05.10.2013 Views

Polymers in Confined Geometry.pdf

Polymers in Confined Geometry.pdf

Polymers in Confined Geometry.pdf

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Diploma Thesis<br />

<strong>Polymers</strong> <strong>in</strong> Conf<strong>in</strong>ed <strong>Geometry</strong><br />

Frederik Wagner<br />

December 31, 2004<br />

Supervisor: Prof. Dr. Erw<strong>in</strong> Frey<br />

Freie Universität Berl<strong>in</strong><br />

Fachbereich Physik<br />

Hahn-Meitner-Institut Berl<strong>in</strong><br />

Abteilung Theorie (SF5)


Contact: frederik.wagner@gmx.de


TMTOWTDI—There’s More Than One Way To Do It<br />

—Larry Wall


Thanks to. . .<br />

First of all I have to thank Prof. Dr. Erw<strong>in</strong> Frey arous<strong>in</strong>g my <strong>in</strong>terest <strong>in</strong> biological<br />

physics and giv<strong>in</strong>g me the opportunity to write my diploma thesis <strong>in</strong> his groups<br />

at the Hahn-Meitner-Institut Berl<strong>in</strong>. Also for very <strong>in</strong>terest<strong>in</strong>g discussions—and<br />

some nice chats <strong>in</strong> between—and the huge amount of help by writ<strong>in</strong>g this thesis.<br />

A big thanks to the small Italian. . . please, don’t kick me. . .<br />

That is, a big thanks to Dottore Gianluca Lattanzi for: <strong>in</strong>troduc<strong>in</strong>g and help<strong>in</strong>g<br />

me with my first steps <strong>in</strong> Monte-Carlo simulations, for be<strong>in</strong>g allowed to drop <strong>in</strong><br />

constantly and go<strong>in</strong>g onto his nerves, for encourag<strong>in</strong>g me, for just hav<strong>in</strong>g fun and<br />

hav<strong>in</strong>g some good talks. . . Hope to see you aga<strong>in</strong> soon 1 . . .<br />

Thanks to Sebastian Fischer as a ‘Leidensgenosse’ and for some <strong>in</strong>terest<strong>in</strong>g<br />

new ideas dur<strong>in</strong>g our canteen-boycott.<br />

And the rest of SF5 for all the nice discussions and coffee breaks. You know<br />

who you are.<br />

For the collaboration on the experimental side, I have to thank Walter Reisner<br />

of Bob Aust<strong>in</strong>s groups at Pr<strong>in</strong>ceton.<br />

Thanks to Dr. Axel Pelster for many <strong>in</strong>terest<strong>in</strong>g lectures and the support<br />

dur<strong>in</strong>g my studies at the Freie Universität Berl<strong>in</strong>.<br />

F<strong>in</strong>ally, I have to thank my parents for so much. . . !<br />

1 “God will<strong>in</strong>g!”<br />

i


Contents<br />

1 Introduction 1<br />

2 Polymer models 5<br />

2.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2.2 Ideal phantom cha<strong>in</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.3 Semi-flexible polymers . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.3.1 Kratky-Porod model . . . . . . . . . . . . . . . . . . . . . 9<br />

2.3.2 Worm-like cha<strong>in</strong> model . . . . . . . . . . . . . . . . . . . . 10<br />

2.4 Stiff polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.4.1 Weakly-bend<strong>in</strong>g rod approximation . . . . . . . . . . . . . 14<br />

2.5 Real cha<strong>in</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

3 <strong>Polymers</strong> <strong>in</strong> conf<strong>in</strong>ed geometry 21<br />

3.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

3.2 Scal<strong>in</strong>g relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

3.2.1 The flexible regime . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.2.2 The stiff regime . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.3 Analytical results . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

3.3.1 Averages by Fourier transformation . . . . . . . . . . . . . 28<br />

3.3.2 Averages by Fourier series . . . . . . . . . . . . . . . . . . 34<br />

3.3.3 Radial distribution function . . . . . . . . . . . . . . . . . 36<br />

4 Simulation methods 37<br />

4.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.2 The Monte-Carlo method . . . . . . . . . . . . . . . . . . . . . . 38<br />

4.3 Sampl<strong>in</strong>g the canonical ensemble . . . . . . . . . . . . . . . . . . 39<br />

4.3.1 Markov cha<strong>in</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

4.3.2 Metropolis algorithm . . . . . . . . . . . . . . . . . . . . . 41<br />

4.3.3 Correlations and error calculation . . . . . . . . . . . . . . 42<br />

4.3.4 Recursive averag<strong>in</strong>g and error calculation . . . . . . . . . . 46<br />

4.3.5 Random number generators . . . . . . . . . . . . . . . . . 47<br />

4.4 Simulation of unconf<strong>in</strong>ed worm-like cha<strong>in</strong>s . . . . . . . . . . . . . 47<br />

4.4.1 Initial part of the simulation . . . . . . . . . . . . . . . . . 50<br />

iii


iv CONTENTS<br />

4.4.2 Self-avoid<strong>in</strong>g cha<strong>in</strong>s . . . . . . . . . . . . . . . . . . . . . . 51<br />

4.5 Simulation of conf<strong>in</strong>ed worm-like cha<strong>in</strong>s . . . . . . . . . . . . . . . 51<br />

4.5.1 Harmonic potential . . . . . . . . . . . . . . . . . . . . . . 52<br />

4.5.2 Hard-wall tubes . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

4.6 Sampled quantities and their relevance . . . . . . . . . . . . . . . 53<br />

5 Simulation results 57<br />

5.1 The unconf<strong>in</strong>ed cha<strong>in</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

5.2 The harmonic potential . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

5.2.1 Undulations . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

5.2.2 Tangent-tangent correlation . . . . . . . . . . . . . . . . . 61<br />

5.2.3 Scal<strong>in</strong>g plot . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

5.2.4 End-to-end distances . . . . . . . . . . . . . . . . . . . . . 69<br />

5.2.5 Radial distribution function . . . . . . . . . . . . . . . . . 70<br />

5.3 The hard walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

6 Conclusion 75<br />

A Discrete worm-like cha<strong>in</strong> correlations 77<br />

B Calculation for the conf<strong>in</strong>ed polymer 81<br />

B.1 Undulation correlations . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

B.2 End-to-end distance correlation . . . . . . . . . . . . . . . . . . . 82<br />

B.3 Parallel end-to-end distance correlation . . . . . . . . . . . . . . . 84<br />

Glossary 87<br />

Bibliography 89


Chapter 1<br />

Introduction<br />

<strong>Polymers</strong> are one of the ma<strong>in</strong> constituents of cells. The cytoskeleton is formed by<br />

a set of prote<strong>in</strong> filaments (microtubules, act<strong>in</strong> filaments, <strong>in</strong>termediate filaments).<br />

It provides mechanic stability and a railway network for <strong>in</strong>tracellular transport.<br />

Genetic <strong>in</strong>formation is also stored <strong>in</strong> a macromolecule, deoxyribonucleic acid<br />

(DNA). As the basic ‘cookbook’ of life it provides the recipe for sequenc<strong>in</strong>g all<br />

the other biopolymers.<br />

DNA is not only carrier of genetic <strong>in</strong>formation, but its mechanic properties<br />

(such as bend<strong>in</strong>g and torsional stiffness) are also essential for how this <strong>in</strong>formation<br />

is processed. For example some processes <strong>in</strong> gene expression are regulated by loop<br />

formation and other conformational changes of DNA.<br />

These few examples are supposed to give a flavor of the manifold role biopolymers<br />

play <strong>in</strong> cellular processes (cf. [3, 5, 6, 14]).<br />

Biopolymers differ from synthetic polymers <strong>in</strong> one important aspect. They are<br />

stiff on length scales relevant for the biophysical processes they are <strong>in</strong>volved <strong>in</strong>.<br />

This stiffness is characterized <strong>in</strong> terms of the persistence length which measures<br />

the length scale (along the backbone) over which correlations of the tangents<br />

(describ<strong>in</strong>g the polymer configuration) decay. For DNA the persistence length is<br />

approximately lp ≈ 50 nm. This is much larger than a typical microscopic length<br />

scale of DNA, e.g. the diameter h ≈ 2 nm:<br />

lp ≫ h semi-flexible polymer.<br />

F-Act<strong>in</strong> is even stiffer, with lp ≈ 17 nm and h ≈ 7 nm. This unique feature<br />

gives rise to a multitude of <strong>in</strong>terest<strong>in</strong>g phenomena genu<strong>in</strong>ely different from those<br />

found for their synthetic cous<strong>in</strong>s (such as polyethylene), where lp is comparable<br />

to the microscopic scale h (cf. [17, 31]).<br />

Our focus <strong>in</strong> this thesis will be on biopolymers <strong>in</strong> conf<strong>in</strong>ed geometry. This is<br />

motivated by rapid grow<strong>in</strong>g <strong>in</strong>terest <strong>in</strong> observ<strong>in</strong>g and manipulat<strong>in</strong>g s<strong>in</strong>ge polymer<br />

cha<strong>in</strong>s <strong>in</strong> biotechnological applications us<strong>in</strong>g micro- and nanofluidic devices.<br />

Us<strong>in</strong>g optical tweezers, it is possible to exert small forces on polymers. By<br />

1


2 CHAPTER 1. INTRODUCTION<br />

(a) 100 nm channel width (b) 200 nm channel width<br />

Figure 1.1: Scann<strong>in</strong>g electron micrograph (SEM) picture of nanochannels. These channels<br />

are used <strong>in</strong> experiments to study the behavior of DNA under <strong>in</strong>creas<strong>in</strong>g conf<strong>in</strong>ement. Courtesy<br />

of W. Reisner et al. [38].<br />

real time optical microscopy on fluorescently dyed polymers a direct observation<br />

method of polymer cha<strong>in</strong>s has become available.<br />

One important task is the localization of b<strong>in</strong>d<strong>in</strong>g sites for restriction and<br />

other DNA b<strong>in</strong>d<strong>in</strong>g enzymes. The restriction sites are traditionally determ<strong>in</strong>ed<br />

by measur<strong>in</strong>g the length of restriction fragments by gel-electrophoresis. Alternatively,<br />

they can be located by us<strong>in</strong>g optical mapp<strong>in</strong>g of stretched DNA molecules<br />

trapped on a surface. To measure the contour-length of a s<strong>in</strong>gle molecule us<strong>in</strong>g<br />

optical techniques directly, it is necessary to extend the polymer such that<br />

a one-to-one mapp<strong>in</strong>g can be established between the spatial position along the<br />

polymer and the position with<strong>in</strong> the genome.<br />

Chip-based devices, that cannot only detect and separate s<strong>in</strong>gle DNA molecules<br />

by size but are able to sequence at the s<strong>in</strong>gle molecule level, would have the<br />

potential to revolutionize biology.<br />

Conf<strong>in</strong>ement elongation of genomic-length DNA has several advantages over<br />

alternative techniques for extend<strong>in</strong>g DNA, such as flow stretch<strong>in</strong>g and/or stretch<strong>in</strong>g<br />

rely<strong>in</strong>g on a tethered molecule. Conf<strong>in</strong>ement elongation does not require the<br />

presence of a known external force because a molecule <strong>in</strong> a nanochannel will rema<strong>in</strong><br />

stretched <strong>in</strong> its equilibrium configuration. Second, it allows for a precise<br />

measurement of length at the s<strong>in</strong>gle molecule level (cf. [44]).<br />

While a number of approaches have been proposed, all of them have <strong>in</strong> common<br />

the conf<strong>in</strong>ement of DNA to nanometer scales, typically 5-200 nm. Conf<strong>in</strong>ement<br />

alters the statistical mechanical properties of DNA. A DNA molecule <strong>in</strong> a<br />

nanochannel (cf. figures 1.1) will extend along the channel axis to a substantial<br />

fraction of its full contour length (cf. figures 1.2). While the study of conf<strong>in</strong>ed<br />

DNA is <strong>in</strong>terest<strong>in</strong>g from a physical perspective, it is also critical for device design,<br />

potentially lead<strong>in</strong>g to new applications of nanoconf<strong>in</strong>ement, for example, the use<br />

of nanochannels to pre-stretch and stabilize DNA before thread<strong>in</strong>g through a


(a) λ-DNA (b) T2-DNA<br />

Figure 1.2: Average <strong>in</strong>tensity of fluorescently dyed DNA (persistence length of undyed DNA<br />

lp ≈ 52 nm and contour-length of λ-DNA L ≈ 16.5 µm and T2-DNA L ≈ 55.8 µm) <strong>in</strong> nanochannels<br />

of decreas<strong>in</strong>g size (aspect ratio near one, width 440 nm down to 30 nm). The DNA stretches<br />

to near full its length upon conf<strong>in</strong>ement to sizes around its persistence length. Courtesy of W.<br />

Reisner et al. [38].<br />

nanopore (cf. [38]).<br />

To turn this idea <strong>in</strong>to a useful device a quantitative theoretical understand<strong>in</strong>g<br />

of polymer conformations <strong>in</strong> conf<strong>in</strong>ed geometry is needed. Previous models (discussed<br />

<strong>in</strong> more detail <strong>in</strong> chapter 2 and section 3.2) were unable to account for the<br />

effect of vary<strong>in</strong>g conf<strong>in</strong>ement over the entire range of scales used <strong>in</strong> nano-devices.<br />

Only asymptotic results valid <strong>in</strong> special limits were available. In this thesis we<br />

try to close this gap.<br />

We will proceed as follows:<br />

• A general <strong>in</strong>troduction to specific models and universal behavior of unconf<strong>in</strong>ed<br />

polymers is given <strong>in</strong> chapter 2. We present a derivation of exactly<br />

known results as well as some scal<strong>in</strong>g arguments. This chapter also serves<br />

as a concise <strong>in</strong>troduction <strong>in</strong>to the field of study and <strong>in</strong>troduces our basic<br />

polymer model, the worm-like cha<strong>in</strong>, used <strong>in</strong> the subsequent analysis.<br />

• In chapter 3 we beg<strong>in</strong> our discussion of polymer configurations conf<strong>in</strong>ed to<br />

tubes (represented <strong>in</strong> terms of a harmonic potential with cyl<strong>in</strong>drical symmetry).<br />

After discuss<strong>in</strong>g some scal<strong>in</strong>g arguments we present an analytical<br />

calculation <strong>in</strong> the limit of strong conf<strong>in</strong>ement. This allows us to give explicit<br />

expressions for the correlation function and end-to-end distances.<br />

• S<strong>in</strong>ce it is not possible to obta<strong>in</strong> analytical expressions for all relevant parameter<br />

ranges it becomes necessary to perform Monte-Carlo simulations<br />

<strong>in</strong> order to arrive at quantitative results. Chapter 4 will give a general <strong>in</strong>troduction<br />

to the basic pr<strong>in</strong>ciples needed for the simulations and a detailed<br />

3


4 CHAPTER 1. INTRODUCTION<br />

description of the methods used for simulat<strong>in</strong>g polymer cha<strong>in</strong>s.<br />

• The discussion of the simulations performed <strong>in</strong> the context of this thesis are<br />

f<strong>in</strong>ally discussed <strong>in</strong> chapter 5. The algorithm developed <strong>in</strong> the preced<strong>in</strong>g<br />

chapter is validated on known results for the unconf<strong>in</strong>ed polymer. A general<br />

scal<strong>in</strong>g behavior is worked out by the systematic <strong>in</strong>vestigation of the<br />

simulation data and a comparison to the analytical results of chapter 3<br />

is given. Different conf<strong>in</strong><strong>in</strong>g geometries are <strong>in</strong>vestigated <strong>in</strong> comparison to<br />

experimental data.<br />

• F<strong>in</strong>ally a summary and a short outlook is given <strong>in</strong> our f<strong>in</strong>al chapter 6.


Chapter 2<br />

Polymer models<br />

This chapter <strong>in</strong>troduces some of the basic models used <strong>in</strong> polymer physics. After<br />

a short motivation of why us<strong>in</strong>g a m<strong>in</strong>imal model—only keep<strong>in</strong>g the most important<br />

features—is possible, we turn to a detailed description of specific realizations<br />

of these models. We present both heuristic scal<strong>in</strong>g arguments and analytical calculations<br />

<strong>in</strong> limit<strong>in</strong>g cases. We conclude by a short discussion of when and how<br />

to <strong>in</strong>clude more realistic effects.<br />

2.1 Motivation<br />

Condensed matter physics has a long stand<strong>in</strong>g tradition of reductionists approaches.<br />

In order to describe the generic properties of a system one tries to<br />

skip as much details as possible and keep only the essential features. How many<br />

details have to be kept for a specific description depends on the length-scale on<br />

which a problem is <strong>in</strong>vestigated and is not always obvious to decide.<br />

Polymer physics is a paradigmatic example for such an approach. A biopolymer<br />

consists of a large number of backbone monomers, i.e. molecules (e.g. DNA<br />

base-pairs) normally consist<strong>in</strong>g of several atoms (e.g. carbon, hydrogen). These<br />

atoms aga<strong>in</strong> consist of electrons and nuclei build of quarks and so on. Naively<br />

one might th<strong>in</strong>k that an exact model should actually start at the lowest possible<br />

level. But this is neither feasible nor necessary s<strong>in</strong>ce there is usually a separation<br />

of time- and length scales between a microscopic description and the mesoscopic<br />

level we are <strong>in</strong>terested <strong>in</strong>.<br />

For many applications it is even possible to disregard all atomistic details and<br />

consider each monomer as a structureless segment of a given length l. This is<br />

motivated by the fact that molecular processes with<strong>in</strong> each monomer relax on<br />

a much shorter time scale than the overall polymer conformations. Interactions<br />

between successive segments are modeled such that its essential properties are<br />

kept, e.g. bend<strong>in</strong>g and torsional stiffness of the thread. The basic models used <strong>in</strong><br />

polymer physics are dist<strong>in</strong>ct at this level, for a thorough discussion we refer to<br />

5


6 CHAPTER 2. POLYMER MODELS<br />

i = 0<br />

R<br />

(a) Cha<strong>in</strong> of segments<br />

i = N<br />

s = 0<br />

R<br />

(b) Smooth space curve<br />

Figure 2.1: The two basic polymer models for cha<strong>in</strong>s with stiffness.<br />

s = L<br />

one of the standard text books, e.g. [9].<br />

How should one then picture a polymer if no unnecessary details are kept?<br />

This depends on the mathematical description one would like to use for the<br />

theoretical analysis. For computer simulations a cha<strong>in</strong> of segments with some<br />

prescribed <strong>in</strong>teractions between the segments is most appropriate (see fig. 2.1(a)).<br />

For analytical treatment one might want to use a cont<strong>in</strong>uous <strong>in</strong>stead of a discrete<br />

model. Here we have to be very careful how to perform such a limit.<br />

Naively on might expect that upon tak<strong>in</strong>g the number of segments to <strong>in</strong>f<strong>in</strong>ity<br />

(N → ∞) and the size of the segments to zero (l → 0) one arrives at a cont<strong>in</strong>uous<br />

smooth space curve. But, this is not true <strong>in</strong> general! As we will discuss <strong>in</strong> more<br />

detail <strong>in</strong> the follow<strong>in</strong>g section, the proper cont<strong>in</strong>uum limit of a flexible cha<strong>in</strong><br />

(when the segments are connected by “free h<strong>in</strong>ges”) is not a smooth space curve<br />

but a fractal which is cont<strong>in</strong>uous but non differentiable at any po<strong>in</strong>t along the<br />

cha<strong>in</strong>. The proper way to describe such a trajectory <strong>in</strong> space has been <strong>in</strong>troduced<br />

by Wiener <strong>in</strong> the aftermath of E<strong>in</strong>ste<strong>in</strong>s discussion of random walks (see e.g. [16]).<br />

It is only if one assigns some <strong>in</strong>ternal stiffness to a cha<strong>in</strong> that a typical conformation<br />

of the polymer will look like a smooth space curve as depicted <strong>in</strong> figure<br />

2.1(b). This will lead us to the model of a worm-like cha<strong>in</strong> <strong>in</strong> section 2.3.2.<br />

This short <strong>in</strong>troduction already shows that even on this abstract level the<br />

problem is very complex. This thesis will be limited only to the static properties<br />

of polymer systems. The area of dynamic properties reveals other <strong>in</strong>terest<strong>in</strong>g and<br />

complex properties and will be the subject of future studies.<br />

2.2 Ideal phantom cha<strong>in</strong><br />

The most basic model describ<strong>in</strong>g the universal properties of a large class of polymers<br />

is the freely jo<strong>in</strong>ted cha<strong>in</strong> (FJC) model. It is a cha<strong>in</strong> consist<strong>in</strong>g of successive<br />

segments ti of length |ti| = l connected by free h<strong>in</strong>ges. As a first step all correlations<br />

between the segments are ignored, which is reasonable only for very flexible<br />

polymers.<br />

One of the ma<strong>in</strong> quantities of <strong>in</strong>terest is the end-to-end distance of a cha<strong>in</strong>


2.2. IDEAL PHANTOM CHAIN 7<br />

configuration<br />

R =<br />

N<br />

ti, (2.1)<br />

i=1<br />

with N ≫ 1. We will be ma<strong>in</strong>ly <strong>in</strong>terested <strong>in</strong> the limit of long cha<strong>in</strong>s.<br />

The first two moments of the end-to-end vector distribution are obta<strong>in</strong>ed<br />

quite readily. The mean of the distribution 〈R〉 must be zero because of spacial<br />

isotropy. The mean square end-to-end distance 〈R 2 〉 is obta<strong>in</strong>ed <strong>in</strong> a straightforward<br />

calculation due to the absence of correlations<br />

R 2 =<br />

N<br />

i=1<br />

N<br />

〈ti · tj〉 =<br />

j=1<br />

N<br />

i=1<br />

N<br />

j=1<br />

δij<br />

t 2 = N t 2 = Nl 2 , (2.2)<br />

where we have used the fact that subsequent segments are uncorrelated. In<br />

summary we have:<br />

〈R〉 = 0, (2.3)<br />

R 2 = Nl 2 . (2.4)<br />

From eq. (2.4) we note that a proper cont<strong>in</strong>uum limit of the FJC needs to take<br />

the limits N → ∞, l → 0 such that Nl 2 stays f<strong>in</strong>ite.<br />

Once we know these moments we can already write down the full probability<br />

distribution of the end-to-end distance by exploit<strong>in</strong>g the central limit theorem.<br />

It states that the limit<strong>in</strong>g distribution of a random variable, which is the sum of<br />

<strong>in</strong>dependent random variables, is a Gaussian 1 with mean 〈R〉 and width 〈R 2 〉<br />

p(R) =<br />

<br />

3<br />

2π 〈R2 3/2 <br />

exp −<br />

〉<br />

3R2<br />

2 〈R2 <br />

. (2.5)<br />

〉<br />

This result br<strong>in</strong>gs us back to the random walk s<strong>in</strong>ce eq. (2.5) may be <strong>in</strong>terpreted<br />

as the Greens function of the diffusion equation with the <strong>in</strong>itial condition p(R, t =<br />

0) = δ(R).<br />

Due to the central limit theorem, universal behavior for a larger class of<br />

flexible polymers is expected. As long as correlation between the segments along<br />

the cha<strong>in</strong> decay on a scale small compared to the total length these correlation<br />

can be subsumed <strong>in</strong> an effective segment length or Kuhn segment b, which is<br />

def<strong>in</strong>ed by the universal relation to hold:<br />

R 2 = b 2 N ⇒ R ∼ bN 1/2 . (2.6)<br />

For the FJC, b = l the Kuhn length is identical to the segment length. For an<br />

actual cha<strong>in</strong> the correlations are renormalized by look<strong>in</strong>g at a larger scale at the<br />

1 Of course the probability distribution function can also be obta<strong>in</strong>ed by calculation from<br />

statistical mechanics.


8 CHAPTER 2. POLYMER MODELS<br />

cha<strong>in</strong>, which is possible only when it is long enough. The end-to-end distance for<br />

this class of polymers scales with a power law <strong>in</strong> the number of segments N with<br />

an exponent ν = 1/2.<br />

This model can be used for analyz<strong>in</strong>g scatter<strong>in</strong>g experiments. In particular <strong>in</strong><br />

small angle scatter<strong>in</strong>g the mean-square radius of gyration R2 <br />

g can be directly<br />

measured. It is def<strong>in</strong>ed by<br />

R 2 g = 1<br />

N<br />

N<br />

i=1<br />

(Ri − Rcm) 2<br />

⇐⇒ R 2 g = 1<br />

2N<br />

N<br />

i=1<br />

N<br />

(Ri − Rj) 2 , (2.7)<br />

where Rj = j i=1 ti is the position vector of a jo<strong>in</strong>t/bead along the cha<strong>in</strong> and<br />

Rcm = 1 N N j=1 Rj is the ‘center of mass’. The mean<strong>in</strong>g of R2 g is that of the<br />

mean-square distance of the monomers to the center of mass or the mean-square<br />

distance between all the beads. It is a general measure for the size of a polymer<br />

because 〈R 2 〉 is not useful for e.g branched polymers. It scales <strong>in</strong> the same way<br />

with the number of segments as the end-to-end distance.<br />

Upon identify<strong>in</strong>g ti with the sp<strong>in</strong> vector si and an external force F with a<br />

magnetic field H, one can take over the results from the statistical mechanics of a<br />

paramagnet to calculate the force extension relation. An <strong>in</strong>terest<strong>in</strong>g result is, that<br />

a polymer behaves as an entropic spr<strong>in</strong>g with a spr<strong>in</strong>g constant proportional to<br />

kBT (see eq. (2.11)) which reflects the entropy dom<strong>in</strong>ated behavior of the system.<br />

In comparison to a solid which expands upon heat<strong>in</strong>g, the polymer cha<strong>in</strong> stiffens.<br />

This allows us to write down the free energy F of the ideal cha<strong>in</strong><br />

j=1<br />

F ∼ kBT R2<br />

, (2.8)<br />

Nb2 which is derived from the end-to-end vector distribution, eq. (2.5), <strong>in</strong> [39]. The<br />

free energy <strong>in</strong>creases when the end-to-end distance becomes lager, s<strong>in</strong>ce the number<br />

of allowed configurations and with it the entropy are reduced.<br />

For mathematical convenience another, more abstract, model is <strong>in</strong>troduced.<br />

The model generalizes the long-scale Gaussian behavior to all scales. For this<br />

Gaussian cha<strong>in</strong> the distance between any two po<strong>in</strong>ts along the cha<strong>in</strong> is assumed<br />

to be Gaussian distributed. The probability distribution of the end-to-end vector<br />

becomes a path-<strong>in</strong>tegral <strong>in</strong> the limit of small bond length b → 0 but fixed total<br />

length L = Nb<br />

<br />

r(L)=R<br />

3L<br />

p(R) = D[r(s)] exp<br />

2 〈R2 〉<br />

r(0)=0<br />

L<br />

0<br />

<br />

2<br />

∂r(s)<br />

ds<br />

, (2.9)<br />

∂s<br />

r(s) denot<strong>in</strong>g the trajectory of the polymer curve. The solution is, of course, eq.<br />

(2.5) which is also the solution of a correspond<strong>in</strong>g free particle diffusion equation.


2.3. SEMI-FLEXIBLE POLYMERS 9<br />

In this path-<strong>in</strong>tegral representation the weight of a particular polymer configuration<br />

is described by an effective Hamiltonian<br />

Hgauss[r(s)] = 3kBT L<br />

2 〈R 2 〉<br />

L<br />

0<br />

ds ˙r(s) 2 . (2.10)<br />

Here a dot <strong>in</strong>dicates the derivative along the contour. This effective Hamiltonian<br />

conta<strong>in</strong>s only entropic contributions, which govern the behavior of the FJC (cf.<br />

[9]). It can be <strong>in</strong>terpreted as the cont<strong>in</strong>uum limit of the energy of a cha<strong>in</strong> of<br />

entropic spr<strong>in</strong>gs with spr<strong>in</strong>g constant (cf. [10])<br />

k = 3kBT<br />

〈R2 . (2.11)<br />

〉<br />

The central limit theorem does not provide any h<strong>in</strong>t when the Gaussian distribution<br />

is applicable, i.e. it does not tell us how large N must be chosen. It<br />

certa<strong>in</strong>ly holds only if the total length L of the polymer is large compared to the<br />

correlation between the segments, which is often not the case for biopolymers.<br />

But still universal features can be extracted by us<strong>in</strong>g a different model. This is<br />

the topic of the follow<strong>in</strong>g section.<br />

2.3 Semi-flexible polymers<br />

2.3.1 Kratky-Porod model<br />

The basic model for polymers show<strong>in</strong>g stiffness along the backbone cha<strong>in</strong> was<br />

<strong>in</strong>troduced by Kratky and Porod <strong>in</strong> [20]. The stiffness can be <strong>in</strong>cluded by a<br />

geometric h<strong>in</strong>drance on the orientation of segments which <strong>in</strong>troduces a preferred<br />

direction. One of the simplest realizations of a constra<strong>in</strong>t is the freely rotat<strong>in</strong>g<br />

cha<strong>in</strong>, where the angle between successive segments is fixed, but the segments<br />

can still rotate freely around one axis.<br />

One of the central quantities of <strong>in</strong>terest is the tangent-tangent correlation<br />

function. It is always of exponential form (cf. [39])<br />

φtt(s, s ′ ) = ˆt(s) · ˆt(s ′ ) <br />

= exp − |s − s′ <br />

|<br />

(2.12)<br />

for a polymer <strong>in</strong> bulk solution. Through this formula the persistence length lp is<br />

<strong>in</strong>troduced as the ma<strong>in</strong> length scale. It depends on the details of the underly<strong>in</strong>g<br />

stiffness mechanism. For lp ≪ L the FJC is recovered—which shows that even<br />

semi-flexible polymers ‘behave’ Gaussian on large length scales, if L is large. The<br />

segments of the freely-rotat<strong>in</strong>g cha<strong>in</strong> are essentially uncorrelated on length scales<br />

much larger than lp. On the other hand for lp ≫ L the polymer is essentially<br />

stiff over the whole length. It then resembles a stiff rod. In the next section<br />

lp


10 CHAPTER 2. POLYMER MODELS<br />

we will give an explicit derivation of these results. Both regimes are convenient<br />

for approximations as we already saw for the Gaussian cha<strong>in</strong> and will see <strong>in</strong> the<br />

section 2.4.1.<br />

The correlation function, eq. (2.12), can be used to calculate the mean-square<br />

end-to-end distance 〈R 2 〉. The details will not be presented here (cf. [15] and<br />

[39]) s<strong>in</strong>ce we will discuss the worm-like cha<strong>in</strong> model <strong>in</strong> more detail, which will<br />

be used <strong>in</strong> the rest of this thesis. Both models give equivalent results.<br />

2.3.2 Worm-like cha<strong>in</strong> model<br />

Unlike the Kratky-Porod model the worm-like cha<strong>in</strong> (WLC) <strong>in</strong>troduces the stiffness<br />

<strong>in</strong> a discrete model by an <strong>in</strong>teraction energy ε between neighbor<strong>in</strong>g bonds<br />

N−1 <br />

H ({ti}) = −ε ti · ti+1 with: |ti| = l|ˆti| = l. (2.13)<br />

i=1<br />

This model is identical to a one-dimensional Heisenberg model for ferromagnets,<br />

where the sp<strong>in</strong>s ˆsi replace the tangents and the exchange <strong>in</strong>teraction J the <strong>in</strong>teraction<br />

energy: HHeisenberg = −J N−1 i=1 ˆsi · ˆsi+1. In the WLC model non-parallel<br />

orientation of successive segments are penalized by a bend<strong>in</strong>g energy.<br />

The discrete model is ideally suited for Monte-Carlo simulations on polymer<br />

systems; see later section 4. In order to be able to compare the discrete model<br />

with the cont<strong>in</strong>uum limit, the tangent-tangent correlation function has to be<br />

calculated to give a relation between the <strong>in</strong>teraction energy ε an the persistence<br />

length lp. This is shown <strong>in</strong> appendix A for two- and three-dimensional embedd<strong>in</strong>g<br />

space.<br />

Here we will first focus on an analytical treatment of the WLC model. For<br />

that purpose it is advantageous to take the cont<strong>in</strong>uum limit to arrive at a real<br />

‘worm-like’ cont<strong>in</strong>uous space curve as <strong>in</strong> figure 2.2. This cont<strong>in</strong>uum description<br />

is a convenient mathematical description on a coarse-gra<strong>in</strong>ed scale. For the cont<strong>in</strong>uum<br />

limit the length of the cha<strong>in</strong> segments must tend to zero, l → 0, as the<br />

number of segments tends to <strong>in</strong>f<strong>in</strong>ity, N → ∞. The <strong>in</strong>teraction energy must tend<br />

R(s) = s<br />

ds t(s)<br />

0<br />

r(s)<br />

t(s)<br />

R(L)<br />

Figure 2.2: The WLC as cont<strong>in</strong>uous space curve r(s).


