31.10.2014 Views

(DPPC) vesicles - dirac

(DPPC) vesicles - dirac

(DPPC) vesicles - dirac

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

A characterization of<br />

the alcohol-induced structural changes of<br />

unilamellar dipalmitoyl-phosphatidylcholine<br />

(<strong>DPPC</strong>) <strong>vesicles</strong><br />

• UNIVERSITAS ROSKILDENSIS •<br />

IN TRANQUILLO MORS • IN FLUCTU VITA<br />

Bitten Plesner and Thomas Hecksher<br />

Physics supervisor: Dorthe Posselt<br />

Chemistry supervisor: Peter Westh<br />

Master Thesis Dissertation<br />

Roskilde University, Denmark<br />

February 2007


Abstract<br />

Title: “A characterization of the alcohol-induced structural changes of unilamellar<br />

dipalmitoyl-phosphatidylcholine (<strong>DPPC</strong>) <strong>vesicles</strong>”<br />

The structural changes of unilamellar dipalmitoylphosphatidylcholine (<strong>DPPC</strong>) <strong>vesicles</strong><br />

in excess water (with a radius of about 60 nm ± 20%) induced by pentanol, hexanol<br />

and heptanol were studied. The working hypothesis was that the alcohol would saturate<br />

the lipid <strong>vesicles</strong>, making the system go from the L β ′ phase to the L βI phase. The<br />

point of saturation and partition coefficient could then be found by isothermal titration<br />

calorimetry (ITC). The transition temperature between the L βI phase and L α phase could<br />

be determined by differential scanning calorimetry (DSC) and the structural differences<br />

of these two phases, notably the bilayer thickness, could be studied by small angle X-ray<br />

scattering (SAXS). The interdigitated phase of lipid <strong>vesicles</strong> is well-documented with a<br />

lot of different inducers, but the combination of <strong>DPPC</strong> and pentanol, hexanol or heptanol<br />

has until now been unexplored (by SAXS).<br />

The partition coefficients were all lower, but still at the same order of magnitude as the<br />

values stated in the literature. The transition temperature was found to decrease as<br />

alcohol was added to the lipid suspension. The phase transition temperature decreased<br />

from 41 ◦ C (pure <strong>DPPC</strong>) to (34 ± 1) ◦ C for <strong>DPPC</strong> and pentanol, (29 ± 1) ◦ C for <strong>DPPC</strong><br />

and hexanol and (35 ± 1) ◦ C for <strong>DPPC</strong> and heptanol. The SAXS spectra modelling<br />

analysis revealed nearly the same values for the lipid bilayer thicknesses for all of the<br />

measurements in the order of 34-38 Å. Adding alcohol mainly seemed to decrease the<br />

bilayer thickness slightly independent of the temperature. And for all alcohols there are<br />

slight tendencies towards a lower bilayer thickness as the temperature is increased. With<br />

respect to the presence of interdigitation our results are inconclusive, we can neither<br />

reject nor confirm for any of the alcohols that our low temperature measurements are in<br />

fact in the L βI phase. According to the literature and our ITC measurements we should<br />

have enough alcohol to induce the phase, but the results from the modelling on the SAXS<br />

data do not allow such a clear-cut conclusion.<br />

The system behaved more complex than initially expected. The DSC measurements<br />

indicate a stable system with a well-defined phase transition, but the ITC and the SAXS<br />

measurements did not. It is still unclear as to why the samples are very sensitive to the<br />

thermal history after adding the alcohol. But using our sample preparation method, we<br />

do get systems which give reproducible SAXS-spectra. The modelling of these spectra<br />

did not completely account for all the features seen in the spectra. They indicate that a<br />

large fraction of the system still has vesicle structure and a smaller fraction of the system<br />

form larger crystal-like aggregates, whose structures are yet unknown to us.


Resumé<br />

Titel: “Karakterisering af alkoholinducerede strukturelle ændringer i unilamellare dipalmitoylfosfatidylcholine<br />

(<strong>DPPC</strong>) vesikler”<br />

Den strukturelle ændring af unilamellare dipalmitoylfosfatidylcholine (<strong>DPPC</strong>) vesikler,<br />

med en radius omkring 50 nm, tilsat pentanol, hexanol eller heptanol er undersøgt. Arbejdshypotesen<br />

var at alkoholen ville mætte vesiklerne og derved få systemet til at gå<br />

fra L β ′-fasen til L βI -fasen. Mætningspunktet og partitionskoefficienten kunne findes med<br />

isotermisk titreringskalorimetri (ITC). Faseovergangstemperaturen mellem L βI -fasen og<br />

L α -fasen bestemmes med differentiel scanningskalorimetri (DSC) og den strukturelle<br />

forskel mellem disse to faser, især bilagstykkelsen, er undersøgt med småvinkelrøntgenspredning<br />

(SAXS). Den interdigiterede fase af lipidvesiklerne er veldokumenteret med<br />

en række forskellige inducere, men kombinationen af <strong>DPPC</strong> og pentanol, hexanol eller<br />

heptanol har hidtil været uudforsket (med SAXS).<br />

Partitionskoefficienterne var alle lavere, omend af samme størrelsesorden, end litteraturværdierne.<br />

Faseovergangstemperaturen faldt fra 41 ◦ C (ren <strong>DPPC</strong>) til (34 ± 1) ◦ C<br />

for <strong>DPPC</strong> og pentanol, (29 ± 1) ◦ C for <strong>DPPC</strong> og hexanol og (35 ± 1) ◦ C for <strong>DPPC</strong> og<br />

heptanol. Analysen af SAXS modeleringen afslørede ensartede bilagstykkelser for alle<br />

målinger i størrelsesordenen 34-38 Å. Tendensen ved tilføjelse af alkohol var en anelse<br />

tyndere bilagstykkelse uafhængigt af temperaturen. Og for alle alkoholer var tendensen<br />

en anelse tyndere bilagstykkelse når temperaturen steg. Vores resultater er ikke entydige<br />

i forhold til tilstedeværelsen af interdigitering, vi kan, for ingen af de benyttede alkoholer,<br />

hverken be- eller afkræfte at lavtemperaturmålingerne rent faktisk finder sted i<br />

den interdigiterede fase. På baggrund af litteraturværdier og ITC-målingerne burde der<br />

være nok alkohol til stede til at inducere fasen, men resultaterne fra modelleringen af<br />

SAXS-dataene lægger ikke op til en entydig konklusion.<br />

Systemet opførte sig mere komplekst end først antaget. DSC-målingerne indikerede et<br />

stabilt system med en veldefineret faseovergang, mens ITC- og SAXS-målingerne indikerede<br />

et mere ustabilt system. Det er stadig uklart, hvorfor prøverne er så følsomme<br />

overfor den termiske historie efter alkoholtilsætning. De rå spektrer indikerer at en stor<br />

del af systemet har vesikelstruktur og en mindre andel af systemet danner store krystallignende<br />

aggregater, hvis struktur endnu er ukendt for os.


Acknowledgements<br />

A number of people deserves acknowledgements for their assistance during the process<br />

of making this thesis.<br />

First of all the supervisors; Associate Professor of physics Dorthe Posselt for patient<br />

guidance, instruction to making the SAXS measurements and helpful discussions on the<br />

experimental work, and Professor Peter Westh for rewarding supervision, fruitful discussions<br />

on the lipid-alcohol system and the interpretation of the ITC experiments as well<br />

as helpful instructions on the DSC and ITC experiments.<br />

We would also like to thank a number of people at and around IMFUFA:<br />

Laboratorian Oda Brandstrup - for her helpfulness in the X-ray laboratory.<br />

Engineer Ib Høst Pedersen - for rapidly fixing miscellaneous lab problems and manufacturing<br />

a special device for the cuvette in order to make the temperature calibration.<br />

Assistant engineer Ebbe Hyldahl Larsen - for manufacturing a sample holder for a<br />

heater/shaker, fixing a leakage in the cuvette as well as a number of other matters.<br />

Assistant engineer Torben Steen Rasmussen - for helping with the Kratky camera set-up.<br />

IT guru Heine Larsen - for linux help and miscellaneous IT help in general.<br />

Ph.D. student Ulf Rørbæk Pedersen - for helping with the Matlab programs and sample<br />

preparation procedure and for numerous discussions about the lipid-alcohol system.<br />

Cand. scient, Ph.D. Christa Trandum for very rewarding discussions on lipid-alcohol<br />

interactions and the use of ITC as well as DSC.<br />

Cand. scient, Ph.D. Brian Igarashi - for helping with the introductory DSC experiments.<br />

Cand. mag. Casper Gaarde Madsen - for English proof reading.<br />

Cand. scient, Ph.D. Kristine Niss - Reading and comments.<br />

Cand. scient Gitte Margrethe Jensen - Reading and comments.<br />

Ph.D. student Bjarke Skipper Petersen - for sharing food, beer, being good company<br />

and participating in numerous jam-sessions.<br />

The computer lab.: Ditte Gundermann, Neslihan Sağlanmak, Martin Graversgaard<br />

Nielsen and others - for being good company and for many cookies, cakes, candy and<br />

fruit (not knowing they shared).<br />

Software used<br />

Matlab [48] (KDE linux) - data processing, modelling.<br />

Kile [35] (KDE linux) - a L A TEXfrontend.<br />

Basic Support for Cooperative Work (BSCW) [4] - online version control of shared documents.<br />

Grace [19] (KDE linux) - a 2d graph plotting tool.<br />

Xfig [86] (KDE linux) - a drawing tool.<br />

GNU Image Manipulation Program (GIMP) [18] (KDE linux, Windows).<br />

Roskilde, February 19th 2007<br />

Thomas Hecksher and Bitten Plesner.


Reading guidelines<br />

The authors of this report are evaluated on different premises. For Bitten this is an<br />

inter-disciplinary chemistry and physics thesis. For Thomas this is just a physics thesis.<br />

Certain parts of the report are therefore dedicated as Bitten’s chemistry thesis and are<br />

solely the work of Bitten. This applies to section 3.1, 4.1 and appendix E. The rest of<br />

the report is a result of the joint effort of Bitten and Thomas.<br />

The reader is urged to read the The course of events appendix at page 111, but maybe<br />

not before reading the report. We chose to write this course of events primarily for the<br />

purpose of other fellow students to make use of our experiences and reflections during<br />

the work of this thesis.<br />

If the reader is familiar with the experimental techinques isothermal titration calorimetry,<br />

differential scanning calorimetry and small angle X-ray scattering, the sections 3.1, 3.2<br />

and 3.3 can be skipped.<br />

Throughout chapters 4 and 5 references will be made to the Matlab source code appendix<br />

showing the actual implementations of the models.<br />

Throughout the report we use the phrases low temperature and high temperature referring<br />

to measurements at 25 ◦ C and 45 ◦ C, repectively.<br />

Certain abbreviations are used over and over again throughout the report. These include<br />

isothermal titration calorimetry (ITC), differential scanning calorimetry (DSC), small angle<br />

X-ray scattering (SAXS), multi lamellar vesicle (MLV), unilamellar vesicle (ULV), dipalmitoylphosphatidylcholine<br />

(<strong>DPPC</strong>), dimyristoylphosphatidylcholine (DMPC). Equally,<br />

references to the various phases of phospholipid <strong>vesicles</strong> are made constantly. The abbreviations<br />

for the frequently referred phases are indicated in table 1.<br />

Abbreviation<br />

L α<br />

L βI<br />

L β ′<br />

P β ′<br />

Phase<br />

Liquid crystal<br />

Interdigitated<br />

Gel<br />

Rippled (gel)<br />

Table 1 Some abbreviations of the phases of phospholipid <strong>vesicles</strong> perturbated by n-alcohols.<br />

Most of the experimental results obtained in this thesis are placed in the appendices due<br />

to the readability of the thesis, not because the experimental work should be put in the<br />

background.<br />

The concentrations are stated differently according to the given system. The concentration<br />

of lipid is stated as a weight ratio, m lipid/ mwater , and the concentration of alcohol in<br />

a lipid suspension is stated as a mole fraction, n lipid/ nalcohol . Here n alcohol refers to total<br />

amount of alcohol in the system, i.e. inside the bilayer and outside in the bulk water.<br />

Otherwise the concentrations are stated in mol per litre M.<br />

The target group of this thesis is primarily physics students and chemistry students with<br />

a special interest in biophysics and biochemistry.


Contents<br />

Contents<br />

vii<br />

I Introduction and Theory 1<br />

1 Introduction 3<br />

1.1 Phospolipids and membranes . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.2 Phases and properties of the <strong>vesicles</strong> . . . . . . . . . . . . . . . . . . . . . 5<br />

1.3 Additives - interactions with small amphiphilic molecules . . . . . . . . . 7<br />

1.4 Interdigitated phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.5 Aim of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.6 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2 Scattering Theory 17<br />

2.1 X-rays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2 The classical (particle) scattering experiment . . . . . . . . . . . . . . . . 20<br />

2.3 X-Ray scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.4 Bragg’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

II Experiments and Results 31<br />

3 Experimental Techniques 33<br />

3.1 ITC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.2 DSC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.3 SAXS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3.4 Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

4 Results 55<br />

4.1 ITC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

4.2 DSC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

4.3 SAXS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

III Analysis 75<br />

5 Modelling 77<br />

5.1 Overview of the models used . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

5.2 Instrumental smearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

5.3 Symmetric (1d) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

vii


viii<br />

Contents<br />

5.4 Symmetric (4g) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

5.5 Asymmetric (1d) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.6 Spherically symmetric (3d) . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

5.7 Multilamellar <strong>vesicles</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.8 Crystalline structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

5.9 Results and summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

6 Discussion 99<br />

7 Conclusion 103<br />

Bibliography 105<br />

A The course of events 111<br />

B Theory 115<br />

C Fourier transform 117<br />

D Experimental series 121<br />

E ITC 123<br />

F DSC 129<br />

G SAXS 133<br />

H Modelling results 135<br />

H.1 Pure <strong>DPPC</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136<br />

H.2 Pentanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

H.3 Hexanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144<br />

H.4 Heptanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

I Matlab source code 157<br />

Index 183


Part I<br />

Introduction and Theory


1 Introduction<br />

The behaviour and characterization of cell membranes are ongoing topics in the reseach<br />

field of biochemistry and biophysics. The cell membrane consists primarily of proteins<br />

and lipids but it also contains a variety of other biological molecules. Among other<br />

purposes the membrane serves to separate the intracellular components from the extracellular<br />

environment and regulate the exits and entries of the cell. The intra membrane<br />

morphology has an effect on the transmembrane processes and hence molecules influencing<br />

the membrane morphology and their mutual interactions are of great interest.<br />

Unilamellar phospholipid membranes are used as a model system for ordinary cell membranes<br />

in vivo, which aside from phospholipids also consist of different kinds of lipids,<br />

proteins and cholesterol.<br />

1.1 Phospolipids and membranes<br />

A lipid consists of a hydrophilic polar head group, and a hydrophobic nonpolar tail. The<br />

polar head group can be either charged or neutral (zwitterion or without charge), and<br />

the tail consists of one or two hydrocarbon chains. Lipids are poorly soluble in water but<br />

they are highly soluble in organic solvents. Most phospholipids are derived from glycerol,<br />

a three-carbon alcohol, and consist of a glycerol backbone, two fatty acid chains, and<br />

phosphorylated alcohol - these phospholipids are called phosphoglycerides, see figure 1.1.<br />

The natural fatty acid chains usually contain an even number of carbon atoms, in animal<br />

cells most frequently 14-24 [17], and may be saturated or unsaturated. The length and<br />

degree of unsaturation of the chains have a strong effect on membrane fluidity.<br />

R<br />

PHOSPHAT<br />

G<br />

L<br />

Y<br />

C<br />

E<br />

R<br />

O<br />

L<br />

FATTY ACID<br />

FATTY ACID<br />

Figure 1.1 The basic structure of a phospholipid. It consists of a phosphat group to which a<br />

chemically functional group, R, is bound, a glycerol group and two identical or two different<br />

fatty acids. R can e.g. represent choline, ethanolamine or serine.<br />

When the phospholipids are dissolved in water their water-insoluble and primarily<br />

hydrophobic nature forces them to assemble rapidly [66]. Those biological phospholipids<br />

which have two long alkyl chains per molecule will primarily aggregate to form bilayers,<br />

they do seldom form spherical micelles 1 . In excess water the micelles are dispersed<br />

1 The nonpolar portion of the molecule has a fixed molecular volume and a maximal length, this will<br />

3


4 Introduction<br />

(c w >40%) and the lipid bilayers are said to be fully hydrated.<br />

The hydrophobic force is the most significant thermodynamically driven force. This<br />

so-called force arises from the unfavourable constraints placed on water as it packs around<br />

a nonpolar hydrocarbon. The force drives the system to adopt a configuration which<br />

minimizes the contact between water and the nonpolar portions of the lipid. The van der<br />

Waals force which represents short, weak attractive forces between adjacent hydrocarbon<br />

chains is another stabilizing factor. The attraction results from interactions between<br />

polarizable electrons (induced dipoles) and compared to the hydrophobic force, the van<br />

der Waals force is a relatively minor stabilizing factor [17]. Forming a bilayer has different<br />

advantages including the possibility of forming a spherical bilayer vesicle which compared<br />

to disks or flat sheets is highly favourable due to the lacking contact by water at the edge.<br />

By forming a vesicle the edge is eliminated, the sphere is also favoured by being smaller<br />

and, hence, entropically preferable to very large bilayer sheets. The bilayers most often<br />

form multilammelar “onion-layered-shaped” <strong>vesicles</strong> with a certain repeat distance. A<br />

single phospholipid membrane consists of a lipid bilayer where the hydrophilic part of the<br />

phospholipids is situated towards the water, and the hydrophobic part, the acyl chains,<br />

is inside the membrane as shown in figure 1.2. The drawing in the lower right corner of<br />

figure 1.2 illustrates a single bilayer. The head group of the lipid is shown as circles, the<br />

tails as curved lines. The grey lamella represents the hydrophobic part and the white<br />

lamella represents the hydrophilic part.<br />

Multilamellar vesicle<br />

Unilamellar vesicle<br />

100 nm<br />

Bilayer normal<br />

5 nm<br />

Hydrophilic<br />

Hydrophobic<br />

Hydrophilic<br />

Figure 1.2 A multilamellar vesicle (MLV) and a unilamellar vesicle (ULV). The multilamellar<br />

vesicle consists of several bilayers. The unilamellar vesicle consists of only one bilayer. The<br />

drawing in the lower right corner illustrates a single bilayer. The head group of the lipid is<br />

shown as circles, the tails as curved lines. The grey lamella represents the hydrophobic part<br />

and the white lamella represents the hydrophilic part. The denoted lengths can vary, but the<br />

values for the bilayer thickness and vesicle diameter denoted here are characteristic for the ULV<br />

<strong>DPPC</strong> <strong>vesicles</strong> used in this project.<br />

A multilamellar vesicle (MLV) can contain up to a thousand bilayers [32]. A unilamellar<br />

vesicle (ULV) consists, as the name reveals, of only a single bilayer and is frequently<br />

called a liposome.<br />

determine the maximal radius of a spherical micelle as well as the number of molecules that can fit<br />

into the micelle. The optimal surface area required by the polar headgroup must also be considered.<br />

For biological phospholipids the area per molecule in a hypothetical spherical micelle is much larger<br />

that the optimal value for headgroup packing and, hence, these amphiphiles do not form stable<br />

spherical micelles [17].


1.2 Phases and properties of the <strong>vesicles</strong> 5<br />

A lot of work has been done on different phospholipids and their response to changes<br />

in temperature, pressure, concentration and chemical solution [5], [14], [38], [53], [70],<br />

[78].<br />

In this work we use the phospholipid di-palmitoyl-sn-phosphatidyl-choline, <strong>DPPC</strong>.<br />

This phospholipid consists of two identical fatty acids, each the length of 16 C-atoms, a<br />

phosphatidyl group and a choline group. The chemical structure of <strong>DPPC</strong> is shown in<br />

figure 1.3.<br />

CH 3<br />

CH 2<br />

H 2 C CH 3<br />

H 2 C<br />

O<br />

C<br />

O<br />

H 2 C<br />

H 2 C<br />

C<br />

O<br />

H<br />

CCH 2<br />

O<br />

O<br />

H 2 C<br />

O<br />

P<br />

O<br />

O<br />

H 3 C<br />

H 3 C<br />

N<br />

CH 2<br />

CH 3<br />

Figure 1.3 <strong>DPPC</strong> is chemically known as 1,2-dipalmitoyl-sn-glycero-3-phosphocholine or 1,2-<br />

dihexadecanoyl-sn-glycero-3-phosphocholine. The numbers 1, 2 and 3 refer to the position<br />

of the fatty acids and the phosphatidyl-choline group on the glycerol. In this case the two<br />

identical fatty acids are placed next to each other and the phosphatidyl-choline group is placed<br />

at the end of the glycerol. The nitrogen is positively charged, whereas one of the oxygens of the<br />

phosphat is negatively charged. This makes the headgroup of the phospholipid a zwitterion.<br />

1.2 Phases and properties of the <strong>vesicles</strong><br />

All lipid bilayers show a first-order phase transition associated with an increased conformational<br />

freedom for the lipid fatty acyl chain [42]. This process is designated to the<br />

”melting” of the hydrocarbon chains. The chain melting temperature differs from phospholipid<br />

to phospholipid and is determined by a balance of competing factors. Below<br />

this phase transition temperature, the membrane is gel like, and above the transition<br />

temperature, the membrane is fluid, liquid crystalline like. The gel phase is denoted L β ′<br />

and the liquid crystalline phase is denoted L α . The phase most relevant for comparison<br />

with biological systems is the L α phase [12], [75].


6 Introduction<br />

Tristram-Nagle & Nagle [78] have done a great deal of work on these phospholipid<br />

systems including phase transitions. For <strong>DPPC</strong> they find that the largest phase transition<br />

enthalpy of MLV is the transition between the L β ′ and L α phase - ∆H ∼ 36 kJ/mol.<br />

Less than half of that amount is due to disordering of the hydrocarbon chains. The rest<br />

can be accounted for by volume expansion as one can see in figure 1.4.<br />

Figure 1.4 A graph showing the volume per lipid (<strong>DPPC</strong>) vs. temperature for <strong>DPPC</strong> bilayers<br />

in excess water (experimental data). [32]<br />

Besides the above mentioned phases the system can also exist in the low temperature<br />

subgel L c phase, and a phase between the gel and liquid phase called ripple phase,<br />

denoted P β ′. Therefore two other phase transitions (in the temperature range 0-60 ◦ C)<br />

exists, the subgel-gel, and the gel-ripple transitions, the latter is also called pre-transition.<br />

The subgel-gel transition exhibits hysteresis, but apparently has a ”true” transition temperature<br />

at 18 ◦ C for <strong>DPPC</strong> [42].<br />

The disordered chains in the L α phase are characterized by gauche conformations,<br />

which are favoured entropically compared to the highly ordered all-trans chain conformation<br />

in the gel state. The gauche conformations are characterized by being energetically<br />

unfavourable compared to the trans conformation, where the CH 2 -groups in the acyl<br />

chain are situated as far away from each other as possible [13], [17]. The two conformations<br />

are illustrated by a Newman projection 2 in figure 1.5.<br />

Numerous techniques indicate that in the gel phase, the hydrocarbon chains of saturated<br />

diacyl phospholipids are predominantly in the all-trans conformation [17].<br />

Sackmann refers to experiments which have shown that the main transition is characterized<br />

by an abrupt increase in the number of gauche conformations, n g . The concentration<br />

of gauche conformations is defined as x g ≡ ng<br />

n , where n is the number of CH 2<br />

segments in the chain. The concentration increases from x g = 0.07 to x g = 0.4 upon<br />

going to the L α phase [66].<br />

In the gel phase the acyl chains of <strong>DPPC</strong> are tilted about 30 ◦ with respect to the<br />

bilayer normal, effectively increasing their cross-sectional area to be compatible with that<br />

of the headgroup; the chains remain in the all-trans configuration. The tilt is a result of<br />

2 The Newman projection views a carbon carbon chemical bond from front to back, front carbon as<br />

a dot and back carbon as a circle. This type of representation makes it easy to assess the torsional<br />

angle between two bonds, one at each carbon atom.


1.3 Additives - interactions with small amphiphilic molecules 7<br />

H<br />

CH 3 CH3 H<br />

CH 3<br />

H<br />

CH 3<br />

CH 3<br />

H<br />

H<br />

H H<br />

CH 3<br />

H H<br />

H<br />

g + = t= g − =<br />

H<br />

H<br />

Figure 1.5 The Newman projection diagram of the gauche conformations, g + and g − , and trans<br />

conformation, t. Adjusted from Gennis [17].<br />

the rather large headgroup of the <strong>DPPC</strong> molecule [66]. The headgroup of <strong>DPPC</strong> requires<br />

a minimum packing of 50 Å 2 . In the L α phase the introduction of gauche configurations<br />

increase the effective chain cross section to about 60-70 Å 2 [17]. Katsaras & Gutberlet<br />

[32] report the interfacial area of <strong>DPPC</strong> in the L α phase to be 68-71Å 2 . Hence, in the L α<br />

phase, the hydrocarbon chains are not tilted. 2 H NMR indicates that the thickness of<br />

the hydrocarbon domain of <strong>DPPC</strong> is 35 Å in the L α phase, compared to 45 Å expected<br />

if the chains were all-trans and oriented along the bilayer normal. The shorter distance<br />

is due to gauche configuration and the chains are basically aligned and perpendicular to<br />

the plane of the bilayer and are not coiled [17].<br />

Bilayer thickness<br />

A common definition of bilayer thickness is D B = 2VL<br />

A , where V L is the volume of a lipid<br />

molecule in the bilayer and A is the interfacial area per lipid; V L has been measured<br />

accurately (relative uncertainty of 0.2%) by a number of groups [54]. The thickness<br />

of lipid bilayers (vide infra for a discussion on various definitions of thickness) is an<br />

important structural quantity for discussing the incorporation of intrinsic membrane<br />

proteins [32]. As mentioned the bilayer membranes are organized assemblies with a<br />

sheet-like structure, they typically have a thickness about 5 nm as shown in figure 1.6.<br />

There is a large uncertainty in the interfacial area A per lipid (<strong>DPPC</strong>), different methods<br />

yield different results. Neutron scattering suggests A = 58Å 2 , and X-ray scattering<br />

suggests A = 71Å 2 [32], however most of the experiments referred in the literature point<br />

towards a value of about 65Å 2 [53].<br />

1.3 Additives - interactions with small amphiphilic molecules<br />

Generally, the molecules added to the bilayer system can be categorized by being entirely<br />

hydrophilic (sugars, salts, DNA, polyethylene glycol), mostly hydrophobic (steroids like<br />

cholesterol and androsten) or amphiphilic (fatty acids, alcohols). It turns out that additives<br />

”which are hydrophilic or amphiphilic will affect the phase behaviour in most<br />

dramatic ways”.[32, page 165]<br />

The literature shows several examples of phospholipids and their interactions with<br />

small amphiphilic molecules under different thermodynamical circumstances [10], [43],<br />

[47], [49], [50], [63], [68], [70], [80], [83], [84]. An example of an amphiphilic molecule<br />

is alcohol. The role of phospholipids and their interactions with especially alcohols has<br />

been of great interest due to the fact that alcohol has an anaesthetic effect [16].


8 Introduction<br />

V L<br />

∼ 5 nm D B = 2·VL<br />

A<br />

V L<br />

Figure 1.6 The bilayer thickness of a unilamellar vesicle. V L is the volume of one phospholipid<br />

and D B represents the lipid bilayer thickness. A is the interfacial area per. <strong>DPPC</strong>. The value<br />

of V L has been measured very accurately, but there is a great uncertainty in the determination<br />

of a value for A. The value varies from 58 Å 2 , measured by neutron diffraction, and 71 Å 2 ,<br />

measured by X-ray scattering [32].<br />

Alcohols<br />

The chemical term alcohol covers any organic compound in which a hydroxyl group, the<br />

hydrogen-oxygen group, is bound to a carbon atom, which in turn is bound to other<br />

hydrogen and/or carbon atoms. A sub-group of alcohols is the n-alcohols, which consist<br />

of the hydroxyl group and an acyl chain. Hexanol is an example of an n-alcohol, see<br />

figure 1.7.<br />

HO<br />

H 2 C<br />

H 2 C<br />

H 3 C<br />

Figure 1.7 The chemical structure of the n-alcohol hexanol. Hexanol consists of an hydroxyl<br />

group and a 6-carbon atom chain.<br />

CH 2<br />

CH 2<br />

CH 2<br />

The alcohol can be described as amphiphilic. Because of its dipol the OH -group<br />

is characterized as hydrophilic and the nonpolar properties of the acyl chain account<br />

for its hydrophobia. The alcohol gets more hydrophobic as the acyl chain length increases.<br />

When pertubating the phospholipid membrane the alcohols are arranged with<br />

the hydrophobic group towards the inner part of the membrane and the hydrophilic part<br />

towards the bulk [39].


1.4 Interdigitated phase 9<br />

1.4 Interdigitated phase<br />

At high temperatures the liquid L α phase is stable. At low temperatures, various structures<br />

are found, dependent on temperature and alcohol concentration. At low concentrations,<br />

we find the transition from the highly ordered subgel phase, L c , via the gel phase,<br />

L β ′, to the rippled phase, P β ′. In all these phases (see figure 1.9) the tails have a tilt<br />

with respect to the bilayer normal. At high concentrations of alcohol the rippled phase<br />

disappears, and the interdigitated phase, L βI , is formed, in which the tails do not show<br />

a tilt [39]. Except for the interdigitated L βI phase, these observations are similar to the<br />

observations of membranes without alcohols.<br />

The hydrophobic thickness of an ordinary bilayer is approximately twice the length<br />

of the hydrophobic tails of a phospholipid, in the interdigitated phase the hydrophobic<br />

thickness is reduced to the sum of the length of the hydrophobic tails of a single<br />

phospholipid and the alcohol [39]. This is illustrated in figure 1.8.<br />

Gel phase<br />

Interdigitated phase<br />

Alcohol concentration<br />

Figure 1.8 The gel phase and the interdigitated phase. In the gel state the hydrophobic<br />

thickness is approximately twice the length of the hydrophobic tails of a phospholipid, in<br />

the interdigitated phase the hydrophobic thickness is approximately reduced to the sum of<br />

the length of the hydrophobic tails of a single phospholipid and the alcohol. Modified from<br />

Kranenburg et al. [39].<br />

A wide variety of amphiphilic compounds are capable of inducing the interdigitated<br />

L βI phase in saturated like-chain phosphatidylcholines [10], [47], [50], [63], [68], [70]. The<br />

longer the lipid chain, the more energitically favourable the interdigitated state becomes<br />

[69]. The ability of a molecule to induce the interdigitated phase does not depend on<br />

the presence or absence of a formal charge while molecules that are totally hydrophobic<br />

such as hexane and benzene do not induce the interdigitated phase [69]. <strong>DPPC</strong> does not<br />

stably interdigitate without an inducer [14]. Methanol through heptanol are known to<br />

induce interdigitation (in DSPC, a 18 carbon-chain lipid). From octanol (and above) the<br />

alcohols do not induce the interdigitated phase [39]. The molecule must displace water<br />

from a particular location in the interfacial region and its nonpolar moiety cannot extend<br />

too deeply into the bilayer interior [50].<br />

In the case of short amphiphilic molecules whose non-polar moieties are not as long<br />

as the <strong>DPPC</strong> hydrocarbon chains, this would potentially cause voids between chains in<br />

the bilayer interior. Since the energy of formation of holes in hydrocarbons is large the<br />

chains must eliminate the voids. To do this the chains could either bend cooperatively<br />

or else interdigitate. The lowest energy phase is the interdigitated phase [50].<br />

Another explanation of interdigitation could be, that the chain-end methyl groups are<br />

slightly recessed from the neighboring headgroups to form an amphipathic cavity. Small<br />

amphipathic molecules are thought to bind within the cavity and shield the hydrophobic


10 Introduction<br />

methyl chain ends from water. As expected for a cavity of defined geometry, and as<br />

observed for many biological processes, there is a sharp cutoff in the ability of longer<br />

alcohols to stabilize interdigitation [14].<br />

Absorption of alcohol in a cell membrane can induce significant changes in the structure<br />

of the membrane. These structural changes of the membrane can influence the<br />

conformation of proteins or other structures embedded in the membrane. Several experiments<br />

and simulations on phospholipids have shown that alcohol can induce the<br />

interdigitated phase, one of these experiments is shown in figure 1.9.<br />

Figure 1.9 Schematic representation of the phase diagram of a phosphatidylcholine/alcohol<br />

mixture as a function of the alcohol concentration and temperature. From Kranenburg et al.<br />

[39].<br />

The analysis by Simon et al. of the interdigitated phase shows that the partition coefficients,<br />

see section 1.4, for the anaesthetics are larger for the interdigitated gel phase<br />

than for the usual bilayer gel phase. They claim that this is due to the fact that the<br />

interdigitated phase increases the number of acyl chains per head group at the bilayer/water<br />

interface by a factor of 2, and thus there are more “sites” where amphiphilic<br />

anaesthetics e.g. alcohol can bind to the membrane [69]. Most of the work done on<br />

n-alcohol as an inducer has been done on the short chain alcohol ethanol [1], [25], [37],<br />

[55], [62]. Kaminoh et al. [31] have shown that short-chain alcohols (with anaesthetic<br />

potency 3 ) decrease the temperature in the main transition of phospholipid membranes<br />

between the solid (rippled, P β ′) and liquid (L α ) states. Using longer chain alcohols, the<br />

potency of the alcohols to decrease the transition temperature increases. But when the<br />

alcohols reach a certain length and when used in low concentrations they start to elevate<br />

the transition temperature. The switch from depression to elevation coincides with the<br />

3 In the late 19th century it was postulated that anaesthetics potency is related to the solubility of the<br />

anaesthetics in an oil phase [69].


1.4 Interdigitated phase 11<br />

cut-off length 4 . Thus, a correlation of anaesthetic potency with the phase transition of<br />

lipid membranes is to be expected [31].<br />

Biphasic effect<br />

Short chain alcohols have two different effects on the transition from the low temperature<br />

gel phase to the high temperature liquid phase, depending on concentration [63], [62],<br />

[74], [47]. At low concentration of alcohol the main phase transition temperature shifts to<br />

a lower temperature, whereas at high concentration this transition temperature shifts to<br />

a higher temperature compared to a pure liquid bilayer. This effect is called the biphasic<br />

effect [39], see figure 1.10.<br />

Long chain alcohol (10 − )<br />

Long chain alcohol (5 − 9)<br />

Short chain alcohol (1 − 4)<br />

Phase transition temperature<br />

Concentration<br />

Figure 1.10 An illustration of the main phase transition temperature as a function of the alcohol<br />

concentration. At low concentration the short chain alcohols (1-3) decrease the main phase<br />

transition temperature, at high concentration the short chain alcohols increase the main phase<br />

transition temperature. The long chain alcohols (4-9) only decrease the main phase transition<br />

temperature, whereas the long chain alcohols (10-?) only increase the main phase transition<br />

temperature. Adapted from Rowe [62].<br />

Kaminoh et al. [31] argue that butanol to octanol depress the main phase transition<br />

temperature, decanol is biphasic, depressing the temperature at low concentration and<br />

elevating at high concentrations. Dodecanol does not have any effect on the transition<br />

temperature, tridecanol and tetradecanol elevate the transition temperature. The low<br />

concentration biphasic effect results from the increasing disorder of the lipid tails which<br />

leads to a lower transition temperature. The high concentration biphasic effect, however,<br />

results from the more tightly packed interdigitated phase, resulting in an increase of the<br />

main phase transition temperature [39]. The longer chain alcohols (above pentanol) show<br />

no biphasic effect [47].<br />

4 Tamura et al. [74] explain the cut-off effect by relating it to the anaesthetic potency. The anaesthetic<br />

potency increases with increasing carbon chain length up to about C 10 -C 12 , where the anaesthetic<br />

activity suddenly disappears. The cutoff chain length coincides with the chain length where the effect<br />

on the phase transition temperature of <strong>DPPC</strong> membrane changes from depression to elevation, see<br />

figure 1.10.


12 Introduction<br />

Partition coefficient<br />

It is clear that partitioning of solutes between the membrane and the aqueous phase is a<br />

very important aspect of the interaction of solutes with membranes. However, determining<br />

the partition coefficients experimentally is still not a very common practice, primarily<br />

due to the experimental difficulties [88].<br />

A partition coefficient is a measure of differential solubility of a compound in two<br />

solvents. For this system the partition coefficient describes the ratio of the solute concentration<br />

in the lipid bilayer and in the bulk water, and its value increases with increasing<br />

chain length of the alcohol, caused by the hydrophobic effect.<br />

The partition coefficient is usually defined in terms of mole fractions. The mole<br />

fraction of a component in a mixture is the relative proportion of the specific molecules<br />

in the mixture, by number of molecules. For each component, i, the mole fraction x i is<br />

the number of moles n i divided by the total number of moles in the system n.<br />

x i = n i<br />

n ,<br />

where n = ∑ j<br />

n j (1.1)<br />

The sum is over all components, including the solvent in the case of a chemical<br />

solution.<br />

The partition coefficient is then defined as<br />

K x = X a,l<br />

X a,w<br />

(1.2)<br />

where the numerator is the mole fraction of alcohol in the lipid bilayer, denoted “a,l”,<br />

and the denominator is the mole fraction of alcohol in the solution, denoted “a,w”. Experimentally<br />

it is very difficult to measure the concentration of alcohol in the membrane<br />

directly. Still using isothermal titration calorimetry the so-called solvent-null method<br />

Zhang & Rowe [88], Rowe et al. [65] and Suurkuusk & Singh [73] have measured values<br />

for the partition coefficient of the partitioning of hexanol and pentanol into <strong>DPPC</strong>.<br />

Simulations can be used as as supplement to experiments as a way of finding an<br />

estimate of the mole fraction. Kranenburg et al. have made mesoscopic simulations on<br />

the interdigitated phase of DSPC in order to determine the partition coefficient and<br />

the fraction of alcohol needed to induce the interdigitated phase. From these simulations<br />

it is found that a ratio of 1:2(X a,b = 0.33) for pentanol:DSPC, hexanol:DSPC and<br />

heptanol:DSPC, respectively, is needed to induce interdigitation [39]. This ratio is significantly<br />

lower than the ratios obtained by different experimental methods, see table 1.1,<br />

which explains why the ratio obtained by simulations is not used further in this work.<br />

The free concentration of the alcohol is higher i.e. the partition coefficient is lower<br />

for the interdigitated phase, L βI , than for the liquid, crystalline phase, L α , even though<br />

the L βI phase is found at lower temperature than the L α phase, see figure 1.9. In the<br />

literature there is disagreement about whether the partition coefficient increases with<br />

increasing temperature or is independent of temperature [73].<br />

Löbbecke & Cevc [47] have experimentally demonstrated an exponential relation between<br />

the threshold concentration for inducing the interdigitated phase and the number<br />

of carbon atoms in the alcohol. They find that the threshold ratio decreases by a factor<br />

3 for each CH 2 added to the n-alcohol. It should be stressed that in order to make such<br />

correlations it is very important to know the concentration of lipid in the solution. The<br />

threshold concentration is highly dependent on the concentration of lipid in the solution.<br />

The difference between alcohol and lipid ratio is listed in table 1.1. The variations in


1.4 Interdigitated phase 13<br />

reported ratios are presumed to be a result of the limitations of the different experimental<br />

methods.<br />

Method Alcohol n l /n a<br />

Fluorescence [63]<br />

DSPC conc: 0,13 mM Ethanol 1:453846<br />

(MLV) Propanol 1:1407<br />

Butanol 1:569<br />

Pentanol 1:138<br />

Hexanol 1:123<br />

Heptanol 1:50<br />

Optical [74]<br />

<strong>DPPC</strong> conc: 1,01 mM Ethanol 1:480<br />

(MLV) Butanol 1:31<br />

Hexanol 1:2(1,79)<br />

Heptanol 1:2(1,9)<br />

DSC [47]<br />

<strong>DPPC</strong> conc: 5 mM Ethanol 1:220<br />

(MLV) Pentanol 1:7<br />

Hexanol 1:2<br />

Fluorescence [47]<br />

<strong>DPPC</strong> conc: 5 mM Ethanol 1:220<br />

(ULV) Pentanol 1:12<br />

Hexanol 1:10<br />

ITC [88]<br />

<strong>DPPC</strong> conc: 70.8 mM Butanol 1:5<br />

(MLV)<br />

Table 1.1 Different methods of measuring the threshold mole ratio for inducing interdigitation in<br />

<strong>DPPC</strong> or DSPC <strong>vesicles</strong>. n l and n a are the amount of matter for lipid and alcohol, respectively.<br />

And it is important to notice that the alcohol threshold ratio is highly dependent on the<br />

lipid concentration. The lower the lipid concentration is, the more alcohol per lipid is<br />

needed to induce the interdigitated phase. This observation can be explained by the fact<br />

that a higher concentration of alcohol is needed to saturate the larger amount of the bulk<br />

water.


14 Introduction<br />

1.5 Aim of the thesis<br />

The aim of this thesis is to investigate and characterize the structural differences between<br />

the alcohol induced interdigitated lipid phase, L βI , and the liquid fluid phase, L α , of<br />

the phospholipid dipalmitoyl-phosphatidylcholine and elucidate the differences associated<br />

with the use of 3 different n-alcohols.<br />

Investigations of the structural changes of <strong>DPPC</strong> ULVs associated with the presence of<br />

large amounts of pentanol, hexanol or heptanol with SAXS are not well-described in the<br />

literature. In fact, we have after a comprehensive search not found preceding examples<br />

of these experiments in the literature of this field of research. The SAXS experiments on<br />

this system have not been reported before and will contribute to the understanding of<br />

lipid and alcohol interactions.<br />

Our working hypothesis is that the alcohol will saturate the lipid <strong>vesicles</strong>, making the<br />

system go from the L β ′ phase to the L βI phase. The point of saturation and partition<br />

coefficient can then be found by isothermal titration calorimetry (ITC). The transition<br />

temperature between the L βI phase and L α phase can be determined by differential<br />

scanning calorimetry (DSC) and the structural differences of these two phases, notably<br />

the bilayer thickness, can be studied by small angle X-ray scattering (SAXS).<br />

1.6 Method<br />

Isothermal titration calorimetry is used for the purpose of getting a value for the threshold<br />

alcohol concentration needed to induce the interdigitation as well as elucidating the<br />

difference between using pentanol, hexanol and heptanol as inducer. Especially hexanol<br />

is well-described in the literature, still the role of heptanol as an inducer remains to be<br />

demonstrated for this specific lipid. The method used is the so-called solvent null method<br />

introduced by Zhang & Rowe [88]. This method provide the opportunity for calculating<br />

the partition coefficient, and on this basis it is possible to suggest the minimum alcohol<br />

to lipid ratio needed to induce the interdigitated phase.<br />

On the basis of the threshold values found in the literature and by use of ITC,<br />

samples of alcohol and <strong>DPPC</strong> in adequate ratios are prepared. Differential scanning<br />

calorimetry (DSC) is used in order to get the phase transition temperature between the<br />

apparent L βI phase and the L α phase as shown in figure 1.9. In addition the appearance<br />

and position of the peak from the DSC scan provide information about the transition<br />

enthalpy and the influence of the added alcohols on both the enthalpy and the phase<br />

transition temperature.<br />

The SAXS spectra obtained on either side of the L βI -L α phase transition temperature<br />

are used to get structural information about the electron density bilayer profile of the<br />

samples and thus a measure for the bilayer thickness. By using and taking existing<br />

models further estimates for the bilayer thickness are obtained. The bilayer thickness<br />

estimates, the temperature at which they are obtained and the related lipid to alcohol<br />

ratio would then provide the basis for conclusions regarding the structural appearance<br />

of the L βI and L α phase, respectively.<br />

The structure of the report<br />

The report is divided into three main parts, “Introduction and Theory”, “Experiments<br />

and Results” and “Analysis”. The first part, chapters 1 and 2, consists of an introduction<br />

to the biochemical system, the aim of the thesis and the scattering theory used. The


1.6 Method 15<br />

second part, chapters 3 and 4, consists of a description of the experimental techniques<br />

and the sample preparation as well as the results obtained with the three different techniques.<br />

The third and final part, chapters 5, 6 and 7, consists of both the analysis of the<br />

experimental results and the conclusion.


