington DC, 1987, pp.
1–21.
306 7 States of polymer systems

Figure 119: Classification and examples of modes of gel formation.

with binding motifs. Alternatively, precursor polymer chains can also be ran-
domly cross-linked along their backbones in a direct way even without specific
pre-functionalization by the action of suitable linkers, as done in the production
of rubbers via vulcanization.
The process of gel formation is a percolation phenomenon. This describes the
buildup of branched and interconnected structures of growing size, eventually
forming the aforementioned system-spanning molecule that turns the material into
a gel at the gel point, corresponding to the percolation threshold, pc.
While this might seem abstract at first glance, percolation phenomena are quite
generic, and we are familiar to the concept from our everyday life. For example,
percolation might happen during a deluge (that may be a result of severe global
overheating). First, before the process, we have a land mass that is interspersed
with separated individual bodies of water. Addition of water will eventually lead to
a reversed system of a continuous lake or ocean that contains some islands of land.
In this example, the percolation threshold is the point at which the first map-span-
ning lake is realized. In another example, a wildfire might or might not spread over
a whole forest, depending on the density and dryness of the trees. If both parame-
ters are too unfortunate (again perhaps as a result of global heating), the fire can
jump from one tree to another, and at a critical point, the percolation threshold, a
first aisle of fire spans over the whole forest. A third example is the one that the
whole world witnessed in the early twenty-twenties: the spreading of a virus. First,
there are local outbreaks that may actually still be confined locally, but if that
chance is missed, at a critical stage, the virus is everywhere, and no local means of
containment are effective anymore.
For polymer gels, percolation can be assessed by statistical modeling of the ge-
lation process based on a lattice model, very much alike the Flory–Huggins mean-
7.2 Polymer networks and gels 307

field theory of polymer thermodynamics. The mean-field approach for percolation


is called the Flory–Stockmayer theory, and we start examining it in detail with
the simplest case: one-dimensional bond percolation.

7.2.2.1 One-dimensional bond percolation


When limited to only one dimension, the only way that monomers or polymer pre-
cursors carrying complementary binding motifs A and B can connect to each other
is in a linear fashion, as shown in the upper sketch of Figure 120. To treat this pro-
cess statistically, we define the fraction of reacted A-groups as p. Consequently, the
fraction of unreacted A-groups is (1 – p). This directly corresponds to the number
density of molecules in the system, ntot(p). We can understand that by the following
thought: as long as no bonds have been formed, the number density of molecules is
unity, which indeed corresponds to ntot(p = 0) = (1 – 0) = 1. During gelation, each
newly formed connection between the monomers or precursor polymers reduces
the number of molecules in the system by 1. Because there is exactly one unreacted
A-group per resulting molecule, the number density of the resulting molecules is
identical to the fraction of unreacted A-groups, so we indeed get

ntot ð pÞ = 1 − p (7:45)

From that, the number-average degree of polymerization, Nn(p), can be determined


as
1 1
Nn = = (7:46)
ntot ð pÞ 1 − p

This value diverges at a percolation threshold of pc = 1. Thus, a one-dimensional


system can reach the percolation threshold by connecting all molecules in the sys-
tem to one single system-spanning macromolecule, but it cannot exceed it.

7.2.2.2 Two-dimensional bond percolation


To add another dimension to the percolation process, we need a multifunctional
molecule that has the possibility to connect to more than to just one other molecule.
Let us consider the condensation (or addition) of monomers with a single A-group
and (f – 1) complementary B-groups, ABf–1. An exemplary structure with f = 3 is
shown in the lower sketch of Figure 120.
In a statistical treatment, we define p to be the fraction of reacted B-groups, so
then p(f – 1) is the fraction of reacted A-groups, whereas the fraction of unreacted
A-groups is 1 – p(f – 1). Again, this latter fraction is identical to the number density
of molecules, ntot(p), as each resulting N-mer has one unreacted A-group:

ntot ð pÞ = 1 − pðf − 1Þ (7:47)

Thus, the number-average degree of polymerization can be expressed as


308 7 States of polymer systems

1 1
Nn ð pÞ = = (7:48)
ntot ð pÞ 1 − pðf − 1Þ

It diverges at a percolation threshold of


1
pc = (7:49)
f −1

Considering the example shown in Figure 120 with a functionality of f = 3, the per-
colation threshold is pc = 0.5. This means that half of all molecules have to be con-
nected to each other to realize a system-spanning molecule and, thus, to reach the
gel point.
These simple examples illustrate to us how bond percolation can be discussed
in a statistical view, and how two key parameters, the percolation threshold, pc,
and the number-average degree of polymerization, Nn(p), can be obtained from
that. In the following, we generalize this discussion.

Figure 120: Schematic representation of one-dimensional (upper sketch) and two-dimensional


(lower sketch) interconnection of mutually reactive AB-type monomers to a linear (upper sketch)
and a branched (lower sketch) macromolecule.

7.2.2.3 Mean-field model of gelation


In a mean-field approach to describe a bond percolation process, f-functional
monomers are placed on an f-functional lattice, quite similar to the conceptual ap-
proach of the Flory–Huggins mean-field theory for polymer thermodynamics. In a
very simple approach, this may be a regular lattice such as one with a square lattice
geometry; in a more complicated approach, we may also work with an irregular ar-
rangement of the monomers, as shown in Figure 121. On such geometries, however,
possible ring formation (“backbiting”) would lead to high mathematical complex-
ity. To avoid this, a dendritic type of lattice, a so-called Bethe lattice, on which
ring formation is precluded by geometry, is used instead. It is displayed for a func-
tionality of f = 3 in Figure 122. On such a lattice, the gelation process can be con-
ceptualized as follows: at a conversion p, a bond between two randomly selected
7.2 Polymer networks and gels 309

adjacent lattice sites has been formed with a probability p. At the percolation
threshold, p = pc, the first fully system-spanning molecule shows up. From this
point on, the system contains two fractions: the gel fraction that is composed of
all molecules that are already part of the system-spanning “infinite cluster,” and
the sol fraction that is comprised of all molecules that have either not yet re-
acted at all or that are part of a non-system-spanning “finite cluster” only.

Figure 121: Bond percolation concept in a lattice-based mean-field view. Random formation of
bonds between adjacent lattice sites corresponds to interconnections of the monomers on them;
this occurs with a conversion of p = 0–1. This conversion also directly corresponds to the
likelihood that a bond has already been formed between two randomly chosen adjacent lattice
sites at the extent of reaction p. At the critical percolation threshold, also named gel point, pc, a
first system-spanning “infinite” molecule occurs.

Figure 122: A dendritic type of lattice, referred to as Bethe lattice or Cayley tree, is specifically
simple to be treated mathematically in a bond-percolation approach. The sketch shows such a
lattice with a functionality of f = 3.

I. The gel point


To determine the gel point, we randomly select a lattice site that hosts a “parent”
molecule. It bonds to one of its neighboring molecules, its “grandparent,” and fur-
ther bonds to each of the remaining f – 1 adjacent sites (“potential children”) with a
probability of p. This means that there are p(f – 1) “parent–children bonds.” Simi-
larly, there are p(f – 1) bonds from each of the “children” to potential “grandchil-
dren,” and so on. From this, two scenarios can be constructed: In the first, the
average number of bonds p(f – 1) is smaller than 1, meaning that the bond probabil-
ity p is smaller than 1/(f – 1). In this case, the “family dynasty” would not survive due
to a lack of sufficient offspring. In the second scenario, the opposite is the case. The
310 7 States of polymer systems

average number of bonds p is bigger than 1, and so the bond probability p is bigger
than 1/(f – 1). Here, each generation produces more children than the one that came
before, resulting in an “infinite family tree.” This corresponds to the creation of a sys-
tem-spanning “infinite” molecule. The transition between both scenarios is the perco-
lation threshold pc, or gel point, that can now be expressed as
1
pc = (7:49)
f −1

II. The sol and gel fraction


There are two distinctive fractions beyond the gel point: the sol and the gel fraction.
To determine the sol fraction, we define Q to be the probability for a molecule on a
randomly chosen lattice site A to be not connected to the gel by neither of its f po-
tential bonds. This is the case when it is either not connected to its neighbor B it-
self, which has a probability of (1 – p), or if it is connected to its neighbor B
(probability of p), but B is not connected to the gel by any of its f – 1 further neigh-
bors. This case would have a probability of Q f−1. This yields the recursive formula

Q = 1 − p + pQ f − 1 (7:50)

The sol fraction is defined as the probability for a randomly chosen lattice site not
to be connected to the gel by any of its f neighbors, neither directly nor indirectly. It
can be formulated as

Psol = Q f (7:51a)

The corresponding gel fraction can then be easily calculated from Psol as

Pgel = 1 − Psol (7:51b)

Below the percolation threshold, the sol fraction always has a value of Psol = 1, and
the gel fraction is always zero, Pgel = 0. Above the gel point, both fractions can be
calculated by eqs. (7.51). This relation is schematically shown in the left schematic
of Figure 123.
Pierre-Gilles de Gennes was the first to realize that the gel fraction evolves like
the order parameter of a second-order phase transition, such as the phase transi-
tion between ferro- and paramagnetic states in metallic solids. The similarity of its
graphical representation, shown in the right schematic of Figure 123, and the evolu-
tion of the gel fraction after the gel point, shown in the left schematic of Figure 123,
are striking.

