381
THE MOLECULAR ORGANIZATION OF ARTIFICIAL LIPID BILAYERS A STUDY OF THEIR DIELECTRIC IMPEDANCE AND SURFACE TENSION by DEREK ROWLAND LAVER A thesis submitted for the degree of DOCTOR OF PHILOSOPHY in the faculty of science of THE UNIVERSITY OF NEW SOUTH WALES 1983

the molecular organization of artificial lipid bilayers a study of

Embed Size (px)

Citation preview

THE MOLECULAR ORGANIZATION OF ARTIFICIAL LIPID BILAYERS

A STUDY OF THEIR DIELECTRIC IMPEDANCE AND SURFACE TENSION

by

DEREK ROWLAND LAVER

A thesis submitted for the degree of

DOCTOR OF PHILOSOPHY

in the faculty of science of

THE UNIVERSITY OF NEW SOUTH WALES

1983

This is to certify that the work embodied in this thesis has not been

previously submitted for the award of a degree in any institution.

• • • •••••••

FEBRUARY, 1983

DEDICATION

To my wife, Deborah

CONTENTS

Acknowledgments ( ; )

Collaboration ( ; ; ; )

Abstract ( i V)

Chapter Index (vii)

Glossary of Major Symbols ( X i )

Appendix A 235

References 240

Publications 247

ACKNOWLEDGEMENTS

My sincerest thanks to Professor Hans Coster for his

guidance, inspiration and constant availability during my years at

laboratory. Professor Hans Coster has been a great supervisor.

sound

his

Dr Robert Ashcroft played a significant part in my "formative"

years of scientific research. I found working with him both enjoyable

and inspiring.

I wish to thank Dr John Smith who has been a good friend and has

shown continuous interest in my welfare over my the years of my research

here.

Dr Joe Wolfe in the last year of my work here has not only helped

me become more aware of the scientific community but has been a willing

source of friendship and advice. He certainly makes "physics fun".

My thanks to Terry Chilcott who has always been willing to share

his technical experience. The quality of his workmanship in designing

and building the impedance measuring system has certainly made

experimental work much less frustrating.

My thanks to David Knowles for lightening my experimental load and

sharing the joys and frustrations of research.

Sue Murray-Jones helped considerably with the proof reading of this

thesis. Sue has been constantly available for organizing those

; i

hard-to-do things and has certainly contributed to the happy atmosphere

in the lab.

I would like to thank Dr Joe D'Arrigo and Dr David Gruen for their

helpful discussions.

Mr Jack Sandall working in the mechanical workshop did a great job

on constructing the membrane supporting apparatus.

Kim Crestani, a student of architecture, removed the burden of

drawing the diagrams. Jane Campbell produced the photographs presented

in this thesis. The quality of their work is certainly a high point in

the presentation of this thesis.

Also wish to thank Toni Benton for accurately (and quickly) typing

the equations and corrections and Rev. Paul Bayliss and Judy Leeds for

proof reading the manuscript.

To my wife, Debbie, who has contributed to much of the typing,

proof reading and funding of this thesis and without whom I would surely

have gone crazy. What can I say?

i i i

COLLABORATION

Several portions of the work described in this thesis was done jointly

with other people.

Dr Hans Coster and Terry Chillcot designed and built the impedance

measuring system used in the experiments described in this thesis.

The software controlling the impedance measuring system was written by

Terry Chil lcot.

The software handling the data and its analysis was written jointly with

Terry Chi llcot.

The discussion on the location of cholesterol in lipid bilayers was

greatly assisted by Dr Robert Ashcroft and Dr John Smith.

The project described in Chapter 9 was a collaborative effort with Dr

John Smith.

The experiments on bilayer tension were done jointly with David Knowles,

who was undertaking an honours project in this topic.

iv

ABSTRACT

The impedance of artificial BLM were measured using the four-terminal

digital technique of Bell, Coster and Smith (1975). Phase angle was

measured to an accuracy of .02°, magnitude to .1% for the frequencies

.003 to 10000 Hz.

Experiments were carried out on bilayers formed using a modified version

of the film drainage method of Mueller, Rudin, Tien and Wescott (1962)

and from solutions containing either egg-lecithin, lecithin -

cholesterol or glycerol monooleate dissolved in one of the n-alkanes

(n-decane to n-hexadecane) or squalene.

Absorption of n-alkanes into lipid bilayers reduced their capacitance

(measured at 1 Hz). This was interpreted as a change in thickness, and

capacitance per unit area was used thus to calculate the area density of

alkane in the membranes.

From the dependence of capacitance on temperature, it was deduced that

egg-lecithin bilayers formed from n-hexadecane solutions of the lipid 0 had negligible alkane concentrations at temperatures less than 30 C.

The alkane concentration in the bilayer increased with increasing I

temperature, decreased with alkane chainlength and was affected by the

composition of the membrane forming solution and the external aqueous

phase.

V

These results were interpreted in terms of a successful statistical

model of n-alkane - lipid bilayers in the liquid crystalline state. The

results obtained were consistent with n-alkane absorption being very

sensitive to the ordering of the acyl chains of the lipids. The

partitioning of n-alkanes into the bilayer, rather than being considered

as a pitfall of the model, was used to probe changes in the ordering of

the lipid acyl chains.

Increasing ion concentration in the external aqueous phase increased the

capacitance of the ionic double layer external to the membrane. The

time-constant of the ionic double layer was found to be equal r~-.,."-

to i. the 1 ~

that of the bilayer interior. The results were consistent with the

presence of a sma11Jof bound surface charge on the bilayer which was

interpreted as ion absorption and was described by the Langmuir

absorption isotherm.

The capacitance and conductance of membranes varied with frequency. The

impedance spectrum so obtained was modelled as a symmetric 4 to 6

layered dielectric structure. Several electrical time-constants were

resolved and attributed to the acetyl and polar head regions of the

bilayer. These analyses implied:

1) A conduction mechanism other than ion migration was operating in

the polar head regions of the bilayer.

I

2) The inclusion of cholesterol in egg-lecithin bilayers altered

the dielectric structure of the acetyl region which indicated the

location of cholesterol in the bilayer. Oxidised and non-oxidised

cholesterol had markedly different effects on the ionic double

layers external to the membrane.

vi

3) The replacement of H20 by 020 in the aqueous phase had no effect

on bilayer structure.

4) Alkane absorption in the hydrophobic region had no detectable

effect on the polar head dielectric substructure.

5) Procaine and benzocaine altered the dielectric structure of the

polar regions of the bilayer. The location of procaine in the

bilayer was in agreement with previous studies.

The activation energy of electrical conduction of bilayer membranes was

independent of membrane conductance over the range .01-l mS/m2.

Electrical conduction was attributed to ionic movement through

sub-microscopic aqueous channels spanning the bilayer.

The bilayer tension was measured using the technique of Coster and

Simons (1968). The bilayer tension of egg-lecithin bilayers decreased

with increasing temperature, whereas that for glycerol monooleate

increased with increasing temperature. This was attributed to

differences in the aggregation of the lipids in the torus component of

the membrane.

Procaine, butanol and pentanol decreased bilayer tension at

anaesthetising concentrations in the aqueous phase.

vii INDEX

Page CHAPTER l - REVIEW

1.1 Introduction 2

1.2 Organization and Function of Biological Membranes 2

1.3 The Physical Properties of Artificial 7 Planar Bimolecular Lipid Membranes (BLM)

1.4 The Dielectric Properties of Artificial BLM 18

CHAPTER 2 - THE DIELECTRIC MODEL OF ARTIFICIAL BIMOLECULAR LIPID MEMBRANES

2. l Introduction 24

2.2 Hydrophobic Region 25

2.3 Polar Regions 32

2.4 Aqueous Phase 34

2.5 Philosophy of Measurement 38

CHAPTER 3 - APPARATUS

3. l Introduction 44

3.2 Membrane Apparatus 46

3.3 Amplifier Assembly 49

3.4 Computor Control 52

CHAPTER 4 - MATERIALS AND METHODS

4. l Introduction 56

4.2 Materials 56

4.3 Making Bilay!:!rS 59 I

4.4 Calibration and Performance of Apparatus 61

4.5 Data Analysis, Reduction and Presentation 67

CHAPTER 5 - LIPID - ALKANE INTERACTIONS IN ARTIFICIAL BLM

5. l Introduction

5.2 Theoretical Considerations

5.3 Materials and Methods

5.4 Results

5.5 Discussion

5.6 Summary

CHAPTER 6 - EFFECT OF EXTERNAL ELECTROLYTE ON THE CAPACITANCE OF LIPID BILAYERS

6. l Introduction

6.2 Preliminary Theoretical Considerations

6.3 Methods

6.4 Results

6.5 Discussion

6.6 Summary

75

79

85

86

89

101

104

107

111

111

114

123

CHAPTER 7 - THE DIELECTRIC STRUCTURE OF THE HYDROPHOBIC-HYDROPHILIC INTERFACE OF EGG-LECITHIN AND GMO BILAYERS

7.1 Introduction

7.2 Methods

7. 3 Results

7.4 Discussion

7.5 Summary

CHAPTER 8 - THE DIELECTRIC STRUCTURE OF ARTIFICIAL BLM I THE EFFECT OF CHOLESTEROL AND n-ALKANE INCLUSION

127

128

129

135

144

II THE EFFECT OF D20/H20 REPLACEMENT IN THE AQUEOUS PHASE

8.1 Introduction

8.2 Methods

8.3 Results

147

150

150

Viii

ix

8.4 Discussion 152

8.5 Summary 156

CHAPTER 9 - THE CONDUCTANCE OF LIPID BILAYERS

9. l Introduction 159

9.2 Materials and Methods 161

9.3 Results 162

9.4 Discussion 167

9.5 Summary 179

CHAPTER 10 - ENERGY OF FORMATION OF LIPID BILAYERS

l O. l Introduction 182

10.2 Theoretical Considerations 185

10.3 Methods 187

10.4 Results 190

10.5 Discussion 192

10.6 Summary 199

CHAPTER 11 - THE EFFECT OF SOME LOCAL ANAESTHETICS ON THE PHYSICAL PROPERTIES OF EGG-LECITHIN BLM

11. l Introduction 202

11. 2 Materials and Methods 204

11. 3 Results 205

11.4 Discussion 209

11. 5 Summary 219

CHAPTER 12 - SUMMARY

12. l Summary 223

12.2 Suggestions for Further Work 233

X

APPENDIX A: DATA STROAGE AND PRESENTATION 235

GLOSSARY OF SYMBOLS xi

a) Alphabetical Symbols

SYMBOL

A

a

a

A.C.