2.3. SEMI-FLEXIBLE POLYMERS 11<br />

to <strong>in</strong>f<strong>in</strong>ity as well, ε → ∞, such that the the energy per segment stays fixed<br />

˜ε = ε<br />

N<br />

The total length L of the polymer must be f<strong>in</strong>ite as well:<br />

!<br />

= fixed. (2.14)<br />

L = Nl ! = fixed. (2.15)<br />

Further, of course, derivatives are def<strong>in</strong>ed as the cont<strong>in</strong>uum limit of a f<strong>in</strong>ite difference<br />

<br />

∂t(s) ti+1 − ti<br />

= lim<br />

.<br />

∂s l→0 l<br />

(2.16)<br />

Here the arc length s, which replaces the <strong>in</strong>dex i, is label<strong>in</strong>g the position along<br />

the contour of the cha<strong>in</strong>. We use partial derivatives to <strong>in</strong>dicate that, <strong>in</strong> general,<br />

the tangent vector may also depend on time. But, s<strong>in</strong>ce we are only concerned<br />

with static properties, we will ignore this <strong>in</strong> the follow<strong>in</strong>g.<br />

Let us start off the Hamiltonian <strong>in</strong> eq. (2.13) which gives after a simple transformation,<br />

where −ti+1 · ti = 1<br />

2 [(ti+1 − ti) 2 − 2l 2 ],<br />

N−1<br />

εl<br />

H[t(s)] = lim<br />

l→0 2<br />

ε,N→∞ i=1<br />

<br />

2 ti+1 − ti<br />

l<br />

, (2.17)<br />

l<br />

by dropp<strong>in</strong>g irrelevant constants. These would only shift the energy zero <strong>in</strong> the<br />

cont<strong>in</strong>uum limit. The sum f<strong>in</strong>ally transforms <strong>in</strong>to an <strong>in</strong>tegral N−1 i=1 l → L<br />

0 ds,<br />

such that<br />

H[t(s)] = κ<br />

L 2 ∂t(s)<br />

ds =<br />

2 ∂s<br />

κ<br />

L 2 ∂ r(s)<br />

ds<br />

2 ∂s2 2 . (2.18)<br />

0<br />

The constra<strong>in</strong>t |ti| = l transformed <strong>in</strong>to |t(s)| = 1 2 , which means that the space<br />

curve is locally <strong>in</strong>extensible 3 . The bend<strong>in</strong>g modulus is def<strong>in</strong>ed as κ = εl = ˜εNl =<br />

˜εL.<br />

2Arc length parametrization enforces the constra<strong>in</strong>t |t(s)| = 1 s<strong>in</strong>ce<br />

ds 2 ∂x 2 2 <br />

2<br />

∂y ∂z<br />

= + + ds<br />

∂s ∂s ∂s<br />

<br />

2 ⇒ |t(s)| 2 = 1<br />

=|t(s)| 2 ŕ ŕ<br />

ŕ<br />

= ŕ<br />

∂r(s) ŕ2<br />

ŕ<br />

ŕ ∂s ŕ<br />

3 When the <strong>in</strong>extensibility constra<strong>in</strong>t is relaxed, then a potential energy for the stretch<strong>in</strong>g<br />

of segments must be <strong>in</strong>troduced, e.g. when the force-extension relation for very large forces<br />

is <strong>in</strong>vestigated, the segments will start stretch<strong>in</strong>g [33]. Normally the backbone segments are<br />

very rigid <strong>in</strong> length so that stretch<strong>in</strong>g requires considerably more energy than bend<strong>in</strong>g, therefore<br />

there is a time-scale separation for the relaxations times, where the stretch<strong>in</strong>g modes are relax<strong>in</strong>g<br />

much faster. That is why normally one can focus on an <strong>in</strong>extensible cha<strong>in</strong> [2].<br />

0


12 CHAPTER 2. POLYMER MODELS<br />

We have now arrived at a description of a stiff polymer <strong>in</strong> terms of a cont<strong>in</strong>uous<br />

(and <strong>in</strong>extensible) space curve r(s). For a given conformation the Hamiltonian<br />

penalizes high absolute curvatures ¨r(s). One may also <strong>in</strong>terpret the polymer<br />

configuration as a trajectory of a particle which is ran through with constant<br />

speed. The position on the curve corresponds to the time.<br />

An alternative way to derive eq. (2.18) is by symmetry considerations. Assum<strong>in</strong>g<br />

the polymer can be represented by a space curve, the form of the Hamiltonian<br />

has to reflect all symmetries of the system, i.e. spatial isotropy, homogeneity and<br />

time-reversal (not <strong>in</strong> static problems). Therefore the second derivative of the<br />

trajectory along the curve squared is the first allowed term. The disadvantage of<br />

this derivation is that it does not expla<strong>in</strong> the microscopic relevance of the bend<strong>in</strong>g<br />

modulus κ.<br />

The relation between the bend<strong>in</strong>g modulus κ and the persistence length lp is<br />

obta<strong>in</strong>ed by calculat<strong>in</strong>g the tangent-tangent correlation function 〈t(s) · t(s ′ )〉. As<br />

shown <strong>in</strong> [40] this is done by exploit<strong>in</strong>g the similarity between the path-<strong>in</strong>tegral of<br />

a free particle <strong>in</strong> quantum mechanics and the conformational part of the partition<br />

function Z represented by the path-<strong>in</strong>tegral<br />

<br />

Z = D[t(s)] δ(|t(s)| − 1) exp{−βH[t(s)]}, (2.19)<br />

where the <strong>in</strong>extensibility constra<strong>in</strong>t |t(s)| = 1 <strong>in</strong>troduces an additional complication.<br />

By analogy with quantum mechanics one arrives at a diffusion equation<br />

for the partition sum. Due to the <strong>in</strong>extensibility constra<strong>in</strong>t this corresponds to<br />

diffusion on the unit sphere. This is similar to the path-<strong>in</strong>tegral representation<br />

of the Gaussian cha<strong>in</strong> <strong>in</strong> eq. (2.9), which was related to the diffusion of a free<br />

particle. Now—<strong>in</strong> contrast to the purely entropic Hamiltonian eq. (2.10)—the<br />

bend<strong>in</strong>g energy is the dom<strong>in</strong>at<strong>in</strong>g contribution <strong>in</strong> the Hamiltonian eq. (2.18).<br />

The relation between the bend<strong>in</strong>g modulus and persistence length is obta<strong>in</strong>ed<br />

by an normal-mode analysis<br />

lp = βκ or κ = lpkBT , (2.20)<br />

which is valid <strong>in</strong> three dimensional embedd<strong>in</strong>g space4 .<br />

With eq. (2.12) the second moment or mean-square end-to-end distance 〈R2 〉<br />

and radius of gyrations R2 <br />

g can be calculated by straightforward <strong>in</strong>tegration (all<br />

odd moments are zero because of the isotropy of the problem). The exact result<br />

4 As shown e.g <strong>in</strong> [19], for arbitrary embedd<strong>in</strong>g space dimension d the relation is<br />

lp =<br />

2κ<br />

kBT (d − 1) .<br />

In the same reference the general relations for R 2 and R 4 can be found—calculated by<br />

Green’s-functions—which have been used to test the simulations (cf. section 5.1).


2.3. SEMI-FLEXIBLE POLYMERS 13<br />

for the mean-square end-to-end distance is (cf. [24])<br />

L<br />

2<br />

2<br />

R (L) = ds t(s)<br />

<br />

L<br />

0<br />

L<br />

ds<br />

0<br />

′ 〈t(s) · t(s ′ )〉<br />

= ds<br />

0<br />

= 2 l 2 <br />

L<br />

p − 1 + e<br />

lp<br />

−L/lp<br />

<br />

= L 2 <br />

L<br />

fD , (2.21)<br />

where the Debye-function fD(x) = 2<br />

x2 (x − 1 + e−x ) has been used. This function<br />

is shown <strong>in</strong> figure 5.1 together with simulation results.<br />

There are two <strong>in</strong>terest<strong>in</strong>g limits<br />

2<br />

R <br />

2Llp for L ≫ lp<br />

=<br />

L2 . (2.22)<br />

for L ≪ lp<br />

The first l<strong>in</strong>e is the flexible limit <strong>in</strong> which the Gaussian model is valid. The<br />

relation is identical to the result obta<strong>in</strong>ed <strong>in</strong> eqs. (2.4) and (2.6). For the WLC<br />

the Kuhn segment length has to be set to<br />

lp<br />

b = 2lp. (2.23)<br />

The other limit is that of a stiff rod, where the result is the fully stretched polymer<br />

of length L.<br />

The radius of gyration is calculated with r(s) − r(s ′ ) = s<br />

s ′ dτ t(τ) us<strong>in</strong>g the<br />

cont<strong>in</strong>uum version of eq. (2.7)<br />

2 1<br />

Rg =<br />

2L2 = l 2 <br />

p<br />

=<br />

L<br />

0<br />

fD<br />

<br />

2lpL<br />

6<br />

L2 12<br />

L<br />

ds ′<br />

r(s) <br />

′ 2<br />

− r(s )<br />

ds<br />

0 <br />

L<br />

− 1 + L<br />

<br />

lp<br />

= bL<br />

6<br />

3lp<br />

(2.24)<br />

for L ≫ lp<br />

, (2.25)<br />

for L ≪ lp<br />

and aga<strong>in</strong> by the relation b = 2lp for the Kuhn segment the result for a flexible<br />

limit is recovered as well as the result for a stiff rod.<br />

Other <strong>in</strong>terest<strong>in</strong>g results can be calculated exactly: The forth moment 〈R4 〉<br />

<strong>in</strong> [40], the pair correlation function 〈r(s) · r(s ′ )〉 <strong>in</strong> [2] and the projection of the<br />

end-to-end vector on the direction of the first tangent <br />

R · ˆt1 as well as the<br />

details for mean-square radius of gyrations R2 <br />

g <strong>in</strong> [15].<br />

Other observations concern<strong>in</strong>g the radial distribution function p(R) are conta<strong>in</strong>ed<br />

<strong>in</strong> section 3.3.3.


14 CHAPTER 2. POLYMER MODELS<br />

0<br />

s − r ||(s)<br />

r(s)<br />

s<br />

r⊥(s)<br />

z-axis<br />

L<br />

Figure 2.3: Parametrization <strong>in</strong> the weakly-bend<strong>in</strong>g rod approximation. The symbols are<br />

expla<strong>in</strong>ed <strong>in</strong> the text.<br />

2.4 Stiff polymers<br />

Up to now we have dealt with the short list of exactly solvable quantities for the<br />

m<strong>in</strong>imal model, eq. (2.18). The model is called m<strong>in</strong>imal s<strong>in</strong>ce only one material<br />

parameter, the bend<strong>in</strong>g stiffness, is used for the description for a complex molecule<br />

like F-Act<strong>in</strong>. The existence of torsion along the cha<strong>in</strong> is neglected (i.e. assumed<br />

to be equilibrated) and possible <strong>in</strong>teractions between torsion and bend<strong>in</strong>g are not<br />

considered.<br />

Hav<strong>in</strong>g the m<strong>in</strong>imal model described by the Hamiltonian eq. (2.18) still does<br />

not allow us to solve problems analytically for all stiffnesses, therefore one has to<br />

resort to approximations <strong>in</strong> the limit of very stiff polymers. In this case <strong>in</strong>stead<br />

of beg<strong>in</strong>n<strong>in</strong>g with a flexible cha<strong>in</strong> the stiff rod limit is used as a start<strong>in</strong>g po<strong>in</strong>t.<br />

2.4.1 Weakly-bend<strong>in</strong>g rod approximation<br />

In a first approximation, we <strong>in</strong>troduce a stiff rod with small fluctuations around<br />

the fully stretched configuration, therefore the name weakly-bend<strong>in</strong>g rod approximation<br />

is used. We choose a preferred direction ˆn and parameterize the curve<br />

<strong>in</strong> the follow<strong>in</strong>g way (see figure 2.3)<br />

r(s) = r⊥(s), s − r||(s) . (2.26)<br />

Fluctuations perpendicular to ˆn are termed undulations and are parametrized by<br />

r⊥(s) = (x(s), y(s)). The parallel component r||(s) is the stored length describ<strong>in</strong>g<br />

how much the length parallel to ˆn is reduced from the arc length s.<br />

The <strong>in</strong>extensibility constra<strong>in</strong>t |t(s)| = |˙r(s)| = 1 implies a relation<br />

2 ˙r||(s) = ˙r⊥(s) 2 + ˙r||(s) 2 . (2.27)<br />

Now the approximation of small changes <strong>in</strong> the undulations |˙r⊥(s)| ≪ 1 is <strong>in</strong>troduced,<br />

which allows us to neglect the term ˙r||(s) 2 <strong>in</strong> the above relation, s<strong>in</strong>ce it<br />

is already fourth order <strong>in</strong> the undulation change. The approximation states that<br />

the polymer stays essentially stretched as depicted <strong>in</strong> figure 2.3.<br />

This leads to the approximative relation<br />

˙r||(s) ≈ 1<br />

2 ˙r⊥(s) 2 . (2.28)


2.5. REAL CHAINS 15<br />

no overhangs allowed<br />

|r⊥(s)| too large<br />

z-axis<br />

Figure 2.4: Overhangs are and large changes <strong>in</strong> |r⊥(s)| are not allowed<br />

Note that this implies that ˙r|| is non-negative or <strong>in</strong> other words that overhangs,<br />

as shown <strong>in</strong> figure 2.4, are not allowed. The derivative<br />

¨r||(s) 2 <br />

1 d˙r⊥(s)<br />

≈<br />

2<br />

2 2 =<br />

ds<br />

˙r⊥(s) · ¨r⊥(s) 2 (2.29)<br />

is already anharmonic and can therefore be neglected. Why this term does not<br />

contribute is easily seen <strong>in</strong> Fourier space. Due to the derivatives this term depends<br />

on the wave-vector as k 6 . S<strong>in</strong>ce <strong>in</strong> stiff polymers high frequency modes are<br />

suppressed by the bend<strong>in</strong>g energy, this term can be neglected. The term ¨r||(s) <strong>in</strong><br />

the relation ¨r(s) 2 = ¨r⊥(s) 2 + ¨r||(s) 2 can be dropped<br />

¨r(s) ≈ ¨r⊥(s). (2.30)<br />

This leads to the bend<strong>in</strong>g energy Hamiltonian <strong>in</strong> the weakly-bend<strong>in</strong>g rod approximation5<br />

Hwb[r⊥(s)] = κ<br />

2<br />

L<br />

0<br />

ds ¨r⊥(s) 2 = κ<br />

2<br />

L<br />

0<br />

ds ¨x(s) 2 + ¨y(s) 2 . (2.31)<br />

Now the <strong>in</strong>extensibility constra<strong>in</strong>t is implicitly, i.e. automatically, fulfilled. The<br />

path-<strong>in</strong>tegral for the partition sum, eq. (2.19), has been simplified, s<strong>in</strong>ce the<br />

<strong>in</strong>extensibility constra<strong>in</strong>t <strong>in</strong> weakly-bend<strong>in</strong>g rod approximation leaves only two<br />

<strong>in</strong>dependent components.<br />

This approximation is also called Monge-representation to <strong>in</strong>dicate the chosen<br />

parametrization. The s<strong>in</strong>gle valued contour is parametrized <strong>in</strong> reference to a<br />

reference l<strong>in</strong>e by the undulations 6 .<br />

2.5 Real cha<strong>in</strong>s<br />

We have dealt so far with idealized cha<strong>in</strong>s represented by segments or a cont<strong>in</strong>uous<br />

space curve, allowed to <strong>in</strong>terpenetrate each other. Therefore these models are<br />

5In classical elasticity theory the energy of a weakly-bend<strong>in</strong>g rod has exactly the same form<br />

as this Hamiltonian (cf. [23]).<br />

6In differential geometry the Monge-representation is used to parametrize s<strong>in</strong>gle-valued<br />

planes by a ‘hight’ over a reference plane.


16 CHAPTER 2. POLYMER MODELS<br />

frequently called phantom cha<strong>in</strong>s. But <strong>in</strong> reality polymers are charged objects <strong>in</strong><br />

a solution surrounded by charges. Therefore the <strong>in</strong>teraction between the polymer<br />

segments is very complex and depends on the solution. But often one is entitled<br />

to replace these <strong>in</strong>teraction by a simplified hard core <strong>in</strong>teraction. The next step <strong>in</strong><br />

the direction of a more realistic model is then to <strong>in</strong>corporate self-avoidance effects<br />

as a steep short-ranged repulsive potential between the segments (of the order of<br />

the microscopic polymer thickness). The polymer is not allowed to <strong>in</strong>tersect with<br />

itself 7 . For detailed description of models with ‘realistic’ features confer e.g. [39].<br />

The <strong>in</strong>fluence of the self-avoidance/excluded volume is visible <strong>in</strong> particular<br />

for flexible/long (i.e. a large ratio of L/lp) cha<strong>in</strong>s s<strong>in</strong>ce they fold back and forth<br />

freely, therefore cross<strong>in</strong>g themselves very often. As a result the polymer will tend<br />

to be more extended than a phantom cha<strong>in</strong> due to the reduced number of allowed<br />

configurations (entropic penalty) <strong>in</strong> the strongly collapsed state. In contrast for<br />

stiff polymers the effect is mostly negligible s<strong>in</strong>ce it is unlikely for them to fold<br />

back onto themselves.<br />

At variance with the description of the ideal phantom cha<strong>in</strong> as a random walk,<br />

the self-avoid<strong>in</strong>g random walk can be used to <strong>in</strong>vestigate the properties of these<br />

cha<strong>in</strong>s. But we want to focus on the derivation of the basic scal<strong>in</strong>g relation for<br />

the end-to-end distance.<br />

The considerations of the “entropic penalty” can be used to derive a scal<strong>in</strong>g<br />

relation similar to eq. (2.6) for the ideal cha<strong>in</strong>. This argument dates back to P.<br />

Flory and is described <strong>in</strong> detail <strong>in</strong> [8] and [39]. It considers a correction to the<br />

free energy of the ideal cha<strong>in</strong>, eq. (2.8), that is due to the excluded volume which,<br />

<strong>in</strong> turn, depends on the number of monomer-monomer contacts. Let v be the<br />

mean volume occupied by a monomer depend<strong>in</strong>g on the segment length b and<br />

the thickness of the cha<strong>in</strong> d. The monomer density φ of a cha<strong>in</strong> of l<strong>in</strong>ear size R<br />

which occupies a sphere of about the volume R 3 is<br />

φ ∼ Nv<br />

. (2.32)<br />

R3 The entropy loss for an exclusion is of order kBT for each degree of freedom<br />

(equipartition theorem). S<strong>in</strong>ce there are N monomers the free energy correction<br />

is<br />

F ∼ kBT v<br />

N 2<br />

. (2.33)<br />

R3 Additionally us<strong>in</strong>g the free energy of the ideal cha<strong>in</strong>, eq. (2.8), one can m<strong>in</strong>imize<br />

the sum of these contributions with respect to the end-to-end distance R. We<br />

obta<strong>in</strong> the scal<strong>in</strong>g result<br />

R ∼ (b 2 v) 1/5 N 3/5 . (2.34)<br />

7 There exists a so called θ-temperature (depend<strong>in</strong>g on the solution) where the segmentsegment<br />

<strong>in</strong>teraction is exactly balanced by effects from the surround<strong>in</strong>g solution. Then the<br />

description of an ideal cha<strong>in</strong> is valid.


2.5. REAL CHAINS 17<br />

Figure 2.5: Snapshot of fluorescently labeled DNA molecules adher<strong>in</strong>g to a two-dimensional<br />

lipid membrane. Courtesy of J. Rädler and B. Maier [32].<br />

The scal<strong>in</strong>g exponent changes to ν = 3/5 for the number of segments. In comparison<br />

to the ideal cha<strong>in</strong> result of ν = 1/2 the polymers swells. The value for<br />

the real cha<strong>in</strong> is <strong>in</strong> good agreement with experiments and simulations which show<br />

a scal<strong>in</strong>g exponent ν ≈ 0.558.<br />

For polymers restricted to an embedd<strong>in</strong>g space of two-dimension this scal<strong>in</strong>g<br />

exponent is found to be ν = 3/4 (e.g. by <strong>in</strong>vestigat<strong>in</strong>g a self-avoid<strong>in</strong>g random<br />

walk). In figure 2.5 snapshots of a fluorescently labeled DNA absorbed on a twodimensional<br />

lipid membrane are shown (cf. [32]). With this experimental set up<br />

measurements on the scal<strong>in</strong>g exponent were performed, which showed a value of<br />

ν = 0.79 ± 0.04 <strong>in</strong> good agreement with the exact result.<br />

As expla<strong>in</strong>ed <strong>in</strong> the references this good result is only due to the lucky cancellation<br />

of errors made by naively summ<strong>in</strong>g up the scal<strong>in</strong>g form of the free energy.<br />

The number of monomer-monomer contacts as well as the entropic energy are<br />

overestimated. The good result is destroyed if one tries to improve these energy<br />

estimations. This argument should only be used “as it stands”.<br />

Depend<strong>in</strong>g on the value of the monomer volume v <strong>in</strong>serted <strong>in</strong> eq. (2.34) three<br />

different prefactors can be obta<strong>in</strong>ed. Us<strong>in</strong>g the the Kuhn segment length b and<br />

a segment thickness/diameter h<br />

v1 ∼ b 3<br />

sphere, (2.35a)<br />

v2 ∼ b 2 h disc, (2.35b)<br />

v3 ∼ bh 2<br />

cyl<strong>in</strong>der. (2.35c)<br />

The first relation assumes that the segments, on average, occupy a sphere of<br />

diameter b, the second assumes a disc of diameter d and thickness h while the<br />

third assumes a cyl<strong>in</strong>der of length b and diameter d. As mentioned <strong>in</strong> [41] the<br />

right average excluded volume to use for a segment is v2. Intuitively this can be<br />

understood by rul<strong>in</strong>g out the other relations: A cyl<strong>in</strong>der would mean that the<br />

segment is strongly fixed and <strong>in</strong> average barely moves around its position. In contrast<br />

a sphere would mean that the segment, fixed <strong>in</strong> the cha<strong>in</strong>, <strong>in</strong> average covers<br />

all orientations. Therefore the usage of a disc is <strong>in</strong>tuitively a good <strong>in</strong>termediate<br />

approach.<br />

Despite these facts, the sphere-picture is normally used <strong>in</strong> the literature as a<br />

general assumption. Insert<strong>in</strong>g <strong>in</strong> eq. (2.34) we obta<strong>in</strong> the standard result<br />

R ∼ bN 3/5 . (2.36)


18 CHAPTER 2. POLYMER MODELS<br />

We will see <strong>in</strong> the follow<strong>in</strong>g that the different possible volumes vi can give significantly<br />

different results for the relevance of self-avoidance as well as the scal<strong>in</strong>g<br />

<strong>in</strong> a conf<strong>in</strong><strong>in</strong>g tube. In this context the important parameter is the relation of<br />

segment length to thickness b/h, called aspect-ratio.<br />

Relevance of self-avoidance <strong>in</strong> static problems<br />

In static problems the self-avoidance does not play a role for all flexible polymers 8 :<br />

• For a system <strong>in</strong> θ-condition we can use the ideal cha<strong>in</strong> model.<br />

• Whether or not self-avoidance has to be taken <strong>in</strong>to account depends on<br />

the number of b<strong>in</strong>ary contacts Nφ between the segments. Only if this<br />

number is large an effect is to be expected. Therefore there exists a critical<br />

number of segments Nc where a cross-over between the description as ideal<br />

phantom cha<strong>in</strong> and as self-avoid<strong>in</strong>g cha<strong>in</strong> is expected. This cross-over will<br />

be <strong>in</strong>vestigated <strong>in</strong> the follow<strong>in</strong>g.<br />

The scal<strong>in</strong>g relation for the end-to-end distance of a free flexible polymer, eq.<br />

(2.6), is valid for a specific range of values of N. But N must not be too small<br />

s<strong>in</strong>ce we still need L ≫ lp. Start<strong>in</strong>g from a specific N ≈ Nc self-avoidance starts<br />

play<strong>in</strong>g a role. Us<strong>in</strong>g eq. (2.6) a scal<strong>in</strong>g relation for Nc can be derived.<br />

The number of b<strong>in</strong>ary contacts is given by (us<strong>in</strong>g eq. (2.32))<br />

1/2 v<br />

Nφ ∼ N<br />

b3 !<br />

= B, (2.37)<br />

with an arbitrary constant B > 1. The number of b<strong>in</strong>ary contacts should be<br />

large for a significant <strong>in</strong>fluence of self-avoidance on the conformation. Solv<strong>in</strong>g for<br />

N yields a scal<strong>in</strong>g relation for the critical number of segments:<br />

Nc ∼ b6<br />

. (2.38)<br />

v2 For N > Nc self-avoidance is relevant.<br />

The N-dependence of R can be split <strong>in</strong> two regimes with a cross-over po<strong>in</strong>t<br />

at Nc which is sketched <strong>in</strong> figure 2.6.<br />

As mentioned above the scal<strong>in</strong>g of Nc depends crucially on the average volume<br />

occupied by a monomer v. The follow<strong>in</strong>g three results are obta<strong>in</strong>ed for the three<br />

8 For the dynamic properties there is an additional effect of dynamic traps. By ergodicity all<br />

configurations are allowed, but some can be very difficult to reach. E.g. when the cha<strong>in</strong> would<br />

have to cross through itself these conformations are close <strong>in</strong> configuration space, but s<strong>in</strong>ce the<br />

segments are impenetrable the polymer would have to uncoil and coil up aga<strong>in</strong> to make the<br />

conformation change. This difficulty does not occur for static problems!


2.5. REAL CHAINS 19<br />

ln R<br />

cases <strong>in</strong>troduced above<br />

1<br />

2<br />

phantom cha<strong>in</strong><br />

Nc<br />

3<br />

5<br />

cross-over<br />

self-avoid<strong>in</strong>g cha<strong>in</strong><br />

ln N<br />

Figure 2.6: Cross-over from ideal-cha<strong>in</strong> to real-cha<strong>in</strong> behavior.<br />

Nc1 ∼ 1, (2.39a)<br />

Nc2 ∼ <br />

b 2<br />

, (2.39b)<br />

h<br />

Nc3 ∼ <br />

b 4<br />

, (2.39c)<br />

h<br />

except for the case of a sphere Nc1, Nc depends on the aspect-ratio b/h.<br />

For a sphere self-avoidance is always important. But follow<strong>in</strong>g [41] the dependence<br />

of Nc2 on the aspect-ratio squared is the most reasonable. E.g. for DNA<br />

the aspect-ratio is about 30 so that Nc2 ≈ 500 − 1000. This value is <strong>in</strong> the range<br />

of the number of segments used <strong>in</strong> the simulations. We do not know the prefactor<br />

<strong>in</strong> the scal<strong>in</strong>g relations hence we cannot be sure whether self-avoidance must be<br />

taken <strong>in</strong>to account. We will see <strong>in</strong> chapter 5, that up to N = 500 segments we<br />

could not detect any <strong>in</strong>fluence of self-avoidance on the static properties.


20 CHAPTER 2. POLYMER MODELS


Chapter 3<br />

<strong>Polymers</strong> <strong>in</strong> conf<strong>in</strong>ed geometry<br />

In this chapter we are go<strong>in</strong>g to <strong>in</strong>vestigate the conformational behavior of polymer<br />

cha<strong>in</strong>s <strong>in</strong> a conf<strong>in</strong><strong>in</strong>g geometry. First a short experimental motivation is given,<br />

after which heuristic scal<strong>in</strong>g relations are derived. This is followed by a more<br />

detailed analytical calculation, start<strong>in</strong>g from the limit of stiff cha<strong>in</strong>s <strong>in</strong> the weaklybend<strong>in</strong>g<br />

rod approximation. In particular we <strong>in</strong>vestigate the cyl<strong>in</strong>drical harmonic<br />

potential as a generic model for conf<strong>in</strong><strong>in</strong>g environments with similar a symmetry.<br />

3.1 Motivation<br />

The <strong>in</strong>terest <strong>in</strong> conf<strong>in</strong>ement of polymers <strong>in</strong>to nano-sized structures emerges ma<strong>in</strong>ly<br />

from biotechnology. As discussed <strong>in</strong> the <strong>in</strong>troduction, there are various applications<br />

encourag<strong>in</strong>g our work on the problem of conf<strong>in</strong>ement.<br />

One direct application is that of conf<strong>in</strong>ement elongation. The advantage is<br />

that upon conf<strong>in</strong><strong>in</strong>g the cha<strong>in</strong> stretches already <strong>in</strong> its equilibrium conformation.<br />

There is no need of a stretch<strong>in</strong>g flow or elongat<strong>in</strong>g external force, which drives the<br />

polymer out of its static conformation <strong>in</strong>to a stretched non-equilibrium conformation.<br />

Here we can try to understand the mechanisms lead<strong>in</strong>g from the unconf<strong>in</strong>ed<br />

to the conf<strong>in</strong>ed static equilibrium conformation.<br />

An application with a large potential impact is the determ<strong>in</strong>ation of the<br />

contour-length L of a polymer by measur<strong>in</strong>g its equilibrium apparent end-toend<br />

distance <strong>in</strong> an conf<strong>in</strong><strong>in</strong>g tube of the order of the persistence length or less.<br />

Up to now gel-electrophoresis is the tool for this sort of measurements, but it has<br />

its problems. It is slow—a measurement takes <strong>in</strong> the order of a day—and it is<br />

relatively imprecise. The advantage of the idea of the conf<strong>in</strong>ement measurement<br />

seems therefore a faster way to obta<strong>in</strong> the data. It is a matter of m<strong>in</strong>utes to<br />

do the actual measurement and it is <strong>in</strong> pr<strong>in</strong>ciple possible to make the results<br />

arbitrarily accurate, by just tak<strong>in</strong>g more pictures of equilibrium configurations.<br />

The basic experiments are performed <strong>in</strong> nano-sized channels with a near<br />

square geometry <strong>in</strong> the group of Bob Aust<strong>in</strong> [38, 44]. They have succeeded<br />

21


22 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

d<br />

ld<br />

r<br />

Figure 3.1: The new scales for conf<strong>in</strong>ed polymers. The Odijk deflection length ld measures the<br />

typical length between successive deflections along the contour. r denotes the distance between<br />

deflection along the tube.<br />

<strong>in</strong> conf<strong>in</strong><strong>in</strong>g fluorescently dyed λ-DNA strands <strong>in</strong>to nano-sized channels seen <strong>in</strong><br />

the figures 1.1. The channels have an aspect ratio between one and two and are<br />

of average size 440 nm down to 30 nm. These experiments provide a picture of<br />

<strong>in</strong>creas<strong>in</strong>gly stretched strands when the channel becomes smaller, which is seen<br />

for relatively long T2-DNA and the shorter λ-DNA <strong>in</strong> the figures 1.2. The value,<br />

which can be read off from the pictures by us<strong>in</strong>g appropriate fitt<strong>in</strong>g functions, is<br />

the apparent end-to-end distance 〈Ra〉, which is the maximal distance between<br />

two monomers along the cha<strong>in</strong>.<br />

Here we will present first by scal<strong>in</strong>g relations then by analytical approximations<br />

on the cyl<strong>in</strong>drical symmetric harmonic potential 1 —model<strong>in</strong>g the tube—<br />

approaches to a solution of the problem of conf<strong>in</strong>ement and the identification<br />

of the relevant scal<strong>in</strong>g parameters. This is <strong>in</strong>vestigated <strong>in</strong> detail—particularly<br />

for the relevant range of parameters for the experimentalists—by simulations <strong>in</strong><br />

chapter 5. The result<strong>in</strong>g “scal<strong>in</strong>g plot” <strong>in</strong> section 5.2.3 can serve as a gauge for<br />

convert<strong>in</strong>g measured apparent end-to-end distances <strong>in</strong>to real contour lengths L.<br />

3.2 Scal<strong>in</strong>g relations<br />

Before giv<strong>in</strong>g a detailed analytical calculation, it is <strong>in</strong>structive to perform a scal<strong>in</strong>g<br />

analysis start<strong>in</strong>g from the relevant length and energy scales. Often it is relatively<br />

easy to f<strong>in</strong>d an <strong>in</strong>tuitive scal<strong>in</strong>g behavior. For an unconf<strong>in</strong>ed polymer the stiffness<br />

of the filament is characterized by the persistence length lp, and the thermal<br />

energies ∼ kBT determ<strong>in</strong>e the typical energy scale. Upon conf<strong>in</strong><strong>in</strong>g a polymer to a<br />

tube of lateral dimension d there is an additional scale, the Odijk deflection length<br />

ld (<strong>in</strong>troduced <strong>in</strong> [37]). It measures the typical distance of successive collisions of<br />

the polymer with the walls of the conf<strong>in</strong><strong>in</strong>g tube. The new parameters d and ld<br />

are shown <strong>in</strong> figure 3.1. A relation between these values can be easily established.<br />

With r⊥ ∼ d and s ∼ ld the transverse and longitud<strong>in</strong>al length scales, respectively,<br />

we f<strong>in</strong>d<br />

kBT ∼ H ∼ (lpkBT ) d2<br />

l 3 d<br />

⇒ l 3 d ∼ d 2 lp , (3.1)<br />

1 The harmonic potential be<strong>in</strong>g aga<strong>in</strong> and aga<strong>in</strong> the “laboratory animal” of the physicist. . .