III. The number-average degree of polymerization at the gel point


A key parameter to define the gelation process is the degree of polymerization at
the gel point. To calculate it, we assume the following: without any bonds present
7.2 Polymer networks and gels 311

Figure 123: The bond percolation yields the sol and gel fraction as characteristic parameters of a
gelation process. At the percolation threshold, also named gel point, pc, a first system-spanning
“infinite” molecule occurs. From that point on, addition of further monomers to this infinite cluster
gradually increases the gel fraction in the system, whereas all residual monomeric or oligomeric
(= non-space-filling) material is part of the sol fraction that steadily decreases in turn. The gel
fraction can be taken as an order parameter to quantify the progress of the gelation process,
similar to the conceptual treatment of second-order phase transitions like the ferromagnetic–
paramagnetic transition in metallic solids.

in the system, the number density of molecules is defined to be unity, ntot(p = 0) = 1.


Each newly formed bond then reduces the total number of molecules by 1. Conse-
quently, the maximum number of bonds that can be formed is f/2. All intermediate
states are calculated according to

f
ntot ð pÞ = 1 − p (7:52)
2

This yields the number-average degree of polymerization as


1 1
Nn ð pÞ = = (7:53)
ntot ð pÞ 1 − p 2f

At the percolation threshold, it is characterized by

1 2ðf − 1Þ
Nn ðpc Þ = = (7:54)
1 − f =2ðf − 1Þ f −2

We see that the number-average degree of polymerization does not diverge at the
gel point, in contrast to the one- and two-dimensional percolation processes treated
earlier. For example, a molecule with a functionality of 3 has a number-average de-
gree of polymerization of 4 at the percolation threshold, Nn(pc) = 4.
312 7 States of polymer systems

IV. The weight-average degree of polymerization at the gel point


The approach to determine the weight-average degree of polymerization at the gel
point is quite similar to the discussion of the sol and gel fraction. Let us define μ as
the average number of molecules that are connected to a randomly chosen lattice
site A by one of its f neighbors. μ is composed of two contributions: first, the direct
neighbor site B, which is connected to A with a probability of p, and second, the
remaining f – 1 neighbors of B, each of which connected to an average number of
monomers μ in turn. This yields the following recursive formula:
p
μ = pð1 + ðf − 1ÞμÞ ) μ = (7:55)
1 − ðf − 1Þp

The weight-average degree of polymerization, Nw, corresponds to the average num-


ber of monomers that belong to a randomly chosen lattice site. This number is com-
posed of 1 for the chosen lattice site itself and μ for each of its f neighbors. Thus, Nw
can be calculated by
1+p
Nw ð pÞ = 1 + f μ = (7:56)
1 − ðf − 1Þp

In contrast to the number-average degree of polymerization, it does diverge at the


gel point, because it puts more emphasis on large clusters/molecules and, conse-
quently, “gets overwhelmed by the infinite one” that arises at the gel point.
In the above discussion, we have (again) successfully simplified a complex
multibody situation, bond percolation during a gelation process, using a mean-field
approach to calculate statistical averages. We have seen how to estimate the perco-
lation threshold, pc, and the number-average as well as the weight-average degrees
of polymerization at this crucial point during the percolation process. Beyond the
gel point, we have calculated the sol and gel fractions, and realized that they evolve
much like in a second-order phase transition as often seen in magnetically active
solids. Such understanding that soft-matter phenomena are fundamentally analo-
gous to seemingly very different hard-matter phenomena have earned Pierre-Gilles
de Gennes the Nobel Prize in physics in 1991.

7.2.2.4 The gel point


The percolation process and the gel point can be monitored using rheology. This is
because rheology is able to distinguish between the viscous, fluid-like and the elas-
tic, solid-like component of a sample, which largely correspond to the sol and gel
fraction. Furthermore, during a gelation process, the continuous buildup of longer
and longer chains can be easily monitored with that method by measuring the raise
of the sample viscosity, η. Most typically, the starting reaction mixture has a low
initial viscosity value, η0, especially when it is made up of low-molar-mass mono-
mers instead of precursor polymer molecules. As soon as the reaction starts, however,
7.2 Polymer networks and gels 313

longer aggregates are formed, and the mixture becomes more viscous. The viscosity
will rise even faster and in a nonlinear fashion as soon as these aggregates become
sufficiently long to entangle with one another, especially as they are also branched.
Once the first system-spanning molecule occurs, at the percolation threshold
pc, the viscosity diverges. Figure 124 summarizes this mechanical signature of a ge-
lation process. Having exceeded the percolation threshold, continuation of the gela-
tion process is now represented in the temporal evolution of the storage and loss
moduli, G′ and G″. Beyond the gel point, the sample is a viscoelastic solid in which
the storage modulus is higher than the loss modulus, G′ > G″; it is no longer able to
flow, but instead, exhibits elasticity. The loss modulus mostly reflects the remain-
ing sol fraction in the sample as well as energy-dissipating entities in the gel net-
work, such as loops and dangling chains. Beyond the gel point, such structures are
more and more interconnected to and within the gel, and so its value gradually de-
creases until it finally vanishes at the point of full cross-linking. The storage modu-
lus reflects the gel fraction and the gel’s mechanical strength; it is composed of a

Figure 124: A gelation process, shown as a sequence of sketches at the bottom of the figure, can
be followed by rheology. Continuous determination of the viscosity will reveal the percolation
threshold, pc, where the viscosity diverges. Beyond the gel point, the evolution of the sol and gel
fractions can be followed by the storage modulus, G’, which represents the gel fraction (and which
is composed of a fraction allocated to cross-links, Gx, and to (trapped) entanglements in the
network gel, Ge), and the loss modulus, G″, which represents the sol fraction. As the cross-linking
process proceeds beyond the gel point, the storage modulus G’ increases due to formation of
further cross-links (that means, it is mainly its Gx contributor which increases), whereas the
fraction of energy-dissipating entities such as noninfinite clusters in the system as well as loops
and dangling chains in the gel network decreases, thereby decreasing G″.
314 7 States of polymer systems

contribution made by (fixed) chain entanglements, Ge, and a contribution made by


interchain cross-links, Gx.
In an even deeper view, frequency-dependent oscillatory shear rheology can reveal
if the percolation threshold has already been crossed, using the Winter–Chambon cri-
terion of a power-law dependence of the relaxation of G at the gel point given by

Gð t Þ ⁓ t − n (7:57)

At the gel point, both the elastic and the viscous parts of the complex modulus,
G’ and G″, scale with that very exponent in a frequency-dependent plot: G’(ω) ⁓
G″ (ω) ⁓ ωn; for G’, and this is shown in Figure 125. The exponent n reflects the
relative amount of cross-linking junctions to the amount of cross-linker. It is one
half, n = 0.5, for stochiometrically balanced end-linking networks, meaning that each
junction is connected to another one by a suitable cross-linker. An excess of cross-
linker leads to a smaller value, n < 0.5, whereas a lack of cross-linker causes a bigger
value, n > 0.5. Precisely, the storage and loss modulus are given by the following
formulas:
ð
πωn
G0 ðωÞ = ω GðtÞ sinðωtÞdt = (7:58a)
ΓðnÞ sinð1=2 πnÞ
ð
πωn
G0 ðωÞ = ω GðtÞ cosðωtÞdt = (7:58b)
ΓðnÞ cosð1=2 πnÞ

According to these formulas, the power-law exponent n reveals information about


the relative values of both moduli: if n = 0.5, both the storage and loss modulus
have the same value, G’ (ω) = G″ (ω). A value of n > 0.5 leads to a bigger loss than

Figure 125: According to the Winter–Chambon relation, Gðt Þ ⁓ t − n , the complex modulus exhibits
an infinite power-law dependence on the measurement frequency at the gel point, with both its
parts scaling according to G’(ω) ⁓ G″(ω) ⁓ ωn. Below the gel point, by contrast, G’ has different
scaling in the Maxwellian free-flowing and in the Kelvin–Voigt leather-like regime, whereas above,
it exhibits an elastic frequency-independent plateau.
7.2 Polymer networks and gels 315

storage modulus, G’(ω) < G″(ω), whereas a value of n < 0.5 entails a bigger storage
than loss modulus, G’(ω) > G″(ω).

7.2.3 Structure of polymer networks and gels

We have seen earlier that polymer gels are basically semi-dilute solutions of poly-
mer chains that are cross-linked together. As such, gels are essentially “shape-sta-
ble fluids”: they consist of more than 90% (often up to 99%) of a fluid medium
supported and held in place by a flexible polymer-network skeleton. In the most
widespread type of gels, that fluid medium is water; those gels are referred to as
hydrogels. The simplified perception of such hydrogels as “shape-stable water”
points to their two most relevant areas of application. First, their ability to take up
large amounts of fluids (here: water) and to undergo swelling, as we have quanti-
fied in Section 5.9.3, makes gels useful as superabsorbers, with a prominent appli-
cation in diapers and other hygiene products. Second, their constitution of more
than 90% water renders them permeable for small substances, whereas their poly-
mer-network skeleton blocks diffusants that are larger than the network mesh size.
This combination makes gels useful for applications as separators and matrixes in
energy-conversion devices, in the analytical sciences, and in the large field of bio-
medicine. For all that, it is crucial to get a notion of the gel-network nano- and
microstructure and about the diffusive dynamics of guest substances that pass
through the gel network; this shall be in the focus of the following two sections.
Let us start with a view on gel structures.