ADC

b

BLM

BZA

BULFIS

C

Cl and C2

PAGE DESCRIPTION

18 area

25 ionic radius

186 head group area in the plane of the bilayer (Chapter 10 only)

19 alternating current

61 ratio of the gain response of voltage and "current" amplifiers

188 area of planar bilayer

62 amplitude ratio of signals arr1v1ng at both input channels of BULFIS

42 amplitude ratio of the voltage across the membrane and series impedance standards

53 analog to digital convertor

172 pore diameter or hydrated ion diameter

1 bimolecular lipid membranes

75 benzyl alcohol

44 Biophysics Ultra Low Impedance Spectrometer

39 capacitance

63 input capacitance of differential amplifiers l and 2 respectively

87 n-tetradecane

87 n-hexadecane

79 area specific n-alkane molar concentration in the bilayer

18 dielectric capacitance of the membrane

c. l

C' m

DAC

DFCM

E

xii 20 capacitance of the ionic double layer

109 binding-ion concentration in "bulk"aqueous phase

27 ion concentration of "ith" species

79 area specific molar concentration of acyl chains in the bilayer "leak"

80 total area specific capacitance of solventless bilayers

80 total area specific capacitance of bilayers containing n-alkanes

38 capacitance of a GC network

188 capacitance of planar bilayer

35 ion concentration in the "bulk" aqueous phase

175 ion concentration in an aqueous pore

35 displacement field

175 ion diffusion coefficient in an aqueous pore

18 direct current

28 ion diffusion coefficient

52 digital to analog convertor

32 double fixed charge membrane

25 electric field

170 activation energy for conduction

186 energy of elastic deformation

187 energy of formation

6 excitation - or excitability inducing modules

39 conductance

38 parallel resistor-capacitor network

28 membrane conductance due to ion species i

131 bi lay.er conductivity due to aqueous pores \

38 conductance of the "Nth" GC network

175 single pore conductance

8 glycerol monooleate

h

I

J. 1

K

k

M

N

n. 1

NFP

NMR

q

R

R

r

Rl and R2

RAM

so

Sl and S2

T

lll

xiii

26 distance from an interface

83 enthalpy of the lipid acyl chains

29 total current

28 electrical current carried by an ion species i

82 partition coefficient of n-alkanes into the bilayer

20 Boltzmann constant

186 area-elastic modulus

142 partition coefficient between the aqueous phase and the polar heads

79 molecular weight

185 micelle aggregation number

78 molecular order parameter of the "ith" carbon atom of the lipid acyl chains

54 normalised fit parameter for raw data

16 nuclear magnetic resonance

189 net hydrostatic pressure

20 electronic charge

82 molar gas constant

188 radius of curvature of bowed membrane (Chapter 10 only)

25 radius vector

63 resistors used for calibrating the differential amplifiers

52 Random Access Memory

83 internal entropy of the lipid acyl chains

189 surface area of water compartments

20 absolute temperature

25 electrostatic self energy of an ion in a medi~m

172 energy of hydration

172 interfacial energy

172 total energy difference

28 externally applied potential

xiv

vm 25 volume

Vo 38 steady state voltage

w 39 angular frequency of the A.C. signal

x2 70 statistical quality of fit parameter

xa 79 alkane mole fraction in the bilayer

x. 27 thermodynamic force driving ion diffusion l

XN 185 Micelle aggregation number

z 40 magnitude of impedance

z 36 ion valency

ZR 67 ratio of impedance of membrane to that of the impedance standard

zs -42 magnitude of the standard impedance

b) Greek Symbols

Cl

y

£ m

£ 0

£ r £ w

e

µ. l

0 µ B

0 µ . l

173 geometrical constant

16 monolayer tension

28 ion partition coefficient

16 surface tension of the oil-water interface

18 membrane thickness

18 Dielectric constant of the hydrophobic region of lipid membranes

20 permittivity of vacuum

22 dielectric constant of a medium

20 dielectric constant of the aqueous phase

78 angle spanned by the C-H bonds of the hydrocarbon chain and the axis perpendicular to the bilayer

20 Debye length in an electrolyte

28 ionic mobility

185 standard chemical potential of lipids in the bilayer

27 standard chemical potential of an ion species, i

-µ. l

0 6µ

0 6µ

V m

0

Orn

T

• 1 and•" D D

185 standard chemical potential of lipids in inverted micelles

27 electrochemical potential of an ion species, i

82 standard chemical potential difference between alkane in the bilayer and torus (Chapter 5 only)

186 standard chemical potential difference between lipids in the torus and the bilayer (Chapter 10 only)

189 volume added to aqueous solution

189 volume displaced by bowed membrane

18 lateral pressure

16 lateral pressure of the acyl chains

16 lateral pressure of the polar heads

30 net charge concentration

37 concentration of bound charge

109 number of possible ion binding sites

142 summation

38 electrical time-constant

40 phase angle of impedance

61 difference between the phase responses of the amplifiers

xv

62 phase difference between signals arriving at both input channels of BULFIS

42 phase difference between A.C. signals across the membrane and series impedance standards

42 phase angle of the standard impedance

27 electrostatic potential

20 membrane surface potential in the absence of an externally applied field

CHAPTER l

REVIEW

1.1 INTRODUCTION

1.2 ORGANIZATION AND FUNCTION OF BIOLOGICAL MEMBRANES

1.3 THE PHYSICAL PROPERTIES OF ARTIFICIAL PLANAR BIMOLECULAR LIPID MEMBRANES (BLM)

1. 31 Reconstitution of Cell Membrane Structure

1.32 BLM Formation and Stability

1.321 Forces Involved in BLM Formation

1.322 The Role of Solvent in Membrane Stability

1.33 The Ultra Structure of Lipid Membranes

1.34 The Pitfalls of Solvent Retention in Artificial BLM

1.35 Molecular Interactions in BLM

1.4 THE ELECTRICAL PROPERTIES OF ARTIFICIAL BLM

1.41 D.C. Characteristics

1.42 A.C. Characteristics

Page

2

2

7

7

9

11

12

13

14

18

18

19

2

1.1 INTRODUCTION

This thesis is mainly concerned with the measurement of the

dielectric properties of planar lipid membranes formed artificially

using the film drainage technique of Mueller, Rudin, Tien and Wescott,

(1962). Part of this work is also concerned with the measurement of the

surface energy of lipid bilayers.

In this chapter a brief outline of the current picture of cell

membranes is presented as well as some structural aspects of artificial

BLM. A major part of this thesis is concerned with the dielectric

properties of BLM. A detailed description of this aspect of BLM is

given in Chapter Two.

1.2 ORGANIZATION AND FUNCTION OF BIOLOGICAL MEMBRANES

A common feature of all living cells is the presence of membrane

structures which delineate the cell from its environment and which

envelop internal organelles. These membranes vary in thickness from 5

to 10 nm.

tell membranes are composed of two broad classes of compounds:

proteins which exist in either extended or globular conformations and

lipids which are amphiphilic molecules which form a bilayer structure.

The total area of the lipid bilayer varies in the range 70-90% of the

total membrane area (Coleman, 1973). The wide range of membrane

3

activity is mainly derived from the protein components whereas the lipid

component provides the supporting matrix and diffusion barrier of the

membrane in which the proteins are either adsorbed or embedded.

The lipids in cell membranes may be divided into three classes:

phospholipids, steroids, and glycolipids. Of these the phospholipids

are the most common type in cell membranes. The phospholipids are

mainly diacyl phosphoglycerides which, depending on the unsaturation of

the acyl chains, are flexible molecules which are either electrically

charged or neutral and have acyl chainlengths in the range 16-24. The

steroids are rigid molecules which when present in bio-membranes act to

reduce membrane fluidity. The most common steroid in the membrane of

mammalian cells is cholesterol. The glycolipids are believed to

function as binding sites for bacterial antigens in cell membranes.

The composition and function of biological membranes varies

considerably between different cells and organelles. For example the

plasma membrane of Schwan cells, which form the myelin sheath around

some nerves, contain only 20% (w/w) protein whereas the membrane of

Halobacterium halobium has 75% (w/w) protein. Typically cell membranes

are composed of 40-60% (w/w) protein.

The current picture of the molecular organization of biological

membranes is based on the fluid mosaic model popularised by Singer and

Nicolson (1972), in which the membrane is considered as a two

dimensional oriented viscous solution of proteins and lipids (see figure

l. l).

On the whole the membrane structure is quite fluid; the lipid and

protein components being free to undergo rapid rotation and diffusion in

Figure 1.1. A schematic drawing summar1s1ng the main features of the fluid mosaic model as envisaged by Israelachvili (1977). From lsraelachvili (1977).

4

the plane of the membrane (Edidin, 1974 and Pope and Cornell, 1978).

However:, the membrane is heterogeneous and "islands" of restricted

mobility may exist which have been associated with the occurrence

of phase separations of membrane components (Edidin, 1974 and Nicolson,

1976).

The stability and organisation of the fluid mosaic structure

of cell membranes chiefly depends on the hydrophobic - hydrophilic

forces between the amphiphilic membrane components and the aqueous

phase (Singer and Nicolson, 1972). The hydrophobic portions of the

membrane components are confined to the membrane interior and the

hydrophilipic parts are in co·ntact with the water. This imposes certain

constraints on the molecular packing of the different membrane components.

If, for example, the polar groups of the lipids have a large cross­

sect i ona 1 area in the p 1 ane of the bilayer re 1 at i ve to that of the

hydrophobic portion of the molecule, then a curved structure would

result (Israelachvili, Mitchell and Ninham, 1976). Further, curved

bilayers would be associated with asymmetric distributions of lipids

across the membrane (Carnie, Israelachvili and Pailthorpe, 1979).

In multicomponent systems like biological membranes the different

components would assemble such that the internal energy of the entire

aggregate is minimized. Figure 1.2 demonstrates how phase separations

of different lipid and protein components could occur in a cell membrane

as a result of different packing constraints imposed on the system.

I • The presence of rigid irregular shaped molecules such as proteins

can produce high energy conformations in the flexible fatty acid chains

of the lipids in the adjacent lipid phase (see fig. l.3). In order

to reduce the internal energy of the system the bilayer will distort in the

{1!) lnt~rmixing (c) Pore

~ .. (~ ! : i : :> >-} -~ •, t I t I I , I 1 , .. , J

,-1--'-,rA-\..)....{.,,·Ji .'-""1--l.,.~r r+ t \ r ' , 1 r :- , . - " I I t t I

~' , ·· t-,._r,-.;; .ii ~ ,._

(d) Integral protein (e) Peripheral protein

Cytochrome oxidase Cytochrome c

Figure l .2. This diagram illustrates how packing constraints in a multicomponent system can lead to phase separations (eg. (d) and (e)) and conversely how phase separations of different membrane components can lead to structural changes in the membrane (eg. (b) and (c)). From Israelachvili et al. (1980).

a

b

d

hydroph1l1c surf ace

hydrophobic surface

/ void region

C

e

Figure 1.3. This diagram illustrates how packing constraints on membrane components can lead to distortion of the bilayer component of the membrane (above) which in turn can lead to protein-protein int2ractions via the lipid bilayer (opposite). From Israelachvili 1977.

5

vicinity of the protein (Israelachvili, 1977, 1978). Differences in the

molecular packing of lipids near rigid proteins can produce environments

hostile to the presence of some lipids and not others.

Separate lipid phases in these boundary regions do exist and have been

detected in spin label probe experiments (eg. Stier and Sackmann, 1973

and Warren, Houslay, Metcalfe and Birdsall, 1975). Perturbations in the

bilayer due to protein inclusion have been found to extend up to 4 lipid

diameters from the protein; thus affecting the conformation of about 140

lipid molecules (Curatolo, Sakura, Small and Shipley, 1977). The

presence of lipid boundary regions can also introduce forces between

adjacent proteins not only in the same manner as meniscus forces are

generated at liquid surfaces but also via entropic forces (Marcelja,

1976). Thus it is easy to visualize how changes in the molecular

organization of the lipid bilayer can regulate enzyme activity in

biological membranes and hence alter membrane function (Sandermann,

1978).