3.2. SCALING RELATIONS 23<br />

us<strong>in</strong>g eq. (2.20) and the Hamiltonian eq. (2.18). This relation may be expla<strong>in</strong>ed<br />

heuristically: The length on which the polymer is stiff along the tube is the<br />

persistence length lp and the two perpendicular directions are conf<strong>in</strong>ed by the<br />

diameter d. Hence a volume of roughly d2lp is occupied, which corresponds to<br />

the volume l 3 d<br />

occupied between two deflections.<br />

In the follow<strong>in</strong>g and <strong>in</strong> particular for the simulation it is convenient to <strong>in</strong>troduce<br />

dimensionless parameters<br />

ɛ = L<br />

which is a small parameter for stiff polymers, and<br />

lp<br />

c = L<br />

ld<br />

the flexibility, (3.2)<br />

the collision parameter, (3.3)<br />

be<strong>in</strong>g a measure for the number of collisions/deflections along the tube.<br />

3.2.1 The flexible regime<br />

A flexible polymer is heavily coiled up <strong>in</strong> bulk solution. The effect of the conf<strong>in</strong><strong>in</strong>g<br />

tube on the end-to-end distance along the tube R||—which will be the ma<strong>in</strong> concern<br />

here—is therefore expected to be quite strong. The polymer conformation<br />

will be straightened and stretched due to excluded volume <strong>in</strong>teractions between<br />

distant segments along the polymer backbone. Note that for an idealized phantom<br />

cha<strong>in</strong> there would be no stretch<strong>in</strong>g at all. Ideal cha<strong>in</strong>s can be represented<br />

by a random walk on all scales (cf. section 2.2) where each of the possible spacial<br />

direction is completely <strong>in</strong>dependent of the others. Therefore the tube does<br />

not have an effect on the end-to-end distance along the tube. Merely the lateral<br />

directions are cut off, be<strong>in</strong>g <strong>in</strong>dependent of the the parallel.<br />

By <strong>in</strong>troduc<strong>in</strong>g a little bit of stiffness—still <strong>in</strong> the range of high flexibility—<br />

the directions are not completely <strong>in</strong>dependent anymore. Then there is a visible<br />

<strong>in</strong>fluence of the tube which has to result from boundary layer effects. The polymer<br />

is stiff on the scale of some segment lengths. Close to the wall the orientation can<br />

happen to be towards the wall. S<strong>in</strong>ce this direction is forbidden the persistence<br />

forces the segment along the tube. This results <strong>in</strong> an <strong>in</strong>crease of R||.<br />

For real self-avoid<strong>in</strong>g cha<strong>in</strong>s we expect a strong <strong>in</strong>fluence of the conf<strong>in</strong><strong>in</strong>g tube.<br />

The polymer cannot be squeezed arbitrarily, due to steric constra<strong>in</strong>ts. When the<br />

polymer size is of the order of the conf<strong>in</strong><strong>in</strong>g tube Rg ≈ d, there are two simple<br />

ways to derive the scal<strong>in</strong>g relation of the end-to-end distance R|| along the tube,<br />

as described <strong>in</strong> [8]. For tubes larger than the polymer size, it behaves as if it<br />

were <strong>in</strong> bulk solution. For smaller tubes the cha<strong>in</strong> effectively stretches out and<br />

behaves as if it were stiff 2 .<br />

2 There is still the possibility that loops and k<strong>in</strong>ks occur. This is an <strong>in</strong>terest<strong>in</strong>g effect and is<br />

currently under <strong>in</strong>vestigation.


24 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

d<br />

blobs<br />

Figure 3.2: ‘Blob’ picture scal<strong>in</strong>g: Inside the blobs of size d the polymer behaves as if it were<br />

free. The self-avoidance is taken <strong>in</strong>to account by treat<strong>in</strong>g the blobs as hard-walled, stack<strong>in</strong>g<br />

them l<strong>in</strong>early.<br />

The derivation of the <strong>in</strong>termediate scal<strong>in</strong>g regime can be done most illustrative<br />

by the “blob picture” <strong>in</strong>troduced by de Gennes. For a tube of diameter d the<br />

polymer can expand up to this diameter as if it were free, us<strong>in</strong>g g < N monomers.<br />

Us<strong>in</strong>g the real-cha<strong>in</strong> end-to-end scal<strong>in</strong>g, eq. (2.34), g can be calculated (cf. figure<br />

3.2)<br />

d ∼ (b 2 v) 1/5 g 3/5 ⇒ g ∼ d 5/3 (b 2 v) −1/3 . (3.4)<br />

The key assumption of the de Gennes scal<strong>in</strong>g argument is, that, due to selfavoidance,<br />

the ‘blobs’ behave as hard spheres. Hence N/g-‘blobs’ must be stacked<br />

l<strong>in</strong>early along the tube:<br />

R|| ∼ N<br />

d ∼ Nv1/3<br />

g<br />

2/3 b<br />

. (3.5)<br />

d<br />

Us<strong>in</strong>g the three different monomer volumes vi <strong>in</strong>troduced <strong>in</strong> section 2.5 we obta<strong>in</strong><br />

R||1 ∼ L <br />

b 2/3<br />

, (3.6a)<br />

d<br />

R||2 ∼ L bh<br />

d2 1/3 , (3.6b)<br />

R||3 ∼ L <br />

h 2/3<br />

. (3.6c)<br />

d<br />

Aga<strong>in</strong> the relation R||1 is mostly found <strong>in</strong> the literature, but the disk-like distribution<br />

of monomers R||2 is actually the right one to use for polymers with thickness<br />

h = 0.<br />

Actually R|| should better be called apparent parallel end-to-end distance Ra||,<br />

as expla<strong>in</strong>ed <strong>in</strong> section 4.6, because as the figure 3.2 suggests, the scal<strong>in</strong>g is valid<br />

for the maximal parallel extension 3 .<br />

Approximation for strong conf<strong>in</strong>ement<br />

When look<strong>in</strong>g at very strong conf<strong>in</strong>ement, a po<strong>in</strong>t is reached where the tube<br />

diameter d is smaller than the segment length b = L/N. In this case, when<br />

3 S<strong>in</strong>ce we are not go<strong>in</strong>g to <strong>in</strong>clude self-avoidance effects, for the reasons discussed on page<br />

18 and <strong>in</strong> section 4.4.2, we do not expect to see the scal<strong>in</strong>g derived here. When stretch<strong>in</strong>g for<br />

flexible cha<strong>in</strong>s is seen, then due to f<strong>in</strong>ite-size effects.


3.2. SCALING RELATIONS 25<br />

b<br />

a<br />

d<br />

R ||<br />

Figure 3.3: M<strong>in</strong>imal extension R || of a cha<strong>in</strong> with stiff segments of size b <strong>in</strong> a tube of diameter<br />

d.<br />

start<strong>in</strong>g from a totally stretched conformation and not allow<strong>in</strong>g ‘directional flips’<br />

between fore and aft—which are very unlikely for cha<strong>in</strong>s with persistence—the<br />

polymer can take a configuration with a m<strong>in</strong>imal end-to-end distance by mak<strong>in</strong>g<br />

a zig-zag configuration along the tube axis. This is calculated by us<strong>in</strong>g the<br />

Pythagoras’ theorem (cf. figure 3.3)<br />

where<br />

Hence:<br />

a = √ b 2 − d 2 = b<br />

m<strong>in</strong> R|| ≈ Na, (3.7)<br />

<br />

1 − 1<br />

2<br />

m<strong>in</strong> R|| = √ b 2 − d 2 ≈ L<br />

2 d<br />

+ O<br />

b<br />

<br />

1 − 1<br />

2<br />

<br />

4<br />

d<br />

. (3.8)<br />

b<br />

<br />

2<br />

d<br />

, (3.9)<br />

b<br />

for d ≪ b. This result is true even, when deal<strong>in</strong>g with an ideal phantom cha<strong>in</strong>,<br />

which should not show any <strong>in</strong>fluence of a conf<strong>in</strong><strong>in</strong>g environment. This argument<br />

is only due to f<strong>in</strong>ite-size effects and particularly <strong>in</strong>terest<strong>in</strong>g <strong>in</strong> connection with<br />

simulations (cf. section 4.4).<br />

3.2.2 The stiff regime<br />

Now turn<strong>in</strong>g to the limit of stiff polymers we can use aga<strong>in</strong> a ‘blob’-picture like<br />

argument to arrive at a scal<strong>in</strong>g relation. We know that the end-to-end distance<br />

of a stiff polymer scales as<br />

R<br />

L ∼<br />

⎧<br />

⎨1<br />

for strong conf<strong>in</strong>ement d → 0<br />

⎩1<br />

− L<br />

for no conf<strong>in</strong>ement d → ∞ and<br />

6lp<br />

L , (3.10)<br />

≪ 1<br />

lp<br />

where the second l<strong>in</strong>e is the square root of the series expansion of eq. (2.21).


26 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

In the spirit of de Gennes’ “blob picture” for flexible cha<strong>in</strong>s, we take the<br />

segments between two deflections as a ‘blob’. We assume that the polymer has<br />

a contour length of ld between two deflections separated a distance r along the<br />

tube (cf. 3.1). Inside these ‘blobs’ the segments behave free as <strong>in</strong> eq. (3.10) with<br />

the substitutions R → r and L → ld, lead<strong>in</strong>g to<br />

r<br />

ld<br />

∼ 1 − α ld<br />

. (3.11)<br />

An arbitrary factor α has been <strong>in</strong>troduced, which cannot be calculated <strong>in</strong> this<br />

scal<strong>in</strong>g argument, but can be used as a fit parameter. Aga<strong>in</strong> assum<strong>in</strong>g the ‘blobs’<br />

to be hard-walled due to self-avoidance, leads to a stack<strong>in</strong>g of L/ld pieces<br />

R ∼ L<br />

ld<br />

r ⇒ R<br />

L<br />

lp<br />

∼ 1 − αld<br />

lp<br />

∼ 1 − α ′<br />

<br />

d<br />

lp<br />

2/3<br />

, (3.12)<br />

us<strong>in</strong>g eq. (3.1), which is also the reason for the use of a second unknown parameter<br />

α ′ . This is our scal<strong>in</strong>g result, which suggests a data collapse for all polymers with<br />

the same ratio of ld/lp or d/lp, respectively.<br />

Stay<strong>in</strong>g <strong>in</strong> the current order of d/lp we can rewrite this result by add<strong>in</strong>g the<br />

higher order terms <strong>in</strong> a geometric series, obta<strong>in</strong><strong>in</strong>g a result which is positive over<br />

the whole range of the parameters and giv<strong>in</strong>g a heuristic <strong>in</strong>terpolation formula<br />

R<br />

L ∼<br />

<br />

1 + α ′<br />

d<br />

lp<br />

2/3 −1<br />

. (3.13)<br />

By chance <strong>in</strong>vestigat<strong>in</strong>g the wrong limit of a flexible cha<strong>in</strong> lp → 0—the scal<strong>in</strong>g<br />

argument is actually only valid for the stiff conf<strong>in</strong>ed cha<strong>in</strong> — we get, co<strong>in</strong>cidently,<br />

the same scal<strong>in</strong>g behavior as derived with the right assumption by de Gennes (cf.<br />

eq. (3.6a))<br />

R|| → L<br />

2/3 lp<br />

. (3.14)<br />

d<br />

This co<strong>in</strong>cidence is artificial and has no real importance, s<strong>in</strong>ce the resummation<br />

is not well controlled.<br />

Alternatively the scal<strong>in</strong>g can be obta<strong>in</strong>ed <strong>in</strong> a bit sturdier way, us<strong>in</strong>g the<br />

asymptotic expansions for a free stiff polymer derived by Yamakawa and Fujii <strong>in</strong><br />

[48]. This is valid as long as s ≪ lp (ˆt(0) is assumed to be fixed along the z-axis)<br />

<br />

2 2<br />

R (s) = s 1 − s<br />

3lp<br />

R(s) · ˆt(0) 2 <br />

= z 2 (s) = s 2<br />

+ O s 2 l −2<br />

p<br />

<br />

<br />

1 − s<br />

lp<br />

, (3.15)<br />

+ O s 2 l −2<br />

p<br />

<br />

. (3.16)


3.3. ANALYTICAL RESULTS 27<br />

The mean-square average deviation from the z-axis (undulation) is thus given by<br />

r⊥(s) 2 = x 2 (s) + y 2 (s) = 2 x 2 (s) = 2<br />

3<br />

s 3<br />

lp<br />

. (3.17)<br />

We can aga<strong>in</strong> <strong>in</strong>troduce the length scale ld, now follow<strong>in</strong>g directly the l<strong>in</strong>e of<br />

reason<strong>in</strong>g of Odijk <strong>in</strong> [37]. When conf<strong>in</strong><strong>in</strong>g the cha<strong>in</strong> <strong>in</strong>to a cyl<strong>in</strong>drical tube of<br />

diameter d ≪ lp, with its axis along the z-direction, the contour length ld can be<br />

calculated by us<strong>in</strong>g eq. (3.17).<br />

As a measure for the distance on the contour between two deflection po<strong>in</strong>ts,<br />

ld can be def<strong>in</strong>ed by assum<strong>in</strong>g that the mean-square average lateral deviation<br />

should be of the same order as the radius of the tube (d/2) 2<br />

r⊥(ld) 2 ∼<br />

2 d<br />

2<br />

⇒ l 3 d ∼ d 2 lp. (3.18)<br />

In this way we reobta<strong>in</strong>ed the important relation eq. (3.1). By this, the scal<strong>in</strong>g for<br />

the end-to-end distance <strong>in</strong> the strongly conf<strong>in</strong>ed regime can be obta<strong>in</strong>ed, stack<strong>in</strong>g<br />

l<strong>in</strong>early L/ld pieces of the polymer of length 〈z(ld) 2 〉 gives<br />

R<br />

L ∼ 〈z(ld) 2 〉/ld ∼ 1 − α ′<br />

d<br />

lp<br />

2/3<br />

, (3.19)<br />

where α ′ is some unknown constant (cf. eq. (3.12)).<br />

The results obta<strong>in</strong>ed here should be valid for all flexibilities as long as the<br />

conf<strong>in</strong>ement is very strong. But for the <strong>in</strong>termediate regime where d ≈ R and<br />

the polymer is semi-flexible lp ≈ L simulations become essential.<br />

3.3 Analytical results<br />

After this prepar<strong>in</strong>g scal<strong>in</strong>g analysis, we address the task of analytical approximations.<br />

In the context of analytical solutions one has to th<strong>in</strong>k about <strong>in</strong> what<br />

complexity a problem is feasible to solve. Every reasonable behav<strong>in</strong>g symmetric<br />

potential is expandable <strong>in</strong> a power series, the first non-trivial term be<strong>in</strong>g harmonic.<br />

These quadratic terms often result <strong>in</strong> Gaussian <strong>in</strong>tegrals, which can be<br />

handled analytically. Already the <strong>in</strong>troduction of slightly more complicated term<br />

renders the <strong>in</strong>tegrals unsolvable and forces to use stronger approximations.<br />

We are go<strong>in</strong>g to conf<strong>in</strong>e a polymer <strong>in</strong> a cyl<strong>in</strong>drical symmetric environment,<br />

hence we <strong>in</strong>troduce a cyl<strong>in</strong>drical symmetric harmonic potential. In the limit of<br />

the weakly-bend<strong>in</strong>g rod, which is only justified when the tube is very narrow<br />

(ld ≪ L and/or lp ≫ L, cf. 2.4.1), some calculation are rendered possible s<strong>in</strong>ce<br />

the <strong>in</strong>extensibility constra<strong>in</strong>t has not to be taken <strong>in</strong>to account explicitly.


28 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

Choos<strong>in</strong>g the preferred direction of the parametrization along the symmetry<br />

axis of the tube, the harmonic potential is just a harmonic term on the undulations.<br />

With eq. (2.31) we arrive at the full Hamiltonian<br />

H[r⊥(s)] = κ<br />

2<br />

= κ<br />

2<br />

L<br />

0<br />

L<br />

0<br />

ds ¨r⊥(s) 2 + γ<br />

2<br />

L<br />

0<br />

ds r⊥(s) 2<br />

ds ¨x(s) 2 + ¨y(s) 2 + γ<br />

2<br />

L<br />

0<br />

ds x(s) 2 + y(s) 2 ,<br />

(3.20)<br />

where γ is the potential strength and characteriz<strong>in</strong>g the degree of conf<strong>in</strong>ement.<br />

3.3.1 Averages by Fourier transformation<br />

We try to obta<strong>in</strong> analytical results by <strong>in</strong>troduc<strong>in</strong>g the Fourier <strong>in</strong>tegrals <strong>in</strong> the<br />

Hamiltonian eq. (3.20) to get rid of the derivatives <strong>in</strong> the <strong>in</strong>tegral. This is an<br />

approximation, s<strong>in</strong>ce the Fourier <strong>in</strong>tegral actually means deal<strong>in</strong>g with an <strong>in</strong>f<strong>in</strong>itely<br />

long polymer. It is ignored to <strong>in</strong>troduce reasonable boundary conditions.<br />

We use the notation<br />

<br />

dk<br />

r⊥(s) =<br />

2π ˜r⊥(k)e iks , (3.21a)<br />

<br />

˜r⊥(k) = ds r⊥(s)e −iks , (3.21b)<br />

lead<strong>in</strong>g to the Hamiltonian<br />

<br />

dk 1 4 2 2<br />

H = κk + γ |˜x(k)| + |˜y(k)|<br />

2π 2<br />

. (3.22)<br />

Us<strong>in</strong>g the equipartition theorem <strong>in</strong> Fourier space (cf. [26]) the result for the<br />

undulation correlation <strong>in</strong> Fourier space is obta<strong>in</strong>ed<br />

〈˜x(k)˜x ∗ (k ′ )〉 = 〈˜y(k)˜y ∗ (k ′ )〉 = 2πδ(k − k ′ ) kBT<br />

κk4 . (3.23)<br />

+ γ<br />

The x and y components are equivalent due to the cyl<strong>in</strong>drical symmetry.<br />

A characteristic length scale can be identified as<br />

<br />

ld := 4 κ<br />

1/4 . (3.24)<br />

γ<br />

Tak<strong>in</strong>g the factor 4 <strong>in</strong>to the def<strong>in</strong>ition is an arbitrary choice s<strong>in</strong>ce we are deal<strong>in</strong>g<br />

with a scale. This scale is a measure of the conf<strong>in</strong>ement and can therefore be<br />

chosen as our Odijk deflection length ld. This will be more apparent when look<strong>in</strong>g<br />

e.g. at the result <strong>in</strong> eq. (3.27). The correlation function is then rewritten to<br />

〈˜x(k)˜x ∗ (k ′ )〉 = 2πδ(k − k ′ )<br />

kBT<br />

κ(k4 + 4l −4<br />

(3.25)<br />

d ).


3.3. ANALYTICAL RESULTS 29<br />

From this equation the correlation function <strong>in</strong> real-space can be obta<strong>in</strong>ed by a<br />

Fourier back-transform. The details of the lengthy calculation can be found <strong>in</strong><br />

appendix B.1. The result is<br />

〈x(s)x(s ′ )〉 = kBT l 3 d<br />

4 √ 2κ exp<br />

<br />

− |s − s′ |<br />

Reduc<strong>in</strong>g for s = s ′ to the mean-square undulation<br />

ld<br />

x 2 = kBT l3 d<br />

8κ<br />

′ |s − s |<br />

cos −<br />

ld<br />

π<br />

<br />

. (3.26)<br />

4<br />

1<br />

=<br />

8<br />

l 3 d<br />

lp<br />

, (3.27)<br />

by us<strong>in</strong>g eq. (2.20). This is <strong>in</strong>dependent of the current position on the polymer,<br />

which is reassur<strong>in</strong>g s<strong>in</strong>ce we have assumed that the polymer is <strong>in</strong>f<strong>in</strong>itely long and<br />

there is translational <strong>in</strong>variance along the symmetry axis of the tube (z-direction).<br />

Here the idea of the deflection length can be depicted a bit clearer. As already<br />

mentioned <strong>in</strong> section 3.2 on page 22, l3 d is approximately the volume occupied by<br />

segments between two deflections. The mean-square undulation scal<strong>in</strong>g is then<br />

given by this volume divided by the length-scale on which the polymer is flexible<br />

lp. This gives the expected relation: 〈x2 〉 ∼ l3 d /lp.<br />

We want to use the harmonic potential to model—<strong>in</strong> first order—other types<br />

of conf<strong>in</strong><strong>in</strong>g potential, <strong>in</strong> particular a hard-wall tube model. The problem of<br />

compar<strong>in</strong>g the two models is, that a harmonic potential only has an energy scale<br />

given by ld through the potential strength γ, but has no <strong>in</strong>tr<strong>in</strong>sic length scale.<br />

In contrast, a hard-walled tube has a length scale given by the tube diameter d,<br />

but no energy scale, s<strong>in</strong>ce the polymer is free, as long it does not hit the tube.<br />

Therefore only a heuristic length-scale comparison is possible, which is already<br />

<strong>in</strong>dicated by eq. (3.1).<br />

From our current calculation it suggests itself to compare the mean-square<br />

undulations 〈r2 ⊥ 〉 with the length scale conf<strong>in</strong><strong>in</strong>g these undulations, the tube<br />

diameter d. Sett<strong>in</strong>g the radius of tube as the conf<strong>in</strong><strong>in</strong>g lateral distance<br />

2 2 d<br />

r⊥ ≈ , (3.28)<br />

2<br />

we obta<strong>in</strong> the relation<br />

l 3 d ≈ d 2 lp, (3.29)<br />

to be compared to eq. (3.1) (cf. also [21, section 9.1]).<br />

In the follow<strong>in</strong>g the derivatives of eqs. (3.26) and (3.27) along the contour will<br />

be needed. Therefore they are presented here:<br />

〈 ˙x(s) ˙x(s ′ )〉 = − ld<br />

2 √ e<br />

2 lp<br />

− |s−s′ |<br />

ld s<strong>in</strong><br />

|s − s ′ |<br />

ld<br />

− π<br />

<br />

4<br />

(3.30)<br />

⇒ ˙x 2 = ld<br />

. (3.31)<br />

4 lp


30 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

〈t(0.5) · t(s/L + 0.5)〉<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

-0.4<br />

-0.2<br />

exact free<br />

ld/L = 10<br />

ld/L = 1<br />

ld/L = 0.5<br />

ld/L = 0.1<br />

ld/L = 0.05<br />

Figure 3.4: Tangent-tangent correlation function 〈t(0.5) · t(s/L + 0.5)〉 for different conf<strong>in</strong><strong>in</strong>g<br />

potentials measured <strong>in</strong> terms of ld/L. For the persistence length we have chosen lp/L = 1.<br />

Tangent-tangent correlation function<br />

From eq. (3.26) the tangent-tangent correlation is easily calculated. The representation<br />

of the tangents <strong>in</strong> the weakly-bend<strong>in</strong>g-rod approximation is<br />

<br />

t(s) ≈ ˙x(s), ˙y(s), 1 − 1<br />

2 2<br />

˙x(s) + ˙y(s) 2<br />

<br />

. (3.32)<br />

Insert<strong>in</strong>g and expand<strong>in</strong>g gives<br />

0<br />

s/L<br />

〈t(s) · t(s ′ )〉 = 1 + 〈 ˙x(s) ˙x(s ′ )〉 + 〈 ˙y(s) ˙y(s ′ )〉<br />

2 ′ 2 2 ′ 2<br />

˙x(s) + ˙x(s ) + ˙y(s) + ˙y(s ) <br />

− 1<br />

2<br />

− 1 2 ′ 2 2 ′ 2 2 ′ 2 2 ′ 2<br />

˙x(s) ˙x(s ) + ˙y(s) ˙x(s ) + ˙y(s) ˙x(s ) + ˙y(s) ˙y(s )<br />

4<br />

<br />

= 1 + 2 〈 ˙x(s) ˙x(s ′ )〉 − 2 ˙x 2 . (3.33)<br />

Only terms to second order are kept, the statistical <strong>in</strong>dependence of the x- and<br />

y-direction, the translational <strong>in</strong>variance of eq. (3.27) and f<strong>in</strong>ally the rotational<br />

symmetry around the z-axis have been used.<br />

Insert<strong>in</strong>g the results eqs. (3.30) and (3.31) gives<br />

〈t(s) · t(s ′ )〉 = 1 − ld<br />

<br />

√2<br />

exp −<br />

2 lp<br />

|s − s′ ′ | |s − s |<br />

s<strong>in</strong> −<br />

ld<br />

ld<br />

π<br />

<br />

+ 1<br />

4<br />

(3.34)<br />

This result is sketched <strong>in</strong> figure 3.4 for lp/L = 1 and different degrees of<br />

conf<strong>in</strong>ement. Obviously, for <strong>in</strong>creas<strong>in</strong>g conf<strong>in</strong>ement, the correlations along the<br />

whole polymer do not decay exponentially anymore (as <strong>in</strong> eq. (2.12)), but reach<br />

0.2<br />

0.4<br />

.


3.3. ANALYTICAL RESULTS 31<br />

a constant non-zero smaller one, that can be calculated, for a distance along the<br />

polymer large compared to the deflection length |s − s ′ | ≫ ld, to<br />

〈t(s) · t(s ′ )〉 = 1 − ld<br />

. (3.35)<br />

2lp<br />

This is due to the restriction of possible orientations. For strong conf<strong>in</strong>ement the<br />

polymer is only allowed to orient along the tube, which gives rise to constant<br />

non-zero values for the tangent-tangent correlation. But there is still a short<br />

range which is of the order ld where the correlation decay as <strong>in</strong> the free case<br />

(exponentially) before ‘notic<strong>in</strong>g’ the <strong>in</strong>fluence of the tube and becom<strong>in</strong>g constant.<br />

This is best seen for ld = 0.1. The short dip below the constant value is due to<br />

the deflection (this is better seen <strong>in</strong> figure 5.4(d) on page 62).<br />

The exact exponential l<strong>in</strong>e from eq. (2.12) is shown <strong>in</strong> figure 3.4 as well and<br />

is quite different from the near free result for ld = 10. This is due to the fact,<br />

that our result (3.34) is consistent with the exact result only up to l<strong>in</strong>ear order<br />

<strong>in</strong> L/lp, which is seen upon expand<strong>in</strong>g for the free limit ld → ∞<br />

〈t(s) · t(s ′ )〉 = 1 − |s − s′ |<br />

lp<br />

<br />

≈ exp − |s − s′ |<br />

lp<br />

+ O(l −1<br />

d )<br />

<br />

,<br />

(3.36)<br />

where <strong>in</strong> the last l<strong>in</strong>e the expansion is resummed to give, <strong>in</strong> the same order, the<br />

expansion of the exponential. This is the same result known for the a semi-flexible<br />

polymer <strong>in</strong> bulk solution, eq. (2.12), up to first order L/lp, i.e. for the stiff limit.<br />

This shows that our condition that L ≪ lp is necessary, but if lp ≫ ld we obta<strong>in</strong><br />

good results too, look<strong>in</strong>g at eq. (3.34). For further discussion, see the section<br />

5.2.2 about the simulation results.<br />

Mean-square end-to-end distance<br />

The mean-square end-to-end distance 〈R 2 (L)〉 can be calculated by a simple but<br />

tedious <strong>in</strong>tegration over the tangent-tangent correlation function. The details can<br />

be found <strong>in</strong> appendix B.2. There we obta<strong>in</strong> the result for the correlation function<br />

of the end-to-end vectors<br />

〈R(s) · R(s ′ )〉 =<br />

<br />

1 − ld<br />

<br />

ss<br />

2 lp<br />

′ + l3 d<br />

2 √ 2 lp<br />

s<br />

− l − e d s<strong>in</strong><br />

s<br />

ld<br />

+ π<br />

4<br />

<br />

<br />

e − |s−s′ <br />

|<br />

′ |s − s |<br />

ld s<strong>in</strong><br />

ld<br />

<br />

s′ ′<br />

− s l − e d s<strong>in</strong> +<br />

ld<br />

π<br />

4<br />

<br />

+ π<br />

<br />

4<br />

+ 1<br />

√ 2<br />

<br />

,<br />

(3.37)<br />

which reduces for s = s ′ = L to the the mean-square end-to-end distance<br />

<br />

2<br />

R (L) = 1 − ld<br />

<br />

L<br />

2 lp<br />

2 + l3 <br />

d<br />

1 −<br />

2 lp<br />

√ <br />

L<br />

− L<br />

l 2 e d s<strong>in</strong> +<br />

ld<br />

π<br />

<br />

. (3.38)<br />

4


32 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

R 2 (L) /L 2<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.01<br />

0.1<br />

1<br />

L/ld<br />

10<br />

lp/L = 0.1<br />

lp/L = 0.5<br />

lp/L = 1<br />

lp/L = 5<br />

lp/L = 10<br />

Figure 3.5: Mean-square end-to-end distance R 2 versus the conf<strong>in</strong>ement strength L/ld.<br />

Shown are curves for different stiffnesses, from the stiff lp/L = 0.1 to the flexible lp/L = 10<br />

regime.<br />

This depends explicitly on two ld/lp and ld/L of the three possible parameters<br />

L, lp and ld (or comb<strong>in</strong>ations thereof). But we expect the ma<strong>in</strong> dependence only<br />

on ld/lp due to our scal<strong>in</strong>g analysis <strong>in</strong> section 3.2.2. It can be seen that this is<br />

actually true also for result above, s<strong>in</strong>ce it is valid only where the weakly-bend<strong>in</strong>g<br />

rod approximation is applicable, which is for lp ≫ L and/or lp ≫ ld. Therefore<br />

eq. (3.38) should only be considered <strong>in</strong> this range 4 .<br />

The limit of large deflection length ld → ∞ corresponds to a free polymer.<br />

Hence we should recover the results for a polymer <strong>in</strong> bulk solution. The series<br />

expansion <strong>in</strong> the stiff regime, up to first order L/lp, of the above result, eq. (3.38),<br />

is identical to that of the bulk solution, eq. (2.21).<br />

Parallel mean-square end-to-end distance<br />

100<br />

1000<br />

<br />

2 2<br />

R (L) = L 1 − L<br />

<br />

, (3.39)<br />

3 lp<br />

In a slightly different way a similar expressions for the end-to-end distance and<br />

correlation function projected onto the symmetry axis can be derived. This will<br />

be denoted by R2 <br />

|| . In our notation this is just the z-component of the end-toend<br />

distance.<br />

Start<strong>in</strong>g with the relation eq. (2.27) and the above result for the undulation<br />

correlation, eq. (3.26), we can use the Wick theorem to arrive at a relation for<br />

4 E.g. for lp/L = 0.5 the part depend<strong>in</strong>g explicitly on L/ld changes the result for the meansquare<br />

end-to-end distance only by 14% for L/ld = 1 and 0.03% for L/ld = 10.