2 Scattering Theory<br />

This chapter follows, more or less, the structure of [3] where gradually more complex<br />

scattering objects are treated - but other sources have been used as well, mostly [76, 81].<br />

The content of the chapter is primarily selected in order to give a theoretical foundation<br />

for the modelling of the SAXS data. But we also try to make a paedagogic presentation for<br />

our target group in which the cases of scattering from one and two electrons, respectively,<br />

are explained in great details. And it reflects, at some degree, how we have come to<br />

understand the scattering theory.<br />

2.1 X-rays<br />

Although the X-ray phenomenon was observed by J. Hittorf, N. Tesla and H. Hertz<br />

before W.C. Röntgen, he is credited as the discoverer of X-rays - since he was the first<br />

to publish in the late nineteenth century (1895). He saw yellow-green light was flickering<br />

from a (fluorescent) plate near his Geisler discharge tube, from which no visible light<br />

could escape. He found that pieces of paper and wood inserted between the plate and<br />

the tube did not cast a shadow, but that metal indeed did. Since then, in the twentieth<br />

century, the theory of X-rays has evolved and X-rays have found many applications. X-<br />

rays were found to be scattered by gases and to give crystal diffraction patterns (1912).<br />

Later, complex molecular structures such as the structure of DNA (1953) was obtained<br />

by X-ray diffraction. And surely, the X-rays have had a tremendous impact on diagnostic<br />

medicine and therapy. In the 1990s an X-ray observatory was built to monitor violent<br />

processes in the distant universe, such as “stars being torn apart by black holes, galactic<br />

collisions, and novas, neutron stars that build up layers of plasma that then explode<br />

into space” [85, X-rays]. Now, in the beginning of the twentyfirst century, scientists and<br />

engineers are working on the X-ray laser for assembling nano-robots.<br />

The wavelengths of X-rays are in the order of Å= 10 −10 m. In quantum mechanics<br />

light is quantized as photons, and the dispersion relation, E = pc, and the De Broglie<br />

wavelength p = h λ<br />

gives the relation between the photon energy (E) and photon wavelength<br />

(λ)<br />

λ = hc<br />

E<br />

=<br />

12.398 keV<br />

E<br />

(2.1)<br />

where h is Planck’s constant, p is the momentum of the photon and c is the speed of<br />

light in vacuum.<br />

Interaction with matter<br />

X-rays interact with matter (mostly the bound electrons of the atom) in different ways.<br />

The most important interactions are absorption, diffraction and reflection. In the case<br />

of photoelectric absorption the photon annihilates and the bound electron is ejected<br />

into the continuum of energy states. The free electron can be ejected in any direction,<br />

which for a macroscopic object means energy loss in terms of heat. There is no classical<br />

17


18 Scattering Theory<br />

interpretation of the absorption. The electron vacancy can be filled by another bound<br />

electron if it is stimulated by, e.g., a photon with an energy equal to the energy gab<br />

between the states and this is called fluorescent emission. If the photon energy exceeds<br />

the energy gab, a less tightly bound electron may be expelled, and is then called an<br />

Auger electron emission.<br />

In the case of linear absorption the intensity, I(z), of the beam is reduced by the same<br />

fraction for every small distance, dz, travelled into the medium dI = −µI(z) where µ is<br />

dz<br />

the linear absorption coefficient. The solution is I(z) = I 0 exp(−µz), and the coefficient<br />

is approximately proportional to Z 4 , Z being the sum of electrons in the atom. This<br />

makes it excellent for imaging (X-ray images) due to the large contrast mainly between<br />

tissue and bone in the body.<br />

Diffraction and reflection of a beam are explained by media with different refractive<br />

index and are the classical interpretations of the scattering process.<br />

Sources<br />

The state of the art X-ray source is a synchrotron, which is a particular type of cyclic<br />

particle accelerator in which the magnetic field (to turn the particles so they circulate)<br />

and the electric field (to accelerate the particles) are carefully synchronized with the<br />

travelling particle beam. Synchrotrons were developed to study high-energy particle<br />

physics [85], but they are used as X-ray sources as well - in fact some synchrotrons are<br />

built only with this purpose in mind. In the synchrotrons the electrons pass so-called<br />

wigglers or undulators and emit very intense X-rays.<br />

The sources have been improved since the fixed X-ray tube. The rotating anode,<br />

1st, 2nd and 3rd generation of the synchrotrons seriously improved the brilliance of the<br />

beam, and the 3rd generation is approximately 10 12 times more brilliant than the fixed<br />

tube. The brilliance of an X-ray beam is determined by several aspects:<br />

⊲ The number of photons per second (more is better)<br />

⊲ The collimation of the beam - how much it spreads out in milli-radian (less is<br />

better)<br />

⊲ The spectral distribution (monochromatic is better) (0.1 % relative bandwith)<br />

⊲ The source area (mm 2 ) (less is better)<br />

Photons per second<br />

Brilliance ∼<br />

(mrad) 2 (mm 2 [3] (2.2)<br />

source area)(0.1% bandwidth)<br />

For the experimental part of this thesis we have used an X-ray tube as source, which we<br />

discuss in section 3.3 on page 41.<br />

The plane wave representation<br />

X-rays are transverse waves, which can be represented by the monochromatic plane wave<br />

vector k which is pointing in the direction of the travelling wave perpendicular to the<br />

electric E and magnetic B fields. It has the magnitude |k| = k = 2π λ<br />

, λ is the wavelength<br />

of the photon.<br />

Ẽ I (r, t) ≡ E 0 exp(i(k · r − ωt))ˆx (2.3)<br />

where ω is the angular frequency, E 0 is the amplitude of the incident electric field Ẽ I<br />

which is a function of time, t, and position, r. The complex notation (exp(iα) = cos(α)+<br />

i sin(α)) is used and the physical electric field is the real part of the complex number<br />

E = Re(Ẽ) which is implied for the rest of the chapter.


2.1 X-rays 19<br />

Coherence and incoherence<br />

Two X-ray beams are coherent if they are parallel and monochromatic. In reality this<br />

idealization is not fulfilled which means that the two beams at some places will make<br />

constructive and destructive interference. Non-parallel beams cause transverse coherence<br />

and polychromatic beams cause longitudinal coherence. The longitudinal coherence<br />

length, l L , is the length between constructive and destructive interference, thus 2l L is the<br />

length between constructive and constructive interference. If one beam has a wavelength,<br />

λ, and the other has a slightly smaller wavelength, λ ′ ≡ λ − δλ, then<br />

2l L = Nλ = (N + 1)λ ′ ⇒ N =<br />

1<br />

λ/λ ′ − 1<br />

(2.4)<br />

where N is the number of wave crests of λ in order to make constructive interference<br />

with the other wave (with wavelength λ ′ ). This gives the longitudinal coherence length<br />

l L =<br />

λ<br />

2 (λ/λ ′ − 1) = λλ ′<br />

2 (λ − λ ′ ) ≈ λ2<br />

2 δλ<br />

ifλ ≈ λ ′ (2.5)<br />

a) Longitudinal coherence length, l L<br />

θ<br />

2l L =N λ<br />

A<br />

λ<br />

λ − δλ<br />

b) Transverse coherence length, l T<br />

B<br />

B<br />

A<br />

θ<br />

R<br />

2 l T<br />

D<br />

λ<br />

S<br />

Figure 2.1 Longitudinal coherence length, l L, and transverse coherence length, l T . a) Two plane<br />

waves are emitted in the same direction. The two waves are out of phase by a factor of π after<br />

the distance l L. b) Two waves with the same wavelength are emitted from the ends of a finite<br />

sized source of a width D, in this work equivalent to the width of the beam.<br />

If a perfect crystal is used for collimation in a synchrotron then δλ λ<br />

means the longitudinal coherence length is about 5µm.<br />

≈ 10−5 , which<br />

The transverse coherence length, l T , is the length between constructive and destructive<br />

interference. Both beams have the same wavelength λ but propagate in slightly different


20 Scattering Theory<br />

directions, θ, e.g. due to the finite size of the entrance slit of the X-ray source. Travelling<br />

the distance l T from the point S along the wavefront of wave A one reaches the point<br />

where the wave A is out of phase with the wave B. Proceeding the distance l T along<br />

the wavefront the waves are in phase again, as 2l T θ = λ, see figure 2.1. If the waves<br />

propagate from the same source but with a distance D from each other, the distance<br />

from the point S to the source is R, then<br />

tan(θ) = λ = D 2l T R ⇒ l T = λR<br />

2D<br />

(2.6)<br />

Thus, the coherence length gives an upper limit on the distances between the scattering<br />

objects. If the scattering centers are separated by distances greater than the coherence<br />

length then we add the intensities rather than the amplitudes.<br />

For the Kratky camera the distance to the sample from the source is about 5cm and the<br />

beam width is about 1cm which gives a transverse coherence length about 3.9Å as the<br />

wavelength is about 1.55Å.<br />

2.2 The classical (particle) scattering experiment<br />

Impact parameter ρ<br />

p 0<br />

Axis<br />

Figure 2.2 Classical scattering of a point particle with initial momentum p 0 and an impact<br />

parameter ρ (figure from Taylor [76, p. 44]).<br />

In the classical scattering experiment a projectile is fired at a target. We can measure<br />

or, otherwise, somehow determine the incident and scattered momentum of the projectile,<br />

but we cannot measure the microscopic details of the event. In particular we cannot<br />

measure the impact parameter, ρ, which is defined as a vector with a direction perpendicular<br />

to an axis chosen to go through the target and has the length magnitude from<br />

the axis to the actual trajectory of the projectile (see figure 2.2). We cannot measure the<br />

impact parameter, because no source is truly point like. We always have a cross sectional<br />

area of the beam of projectiles. The problem is then to extract as much information as<br />

possible about the target given these limitations.<br />

The outcome of a single experiment shows us if the projectile has hit the target or not.<br />

The projectile has a different momentum if it has hit the target, and it is undeviated and<br />

has the same momentum if it has missed the target. As the experiment is repeated many<br />

times with the same incident momentum and random impact parameters, it makes sense<br />

to talk about the average number of incident projectiles per unit area (n i ) perpendicular<br />

to the incident beam. The number of scattered projectiles (N s ) by the cross sectional<br />

area of the target (σ) is then<br />

N s = n i σ (2.7)<br />

This actually allows us to determine the σ since we, in principle, can count n i and<br />

N s . And in this case N s is the total count of scattered projectiles. If you only count


2.3 X-Ray scattering 21<br />

the scattered projectiles at a certain solid angle, N s (∆Ω), usually, it is only a part of σ,<br />

which is the source of those projectiles. Thus, σ is a function of ∆Ω:<br />

N s (∆Ω) = n i σ(∆Ω) (2.8)<br />

[8].<br />

In the case of a target with a cross-sectional area larger than the incident beam,<br />

the scattered intensity is proportional to the incident intensity. And in the case of a<br />

target with a cross-sectional area smaller than the incident beam, the scattered intensity<br />

is proportional to the incident flux[3, p. 261]. In the latter case the differential cross<br />

section can be defined as<br />

dσ (Number of X-rays scattered per second into ∆Ω)<br />

≡ (2.9)<br />

dΩ (Incident flux)(∆Ω)<br />

The total scattering cross section is found by integrating equation 2.9 over all of space.<br />

The differential cross section is perhaps the most important quantity in scattering theory<br />

since it is the meeting point between theory and experiment.<br />

2.3 X-Ray scattering<br />

In a scattering experiment the incident beam consists of either light, X-rays, neutrons<br />

or other probes and is directed at the sample. In the case of X-ray scattering most of<br />

the photons pass through the sample undeviated and a few are scattered once from the<br />

scattering volume. Mainly there are three types of scattering experiments in which you<br />

get different information about the sample:<br />

⊲ measuring the dependence on angle of the average scattered intensity (’static scattering’)<br />

yields structural information. Shapes and internal structure of individual<br />

particles in a dilute system, and spatial, positional correlations in a concentrated<br />

system.<br />

⊲ analysis of the time dependence of fluctuations in the scattered radiation yields dynamic<br />

information. The Brownian movement and/or how the shape/configuration<br />

fluctuate in time.<br />

⊲ measuring the absolute magnitude of the scattered intensity (time or frequency<br />

average) yields mass or molecular weight of the scattering objects.<br />

Weak scattering (also called kinematical scattering, which essentially means a gentle<br />

probe that does not destroy or alter the target) is the case when the sample is a perturbation<br />

to the incident beam, which means the Born approximation holds. But the weak<br />

interaction also means a weak signal, which is then compensated by an intense beam.<br />

As the time of interaction for the X-rays with the sample is small compared to the motion<br />

time of the particles within the sample, the measurements are series of ”snapshots”<br />

which are averaged.<br />

In this X-ray scattering section, we have chosen a structure similar to [3], in which<br />

the complexity of the scattering objects are gradually increased, starting with just one<br />

electron, then treating two electrons, the electrons in an atom, a molecule, a small crystal<br />

and a lipid bilayer.<br />

One free electron<br />

The electric field, E I (t ′ ) = E 0 exp(iωt ′ )ˆx, of the incident plane wave travelling along the<br />

ẑ axis is the driving force. Therefore, the equation of motion for the free electron (placed


22 Scattering Theory<br />

at origo) is<br />

m e a(t ′ ) = −eE I (t ′ ) (2.10)<br />

where m e and e is mass and charge of the electron, respectively. The oscillating electron<br />

can be considered an accelerated dipole, ¨p, if the amplitude of the oscillation, |E I | is<br />

much smaller than the distance to the point of observation, r, see figure 2.3.<br />

x<br />

n<br />

θ<br />

r<br />

E<br />

B<br />

R<br />

ṗ .<br />

Ψ<br />

z<br />

ϕ<br />

y<br />

Figure 2.3 The accelerated dipole, ¨p, located at origo, radiates a spherical wave. At this<br />

snapshot the observation point, r, the propagation vector, ˆk, the radiated electric field, E R,<br />

and the radiated magnetic field, B, are sketched. (figure from [3, p. 269])<br />

The observation is performed at time t, which is r / c later, so t = t ′ + r / c . The radiated<br />

electric field, E R (r, t), is derived from Maxwell’s equations along the following path: the<br />

current density J −→ the retarded vector potential A −→ the magnetic field B −→ the<br />

electric field E = cB × ˆr. The details of this derivation are explained in [3, p. 267-271]<br />

and the result is<br />

E R (r, t) = 1<br />

4πǫ 0<br />

1<br />

c 2 r (¨p(t′ ) × ˆr) × ˆr = 1<br />

4πǫ 0<br />

−e<br />

c 2 r (a(t′ ) × ˆr) × ˆr (2.11)<br />

since ¨p(t ′ ) = d2<br />

dt ′2 p = −ea(t ′ ).<br />

Isolating a(t ′ ) in (2.10) inserting that into (2.11) then the radiated field in terms of<br />

incident field becomes<br />

E R (r, t) =<br />

e 2<br />

4πǫ 0 m e c 2 1<br />

r (E I(t ′ ) × ˆr) × ˆr (2.12)<br />

And since E I (t) = E 0 exp(iω(t − r / c ))ˆx = E 0 exp(iωt − k · r)ˆx the radiated field<br />

becomes<br />

(<br />

e 2 ) ( )<br />

E0 exp(i(ωt − k · r))<br />

E R (r, t) =<br />

4πǫ 0 m e c 2 (ˆx × ˆr) × ˆr (2.13)<br />

r


2.3 X-Ray scattering 23<br />

or written as the ratio between radiated and incident<br />

( )<br />

|E R (r, t)| exp(−ik · r)<br />

E 0 exp(iωt) = (b T)<br />

(− cos(ψ)) (2.14)<br />

r<br />

where b T is the Thomson scattering length<br />

b T =<br />

e 2<br />

4πǫ 0 m e c 2 = 2.82 × 10−5 Å (2.15)<br />

It is worth noting the sign of the last term, which means that the radiated beam is π<br />

out of phase with respect to the incident beam. This is the same as the E R is pointing<br />

opposite to ¨p along the x-axis in figure 2.3. cosψ is valid when the measuring point is<br />

in a plane parallel to the incident plane of polarization. Otherwise, the factor is 1 if the<br />

measuring point is in a plane perpendicular to the incident plane of polarization, since<br />

full acceleration is seen from the observation point.<br />

The intensity, I, is the absolute square of the scattered electric field amplitude<br />

I ∝ |E| 2 (2.16)<br />

which means that the ratio between the scattered intensity, I S , and the incident intensity,<br />

I I is<br />

I S<br />

= |E R| 2 r 2 ∆Ω<br />

I I |E I | 2 (2.17)<br />

σ I<br />

where r 2 ∆Ω is the area of the detector and σ I is the cross sectional area of the incident<br />

beam.<br />

Combining (2.17) and (2.14) gives the differential cross sectional area of Thomson<br />

scattering<br />

dσ<br />

dΩ =<br />

I S<br />

(I I /σ I )∆Ω = |E R| 2 r 2<br />

|E I | 2 = b 2 T cos 2 ψ = b 2 TP (2.18)<br />

where P is the polarization factor for the intensity. The Thomson cross section can then<br />

be found by integration<br />

∫ ∫<br />

σ T = dσ = b 2 T cos 2 (ψ)dΩ = 8π 3 b2 T = 0.665 barn (2.19)<br />

where 1 barn = 10 −24 cm 2 is a typical measuring unit for cross sections. Hence, the cross<br />

section for Thomson scattering is a constant and does not depend on energy.<br />

In the case of an unpolarized incident beam, the polarization vector can point along<br />

any direction in the x-y plane, and the polarization factor is averaged to be 1 / 2 (1+cos 2 ψ).<br />

In order to summarize, the polarization factor, P , for the Thomson differential cross<br />

section is<br />

⎧<br />

1 measuring point in a plane perpendicular<br />

⎪⎨<br />

to the incident plane of polarization<br />

P = cos 2 ψ measuring point in a plane parallel<br />

(2.20)<br />

to the incident plane of polarization<br />

⎪⎩ 1/ 2 (1 + cos 2 ψ) an unpolarized source<br />

For small angle scattering P ≈ 1.


24 Scattering Theory<br />

At this point we may wonder; how on Earth can we detect the scattering, when the<br />

intensity ratio is in the order of 10 −24 just a few centimeters away from the electron? A<br />

part of the answer to that question is that the amount of electrons in a sample is in the<br />

order of Avogadro’s number, but it is not the whole answer, since the arrangement of the<br />

electrons is crucial to whether or not the intensity becomes measurable.<br />

Two free electrons<br />

The electron seems to be a structureless particle, hence the most simple structure must<br />

be two electrons. If the first electron is placed at origo, the scattered waves from this<br />

electron are out of phase. The two free electrons are situated at a fixed distance r between<br />

each other and the displacement vector is r. The phase lag (behind) of the incident wave<br />

between the two electrons are φ in = −k · r (marked as z on figure 2.4), and the phase<br />

lag (ahead) of the scattered wave between the two electrons are φ out = k ′ ·r 1 . This gives<br />

the combined phase lag between the incident and scattered wave of the second electron<br />

φ = φ in + φ out = −k · r + k ′ · r = −q · r, which defines the scattering vector,<br />

q ≡ k − k ′ (2.21)<br />

which is sometimes, suggestively, written as the momentum transfer vector q ≡ k−k ′ .<br />

The geometry of figure 2.4 shows that for Thomson scattering |k ′ | = |k|<br />

|q| = |k| sin(θ) + |k ′ | sin(θ) = 2|k| sin(θ) = 4π λ<br />

sin(θ) (2.22)<br />

In the case of Compton scattering |k ′ | < |k| equation (2.22) does not hold. Looking at<br />

one particular q is the same as looking at one particular angle. The wavelength of the<br />

X-ray is inversely proportional to the scattering vector.<br />

k’<br />

r<br />

q<br />

k<br />

z<br />

k’<br />

2θ<br />

λ<br />

k<br />

Figure 2.4 a) The vertical straight lines represent the plane wave crests each separated by λ and<br />

the wave vector k indicates the direction of propagation. The electron in the middle is placed<br />

at origo and the other electron is displaced by r. b) The two plane wave vectors are placed in<br />

a suggestive way, such that the definition of the scattering vector q becomes apparent. The<br />

scattering angle is defined as 2θ such that the sine term in equation (2.22) becomes θ.<br />

Ignoring the polarization, the form factor, f , for two electrons separated by r is<br />

f two electrons (q) = −b T exp(i(q · 0)) − b T exp(i(q · r)) = −b T (1 + exp(i(q · r))) (2.23)<br />

1 We could equally have defined it as φ = q · r and we drop the minus as many textbooks do.


2.3 X-Ray scattering 25<br />

which in the case of parallel q and r vector gives the intensity<br />

I(q) ‖ = 2b 2 T(1 + cos(qr)) (2.24)<br />

and an orientational average over all different angles gives<br />

(<br />

〈I(q)〉 or. = 2b 2 T (1 + 〈exp(i(q · r))〉) = 2b2 T 1 + sin(qr) )<br />

qr<br />

(2.25)<br />

where the intermediate steps of<br />

〈exp(iq · r)〉 orientation = sin(qr)<br />

qr<br />

(2.26)<br />

are skipped, see Als-Nielsen & McMorrow [3, p. 111]. This shows that the arrangement<br />

of the electron is crucial to the diffraction pattern. Now, it’s important not to forget that<br />

q is a function of both angle as well as λ, which means that for a polychromatic beam<br />

the detected intensity I for one particular q gets contributions from different angles.<br />

Secondary scattering occurs when a photon is scattered by an electron and afterwards<br />

hit another electron, and is detected subsequently. The detected energy and position of<br />

this electron would contribute to the spectrum with a so to speak misleading information<br />

about the geometry of the system.<br />

4<br />

q and r parallel<br />

Orientational average<br />

3<br />

Intensity / b T<br />

2<br />

2<br />

1<br />

0<br />

0 0.5 1 1.5 2<br />

q / [2π / r]<br />

Figure 2.5 Two diffraction patterns of two electrons placed the distance r from each other but<br />

with different angles in respect to the incident beam. The solid line represents the diffraction<br />

pattern where r is parallel to q and the intensity given by equation (2.24). The dotted line<br />

represents the diffraction pattern where r is randomly oriented and an average of all angles<br />

between r and q and the intensity is given by equation (2.25).<br />

There is a periodic variation in the calculated intensities, reflected in the scattered<br />

waves being either out of phase (minimum) or in phase (maximum). By measuring the<br />

intensities as a function of the scattering vector q the diffraction pattern is obtained, and<br />

when subsequently fitting r the structure of the two-electron system is obtained.


26 Scattering Theory<br />

When extending the model to comprising more than two electrons, all the elastic<br />

scattering amplitudes must be added up and<br />

∑<br />

N<br />

f many electrons (q) = −b T exp(iq · r j ) (2.27)<br />

where r j is every displacement vector between the electrons.<br />

Defining the form factor, equation 2.27, is the first step in the direction of developing<br />

form factors for more complex systems such as atoms, molecules and finally the lipid<br />

bilayer.<br />

An atom<br />

Depending on its atomic number the atom contains a number of electrons. The electrons<br />

are represented by wavefunctions ψ(r) and the electron propability density ρ(r) = |ψ(r)| 2<br />

and the propability ρ(r)dr of finding an electron in the volume element dr. The form<br />

factor of an atom, f atom , is then defined as the contributions to the scattering field from<br />

each volume element<br />

j=1<br />

f atom (q) =<br />

=<br />

∫<br />

ρ(r)exp (iq · r)dr (2.28)<br />

V<br />

{<br />

Z if |q| → 0<br />

(2.29)<br />

0 if |q| → +∞<br />

The limit for the form factor as |q| → 0 equals the total number of electrons in<br />

the atom, Z, as the phase factor approaches zero. And the limit for the form factor as<br />

|q| → +∞ equals zero as there is destructive interference between the waves scattered<br />

from the different electrons in the atom and the radiation waves, as the latter are small<br />

compared to the atom (for further explanation, see Als-Nielsen & McMorrow [3, p. 112]).<br />

The right hand side of equation 2.28 is recognized as a Fourier transform, hence the<br />

form factor is the Fourier transform of the electron distribution ρ of the given sample.<br />

A molecule<br />

Upon defining the form factor for an atom comes the definition of the form factor for a<br />

molecule containing a number of atoms. The molecule form factor is composed of atomic<br />

form factors, fj atom (q), located at r j .<br />

f molecule (q) =<br />

N∑<br />

j=1<br />

f atom<br />

j (q)exp (iq · r j ) (2.30)<br />

The atomic form factor differs for each of the atoms in the molecule, and in order<br />

to determine the atomic arrangement of the molecule one has to have an idea of the<br />

structure before interpreting the measured data. Molecules with the same number of<br />

atoms can differ radically in structure.<br />

In case of measuring geometric properties of e.g. large biomolecules it is most often<br />

necessary to dissolve these in an appropriate solvent. The electron density is relative to<br />

the solvent<br />

∆ρ(r) = ρ(r) − ρ solvent (2.31)


2.3 X-Ray scattering 27<br />

where ρ(r) is the actual electron density at r. In order to correct for the contribution<br />

from the solvent a so-called background spectrum is obtained. This spectrum specifies the<br />

contributions from the solvent and other possible contributors within the experimental<br />

setup.<br />

This relative electron density gives rise to the possibility of contrast variation, a topic<br />

of which neutron scattering is especially usefull, since the scattering length of neutrons<br />

depends strongly on the isotope. The scattering length for deuterium, e.g., is much larger<br />

than hydrogen, whereas the X-ray scattering lengths for these two isotopes are the same.<br />

A small crystal<br />

In this context a crystal is defined as a crystalline material which is periodic in space [3].<br />

The crystal structure is specified as a lattice of points in space, reflecting the symmetry<br />

of the crystal and a unit cell structure, referring to which atoms to associate with each<br />

lattice sites. The scattering amplitude for a crystal is then the product of the unit cell<br />

structure factor and the lattice sum<br />

f crystal (q) = f basis (q) ∑ r l<br />

exp (i(q · (r l + r))) (2.32)<br />

=<br />

Basis<br />

{ }} {<br />

f basis (q)exp(iq · r) ∑ r l<br />

exp(iq · r l )<br />

} {{ }<br />

Lattice sum<br />

(2.33)<br />

where r l = n 1 a 1 + n 2 a 2 + n 3 a 3 is the lattice vector with a as the basis in real space,<br />

and r is the position vector of the unit cell structure relative to the lattice. The unit cell<br />

structure factor f basis could be an atom, a molecule, or possibly multiple combinations<br />

of both. The lattice sum usually consists of a huge amount of terms which only blow up<br />

as q takes on special positions in q-space (also called the reciprocal space). This happens<br />

when q · r l = 2π × integer and is called the Laue condition<br />

q = g = ha ∗ 1 + ka∗ 2 + la∗ 3 (2.34)<br />

where g is the reciprocal lattice vector, h, k, l are the Miller indices, and a ∗ is the basis<br />

vector in the reciprocal space. So, for every lattice in real space you can construct a lattice<br />

in reciprocal space which gives constructive interference 2 . Conversely, if the scattering<br />

objects in the sample are dispersed (no characteristic lengths) there is no structure factor<br />

- it is unity.<br />

For every h, k, l-plane in real space there is a lattice point in the reciprocal space.<br />

The distances between the planes are given by<br />

d hkl =<br />

2π<br />

|g hkl |<br />

(2.35)<br />

As an example consider a hexagonal structure with the lattice vector<br />

⎛ √ ⎞ ⎛<br />

3a/2<br />

r n = n 1 a 1 + n 2 a 2 + n 3 a 3 = n 1<br />

⎝ a/2 ⎠ + n 2<br />

⎝ −√ ⎞ ⎛<br />

3a/2<br />

a/2 ⎠ + n 3<br />

⎝ 0 0<br />

0<br />

0<br />

c<br />

⎞<br />

⎠ (2.36)<br />

2 The Laue condition is equivalent to Bragg scattering.


28 Scattering Theory<br />

with the two lattice constants a and c (constructed such that |a 1 | = |a 2 | = a and |a 3 | = c).<br />

The reciprocal lattice basis becomes<br />

a ∗ 1 =<br />

a ∗ 2 =<br />

a ∗ 3 =<br />

2π<br />

a 1·(a 2×a 3) a 2 × a 3 =<br />

2π<br />

a 1·(a 2×a 3) a 3 × a 1 =<br />

2π<br />

a 1·(a 2×a 3) a 1 × a 2 =<br />

⎛<br />

4π<br />

√ ⎝<br />

3a2 c<br />

⎛<br />

4π<br />

√ ⎝<br />

3a2 c<br />

⎛<br />

4π<br />

√ ⎝<br />

3a2 c<br />

ac/2<br />

√<br />

3ca/2<br />

0<br />

−ac/2<br />

√<br />

3ac/2<br />

0<br />

0<br />

0<br />

√<br />

3a 2 /2<br />

⎞<br />

⎠ = 2π a<br />

⎞<br />

⎠ = 2π a<br />

⎞<br />

⎠ = 2π c<br />

⎛<br />

⎝ 1/√ 3<br />

1<br />

0<br />

⎛<br />

⎞<br />

⎝ −1/√ 3<br />

1<br />

0<br />

⎛ ⎞<br />

⎝ 0 0<br />

1<br />

⎠ (2.37)<br />

⎞<br />

⎠ (2.38)<br />

⎠ (2.39)<br />

and the reciprocal lattice vector is<br />

⎛<br />

g hkl = ha ∗ 1 + ka ∗ 2 + la ∗ 3 = 2π ⎝<br />

√<br />

3(h − k)a<br />

−1<br />

(h + k)a −1<br />

lc −1<br />

⎞<br />

⎠ (2.40)<br />

with the length<br />

|g hkl | = √ g hkl · g hkl =<br />

√<br />

4π 2 ( ( √3(h−k)<br />

a<br />

√<br />

= 2π<br />

) )<br />

2 (<br />

+<br />

h+k<br />

) 2 (<br />

a + l<br />

) 2<br />

c<br />

h 2 +k 2 −hk<br />

(a/2) 2<br />

(2.41)<br />

+ l2<br />

c 2 (2.42)<br />

The lattices in direct space and reciprocal space are sketched in figure 2.36.<br />

If one suspects hexagonal rods (l = 0), one should look for q-peaks with the following<br />

distances<br />

2π<br />

|g hk0 |<br />

|g 100 | =<br />

√ ( )<br />

h 2 +k 2 −hk<br />

(a/2)<br />

+ 02<br />

2 c 2<br />

2π√ (<br />

1 2 +0 2 −0<br />

(a/2) 2 + 02<br />

c 2 ) = √ h 2 + k 2 − hk = √ 1, √ 3, √ 4, √ 7, √ 9, · · · (2.43)<br />

for different combinations of h and k in terms of the first peak, i.e. the first peak is<br />

|q ′ | = √ 1|g 100 |, the next is |q ′′ | = √ 3|q ′ | and the next is |q ′′′ | = √ 4|q ′ | and so forth.<br />

2.4 Bragg’s law<br />

Fulfilling the requirements for the Laue condition and referring to the relations in equation<br />

2.22 Bragg’s law is obtained.<br />

nλ = 2d sin(θ) (2.44)<br />

where n is an integer and d is the repeat spacing of the lattice planes. Combining (2.44)<br />

with (2.22) gives the Bragg plane distance<br />

d = 2π<br />

|q ′ | n (2.45)


d 11<br />

d 10<br />

g 11<br />

2.5 Summary 29<br />

g 01<br />

g 10<br />

Figure 2.6 The figure on the left-hand side shows a schematic drawing of the hexagonal type II.<br />

The characteristic repeat distances are denoted d 10 and d 11 where d 10 = 2π<br />

2π<br />

|g 10<br />

and d11 = .<br />

| |g 11 |<br />

The figure on the right-hand side shows the diagram of the reciprocal lattice. The diffraction<br />

planes are perpendicular to the reciprocal lattice vectors g 10 and g 11, the basis vectors are<br />

shown in red. Both figures adapted from [24].<br />

k<br />

k’<br />

d<br />

θ<br />

θ<br />

Figure 2.7 Bragg scattering, adapted from [3].<br />

where n is an integer.<br />

Laue condition is not true for Compton scattering, which gives a smoothly varying<br />

background. For bound electrons with initial momentum the Compton cross section gives<br />

the momentum distribution. The Compton scattering length is approximately 137 times<br />

larger than the Thomson scattering length.<br />

2.5 Summary<br />

Experimentally, both Thomson and Compton scattering are observed, but the classical<br />

theory can only account completely for Thomson scattering. 3 Even so, there are several<br />

reasons why the classical treatment is valid and sometimes sufficient:[81]<br />

3 Please refer to appendix B, page 115, for a short outline of the quantum mechanical treatment of the<br />

scattering process.


30 Scattering Theory<br />

⊲ Correct wave mechanical treatment shows that the sum of intensities from Thomson<br />

and Compton scattering is close to the sum of individual classical electron<br />

intensities.<br />

⊲ The polarization is given correctly for both Thomson and Compton scattering (the<br />

polarization is the classical representation of the quantum mechanical spin).<br />

⊲ Plane waves can be superpositioned in order to get a normalizable wavepacket.<br />

In the elastic scattering theory the incident beam is considered a plane wave which<br />

makes the electrons of the sample oscillate and radiate spherical waves. The diffraction<br />

pattern then depends on the arrangement of the electrons. The diffraction is governed<br />

by a reciprocal law: The scattering angle becomes smaller as the dimensions of the<br />

scattering object become larger with respect to the wavelength of the incident beam. If<br />

the system consists of many electrons it makes sense to talk about the electron density<br />

and replace the sum with an integral. The intensity is thus integrated and by convention,<br />

the system is divided into identical subsystems (form factors), and the arrangement of<br />

these unit cells gives rise to terms labelled structure factor. Apart from the polarization<br />

factor the integrated intensity sometimes gives rise to additional terms which depend on<br />

the scattering angle, θ. This is referred to as the Lorentz factor, but the textbooks on<br />

X-ray scattering ([3, p. 146], [81, p. 44], [45, p. 206]) leave vague impressions on what<br />

exactly the definition of the Lorentz factor is. They all agree that it depends highly<br />

on geometry. [3] states that for a small crystallite proper integration over k ′ and θ in<br />

the case of slightly incoherent incident and/or scattered beam a factor of 4 1<br />

sin(2θ) ≈ 1 q is<br />

added to the differential cross section. According to [45] the same expression is added as<br />

a result of rotating the crystal with a constant angular velocity. Both [45] and [81] refer<br />

to the angel-dependent terms as the polarization-Lorentz factor (or correction).<br />

4 Valid approximation for small angles.


Part II<br />

Experiments and Results


3 Experimental Techniques<br />

This chapter is intended to introduce the experimental techniques used. We have chosen<br />

to review the experimental equipment thoroughly as the instrumental setup often is one<br />

of the keys to understanding the potentials and limits of the conducted experiments.<br />

Especially with regard to the SAXS experiments the experimental setup has a significant<br />

impact on the evaluation of the obtained data and the further modelling process.<br />

3.1 ITC<br />

Isothermal Titration Calorimetric (ITC) is a very powerful tool to determine thermodynamic<br />

parameters of various binding processes by directly measuring the heat absorbed<br />

or generated during the binding processes. ITC has the great advantage that all thermodynamic<br />

quantities such as stoichiometry, binding constants, Gibbs free energy change,<br />

enthalpy change, and entropy change can be obtained at the same time from only one<br />

experiment [71].<br />

The ITC experiments were carried out using two different calorimeters, a VP-ITC<br />

instrument and a MSC-ITC instrument from both Microcal, Inc. (Northampton, MA).<br />

The data were primarily analysed using ORIGIN software from Microcal, Inc., Amherst,<br />

MA. Both instruments are ultrasensitive isothermal calorimeters and use a cell feedback<br />

network to differentially measure and compensate for heat produced or absorbed between<br />

the sample and reference cell. The temperature difference between the two cells<br />

is measured thermoelectrically while another device measures the temperature difference<br />

between the cells and the jacket, see figure 3.1. When a chemical reaction in the sample<br />

cell occurs, heat is either generated or absorbed, resulting in a temperature difference<br />

between the two cells.<br />

If an exothermic reaction occurs, the temperature of the sample cell increases upon an<br />

injection from the syringe. This causes a decrease in the feedback power to the sample<br />

cell for the purpose of maintaining an equal temperature between the two cells. The<br />

opposite occurs in case of an endothermic reaction; the feedback power increases in order<br />

to keep the temperature constant.<br />

Measuring the partition coefficient<br />

Zhang & Rowe [88] have used ITC to measure the heat of binding and the partition coefficient<br />

of n-butanol into multilamellar <strong>vesicles</strong> of <strong>DPPC</strong>. Zhang & Rowe [88] demonstrate<br />

that ITC can be used to give a direct measure of the X a,b , see equation 1.2, associated<br />

with the phase transition from the gel phase, L β ′, to the interdigitated phase L βI by<br />

applying the so-called solvent null method.<br />

Applying the method to the investigated system in this work would yield additional<br />

information about the phase diagram, see figure 1.9. By use of this method it would be<br />

possible to suggest an alcohol to lipid ratio for the phase transition from the L β ′ to the<br />

L βI . Besides this ratio the partition coefficient, the free energy of transfer, the enthalpy<br />

33


34 Experimental Techniques<br />

Alcohol in<br />

reaction cell<br />

Lipid/alcohol in<br />

stirring syringe<br />

∆T 1<br />

∆T 2<br />

Jacket<br />

feedback<br />

Sample cell<br />

feedback<br />

(a) Diagram of the ITC cell and the<br />

syringe. For the experiments the<br />

lipid/alcohol suspension is placed in<br />

the injection syringe and the alcohol<br />

solution is placed in the ITC sample<br />

cell. The syringe rotates during the<br />

experiment. The end of the syringe<br />

has been adapted to provide continuous<br />

mixing in the ITC cell. The<br />

plunger is computer-controlled and injects<br />

precise volumes of the lipid/alcohol<br />

solution.<br />

(b) A diagram of the heat flow between<br />

the reference cell and the sample<br />

cell. During the experiment the<br />

coinshaped cells are kept at a constant<br />

temperature and ∆T 1 , the temperature<br />

difference between the sample<br />

and reference cell, is kept at a constant<br />

value. ∆T 2 is the temperature<br />

difference between surroundings and<br />

the reference cell. The two cells are<br />

connected to the outside by narrow<br />

tubes through which the lipid/alcohol<br />

suspension is injected and subsequently<br />

the syringe is immersed.<br />

These cells are kept in an adiabatic<br />

environment. Adapted from [51].<br />

Figure 3.1 Schematic drawing of the ITC setup.<br />

of transfer and the entropy of transfer can be calculated for each experimental series.<br />

These experiments should hence confirm that the ratio of the lipid/alcohol suspensions<br />

used in the SAXS experiments was high enough to induce the interdigitated phase.<br />

The method can be used to determine the concentration of free alcohol at any alcohollipid<br />

composition. In this work the solvent-null method has been used to examine the<br />

pentanol/hexanol/heptanol-<strong>DPPC</strong> vesicle system. The interactions between <strong>DPPC</strong> and<br />

the three different alcohols were studied by placing the lipid suspension in the syringe<br />

and the alcohol solutions in the sample cell. This was preferred over placing lipid in the<br />

cell and alcohol in the syringe, as alcohol has a very large heat of dilution whereas the<br />

heat of dilution of the lipid solution is negligible [21], [88].<br />

The principle behind the method is that when the alcohol concentration in the cell<br />

exceeds the free alcohol in the alcohol-<strong>DPPC</strong> solution being injected, the measured signal<br />

is correlated with the resultant partitioning of alcohol to the <strong>vesicles</strong>, and an endothermic<br />

signal is measured. Similarly if the alcohol concentration in the cell is below the free


3.1 ITC 35<br />

concentration, the heat signal correlates to the release of bound alcohol from the lipid<br />

<strong>vesicles</strong>, and an exothermic signal is measured. Therefore, when the heat signal is zero,<br />

there is no net binding/release and the alcohol concentration in the cell equals the free<br />

alcohol concentration in the syringe.<br />

The raw data are plotted as the heat flow in µcal/sec as a function of time in seconds,<br />

causing the raw data to consist of a series of peaks of heat flow, each peak referring<br />

to an injection. Integrating these peaks with respect to time gives the total heat effect<br />

per injection. The profile of the plot of these effects as a function of the molar ratio<br />

can be used in the analysis of the thermodynamic parameters of the interaction under<br />

investigation. Pentanol stock solution and <strong>DPPC</strong> ULVs were premixed and kept shaken<br />

at 850 rpm at 52 ◦ C for at least 48 hours prior to using. The <strong>DPPC</strong>-pentanol total<br />

molar ratio was approximately equal to 2:1. The pentanol solution and the pentanollipid<br />

suspension were degassed prior to being loaded into the 1.3 mL reaction cell and<br />

the 100 µL syringe, respectively. After equilibrating the system to 25 ◦ C, 4 injections of 5<br />

µL are titrated into the cell. Hexanol stock solution and <strong>DPPC</strong> ULVs were premixed and<br />

kept shaken at 850 rpm at 52 ◦ C for at least 48 hours prior to using. The <strong>DPPC</strong>-hexanol<br />

total molar ratio was approximately equal to 3:1 for all the experimental series. The<br />

hexanol solution and the hexanol-lipid suspension were degassed prior to being loaded<br />

into the 1.3 mL reaction cell and the 100 µL syringe, respectively. After equilibrating the<br />

system to 25 ◦ C, 4 injections of 5 µL are titrated into the cell. Heptanol stock solution<br />

and <strong>DPPC</strong> ULVs were premixed and kept shaken at 850 rpm at 52 ◦ C for at least 48<br />

hours prior to using. The <strong>DPPC</strong>-heptanol total molar ratio was approximately equal<br />

to 6:1. The hexanol solution and the heptanol-lipid suspension were degassed prior to<br />

being loaded into the 1.3 mL reaction cell and the 100 µL syringe, respectively. After<br />

equilibrating the system to 25 ◦ C, 4 injections of 5 µL are titrated into the cell.<br />

A typical example of the ITC experiments is shown in figure 3.2.<br />

C cell<br />

> C syringe<br />

C cell<br />

< C syringe<br />

21<br />

20<br />

20.5<br />

19.5<br />

19<br />

µcal/sec<br />

20<br />

µcal/sec<br />

18.5<br />

19.5<br />

18<br />

17.5<br />

19<br />

0 500 1000 1500<br />

Time (sec)<br />

17<br />

0 500 1000 1500<br />

Time (sec)<br />

(a) An endothermic reaction, indicated<br />

by the peaks pointing upwards. The free<br />

concentration of hexanol in the syringe<br />

is lower than the hexanol concentration<br />

in the cell, see section 3.1.<br />

(b) An exothermic reaction, indicated by<br />

the peaks pointing downwards. The free<br />

concentration of hexanol in the syringe<br />

is higher than the hexanol concentration<br />

in the cell, see section 3.1.<br />

Figure 3.2 Raw ITC data from two scans of 10.5 mM hexanol in 45.5 mM <strong>DPPC</strong> suspension in<br />

the syringe and a hexanol concentration in the cell of 1 mM and 10 mM, respectively.


36 Experimental Techniques<br />

3.2 DSC<br />

Differential Scanning Calorimetry (DSC) is a measuring technique to determine e.g.<br />

specific heat capacity, phase transition temperatures or the associated change in specific<br />

enthalpy.<br />

The scan rate ( T) ˙ is constant during the experiment and heat flowing ( ˙Q) into the<br />

sample is measured as a function of temperature (T ). During a first order transition the<br />

sample temperature remains constant during the whole transition, still the heating power<br />

is continuously supplied. The isobaric heat capacity ( dH<br />

dT<br />

) at this point is (infinitely) high<br />

and results in a jump on a (T, ˙Q) graph. But since the scan rate is not infinitely low,<br />

generally, the system does not have enough time to equilibrate and the jump becomes<br />

a slope or peak instead, resulting in the characteristic DSC scan-graphs, where the heat<br />

flow is plotted as a function of the temperature.<br />

The DSC scans were run prior to the SAXS experiments to get an idea of where<br />

to find the phase transition temperature for the interdigitated gel phase to the liquid,<br />

crystalline phase transition. Two different DSCs have been used for the experiments, a<br />

scal-1 microcalorimeter produced by Scal Co, Pushchino, Russia, and a DSC7 produced<br />

by Perkin Elmer, USA.<br />

The compensating DSC<br />

In general there are three different types of DSC instruments; heat-exchanging calorimeters,<br />

temperature-changing calorimeters and compensating calorimeters [22]. Amongst<br />

the compensating calorimeters the power compensating and the heat compensating DSC<br />

are found. The name ”compensating” refers to the fact that the temperature changes of<br />

the sample due to e.g. a phase transition are compensated by adding or eliminating an<br />

equally high, well-known amount of energy in the form of heat or electric energy of the<br />

opposite sign. In an ideal compensating DSC each ∆T signal appearing between sample<br />

and reference sample would immediately be compensated by a corresponding change in<br />

the heating power [27].<br />

The adiabatic scanning calorimeter<br />

The scal-1 microcalorimeter is a differential adiabatic 1 scanning calorimeter with a cylindertype<br />

measuring system. In this calorimeter the heating power is supplied according to a<br />

given, preset temperature program, and the compensation of heat is measured with the<br />

aid of electric energy [27]. The measuring system, the two sample containers, with the<br />

temperature sensors determine the temperature inside the system. When the measuring<br />

system is ideally symmetrical, equally high heat-flow rates flow into the cells; the differential<br />

temperature signal is then zero. If this steady-state equilibrium is disturbed by a<br />

sample transition, a differential signal is generated which is proportional to the difference<br />

between the heat-flow rates flowing to the sample and to the reference. Heat losses are<br />

minimized by adapting the temperature of the surroundings as well as possible to the<br />

temperature of the sample. In practice the heating power (the feed heat) is often given<br />

and the resulting heating rate (determined by the temperature difference between the<br />

two cells) is measured. Heat losses are minimized by the adiabatic jacket [22].<br />

In the calorimeter the reference and the sample calorimetric cells are made of glass,<br />

since glass has a high chemical resistance to the majority of chemicals, and the calorimeter<br />

is designed for studying liquids of various compositions. The internal and external<br />

1 adiabatic refers to a prevention of the heat exchange between measuring system (sample) and surroundings<br />

[7].