Figure 126: Spatial inhomogeneities and connectivity defects in polymer networks and gels. On
scales of about Ξ = 10–100 nm, polymer-network gels usually feature a rather inhomogeneous
spatial distribution of their cross-linking and polymer-segmental density. Furthermore, on scales of
about ξ = 1–10 nm, polymer networks commonly display local structural defects such as loose
dangling chains and loops, along with a nonuniform distribution of the network mesh sizes.
316 7 States of polymer systems

An intuitive imagination of a polymer-network gel is that of a regular, monodis-


perse, and uniform array of meshes with a size denoted ξ, typically spanning a
range of some few nanometers. This imagination, however, is incorrect in almost all
cases. Instead, gels usually display a rich variety of structural imperfections and ir-
regularities, as illustrated in Figure 126.
At the local scale of the gel-network meshes, in the range of about ξ = 1–10 nm,
we first have to consider a polydisperse distribution of network mesh sizes, featur-
ing small ones, large ones, and all kinds of sizes in between. On top of that, poly-
mer networks commonly display structural defects on these scales, such as loose
dangling chains connected to the network only with one extremity and loops con-
nected to the network with two extremities but not connecting to further parts of
it, thereby not being able to store elastic energy upon deformation.
At larger scales, in the range of about Ξ = 10–100 nm, polymer-network gels
usually feature quite pronounced spatial inhomogeneity of their cross-linking and
polymer-segmental density. We have touched upon that already in Section 6.4.1,
and also detailed about the experimental assessment of these structures by light
scattering in Section 6.4.2. These inhomogeneous structures form due to two rea-
sons. First, in the case of gel formation by random interconnection of preexisting
precursor polymer chains, thermal fluctuations in the system will always cause mo-
mentary nonhomogeneous distribution of the chain material in the system, and
upon cross-linking, these random temporal fluctuations are “frozen-in” into the
network structure to becoming permanent spatial fluctuations. Second, in the case
of gel formation by polymerization of multifunctional monomers (or of a mixture of
bifunctional monomers and some multifunctional ones acting as cross-linkers), in
the early stage of the reaction, multiple cyclization and self-cross-linking will take
place, leading to formation and growth of local nanogel clusters. At a later stage of
the process, these clusters will become interconnected to form a space-filling gel,
which will then still resemble the inhomogeneous distribution of the material from
the early stage of the process.
Some gels even display further structural inhomogeneity on scales of 100–
1,000 nm and beyond, referred to as porosity. Such pores in gels can be formed
due to multiple reasons, including gas bubbles in the system that emerge due to
formation of a gaseous by-product, the intentional use of a porogen, or microphase
separation of a thermally responsive gel due to the heat of reaction during the gela-
tion process.
The inhomogeneous structures of gels have multiple consequences on their
properties. Most technically relevant is their effect on the gel mechanics. We have
learned how to appraise the elastic modulus of a polymer network (and therewith
also of a gel) in Section 5.9.2, in the context of the rubber elasticity model. That
model, however, assumes a monodisperse and uniform network mesh topology. In
a more realistic polydisperse array of meshes, though, the experimentally assessed
modulus will be some kind of average. Even more, in the case of presence of spatial
7.2 Polymer networks and gels 317

inhomogeneities, the locally densely cross-linked nanogel clusters essentially do


not store elastic energy internally, but instead, just act like one large cross-linking
node each. That means that all the cross-linking points “buried” inside these clus-
ters are “lost” and do not contribute to elastic energy storage, which is why the
elastic modulus of gels is often a lot smaller than (typically just a few percent of)
what might be expected based on their pure chemical amount of cross-linker.

7.2.4 Dynamics of polymer networks and gels

The polymer network mesh structure (and all the inhomogeneity that it exhibits)
renders gels interesting as separator materials that let pass substances smaller than
the mesh size while blocking diffusants larger than the meshes. Based on that
premise, it is logical that depending on the probe size and flexibility, different re-
gimes and mechanisms of diffusion through a polymer-network gel matrix can be
identified and presumed.
In one scenario, the diffusion of small molecules through the network mesh
structure is regarded. For that type of small-probe diffusion, three classes of models
have been developed, all predicting a marked obstruction exerted by the polymer
network on the diffusion of the small molecules as compared to their diffusion in
the absence of the polymer matrix, which is commonly addressed to (i) reduction of
the free volume for the probe to diffuse into, (ii) hydrodynamic interactions be-
tween the probes and the gel-network matrixes, and (iii) an increased path length
for the probe motion due to obstacles set by the gel-network matrix; in addition,
combinations of these effects have been discussed.
A much more complicated situation is encountered when larger probes, such as
macromolecules or colloidal particles, shall diffuse through a gel network, which in
fact requires the precise consideration of the probe flexibility, its fractal dimension,
its size compared to the network mesh size, and the ratio of its characteristic diffu-
sion time to the characteristic time for network rearrangement, as illustrated and
summarized in Figure 127. We can simplify that distinction by narrowing it down to
just two cases: rigid mesoscopic probes that can pass the network meshes only if
these are still not too small for the probe, and flexible mesoscopic probes that may
find mechanisms to move though the network meshes even if these are inherently
smaller than the probe.
The prime example of rigid mesoscopic probes are colloidal hard spheres. The
diffusion of those in polymer matrixes has been appraised originally by De Gennes
for the case of uncross-linked semi-dilute polymer solutions, which was later re-
ported by Langevin and Rondelez,105 by considering the increase of the friction

 Polymer 1978, 19(8), 875–882.


318 7 States of polymer systems

Figure 127: Diffusion of probe particles in polymer-network matrixes as a function of the relative
size of the diffusing species (R) to the network mesh size (ξ) and the fractal dimension of the
diffusing species, related to its rigidity or flexibility. Figure reprinted with permission from
Macromolecules 2002, 35(21), 8111–8121. Copyright 2002 American Chemical Society.

coefficient in a form of f0/f ⁓ exp(–(r/ξ)δ), with r being the probe radius and ξ being
the screening length that corresponds to the (here transient) network mesh size.
This notion is based on a simple estimation of the energy of activation for slipping
through the meshes. A further argument by Cukier suggested δ to be unity.106 The
same type of formula has also been found experimentally and later been derived
theoretically by George D. Phillies, who suggested a universal scaling equation of
type D ⁓ exp(–(r)α) to describe the translational diffusion coefficient, D, of various
types of probes such as globular particles, linear chains, or star-shaped polymers in
macromolecular matrixes over the entire concentration range from dilute to concen-
trated solution. In an alternative assessment, the motion of free flexible polymer
chains through a polymer network gel may also be quantified based on a reptation
mechanism along with scaling arguments, as we have treated it in Section 7.1.4.3.

 Macromolecules 1984, 17(2), 252–255.


Questions to Lesson Unit 22 319

Questions to Lesson Unit 22

(1) What holds for unbound polymer chains according to the definition of a fish-
net-type gel?
a. Per definition, a polymer fishnet-type gel consists of crosslinked network
chains, unbound chains, and solvent.
b. The definition of the gel includes all polymer chains, the ones included in
the polymer network and unbound ones.
c. Only branched polymer chains that are not part of the network but contrib-
ute to the sample’s elasticity are part of the definition.
d. Unbound chains are not included in the definition of the gel.

(2) What is the difference between the sol and the gel fraction?
a. The gel fraction only includes the system-spanning network(s), whereas
every network of smaller size is included in the sol fraction. Unbound
chains are not included in neither of these.
b. The gel fraction only includes the system-spanning network(s), whereas
every network of smaller size is included in the sol fraction together with
currently unbound chains.
c. The gel fraction includes every network in the sample, whereas the sol frac-
tion only includes the unbound polymer chains.
d. The gel fraction only consists of networks that are permanently cross-
linked, whereas the sol fraction consists of the reversibly cross-linked ones.

(3) What does the phenomenon of percolation in polymer science refer to?
a. The buildup of a system-spanning network
b. The separation of the gelled part from the still-liquid part of a polymer
solution
c. The filtration of a polymer solution
d. The formation of polymer coils in a polymer solution

(5) What is the percolation threshold?


a. The point at which percolation is completed.
b. The point at which percolation reaches a critical value – the first system
spanning network is formed.
c. The point at which percolation is at its highest rate.
d. The point at which percolation is possible due to a high-enough polymer
concentration.
320 7 States of polymer systems

(6) What is another popular simple model of percolation next to bond percolation?
a. String percolation, where adjacent lattice sites can be connected by linear
sting-like structures.
b. Site percolation, where all lattice sites are unoccupied at first and network
formation depends on whether the adjacent lattice sites get occupied or not.
c. Branched percolation, where dendritic networks are formed starting from a
single percolation center.
d. Cluster percolation, where inhomogeneous clusters form due to inhomoge-
neous conversion of the network structure.

(7) What does not happen when the percolation threshold pc is reached?
a. A system spanning network is formed.
b. The gel point is reached.
c. The gel fraction reaches a value of 1.
d. The viscosity diverges.

(8) Organize the following polymer network inhomogeneities according to their


size from large to small:
① Loose dangling chains, loops, and diverse mesh sizes
② Pores, formed by gaseous by-products or phase separation during gelation
③ Clusters due to multifunctional monomers
a. ② > ① > ③
b. ③ > ② > ①
c. ① > ② > ③
d. ② > ③ > ①
7.3 Glassy and crystalline polymers 321

LESSON 23: POLYMERS IN BULK SOLID STATES


The former lesson unit has introduced an important solid state of polymers: the gel. In that
state, the mechanical appearance of a specimen is form-stable and elastic, meaning that the
storage part of the complex modulus is greater than the loss part. The composition of a sam-
ple in that state, however, is mostly constituted by a marked amount of immobilized low-
molar-mass swelling medium. In the following, we will focus on solid states that are composed
of a polymer in its bulk only, thereby featuring much higher moduli than gels. Depending on
the regularity of microscopic packing of the material in these bulk solid states, we speak
about a glassy or a crystalline polymer. In this lesson unit, we learn what the differences and
characteristics of these states are, how they form, and what properties they have.