In recent years it has been increasingly realized that the lipid

bilayer plays an important role in the functioning of biological

membranes. Membrane bound enzymes suffer a loss of activity when

removed from their membrane environment; the activity only being

restored upon reconstituting the enzyme with a lipid phase of similar

physical properties (Caffrey and Feigenson, 1981). The activity of

membrane-bound enzymes in reconstituted membranes is dependent on the

composition of the lipid bilayer phase (Coleman, 1973) indica\ing that j

there are certain structural requirements for enzyme activity. As yet

no specific lipid requirement has been convincingly demonstrated

(Sandermann, 1978). So it seems that the role of the lipid bilayer is

mainly to provide an environment sympathetic to the function of proteins

(Israelachvili, Marcelja and Horn 1980). The function of membrane-bound

6

polypeptides is affected by the mobility of the fatty acid carbon chains

of the lipids as well as the state of the bilayer interface (eg. the

degree of hydration, surface charge, hydrogen bonding etc.) (Sandermann,

1978).

The structure and function of cell membranes is sensitive to

changes in their internal end external environment (Nicolson, 1976).

Phase transitions and phase separations of membrane components can be

triggered by the binding of divalent cations, changes in temperature and

pH (Jacobson and Papahadjopolous, 1975), changes in the ionic strength

(MacDonald, Simon and Baer, 1976) or the presence of foreign corapounds

such as anaesthetics (Lee, 1978), cholesterol or proteins (Birrel and

Griffith, 1976). The close proximity of another membrane is known to

alter the organization of proteins in thylakoid membranes (Staehelin and

Arntzen, 1979). The various mechanisms for these responses to

environmental changes on membrane structure were considered by

Israelachvili (1978).

The presence of raembrane-soluble raolecules such as short chain

n-alkanes inactivate excitation inducing modules (EIM) in nerve axon.

Haydon, Hendry, Levinson and Requena (1977) correlated the inactivation

of the EIM with changes in the structure of the lipid bilayer induced by

the alkane molecules. Many studies have correlated the potency of

membrane-modifying drugs to their hydrophobicity and hence postulate

that the site of action of these drugs is the lipid bilayer rather than

specific proteins. To examine the validity of this hypothesis part of

this thesis will investigate the effect of membrane soluble drugs such

as n-alkanes, alkanols and aminobenzoic acid esters on the properties

of lipids bilayers.

7

1.3 THE PHYSICAL PROPERTIES OF ARTIFICIAL PLANAR BIMOLECULAR LIPID

MEMBRANES (BLM)

1.31 Reconstitution of Cell Membrane Structure

The study of the cell membrane structure reconstituted in vitro has

begun only recently with the work of Mueller, Rudin, Tien and Wescott

(1962). Though the significance of such studies was realized much

earlier by Langmuir and Waugh (1938). The large area and planar

geometry of these structures allowed easy access to the aqueous phases

on both sides of the membrane and thus proved to be convenient membrane

models in mechanical, electrical and permeability studies.

Basically the technique of Mueller et al. (1962) was to disperse

the lipid components of the membrane in a hydrophobic solvent and form a

film of this solution across a circular support submerged in an aqueous

solution. The surface active lipid components form monolayers at each

oil-water interface. The hydrophobic solution drains away from between

the two monolayers allowing them to form a bilayer (see figure 1.4).

The formation of the bilayer is spontaneous and is usually complete

within 5-30 minutes. The film drainage technique of Mueller et al.

(1962) is the most common technique used to generate lipid bilayer

membranes, though modifications on this basic process are many and

varied depending on the specific study at hand.

Takagi et al (1965) and also more recently Montal and Mueller il972)

formed asymmetric lipid bilayers by apposing two monolayers across a

circular support. This was done by spreading a monolayer of two

different lipid mixtures at the air-water interface of two separated

compartments. By alternately raising the water levels of each

AQUEOUS PHASE

)

Figure 1.4. The formation of a lipid bilayer from apposing lipid monolayers at two oil-water interfaces. The arrows indicate the relative magnitude of the compressive Van der Waals force (between the adjacent water phases) on the lipid film. The action the this force tends to squeeze the solvent from between the monolayers.

8

compartment past the level of a circular hole in the septum both

monolayers were apposed in a bilayer structure (see figure 1.5).

Generally this technique can only be used to generate lipid bilayers of

relatively small area.

Bilayers have been formed from a wide variety of lipid mixtures

~ome examples are given in Goldup, Ohki and Danielli, 1970); the

main lipids being mixtures extracted from bovine brain and eggs as well

as isolated lipids such as phosphatidyl ethanolamine, sphingomyelin and

phosphatidyl serine. In more recent work there has been a move away

from natural lipids to the study of bilayers formed from mono- and

diglycerides which are rarely found in biological membranes. The most

coramonly used monoglyceride is glycerol monooleate (GMO). The advantage

of using monoglycerides is that the chemical composition is well

defined, they form stable bilayers and are fairly soluble in hydrophobic

solvents in comparison to natural lipids. Considerable insight into

amphiphile interactions in lipid bilayers have been derived from studies

on bilayers formed from this class of compounds.

Another molecule coramonly used in bilayer studies is cholesterol.

Pure cholesterol does not form bilayers but forms stable bilayers with

mixtures of other lipids. Oxidised cholesterol, though not chemically

well characterised, has proved convenient in artificial bilayer studies

as it enhances bilayer stability. Much of the data presented in this

thesis has been obtained from egg-lecithin bilayers containing oxidised l

cholesterol. The term oxidised refers to the fact that preparation of

this substance involves bubbling oxygen through cholesterol-containing

solutions (see Tien and Dawidowizc, 1966). It has been found that

oxidised cholesterol is a mixture of different steroids; some of which

have been isolated (Feiser and Feiser, 1959). Some of the known

Ll l.J .:-,.,,·t~t'. ........ ''J I_ ! ;"

I ' ~\ ... 1· L:: ~1 .._· -· ·

' ANNU LUS (l 0 _,

F ---...:.;:,·

AIR 50LVEN f f r- )

a i 1

SOLVENT'\:

1-~TER ~ ! f

\___ / ~ 1 : ... l

lLR\ ··t111 mr\~ iw.7\ ~ ; ~ . ~ ~ '.

R. ~ J ~ :> •

Figure 1.5. Th.e process of forming a asymmetric bilayers using the method of mbnolayer apposition first used by f•lontal and f,iueller (1972). The apposing monolayers are positioned across the aperture by raising the water levels in each compartment. From White et al . (1976).

9

oxidation products of cholesterol have been found to exist in living

tissue and therefore it is not unreasonable to use this steroid mixture

in artificial BLM to model the bilayer component of living membranes

(Tien, Carbone and Dawidowicz, 1966).

Most commonly used solvents in bilayer generation in earlier

bilayer studies were mixtures of hydrocarbons (eg. n-decane or

n-tetradecane) with chloroform - methanol solutions or mineral oil

mixtures. The recent trend is toward more well defined systems which

only contain a single component solvent of which the most common are

compounds from the n-alkane series between n-decane and n-hexadecane.

Squalene is a common solvent used in forming bilayers of GMO as it does

not, apparently, affect membrane structure (Simon, Lis, MacDonald and

Kauffman, 1977 and White, 1978).

1.32 Formation and Stability of BLM

1.321 Forces Involved in BLM Formation

Lipid molecules become aligned in monolayer aggregates at each

oil-water interface of the film in order to minimise the interfacial

energy arising from the Born repulsive forces between the hydrophobic

acyl chains of the lipids and the water molecules and the charged groups

in the polar heads of adjacent lipids (see figure 1.4 and 1.6). The

film then thins over a period of a few minutes as the lipid solution

drains away under the action of its buoyancy in water.

The transition from the thick film (approximately lOum thick) to

the bilayer state occurs rapidly and the mechanism for this is not well

understood. The spontaneous transition from the thick film to the

:/> / •·. ,, .

·--:'

/

·/ /

/

Figure 1.6. The formation of a symmetrical bilayer using the film drainage method of Mueller et al. (1972). From White et al. (1976).

10

bilayer indicates that the bilayer is a lower energy structure. It was

pointed out by Danielli (1966), that the energy of the oil-lipid

interface of the thick film could contribute significantly to the total

energy of the film. The bilayer would not have such an interface and

therefore should have a lower energy. More recent measurements of the

contact angles between thick films and bilayers (Pagano and Thompson,

1967 and Andrews et al., 1970) indicate that the energy difference

between the thick film and the bilayer is only .1% of the energy of

formation. It was then postulated that for small film thicknesses that

the Van der Waals attractive forces between adjacent water phases is a

small but significant compressive force on the bilayer which accelerates

the removal of solvent from between apposing monolayers. The size of

this attractive Van Der Waals force has been calculated by Ninham and

Parsegian (1970) and was found to rapidly increase with decreasing

membrane thickness. The magnitude of this compressive force is shown in

table l. 1. This positive feed-back situation leads to a squeezing of

the remaining bulk solution between the monolayers in a zipper action.

Another possible contributing factor to the bilayer formation could

be the partial ordering of the alkane molecules in the partly thinned

bilayer compared with the molecules in the torus. The initial formation

of bilayer is possibly initiated by random thermal fluctuations in

membrane thickness (Tien and Dawidowicz, 1966).

The main force opposing the continued thinning of lipid bilayers is

the short range repulsive forces between the atomic orbitals of adjacent

molecules causing the lipids in the bilayer to behave like "hard discs"

of a finite radius. The osmotic and viscous drag forces tend to impede

the flow of hydrophobic solvent out of the bilayer during thinning.

TABLE l. l

VAN DER WAALS FORCE BETWEEN WATER LAYERS SEPARATED BY A HYDROCARBON FILM

FILM THICKNESS,o (nm)

1000

100

10

5

VAN DER WAALS FORCE (N/m 2 )

2. 3 . 10- I

3.1 . 10 2

2.5 . 10'

3.4. 10 5

The transverse pressure on the bilayer ar1s1ng from the Van der Waals attractive forces between the water phases on either side of the membrane. The table is taken from Ashcroft (1979).

11

1.322 The Role of Solvent in Membrane Stability

The role of solvents in the formation of bilayer membranes was

mainly thought to provide an oil-water interface at which the small

concentration of lipids in the oil phase can adsorb. However, it has

become quite apparent that the solvent also plays a role in membrane

stability. Solvents that are water soluble or volatile such as

chloroform or short chainlength n-alkanes are not wholly confined to the

interior of the bilayer. Bilayers formed from such solvents become

unstable when the solvent concentration in the membrane becomes small.

It was postulated that the solvent molecules act as a "filler" of

structural defects present in lipid bilayers (see Fettiplace et

al., 1975).

BilayerJormation using the monolayer apposition technique only

requires solvents during the formation of the monolayers at each water

air interface. In principal this technique requires no solvents during

the formation of the bilayer. However, it was found that stable

bilayers would only form when small amounts of solvent were present in

the monolayers or when the membrane support was precoated with a

hydrophobic grease {Benz, Frohlich, LaUger and Montal, 1975). An

explanation for this was offered by White, Petersen, Simon and Yafuso

(1976) whereby a 11 bulk 11 lipid solution at the bilayer septum border was

needed to minimise the surface free energy of the lipid film septum

arrangement., Thus the solvent molecules act as a "filler" in the I

transition region between the bilayer and the much thicker septum.

12

1.33 The Ultra Structure of Lipid Membranes

Bilayer membranes formed using the film drainage method of Mueller

et al. (1962) have four main components (ie. lipid, solvent, water,

ions) which are present is three separate phases. These are:

a) the bilayer phase consisting of two apposing monolayers of lipid

which bridges an aperture in the septum and separates two water phases.

b) the torus phase which contains the bulk solution that was

displaced from the bilayer during thinning and which forms an annular

boundary between the very thin (5 nm)

septum (approximately 10 um).

bilayer and the much thicker

c) the aqueous phase which contains solvated ions as well as small

amounts of dispersed lipid and solvent.