3.2 DSC 37<br />

thermostat shield and the cold thermostat are cylindrically shaped, made of aluminium<br />

alloy and placed in a hermetic housing, see figure 3.3.<br />

Sample cell<br />

Reference cell<br />

External thermostatic shield<br />

Internal thermostatic shield<br />

Temperature difference sensor<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000<br />

1111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

0000000000000000000000000000000<br />

1111111111111111111111111111111<br />

Cold thermostat<br />

Heating elements for internal<br />

and external shield<br />

Heating element for calorimetric cells<br />

Temperature difference sensor<br />

Temperature difference sensor<br />

Figure 3.3 A schematic drawing of the scal-1 microcalorimeter. The reference and sample<br />

cells are made of glass and they are placed in a low-melting alloy. The internal and external<br />

thermostat shield and the cold thermostat have the shape of cylinders, are made of aluminium<br />

alloy and placed in a hermetic housing. Adapted from Senin et al. [67]<br />

Before the measurements begin the sample and the reference are degassed for 5 minutes<br />

prior to being aspirated into the syringe. The syringe containing the 500 µL aqueous<br />

lipid-alcohol suspension is immersed into the cell. The reference cell is filled with degassed<br />

water. The cells are filled at room temperature and the calorimeter starts the<br />

cooling process. When the temperature in the cells is about 6 ◦ C, the instrument switches<br />

automatically to controlled heating. The internal and external thermostatic shields will<br />

have the same temperature as the cells. The heating power of the two cells is set by<br />

a voltage source, and this power is proportional to the rate of their controlled heating.<br />

The rate of the controlled heating is linear through the whole temperature range. The<br />

temperature difference between the two cells is controlled by the temperature difference<br />

sensors, and a compensation regulator distributes the measured current via other thermopiles<br />

in order to keep the temperature difference between the two cells at a minimum.<br />

The compensating current is detected and used in the further processing. When the<br />

temperature in the cells reaches the preset value, the heaters of the shields are switched<br />

off and the instrument turns to uncontrolled cooling. The shields and the cells are cooled<br />

to the temperature of the cold thermostat, which is kept at a temperature of ∼ 0 ◦ C.<br />

When the temperature of the internal thermostatic shield is 6 ◦ C, the controlled heating<br />

is switched on again [67].<br />

The instrument can be programmed to measure several identical scans successively,<br />

which makes it possible to check the precision of the scans as well as a possible time<br />

dependence. The measurements are performed at a pressure of 3 atmospheres and a scan<br />

takes approximately 3 hours 2 . Using different scanrates result in different baselines, see<br />

2 At a scanrate of 1 K/min.


38 Experimental Techniques<br />

appendix F.<br />

Power compensating DSC<br />

The DSC7 is a disk type power compensating DSC with an isoperibolic 3 mode of operation.<br />

The measuring system consists of two platinum-iridum alloy microfurnaces, which<br />

contain one temperature sensor and two heating elements. Both furnaces are positioned<br />

in an aluminium block kept - thermally decoupled - at a constant temperature, the<br />

measuring range extends from -90 ◦ C 4 to 725 ◦ C.<br />

Into each of the two furnaces a small pan of aluminium containing the sample and<br />

the reference, respectively, is mounted. The sample pans are made of aluminium as<br />

they need to have a very high thermal conductivity. Using aluminium pans lower the<br />

measuring area from 725 ◦ C to lower than 660 ◦ C, the melting temperature of aluminium.<br />

The microfurnaces are be covered with lids of platinum-iridium.<br />

Sample<br />

Reference<br />

*<br />

* * * *<br />

* * *<br />

o o o o o<br />

o o o o<br />

P T T P<br />

S S R R<br />

T(t) − program<br />

P − control<br />

T − amplicification<br />

m P T<br />

Figure 3.4 A schematic drawing of a power compensating DSC. Each furnace is provided with<br />

heating elements and a temperature sensor. φ m is the heat flow and P is the power. Adapted<br />

from Hemminger [22].<br />

A control circuit supplies the same heating power to both furnaces in order to change<br />

their mean temperature according to the preset temperature-time program. In case of<br />

ideal symmetry (the sample and reference are similar) the temperature of both furnaces<br />

is the same. When a sample reaction takes place, the symmetry breaks and causes a<br />

temperature difference between the two microfurnaces.<br />

This temperature difference is at the same time the measured signal and input signal of<br />

a second proportional control circuit, which tries to compensate the reaction heat-flow<br />

rate by increasing or decreasing an additional heating power. The compensating heating<br />

3 isoperibol refers to constant temperature surroundings with the temperature of the measuring system<br />

possibly differing from this.<br />

4 the block could reach -175 ◦ C if it was cooled with fluid nitrogen.


3.2 DSC 39<br />

power is proportional to the measured temperature difference. The time integral over<br />

the compensating heating power is proportional to the heat, which was consumed or<br />

released in the sample[22]. This information is sent to an output device, a computer,<br />

and results in a plot of the heat flow between the reference and sample cell as a function<br />

of temperature. Whether the peak is above the baseline, positive, or below the baseline,<br />

negative, depends on whether it is an endothermic or exothermic process.<br />

35<br />

<strong>DPPC</strong> : hexanol<br />

Scanrate: 5 K / min<br />

Heat flow [mW]<br />

30<br />

25<br />

20<br />

10 20 30 40 50 60<br />

Temperature [°C]<br />

Figure 3.5 A DSC scan of <strong>DPPC</strong> with hexanol, ratio 1:2. Scanrate 5 K/min. Perkin Elmer.<br />

Two separately controlled electrical heaters are connected to each microfurnace. Heating<br />

block A supplies both cells with a constant heat flow, depending on the scanrate. If<br />

the content of the two cells is identical then a plot of the temperature as a function of<br />

the feed heat would result in a straight line without any peaks [22], and heating block<br />

B will not contribute to the total heat supply to the cells. If the content of the cells<br />

differ, then heat block A will supply both cells with the constant specific amount of heat<br />

defined by the chosen scanrate and during e.g. a phase transition in the sample cell heat<br />

block B will supply the sample cell with heat equivalent to the heat of transition. The<br />

two heating blocks, A and B, are decoupled and serve basicly different purposes.<br />

Sample<br />

Reference<br />

* *<br />

* * * *<br />

*<br />

A<br />

B<br />

*<br />

*<br />

T<br />

O O<br />

O O O<br />

O O<br />

A<br />

B<br />

O O<br />

O O<br />

Figure 3.6 A rough sketch of the heating system i the DSC7, Perkin Elmer.<br />

For the DSC7 instrument small pans measuring 5 mm across are used. 30 µL of the<br />

sample is placed in the sample pan by using a pipette, a lid is placed on top of the pan


40 Experimental Techniques<br />

and the lid is pressed down the pan using a specially designed pan-sealer, see figure 3.7.<br />

Figure 3.7 A photo of the lid, the pan before and after the sealing. During the sealing the outer<br />

edges of the lid and the pan are cut off before being used in the DSC7.<br />

Afterwards the pan is placed in the sample cell. The reference consists of a matched<br />

aluminium sample pan containing distilled water. It is placed in the reference cell of the<br />

instrument. A scan takes between 5 minutes and 30 minutes depending on the chosen<br />

scanrate. During the experiments the pressure has not been regulated, the experiments<br />

have been performed at ambient pressure. In appendix F there is a figure showing the<br />

baselines at different temperatures.


3.3 SAXS 41<br />

3.3 SAXS<br />

The Kratky camera is the most widely used long-slit block camera. The most characteristic<br />

feature of the camera is the block collimation. The original camera was designed by<br />

Otto Kratky, hence the name Kratky-camera, in the 1950s and redesigned in the 1970s<br />

[15]. The Kratky camera at IMFUFA, see figure 3.8, is a modified Compact Camera<br />

produced by M. Braun-Graz Systems Gmbh. The measuring equipment is situated in<br />

a thermostatically controlled, lead isolated room with a constant room temperature of<br />

20 ◦ C. This camera is well-suited for measurements on randomly oriented molecules and<br />

objects in solution, and covers a range of scattering vectors q down to 0.025 Å −1 .<br />

Figure 3.8 A schematic drawing showing the most important components of the Kratky compact<br />

camera. From the detector to the X-ray tube, the camera measures about 55 cm, (drawing<br />

adapted from [11]).<br />

The camera<br />

The camera is mounted on top of the source, the X-ray tube. The camera head, beam<br />

length and beam width can be adjusted easily when calibrating by turning the knobs<br />

shown in figure 3.8. To reduce air scattering the chamber is evacuated (P < 0.3 Torr<br />

≈ 2 mPa). As a safety mechanism the shutter only opens when the pressure is low.<br />

This ensures that the lead glass plate on top of the camera is mounted air-tight after a<br />

possible maintenance operation or adjustment carried out inside the chamber. A beam<br />

stop is situated just in front of the detector to protect it from the direct beam.<br />

Source<br />

The source is a sealed X-ray tube (Phillips Fine line) which is cooled by circulating water.<br />

The X-ray generator (Phillips PW 1729) is for our measurements provided with a voltage<br />

of 40 kV and current of 40 mA. As the cathode is heated, the electrons are vapourized<br />

and accelerated towards the copper anode, called the target. The electrons interact<br />

with the target in two ways in order to create X-rays: bremsstrahlung and fluorescence.<br />

Bremsstrahlung (or “braking radiation”) occurs as the electrons are decelerated in the<br />

electric fields of the nuclei. The energy loss is emitted as energetic photons, which form a


42 Experimental Techniques<br />

continuous spectrum. The other source of X-rays is fluorescence. The incoming electrons<br />

excite the bound electrons of the target. When they decay, photons are emitted. This<br />

radiation spectrum consists of the characteristic lines of copper. The copper spectrum<br />

shows primarily radiation from the two electron transitions n = 2 → n = 1, (K α , λ = 1.55<br />

Å), n = 3 → n = 1, (K β , λ = 1.41 Å), see figure 3.9.<br />

Log (Intensity)<br />

K α<br />

K β<br />

Bremsstrahlung<br />

Energy<br />

Figure 3.9 The low energy radiation K α originates from the transition from the L-shell to the<br />

K-shell of the copper atom, whereas the high energy radiation K β originate from the transition<br />

from the M-shell to the K-shell.<br />

Monochromator<br />

Generally, the intensity of a beam decreases as it passes through an absorbing medium. 5<br />

The absorption spectrum of nickel has a top between Cu-K α and Cu-K β , thus inserting<br />

the nickel filter dampens the Cu-K β peak strongly, and the Cu-K α peak lightly, resulting<br />

in a nearly monochromatic beam.<br />

Block collimation<br />

In the camera, the primary beam is guided by three bodies, two blocks and one edge<br />

[40]. The edge is situated as an entrance slit, S, in front of the X-ray tube and controls<br />

the height of the beam. The entrance slit, the middle block, B 1 , and the second block,<br />

B 2 , also called the bridge, have to be accurately aligned with respect to the centerline,<br />

H, and in the direction perpendicular to the paper. If this is done correctly there will<br />

be no parasitic (secondary) scattering above the centerline, H [8]. The block collimation<br />

principle of the camera is shown in figure 3.10.<br />

In figure 3.10(c) and figure 3.10(d) incorrect setups of the Kratky camera are shown. In<br />

figure 3.10(c) B 2 is placed too low, resulting in scattering by edge k 2 which illuminates<br />

region S, whereas in figure 3.10(d) B 2 is placed too high, resulting in parasitic scattering<br />

above the primary beam. In the correct setup of the Kratky camera, the primary beam, p,<br />

forms a triangle shaped intensity distribution in the vertical direction in the registration<br />

plane, R.<br />

5 In a linear, spatially isotropic, absorbing medium the intensity as a function of distance is I(x) =<br />

I 0 exp[−µx], where I 0 is the initial intensity (at x = 0) and µ is the linear absorption coefficient (as<br />

explained in Chapter 2).


3.3 SAXS 43<br />

(a) A schematic view of the collimation system.<br />

The bridge B 2 is put on top of the<br />

side-bars of the U -shaped body the central<br />

part of which is the middle block B 1 .<br />

(b) The correct arrangement of the block<br />

collimation camera. The section along the<br />

middle axis of the camera, perpendicular to<br />

the plane of the primary-beam. The vertical<br />

scale is multifold stretched compared to<br />

the horizontal one. S is the entrance slit,<br />

B 1 and B 2 are the two blocks used for collimation.<br />

R is the plane of registration and<br />

p is the primary beam.<br />

(c) Incorrect Kratky setup as the second<br />

collimation block, O 2 , is too low.<br />

(d) Incorrect Kratky setup as the second<br />

collimation block, O 2 , is too high.<br />

Figure 3.10 The principle of block collimation. Note that right and left is reversed in<br />

comparison to figure 3.8. (All four schematic drawings are from [40])<br />

The block collimation increases the intensity of the beam, but it also increases the smearing<br />

(see figure 5.2) of the beam. 6<br />

Sample holder<br />

The sample holder, shown in figure 3.11, is placed inside the middle of the Kratky camera.<br />

It is temperature controllable (heated from below) and holds the cuvette (glass tube) with<br />

the sample perpendicular to the beam.<br />

The sample holder can be adjusted by turning the knobs. It is possible to adjust the<br />

cuvette both perpendicular and parallel to the beam (height and tilt angle). The best<br />

arrangement of the sample is found by adjusting and subsequently measuring. As the<br />

6 Smearing of the beam is one of the biggest problems using line focusing cameras instead of pinhole<br />

cameras, the topic will be discussed further in section 5.2.


44 Experimental Techniques<br />

Figure 3.11 A schematic drawing of the sample holder (from [11]).<br />

largest intensity is found, the sample holder is kept at this position during the experiments.<br />

The dimensions of the glass tube and its holder are shown in figure 3.12.<br />

Figure 3.12 The dimensions of the cuvette (the sliding gauge is in millimetres). The lids of<br />

both ends can be unscrewed in order to inject the sample fluid.<br />

Detector<br />

The detector is a MBraun Position Sensitive Detector which detects both energy and<br />

position of the scattered photons. The detector consists of a platinum wire encased in a<br />

chamber filled with an argon/methane mixture kept at about 7 to 8 bar. As the photons<br />

enter the chamber, they ionize the gas atoms, and the electrons from this ionization<br />

process are accelerated towards the anode, which consists of wires kept at a potential of<br />

about 3.6 kV relative to ground.<br />

In the close vicinity of the anode wire the electrons gain enough energy to start an<br />

electron avalanche process, resulting in an induced charge distribution on the cathodes,<br />

which amplify the charge signal [87]. The size of this signal is proportional to the energy<br />

of the photon [2]. The distribution of the energy of the incoming photons and the energy<br />

window used is shown in figure 3.13.<br />

The position of the photon is detected by a backgammon technique, see figure 3.14,<br />

in which this technique and the main components of the detector are shown.<br />

In linear counters the anode wire usually has a high resistance and the position of<br />

the event can e.g. be determined by the division of the charge collected at the two ends.<br />

The detector systems form an RC circuit and due to the dependence of the resistance on


3.3 SAXS 45<br />

2<br />

1.5<br />

Energy spectrum<br />

Lorentz distribution<br />

Counts per second<br />

1<br />

0.5<br />

0<br />

200 300 400 500 600 700<br />

Channel number<br />

Figure 3.13 The energy spectrum for the multilamellar <strong>DPPC</strong> measurement. The vertical<br />

lines frame indicates the energy window used. The red line is a Lorentz distribution,<br />

L = 1.53 · [( ch.nr.-459 ) 2 + 1], in which 459 specifies the location of the peak of the distribution.<br />

92<br />

the position the rise time of the signals depends on position. By appropriate electronic<br />

processing of the signals the position can be determined and the energy can be obtained<br />

from the sum of the two signals [8].<br />

Calibration<br />

q-calibration<br />

When performing a scattering experiment the outcome is the scattering intensity as a<br />

function of the detector’s channel numbers. To convert the detector channel number into<br />

a q-vector value, a q-calibration is made. The scattering intensity is measured along the<br />

anode wire and picked up in 1024 channels. The calibration is made with silver behanate<br />

(AgBH), chosen because of its well-defined repeat distance, 58.380 Å [26]. As seen in<br />

figure 3.15 there is a repeating pattern in the SAXS spectrum of silver behanate.<br />

Combining this information with the measurements of the direct beam and Bragg’s<br />

law, equation 2.44, it is possible to find the q-value as a function of the channel number.<br />

Reading the value for the direct beam and the observed Bragg peaks gives<br />

q(c) = 7.914 · 10<br />

−4 Å−1<br />

ch.no. · c − 0.185Å−1 (3.1)<br />

where q can be found as a function of the channel number. This calibration is used<br />

in the further data processing.


46 Experimental Techniques<br />

(a) Subfigure a) shows a schematic drawing<br />

of the non-encoding cathode (1), the<br />

anode wire (2) and the encoding cathode<br />

(3), and subfigure b) shows the backgammon<br />

cathode. The broken circle indicates<br />

a spot of the induced charge.<br />

(b) A schematic diagram of the electronic<br />

setup of the detector.<br />

Figure 3.14 An overview of the electronic equipment inside the detector. From [87].<br />

3.5 x 104 Channel number<br />

3<br />

2.5<br />

2<br />

Intensity (a.u)<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

0 200 400 600 800 1000 1200<br />

Figure 3.15 The measured intensity of scattering from AgBH.


3.3 SAXS 47<br />

Temperature calibration<br />

The sample is heated from below, which leads to a temperature gradient from the heater<br />

to the sample, and by calibrating the system a connection between the temperature set<br />

by the machine to the absolute temperature of the sample is found.<br />

The temperature calibration was done with a negative temperature coefficient (NTC)<br />

(semiconductor) resistor, of which the resistance was calibrated as a function of the<br />

absolute temperature with an error of 0.1 K using a Ametek D55SE 7 . The temperature<br />

dependence (given a certain temperature range) of the resistance of the semiconductor is<br />

a constant times a Boltzmann factor with a well-known activation energy, E ∼ 0.297 eV<br />

(suggested by the manufacturer), see figure 3.16.<br />

( ) E<br />

R(T) = R ∞ exp<br />

(3.2)<br />

k B T<br />

R(T) of NTC resistor<br />

12<br />

10<br />

Calibration of NTC resistor<br />

Fit: R(T) = 0.105 · exp( 0.295 / (k B<br />

T))<br />

Resistance [kΩ]<br />

8<br />

6<br />

4<br />

2<br />

0<br />

20 30 40 50 60<br />

Absolute temperature [°C]<br />

Figure 3.16 The resistance of the NTC resistor as a function of the absolute temperature.<br />

It is worth noting that, unlike metals, the resistance of the semiconductor decreases<br />

as the temperature increases (hence the name, NTC). This is because the effect of an<br />

increased amount of charge carriers in the conduction band dominates the effect of electrical<br />

resistance due to the scattering of charge carriers by the ion lattice - the increase<br />

of charge carriers giving a positive (and the latter giving a negative) contribution to the<br />

electrical conductivity.<br />

The NTC resistor was placed inside the cuvette and connected to an Ohm-meter<br />

outside the Kratky camera. Different calibration measurements were performed, varying<br />

the content of the cuvette and with or without evacuated chamber. The cuvette contained<br />

either air or light-fluid silicone oil. Light-fluid silicone oil was chosen over water, as we<br />

did not dare to test insulatory properties of the coating of the NTC resistor and the<br />

7 Jofra Instruments, DK-3520 Farum


48 Experimental Techniques<br />

wiring. We waited about 30 minutes between the measurements and it seemed enough<br />

to obtain a steady state (settled temperature).<br />

Temperature calibration, T(T S<br />

)<br />

60<br />

Absolute temperature [°C]<br />

50<br />

40<br />

30<br />

20<br />

Airfilled cuvette<br />

Fittet line (linear regression):<br />

T = 0.847T S<br />

+ 45.0 K<br />

Cuvette with silicone oil<br />

Fittet line (linear regression):<br />

T = 0.794T S<br />

+ 60.5 K<br />

20 30 40 50 60 70<br />

Set temperature [°C]<br />

Figure 3.17 Temperature calibration with evacuated chamber and a cuvette filled with either<br />

air or silicone oil. The correlation coefficient for □ is 0.9999754, and the correlation coefficient<br />

for ∗ is 0.9999929.<br />

The correlation between the absolute and the set temperature was found to be linear,<br />

see figure 3.17. The figure shows that the presence of silicone oil causes a smaller slope<br />

because of a higher thermal conductivity.<br />

Matlab Code<br />

☞ page 180 T(T S ) air = 0.847T S + 45.0 K<br />

T(T S ) oil = 0.794T S + 60.5 K<br />

The difference between the set temperature and the measured temperature is especially<br />

distinct at high temperatures where a set temperature of 50 ◦ C corresponds to an<br />

actual temperature of about 44 ◦ C.


3.4 Sample preparation 49<br />

3.4 Sample preparation<br />

First, here is a short version of the experimental procedure for preparation of ULV. Later<br />

in this section some of the steps in the procedure are elaborated.<br />

Preparation of MLVs The lipid powder 8 is weighed directly into the glass vial (8 mL)<br />

and boiled, lukewarm milli q water 9 is added. A 5 weight percentage (w%) is<br />

endeavoured. The sample holder is sealed with parafilm. Using an Eppendorf<br />

Thermomixer comfort the sample is consistently shaken and held at a constant<br />

temperature, 10 degrees above the phase transition temperature, for two hours.<br />

The liquid sample is now usually homogeneous and milk-white, if not so the sample<br />

is exposed to ultrasound for 10 minutes also at 10 degrees above the phase transition<br />

temperature. The ultrasound procedure was seldom required.<br />

Preparation of ULVs, extruding By following the directions given in the “Lipex Extruder”<br />

manual the extruder is assembled. A combination of filters is used, first<br />

a polycarbonate filter, then a drain filter, then a polycarbonate filter and finally<br />

another drain filter. The filters are wet by milli q water and water is extruded<br />

three times at a pressure of 20 bar before the sample is extruded. The sample is<br />

extruded 10 times at 20 bar, the glass vial and the pipette are changed during the<br />

last three extrusions, for further explanations see later in this section.<br />

Refining the sample, centrifugation After the extrusions the sample is centrifuged. The<br />

sample is rotated by 19.500 rounds per minute at a temperature of 20 ◦ C for 20<br />

minutes. After the centrifugation the sample is separated into 3 phases, the top<br />

phase is water, the middle phase is unilamellar <strong>vesicles</strong> and the bottom phase is<br />

multilamellar <strong>vesicles</strong>. By use of a pipette the two top phases are transferred to<br />

a glass vial. In [59, p. 70] it was found that this procedure removes MLVs which<br />

have survived the extrusion such that the number of scattered photons from MLVs<br />

are halved, approximately.<br />

Concentration of lipid suspension<br />

For the gravimetric determination of the lipid concentration 4 small metal pans (diameter<br />

of 5 mm) are numbered and weighed 4 times each. Each of the 4 times 4 weights are<br />

noted. A drop of the ULV-solution is placed at one of the cups and a stop watch is<br />

started and the weight is noted every 15 seconds for two minutes. This procedure is<br />

repeated for the last three cups. Finally the 4 cups are placed in an oven at 50 ◦ C. After<br />

a couple of days the 4 dehydrated cups are weighed 4 times each, again. The weights are<br />

noted. By plotting these data it is possible to determine the lipid concentration. The<br />

linear correlation between the time and the weight of the lipid drop is extrapolated to<br />

zero, see figure 3.18. The weight at t=0 is equivalent to the weight of the lipid drop.<br />

By combining these informations the concentration of the lipid suspension is found.<br />

Even if a lipid concentration of 5 % is endeavoured, the centrifugation process makes it<br />

impossible to end up with the same lipid concentration in each experimental series. The<br />

concentrations vary from 1.62 (first experimental series) to 4.59, the average standard<br />

being around 3.4, see appendix D.<br />

The aqueous, pure lipid sample is kept in a container, shaken constantly at 10 degrees<br />

above the phase transition temperature until the alcohol is added. A SAXS experiment<br />

is performed with the pure lipid sample.<br />

8 Purchased from Avanti Polar Lipids Inc. www.avantilipids.com<br />

9 Arium® 611 Ultrapure Water Systems from Sartorius AG


50 Experimental Techniques<br />

52.2<br />

Raw data<br />

y=-0.006 (mg/sec)x+52.21 (mg)<br />

52<br />

m [mg]<br />

51.8<br />

51.6<br />

51.4<br />

50 100<br />

Time [sec]<br />

Figure 3.18 An example of the determination of the gravimetric determination of the lipid<br />

concentration. The weight is plotted as a function of time. The measurement after 15 seconds<br />

might not be as precise as the other 7 measurements as it takes 10-20 seconds for the weighing<br />

machine to stabilize.<br />

Adding alcohol<br />

The amount of alcohol added depends on the experiment performed, although the lipid<br />

to alcohol ratio is the same, two alcohol molecules to one lipid molecule. This ratio was<br />

chosen on the basis of the literature values in table 1.1. The alcohol to lipid ratios are<br />

highly dependent on the lipid concentration, the lower the lipid concentration the higher<br />

alcohol to lipid ratio is needed to induce the interdigitation. As to the concentrations<br />

of lipid used in this work only one experiment can be used in strict comparison. Zhang<br />

& Rowe [88] use a lipid concentration of 70.8 mM and report interdigitation from the<br />

n-alcohol butanol at a 1:5 total molar ratio. Using this information along with the<br />

correlation reported by Löbbecke & Cevc [47], see section 1.4, the <strong>DPPC</strong> to pentanol<br />

total molar ratio should be more than 1:1.7. And as less and less alcohol is needed to<br />

induce interdigitation the longer the n-alcohols are, a fixed total molar ratio of 1:2, <strong>DPPC</strong><br />

to alcohol, was used.<br />

Different amounts of alcohol are needed for the pentanol, hexanol and heptanol experiments<br />

respectively. The procedure is the same, though. An Eppendorf tube is placed<br />

on the weight and by using a syringe a specific amount of alcohol 10 is added to the side<br />

of the tube and the weight is noted. Afterwards the tube is filled with the sample and<br />

the final weight is noted. The tube is sealed with parafilm.<br />

The Eppendorf tube with lipid and alcohol is placed in the tempered shaker and is<br />

shaken for at least 48 hours before the SAXS experiment is performed. The temperature<br />

of the shaker is subsequently set to 52 ◦ C, 10 ◦ C above the main phase transition<br />

temperature.<br />

10 Purchased from Aldrich-Chemie, D-7924 Steinheim


3.4 Sample preparation 51<br />

Vesicle extrusion and Polydispersity<br />

The extrusion procedure primarily involves pushing a solution of large MLVs through<br />

polycarbonate membrane filters several times. The polycarbonate filters have roughly<br />

cylindrical pores with a diameter of 100 nm and a length of 6 µm. By pushing the MLVs<br />

through the pores, they break up into smaller <strong>vesicles</strong>. The extrusion process can be<br />

explained in terms of blowing bubbles through pores, see figure 3.19.<br />

A<br />

P 1<br />

B<br />

P 1<br />

P 2<br />

P 0<br />

l<br />

d<br />

C<br />

P 0<br />

Figure 3.19 (A) Multilamellar <strong>vesicles</strong> are extruded through the polycarbonate membrane filters<br />

resulting in unilamellar <strong>vesicles</strong> whose size is comparable to the pore size, see figure 3.20. l is<br />

the length and d is the diameter of pores in the polycarbonate filter. (B) A small fragment<br />

of the vesicle is pulled into the pore. This fragment of the vesicle has the same radius as the<br />

pore, whereas the large fragment outside the pore has a radius comparable with the original<br />

radius of the vesicle. (C) Schematic diagram of the vesicle extrusion. The force due to the<br />

applied pressure is balanced by a force caused by line tension around the neck of the vesicle.<br />

The difference between P 0 and P 1 is the applied pressure, P 2 is the pressure inside the vesicle.<br />

Adapted from [58].<br />

Patty & Frisken [58] have modelled the vesicle size (radius), R v as a function of pore<br />

size (radius), R p and the extrusion pressure, P .<br />

√ √<br />

Rp γ l<br />

R v (P) = A<br />

2P + B or R v γl<br />

= a + b (3.3)<br />

R p 2PR p<br />

where γ l is the lysis tension and A and B (or a and b) are fitting parameters. The model<br />

agrees with their experimental results, which unfortunately is restricted to 1-palmitoyl-<br />

2-oleoyl-sn-glycero-3-phosphatidylcholine (POPC). POPC has asymmetric tail lengths,<br />

one of 16 and one of 18. Nevertheless, these results were used since it was not possible<br />

to find similar results for <strong>DPPC</strong> (but the reader should keep this in mind). For POPC<br />

they find a = A = 0.61 ± 0.04, b = B R p<br />

= 1.09 ± 0.02 and γ l = (7.4 ± 0.04) mN/m.<br />

The vesicle size does not change much as the extrusion pressure is increased from 20<br />

to 30 bar<br />

R v (P = 20 bar) ≈ 604 Å and R v (P = 30 bar) ≈ 593 Å (3.4)<br />

using equation 3.3 and R p = 50 nm. In our preparation procedure, we have certainly<br />

stayed within this range, trying to stay at 20 − 21 bar and only higher pressure when we


52 Experimental Techniques<br />

Figure 3.20 The average size and polydispersity of the <strong>vesicles</strong> in a sample with a 5% lipid-water<br />

mass ratio calculated from the model in equation 3.3 and a reported 20% standard deviation<br />

[58].<br />

had trouble extruding (when nothing happened as we tried to extrude). The polydispersity<br />

of the vesicle size was determined by light scattering [58], and Patty & Frisken<br />

[58] find a standard deviation of about 20% for the smaller pore diameters (50 nm) and<br />

about 30% for the larger pore diameters (200 nm). According to the LIPEXExtruder<br />

Manual [44] the vesicle size should be ±20% of the filter pore size. This was verified by<br />

light scattering on their test system, egg phosphatidylcholine (EPC) <strong>vesicles</strong> 11 . But the<br />

manual does not say anything about the extrusion pressure used.<br />

If all ULVs are dispersed, the characteristic distance, d, between two <strong>vesicles</strong> can be<br />

estimated from<br />

d ULV = 3√ v = 3 √<br />

V<br />

N V<br />

(3.5)<br />

where v is the volume per vesicle, i.e. the fraction between the total volume V of the<br />

sample (lipid and water) and the number of <strong>vesicles</strong> N V . N V must be the ratio between<br />

the total number of lipid molecules, N l , and the number of lipid molecules per vesicle,<br />

N mpv<br />

N V = N l<br />

N mpv<br />

(3.6)<br />

11 The fatty-acid composition of egg lecithin prepared by the cadmium chloride method has about 30%<br />

of C16, 55% of C18 and 15% of C20 and C22 acids; approximately 39% and 45%, respectively, of the<br />

total acids were saturated [41].


3.4 Sample preparation 53<br />

d<br />

d<br />

1.6 µm<br />

Figure 3.21 The characteristic distance between two ULVs in a sample with a 5% lipidwater<br />

mass ratio, calculated using equation 3.9 with the following values: A l = 50 Å 2 [32],<br />

r ULV = 60 nm (see estimate, equation 3.4), m l = 255 mg (measured), V = 5.4 mL (measured),<br />

N A = 6.0221 · 10 23 mole −1 [85], M l = 734.1 kg/mole [32]. It is about 14 times larger than the<br />

vesicle diameter. The number of <strong>vesicles</strong> are about N V = 1.1 · 10 12 using equation 3.6.<br />

with a characteristic radius r ULV . The number of lipid molecules are<br />

N l = n l N A = m l<br />

M l<br />

N A (3.7)<br />

where n l , N A , m l and M l are the amount of lipid molecules, Avogadro’s number, the<br />

total lipid mass and the molar mass of the lipid, respectively. The number of molecules<br />

per vesicle can be estimated from the total (of a single) vesicle surface area, A, and the<br />

area pr. lipid molecule, A l ,<br />

N mpv = A A l<br />

= 2 · 4πr2 ULV<br />

A l<br />

(3.8)<br />

assuming that the inner and outer surface areas are roughly the same. Combining 3.5,<br />

3.6, 3.7 and 3.8 gives<br />

√<br />

8πr 2<br />

d ULV = 3 ULV V M l<br />

(3.9)<br />

N A m l A l<br />

In figure 3.21 the distances are sketched. This is of course a rough estimate, but there is<br />

no doubt that d ULV is much larger than the coherence length of the Cu − K α X-rays.


4 Results<br />

In this chapter we present the results obtained with ITC, DSC and SAXS. The results<br />

in section 4.2 and 4.3 are denoted with a series number, the details about this specific<br />

experimental series can be found in appendix D.<br />

4.1 ITC<br />

High sensitive isothermal calorimetry has become more and more popular for studying<br />

heats of reaction observed upon the incorporation or partitioning of molecules into membranes<br />

[34]. There are several examples in the literature of the use of ITC in determining<br />

different thermodynamical properties of the lipid-alcohol system [65], [73], [82], [83], [88].<br />

Experiments<br />

Eight different kinds of experiment series were conducted, two with pentanol, five with<br />

hexanol and one with heptanol. All the experiments were conducted at 25 ◦ C. The samples<br />

differed mutually in the lipid/alcohol ratio. A common characteristic of all the series<br />

was that at least 10 different runs were made per series varying the cell concentration<br />

of the cell, gradually approaching and passing the free concentration of alcohol in the<br />

syringe. Several introductory experiments were carried out prior to the eight experiment<br />

series represented. Blank injections of pure ULV of <strong>DPPC</strong> were made to measure the<br />

heat of dilution. The resulting heat release was insignificant compared to the heat of<br />

dilution from the alcohols.<br />

After conducting an experiment the heat of each peak is integrated to find the enthalpy<br />

originating from each injection. The integrated heat from the second injection is<br />

used in the further data processing. From each experiment the outcome was consequently<br />

one set of simultaneous data of the enthalpy and the concentration of alcohol in the cell.<br />

The integrated heats of reaction are plotted as a function of the alcohol concentration in<br />

the cell, and a linear function is used to describe the correlation. An example of such a<br />

correlation can be seen in figure 4.1.<br />

The data of all the eight series have been processed, and the results are found in<br />

appendix E. The calculated free alcohol concentration, C alc,free , equivalent to the intersection<br />

of the graph with the axis of abscissa, is used to calculate the molar partition<br />

coefficient. The amount of alcohol in the lipid is calculated on the basis of the total concentration<br />

of alcohol in the lipid/alcohol suspension and the free alcohol concentration:<br />

C alc, total = C alc,free + C alc,lipid<br />

The concentration of alcohol in the lipid, C alc,lipid , is then found by using the injected<br />

volume, called V lipid :<br />

55


56 Results<br />

250<br />

66.6 mM pentanol in 45.5 mM lipid<br />

200<br />

cal/mole of injectant<br />

150<br />

100<br />

50<br />

0<br />

−50<br />

−100<br />

Raw data<br />

y=30.6 x − 1600<br />

−150<br />

50 52 54 56 58 60 62<br />

Alcohol concentration in reaction cell/ mM<br />

Figure 4.1 The integrated heat contributions as a function of the concentration of alcohol in the<br />

reaction cell. The free alcohol concentration, equivalent to the intersection of the graph with<br />

the axis of abscissa, is found to be 53.9 mM. The total alcohol concentration in the syringe is<br />

66.6 mM. The concentration of alcohol in the lipid is then 66.6 − 53.9 = 12.7mM. Each point is<br />

equivalent to an experimental series.<br />

C alc,lipid = C alc,total − C alc,free<br />

C alc,lipid = n alc,lipid<br />

V lipid<br />

n alc,lipid = C alc,lipid · V lipid (4.1)<br />

The amount of alcohol in the lipid is found as the concentration of alcohol in the lipid<br />

suspension and the injected volume is known. The amount of lipid, n <strong>DPPC</strong> , free alcohol,<br />

n alc,free , and water, n water , is calculated in the same way:<br />

n <strong>DPPC</strong> = C <strong>DPPC</strong> · V lipid<br />

n alc,free = C cell · V lipid<br />

n water = V lipid · ρ water<br />

18.02<br />

The factor 18.02 is the molar mass of water. n alc,free is the concentration of alcohol<br />

in the cell. The partition coefficient can then be calculated, since it is defined as:<br />

K x = X a,b<br />

X a,w<br />

(4.2)


4.1 ITC 57<br />

where X a,b is the mole fraction of alcohol in the lipid bilayer given by<br />

n alc,lipid<br />

X a,b =<br />

n alc,lipid + n <strong>DPPC</strong><br />

and X a,w is the mole fraction of alcohol in water given by<br />

X a,w =<br />

n alc,free<br />

n alc,free + n water<br />

The partition coefficients for all the experiments are listed in table 4.1. The figures<br />

marked by stars refer to experiments with samples which have been kept at room<br />

temperature for 24 hours.<br />

n-Alcohol Lipid K x Lipid:Alcohol<br />

1 Pentanol (28.7 mM) <strong>DPPC</strong> (53.5 mM) 261 8.5:1<br />

2 Pentanol (66.6 mM) <strong>DPPC</strong> (45.5 mM) 225 3.5:1<br />

3 Hexanol (10.5 mM) <strong>DPPC</strong> (45.5 mM) 367 18:1<br />

4 Hexanol (15.1 mM) <strong>DPPC</strong> (54.3 mM) 2553* 4.8:1<br />

5 Hexanol (21.6 mM) <strong>DPPC</strong> (54.3 mM) 529 6.7:1<br />

6 Hexanol (22.6 mM) <strong>DPPC</strong> (54.3 mM) 749/ 3546* 5.3/3*:1<br />

7 Hexanol (31.3 mM) <strong>DPPC</strong> (53.5 mM) 986 3:1<br />

8 Heptanol (7.3 mM) <strong>DPPC</strong> (45.5 mM) 2166/1713* 9.4/10.3*:1<br />

Table 4.1 The first column represents the used alcohol and the related concentration in the<br />

syringe, the second column represents the lipid concentration, the third column the calculated<br />

partition coefficient and the fourth column represents the lipid to alcohol ratio within the lipid<br />

membrane. The literature values are found on page 59.<br />

There are two main issues related to the calculated partition coefficient, the phase<br />

dependence and the alcohol-chain dependence.<br />

During the experimental process some characteristic features of the different lipidalcohol<br />

systems were discovered. All the samples were kept at 52 ◦ C until used for the<br />

experiments. Normally, an experiment series were conducted within 10 hours, the results<br />

were reproduceable within this period of time. While the experiments were carried<br />

out, the samples were kept at room temperature. After keeping the samples at this<br />

temperature for 24 hours, some of the experiments were conducted again. These ITC<br />

results differed systematically from the results obtained the previous day.<br />

The ITC experiments for pentanol and hexanol revealed that the free concentration<br />

of alcohol in the lipid/alcohol suspension had decreased, entailing that a clearly larger<br />

amount of alcohol had partitioned into the lipid <strong>vesicles</strong> after these have been kept at 25 ◦ C<br />

for 24 hours. Hence for these measurements a larger partition coefficient was calculated<br />

(marked by (*) in table 4.1 and E.2). The significantly larger partition coefficient could<br />

be a sign of evaporation of the free alcohol from the sample or a sign of the appearance<br />

of another lipid phase. According to Zhang & Rowe [88] the partition coefficient for the<br />

L βI phase is lower than the partition coefficient for the L β ′ phase and much lower than<br />

the partition coefficient for the L α phase, see table 4.2. The values for K x in general<br />

demonstrate that the partition coefficient for the liquid crystalline phase is significantly<br />

larger than for the three other phases. It is a reasonable assumption that the lipid <strong>vesicles</strong><br />

are primarily in the interdigitated phase when the main experiments are conducted.<br />

This is also reflected in the measured partition coefficients, which are similar to the


58 Results<br />

values found in the literature. There is most likely a certain equilibrium between the<br />

amount of <strong>vesicles</strong> in the interdigitated phase and the amount of <strong>vesicles</strong> in the ripple<br />

phase, dominated by the interdigitated phase. After being kept at room temperature<br />

for 24 hours the part of <strong>vesicles</strong> in the L β ′ phase have increased. As the partition<br />

coefficient for alcohol into <strong>vesicles</strong> of this type is higher than the partition coefficient<br />

for the interdigitated phase, the total measured partition coefficient consequently will be<br />

higher.<br />

Löbbecke & Cevc [47] allowed the samples of <strong>DPPC</strong> and alcohol to equilibrate at<br />

room temperature for more than a week, at least, to make sure the lipids were in the<br />

interdigitated phase. In our work a DSC scan was run on samples, which had been<br />

equilibrating for a week. These scans were not a source of concern as the scans did not<br />

differ from the scans obtained the previous week. For this reason the samples have only<br />

been equilibrating for 48 hours.<br />

The ITC results for the heptanol samples kept at room temperature for 24 hours<br />

differed from those obtained with pentanol and hexanol samples. These results indicated<br />

that the free heptanol concentration in the heptanol/<strong>DPPC</strong> suspension increased and<br />

hence the partition coefficient for the <strong>DPPC</strong>/heptanol system decreased. Still the partition<br />

coefficient for heptanol into <strong>DPPC</strong> is much larger than the comparable experiments<br />

with hexanol and pentanol. But the lower partition coefficients obtained after keeping<br />

the samples at room temperature for 24 hours indicate a difference in the interactions<br />

between this alcohol and <strong>DPPC</strong> compared to the lipid interactions with pentanol and<br />

hexanol. This could indicate a quite long equilibration time for heptanol into the <strong>DPPC</strong><br />

vesicle. Only one example of comparable experiments with lipids and heptanol have<br />

been found in the literature [30]. These measurements were conducted on multilamellar<br />

<strong>vesicles</strong> and in the liquid crystalline phase, both differing from the conditions used in<br />

this work. Nevertheless the values are of the same order of magnitude. The partition<br />

coefficients are in accordance with the partition coefficients found in the literature, see<br />

table 4.2. The values for the partition coefficient obtained in this work are lower than<br />

those obtained by Kamaya et al. [30], primarily owing to the fact that the partition coefficients<br />

are obtained in the L α phase and the fact that they use MLVs as opposed to<br />

ULVs. Using MLVs instead of ULVs gives rise to a larger partition coefficient [73]<br />

In general the partition coefficient for a specific phospholipid system increases with<br />

increasing alcohol length [73]. This tendency is also reflected in the results obtained in<br />

this work. The partition coefficients averaged increase by a factor 3 as a CH 2 group<br />

is added to the n-alcohol. This observation is in relation to values for the partition<br />

coefficient obtained by Katz & Diamond [33], Kamaya et al. [30] and Ly & Longo [46].<br />

These groups obtained values for the partition coefficient of DMPC, SOPC, and <strong>DPPC</strong>,<br />

respectively. All series of partition coefficients increased approximately by a factor 3 as<br />

the chain length of the n-alcohols was increased by a methyl group. These relations all<br />

follow Traube’s Rule 1 as it is interpreted by Ly & Longo [46]. Ly & Longo [46] state that<br />

the difference between measuring values for e.g. propanol, butanol and pentanol roughly<br />

follows Traube’s Rule of a factor 3 indicating that alcohol partitioning actually can be<br />

described by Traube’s rule. The increase in partition coefficient with alcohol chain length<br />

can also be related to the solubility of the alcohol in water. The longer the alcohol, the<br />

lower the solubility and the higher the hydrophobia.<br />

1 Traube’s Rule predicts that for each CH 2 group in an alcohol molecule, three-times-lower alcohol<br />

concentration will be required to reach the same interfacial surface tension between a hydrophobic<br />

and water/alcohol solution [77]


4.2 DSC 59<br />

Lipid n-Alcohol Phase T ( ◦ C) K x Method Reference<br />

<strong>DPPC</strong> Hexanol L β ′ (-) 100 Solvent-null Janes [29]<br />

<strong>DPPC</strong> Hexanol L β ′ (-) 1000 Solvent-null Janes [29]<br />

<strong>DPPC</strong> Hexanol L α (-) 3500 Solvent-null Janes [29]<br />

DMPC Hexanol L α 25 1342 Solvent-null Suurkuusk [73]<br />

<strong>DPPC</strong> Butanol L βI 15 70 Solvent-null Zhang [88]<br />

<strong>DPPC</strong> Butanol L β ′ 35 138 Solvent-null Zhang [88]<br />

<strong>DPPC</strong> Butanol L α 50 180 Solvent-null Zhang [88]<br />

<strong>DPPC</strong> Hexanol L α 45 839 Solvent-null Rowe [65]<br />

<strong>DPPC</strong> Pentanol L α 41 278 T c depression Kamaya [30]<br />

<strong>DPPC</strong> Hexanol L α 41 963 T c depression Kamaya [30]<br />

<strong>DPPC</strong> Heptanol L α 41 4060 T c depression Kamaya [30]<br />

Table 4.2 Partition coefficients for some relevant alcohol-lipid systems reported in the literature.<br />

K x is the mole fraction partition coefficient.<br />

Increasing the length of the partitioning alcohol cause a practically constant decrease<br />

in the absolute value of ∆∆G ◦ of around 3 kJ per one carbon chain, when the lipid/alcohol<br />

system is investigated in the same lipid phase [30]. This decrease in free energy<br />

is also seen in this work, by increasing the n-alcohol length from pentanol to hexanol<br />

an average increase in ∆G of 2.4 kJ per carbon chain is found. Going from hexanol to<br />

heptanol an increase in the free energy of 3.0 kJ is found, see table E.2.<br />

The partition coefficients for <strong>DPPC</strong> and hexanol is expected to be larger than the<br />

partition coefficients for DMPC and hexanol as <strong>DPPC</strong> is more hydrophobic or lipophilic<br />

than DMPC [73].<br />

4.2 DSC<br />

In the literature there are several examples of how DSC is used to measure the phase<br />

transition temperature between the interdigitated phase, L βI , and the liquid, crystalline<br />

phase, L α [47], [63], [88]. The specific phase transition temperature differs according to<br />

the lipid and alcohol as a result of which a DSC scan has been run prior to each SAXS<br />

experiment. It is important to know the phase transition temperature, as the SAXS<br />

instrument must be set correctly in order to measure in either the interdigitated or the<br />

fluid phase of the lipid. Directly from the shape of the DSC scan it is possible to determine<br />

whether the expected phase transition has been reached and measured. When the alcohol<br />

threshold concentration is reached one can see that the small “shoulder” (referring to<br />

the pretransition) in the DSC scan disappears. The interdigitated phase is induced at<br />

the same concentration as the pretransition disappears [79]. The width of the peak is<br />

another basis for qualitatively comparison, as presence of enough alcohol to induce the<br />

interdigitated phase is reflected in a broader peak [88]. In simulations it is found that<br />

the main phase transition temperature does not depend on the alcohol concentration if<br />

the interdigitated phase is reached [39]. Experimentally it is found that increasing the<br />

alcohol concentration at a distinct temperature can cause the interdigitated phase to coexist<br />

with either the ripple or the gel phase, depending on temperature. This co-existence<br />

is probably due to an inhomogeneous distribution of the alcohol in the bilayer, where<br />

the alcohol molecules aggregate in certain regions cause small domains of interdigitation<br />

[39].