7.3 Glassy and crystalline polymers

7.3.1 The glass transition

7.3.1.1 Fundamentals
To this point, we have extensively considered polymers with flexible conforma-
tions, as realized in the states of a melt, a solution, or a gel. We now switch gears
and consider the opposite: polymers in a fixed, nonflexible conformation, as real-
ized in bulk solid states. Depending on whether the fixed microscopic chain confor-
mation is disordered and amorphous or ordered and regular (at least in part),
we distinguish between polymers in a glassy or in a crystalline state (to be precise,
it is actually a partially crystalline state for polymers). Unlike low-molar-mass mole-
cules that predominantly crystallize in an ordered fashion upon decrease of temper-
ature, polymers often undergo a glassy solidification named vitrification instead.
This is because their usually nonregular primary constitution as well as the polymer
chain random coiling and potential chain entanglement prevent many polymers to
build up a regular crystal structure. Instead, such polymers rather “freeze” as an
amorphous structure as soon as the environmental energy is too low to allow for
segmental mobility. This point is called the glass-transition temperature, Tg,
which is actually more a temperature range than a defined single temperature.
Figure 128 shows the position of the glass-transition regime on the modulus
versus time-or-temperature representation that we have already encountered in
Section 5.7.3 (see Figure 83), when we learned about the mechanical spectrum of a
polymer. In that transition regime, a polymer exhibits sequential activation of
Rouse or Zimm modes, thereby transitioning from a completely frozen state without
any kind of chain-backbone dynamics to a state in which such dynamics is acti-
vated. In the former state, chain deformation can only be accommodated on an
atomic scale in a sense that bonds are stretched or bonding angles are distorted,
which comes along with an energy-elastic response captured by values of the elastic
322 7 States of polymer systems

Figure 128: Schematic representation of a polymer’s full mechanical spectrum. Below the glass-
transition temperature, Tg, the energy is insufficient even for motion of only single chain segments,
and so the amorphous polymer structure is frozen in time. Often, the glass transition happens
gradually over a glass-transition regime rather than at a specific sharp temperature.

modulus in the gigapascal range. In the latter state, by contrast, chain deformation
can be accommodated by much easier bond torsion, which comes along with an en-
tropy-elastic response captured by values of the elastic modulus in the megapascal
range. In the transition range in between these fundamentally different states, the
polymer sample has both already activated “fluid” and still deactivated “solid” chain
dynamics, and as a result, both the storage and the loss part of the complex modulus
are relevant, which ideally results in mechanical characteristics with G’ ≈ G″, that is,
tan δ ≈ 1.107

Due to the fundamental change of the polymer mechanical characteristics in the glass-transition
regime, the glass-transition temperature is in fact one of the most relevant material parameters.
There is one remarkable and tragic example of material failure caused by ignorance of that fun-
damental change. On the morning of January 28, 1986, the spacecraft Challenger was supposed
to launch from Kennedy Space Center to a 6-day space mission. Temperatures on that January
morning were low, and so rubbery gaskets in the fuel system were in a glassy state instead of a

 To be precise, this line of argument actually only holds for materials that directly transition
from a solid-like elastic glass, where tan δ < 1, to a liquid-like viscous melt, where tan δ > 1, so that
the intermediate point with tan δ = 1 marks the transition. According to Figure 128, however, sam-
ples with entangled or cross-linked chains instead show a transition from one solid state to another
solid state, namely between an energy-elastic glass to an entropy-elastic rubber. In those, tan δ is
not necessarily 1 at the transition point, because then, G′ > G″ in both the glassy and the rubbery
realm, that means, both below and above the glass transition. But at least, G″ approaches G′ during
the transition in that case, as also shown in Figure 128. Hence, in that case, tan δ is always smaller
than 1, but at least it is “minimally small” (i.e., maximal) at the point of closest proximity of G′ and
G″, which thereby marks the glass-transition.
7.3 Glassy and crystalline polymers 323

rubbery one, thereby having lost their ability to seal. The result was fuel leakage, leading to an
explosion that disrupted the vessel 73 s after liftoff at about 15 km altitude. Seven astronauts’
lives on board were lost on that day, including the one of a school teacher who had planned to
give classes from space. This catastrophe shows us how (literally) vital it is to be aware of the
fundamentally different properties of a polymer material in the glassy versus the rubbery state,
and what role temperature plays in the transition between the two.

It is important to note that the glass transition is not a phase transition phenome-
non, despite it is being called a transition. A phase transition is defined as a change
in the microstructure of a sample that entails an observable change of the macro-
scopic properties. During the glass transition, however, the chain segmental dynam-
ics is frozen, but the microstructure of the sample does not change. Hence, even
though the glass transition has some phenomenological characteristics that also
second-order phase transitions have, it is actually not a phase transition but a
strictly kinetic phenomenon! That phenomenon may in fact be expected to happen
for all matter, that means, not only for polymers, as all matter may generally be
subject to the exact same principle of getting to a point where its dynamics is frozen
while its structure is still amorphous. For that to happen, we must simply have a
situation at hand where during the drop of temperature, crystallization is precluded
to happen somehow, so that we can get to a point where the temperature eventually
gets so low that all dynamics is frozen while the sample still has an amorphous
structure. This is the case if the rate of cooling is so fast that the sample has no time
to crystallize underway, meaning that the molecules do not manage to arrange
themselves on a regular crystalline lattice before their dynamics is already frozen.
As classical molecular matter is commonly composed of rather small molecules,
though, these in fact can do so in many cases, such that no glass transition but in-
stead normal crystallization is seen usually, unless we cool down really fast, for ex-
ample, by shock-freezing in liquid nitrogen, helium, or argon. In soft matter, by
contrast, the large size and slow dynamics of the building blocks is a great drag or
even impairment for crystallization to happen, such that glassy solidification hap-
pens in many cases even at moderate cooling rates. Here, we simply pass the point
of potential crystallization too quickly and thereby then go to a state of frozen dy-
namics but still amorphous structure. This is especially pronounced for polymers,
as in those, it is not the coils as a whole that must be arranged on lattice positions,
but the segments in them, and due to their mutual connectivity, which is often
even irregular, such regular arrangement on crystalline lattice positions is either
pretty impaired or even completely impossible. This is why the glass transition is
observed quite frequently (and why crystallization is observed quite rarely) in poly-
mer materials.
324 7 States of polymer systems

7.3.1.2 Conceptual grounds of the glass transition


We can get a conceptual microscopic understanding of the glass-transition phe-
nomenon from two viewpoints. The first viewpoint has a kinetic and energetic
focus, as we have also partly argued just above already. Segmental dynamics in
polymer chains takes time; we have quantified that in the concept of relaxation
modes in Section 3.6.4. The higher the mode index, the more segments move in a
cooperative fashion, and the longer does it take for that subsection of a chain to get
displaced over a distance at least equal to its own size. And this is even more true
when the temperature is low, because dynamics is generally thermally activated.
In this conception, we may view the glass-transition temperature to be the one
below which even the displacement of just single monomer units in a chain over a
distance equal to their own size requires infinitely long time. In an alternative con-
ception, we may also address an energy argument for the same matter: chain seg-
mental motion requires bond torsion along the polymer backbone, and this has
energies of activation in the range of some few multiples of RT (or kBT) at room tem-
perature (see Figure 14 in Section 2.1.1). Even though these are low activation ener-
gies, at very low temperatures even those cannot be overcome, so that any sort of
chain dynamics is frozen – or takes infinity long times if we translate the activation
energy into a rate constant based on an Arrhenius-type equation.
A second viewpoint on the glass transition has a focus on the sample volume.
Most of the volume of a material at high temperature is constituted by free volume,
that is, by empty space between its molecules. Now imagine a polymer melt that is
cooled down. The cooling will reduce the volume, and according to what we have
just said, this is first predominantly constituted by reduction of the free volume be-
tween the chain segments. At some point, the free volume is so low that the chain
segments cannot move anymore, as they do no longer have enough empty space
around to move into. This point is the glass transition, and the temperature at
which it happens is the glass-transition temperature. At further cooling, further
shrinkage of the sample does then only occur due to thermal contraction of the
chain material itself, but no longer of the residual small free volume in between,
which then stays constant.
The latter conception of the glass transition is visualized in Figure 129, which dis-
plays the relation between a polymer compound’s specific volume and temperature.
The volume of the chains themselves, Vp, increases constantly with temperature ac-
cording to the thermal expansion coefficient, α. Above the glass-transition tempera-
ture Tg, there is an additional free volume Vf that also shows thermal expansion, and
that expansion is actually a lot more pronounced than the underlying one of the
chain material itself. Below the glass-transition temperature Tg, however, the free vol-
ume Vf,g is only minimal and constant, so that here, thermal expansion is constituted
only by the contribution of the chain material itself, Vp. We may express this matter
via the fraction of the free volume in the system, f = Vf /V, as follows:
7.3 Glassy and crystalline polymers 325

Figure 129: The temperature-dependent specific volume of a polymer sample is a sum of the
inherent volume of the polymer chain segmental material itself, Vp, and the free volume in between
the chains and segments, Vf. Vp steadily increases by the usual thermal expansion of the chain
material. Vf adds upon that a much greater increase of the total sum volume at high temperatures.
By contrast, at low temperatures, below the glass-transition temperature, Tg, the free volume in the
glassy state Vf,g stays constant, such that thermal expansion (upon heating) or contraction (upon
cooling) in the low-T range is due to expansion/contraction of the polymer chain material itself
only. Picture redrawn from B. Tieke: Makromolekulare Chemie, Wiley VCH, 1997.

 
f = fg + T − Tg αf (7:59)