3.3. ANALYTICAL RESULTS 33<br />

the projected length correlation<br />

˙R||(s) ˙ R||(s ′ ) = 1 − ˙r||(s) 1 − ˙r||(s ′ ) <br />

= 1 − 2 ˙x 2 + 1<br />

2 ′ 2<br />

˙x(s) ˙x(s )<br />

4<br />

+ ˙y(s) 2 ˙y(s ′ ) 2<br />

+ ˙x(s) 2 ˙y(s ′ ) 2 + ˙y(s) 2 ˙x(s ′ <br />

)<br />

2<br />

= 1 − 2 ˙x 2 + ˙x 2 2 + 〈 ˙x(s) ˙x(s ′ )〉 2 . (3.40)<br />

It has been used that the x- and y-directions are uncorrelated an that the system<br />

is cyl<strong>in</strong>drical symmetric.<br />

Now aga<strong>in</strong> a tedious <strong>in</strong>tegration—the details are found <strong>in</strong> appendix B.3—<br />

results <strong>in</strong><br />

<br />

R||(s)R||(s ′ ) <br />

= 1 − ld<br />

<br />

ss<br />

2 lp<br />

′ + r||(s)r||(s ′ ) <br />

<br />

= 1 − ld<br />

+<br />

2 lp<br />

l2 d<br />

16 l2 <br />

ss<br />

p<br />

′ + l3 d<br />

32 l2 m<strong>in</strong>(s, s<br />

p<br />

′ )<br />

<br />

2 − cos 2 |s − s′ <br />

|<br />

+ l4 d<br />

128 l 2 p<br />

2s<br />

− l + e d<br />

− e −2 |s−s′ |<br />

ld <br />

2 − cos 2s<br />

ld<br />

<br />

ld<br />

<br />

2s′<br />

− l + e d 2 − cos 2s′<br />

ld<br />

(3.41)<br />

<br />

− 1 .<br />

As <strong>in</strong>dicated <strong>in</strong> the first l<strong>in</strong>e, the first two summands are miss<strong>in</strong>g if the stored<br />

length correlation function r||(s)r||(s ′ ) is exam<strong>in</strong>ed.<br />

For s = s ′ = L this reduces to mean-square parallel end-to-end distance<br />

R||(L) 2 =<br />

<br />

1 − ld<br />

+<br />

2 lp<br />

l2 d<br />

16 l2 p<br />

+ l3 d<br />

32 l2 L +<br />

p<br />

l4 d<br />

64 l2 p<br />

<br />

L 2<br />

<br />

e −2L/ld<br />

<br />

2 − cos 2L<br />

<br />

− 1 .<br />

ld<br />

(3.42)<br />

The problem with these results is, that they diverge <strong>in</strong> the limit of no conf<strong>in</strong>ement,<br />

that is for ld → ∞<br />

r||(s)r||(s ′ ) → ∞, (3.43)<br />

which is unfortunately not a sane result.<br />

An approximate derivation of the mean parallel end-to-end distance<br />

A direct way to calculate the mean projected length R||(L) is the <strong>in</strong>tegral<br />

L =<br />

R||(L)<br />

0<br />

<br />

ds 1 + ˙r⊥(s) 2 , (3.44)


34 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

measur<strong>in</strong>g the contour length of a space curve <strong>in</strong> the usual way (cf. [18]). Expand<strong>in</strong>g<br />

the square root <strong>in</strong> the weakly-bend<strong>in</strong>g rod limit |˙r⊥| ≪ 1<br />

L ≈ R||(L) + 1<br />

2<br />

= R||(L) + 1<br />

2<br />

R||(L)<br />

0<br />

R||(L)<br />

0<br />

ds ˙r⊥(s) 2<br />

ds ˙x(s) 2 + ˙y(s) 2 , (3.45)<br />

and tak<strong>in</strong>g the ensemble average, where we approximate the average over the<br />

<strong>in</strong>tegral by the <strong>in</strong>tegral of the average. Us<strong>in</strong>g the symmetry of the problem and<br />

eq. (3.27) gives<br />

The result f<strong>in</strong>ally is<br />

R||(L) =<br />

L ≈ R||(L) + 1<br />

2<br />

= <br />

<br />

R||(L)<br />

L<br />

1 + ld<br />

4 lp<br />

〈R ||(L)〉<br />

0<br />

1 + ld<br />

4 lp<br />

ds 2 ˙x 2<br />

<br />

. (3.46)<br />

<br />

≈ L 1 − ld<br />

+<br />

4 lp<br />

l2 d<br />

16 l2 p<br />

<br />

+ · · · , (3.47)<br />

with an approximation for the stiff/conf<strong>in</strong>ed limit ld/lp ≪ 1. This result is up to<br />

first order identical to the square-root of eq. (3.42).<br />

3.3.2 Averages by Fourier series<br />

Instead of us<strong>in</strong>g Fourier <strong>in</strong>tegrals, which supposes the approximation that the<br />

polymer is <strong>in</strong>f<strong>in</strong>itely long, Fourier series can be used. This means fix<strong>in</strong>g specific<br />

boundary condition on the polymer5 . We are go<strong>in</strong>g to use “tracked ends” x(0) =<br />

y(0) = x(L) = y(L) = 0, where the end po<strong>in</strong>ts are fixed on but free to move<br />

along the tube-axis, s<strong>in</strong>ce the eigenfunctions for free ends render the calculations<br />

to tedious. The “tracked ends” lead to a normal Fourier-s<strong>in</strong>us transformation,<br />

with the notation<br />

r⊥(s) = 2 <br />

hk s<strong>in</strong>(ks), (3.48a)<br />

L<br />

with eigenvalues k = n π<br />

L<br />

k<br />

L<br />

hk =<br />

0<br />

(n ∈ N).<br />

ds r⊥(s) s<strong>in</strong>(ks), (3.48b)<br />

5 Different boundary condition <strong>in</strong> this problem means, us<strong>in</strong>g a different set of eigenfunctions<br />

of the operator ∇ 4 , which are not just s<strong>in</strong>es, but comb<strong>in</strong>ation of the s<strong>in</strong>es, cos<strong>in</strong>es and their<br />

hyperbolic counterparts. The eigenfunctions are listed <strong>in</strong> [46].


3.3. ANALYTICAL RESULTS 35<br />

x 2 (s/L) <br />

1.5e-05<br />

1e-05<br />

5e-06<br />

0<br />

0<br />

0.2<br />

Integral ansatz<br />

Series ansatz<br />

0.4<br />

Figure 3.6: Comparison of the mean-square undulations along the polymer x 2 (s/L) versus<br />

the contour length s/L obta<strong>in</strong>ed by Fourier <strong>in</strong>tegral and Fourier series (‘tracked ends’)<br />

calculation.<br />

Insert<strong>in</strong>g this transformation <strong>in</strong>to the Hamiltonian eq. (3.20) and us<strong>in</strong>g the<br />

equipartition theorem we arrives at (<strong>in</strong> comparison to eq. (3.25))<br />

s/L<br />

L<br />

〈xkxk ′〉 = 〈ykyk ′〉 = δkk ′<br />

2<br />

lp<br />

0.6<br />

0.8<br />

1<br />

.<br />

k4 −4<br />

(3.49)<br />

+ 4ld The back-transform—which is very tedious, and is not presented here—can<br />

be calculated exactly. We obta<strong>in</strong><br />

〈x(s)x(s ′ )〉 = L2<br />

2<br />

ɛ<br />

c 3<br />

1<br />

s<strong>in</strong>h <br />

c cos c − cosh c s<strong>in</strong> c<br />

cos 2c − cosh 2c<br />

· cosh(c − sc) s<strong>in</strong>(c − sc) s<strong>in</strong>h s ′ c cos s ′ c<br />

+ s<strong>in</strong>h(c − sc) cos(c − sc) cosh s ′ c s<strong>in</strong> s ′ c<br />

+ s<strong>in</strong>h c cos c + cosh c s<strong>in</strong> c <br />

· cosh(c − sc) s<strong>in</strong>(c − sc) cosh s ′ c s<strong>in</strong> s ′ c<br />

− s<strong>in</strong>h(c − sc) cos(c − sc) s<strong>in</strong>h s ′ c cos s ′ c<br />

<br />

1<br />

<br />

,<br />

(3.50)<br />

which is valid for s > s ′ , us<strong>in</strong>g besides the dimensionless parameters ɛ and c,<br />

the abbreviations sc = s/ld and s ′ c = s ′ /ld. For further calculations this is a<br />

very <strong>in</strong>convenient expression. To obta<strong>in</strong> more results an algebraic program like<br />

mathematica should be used.<br />

For s = s ′ the mean-square undulations are obta<strong>in</strong>ed, which is compared to<br />

eq. (3.27) <strong>in</strong> figure 3.6. The results for the <strong>in</strong>ner part of the polymer are identical.<br />

Due to the boundary conditions, the result obta<strong>in</strong>ed by Fourier series has to<br />

go to zero at the boundaries, <strong>in</strong> contrast to the Fourier <strong>in</strong>tegral solution, which<br />

stays constant. In the range of validity of the weakly-bend<strong>in</strong>g rod approximation


36 CHAPTER 3. POLYMERS IN CONFINED GEOMETRY<br />

the differences between the approaches will not effect the results of calculations,<br />

besides m<strong>in</strong>or boundary effects. Therefore the series ansatz is not further <strong>in</strong>vestigated.<br />

The calculation become very tedious.<br />

3.3.3 Radial distribution function<br />

The radial distribution function (RDF) p(R) is another <strong>in</strong>terest<strong>in</strong>g quantity. It is<br />

the probability distribution of the end-to-end distance vector R that <strong>in</strong>corporates<br />

more <strong>in</strong>formation than only the moments. It is def<strong>in</strong>ed as<br />

p(R) = 〈δ(R − |R|)〉 . (3.51)<br />

All configurations with the specified end-to-end distance |R| = R are counted.<br />

For the free polymer this quantity can be calculated—or at least approximated—<br />

over the whole stiffness range. The result for the Gaussian cha<strong>in</strong> was shown <strong>in</strong><br />

eq. (2.5). For other regions confer [7] and [47].<br />

The calculations to obta<strong>in</strong> the RDF are quite <strong>in</strong>volved for the free polymer.<br />

For a conf<strong>in</strong>ed polymer this becomes even more difficult. The advantage of the<br />

polymer <strong>in</strong> bulk solution is, that the problem can be formulated by exclusively<br />

us<strong>in</strong>g tangent-vectors describ<strong>in</strong>g the segments, <strong>in</strong>dependent of the embedd<strong>in</strong>g<br />

space. By <strong>in</strong>troduc<strong>in</strong>g a conf<strong>in</strong><strong>in</strong>g environment a non-local <strong>in</strong>teraction term has<br />

to be <strong>in</strong>troduced, s<strong>in</strong>ce the problem is now dependent on the actual positions of<br />

the cha<strong>in</strong> jo<strong>in</strong>ts, which are given as <strong>in</strong>tegrals over the tangents. In Fourier space<br />

problems arise due to higher order modes.<br />

Therefore we do without an analytical approach and resort to simulation<br />

results discussed <strong>in</strong> section 5.2.5.


Chapter 4<br />

Simulation methods<br />

This chapter starts with a review of the Monte-Carlo method <strong>in</strong> general and how<br />

to calculate partition sums <strong>in</strong> the canonical ensemble. The Metropolis algorithm<br />

is <strong>in</strong>troduced as a method to implement the idea of Markov cha<strong>in</strong>s <strong>in</strong> computer<br />

programs. In this context the problem of correlated sampl<strong>in</strong>g and the importance<br />

of the quality of random number generators are discussed.<br />

Then we turn to the discussion of these methods <strong>in</strong> the context of the wormlike<br />

cha<strong>in</strong> model <strong>in</strong> bulk solution and <strong>in</strong> conf<strong>in</strong>ed environments. F<strong>in</strong>ally we give<br />

a list of the important observables which are obta<strong>in</strong>ed <strong>in</strong> simulations and where<br />

they are used.<br />

4.1 Motivation<br />

As <strong>in</strong> most cases <strong>in</strong> physics there is no way of exploit<strong>in</strong>g all <strong>in</strong>terest<strong>in</strong>g parameter<br />

values of a given problem analytically. So one has to resort to simulation<br />

techniques [4, 13, 22]. We saw <strong>in</strong> the chapters 2 and 3 that analytical results are<br />

limited to either the very stiff and flexible polymer, or the strongly conf<strong>in</strong>ed and<br />

totally free environment.<br />

For problems from biotechnology it is very important to obta<strong>in</strong> quantitative<br />

results particularly <strong>in</strong> the <strong>in</strong>termediate regime—where up to now only qualitative<br />

scal<strong>in</strong>g arguments were available (cf. section 3.2). The problem of flexible DNA<br />

conf<strong>in</strong>ed to nano-channels with lateral dimensions <strong>in</strong> the order of the persistence<br />

length, as <strong>in</strong>troduced <strong>in</strong> section 3.1, is just <strong>in</strong> this complicated parameter regime.<br />

Also the properties of semi-flexible F-Act<strong>in</strong> filaments, with a contour-length of<br />

about the persistence length, are <strong>in</strong> this parameter range.<br />

We therefore have to rely on simulation techniques. The Monte-Carlo simulation<br />

is the method of choice, s<strong>in</strong>ce it allows to effectively sample the important<br />

conformations of the polymers. The problem of other methods, e.g. molecular dynamics<br />

(MD) simulations, is the large difference <strong>in</strong> time-scales. Typical changes<br />

on the monomer level are much faster than conformational changes of the whole<br />

37


38 CHAPTER 4. SIMULATION METHODS<br />

polymer with N ≫ 1 monomers. S<strong>in</strong>ce <strong>in</strong> MD simulation the smallest time-scale<br />

of the model must be simulated, it cannot be effectively applied to our problem.<br />

We therefore concentrate on the Monte-Carlo simulation <strong>in</strong> the follow<strong>in</strong>g sections.<br />

For a very good <strong>in</strong>troduction also refer to [35].<br />

4.2 The Monte-Carlo method<br />

In statistical physics ensemble averages of fluctuat<strong>in</strong>g quantities are of special<br />

<strong>in</strong>terest. To obta<strong>in</strong> these averages, <strong>in</strong>tegrals of the follow<strong>in</strong>g form have to be<br />

evaluated (schematically for one dimension)<br />

〈A〉 p =<br />

+∞<br />

−∞<br />

dx A(x)p(x) . (4.1)<br />

<br />

φ(x)<br />

This example could be an average of an observable A(x) (“score function”)<br />

weighted by the probability distribution p(x), where the <strong>in</strong>tegral corresponds—<strong>in</strong><br />

the language of statistical mechanics—to a ‘sum’ over all states. But the method<br />

expla<strong>in</strong>ed <strong>in</strong> the follow<strong>in</strong>g is applicable to every <strong>in</strong>tegral presentable <strong>in</strong> this form.<br />

When eq. (4.1) is not analytically solvable then one has to resort to numerical<br />

methods. The direct way would be to divide the <strong>in</strong>tegration range—which can<br />

be of any dimension—<strong>in</strong>to n equidistant <strong>in</strong>tervals and evaluate the sum over<br />

{φ(xi)} on the discrete set of po<strong>in</strong>ts {xi} (Riemann <strong>in</strong>tegral). In this way the<br />

error of the approximative result <strong>in</strong> this determ<strong>in</strong>istic algorithm reduces with<br />

n −1/d depend<strong>in</strong>g on the dimensionality d. Therefore this method becomes very<br />

<strong>in</strong>effective for higher dimensions d. But <strong>in</strong> particular high-dimensional <strong>in</strong>tegrals<br />

is what we are calculat<strong>in</strong>g <strong>in</strong> statistical mechanical problems, therefore we need<br />

a smart way to <strong>in</strong>crease the efficiency of the numerical <strong>in</strong>tegration.<br />

The basic idea of Monte-Carlo 1 evaluation of an <strong>in</strong>tegral is to choose the n<br />

discretization po<strong>in</strong>ts/sample po<strong>in</strong>ts {xi} randomly out of the probability distribution<br />

p(x) such that the sum has to be performed only over the observable<br />

quantity at the so generated po<strong>in</strong>ts {A(xi)}. This algorithm has the advantage<br />

of a statistical error proportional to n −1/2 , <strong>in</strong>dependent of the dimension which<br />

outperforms the determ<strong>in</strong>istic calculation for n ≥ 3. But both algorithms have<br />

the problem, that they are very <strong>in</strong>efficient, when the problem under <strong>in</strong>vestigation<br />

has a large <strong>in</strong>herent variance, due to the fact that the computational time grows<br />

proportional to n and the variance is <strong>in</strong>versely proportional to √ n.<br />

This problem is seen <strong>in</strong> the follow<strong>in</strong>g way: If the probability distribution of<br />

the random numbers is such, that the important region of the observable is poorly<br />

sampled and therefore the variance is large, a huge number of samples may be<br />

needed to get a reliable result. It may happen, that the <strong>in</strong>terest<strong>in</strong>g region is<br />

1 The name ought to br<strong>in</strong>g the cas<strong>in</strong>os of Monte-Carlo to m<strong>in</strong>d.


4.3. SAMPLING THE CANONICAL ENSEMBLE 39<br />

barely hit at all dur<strong>in</strong>g the numerical evaluation procedure. Therefore so called<br />

variance reduction techniques are needed.<br />

The most prom<strong>in</strong>ent idea is that of importance sampl<strong>in</strong>g. By replac<strong>in</strong>g the<br />

<strong>in</strong>itial probability distribution p(x) for the random numbers by an importance<br />

density q(x)—which should be chosen such that the poorly sampled regions are<br />

hit more frequently, otherwise the problem can be made worse—one can reduce<br />

the need for a large number of samples significantly! When us<strong>in</strong>g the importance<br />

density an additional step <strong>in</strong> the evaluation has to be performed: S<strong>in</strong>ce the<br />

probability distribution of the random samples has been changed to q(x) the<br />

score function has to be adjusted to Aq(x) = A(x)f(x)/q(x). It can be shown<br />

that this results <strong>in</strong> the same value for the desired average of the <strong>in</strong>tegral but the<br />

variance can be reduced by orders of magnitudes.<br />

The nice idea of simple Monte-Carlo <strong>in</strong>tegration can reduce the error <strong>in</strong> high<br />

dimensional <strong>in</strong>tegrals, but without the importance sampl<strong>in</strong>g method still a large<br />

number of problems would not be solvable <strong>in</strong> a reasonable amount of comput<strong>in</strong>g<br />

time!<br />

4.3 Sampl<strong>in</strong>g the canonical ensemble<br />

In our problem of sampl<strong>in</strong>g averages for polymers we are work<strong>in</strong>g <strong>in</strong> a canonical<br />

ensemble. Hence we have to deal with the probability distribution <strong>in</strong> the canonical<br />

ensemble for a given configuration/micro state C<br />

P (C) = 1<br />

Z(β) exp −βH(C) , (4.2)<br />

where H(C) is the energy of the micro state and β−1 = kBT is the temperature.<br />

The normalization constant is the canonical partition sum<br />

Z(β) = <br />

exp −βH(C) , (4.3)<br />

C<br />

the sum over all micro states weighted by the Boltzmann weight. In the case of<br />

cont<strong>in</strong>uous states the sum becomes an <strong>in</strong>tegral.<br />

Typically one deals with a large number of particles <strong>in</strong> a system, but already<br />

a system consist<strong>in</strong>g of N = 100 particles with two possible states per particle<br />

would <strong>in</strong>volve a number of 2 100 configurations which is not directly summable<br />

<strong>in</strong> a lifespan of an average physicist on a typical computer system (now imag<strong>in</strong>e<br />

a system of N ≈ 10 23 particles. . . ). Even simulat<strong>in</strong>g this system by the<br />

Monte-Carlo approach of choos<strong>in</strong>g n micro states by equal probability (out of a<br />

flat/uniform probability distribution) is not effective s<strong>in</strong>ce most micro states have<br />

very low probability 2 . Therefore we have the problem described above, where the<br />

2 E.g. the energy landscape of a high dimensional problem is ma<strong>in</strong>ly flat, but has some<br />

narrow deep m<strong>in</strong>ima. In this case it is very unlikely that the m<strong>in</strong>ima are hit. But s<strong>in</strong>ce they<br />

are govern<strong>in</strong>g the behavior of the system, they should be sampled effectively.


40 CHAPTER 4. SIMULATION METHODS<br />

variance will be large, s<strong>in</strong>ce we barely hit contribut<strong>in</strong>g configurations.<br />

Introduc<strong>in</strong>g the importance sampl<strong>in</strong>g method aga<strong>in</strong> and choos<strong>in</strong>g random<br />

states directly out of the distribution eq. (4.2) would be smart, but than we<br />

are trapped <strong>in</strong> a vicious circle s<strong>in</strong>ce the partition sum Z, which would be needed<br />

for that, was what we had to calculate <strong>in</strong> first place. But we were not able to do<br />

that. The solution to that problem is found by us<strong>in</strong>g Markov cha<strong>in</strong>s.<br />

4.3.1 Markov cha<strong>in</strong>s<br />

The method of Markov cha<strong>in</strong>s <strong>in</strong> connection with Monte-Carlo simulations was<br />

<strong>in</strong>troduced by N. Metropolis <strong>in</strong> [34] together with an algorithm, <strong>in</strong>troduced <strong>in</strong><br />

the next section 4.3.2. The concept is to generate a sequence of successive configuration<br />

Ci which sample the configuration space <strong>in</strong> a representative manner.<br />

Each step <strong>in</strong> this cha<strong>in</strong> only depends on the present configuration. This already<br />

is the Markov property.<br />

A specific cha<strong>in</strong> of n configurations {C0, . . . , Cn} out of the whole configuration<br />

space occurs with the jo<strong>in</strong>t probability P (C0, . . . , Cn). Writ<strong>in</strong>g this down as<br />

a sequence of events gives a product of conditional probabilities gives<br />

P (C0, . . . , Cn) = P (Cn|Cn−1, . . . , C0) · P (Cn−1|Cn−2, . . . , C0)<br />

· · · P (C2|C1, C0) · P (C1|C0) · P (C0)<br />

n<br />

= P (C0) P (Ci|Ci−1), (4.4)<br />

i=1<br />

where <strong>in</strong> the last l<strong>in</strong>e the Markovian property has already been <strong>in</strong>troduced (cf.<br />

[45]). Markovian property means that the next step only depends on the current<br />

state of the system, the past has no <strong>in</strong>fluence on future steps 3 . For the conditional<br />

probabilities this results <strong>in</strong> a dependence only on the previous configuration<br />

P (Ck|Ck−1, . . . , C0) = P (Ck|Ck−1). (4.5)<br />

The advantage of this property is that we can <strong>in</strong>troduce (for a time homogeneous<br />

Markov cha<strong>in</strong>) a fixed set of transition rates Wij between every two<br />

states Cj → Ci of the configuration space irrespective of all the other states. This<br />

rates form a transition matrix with n × n elements. The probabilities of the<br />

configurations than follow a discrete time master equation, which should have a<br />

equilibrium probability distribution {πi = π(Ci)} (where Ci is some configuration<br />

of the allowed space) for our static problem. By demand<strong>in</strong>g detailed balance<br />

Wijπj = Wjiπi, (4.6)<br />

3 A famous example is the Markovian Scopa player: This player plays his cards only on basis<br />

of the last one played. He does not remember what has been played before. . .<br />

(Scopa is a famous Italian card game)


4.3. SAMPLING THE CANONICAL ENSEMBLE 41<br />

it is ensured that the system evolves towards equilibrium. So there is a direct<br />

balance between every two states. There are other ways to ensure the evolution—<br />

e.g. circular balance between all the states—<strong>in</strong>to an equilibrium distribution, but<br />

detailed balance is the easiest. The most important po<strong>in</strong>t is, that there is noth<strong>in</strong>g<br />

like an absorb<strong>in</strong>g state. In the follow<strong>in</strong>g sections we will only deal with the<br />

detailed balance condition.<br />

Know<strong>in</strong>g this detailed balance condition, it is easy to construct the transition<br />

probabilities for the canonical ensemble, s<strong>in</strong>ce we know the equilibrium distribution,<br />

eq. (4.2),<br />

Wij<br />

Wji<br />

= πi<br />

πj<br />

= exp (−β∆E) , (4.7)<br />

where ∆E = H(Ci) − H(Cj) is the energy difference between the configurations.<br />

The normaliz<strong>in</strong>g factor (the partition sum) Z drops out! We only need ratios<br />

of equilibrium probabilities which we know to be the Boltzmann weights. This<br />

f<strong>in</strong>ally renders the desired evaluations possible.<br />

The question which rema<strong>in</strong>s is how to construct an algorithm which ‘moves’<br />

us through configuration space with this transition probabilities, s<strong>in</strong>ce stor<strong>in</strong>g the<br />

transition matrix W as a whole is not possible for large configurations spaces.<br />

4.3.2 Metropolis algorithm<br />

The Metropolis algorithm is a recipe how to generate a Markov cha<strong>in</strong> for an <strong>in</strong>itial<br />

configuration C0 → C1 → · · · → Ck−1 → Ck → · · · (cf. [34]). If we choose—as<br />

will be described <strong>in</strong> more detail later—a worm-like cha<strong>in</strong> as an example, then the<br />

algorithm would consist of three ma<strong>in</strong> steps:<br />

1. Choose one of the N discrete cha<strong>in</strong> segments at random.<br />

2. Change this segment <strong>in</strong>to a new one by an arbitrary rotation. Be careful<br />

to have the back-rotation possible as well, to ensure detailed balance!<br />

3. Check if this move is accepted or rejected.<br />

These three steps are iterated over for as long as the desired accuracy for the<br />

value to calculate is achieved. In this context one has to bear <strong>in</strong> m<strong>in</strong>d, that<br />

successive steps may be correlated so that one should wait a specific time before<br />

us<strong>in</strong>g a configuration for sampl<strong>in</strong>g. This problems will be addressed later.<br />

The last step is where the above transition probability comes <strong>in</strong>to play. When<br />

start<strong>in</strong>g off a configuration Ck a test configuration Ct is generated, for which we<br />

can calculate the energy difference ∆E(Ct) = H(Ct) − H(Ck). With this the<br />

transition is accepted with the probability—us<strong>in</strong>g eq. (4.7):<br />

p = m<strong>in</strong> 1, exp(−β∆E) . (4.8)


42 CHAPTER 4. SIMULATION METHODS<br />

This means <strong>in</strong> terms of energy: If the energy is lowered the system evolves to the<br />

preferred lower energy state. If the energy is <strong>in</strong>creased, the system moves only<br />

with a Boltzmann factor <strong>in</strong>to that less preferred region. But one has to allow for<br />

this movement “up the energy-hill” s<strong>in</strong>ce the system is fluctuat<strong>in</strong>g around the<br />

equilibrium configuration.<br />

In an computer algorithmic form this consists of either accept<strong>in</strong>g the move<br />

right away s<strong>in</strong>ce ∆E is lowered. Or a random number of a uniform deviate has<br />

to be drawn which must lie below the transition probability exp(−β∆E).<br />

So the new state is:<br />

<br />

Cr accepted with probability: p<br />

Ck+1 =<br />

. (4.9)<br />

Ck rejected with probability: 1 − p<br />

Instead of the usual non-differential acceptance probability, eq. (4.8), other<br />

schemes are possible—like the Glauber dynamics—but we will not go <strong>in</strong>to the<br />

details of this but stick we the equations derived above.<br />

The condition of detailed balance is a sufficient but not necessary condition<br />

for the Monte-Carlo scheme to work properly. It can be shown that e.g. sequential<br />

updat<strong>in</strong>g of the particles—after a particle is moved, <strong>in</strong> the next trials this particle<br />

is skipped, so detailed balance is obviously not satisfied—gives a work<strong>in</strong>g MCscheme<br />

as well (see [13]).<br />

4.3.3 Correlations and error calculation<br />

When us<strong>in</strong>g the Metropolis algorithm one has to assure that the simulation is<br />

explor<strong>in</strong>g the configuration space. There are several problems which have to be<br />

taken <strong>in</strong>to account. Normally two configurations which are only separated by<br />

some small number of MC steps are “quite close” together <strong>in</strong> configuration space<br />

(<strong>in</strong> particular when be<strong>in</strong>g close to a steep energy m<strong>in</strong>imum, a large number of MC<br />

steps are rejected, so that one has to be careful to sample the fluctuation around<br />

the m<strong>in</strong>imum efficiently). When tak<strong>in</strong>g a sample configuration too frequently (<strong>in</strong><br />

the worst case every step), one has an ensemble of samples which are correlated.<br />

This asks for careful treatment of error estimation and for a smart way of choos<strong>in</strong>g<br />

trial steps.<br />

Choos<strong>in</strong>g a trial Monte-Carlo step must be done <strong>in</strong> the follow<strong>in</strong>g way:<br />

• Make as large steps as possible to explore all configuration space without<br />

hav<strong>in</strong>g too many steps rejected, s<strong>in</strong>ce large changes <strong>in</strong> energy ∆E are not<br />

very probable to be accepted.<br />

• On the other hand the steps should be small enough to have an overall<br />

high enough acceptance rate. But choos<strong>in</strong>g the step size to small—which<br />

makes the acceptance rate large—will not drive the simulation far through<br />

configuration space.