60 Results<br />

DSC on water<br />

Several experiments on the DSC7 instrument were conducted with water in both pans.<br />

These experiments were made in order to estimate the uncertainty of the heat flow and<br />

find out whether there was a correlation between the baseline and used scanrate. As the<br />

DSC experiments with the lipid/alcohol system are carried out at different scanrates,<br />

this correlation would be important in the further data processing. All the DSC scans<br />

on water were conducted after a so-called burn-off, where the furnaces were heated to<br />

600 ◦ C to burn off all impurities in the furnaces.<br />

DSC on water in both sample and reference cells<br />

Source data<br />

Linear fit to baseline above settling interval<br />

80<br />

80<br />

Heat flow / mW<br />

60<br />

40<br />

5 °C / min<br />

10 °C / min<br />

15 °C / min<br />

20 °C / min<br />

25 °C / min<br />

30 °C / min<br />

35 °C / min<br />

40 °C / min<br />

Heat flow / mW<br />

60<br />

40<br />

5 °C / min<br />

Linear fit to 5 °C / min<br />

20 °C / min<br />

Linear fit to 20 °C / min<br />

40 °C / min<br />

Linear fit to 40 °C / min<br />

20<br />

20<br />

0<br />

10 20 30 40 50 60<br />

Temperature / °C<br />

0<br />

10 20 30 40 50 60<br />

Temperature / °C<br />

Residual between source data and linear fit<br />

1<br />

SDOM of the heat flow<br />

as a function of the scanrate<br />

Heat flow / mW<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

20 30 40 50<br />

5 °C / min<br />

20 °C / min<br />

40 °C / min<br />

Temperature / °C<br />

SDOM of heat flow / mW<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 />

SDOM<br />

Linear fit to SDOM<br />

0<br />

0 10 20 30 40<br />

Scanrate / °C pr. minute<br />

Figure 4.2 The four graphs show different aspects of DSC performed on water in both sample<br />

and reference cells. The upper-left figure shows that the settling interval increases as the<br />

scanrate increases. The baselines are rising as function of the temperature and they seem to<br />

be less straight as the scanrate increases. This is indeed the case as shown in the upper-right<br />

corner and the lower left corner in which just three scanrates are treated (in order not to overfill<br />

the graphs). The upper-right graph shows the linear fit above the settling interval and the<br />

lower-left show the residual between the source data and the fit. The lower-right graph shows<br />

that the SDOM of the heat flow increases as the scanrate increases, see table 4.3.<br />

From the residuals between the source data and the linear fit in figure 4.2 a more or less<br />

systematic deviation is found. This deviation is clearly not dependent on the scanrate<br />

as the same correlation is found for 3 different scanrates. We can not explain the reason<br />

for the systematic deviation.<br />

Looking at the low temperature end of the scan in figure 4.3 one can recognize the<br />

same settling interval as seen in the water experiment, see figure 4.2. Even though the


4.2 DSC 61<br />

Scanrate, ◦ C/min 5 10 15 20 25 30 35 40<br />

SDOM of heat flow, mW 0.19 0.28 0.46 0.54 0.65 0.71 0.71 0.81<br />

Table 4.3 The estimated uncertainties of the heat flow. Extrapolation to 0 ◦ C/min gives an<br />

uncertainty of 0.15 mW as indicated in the lower right corner in figure 4.2.<br />

DSC on MLV (<strong>DPPC</strong>)<br />

with different scanrates<br />

Heat flow / mW<br />

40<br />

20<br />

5 °C / min<br />

10 °C / min<br />

15 °C / min<br />

20 °C / min<br />

25 °C / min<br />

30 °C / min<br />

40<br />

40 45 50<br />

10 20 30 40 50 60<br />

Temperature / °C<br />

Transition temperature as a function of scanrate<br />

Transition time as a function of scanrate<br />

48<br />

0.6<br />

Transition temperature / °C<br />

46<br />

44<br />

42<br />

Transition temperature<br />

estimated peak<br />

Linear fit, Temp =<br />

0.27 minute · Scanrate +<br />

41 °C<br />

Transition time / minutes<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Transition time estimated from<br />

transition temperature interval<br />

40<br />

0 5 10 15 20 25 30<br />

Scanrate / °C pr. minute<br />

0<br />

0 5 10 15 20 25 30<br />

Scanrate / °C pr. minute<br />

Figure 4.3 The top figure shows the source data of DSC measurements performed on multilamellar<br />

<strong>DPPC</strong> <strong>vesicles</strong> with a concentration of (5.12 ±0.02) %. The transition temperature and<br />

time are then plotted against the scanrate in the two lower graphs. The transition temperature<br />

decreases as the scanrate decreases, and the transition temperature vs. the scanrate seems to be<br />

linear and gives a “true” transition temperature at about 41 ◦ C for an infinitely slow scanrate.


62 Results<br />

transition temperature decreases as the scanrate decreases, the transition time actually<br />

shows an increasing tendency as the scanrate decreases (one should not be fooled by the<br />

narrower peaks, since the transition time is transition temperature interval divided by<br />

the scanrate). The smaller temperature gradients for low scanrates apparently slow down<br />

the transition process. The system does not seem to reach a state of equilibrium after<br />

the main transition, which is indicated by the wobbly post-transition baselines - opposed<br />

to the much nicer baselines before the transition in figure 4.3 (on the left). The expected<br />

pretransition is not convincingly detected.<br />

DSC error sources<br />

There were a lot of error sources. Some of which occur randomly, and some do not.<br />

The three worst error sources were lack of thermal equilibrium, lack of thermal contact<br />

between the heater and the pans and water condensating inside the apparatus. It was<br />

hard to tell which error source was to blame in the specific case. The presented data is<br />

only a small portion of the measurements performed. And due to the error sources, a lot<br />

of data was rejected. We have not had an objective criterion on when to reject data from<br />

our DSC measurements. It was almost a reversed situation. We have had an objective<br />

criterion on when to accept data: our results had to be reproducible.<br />

The error sources possibly include:<br />

• Pollution from the dry air fed into the sample holder chamber<br />

• Water condensating on top of the sample holder (and other places)<br />

• Pollution inside the sample holder<br />

• Mechanical deformation of the pans<br />

Enthalpy of transition<br />

Several DSC scans were conducted on the different lipid/alcohol solutions varying the<br />

alcohol concentration. The pure <strong>DPPC</strong> main phase transition was also measured. The<br />

data processing of these scans served two purposes, determining the phase transition<br />

temperature and calculating the phase transition enthalpy. The former is used with<br />

regards to the setting of the SAXS instrument and the latter is primarily used as a basis<br />

for comparison with comparable values found in the literature.<br />

This value for the phase transition temperature is in accordance with values reported<br />

in the literature. Löbbecke & Cevc [47] find the main transition enthalpy of <strong>DPPC</strong> to<br />

be 43 kJ/mole. According to Löbbecke & Cevc [47] the phase transition temperature<br />

decreases with increasing alcohol concentration and the enthalpy of transition decreases<br />

at low alcohol concentration and increases with high alcohol concentration, relative to<br />

the main transition enthalpy of pure <strong>DPPC</strong>. This tendency was also recognized in this<br />

work, but as the measurements were primarily conducted to find the phase transition<br />

temperature, the enthalpies have not been calculated systematically.<br />

The enthalpies presented in figure 4.4 are all found by determining the area between<br />

the peak and the baseline, principle shown in figure F.3, and the results are represented<br />

graphically in figure 4.3. The location of the baseline and the peak have been estimated<br />

and as these are determined subjectively, they will contribute to the uncertainties of<br />

the final integral. The calculated enthalpies should be constant and independent of the<br />

scanrate, and should not decrease with increasing scanrate.


4.2 DSC 63<br />

50<br />

Main transition enthalpy<br />

Transition enthalpy/ kJ/mol<br />

45<br />

40<br />

35<br />

30<br />

25<br />

Raw data<br />

y=−0.82x+45.3<br />

20<br />

0 5 10 15 20 25 30<br />

Scanrate/ °C per minute<br />

Figure 4.4 The main transition enthalpy of <strong>DPPC</strong> as a function of scanrate. As seen on figure<br />

4.3 the sizes of the peaks are varying with scanrate in which case the enthalpies of transition<br />

also vary with scanrate. The “true” phase transition enthalpy is found by extrapolation to<br />

45 ± 2 kJ/mole.<br />

Temperature of transition<br />

The scans show that the phase transitions decrease when alcohol is added to the lipid<br />

solution. This behaviour is also reported in the literature [47], [64]. The phase transition<br />

temperature decreases with increasing alcohol concentration (results from the scal-1<br />

microcalorimeter, not shown). In a pure <strong>DPPC</strong> lipid solution the phase transition temperature<br />

is 41 ◦ C, see figure 4.3. By adding pentanol the temperature decreases, and the<br />

measured phase transition temperature is around 35 ◦ C. Adding hexanol decreases the<br />

phase transition temperature even more drastically, and the measured temperature is<br />

around 29 ◦ C, see figure 4.5. For heptanol the phase transition temperature is around<br />

35 ◦ C. The phase transition temperatures are found by reading the top position of the<br />

peak. The uncertainties are estimated to be ±1 ◦ C. This value is based on the results<br />

from several performed DSC scans.<br />

From the scan in figure 4.5 one can see that no pretransition has been detected<br />

and that the peaks have widened compared to the pure <strong>DPPC</strong> main phase transition.<br />

Both observations indicating the presence of an interdigitated phase. The enthalpy of<br />

transition has been calculated for the four scans represented in figure 4.5. The enthalpy<br />

has been calculated as the integral of the peak, see figure F.3. The four scans all had<br />

a phase transition enthalpy of 62-65 kJ/mole. This enthalpy is about 20 % higher than<br />

the values reported in the literature [47]. Still the enthalpy is known to increase with<br />

increasing concentration of alcohol in the lipid.<br />

A test was made to investigate whether the lipid/alcohol suspension suffered damage<br />

when being heated and cooled rapidly, which is the case when conducting a DSC scan.<br />

This was at the same time a test to check whether the phase transition temperature<br />

increased or decreased when the sample was repeatedly heated and cooled. 12 successive<br />

scans were conducted and plotted on top of each other, see figure 4.6.<br />

The DSC scans reveal that the <strong>DPPC</strong>/hexanol suspension is quite stable. The sus-


64 Results<br />

<strong>DPPC</strong> : pentanol<br />

<strong>DPPC</strong> : pentanol<br />

12<br />

Scanrate: 2 K / min<br />

14<br />

Scanrate: 5 K / min<br />

11.5<br />

13<br />

Heat flow [mW]<br />

11<br />

10.5<br />

Heat flow [mW]<br />

12<br />

10<br />

11<br />

9.5<br />

10<br />

10 20 30 40 50 60<br />

Temperature [°C]<br />

10 20 30 40 50 60<br />

Temperature [°C]<br />

<strong>DPPC</strong> : hexanol<br />

<strong>DPPC</strong> : hexanol<br />

24<br />

Scanrate: 2 K / min<br />

35<br />

Scanrate: 5 K / min<br />

23<br />

Heat flow [mW]<br />

22<br />

21<br />

Heat flow [mW]<br />

30<br />

25<br />

20<br />

19<br />

10 20 30 40 50 60<br />

Temperature [°C]<br />

20<br />

10 20 30 40 50 60<br />

Temperature [°C]<br />

Figure 4.5 DSC scans for series 9, showing <strong>DPPC</strong> with pentanol and hexanol, respectively.<br />

The upper-left figure shows <strong>DPPC</strong> and pentanol, ratio 1:2, there is a sharp peak at 34 ◦ C. The<br />

upper-right figure shows <strong>DPPC</strong> and pentanol, ratio 1:2, there is a sharp peak at 35 ◦ C. The<br />

lower-left figure shows <strong>DPPC</strong> and hexanol, ratio 1:2, there is a sharp peak at 28.5 ◦ C. The<br />

lower-right figure shows <strong>DPPC</strong> and hexanol, ratio 1:2, there is a sharp peak at 29.5 ◦ C. The two<br />

“ghost-peaks” in the last part of the scans of pentanol are most likely due to impurities in the<br />

apparatus. These DSC scans have been obtained by using the Perkin-Elmer DSC7 instrument.<br />

pension responded similarly to being heated, even after being heated to 80 ◦ C and cooled<br />

to 10 ◦ C several times. The position of the peak and hence the phase transition temperature<br />

does not change throughout the 12 scans, indicating that the phase transition from<br />

the L βI phase to the L α phase is fast, reversible and that the transition temperature is<br />

stable. The same results apply for the samples with pentanol and heptanol, see appendix<br />

F.


4.2 DSC 65<br />

12000<br />

<strong>DPPC</strong> : hexanol<br />

10000<br />

9500<br />

scan 1<br />

scan 12<br />

8000<br />

mV<br />

9000<br />

mV<br />

6000<br />

8500<br />

38 38.5 39<br />

Temperature [°C]<br />

4000<br />

2000<br />

0<br />

36 38 40 42 44 46 48 50<br />

Temperature [°C]<br />

Figure 4.6 12 successive scans are plotted on top of each other. There is practically no<br />

difference between the 12 scans and the location of the peak does not differ. A lipid to hexanol<br />

ratio of 1:17, w = 0.05% and a scanrate of 1K/min was used. There is an average standard<br />

deviation of 174.3 relative to the first scan. These scans have been obtained by using the scal-1<br />

microcalorimeter.


66 Results<br />

4.3 SAXS<br />

Introductory experiments<br />

Several introductory experiments were made varying different parameters i.e. the sample<br />

preparation procedure 2 in order to elaborate on the sensitivity of the system. During<br />

these experiments it was found that there was no difference in the obtained spectra<br />

when keeping the alcohol-containing samples shaking at 32 ◦ C instead of at 52 ◦ C. The<br />

first series of the SAXS experiments consisted of experiments conducted at 4 different<br />

temperatures - two experiments below and two experiments above the phase transition<br />

temperature. The two experiments obtained in the apparent L βI phase are practically<br />

indistinguishable just like the two experiments obtained in the L α phase, see figure 4.7.<br />

<strong>DPPC</strong> : hexanol<br />

Serie 7<br />

8<br />

<strong>DPPC</strong> : hexanol<br />

Serie 7<br />

8<br />

25 °C<br />

35 °C<br />

40 °C<br />

45 °C<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q [Å -1 ]<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q [Å -1 ]<br />

(a) The two datasets obtained at<br />

25 ◦ C and 30 ◦ C are practically indistinguishable.<br />

(b) The two datasets obtained at<br />

40 ◦ C and 45 ◦ C are practically indistinguishable.<br />

Figure 4.7 The raw data (with subtracted background) from the SAXS experiments with series<br />

07, <strong>DPPC</strong>:hexanol 1:2. The temperatures are chosen given the DSC results.<br />

The results led to the conclusion that one experiment in each lipid phase was sufficient.<br />

Besides wearing the Kratky camera and its associated equipment lesser and saving time<br />

on the experiment per se, the decision also had the advantage that the sample was exposed<br />

to X-rays for a shorter period of time, causing less possible damage to the <strong>vesicles</strong>. The<br />

results also show a noticeable difference between the SAXS spectra obtained in the two<br />

phases, reflecting the assumption of a measurable structural difference between the two<br />

phases, see also figure 4.8.<br />

Matlab Code<br />

☞ page 157<br />

In order to get measurements with enough counts to obtain good statistics the SAXS<br />

experiments last 18 hours. To see if there is any temporal development this has been<br />

divided into three measurements of six hours each. In figure 4.9 these three spectra from<br />

the same SAXS experiment are represented. The photon counts of the three spectra are<br />

added and divided by the total measuring time (equivalent to averaging the intensities).<br />

The averaged intensity is then normalized (see equation 5.32) for the modelling.<br />

2 For further details see appendix D.


4.3 SAXS 67<br />

<strong>DPPC</strong> : Pentanol<br />

Series 15<br />

10<br />

<strong>DPPC</strong> : Hexanol<br />

Series 12<br />

8<br />

25°C<br />

45°C<br />

8<br />

45 °C<br />

25 °C<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

Figure 4.8 Spectra obtained at 25 ◦ C and at 45 ◦ C.<br />

Normally we have been doing two measurements in a row, leaving the sample to be<br />

exposed to radiation for 36 hours. This long period of exposure will unquestionably<br />

have an effect on the sample. We have shown that there is not detectable effect within<br />

the first 18 hours. But during the experimental work with the Kratky camera we have<br />

experienced more than once that an amount of lipid <strong>vesicles</strong> had aggregated in the glass<br />

tube just where the beam passes through the sample holder when the sample had been<br />

exposed to radiation for more than 36 hours. This shows that the sample can be affected<br />

by long-term beam exposure.<br />

8<br />

<strong>DPPC</strong> : hexanol<br />

Series 14<br />

6<br />

6 hours<br />

12 hours<br />

18 hours<br />

Intensity [a.u.]<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q [Å -1 ]<br />

Figure 4.9 Three spectra obtained at the same temperature and recorded during 0-6, 6-12 and<br />

12-18 hours of measuring, respectively. The spectra are averaged and used in the following data<br />

processing.<br />

The three spectra in figure 4.9 are practically indistinguishable, demonstrating that there<br />

are no significant changes within the sample during the time of the 18 hour long experiment,<br />

and justifying the averaging. The stability of the structure also indicates that the<br />

beam does not alter the chemistry of the sample.


68 Results<br />

The SAXS spectra used for the data modelling are all obtained by measuring 3 times 6<br />

hours at different temperatures, starting at the low temperatures. The DSC experiments<br />

already indicated that the system did not exhibit notable hysteresis going from the high<br />

temperature phase to the low temperature phase, but we checked that there was no<br />

difference in starting the experiment in the low temperature area instead of the high<br />

temperature area. Two experiments with the same sample were made to verify this. The<br />

first experiment started in the low temperature range going to the high temperature<br />

range, denoted “upscan”, the second experiment started in the high temperature area<br />

going down to the low temperature area, denoted “downscan”. There was no difference<br />

between the two experiments. The results are represented in appendix G. These results<br />

are in good agreement with the results obtained with DSC.<br />

Background<br />

In the data processing it is necessary to correct for non-system (air, solvent and cuvette)<br />

scattering called the background. One could argue that this actually is a part of the<br />

modelling, but since this correction is made before the description of spectra characteristics<br />

we decided to include it in this chapter. The background is measured with a<br />

sample of pure water as shown in figure 4.10. Water is chosen over a water/n-alcohol<br />

solution since the large partition coefficients suggest that most of the alcohol is in the<br />

bilayer. Even so, the bilayer will eventually get saturated and there will be a considerable<br />

amount n-alcohol in the solvent. This was realized too late (after the modelling) and<br />

we hope it does not compromise our conclusions. Throughout the whole experiment the<br />

Intensity / counts pr. second<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Measured<br />

Background<br />

Measured with subtracted background<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Figure 4.10 The measured, background and measured with subtracted background intensity<br />

spectrum. The sample is <strong>DPPC</strong> <strong>vesicles</strong> perturbated by hexanol, series 14 at 45 ◦ C. The<br />

background is almost constant as a function of q. Both sample and background measurement<br />

lasted 18 hours and had a photon count N of 1.3 · 10 7 and 6.8 · 10 6 , respectively.


4.3 SAXS 69<br />

lipid/alcohol sample absorb a far greater number of photons than the water sample. At<br />

high q-values, q = 0.4Å −1 to q = 0.5Å −1 , the modelled I(q) is nearly zero, and hence<br />

the transmission, T , is calculated at high q-values<br />

∫ q2<br />

q<br />

T =<br />

1<br />

I measured (q)dq<br />

∫ q2<br />

(4.3)<br />

q 1<br />

I solvent (q)dq<br />

The transmission calculated with equation 4.3 from q 1 = 0.45 Å −1 to q 2 = 0.49 Å −1<br />

is found to usually be about T = 1.2. The intensity as function of the q values of an<br />

experiment used in the data processing is then,<br />

I sample (q) = I measured (q) − T · I solvent (q). (4.4)<br />

The transmission is calculated during the data processing of each spectrum and hence,<br />

the transmission varies slightly from experiment to experiment.<br />

Matlab Code<br />

☞ page 157<br />

8<br />

Intensity / a.u.<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

7.5<br />

7<br />

6.5<br />

6<br />

5.5<br />

5<br />

1<br />

0<br />

0.025 0.05 0.075 0.1 0.125 0.15<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Figure 4.11 A sample of <strong>DPPC</strong> <strong>vesicles</strong> perturbated by hexanol, series 14 at 45 ◦ C. The<br />

absolute uncertainty of the photon count is √ N and the relative uncertainty is √ N<br />

= √ 1<br />

N N<br />

(= 0.94% as a minimum relative uncertainty for this particular case).<br />

The spectra used<br />

One of the imposed conditions of using a spectrum in the modelling process was that the<br />

spectrum could be reproduced. The experimental progress was delayed repeatedly due to<br />

several unexpected difficulties with the Kratky camera, primarily. Among the difficulties<br />

was the break down of the air condition in the lead-isolated room, numerous software<br />

breakdowns and leaks in the cuvette. Still, the spectra for all six different experiments<br />

was reproduced, see figure 4.12. For larger versions of the graphs the reader is referred<br />

to appendix H, page 135.


70 Results<br />

<strong>DPPC</strong> : Pentanol<br />

25 °C<br />

8<br />

<strong>DPPC</strong> : Pentanol<br />

45 °C<br />

8<br />

Series 14<br />

Series 15<br />

Series 13<br />

Series 15<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q [Å -1 ]<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

10<br />

<strong>DPPC</strong> : Hexanol<br />

25 °C<br />

10<br />

<strong>DPPC</strong> : Hexanol<br />

45 °C<br />

8<br />

Series 12<br />

Series 15<br />

8<br />

Series 14<br />

Series 12<br />

Intensity [a.u.]<br />

6<br />

4<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

2<br />

0<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

10<br />

<strong>DPPC</strong> : Heptanol<br />

25 °C<br />

10<br />

<strong>DPPC</strong> : Heptanol<br />

45 °C<br />

8<br />

Series 10<br />

Series 15<br />

8<br />

Series 10<br />

Series 15<br />

Intensity [a.u.]<br />

6<br />

4<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

Figure 4.12 The raw data from the SAXS experiments with all the alcohols used. The spectra<br />

have been normalized according to equation 5.32 - we could not expect the intensities to be the<br />

exact same, since the refining process leaves varying amounts of <strong>vesicles</strong> and thus scattering<br />

centers from series to series.


4.3 SAXS 71<br />

Some features characterize the raw data plots. All the spectra display an envelope curve<br />

reflecting the main structure of the lipids in the suspension, the unilamellar <strong>vesicles</strong>. The<br />

spectrum for unilamellar <strong>vesicles</strong> are found in figure 4.13. The envelope curve is fairly<br />

10<br />

<strong>DPPC</strong><br />

Unilamellar <strong>vesicles</strong>, 25 °C<br />

<strong>DPPC</strong><br />

Unilamellar <strong>vesicles</strong>, 45 °C<br />

8<br />

8<br />

Intensity [a.u.]<br />

6<br />

4<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q [Å -1 ]<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

Figure 4.13 SAXS spectra of unilamellar <strong>vesicles</strong> of <strong>DPPC</strong>, measured at two different temperatures.<br />

smooth and does not display the indent around q = 0.05Å −1 as opposed to the spectra<br />

obtained with <strong>DPPC</strong> and alcohol. Besides the main feature, the envelope curve, at least<br />

three peaks, see close-up in figures 4.14, on the graphs are found at low q-values for low as<br />

well as high temperature measurements. On the raw data for the low temperature measurements<br />

of heptanol/<strong>DPPC</strong> these peaks are not as pronounced. A common feature of<br />

all the low temperature measurements compared to the high temperature measurements<br />

is the fact that the second peak is moving towards higher q-values as the temperature is<br />

raised. The first peak, however, is practically not moving as the temperature is raised,<br />

except from the pentanol/<strong>DPPC</strong> measurements where the peak moves towards higher<br />

q-values as the temperature increases. For the low temperature pentanol/<strong>DPPC</strong> and<br />

hexanol/<strong>DPPC</strong> measurements one can also find a very broad peak around q = 0.4Å −1 .<br />

The peak is also found at the pure ULV <strong>DPPC</strong> low temperature measurement, see figure<br />

4.13, but it is not found in any of the high temperature measurements.<br />

The location of the peaks is also differing. For the low temperature pentanol/<strong>DPPC</strong><br />

measurements the first peak is located at lower q-values than for the high temperature<br />

measurements with the same alcohol. The opposite apply to the location of the second<br />

peak. The third peak does not change location as the temperature increases. For the<br />

hexanol/<strong>DPPC</strong> measurements the peaks do not change location with temperature. For<br />

the heptanol/<strong>DPPC</strong> measurements only the second peak re-locates, it moves to higher<br />

q-values as the temperature increases. The graphs flatten differently depending on the<br />

experimental temperature. The low temperature measurements for pentanol/<strong>DPPC</strong> and<br />

hexanol/<strong>DPPC</strong> seem to flatten faster than the high temperature measurements. The low<br />

temperature and high temperature heptanol/<strong>DPPC</strong> measurements do not seem to differ<br />

in this way.<br />

As earlier mentioned an imposed condition in using a spectrum in the further data<br />

modelling was that it was reproducible. In figure 4.12 examples of low temperature<br />

hexanol/<strong>DPPC</strong> spectra are represented, but two other spectra of low temperature hexanol/<strong>DPPC</strong><br />

are worth mentioning. There are several characteristic features in the spectra


72 Results<br />

8<br />

<strong>DPPC</strong> : Pentanol<br />

45 °C<br />

Series 13<br />

7.5<br />

Intensity [a.u.]<br />

7<br />

6.5<br />

6<br />

5.5<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

q [Å -1 ]<br />

10<br />

<strong>DPPC</strong> : Hexanol<br />

25°C<br />

9<br />

Series 12<br />

Intensity [a.u.]<br />

8<br />

7<br />

6<br />

5<br />

4<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

q [Å -1 ]<br />

<strong>DPPC</strong> : Heptanol<br />

25 °C<br />

9<br />

Series 15<br />

Intensity [a.u.]<br />

8<br />

7<br />

6<br />

5<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

q [Å -1 ]<br />

Figure 4.14 The main three peaks displayed at low q-values. The values in table 4.4 are found<br />

by reading the q-values in a close-up.


4.3 SAXS 73<br />

Low temperature<br />

High temperature<br />

1. Peak 2. Peak 3. Peak 1. Peak 2. Peak 3. Peak<br />

[Å −1 ] [Å −1 ] [Å −1 ] [Å −1 ] [Å −1 ] [Å −1 ]<br />

Pentanol 0.038 (14) 0.093 (14) 0.105 (14) 0.043 (13) 0.088 (13) 0.106 (13)<br />

0.038 (15) 0.090 (15) 0.108 (15) 0.043 (15) 0.089 (15) 0.106 (15)<br />

Hexanol 0.0432 (12) 0.09 (12) 0.106 (12) 0.0425 (12) 0.088 (12) 0.106 (12)<br />

0.0417 (14) 0.089 (14) 0.1061 (14) 0.0425 (14) 0.091 (14) 0.106 (14)<br />

Heptanol 0.043 (15) 0.082 (15) 0.106 (15) 0.0434 (15) 0.0908 (15) 0.106 (15)<br />

0.0415 (10) 0.085 (10) 0.0434 (10) 0.0908 (10) 0.106 (10)<br />

Table 4.4 The location of the three peaks. The values are found by reading from close-ups of<br />

the spectra, see figure 4.14. The numbers in the brackets refer to the series. The figures in the<br />

brackets refer to the series.<br />

10<br />

<strong>DPPC</strong> : hexanol<br />

25 °C<br />

8<br />

Series 9<br />

Series 13<br />

Intensity [a.u.]<br />

6<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

Figure 4.15 The raw spectra of hexanol/<strong>DPPC</strong> obtained at 25 ◦ C. There is a certain repeatdistance<br />

between the peaks, ∆q≃0.12 Å −1 , equivalent to a distance of around 52 Å.<br />

of the low temperature measurements of hexanol/<strong>DPPC</strong> series 9 and series 13. Unlike<br />

the spectra in figure 4.12 these spectra seem to demonstrate the presence of multilamellar<br />

<strong>vesicles</strong> as there is a certain repeat distance, ∆q≃0.12 Å −1 , between the peaks. The first<br />

and largest peak is located at high q-values like the other low temperature hexanol/<strong>DPPC</strong><br />

measurements.


74 Results<br />

MAX-Lab measurements<br />

In late January an opportunity came by to do measurements on a synchrotron 3 . After<br />

preparing the samples, one pure <strong>DPPC</strong>, one with hexanol and one with heptanol, as<br />

described in section 3.4, the samples were kept at room temperature until the measurements<br />

were performed, that is 24 to 48 hours after leaving the thermomixer. The spectra<br />

obtained from the synchrotron measurements differ from the spectra obtained with the<br />

Kratky camera. The three peaks at low q-values were not detected, but the envelope<br />

curve was the same as what we see. Except from the thermal history (being kept at<br />

room temperature), the samples have been treated like all the other samples used for<br />

this work.<br />

3 Performed by Dorthe Posselt at the MAX-lab in Lund, Sweden


Part III<br />

Analysis


5 Modelling<br />

This chapter contains an overview of the models used, introductions to each of the models<br />

and a summary of the modelling results. Please refer to appendix H, page 135, for an<br />

exhaustive list of model graphs and fitted parameters.<br />

Before we begin, it is enlightening to take one step back and look more generally at the<br />

modelling process. Often, physical modelling can be expressed as<br />

ˆKx = y (5.1)<br />

where ˆK is a model operating on some modelling parameters x giving the output y,<br />

which can be compared with the measured data obtained by experiments. The direct<br />

problem would be to evaluate y given ˆK and x, but it leaves two other indirect problems:<br />

to determine ˆK or x (given x and y, or, ˆK and y, accordingly). Both types of problems<br />

are found in the various disciplines of physics.<br />

Direct problems are often well-posed, which means that ([85, well-posed problem])<br />

• a solution exists,<br />

• the solution is unique, and<br />

• the solution depends continuously on the data, in some reasonable topology.<br />

Conversely, the indirect problems are often ill-posed and hence lack some or all of the<br />

mentioned properties (existence, uniqueness and stability).<br />

An example of a well-posed problem could be to determine the temperature field, y,<br />

after a period of time given Fourier’s law of heat conduction, ˆK, with specified initial<br />

conditions, x. By contrast the backwards heat equation, deducing an initial temperature<br />

field from the final field is ill-posed in that the solution, x, is highly sensitive to changes<br />

in the final data y [85, well-posed problem], and the solution might not be unique. So,<br />

in this particular case both the uniqueness and stability conditions are violated.<br />

Actually our indirect problem is to determine both ˆK and x given y, since we have to<br />

find a suitable electron distribution model and, then, determine the parameters of that<br />

specific model.<br />

5.1 Overview of the models used<br />

The elastic scattering theory is the basis of the models. It states that the differential<br />

cross section depends on the density and the arrangement of the electrons as described<br />

in chapter 2. The form factor (f bilayer model ) is proportional to the Fourier transform of<br />

the relative electron density distribution, ρ(r):<br />

f bilayer model (q) ∝ F{ρ(r)} ≡<br />

1<br />

(2π) 3/2 ∫u.c.<br />

ρ(r)exp(−iq · r)d 3 r (5.2)<br />

77


78 Modelling<br />

where q is the momentum transfer vector and the unit cell (u.c.) is the bilayer of a<br />

vesicle.<br />

Several aspects of the experimental setup, including the cuvette glass, air and solvent<br />

scattering should be taken take of, since the background has been subtracted (see section<br />

4.3, page 68). But other aspects such as secondary scattering and inelastic scattering are<br />

completely disregarded.<br />

Looking at one particular q-value, the scattered intensity measured gets contributions<br />

from all over the scattering volume. This smearing of the intensity spectrum is based on<br />

the geometry of the experimental setup (see section 3.3, page 41) and is explained further<br />

in section 5.2. This correction is very important as the Kratky camera has a slit-shaped<br />

beam, hence it has been applied to all of the models used:<br />

∫<br />

I smeared (q) = P(r)I p (q ′ )d 3 r (5.3)<br />

V<br />

where V is the scattering volume and I p (q ′ ) is the scattered intensity from a point shaped<br />

beam and P(r) is the intensity profile of the beam.<br />

Although we try to keep the amount of MLVs to a minimum, both ULVs and MLVs are<br />

suspended in the solvent. And since MLVs scatter more than ULVs the sample does<br />

not have to be polluted with many MLVs in order to see their presence on the intensity<br />

spectrum. Therefore, we include MLVs in our modelling. We assume that the <strong>vesicles</strong><br />

are dispersed in the sample so no net interference occurs between the <strong>vesicles</strong>, see figure<br />

3.21. The intensity from each of the <strong>vesicles</strong> can then be added<br />

I p (q) ∝ I ULV (q) + A b I MLV (q) (5.4)<br />

where the factor A b is the relative amount of scattered photons by MLVs in the sample.<br />

The models representing the electron density are formulated in direct space and are<br />

Fourier transformed to the inverse space to comprise the form factors of the model. The<br />

models are based on Gaussian functions, which have the advantage of having analytical<br />

Fourier transforms. This is not an idea of ours - others have used Gaussian functions<br />

before us: [6], [9], and see figure 5.1. It does make sense to use such coarse-grained<br />

functions to represent the density of electrons of large assemblies of molecules since the<br />

time of exposure during the measurement is about 18 hours - all dynamics on smaller<br />

time scales are completely smeared out in time<br />

I ULV (q) ∝ 〈|f bilayer model (q)| 2 〉 and I MLV (q) ∝ 〈|f bilayer model (q)| 2 S(q)〉 (5.5)<br />

where 〈···〉 is the time average of all the <strong>vesicles</strong> (the time average will be implied throughout<br />

the whole chapter) and S(q) is the structure factor [56], [89].<br />

We have used 6 models in our analysis of the SAXS data as they all contribute in<br />

different ways to the understanding of the modelled system. Clearly the 6 models attend<br />

to slightly different features in the interpretation of the SAXS spectra and hence confirm<br />

or invalidate the features characterizing the alcohol perturbated <strong>DPPC</strong> <strong>vesicles</strong>. We have<br />

applied all models to data of both low and high temperature phases in order to see which<br />

model was suited to which phase.<br />

sym and sym-mlv are similar, the latter differs from the former by correcting for the<br />

presence of MLVs. 3d and 3d-mlv share the same similarity. The symmetric model sym


5.2 Instrumental smearing 79<br />

Figure 5.1 Bilayer electron density profile and a propability profile of the bilayer constituents<br />

[53]. (a) shows the probability density distribution functions for the different components of<br />

the bilayer. (b) shows the electron density profile, the dotted line from simulations and the<br />

solid line from X-ray studies. (c) shows two volumetric pictures.<br />

is the most commonly used model for describing the electron density profile. Taking<br />

this one dimensional model further to three dimensions lead to the symmetric three<br />

dimensional model 3d-sym. The 3d-sym model fits the spectra as if the scattering object<br />

were a spherical vesicle also taking into account the distribution of vesicle radius.<br />

The number of parameters differ from model to model. The sym and 3d-sym models<br />

have 4 parameters, the asym and sym-4g models have 5 parameters and the two models<br />

which also correct for the presence of multilamellar <strong>vesicles</strong>, sym-mlv and 3d-mlv both<br />

have 6 parameters. Choosing the number of fitting parameters is a tricky balancing act.<br />

The more fitting parameters the broader the parameter space yielding a greater the risk<br />

of finding a local minimum than a global minimum. But more fitting parameters might<br />

also do the difference in finding a good fit for your raw data as there simply are more<br />

parameters to turn on so to speak. The general guideline is that the more parameters<br />

the less certain is the fitted values of the single parameter. We have tried to keep the<br />

numbers of parameters low in order to get a better determination of the single parameters<br />

as especially one of these parameters, z h is crucial in our analysis of the SAXS spectra.<br />

In table 5.1 we have summarized the number of parameters used, the names we use as<br />

references and the conceptual basis for each of the models. In the following sections we<br />

describe each of the models (and their motivations) and present the results in the end of<br />

the chapter.<br />

5.2 Instrumental smearing<br />

The scattered intensity measured (at one particular q value) gets contributions from all<br />

over the scattering volume, hence the name smearing. Compared to pinhole collimation<br />

the block collimation of the Kratky camera provides a higher intensity beam, but one<br />

has to take into account that photons arriving to a certain point at the detector can


80 Modelling<br />

Name Parameters Conceptual basis<br />

sym 4 symmetric bilayer electron profile<br />

with three Gaussian functions<br />

sym-4g 5 symmetric bilayer electron profile<br />

with four Gaussian functions<br />

asym 5 asymmetric bilayer electron profile<br />

sym-mlv 6 symmetric bilayer electron profile<br />

and scattering from both ULV and MLV <strong>vesicles</strong><br />

3d 4 spherically symmetric bilayer electron profile<br />

3d-mlv 6 spherically symmetric bilayer electron profile<br />

and scattering from both ULV and MLV <strong>vesicles</strong><br />

Table 5.1 A summary of the models applied.<br />

Figure 5.2 A schematic drawing of 2-dimensional smearing due to the slit-shaped beam (adapted<br />

from [57]).<br />

originate from a range of scattering angles, see figure 5.2. This corresponds to a range<br />

of q-values and the measured intensity at the detector becomes<br />

Matlab Code<br />

☞ page 174<br />

∫<br />

∫<br />

I smeared (q) = P(r)I p (q ′ )d 3 r ≈<br />

V<br />

P(x)I p (q ′ )dx (5.6)<br />

where V is the scattering volume and I p (q ′ ) = √ q 2 + x 2 is the scattered intensity from<br />

a point shaped beam and P(x) is the horizontal propability density profile of the beam.<br />

Here, we ignore the smearing along the y and z axes, since the beam is much wider than<br />

the height of the beam and the depth of the sample, see figure 5.3.<br />

It should be noted that the axis of abscisses is different in figure 5.3(a) and 5.3(b).<br />

The beam width is much larger than the beam height, justifying that we only consider<br />

smearing from the width of the beam.


5.3 Symmetric (1d) 81<br />

350<br />

3000<br />

300<br />

2500<br />

250<br />

2000<br />

G(x) [a.u.]<br />

200<br />

G(y) [a.u.]<br />

1500<br />

150<br />

1000<br />

100<br />

500<br />

50<br />

0<br />

0.1 0.2 0.3 0.4 0.5 0.6<br />

x [Å −1 ]<br />

0.05 0.1 0.15 0.2 0.25 0.3<br />

y [Å −1 ]<br />

(a) The intensity profile along the<br />

horizontal axis, x. The measured<br />

beam is the spots, the fitted Gaussian<br />

is the line. The fitting parameters<br />

are a=346.60, b=0.31127 and<br />

c=0.14681<br />

(b) The intensity profile along the<br />

vertical axis, y. The measured<br />

beam is spots, the fitted Gaussian<br />

is the line. The fitting parameters<br />

are a=2398.9, b=0.13386 and<br />

c=0.0056663<br />

Figure 5.3 Two Gaussian functions have been used to model the beam G(x) = a·exp<br />

“−<br />

”.<br />

(x−b)2<br />

c 2<br />

In order to find the beam profile in both the horizontal (x axis) and vertical (y axis) direction a<br />

measurement of the direct beam is performed by rotating the detector and using a filter instead<br />

of a beam stop.<br />

5.3 Symmetric (1d)<br />

One of the most commonly used models for the bilayer electron density profile is a<br />

symmetric lipid bilayer electron density profile in which the bilayer is divided into the<br />

three areas: two head group parts and one tail part [6], [9], see figure 5.1. In this<br />

one-dimensional model there is no representation of actual <strong>vesicles</strong>, but the bilayers are<br />

represented by infinitely large sheets which have no preferred orientation - the effect of<br />

the membrane curvature is therefore neglected in this model.<br />

Compared to the electron density of the solvent (primarily water), the electron density<br />

is larger for the head group and smaller for the acyl chains (tails). The electron density,<br />

ρ(z), of these three areas can be regarded as a sum of Gaussian functions. A normalized<br />

Gaussian function is used to describe each of the three areas<br />

where<br />

ρ(z) = ∑ n<br />

A n<br />

√<br />

2πσ<br />

2 n<br />

exp<br />

(− (z − z n) 2 )<br />

2σ 2 n<br />

(5.7)<br />

√A n<br />

is the amplitude, z 2πσ 2 n is the distance from the one headgroup to the middle<br />

n<br />

of the bilayer and σ n is the spread. The Fourier transform of 5.7 is 1<br />

F{ρ(z)} = ∑ n<br />

A n<br />

2<br />

(<br />

exp − 1 )<br />

2 q2 σn<br />

2 exp(iqz n ) (5.8)<br />

where A n , z n and σ n are parameters, which can be determined by a fitting procedure.<br />

1 The derivation is found in appendix C.


82 Modelling<br />

In detail, the three parts are<br />

ρ sym (z) = ρ head group inside (z) + ρ head group outside (z) − ρ tail (z) (5.9)<br />

where<br />

ρ h inside = √ ρ0A exp 2πσ 2<br />

h<br />

ρ h outside = √ ρ0A exp 2πσ 2<br />

h<br />

(− (z−z h) 2<br />

2σh<br />

2<br />

(− (z+z h) 2<br />

2σ 2 h<br />

( )<br />

ρ tail = − √ ρ0<br />

exp − z2<br />

2πσ 2 2σ 2<br />

t<br />

t<br />

)<br />

)<br />

(5.10)<br />

Matlab Code<br />

☞ page 168<br />

The first two parts of equation 5.9 and hence 5.10 refer to the electron density of the head<br />

group, cf. the label h, whereas the last group describes the tail region, labelled t. The<br />

label inside and outside only refer to the fact that the two headgroups are situated on the<br />

inside and outside of the bilayer of the vesicle, respectively. This naming convention goes<br />

for all of the models even though it really does not make sense in the one-dimensional<br />

models. ρ 0 has the dimension length times the electron density, z h and σ have the<br />

dimension length and A is dimensionless. z h represents half the distance between the<br />

areas with high electron density, the head groups of the phospholipids, and therefore z h<br />

can be used as a measure for the thickness of a lipid monolayer. σh 2 is the variance of the<br />

two first Gauss functions in equation 5.9 and refers to the propagation of electrons along<br />

the normal of the headgroup. σt 2 is the variance of the last Gaussians in equation 5.9 and<br />

refers to the propagation of electrons along the normal of the tails. The dimensionless<br />

parameter A specify the weighing of the three Gaussians.<br />

By using equation 5.8 and equation 5.9 the symmetric form factor becomes<br />

f sym (q) = 1 2q<br />

(<br />

Aexp<br />

(<br />

− 1 2 q2 σ 2 h<br />

)<br />

cos(qz h ) − 1 2 exp(q2 σ 2 t )<br />

)<br />

(5.11)<br />

The factor of 1/q is added to mimic the orientational averaging of the sheets, which<br />

automatically appears in the three-dimensional model in section 5.6. Please note that<br />

this is what others apply to the intensity (which then is squared, 1/q 2 ) and refer to as<br />

the Lorentz factor.<br />

The model is shown in figure 5.4. The intensity was calculated as the absolute square<br />

of the form factor (equation 5.5) and the smeared intensity was calculated using equation<br />

5.6 with the beam profile in figure 5.3. 32 steps were used in the numerical integration 2<br />

of the smearing for this model and all of the other models. Above the beam stop cutoff<br />

(at about 0.025 Å −1 ) there was no significant difference using more than 32 steps as<br />

indicated by the dotted line with 10000 steps in figure 5.5.<br />

2 Using the built-in Matlab function trapz.


5.3 Symmetric (1d) 83<br />

Figure 5.4 The form factor and both point scatterer and smeared intensity for the symmetric<br />

bilayer electron density profile model. For the form factor, equation 5.11 was used with the<br />

parameters A = 0.5, z h = 17 Å, σ h = 4 Å and σ t = 6 Å.<br />

Figure 5.5 The smeared intensity for the symmetric bilayer electron density profile model with<br />

varying amount of steps in the numerical integration done in equation 5.6.


84 Modelling<br />

5.4 Symmetric (4g)<br />

Matlab Code<br />

☞ page 173<br />

On the basis of the electron density profiles of the interdigitated phase found in the<br />

literature, e.g. [60], [80], we have expanded the symmetric (1d) model to consist of 4<br />

Gaussians, two Gaussians representing the tail region and two Gaussians representing the<br />

head region, see figure 5.6. The form factor for the sym-4g model is based on equation<br />

5.8 and is a sum of 4 contributions.<br />

( ( )<br />

f sym-4g (q) = 1<br />

2q<br />

Aexp − q2 σh<br />

2<br />

2<br />

(exp(−iqz h ) + exp(iqz h ))<br />

( )<br />

)<br />

− exp − q2 σ 2 t<br />

2<br />

(exp(−iqz t ) + exp(iqz t ))<br />

z t represents the distance from the bilayer centre to the peak of the Gaussians representing<br />

the tail region. Presuming that the phospholipid chains do interdigitate the tail<br />

region of the electron density profile will differ from the profile represented in figure 5.1.<br />

The electron density will increase in the tail region and the distance from the two head<br />

groups will decrease. The sym-4g model takes into account that the electron density<br />

probably will increase in the middle of the lipid bilayer.<br />

Figure 5.6 The form factor and both point scatterer and smeared intensity for the symmetric<br />

bilayer electron density profile model with 4 Gauss functions. For the form factor, equation<br />

5.12 was used with the parameters A = 1.5, z h = 20 Å, σ h = 4 Å, z t = 3.5 Å and σ t = 2 Å. The<br />

intensity was calculated as the absolute square of the form factor and the smeared intensity<br />

was calculated using equation 5.6 with the beam profile in figure 5.3.