Here, f is the just introduced free volume fraction, fg is the constant free volume
fraction in the glassy state, and αf is the thermal expansion coefficient of the free
volume.
We can use these relations to again derive the Williams–Landel–Ferry (WLF)
equation for time–temperature superposition of rheology data that we have al-
ready derived from a slightly different perspective in Section 5.7.3. Here, we start
from the Doolittle equation for the sample viscosity:

V
ln η = ln A + B (7:60)
Vf

This expression was found empirically, but it can also be regarded illustratively: V
refers to the volume of the macromolecules or particles in a soft matter sample,
whereas Vf is that of a corresponding void where these can move if dynamics is not
frozen. If this void volume, Vf, is larger than the molecule/particle volume V, then
this motion is easy and the viscosity η is low, and vice versa. In this picture, the
Doolittle equation is like an Arrhenius or Eyring equation that relates the thermal
energy in the system to that required to overcome a certain barrier.
Use of the free volume fraction f = Vf/V turns this into

ln ηðT Þ = ln A + Bð1=f Þ (7:61a)

At the glass-transition temperature, we can write


326 7 States of polymer systems

   
ln η Tg = ln A + B 1=fg (7:61b)

Taking the ratio of these two expressions and replacing f according to eq. (7.59)
yields
!
ηðT Þ 1 1
ln   = B  − (7:62)
η Tg fg + αf T − Tg fg

We can substitute the viscosity η by the relaxation time τ according to τ = η/G to


generate
  
ηðT Þ τðT Þ τðT Þ − B=2:303fg T − Tg
ln   = ln   ≈ log   = log aT =     (7:63)
η Tg τ Tg τ Tg fg =αf + T − Tg

Here, aT is the shift factor for time–temperature superposition as encountered in


Section 5.7.3.
Empirical findings show that most polymers exhibit universal values of fg = 0.025
and αf = 4.8 × 10−4·K−1. This means that universally 2.5% of the available volume is
free in the glassy state, whereas 97.5% of the volume is occupied by the polymer
chain material.
The latter finding stands in contrast to a different approach and prediction by
Simha and Boyer. In their perception, both the glass and melt are amorphous and
should therefore have the same volume at the absolute zero point, meaning their ther-
mal expansion lines should meet there, as shown by the dashed lines in Figure 130.
This perception then implies a free volume of 11.3% in the glassy state, as also shown
geometrically in Figure 130, which is about the fivefold of the prediction in the WLF
treatment. An explanation for this deviation might be that the WLF treatment has a
sole focus on volume effects, whereas there is in fact also an energy of activation con-
tribution that is lumped into volume effects in that notion; as a result, the volume
confinement in the glassy state is overestimated in the WLF assessment, whereas in
fact, the glassy arrest is only in part due to lack of free volume, and in other part due
to lack of thermal activation of dynamics.
The latter thought is in line with our initial conceptual consideration about the
two factors contributing to the glass transition. The first is the temperature depen-
dence of the free volume, Vf, as just discussed. The second is a temperature depen-
dence of the rate of overcoming the energy of activation, Ea, of the chain segmental
motion, as discussed further. Both are operative, and as a result, an Arrhenius plot
for a glass transition may deviate from linearity, as is illustrated in Figure 131. This
figure shows the logarithmic inverse time of occurrence of a glass transition, 1/τg = ωg,
and an additional potential secondary transition in plots of G versus t, recorded at
many different temperatures. The secondary transition, which is due to potential
additional freezing of chain side-group rotations in the glassy state upon further cool-
ing, exhibits a simple Arrhenius relation with a shallow slope due to its single
7.3 Glassy and crystalline polymers 327

Figure 130: Temperature-dependent specific volume of a polymer sample (solid line) and estimates
of its composition of a 2.5% free volume, as it results from a WLF-based conception (dotted line), or
alternatively, an 11.3% free volume, as it results from the Simha–Boyer expectation that the
volumes of the glass and melt should coincide at the absolute zero point (dashed lines), since both
these states share the same amorphous structure.

contribution of the energy of activation, Ea. The glass transition deviates from this
linearity, especially at low temperatures, where it displays a much steeper slope. This
is because in this regime, both the free volume, Vf, and the energy of activation, Ea,
contribute. At high temperatures (T ! ∞) both converge, as there is enough free vol-
ume then, such that only the energy-of-activation contribution matters.
Glass transitions are nonequilibrium kinetic processes. This can be readily visu-
alized by the two examples shown in Figure 132. The two panels show plots of the
specific volume as a function of temperature for two different phenomena. Panel
(A) visualizes an aging process: here, a sample is cooled down from an initial tem-
perature that is higher than Tg to a temperature below Tg. The glass transition is
clearly visible from the bend in the graph. However, even after the cooling has been
stopped, the sample further reduces its specific volume – a phenomenon named
aging. The occurrence of that aging shows that the sample has not been at equilib-
rium before, even though the glassy state was already reached. Panel (B) displays
the dependence of Tg on the rate of cooling. For different cooling rates, denoted q1
and q2, different glass-transition temperatures are obtained. Again, this ambiguity
shows that the glass transition is not an equilibrium transition. Both the phenom-
ena shown in Figure 132 therefore demonstrate that the glass transition has various
kinds of time dependences; this proves it to be a kinetic phenomenon and not an
equilibrium phase transition.
328 7 States of polymer systems

Figure 131: Schematic representation of the logarithmic inverse time of occurrence of a glass
transition, 1/τg = ωg, and a potential additional secondary transition in plots of G versus t,
recorded at many different temperatures, T. The secondary transition exhibits a single contribution
of the energy of activation, Ea, whereas the glass transition has a second contribution stemming
from the free volume, Vf, that is exacerbated at low temperature (right end of the 1/T axis). At high
temperature (left end of the 1/T axis), both graphs converge, as Vf is then large enough.

Figure 132: The kinetic nature of the glass transition is seen by aging processes of glassy samples
(A) and by the dependence of the glass transition and the temperature of its occurrence on the
cooling/heating rate, q (B). Picture redrawn from J. M. G. Cowie: Chemie und Physik der
synthetischen Polymeren, Verlag Vieweg, 1996.
7.3 Glassy and crystalline polymers 329

7.3.1.3 Structure–property relations for Tg


The glass-transition temperature is connected to the polymer sample’s structure by
a set of structure–property relations. For example, bulky or polar substituents hin-
der the main chain rotation. These types of polymers are therefore more easily fro-
zen in motion than their more flexible counterparts. Indeed, polyethylene (PE)
exhibits a glass-transition temperature as low as Tg(PE) = 188 K, whereas by adding
a bulky benzene side group to each monomeric unit, the glass-transition tempera-
ture of the resulting polymer polystyrene (PS) is raised by a factor of 2, to Tg
(PS) = 373 K. Polypropylene (PP) exhibits a glass-transition temperature in between,
at Tg(PP) = 253 K. Substitution of the unpolar methyl groups by polar nitrile groups
yields polyacrylonitrile (PAN); the glass-transition temperature is then raised to Tg
(PAN) = 378 K. Cross-linking of polymer chains also makes the system more rigid,
thereby increasing Tg, whereas branching decreases Tg due to a lowered packing
density and a resulting higher free volume as compared to a non-branched sample.
Other means of artificially lowering the glass-transition temperature include the
use of so-called softeners. These can either be chemical, through the inclusion of
flexible comonomers, or structural, by using other architectures like side chains.

7.3.1.4 Experimental assessment of Tg


All the preceding discussion shows us how central the glass-transition temperature
is. But how can it be assessed experimentally? There are many possible answers. At
the beginning of this section, we have assessed the glass-transition regime based
on the mechanical spectrum of a polymer sample in Figure 128, wherein which the
glass-transition temperature was in fact also marked on the abscissa. If the sample
under consideration shows a rubbery elastic regime behind the glass transition, as
it is sketched in Figure 128, the parameter tan δ is a proper quantity to capture this
numerically: it is low in both the energy-elastic glassy state below Tg and in the en-
tropy-elastic rubbery state above Tg, whereas it is high in the glass-transition range
at and around Tg, because in that range, both still-frozen and already-activated
main-chain dynamics is present in the sample, such that both G’ and G″ are rele-
vant. The point where their ratio, tan δ, is maximal is therefore reliably seen to be
an assessment of the glass-transition temperature Tg.108 Alternatively, a little later
in our discussion, we have assessed the glass transition by the temperature-depen-
dent change of the sample (specific) volume, as displayed in Figures 129, 130, and
132, wherein which the glass-transition temperature was also marked on the ab-
scissa at the point where the volume change had a kink or bend. In addition to
these two methods, there are several further ways to assess that temperature. A

 This assessment is actually only applicable if we deal with a plain polymer material. In practice,
however, materials are often modified (e.g., enforced) by fillers. Such additives drastically change tan
δ, thereby potentially rendering it impractical to serve as a marker of the glass transition.
330 7 States of polymer systems

practical problem, however, is that all these different assessments lead to different
values of Tg even for the absolute same sample. This ambiguity shows us again that
the glass transition is not a sharp and well-defined phase transition, and so in prac-
tice, giving a number for Tg should always come along with telling how exactly it
was determined.