4.3. SAMPLING THE CANONICAL ENSEMBLE 43<br />

Therefore for each simulation the acceptance rate must be adjusted before the<br />

sampl<strong>in</strong>g is started. Acceptance rates <strong>in</strong> the range 25–50% are usually giv<strong>in</strong>g<br />

good results.<br />

Some additional po<strong>in</strong>ts have to considered. When start<strong>in</strong>g a simulation from<br />

an arbitrary <strong>in</strong>itial configuration, one has to let the system equilibrate for “some<br />

time” (the number of MC-steps needed for equilibration, has to determ<strong>in</strong>ed <strong>in</strong>dependently)<br />

before the sampl<strong>in</strong>g is started. Otherwise non-equilibrium configurations<br />

are sampled for an equilibrium average.<br />

Furthermore ergodicity has to be ensured: The trial steps have to be chosen<br />

such that the whole configuration space is covered, if the simulation would be<br />

runn<strong>in</strong>g long enough. In this context one has to be particularly sure, that the<br />

system does not get trapped <strong>in</strong> local m<strong>in</strong>ima, show<strong>in</strong>g quasi ergodicity.<br />

Statistical error<br />

Hav<strong>in</strong>g a set of uncorrelated data po<strong>in</strong>ts {x1, . . . , xn}—by experiment or simulation—a<br />

measure for the error of the average over the samples<br />

〈x〉 = 1<br />

n<br />

n<br />

xi, (4.10)<br />

is calculated through the statistical variance, be<strong>in</strong>g the mean-square deviation of<br />

the data po<strong>in</strong>ts from the mean<br />

σ(x) 2 = 1<br />

n − 1<br />

i=1<br />

n <br />

xi − 〈x〉 2 . (4.11)<br />

This actually represents the width of the distribution of data po<strong>in</strong>ts. S<strong>in</strong>ce we<br />

expect the average to converge to the exact value xs <strong>in</strong> the limit n → ∞ the<br />

estimated error of the mean is given by<br />

i=1<br />

∆x = σ(x)<br />

√ n . (4.12)<br />

But eq. (4.11) only holds as long as the samples taken are uncorrelated, s<strong>in</strong>ce<br />

otherwise the error calculated is to small compared to the real one—the samples<br />

look ‘similar’ as long as they are correlated. The aim is to have a measure after<br />

how many MC-steps samples must be taken to have them uncorrelated or to<br />

extract a real error estimate from correlated data. Three possible solutions are<br />

presented <strong>in</strong> the follow<strong>in</strong>g.<br />

Correlation function to make <strong>in</strong>dependent samples Up to now there was<br />

no real time evolution <strong>in</strong> Monte-Carlo s<strong>in</strong>ce steps are generated mechanically <strong>in</strong><br />

fixed <strong>in</strong>tervals which have a priori noth<strong>in</strong>g to do with the real time evolution


44 CHAPTER 4. SIMULATION METHODS<br />

of the problem. As described <strong>in</strong> [4] the dynamic <strong>in</strong>terpretations of Monte Carlo<br />

averag<strong>in</strong>g can be <strong>in</strong>troduced <strong>in</strong> terms of a master equation. Besides be<strong>in</strong>g able<br />

to use this description to calculate correlations this is the start<strong>in</strong>g po<strong>in</strong>t for MC<br />

simulations of dynamic processes.<br />

By this <strong>in</strong>terpretation we write for the probability of hav<strong>in</strong>g a configuration<br />

x at step i as hav<strong>in</strong>g it at time t: p(xi) ≡ p(x, t).<br />

With this dynamic <strong>in</strong>terpretation, the direct way to f<strong>in</strong>d out after how many<br />

steps the samples are uncorrelated is to measure the time dependence of a suitable<br />

correlation function. This correlation function normally decays as an exponential<br />

<strong>in</strong> time<br />

<br />

φ(t) = exp − |t|<br />

<br />

, (4.13)<br />

τ<br />

where the time constant τ gives the time when the correlations are decayed to<br />

1/e of the <strong>in</strong>itial value 4 .<br />

Approximation for the correlation time Besides measur<strong>in</strong>g the, <strong>in</strong> the<br />

above dynamic <strong>in</strong>terpretation, time-dependent correlation function directly, there<br />

is an alternative way to get an estimate for it by sampl<strong>in</strong>g different averages (cf.<br />

[4, 22]). Hav<strong>in</strong>g a large quantity n ≫ 1 of samples xi of an observable x we use<br />

the statistical error<br />

and take the mean of the square<br />

(δx) 2 = 1<br />

n 2<br />

δx = 1<br />

n<br />

n xi − 〈x〉 2 <br />

i=1<br />

= 1 2<br />

x<br />

n<br />

<br />

− 〈x〉<br />

2<br />

1 + 2<br />

n <br />

xi − 〈x〉 <br />

i=1<br />

+ 2<br />

n 2<br />

i=1<br />

n<br />

n<br />

i=1 j=i+1<br />

〈xixj〉 − 〈x〉 2<br />

(4.14)<br />

n<br />

<br />

1 − i<br />

<br />

n<br />

<br />

〈x0xi〉 − 〈x〉 2<br />

〈x2 〉 − 〈x〉 2<br />

<br />

. (4.15)<br />

In the second l<strong>in</strong>e the time <strong>in</strong>variance of the correlation <strong>in</strong> equilibrium 〈xixj〉 =<br />

〈x0xj−i〉 and a change of the summation <strong>in</strong>dex j → j −i has been used. The time<br />

<strong>in</strong>variance only holds if an <strong>in</strong>itial transient regime, where the system ‘forgets’ its<br />

<strong>in</strong>itial configuration, is omitted.<br />

Now <strong>in</strong>troduc<strong>in</strong>g the ‘time’ ti = i δt, where δt is the time unit used, e.g.<br />

δt = 1 for a sample every MC step or δt = N for one MC step per polymer<br />

4For the worm-like cha<strong>in</strong> model described <strong>in</strong> section 4.4 a good correlation function to<br />

measure is the tangent-tangent correlation function<br />

φ(t) = <br />

〈ti(0) · ti(t)〉 − 〈ti〉 2<br />

,<br />

i<br />

which shows reasonable decay<strong>in</strong>g behavior <strong>in</strong> the case of an unconf<strong>in</strong>ed polymer, but fails to be<br />

usable when conf<strong>in</strong><strong>in</strong>g the molecule, due to correlations by the tube.


4.3. SAMPLING THE CANONICAL ENSEMBLE 45<br />

(where the first would not be very effective which will be expla<strong>in</strong>ed later <strong>in</strong><br />

more detail). Transform<strong>in</strong>g the sum <strong>in</strong>to an <strong>in</strong>tegral—which is allowed when the<br />

“correlation time” between successive states is s large compared to δt—than gives<br />

(T = tn = n δt be<strong>in</strong>g the total sampl<strong>in</strong>g time)<br />

2<br />

(δx) = 1 2<br />

x<br />

n<br />

<br />

− 〈x〉<br />

2<br />

1 + 1<br />

δt<br />

T<br />

0<br />

<br />

dt 1 − t<br />

<br />

φx(t) , (4.16)<br />

T<br />

with the normalized autocorrelation function (or “l<strong>in</strong>ear relaxation function”)<br />

φx(t) =<br />

〈x(0)x(t)〉 − 〈x〉2<br />

〈x 2 〉 − 〈x〉 2 . (4.17)<br />

This function is normalized φx(0) = 1 and decays to zero φx(t → ∞) = 0, s<strong>in</strong>ce<br />

the samples of x should become uncorrelated when they are separated by a large<br />

time <strong>in</strong>terval.<br />

Sett<strong>in</strong>g a time-scale<br />

∞<br />

τx = dt φx(t), (4.18)<br />

0<br />

<strong>in</strong> which we assume the correlation function φx(t) to decay essentially to zero,<br />

with τx ≪ T which just means sampl<strong>in</strong>g long enough. This also means that φx(t)<br />

is essentially nonzero only for t ≪ T , so that the term t/T <strong>in</strong> eq. (4.16) can be<br />

neglected and the <strong>in</strong>tegration extended to the upper limit of <strong>in</strong>f<strong>in</strong>ity<br />

2<br />

(δx) = 1 2<br />

x<br />

n<br />

<br />

− 〈x〉<br />

2<br />

1 + 2 τx<br />

<br />

. (4.19)<br />

δt<br />

This shows that the error over correlated data must be enhanced <strong>in</strong> comparison<br />

to the error calculated over uncorrelated data (〈x2 〉−〈x〉 2 )/n (which would mean<br />

sett<strong>in</strong>g τx = 0 here). The factor by which this error is enhanced is also called<br />

statistical <strong>in</strong>efficiency. The time τx can then be seen as the autocorrelation time.<br />

In the case of sampl<strong>in</strong>g <strong>in</strong> steps much larger than the correlation time δt ≫ τx<br />

the result for uncorrelated sampl<strong>in</strong>g is reobta<strong>in</strong>ed, s<strong>in</strong>ce the parenthesis is unity.<br />

On the other hand us<strong>in</strong>g τx ≪ δt results <strong>in</strong> a form <strong>in</strong>dependent of the sampl<strong>in</strong>g<br />

step size<br />

2<br />

(δx) = 2 τx 2<br />

x − 〈x〉 2 , (4.20)<br />

T<br />

because the unity term <strong>in</strong> the parenthesis can be omitted. This means that it<br />

is not useful to <strong>in</strong>crease the number of correlated samples by reduc<strong>in</strong>g the time<br />

between two samples. In this regime there is no change <strong>in</strong> the error. This would<br />

only <strong>in</strong>crease the simulation time by tak<strong>in</strong>g more (useless) samples!<br />

We f<strong>in</strong>ally have a method at hand to calculate the autocorrelation time τx<br />

without directly sampl<strong>in</strong>g and <strong>in</strong>tegrat<strong>in</strong>g it. By measur<strong>in</strong>g the mean 〈x〉 and<br />

mean-square 〈x 2 〉 of x as well as the statistical error 〈(δx) 2 〉 the autocorrelation<br />

time is obta<strong>in</strong>ed by rearrang<strong>in</strong>g eq. (4.19) or (4.20).


46 CHAPTER 4. SIMULATION METHODS<br />

Block<strong>in</strong>g method There are situations, when a set of samples has been obta<strong>in</strong>ed,<br />

but it has not been taken care of measur<strong>in</strong>g the correlation time. When<br />

the whole set of sampl<strong>in</strong>g po<strong>in</strong>ts is available a different approach can be made to<br />

extract an useful error estimate. It is a block<strong>in</strong>g method and therefore related to<br />

real space renormalization (cf. [13, 35] and the orig<strong>in</strong>al paper [12]).<br />

The idea is to calculate the error over a successively smaller set of sample<br />

po<strong>in</strong>ts and observe how this error behaves. The first step is to calculate the error<br />

over the whole set of po<strong>in</strong>ts. Then the set is reduced to half its size by tak<strong>in</strong>g<br />

the mean of two successive po<strong>in</strong>t as new po<strong>in</strong>ts for the smaller sequence. Aga<strong>in</strong><br />

the error is calculated over this new set.<br />

By repeat<strong>in</strong>g this until the set is too small, a sequence of errors is obta<strong>in</strong>ed,<br />

which is first <strong>in</strong>creas<strong>in</strong>g and then should show a plateau (at the end fluctuations<br />

become very large, so that the last po<strong>in</strong>ts are unreliable). It can be shown (see<br />

[12], also for further details and plots) that the real uncorrelated error corresponds<br />

to the value of the plateau. If there is no plateau visible, than too few sample<br />

po<strong>in</strong>ts are available to uncorrelate the error by this method and one can just<br />

get a lower border for the error, which is given by the largest error value <strong>in</strong> the<br />

sequence.<br />

4.3.4 Recursive averag<strong>in</strong>g and error calculation<br />

In the simulations a lot of averages and errors thereof have to be calculated. If<br />

this is done at the end of the program run all the sample data has to be stored.<br />

This can result <strong>in</strong> a large memory usage. The smart way is to calculate the<br />

averages and errors recursively from the last value and the just sampled one.<br />

The mean <strong>in</strong> the Nth step is denoted by (where xi is the sample taken <strong>in</strong><br />

MC-step i)<br />

〈x〉 N = 1<br />

N<br />

N<br />

xi. (4.21)<br />

Insert<strong>in</strong>g this <strong>in</strong> the same formula written down for N +1, results <strong>in</strong> the recursive<br />

equation for the average<br />

i=1<br />

〈x〉 N+1 = N<br />

N + 1 〈x〉 1<br />

N +<br />

N + 1 xN+1. (4.22)<br />

To equate the relation for the mean-square deviation 〈∆x 2 〉 N = 〈x 2 〉 N − 〈x〉 2<br />

N<br />

the result eq. (4.22) has to be used besides the same for mean-square average<br />

x 2 <br />

N+1<br />

N 2<br />

= x<br />

N + 1<br />

1<br />

+ N N + 1 x2N+1. (4.23)<br />

Insert<strong>in</strong>g and after some rearrang<strong>in</strong>g of terms the results is<br />

(N + 1) ∆x 2<br />

N+1 = N ∆x 2 N 2. + 〈x〉N − xN+1 (4.24)<br />

N N + 1


4.4. SIMULATION OF UNCONFINED WORM-LIKE CHAINS 47<br />

So the average and the error are easily calculated with m<strong>in</strong>or storage overhead.<br />

The only drawback is, when a block<strong>in</strong>g technique has to be used to free the<br />

samples of the ‘omnipresent’ correlation <strong>in</strong> MC simulations, then all the averages<br />

have to be stored anyway.<br />

4.3.5 Random number generators<br />

The quality of Monte-Carlo simulations depends crucially on the quality of the<br />

pseudo-random numbers used. The quality of the l<strong>in</strong>ear congruential generators,<br />

as for example implemented <strong>in</strong> most standard libraries, have poor statistical<br />

quality and a very small period. Other generators have to be used. There is a<br />

large list of different types but most of them fail to pass statistical tests rather<br />

spectacularly. There is no exact measure for the quality of a generator but there<br />

exists a list of structural mathematical and statistical tests which give a ‘feel<strong>in</strong>g’<br />

for the quality.<br />

A good overview over the important details of random number generators and<br />

examples is given <strong>in</strong> the review [28].<br />

A fast generator with sufficient quality is based on the l<strong>in</strong>ear feedback shift<br />

register or Tausworthe generator (see [27] and [29]). The implementation from<br />

the GNU Scientific Library 5 has been used <strong>in</strong> our simulation. The advantage of<br />

this generator, besides its structural mathematical qualities, is its speed. The<br />

random number generation consumes up to one forth of the computational time<br />

of our simulations. Test with different generator have been performed, but the<br />

Tausworthe generator seems to be good enough.<br />

If for some reason an even better generator is needed, the RANLUX generator<br />

described <strong>in</strong> [30] should be used, which can provide, by choice of a numerical<br />

constant, uncorrelated numbers down to the last digit of the used float<strong>in</strong>g po<strong>in</strong>t<br />

variable. This has been theoretically justified by chaos theory. The quality of<br />

this generator goes on costs of computational speed which is up to 20 times lower<br />

than <strong>in</strong> the aforementioned generator.<br />

4.4 Simulation of unconf<strong>in</strong>ed worm-like cha<strong>in</strong>s<br />

Simulation of semi-flexible polymers are based on the discrete worm-like cha<strong>in</strong><br />

model as <strong>in</strong>troduced <strong>in</strong> section 2.3.2. The basic Hamiltonian is given by eq. (2.13),<br />

which can be directly used to perform simulations<br />

N−1 <br />

e = βH ({ti}) = −k ˆti · ˆti+1 with k = βεl 2 , (4.25)<br />

i=1<br />

5 http://www.gnu.org/software/gsl/manual/gsl-ref 17.html#SEC262 (17.11.2004)


48 CHAPTER 4. SIMULATION METHODS<br />

where the prefactor k is related to the persistence length lp by eqs. (A.8) and<br />

(A.10). As <strong>in</strong>put parameter for the simulation the flexibility ɛ is taken so that<br />

everyth<strong>in</strong>g is measured <strong>in</strong> units of total length L (cf. eq. (3.2)). All output is<br />

normalized by L as well.<br />

This Hamiltonian is used to perform an off-lattice simulation where the polymer<br />

can orient without any restriction (cf. [25]).<br />

In a straightforward algorithm the polymer is represented by just its N segments,<br />

where the positions of the jo<strong>in</strong>ts can be ignored as long as the polymer is<br />

unconf<strong>in</strong>ed. The basic steps are:<br />

1. perform a random movement on a randomly chosen segment<br />

2. calculate the change <strong>in</strong> configuration energy ∆e<br />

3. accept/reject the move correspond<strong>in</strong>g to the Metropolis algorithm<br />

(4. take samples after a specific number of steps)<br />

This three (four) basic steps are the essential part of the MC scheme. They will<br />

be discussed now <strong>in</strong> more detail.<br />

Step 1: The trial movement <strong>in</strong> the n-th MC-step is chosen by draw<strong>in</strong>g a<br />

random number m to specify a segment to change ˆt (n)<br />

m → ˜t (n)<br />

m . The orientation<br />

is changed by a multiplication with a rotational matrix R l k<br />

, which is chosen out<br />

of a set that has been calculated <strong>in</strong> the <strong>in</strong>itial phase of the simulation. The<br />

superscript l numbers one of the d spacial dimensions and k the specific matrix<br />

chosen from the set. It turned out to be optimal hav<strong>in</strong>g 20 d equidistant discrete<br />

matrices (20 for each dimension) out of a maximal rotation angle ±δα/2. This<br />

angels must be tuned to give an overall acceptance rate of 25–50%.<br />

Detailed balance is obviously guaranteed: choos<strong>in</strong>g a rotation matrix by first<br />

choos<strong>in</strong>g one direction at random and than aga<strong>in</strong> randomly a specific rotation 6 .<br />

Thus this step is subsumed by—with random numbers m = 1, . . . , N, l = 1, . . . , d<br />

and k = 1, . . . , 20—<br />

˜t (n)<br />

i<br />

=<br />

<br />

Rl k · ˆt (n)<br />

m for i = m<br />

ˆt (n)<br />

. (4.26)<br />

i for i = m<br />

One segment m is changed, where the orientation of all others is kept fixed as<br />

shown <strong>in</strong> figure 4.1.<br />

An alternative approach to the simulation <strong>in</strong> two dimension is by not<strong>in</strong>g that<br />

the polymer configuration is fully specified by know<strong>in</strong>g the angles θi between each<br />

segment and a fixed global direction (see figure 4.1). The simulation can than<br />

be build on trial moves chang<strong>in</strong>g these angles, i.e. tak<strong>in</strong>g a random change on<br />

6 E.g. it would be wrong to use <strong>in</strong> each step a rotation <strong>in</strong> all d-direction, when apply<strong>in</strong>g the<br />

rotation always <strong>in</strong> the same order. This is due to the non-commutativity of rotations, which<br />

breaks detailed balance <strong>in</strong> this case. But choos<strong>in</strong>g the order of application randomly, aga<strong>in</strong><br />

obeys detailed balance, but has an unnecessary calculation overhead.


4.4. SIMULATION OF UNCONFINED WORM-LIKE CHAINS 49<br />

ˆt1<br />

θ1<br />

ˆt2<br />

θ2<br />

ˆtm<br />

˜tm<br />

Figure 4.1: Trial move <strong>in</strong> MC-algorithm for the WLC. The randomly chosen segment m is<br />

changed, whereas all other tangents are kept fixed.<br />

posit<strong>in</strong>g m with δθ where the angle θm → ˜ θm = θm + δθ must be changed. We<br />

tested this algorithm but it did not show any major advantage neither <strong>in</strong> accuracy<br />

nor speed (one has to calculate trigonometrical functions <strong>in</strong> every step). S<strong>in</strong>ce it<br />

is not easily generalized <strong>in</strong>to three dimensions the preference has been given to<br />

the first algorithm, s<strong>in</strong>ce it can supply a general code for arbitrary dimensions.<br />

Step 2: For the chosen trial move the change <strong>in</strong> configuration energy ∆e must<br />

be calculated. The advantage of the algorithm is, that this energy change is given<br />

by a simple term, s<strong>in</strong>ce the orientation on only two jo<strong>in</strong>ts is changed (see aga<strong>in</strong><br />

figure 4.1)<br />

∆e = −k<br />

N<br />

i=1<br />

˜t (n)<br />

i<br />

<br />

= −k ˜t (n)<br />

m − ˆt (n)<br />

<br />

m ·<br />

· ˜t (n)<br />

i+1 − ˆt (n)<br />

i<br />

θm<br />

˜θm<br />

ˆtm+1<br />

· ˆt (n)<br />

<br />

i+1<br />

<br />

ˆt (n)<br />

m+1 + ˆt (n)<br />

<br />

m−1<br />

θm+1<br />

ˆtm+1<br />

θm+1<br />

ˆtN<br />

θN<br />

ˆtN<br />

θN<br />

for: i = 1, . . . , N, (4.27)<br />

which can be easily implemented by us<strong>in</strong>g two dummy segments with length zero<br />

ˆt0 = ˆtN+1 = 0.<br />

Step 3: Accord<strong>in</strong>g to the Metropolis algorithm <strong>in</strong>troduced <strong>in</strong> section 4.3.2<br />

this move is f<strong>in</strong>ally accepted, if the total configuration energy is reduced ∆e < 0.<br />

In the case of an energy <strong>in</strong>crease, it is accepted correspond<strong>in</strong>g to the Boltzmann<br />

weight x < exp(−∆e) by draw<strong>in</strong>g another random number 0 ≤ x < 1. Otherwise<br />

the move is rejected<br />

ˆt (n+1)<br />

i<br />

=<br />

˜t (n)<br />

i<br />

accepted<br />

. (4.28)<br />

ˆt (n)<br />

i rejected<br />

Step 4 is set <strong>in</strong> brackets, s<strong>in</strong>ce it is not useful to take a samples from the<br />

configuration dur<strong>in</strong>g every MC-step. Two successive configurations are normally<br />

strongly correlated. So we estimate the correlation time τ of one observable (see<br />

page 44), which is taken as the sample step size, after which a configuration is<br />

taken. This correlation time is also used to determ<strong>in</strong>e the error estimate on the<br />

averages.


50 CHAPTER 4. SIMULATION METHODS<br />

4.4.1 Initial part of the simulation<br />

Before actually start<strong>in</strong>g the ma<strong>in</strong> Monte-Carlo sampl<strong>in</strong>g loop first a set of th<strong>in</strong>gs<br />

have to be performed <strong>in</strong> the <strong>in</strong>itial phase of the program:<br />

• As mentioned before the acceptance rate has to be adjusted such that one is<br />

sampl<strong>in</strong>g the whole configuration space (mak<strong>in</strong>g the steps sufficiently large)<br />

but still hav<strong>in</strong>g the steps accepted frequently enough (mak<strong>in</strong>g sufficiently<br />

small steps for ∆e not becom<strong>in</strong>g too large). This has to be tuned for<br />

each simulation s<strong>in</strong>ce the rate depends on the parameters of the system.<br />

The variable to be tuned here is the maximal rotation angle δα. In our<br />

simulations this is done <strong>in</strong> the first 1000N 2 steps, where this angle is de- or<br />

<strong>in</strong>creased to have the acceptance rate as desired. A value of 25-50% seems<br />

to be an reasonable range. Fix<strong>in</strong>g the acceptance rate is the first important<br />

step <strong>in</strong> a simulation.<br />

• Indicated <strong>in</strong> step 4 sampl<strong>in</strong>g should be done <strong>in</strong> effective steps—not tak<strong>in</strong>g<br />

too many samples which do not <strong>in</strong>crease the accuracy of the f<strong>in</strong>al averages.<br />

We therefore have to know an estimate for the correlation time τR of the<br />

observables. Then samples can be taken <strong>in</strong> steps of this correlation time<br />

which assures, by eq. (4.19), to improve the averages. This value of τR is<br />

f<strong>in</strong>ally used to extract the error estimate by the same formula.<br />

Only one correlation time τR will be used as step size. We calculate it,<br />

as <strong>in</strong>dicated, for the end-to-end distance s<strong>in</strong>ce our ma<strong>in</strong> <strong>in</strong>terest is <strong>in</strong> this<br />

value or other ‘types’ of end-to-end distances (see section 4.6), which are<br />

assumed to show correlations of the same order. As shown <strong>in</strong> section 4.3.3<br />

(page 45) we need the values of 〈R〉, 〈R 2 〉 and 〈(δR) 2 〉 to obta<strong>in</strong> τR. This<br />

is the second task to be fulfilled <strong>in</strong> our simulation.<br />

S<strong>in</strong>ce we do not know the correlation time beforehand—which actually must<br />

be determ<strong>in</strong>ed self-consistently—the mentioned averages are sampled <strong>in</strong><br />

1000 <strong>in</strong>dependent (uncorrelated) runs by restart<strong>in</strong>g with different random<br />

<strong>in</strong>itial configurations. In each run 10N 2 MC-steps are performed, where<br />

samples are taken <strong>in</strong> steps of N (s<strong>in</strong>ce the correlation time somehow depends<br />

on N, the number of segments 7 ). We are <strong>in</strong>terested <strong>in</strong> the static equilibrium<br />

properties of the system, so attention must be paid to the <strong>in</strong>itial transient<br />

phase, where the polymer relaxes from the arbitrary <strong>in</strong>itial non-equilibrium<br />

configuration to the equilibrium behavior—which is measured by a different<br />

(non-equilibrium) correlation time. We observed this relaxation to be very<br />

fast. To be on the sure side 10N 2 MC-steps are performed after the <strong>in</strong>itial<br />

condition has been chosen, before the sampl<strong>in</strong>g is started.<br />

7 Actually the number of possible configurations scales exponentially ∼ A N , assum<strong>in</strong>g each<br />

segment to have A possible configurations.


4.5. SIMULATION OF CONFINED WORM-LIKE CHAINS 51<br />

• Now all prerequisites are obta<strong>in</strong>ed to start the ma<strong>in</strong> simulation. For that<br />

purpose a new random <strong>in</strong>itial configuration is chosen and 2000 τR MC-steps<br />

are waited for the cha<strong>in</strong> to equilibrate, after that the ma<strong>in</strong> sampl<strong>in</strong>g loop<br />

as described above can be started.<br />

A f<strong>in</strong>al remark concern<strong>in</strong>g the numerics: Due to a possible accumulation of<br />

numerical <strong>in</strong>accuracies dur<strong>in</strong>g the float<strong>in</strong>g po<strong>in</strong>t operations on the tangents, it<br />

must be remembered to renormalize the tangents from time to time.<br />

4.4.2 Self-avoid<strong>in</strong>g cha<strong>in</strong>s<br />

A step to more realistic simulation has also been performed by <strong>in</strong>troduc<strong>in</strong>g excluded<br />

volume effect. This has been done <strong>in</strong> two ways. The first straightforward<br />

way is to put a sphere, of diameter correspond<strong>in</strong>g to the polymer diameter, around<br />

the bead positions, which are <strong>in</strong>teract<strong>in</strong>g by e.g. a hard-core potential. The problem<br />

here is, that by us<strong>in</strong>g very th<strong>in</strong> polymers as a model, the discretization N<br />

must be chosen very large for not hav<strong>in</strong>g “empty spaces” <strong>in</strong> between the beads<br />

(the model without spac<strong>in</strong>gs between the beads is called shish-kebab model; cf.<br />

[10]), which would not take self-avoidance rigorously <strong>in</strong>to account. Besides the<br />

normal <strong>in</strong>crease of the number of configurations by <strong>in</strong>creas<strong>in</strong>g N the computational<br />

time <strong>in</strong>creases additionally ∼ N 2 (for a naive algorithm. For smarter<br />

algorithms there is at least a N log N dependence.) due to the number of tests<br />

for overlaps.<br />

A way to keep the number of required segments smaller is to use tube-like<br />

segments of the the size of the polymer diameter as basic segments. Then selfavoidance<br />

is always rigorous. To check if two segments <strong>in</strong>tersect <strong>in</strong> a MC-step, an<br />

algorithm is needed which calculates the m<strong>in</strong>imal distance between to l<strong>in</strong>e segments.<br />

These k<strong>in</strong>d of problems are common <strong>in</strong> 3D graphics programm<strong>in</strong>g (i.e. for<br />

computer games). A fast algorithm can be found <strong>in</strong> [42] which we implemented.<br />

In this model the polymer is a space curve with thickness, therefore we called it<br />

spaghetti model.<br />

It turned out that self-avoidance effects are not observable for the cha<strong>in</strong>s<br />

simulated. This is due to the relatively small number of segments N used. As<br />

described <strong>in</strong> section 2.5 on page 18, there is a critical number Nc below which, even<br />

for flexible polymers, self-avoidance is irrelevant. For stiff/semi-flexible polymers<br />

self-avoidance should not have an <strong>in</strong>fluence at all, s<strong>in</strong>ce backbends are not likely.<br />

In the future, we are able to perform simulation of longer cha<strong>in</strong>s. With these<br />

consideration, we therefore have the basic algorithm already at hand.<br />

4.5 Simulation of conf<strong>in</strong>ed worm-like cha<strong>in</strong>s<br />

As it is the ma<strong>in</strong> subject of this thesis we now turn to the simulation of worm-like<br />

cha<strong>in</strong>s <strong>in</strong> conf<strong>in</strong><strong>in</strong>g tube-like environments. The ‘environment’ discussed <strong>in</strong> most


52 CHAPTER 4. SIMULATION METHODS<br />

detail will be the cyl<strong>in</strong>drical symmetric harmonic potential as already <strong>in</strong>troduced<br />

<strong>in</strong> eq. (3.20). The accuracy of this approximation and how to relate it to the<br />

other potentials will be discussed <strong>in</strong> section 5.<br />

4.5.1 Harmonic potential<br />

The configuration energy function ∆e correspond<strong>in</strong>g to eq. (4.25) is given by the<br />

discretized version of eq. (3.20)<br />

N−1 <br />

∆e = βH ({ti}) = −k ˆti · ˆti+1 + g<br />

2<br />

i=1<br />

N 2<br />

xi + y 2 i , (4.29)<br />

such that we now need the tangents and the positions of the beads <strong>in</strong> the simulation.<br />

The latter can be directly calculated from the tangents, but stor<strong>in</strong>g them<br />

is faster.<br />

We have to relate the prefactor g to the Odijk deflection length ld by discretiz<strong>in</strong>g<br />

the harmonic potential part of eq. (3.20). Additionally one has to use<br />

dimensionless units <strong>in</strong> the simulation. Hence all positions are measured <strong>in</strong> units<br />

of the segment length l<br />

γ<br />

2<br />

L<br />

0<br />

ds x(s) 2 + y(s) 2 = γl3<br />

2<br />

→ γl3<br />

2<br />

<br />

g/2β<br />

Nl<br />

0<br />

ds<br />

l<br />

i=0<br />

x(s)<br />

l<br />

2<br />

<br />

2<br />

y(s)<br />

+<br />

l<br />

N 2<br />

xi + y 2 i , (4.30)<br />

where xi, yi are normalized by the segment length l = L/N. Upon us<strong>in</strong>g the<br />

one f<strong>in</strong>ds<br />

def<strong>in</strong>ition ld := (4κ/γ) 1<br />

4 ⇒ γ = 4κ/l 4 d<br />

g = βγl 3 = 4βκl3<br />

l 4 d<br />

i=0<br />

(2.20)<br />

= 4lpl 3<br />

l 4 d<br />

= 4c4<br />

N 3 . (4.31)<br />

ɛ<br />

The actual parameter used for the simulations is the collision parameter c as<br />

def<strong>in</strong>ed <strong>in</strong> eq. (3.3).<br />

With these prerequisites the simulation can be implemented straightforwardly.<br />

Only one additional MC move must be added s<strong>in</strong>ce the space is not isotropic<br />

anymore. Now global movements of the polymer perpendicular to the tube axis<br />

must be attempted. This can be implemented by us<strong>in</strong>g a virtual segment ‘zero’.<br />

If this is randomly selected, all positions are shifted by a random amount out of<br />

±∆r⊥. The detailed balance condition is therefore obeyed. This maximum shift<br />

is, along with the maximal rotation angle δα, adjusted <strong>in</strong> the <strong>in</strong>itial simulation<br />

phase. The z-position of the first bead can be kept fixed, s<strong>in</strong>ce there is still the<br />

rotational symmetry around the tube/z-axis.


4.6. SAMPLED QUANTITIES AND THEIR RELEVANCE 53<br />

4.5.2 Hard-wall tubes<br />

Hard-wall tubes with circular and quadratic cross section have been simulated.<br />

With these potentials no additional potential energy has to be calculated, but <strong>in</strong><br />

every MC-step it has to be checked if any of the beads crosses the border of the<br />

allowed space. This can also be stated <strong>in</strong> terms of a <strong>in</strong>f<strong>in</strong>ite energy barriers<br />

<br />

∞ for |r⊥| ><br />

Vtube(r⊥) =<br />

d<br />

2 , (4.32a)<br />

Vchannel(r⊥) =<br />

0 else<br />

<br />

∞ for |x| > w<br />

2<br />

0 else<br />

or |y| > w<br />

2<br />

, (4.32b)<br />

where d is the tube diameter and w the width of the square channel.<br />

Close to the walls approximately only half of the moves are accepted. The<br />

adjustment of the acceptance rate has to be done with care. Therefore the <strong>in</strong>itial<br />

configuration is put fully stretched parallel to the tube-axis, very close to the<br />

walls (<strong>in</strong> the channel close to a corner, where even less moves are accepted). The<br />

acceptance rate is then adjusted by lett<strong>in</strong>g the configuration evolve freely and<br />

putt<strong>in</strong>g it back to the wall from time to time, s<strong>in</strong>ce it is mov<strong>in</strong>g to the center<br />

of the tube. In this way we can be sure, that we are still accept<strong>in</strong>g movements<br />

close to the boundaries and are not stick<strong>in</strong>g to the wall. For the ma<strong>in</strong> sampl<strong>in</strong>g<br />

part of the simulation, the polymer is then set <strong>in</strong>itially totally stretched along<br />

the tube-axis. The <strong>in</strong>itial configuration is decay<strong>in</strong>g very fast, so that this specific<br />

chosen configuration does not matter.<br />

4.6 Sampled quantities and their relevance<br />

Dur<strong>in</strong>g the simulation run a whole bunch of observables are sampled. Some of<br />

them make only sense when conf<strong>in</strong><strong>in</strong>g the cha<strong>in</strong>, as described <strong>in</strong> the previous<br />

section. This is <strong>in</strong>dicated at the correspond<strong>in</strong>g location. It follows a list of the<br />

observables, how they are calculated and their relevance <strong>in</strong> experiments:<br />

mean (square) end-to-end distance 〈R〉, 〈R 2 〉:<br />

The end-to-end distance is given by the sum over all tangents<br />

R =<br />

N<br />

ˆti. (4.33)<br />

From this formula both averages can be obta<strong>in</strong>ed straightforwardly.<br />

i=1<br />

This is the quantity of <strong>in</strong>terest <strong>in</strong> our comparison to analytical results obta<strong>in</strong>ed<br />

<strong>in</strong> chapter 3. In particular, we are go<strong>in</strong>g to <strong>in</strong>vestigate the scal<strong>in</strong>g<br />

properties <strong>in</strong> section 5.2.3.