The physical properties of the bilayer membranes are not intrinsic

to the bilayer phase per se but rather are a consequence of the

equilibrium that exists among these three phases in the bilayer system.

The components of biological membranes are also in equilibrium with

their intra- and extracellular environments. However, it does not

follow that the kinetics of the equilibrium that exists for biological

membranes are the same for the model membranes studied here. For

example the restin~ tension of the plasma membrane of Rye protoplasts j

arises from the equilibrium between lipids in the membrane reservoir

(probably in the form of large vesicles) and the plasma membrane (Wolfe

and Steponkus, 1981), whereas some other membranes, such as the plasma

membrane of erythrocytes have no detectable membrane reservoir. The

resting tension in artificial BLM arises from the equilibrium between

13

lipids dispersed in the torus (probably in the form of inverted

micelles) and the bilayer (Gruen and Wolfe, 1982). Differences such as

these must be born in mind when comparing phenomena witnessed in

artificial BLM and living membranes.

1.34 The Pitfalls of Solvent Retention in Artificial BLM

A consequence of the thermodynamic equilibrium between the torus

and bilayer phases in artificial BLM is that a significant amount of

alkane solvent is present in both the bilayer phase and in small lenses

of solvent distributed across the bilayer interior (see figure 1.7).

These solvent lenses "microlenses" scatter light strongly and have a

large mass compared to the surrounding bilayer (White and Thompson

1973). These microlenses have proved disastrous for analytical and

optical studies. This problem has been reviewed by Fettiplace et al.

(1975).

The presence of n-alkanes in BLM poses additional problems as

changes in the concentration of alkanes in the bilayer due to

environmental changes render it difficult to interpret changes induced

in the molecular organization of lipid bilayers. For example, the

effect of benzyl alcohol on membrane thickness was interpreted by

Ashcroft, Coster and Smith (1977) as a change in lipid head group area.

However, Ebihara, Hall, MacDonald, McIntosh and Simon (1979) interpreted

this data in terms of a change in the solvent retention of the bilayer

induced by the adsorption of benzyl alcohol.

Furthermore, biological membranes do not contain extraneous

molecules such as the n-alkanes and even more important, the presence of

Figure 1.7. This diagram shows what is referred to in this thesis as a microlense. The alkane in the microlenses and torus is in equilibrium with the alkane in the bilayer. After White (1977).

14.

n-alkanes in living membranes significantly alters membrane function and

structure (Haydon et al., 1977). Also the ionic conduction properties

of membrane bound polypeptides such as Gramicidin are altered by the

presence of n-alkanes in the hydrophobic interior of reconstituted

membranes ( Hendry, Urban and Haydon, 1978).

Thus efforts have been made to form bilayers which contain

insignificant solvent concentrations. White (1978) found that squalene,

though present in the torus could not partition into the bilayer phase

of bilayers formed from GMO, presumably because of its large molecular

dimensions compared to that of the bilayer. Longer chainlength

n-alkanes such as n-hexadecane are believed to have similarly low

solubility characteristics in bilayers formed from egg-lecithin (Haydon

et al., 1977).

Part of this thesis will be involved with characterizing the alkane

solubility properties of lipid bilayers in order to identify changes in

the lipid ordering from variations in the alkane solubility as well as

distinguishing solvent related properties of lipid bilayers to those

more relevant to living membranes.

1.35 Molecular Interactions in BLM

The conformation of lipid molecules in bilayers will be such that

the total Gibbs-free-energy of the bilayer-torus, aqueo~s solution 1

system is a minimum. The lipid molecules have a uniform packing density

throughout the bilayer and behave as a compressed liquid. The partial

molar volume of the lipids in the bilayer is constant and remains

independent of molecular conformation (ie. the lipid molecules are

non-compressible) as the presence of voids in a bilayer composed of

15 ·

flexible molecules such as lipids is energetically unfavourable (eg see

Fettiplace et al., 1971 and Gruen, 1980a).

Therefore the thickness of solventless* lipid bilayers is inversely

proportional to the cross sectional area of the lipid molecules in the

plane of the bilayer. The interfacial energy for the hydrocarbon-water

interface is 50 mJ/m 2 • However the free energy associated with the

total membrane-water interface is much lower (in the range .5-5mJ/m 2;

see Goldup et al., 1970). This makes the self-assembly of lipids into

bilayer structures energetically favourable.

The acyl chains are anchored at one end to the the polar head

groups at the bilayer-water interface which causes them to be partially

aligned in the bilayer. The Van der Waals attractive forces between

adjacent hydrocarbon chains favours a small lipid area per molecule in

the plane of the membrane (Gruen, 1980a).

The lowest energy configuration of the lipid fatty acid chains is a

random coil similar to that of n-alkanes in a liquid. The lateral

pressure in the lipid bilayer favours extended (all trans carbon-carbon

bonds) configurations which produce small lateral molecular cross

sections at the bilayer water interface (ie. minimising the energy due

to the oil-water interface). However, decreasing the molecular area at

the bilayer surface increases the order and the internal energy of the

fatty acid chains of the lipid molecules.

* This relationship does not hold when bilayers contain hydrophobic

molecules that do not contribute to the total surface area of the

bilayer.

16

Measurements of the order parameters of the acyl chains of

deuterated dipalmitoyl phosphatidylcholine using NMR techniques

(Stockton and Smith, 1976) found that the acyl chains are in an ordered

state near the hydrophobic-hydrophilic interface. However, near the

bilayer midplane they are disordered, behaving like an alkane liquid.

0 Computer modelling of this system by Marcelja (1974) and later by ·~

Gruen (1980a) found that the order profile of the acyl chains across the . er

~'J''),~ hydrophobic interior of these lipid bilayers is consistent with a

tf lateral pressure of 27 rnN/m. At equilibrium the following equation (

holds:

1. l

Where from left to right the terms are the lateral pressure of the

acyl chains and polar groups, the surface tension of the water-oil

interface (50 mN/m) and the surface tension of the bilayer-water

interface (which is negligible; see Chapter 10). From inspection of

equation 1.1 it seems that a significant contribution to the lateral

pressure arises from the polar head-group interactions in the bilayer,

22 mN/m for the polar heads as compared to 27 mN/m for the acyl chains.

The hydrophilic portion of the egg-lecithin molecule consists of a

choline phosphate electric dipole. In principle,the axis of the choline

phosphate dipole can take on any configuration between 0°and 90° to the

plane of the bilayer. The electrostatic interaction between adjacent

lipid molecules is a function of the lipid cross sectional area, the

relative orientation of the charged dipoles and the dielectric nature of

the material separating them. The internal energy of the charged

dipoles comes from the internal energy of the chemical bonds, energy in

17

the electrostatic field of the dipole (ie. the Born self energy), and

the dipole electrostatic potential between adjacent lipids. NMR studies

of Buldt, Gally, Seelig, Seelig and Zaccai (1979) indicate that the axis

of the choline phosphate dipole of dipalmitoyl phosphatidylcholine in

multilamella preparations is oriented parallel to the plane of the

bilayer though this has not yet been conclusively validated for planar

lipid bilayers. If this were so then the electrostatic force between

adjacent lipids is attractive. A simple calculation* of the maximum

lateral pressure that would arise from electrostatic attraction is

approximately 10 mN/m. However electrostatic screening effects of ions

in the external electrolyte could reduce the head-group interactions.

This possibility will be investigated latter in this thesis.

The net head-group interactions between adjacent lipids are

repulsive. The repulsive force may derived from water structuring in

the hydration shells of lipid polar groups. This effect has already

been postulated for ions in solution (Bockris and Reddy, 1970). NMR

studies on deuterated water has shown that that ten water molecules per

lipid to not freeze at 0°c in dipalmitoyl phosphatidylcholine bilayers

presumably as a result of the inhomogeneous structuring which prevents

hydrogen bonding between water molecules (see review of Pope et al.,

1978). Each lipid molecule appeared to alter the structuring of some 20

water molecules.

* This value was derived by calculating the electrostatic attractive

force between adjacent dipoles. From the number of dipoles along a

metre of bilayer surface an approximation of the total lateral

pressure was made.

18

Mechanical studies of lipid monolayers at the air-water interfaces

have measured the area per lipid as a function of lateral pressure. The

lateral pressure - lipid area relationship (n-A curve) of egg-lecithin

monolayers is shown in figure 1.8. The relationship is very steep for 0

lipid areas in the plane of the monolayer less than 65 A2 which

indicates that the lipids in bilayer aggregates are quite compressed.

Thus in lipid bilayers the area per molecule and hence bilayer thickness

will be very insensitive to changes in lateral pressure.

1.4 ELECTRICAL PROPERTIES OF ARTIFICIAL BLM

1.41 D.C. Characteristics

The most striking feature of lipid bilayer membranes is their high

electrical resistance; measurements of this parameter vary from

10 2to l0 5 ohms/m 2 • Membrane conductance has been measured as a function

of electrolyte concentration, pH, and cation binding (see review by

Goldup et al., 1970). However, measurements of the bilayer resistance

are never very reproducible and large variations in electrical

conductivity have been reported even on identical bilayer systems. For

example LaUger, Lesslauer, Marti and Richter (1967) found the resistance

of lecithin - n-decane bilayer was generally in the range l0 2 to 10~

ohm/m 2 whereas Hanai, Haydon and Taylor (1965c) found the resistance of

the same bilayers to be much higher and attributed the

lower values reported elsewhere to border "leakage". Hanai et al.

(1965c) demonstrated a linear relationship between bilayer area and

conductivity though the authors noted that this result is difficult to

obtain as border leakage often varied during changes in the membrane

area ( see also Miyamoto and Thompson, 1967). Van Zutphen and Van

~o

20

,u) area/molecule I.I\

Figure 1.8. The lateral pressure in an egg-lecithin bilayer shown as a function of lipid head group area in a monolayer at an air-water interface at 21°C. Note that for lipid ~rea~ typical of that found in lipid bilayer (6SK 2 ), the molecular area i;1 -.::ie plane of the monolayer is insensitive to changes in lateral pressure.

19

Deenen (1967) found that the resistance of egg-lecithin bilayers could

be reduced a hundred fold by adding trace amounts of lysolecithin to the

bilayer forming mixture. Israelachvili et al. (1980) pointed out that

the shape of these molecules makes it an ideal lipid for pore formation

in lipid membranes.

It is of interest to note that the resistance of lipid bilayers,

though very high, is a factor of 10''-10 16 less than that predicted from

theoretical calculations of "naked" ion translocation through the

hydrophobic interior of the bilayer (see Chapter 9). Various

alternatives to "naked" ion translocation were proposed by Parsegian

(1969) and MacDonald (1976) which will be considered in later sections.

1.42 A.C. Characteristics

Studies of the alternating current electrical characteristics of

lipid bilayer membranes have obtained values of membrane capacitance in

the range 3-8 mF/m 2 depending on the bilayer composition. The

capacitance measurements of lipid bilayers, in contrast to the

electrical resistance measurements, are very reproducible. The

dielectric capacitance of the bilayer membranes can be related to their

thickness by the following equation:-

Where 11 0 11 is the membrane thickness and 11 e: 11 is the dielectric m

constant of the region containing the fatty acid chains. Hanai, Haydon

and Taylor (1965b) suggested that the dielectric constant of the

hydrophobic interior of lipid bilayers should be similar to that

20

measured in alkane liquids (in the range 2-2.2). However, Ohki (1968)

pointed out that if the fatty acid chains of the lipids were aligned

perpendicular to the plane of the bilayer then the dielectric constant

of the bilayer would be appreciably higher than that of hydrocarbon

liquids. However, the more recent calculations of Huang and Levitt

( 1977) obtained values of the dielectric constant in the range 2. 1 to

2.2. Comparative optical and electrical studies of Tien and Diana

( 1967) placed the value of the dielectric constant of the hydrophobic

region in the range 1.6-3.8.