7.3.2 Polymer crystallization

7.3.2.1 Fundamentals
As an exemption to the general rule that polymers do not crystallize, some polymers
do in fact show partial crystallinity before the glass-transition regime. Here, we can
observe a true phase transition. However, polymers only crystallize under certain
preconditions. First, they have to be composed of strictly linear chains to be able to
align themselves in a regular fashion. Side chains, branching, or other complex pri-
mary architectures prevent this. Second, the monomer sequence needs to be well
defined in view of the monomer-to-monomer connectivity and even in view of its
stereochemistry, meaning that crystallization is only possible for isotactic or syn-
diotactic homopolymers. Third, intermolecular interactions that stabilize the
polymer chain against the entropically unfavorable packing in a crystal aid the
crystallization process as well.
Polymer crystallinity is best characterized by x-ray diffraction. Here, a dry sam-
ple is exposed to x-rays that are reflected when they hit a crystalline scattering cen-
ter. If these centers are arranged symmetrically with a separation d, part of the
incoming beam is deflected by an angle of 2θ, producing a reflection spot in the
diffraction pattern. A sketch of an x-ray diffraction experiment is shown in Figure
97 in Section 6.1. The Bragg equation quantifies the relation between the lattice
constant, d, the scattering angle, θ, and the wavelength, λ:

2d sin Θ = nλ with n = 1; 2; 3; . . . (7:64)

A typical result for a polymeric sample reveals two types of x-ray refractions: a dif-
fuse halo and discrete reflexes on top. This means that the sample is composed of
both amorphous and crystalline domains, a phenomenon that is called partial
crystallinity.

7.3.2.2 Partial crystalline microstructures


An initial description of a polymer’s partially crystalline microstructure was that of
a fringed micelle structure. In this picture, crystalline domains are imagined to be
randomly distributed throughout the sample and connected by amorphous areas in
between, as shown in Figure 133. It was soon realized, however, that this is often
not the case. A noticeable example is native cellulose, whose partial crystallinity
7.3 Glassy and crystalline polymers 331

Figure 133: Initial conceptual picture of a partially crystalline polymer morphology named micellar
fringes; this is, however, very often untrue. Picture redrawn from B. Tieke: Makromolekulare
Chemie, Wiley VCH, 1997.

matches the structure of a fringed micelle. Most other crystalline polymer struc-
tures, though, are different.
How can we imagine polymer crystalline structures to look like then? To answer
this question, let us orient ourselves on a related kind of matter: long alkane oligom-
ers. How do these crystallize? They do so in the form of unfolded all-trans chains
arranged next to each other in lamellar platelets, whereby the oligomer-chain
end-groups sit on the frontal faces. By simple analogy, we may imagine a similar
kind of crystalline structures for longer polymers such as PE as well. However, in
that case, to fully unfold a polymer to an all-trans conformation in order to ar-
range it into such a hypothetic lamellar assembly, we need to overcome an ex-
tremely high entropy barrier. In addition, irregularities along the chain backbone
as well as entanglements with other chains prevent such perfect lamellar arrange-
ments to be formed over the full chain contour. Hence, we may presume that only
parts of the chain backbones are arranged in lamellar assemblies, whereas other
parts find themselves on the frontal faces, just like the end-groups of alkane
oligomers do in their lamellar crystalline phases. In experiments, it is commonly
found that the chain-folded lamellae of many kinds of polymers are roughly
10 nm thick. The reason is a free-energy balance, as shown in Figure 134. The
straightening of a polymer chain from a random coil in a polymer melt carries an
entropy penalty, TΔS, that needs to be overcome. The attachment of the straight-
ened chain segment to the crystalline domain, by contrast, carries an energy gain,
ΔH, that is comparable to the lattice energy that low-mass molecules gain when
they crystallize. Once the latter energy gain outweighs the first cost, crystalliza-
tion is possible. If the growing polymer crystal is too thick, however, the entropy
penalty is larger than the subsequent energy gain so that the total free-energy
332 7 States of polymer systems

Figure 134: Conceptual growth mechanism of a polymer crystal. First, a polymer chain segment
must be straightened and elongated from a random coil, which carries an entropy penalty of TΔS.
In a second step, the straightened chain segment can attach to a crystalline domain, which carries
an energy gain, ΔH. The overall free-energy balance is most favorable for lamellae that are 10 nm
tall. Picture redrawn from R. A. L. Jones: Soft Condensed Matter, Oxford University Press, 2002.

balance is unfavorable, and crystallization cannot happen. On the other hand, a


crystal that is too thin causes a too low energy gain in the second step, thereby
again producing a disfavorable free-energy balance. The optimum crystal size that
leads to an overall favorable energy balance is around 10 nm. This corresponds to
roughly 50 all-trans monomeric units.
The chain-folded lamellae constitute the basic building block for further higher
order crystal morphologies. The simplest kind of structure are platelets of about
10–20 μm width and, due to the aforementioned free-energy reasons, 10 nm height.
These structures are usually formed when crystals are grown from dilute solution,
and they may be seen as the closest match to a hypothetic “single crystal” of a poly-
mer. A schematic representation of such a polymer crystal platelet is shown in
Figure 135A.
In such platelets, the chain lamellae arrange themselves next to each other,
which requires backfolding of the polymer chains on the surface. For that, we need
about five gauche C–C bonds, creating a small amorphous domain. Depending on the
regularity of the 50 all-trans and 5 gauche bond sequences, the platelets have a more
or less ordered structure, as shown in Figure 135B. A very regular sequence leads to
strong adjacent backfolding, whereas a less regular sequence produces weak adja-
cent backfolding. If the ratio of 50:5 is missed by a bigger margin, then a chain might
not fold back to an adjacent other chain, but a random other one further away,
which is referred to as the switchboard model. This arrangement can still lead to
crystalline platelets, but the amorphous domain is much bigger in this case.
When multiple platelets find themselves stacked on top of each other and sepa-
rated by amorphous regions, they constitute stapled lamellae, as shown in Figure 136.
Such structures are commonly obtained upon crystallization from concentrated
7.3 Glassy and crystalline polymers 333

Figure 135: Chain-folded lamellae can form crystalline platelets. Panel (A) shows a schematic
representation of this structure. (B) Backfolding on the platelet surface requires about 5 gauche C–
C bonds for every 50 all-trans monomeric units. Depending on the regularity of this alternating
sequence, the adjacent backfolding can be either strong, weak, or irregular. Pictures redrawn from
B. Tieke: Makromolekulare Chemie, Wiley VCH, 1997.

Figure 136: Further higher order assembly of chain-folded lamellar subunits in the form of stapled
lamellae. Picture redrawn from J. M. G. Cowie: Chemie und Physik der synthetischen Polymeren,
Verlag Vieweg, 1996.

solution or melt. Their formation can be explained by the solidification model by


Erhard W. Fischer and Manfred Stamm, which presumes that crystallization hap-
pens by partial alignment of sections of the coils, thereby forming planes of mu-
tually close and regular arrangement bridged and linked by amorphous chain
sections in between. The advantage of that solidification mechanism is that it
does not rely on far-distant diffusive motion of the material.
Alternatively, the chain-folded lamellae can also be arranged in a radial fibrillar
fashion. The resulting structure is called a spherulite, whose fibrils are again com-
posed of chain-folded lamellae separated by layers of amorphous domains, as de-
picted in Figure 137. Such structures are also obtained upon crystallization from
concentrated solution or melt, with a growth mechanism as shown in Figure 138.
All the above examples show that polymer crystals are just partial, and all
these crystalline samples still contain substantial degrees of amorphous parts. The
334 7 States of polymer systems

Figure 137: Further higher order assembly of chain-folded lamellar subunits in the form of fibrillar
spherulites. Picture redrawn from J. M. G. Cowie: Chemie und Physik der synthetischen Polymeren,
Verlag Vieweg, 1996.

Figure 138: Schematic of the growth of a fibrillar spherulite.

Figure 139: (A) Fast cooling rates lead to kinetically trapped polymer coils that cannot rearrange
anymore. This effect massively disfavors a high degree of crystallinity. (B) Slow cooling rates that
allow for sufficient time for the polymer chains to arrange themselves into ordered structures allow
for a high degree of crystallinity. Picture redrawn from B. Vollmert: Grundriss der
Makromolekularen Chemie, Springer, 1962.
7.3 Glassy and crystalline polymers 335

overall degree of crystallinity within a solid polymer is highly dependent on the


rate of crystallization, which experimentally corresponds to the rate of cooling of
the sample. We have learned already that also the position of the glass-transition
temperature can be influenced this way (see Figure 132). During crystallization, the
polymer chains need a lot of time to arrange themselves in lamellae, and even lon-
ger to accomplish backfolding. When the rate of cooling is too fast, the glass-transi-
tion temperature is undercut before such an arrangement can come to pass. The
coil shape is then kinetically trapped in the glassy state with no or just little crystal-
line domains, as shown in Figure 139A. Below Tg there is not enough energy for
even single segmental motion, prohibiting the formation of further ordered struc-
tures. By contrast, hypothetic polymer single-crystal lamellae can only be assumed
to grow by slowly cooling down to, but not under, the glass-transition temperature.
That way, the polymer chains have enough time as well as enough energy to ar-
range themselves into chain-folded lamellae and/or higher order crystal morphologies,
as shown in Figure 139B. In general, the kinetic signature of polymer crystallization is
so important that the dominant crystal structure is not the one that has the most
favorable thermodynamic signature (i.e., the lowest free energy), but the most fa-
vorable kinetic signature (i.e., the highest rate of formation).