54 CHAPTER 4. SIMULATION METHODS<br />

For stiff and/or conf<strong>in</strong>ed polymers this is also a quantity, which can be<br />

experimentally measured on fluorescently dyed polymers (e.q. [44, 38]).<br />

mean (square) parallel end-to-end distance <br />

2<br />

R|| , R|| :<br />

The parallel end-to-end distance is the component of the end-to-end distance<br />

along the conf<strong>in</strong><strong>in</strong>g ‘tube’ (here z-axis). So this only makes sense <strong>in</strong><br />

connection with the simulation <strong>in</strong> conf<strong>in</strong>ed geometries. In the unconf<strong>in</strong>ed<br />

case the simple relation R2 <br />

2<br />

|| = 〈R 〉 /3 must hold.<br />

The parallel end-to-end distance is given by<br />

R|| =<br />

N<br />

i=1<br />

<br />

ˆti . (4.34)<br />

z<br />

This quantity should co<strong>in</strong>cide with the end-to-end distance <strong>in</strong> the very<br />

strong conf<strong>in</strong>ed limit, due to the orientation along the axis. It is more of<br />

theoretical <strong>in</strong>terest, s<strong>in</strong>ce it can be used to expla<strong>in</strong> some <strong>in</strong>terest<strong>in</strong>g global<br />

features (cf. section 5.2.4).<br />

apparent end-to-end distances 〈Ra〉, 〈R2 a〉, <br />

2<br />

Ra|| , Ra|| :<br />

Another end-to-end distance is important, when experiments of flexible<br />

polymers <strong>in</strong> conf<strong>in</strong>ed environments are <strong>in</strong>vestigated (e.g. [44, 38]). S<strong>in</strong>ce on<br />

fluorescently dyed polymers, the real ends are difficult to discern from the<br />

rest of the cha<strong>in</strong>. Only a blurry picture is seen as depicted <strong>in</strong> figure 1.2.<br />

Therefore the apparent end-to-end distance is <strong>in</strong>troduced, which is the maximal<br />

distance between two beads along the cha<strong>in</strong>. The distance between<br />

two beads i, j at positions Ri, Rj is def<strong>in</strong>ed as<br />

dij = Rj − Ri<br />

(4.35)<br />

so that <strong>in</strong> the same spirit as before, apparent and the apparent parallel<br />

end-to-end distance are def<strong>in</strong>ed as<br />

mean-square (parallel) radius of gyration R 2 g<br />

Ra = max |dij| , (4.36a)<br />

i,j<br />

Ra|| = max<br />

i,j (dij)z. (4.36b)<br />

A scal<strong>in</strong>g analysis and comparison to experimental data is provided <strong>in</strong> section<br />

5.2.3.<br />

<br />

2 , Rg|| :<br />

The radius of gyration can be calculated directly as given <strong>in</strong> eq. (2.7).<br />

Where the form with the center of mass Rcm is preferred s<strong>in</strong>ce it only<br />

<strong>in</strong>volves order 2N calculations not order N 2 .


4.6. SAMPLED QUANTITIES AND THEIR RELEVANCE 55<br />

This value has importance for experiments on conformations of unconf<strong>in</strong>ed<br />

polymers. When e.g. a fluorescently dyed polymer shows spots of strong<br />

<strong>in</strong>tensity (which are not resolved by the camera tak<strong>in</strong>g the picture), one<br />

can weighten the configuration with the <strong>in</strong>tensity, lead<strong>in</strong>g to the radius of<br />

gyration obta<strong>in</strong>ed through the fluorescent “mass” (cf. [32]).<br />

mean-square undulations 〈x 2 (s)〉, 〈x 2 〉, 〈y 2 〉:<br />

This values for the undulation makes sense only <strong>in</strong> simulations of conf<strong>in</strong>ed<br />

polymers, s<strong>in</strong>ce it relies on hav<strong>in</strong>g a f<strong>in</strong>ite system <strong>in</strong> the lateral direction to<br />

give f<strong>in</strong>ite averages. The sampl<strong>in</strong>g is straightforward by the perpendicular<br />

distance of the beads from the axis.<br />

The function 〈x 2 (s)〉 gives the mean-square undulation for each bead whereas<br />

the values 〈x 2 〉 and 〈y 2 〉 are additionally averaged over all beads.<br />

These results can also be compared with our results from chapter 3. This<br />

will be done <strong>in</strong> section 5.2.1.<br />

tangent-tangent correlation function ˆt(0.5) · ˆt(s + 0.5) :<br />

The tangent-tangent correlation function is sampled us<strong>in</strong>g a fixed reference<br />

po<strong>in</strong>t <strong>in</strong> the middle of the polymer, to reduce boundary effects as much as<br />

possible. We want to be able to compare the results with analytical results<br />

obta<strong>in</strong>ed for different boundary conditions (see section 5.2.2). The function<br />

is obta<strong>in</strong>ed directly by us<strong>in</strong>g the simulated tangents.<br />

radial distribution function p(R) = p(|R|):<br />

Instead of only sampl<strong>in</strong>g an average, sampl<strong>in</strong>g a probability distribution<br />

needs some additional steps. The range of possible values—here 0 ≤ R ≤<br />

L = 1—has to be divided <strong>in</strong>to b<strong>in</strong>s. In each sampl<strong>in</strong>g step the value of R<br />

hits one of these b<strong>in</strong>s, where the correspond<strong>in</strong>g counter is <strong>in</strong>creased. S<strong>in</strong>ce<br />

we do not obta<strong>in</strong> a whole distribution <strong>in</strong> one sampl<strong>in</strong>g step, we are forced<br />

to f<strong>in</strong>d a different way for the error calculation of the distribution po<strong>in</strong>ts.<br />

Therefore, dur<strong>in</strong>g each sampl<strong>in</strong>g step of τR MC-steps, 500 samples are taken<br />

for the averages to get a rough distribution. This is then used to obta<strong>in</strong> an<br />

averaged distribution with an error estimate.<br />

A value of 100 b<strong>in</strong>s gives a smooth look<strong>in</strong>g probability distribution. Only<br />

for strongly conf<strong>in</strong>ed or very stiff polymers the number of b<strong>in</strong> should be<br />

<strong>in</strong>creased around the fully stretched regime.<br />

The same sort of distribution function is obta<strong>in</strong>ed for R||. Both results are<br />

shown <strong>in</strong> section 5.2.5.<br />

In pr<strong>in</strong>ciple these distribution can be compared to experimental results.<br />

This work is currently <strong>in</strong> progress.<br />

All sampl<strong>in</strong>g is done us<strong>in</strong>g the recursive method <strong>in</strong>troduced <strong>in</strong> section 4.3.4 <strong>in</strong><br />

order to avoid large arrays of data or <strong>in</strong>-between stor<strong>in</strong>g and averag<strong>in</strong>g to reduce


56 CHAPTER 4. SIMULATION METHODS<br />

the memory usage. Therefore the memory usage is only <strong>in</strong> the order of some<br />

MBytes.


Chapter 5<br />

Simulation results<br />

In this chapter we are go<strong>in</strong>g to explore the scal<strong>in</strong>g behavior of the worm-like cha<strong>in</strong><br />

upon conf<strong>in</strong>ement, us<strong>in</strong>g the Monte-Carlo algorithm discussed <strong>in</strong> the preced<strong>in</strong>g<br />

chapter 4. After validat<strong>in</strong>g the simulation with exactly known results <strong>in</strong> bulk<br />

solution (cf. chapter 2), we turn to the <strong>in</strong>fluence of a harmonic potential with<br />

cyl<strong>in</strong>drical symmetry on the polymer configuration. The undulations and the<br />

tangent-tangent correlation function are compared to analytical results (cf. chapter<br />

3). After becom<strong>in</strong>g familiar with the <strong>in</strong>fluence of the conf<strong>in</strong><strong>in</strong>g environment,<br />

we explore the scal<strong>in</strong>g behavior of the mean-square end-to-end distance 〈R 2 〉 with<br />

the ratio of the collision parameter to the persistence length c/ɛ.<br />

We also go beyond an analysis of the moments by study<strong>in</strong>g the full probability<br />

distribution function of the end-to-end distance. It turns out that there is a<br />

cross over from the well known unconf<strong>in</strong>ed distribution to the conf<strong>in</strong>ed regime,<br />

strongly peaked towards full stretch<strong>in</strong>g, with an <strong>in</strong>termediate bimodal distribution<br />

function, as one decreases the channel width.<br />

F<strong>in</strong>ally we have also started to explore other types of conf<strong>in</strong><strong>in</strong>g potentials, <strong>in</strong><br />

particular hard-wall potentials with square and cyl<strong>in</strong>drical cross section.<br />

Before start<strong>in</strong>g the <strong>in</strong>vestigation, some prelim<strong>in</strong>ary remarks are <strong>in</strong> order:<br />

• In this chapter all values are measured <strong>in</strong> units of the total length L.<br />

• In the plots the statistical error is smaller than the symbols if not stated<br />

otherwise.<br />

• All simulations were performed without us<strong>in</strong>g self-avoidance effects. We<br />

have conv<strong>in</strong>ced ourselves that such effects are negligible for flexible filaments<br />

up to a length of at least N = 500 segments.<br />

• The simulations were performed on a standard computer desktop system<br />

(Intel Pentium IV processor, various frequencies). A typical simulation for<br />

fixed segment size N, flexibility ɛ and collision parameter c takes about one<br />

hour for N = 50 and up to the order of a week for N = 1000 segments.<br />

57


58 CHAPTER 5. SIMULATION RESULTS<br />

R 2 <br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0<br />

1<br />

2<br />

3<br />

4<br />

5<br />

ɛ<br />

6<br />

7<br />

exact<br />

simulation<br />

Figure 5.1: Comparison of the simulation for a polymer <strong>in</strong> bulk solution c = 0 to the exact<br />

formula eq. (2.21). The dependence of R 2 on the flexibility ɛ is shown.<br />

5.1 The unconf<strong>in</strong>ed cha<strong>in</strong><br />

Before one can start us<strong>in</strong>g a simulation algorithm to produce data for a situation<br />

where no analytical comparison is available, it should be validated by compar<strong>in</strong>g<br />

the results with well known limit<strong>in</strong>g cases where analytical results are available.<br />

This is the case for a polymer <strong>in</strong> bulk solution, where the dependence of the<br />

second moment of the end-to-end vector distribution on the persistence length is<br />

well known (cf. eq. (2.21)). This limit can be simulated us<strong>in</strong>g our algorithm for<br />

a collision parameter c = 0. The results <strong>in</strong> figure 5.1 show that there is perfect<br />

agreement with<strong>in</strong> the statistical error.<br />

An additional check can be done with the tangent-tangent correlation function<br />

which should show simple exponential behavior. In the logarithmic plot 5.2 this<br />

results <strong>in</strong> a straight l<strong>in</strong>e. The simulation data for e.g. lp = 10 aga<strong>in</strong> fits eq. (2.12)<br />

perfectly with<strong>in</strong> the errorbars.<br />

In addition, the radial distribution function has also been checked aga<strong>in</strong>st<br />

known results, such that the algorithm—at least concern<strong>in</strong>g the bend<strong>in</strong>g energy—<br />

is fully validated.<br />

5.2 The harmonic potential<br />

A convenient start<strong>in</strong>g po<strong>in</strong>t for the discussion of conf<strong>in</strong>ement on the conformation<br />

of a polymer is a harmonic potential with cyl<strong>in</strong>drical symmetry. This has the<br />

advantage that we can compare with analytical results at least <strong>in</strong> the stiff limit<br />

(see section 3.3).<br />

8<br />

9<br />

10


5.2. THE HARMONIC POTENTIAL 59<br />

ˆt(0.5) · ˆt(s + 0.5) <br />

1<br />

0.1<br />

0.01<br />

0.001<br />

-0.5<br />

-0.4<br />

-0.3<br />

-0.2<br />

exact<br />

simulation<br />

-0.1 0<br />

s − 0.5<br />

0.1<br />

Figure 5.2: Comparison of the tangent-tangent correlation function from eq. (2.12) to simulation<br />

data for c = 0. The reference po<strong>in</strong>t of the correlation function is fixed <strong>in</strong> the middle of<br />

the polymer, where the zero of the plot is set.<br />

5.2.1 Undulations<br />

We start our discussion with the effect of the conf<strong>in</strong><strong>in</strong>g environment on the meansquare<br />

transverse displacements.<br />

Upon conf<strong>in</strong><strong>in</strong>g we observed that undulations are more and more suppressed,<br />

as one would also expect <strong>in</strong>tuitively. In particular we have derived a result for the<br />

mean-square undulations, eq. (3.27), that was already shown <strong>in</strong> figure 3.6 along<br />

with the result obta<strong>in</strong>ed by a Fourier series ansatz.<br />

The relative deviation of the simulation data for the mean distance between<br />

the cha<strong>in</strong> jo<strong>in</strong>ts and the tube axis 〈x 2 s(s)〉 from the analytical result 〈x 2 a〉, obta<strong>in</strong>ed<br />

<strong>in</strong> eq. (3.27), is shown <strong>in</strong> figure 5.3. For stiff conf<strong>in</strong>ed polymers the value obta<strong>in</strong>ed<br />

<strong>in</strong> the simulations is identical to the undulations as parametrized <strong>in</strong> the weaklybend<strong>in</strong>g<br />

rod limit. The po<strong>in</strong>ts are taken from simulations with N = 50 segments 1<br />

for a value of ɛ = 0.1 and conf<strong>in</strong>ement c = 1 and c = 10.<br />

In the case of strong conf<strong>in</strong>ement, c = 10, the simulation co<strong>in</strong>cides with the<br />

analytical value of 〈x 2 a〉 = 1.25 · 10 −5 over most of the polymer length. The<br />

<strong>in</strong>fluence of the tube is seen as the overall small value of the average undulation<br />

along the tube. The polymer segments are not allowed to deviate much from the<br />

axis because the energy penalty becomes to high.<br />

S<strong>in</strong>ce the error is smaller than the visible po<strong>in</strong>t size, the deviation from the<br />

constant analytical result is due to the effect of the boundary conditions used<br />

<strong>in</strong> the analytical calculations. In the simulation the ends are free, such that<br />

there is a section of the cha<strong>in</strong> at its ends—a length of the order of the Odijk<br />

deflection length 1/c—which is free to explore a larger distance from the tube<br />

axis. Increas<strong>in</strong>g the conf<strong>in</strong>ement reduces the effect of the boundary conditions,<br />

1 therefore 50 data po<strong>in</strong>t for the discrete undulation function<br />

0.2<br />

0.3<br />

0.4<br />

0.5


60 CHAPTER 5. SIMULATION RESULTS<br />

− 1<br />

x 2 s (s) / x 2 a<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0<br />

0.2<br />

0.4<br />

c = 1<br />

c = 10<br />

Figure 5.3: Mean-square undulation x2 (s) along the polymer parametrized by the contourlength<br />

s. Shown is the relative deviation of the simulation data x2 s(s) from the analytical<br />

results x2 <br />

a (cf. eq. (3.27)). The simulation data was obta<strong>in</strong>ed for a cha<strong>in</strong> of N = 50 segments,<br />

flexibility ɛ = 0.1 and collision parameter c = 1 and c = 10, respectively.<br />

s<strong>in</strong>ce the cha<strong>in</strong> is deflected earlier along its contour.<br />

In contrast, for weak conf<strong>in</strong>ement (c 1) the tube does not constra<strong>in</strong> the<br />

undulations anymore but only the overall center of mass motion, which is also<br />

seen <strong>in</strong> figure 5.3 for the collision parameter c = 1. Then the shape of the<br />

mean-square displacement 〈x 2 (s)〉 is parabolic and <strong>in</strong>dependent of c. The overall<br />

deviation from the analytical result starts to <strong>in</strong>crease, due to the center of mass<br />

motion.<br />

F<strong>in</strong>ally for c ≪ 1, 〈x 2 (s)〉 is not a proper measure of the polymer configuration<br />

s<strong>in</strong>ce the center of mass is free to diffuse <strong>in</strong> space. In this case of unconf<strong>in</strong>ed<br />

polymers the measured value <strong>in</strong> the simulation is not identical to the undulations<br />

2 . Hence we need to study another set of quantities to explore the difference<br />

<strong>in</strong> behavior between conf<strong>in</strong>ement and cha<strong>in</strong>-stiffness. This is done <strong>in</strong> the next<br />

sections.<br />

In summary we can classify the conf<strong>in</strong>ement <strong>in</strong>to three regimes:<br />

1. weak conf<strong>in</strong>ement c ≪ 1: The polymer is completely free.<br />

2. <strong>in</strong>termediate conf<strong>in</strong>ement c ≈ 1: The center of mass diffusion and global<br />

rotations are restricted by the tube.<br />

3. strong conf<strong>in</strong>ement c ≫ 1: Additionally <strong>in</strong>ternal bend<strong>in</strong>g modes are restricted.<br />

2 The parametrization of stiff free polymers by the weakly-bend<strong>in</strong>g rod approximation is<br />

anyway possible, s<strong>in</strong>ce for free polymers global rotations can be taken <strong>in</strong>to account by the<br />

trivial phase-space factor 4πR 2 .<br />

s<br />

0.6<br />

0.8<br />

1


5.2. THE HARMONIC POTENTIAL 61<br />

5.2.2 Tangent-tangent correlation<br />

Now we turn to <strong>in</strong>vestigate the tangent-tangent correlation function def<strong>in</strong>ed as<br />

φtt(s, s ′ ) = 〈t(s) · t(s ′ )〉 . (5.1)<br />

This allows us to extract some more <strong>in</strong>formation on the universal properties of<br />

our polymer system.<br />

The <strong>in</strong>fluence of the tube on the tangent-tangent correlation can be divided<br />

<strong>in</strong>to three regimes, when the important length scale is aga<strong>in</strong> the Odijk deflection<br />

length 1/c:<br />

1. For segments which are close enough along the cha<strong>in</strong> the <strong>in</strong>fluence of the<br />

tube is negligible. Then, the tangent-tangent correlation decays exponentially<br />

as if for a free polymer until the segments get deflected by the ‘tube’.<br />

The free behavior is seen with<strong>in</strong> the range 0 ≤ |s − s ′ | 1/c.<br />

2. In the range of |s−s ′ | ≈ 1/c the first deflection of the segments takes place.<br />

This is seen <strong>in</strong> a <strong>in</strong>termediate flatten<strong>in</strong>g of the correlation function, or, for<br />

strong conf<strong>in</strong>ement, <strong>in</strong> a local m<strong>in</strong>imum (see figs. 5.4).<br />

This feature <strong>in</strong>dicates that the fluctuations of the tangent-tangent correlations<br />

around the mean 〈δt(s) · δt(s ′ )〉 = 〈t(s) · t(s ′ )〉 − 〈t〉 2 are anticorrelated<br />

<strong>in</strong> this range, due to the deflections.<br />

3. For large distances along the cha<strong>in</strong> |s − s ′ | ≫ 1/c the correlation function<br />

depends only on the ratio c/ɛ as expected from eq. (3.35) and becomes<br />

constant except for boundary effects. This is the universal behavior we<br />

expect for strong conf<strong>in</strong>ement.<br />

The three regimes can be identified by <strong>in</strong>spect<strong>in</strong>g the plots <strong>in</strong> figure 5.4.<br />

The behavior of the tangent-tangent correlation for fixed stiffness ɛ = 0.1 is<br />

shown <strong>in</strong> figure 5.4(a) where the purely exponential behavior is seen for quasifree<br />

cha<strong>in</strong>s with c 1, i.e. on average less than one collision with the tube. Upon<br />

<strong>in</strong>creas<strong>in</strong>g the collision parameter, the correlation function develops a plateau.<br />

Figure 5.4(d) clearly displays all three regimes. The analytical results capture<br />

the exponential decay as well as the saturation of the correlation function but<br />

not the f<strong>in</strong>al decay at the ends. The deviation at the boundary is due to the<br />

free ends of the simulated cha<strong>in</strong>. The overview plot figure 5.4(a) shows that<br />

boundary effects become less pronounced as the conf<strong>in</strong>ement becomes stronger.<br />

Then the behavior is dom<strong>in</strong>ated by the constant value of strong conf<strong>in</strong>ement,<br />

which depends only on the scal<strong>in</strong>g parameter c/ɛ as expected from eq. (3.35).<br />

The behavior of the correlation function for <strong>in</strong>termediate conf<strong>in</strong>ement with<br />

c ≈ 3 is shown <strong>in</strong> 5.4(c). Here we observe a strong deviation from the analytical<br />

result, which is due to the boundary conditions. In this regime the specific<br />

choice of boundary conditions for analytical calculations is essential for the correct


62 CHAPTER 5. SIMULATION RESULTS<br />

1<br />

0.99<br />

0.98<br />

0.97<br />

0.96<br />

-0.5<br />

φtt(0.5, s + 0.5)/ t 2<br />

1<br />

0.99<br />

0.98<br />

0.97<br />

-0.5<br />

-0.4<br />

-0.3<br />

-0.2<br />

c = 0.1<br />

c = 1<br />

c = 3<br />

c = 5<br />

c = 10<br />

c = 20<br />

c = 100<br />

-0.1 0<br />

s − 0.5<br />

0.1<br />

(a) Normalized φtt for ɛ = 0.1 and a set of <strong>in</strong>creas<strong>in</strong>g collision<br />

parameters c <strong>in</strong>dicated <strong>in</strong> the graph.<br />

-0.25<br />

1<br />

0.99<br />

0.98<br />

0.97<br />

0.96<br />

0.95<br />

-0.5<br />

0<br />

-0.25<br />

0<br />

0.25<br />

0.2<br />

(b) same as <strong>in</strong> (a) for c = 5 · 10 −4<br />

0.25<br />

(c) same as <strong>in</strong> (a) for c = 3<br />

0.5<br />

1<br />

0.99<br />

-0.5<br />

-0.25<br />

0.3<br />

0.5<br />

0<br />

0.4<br />

0.5<br />

0.25<br />

(d) same as <strong>in</strong> (a) for c = 10<br />

Figure 5.4: Normalized tangent-tangent correlation function along a cha<strong>in</strong> with flexibility<br />

ɛ = 0.1. The reference po<strong>in</strong>t is fixed <strong>in</strong> the middle of the cha<strong>in</strong>, therefore the graphs must<br />

be symmetric. N = 50 segments were used <strong>in</strong> the simulations. Additionally, <strong>in</strong> (b)–(d) the<br />

analytical result eq. (3.34) is shown.<br />

0.5


5.2. THE HARMONIC POTENTIAL 63<br />

behavior of the tangent-tangent correlation function. The actual polymer <strong>in</strong> the<br />

simulation has free ends and is therefore much less constra<strong>in</strong>ed <strong>in</strong> comparison to<br />

the analytical calculation where the fluctuation of the boundaries are suppressed.<br />

The <strong>in</strong>fluence of the free ends was already seen on the undulations <strong>in</strong> figure 5.3<br />

as a large <strong>in</strong>crease of the deviation from the analytical result. But also <strong>in</strong> figure<br />

5.4(c) a rem<strong>in</strong>iscence of anticorrelations of the segments can be seen as a small<br />

shoulder at |s − 0.5| ≈ 1/c = 1/3.<br />

Figure 5.4(b) shows, for a collision parameter c = 5 · 10 −4 , an effectively free<br />

polymer. The analytical result fits the data perfectly over the whole polymer<br />

length, except for a very small deviation at the ends. This behavior is expected<br />

s<strong>in</strong>ce the analytical calculation is performed <strong>in</strong> the weakly-bend<strong>in</strong>g rod limit,<br />

which gives the correct result for unconf<strong>in</strong>ed stiff polymers up to first order <strong>in</strong> ɛ,<br />

as shown <strong>in</strong> the discussion lead<strong>in</strong>g to eq. (3.36). Therefore the small deviation<br />

are only due to the limits of the weakly-bend<strong>in</strong>g rod approximation (terms of<br />

order ɛ 2 ) and not due to boundary effects.<br />

F<strong>in</strong>ite-size effects have only a very small <strong>in</strong>fluence on the results obta<strong>in</strong>ed for<br />

the tangent-tangent correlation. The overall shape is not affected by <strong>in</strong>creas<strong>in</strong>g<br />

the number of segments N. Only small global shifts of the whole correlation<br />

function can be observed. This has only an effect on <strong>in</strong>tegral quantities as the<br />

end-to-end distance, which will be discussed <strong>in</strong> the follow<strong>in</strong>g section.<br />

After these prelim<strong>in</strong>ary <strong>in</strong>vestigations towards the validity of the dependence<br />

of the observables on the scal<strong>in</strong>g parameter c/ɛ, we now turn to the systematic<br />

<strong>in</strong>vestigation us<strong>in</strong>g a scal<strong>in</strong>g plot.<br />

5.2.3 Scal<strong>in</strong>g plot<br />

We have already discussed the <strong>in</strong>fluence of conf<strong>in</strong>ement on the undulations and<br />

the tangent-tangent correlation. Now we will turn to quantities, which are of<br />

particular <strong>in</strong>terest <strong>in</strong> experiments, where it is easier to measure the end-to-end<br />

distance and related quantities than the very small undulation or a correlation<br />

function of the tangents.<br />

We want to present the data <strong>in</strong> a compact form, which works out the scal<strong>in</strong>g<br />

properties expected by our <strong>in</strong>vestigation up to this po<strong>in</strong>t. In a suitable plot<br />

we expect a data collapse. Look<strong>in</strong>g back at eqs. (3.12) and (3.38) we see that<br />

plott<strong>in</strong>g the mean-square end-to-end distance 〈R 2 〉 versus c/ɛ is expected to show<br />

the data collapse. But before turn<strong>in</strong>g to the plot we first give an overview over<br />

the various parameter regimes and than try to f<strong>in</strong>d out <strong>in</strong> which range the scal<strong>in</strong>g<br />

is expected to start, s<strong>in</strong>ce it is not universal for all parameters as noted <strong>in</strong> the<br />

preced<strong>in</strong>g section.


64 CHAPTER 5. SIMULATION RESULTS<br />

Comparison of analytic and simulation results<br />

Even if one measures all length <strong>in</strong> units of the total polymer length L, the meansquare<br />

end-to-end distance is still a function of two length scales, the tube diameter<br />

d and the persistence length lp. As discussed <strong>in</strong> section 3.2 this suggests to<br />

<strong>in</strong>troduce two dimensionless parameters, ɛ = L/lp as a measure of the polymer<br />

stiffness and c = L/ld count<strong>in</strong>g the number of collision of the polymer with the<br />

tube and be<strong>in</strong>g related to the tube diameter through the relation <strong>in</strong> eq. (3.1).<br />

In the discussion we have to dist<strong>in</strong>guish the limit<strong>in</strong>g cases of stiff (ɛ ≪ 1) and<br />

flexible polymers (ɛ ≫ 1). In the stiff limit the ma<strong>in</strong> parameter characteriz<strong>in</strong>g<br />

the <strong>in</strong>fluence of the conf<strong>in</strong>ement on the polymer conformation is the collision<br />

parameter c. For c ≪ 1 the polymer is unconf<strong>in</strong>ed and behaves as a free polymer<br />

with 〈R 2 〉 ∼ L 2 (cf. eq. (2.22)). For c ≈ 1 undulations for the filament are not<br />

yet restricted but the overall rotation of the filament is h<strong>in</strong>dered by the walls.<br />

For c ≫ 1 there is strong conf<strong>in</strong>ement with many collisions of the polymer with<br />

the wall.<br />

In the flexible limit the critical scales to compare are the radius of gyration<br />

Rg and the tube diameter. As noted by de Gennes the scal<strong>in</strong>g behavior of the<br />

unconf<strong>in</strong>ed cha<strong>in</strong>, 〈R 2 〉 ∼ lpL, starts to change once the mean-square end-to-end<br />

distance becomes comparable to the tube diameter d. Then the polymer starts<br />

to stretch due to the excluded volume <strong>in</strong>teraction of distant segments along the<br />

cha<strong>in</strong>, such that 〈R 2 〉 ∼ Lc/ɛ. Here we consider phantom cha<strong>in</strong>s where selfavoidance<br />

effects are neglected. Hence, as discussed <strong>in</strong> section 3.2.1 one expects<br />

no effect at all of conf<strong>in</strong>ement on the polymer configuration. A stretch<strong>in</strong>g for a<br />

cha<strong>in</strong> consist<strong>in</strong>g of segments is then only due to the f<strong>in</strong>ite segment length.<br />

Even for flexible cha<strong>in</strong>s stiffness effects will become important once the tube<br />

diameter becomes comparable with the persistence length. Then the same arguments<br />

as <strong>in</strong> the stiff case apply.<br />

From the forego<strong>in</strong>g discussion we can expect to f<strong>in</strong>d universal behavior for<br />

stiff cha<strong>in</strong>s, when the cha<strong>in</strong> starts to be deflected, c 1. S<strong>in</strong>ce we expect to have<br />

a dependence on the scal<strong>in</strong>g parameter c/ɛ this means<br />

c<br />

ɛ ɛ−1 . (5.2)<br />

On the other hand for flexible cha<strong>in</strong>s the relevant condition is that the tube<br />

diameter is of the order of the segment length or less d lp which leads through<br />

eq. (3.1) to<br />

c<br />

1 (5.3)<br />

ɛ<br />

for the scal<strong>in</strong>g parameter. We can summarize this conditions for the universal<br />

scal<strong>in</strong>g regime <strong>in</strong><br />

c<br />

<br />

<br />

max(1, ɛ<br />

ɛ univ<br />

−1 ). (5.4)<br />

For flexible polymers we therefore expect to have a longer scal<strong>in</strong>g regime.