LaUger et al. ( 1967) and Everitt and Haydon (1968),

Gouy-Chapman theory applied to the bilayer-solution interface,

using

showed

that during capacitance measurements part of the externally applied

potential appears across the aqueous phase adjacent to the membrane.

The subsequent redistribution of ions at the membrane-solution

interfaces gives rise to ionic double layers that have capacitances

which act in series with the dielectric capacitance of the bilayer. The

exact analytical solution for the A.C. impedance of the double layer was

calculated by Smith (1977) who predicted that the time-constant of the

ionic double layer was equal to that of the membrane. The capacitance

of the double layer, CDL' for the case of small applied potential

differences was shown to be given by the following equation:

e: e: c0L = ~ w coshtffi} 1.3

Where e:w is the dielectric constant of water.

In effect, the double layer capacitance is equal to what is

expected frora a slice of electrolyte one "hypothetical" Debye length

thick. The "hypothetical" Debye length here refers to the Debye length

21

of a bulk solution phase with an ion concentration equal to that in the

plane of fixed charges.

The predictions of Everitt and Haydon (1968) were latter verified

experimentally by White (1973) from measurements of the salt dependent

capacitance of GMO - n-decane bilayers. Earlier measurements of Hanai

et al. (1964) reported that ionic strength had no effect on the

capacitance of egg-lecithin - n-decane bilayers which was attributed to

absorbed surface charge in these bilayers. However, more recent

measurements of Coster and Smith (1974) detected a significant variation

in bilayer capacitance of egg-lecithin bilayers which were formed using

n-tetradecane solvent. The effects of varying ion concentration in the

external electrolyte on bilayer capacitance will be examined in later

sections of this thesis.

It has long been known that the heterogeneity of the dielectric

structure due to the different chemical nature of the hydrophobic and

hydrophilic regions of lipid bilayers would influence the optical and

electrical properties of the film. The impedance of this heterogeneous

structure should exhibit a dispersion with the frequency of the

externally applied potential (Hanai et al., 1965a). Though some

indirect evidence for a dispersion in bilayer capacitance existed

(Coster and Simons, 1970 and Clowes, Cherry and Chapman, 1971) it was

commonly believed that the bilayer capacitance was independent of

frequency {Hanai et al., 1964).

Coster and Smith (1974) using a novel four terminal digital

impedance @easuring technique documented in Bell, Coster and Smith

(1975) demonstrated a small (3%) dispersion in the capacitance of

egg-lecithin bilayers over the frequency range 1-90 Hz. These results

22 ·

were consistent with a region of low dielectric constant (the

hydrophobic interior) sandwiched between two regions of higher

dielectric constant with £r in the range 20-40 (the hydrophilic region).

By extending the frequency range of the impedance measurements Ashcroft,

Coster and Smith (1977) resolved two electrically distinct polar regions

which were associated with the regions containing the choline phosphate

dipoles and the glycerol region of the egg-lecithin molecules. This

thesis will employ comparative studies on GMO and egg-lecithin

bilayers to test the hypothesis of Ashcroft et al. (1977) and so

characterize the dielectric structure of the hydrophobic-hydrophilic

interface of egg-lecithin and GMO bilayers.

CHAPTER 2

THE DIELECTRIC MODEL OF ARTIFICIAL BIMOLECULAR LIPID MEMBRANES.

2. l INTRODUCTION

2.2 HYDROPHOBIC REGION

2. 21 Ion Self Energy

2.22 Membrane Conductance

2.23 Membrane Capacitance

2.3 POLAR REGIONS

2.4 AQUEOUS PHASE

2.41 Effect of Unstirred Regions

2.42 Effect of Ionic Double Layers

2.5 PHILOSOPHY OF MEASUREMENT

2.51 Time Domain vs. Frequency Domain

2.52 Four Terminal Digital Impedance Measuring Technique

page

24

25

25

27

29

32

34

34

35

38

38

41

23

24 -

2.1 INTRODUCTION

This chapter is concerned with the quantitative examination of the

dielectric model of egg-lecithin bilayers as well as the principles

involved in the impedance - measuring techniques used in elucidating

bilayer structure.

For simplicity, in the analysis which follows, monovalent ions are

considered to be the carriers of electrical current through the bilayer.

The equilibrium distribution of ions throughout the membrane and

solution will determine the relative electrical conduction properties of

different regions in the membrane system.

For the purposes of this study the ions are considered to exist in

three chemically distinct regions; the aqueous phase external to the

membrane, the hydrophobic region containing the acyl chains and the

polar heads forming a region with fixed charges between the hydrophobic

region and the aqueous phase. In this thesis the electrical properties

of the bilayer were modelled by a series combination of parallel

resistor-capacitor networks; each simulating the dielectric properties

of different regions in the bilayer (see figure 2.1).

In later chapters of this thesis the dielectric substructure of

membrane - electrolyte systems will be deduced by modelling the i

impedance dispersion of lipid bilayers to this equivalent circuit. The

interpretation of the data requires an understanding of the charge

storage and conductive mechanisms in these bilayer systems. The

dielectric and conductive properties of these regions are considered

separately.

AOUEOJS SOLUTION

Cp

AC YL CHAIN REGION (HYDROPHOBIC)

GH

REGION

CHOLINE -PHOSPHATE REGION

( POLAR HEAD)

Figure 2.1. The equivalent circuit used to model the dielectric properties of lipid bilayers in this thesis. The electrically distinct regions detected by Ashcroft, Coster and Smith (1981) are shown (ie. the hydrophobic the acetyl and the polar head regions). After Coster and Smith ( 197 4) .

25

2.2 THE HYDROPHOBIC REGION OF BIMOLECULAR LIPID MEMBRANES.

2.21 Ion Self Energy

The hydrophobic region of bimolecular lipid membranes is here

defined as the region containing the acyl chains of the lipid molecules.

This region is treated as a uniform thin slab of dielectric,

approximately 2.7 - 5 nm thick and with a dielectric constant of

2.1 - 2.2 (Huang and Levitt, 1977) which is in equilibrium with a binary

electrolyte of rnonovalent ions. The partitioning of ions between the

two phases is determined by the potential energy difference of the ions

in the hydrophobic region of the bilayer and aqueous phase.

Born (1920) calculated the ion self energy "U" from the total

energy stored in the electrostatic field. The energy stored in the

electrostatic field, dU, in the volume element, dVm, is then given by:

dU = ½ £ £ E2 .dVm o r

For a radially symmetric field

E(r) = q/4n£ £ r 2 o r

2.1

2.2

Where ''E" is the electrostatic field in the volume element "dV" m

and "q" is the electronic charge. For an infinite medium of dielectric

constant E the total energy of the ion (ionic radius, a) is then equal r

to:

U = '-"foo q2 2 4n£ £ a o r

1 g2 r2"" dr = 8n£ £

o r 2.3

26

When an ion is present near the planar interface of two dielectric

media of differing dielectric constants the additional polarisation

charge induced at the interface will alter the electrostatic field of

the ion in the dielectric medium. Calculation of the "self energy"

using the Born method is then more difficult as the electrostatic field

is no longer radially symmetric.

A much easier approach to this problem is to use the method of

electrostatic images (eg. see Lorrain and Corson, 1970). If the

dielectric interface is located at x=O and the ion is in medium 2 at x=h

(refer to figure 2.2) then the electrostatic field in medium 2 is

identical to that produced by the original charge in an infinite medium

with dielectric constant, £2, plus an image charge at x=-h. The

electrostatic self energy of the ion in medium 2 is given by equations

2.4 to 2.5.

2.4

where 2.5

The "self energy" of an ion near the interface of medium 2 is equal

to the "self energy11 if medium 2 were infinite, plus an extra term

associated with the electrostatic potential energy due to the proximity

of the image charge.

Thus the electrostatic self energy is a smoothly varying function

of position across the dielectric discontinuity. Strictly speaking the

method of electrostatic images presented here is only valid for point

charges. However this approach is a useful approximation for the case

where the charge is several ionic radii from the dielectric interface.

MEDIUM 1

0 Q'

X= -h X=O

MEDIUM 2

-Q

X =h

Figure 2.2. An ion near a dielectric discontinuity in a medium will posses an electrostatic self energy which differs from that in an infinite medium. The effect of the dielectric interface at X=O is identical to that of an image charge (Q') located in at X=-h.

27 ,.

Using equations 2.4 to 2.5 one can calculate the partition

coefficient of an ion as a function of position using the Boltzmann

equation. For a monovalent ion with a radius .2nm the partitioning

between the aqueous phase (£w=80) and the bilayer interior (~=2) is -30 approximately 10 . Thus the ion concentration in the hydrophobic phase

must be exceedingly small.

2.22 Membrane Conductance

The Nernst-Planck equations have been widely used in the analysis

of ion flows through membranes. The dielectric and conductive

properties of the hydrophobic region can be derived from solutions of

the Nernst-Planck equations.

The definition of the electrochemical potential for the case of a

monovalent ion species, i, in dilute solutions with a charge, q, is :

µ. = µ? + kT lnc. + q~ 2.6 1 1 1

The thermodynamic force driving the diffusion process for a given

ion is equal to the negative gradient of the electrochemical potential:

2.7

When the ion is more than a few ionic radii from the dielectric

interface dµ~/dx is approximately zero. 1

28.

The electrical current carried by an ion species,i, is given by:

J. = q D.c.X./kT l l l l

2.8

Where Di is the ion diffusion coefficient, substituting Xi from

equation 2.7 we get:

de. l

Ji = -q Di dX 2.9

Previous calculations based on equations 2.4 and 2.5 showed that

the ion concentrations in the hydrophobic region are very small.

Therefore the electric field in the membrane is independent of position

( See Goldman, 1943). Hence:

~ = constant = i dx u 2.10

Where 11 V11 is the externally applied potential and 11 611 is the

membrane thickness. Provided, qV/kT « 1, then the ion concentration

throughout the membrane is uniform. Then equation 2.9 reduces to:

Hence

or

where the

-q 2 D.c.V l l Ji = --,k...,,,T=-o-

the membrane conductance is

G. l

= q2 D.c./kTo l l

G. l

= qµ. y c./6 l p l

ionic mobility, µi = D.q/kT l

2 .11

given by:

2.12

2.13

2.14

29 ·

Neumcke and Lauger (1969) extended this analysis to the case where

the standard chemical potential was position dependent. The self energy

of the ions in the hydrophobic interior included terms which accounted

for the effects of the dielectric discontinuity at the membrane aqueous

interface. The results obtained by these authors showed a non-linear

voltage - current (V-1) property which, for suitable values of the

parameters, was similar to the V-1 characteristics of lipid bilayers

measured in previous studies (Hanai et al., 1964).

2.23 Membrane Capacitance

i) Capacitance Due to Charge Storage

Electrical work can be stored within the hydrophobic region of the

membrane in the form of either space charge due to unequal cation and

anion concentration profiles within the membrane or as dielectric charge

storage due to interfacial polarisation at discontinuities in the

dielectric properties of the medium.

The dielectric, area-specific, membrane capacitance, CD, can be

calculated using the following expression:

CD=££ /6 2.15 o r

The non-dielectric charge storage in the membrane can only be

calculated from the exact solutions to the Nernst-Planck equations.