7.3.2.3 Experimental assessment of the degree of crystallinity


Apart from that conceptual understanding, how can we actually measure the de-
gree of crystallinity experimentally? There are two common ways. The first way is
based on the notion that the densities of the tightly packed crystalline and the
loosely packed amorphous parts of a partially crystalline polymer are different. We
may state that

V = Vcryst + Vamorph and m = mcryst + mamorph (7:65)

so that

Vρ = Vcryst ρcryst + Vamorph ρamorph (7:66)

This yields

Vcryst ρ − ρamorph
ϕcryst = = (7:67)
V ρcryst − ρamorph

and

mcryst Vcryst ρcryst ρcryst Vcryst ρcryst ρcryst ρ − ρamorph


wcryst = = = · = · ϕcryst = ·
m Vρ ρ V ρ ρ ρcryst − ρamorph
(7:68)
336 7 States of polymer systems

The latter pair of formulae allows us to determine the degree of crystallinity by mea-
surement of the sample density and knowledge of the densities in a purely crys-
talline and a purely amorphous reference. The first one can be calculated by
crystallographic means if we know the elementary cell, which we can determine
by knowledge of the chemical primary structure of the regular crystalline parts.
The second one can be assessed by measurement of the density of a glassy refer-
ence sample in which crystallization is suppressed, for example, by fast shock-
freezing.
A second way of experimental assessment is x-ray scattering. The scattering
profile of a partially crystalline polymer sample is composed of both distinct peaks
(caused by the crystalline domains) on top of a diffuse halo (caused by the amor-
phous parts). The degree of crystallinity is then simply determined from the areas
underneath these parts of the scattering curve:

Acryst:
wcryst: = (7:69)
Acryst + Aamorph

7.3.2.4 Structure–property relations for Tm


In addition to the abovementioned rate of cooling, the degree of crystallinity and
also the temperature where crystallization happens are also affected by chemical
factors, meaning by the type and constitution of the polymer backbone. In general,
on the one hand, a backbone with side groups that undergo strong interactions fa-
vors crystallization and entails a high Tm, but on the other hand, if that backbone is
not flexible enough (which it may be if the side groups interact too strongly), crys-
tallization is hampered, because then the polymer has a hard time to arrange itself
into the above-described chain lamellae. Yet on the other hand, if the chain back-
bone is too flexible, there are many different conformations possible in a melt state,
and giving up that liberty upon crystallization comes along with high-entropy cost
that hampers crystallization and enforces very low temperatures for it to eventually
become possible. Hence, a specific intermediate chain flexibility – not too little but
also not too high – is best for good crystallizability. Further factors affecting the
ability for polymer crystallization, and with that also the temperature point where
that happens, are alike those that have the same effect on the glass transition, as
we have already discussed in Section 7.3.1.3. This might be the presence of comono-
mer units or specifically branching, because branches act like “a solvent chemically
connected to the polymer.” As in fact both the glass-transition temperature and the
crystallization temperature are equally affected by these (and further) factors,
they are actually coupled, as expressed by the empirical Beaman–Bayer rule:
Tg ≈ 2/3 Tm.
7.3 Glassy and crystalline polymers 337

7.3.3 Comparison of the glass and the crystallization transition

To close this chapter, let us have a comparative look on the thermodynamic charac-
teristics of the glass transition and the crystal–melt transition of polymers, as
summarized in Figure 140. The uppermost row of schematic plots displays the
temperature course of the central state function, Gibbs’ free energy, G, during
these two processes. We see that whereas it exhibits a kink during the crystal–
melt transition at the melting temperature Tm, thereby denoting it to be a true
first-order phase transition, it is steady and smooth at the glass-transition temper-
ature Tg, similar to what second-order phase transitions show. For both, actual
discontinuities are seen in the first and/or second derivatives of the free energy,
as assessed by the temperature-dependent volume V (⁓ ∂G/∂p) and heat capacity
cp (⁓ ∂2G/∂T2) shown in the middle and lowermost rows. The curves that we have
discussed most in this chapter (and that are the easiest to be determined experi-
mentally) are the V(T) representations in the middle row, which actually show the
difference between the crystal–melt and the glass transition most strikingly. In
the crystal–melt transition, the system transitions from an ordered and perfectly
packed state at temperatures below Tm to an unordered and less packed one at
temperatures above Tm, thereby causing a distinct jump in V. With that, we have a
discontinuity in a first derivative of the thermodynamic potential G, as V ⁓ ∂G/∂p,
thereby indicating a first-order phase transition to be present. By contrast, during
the glass transition, our system transitions from a state with frozen dynamics at
temperatures below Tg to one with activated dynamics at temperatures above Tg,
which comes along with different thermal expansion coefficients on these two
sides, thereby only causing a kink or bend in the course of V(T). In this case, an
actual jump is seen only in one of the second derivatives of G, like in the heat ca-
pacity cp (⁓ ∂2G/∂T2).109 Although this signature is typical for second-order phase
transitions and might therefore indicate such one to be present here, the glass
transition does not show any microstructural change along with that, with a fun-
damental principle of any phase transition; it is therefore not an actual phase
transition but a kinetic phenomenon only.
For further comparison, Figure 141 summarizes the characteristic change of the
sample volume (which may also be assessed by its specific volume) of a purely
glassy, a partially crystalline, and a hypothetic purely crystalline polymer upon

 In practice, datasets like those shown in the lowermost row of Figure 140 are probed by differ-
ential scanning calorimetry, DSC. Like any technique, of course, that method has practical limita-
tions in resolution, and in a real experiment, we are also bound to practical constraints like the
need of conducting an experiment in a finite time, with no possibility to perform an extensively
slow heating or cooling ramp that allows for long thermal equilibration of the sample all the way
on its course. Hence, the dataset on the left of the lowest row in Figure 140 practically more looks
like the Greek letter lambda, whereas the dataset on the right practically looks like a step only.
338 7 States of polymer systems

Figure 140: Characteristics of the glass and crystallization transitions of polymers. Whereas
a crystallization process is a true first-order phase transition, accompanied by corresponding
discontinuities in the derivatives of the thermodynamic potential G, such as the volume, V (⁓ ∂G/∂p)
and the specific heat capacity, cp (⁓ ∂2G/∂T2), the glass transition is only a “pseudo”-second-order
phase transition. It exhibits some characteristics that also second-order phase transitions have, but
is actually a purely kinetic phenomenon.

heating. The glassy polymer shows the transition from a glassy to a melt state that
we have discussed in Section 7.3.1.2. The hypothetic purely crystalline polymer
would show a sharp first-order melting-phase transition as also sketched in the
middle left schematic of Figure 140; it would also show no glass transition below
the melting point, as its hypothetic perfect crystalline order in that domain would
not leave any amorphous material that could undergo a glass transition. An actual
sample, which has both crystalline and amorphous parts, first shows a glass-to-
melt transition of its amorphous parts followed by melting of its crystalline parts.
7.3 Glassy and crystalline polymers 339

Figure 141: Temperature-dependent (specific) volume of a fully amorphous polymer that only shows
a glassy state in the cold (blue curve), an hypothetic fully regular polymer that shows a perfect
crystalline state in the cold (green curve), and an actual polymer that shows both partial
crystallization upon cooling, followed by further vitrification of its remaining amorphous parts
upon further cooling (black curve).

This melting is not sharp but spans a range, as the crystalline domains in the sam-
ple are diverse in size and morphology, each showing an own melting temperature,
such that the overlay smears to a broad melting range. In the range between Tg and
Tm, the sample exhibits particularly interesting and favorable properties, as it has
both solid domains (these are the still unmolten crystalline domains) and liquid do-
mains (these are the already dynamically activated amorphous domains). As a re-
sult, the sample specimen is form-stable like a solid but also dissipative like a fluid,
thereby exhibiting good impact resistance.
340 7 States of polymer systems

Questions to Lesson Unit 23

(1) What is the difference between the glassy and crystalline states?
a. A sample in a glassy state is always transparent, whereas in a crystalline
state it is opaque.
b. A glassy state consists of amorphous structures, whereas a crystalline state
is ordered.
c. Glassy samples are often more brittle, whereas crystalline samples have bet-
ter mechanical durability.
d. A glassy sample has a lower melting point than a crystalline sample.

(2) How do polymers usually solidify?


a. They usually solidify in a crystalline state because it is the energetically
most favorable structure.
b. They usually solidify in a crystalline state because it allows the closest
packing of chain segments.
c. They usually solidify in a glassy state because the disordered chains “freeze”
when lowering the temperature, which is called vitrification.
d. They usually solidify in a glassy state because it is the entropically most fa-
vorable structure.

(3) What holds for the glass transition of a sample with entanglements and cross-
links?
a. The glass transition is the transition from a glassy state with elastic proper-
ties to a viscous melt.
b. The glass transition is the transition from a glassy state with energy-elastic
properties to an entropic-elastic rubber state.
c. The glass transition is the transition from a less-ordered glassy to a higher
ordered crystalline state.
d. The glass transition does not take place in entangled polymers.

(4) Which statement regarding the glass-transition temperature is correct?


The glass-transition temperature ______
a. is actually a temperature range in most cases.
b. marks the point of transition from an amorphous glassy state to an ordered
crystalline one.
c. is only noticeable in spectra when a slow cooling process is provided during
measurement.
d. only exists for highly cross-linked polymer samples.
Questions to Lesson Unit 23 341

(5) Is the glass transition a phase transition?


a. Yes, it is a phase transition of first order.
b. Yes, it is a phase transition of second order.
c. No, but it has similar characteristics to a second-order phase transition.
d. None of the above statements is applicable.

(6) What is not an important parameter when talking about the concept of glass
transition?
a. The different contributions to the sample’s volume.
b. The relaxation dynamics of chains and their segments.
c. The activation energies for chain segment’s displacements.
d. The entanglement molar mass.

(7) How does the free volume change with increasing temperature?
a. For T < Tg the free volume is constant, and for T > Tg it increases linearly.
b. For T < Tg the free volume is constant, and for T > Tg it increases expo-
nentially.
c. For T < Tg the free volume decreases exponentially, and for T > Tg it is
constant.
d. For T < Tg the free volume decreases linearly, and for T > Tg it is constant.