5.2. THE HARMONIC POTENTIAL 65<br />

R 2 <br />

1<br />

0.99<br />

0.98<br />

0.97<br />

0.96<br />

1<br />

10<br />

100<br />

c/ɛ<br />

1000<br />

ɛ = 0.01<br />

ɛ = 0.05<br />

ɛ = 0.1<br />

ɛ = 0.18<br />

10000<br />

Figure 5.5: Mean-square end-to-end distance R 2 versus the scal<strong>in</strong>g variable c/ɛ for different<br />

flexibilities ɛ <strong>in</strong> the stiff regime. Along with the simulation data for cha<strong>in</strong>s with N = 50<br />

segments, the analytical result from eq. (3.38) and the free result eq. (2.21) is shown.<br />

With these consideration at hand we start the <strong>in</strong>vestigation of the simulation<br />

result <strong>in</strong> the stiff limit as shown <strong>in</strong> figure 5.5. Here the mean-square end-to-end<br />

distance 〈R 2 〉 is plotted as a function of the scal<strong>in</strong>g parameter c/ɛ for different<br />

flexibilities ɛ. In addition to the simulation data the analytical results obta<strong>in</strong>ed<br />

<strong>in</strong> eq. (3.38) are shown (non-constant l<strong>in</strong>es) as well as correspond<strong>in</strong>g results for<br />

a free polymer c = 0 (constant l<strong>in</strong>es) from eq. (2.21).<br />

We observe that the simulation data starts to deviate at c/ɛ ≈ ɛ −1 from the<br />

free result represented by the horizontal l<strong>in</strong>es. This is the behavior expected<br />

by our scal<strong>in</strong>g arguments (see eq. (5.2)). Both <strong>in</strong> the unconf<strong>in</strong>ed and strongly<br />

conf<strong>in</strong>ed regime the simulation data are well described by our analytical results.<br />

This is due to the nature of the weakly-bend<strong>in</strong>g rod approximation which is valid<br />

for strong conf<strong>in</strong>ement as well as for very stiff cha<strong>in</strong>s.<br />

For the <strong>in</strong>termediate regime, e.g. for a flexibility ɛ = 0.1 <strong>in</strong> the range 1 <br />

c/ɛ 100, there is a strong deviation from the the analytical results, which is<br />

not expected from the assumptions lead<strong>in</strong>g to the weakly-bend<strong>in</strong>g rod approximation.<br />

But <strong>in</strong> the context of the discussion of the tangent-tangent correlation<br />

function <strong>in</strong> the preced<strong>in</strong>g section we understand this deviation. Remember<strong>in</strong>g the<br />

discussion of figure 5.4(c) for the parameter c = 3 (correspond<strong>in</strong>g to the scal<strong>in</strong>g<br />

parameter c/ɛ = 30), these deviations are aga<strong>in</strong> due to the free boundaries <strong>in</strong><br />

the simulation. The mean-square end-to-end distance is an <strong>in</strong>tegral quantity of<br />

the tangent correlation function (cf. section 3.3.1), therefore these deviations are<br />

most prom<strong>in</strong>ent <strong>in</strong> the end-to-end distance. S<strong>in</strong>ce the tangent correlation function<br />

for the simulation decays stronger than the analytical result the end-to-end<br />

distance shows values which are too small <strong>in</strong> comparison to eq. (3.38).<br />

Now let us turn to the flexible limit, L ≫ lp. Then we expect from eq. (5.3)<br />

the scal<strong>in</strong>g range to be larger, which can be seen <strong>in</strong> figure 5.6. Here the mean-


66 CHAPTER 5. SIMULATION RESULTS<br />

R 2 <br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.1<br />

1<br />

10<br />

c/ɛ<br />

100<br />

ɛ = 0.1<br />

ɛ = 0.5<br />

ɛ = 1<br />

ɛ = 5<br />

ɛ = 10<br />

ɛ = 50<br />

ɛ = 300<br />

Figure 5.6: Scal<strong>in</strong>g plot for the mean-square end-to-end distance R 2 versus the scal<strong>in</strong>g<br />

parameter c/ɛ for various flexibilities ɛ from the stiff to the flexible polymer limit.<br />

square end-to-end distance 〈R 2 〉 as function of the scal<strong>in</strong>g parameter c/ɛ is shown<br />

for the whole range of flexibilities ɛ, from the stiff up to the flexible limit. It can<br />

be seen that a universal behavior is visible to start around c/ɛ 5. Tak<strong>in</strong>g a<br />

lower envelope of the data reveals the expected scal<strong>in</strong>g.<br />

The deviations from a lower envelope <strong>in</strong> this range can be expla<strong>in</strong>ed by f<strong>in</strong>itesize<br />

effects of the simulated cha<strong>in</strong>s. Already for ɛ ≈ 5 differences from the analytic<br />

result eq. (3.38) can be observed. This is explicitly shown <strong>in</strong> figure 5.7 for the<br />

mean-square end-to-end distance 〈R 2 〉 versus the scal<strong>in</strong>g parameter c/ɛ for a<br />

flexibility ɛ = 5. The simulation data is obta<strong>in</strong>ed by <strong>in</strong>creas<strong>in</strong>g the number N of<br />

the segments up to 1000. S<strong>in</strong>ce a simulation for cha<strong>in</strong>s of length N = 1000 takes<br />

<strong>in</strong> the order of a week per data po<strong>in</strong>t, only s<strong>in</strong>gle po<strong>in</strong>ts have been simulated.<br />

The <strong>in</strong>set gives a zoomed view of the area where the major effect of f<strong>in</strong>ite-size is<br />

seen. The data obta<strong>in</strong>ed for N = 1000 is seen to fit the analytic result perfectly.<br />

The data converges, <strong>in</strong> this range, for decreas<strong>in</strong>g segments size to the analytic<br />

result.<br />

These f<strong>in</strong>ite-size deviations <strong>in</strong>crease for more flexible cha<strong>in</strong>s: The condition<br />

that the segment length l = 1/N should be at least about the size of the persistence<br />

length or smaller—otherwise the persistence length is artificially <strong>in</strong>creased<br />

due to f<strong>in</strong>ite-size effects—forces us to use more segments for flexible cha<strong>in</strong>s.<br />

Therefore deviations due to f<strong>in</strong>ite-size are <strong>in</strong>creas<strong>in</strong>g even more. But us<strong>in</strong>g much<br />

more segments than N = 1000 is problematic due to the simulation time needed.<br />

The form of the mean-square end-to-end distance for very flexible cha<strong>in</strong> is<br />

given by the limit of the ideal cha<strong>in</strong> without persistence, the freely jo<strong>in</strong>ted cha<strong>in</strong><br />

(FJC). For c/ɛ 1 the function stays effective constant and for c/ɛ 1 the<br />

function behaves like the approximation for strong conf<strong>in</strong>ement eq. (3.9), only<br />

1000


5.2. THE HARMONIC POTENTIAL 67<br />

R 2 /L 2<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

analytic<br />

N = 50<br />

N = 100<br />

N = 500<br />

N = 1000<br />

1<br />

1<br />

Figure 5.7: Mean-square end-to-end distance R 2 versus the scal<strong>in</strong>g parameter c/ɛ for different<br />

levels of discretization N (flexibility ɛ = 5) and the analytic result eq. (3.38). The <strong>in</strong>set<br />

shows a zoomed region, where the <strong>in</strong>fluence of f<strong>in</strong>ite-size effects is seen best.<br />

10<br />

10<br />

c/ɛ<br />

the prefactor <strong>in</strong> the term (d/lp) 2 has to be adjusted 3 .<br />

Comparison to the experiment<br />

Now we turn to the discussion of how the simulations compare to experiments.<br />

The correspond<strong>in</strong>g experiments were performed by Walter Reisner <strong>in</strong> Bob Aust<strong>in</strong>s<br />

group at Pr<strong>in</strong>ceton. The details of the experiment are discussed <strong>in</strong> section 3.1<br />

and <strong>in</strong> the references [38, 44].<br />

For the experimental measurements on flexible cha<strong>in</strong>s the most important<br />

quantity to <strong>in</strong>vestigate is the the apparent end-to-end distance 〈Ra〉 which measures<br />

the largest possible distance between two segments along the cha<strong>in</strong>. In<br />

experiments the polymer under <strong>in</strong>vestigation is marked with some fluorescent<br />

dye, such that only a blurry l<strong>in</strong>e is visible, as seen for DNA <strong>in</strong> the figures 1.2.<br />

To resolve the actual position of the end po<strong>in</strong>t of a polymer is experimentally<br />

very hard to achieve. One could imag<strong>in</strong>e us<strong>in</strong>g molecular biology technology to<br />

specifically label certa<strong>in</strong> regimes along a DNA strand. But this is not how experiments<br />

<strong>in</strong> the Aust<strong>in</strong> group are performed. They use TOTO-1 fluorescent dye<br />

to uniformly label the whole filament. In such a situation the proper analytical<br />

quantity to compare to the theory is the apparent end-to-end distance 〈Ra〉.<br />

3 For the FJC stretch<strong>in</strong>g of the cha<strong>in</strong> is only due to the f<strong>in</strong>ite segment length, and the fact that<br />

we did not allow for backflips. If we would have allowed for those backflips the model would,<br />

<strong>in</strong> the stiff limit, reduce to an “Is<strong>in</strong>g cha<strong>in</strong>”, where each segment may either po<strong>in</strong>t parallel or<br />

anti-parallel to the tube axis. Then, of course, the average end-to-end distance would aga<strong>in</strong><br />

be zero. S<strong>in</strong>ce we did not allow for backflips the anti-parallel segment configurations are not<br />

allowed mak<strong>in</strong>g our system non ergodic. This is still an <strong>in</strong>terest<strong>in</strong>g and useful limit s<strong>in</strong>ce it<br />

allows us to study the effect of stiff cha<strong>in</strong> segments which—<strong>in</strong> a real system—would neither be<br />

allowed to backflip.<br />

100<br />

100


68 CHAPTER 5. SIMULATION RESULTS<br />

〈Ra〉<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.1<br />

1<br />

10<br />

c/ɛ<br />

experimental data<br />

Odijk scal<strong>in</strong>g (fit)<br />

ɛ = 0.1<br />

ɛ = 0.5<br />

ɛ = 1<br />

ɛ = 5<br />

ɛ = 10<br />

ɛ = 50<br />

ɛ = 300<br />

Figure 5.8: Scal<strong>in</strong>g plot for the apparent end-to-end distance 〈Ra〉 as a function of the scal<strong>in</strong>g<br />

parameter c/ɛ. Beside the simulation data for various flexibilities ɛ and the Odijk scal<strong>in</strong>g<br />

relation, eq. (3.12), with fitted parameter, experimental data for λ-DNA (see text and [38]) is<br />

shown.<br />

The experimental data obta<strong>in</strong>ed for λ-DNA is shown <strong>in</strong> figure 5.8 <strong>in</strong> comparison<br />

to our simulation data <strong>in</strong> the cyl<strong>in</strong>drical harmonic potential. In the<br />

experiment the geometric average of the channel width is taken as the effective<br />

tube diameter, which can be converted approximately—s<strong>in</strong>ce the prefactor <strong>in</strong> the<br />

relation is not exactly known, it can only be fitted—<strong>in</strong>to a collision parameter<br />

by tak<strong>in</strong>g eq. (3.29) literally. The Odijk scal<strong>in</strong>g relation, eq. (3.12), is also fitted<br />

<strong>in</strong>to the plot (α ′ = 0.361).<br />

Although the geometry of the experiment is different from the simulations<br />

performed and the conversion of the channel width to the conf<strong>in</strong><strong>in</strong>g scale of the<br />

harmonic potential is not rigorous, the data fits very well <strong>in</strong>to the trend seen <strong>in</strong><br />

the plot. Despite the good trend, a more quantitative comparison is not possible.<br />

But the deviation <strong>in</strong> the range c/ɛ 2 can be understood:<br />

100<br />

1000<br />

• There is some effect of the specific conf<strong>in</strong><strong>in</strong>g geometry used <strong>in</strong> the simulation.<br />

We used a harmonic potential and the experimental data was obta<strong>in</strong>ed<br />

<strong>in</strong> a rectangular channel. But this effect is expected to be small and we<br />

already have a relatively good relation by us<strong>in</strong>g eq. (3.1) to compare the<br />

geometries. See also section 5.3 for more details.<br />

• Us<strong>in</strong>g a large number of segments for flexible cha<strong>in</strong>s could br<strong>in</strong>g us <strong>in</strong>to<br />

the regime where self-avoidance due to excluded volume effects f<strong>in</strong>ally becomes<br />

important <strong>in</strong> static problems (cf. discussion on page 18). Check<strong>in</strong>g<br />

this would require a large amount of computer time. Already the simulations<br />

for large N were slow due to the f<strong>in</strong>e discretization. Introduc<strong>in</strong>g<br />

self-avoidance <strong>in</strong> a straightforward way <strong>in</strong>creases the simulation time to an<br />

unacceptable length. To <strong>in</strong>vestigate this, a further development of a novel<br />

and fast algorithm is needed.


5.2. THE HARMONIC POTENTIAL 69<br />

<br />

R 2 and R 2 ||<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.1<br />

R 2 <br />

<br />

2 R|| 1<br />

Figure 5.9: The ‘dip’-feature <strong>in</strong> the end-to-end<br />

<br />

distance. The mean-square and mean-square<br />

, are shown versus the collision parameter c for the<br />

parallel end-to-end distance, R 2 and R 2 ||<br />

flexible regime ɛ = 10.<br />

10<br />

c<br />

100<br />

1000<br />

• In particular for biopolymers electrostatic effects play an important role.<br />

E.g. DNA is a highly charged polyelectrolyte, where the electrostatic repulsion<br />

will have a stretch<strong>in</strong>g effect, lead<strong>in</strong>g to a second source of self-avoidance<br />

(cf. [36]), beside the excluded volume effect.<br />

In our simulations electrostatic effects are not taken <strong>in</strong>to account. Therefore<br />

an extra repulsion between the segments is expected, significantly effect<strong>in</strong>g<br />

the flexible regime. This is expected to shift our results <strong>in</strong> a way, that the<br />

trend of the experimental data <strong>in</strong> the <strong>in</strong>termediate regime is fitted better.<br />

Simulations of polyelectrolytes need special complex techniques. Implementation<br />

of these effects is an <strong>in</strong>terest<strong>in</strong>g task for future projects.<br />

It can be noted by <strong>in</strong>vestigat<strong>in</strong>g figure 5.8, that chang<strong>in</strong>g the observable from<br />

the mean-square end-to-end distance to the apparent end-to-end distance does<br />

not change our discussion about the scal<strong>in</strong>g regime lead<strong>in</strong>g to eq. (5.4).<br />

The goal of this scal<strong>in</strong>g plot is to use it as a gauge plot for the experiments, to<br />

convert the apparent length measured, <strong>in</strong>to the real contour length of the polymer<br />

strand.<br />

5.2.4 End-to-end distances<br />

In order to understand the conformational changes which take place upon conf<strong>in</strong>ement,<br />

we are go<strong>in</strong>g to <strong>in</strong>vestigate the differences between the full end-to-end<br />

distance and its component along the tube axis. The def<strong>in</strong>ition of these observables<br />

was given <strong>in</strong> section 4.6. The <strong>in</strong>vestigation will expla<strong>in</strong> a global feature<br />

qualitatively, only visible for more flexible cha<strong>in</strong>s, which is different between these<br />

two observables.


70 CHAPTER 5. SIMULATION RESULTS<br />

R<br />

R ||<br />

Figure 5.10: Behavior of the end-to-end distances of flexible cha<strong>in</strong>s upon conf<strong>in</strong>ement.<br />

A representative example of these observables for a flexible cha<strong>in</strong> ɛ = 10 is<br />

shown <strong>in</strong> figure 5.9, where the mean-square and mean-square parallel end-to-end<br />

distance, 〈R2 〉 and R2 <br />

|| , are plotted as a function of the collision parameter c.<br />

The full end-to-end distance exhibits a ‘dip’ (local m<strong>in</strong>imum) for c ≈ 1, when the<br />

cha<strong>in</strong> beg<strong>in</strong>s to notice the tube through deflections. This dip is absent for R2 <br />

|| .<br />

The two limits, c → 0 and c → ∞, are easily understood: In the limit of<br />

an unconf<strong>in</strong>ed cha<strong>in</strong> (c < 1) the average of the z-component is exactly a third<br />

of the full average R2 <br />

2<br />

|| = 〈R 〉 /3. In the opposite limit, when the conf<strong>in</strong>ement<br />

becomes larger (c > 20), both observables, 〈R2 〉 and R2 <br />

|| , behave identically<br />

s<strong>in</strong>ce the polymer is oriented ma<strong>in</strong>ly along the tube. Therefore the end-to-end<br />

distance is nearly parallel to the tube axis.<br />

The dip—which has also been observed without explanation <strong>in</strong> [43]—can be<br />

expla<strong>in</strong>ed as illustrated <strong>in</strong> figure 5.10. Upon conf<strong>in</strong>ement of the flexible cha<strong>in</strong>,<br />

the <strong>in</strong>itially randomly oriented end-to-end distance is h<strong>in</strong>dered <strong>in</strong> the perpendicular<br />

directions. This implies that first the end-to-end distance is changed<br />

predom<strong>in</strong>antly perpendicular to the tube axis, only decreas<strong>in</strong>g the value of 〈R2 〉<br />

but leav<strong>in</strong>g R2 <br />

|| unchanged. Formation of the dip should start, as seen <strong>in</strong> figure<br />

5.9, at a value of c 1. Upon further conf<strong>in</strong>ement the end-to-end distance<br />

slowly starts giv<strong>in</strong>g way <strong>in</strong> the direction along the tube, f<strong>in</strong>ally also <strong>in</strong>creas<strong>in</strong>g<br />

the parallel component.<br />

The behavior of the end-to-end distances upon conf<strong>in</strong>ement can also be seen<br />

<strong>in</strong> the shape of the radial distribution functions (see next section 5.2.5).<br />

The dip is present <strong>in</strong> all end-to-end distances <strong>in</strong>troduced <strong>in</strong> section 4.6, as it<br />

is absent for all parallel end-to-end distances.<br />

5.2.5 Radial distribution function<br />

In this section we are go<strong>in</strong>g to <strong>in</strong>vestigate the probability distribution of the<br />

end-to-end distance or radial distribution function (RDF) (see section 3.3.3 for<br />

a discussion of the RDF <strong>in</strong> the unconf<strong>in</strong>ed case). Figures 5.11 show some representative<br />

numerical results for a cha<strong>in</strong> with flexibility ɛ = 10 and for <strong>in</strong>creas<strong>in</strong>g<br />

collision parameter c.<br />

The RDF for the end-to-end distance R is shown <strong>in</strong> figure 5.11(a) and for<br />

δR ||<br />

δR<br />

d


5.2. THE HARMONIC POTENTIAL 71<br />

p(R)<br />

p(R ||)<br />

4<br />

2<br />

0<br />

0<br />

4<br />

2<br />

0<br />

0<br />

c = 2<br />

c = 3<br />

c = 2<br />

c = 3<br />

0.2<br />

0.2<br />

0.4<br />

c = 6<br />

c = 10<br />

R<br />

0.6<br />

(a) RDF for R = |R|<br />

0.4<br />

c = 6<br />

c = 10<br />

R ||<br />

0.6<br />

(b) RDF for R || = |R |||<br />

c = 15<br />

c = 20<br />

c = 15<br />

c = 20<br />

0.8<br />

0.8<br />

c = 60<br />

c = 60<br />

Figure 5.11: Comparison of the probability distributions of the end-to-end distances for a<br />

cha<strong>in</strong> of flexibility ɛ = 10 and <strong>in</strong>creas<strong>in</strong>g collision parameter c.<br />

1<br />

1


72 CHAPTER 5. SIMULATION RESULTS<br />

the parallel end-to-end distance R|| <strong>in</strong> figure 5.11(b). For an unconf<strong>in</strong>ed cha<strong>in</strong><br />

(c 2), p(R) is peaked around R = 0.4 and shows Gaussian behavior, whereas<br />

p(R||) is peaked at R|| = 0. This is simply due to a phase space factor of 4πR 2<br />

<strong>in</strong>cluded <strong>in</strong> p(R) result<strong>in</strong>g from sampl<strong>in</strong>g all directions of R. Such a phase space<br />

factor is absent from p(R||) which samples the projection of R on the tube axis.<br />

For <strong>in</strong>termediate values of c (6 c 20) the RDF p(R) shows an <strong>in</strong>terest<strong>in</strong>g<br />

bimodal distribution. If the collision parameter c is <strong>in</strong>creased from c ≈ 2 the peak<br />

is shifted towards smaller values of R such that the distribution is significantly<br />

skewed. Then, at about c ≈ 6 a second peak at large values of R develops, which<br />

then grows at the expense of the peak at low values of R.<br />

This feature is closely related to the ‘dip’ discussed <strong>in</strong> the previous section.<br />

The <strong>in</strong>itial drift of the peak towards smaller values is due to the restriction<br />

of the polymers’ configuration perpendicular to the tube axis. Such a behavior<br />

is expected from a l<strong>in</strong>ear response argument. When the conf<strong>in</strong>ement becomes<br />

stronger, the end-to-end distance evades the transverse compression by<br />

also stretch<strong>in</strong>g along the tube axis. As a result the RDF p(R) exhibits two<br />

peaks, one at small and the other at large values of R. This quantifies the heuristic<br />

arguments for the dip of 〈R 2 〉 <strong>in</strong> the previous section.<br />

For comparison the parallel end-to-end distance distribution is shown <strong>in</strong> figure<br />

5.11(b) for the same set of parameters. Obviously the average is grow<strong>in</strong>g<br />

monotonously to larger values, not show<strong>in</strong>g any bimodal behavior. We have already<br />

eluded to the fact that p(R||) is peaked at R|| = 0 for weak conf<strong>in</strong>ement<br />

(due to phase space effects). Hence, <strong>in</strong>creas<strong>in</strong>g the conf<strong>in</strong>ement can only result<br />

<strong>in</strong> a shift of this peak to larger values.<br />

F<strong>in</strong>ally for narrow ‘tubes’ (c 30) both RDFs show exactly the same shape<br />

s<strong>in</strong>ce the end-to-end distance is now oriented parallel to the tube-axis.<br />

5.3 The hard walls<br />

In the f<strong>in</strong>al chapter we turn to the <strong>in</strong>vestigation of prelim<strong>in</strong>ary simulation results<br />

concern<strong>in</strong>g the conformations of polymers conf<strong>in</strong>ed by hard walls. In particular,<br />

we are go<strong>in</strong>g to analyze the scal<strong>in</strong>g of the end-to-end distance similar as <strong>in</strong> section<br />

5.2.3. We study two different cross sections of the tube, a circular and square<br />

shaped channel, and leave other geometries for future <strong>in</strong>vestigations.<br />

S<strong>in</strong>ce we are ma<strong>in</strong>ly <strong>in</strong>terested <strong>in</strong> compar<strong>in</strong>g with the experimental data (cf.<br />

section 5.2.3) we concentrate on the mean apparent end-to-end distance 〈Ra〉.<br />

The simulation data for the two different channel cross sections are shown <strong>in</strong> figures<br />

5.12 and 5.13. In both cases one f<strong>in</strong>ds scal<strong>in</strong>g behavior with scal<strong>in</strong>g variables<br />

(dɛ) −1 and (wɛ) −1 , respectively. Here d is the tube diameter for a circular cross<br />

section and w the channel width for a square cross section.<br />

Compar<strong>in</strong>g both plots with each other and to figure 5.6 shows that the overall<br />

behavior is identical and that there appears to be no major <strong>in</strong>fluence of the chosen


5.3. THE HARD WALLS 73<br />

〈Ra〉<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.01<br />

0.1<br />

1<br />

(dɛ) −1<br />

10<br />

experimental data<br />

Odijk scal<strong>in</strong>g (fit)<br />

0.5<br />

1<br />

5<br />

10<br />

50<br />

100<br />

1000<br />

10000<br />

Figure 5.12: Mean apparent end-to-end 〈Ra〉 versus the scal<strong>in</strong>g parameter (dɛ) −1 <strong>in</strong> the<br />

circular cross section tube with diameter d. Besides simulation data for various flexibilities ɛ,<br />

experimental data and the correspond<strong>in</strong>g Odijk scal<strong>in</strong>g fit are shown.<br />

〈Ra〉<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.01<br />

0.1<br />

1<br />

(wɛ) −1<br />

10<br />

experimental data<br />

Odijk scal<strong>in</strong>g (fit)<br />

ɛ = 0.5<br />

ɛ = 5<br />

ɛ = 10<br />

ɛ = 50<br />

ɛ = 300<br />

Figure 5.13: Mean apparent end-to-end 〈Ra〉 versus the scal<strong>in</strong>g parameter (dɛ) −1 <strong>in</strong> the square<br />

cross section tube with width w. Besides simulation data for various flexibilities ɛ, experimental<br />

data and the correspond<strong>in</strong>g Odijk scal<strong>in</strong>g fit are shown.<br />

100<br />

1000


74 CHAPTER 5. SIMULATION RESULTS<br />

geometry. The ma<strong>in</strong> features of the behavior of the polymer <strong>in</strong> the conf<strong>in</strong><strong>in</strong>g ‘tube’<br />

are universal. The relevant <strong>in</strong>gredient is the rotational symmetry around the tube<br />

axis, but the details, cont<strong>in</strong>uous symmetry for the circular tube and harmonic<br />

potential or fourfold discrete symmetry for the square tube, are irrelevant. The<br />

only relevant scale is the deflection length ld! The difference <strong>in</strong> slope between the<br />

data <strong>in</strong> the harmonic potential scal<strong>in</strong>g plot and the data presented here is only<br />

due to the relation convert<strong>in</strong>g the relevant “<strong>in</strong>put parameters” d and ld by eq.<br />

(3.1).<br />

The details <strong>in</strong> the shape of the scal<strong>in</strong>g function for hard and soft walls are,<br />

however, quite different. For soft walls discussed <strong>in</strong> section 5.2.3 we were able to<br />

fit the Odijk scal<strong>in</strong>g result to both the experimental data and the Monte-Carlo<br />

results. Here this seems not to be possible. The Odijk curve fits the experimental<br />

data but misses the sharp <strong>in</strong>crease of the Monte-Carlo data for hard walls <strong>in</strong> the<br />

limit of strong conf<strong>in</strong>ement.<br />

This raises the question on the actual form of the <strong>in</strong>teraction potential between<br />

the polymer an the conf<strong>in</strong><strong>in</strong>g walls. Us<strong>in</strong>g a hard wall potential one assumes that<br />

there are steric <strong>in</strong>teractions only. This is not necessarily true s<strong>in</strong>ce we are deal<strong>in</strong>g<br />

with a charged system. DNA is a polyelectrolyte and therefore we can only hope<br />

to describe this problem by an effective potential. If such a construction of an<br />

effective potential is not possible, an explicit simulation of the charged polymer<br />

and the charges <strong>in</strong> solution must be performed. This is an <strong>in</strong>terest<strong>in</strong>g future<br />

project but beyond the scope of this thesis.<br />

With the simulation program used to obta<strong>in</strong> all data here, many <strong>in</strong>terest<strong>in</strong>g<br />

properties can be discussed. Here our first task was to have a comparison to the<br />

analytical theory to validate the simulation and then to obta<strong>in</strong> experimentally<br />

relevant data presented here. It provides a solid basis for future work.


Chapter 6<br />

Conclusion<br />

Emerg<strong>in</strong>g from bio- and nanotechnology, the <strong>in</strong>terest <strong>in</strong> polymers <strong>in</strong> conf<strong>in</strong><strong>in</strong>g<br />

geometries is rapidly grow<strong>in</strong>g. The property of tube-like environments of lateral<br />

dimension on the nanometer scale forces polymers to extend to a substantial<br />

fraction of its contour length, which has direct applications <strong>in</strong> bioeng<strong>in</strong>eer<strong>in</strong>g,<br />

e.g. for the measurement of contour lengths.<br />

The goal <strong>in</strong> this thesis was to get a deeper understand<strong>in</strong>g of the conformational<br />

behavior of polymers <strong>in</strong> conf<strong>in</strong><strong>in</strong>g environments, <strong>in</strong> particular tube like structures.<br />

Along with the exploration of analytically feasible limits of the polymer conformation,<br />

Monte-Carlo simulations were essential to <strong>in</strong>vestigate the experimental<br />

relevant parameter regimes, <strong>in</strong> particular for experiments on DNA. With a better<br />

understand<strong>in</strong>g of the subtle <strong>in</strong>terplay between cha<strong>in</strong> flexibility and degree of<br />

conf<strong>in</strong>ement this should lead to the identification of universal scal<strong>in</strong>g properties.<br />

Investigations of the scal<strong>in</strong>g properties of conf<strong>in</strong>ed polymers were pioneered by<br />

de Gennes. He considered the comb<strong>in</strong>ed effect of conf<strong>in</strong>ement and self-avoidance<br />

on the conformation of flexible polymers. The limit of stiff cha<strong>in</strong>s was later explored<br />

by Odijk. Further theoretical progress <strong>in</strong> analytical and numerical studies<br />

was ma<strong>in</strong>ly on the conf<strong>in</strong>ement free energy, not directly relevant to current experimental<br />

<strong>in</strong>vestigations <strong>in</strong> nanofluidics. This lack of theoretical analysis as well<br />

as recent experimental results of the Aust<strong>in</strong> group at Pr<strong>in</strong>ceton motivated this<br />

thesis.<br />

We were able to get an advanced understand<strong>in</strong>g of the behavior of polymers<br />

<strong>in</strong> conf<strong>in</strong><strong>in</strong>g geometries, extend<strong>in</strong>g the established scal<strong>in</strong>g arguments to the limits<br />

of stiff cha<strong>in</strong>s and strong conf<strong>in</strong>ement relevant for recent experimental <strong>in</strong>vestigations.<br />

The ma<strong>in</strong> focus, however, was on f<strong>in</strong>d<strong>in</strong>g a quantitative description of the<br />

conformational behavior of polymers us<strong>in</strong>g both analytical treatment if feasible<br />

and simulation techniques to fully explore parameter space. This lead to the<br />

follow<strong>in</strong>g specific results.<br />

Analytical expressions were derived for the cyl<strong>in</strong>drical symmetric harmonic<br />

potential <strong>in</strong> the weakly-bend<strong>in</strong>g rod approximation. By construction this approximation<br />

is valid for stiff polymers. But, it turns out to be valid also <strong>in</strong> the<br />

75


76 CHAPTER 6. CONCLUSION<br />

strong conf<strong>in</strong><strong>in</strong>g limit for semi-flexible to flexible polymers. In particular we could<br />

derive an explicit expression for the mean-square end-to-end distance 〈R 2 〉. It<br />

shows that <strong>in</strong>stead of depend<strong>in</strong>g on both the persistence length and the Odijk<br />

length (or tube diameter), it depends on a s<strong>in</strong>gle scal<strong>in</strong>g parameter c/ɛ, the ratio<br />

of the conf<strong>in</strong>ement strength to the flexibility of the polymer, only.<br />

By off-lattice Monte-Carlo simulations we were able to verify this scal<strong>in</strong>g<br />

behavior over a broad parameter range. This allowed us to verify the validity of<br />

the weakly-bend<strong>in</strong>g rod approximation <strong>in</strong> the strong conf<strong>in</strong>ement limit. It is found<br />

that polymers (for all flexibilities) stretch out to nearly their full contour length<br />

upon conf<strong>in</strong>ement to lateral dimensions smaller than their persistence length.<br />

These results obta<strong>in</strong>ed for the cyl<strong>in</strong>drical harmonic potential were compared<br />

to experimental data obta<strong>in</strong>ed for λ-DNA conf<strong>in</strong>ed <strong>in</strong> channels with a square<br />

cross section. The overall trend of the data for the apparent end-to-end distance<br />

is well described by the Monte-Carlo data. There are some systematic deviations<br />

h<strong>in</strong>t<strong>in</strong>g at additional effects not accounted for by a worm-like cha<strong>in</strong> model.<br />

Further prelim<strong>in</strong>ary simulation for different geometries with hard-walls were<br />

performed as well. The results show a similar behavior as those for the harmonic<br />

potential. This implies that the universal behavior of the end-to-end distance is<br />

not sensitive to the specific details (e.g. channel shape) of the conf<strong>in</strong><strong>in</strong>g geometry<br />

used. As the key features of the conf<strong>in</strong><strong>in</strong>g potentials we have identified the collision<br />

parameter characteriz<strong>in</strong>g the lateral extension of the channel and, of course,<br />

the overall symmetry of the channel.<br />

There are also some limits of our <strong>in</strong>vestigations, clearly seen <strong>in</strong> the deviations<br />

of the simulation data from the experimental results <strong>in</strong> the weak to <strong>in</strong>termediate<br />

conf<strong>in</strong>ement regime for long semi-flexible polymers. This h<strong>in</strong>ts towards<br />

additional physical effects not conta<strong>in</strong>ed <strong>in</strong> the worm-like cha<strong>in</strong> model, e.g. the<br />

polyelectrolyte nature of DNA and result<strong>in</strong>g self-avoidance effects along the backbone.<br />

We were able to rule out that effects of simple excluded volume could<br />

expla<strong>in</strong> the deviations. Model<strong>in</strong>g electrostatic effects leads to the theory of polyelectrolytes,<br />

which ask for special sophisticated simulation techniques not easily<br />

implemented. The development of an effective algorithm to simulate polyelectrolytes<br />

<strong>in</strong> nanochannels is an <strong>in</strong>terest<strong>in</strong>g task for future projects.<br />

In addition, there is room for improv<strong>in</strong>g the algorithm by <strong>in</strong>troduc<strong>in</strong>g more<br />

complex Monte-Carlo moves. This would enable us to explore the behavior of<br />

cha<strong>in</strong>s with a larger number of segments or even different topologies of the conf<strong>in</strong><strong>in</strong>g<br />

environment.