5.5 Asymmetric (1d) 85<br />

5.5 Asymmetric (1d)<br />

Studies of lipid bilayers have indicated that the electron density profile is not completely<br />

symmetric. Brzustowicz & Brunger [9] have shown an asymmetric electron density profile<br />

of the lipid bilayer through model fitting and direct calculation of the form factor of<br />

unilamellar <strong>vesicles</strong> of SOPS. Hirai et al. [23] have also observed bilayer asymmetry in<br />

a mixture of small unilamellar <strong>DPPC</strong> <strong>vesicles</strong> with monosiaganglioside. Brzustowicz &<br />

Brunger [9] use SOPS over <strong>DPPC</strong> arguing that SOPS has a larger difference in scattering<br />

length compared with buffer and that the lipid for this reason is easier to do experiments<br />

on. SOPS differs primary from <strong>DPPC</strong> by having two different acyl chains, one with a<br />

length of 14 C-atoms and one with the length of 18 C-atoms, in addition <strong>DPPC</strong> contains<br />

a choline group whereas SOPS contain a serine group. To describe the asymmetric<br />

electron density profile the two head group Gaussians must be different. Brzustowicz &<br />

Brunger [9] find that the electron density profile of the two headgroups are described by<br />

the below-mentioned parameters<br />

ρ headgroup inside (z) =<br />

ρ headgroup outside (z) =<br />

ρ tail =<br />

)<br />

(1.18 ± 0.03)exp<br />

(− (z+(22.4±0.42)Å)2<br />

2((4.87±0.13)Å) 2 )<br />

(2.78 ± 0.04)exp<br />

(− (z+(19.6±0.46)Å)2<br />

2((2.11±0.06)Å) 2<br />

(<br />

− exp<br />

)<br />

z 2<br />

2((7.76±0.07)Å) 2<br />

(5.12)<br />

(5.13)<br />

(5.14)<br />

We have tried to use these constants, leaving the same four parameters as in the case of<br />

the symmetric model. But this fitted poorly to our data and we decided to discard the<br />

constants and we have developed another model to take the asymmetry into account. In<br />

our asymmetric model we added an amplitude parameter to the symmetric model leaving<br />

the spread of the head groups the same<br />

ρ headgroup inside (z) =<br />

ρ headgroup outside (z) =<br />

ρ tail =<br />

√A 1<br />

exp 2πσ 2<br />

h<br />

√A 2<br />

exp 2πσ 2<br />

h<br />

(<br />

− exp<br />

(− (z−z h) 2<br />

2σh<br />

2<br />

(− (z+z h) 2<br />

z 2<br />

2σ 2 t<br />

2σ 2 h<br />

)<br />

)<br />

)<br />

(5.15)<br />

(5.16)<br />

(5.17)<br />

Using equations 5.8, 5.9 and 5.15 gives the form factor of the asymmetric model<br />

f asym (q) = 1 ( (<br />

exp − q2 σ 2 )<br />

(<br />

h<br />

(A 1 exp(−iqz h ) + A 2 exp(iqz h )) − exp − q2 σ 2 ))<br />

t<br />

2q 2<br />

2<br />

(5.18)<br />

with the five parameters A 1 , A 2 , z h , σ h and σ t . Figure 5.7 shows the form factor, intensity<br />

from both point scatterer and smeared. The intensity was calculated as the absolute<br />

square of the form factor and the smeared intensity was calculated using equation 5.6<br />

with the beam profile in figure 5.3.<br />

Matlab Code<br />

☞ page 169


86 Modelling<br />

Figure 5.7 The form factor and both point scatterer and smeared intensity for the asymmetric<br />

bilayer electron density profile model. For the form factor, equation 5.18 was used with the<br />

parameters A 1 = 1, A 2 = 1.5, z h = 17 Å, σ h = 4 Å and σ t = 6 Å.


5.6 Spherically symmetric (3d) 87<br />

5.6 Spherically symmetric (3d)<br />

Since the bilayers are shells it would be tempting to make a spherically symmetric model,<br />

since you could be lucky to capture the features that the curvature is responsible for. We<br />

developed the model ourselves, but later realized that Brzustowicz & Brunger [9] had<br />

used it before us.<br />

I 3d (q) =<br />

∫ ∞<br />

0<br />

P(R)I 3d (q, R)dR (5.19)<br />

where P(R) is the propability density of a vesicle in the sample having a radius between<br />

R and R + dR. We have estimated P(R) in section 3.4 and intensity for a vesicle with<br />

radius R is<br />

I 3d (q, R) = |f 3d (q, R)| 2 (5.20)<br />

In the fitting procedure equation 5.19 was evaluated with 64 steps in the (trapezoidal)<br />

numerical integration. Below 64 steps the graph becomes noticeably wobbly, and there<br />

is no difference between using 64 and more (256) steps. Even so, the intensity graph of<br />

this model does not become as smooth as the one-dimensional models.<br />

Again, the form factor is the Fourier transform of the electron density distribution<br />

Matlab Code<br />

☞ page 172<br />

f 3d (q, R) ∝ F{ρ(r, R)} ≡<br />

1<br />

(2π) 3/2 ∫u.c.<br />

ρ(r, R)exp(−iq · r)d 3 r (5.21)<br />

since the integrand is independent of the azimuthal angle φ, you get<br />

and equation 5.21 becomes 3<br />

f 3d (q, R) ∝<br />

d 3 r = 2πr 2 sinθ dθ dr and q · r = qr cosθ<br />

=<br />

=<br />

= 1 q<br />

∫∫<br />

1<br />

ρ(r, R)exp(−iqr cosθ)2πr 2 sin θ dθ dr (5.22)<br />

(2π) 3/2 ∫<br />

1 ∞ ∫ π<br />

√ ρ(r, R)r 2 exp(−iqr cosθ)sin θ dθ dr (5.23)<br />

2π<br />

0<br />

∫<br />

1 ∞<br />

√<br />

2π<br />

√<br />

2<br />

π<br />

0<br />

∫ ∞<br />

0<br />

0<br />

ρ(r, R)r 2 2 sin(qr)<br />

qr<br />

dr (5.24)<br />

ρ(r, R)r sin(qr)dr (5.25)<br />

if the electron density function ρ(r, R) is spherically symmetric. One has to pick a<br />

function, which makes the integral converge. It is important to notice the factor 1 q is<br />

independent of which ρ(r, R) we choose. In the case of a Gaussian 4 the integration over<br />

r becomes[9]<br />

f 3d (q, R) = σ (<br />

r<br />

q exp − q2 σr<br />

2 ) (R<br />

sin(qR) + σ<br />

2<br />

2<br />

r q cos(qR) ) (5.26)<br />

Matlab Code<br />

☞ page 171<br />

3 R h i<br />

π<br />

0 exp(−iqr cos θ)sinθ dθ = exp(−iqr cos θ) π<br />

iqr = exp(iqr)−exp(−iqr)<br />

0<br />

iqr<br />

exp(iψ)−exp(−iψ)<br />

2i<br />

4 ρ(r, R) =<br />

1 √ 2πσ 2 r<br />

and d exp(−iqr cos θ) = exp(−iqr cos θ)iqr sin θ<br />

dθ ”<br />

exp<br />

“− (r−R)2<br />

2σ 2 r<br />

= 2 sin(qr)<br />

qr , since sin(ψ) =


88 Modelling<br />

in which the peak of the spherically symmetric Gauss function is at R, and the spread<br />

is σ r .<br />

The electron density profile used in the symmetric (1d) model has been the basis<br />

for the symmetric (3d) model, as the sym-1d model was the 1d model, which fitted the<br />

spectra best. The underlying idea of extending the sym-1d model from 1 dimension to 3<br />

dimensions was to take into account the variation in vesicle radius.<br />

The form factor of the 3d model differs notably from the other form factors. Imagining<br />

it is was possible to make a sample with <strong>vesicles</strong> of only one radius then the intensity<br />

would look like the intensity in figure 5.8.<br />

Figure 5.8 The form factor and both point scatterer and smeared intensity for the spherically<br />

symmetric bilayer electron density profile model. The form factor was calculated using equation<br />

5.26 with the parameters A = 0.5, z h = 17 Å, σ h = 6 Å, σ t = 6 Å and R = 600 Å, i.e. a<br />

single vesicle radius. The intensity was calculated using equation 5.19 (with a vesicle radius<br />

distribution) and the smeared intensity was calculated using equation 5.6 with the beam profile<br />

in figure 5.3.<br />

5.7 Multilamellar <strong>vesicles</strong><br />

During the sample preparation most of the multilamellar <strong>vesicles</strong> have been removed,<br />

but even if there is only a small percentage left in the final solution it is visible in the<br />

SAXS experiment. Unlike ULV the MLV Bragg-scatter much more due to the lamellar<br />

repeat distance. In modelling the SAXS data the presence of MLV has to be taken into<br />

account. Still assuming no interference between the <strong>vesicles</strong> in the solution the equation<br />

describing the measured intensity from the ULV is added the measured intensity deriving<br />

from the scattering from MLV. The total measured intensity is then<br />

I p (q) ∝ I ULV (q) + A b I MLV (q) (5.27)


5.7 Multilamellar <strong>vesicles</strong> 89<br />

In figure 5.9 the result of a SAXS experiment on a suspension containing multilamellar<br />

<strong>vesicles</strong> is shown. This spectrum clearly differs from the spectra represented in figure 4.12<br />

by displaying several distinct peaks. These peaks are found with the same ∆q between<br />

them, which indicates the repeating structure characterizing the multilamellar <strong>vesicles</strong>.<br />

By using equation 2.45 the distance between the two lipid bilayers is found.<br />

40<br />

Multilamellar <strong>vesicles</strong><br />

25 °C<br />

Raw data<br />

Structure factor<br />

30<br />

Intensity [a.u.]<br />

20<br />

10<br />

0<br />

0 0.1 0.2 0.3 0.4<br />

q [Å -1 ]<br />

Figure 5.9 The raw data from a SAXS experiment on MLV of <strong>DPPC</strong>. The distance between<br />

the two peaks is ∆q = (0.1003 ± 0.0005)Å −1 , corresponding to a distance of d = 2π = 62.6 Å.<br />

∆q<br />

This distance is equivalent to a lamellar repeat distance [53].<br />

The structure factor used in the description of the intensity contribution from the<br />

scattering of multilamellar <strong>vesicles</strong> can be approximated as a sum of Gaussians<br />

S(q) = ∑ A n<br />

√ exp<br />

(− (q − nq 0) 2 )<br />

n=1 2πσ<br />

2<br />

b<br />

2σb<br />

2 (5.28)<br />

The fitted structure factor in figure 5.9 is a sum of three Gaussians<br />

1<br />

S(q) = √<br />

(exp 2πσ 2<br />

b<br />

(− (q−q0)2<br />

2σ 2 b<br />

+ 0.1Aexp<br />

)<br />

+ Aexp<br />

(− (q−3q0)2<br />

2σ 2 b<br />

(− (q−2q0)2<br />

) )<br />

2σ 2 b<br />

)<br />

(5.29)<br />

Matlab Code<br />

☞ page 174<br />

where A = 0.22, q 0 = 0.1003Å −1 and σ b = 0.0128Å.<br />

There were some experimental difficulties related to the SAXS experiments on the multilamellar<br />

<strong>vesicles</strong> at 25 ◦ C. Normally a spectrum was obtained by letting the Kratky<br />

camera measure for 6 hours in a row, but with a sample of the multilamellar <strong>vesicles</strong> of<br />

<strong>DPPC</strong> this was not possible. After a short period of time (half an hour) the <strong>vesicles</strong> were<br />

no longer randomly distributed within the suspension. They fell down the bottom of the<br />

glass tube. A way to circumvent this problem would be to dilute the lipid suspension, but<br />

then the experimental time would have to be increased. Instead the cuvette was taken<br />

out of the camera every hour and shaken until the suspension was milk-white again. The<br />

experiment was subsequently resumed and repeating this procedure over a period of 10<br />

hours a spectrum of MLV of <strong>DPPC</strong> was obtained. The experiment with the same sample


90 Modelling<br />

at 45 ◦ C, however, was carried out without a hitch. The fitted structure factor is plotted<br />

with the raw data in figure 5.10<br />

30<br />

Multilamellar <strong>vesicles</strong><br />

45 °C<br />

25<br />

Raw data<br />

Structure factor<br />

Intensity [a.u.]<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 0.1 0.2 0.3 0.4<br />

q [Å -1 ]<br />

Figure 5.10 The raw data from a SAXS experiment on MLV of <strong>DPPC</strong>. The distance between<br />

the two peaks is ∆q = (0.096 ± 0.0005)Å −1 , corresponding to a distance of d = 2π =65.5 Å.<br />

∆q<br />

This distance is equivalent to a lamellar repeat distance [53].<br />

The fitted structure factor in figure 5.10 is a sum of two Gaussians<br />

Matlab Code<br />

1<br />

☞ page 174 S(q) = √<br />

(exp<br />

(− (q − q 0) 2 )<br />

+ Aexp<br />

(− (q − 2 · q 0) 2 ))<br />

2πσ<br />

2<br />

b<br />

2σ 2 b<br />

2σ 2 b<br />

(5.30)<br />

where A = 0.3, q 0 = 0.096Å −1 and σ b = 0.0159Å.<br />

There is a danger of modelling the structure factor as an intensity and not an amplitude.<br />

The intensity is by definition a positive quantity whereas the amplitude can be negative<br />

as well as positive. Modelling the intensity does not take into account the possible negative<br />

amplitudes of structure factor and hence their contribution to the total scattering<br />

intensity.<br />

5.8 Crystalline structure<br />

None of the models could capture the feature of the small peaks at low q-values, see<br />

table 4.4. The peaks are possibly Bragg reflections (which we unfortunately realized<br />

after the modelling) and we have tried to match them with the Bragg peak of hexagonal<br />

structures, since this have been observed in perturbated lipid systems as well as high<br />

temperature (above 70-80 ◦ C, the H II phase) lipid systems before - the existence of the<br />

H II phase is well documented [20], [36], [61].<br />

In section 2.3 page 27 the relation between q-values and the location of the peak in<br />

a hexagonal lattice structure is described. The relation between the q-values in the<br />

hexagonal structure does not apply conclusively to the location of the peaks in the spectra<br />

obtained in this work. Only if the first peak q ′ was located at 0.053 Å −1 instead of around<br />

0.042 Å −1 the next two peaks actually fits q ′′ = √ 3q ′ and q ′′′ = √ 4q ′ , reasonably. But the<br />

difference is too great 0.053Å−1 −0.042Å −1<br />

≈ 26% to be within experimental uncertainty.<br />

0.042Å −1<br />

Then one or more of the peaks could refer to scattering from the hexagonal rods (l ≠ 0)


5.8 Crystalline structure 91<br />

Figure 5.11 The point scatterer intensity and smeared intensity for the symmetric bilayer<br />

electron density profile model (sym-mlv) with multilamellar correction. The intensity was<br />

calculated in equation 5.27 and 5.5 using the parameters A = 0.5, z h = 17 Å, σ h = 4 Å,<br />

σ t = 6 Å, A b = 0.05 and q 0 = 0.1Å −1 . The smeared intensity was calculated using equation 5.6<br />

with the beam profile in figure 5.3.<br />

with a characteristic length. Another explanation of one of the peaks could be the<br />

presence of multilamellar <strong>vesicles</strong>. As seen in figure 5.9 and 5.10 the presence of MLVs<br />

gives rise to a distinct peak around q = 0.1Å −1 . And, furthermore, the distances could<br />

be larger than what we can observe - with a peak hidden by the beam stop. Nevertheless,<br />

none of these explanations could account for the peaks to indicate a hexagonal lattice<br />

structure. Actually, the small peaks at the spectra between q = 0.1Å −1 and q = 0.2Å −1<br />

could be indicative of a more complicated structure than the presence of the hexagonal<br />

lattice. Similar peaks are found in Katsaras & Gutberlet [32, page 258] as an example<br />

of the presence of the complex cubic phase.<br />

Retrospectively, we should have tried to separate the smooth envelope curve and the<br />

peaks and model them individually. If the models were developed so they could account<br />

for the peaks and the envelope curve individually then the models used would probably fit<br />

a lot better. It would also assess the value of the bilayer thickness with less uncertainty.<br />

Developing a model which could fit the peaks would partly give information about the<br />

underlying structure giving rise to the peaks, partly indicate how much of the total<br />

structure were in fact <strong>vesicles</strong> and how much were crystalline structure.


92 Modelling<br />

Form factor<br />

Intensity from point scatterer<br />

Intensity (a.u.)<br />

Electron density (a.u.)<br />

-40 -30 -20 -10 0 10 20 30 40<br />

z / Å<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Figure 5.12 The point scatterer intensity and smeared intensity for the spherical symmetric<br />

bilayer electron density profile model (3d-mlv) with multilamellar correction. The intensity<br />

was calculated in equation 5.27 and 5.5 using the parameters A = 0.5, z h = 17 Å, σ h = 4 Å,<br />

σ t = 6 Å, A b = 0.02 and q 0 = 0.1Å −1 . The smeared intensity was calculated using equation 5.6<br />

with the beam profile in figure 5.3.<br />

5.9 Results and summary<br />

The quality of the fits is evaluated by a function, sse (sum of squares due to error) based<br />

on the least-squares method, which minimizes the sum of the squares of the divergence:<br />

sse(x) = ∑ i<br />

[I measured (q i ) − I model (x, q i )] 2 (5.31)<br />

Matlab Code<br />

☞ page 177<br />

Here (q i , I measured (q i )) is the i’th point of measurement and I model (x, q i ) is the intensity<br />

calculated by the model, which is a function of q i and the fitted parameters, x. The<br />

value of sse determines the quality of the fit (lower the value → better the fit). Please<br />

note, that both I measured and I model are normalized before the subtraction such that<br />

∫ q2<br />

I(q)dq = 1 (5.32)<br />

q 1<br />

where q 1 and q 2 are the lower and upper limit on our SAXS spectra.<br />

We have used an altered version 5 of the fitting function in the MATLAB add-on package,<br />

nlinfit, which uses a Gauss-Newton algorithm to minimize the sse function. The<br />

method is an effective way to solve non-linear least-squares problems. It calculates the<br />

gradient from partial derivatives and moves towards minimum. However in using this<br />

method you might end up in one of the, possibly many, local minima and not the global<br />

minimum.<br />

5 We have modified the script in order to fit our needs.


5.9 Results and summary 93<br />

We do not only use the sse as criterion for a good model fit. We require a reasonable<br />

correlation between the parameters as well. Combining this with simply looking at a<br />

plot of the modelled curve with the raw data determined whether the fitting procedure<br />

should be repeated. Besides looking at the model fitted with the raw data and the sse<br />

value the parameters from the model should also be realistic. Some constraints were<br />

embodied in the fitting procedure making sure e.g. that the calculated amount of MLVs<br />

in the sample was not negative. Most of the evaluations of the model parameters were<br />

based on subjective judgements. The values for the bilayer thickness, 2z h , the deviation<br />

of the head group, σ h , and tail, σ t , as well as z t (sym-4g model) should all be positive.<br />

z h should be between 5 Å and 30 Å, the deviations between 1 Å and 15 Å and z t (sym-4g<br />

model) should be between 0 and 5 Å. It is important to stress that these ranges are all<br />

based on either experience or literature values.<br />

Generally a model fit can be discarded if there is an unrealistic correlation between the<br />

model parameters. First of all the deviations of the tail and head regions should not<br />

differ significantly from each other, and the values for the deviations should be lower<br />

than the values for z h . If the deviations are higher than z h then the electron density<br />

profile would have much broader peaks, and the peak (z h ) would be determined with<br />

high uncertainty.<br />

The result of a good fit selection process (to narrow our analysis down) is listed in table<br />

5.2. Please refer to appendix H (p. 135) for a complete list of model graphs and fitted<br />

parameters. One of the good model fits can be found in figure 5.13. This fit shows one<br />

of the best fits to the sym-mlv model with an sse of only 8.6.<br />

Figure 5.13 also demonstrates the great difference in the modelled electron density<br />

profile compared to the relatively small difference of the fit to the measured data. The<br />

model fits plotted with the raw spectrum clearly display that none of the models take the<br />

distinct peaks at low q-values into account. The best fits are from the sym-mlv model<br />

and 3d-mlv. Even though these two models have almost the same fitting parameters<br />

the electron density profiles differ. The bilayer thickness directly read from the electron<br />

density profile is lower for the 3d-mlv model than for the sym-mlv model.<br />

In appendix H it is important to notice the absurdly large parameter uncertainties especially<br />

related to the sym-4g model, e.g. for the fit of heptanol, high temperature, series 10<br />

the sse is 16.9, but the value for z h is 0.2ű8000Å. A large parameter uncertainty means<br />

that the model is insensitive to changes of that parameter (in the particular minimum<br />

found). 6 These parameters have been excluded from the further analysis.<br />

The bilayer thickness (2z h ) is an average of billions of <strong>vesicles</strong> and you have to take both<br />

z h and σ h into account as they interdepend. The criteria for the selection of z h -values<br />

are sse values under 40, and a reasonably low deviation of the head group Gaussian<br />

σ h . From the models listed in table 5.2 the z h values are found. The parameter values<br />

(z h and σ h ) are listed in table 5.3. sym-mlv has the best over-all fit and the parameter<br />

values (z h and σ h ) are listed i table 5.4. The values in the two tables show that adding<br />

alcohol to the sample does not change the z h or σ h values much taking the uncertainties<br />

into account. They also show that most of the fits overlap in their confidence intervals<br />

between the low temperature and high temperature which means we cannot argue that<br />

there is a difference in the bilayer thickness for the two phases.<br />

6 The uncertainties are calculated on the basis of partly the residuals, partly the Jacobian (the Jacobian<br />

is the partial derivatives of each the parameters (and for each of the data points) with respect to the<br />

sse).


94 Modelling<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=17.3<br />

Asym model, sse=18.8<br />

Sym-mlv model, sse=8.6<br />

3d model, sse=17.2<br />

3d-mlv model, sse=8.6<br />

Sym-4g model, sse=16.9<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure 5.13 The fit of <strong>DPPC</strong> with heptanol, high temperature (45 ◦ C), Series 10<br />

Generally, the models with MLV correction fit the data best (lowest sse), this goes for<br />

the pure <strong>DPPC</strong> samples as well as the samples added alcohol. But these models tend<br />

to have an increased deviation for the tail region, σ t . Another general and interesting<br />

result is that the high temperature measurements are fitted better than the low temperature<br />

measurements. This could partially be explained as a consequence of the samples<br />

containing more multilamellar <strong>vesicles</strong> in the high temperature phases than in the low<br />

temperature phases. It seems to be the case and is indicated by the generally larger A b<br />

values.<br />

Evaluating the models individually, sym-mlv had the best fits overall. The 3d-mlv model<br />

fitted the pentanol and hexanol data best, and the sym-4g model fitted hexanol data<br />

best. The fact that the one-dimensional model with correction for the presence of MLVs,<br />

sym-mlv, overall fits the data best can indicate that most of the system do consist of<br />

symmetric lipid bilayers. It also indicates that you need to take the presence of MLVs<br />

into account and that all the MLVs have not been removed during the sample preparation


5.9 Results and summary 95<br />

Low temperature High temperature<br />

Pure sym (20)<br />

sym-mlv (10.2) sym-mlv (4.8)<br />

3d (20.3)<br />

3d-mlv (10.3) 3d-mlv (4.8)<br />

sym-4g (20.1)<br />

Pentanol sym (13)<br />

asym (12.5)<br />

sym-mlv (10.4/15.3)<br />

3d (12.8)<br />

3d-mlv (12.8) 3d-mlv (10.4/15.3)<br />

Hexanol sym (33.5) sym (18.5/15.8)<br />

asym (39) asym (23.1)<br />

sym-mlv (13.9/14.8)<br />

3d (33.8) 3d (18.5/15.8)<br />

3d-mlv (13.9/14.8)<br />

sym-4g (33.3) sym-4g (18.5/15.8)<br />

Heptanol sym (41)<br />

asym (18.8)<br />

sym-mlv (9.58) sym-mlv (8.61/18)<br />

3d (42)<br />

3d-mlv (15/35.7) 3d-mlv (8.58/18)<br />

sym-4g (41.6)<br />

Table 5.2 The models used for interpreting the SAXS data. The sse values are listed in<br />

brackets. Low and high temperature refer to 25 ◦ C and 45 ◦ C, respectively.<br />

Low temperature (25 ◦ C) High temperature (45 ◦ C)<br />

Alcohol z h / Å σ h / Å z h / Å σ h / Å<br />

None 19.2 ± 0.8 6.6 ± 0.5 (5) 18.8 ± 0.2 3.8 ± 2 (2)<br />

Pentanol 18.2 ± 0.9 6.7 ± 0.7 (3) 16.5 ± 0.4 4.5 ± 0.4 (4)<br />

Hexanol 18.1 ± 1.3 5.8 ± 0.8 (5) 17.6 ± 1.3 4.2 ± 0.7 (11)<br />

Heptanol 19.4 ± 0.5 5.3 ± 0.5 (6) 18.1 ± 0.3 3.8 ± 0.3 (5)<br />

Table 5.3 The averaged values for z h and σ h from the different models. The figures in brackets<br />

refer to the number of models used for the averaging.<br />

Low temperature (25 ◦ C) High temperature (45 ◦ C)<br />

Alcohol z h / Å σ h / Å z h / Å σ h / Å<br />

None 17.9 ± 0.4 7.7 ± 0.4 (1) 18.8 ± 0.2 3.8 ± 0.2 (1)<br />

Pentanol 18 ± 5 6 ± 2 (1) 16.5 ± 0.4 4.5 ± 0.4 (2)<br />

Hexanol 16 ± 2 10 ± 2 (1) 18.0 ± 0.5 4.0 ± 0.4 (2)<br />

Heptanol 17.1 ± 0.5 8.0 ± 0.5 (1) 18.0 ± 0.4 3.7 ± 0.4 (2)<br />

Table 5.4 The averaged values for z h and σ h from the sym-mlv model. The figures in brackets<br />

refer to the number of data sets used for the averaging.


96 Modelling<br />

procedure.<br />

Additionally, based on prior experiences with modelling electron density profiles of similar<br />

systems [59], we found it necessary to extend the models to take into account the presence<br />

of MLVs. The sym-mlv and 3d-mlv are two models which overall fit the best, this stresses<br />

the importance of embedding the structure factor in the modelling.<br />

For the asym model the parameters A 1 , A 2 and z h are often stated with great uncertainty.<br />

For the spherical symmetric model the parameter σ t is generally stated with large<br />

uncertainties. When extending the models to correct for the presence of MLVs the fitted<br />

parameters z h , σ h and σ t from the “non”-extendend model are often changing upon the<br />

new fitting procedure. Still the parameters from fitting the models sym-mlv and 3d-mlv<br />

are frequently very similar. Also for the asym model it is noted that mostly there are no<br />

difference in the values of the two amplitudes, and this indicates that the lipid bilayer is<br />

relatively symmetrical. The fitted values from the asym model are overall fairly similar<br />

to those obtained with the sym model.<br />

The asymmetric model asym has been used for similar systems previously [9], [59] with<br />

good results. This model accounts for possible asymmetry in the bilayer, a feature one<br />

could imagine being more pronounced as the amount of partitioned alcohol increased.<br />

However, the modelling results did far from convincingly detect the presence of any<br />

asymmetry within the lipid bilayer.<br />

This alcohol induced interdigitated phase of <strong>DPPC</strong> has not been modelled previously,<br />

but based on electron density profiles in the literature, [60], [80], describing the ethanol<br />

induced interdigitated phase we have developed the sym-4g model. The results from the<br />

sym-4g model did not support the hypothesis that the spectra obtained at low temperature<br />

were interdigitated, at least not if taking the electron profiles presented by [60], [80]<br />

as being valid.<br />

Regarding the sensitivity of the parameters we have not conducted a sensitivity analysis<br />

per se. Still from the uncertainties of the parameters given in the fitting process some<br />

conclusions regarding the sensitivity of the parameters can be made. The modelling<br />

results from especially the sym-4g model and the asym state values of the different fitting<br />

parameters with absurdly large uncertainties. These uncertainties indicate that the<br />

overall fitting of the spectrum and the sse value is not much affected by changing this<br />

specific parameter. Mainly the large uncertainties centre on the amplitudes (asym) and<br />

the value for z h and z t (sym-4g).<br />

The comparable values found in the literature mostly relate to the bilayer thickness,<br />

which for pure <strong>DPPC</strong> depends on temperature - the thickness decreases with increasing<br />

temperature. Within the temperature range used in this work, the value for 2z h should<br />

change from 44.2 Å (20 ◦ C) to 38.3 Å (50 ◦ C) [53]. In general the thickness of a lipid bilayer<br />

decreases as the concentration of alcohol is increased [47], [59]. If the interdigitated phase<br />

is obtained the bilayer thickness is decreased significantly from 2z h = 42 Å to 2z h = 30 Å<br />

(<strong>DPPC</strong> and ethanol, [68]). The bilayer thickness of the interdigitated phase of <strong>DPPC</strong><br />

depends on the alcohol used, the longer the chain, the larger z h , [1]. This can be explained<br />

by referring to the fact that the thickness of an interdigitated bilayer is the sum of the<br />

the lipid chains and the alcohol chains, hence when lengthening the alcohol, the sum<br />

of alcohol-chain and lipid chain will increase. By increasing the used alcohol with a<br />

methylgroup the fully interdigitated lipid bilayer thickness would increase by about 1Å<br />

[1].


5.9 Results and summary 97<br />

2z h Lipid phase Inducer Reference<br />

42/30 Å L β ′/L βI Ethanol Simon & McIntosh [68]<br />

62/49 Å L β ′/L βI Ethanol Vierl et al. [80]<br />

63/49 Å L β ′/L βI Ethanol Nambi et al. [55]<br />

43/30 Å L β ′/L βI Ethanol Simon et al. [69]<br />

49/41 Å L β ′/L βI Ethanol Adachi et al. [1]<br />

45/19 Å L β ′/L βI Ethanol Mou et al. [52]<br />

30 Å L βI Chlorpromazine McIntosh et al. [50]<br />

30 Å L βI not mentioned Veiro et al. [79]<br />

44.2Å L β ′ None Nagle & Tristram-Nagle [53]<br />

38.8 Å L α None Nagle & Tristram-Nagle [53]<br />

Table 5.5 Chosen z h from the literature, <strong>DPPC</strong> and ethanol.<br />

The bilayer thicknesses in table 5.5 are varying from 42-63 Å in the L β ′ and from<br />

19-49 Å in the L βI . It should be noted that all the SAXS spectra have been obtained<br />

with <strong>DPPC</strong> and ethanol as an inducer. Therefore the bilayer thicknesses are not directly<br />

comparable to the modelled thicknesses in this work. The values from the literature reveal<br />

great variation in the obtained values for the distance between the two head groups, still<br />

there is a trend to measuring lower values for the lipid bilayer thickness in the L βI phase.<br />

There is an overlap of 7 Å between values for the L βI phase and the L β ′ phase, within<br />

this range it is impossible to determine whether the lipid is in one or the other phase.<br />

The modelled values for the bilayer thickness obtained in this work are just below this<br />

overlapping range, both for the pure and perturbated <strong>vesicles</strong>.


6 Discussion<br />

Are the <strong>vesicles</strong> alcohol saturated?<br />

The ITC experiments give us ambiguous results and hence no clear-cut determination of<br />

whether the <strong>vesicles</strong> are alcohol saturated or not. The alcohol actually getting into the<br />

bilayer, however, is supported by the SAXS results which indicate the greater part of the<br />

system still has vesicle structure. Thus, we believe the heat exchange does dominantly<br />

stem from alcohol molecules moving into and out of the bilayer. And, if the temporal<br />

change of the partition coefficients can be explained by water vapourization the ITC<br />

experiments show that the <strong>vesicles</strong> are saturated after 24 hours - and still are after two<br />

days. However, the evidence is not that compelling as we do not know anything about<br />

the spooky sub-system.<br />

Are the <strong>vesicles</strong> interdigitated?<br />

The DSC scans have three indications of the lipid bilayers being interdigitated: the main<br />

phase transition temperature decreases with increasing alcohol concentration, the peak<br />

broadens (and has a higher enthalpy) and the small peak indicating the pretransition is<br />

not detected. Thus the DSC results lead us to believe that we have reached the L α -L βI<br />

phase transition. The SAXS spectra show a small but apparently unmistakable difference<br />

between the low and high temperature measurements. Still, the modelling of the SAXS<br />

spectra did not allow us to conclude that the bilayer thicknesses had changed notably.<br />

The modelling results show that adding alcohol do influence the structure of the lipid<br />

<strong>vesicles</strong>. This is primarily reflected in the difference in the obtained spectra from the pure<br />

<strong>DPPC</strong> suspension and the suspension with <strong>DPPC</strong> and alcohol. From the modelling of<br />

the lipid bilayer thickness no clear-cut conclusions could be made regarding the specific<br />

lipid phase. Still, the results of the modelling, i.e. the electron density profile as well<br />

as the values for z h , show that adding alcohol indeed do have an impact on the lipid<br />

membrane. We cannot rule out the possibility of domain interdigitation. The generally<br />

low z h values actually support this hypothesis. An average of interdigitated and noninterdigitated<br />

<strong>vesicles</strong> could possibly produce the SAXS spectra we see. The modelling<br />

results reflect an average over all the <strong>vesicles</strong> in the suspension, and one could argue that<br />

a part of the <strong>vesicles</strong> was interdigitated and another part was not fully interdigitated. In<br />

the modelling we did not take the different bilayer thicknesses into account, still from the<br />

modelled electron density profiles, appendix H, it looks like there is a slight distribution<br />

of the lipid head group - head group length, as the spread (σ h ) is quite large. This<br />

distribution could cover up the presence of two different lipid bilayer thicknesses.<br />

Another issue in the discussion of whether we have obtained the L βI phase concerns<br />

whether we have added enough alcohol to the lipid suspension. The total molar lipid<br />

to alcohol ratio was 1:2 for all the alcohols used. According to the values found in the<br />

literature, and found by ITC, this should be enough to induce the interdigitated phase.<br />

99


100 Discussion<br />

As we have not found examples of experiments with this specific lipid concentration<br />

in the literature (most experiments are performed with lower lipid concentration) we<br />

needed to estimate the total molar lipid to alcohol ratio in this work. We chose the same<br />

total molar lipid to alcohol ratio for all the alcohols even though we knew the partition<br />

coefficients are higher for longer n-alcohol.<br />

The enthalpy of transition calculated from the DSC results is generally higher (around<br />

20%) than the literature values. This could indicate that the transition enthalpy might<br />

cover more than just the transition from the L βI phase to the L α phase. As mentioned<br />

earlier the SAXS spectra display peaks which could be interpreted as the presence of a<br />

more complex structure than the L βI phase and the L α phase. A transition within this<br />

structure could contribute to the phase transition enthalpy. As we have not determined<br />

the structure for the possibly crystalline aggregates, we cannot comment on the size of<br />

the contribution to the phase transition enthalpy.<br />

Where is the alcohol?<br />

By letting the system settle for 48 hours at 10 ◦ C above the main phase transition temperature<br />

before beginning the experiments we expect the alcohol to be evenly distributed<br />

throughout the sample during measurements. And we do certainly not expect a situation<br />

with completely alcohol-free <strong>vesicles</strong> and alcohol-saturated <strong>vesicles</strong> side by side. However,<br />

we do not have any evidence to support these assumptions, so how can we tell where<br />

the alcohol is? The modelling of the SAXS data should give us a hint, but none of the<br />

models distinguished themselves in fitting the L βI phase or the L α phase better. Still,<br />

from the experiments in this work the only indicator of where the alcohols are located<br />

in the membrane is the modelled electron density profiles. By comparing the electron<br />

density profile of the pure lipid bilayer to the profile of the perturbated lipid bilayer one<br />

can see that the spread of the head group region is lower in the modelled profile of the<br />

pure lipid bilayer than in the perturbed lipid bilayer. This shows that the location of<br />

the head groups is more well-defined for the pure lipid bilayer. Adding alcohol might not<br />

result in a distinct decrease in the lipid bilayer thickness (insofar as our modelling) but it<br />

definitely has an effect on the location of the phospholipid head group. This is reflected<br />

in the broadening of the spread of the Gaussians representing the lipid head groups. The<br />

broadening of the head group could also be interpreted as a sign of the location of the<br />

alcohol, i.e. the alcohol could be placed with the hydrophobic tail along the fatty acid<br />

chains of the phospholipid and the hydrophilic head group could be located next to the<br />

much larger phosphatidylcholine group. The acyl chain of the alcohols is fairly long and<br />

it would be energetically favourable for the alcohol to place its tail down the lipid membrane.<br />

For lipid membranes perturbed by a lower concentration of alcohol, this location<br />

has been demonstrated using MD-simulations [59].<br />

For the three n-alcohols used there are three different solubilities in water, decreasing<br />

values with increasing acyl chain length. When the concentration of alcohol in the<br />

lipid suspension approaches this concentration, the risk of unstabilizing the lipid-alcohol<br />

system increases. For both hexanol and heptanol this limit is formally crossed, still more<br />

alcohol can be dissolved in a lipid suspension than in pure water as the alcohol partitions<br />

into the membrane and in doing so it does not contribute to the bulk concentration of<br />

alcohol. If a larger amount of alcohol is added (than we have used) the alcohol would<br />

probably aggregate and phase separate from the lipid suspension but there is a risk that<br />

the lipid molecules of the <strong>vesicles</strong> simply will be dissolved in a solution containing alcohol<br />

and water. This is not the case in this work as the SAXS spectra show signs of distinct


structures within the lipid suspension. Even so, by working with such a large alcohol<br />

concentration the possibility increases that the lipid molecules form other structures than<br />

bilayers. We could not have avoided this risk as the lipid concentration had to be at least<br />

5 % in order to get a reasonable signal to noise ratio in the SAXS experiments, and we<br />

had to add enough alcohol to be certain that the interdigitated phase could be obtained.<br />

101


102


7 Conclusion<br />

With respect to interdigitation our results are inconclusive, we can neither reject nor<br />

confirm for any of the alcohols that our low temperature measurements are in fact in<br />

the L βI phase. According to the literature and our ITC measurements we should have<br />

enough alcohol to induce the phase, but the results from the modelling on the SAXS<br />

data do not allow such a clear-cut conclusion.<br />

The system behaved more complex than initially expected. The DSC measurements<br />

indicate a stable system with a well-defined phase transition, but the ITC and the SAXS<br />

measurements did not. It is still unclear as to why the samples are very sensitive to the<br />

thermal history after adding the alcohol. But using our sample preparation method, we<br />

do get systems which give reproducible SAXS-spectra. The modelling of these spectra<br />

did not completely account for all the features seen in the spectra. They indicate that a<br />

large fraction of the system still has vesicle structure and a smaller fraction of the system<br />

form larger crystalline aggregates, whose structures are yet unknown to us.<br />

Numerical results and trends<br />

The transition temperature was found to decrease as alcohol was added to the lipid<br />

suspension. The phase transition temperature decreased from (41 ± 1) ◦ C (pure <strong>DPPC</strong>)<br />

to (34 ±1) ◦ C for <strong>DPPC</strong> and pentanol, (29 ±1) ◦ C for <strong>DPPC</strong> and hexanol and (35 ±1) ◦ C<br />

for <strong>DPPC</strong> and heptanol. The threshold ratio for inducing the low temperature phase was<br />

found to be lower than the ratios stated in the literature. The SAXS spectra modelling<br />

analysis revealed nearly the same values for the lipid bilayer thicknesses for all of the<br />

measurements in the order of 34-38 Å. Adding alcohol mainly seemed to decrease the<br />

bilayer thickness slightly independent of the temperature. And for all alcohols there are<br />

slight tendencies towards a lower bilayer thickness as the temperature is increased.<br />

Perspectives<br />

There are several threads to pick up after this project. More SAXS experiments would<br />

either confirm or reject the idea of a crystalline sub-system. If it could be confirmed, one<br />

could look for an equilibrium in which a larger part or the whole system is a crystalline<br />

structure to see if it is an independent phase. Even in the case of a negative confirmation,<br />

it would be interesting to perform light scattering experiments on the system in order to<br />

find the aggregate size distribution.<br />

Neutron scattering (supported by simulations, e.g. Molecular dynamics) on the system<br />

using deuteriated alcohol might reveal the preferred location of the alcohol within the<br />

membrane.<br />

103


104


Bibliography<br />

[1] Adachi, T., Takahashi, H., Ohki, K. & Hatta, I. [1995]. Interdigitated Structure of<br />

Phospholipid-Alcohol Systems Studied by X-Ray Diffraction, Biophysical Journal<br />

68: 1850–1855.<br />

[2] Agarwal, B. [1991]. X-ray Spectroscopy, Kluwer Academic Publichers, Dordrecht.<br />

[3] Als-Nielsen, J. & McMorrow, D. [2001]. Elements of Modern X-Ray Physics, John<br />

Wiley and Sons, Ltd, West Sussex, England.<br />

[4] Basic Support for Cooperative Work (BSCW) - online version control of shared<br />

documents [2006].<br />

URL: http://bscw.fit.fraunhofer.de/<br />

[5] Biltonen, R. L. [1990]. A statistical-thermodynamic view of cooperative structural<br />

changes in phospholipid bilayer membranes: their potential role in biological function,<br />

J. Chem. Thermodynamics 22: 1–19.<br />

[6] Bouwstra, J. A., Gooris, G. S., Bras, W. & Talsma, H. [1993]. Small angle X-ray<br />

scattering: possibilities and limitations in characterization of <strong>vesicles</strong>, Chemistry<br />

and Physics of Lipids 64: 83–98.<br />

[7] Brown, M. E. [1998]. Handbook of thermal analysis and calorimetry, vol 1, Elsevier,<br />

Amsterdam.<br />

[8] Brumberger, H. [1995]. Modern Aspects of Small-Angle Scattering, Kluwer Academic<br />

Publichers, Dordrecht.<br />

[9] Brzustowicz, M. R. & Brunger, A. T. [2005]. X-ray scattering from unilamellar lipid<br />

<strong>vesicles</strong>, Journal of Applied Crystallography 38: 126–131.<br />

[10] Bánó, M. & Pajdalová, O. [1999]. Interaction of the local anaesthetic heptacaine with<br />

dipalmitoylphosphatidylcholine lipsomes: A densimetric study, Biophysical Chemistry<br />

80: 53–66.<br />

[11] Bøggild, S., Nielsen, M. H. & Petersen, B. W. [1996]. Røntgendiffraktion af<br />

blandinger af symmetriske diblockcopolymerer, IMFUFA, Roskilde Universitetscenter.<br />

[12] Cevc, G. [1991]. Isothermal lipid phase transitions, Chemistry and Physics of Lipids<br />

57: 293–307.<br />

[13] Cevc, G. & Marsh, D. [1987]. Phospholipid Bilayers, John Wiley and Sons, New<br />

York.<br />

[14] Channareddy, S. & Janes, N. [1999]. Direct Determination of Hydration in the<br />

Interdigitated and Ripple Phases of Dihexadecylphosphatidylcholine: Hydration<br />

of a Hydrophobic Cavity at the Membrane/Water Interface, Biophysical Journal<br />

77: 2046–2050.<br />

[15] Feigin, L. & Svergun, D. I. [1987]. Structure Analysis by Small-Angle X-ray and<br />

Neutron Scattering, Plenum Press, New York.<br />

[16] Frangopol, P. T. & Mihailescu, D. [2001]. Interactions of some local anesthetics and<br />

alcohols with membranes, Colloids and Surfaces B: Biointerfaces 22: 3–22.<br />

105


106 Bibliography<br />

[17] Gennis, R. B. [1989]. Biomembranes. Molecular Structure and Function, Springer-<br />

Verlag, New York.<br />

[18] GNU Image Manipulation Program (GIMP) [2006].<br />

URL: http://www.gimp.org/<br />

[19] Grace - a 2d graph plotting tool [2006].<br />

URL: http://plasma-gate.weizmann.ac.il/Grace/<br />

[20] Grage, S. L., Gauger, D. R., Selle, C., Pohle, W., Richter, W. & Ulrich, A. S.<br />

[2000]. The Amphiphilic Drug Flufenamic Acid can Induce a Hexagonal Phase<br />

in DMPC: a Solid State 31 P- and 19 F-NMR Study, Physical Chemistry Chemical<br />

Physics 2: 4574–4579.<br />

[21] Hallén, D., Nilsson, S.-O., Rothschild, W. & Wadsö, I. [1986]. Enthalpies and heat<br />

capacities for n-alcan-1-ols in H 2 O and D 2 O, J. Chem. Thermodynamics 18: 429–<br />

442.<br />

[22] Hemminger, W. [1994]. Calorimetric Methods, in V. B. F. Mathot (ed.), Calorimetry<br />

and Thermal Analysis of Polymers, Hanser, New York, pp. 17–47.<br />

[23] Hirai, M., Iwase, H., Hayawa, T., Koizumi, M. & Takahashi, H. [2003]. Determination<br />

of Asymmetric Structure of Gangloiside-<strong>DPPC</strong> Mixed Vesicle Using SANS,<br />

SAXS and DLS, Biphysical Journal 85: 1600–1610.<br />

[24] Holm, J. K. [2004]. Stucture and Structural Flexibility of Chloroplast Thylakoid<br />

Membranes, Ph.D. Thesis, IMFUFA, Roskilde Universitetscenter.<br />

[25] Holte, L. L. & Gawrisch, K. [1997]. Determining Ethanol Distribution in Phospholipid<br />

Multilayers with MAS-NOESY Spectra, Biochemistry 36: 4669–4674.<br />

[26] Huang, T. C., Toraya, H., Blanton, T. N. & Wu, Y. [1993]. X-ray powder diffraction<br />

analysis of silver behanate, Journal of Applied Crystalography 26: 180–184.<br />

[27] Höhne, G., Hemminger, W. & Flammersheim, H.-J. [1996]. Differential Scanning<br />

Calorimetry, Springer, Heidelberg.<br />

[28] James, J. [2002]. A Student’s Guide to Fourier Transforms, Cambridge University<br />

Press, Cambridge.<br />

[29] Janes, N., Hsu, J. W., Rubin, E. & Taraschi, T. R. [1992]. Nature of Alcohol and<br />

Anesthetic Action on Cooperative Membrane Equilibria, Biochemistry 31: 9467–<br />

9472.<br />

[30] Kamaya, H., Kaneshina, S. & Ueda, I. [1981]. Partition equilibrium of inhalation<br />

anaesthetics and alcohols between water an membranes of phospholipids of varying<br />

acyl chain lengths, Biochim. Biophys. Acta 646: 135–142.<br />

[31] Kaminoh, Y., Nishimura, S., Kamaya, H. & Ueda, I. [1992]. Alcohol interaction<br />

with high entropy states of macromolecules: critical temperature hypothesis for<br />

anesthesia cutoff, Biochimica et Biophysica Acta 1106: 335–343.<br />

[32] Katsaras, J. & Gutberlet, T. [2000]. Lipid Bilayers. Structure and Interactions,<br />

Springer, New York.<br />

[33] Katz, Y. & Diamond, J. M. [1974]. Thermodynamic Constants for Nonelectrolyte<br />

Partition between Dimyristoyl Lecithin and Water, J. Membrane Biol. 17: 101–120.<br />

[34] Kemp, R. B. [1999]. Handbook of thermal analysis and calorimetry, vol 4, Elsevier,<br />

Amsterdam.<br />

[35] Kile - a L A TEXfrontend [2006].<br />

URL: http://kile.sourceforge.net/<br />

[36] Kinoshita, K. & Yamazaki, M. [1997]. Phase Transition Between Hexagonal II<br />

(H II ) and Liquid-Crystalline Phase Induced by Interaction Between Solvents and<br />

Segments of the Membrane Surface of Dioleoylphosphatidylethanolamine, Biochimica<br />

et Biophysica Acta pp. 199–206.