(8) What shows that the glass transition is a kinetic process far away from
equilibrium?
a. Even beyond the glass-transition temperature, the specific volume of the
sample further decreases with time even without further cooling; this is
called aging.
b. The glass-transition temperature is independent of the rate of cooling and
just depends on the type of polymer.
c. As soon as the glass-transition temperature is reached, the sample remains
in the glassy state unless the temperature is changed.
d. The vitrification actually starts at temperatures higher than the glass-transi-
tion temperature and is completed when the glass-transition temperature is
reached.
8 Closing remarks

The target of this book is to deliver an understanding of the relations between


structure and properties of polymers, thereby bridging the fields of polymer
chemistry, which focuses on the primary molecular-scale structure of individual
polymer chains, and polymer engineering, which focuses on the macroscopic
properties of polymer materials. With that arc, molecular parameters are ratio-
nally connected to macroscopic function. As a fundament for this bridge, the
first chapters of this book have considered the shape of a single ideal polymer coil
using models such as the phantom chain model and the Kuhn model. From that, we
have recognized that the statistics of the ideal chain are those of a random walk,
and we have used this notion to quantify the free energy and the mechanical char-
acteristics upon deformation of such ideal chains. We have then moved further to
incorporate chain–chain and chain–solvent interactions into our model, thereby fo-
cusing on real polymer chains; this enabled us to categorize solvents based on their
ability to dissolve and swell a polymer. We have realized that when all interactions
exactly balance each other, the real chain acts like an ideal chain, and we have
called this quasi-ideal state the Θ-state. In a next step, we have developed a thermody-
namic theory for polymer solutions, the Flory–Huggins theory, based on a mean-field
approach, and with it we were able to construct a full phase diagram of a polymer
solution or a polymer blend. Furthermore, we have learned about two conceptual ap-
proaches to model the dynamics and motion of polymer chains: the Rouse and the
Zimm model.
While doing all of the above, we have often encountered striking similarities of
the concepts that we have employed to those known from elementary physical chem-
istry. These similarities appeared both on the conceptual and on the mathematical
side of our argumentations. Conceptually, for example, the description of an ideal
polymer chain and an ideal gas was based on the same assumptions. The same re-
semblance was also retained for real polymer chains and real gases, which are both
subject to mutual and environmental interactions. As a consequence, both materials
display quasi-ideal states at a very specific temperature, the Θ-temperature for poly-
mer chains and the Boyle temperature for gases. Mathematically, we have seen simi-
larities between the end-to-end distance distribution within a Gaussian polymer coil,
the Boltzmann and the Maxwell–Boltzmann distributions of kinetics of gas particles,
and the electron density distribution in an s-orbital. We have also seen how random-
walk statistics that describes the diffusive motion of a particle or molecule also ap-
plies to the statistical treatment of the shape of ideal polymer chains.
Methodically, we have touched upon two fundamental approaches that are used
in polymer physical chemistry. First, we have learned how a multibodied complex
system can be simplified by only considering average values. This is the mean-field
approach. Second, we are now familiar with the concept of scaling laws. Realizing

https://doi.org/10.1515/9783110713268-008
344 8 Closing remarks

that polymers are self-similar objects, we can rescale our conception of them to differ-
ent length-scales without having to change the mathematical description of them, as
it is done in the blob concept.
When we reconsider the conceptualized goal of this book, to build a bridge be-
tween polymer chemistry and polymer engineering, we may conclude that the first
four chapters of this book have laid its pillars. Based upon that, the actual construc-
tion of the bridge has been outlined in the fifth chapter of this book. In this central
chapter, we could rationally and quantitatively understand why many of the poly-
meric materials that we encounter in our everyday lives behave the way they do.
Our prime focus in this consideration was on those properties of polymers that have
made greatest impact on our lives in the past decades, which is their mechanical
properties, spanning from viscous flow to elastic snap, along with their time and
temperature dependence. To complete our picture, this book has then added a sixth
chapter on scattering methods as a prime experimental platform to characterize
nano- and microstructures of polymer systems. A final seventh chapter eventually
showed us how the conceptual grounds of the preceding ones apply to some rele-
vant states of appearance of polymer samples and systems. With that, we have
closed the circle of our endeavor of deriving relations between the structure of a
polymer system and its most relevant properties.
Index
active microrheology 181 effective interactions 74
adaptivity 26 Einstein–Smoluchowski equation 96
anionic polymerization 13 elastic modulus 160
atactic 8 elastic scattering 257
athermal solvent 77 elastic solid 159
attractive interactions 72 endothermic mixing 119
autocorrelation 264 end-to-end distance 35
energy of mixing 20
bad solvent 78 entanglement 20
Bingham fluid 176 entanglement molar mass 226
binodal line 137 entanglement time 210
biopolymers 3 entropic spring constant 59
blob concept 62 entropy elasticity 59
blob size 62 entropy of mixing 20
block copolymer 124 excluded volume 71
Boltzmann distribution 43 exothermic mixing 119
Boltzmann superposition principle 199
bonding angle 32 Flory exponent 65
Bragg diffraction 244 Flory theory 85
Bragg equation 245 flow 20
flow curve 175
caoutchouc 23 form factor 247
Casson fluid 176 fractal dimension 65
cellulose 23 free volume 201
centrifugal-average molar mass 9 freedom 60
chain-growth 11 freely-jointed chain 37
chain-transfer agent 13 free-radical polymerization 11
characteristic ratio 40
cohesive energy 21, 126
Gaussian chain 57
coil expansion 83
glass transition temperature 1
colligative properties 10
good solvent 77
colloidal 20
Guinier function 251
constraint release 239
Guinier region 250
creep 20
creep compliance 169
creep test 168 hard-sphere repulsion 75
Hooke’s law 160
hydrodynamic interactions 100, 283
De Broglie wavelentgh 243
hydrodynamic radius 97
deformation blob 90
degree of swelling 227
demixing 131 ideal chain 31
diffusion coefficient 97 ideal mixture 119
dynamic experiment 169 intensity autocorrelation function 264
dynamic light scattering 264 interaction parameter 119

https://doi.org/10.1515/9783110713268-009
346 Index

irregularity 7 polydispersity index 11


isotactic 7 polyethylene 5
polymer blend 114
Kelvin–Voigt model 191 polymer solution 114
Kuhn length 42 polymerization 5
polypropylene 32
lattice model 113 polystyrene 32
Lennard–Jones potential 72 polyvinylchloride 32
living polymerization 13 Porod’s law 252
loss modulus 171 power law 1
loss tangent 171 precipitation 19, 78
lower critical solution temperature 123 primitive path 234
properties 1
macroconformation 32
macromolecule 3 quasi-living polymerization 13
Mark–Houwink–Sakurada equation 287
Maxwell model 184 radius of gyration 35
Mayer f-function 75 Raman scattering 257
mean-field theory 112 random chain 37
mean-square 38 random coil 32
mean-square displacement 96 random walk 38
metastable 135 rational design 1
microconformation 32 Rayleigh ratio 247
microphase separation 124 Rayleigh scattering 257
monomer 3 real chain 71
reference temperature 203
Newtonian flow 174 regular solution 114
Newtonian fluid 20 relaxation 20
Newton’s law 160 relaxation experiment 168
nonsolvent 78 relaxation modes 102
nucleation and growth 136 relaxation modulus 168
number-average molar mass 9 relaxation time 21, 185
relaxation time spectrum 100
osmotic pressure 10, 148 reptation 211, 234
reptation time 211
passive microrheology 181 repulsive interactions 72
persistence length 46 responsiveness 26
phantom chain 37 retardation time 192
phase diagram 131 rheology 159
Poisson distribution 13 root-mean-square 38
Poisson effect 162 rotational isomeric states 43
Poisson’s ratio 162 Rouse model 98
poly(methylmethacrylate) 40 Rouse time 99
poly(sodium acrylate) 40 rubbery plateau 218
polyaddition 11 rubbery-elastic plateau 210
polyamide 5
polycondensation 6 scaling discussion 64
polydispersity 9 scaling law 1
Index 347

scattering intensity 243 swelling 227


scattering vector 245 syndiotactic 7
Schulz–Flory distribution 12
Schulz–Zimm distribution 11 tacticity 7
screening 106 theta-solvent 77
second virial coefficient 150 theta-state 76
segmental lentgh 32 time–temperature superposition 202
self-assembly 26 torsion angle 32
self-avoiding walk 71 tube concept 234
self-similar 64
sensitivity 26 ultracentrifugal analysis 11
shear modulus 162 universality 1
shear rheology 180 uphill diffusion 143
shear-thickening 176 upper critical solution temperature 140
shear-thinning 175
shift factor 202 Van’t Hoff’s law 149
soft condensed matter 1 vinyl 7
soft matter 20 virial series expansion 150
specificity 1 viscoelastic liquid 172
spinodal decomposition 136 viscoelastic solid 172
spinodal line 137 viscoelasticity 21
static light scattering 10, 255 viscosity 161
Staudinger index 284 viscous liquid 159
step-growth 11
Stokes–Einstein equation 97 weight-average molar mass 9
storage modulus 171
Williams–Landel–Ferry equation 203
strain 160
strain-induced hardening 226
yield stress 176
stress 160
Young’s modulus 160
stress relaxation 168
structural viscosity 175
zero shear viscosity 145
structure 1
Zimm model 98
structure factor 247
Zimm plot 251
structure–property relation 34, 45, 48, 58
Zimm time 101
subdiffusion 104
superdiffusion 104

You might also like