Appendix A<br />

Discrete worm-like cha<strong>in</strong><br />

correlations<br />

For the cont<strong>in</strong>uum limit of the WLC we were able to relate the <strong>in</strong>teraction energy<br />

κ and the persistence length lp <strong>in</strong> eq. (2.20). Now we will try to obta<strong>in</strong> similar<br />

relations for ε and lp for the discrete WLC <strong>in</strong> two- and three-dimensional embedd<strong>in</strong>g<br />

space. The Hamiltonian for the discrete WLC is is given by the sum <strong>in</strong> eq.<br />

(2.13). Sett<strong>in</strong>g k = βεl2 , the configuration part of the partition sum is calculated<br />

by<br />

<br />

<br />

N<br />

N−1 <br />

Z(k) = dˆti exp k ˆti · ˆti+1 . (A.1)<br />

i=1<br />

To solve this <strong>in</strong> two dimensions we use the <strong>in</strong>extensibility constra<strong>in</strong>t and<br />

<strong>in</strong>troduce the angle between two successive segments as ˆti · ˆti+1 = cos θi. Then<br />

the <strong>in</strong>tegration variable changes dˆti → dθi<br />

Z2D(k) =<br />

+π<br />

= 2π<br />

dθN<br />

−π<br />

+π<br />

−π<br />

N−1<br />

i=1<br />

<br />

<br />

dθi exp k<br />

i=1<br />

dθ exp (k cos θ)<br />

N−1 <br />

i<br />

N−1 cos θi<br />

<br />

. (A.2)<br />

Transform<strong>in</strong>g this us<strong>in</strong>g [1, (9.6.19)] the <strong>in</strong>tegral is just a modified Bessel function<br />

of the first k<strong>in</strong>d I0(x)<br />

Z2D(k) = (2π) N I0(k) N−1 . (A.3)<br />

Introduc<strong>in</strong>g local stiffnesses ki <strong>in</strong>stead of a uniform stiffness k the calculation can<br />

be repeated to give<br />

N−1 <br />

βH({ki}, {ˆti}) = − ki ˆti · ˆti+1 ⇒ Z2D({ki}) = (2π) N<br />

i=1<br />

77<br />

N−1 <br />

i=1<br />

I0(ki). (A.4)


78 APPENDIX A. DISCRETE WORM-LIKE CHAIN CORRELATIONS<br />

The persistence length lp was def<strong>in</strong>ed through the spatial tangent-tangent<br />

correlation function <strong>in</strong> eq. (2.12), which is for the discrete model<br />

<br />

<br />

ˆtm · ˆtn = exp −|m − n| l<br />

<br />

. (A.5)<br />

To derive the relation between k and lp the correlation function is calculated<br />

us<strong>in</strong>g eq. (A.4)<br />

<br />

1 N<br />

<br />

ˆtm · ˆtn =<br />

dˆti ˆtm · ˆtn<br />

−βH({ki}) <br />

e <br />

Z({ki})<br />

<br />

i=1<br />

{ki=k}<br />

<br />

1 N<br />

<br />

=<br />

dˆti ˆtm · ˆtm+1<br />

ˆtm+1 · ˆtm+2 · · ·<br />

Z({ki})<br />

i=1<br />

· · · <br />

<br />

ˆtn−1 · ˆtn<br />

−βH({ki}) <br />

e <br />

<br />

{ki=k}<br />

1 ∂<br />

=<br />

Z({ki})<br />

|m−n| <br />

Z({ki}) <br />

<br />

∂km · · · ∂kn−1 {ki=k}<br />

(A.3)<br />

(A.4) 1<br />

=<br />

I0(k) N−1<br />

∂ |m−n|<br />

<br />

N−1 <br />

<br />

I0(ki) , (A.6)<br />

∂km · · · ∂kn−1 <br />

i=1 {ki=k}<br />

where <strong>in</strong> the second l<strong>in</strong>e <br />

ˆti · ˆti+2 = ˆti · ˆti+1<br />

ˆti+1 · ˆti+2 has been used, which<br />

can be seen by writ<strong>in</strong>g the product of tangents as the cos<strong>in</strong>es, us<strong>in</strong>g trigonometric<br />

theorems and the properties of the averages. For the 3D case below it is shown<br />

<strong>in</strong> [24, p. 377].<br />

Now us<strong>in</strong>g [1, (9.6.28)] <strong>in</strong> the form I ′ 0(k) = I1(k) gives<br />

|m−n|<br />

I1(k)<br />

ˆtm · ˆtn =<br />

I0(k)<br />

which results <strong>in</strong> the desired relation <strong>in</strong> 2D<br />

−1 I1(k)<br />

2D: lp = −l ln<br />

I0(k)<br />

lp<br />

!<br />

= (A.5), (A.7)<br />

. (A.8)<br />

For a given value of lp, k has to be determ<strong>in</strong>ed numerically, s<strong>in</strong>ce this formula is<br />

not <strong>in</strong>vertible <strong>in</strong> a closed form.<br />

A similar calculation, follow<strong>in</strong>g the above l<strong>in</strong>es, can be done for three dimension.<br />

By substitut<strong>in</strong>g dˆti = dφi dcos θi, we obta<strong>in</strong> for the partition function<br />

Z3D({ki}) = (4π) N<br />

N−1 <br />

i=1<br />

s<strong>in</strong>h ki<br />

, (A.9)<br />

ki


which, after tak<strong>in</strong>g all the derivatives of the partition sum and solv<strong>in</strong>g for the lp,<br />

results <strong>in</strong><br />

3D:<br />

<br />

lp = −l ln coth(k) − 1<br />

−1 ,<br />

k<br />

(A.10)<br />

for the desired relation. Some parts of the calculations are presented <strong>in</strong> [11] for<br />

the equivalent problem of the Heisenberg model for ferromagnets. This result<br />

must be <strong>in</strong>verted numerically as well.<br />

79


80 APPENDIX A. DISCRETE WORM-LIKE CHAIN CORRELATIONS


Appendix B<br />

Calculation for the conf<strong>in</strong>ed<br />

polymer<br />

B.1 Undulation correlations<br />

The Fourier back transform of the undulation correlation function, eq. (3.25),<br />

<strong>in</strong>to real space is the <strong>in</strong>tegral<br />

〈x(s)x(s ′ <br />

dk dk′<br />

)〉 = eiks<br />

2π 2π e−ik′ s ′<br />

〈˜x(k)˜x ∗ (k ′ )〉<br />

<br />

dk kBT e<br />

=<br />

2π κ<br />

ik(s−s′ )<br />

k4 + 4l −4 , (B.1)<br />

d <br />

<strong>in</strong>tegrat<strong>in</strong>g out the δ-function <strong>in</strong> eq. (3.25).<br />

For evaluation of the rema<strong>in</strong><strong>in</strong>g <strong>in</strong>tegral, we can use the complex analysis. We<br />

have to sum over the residues of the <strong>in</strong>tegral kernel, hence we need to f<strong>in</strong>d the<br />

poles of ˜ f(k). Expand<strong>in</strong>g the denom<strong>in</strong>ator, the zeros are found<br />

˜f(k)<br />

k 4 + 4l −4<br />

d = k 2 + i2l −2<br />

2 −2<br />

d k − i2ld <br />

= k − (1 − i)l −1<br />

−1<br />

d k + (1 − i)ld · k + (1 + i)l −1<br />

−1<br />

d k − (1 + i)ld . (B.2)<br />

These poles, as well as the path of <strong>in</strong>tegration, are shown <strong>in</strong> figure B.1. The<br />

<strong>in</strong>tegration over the real axis is done, for s > s ′ , by clos<strong>in</strong>g the contour <strong>in</strong> the<br />

upper half complex plane, where the exponential suppresses any contribution of<br />

the closed path.<br />

The Fourier transformation is then given by<br />

〈x(s)x(s ′ )〉 = kBT<br />

2πκ 2πi<br />

<br />

81<br />

Res ˜ f(k) + Res<br />

k=−1+i<br />

˜ f(k)<br />

k=1+i<br />

<br />

. (B.3)


82 APPENDIX B. CALCULATION FOR THE CONFINED POLYMER<br />

(−1 + i)l −1<br />

d<br />

(−1 − i)l −1<br />

d<br />

Im k<br />

(1 + i)l −1<br />

d<br />

(1 − i)l −1<br />

d<br />

path of <strong>in</strong>tegration<br />

for s > s ′<br />

Re k<br />

path of <strong>in</strong>tegration<br />

for s < s ′<br />

Figure B.1: Complex plane of the function ˜ f(k) with <strong>in</strong>dicated poles.<br />

Insert<strong>in</strong>g the residues, we obta<strong>in</strong> for s > s ′<br />

〈x(s)x(s ′ )〉 = kBT l3 ⎧ ⎫<br />

s − s′<br />

s − s′<br />

⎪⎨ exp i exp −i ⎪⎬ <br />

d ld<br />

ld<br />

s − s′<br />

+<br />

exp −<br />

8κ ⎪⎩<br />

1 + i<br />

1 − i ⎪⎭<br />

ld<br />

= kBT l3 d<br />

4 √ 2κ exp<br />

′<br />

s − s′ s − s<br />

− cos −<br />

ld<br />

ld<br />

π<br />

<br />

(B.4)<br />

4<br />

which—by additionally clos<strong>in</strong>g the contour for s < s ′ <strong>in</strong> the lower half plane—<br />

results <strong>in</strong><br />

〈x(s)x(s ′ )〉 = kBT l3 d<br />

4 √ 2κ exp<br />

<br />

− |s − s′ ′ | |s − s |<br />

cos −<br />

ld<br />

ld<br />

π<br />

<br />

. (B.5)<br />

4<br />

This is the result <strong>in</strong> eq. (3.26) used for further calculations.<br />

B.2 End-to-end distance correlation<br />

Here we present the details of the calculation for the mean-square end-to-end<br />

distance 〈R2 (L)〉. We start by calculat<strong>in</strong>g the distance between a po<strong>in</strong>t r(s) on<br />

the contour and the orig<strong>in</strong> of the polymer r(0), given by the <strong>in</strong>tegral<br />

R(s) = r(s) − r(0) =<br />

s<br />

0<br />

dτ t(τ). (B.6)


B.2. END-TO-END DISTANCE CORRELATION 83<br />

s ′<br />

τ ′<br />

£¤£¤£¤£¤£¤£<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢<br />

¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢<br />

£¤£¤£¤£¤£¤£<br />

£¤£¤£¤£¤£¤£<br />

£¤£¤£¤£¤£¤£<br />

¥¤¥¤¥¤¥¤¥¤¥<br />

¥¤¥¤¥¤¥¤¥¤¥<br />

¥¤¥¤¥¤¥¤¥¤¥ ¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

2<br />

¥¤¥¤¥¤¥¤¥¤¥ £¤£¤£¤£¤£¤£<br />

1<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢<br />

¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢<br />

£¤£¤£¤£¤£¤£<br />

£¤£¤£¤£¤£¤£<br />

£¤£¤£¤£¤£¤£<br />

¥¤¥¤¥¤¥¤¥¤¥<br />

¥¤¥¤¥¤¥¤¥¤¥<br />

¥¤¥¤¥¤¥¤¥¤¥ ¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

¥¤¥¤¥¤¥¤¥¤¥ £¤£¤£¤£¤£¤£<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢<br />

¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢¡¢<br />

£¤£¤£¤£¤£¤£<br />

£¤£¤£¤£¤£¤£<br />

£¤£¤£¤£¤£¤£<br />

¥¤¥¤¥¤¥¤¥¤¥<br />

¥¤¥¤¥¤¥¤¥¤¥<br />

¥¤¥¤¥¤¥¤¥¤¥ ¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡<br />

¥¤¥¤¥¤¥¤¥¤¥ £¤£¤£¤£¤£¤£<br />

§¤§¤§¤§¤§¤§¤§¤§<br />

¦¤¦¤¦¤¦¤¦¤¦¤¦¤¦<br />

Figure B.2: Integration ranges for the <strong>in</strong>tegral eq. (B.8).<br />

Us<strong>in</strong>g this, the end-to-end distance correlation function can be calculated by an<br />

<strong>in</strong>tegral over tangent-tangent correlation function, eq. (3.34),<br />

〈R(s) · R(s ′ s<br />

)〉 = dτ<br />

=<br />

=<br />

=<br />

s<br />

0<br />

s<br />

0<br />

0<br />

dτ<br />

dτ<br />

<br />

1 − ld<br />

2 lp<br />

<br />

3<br />

s ′<br />

dτ<br />

0<br />

′ t(τ) · t(τ ′ )<br />

s ′<br />

dτ<br />

0<br />

′ 〈t(τ) · t(τ ′ )〉<br />

s ′<br />

dτ<br />

0<br />

′<br />

<br />

1 − ld<br />

<br />

√2 −<br />

e<br />

2 lp<br />

|τ−τ′ <br />

|<br />

′ |τ − τ |<br />

ld s<strong>in</strong> −<br />

ld<br />

π<br />

<br />

+ 1<br />

4<br />

<br />

ss ′ − ld<br />

s s ′<br />

√2 dτ dτ<br />

lp 0 0<br />

′ K(|τ − τ ′ |). (B.7)<br />

The <strong>in</strong>tegral kernel has been abbreviated to K(x) = e −x/ld s<strong>in</strong> (x/ld − π/4). The<br />

<strong>in</strong>tegral over the absolute value <strong>in</strong> the last l<strong>in</strong>e has to be split. For s ≥ s ′ the<br />

<strong>in</strong>tegration range is shown <strong>in</strong> figure B.2 and is therefore calculated as<br />

s<br />

0<br />

dτ<br />

1<br />

<br />

s ′<br />

dτ<br />

0<br />

′ K(|τ − τ ′ s ′ τ<br />

|) = dτ dτ<br />

0 0<br />

′ K(τ − τ ′ s ′<br />

) +<br />

s<br />

0<br />

+<br />

s ′<br />

s ′<br />

dτ dτ<br />

0<br />

′ K(τ − τ ′ )<br />

s<br />

τ<br />

2<br />

<br />

dτ<br />

s ′<br />

dτ<br />

τ<br />

′ K(τ ′ − τ)<br />

, (B.8)<br />

<br />

3<br />

<br />

where the numbers denote the correspond<strong>in</strong>g areas <strong>in</strong> the figure. As an example


84 APPENDIX B. CALCULATION FOR THE CONFINED POLYMER<br />

the first <strong>in</strong>tegral is calculated here, the others are evaluated <strong>in</strong> the same way,<br />

s s ′<br />

s (τ−s ′ )/ld<br />

dτ<br />

dτ dx e −x <br />

s<strong>in</strong> x − π<br />

<br />

4<br />

s ′<br />

0<br />

dτ ′ K(τ − τ ′ ) = −ld<br />

s ′<br />

s<br />

= ld<br />

√2<br />

= l2 d √2<br />

= l2 d<br />

2<br />

dτ<br />

<br />

e<br />

τ/ld<br />

s ′<br />

(s−s ′ )/ld<br />

τ−s′<br />

− ld s<strong>in</strong><br />

0<br />

<br />

1<br />

√2 − e s−s′<br />

ld s<strong>in</strong><br />

s<br />

− l + e d s<strong>in</strong><br />

s<br />

ld<br />

τ − s′<br />

ld<br />

dy e −y s<strong>in</strong> y −<br />

τ<br />

− l − e d s<strong>in</strong> τ<br />

ld<br />

s ′ /ld<br />

s/ld<br />

′ s − s<br />

+<br />

ld<br />

π<br />

<br />

4<br />

′ s<br />

− π<br />

<br />

s′<br />

− l − e d s<strong>in</strong><br />

4<br />

ld<br />

<br />

dz e −z s<strong>in</strong> z<br />

<br />

− π<br />

<br />

. (B.9)<br />

4<br />

The trigonometric relations used can be found <strong>in</strong> [1]. Calculat<strong>in</strong>g the rema<strong>in</strong><strong>in</strong>g<br />

<strong>in</strong>tegrals and comb<strong>in</strong><strong>in</strong>g the results gives<br />

〈R(s) · R(s ′ <br />

)〉 = 1 − ld<br />

<br />

ss<br />

2 lp<br />

′ + l3 d<br />

2 √ <br />

e<br />

2 lp<br />

− |s−s′ <br />

|<br />

′ |s − s |<br />

ld s<strong>in</strong> +<br />

ld<br />

π<br />

<br />

4<br />

<br />

s<br />

− s<br />

l − e d s<strong>in</strong> +<br />

ld<br />

π<br />

<br />

s′ ′<br />

− s l − e d s<strong>in</strong> +<br />

4<br />

ld<br />

π<br />

<br />

+<br />

4<br />

1<br />

(B.10)<br />

√<br />

2<br />

valid for all s, s ′ .<br />

From this result the mean-square end-to-end distance is obta<strong>in</strong>ed <strong>in</strong> the l<strong>in</strong>es<br />

follow<strong>in</strong>g eq. (3.37).<br />

B.3 Parallel end-to-end distance correlation<br />

The mean-square end-to-end distance is calculated <strong>in</strong> detail, by <strong>in</strong>sert<strong>in</strong>g the<br />

derivatives of the average undulations eqs. (3.30) and (3.31) <strong>in</strong> eq. (3.40)<br />

˙R||(s) ˙ R||(s ′ ) = d2<br />

ds ds ′<br />

R||(s)R||(s ′ ) <br />

= 1 − ld<br />

+<br />

2 lp<br />

l2 d<br />

16 l2 p<br />

+ l2 d<br />

8 l2 e<br />

p<br />

−2 |s−s′ |<br />

ld s<strong>in</strong> 2<br />

′ |s − s |<br />

−<br />

ld<br />

π<br />

4<br />

<br />

=〈 ˙r ||(s) ˙r ||(s ′ )〉<br />

<br />

. (B.11)<br />

An <strong>in</strong>tegration for s ≥ s ′ yields the desired result for the projected length correlation<br />

<br />

R||(s)R||(s ′ ) s s ′<br />

= dτ dτ<br />

0 0<br />

′ R||(s) ˙ ˙ R||(s ′ ) <br />

<br />

= 1 − ld<br />

+<br />

2 lp<br />

l2 d<br />

16 l2 <br />

ss<br />

p<br />

′ + l2 d<br />

8 l2 s s ′<br />

dt dt<br />

p 0 0<br />

′ K(|τ − τ ′ |), (B.12)


B.3. PARALLEL END-TO-END DISTANCE CORRELATION 85<br />

with K(x) = e −2x/ld s<strong>in</strong> 2 (x/ld − π/4)). The <strong>in</strong>tegration over K(|τ − τ ′ |) has to<br />

be treated as <strong>in</strong> eq. (B.8).<br />

Evaluat<strong>in</strong>g only the first <strong>in</strong>tegral explicitly<br />

τ<br />

0<br />

dτ ′ K(τ − τ ′ ) = 1<br />

2<br />

τ<br />

dτ<br />

0<br />

′ τ−τ′<br />

−2 l e d<br />

0<br />

and perform<strong>in</strong>g the second <strong>in</strong>tegration gives<br />

s ′<br />

0<br />

dτ<br />

τ<br />

0<br />

dτ ′ K(τ − τ ′ ) = ld<br />

8<br />

<br />

1 − cos<br />

ld<br />

′ τ − τ<br />

2 −<br />

ld<br />

π<br />

<br />

4<br />

= − ld<br />

dx e<br />

4 2τ/ld<br />

−x (1 − s<strong>in</strong> x)<br />

= ld<br />

<br />

2τ<br />

− l 1 + e d −2 + s<strong>in</strong><br />

8<br />

2τ<br />

+ cos 2τ<br />

<br />

, (B.13)<br />

s ′ <br />

2τ<br />

− l dτ 1 + e d −2 + s<strong>in</strong><br />

0<br />

2τ<br />

+ cos<br />

ld<br />

2τ<br />

ld<br />

= l2 2s ′ /ld<br />

d<br />

dτ<br />

16 0<br />

1 − 2e −x + e −x (s<strong>in</strong> x + cos x) <br />

= l2 d<br />

16<br />

2s ′<br />

ld<br />

2s′<br />

− l + e d<br />

ld<br />

<br />

2 − cos 2s′<br />

ld<br />

<br />

ld<br />

<br />

2s′<br />

− l + e d 2 − cos 2s′<br />

ld<br />

<br />

<br />

− 1 . (B.14)<br />

The rema<strong>in</strong><strong>in</strong>g <strong>in</strong>tegrals are treated equivalently. F<strong>in</strong>ally we obta<strong>in</strong> the follow<strong>in</strong>g<br />

result valid for all s and s ′<br />

<br />

R||(s)R||(s ′ ) <br />

= 1 − ld<br />

<br />

ss<br />

2 lp<br />

′ + r||(s)r||(s ′ ) <br />

<br />

= 1 − ld<br />

+<br />

2 lp<br />

l2 d<br />

16 l2 <br />

ss<br />

p<br />

′ + l3 d<br />

32 l2 m<strong>in</strong>(s, s<br />

p<br />

′ )<br />

+ l4 d<br />

128 l2 <br />

− e<br />

p<br />

−2 |s−s′ <br />

|<br />

ld 2 − cos 2 |s − s′ <br />

|<br />

(B.15)<br />

ld<br />

<br />

2s<br />

− l + e d 2 − cos 2s<br />

<br />

<br />

− 1 .<br />

This result is used <strong>in</strong> the l<strong>in</strong>es follow<strong>in</strong>g eq. (3.41).


86 APPENDIX B. CALCULATION FOR THE CONFINED POLYMER


Glossary<br />

L filament length<br />

lp<br />

persistence length<br />

ɛ = L/lp<br />

flexibility<br />

ld<br />

Odijk deflection length<br />

c = L/ld<br />

collision parameter<br />

s arc length<br />

r(s) trajectory<br />

r⊥(s) = (x(s), y(s)) undulations<br />

r||(s) stored length<br />

ti, t(s)<br />

ˆti<br />

tangent, segment<br />

normalized tangent<br />

l = |ti| segment length<br />

φtt(s, s ′ ) = 〈t(s) · t(s ′ )〉 tangent-tangent correlation function<br />

b Kuhn segment length<br />

N number of segments<br />

v monomer volume<br />

β = (kBT ) −1 <strong>in</strong>verse temperature<br />

H Hamiltonian<br />

Z partition sum<br />

κ bend<strong>in</strong>g modulus<br />

ε <strong>in</strong>teraction energy<br />

d tube diameter (or dimension)<br />

w tube width<br />

γ harmonic potential strength<br />

R(s) distance to current position<br />

R(L) end-to-end distance<br />

p(R) radial distribution function (RDF)<br />

Rg(L) radius of gyrations<br />

R||(L), Ra(L) parallel, apparent end-to-end distance<br />

〈R〉, <br />

R|| , 〈Ra〉 mean (parallel, apparent) end-to-end distance<br />

〈R2 〉, R2 <br />

2<br />

|| , 〈Ra〉 mean-square (parallel, apparent) end-to-end dist.<br />

<br />

2 Rg mean-square radius of gyration<br />

〈x2 (s)〉 mean-square undulation<br />

87


88 APPENDIX B. CALCULATION FOR THE CONFINED POLYMER


Bibliography<br />

[1] M. Abramowitz and I. A. Stegun, editors. Pocketbook of Mathematical Functions.<br />

Verlag Harri Deutsch, 1984.<br />

[2] S. R. Aragón and R. Pecora. Dynamics of wormlike cha<strong>in</strong>s. Macromolecules,<br />

18:1868–1875, 1985.<br />

[3] R. Aust<strong>in</strong>. Nanopores: The art of suck<strong>in</strong>g spaghetti. Nature Materials,<br />

2:567–568, 2003.<br />

[4] K. B<strong>in</strong>der and D. W. Heermann. Monte Carlo Simulation <strong>in</strong> Statistical<br />

Physics: An Introduction. Spr<strong>in</strong>ger-Verlag Berl<strong>in</strong> Heidelberg, 2002.<br />

[5] C. Bustamante, Z. Bryant, and S. B. Smith. Ten years of tension: s<strong>in</strong>glemolecule<br />

DNA mechanics. Nature, 421:423–427, 2003.<br />

[6] C. Bustamante, J. C. Macosko, and G. J. Wuite. Grabb<strong>in</strong>g the cat by the<br />

tail: manipulat<strong>in</strong>g molecules one by one. Nature Reviews Molecular Cell<br />

Biology, 1:130–136, 2000.<br />

[7] H. E. Daniels. The statistical theory of stiff cha<strong>in</strong>s. Proceed<strong>in</strong>gs of the Royal<br />

Society of Ed<strong>in</strong>burgh, 63A:290–311, 1952.<br />

[8] P.-G. de Gennes. Scal<strong>in</strong>g Concepts <strong>in</strong> Polymer Physics. Cornell University<br />

Press, 1979.<br />

[9] M. Doi. Introduction to Polymer Physics. Clarendon Press, Oxford, 1996.<br />

[10] M. Doi and S. F. Edwards. The Theory of Polymer Dynamics. Clarendon<br />

Press, Oxford, 1986.<br />

[11] M. E. Fisher. Magnetism <strong>in</strong> one-dimensional systems — the Heisenberg<br />

model for <strong>in</strong>f<strong>in</strong>ite sp<strong>in</strong>. American Journal of Physics, 32:343–346, 1963.<br />

[12] H. Flyvbjerg and H. G. Petersen. Error estimates on averages of correlated<br />

data. The Journal of Chemical Physics, 91:461–466, 1989.<br />

[13] D. Frenkel and B. Smit. Understand<strong>in</strong>g Molecular Simulation, From Algorithms<br />

to Applications. Academic Press, 2002.<br />

89


90 BIBLIOGRAPHY<br />

[14] E. Frey. Physics <strong>in</strong> cell biology: On the physics of biopolymers and molecular<br />

motors. ChemPhysChem, 3:270–275, 2002.<br />

[15] E. Frey and T. Franosch. Statistische Physik — Biologische Physik und Soft<br />

Matter. Lecture notes (Freie Universität Berl<strong>in</strong>), WS 2002/03.<br />

[16] E. Frey and K. Kroy. Brownian motion: paradigm of soft matter & biological<br />

physics. Review submitted to Annalen der Physik, 2004.<br />

[17] E. Frey, K. Kroy, and J. Wilhelm. Viscoelasticity of biopolymer networks and<br />

statistical mechanics of semiflexible polymers. In S. Malhotra and J. Tuszynski,<br />

editors, Advances <strong>in</strong> Structural Biology, volume 5, pages 135–168. Jai<br />

Press, 1998.<br />

[18] R. Granek. From semi-flexible polymers to membranes: Anomalous diffusion<br />

and reptation. Journal de Physique II, 7:1761–1788, 1997.<br />

[19] H. Kle<strong>in</strong>ert. Path Integrals <strong>in</strong> Quantum Mechanics, Statistics, Polymer<br />

Physics and F<strong>in</strong>ancial Markets. World Scientific, third edition, 2004.<br />

[20] O. Kratky and G. Porod. Röntgenuntersuchung gelöster Fadenmoleküle.<br />

Recueil Travaux Chimiques (Pays Bas), 68:1106–1122, 1949.<br />

[21] K.-D. R. Kroy. Viskoelastizität von Lösungen halbsteifer Polymere. PhD<br />

thesis, Technische Universität München, 1998.<br />

[22] D. P. Landau and K. B<strong>in</strong>der. A Guide to Monte Carlo Simulations <strong>in</strong> Statistical<br />

Physics. Cambridge University Press, 2000.<br />

[23] L. D. Landau and E. M. Lifschitz. Lehrbuch der theoretischen Physik,<br />

Band VII, Elastizitätstheorie. Akademie Verlag Berl<strong>in</strong>, 1965.<br />

[24] L. D. Landau and E. M. Lifschitz. Lehrbuch der theoretischen Physik,<br />

Band V, Statistische Physik Teil 1. Akademie Verlag Berl<strong>in</strong>, 1979.<br />

[25] G. Lattanzi, T. Munk, and E. Frey. Transverse fluctuations of grafted polymers.<br />

Physical Review E, 69:021801, 2004.<br />

[26] M. le Bellac. Quantum and Statistical Field Theory. Oxford University Press,<br />

1991.<br />

[27] P. L’Ecuyer. Maximally equidistributed comb<strong>in</strong>ed Tausworthe generators.<br />

Mathematics of Computation, 65:203–213, 1996.<br />

[28] P. L’Ecuyer. Random Number Generation, chapter 4 of the Handbook on<br />

Simulation, Jerry Banks Ed., pages 93–137. Wiley, 1998.


BIBLIOGRAPHY 91<br />

[29] P. L’Ecuyer. Tables of maximally-equidistributed comb<strong>in</strong>ed LSFR generators.<br />

Mathematics of Computation, 68:261–269, 1999.<br />

[30] M. Lüscher. A portable high-quality random number generator for lattice<br />

field theory simulations. Computer Physics Communications, 79:100–110,<br />

1994.<br />

[31] F. C. MacK<strong>in</strong>tosh and C. F. Schmidt. Microrheology. Current Op<strong>in</strong>ions <strong>in</strong><br />

Colloid & Interface Science, 4:300–307, 1999.<br />

[32] B. Maier and J. O. Rädler. Conformation and self-diffusion of s<strong>in</strong>gle DNA<br />

molecules conf<strong>in</strong>ed to two dimensions. Physical Review Letters, 82:1911–<br />

1914, 1999.<br />

[33] J. F. Marko and E. D. Siggia. Stretch<strong>in</strong>g DNA. Macromolecules, 28:8759–<br />

8770, 1995.<br />

[34] N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, and<br />

E. Teller. Equation of state calculations by fast comput<strong>in</strong>g mach<strong>in</strong>es. The<br />

Journal of Chemical Physics, 21:1087–1092, 1953.<br />

[35] K. P. N. Murthy. Monte Carlo Methods <strong>in</strong> Statistical Physics. Universities<br />

Press (India) Private Ltd, 2004.<br />

[36] P. Nelson. Biological Physics: Energy, Information, Life. W. H. Freeman &<br />

Co, 2003.<br />

[37] T. Odijk. On the statistics and dynamics of conf<strong>in</strong>ed or entangled stiff<br />

polymers. Macromolecules, 16:1340–1344, 1983.<br />

[38] W. Reisner et al. Statics and dynamics of s<strong>in</strong>gle DNA molecules conf<strong>in</strong>ed <strong>in</strong><br />

nanochannels. Unpublished.<br />

[39] M. Rub<strong>in</strong>ste<strong>in</strong> and R. H. Colby. Polymer Physics. Oxford University Press,<br />

2003.<br />

[40] N. Saitô, K. Takahashi, and Y. Yunoki. The statistical mechanical theory of<br />

stiff cha<strong>in</strong>s. Journal of the Physical Society of Japan, 22:219–226, 1967.<br />

[41] D. W. Schaefer, J. F. Joanny, and P. P<strong>in</strong>cus. Dynamics of semiflexible<br />

polymers <strong>in</strong> solution. Macromolecules, 13:1280–1289, 1980.<br />

[42] P. J. Schneider and D. H. Eberly. Geometric Tools for Computer Graphics.<br />

Morgan Kaufmann Publishers, 2002.<br />

[43] C.-Y. Shew. Conformational behavior of a s<strong>in</strong>gle polymer cha<strong>in</strong> conf<strong>in</strong>ed<br />

by a two-dimensional harmonic potential <strong>in</strong> good solvents. The Journal of<br />

Chemical Physics, 119:10428–10437, 2003.


92 BIBLIOGRAPHY<br />

[44] J. O. Tegenfeldt et al. The dynamics of genomic-length DNA molecules<br />

<strong>in</strong> 100-nm channels. Proceed<strong>in</strong>gs of the National Academy of Sciences,<br />

101:10979–10983, 2004.<br />

[45] N. van Kampen. Stochastic Processes <strong>in</strong> Physics and Chemistry. North-<br />

Holland Personal Library, Elsevier Science B.V., 2001.<br />

[46] C. H. Wigg<strong>in</strong>s et al. Trapp<strong>in</strong>g and wiggl<strong>in</strong>g: Elastohydrodynamics of driven<br />

microfilaments. Biophysical Journal, 74:1043–1060, 1998.<br />

[47] J. Wilhelm and E. Frey. Radial distribution function of semiflexible polymers.<br />

Physical Review Letters, 77:2581–2584, 1996.<br />

[48] H. Yamakawa and M. Fujii. Wormlike cha<strong>in</strong>s near the rod limit: Path <strong>in</strong>tegral<br />

<strong>in</strong> the WKB approximation. The Journal of Chemical Physics, 59:6641–6644,<br />

1973.


:wq

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!