However, an upper limit to the non-dielectric capacitance was calculated

by Ashcroft (1979) using a modification of the method of Neumcke, Walz

and Lauger (1970). It was assumed that the net space-charge was a

linear function of position with a maximum at the membrane midplane

30

equal to 50% of the total ion concentration (see fig 2.3). The voltage

across the membrane, V, can then be calculated by twice integrating

Poisson's equation which gives the following expression:

Uhere cris the net charge concentration at the membrane midplane.

An upper limit to the charge storage in ion profiles can be made by

choosing a large ion concentration in the membrane, say 1~7 mole/m ~ and

d = 6 nm then Vis 10- 15 volts. The total charge stored, calculated by

integrating the space-charge over volume, is 10- 37 coulomb. Hence the

capacitance due to the presence of this space charge distribution is

10- 22 F/m2 which is negligible compared to the dielectric capacitance of

the membrane (5.10- 7 F/m2 ).

The membrane capacitance and conductance have been examined here

using the time independent Nernst-Planck equations. However, the

membrane impedance is calculated from measurements of A.C. voltage

signals. The general solution of the Nernst-Planck equation in the

presence of a sinusoidal displacement from the equilibrium has been

presented by Smith (1977). His calculations show that the Goldman and

time-independent approximations are valid for the derivation of the

capacitance and conductance of the hydrophobic region.

ii) Phenoffienological Capacitance Due to Time Varying Resistances

The analysis of the data in this thesis relies on the assumption

that the dispersion in the measured capacitance and conductance of lipid

bilayers is due solely to their non-homogeneous dielectric structure.

R~/aflve Difference between Gofton &

100¾

An1on fIJ 3/t Corcenfrat,ons 0

·o

Figure 2.3. The assumed relative difference between cation and anion concentrations in a membrane used to attain an upper limit to non-dielectric charge storage within a membrane (see text).

6

31

Mauro (1961) showed that system with a time-variant resistance

could exhibit an additional phenomenological A.C. impedance; the real

and imaginary parts showing a dispersion with frequency. A time varying

resistance will manifest itself whenever the steady state V-1

characteristics are non-linear and when there is a finite delay between

the application of an external potential and the current response to

that signal. At high frequencies when a time-varying system has

insufficient time to respond to perturbations from steady state the

impedance is equal to the "cord" impedance (refer to figure 2.4).

However, at lower frequencies the impedance is equal to the slope of the

V-1 curve.

The observation of a non-linear, D.C. V-1 steady state

characteristic in egg-lecithin bilayers is necessary, though

insufficient evidence for a time-dependent resistance for BLM. It is

important, therefore, to estimate the effects of a possible time

dependent membrane resistance on the A.C. impedance results measured at

different frequencies. Measurements of Hanai et al. (1964) show that it

is linear within the range of applied ;:>otentials ± 50 mV. The

measurements in this thesis employed A.C. potentials less than 15 mV

with no D.C bias. For such displacements in potential-difference the

BLM were found to have linear V-1 characteristics to within the

precision of the potential-difference ~easuring technique* (±.1%).

to detect * The impedance measuring system could be used

non-linearities in the V-1 response of the membrane. This was done

voltage signals by measuring the relative distortion in sinusoidal

appearing across the membrane (see section 3.42).

Figure 2.4. A non-linear V-1 characteristic curve. When the cord and slope resistances differ, the system may exhibite a phenomenological impedance.

32 "

The anomalous impedance dispersion that would occur if the bilayer

impedance was indeed both time-varying and non-linear could be ignored.

From the above considerations it is valid, for small voltage

signals, to consider the hydrophobic region of egg-lecithin bilayers as

an ideal capacitor shunted by an ohmic resistor. The capacitance of the

hydrophobic region will then be equal to that of a parallel plate

capacitor filled with a medium with a dielectric constant equal to that

of the acyl chain region of the BLM, having a plate separation equal to

the thickness of that region.

2.3 POLAR REGION

The polar region, here, is defined as the region containing the

non-acyl chain chemical moieties at the hydrophobic - hydrophilic

interface of lipid bilayers. In the case of egg-lecithin that would

include the glycerol and choline phosphate moieties of the molecule.

The alignment of the choline phosphate dipoles at the

membrane-solution interface creates a region of fixed charges believed

to have a dielectric constant in the range 20-40 (Coster and Smith, 1974

and Ashcroft, 1979) and a thickness, depending on the dipole

orientation, between .5 and 1.1 nm. The dielectric capacitance of the

choline phosphate region should then be in the range .2-.8 F/m 2 •

Theoretical calculations of the impedance of the double fixed

charge membrane (DFCM) by Mauro (1962) showed that an additional

capacitance could arise from space charge at the junction of the

33

positive and negative sheets of fixed charge. Mauro pointed out that

the choline phosphate groups may form a DFCM and that the capacitance of

the polar region would be the parallel combination of the dielectric and

'Mauro' capacitances.

Subsequent examination of the electrical properties of the polar

region by raeans of their electrophoretic raobility (Hanai et al., 1965a)

indicated that the choline phosphate dipole of the egg-lecithin

molecules in bilayers is oriented parallel to the plane of the membrane.

More recent NMR studies in multilayers (eg. Seelig, Gally and

Wohlgemuth, 1977) show that the axis of the choline phosphate group is

parallel to the plane of the bilayer and has considerable rotational

mobility. In the light of this it appears unlikely that the theory of

the DFCM can be successfully applied to the polar region of egg-lecithin

bilayers. In any case Coster (1973) showed that even if the polar head

region of egg-lecithin bilayers could be modelled as a DFCM the effect

of the 'Mauro' capacitance is small compared to the dielectric

capacitance of the polar heads over the frequency range of impedance

measurements employed in this thesis.

34 ,,

2.4 AQUEOUS PHASE

2.41 The Effect of Unstirred Regions

Significant contributions to the merabrane impedance can arise from

unstirred layers and ionic double layers in the electrolyte adjacent to

the membrane.

Passing a steady current across a membrane - electrolyte boundary

when ion transport numbers in the membrane and electrolyte are

different can lead to a perturbation in the ion concentration profiles

in the regions adjacent to the BLM whereby the ion concentrations are

either increased or decreased (see figure 2.5). If the electrolyte was

perfectly stirred right up to the membrane boundary then this effect

\'IOU l d not occur.

The perturbations in the ion concentration profiles in the

unstirred regions adjacent to the membrane will introduce a dispersion

in capacitance and conductance. Measureraents of the effect of unstirred

layers on the zero-current conductance of glycerol monooleate

(Gf,iO) - n-decane bilayers containing valinomycin (an ionophore

selective for potassiura) ~as obtained by Ciani, Gambali, Gliozzi and

Rolandi (1975). This study revealed that the effect of unstirred layers

on the measured raembrane conductance, where the ionic strength in th~ I

4 aqueous phase exceeded 10 M/m 3 , could be ignored provided the membrane

conductance was less than 0.1S/m 2 • All the results reported in this

thesis were obtained from bilayers for which this condition was

applicable.

1---------~ !/1 , . --· =t .-------- -- I iv t c-=-:-.-:::====---------)C ----~~----~---·-· __ ) i

C: ::,

-0

t = t 'k 'Cl

--x:::-L X=--8 x=·b X=L

Un s t i r red M cm bra P e Un s t i r r e d S t i r r ;, j

------>-solvent flow

------1:-· e I e c t ric current

------~ electroosmotic f !ow

Figure 2.5. A discontinuity in ion transport number {represented by the thick arrows) at a membrane solution interface can give rise to changes in the ion concentrations adjacent to the bilayer. If the solutions were stirred up the the membrane interface then the ion concentrations would be uniform in the aqueous phase.

,,

Sraith (1977) calculated the effects of unstirred layers on the low

frequency membrane - electrolyte impedance. It was found that the

capacitance of a membrane bounded by unstirred layers could be

significantly higher than its dielectric capacitance. However, for the

lipid bilayers studied in this thesis the effect of unstirred layers was

expected to be negligible.

2.42 Ionic Double Layers

When a potential difference is applied across a

membrane - electrolyte system a portion of the _applied electric field

will appear in the external solution. This is a consequence of the fact

that the displacement field, D, due to the applied potential difference,

is continuous across the membrane solution interface. However, the

movement of ions, external to the membrane, in this field will screen

the electric field frora the bulk electrolyte phase. This produces a net

space-charge at the membrane solution interface. This region of net

charge will be referred to as the Gouy-Chapman ionic double layer. The

potential as a function of position through such a system under an

applied potential, V, is shown in figure 2.6. Lauger et al. ( 1967)

calculated the capacitance of a neutral membrane solution system by

calculating the derivative of net space charge with respect to the

applied voltage. This was done as follows.

The space charge as a function of position in the aqueous phase ~s f

given by:

\there (refer to figure 2.6)

<1> 0 (x) = (tjJ(x) - t/l(oo))fkT

2.17

2.18

36

The position dependent potential can then be calculated from

Poisson's equation which with planar symmetry becomes:

~= ~-= ~ dx 2 dx £ £ o r 2.19

It can be shown that equation 2.21 is obtained by solving equations

2.18 and 2.19 and applying the following boundary conditions.

where

and

- V iµ{oo) - 2

= y_ + 2kT ln {1 + tanh(a/2).e-X/A} IJi(x) 2 q 1 - tanh(a/2).e-x/A

V a= (2 - iµ(o))/2kT

which is the debye length in the bulk aqueous phase.

2.20

2.21

2.22

2.23

One can

further simplify this equation by applying the boundary conditions and

using the Goldman approximation.

The total capacitance of the membrane electrolyte system, I

now be calculated by integrating the space charge over all

dividing by the applied voltage.

Hence:

2.24

2.25

C. can m' 11 x11 and

2.26

-'------:\/--(co J

V

1 X= -6 x=O

BULK AQUEOJS PHASES J

j

Figure 2.6. The spatial variation of potential across a membrane when there is an externally applied potential, V. A significant fraction of the potential appears across the external electrolyte due to the presence of ionic double layers at the membrane solution interface.

37

Everitt and Haydon (1968) used the same theoretical analysis to

calculated the capacitance of a charged membrane. If the presence of a

surface bound charge, 0 b , gives rise to a potential at the membrane's

surface, ~0• in the absence of an externally applied field, then:

v .. £ £w cosh(q~ /kf)

0 0

2.27

~Jhere ~ = 2kT sinh- 1{qobA/2kT£ Ew) 0 q 0

2.28

When 11 ~ 11 is large or 11 A II is sma 11 equation 2. 27 reduces to 0

equation 2. 15. Therefore neglecting the presence of ion double layers

when calculating membrane thickness from the total membrane capacitance,

using equation 2. 15, is only valid when the bilayer has a large net

surface charge or when the ion concentration in the aqueous phase is

high (>. l Molar).

The rigorous treatment of the impedance of membrane solutions

systems has been made by Smith (unpublished data). The solutions to the

steady state Nernst-Planck equations for the membrane - solution

interface allowed calculation of the capacitance and conductance of the

ionic double layers external to the membrane. Smiths results indicated

that the ionic double layer has a electrical time constant equal to that

of the membrane irrespective of the external ion concentrations in the

bulk phase. This region will therefore not effect the relative

dispersion in membrane impedance. However, th~ capacitance at each l

frequency will be somewhat reduced. This surprising result will be

discussed later in this thesis (see Chapter 7).