[37] Komatsu, H., Guy, P. T. & Rowe, E. [1993]. Effect of unilamellar vesicle size on<br />

ethanol-induced interdigitation in dipalmitoylphosphatidylcholine, Chemistry and<br />

Physics of Lipids 65: 11–21.<br />

[38] Kranenburg, M. & Smit, B. [2005]. Phase Behavior of Model Lipid Bilayers, J. Phys.<br />

Chem. B 109: 6553–6563.<br />

[39] Kranenburg, M., Vlaar, M. & Smit, B. [2004]. Simulating Induced Interdigitation<br />

in Membranes, Biophysical Journal 87: 1596–1605.<br />

[40] Kratky, O. & Stabinger, H. [1984]. X-ray small angle camera with block-collimation<br />

system an instrument of colloid research, Colloid and Polymer Science 262(5): 345–<br />

360. Springer-Verlag GmbH.<br />

[41] Lea, D. N. R. . C. H. [1956]. Composition of Egg Phospholipids, Nature (Letters to<br />

Nature) 177: 1129 – 1130.<br />

[42] Lee, A. G. [1983]. Lipid Phase Transitions and Mixtures, in R. C. Aloia (ed.),<br />

Membrane Fluidity in Biology, Academic Press, New York, pp. 43–88.<br />

[43] Lewis, B. & Engelman, D. [1983]. Lipid Bilayer Thickness Varies Linearly with Acyl<br />

Chain Length in Fluid Phosphatidylcholine Vesicles, Journal of Molecular Biology<br />

166: 211–217.<br />

[44] Lipexextruder operating manual [2005].<br />

URL: http://www.nothernlipids.com<br />

[45] Luger, P. [1980]. Modern X-Ray Analysis on Single Crystals, Walter de Gruyter,<br />

Berlin.<br />

[46] Ly, H. V. & Longo, M. L. [2004]. The Influence of Short-Chain Alcohols on Interfacial<br />

Tension, Mechanical Properties, Areal/Molecule, and Permeability of Fluid Lipid<br />

Bilayers, Biophysical Journal 87: 1013–1033.<br />

[47] Löbbecke, L. & Cevc, G. [1995]. Effects of short-chain alcohols on the phase behavior<br />

and interdigitation of phosphatidylcholine bilayer membranes, Biochimica et<br />

Biophysica Acta 1237: 59–69.<br />

[48] Matlab [2006].<br />

URL: http://www.mathworks.com/products/matlab/<br />

[49] May, S. [2000]. Theories on structural perturbations of lipid bilayers, Current Opinion<br />

in Colloid and Interface Science 5: 244–249.<br />

[50] McIntosh, T., McDaniel, R. & Simon, S. [1983]. Induction of an interdigitated gel<br />

phase in fully hydrated phosphatidylcholine bilayers, Biochimica et Biphysica Acta<br />

73: 109–114.<br />

[51] MicroCalorimetry - ultrasensitive calorimetry for the life sciences [2006].<br />

URL: http://www.microcal.com/<br />

[52] Mou, J., Yang, J., Huang, C. & Shao, Z. [1994]. Alcohol Induces Interdigitated<br />

Domains in Unilamellar Phosphatidylcholine Bilayers, Biochemistry 33: 9981–9985.<br />

[53] Nagle, J. F. & Tristram-Nagle, S. [2000]. Structure of lipid bilayers, Biochimica et<br />

Biophysica Acta 1469: 159–195.<br />

[54] Nagle, J. & Wiener, M. C. [1988]. Structure of fully hydrated bilayer dispersions,<br />

Biochimica et Biophysica Acta 942: 1–10.<br />

[55] Nambi, P., Rowe, E. & McIntosh, T. J. [1988]. Studies of the Ethanol-Induced Interdigitated<br />

Gel Phase in Phosphatidylcholine Using the Fluorophore 1,6-Diphenyl-<br />

1,3,5-hexatriene, Biochemistry 27(1988): 9175–9182.<br />

[56] Pabst, G., Rappolt, M., Amenitsch, H. & Laggner, P. [2000]. Structural information<br />

from multilamellar liposomes at full hydration: Full q-range fitting with high quality<br />

x-ray data, Physical Review E 62: 4000–4009.<br />

107


108 Bibliography<br />

[57] Papadakis, C. M. [2004]. Structure and Dynamics of Symmetric Diblock Copolymers,<br />

Ph.D. Thesis, IMFUFA, Roskilde Universitetscenter.<br />

[58] Patty, P. J. & Frisken, B. J. [2003]. The Pressure-Dependence of the Size of Extruded<br />

Vesicles, Biophysical Journal 85: 996–1004.<br />

[59] Pedersen, U. R. [2005]. Stukturelle ændringer af en phospholipidmembran perturberet<br />

med n-alkohol undersøgt med MD-simuleringer og SAXS, Kombineret fysik og kemispeciale,<br />

Roskilde Universitetscenter.<br />

[60] Ranck, J. L., Keira, T. & Luzzaati, V. [1977]. A novel packing of the hydrocarbon<br />

chains in lipids. The low temperature phases of dipalmitoyl phosphatidyl-glycerol,<br />

Biochimica et Biophysica Acta 488: 432–441.<br />

[61] Rappolt, M., Bringezu, A. & Lohner, K. [2003]. Mechanism of the Lamellar/Inverse<br />

Hexagonal Phase Transition Examined by High Resolution X-Ray Diffraction,<br />

Biophysical Journal 84: 3111–3122.<br />

[62] Rowe, E. S. [1983]. Lipid Chain Length and Temperature Dependence of Ethanol-<br />

Phosphatidylcholine Interactions, Biochemistry 22: 3299–3305.<br />

[63] Rowe, E. S. & Campion, J. M. [1994]. Alcohol Induction of Interdigitation in Distearoylphosphatidylcholine:<br />

Flourescence Studies of Alcohol Chain Length Requirements,<br />

Biophysical Journal 67: 1888–1895.<br />

[64] Rowe, E. S. & Cutera, T. A. [1990]. Differential Scanning Calorimetric Studies of<br />

Ethanol Interactions with Distearoylphosphatidylcholine: Transition to the Interdigitated<br />

Phase, Biochemistry 29: 10398–10404.<br />

[65] Rowe, E. S., Zhang, F., Leung, T. W., Parr, J. S. & Guy, P. T. [1998]. Thermodynamics<br />

of Membrane Partioning for a Series of n-Alcohols Determined by Titration<br />

Calorimetry: Role of Hydrophobic Effects, Biochemistry 37: 2430–2440.<br />

[66] Sackmann, E. [1983]. Physical Foundation fo the Molecular Organization and Dynamics<br />

of Membranes, in W. Hoppe, W. Lohmann, H. Markl & H. Ziegler (eds),<br />

Biophysics, Springer-Verlag, New York, pp. 425–457.<br />

[67] Senin, A. A., Potekhin, S. A., Tiktopulo, E. I. & Filomonov, V. V. [2000]. Differential<br />

scanning microcalorimeter scal-1, Journal of Thermal Analysis and Calorimetry<br />

62: 153–160.<br />

[68] Simon, S. & McIntosh, T. [1984]. Interdigitated hydrocarbon chain packing causes<br />

the biphasic transition behavior in lipid/alcohol suspensions, Biochimica et Biophysica<br />

Acta 773: 169–172.<br />

[69] Simon, S., McIntosh, T. & Hines, M. [1986]. The Influence of Anesthetics on the<br />

Structure and Thermal Properties of Saturated Lecithins, in S. Miller & K. W.<br />

Roth (eds), Molecular and Cellular Mechanisms of Anesthetics, Plenum Publishing<br />

Corporation, New York, pp. 297–308.<br />

[70] Slater, J. L. & Huang, C.-H. [1988]. Interdigitated bilayer membranes, Prog. Lipid.<br />

Res. 27: 325–359.<br />

[71] Sorai, M. [2005]. Comprehensive Handbook of Calorimetry and Thermal Analysis,<br />

John Wiley, West Sussex, England.<br />

[72] Spiegel, M. R. [1996]. Mathematical Handbook, Shaum’s Outline Series, McGraw-<br />

Hill.<br />

[73] Suurkuusk, M. & Singh, S. K. [1998]. Microcalorimetic study of the interaction of<br />

1-hexanol with dimyristoylphosphatidylcholine <strong>vesicles</strong>, Chemistry and Physics of<br />

Lipids 94: 119–138.<br />

[74] Tamura, K., Kaminoh, Y., Kamaya, H. & Ueda, I. [1991]. High pressure atagonism<br />

of alcohol effects on the main phase-transition temperature of phospholipid<br />

membranes: biphasic response, Biochimica et Biophysica Acta 1066: 219–224.


109<br />

[75] Tardieu, A. & Luzzati, V. [1973]. Structure and Polymorphism of the Hydrocarbon<br />

Chains of Lipids: A Study of Lecithin-Water Phases, Journal of Molecular Biology<br />

75: 711–733.<br />

[76] Taylor, J. R. [1972]. Scattering Theory: The Quantum Theory on Nonrelativistic<br />

Collisions, John Wiley and Sons, Inc, Canada.<br />

[77] Traube, J. [1891]. Ueber die Capillaritätsconstanten organischer Stoffe in wässerigen<br />

Lösungen, Journal of Molecular Biology 265: 27–55.<br />

[78] Tristram-Nagle, S. & Nagle, J. F. [2004]. Lipid bilayers: thermodynamics, structure,<br />

fluctuations and interactions, Chemistry and Physics of Lipids 127(1): 3–14.<br />

[79] Veiro, J. A., Nambi, P., Herold, L. L. & Rowe, E. [1987]. Effect of n-alcohol and<br />

glycerol on the pretransition of dipalmitoylphosphatidylcholine, Biochimica et Biophysica<br />

Acta 900: 230–238.<br />

[80] Vierl, U., Löbbecke, L., Nagel, N. & Cevc, G. [1994]. Solute Effects on<br />

the Colloidal and Phase Behaviour of Lipid Bilayer Membranes: Ethanol-<br />

Dipalmitoylphophatidylcholine Mixtures, Biophysical Journal 67: 1067–1079.<br />

[81] Warren, B. E. [1969]. X-Ray Diffraction, Addison-Wesley, New York, USA.<br />

[82] Westh, P. & Trandum, C. [1999]. Thermodynamics of alcohol-lipid bilayer interactions:<br />

application of a binding model, Biochimica et Biophysica Acta 1421: 261–272.<br />

[83] Westh, P., Trandum, C., Jørgensen, K. & Mouritsen, O. G. [1999]. A Calorimetric<br />

Investigation of the Interaction of Short Chain Alcohols with Unilamellar DMPC<br />

Liposomes, J. Phys. Chem. B 103: 4751–4756.<br />

[84] Westh, P., Trandum, C. & Koga, Y. [2001]. Binding of small alcohols to a lipid<br />

bilayer membrane: does the partitioning coefficient express the net affinity?, Biophysical<br />

Chemistry 89: 53–63.<br />

[85] Wikipedia, International Web-based Free-content Encyclopedia Project [2006].<br />

URL: http://wikipedia.org<br />

[86] Xfig - a drawing tool [2006].<br />

URL: http://www.xfig.org/<br />

[87] Zahorowski, W., Mitternacht, J. & Wiech, G. [1991]. Application of a positionsensitive<br />

detector to soft x-ray emission spectroscopy, Meas. Sci. Technol. 2: 602–<br />

609.<br />

[88] Zhang, F. & Rowe, E. [1992]. Titration Calorimetric and Differential Scanning<br />

Calorimetric Studies of the Interactions of n-Butanol with Several Phases of Dipalmitoylphosphatcholine,<br />

Biochemistry 31: 2005–2011.<br />

[89] Zhang, R., Suter, R. M. & Nagle, J. F. [1994]. Theory of the structure factor of<br />

lipid bilayers, Physical Review E 50: 5047–5060.


110


A<br />

The course of events<br />

The work with this thesis started in early January 2006. Inspired by the work of Ulf<br />

Rørbæk Pedersen ([59]) we decided to work with long chain alcohols (5-7 carbon atoms)<br />

and their influence on phospholipids with a chainlength of 14-18. The prime experimental<br />

method should be Small-Angle X-ray Scattering (SAXS) as we had the Kratky camera<br />

at our disposal at the university. Over the following weeks we were taught how to do the<br />

sample preparation and did some pilot experiments with DMPC in order to reproduce<br />

some of the experiment in [59]. After a thorough review of a lot of the literature in<br />

this field of research and a number of discussions on possible subjects, we decided to<br />

investigate the alcohol induced interdigitated phase of <strong>DPPC</strong>. In the literature there were<br />

examples of pentanol and hexanol as inducers, whereas heptanol never had been reported<br />

as being an inducer. SAXS experiments could be used in the structural investigation<br />

of both the interdigitated phase and the liquid crystalline phase of <strong>DPPC</strong>. These two<br />

phases differ most significantly in the thickness of the bilayer, reflected in the electron<br />

density profile through the bilayer, which can be measured by SAXS. By choosing the<br />

three alcohols, pentanol, hexanol and heptanol, we had the possibility of investigating<br />

the chain-dependence of the bilayer thickness, rechecking the results obtained in the<br />

literature with pentanol and hexanol as inducers for the interdigitated phase as well as<br />

heptanol as non-inducer.<br />

The first pilot experiments with <strong>DPPC</strong> and pentanol, hexanol and heptanol, respectively,<br />

were conducted in mid-March 2006. We did a lot of pilot experiments with<br />

different lipid to alcohol ratios while reviewing the literature for references, which could<br />

serve as a basis for our experiments. Quite early in the working process we decided to<br />

use DSC in order to locate the phase transition temperature for the transition between<br />

the interdigitated phase and the liquid crystalline phase. In the beginning of April some<br />

introductory DSC scans were conducted with the help of Brian Igarashi (The ”Glass and<br />

Time” group at IMFUFA, University of Roskilde, Denmark) at the instrument belonging<br />

to the research centre ”Glass and Time”.<br />

In the beginning of May we began using the scal-1 microcalorimeter. This instrument<br />

is very sensitive and only very dilute (0.05 w%) solutions could be used. As the solutions<br />

were so dilute a correspondingly high lipid to alcohol ratio was used in order to make<br />

sure, that the lipids were interdigitated prior to performing the experiments. These<br />

experiments convinced us that the samples were stable with regards to temperature<br />

and that the phase transition temperature was unaffected by cooling and heating the<br />

sample. We ran a number of DSC scans on lipid solutions containing different amounts<br />

of alcohol. These scans showed that the phase transition temperature decreased with<br />

increasing alcohol concentration as predicted in the literature. The results from the<br />

scal-1 DSC supported our hypothesis. But the measured phase transition temperatures<br />

were found in a system 100 times more dilute than the solution used for the SAXS<br />

experiments. For this reason we turned to another, less sensitive DSC instrument, the<br />

DSC7 in mid-August.<br />

111


112 The course of events<br />

In June and July we calibrated the Kratky camera with regards to beam, q-vector and<br />

temperature. The temperature calibration (performed with a NTC-resistor) was done<br />

with both silicone oil and air in the cuvette. At first only the calibration with air was<br />

done, but as silicone oil after all has a more similar consistency to the lipid suspension<br />

the calibration was also done with this liquid.<br />

The DSC experiments performed on the DSC7 showed that the phase transition<br />

temperature of <strong>DPPC</strong> decreased when adding alcohol to the sample. The results did not<br />

differ notably from the results obtained with the DSC7. During the following months no<br />

SAXS experiments were performed without a preceding DSC scan. Besides the location of<br />

the phase transition temperature the shape of the peak also contributed to our conviction<br />

that the <strong>vesicles</strong> below the phase transition temperature were interdigitated.<br />

The underlying idea behind making use of ITC was that using lipids in the L β ′ and<br />

adding alcohol while measuring the reaction heat would reveal when the lipid membrane<br />

was saturated and hence interdigitated. By doing the experiment at a constant temperature<br />

below the phase transition temperature, the bilayers should turn from the gel<br />

phase, (L β ′), to the interdigitated phase as alcohol was added to the suspension. Different<br />

experimental procedures were tested before the solvent null was chosen. The first<br />

experiments were conducted with the lipid suspension in the reaction cell and alcohol<br />

dissolved in water was added from the syringe. This experimental method was skipped<br />

due to several reasons e.g. a too large amount of lipid was used for each experiment<br />

(more than 1 mL lipid suspension per experiment) and the large contribution to the total<br />

heat from the heat of dilution of the alcohol. In mid-December 2006 the solvent-null<br />

method was conducted with <strong>DPPC</strong> and hexanol, and subsequently the experiments with<br />

pentanol and heptanol were carried out.<br />

In the autumn of 2006 the SAXS experiments were done. During this period of<br />

time we ran into several problems with the samples and experimental equipment. For<br />

the samples especially the <strong>DPPC</strong> suspensions with heptanol were unstable in the sense<br />

that they turned gel-like after being kept at room temperature for only a few hours.<br />

Some of the samples even turned gel-like at 52 ◦ C. There were no systematism in these<br />

observations. But every time the one or more of the samples turned into a gel-like liquid<br />

we discarded the sample series and prepared a new quantity of ULV of <strong>DPPC</strong>.<br />

The SAXS spectra of all the different measurements (all three alcohols, low and high<br />

temperature) were reproduced after a number of experimental series, which mostly were<br />

discarded by computer-connection problems. When the computer lost the connection<br />

to the temperature-regulator, it froze and stopped counting the incoming photons, the<br />

sample was still exposed to radiation though. And as we assigned importance to treating<br />

the samples the same, we chose not to use spectra from samples, which had been exposed<br />

to radiation longer than the necessary 2 times 18 hours.<br />

Before reproducing the spectra we only used the models from [59] to fit to our data.<br />

We did not want to spend a lot of time working and developing the modelling procedure<br />

before we were sure the spectra could be reproduced. In the process of obtaining the<br />

SAXS spectra we repeatedly compared the spectra obtained in the low and high temperature<br />

region to see if they differed. We expected them to differ as we assumed the spectra<br />

to reflect two different lipid phases. As the spectra obtained in the low temperature and<br />

high temperature region indeed differed, we continued the measurements.<br />

After reproducing the spectra we developed and expanded the models (primarily<br />

adapted from [59]) to cover the interdigitated phase as well as the liquid fluid phase.<br />

These models did fit the main features of the raw data spectra but none of the models<br />

took the small, distinct peaks into account. At first the peaks were considered as noise


on the spectra. Not until all the SAXS spectra had been reproduced it became clear that<br />

it was in fact not noise but distinct reproducible peaks, which we subsequently tried to<br />

fit to a hexagonal lattice structure.<br />

113


114


B<br />

Theory<br />

The nonrelativistic quantum mechanical scattering experiment<br />

The scattering experiment has played an important role in the development of quantum<br />

theory and is perhaps the most important technique. Rutherford discovered the nucleus<br />

with α-particles (α+ 14 N → p+ 17 O) and Franck-Hertz discovered the existence of atomic<br />

energy levels by observation of electrons scattering off mercury vapor.<br />

The timedependent Schrödinger equation i d |Ψ〉 = H|Ψ〉 has the general solution |Ψ〉 =<br />

dt<br />

U(t) |ψ〉 = exp(−iHt) |ψ〉 where U(t) is the evolution operator and H is the Hamiltonian,<br />

which consists of the Hamiltonian for the free particle, H 0 = p 2 /2m, and timeindependent<br />

potential, V (x). U 0 (t) ≡ exp(−iH 0 t)<br />

A long time before and a long time after 1 the collision the particles can be considered<br />

as free particles. Accordingly, the actual wave function will tend to the free wave function<br />

asymptotes.<br />

U(t)|ψ〉 −−−−→<br />

t→−∞ U0 (t)|ψ in 〉 (B.1)<br />

U(t)|ψ〉 −−−−→<br />

t→+∞ U0 (t)|ψ out 〉 (B.2)<br />

Figure B.1 A figure showing a classical interpretation of the Møller operators (figure from<br />

Taylor [76, p. 30]).<br />

The actual wave function |ψ〉 is connected to the scattering asymptotes (|ψ in 〉 and<br />

|ψ out 〉) by the isometric Møller operators (Ω + and Ω − )<br />

|ψ〉 = lim<br />

t→−∞ U(t)† U 0 (t)|ψ in 〉 ≡ Ω + |ψ in 〉<br />

|ψ〉 = lim<br />

t→+∞ U(t)† U 0 (t)|ψ out 〉 ≡ Ω − |ψ out 〉<br />

(B.3)<br />

(B.4)<br />

1 On the atomic scale, for the actual scattering process 10 −10 seconds are considered very slow.<br />

115


116 Theory<br />

If we define the scattering operator, S ≡ Ω † − Ω +, this gives the relation between the<br />

asymptotes, since |ψ〉 = Ω + |ψ in 〉 = Ω − |ψ out 〉<br />

|ψ out 〉 = Ω † − |ψ〉 = Ω† − Ω +|ψ in 〉 = S|ψ in 〉 (B.5)<br />

All scattering propabilities can be calculated from S, e.g. the propability of scattering<br />

from the initial (combined) state |ψ in 〉 = |i〉 to the final (combined) state |ψ out 〉 = |f〉 is<br />

proportional to |〈f|S|i〉| 2 and the differential cross section is, Als-Nielsen & McMorrow<br />

[3, p. 265]<br />

( ) 2 ∫<br />

dσ V<br />

dΩ = 1<br />

2π 4 c 4 |〈f|S|i〉| 2 E f δ(E f − E i )dE f<br />

(B.6)<br />

where E f and E i are the energy of final and initial state, respectively. Thus S contains<br />

all the information of experimental interest - unfortunately, S is not easily calculated.


C<br />

Fourier transform<br />

To model the electron density profile of the lipid bilayer from the measured intensity it<br />

is necessary to Fourier transform the Gaussians representing the three different electron<br />

density profiles to find the form factors. Basicly the Fourier transform of a given function<br />

f(x) is defined by [28]<br />

F(q) =<br />

∫ ∞<br />

and the inverse Fourier transform as<br />

−∞<br />

f(x)exp(iqx)dx,<br />

(C.1)<br />

f(x) = 1 ∫ ∞<br />

F(q)exp(−iqx)dq.<br />

2π −∞<br />

(C.2)<br />

The Fourier transform is linear, since if f(x) and g(x) have Fourier transforms F(x)<br />

and G(x) respectively then<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

[c 1 f(x) + c 2 g(x)] exp(iqx)dx = c 1 f(x)exp(iqx)dx<br />

−∞<br />

∫ ∞<br />

+c 2 g(x)exp(iqx)dx<br />

−∞<br />

= c 1 F(q) + c 2 G(q). (C.3)<br />

This is useful as the three electron densities have different form factors which must<br />

be added.<br />

The Fourier transform of a shifted function is<br />

F shift (q) =<br />

Substituting u = x − x 0 gives<br />

∫ ∞<br />

−∞<br />

f(x − x 0 )exp(iqx)dx<br />

(C.4)<br />

F shift (q) =<br />

∫ ∞<br />

−∞<br />

= exp(iqx 0 )<br />

f(u)exp(iq(u + x 0 ))du<br />

∫ ∞<br />

−∞<br />

f(u)exp(iqu)du<br />

(C.5)<br />

(C.6)<br />

= exp(iqx 0 )F(q) (C.7)<br />

i.e. the factor exp(iqx 0 ) is multiplied onto the Fourier transform of the original<br />

function f(x). This result is commonly known as the Shift theorem.<br />

The evaluation of equation C.1 is simplified as<br />

117


118 Fourier transform<br />

F(q) =<br />

=<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

f(x)cos(qx)dx + i f(x)sin(qx)<br />

−∞<br />

f(x)cos(qx)dx<br />

(C.8)<br />

(C.9)<br />

as the integral of an antisymmetic function on a symmetric interval is zero, and the<br />

product of the symmetric Gaussian and the antisymmetric sine function is antisymmetric.<br />

The Gaussian<br />

g(x) = Aexp(− x2<br />

σ 2 )<br />

(C.10)<br />

can be Fourier transformed analytically by use of C.1 and C.8,<br />

G(q) =<br />

∫ ∞<br />

0<br />

= 2A<br />

= 2A<br />

= 2A<br />

)<br />

Aexp<br />

(− x2<br />

σ 2 cos(xq)dx<br />

∫ ∞<br />

0<br />

∫ ∞<br />

0<br />

∫ ∞<br />

0<br />

= 2Aexp<br />

Re<br />

(exp(− x2<br />

Re<br />

(exp(− x2<br />

(<br />

Re<br />

(<br />

− qσ 2<br />

(<br />

= 2Aexp − qσ 2<br />

(<br />

= 2Aexp − qσ ) ( 2<br />

Re σ<br />

2<br />

σ 2 )exp(iqx) )<br />

dx<br />

σ 2 + iqx) )<br />

dx<br />

) 2<br />

)<br />

exp<br />

(− x2<br />

σ 2 − iqσ − ( qσ 2 e )2 dx<br />

) 2<br />

∫ ∞<br />

)<br />

Re<br />

(exp(− x2<br />

0 σ 2 − iqσ 2 )2 dx<br />

) 2<br />

(∫ ∞<br />

)<br />

Re σ exp(−z 2 ) dz<br />

0<br />

√ ) π<br />

2<br />

(C.11)<br />

(C.12)<br />

(C.13)<br />

(C.14)<br />

(C.15)<br />

(C.16)<br />

(C.17)<br />

= σA √ ( qσ<br />

π exp −<br />

2<br />

) 2<br />

, (C.18)<br />

which is also a Gaussian but with amplitude σA √ π and a width parameter of 2σ −1 .<br />

The substitution<br />

z = x σ − iqσ 2 ,<br />

(C.19)<br />

was made, yielding dx = σdz and the square<br />

was completed.<br />

(<br />

− x2 x<br />

) 2 ( qσ<br />

) 2<br />

σ 2 + iqx = − σ − iqσ − (C.20)<br />

2 2


119<br />

Electron density to form factor<br />

The electron density is defined as a sum of Gaussians<br />

ρ(z) = ∑ A n<br />

√ exp<br />

(− (z − z n) 2 )<br />

n 2πσ<br />

2 n<br />

2σn<br />

2<br />

(C.21)<br />

The form factor is the Fourier transform of the electron density, using the shift theorem<br />

equation C.5,<br />

and the fact that, [72],<br />

F{ρ(z)} = F{ρ(z − z n )} = F{ρ(z)} exp(iqz n )<br />

F { exp(−bx 2 ) } = 1 2<br />

which in this case corresponds to<br />

√ ( ) π<br />

b exp − α2<br />

4b<br />

(C.22)<br />

(C.23)<br />

F<br />

{ }<br />

A<br />

√<br />

2πσ<br />

2 exp(−(2σ2 ) −1 z 2 )<br />

= 1 √ (<br />

A π<br />

√<br />

2 2πσ<br />

2 (2σ 2 ) −1 exp q 2 )<br />

−<br />

4 · (2σ 2 ) −1<br />

= A 2 exp(−1 2 q2 σ 2 )<br />

Yielding, with equation C.3, the form factor<br />

F{ρ(z)} = ∑ (<br />

A n<br />

2 exp − 1 )<br />

2 q2 σn<br />

2 exp(iqz n ) = F(q)<br />

n<br />

(C.24)<br />

(C.25)


120


D<br />

Experimental series<br />

When a new series of experiments was planned, a new quantity of lipid solution was<br />

prepared. Whenever a new quantity was made, it was numbered consecutively. The last<br />

three series were only made to reproduce results obtained with previous series.<br />

Series 7 <strong>DPPC</strong> with hexanol. It was intended to check whether there was a difference<br />

in measuring from 25 ◦ C to 45 ◦ C, the so-called up-scan, and from 45 ◦ C to 25 ◦ C,<br />

the so-called down-scan. After preparing the samples, different subsamples were<br />

made containing different amounts of hexanol, equivalent to a ratio of 1:2, 1:4,<br />

1:6 and 1:9, and they were kept in the Eppendorph Thermomixer at 30 ◦ C. The<br />

DSC scans showed that the phase transition temperature decreased with increasing<br />

hexanol concentration. The SAXS experiments were performed using the 1:2<br />

ratio <strong>DPPC</strong>:hexanol sample. Using a MTC-script made sure that the spectra were<br />

measured at 25 ◦ , 30 ◦ , 40 ◦ and 45 ◦ , respectively. The selected temperatures were<br />

chosen given the DSC results. Concentration: 1.62±0.061 w%.<br />

Series 8 <strong>DPPC</strong> with pentanol. After preparing the samples, different subsamples were<br />

made containing different amounts of pentanol, equivalent to a ratio of 1:2, 1:4,<br />

and they were kept in the Eppendorph Thermomixer at 30 ◦ C. The DSC scans<br />

showed that the phase transition temperature decreased with increasing hexanol<br />

concentration. The SAXS experiments were performed using the 1:2 lipid:pentanol<br />

sample. An MTC-script was made, making sure the spectra was measured at 25 ◦ ,<br />

30 ◦ , 40 ◦ and 45 ◦ , respectively. These temperatures were chosen given the DSC<br />

results. Concentration: 3.07±0.026 w%.<br />

Series 9 <strong>DPPC</strong> with pentanol and hexanol. After preparing the samples, different subsamples<br />

were made containing different amounts of pentanol and hexanol, equivalent<br />

to a ratio of 1:2, and they were kept in the Eppendorph Thermomixer at 52 ◦ C.<br />

The DSC scans showed that the phase transition temperature decreased drastically<br />

in the samples containing hexanol. The pentanol samples also decreased the<br />

phase transition temperature. The SAXS experiments were performed on the 1:2<br />

<strong>DPPC</strong>:pentanol sample first. An MTC-script was made, making sure the spectra<br />

was measured at 25 ◦ , 30 ◦ , 40 ◦ and 45 ◦ , respectively. These temperatures were<br />

chosen given the DSC results. The SAXS experiments were performed on the 1:2<br />

<strong>DPPC</strong>:hexanol was measured at 20 ◦ , 25 ◦ , 35 ◦ and 40 ◦ , respectively. These temperatures<br />

were chosen given the DSC results. Concentration: 4.32±0.063 w%.<br />

Series 10 <strong>DPPC</strong> with heptanol. After preparing the samples heptanol was added, equivalent<br />

to a ratio of 1:2, and they were kept in the Eppendorph Thermomixer at<br />

52 ◦ C. The DSC scans showed that the phase transition temperature was around<br />

35 ◦ C. The SAXS experiments were performed at 25 ◦ C, 30 ◦ C, 40 ◦ C and 45 ◦ C.<br />

The heptanol sample was gellike after the last SAXS experiment. Concentration:<br />

3.56±0.021 w%.<br />

Series 11 Pure <strong>DPPC</strong>. After preparing the multilamellar <strong>vesicles</strong> several DSC scans were<br />

performed, and a phase transition temperature of 41 ◦ C was found. SAXS exper-<br />

121


122 Experimental series<br />

iments were performed, identifying the repeating structure of the lipid bilayers<br />

inside the MLV. Concentration: ∼5.12 w%<br />

Series 12 <strong>DPPC</strong> with pentanol, hexanol and heptanol. After preparing the samples<br />

pentanol, hexanol or heptanol was added, equivalent to a ratio of 1:2 (1:3 for<br />

heptanol), and the samples were kept in the Eppendorph Thermomixer until used.<br />

The SAXS experiments were only performed on the pentanol and hexanol samples<br />

as the heptanol sample turned gellike and was partly evaporated, probably due to<br />

a too high alcohol concentration. The pentanol SAXS experiments were performed<br />

at 25 ◦ C, 30 ◦ C, 40 ◦ C and 45 ◦ C. The hexanol SAXS experiments were performed at<br />

25 ◦ C and 35 ◦ C. Concentration: 2.1±0.187 w%.<br />

Series 13 <strong>DPPC</strong> with pentanol, hexanol and heptanol. After preparing the samples pentanol,<br />

hexanol or heptanol was added, equivalent to a ratio of 1:2, and the samples<br />

were kept in the Eppendorph Thermomixer until used. The SAXS experiments<br />

were only performed on the pentanol and hexanol samples as the heptanol sample<br />

turned gellike and was partly evaporated. The pentanol SAXS experiments were<br />

performed at 25 ◦ C and 45 ◦ C. The hexanol SAXS experiments were performed at<br />

25 ◦ C and 35 ◦ C. Concentration: 4.59±0.046 w%.<br />

Series 13b Concentration: 2.22±0.040 w%. Used for ITC.<br />

Series 14 <strong>DPPC</strong> with pentanol, hexanol and heptanol. After preparing the samples<br />

pentanol, hexanol or heptanol was added, equivalent to a ratio of 1:2, and the<br />

samples were kept in the Eppendorph Thermomixer until used. The pentanol SAXS<br />

experiments were performed at 25 ◦ C and 45 ◦ C. The hexanol SAXS experiments<br />

were performed at 25 ◦ C and 45 ◦ C. The heptanol SAXS experiments were performed<br />

at 25 ◦ C and 45 ◦ C. Concentration: 4.29±0.039 w%.<br />

Series 14b Concentration: 4.77±0.034 w%. Used for ITC.<br />

Series 15 <strong>DPPC</strong> with pentanol, hexanol and heptanol. After preparing the samples<br />

pentanol, hexanol or heptanol was added, equivalent to a ratio of 1:2, and the<br />

samples were kept in the Eppendorph Thermomixer until used. The pentanol SAXS<br />

experiments were performed at 25 ◦ C and 45 ◦ C. The hexanol SAXS experiments<br />

were performed at 25 ◦ C and 45 ◦ C. The heptanol SAXS experiments were performed<br />

at 25 ◦ C and 45 ◦ C. Concentration: 3.90±0.024 w%.<br />

Series 15b Concentration: 4.52±0.054 w%. Used for ITC introductory experiments.<br />

Series 16 Concentration: 4.43±0.017 w%. Used for ITC introductory experiments.<br />

Series 17b Concentration: 3.93±0.038 w%. Kept at 25 ◦ C until alcohol was added. To<br />

this lipid suspension pentanol and hexanol, respectively was added. The lipidalcohol<br />

suspension was kept shaken at 52 ◦ C until used for ITC experiments.<br />

Series 18b Concentration: 3.99±0.019 w%. Kept at 25 ◦ C until alcohol was added. To<br />

this lipid suspension pentanol and hexanol, respectively was added. The lipidalcohol<br />

suspension was kept shaken at 52 ◦ C until used for ITC experiments.<br />

Series 19b Concentration: 3.34±0.060 w%. Kept at 25 ◦ C until alcohol was added. To<br />

this lipid suspension hexanol and heptanol, respectively was added. The lipidalcohol<br />

suspension was kept shaken at 52 ◦ C until used for ITC experiments.<br />

Series 20 Pure <strong>DPPC</strong>. Made for SAXS experiments on ULV <strong>vesicles</strong>. Concentration:<br />

∼4.8 w%.


E<br />

ITC<br />

The partition coefficient is defined as:<br />

K x = X a,b<br />

X a,w<br />

(E.1)<br />

where<br />

n alc,lipid<br />

X a,b =<br />

n alc,lipid + n <strong>DPPC</strong><br />

n alc,free<br />

X a,w =<br />

n alc,free + n water<br />

The used n-alcohols, lipid and the free alcohol concentration calculated from the linear<br />

n-Alcohol Lipid Free alcohol conc. K x<br />

1 Pentanol (28.7 mM) <strong>DPPC</strong> (53.5 mM) 22.5 mM 261<br />

2 Pentanol (66.6 mM) <strong>DPPC</strong> (45.5 mM) 53.9 mM 225<br />

3 Hexanol (10.5 mM) <strong>DPPC</strong> (45.5 mM) 7.9 mM 367<br />

4 Hexanol (15.1 mM) <strong>DPPC</strong> (54.3 mM) 3.7 mM* 2553*<br />

5 Hexanol (21.6 mM) <strong>DPPC</strong> (54.3 mM) 13.5 mM 529<br />

6 Hexanol (22.6 mM) <strong>DPPC</strong> (54.3 mM) 11.8 mM /3.9 mM* 749/ 3546*<br />

7 Hexanol (31.3 mM) <strong>DPPC</strong> (53.5 mM) 13.8 mM 986<br />

8 Heptanol (7.3 mM) <strong>DPPC</strong> (45.5 mM) 2.5 mM/ 2.9 mM* 2166/1713*<br />

Table E.1 The first column represents the used alcohol and the related concentration in the<br />

syringe. The second column represents the lipid concentration, the third column the calculated<br />

free alcohol concentration and the fourth column the calculated partition coefficient. The<br />

figures marked by stars are explained in section 4.1.<br />

correlation in figures E.1 and E.2 and the partition coefficients are shown i table E.1. All<br />

the experimental series have been conducted using the MSC-ITC instrument, but hexanol<br />

no. 3, which has been conducted on the VP-ITC instrument. For every plot a measure<br />

for the standard deviation is made by using the definition for standard deviation:<br />

S = √ 1 ∑i=n<br />

(y − ỹ)<br />

n − 1<br />

2 (E.2)<br />

where y is the measured heat, ỹ is the fitted value for the heat and n is the number of<br />

measurements.<br />

i=1<br />

123


124 ITC<br />

80<br />

28.7 mM pentanol in 54.3 mM lipid (test 4)<br />

250<br />

66.6 mM pentanol in 45.5 mM lipid<br />

60<br />

200<br />

cal/mole of injectant<br />

40<br />

20<br />

0<br />

−20<br />

cal/mole of injectant<br />

150<br />

100<br />

50<br />

0<br />

−50<br />

−40<br />

−100<br />

−60<br />

18 20 22 24 26 28 30<br />

Alcohol concentration in reaction cell/ mM<br />

−150<br />

50 52 54 56 58 60 62<br />

Alcohol concentration in reaction cell/ mM<br />

(a) The free alcohol concentration is found<br />

to be 22.5 mM. y=9.1·x-203.24, S=9.65.<br />

(b) The free alcohol concentration is found<br />

to be 53.9 mM. y=30.6·1.6·10 3 , S=13.9.<br />

2000<br />

31.3 mM hexanol in 53.5 mM lipid<br />

150<br />

10.5 mM hexanol in 45.5 mM lipid<br />

1500<br />

100<br />

cal/mole of injectant<br />

1000<br />

500<br />

0<br />

−500<br />

cal/mole of injectant<br />

50<br />

0<br />

−50<br />

−100<br />

−1000<br />

−150<br />

−1500<br />

10 12 14 16 18 20<br />

Alcohol concentration in reaction cell/ mM<br />

−200<br />

5 6 7 8 9 10<br />

Alcohol concentration in reaction cell/ mM<br />

(c) The free alcohol concentration is found to<br />

be 13.8 mM. y=316.2·x-4.4·10 3 , S=31.3.<br />

(d) The free alcohol concentration is found<br />

to be 7.9 mM. y=57.5 ·x-456.6, S=4.7.<br />

Figure E.1 Results from two pentanol experimental series and two hexanol experimental series.<br />

The graphs represent the integrated heat contributions as functions of the concentration of<br />

alcohol in the reaction cell. The free alcohol concentration is equivalent to the intersection of<br />

the graph with the axis of abscisses. Each point is equivalent to an experimental series.


125<br />

1000<br />

15.1 mM hexanol in 54.3 mM lipid<br />

300<br />

21.6 mM hexanol in 54.3 mM lipid (test 7)<br />

800<br />

200<br />

600<br />

cal/mole of injectant<br />

400<br />

200<br />

0<br />

−200<br />

cal/mole of injectant<br />

100<br />

0<br />

−100<br />

−400<br />

−200<br />

−600<br />

0 2 4 6 8 10<br />

Alcohol concentration in reaction cell/ mM<br />

−300<br />

10 11 12 13 14 15 16 17<br />

Alcohol concentration in reaction cell/ mM<br />

(a) The free alcohol concentration is found<br />

to be 3.7 mM. y=174.2·x-544.9, S=84.4.<br />

(b) The free alcohol concentration is found<br />

to be 13.5 mM. y=75.9·x-1.0·10 3 , S=24.3.<br />

2000<br />

22.6 mM hexanol in 53.5 mM lipid<br />

400<br />

7.3 mM heptanol in 45.5 mM lipid (test 1)<br />

1500<br />

300<br />

cal/mole of injectant<br />

1000<br />

500<br />

0<br />

−500<br />

cal/mole of injectant<br />

200<br />

100<br />

0<br />

−100<br />

−200<br />

−1000<br />

−300<br />

−1500<br />

0 5 10 15 20<br />

Alcohol concentration in reaction cell/ mM<br />

−400<br />

1 1.5 2 2.5 3 3.5 4<br />

Alcohol concentration in reaction cell/ mM<br />

(c) The free alc. conc: 11.6 mM(*) and<br />

3.9 mM (o). y(*)=123.7·x-1.5·10 3 , S=257.1.<br />

y(o)=262.5·x-1.0·10 3 , S=68.4.<br />

(d) The free alc. conc: 2.5 mM(*) and<br />

2.9 mM(o). y(*)=230.8 ·x-566.1, S=26.7.<br />

y(o)=214.8·x-614.5, S=19.1.<br />

Figure E.2 Results from three hexanol experiment series and one heptanol series. The graphs<br />

represent the integrated heat contributions as functions of the concentration of alcohol in the<br />

reaction cell. The free alcohol concentration is equivalent to the intersection of the graph with<br />

the axis of abscisses. Each point is equivalent to an experimental series.


126 ITC<br />

Thermodynamical reflections on the ITC results<br />

Through the use of the partition coefficient several thermodynamical parameters can be<br />

calculated, since the free Gibbs energy is defined as<br />

− ∆G 0 transfer = R · T · ln(K x)<br />

∆G 0 transfer = ∆H 0 transfer − T∆S 0 transfer<br />

(E.3)<br />

(E.4)<br />

Here R is the universal gas constant, T is the temperature in Kelvin, ∆H 0 transfer is<br />

calculated for the transfer of alcohol into the vesicle. The ∆Htransfer 0 was calculated from<br />

the slope of the solvent-null curve, see figure 4.1, and the amount of alcohol transferred<br />

to the lipid-vesicle was obtained from equation 4.2 1 . ∆S 0 transfer is the entropy for the<br />

transfer. The thermodynamic parameters are listed in table E.2.<br />

n-Alcohol K x ∆G 0 transfer ∆H 0 transfer ∆S 0 transfer<br />

(kJ/mol alc.) (kJ/mol alc.) (J/(mol · K) alc.)<br />

1 Pentanol 261 -13.8 0.32 47.5<br />

2 Pentanol 225 -13.4 0.46 46.6<br />

3 Hexanol 367 -14.7 4.3 63.7<br />

4 Hexanol 2553* -19.5* 3.0* 75.2*<br />

5 Hexanol 524 -15.5 2.1 59.3<br />

6 Hexanol 749/ 3546* -16.4/-20.27* 2.7/3.25* 64.2/78.9*<br />

7 Hexanol 986 -17.1 3.44 68.9<br />

8 Heptanol 2166/1713* -19.0/-18.5* 9.1/9.28* 94.47/93.0*<br />

Table E.2 Partition coefficients and thermodynamic parameters calculated from solvent-null<br />

experiments.<br />

As an alcohol is mixed with a lipid membrane the configurational entropy increases<br />

resulting in a stabilization with regard to an energetic standpoint. This is due to an additional<br />

entropic term, the configurational entropy, which is competing with the thermal<br />

entropy in changing the free energy of the lipid membrane [29]. Neither the configurational<br />

nor the thermal entropy can be measured directly. The former is deduced from<br />

the partitioning, the latter from the enthalpy change of the equilibrium. A membrane<br />

equilibrium sensitive to partitioning of alcohols exhibits a large change in the configurational<br />

entropy and a smaller change in the enthalpy, which is exactly the case in this<br />

work.<br />

As the alcohol partitions differently into the the L βI , the P β ′ and the L α phase, the<br />

free energy, ∆G, will vary accordingly [29]. This is also seen in the results from this work,<br />

table E.2. Here the free energy for transferring alcohol into the interdigitated phase is<br />

higher (less negative) than the free energy for transferring alcohol into the ripple phase.<br />

Another contribution to the increase in entropy is entropy gain of the water from<br />

the hydration layer and of the acyl chains, as well as the head group disordering due<br />

to the partitioning of the alcohol molecules [73]. This gain in entropy will continue<br />

until the <strong>vesicles</strong> are saturated. This can explain the larger entropy associated with the<br />

1 ∆H 0 transfer is calculated as ∆H= α·4.186·C lipid·V sample<br />

n alc,lipid<br />

. Here α is the slope from the solvent-null curve,<br />

and 4.186 is the universal gas constant.


hexanol and pentanol samples, which were kept at room temperature for 24 hours prior<br />

to the experiments. The <strong>vesicles</strong> in these samples were not saturated when the main<br />

experiments were conducted.<br />

Classical hydrophobic interactions are also characterised by positive entropic driven<br />

forces. And this indicates the importance of the water structure around the non-polar<br />

moieties involved in the partitioning process. As the acyl chains of the alcohol are moved<br />

from the aqueous to the lipid environment hydration water is released and this causes a<br />

growth in entropy. In addition the acyl chains of the alcohol most likely have a higher<br />

degree of mobility when in the membrane than in water [73].<br />

127


128


F<br />

DSC<br />

Several scans using the scal-1 microcalorimeter with different scanrates resulted in figure<br />

F.1.<br />

Figure F.1 The baseline of the scal-1 microcalorimeter obtained with different scanrates.<br />

129


130 DSC<br />

18000<br />

<strong>DPPC</strong> : pentanol<br />

16000<br />

scan 1<br />

scan 11<br />

14000<br />

mV<br />

12000<br />

10000<br />

8000<br />

6000<br />

10000<br />

36 38 40 42 44 46 48 50<br />

Temperature [°C]<br />

<strong>DPPC</strong> : heptanol<br />

8000<br />

scan 1<br />

scan 9<br />

6000<br />

mV<br />

4000<br />

2000<br />

0<br />

36 38 40 42 44<br />

Temperature [°C]<br />

Figure F.2 Both figures show 12 successive scans plotted on top of each other. There is<br />

practically no difference between the 12 scans and the location of the peak does not differ. A<br />

lipid to pentanol/heptanol ratio of 1:2, 0.05 weight % lipid and a scanrate of 1 K/min was used.<br />

There is an average standard deviation of 1082.4 on the pentanol/<strong>DPPC</strong> scan, and an average<br />

standard deviation of 382.5 on the heptanol/<strong>DPPC</strong> scan, both relative to the first scan. These<br />

scans have been obtained by using the scal-1 microcalorimeter.