38

2.5 PHILOSOPHY OF MEASUREMENT

2.51 Time Domain vs. Frequency Domain

Thus far, techniques for measuring membrane impedance have fallen

into two main categories; measurements in the time domain and

measurements in the frequency doraain ( each being the Fourier transform

of the other) .

i) Time Doraain

This method of measuring membrane impedance involves introducing a

step change in voltage or current across the bilayer and measuring the

subsequent time-course of the current or voltage signal. This is

equivalent to simultaneously applying a whole spectrum of A.C. voltage

signals across the bilayer. However the choice of the relative

amplitudes of the different frequency components of the voltage and

current steps are restricted to that obtained from fourier transform of

the time-course of the applied voltage or current.

The voltage response of a single parallel resistor-capacitor (GC)

network to a current step is of the form:

where:

V(t) = voltage across the GC network at time t

VO = voltage at t= 00

2.29

2.30

39

T1 = the electrical time-constant of the network

c1 = the capacitance of the GC network

G1 = the conductance of the GC network

Uhen r~ such arrangements of GC elements in series (see figure 2.7)

are subjected to a current step.

0 It can be seen that, Vn,

N -t/T L V0 (1 - e n)

n=l n

the steady state voltage across

2.31

each

element will determine the maximum contribution of each element to the

total potential across the bilayer. It is seen that in latter sections

(see Chapter 7 and Chapter 9) that the conductance of the polar head

regions is at least three to five orders of magnitude higher than the

hydrophobic region. Therefore the polar heads will only contribute

.001-.1% of the total measured signal. further, these measurements of

V(t) in the time interval between the step current pulse and t<.0001

second have to be made to accurately determine the time constant of the

polar head regions.

ii) Frequency Domain

The Maxwell-Wagner dispersion in capacitance, C, and conductance,

G, of a network of N GC elements in series can be derived from the

following recursive relations:

C n+ 1 ( w) = _w_2 c_n_c_n_-l_(_2_c_n +_c_n_-_1_) _+_4_Cn_-_l_(_cn_)_2_+_2c_n_G_~_-_l

For n=2 to N-1 (2Gn+Gn-1)2+w2(2cn+Cn-l)2 2.12

2GnGn_ 1(2Gn+Gn-l) + w2(2GnCn-l+4Gn-l(Cn) 2) =------------------ 2.33

where and 2.34

0

c,

Figure 2.7. The circuit network used to model the lipid bilayer and electrolyte in this study. The circuit consists a number of parallel resistor - capacitor elements in series. The capacitance and conductance spectrum of a two element network of this type is shown in figure 2.8

0

5·0 C 1a3

4,9

lJ_ l/) 1rl E 4.8 E

Lu

~ (._)

<: ~ t..7 ~ ~ (._) - 1 :::, (J

~ ~ "( 46 8 (._)

4.5 10-1

10 10 1 10 100 1000 FREQUENCY Hz

Figure 2.8. The theoretical capacitance and conductance spectrum of a two element impedance network of the type shown in figure 2.7. The resistance and capacitance values of the two series elements are:

Cl=5mF and Gl=lmS C2=300mF and G2=5S

40

The frequency dependence of a two element circuit of the type shown

in figure 2.7 is shown in figure 2.8. G(w) and C(w) can be derived

from measurements of the total impedance Z(w) and phase •(w) angle of

the multi-element circuit using the following equations:

C(w) = sin i{w) w Z w) 2.35

2.36

Measurements of Z(w) and •(w) have been made using and A.C.

impedance measuring bridge (see Hanai et al., 1964 and White 1970). A

balance is obtained when the phase angle and amplitude in the known and

unknown arms of the bridge are equal. The main difficulty with this

method is the detection of balance as the voltage across the membrane

must remain small in order to remain in the linear portion of the V-1

characteristics of the bilayer (V<50 mV). The long times required to

obtain a balance at frequencies less than 10 Hz make this null detection

method impractical for ultra-low frequency impedance measurements.

Further, such bridge measurements are restricted to 2 (or sometimes 3)

terminal methods where the electrode - solution impedances make it

difficult to extract the membrane impedances from the total impedance.

This is particularly so at low frequencies (eg.

1970).

Coster and Simons,

Measurements pf membrane impedance in the frequency domain are l

superior to that in the time domain because there is no necessity for

high time resolution during bridge measurements. Further, noise

reduction techniques in A.C. measurements is superior to those in D.C.

measurements.

41

2.52 Four Terminal Digital lrapedance Measuring Technique

A digital impedance measuring technique was developed by Coster and

Smith (1974) to measure membrane impedance in the frequency range .lHz 0

to 100 Hz with a resolution of .3% amplitude and .02 phase angle. This

method overcame the problems inherent in A.C. null detection techniques

at these low frequencies.

The digital impedance measuring technique directly measures the

phase difference and amplitude ratio of the current and voltage signals

occurring across the bilayer and a known standard hard-wire network

connected in series with the membrane. This is done by simultaneously

measuring the time course of the voltage signal developed across the

membrane and standard impedance network. This eliminates the need to

detect a balance. A schematic diagram showing the principle of the

impedance measuring system is shown in figure 2.9. Greater detail of

the impedance measuring system "BULFIS'' is given in chapter 4.

Potential measurements in the frequency and time domain require an

electrical connection between the measuring device and the aqueous phase

near the membrane. Two different electrode configurations have been

used for sampling membrane potential; the two terminal and the four

terminal method.

The two terminal method involves connecting the potential measuring

device to the current injecting electrodes. Correct calculation of the

membrane potential necessitates a correction for the frequency dependent

electrode-solution potential in series with the membrane potential. The

effect of the electrode solution potential can be allowed for (eg see

Coster and Simons, 1970) but requires highly accurate measurements as

Rs

Cs

Figure 2.9. The main features of the four terminal impedance measuring method. The potential measuring electrodes do not inject any significant ~urient. The current is calculated from the potential developed across an accurately known impedance in series with the membrane.

42

the vectorial subtraction procedure is extremely sensitive to small

errors.

The four terminal method employs two pairs of electrodes; one pair

for injecting current and a separate pair for measuring membrane

potential. Provided the input impedance of the voltmeter is

sufficiently high (10 13 ohms} the potential at the electrode solution

interface is equivalent to the electrostatic (zero current} value. The

effect of the electrode-solution potential then cancels when using an

identical pair of potential electrodes. Thus the four terrainal method

eliminates the need to correct for the electrode - solution impedance.

The impedance of the membrane can be derived from the phase

difference, ~D, and amplitude ratio, AR', between the voltage signals

appearing across the me~brane and the impedance standards using the

following equations:

2.37

\Jhere "~s" and "Zs" are the phase and magnitude of the standard

impedance.

CHAPTER 3

APPARATUS

3.1 INTRODUCTION

3.2 MEMBRANE APPARATUS

3. 3 AMPLIFIER ASSEMBLY

3.31 General Principles of the Design

page

44

46

49

49

3.32 Amplifier Assembly; Mechanical Construction 50

3.33 Amplifier Electronics 51

3.4 COMPUTER CONTROL

3.41 BULFIS Hardware

3.411 Progranvnable Signal Generator

3.412 Data Acquisition Boards

3.42 BULFIS Software

52

52

52

53

53

43

44

3.1 INTRODUCTION

The function of the various components of the experimental

apparatus fell into four main categories. Photographs of the overall

experimental setup are shown in figures 3.1 and 3.2. The experimental

apparatus consisted of:-

a) an assembly to contain the membrane and its electrolyte environment

for impedance measurements·.

b) Four Ag/AgCl electrodes, a voltage divider network, standard

impedances and two differential amplifiers (for amplifying the small

voltage signals developed across the membrane and standard impedance)

which were housed in a Teflon frame.

c) a Faraday cage mounted on an antivibration platform which contained

the amplifiers and bilayer apparatus.

d) the Biophysics Ultra Low Frequency Impedance Spectrometer (BULFIS)

which embodied an LSI 11 microprocessor connected on line to the

experiment. It controlled the generation of the sine wave, acquisition

of raw data (voltage and current signals) and computation of membrane

impedance.

The hardware and software for BULFIS was developed by T. Chilcott

of this laboratory, based on an earlier digital four terminal impedance

measuring system described by Bell, Coster and Smith (1975).

A

B

Figure 3.1. The rrarbrane apparatus. A photograph depicting the cell in which the rrarbrane ~'/as fonred and its relation with the electronic apparatus.

A) Bi nocu l ar microscope for vi e-,i ng the rrarbr ane

B) Arplifier asserbly

C) Memrane apparatus

C

A

C

Figure 3.2. The general layout of the experiirental apparatus shONing the:

A) Biophysics Ultra LON Frequency Irrpedance Spectrareter

B) Faraday cage housing the rrarbrane apparatus

C) Anti-vibration platform

45

The experimental design was aimed at performing accurate impedance

measurements of lipid bilayers which have stable mechanical and

electrical properties in an accurately monitored environment. Some

factors taken into consideration during the design of this apparatus

were:-

a) Minimising electrical noise and mechanical vibration in the membrane

environment.

b) Minimising the effect of any stray capacitance or current shunt paths

on the measured membrane impedance.

c) Eliminating the probability of surface active or extraneous

hydrophobic substances (from other experiments conducted with the

apparatus) contaminating the membrane. This was done by ensured easy

cleaning of all non-disposable components.

The precise impedance measurements needed in these experiments

placed heavy demands on the experimental apparatus. A significant part

of the work in this thesis was devoted to overcoming the problems

inherent in accurate measurements of high impedance. Technical details

of apparatus was only given in this thesis where such detail is required

to clarify the validity/feasibility of experiments or any results

obtained.

46

3.2 MEMBRANE APPARATUS

The apparatus containing the membrane and electrolyte is shown in

figures 3.2 and 3.3.

Black lipid bilayers (membranes) were formed across a 1-2 mm

aperture in a polycarbonate septum dividing two compartments which

contained electrolyte. A plexiglass vial formed the inner compartment

which was suspended in a small glass beaker. The polycarbonate septum

was fixed across a hole milled into the side of the plexiglass vial.

The adhesive used for this purpose was a solution of plexiglass in

ethylene dichloride. This was used as it ensured a water-tight,

mechanically strong, non-contaminating seal between the polycarbonate

and plexiglass.

A countersunk hole was milled into the polycarbonate septum, while

mountea on a brass block, using the tip of a highly polished 5 mm

diameter drill bit. Any rag that remained at the edge of the hole was

removed by gently abrading it with a piece of fine emery cloth. Water

was the only lubricant used while machining the plexiglass vial. This

was to circu~vent the possibility of trace amounts of machine oil being

absorbed into the plexiglass and contaminating membranes in future

experiments. The vial and septum were then thoroughly cleaned in

alcohol and distilled water before use.

Heat was added or removed from the electrolyte via a solid state

Peltier (Cambion 601-4000) which was raounted in thermal contact with the

base of the glass beaker. The Peltier device was operated from a

variable (0-SV) D.C. power supply. The power supply was remote from the

1 2

13

15

Figure 3.3. A schematic diagram showing details of the apparatus upon which the bilayer membranes were formed.

KEY

l Quartz fibre, light pipe for illumination 2 Binocular microscope 3 Aperture upon which bilayers are formed 4 Polycarbonate septum insert 5 Plexiglass holder 6 Plexiglass inner chamber 7 Voltage measuring Ag/AgCl electrodes 8 Current injecting Ag/AgCl electrodes 9 Vinyl tube for hydrostatic pressure adjustment 10 Temperature transducer 11 Glass beaker 12 enclosure for temperature transducer 13 Aqueous phase 14 Peltier device 15 Heat sink

47

rest of the experimental apparatus, to avoid problems of A.C.

pickup by voltage and current measuring amplifiers.

signal

In most experiments an air-cooled copper heat sink under the

Peltier device was sufficient as a source, or sink of, heat which was

pumped into, or out of, the membrane chamber via the Peltier device.