131<br />

14<br />

Enthalpy of the phase transition<br />

heat flow [mW]<br />

13.5<br />

13<br />

12.5<br />

12<br />

11.5<br />

11<br />

10.5<br />

10<br />

9.5<br />

100 200 300 400 500 600 700 800<br />

time [sec]<br />

Figure F.3 The phase transition enthalpy of pentanol/<strong>DPPC</strong>, series 09. Lipid to alcohol ratio<br />

1:2. The blue line is the raw DSC data, the red line is the raw DSC data added the uncertainty<br />

calculated and represented in figure 4.2. For this particular scan the calculated enthalpy is<br />

64.5 ± 11 kJ/mol. The vertical lines represent the interval of the phase transition.


132


G<br />

SAXS<br />

Two different experimental series with <strong>DPPC</strong> and hexanol were conducted to show that<br />

there were no difference between starting the measurements in the low temperature area<br />

and the high temperature area.<br />

<strong>DPPC</strong> : hexanol<br />

Serie 7<br />

Intensity [a.u.]<br />

8<br />

6<br />

4<br />

25 °C (downscan)<br />

25 °C (upscan)<br />

30 °C (downscan)<br />

30 °C (upscan)<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

q [Å -1 ]<br />

Figure G.1 Low temperature.<br />

133


134 SAXS<br />

8<br />

<strong>DPPC</strong> : hexanol<br />

Serie 7<br />

6<br />

40 °C (downscan)<br />

40 °C (upscan)<br />

45 °C (downscan)<br />

45 °C (upscan)<br />

Intensity [a.u.]<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

q [Å -1 ]<br />

Figure G.2 High temperature.


H<br />

Modelling results<br />

The electron density profile graphs have been scaled such that every graph has the same<br />

amplitude for the inner head group Gauss function.<br />

135


136 Modelling results<br />

H.1 Pure <strong>DPPC</strong><br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=20.2<br />

Asym model, sse=20.4<br />

Sym-mlv model, sse=10.2<br />

3d model, sse=20.3<br />

3d-mlv model, sse=10.3<br />

Sym-4g model, sse=20.1<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.1 Low temperature (25 ◦ C), Series 20


H.1 Pure <strong>DPPC</strong> 137<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=20.2<br />

Asym model, sse=20.4<br />

Sym-mlv model, sse=10.2<br />

3d model, sse=20.3<br />

3d-mlv model, sse=10.3<br />

Sym-4g model, sse=20.1<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.2 High temperature (45 ◦ C), Series 20


138 Modelling results<br />

series 20 series 20<br />

sym<br />

sse 20.2 13<br />

A 0.88 ± 0.07 0.5 ± 1<br />

z h / Å 20.1 ± 0.9 10 ± 60<br />

σ h / Å 5.6 ± 0.5 7 ± 20<br />

σ t / Å 6 ± 2 9 ± 200<br />

asym<br />

sse 20.4 25.9<br />

A1 0.6 ± 4 0.5 ± 30<br />

A2 0.6 ± 4 0.5 ± 30<br />

z h / Å 17.5 ± 0.6 20 ± 40<br />

σ h / Å 7.4 ± 0.7 7 ± 10<br />

σ t / Å 14 ± 1 9 ± 100<br />

sym-mlv<br />

sse 10.2 4.78<br />

A 0.6 ± 0.01 0.589 ± 0.004<br />

z h / Å 17.9 ± 0.4 18.8 ± 0.2<br />

σ h / Å 7.7 ± 0.4 3.8 ± 0.2<br />

σ t / Å 17.6 ± 0.2 5.1 ± 0.3<br />

A b 0.019 ± 0.001 0.0156 ± 0.0005<br />

q 0 / Å −1 0.0928 ± 0.0006 0.0936 ± 0.0004<br />

3d<br />

sse 20.3 13<br />

A 0.9 ± 0.1 0.7 ± 10<br />

z h / Å 20.1 ± 0.9 10 ± 60<br />

σ h / Å 5.6 ± 0.5 7 ± 20<br />

σ t / Å 6 ± 2 9 ± 200<br />

3d-mlv<br />

sse 10.3 4.76<br />

A 1.14 ± 0.08 0.79 ± 0.03<br />

z h / Å 17.9 ± 0.6 18.8 ± 0.2<br />

σ h / Å 8.4 ± 0.6 3.8 ± 0.2<br />

σ t / Å 21.7 ± 0.5 5.1 ± 0.3<br />

A b 0.035 ± 0.001 0.0156 ± 0.0005<br />

q 0 / Å −1 0.0921 ± 0.0004 0.0936 ± 0.0004<br />

sym-4g<br />

sse 20.1 12.7<br />

A 1.8 ± 0.2 1.03 ± 0.01<br />

z h / Å 20 ± 1 0.2 ± 5000<br />

σ h / Å 5.5 ± 0.7 10 ± 70<br />

z t / Å 4 ± 3 5 ± 0.1<br />

σ t / Å 3 ± 7 6.1 ± 0.6


H.2 Pentanol 139<br />

H.2 Pentanol<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=13.0<br />

Asym model, sse=12.5<br />

Sym-mlv model, sse=12.4<br />

3d model, sse=12.8<br />

3d-mlv model, sse=12.8<br />

Sym-4g model, sse=12.8<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.3 Low temperature (25 ◦ C), Series 14


140 Modelling results<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=35.1<br />

Asym model, sse=34.5<br />

Sym-mlv model, sse=35.2<br />

3d model, sse=48.8<br />

3d-mlv model, sse=35.1<br />

Sym-4g model, sse=35.1<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.4 Low temperature (25 ◦ C), Series 15. Please notice the sym is hidden behind sym-4g.


H.2 Pentanol 141<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=15.0<br />

Asym model, sse=15.1<br />

Sym-mlv model, sse=10.4<br />

3d model, sse=15.0<br />

3d-mlv model, sse=10.4<br />

Sym-4g model, sse=16.7<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.5 High temperature (45 ◦ C), Series 13.


142 Modelling results<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=19.5<br />

Asym model, sse=23.8<br />

Sym-mlv model, sse=15.3<br />

3d model, sse=19.5<br />

3d-mlv model, sse=15.3<br />

Sym-4g model, sse=19.4<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.6 High temperature (45 ◦ C), Series 15.


H.2 Pentanol 143<br />

low temp.<br />

high temp.<br />

series 14 series 15 series 13 series 15<br />

sym<br />

sse 13 35.1 15 19.5<br />

A 0.79 ± 0.07 0.55 ± 0.04 0.5 ± 0.002 0.5 ± 0.1<br />

z h / Å 18 ± 1 13 ± 2 15 ± 6 15 ± 10<br />

σ h / Å 6.9 ± 0.6 9 ± 1 6 ± 2 6 ± 4<br />

σ t / Å 6 ± 2 13.7 ± 0.3 8 ± 20 9 ± 30<br />

asym<br />

sse 12.5 34.5 15.1 23.8<br />

A1 0.74 ± 0.04 0.51 ± 0.06 0.5 ± 100 0.5 ± 30<br />

A2 0.9 ± 0.06 0.5 ± 0.2 0.5 ± 100 0.5 ± 30<br />

z h / Å 18.6 ± 0.9 10 ± 20 15 ± 9 17 ± 2<br />

σ h / Å 6.4 ± 0.5 10 ± 6 6 ± 4 4.5 ± 0.8<br />

σ t / Å 5 ± 1 14 ± 4 8 ± 30 6 ± 3<br />

sym-mlv<br />

sse 12.4 35.2 10.4 15.3<br />

A 0.5 ± 1 0.8 ± 0.3 0.504 ± 0.003 0.532 ± 0.01<br />

z h / Å 8 ± 300 18 ± 5 17.3 ± 0.4 15.6 ± 0.3<br />

σ h / Å 10 ± 60 6 ± 2 3.7 ± 0.3 5.3 ± 0.5<br />

σ t / Å 10 ± 50 6 ± 6 4.8 ± 0.7 10 ± 2<br />

A b 0.004 ± 0.001 0.0011 ± 0.0007 0.0129 ± 0.0009 0.013 ± 0.001<br />

q 0 / Å −1 0.077 ± 0.002 0.122 ± 0.01 0.0854 ± 0.0007 0.085 ± 0.001<br />

3d<br />

sse 12.8 48.8 15 19.5<br />

A 0.67 ± 0.06 0.8 ± 0.4 0.7 ± 1 0.8 ± 2<br />

z h / Å 18 ± 1 13 ± 7 15 ± 6 15 ± 9<br />

σ h / Å 6.9 ± 0.6 10 ± 4 6 ± 2 6 ± 4<br />

σ t / Å 6 ± 2 17 ± 3 8 ± 20 9 ± 30<br />

3d-mlv<br />

sse 12.8 35.1 10.4 15.3<br />

A 0.67 ± 0.07 0.7 ± 0.3 0.66 ± 0.05 1.03 ± 0.08<br />

z h / Å 17 ± 3 18 ± 4 17.3 ± 0.4 15.6 ± 0.3<br />

σ h / Å 7 ± 1 6 ± 2 3.7 ± 0.3 5.3 ± 0.5<br />

σ t / Å 6 ± 3 6 ± 6 4.8 ± 0.7 10 ± 2<br />

A b 0.0008 ± 0.0005 0.0012 ± 0.0007 0.0129 ± 0.0009 0.013 ± 0.001<br />

q 0 / Å −1 0.137 ± 0.007 0.122 ± 0.009 0.0855 ± 0.0007 0.085 ± 0.001<br />

sym-4g<br />

sse 12.8 35.1 16.7 19.4<br />

A 1 ± 3 1.1 ± 0.2 1 ± 7e − 05 1.009 ± 0.008<br />

z h / Å 7 ± 500 13 ± 4 1 ± 50 0.2 ± 10000<br />

σ h / Å 10 ± 90 9 ± 3 10 ± 3 9 ± 300<br />

z t / Å 0.8 ± 3000 0.5 ± 30000 0.05 ± 2 4.7 ± 0.5<br />

σ t / Å 10 ± 100 10 ± 1000 10 ± 3 6 ± 0.7


144 Modelling results<br />

H.3 Hexanol<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=72.5<br />

Asym model, sse=69.7<br />

Sym-mlv model, sse=39.6<br />

3d model, sse=73.4<br />

3d-mlv model, sse=39.5<br />

Sym-4g model, sse=71.8<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.7 Low temperature (25 ◦ C), Series 12


H.3 Hexanol 145<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=33.5<br />

Asym model, sse=39.0<br />

Sym-mlv model, sse=25.1<br />

3d model, sse=33.8<br />

3d-mlv model, sse=25.2<br />

Sym-4g model, sse=33.3<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.8 Low temperature (25 ◦ C), Series 15


146 Modelling results<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=18.5<br />

Asym model, sse=18.7<br />

Sym-mlv model, sse=13.9<br />

3d model, sse=18.5<br />

3d-mlv model, sse=13.9<br />

Sym-4g model, sse=18.5<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.9 High temperature (45 ◦ C), Series 14


H.3 Hexanol 147<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=15.8<br />

Asym model, sse=23.1<br />

Sym-mlv model, sse=14.8<br />

3d model, sse=15.8<br />

3d-mlv model, sse=14.8<br />

Sym-4g model, sse=15.8<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.10 High temperature (45 ◦ C), Series 12


148 Modelling results<br />

series 12 series 15 series 14 series 12<br />

sym<br />

sse 72.5 33.5 18.5 15.8<br />

A 0.77 ± 0.03 0.76 ± 0.02 0.501 ± 0.003 0.56 ± 0.02<br />

z h / Å 19.9 ± 0.5 19.2 ± 0.4 16 ± 2 18 ± 2<br />

σ h / Å 4.3 ± 0.4 4.4 ± 0.4 5 ± 0.9 5 ± 0.9<br />

σ t / Å 5 ± 1 4.7 ± 0.9 6 ± 3 5 ± 2<br />

asym<br />

sse 69.7 39 18.7 23.1<br />

A1 0.5 ± 200 0.7 ± 20 0.5 ± 30 0.5 ± 20<br />

A2 0.5 ± 200 0.7 ± 20 0.5 ± 30 0.5 ± 20<br />

z h / Å 16 ± 1 18.5 ± 0.8 16 ± 4 18.4 ± 0.5<br />

σ h / Å 8 ± 1 5.9 ± 0.5 5 ± 2 4.4 ± 0.4<br />

σ t / Å 15.1 ± 0.7 4.7 ± 1 7 ± 5 4.6 ± 0.6<br />

sym-mlv<br />

sse 39.6 25.1 13.9 14.8<br />

A 0.5 ± 0.01 0.51 ± 0.07 0.509 ± 0.003 0.573 ± 0.007<br />

z h / Å 16 ± 2 10 ± 30 17.6 ± 0.4 18.5 ± 0.5<br />

σ h / Å 10 ± 2 11 ± 9 3.7 ± 0.4 4.3 ± 0.4<br />

σ t / Å 19 ± 1 14 ± 8 4.8 ± 0.7 4.8 ± 0.8<br />

A b 0.016 ± 0.003 0.0032 ± 0.0007 0.0127 ± 0.001 0.0057 ± 0.0009<br />

q 0 / Å −1 0.089 ± 0.001 0.115 ± 0.003 0.084 ± 0.0009 0.084 ± 0.002<br />

3d<br />

sse 73.4 33.8 18.5 15.8<br />

A 0.9 ± 0.1 0.8 ± 0.09 0.6 ± 0.2 0.6 ± 0.1<br />

z h / Å 19.9 ± 0.5 19.2 ± 0.4 16 ± 2 18 ± 2<br />

σ h / Å 4.3 ± 0.4 4.4 ± 0.3 5 ± 0.9 5 ± 0.9<br />

σ t / Å 5 ± 1 4.7 ± 0.9 6 ± 3 5 ± 2<br />

3d-mlv<br />

sse 39.5 25.2 13.9 14.8<br />

A 1 ± 0.2 0.7 ± 1 0.65 ± 0.06 0.64 ± 0.06<br />

z h / Å 16 ± 2 10 ± 30 17.6 ± 0.4 18.5 ± 0.5<br />

σ h / Å 10 ± 2 11 ± 9 3.7 ± 0.4 4.3 ± 0.4<br />

σ t / Å 19 ± 1 15 ± 8 4.8 ± 0.7 4.8 ± 0.8<br />

A b 0.016 ± 0.003 0.0033 ± 0.0007 0.0126 ± 0.001 0.0057 ± 0.0009<br />

q 0 / Å −1 0.089 ± 0.001 0.114 ± 0.003 0.084 ± 0.0009 0.084 ± 0.002<br />

sym-4g<br />

sse 71.8 33.3 18.5 15.8<br />

A 1.57 ± 0.07 1.55 ± 0.07 1.003 ± 0.006 1.13 ± 0.03<br />

z h / Å 20.2 ± 0.6 19.6 ± 0.7 16 ± 2 18 ± 2<br />

σ h / Å 4 ± 0.6 4.2 ± 0.6 5 ± 1 5 ± 0.9<br />

z t / Å 3.5 ± 0.8 3.2 ± 1 0.1 ± 3000 0.09 ± 2000<br />

σ t / Å 2 ± 4 2 ± 6 6 ± 50 5 ± 40


H.3 Hexanol 149<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=212<br />

Asym model, sse=217<br />

Sym-mlv model, sse=37.1<br />

3d model, sse=212<br />

3d-mlv model, sse=37.1<br />

Sym-4g model, sse=211<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

-40 -20 0 20 40<br />

Electron density / a.u.<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.11 Low temperature (25 ◦ C), Series 9.


150 Modelling results<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=120<br />

Asym model, sse=124<br />

Sym-mlv model, sse=37.9<br />

3d model, sse=120<br />

3d-mlv model, sse=38.3<br />

Sym-4g model, sse=120<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.12 Low temperature (25 ◦ C), Series 13.


H.3 Hexanol 151<br />

low temp.<br />

series 9 series 13<br />

sym<br />

sse 212 120<br />

A 0.65 ± 0.04 0.66 ± 0.06<br />

z h / Å 10 ± 4 13 ± 3<br />

σ h / Å 11 ± 2 9 ± 1<br />

σ t / Å 4.9 ± 0.7 4.9 ± 1<br />

asym<br />

sse 217 124<br />

A1 0.7 ± 20 0.6 ± 50<br />

A2 0.7 ± 20 0.6 ± 50<br />

z h / Å 13 ± 3 1 ± 1000<br />

σ h / Å 9 ± 2 10 ± 100<br />

σ t / Å 4.5 ± 0.9 7 ± 1<br />

sym-mlv<br />

sse 37.1 37.9<br />

A 0.86 ± 0.05 0.6 ± 0.2<br />

z h / Å 15.3 ± 0.4 15 ± 2<br />

σ h / Å 2.2 ± 0.5 4 ± 2<br />

σ t / Å 5 ± 1 7 ± 9<br />

A b 0.137 ± 0.004 0.063 ± 0.002<br />

q 0 / Å −1 0.1174 ± 0.0002 0.1193 ± 0.0003<br />

3d<br />

sse 212 120<br />

A 0.29 ± 0.03 0.37 ± 0.02<br />

z h / Å 10 ± 4 13 ± 3<br />

σ h / Å 11 ± 2 9 ± 1<br />

σ t / Å 4.9 ± 0.7 4.9 ± 1<br />

3d-mlv<br />

sse 37.1 38.3<br />

A 1.9 ± 0.1 1.6 ± 0.06<br />

z h / Å 15.3 ± 0.4 14.9 ± 0.1<br />

σ h / Å 2.2 ± 0.5 3.9 ± 0.2<br />

σ t / Å 5 ± 1 13 ± 0.4<br />

A b 0.137 ± 0.004 0.076 ± 0.003<br />

q 0 / Å −1 0.1174 ± 0.0002 0.1208 ± 0.0003<br />

sym-4g<br />

sse 211 120<br />

A 1.3 ± 0.2 1.4 ± 0.3<br />

z h / Å 11 ± 7 14 ± 6<br />

σ h / Å 11 ± 3 8 ± 2<br />

z t / Å 3 ± 2 3 ± 4<br />

σ t / Å 3 ± 6 3 ± 9


152 Modelling results<br />

H.4 Heptanol<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=41.8<br />

Asym model, sse=72.8<br />

Sym-mlv model, sse=9.6<br />

3d model, sse=42.0<br />

3d-mlv model, sse=15.0<br />

Sym-4g model, sse=41.8<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.13 Low temperature (25 ◦ C), Series 10


H.4 Heptanol 153<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=41.0<br />

Asym model, sse=121<br />

Sym-mlv model, sse=23.3<br />

3d model, sse=42.0<br />

3d-mlv model, sse=35.7<br />

Sym-4g model, sse=41.6<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.14 Low temperature (25 ◦ C), Series 15


154 Modelling results<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=17.3<br />

Asym model, sse=18.8<br />

Sym-mlv model, sse=8.6<br />

3d model, sse=17.2<br />

3d-mlv model, sse=8.6<br />

Sym-4g model, sse=16.9<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.15 High temperature (45 ◦ C), Series 10


H.4 Heptanol 155<br />

Intensity / a.u.<br />

Measured<br />

Sym model, sse=35.3<br />

Asym model, sse=43.6<br />

Sym-mlv model, sse=18.0<br />

3d model, sse=35.2<br />

3d-mlv model, sse=18.0<br />

Sym-4g model, sse=34.3<br />

0.1 0.2 0.3 0.4 0.5<br />

q / Å -1<br />

Electron density / a.u.<br />

-40 -20 0 20 40<br />

Sym model<br />

Asym model<br />

Sym-mlv model<br />

3d model<br />

3d-mlv model<br />

Sym-4g model<br />

z / Å<br />

Figure H.16 High temperature (45 ◦ C), Series 15


156 Modelling results<br />

low temp.<br />

high temp.<br />

series 10 series 15 series 10 series 15<br />

sym<br />

sse 41.8 41 17.3 35.3<br />

A 0.6 ± 0.3 0.55 ± 0.01 0.5 ± 0.1 0.5 ± 2<br />

z h / Å 17 ± 9 17.8 ± 0.5 10 ± 30 10 ± 200<br />

σ h / Å 7 ± 3 7 ± 0.6 6 ± 10 7 ± 60<br />

σ t / Å 9 ± 30 15.2 ± 0.7 9 ± 80 9 ± 500<br />

asym<br />

sse 72.8 121 18.8 43.6<br />

A1 0.5 ± 40 0.5 ± 40 0.5 ± 100 0.5 ± 100<br />

A2 0.5 ± 40 0.5 ± 40 0.5 ± 100 0.5 ± 100<br />

z h / Å 20 ± 30 15 ± 2 18.6 ± 0.2 10 ± 20<br />

σ h / Å 7 ± 10 7 ± 3 4.6 ± 0.3 6 ± 6<br />

σ t / Å 9 ± 80 12 ± 4 3.9 ± 0.3 8 ± 50<br />

sym-mlv<br />

sse 9.58 23.3 8.61 18<br />

A 0.515 ± 0.008 0.50001 ± 1e − 04 0.52 ± 0.003 0.561 ± 0.006<br />

z h / Å 17.1 ± 0.5 2 ± 8 18 ± 0.3 17.9 ± 0.4<br />

σ h / Å 8 ± 0.5 13.6 ± 0.6 3.6 ± 0.3 3.7 ± 0.4<br />

σ t / Å 17.6 ± 0.2 13.7 ± 0.6 4.9 ± 0.5 4.8 ± 0.7<br />

A b 0.044 ± 0.002 0.033 ± 0.002 0.0169 ± 0.0007 0.026 ± 0.001<br />

q 0 / Å −1 0.0824 ± 0.0002 0.0796 ± 0.0005 0.088 ± 0.0005 0.0888 ± 0.0005<br />

3d<br />

sse 42 42 17.2 35.2<br />

A 0.8 ± 1 0.82 ± 0.09 0.7 ± 5 0.7 ± 30<br />

z h / Å 17 ± 9 20.3 ± 0.5 10 ± 30 10 ± 200<br />

σ h / Å 7 ± 3 4.4 ± 0.5 6 ± 10 7 ± 60<br />

σ t / Å 9 ± 20 5 ± 1 9 ± 80 9 ± 500<br />

3d-mlv<br />

sse 15 35.7 8.58 18<br />

A 0.91 ± 0.06 0.89 ± 0.09 0.71 ± 0.04 0.73 ± 0.07<br />

z h / Å 20.1 ± 0.2 20.3 ± 0.3 18 ± 0.3 17.9 ± 0.4<br />

σ h / Å 3.9 ± 0.2 4.1 ± 0.3 3.6 ± 0.3 3.7 ± 0.4<br />

σ t / Å 4.9 ± 0.5 5 ± 0.7 4.9 ± 0.5 4.8 ± 0.7<br />

A b 0.027 ± 0.001 0.011 ± 0.001 0.0169 ± 0.0007 0.026 ± 0.001<br />

q 0 / Å −1 0.0822 ± 0.0004 0.081 ± 0.001 0.088 ± 0.0005 0.0888 ± 0.0005<br />

sym-4g<br />

sse 41.8 41.6 16.9 34.3<br />

A 1 ± 100 1.39 ± 0.05 1.009 ± 0.004 1.029 ± 0.008<br />

z h / Å 10 ± 9000 20.6 ± 0.8 0.2 ± 8000 0.2 ± 6000<br />

σ h / Å 8 ± 1000 4.2 ± 0.8 10 ± 200 10 ± 90<br />

z t / Å 8 ± 200 4 ± 1 4.9 ± 0.1 4.8 ± 0.1<br />

σ t / Å 7 ± 1000 3 ± 4 5.6 ± 0.6 5.1 ± 0.6


I<br />

Matlab source code<br />

RUCSAXS.M<br />

100 % Process RUCSAXS data . The f uncti on loads the data f i l e s , s u b t r a c t s<br />

% background , converts from channel to q values , w r i t e s processed<br />

% f i l e s ( explained l a t e r ) , d i s p l a y s graphs and returns a [ q , c ] matrix<br />

% ( q ~ q−v e c t o r magnitude , c ~ photon count )<br />

% In order to s u b t r a c t the background , the transmission f a c t o r has to<br />

105 % be c a l c u l a t e d . The sample and background should coincide at high q−v a l u e s .<br />

%<br />

% [ q , c ] = rucsaxs ( fn , sb =1, wf=1, sg =1)<br />

%<br />

% fn : filename , ( p a t t e r n s allowed , e . g . fn = ’07 hex2_1 . p ∗ ’)<br />

110 % i f more than one f i l e matches the pattern , the s p e c t r a are added<br />

% up to one spectrum , which i s processed . The measuring time i s<br />

% looked up in the INF−f i l e . I f no INF−f i l e i s found , the user i s<br />

% prompted f o r measuring time .<br />

%<br />

115 % sb : s u b t r a c t background ( the background f i l e i s s p e c i f i e d in another f i l e )<br />

% (1=yes , 0=no , d e f a u l t =1)<br />

%<br />

% wf : write f i l e s (1=yes , 0=no , d e f a u l t =1), the w r i t t e n f i l e s are :<br />

% [ fn ] . dat : s p e c t r a in count pr . sec pr . channel<br />

120 % [ fn ] . norm . dat : normalized s p e c t r a<br />

% [ fn ] . i n f o . dat : ’ [ n tmt davg ch_interval bmt trans i n t e g r a l e ] ’<br />

% Warning : the f i l e s are overwritten , i f they e x i s t !<br />

%<br />

% sg : show graphs (1=yes , 0=no , d e f a u l t =1)<br />

125 %<br />

% Please note t h a t you have to s p e c i f y s e v e r a l v a r i a b l e s in the<br />

% " r u c s a x s _ s e t t i n g s .m’ f i l e , t h i s i n c l u d e s the q−c a l i b r a t i o n data and<br />

% r e f e r e n c e s to a background sprectrum .<br />

%<br />

130 % q_pr_c : q pr . channel<br />

% q0 : the q v e c t o r magnitude of the d i r e c t beam<br />

% bfn : background s p e c t r a name<br />

%<br />

% Example : [ q , c]= rucsaxs ( ’07 hex2_1 . p ∗ ’ ) ;<br />

135 %<br />

f u n c t i o n [ q , cmb ] = rucsaxs ( fn , v arargin )<br />

%%%%%%%%%%%%%%%%%%%%%%%<br />

140 %%% Check i nput %%%<br />

%%%%%%%%%%%%%%%%%%%%%%%<br />

157


158 Matlab source code<br />

% Default return v a l u e s<br />

q = 0 ;<br />

145 cmbn = 0 ;<br />

% Default f uncti on parameters<br />

sb = 1 ; % yes , s u b t r a c t background<br />

wf = 1 ; % yes , write output f i l e s<br />

150 sg = 1 ; % yes , show graphs<br />

% Number of o p t i o n a l v a r i a b l e s<br />

nv = length ( v arargin ) ;<br />

i f nv >= 3 sg = cell2mat ( v arargin ( 3 ) ) ; end<br />

155 i f nv >= 2 wf = cell2mat ( v arargin ( 2 ) ) ; end<br />

i f nv >= 1 sb = cell2mat ( v arargin ( 1 ) ) ; end<br />

%%%%%%%%%%%%%%%%%%%%%%%<br />

%%% Find f i l e s %%%<br />

160 %%%%%%%%%%%%%%%%%%%%%%%<br />

165<br />

% Get f i l e parts from [ fn ]<br />

[ path , name , ext , versn ] = f i l e p a r t s ( fn ) ;<br />

i f isempty ( path ) path = ’ . ’ ; end<br />

pattern = [ path f i l e s e p name ext ] ;<br />

r u c s a x s _ f i l e s = d i r ( pattern ) ;<br />

f c = length ( r u c s a x s _ f i l e s ) ;<br />

i f ( f c == 0) error ( s p r i n t f ( ’ F i l e ␣ not ␣ found : ␣"%s " ’ , fn ) ) ; end<br />

170 d a t a f i l e s = [ ] ;<br />

tmt = 0 ; % Total measuring time<br />

% Turn on diary ( l o g )<br />

i f wf<br />

175 t_name = name ;<br />

i f t_name == ’ ∗ ’ t_name = ’ a l l ’ ; end<br />

dfn = [ path f i l e s e p t_name ext ’ . log ’ ] ;<br />

i f f c > 1 dfn = [ path f i l e s e p t_name ’ . log ’ ] ; end<br />

d iary ( dfn ) ;<br />

180 end % i f<br />

d isp ( s p r i n t f ( ’%d␣ f i l e ( s ) ␣ found . ’ , f c ) ) ;<br />

%%%%%%%%%%%%%%%%%%%%%%%<br />

185 %%% Load s e t t i n g s %%%<br />

%%%%%%%%%%%%%%%%%%%%%%%<br />

% Please read the r u c s a x s _ s e t t i n g s .m f i l e f o r information<br />

% about each of the parameters<br />

190 [ q_pr_ch , q0 , bfn , bmt , ch_first , ch_last , sbx , dgpfn ] = rucsaxs_settings ;<br />

ch_interval=f l o o r ( ch_last−c h _ f i r s t +1);<br />

ch_pr_q=1/q_pr_ch ;<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%


159<br />

195 %%% Read INF f i l e s %%%<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

f o r i =1: f c<br />

rucsaxs_fn = [ path f i l e s e p r u c s a x s _ f i l e s ( i ) . name ] ;<br />

200 d a t a f i l e . filename = rucsaxs_fn ;<br />

d a t a f i l e . no = 0 ;<br />

d a t a f i l e . abs_temperature = 0 ;<br />

d a t a f i l e . delay = 0 ;<br />

d a t a f i l e . measuring_time = 0 ;<br />

205 d a t a f i l e s = [ d a t a f i l e s ; d a t a f i l e ] ;<br />

end<br />

pattern = [ path f i l e s e p name ’ . INF ’ ] ;<br />

i n f o _ f i l e s = d i r ( pattern ) ;<br />

210 count = length ( i n f o _ f i l e s ) ;<br />

f o r i =1: count<br />

info_fn = [ path f i l e s e p i n f o _ f i l e s ( i ) . name ] ;<br />

[ i f p a t h ifname i f e x t i f v e r s n ] = f i l e p a r t s ( info_fn ) ;<br />

215 % Read the contents of the i nput INF f i l e<br />

contents = textread ( info_fn , ’%s ’ , ’ whitespace ’ , ’ \n ’ ) ;<br />

n = length ( contents ) ;<br />

% Find data<br />

220 ln = 0 ; % S c r i p t data l i n e<br />

f o r j =1:n<br />

s t r = cell2mat ( contents ( j ) ) ;<br />

% Determine channel count f o r p o s i t i o n and energy spectrum<br />

225 i f f i n d s t r ( str , ’ [ channels ] ’ ) > 0 & j+2 0<br />

[T,R] = s t r t o k ( str , ’ : ’ ) ;<br />

[ ll_energy , c , e , ni ] = s s c a n f ( s t r r e p (R, ’ : ␣ ’ , ’ ’ ) , ’%d ’ ) ;<br />

end % i f<br />

i f f i n d s t r ( str , ’PSD1␣Upper␣ Limit : ’ ) > 0<br />

240 [T,R] = s t r t o k ( str , ’ : ’ ) ;<br />

[ ul_energy , c , e , ni ] = s s c a n f ( s t r r e p (R, ’ : ␣ ’ , ’ ’ ) , ’%d ’ ) ;<br />

end % i f<br />

i f ( f i n d s t r ( str , ’No ’ ) > 0 & f i n d s t r ( str , ’Temp ’ ) > 0 & . . .<br />

245 f i n d s t r ( str , ’ Hold ’ ) > 0 & f i n d s t r ( str , ’ Real␣Time ’ ) > 0)<br />

ln = j +1;<br />

break ;


160 Matlab source code<br />

250<br />

end % i f<br />

end % f o r<br />

% I t e r a t e s c r i p t data l i n e s<br />

k = ln ;<br />

while k


161<br />

305<br />

[ t_path , t_name , t_ext , versn ] = f i l e p a r t s ( dfn ) ;<br />

s=load ( [ t_path f i l e s e p t_name s t r r e p ( t_ext , ’P ’ , ’E ’ ) ] ) ;<br />

e=e+s ( 2 : cc_energy +1);<br />

end % f o r<br />

t f c = sum( c ) ; % t o t a l f oton count<br />

d isp ( s p r i n t f ( ’ Total ␣ foton ␣ count :%14d␣ (%0.1 e ) ’ , tfc , t f c ) ) ;<br />

q=transpose ( ( 1 : length ( c ))/ ch_pr_q−q0 ) ;<br />

310 c_err=s q r t ( c ) ;<br />

r p i=c/tmt ; % Raw p o s i t i o n i n t e n s i t y<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

%%% S u b t r a c t i n g background %%%<br />

315 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

load_bg=load ( bfn ) ;<br />

bg=rot90 ( z e r o s (1 , ch_interval ) , −1);<br />

bg_err=bg ;<br />

320 f o r i =1: ch_interval<br />

bg ( i )=load_bg ( i+c h _ f i r s t ) ;<br />

bg_err ( i )= s q r t ( bg ( i ) ) ;<br />

end<br />

bg=bg/bmt ;<br />

325 bg_err=bg_err /bmt ;<br />

% Calculate transmission<br />

trans_bp =752; %%% f i r s t channel to use<br />

trans_np=50; %%% number of p o i n t s<br />

330 trans_ep=trans_bp+trans_np ; %%% end point<br />

i n t=trans_bp : trans_ep ;<br />

trans=mean( c ( i n t )/ tmt)/mean( load_bg ( i n t )/bmt ) ;<br />

d isp ( s p r i n t f ( ’ Transmission : ␣␣%6.2 f ␣ ( channels ␣%3d␣ through ␣%3d) ’ , trans , . . .<br />

335 trans_bp , trans_ep ) ) ;<br />

s t r = s p r i n t f ( [ ’The␣ transmission ␣ i s ␣ the ␣ r a t i o ␣between ␣ the ␣ average␣ ’ . . .<br />

’ i n t e n s i t y ␣ pr . ␣ channel ␣ of ␣ the ␣ current ␣ data ␣ s e t ␣and␣ ’ . . .<br />

’ the ␣background ␣ data ␣ s e t . ’ ] ) ;<br />

%disp ( s t r ) ;<br />

340<br />

% c a l c u l a t i n g c minus background<br />

cmb=c ( c h _ f i r s t : ch_last )/ tmt−bg∗ trans ;<br />

cmb_err=(( c_err ( c h _ f i r s t : ch_last )/ tmt).^2+ bg_err . ^ 2 ) . ^ 0 . 5 ;<br />

345 % C a l c u l a t i n g the normalized spectrum with the trapez −f uncti on<br />

% i n t e g r a l e=sum(cmb )∗( q(2)−q (1))<br />

i n t e g r a l e=trapz ( q ( c h _ f i r s t : ch_last ) ,cmb ) ;<br />

cmbn=cmb/ i n t e g r a l e ;<br />

cmbn_err=cmb_err/ i n t e g r a l e ;<br />

350<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

%%% Save data f i l e s %<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%


162 Matlab source code<br />

355 [ path2 name2 ext2 versn2 ] = f i l e p a r t s ( d a t a f i l e s ( 1 ) . filename ) ;<br />

base = [ path f i l e s e p name2 ext ] ; % base path & f i lename<br />

i f name == ’ ∗ ’ name = ’ a l l ’ ; end<br />

i f f c > 1 base = [ path f i l e s e p name ] ; end<br />

360 e_vek=transpose ( [ 1 : 1 : length ( e ) ] ) ;<br />

i f wf<br />

% Save ( q , c ) " raw " data to f i l e<br />

s a v e f i l e =[ base ’ . dat ’ ] ;<br />

365 out = [ q ( c h _ f i r s t : ch_last ) cmb cmb_err ] ;<br />

save ( s a v e f i l e , ’ out ’ , ’−ASCII ’ ) ;<br />

% Save the normalized spectrum<br />

s a v e f i l e =[ base ’ . norm . dat ’ ] ;<br />

370 out = [ q ( c h _ f i r s t : ch_last ) cmbn cmbn_err ] ;<br />

save ( s a v e f i l e , ’ out ’ , ’−ASCII ’ ) ;<br />

% Save div info , computer readable<br />

s a v e f i l e =[ base ’ . i n f o . dat ’ ] ;<br />

375 out=[n tmt ch_interval bmt trans i n t e g r a l e ] ;<br />

save ( s a v e f i l e , ’ out ’ , ’−ASCII ’ ) ;<br />

% Save info , counts pr second f o r p l o t counts pr sec /q<br />

s a v e f i l e =[ base ’ . inten . dat ’ ] ;<br />

380 out=[q ( c h _ f i r s t : ch_last ) r p i ( c h _ f i r s t : ch_last ) bg ] ;<br />

save ( s a v e f i l e , ’ out ’ , ’−ASCII ’ ) ;<br />

%Save info , energy spectrum<br />

s a v e f i l e =[ base ’ . energy . dat ’ ] ;<br />

385 out=[e_vek e ] ;<br />

save ( s a v e f i l e , ’ out ’ , ’−ASCII ’ ) ;<br />

% turn diary o f f diary<br />

390 end % i f<br />

395<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

%%% P l o t t i n g graphs %%%<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

legend_str = s t r r e p (name2 , ’_ ’ , ’− ’ ) ;<br />

i f f c > 1 legend_str = s t r r e p (name , ’_’ , ’− ’ ) ; end<br />

% Raw data p l o t<br />

400 i f sg<br />

f i g u r e (105)<br />

hold on<br />

p l o t ( rpi , ’ bx ’ , ’ MarkerSize ’ , 2 . 0 )<br />

405 x l a b e l ( ’ Channel ’ ) ;<br />

y l a b e l ( ’ I n t e n s i t y ␣/␣Counts ␣ pr . ␣ second ’ ) ;


163<br />

t i t l e ( ’Raw␣ data ␣ spectrum ’ ) ;<br />

legend ( [ legend_str ’ ␣raw␣ data ’ ] ) ;<br />

410 % Setup f i g u r e<br />

s e t ( gca , ’ XGrid ’ , ’ on ’ ) ;<br />

s e t ( gca , ’ Units ’ , ’ centimeters ’ ) ;<br />

s e t ( gca , ’ Position ’ , [ 3 , 2 , 1 1 , 8 ] ) ;<br />

v = a x i s ;<br />

415 a x i s ( [ 1 , cc_position , 0 , v ( 4 ) ] ) ;<br />

lower_limit_line = l i n e ( [ c h _ f i r s t c h _ f i r s t ] , [ 0 v ( 4 ) ] ) ;<br />

s e t ( lower_limit_line , ’ LineStyle ’ , ’− ’ ) ;<br />

s e t ( lower_limit_line , ’ Color ’ , ’ k ’ ) ;<br />

upper_limit_line = l i n e ( [ ch_last ch_last ] , [ 0 v ( 4 ) ] ) ;<br />

420 s e t ( upper_limit_line , ’ LineStyle ’ , ’− ’ ) ;<br />

s e t ( upper_limit_line , ’ Color ’ , ’ k ’ ) ;<br />

i f wf<br />

p r i n t ( ’−r300 ’ , ’−depsc ’ , [ base ’ . raw . c o l o r . eps ’ ] ) ;<br />

425 p r i n t ( ’−r300 ’ , ’−deps ’ , [ base ’ . raw . eps ’ ] ) ;<br />

end % i f<br />

430<br />

435<br />

hold<br />

end<br />

o f f<br />

% Plot r p i and bg without error bars<br />

i f sg<br />

f i g u r e ( 1 1 0 ) ;<br />

hold on<br />

p l o t ( q ( c h _ f i r s t : ch_last ) , bg , ’ k . ’ , ’ MarkerSize ’ , 4 . 0 ) ;<br />

%errorbar ( q , bg , bg_err )<br />

p l o t ( q ( c h _ f i r s t : ch_last ) , r p i ( c h _ f i r s t : ch_last ) , ’ bx ’ , ’ MarkerSize ’ , 2 . 0 ) ;<br />

%errorbar ( q , c , c_err )<br />

440 x l a b e l ( ’ q␣ [ ␣Å^{␣ −1}] ’ ) ;<br />

y l a b e l ( ’ I n t e n s i t y ␣ [ counts ␣ pr . ␣ second ] ’ ) ;<br />

t i t l e ( ’ Background ␣ spectrum␣and␣raw␣ data ␣ spectrum ’ ) ;<br />

legend ( [ legend_str ’ ␣ background ’ ] , [ legend_str ’ ␣raw␣ data ’ ] ) ;<br />

445 % Setup f i g u r e<br />

s e t ( gca , ’ XGrid ’ , ’ on ’ ) ;<br />

s e t ( gca , ’ Units ’ , ’ centimeters ’ ) ;<br />

s e t ( gca , ’ Position ’ , [ 3 , 2 , 1 1 , 8 ] ) ;<br />

v = a x i s ;<br />

450 a x i s ( [ 0 , 0.55 , 0 , v ( 4 ) ] ) ;<br />

i f wf<br />

p r i n t ( ’−r300 ’ , ’−depsc ’ , [ base ’ . cb . c o l o r . eps ’ ] ) ;<br />

p r i n t ( ’−r300 ’ , ’−deps ’ , [ base ’ . cb . eps ’ ] ) ;<br />

455 end % i f<br />

hold<br />

end<br />

o f f


164 Matlab source code<br />

460 % Plot cmb without error bars<br />

i f sg<br />

f i g u r e (100)<br />

hold on<br />

465 p l o t ( q ( c h _ f i r s t : ch_last ) , cmb , ’ bx ’ , ’ MarkerSize ’ , 2 . 0 ) ;<br />

%errorbar ( q , cmb , cmb_err , ’ . ’ )<br />

x l a b e l ( ’ q␣/␣Å^{␣−1} ’ ) ;<br />

y l a b e l ( ’ I n t e n s i t y ␣/␣ counts ␣ pr . ␣ second ’ ) ;<br />

t i t l e ( ’ Spectrum ␣ with ␣ subtracted␣background ’ ) ;<br />

470<br />

% Setup f i g u r e<br />

s e t ( gca , ’ XGrid ’ , ’ on ’ ) ;<br />

s e t ( gca , ’ Units ’ , ’ centimeters ’ ) ;<br />

s e t ( gca , ’ Position ’ , [ 3 , 2 , 1 1 , 8 ] ) ;<br />

475 v = a x i s ;<br />

a x i s ( [ 0 , 0.55 , 0 , v ( 4 ) ] ) ;<br />

legend ( legend_str ) ;<br />

i f<br />

wf<br />

480 p r i n t ( ’−r300 ’ , ’−depsc ’ , [ base ’ . cmb . c o l o r . eps ’ ] ) ;<br />

p r i n t ( ’−r300 ’ , ’−deps ’ , [ base ’ . cmb . eps ’ ] ) ;<br />

i f i s u n i x<br />

cmd = s p r i n t f ( ’ gracebat ␣−settype ␣xydy␣%s ␣−param␣%s ␣−s a v e a l l ␣%s ’ , . . .<br />

[ base ’ . dat ’ ] , dgpfn , [ base ’ . cmb . agr ’ ] ) ;<br />

485 % [ status , r e s u l t ] = system (cmd ) ;<br />

end % i f<br />

end % i f<br />

490 end<br />

hold<br />

o f f<br />

% Plot energy spectrum<br />

i f sg<br />

f i g u r e (120)<br />

495 hold on<br />

p l o t ( e , ’ bx ’ , ’ MarkerSize ’ , 2 . 0 ) ;<br />

x l a b e l ( ’ Energy ␣ channel ’ ) ;<br />

y l a b e l ( ’ Total ␣ counts ’ ) ;<br />

500 t i t l e ( ’ Energy ␣ spectrum ’ ) ;<br />

% Setup f i g u r e<br />

s e t ( gca , ’ XGrid ’ , ’ on ’ ) ;<br />

s e t ( gca , ’ Units ’ , ’ centimeters ’ ) ;<br />

505 s e t ( gca , ’ Position ’ , [ 3 , 2 , 1 1 , 8 ] ) ;<br />

v = a x i s ;<br />

a x i s ( [ 1 , cc_energy , 0 , v ( 4 ) ] ) ;<br />

legend ( legend_str ) ;<br />

510 lower_limit_line = l i n e ( [ ll_energy ll_energy ] , [ 0 v ( 4 ) ] ) ;<br />

s e t ( lower_limit_line , ’ LineStyle ’ , ’− ’ ) ;<br />

s e t ( lower_limit_line , ’ Color ’ , ’ k ’ ) ;


165<br />

upper_limit_line = l i n e ( [ ul_energy ul_energy ] , [ 0 v ( 4 ) ] ) ;<br />

s e t ( upper_limit_line , ’ LineStyle ’ , ’− ’ ) ;<br />

515 s e t ( upper_limit_line , ’ Color ’ , ’ k ’ ) ;<br />

i f wf<br />

p r i n t ( ’−r300 ’ , ’−depsc ’ , [ base ’ . e . c o l o r . eps ’ ] ) ;<br />

p r i n t ( ’−r300 ’ , ’−deps ’ , [ base ’ . e . eps ’ ] ) ;<br />

520 end % i f<br />

hold<br />

end<br />

o f f<br />

525 q = q ( c h _ f i r s t : ch_last ) ;<br />

RUCSAXS_SETTINGS.M<br />

f u n c t i o n<br />

[ q_pr_ch , q0 , bfn , bmt , ch_first , ch_last , sbx , dgpfn ] = rucsaxs_settings<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

%%% Data processing parameters %<br />

529 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

% q−c a l i b r a t i o n constants<br />

q_pr_ch=8.0318 e −4;<br />

% d i r e c t beam<br />

534 q0 =0.1759;<br />

539<br />

% Background spectrum f i l e name<br />

bfn=’ /home/ grupper/ alkohol / rucsaxs/background /background . dat ’ ;<br />

bmt=64800; % background spectrum measuring time<br />

% Channels to use<br />

c h _ f i r s t = 250;<br />

ch_last = 900;<br />

544 % Beam p r o f i l e along x−a x i s<br />

% ( h o r i z o n t a l a x i s perpendicular to beam f l u x )<br />

% f i t t e d to a gaussian f uncti on with a spread of<br />

sbx = 0 . 1 4 6 8 1 ;<br />

% centered around 0!<br />

549<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

%%% Graph parameters %<br />

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />

554 % Default Grace parameters f i lename<br />

dgpfn = ’ /home/ grupper/ alkohol / s c r i p t s / d e f a u l t . grace . par ’ ;<br />

DSC7.M<br />

% Process data f i l e ( s ) from the DSC7 equipment in øSren Hvidts l a b o r a t o r y .<br />

% The s c r i p t reads the [ i f n ] . d7 f i l e , p l o t s the data and<br />

% ( o p t i o n a l l y ) w r i t e s an [ i f n ] . dat f i l e containing the (T, J ) data s e t .<br />

559 %


166 Matlab source code<br />

% f uncti on [T, J ] = dsc7 ( i f n , s t r = ’ ’ , sg =1, wf=1, pr=0)<br />

% i f n − i nput f i l e name , p a t t e r n s are allowed<br />

% s t r − t i t l e s t r i n g of the graph produced . ( d e f a u l t : ’ ’ )<br />

% sg − show graphs (1=yes , 0=no , d e f a u l t : 1)<br />

564 % wf − write output f i l e , [ i f n ] . dat (1=yes , 0=no , d e f a u l t : 1)<br />

% containing (T, J) data s e t .<br />

% pr − pans reversed ( Sample and r e f e r e n c e pans reversed )<br />

%<br />

% Example :<br />

569 %<br />

% dsc7 ( ’ hex2 ∗ ’ , ’DSC on ULVs (<strong>DPPC</strong>) pertubated by hexanol ’ , 1 , 0 ) ;<br />

% Shows a (T, J) graph with the t i t l e above , and does not write the output f i l e .<br />

574<br />

f u n c t i o n [T, J ] = dsc7 ( ifn , v arargin )<br />

% Default f uncti on parameters<br />

J = 0 ;<br />

T = 0 ;<br />

t i t l e _ s t r = ’ ’ ;<br />

579 legend_str = ’ ’ ;<br />

sg = 1 ; % yes , show graphs<br />

wf = 1 ; % yes , write output f i l e s<br />

pr = 0 ; % no , pans are not reversed<br />

graph_y_max = 0 ;<br />

584 graph_y_min = 0 ;<br />

% Number of o p t i o n a l v a r i a b l e s<br />

nv = length ( v arargin ) ;<br />

i f nv >= 4 pr = cell2mat ( v arargin ( 4 ) ) ; end<br />

589 i f nv >= 3 wf = cell2mat ( v arargin ( 3 ) ) ; end<br />

i f nv >= 2 sg = cell2mat ( v arargin ( 2 ) ) ; end<br />

i f nv >= 1 t i t l e _ s t r = cell2mat ( v arargin ( 1 ) ) ; end<br />

% Get f i l e parts from i f n<br />

594 [ path , name , ext , versn ] = f i l e p a r t s ( i f n ) ;<br />

i f isempty ( path ) path = ’ . ’ ; end<br />

ext = ’ . d7 ’ ;<br />

pattern = [ path f i l e s e p name ext ] ;<br />

f i l e s = d i r ( pattern ) ;<br />

599 count = length ( f i l e s ) ;<br />

i f count == 0<br />

ext = ’ . D7 ’ ;<br />

pattern = [ path f i l e s e p name ext ] ;<br />

f i l e s = d i r ( pattern ) ;<br />

604 count = length ( f i l e s ) ;<br />

end % i f<br />

i f count == 0<br />

error ( s p r i n t f ( ’ F i l e ␣ not ␣ found␣ using ␣ pattern : ␣%s ’ , pattern ) ) ;<br />

609 end % i f<br />

f o r i =1: count<br />

T = 0 ;


167<br />

614<br />

J = 0 ;<br />

fn = [ path f i l e s e p f i l e s ( i ) . name ] ;<br />

d isp ( s p r i n t f ( ’ Current ␣ f i l e : ␣%s ’ , fn ) ) ;<br />

% Read i nput f i l e<br />

619 contents = textread ( fn , ’%s ’ , ’ whitespace ’ , ’ \n ’ ) ;<br />

measurement_count = str2num ( char ( contents ( 8 ) ) ) ;<br />

scan_rate = str2num ( char ( contents ( 9 ) ) ) ;<br />

T_start = str2num ( char ( contents ( 7 ) ) ) ;<br />

T_final = str2num ( char ( contents ( 1 3 ) ) ) ;<br />

624 Time_at_T_start = str2num ( char ( contents ( 1 5 ) ) ) ;<br />

J_axis_max = str2num ( char ( contents ( 1 4 ) ) ) ;<br />

metainfo = char ( contents ( 1 ) ) ;<br />

id = s t r c a t ( metainfo ( 2 : 3 0 ) ) ;<br />

J_factor = J_axis_max / 10000;<br />

629 i f pr<br />

J_factor = −1 ∗ J_factor ;<br />

end<br />

i f length ( t i t l e _ s t r ) == 0<br />

t i t l e _ s t r = id ;<br />

634 end<br />

d isp ( s p r i n t f ( ’ ID : ␣"%s "\ t \ tScanrate : ␣%3d␣ ’ , id , scan_rate ) ) ;<br />

% Process data<br />

time = ( T_final − T_start ) / scan_rate ;<br />

639 x = 58:57+ measurement_count ;<br />

T = T_start + (x−58) ∗ ( T_final − T_start ) / measurement_count ;<br />

f o r j =58:57+measurement_count<br />

J ( j −57) = ( str2num ( char ( contents ( j ) ) ) − J ( 1 ) ) ∗ J_factor ;<br />

end<br />

644 Tmin = min (T) ; Tmax = max(T) ; deltaT = Tmax−Tmin ;<br />

Jmin = min ( J ) ; Jmax = max( J ) ; deltaJ = Jmax−Jmin ;<br />

i f sg<br />

% Plot data<br />

649 f i g u r e ( 1 ) ; hold on ; p l o t (T, J , ’ k . ’ , ’ MarkerSize ’ , 4 . 0 ) ; hold o f f ;<br />

s e t ( gca , ’ XGrid ’ , ’ on ’ ) ;<br />

x l a b e l ( ’ Temperature␣/␣\ circC ’ ) ;<br />

y l a b e l ( ’ Heat␣ flow ␣/␣mW’ ) ;<br />

t i t l e ( t i t l e _ s t r ) ;<br />

654 s e t ( gca , ’ Units ’ , ’ centimeters ’ ) ;<br />

s e t ( gca , ’ Position ’ , [ 3 , 2 , 1 0 , 8 ] ) ;<br />

i f length ( legend_str ) == 0<br />

legend_str = char ( id ) ;<br />

e l s e<br />

659 legend_str = char ( legend_str , id ) ;<br />

end<br />

% Set a x i s<br />

v = a x i s ;<br />

664 i f Jmax > graph_y_max<br />

graph_y_max = ( f l o o r (Jmax /10) + 1) ∗ 10;


168 Matlab source code<br />

end<br />

i f Jmin < graph_y_min<br />

graph_y_min = f l o o r ( Jmin /10) ∗ 10;<br />

669 end<br />

a x i s ( [ v ( 1 ) , v ( 2 ) , graph_y_min , graph_y_max ] ) ;<br />

i f count == 1<br />

tx = deltaT /8 + Tmin ;<br />

674 ty = Jmax − 1∗ deltaJ /8;<br />

text ( tx , ty , s p r i n t f ( ’ Scanrate : ␣%d␣K␣/␣min . ’ , scan_rate ) ) ;<br />

679<br />

ty = Jmax − 2∗ deltaJ /8;<br />

text ( tx , ty , s p r i n t f ( ’ Delay ␣ at ␣T−s t a r t : ␣%d␣min . ’ , Time_at_T_start ) ) ;<br />

ty = Jmax − 3∗ deltaJ /8;<br />

text ( tx , ty , s p r i n t f ( ’ ID : ␣%s ’ , id ) ) ;<br />

ty = Jmax − 4∗ deltaJ /8;<br />

684 f n s t r = s t r r e p ( fn , ’_’ , ’− ’ ) ;<br />

text ( tx , ty , s p r i n t f ( ’ F i l e : ␣%s ’ , f n s t r ) ) ;<br />

end % i f<br />

end % i f<br />

689 % Write output f i l e<br />

i f wf<br />

data = [ transpose (T) , transpose ( J ) ] ;<br />

[ path , name , extension , versn ] = f i l e p a r t s ( fn ) ;<br />

694 % Set output f i lename<br />

ofn = [ path f i l e s e p name ’ . dat ’ ] ;<br />

save ( ofn , ’ data ’ , ’−ASCII ’ ) ;<br />

end % i f<br />

end % f o r<br />

699<br />

i f sg<br />

legend ( legend_str ) ;<br />

end % i f<br />

F_SYM.M<br />

% Returns the form f a c t o r from a 1−dimensional symmetric<br />

% e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

704 %<br />

% f_sym ( q , [A, zh , sh , s t ] , [ ] )<br />

%<br />

% q ~ s c a t t e r i n g v e c t o r<br />

% A ~ amplitude<br />

709 % zh ~ center of gauss peak<br />

% sh ~ spread of head gauss f uncti on<br />

% s t ~ spread of t a i l gauss f uncti on<br />

%<br />

714 f u n c t i o n f = f_sym (q , parameters , constants )


169<br />

A = parameters ( 1 ) ;<br />

zh = parameters ( 2 ) ;<br />

sh = parameters ( 3 ) ;<br />

719 s t = parameters ( 4 ) ;<br />

f = (A ∗ ( fourier_gauss (q , zh , sh ) + fourier_gauss (q,−zh , sh ) ) − fourier_gauss (q , 0 , s t ) ) . / q ;<br />

I_SYM.M<br />

% Returns the i n t e n s i t y from a point s c a t t e r e r f o r<br />

% a symmetric e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

%<br />

724 % I_sym(q , [A, zh , sh , s t ] , [ ] )<br />

%<br />

% q ~ s c a t t e r i n g v e c t o r<br />

% A ~ amplitude<br />

% zh ~ center of gauss peak<br />

729 % sh ~ spread of head gauss f uncti on<br />

% s t ~ spread of t a i l gauss f uncti on<br />

%<br />

734<br />

f u n c t i o n I = I_sym(q , parameters , constants )<br />

f = f_sym (q , parameters , constants ) ;<br />

I = f . ∗ conj ( f ) ;<br />

F_ASYM.M<br />

% Returns the form f a c t o r from a 1−dimensional asymmetric<br />

% e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

%<br />

739 % f_asym(q , [ A1, A2, zh , sh , s t ] , [ ] )<br />

%<br />

% Inner Gauss f uncti on<br />

% A1 ~ amplitude f o r gauss f uncti on<br />

% zh ~ center of gauss peak<br />

744 % sh ~ spread of head gauss f uncti on<br />

%<br />

% Outer Gauss f uncti on<br />

% A2 ~ amplitude f o r gauss f uncti on<br />

%<br />

749 % Tail<br />

% s t ~ spread of t a i l gauss f uncti on<br />

f u n c t i o n f = f_asym (q , parameters , constants )<br />

754 A1 = parameters ( 1 ) ;<br />

A2 = parameters ( 2 ) ;<br />

zh = parameters ( 3 ) ;<br />

sh = parameters ( 4 ) ;<br />

s t = parameters ( 5 ) ;<br />

759<br />

f = (A1∗ fourier_gauss (q,−zh , sh ) + A2∗ fourier_gauss (q , zh , sh ) − fourier_gauss (q , 0 , s t ) ) . / q ;


170 Matlab source code<br />

I_ASYM.M<br />

% Returns the i n t e n s i t y from a point s c a t t e r e r f o r<br />

% an asymmetric e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

%<br />

% I_asym( q , [ A1, A2, zh , sh , s t ] , [ ] )<br />

764 %<br />

% Inner Gauss f uncti on<br />

% A1 ~ amplitude f o r gauss f uncti on<br />

% zh ~ center of gauss peak<br />

% sh ~ spread of head gauss f uncti on<br />

769 %<br />

% Outer Gauss f uncti on<br />

% A2 ~ amplitude f o r gauss f uncti on<br />

%<br />

% Tail<br />

774 % s t ~ spread of t a i l gauss f uncti on<br />

f u n c t i o n I = I_asym(q , parameters , constants )<br />

f = f_asym (q , parameters , constants ) ;<br />

779 I = f . ∗ conj ( f ) ;<br />

I_SYM_MLV.M<br />

779 % Returns the i n t e n s i t y from a point s c a t t e r e r f o r<br />

% a symmetric e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

%<br />

% I_sym_mlv ( q , [A, zh , sh , st , Ab , q0 ] , [ phase ] )<br />

%<br />

784 % A ~ amplitude f o r gauss f uncti on<br />

% zh ~ center of gauss peak<br />

% sh ~ spread of head gauss f uncti on<br />

% s t ~ spread of t a i l gauss f uncti on<br />

%<br />

789 % M u l t i l a m e l l a r c o r r e c t i o n<br />

% Ab ~ the r e l a t i v e amount of m u l t i l a m e l l a r v e s i c l e s ( mlv/ u l v )<br />

% q0 ~ F i r s t bragg peak of mlv<br />

%<br />

% phase ~ 1 ( low temp ) or 2 ( high temp )<br />

794<br />

f u n c t i o n I = I_sym_mlv(q , parameters , constants ) ;<br />

A = parameters ( 1 ) ;<br />

zh = parameters ( 2 ) ;<br />

799 sh = parameters ( 3 ) ;<br />

s t = parameters ( 4 ) ;<br />

Ab = parameters ( 5 ) ;<br />

q0 = parameters ( 6 ) ;<br />

phase = constants ( 1 ) ;<br />

804<br />

Isym = I_sym(q , [A, zh , sh , s t ] , [ ] ) ;<br />

I = Isym . ∗ (1 + Ab. ∗ S_mlv(q , [ q0 ] , [ phase ] ) ) ;


171<br />

I_ASYM_MLV.M<br />

% Returns the i n t e n s i t y from a point s c a t t e r e r f o r<br />

% an asymmetric e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

809 %<br />

% I_asym_mlv( q , [A, zh , sh , st , Ab , q0 ] , [ A_f , zh_f , sh_f , phase ] )<br />

%<br />

% Inner Gauss f uncti on<br />

% A ~ amplitude f o r gauss f uncti on<br />

814 % zh ~ center of gauss peak<br />

% sh ~ spread of head gauss f uncti on<br />

%<br />

% Outer Gauss f uncti on<br />

% A_f ~ amplitude weighing f a c t o r<br />

819 % zh_f ~ peak weighing f a c t o r<br />

% sh_f ~ spread weighing f a c t o r<br />

%<br />

% Tail<br />

% s t ~ spread of t a i l gauss f uncti on<br />

824 %<br />

% M u l t i l a m e l l a r c o r r e c t i o n<br />

% Ab ~ the r e l a t i v e amount of m u l t i l a m e l l a r v e s i c l e s ( mlv/ u l v )<br />

% q0 ~ F i r s t bragg peak of mlv<br />

%<br />

829 % phase ~ 1 ( low temp ) or 2 ( high temp )<br />

f u n c t i o n I = I_asym_mlv(q , parameters , constants ) ;<br />

A = parameters ( 1 ) ;<br />

834 zh = parameters ( 2 ) ;<br />

sh = parameters ( 3 ) ;<br />

s t = parameters ( 4 ) ;<br />

Ab = parameters ( 5 ) ;<br />

q0 = parameters ( 6 ) ;<br />

839 A_f = constants ( 1 ) ;<br />

zh_f = constants ( 2 ) ;<br />

sh_f = constants ( 3 ) ;<br />

phase = constants ( 4 ) ;<br />

844 Iasym = I_asym(q , [A, zh , sh , s t ] , [ A_f , zh_f , sh_f ] ) ;<br />

I = Iasym . ∗ (1 + Ab. ∗ S_mlv(q , [ q0 ] , [ phase ] ) ) ;<br />

F_3D.M<br />

% Returns the form f a c t o r from a 3−dimensional symmetric<br />

% e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

% ( from : Brzustowicz & Brunger )<br />

849 %<br />

% f_3d_bb(q , [A, zh , sh , s t ] , [ r0 ] )<br />

%<br />

% A ~ amplitude f o r gauss f uncti on<br />

% zh ~ center of gauss peak<br />

854 % sh ~ spread of head gauss f uncti on<br />

% s t ~ spread of t a i l gauss f uncti on


172 Matlab source code<br />

% r0 ~ v e s i c l e radius<br />

%<br />

859 f u n c t i o n f = f_3d_bb (q , parameters , constants )<br />

A = parameters ( 1 ) ;<br />

zh = parameters ( 2 ) ;<br />

sh = parameters ( 3 ) ;<br />

864 s t = parameters ( 4 ) ;<br />

r0 = constants ( 1 ) ;<br />

f = (A∗ fourier_spheric al _gauss (q , r0−zh , sh ) . . .<br />

+ A∗ fourier_spheric a l_gauss (q , r0+zh , sh ) . . .<br />

869 − fourier_spheric al _gauss (q , r0 , s t ) ) . / q ;<br />

I_3D.M<br />

869 % Returns the i n t e n s i t y from a point s c a t t e r e r f o r<br />

% a 3d symmetric e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

%<br />

% I_3d ( q , [A, zh , sh , s t ] , [ ] )<br />

%<br />

874 % q ~ s c a t t e r i n g v e c t o r<br />

% A ~ amplitude<br />

% zh ~ center of gauss peak<br />

% sh ~ spread of head gauss f uncti on<br />

% s t ~ spread of t a i l gauss f uncti on<br />

879 %<br />

f u n c t i o n I = I_3d(q , parameters , constants )<br />

R0 = 600; % Å<br />

884 R_s = 120; % Å<br />

R_start = R0−R_s∗2;<br />

R_end = R0+R_s∗2;<br />

% checked with 8 , 16 , 32 , 64 , 128 , 256<br />

889 % above 64 , a l l graphs are i d e n t i c a l<br />

R_steps = 64;<br />

R_stepsize = (R_end − R_start )/ R_steps ;<br />

R = R_start : R_stepsize : R_end ;<br />

894 R_dist = gauss (R, R0 , R_s ) ;<br />

N = length ( q ) ;<br />

I = z e r o s (1 ,N) ;<br />

899 % I t e r a t e v e s i c l e radius<br />

f o r j =1:R_steps<br />

f = f_3d (q , parameters , [R( j ) ] ) ;<br />

I_R = f . ∗ conj ( f ) ;<br />

I = I + R_dist ( j ) ∗ I_R ;<br />

904 end ;


173<br />

I_3D_MLV.M<br />

904 % Returns the i n t e n s i t y from a point s c a t t e r e r f o r<br />

% a 3d e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r .<br />

%<br />

% I_3d_mlv ( q , [A, zh , sh , st , Ab , q0 ] , [ phase ] )<br />

%<br />

909 % A ~ amplitude f o r gauss f uncti on<br />

% zh ~ center of gauss peak<br />

% sh ~ spread of head gauss f uncti on<br />

% s t ~ spread of t a i l gauss f uncti on<br />

%<br />

914 % M u l t i l a m e l l a r c o r r e c t i o n<br />

% Ab ~ the r e l a t i v e amount of m u l t i l a m e l l a r v e s i c l e s ( mlv/ u l v )<br />

% q0 ~ F i r s t bragg peak of mlv<br />

%<br />

% phase ~ 1 ( low temp ) or 2 ( high temp )<br />

919<br />

f u n c t i o n I = I_3d_mlv(q , parameters , constants ) ;<br />

A = parameters ( 1 ) ;<br />

zh = parameters ( 2 ) ;<br />

924 sh = parameters ( 3 ) ;<br />

s t = parameters ( 4 ) ;<br />

Ab = parameters ( 5 ) ;<br />

q0 = parameters ( 6 ) ;<br />

phase = constants ( 1 ) ;<br />

929<br />

I3d = I_3d(q , [A, zh , sh , s t ] , [ ] ) ;<br />

I = I3d . ∗ (1 + Ab. ∗ S_mlv(q , [ q0 ] , [ phase ] ) ) ;<br />

F_SYM_4G.M<br />

% Returns the form f a c t o r from a 1−dimensional symmetric<br />

% e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r using 4 gauss f u n c t i o n s .<br />

934 %<br />

% f_sym_4g( q , [A, zh , sh , zt , s t ] , [ ] )<br />

%<br />

% q ~ s c a t t e r i n g v e c t o r<br />

% A ~ amplitude<br />

939 % zh ~ center of gauss peak f o r the heads<br />

% sh ~ spread of head gauss f uncti on<br />

% z t ~ center of gauss peak f o r the t a i l s<br />

% s t ~ spread of t a i l gauss f uncti on<br />

%<br />

944<br />

f u n c t i o n f = f_sym_4g(q , parameters , constants )<br />

A = parameters ( 1 ) ;<br />

zh = parameters ( 2 ) ;<br />

949 sh = parameters ( 3 ) ;<br />

zt = parameters ( 4 ) ;<br />

s t = parameters ( 5 ) ;


174 Matlab source code<br />

f = (A ∗ ( fourier_gauss (q , zh , sh ) + fourier_gauss (q,−zh , sh ) ) . . .<br />

954 − fourier_gauss (q,−zt , s t ) − fourier_gauss (q , zt , s t ) ) . / q ;<br />

I_SYM_4G.M<br />

954 % Returns the form f a c t o r from a 1−dimensional symmetric<br />

% e l e c t r o n d e n s i t y p r o f i l e f o r a b i l a y e r using 4 gauss f u n c t i o n s .<br />

%<br />

% I_sym_4g ( q , [A, zh , sh , zt , s t ] , [ ] )<br />

%<br />

959 % q ~ s c a t t e r i n g v e c t o r<br />

% A ~ amplitude<br />

% zh ~ center of gauss peak f o r the heads<br />

% sh ~ spread of head gauss f uncti on<br />

% z t ~ center of gauss peak f o r the t a i l s<br />

964 % s t ~ spread of t a i l gauss f uncti on<br />

%<br />

f u n c t i o n I = I_sym_4g(q , parameters , constants )<br />

969 f = f_sym_4g (q , parameters , constants ) ;<br />

I = f . ∗ conj ( f ) ;<br />

S_MLV.M<br />

% Returns the s t r u c t u r e f a c t o r f o r m u l t i l a m e l l a r v e s i c l e s .<br />

%<br />

% S_mlv(q , [ q0 ] , [ phase ] )<br />

%<br />

974 % q0 ~ Bragg peak of mlv<br />

% sb ~ spred of Bragg peak<br />

% phase ~ 1 ( low temp ) or 2 ( high temp )<br />

979<br />

f u n c t i o n S = S_mlv(q , parameters , constants )<br />

q0 = parameters ( 1 ) ;<br />

phase = constants ( 1 ) ;<br />

switch phase<br />

984 case {1} % low temperature<br />

sb = 0 . 0 1 2 8 ; % f i t t e t to ( q , I ) of MLV sample<br />

S = gauss (q , q0 , sb ) + 0.22∗ gauss (q ,2∗ q0 , sb ) + 0.022∗ gauss (q ,3∗ q0 , sb ) ;<br />

case {2} % high temperature<br />

sb = 0 . 0 1 5 9 ; % f i t t e t to ( q , I ) of MLV sample<br />

989 S = gauss (q , q0 , sb ) + 0.3∗ gauss (q ,2∗ q0 , sb ) ;<br />

otherwise<br />

error ( ’ Phase ␣ parameter ␣must␣be␣ e i t h e r ␣1␣ ( low␣temp) ␣ or ␣2␣ ( high ␣temp) ’ ) ;<br />

end ;<br />

SMEARING.M<br />

% Returns the smeared i n t e n s i t y<br />

%<br />

994 % smearing ( I_func , q , parameters , s t e p s )


175<br />

%<br />

% I_func ~ I n t e n s i t y f uncti on from a point s c a t t e r e r model<br />

% q ~ s c a t t e r i n g v e c t o r<br />

% parameters ~ Parameters f o r the i n t e n s i t y f uncti on<br />

999 % s t e p s ~ The number of i n t e g r a t i o n s t e p s<br />

%<br />

f u n c t i o n I_smeared = smearing ( I_func , q , parameters , constants )<br />

1004 %%%%%%%%%%%%%%%%%%%%%%%<br />

%%% Load s e t t i n g s %%%<br />

%%%%%%%%%%%%%%%%%%%%%%%<br />

% Please read the r u c s a x s _ s e t t i n g s .m f i l e f o r information<br />

1009 % about each of the parameters<br />

[ q_pr_ch , q0 , bfn , bmt , ch_first , ch_last , sbx , dgpfn ] = rucsaxs_settings ;<br />

% " s t e p s " found by t r y i n g 2 ,4 ,8 ,16 ,32 ,10000<br />

% There was no d i f f e r e n c e between 32 and 10000 above q =0.025 (beam stop ) .<br />

1014 s t e p s = 32;<br />

% I n t e r v a l along x−a x i s<br />

x = −2∗sbx :4∗ sbx / s t e p s :2∗ sbx ;<br />

1019 % Beam p r o f i l e along x−a x i s<br />

Px = transpose ( gauss (x , 0 , sbx ) ) ;<br />

N = length ( q ) ;<br />

f o r i =1:N<br />

1024 I_p = f e v a l ( I_func , s q r t (q ( i ) . ^ 2 + x . ^ 2 ) , parameters , constants ) ’ ;<br />

I_smeared ( i ) = trapz (x , Px . ∗ I_p ) ;<br />

end<br />

FIT.M<br />

% Fit m u l t i p l e models to m u l t i p l e data s e t s .<br />

%<br />

% f uncti on r e s u l t s = f i t ( saxs_data , max_iterations ) ;<br />

1029 %<br />

% saxs_data ~ 1xN s t r u c t array<br />

% data : [ 3 x651 double ]<br />

% a l c o h o l : n<br />

% phase : n<br />

1034 % s e r i e s : n<br />

% modeling_results : [ 1 x1 s t r u c t ]<br />

%<br />

% Each model has the f o l l o w i n g s t r u c t u r e<br />

% modeling_results . ( model )<br />

1039 % parameters : [ Nx1 double ]<br />

% converged : 0 or 1<br />

% r e s i d u a l s : [561 x1 double ]<br />

% jacobian : [561xN double ]<br />

% sse : d<br />

1044 % i t e r a t i o n s : n


176 Matlab source code<br />

% time : d<br />

% constants : [Mx1 double ]<br />

% l a b e l s : {Nx1 c e l l }<br />

% l a t e x _ l a b e l s : {Nx1 c e l l }<br />

1049 %<br />

1054<br />

f u n c t i o n r e s u l t s = f i t ( saxs_data , max_iterations ) ;<br />

r e s u l t s = saxs_data ;<br />

% data s e t s<br />

N = length ( saxs_data ) ;<br />

% i t e r a t e data s e t s<br />

1059 f o r i =1:N<br />

% do nothing , i f there i s no data<br />

i f saxs_data ( i ) . data (1 ,1) == 0<br />

continue ;<br />

1064 end<br />

1069<br />

a l c o h o l = alcohol_str ( saxs_data ( i ) . a l c o h o l ) ;<br />

s e r i e s = saxs_data ( i ) . s e r i e s ;<br />

phase = phase_str ( saxs_data ( i ) . phase ) ;<br />

log_str = ’ \ nAlcohol : ␣%s , ␣Phase : ␣%s , ␣ S e r i e s :%3d ’ ;<br />

d isp ( s p r i n t f ( log_str , alcohol , phase , s e r i e s ) ) ;<br />

q = saxs_data ( i ) . data ( 1 , : ) ;<br />

1074 I = saxs_data ( i ) . data ( 2 , : ) ;<br />

model_names = fieldnames ( saxs_data ( i ) . modeling_results ) ;<br />

M = length ( model_names ) ;<br />

1079 % I t e r a t e models<br />

f o r j =1:M<br />

% current model<br />

model = saxs_data ( i ) . modeling_results . ( model_names { j } ) ;<br />

1084 % current model parameters<br />

parameters = model . parameters ;<br />

par_count = length ( parameters ) ;<br />

% do nothing i f there i s no parameters<br />

1089 i f parameters (1) == 0<br />

continue ;<br />

end<br />

% current model constants<br />

1094 constants = model . constants ;<br />

log_str = ’ F i t t i n g ␣ to ␣model : ␣%s ’ ;<br />

d isp ( s p r i n t f ( log_str , model_names { j } ) ) ;


177<br />

1099 % or i f the f i t t i n g already<br />

% has been done b e f o r e and has converged<br />

i f model . converged == 1<br />

log_str = ’ Modeling␣ has ␣ converged ␣ before , ␣ skipping ␣model . ’ ;<br />

d isp ( s p r i n t f ( log_str ) ) ;<br />

1104<br />

par_str = ’ ’ ;<br />

f o r k = 1 : par_count<br />

pattern = [ ’ ␣ ’ model . l a b e l s {k} ’=%.3g , ’ ] ;<br />

i f k == par_count<br />

1109 pattern = [ ’ ␣ ’ model . l a b e l s {k} ’=%.3g ’ ] ;<br />

end<br />

par_str = [ par_str s p r i n t f ( pattern , parameters ( k ) ) ] ;<br />

end<br />

d isp ( s p r i n t f ( ’SSE : ␣%.8g , ␣ Parameters : ␣[% s ] \ n ’ , model . sse , par_str ) ) ;<br />

1114<br />

continue ;<br />

end<br />

% parameter count<br />

1119 P = length ( parameters ) ;<br />

% g e t s t a r t time<br />

time_before = c l o c k ;<br />

1124 [ fitted_parameters , r e s i d u a l s , jacobian , i t e r a t i o n s , error_code ] = . . .<br />

n l i n f i t (q , I , model_names { j } , parameters , constants , max_iterations ) ;<br />

1129<br />

% g e t end time<br />

process_time = etime ( clock , time_before ) ;<br />

r e s u l t s ( i ) . modeling_results . ( model_names { j } ) . time = . . .<br />

r e s u l t s ( i ) . modeling_results . ( model_names{ j } ) . time + process_time ;<br />

r e s u l t s ( i ) . modeling_results . ( model_names { j } ) . r e s i d u a l s = r e s i d u a l s ;<br />

r e s u l t s ( i ) . modeling_results . ( model_names { j } ) . jacobian = jacobian ;<br />

1134 r e s u l t s ( i ) . modeling_results . ( model_names { j } ) . parameters = fitted_parameters ;<br />

r e s u l t s ( i ) . modeling_results . ( model_names { j } ) . s s e = transpose ( r e s i d u a l s )∗ r e s i d u a l s ;<br />

r e s u l t s ( i ) . modeling_results . ( model_names { j } ) . i t e r a t i o n s = . . . ;<br />

r e s u l t s ( i ) . modeling_results . ( model_names{ j } ) . i t e r a t i o n s + i t e r a t i o n s ;<br />

1139 i f error_code == 0<br />

r e s u l t s ( i ) . modeling_results . ( model_names{ j } ) . converged = 1 ;<br />

e l s e<br />

r e s u l t s ( i ) . modeling_results . ( model_names{ j } ) . converged = 0 ;<br />

end<br />

1144 end % models<br />

end % data<br />

NLINFIT.M<br />

% NLINFIT Nonlinear l e a s t −squares data f i t t i n g by the Gauss−Newton method .<br />

%<br />

% NLINFIT(X,Y, ’MODEL’ ,BETA0, maxiter ) f i n d s the c o e f f i c i e n t s of the nonlinear


178 Matlab source code<br />

% f uncti on d e s c r i b e d in MODEL. MODEL i s a user s u p p l i e d f uncti on having<br />

1149 % the form y = f ( beta , x ) . That i s MODEL returns the p r e d i c t e d values of y<br />

% given i n i t i a l parameter esti mates , beta , and the independent v a r i a b l e , X.<br />

% [BETA,R, J ] = NLINFIT(X,Y, ’MODEL’ ,BETA0) returns the f i t t e d c o e f f i c i e n t s<br />

% BETA the r e s i d u a l s , R, and the Jacobian , J , f o r use with NLINTOOL to<br />

% produce error esti mates on p r e d i c t i o n s .<br />

1154 %<br />

% B.A. Jones 12−06−94.<br />

% Copyright ( c ) 1993−98 by The MathWorks , Inc .<br />

% $Revision : 2.10 $ $Date : 1997/11/29 01:46:10 $<br />

%<br />

1159 % Modifies by Ulf øæRrbk Pedersen , may 2005<br />

% Modificati ons by Thomas Hecksher , january 2007<br />

%<br />

f u n c t i o n [ beta , r , J , iterations_used , error_code ] = . . .<br />

1164 n l i n f i t (X, y , model , beta0 , constants , maxiter )<br />

error_code = 0 ; % OK<br />

iterations_use d = 0 ;<br />

1169 n = length ( y ) ;<br />

i f min ( s i z e ( y ) ) ~= 1<br />

error ( ’ Requires ␣a␣ v ector ␣ second␣ input␣argument . ’ ) ;<br />

end<br />

y = y ( : ) ;<br />

1174<br />

p = length ( beta0 ) ;<br />

beta0 = beta0 ( : ) ;<br />

J = z e r o s (n , p ) ;<br />

1179 beta = beta0 ;<br />

betanew = beta + 1 ;<br />

i t e r = 0 ;<br />

b e t a t o l = 1.0E−4;<br />

r t o l = 1.0E−4;<br />

1184 s s e = 1 ;<br />

s s e o l d = s s e ;<br />

par_count = length ( beta ) ;<br />

while ( norm ( ( betanew−beta ) . / ( beta+s q r t ( eps ) ) ) > b e t a t o l | . . .<br />

1189 abs ( sseold −s s e )/( s s e+s q r t ( eps ) ) > r t o l ) & i t e r < maxiter<br />

i f i t e r > 0 ,<br />

beta = betanew ;<br />

end<br />

1194 i t e r = i t e r + 1 ;<br />

y f i t = normalize (X, transpose ( f e v a l ( ’ smearing ’ , model ,X, beta , constants ) ) ) ;<br />

r = y − y f i t ;<br />

s s e o l d = r ’ ∗ r ;<br />

1199 par_str = ’ ’ ;<br />

f o r i = 1 : par_count


179<br />

pattern = ’ ␣%.3g , ’ ;<br />

i f i == par_count<br />

pattern = ’ ␣%.3g ’ ;<br />

1204 end<br />

par_str = [ par_str s p r i n t f ( pattern , beta ( i ) ) ] ;<br />

end<br />

d isp ( s p r i n t f ( ’ I t e r ␣%4d , ␣SSE␣%.8g , ␣Par␣[% s ] ’ , i t e r , sseold , par_str ) ) ;<br />

1209 f o r k = 1 : p ,<br />

d e l t a = z e r o s ( s i z e ( beta ) ) ;<br />

d e l t a (k ) = s q r t ( eps )∗ beta ( k ) ;<br />

yplus = normalize (X, transpose ( f e v a l ( . . .<br />

’ smearing ’ , model ,X, beta+delta , constants ) ) ) ;<br />

1214 J ( : , k ) = ( yplus − y f i t )/( s q r t ( eps )∗ beta ( k ) ) ;<br />

end<br />

1219<br />

1224<br />

Jplus = [ J ; 1 . 0 E−2∗eye (p ) ] ;<br />

r p l u s = [ r ; z e r o s (p , 1 ) ] ;<br />

% Levenberg −Marquardt type adjustment<br />

% Gauss−Newton s t e p −> J\ r<br />

% LM s t e p −> inv (J ’ ∗ J+constant ∗ eye ( p ))∗ J ’ ∗ r<br />

step = Jplus \ r p l u s ;<br />

betanew = beta + step ;<br />

betanew = abs ( betanew ) ; % only accept p o s i t i v e parameter v a l u e s<br />

yfitnew = normalize (X, transpose ( f e v a l ( . . .<br />

1229 ’ smearing ’ , model ,X, betanew , constants ) ) ) ;<br />

rnew = y − yfitnew ;<br />

s s e = rnew ’ ∗ rnew ;<br />

i t e r 1 = 0 ;<br />

while s s e > s s e o l d & i t e r 1 < 12<br />

1234 step = step / s q r t ( 1 0 ) ;<br />

betanew = beta + step ;<br />

betanew = abs ( betanew ) ; % only accept p o s i t i v e parameter v a l u e s ;<br />

yfitnew = normalize (X, transpose ( f e v a l ( . . .<br />

’ smearing ’ , model ,X, betanew , constants ) ) ) ;<br />

1239 rnew = y − yfitnew ;<br />

s s e = rnew ’ ∗ rnew ;<br />

i t e r 1 = i t e r 1 + 1 ;<br />

end<br />

end<br />

1244<br />

iterations_used = i t e r ;<br />

i f<br />

1249 end<br />

i t e r == maxiter<br />

error_code = 1 ; % I n t e g r a l did not converge<br />

T_SET.M<br />

1249 f u n c t i o n out=T_set (T_abs ) ;<br />

% C a l c u l a t e s the temperature to be s e t from the d e s i r e d temperature .


180 Matlab source code<br />

% The c a l i b r a t i o n was done in August with a s i l i c o n e o i l i n s i d e the c u v e t t e .<br />

%<br />

% f uncti on T_set (T_abs )<br />

1254 % T_abs i s the d e s i r e d temperature ( Celcius ) .<br />

% T_set return the s e t temperature ( Celcius ) .<br />

%<br />

% example : T = T_set ( 3 0 ) ; % returns 32.4254<br />

1259 out= ( ( 2 7 3 . 1 5 + T_abs − 60.459) / 0.79421) − 2 7 3 . 1 5 ;<br />

T_ABS.M<br />

1259 f u n c t i o n out = T_abs( T_set ) ;<br />

% C a l c u l a t e s the a b s o l u t e temperature from the s e t temperature .<br />

% The c a l i b r a t i o n was done in August with a s i l i c o n e o i l i n s i d e the c u v e t t e .<br />

%<br />

% f uncti on T_abs( T_set )<br />

1264 % T_set i s the s e t temperature ( Celcius ) .<br />

% T_abs i s the returned a b s o l u t e temperature ( Celcius ) .<br />

%<br />

% example : T = T_abs ( 3 2 . 4 ) ; % returns ~30<br />

1269 out = 0.79421 ∗ ( T_set + 273.15) − 273.15 + 6 0 . 4 5 9 ;<br />

LIPID_WATER_MASS_RATIO.M<br />

1269 % out=lipid_water_mass_ratio ( numberOfDropTimeToUse )<br />

%<br />

% pans . dat : weight of emty t r a y s<br />

% drops . dat : vap . data op drops<br />

% l i p i d . dat : end weight of t r a y s<br />

1274 %<br />

% Example : [ wp wp_err]= lipid_water_mass_ratio (3)<br />

%<br />

1279<br />

f u n c t i o n out=lipid_water_mass_ratio ( numberOfDropTimeToUse )<br />

%%% Loading data<br />

pans=transpose ( load ( ’ pans . dat ’ ) ) ;<br />

drops=transpose ( load ( ’ drops . dat ’ ) ) ;<br />

l i p i d=transpose ( load ( ’ l i p i d . dat ’ ) ) ;<br />

1284 [ numberOfTrays , pansncol]= s i z e ( pans ) ;<br />

[ numberOfTrays , l i p i d n c o l ]= s i z e ( l i p i d ) ;<br />

[ numberOfTrays , numberOfDropTimes]= s i z e ( drops ) ;<br />

%%% Running through t r a y s<br />

1289 f o r i =1:numberOfTrays<br />

[ ’ ␣␣␣−−−␣ tray ␣number␣ ’ num2str ( i ) ’ ␣−−−’ ]<br />

%%% g e t t i n g tray weight<br />

pans_weight=mean( pans ( i , : ) ) ;<br />

1294 pans_weight_err=std ( pans ( i , : ) ) ;<br />

lipid_weight=mean( l i p i d ( i , : ) ) ;<br />

lipid_weight_err=std ( l i p i d ( i , : ) ) ,


181<br />

1299<br />

lw=lipid_weight −pans_weight ;<br />

lw_err=s q r t ( pans_weight_err^2+lipid_weight_err ^ 2 ) ;<br />

%%% g e t t i n g the weight from drops data<br />

c l e a r i n t ; i n t=numberOfDropTimes−numberOfDropTimeToUse+1:numberOfDropTimes ;<br />

x=i n t ;<br />

[ par , s s ]= p o l y f i t (x , drops ( i , i n t ) , 1 ) ;<br />

1304 drops_weight=par (2)− pans_weight ;<br />

1309<br />

%%% C a l c u l a t i n g wheight %<br />

wp( i )=lw/ drops_weight ;<br />

wp_err ( i )=lw_err/drops_weight ;<br />

%%% p l o t data<br />

f i g u r e ( 1 ) ; hold on ; p l o t ( pans ( i , : ) , ’ x−’ ) ; hold o f f<br />

f i g u r e ( 2 ) ; hold on ; p l o t ( l i p i d ( i , : ) , ’ x−’ ) ; hold o f f<br />

f i g u r e ( 3 ) ; hold on ; p l o t ( 1 : numberOfDropTimes , drops ( i , : ) , ’ o−’ , [ 0 x ] , . . .<br />

1314 p o l y v a l ( par , [ 0 x ] ) , ’ x : ’ ) ; hold o f f<br />

end<br />

%%% o v e r a l l weight %<br />

WP=mean(wp)<br />

1319 WP_err=std (wp)/ s q r t ( numberOfTrays)<br />

f i g u r e ( 4 ) ;<br />

hold on<br />

errorbar ( 1 : numberOfTrays , wp, wp_err ) ;<br />

p l o t ( [ 1 numberOfTrays ] ,WP∗[1 1 ] , ’−−’ ) ;<br />

1324 p l o t ( [ 1 numberOfTrays ] , (WP+WP_err ) ∗ [ 1 1 ] , ’ : ’ ) ;<br />

p l o t ( [ 1 numberOfTrays ] , (WP−WP_err ) ∗ [ 1 1 ] , ’ : ’ ) ;<br />

hold o f f<br />

1329<br />

out=[WP; WP_err ] ;<br />

%%% Saving data<br />

savedata =[WP WP_err ] ;<br />

s a v e f i l e =[ ’ kons . dat ’ ] ;<br />

d isp ( [ ’ Saving ␣ data ␣ to ␣ f i l e : ␣ ’ s a v e f i l e ] ) ;<br />

1334 save ( s a v e f i l e , ’ savedata ’ , ’−ASCII ’ ) ;<br />

ALCOHOL_LIPID_RATIO<br />

1334 f u n c t i o n out=a l c o h o l _ l i p i d _ r a t i o ( acl , ma, ms , lwr )<br />

% C a l c u l a t e s the alcohol −l i p i d r a t i o<br />

% (<strong>DPPC</strong> i s i mplied and the output has the same unit as ms i nput ) .<br />

%<br />

% f uncti on a l c o h o l _ l i p i d _ r a t i o ( acl , ma, ms, lwr )<br />

1339 % a c l − a l c o h o l chain l e n g t h<br />

% ma − mass a l c o h o l<br />

% ms − mass sample ( water + l i p i d )<br />

% lwr − l i p i d water r a t i o<br />

1344 % Assignments<br />

M_lipid = 7 3 4 . 1 ; % g/mol (<strong>DPPC</strong>)<br />

M_alc = M_alcohol ( a c l ) ;


182 Matlab source code<br />

% C a l c u l a t i o n<br />

1349 out = ( M_lipid ∗ ma ∗ (1 + lwr ) ) / (M_alc ∗ lwr ∗ ms ) ;<br />

% Output<br />

%disp ( s p r i n t f ([ ’%1.3 e ’ ] , out ) ) ;


Index<br />

Adiabatic Scanning Calorimeter 36<br />

Anaesthetic potency 10<br />

Asymmetric 1d model 85<br />

Atomic Form Factor 26<br />

Backgammon technique 44<br />

Background 68<br />

Baseline 60<br />

Bilayer 3, 4, 5, 6, 9, 10, 11, 12, 14<br />

Bilayer normal 6<br />

Bilayer thickness 7, 8, 93<br />

Biphasic effect 11<br />

Block collimation 42<br />

Bragg reflections 90<br />

Bragg’s Law 28<br />

Brilliance 18<br />

Centrifuge 49<br />

Coherence 19<br />

Compensating DSC 36<br />

Compton scattering 24, 29<br />

Compton scattering length 29<br />

Cuvette 43<br />

De Broglie wavelength 17<br />

Detector 41, 44<br />

Di-palmityol-Phosphatidyl-Choline (<strong>DPPC</strong>)<br />

5<br />

Differential Scanning Calorimetry (DSC) 36,<br />

59<br />

DSC 7 38<br />

Electron density profile 14, 79<br />

Electron propability 26<br />

Extruder 49, 51<br />

Gravimetric determination 49<br />

H II 90<br />

Hexagonal structure 90<br />

Hydrophobic effect 12<br />

Hydrophobic force 4<br />

Incoherence 19<br />

Isothermal Titration Calorimetry (ITC) 12,<br />

33, 55<br />

Kratky camera 41<br />

L α 6, 7, 9, 10, 12, 14<br />

L c 9<br />

L β ′ 9<br />

L βI 9, 14<br />

Laue Condition 27<br />

Lipid 3<br />

Longitudinal Coherence length 19<br />

Lorentz factor 82<br />

Main phase transition temperature 11, 59<br />

Microfurnace 38<br />

Miller indices 27<br />

Mole fractions 12<br />

Monochromator 42<br />

MSC-ITC 33<br />

Multilamellar vesicle (MLV) 4<br />

P β ′ 9, 10<br />

Pan (DSC) 39<br />

Partition coefficient 10, 12, 14, 33<br />

Phase diagram 10<br />

Phase transition 5, 6, 11<br />

Phospholipid 3, 4, 5, 7, 8, 9, 10<br />

Photoelectric absorption 17<br />

Photon 17<br />

Polydispersity 52<br />

Power compensating DSC 38<br />

Pretransition 59, 62<br />

q-calibration 45<br />

Sample preparation 49<br />

Scal-1 microcalorimeter 36<br />

Scanrate 61<br />

Scattering cross section 21<br />

Simulation 12<br />

Smearing 78, 79<br />

Solvent-null method 12, 14, 33<br />

Spherically symmetric 3d model 87<br />

Structure factor 89<br />

183


184 Index<br />

Symmetric 1d model 81<br />

Symmetric 4g model 84<br />

Synchrotron 18<br />

Temperature calibration 47<br />

Transition temperature 63<br />

Thermostatic shield 37<br />

Thomson scattering 24<br />

Thomson scattering length 23, 29<br />

Threshold concentration 12, 59<br />

Transition Enthalpy 62<br />

Transverse Coherence length 19<br />

Unilamellar vesicle (ULV) 4, 8<br />

Unit Cell Structure Factor 27<br />

van der Waals force 4<br />

VP-ITC 33<br />

X-ray source 41

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!