Experiments involving measurements at more that 5°C below ambient

temperature employed a water-cooled brass heat sink. This improved the

efficiency of the Peltier device. However, mechanical vibration

transmitted through the water pipes feeding the heat sink reduced

membrane stability.

wherever possible.

Therefore the air-cooled heat sink was used

The temperature was monitored (±.5°C) using

temperature transducer Analog Devices AD 590KF).

mounted in a small glass vial which was suspended in

adjacent to the membrane.

a

This

the

solid state

device was

electrolyte

A syringe connected to the beaker by a length of vinyl tube was

used to equalize the hydrostatic pressure across the membrane by adding

or removing small quantities of aqueous solution.

Four silver-chloride coated silver (Ag/AgCl) electrodes, suspended

from fixed brass rods, were immersed in the aqueous solutions on either

side of the ~embrane. Two of these electrodes were for injecting

current though the membrane and two for measuring the pot~ntial

developed across the membrane. The potential-measuring electrodes were

suspended so that they were close to the membrane. This was to reduce

the voltage drop that occurred across the layer of electrolyte between

the membrane and the ~otential measuring electrodes.

48

These electrodes were made by electrolytic deposition of silver

chloride onto a length of silver wire immersed in a l Molar potassium

chloride electrolyte. The main advantage of using Ag/AgCl electrodes in

measuring potential differences is that they have a well defined

electrode-solution potential.

The plexiglass chamber, septum and Ag/AgCl electrodes were

discarded after they had been in contact with certain contaminating

chemicals (where their effect on BLM may have been the subject of

investigations in prior experiments).

The membrane was viewed in white light under a 10-40x binocular

microscope. Good visibility was required to monitor changes in the

me~brane geometry due to bowing as the reflectivity coefficient of

bilayer membranes is very low. The illumination was produced by a 100

watt quartz-halogen light bulb operating from a remote 12 V D.C. power

supply and variac. The light was channelled through quartz fibre optics

to the membrane. This provided a cool source of illumination very close

to the membrane which could be easily adjusted for maximum visibility

and which did not introduce A.C. pickuµ effects associated with normal

A.C. driven microscope type illumination.

The membrane and amplifier apparatus were placed in an earthed

steel Faraday cage which reriuced the extraneous electrical noise induced

in the electrical circuitry of the apparatus. All power supplies

operating from 50 Hz mains power supplies were well removed from the

Faraday cage.

49

The Faraday cage was mounted on an antivibration platform which

consisted of two stages in series. The first consisted of a 160 Kg

cement block mounted on six springs resting in a viscous bath of silicon

oil. The second stage was a 10 Kg steel slab which was supported by

four soft rubber balls. The time-constants for the two stages of the

anti-vibration platform were l second and .3 second respectively. This

arrangement greatly attenuated any mechanical vibration that was

transmitted from the floor to the me@brane.

3. 3 A~1PLIFIER ASSEMBLY

3.31 General Design Principles

The construction of the amplifier assembly was designed to overcome

many of the problems inherent in high impedance electrical measurements.

All high impedance current pathways had to be rigidly supported, well

separated, with all insulating surfaces being accessible for easy

cleaning. This was necessary to reduce any capacitive or leak current

paths to ground which could otherwise be significant considering the

small currents flowing through the membrane. The length of all high

im?edance current pathways was kept to a minimum to reduce the

extraneous electrical noise in the circuitry. The amplifier assembly

had to be compact to enable it to be located close to the membrane

apparatus. This reduced the length of the high impedance leads between

the merabrane and the amplifiers.

It was imperative that the design of this apparatus was such that

its electrical contact with the @embrane the electrolyte did not affect

the membrane properties. For example the maximura potential difference

A

A

C B

Figure 3.4. The a,plifier asserbly which houses:

A) the different ial ~lifier

B) the ir,1pedance standards

C) brass temiinals for supporting Ag/AgCl electrodes

50

developed across the membrane had to be less than 30mV in order to

remain in the linear region of the membranes voltage current

characteristics. If the signal greatly exceeded this value, the

membrane potential difference could swing into the non-linear part of

its V-1 characteristics.

3.32 Amplifier Assembly; Mechanical Construction

The amplifier assembly provided an aluminium and Teflon framework

supporting two differential amplifiers (each enclosed in separately

earthed boxes), four brass rods {which held the Ag/AgCl electrodes), a

voltage divider and a network of resistors and capacitors which were

used as standard impedances in series with the membrane. Al 1 high

impedance current paths were supported on Teflon inserts (see figure

. 3.4).

A 5 V (peak-to-peak) sine wave signal, generated by BULFIS, passed

through a voltage divider (mounted at the back of the amplifier frame)

which attenuated the signal to 10 mV. The divider was designed to have

a frequency response that filtered out any high frequency extraneous

noise that may have been induced (by coupling to the rest of the digital

circuitry) in the current line from BULFIS. The current then passed

through the electrolyte and membrane via two rigi;d .. brass rods and

Ag/AgCl electrodes. The return path to ground included the network of

standard resistors and capacitors. The voltage developed. across this

network was used to calculate the current flowing through tte ~mbrane.

51

3.33 Amplifier Electronics

Determination of bilayer substructure from the dispersion of its

impedance with frequency requires high accuracy measurements of

impedances in the order of l0 9 -l0' 0 ohms (±.1% amplitude and ±.02° phase

angle). Therefore it was essential that both of the amplifiers had very

low distortion characteristics over the frequency

range .003 Hz to 10 KHz.

The amplifier circuitry consisted of three stages; an input buffer,

a differential high gain amplifier and an output line driver.

The purpose of the first stage was to reduce impedance of the

current path from 10' 3ohms to 2 K ohms. This was achieved by means of a

FET-input electrometer operational amplifier (unity gain, Analog Devices

AD 515) which was used as an input buffer for the differential

amplifier. Across the input of any FET-input amplifier a small bias

current is present. The electronic design must minimise this as it

could cause significant D.C. potentials to develop across a lipid

bilayer membrane because it normally has such a high impedance. D.C.

offsets of the membrane potential difference had to be avoided to ensure

that the A.C. measurements were made in the linear portion of the V-1

characteristics. The very high input impedance (l0 13 ohms) and ulta-low

input bias current (100 fA) of the AD 515 caused no significant

perturbation in the electrical potential across the membrane.

The second stage of the amplifier employed a precision

instrumentation amplifier (gain 300, Analog Devices AD 521) to amplify

the 10 r.iV differential inj)ut signal. Its high cor,1mon mode rejection

52

ratio and its nearly ideal output characteristics were well suited for

this high gain application.

The output stage used an operational amplifier unity gain,

National LN 355) to drive the low impedance co-axial cable between the

amplifiers and BULFIS.

The circuit diagram of the amplifiers and the important features

of the amplifier input and output characteristics are shown in figure

3.5 and table 3.1.

3.4 COMPUTER CONTROL

3.41 BULFIS Hardware

BULFIS is the Biophysics Ultra Low rrequency Impedance Spectrometer

which consists of an LSI 11 mini-computer, a programmable signal

generator and two transient recorder boards for rapid data acquisition.

The microprocessor initiated signal generation/sampling. Real time was

generated by a crystal controlled clock which provided 50 clocking rates

between 1MHz and .0426 Hz in steps increasing by a factor of 1.416.

Figure 3.6 shows a block diagram depicting the main aspects of the

BULFIS systera.

3.411 Programmable Signal Generator

A sine wave voltage signal was digitally synthesized as follows. A

sine wave table held in random access memory (RAM) was read into a

twelve bit digital-to-analog convertor (DAC). The frequency of the

+ 15v

Rg

INPUT-ve

-15v

100 K.rL

AD 521

EXTERNAL OFFSET

Figure 3.5. The circuit diagram of the differential amplifiers designed to amplify the small voltage signal across the bilayer and standard · series impedances. The arrows indicate the different input and output lines to the circuit board.

TABLE 3.1

AMPLIFIER CHARACTERISTICS

Characteristic

Input impedance

Common mode Rejection Ratio

Linearity

Gain

Slew rate (maximum rate of change of output voltage)

Input bias current

Minimum Requirement

> 10 ohms <10 pF

>60 dB

• 1 %

>106 V/sec

< 1 pA

Amplifier Performance

10 ohms 10 pF

110 dB < 100 Hz .60 dB < 10 KHz

• 1 %

300

5 . 106 V/sec

.03 pA

Table 3.1. The input and output characteristics of the amplifiers used to amplify the small voltage signals produced across the membrane. The table also shows the minimum amplifier performance needed to meet the experimental requirements.

V D.U OUTPUT

LSI II MICRO­PROCES9JR

10 MHz CLOCK

FREQUENCY DIVIDER

MAGNETIC DISC

,--------7 - SIGNAL : GENERATOR I i BOARD 1

I I I 0.A.C I

I .______,,,..------,-----.& I I

~boqram I RA.M. ,~ I

FILTERS I L _____ . __ J

to BLM ---------

;~-=~----=~-----=~---~_._=-~iBLM I I I MEMORY A D. c. I I I I I I I : TRANSIENT( RECORDER I 1 BOARDS 2 OFF} 1 L________________ J

Figure 3.6. A block diagram depicting the signal generation and data acquisition of the Biophysics Ultra Low Frequency Impedance Spectrometer.

53

ensuing sinusoidal signal was determined by the clock ?Ulse rate from

the crystal controlled clock; the basic clock rate being divided by a

series of programmable dividers. The signal was then filtered by a

series of digital filters. The generator had an accuracy of twelve bits

and a linearity of ± .5 of the least significant bit (this was

approximately .012% ). The frequency was accurate to one part in 10 5 •

3.412 Data Acquisition Boards

Two separate phase locked analog-to-digital convertors (ADC)

sampled the amplified voltage signals returning from the membrane and

series impedance. The data was then stored in random access memory

(RAM), signal averaged (at all but the lowest frequencies used, multiple

cycles of the A.C. signal were sampled) on board to reduce any random

noise and then dumped into the computer memory.

3.42 BULFIS Software

A program was written in machine language for the LSI 11

microprocessor for precise impedance measurements. The program

initialized a data file for the impedance data and then called up a

separate data file (frequency file) which contained the information

required to program the signal generator and data acquisition boards.

The computer program gave the signal generator the correct signal ·,

amplitude, frequency, offset and filtering. The appropriate signal was

then generated. Signal sampling was triggered by the program. The

sampling rate and the number of signal periods to be averaged was passed

across to the transient recorder boards. When the data had been

acquired and stored in memory the program used a fast least squares

54

fitting routine to fit a theoretical sinusoidal wave to the voltage and

current data in memory and returned the amplitude ratio, the phase

difference and the normalized fit parameter (NFP) for the data sampled

at each channel.

The NrP was calculated by summing the square of the differences

between all the experimental and theoretical data points, and was

nor~alized by dividing by the sum of the squares of the experimental

data points. Ideally the NFP should be zero, and minimization of this

parameter was used as the criteria for accepting the subsequent

impedance calculations as valid. Consequently the sum of the NFP of

both channels was stored along with the amplitude ratio· and phase

difference between the two channels in the impedance data file. These

parameters were considered representative of the raw data.

The program displayed the corrected impedance data on a visual

display unit. The form of this output and the computer print out of the

data and frequency files are shown in Appendix A.

CHAPTER 4

MATERIALS AND METHODS

4. l INTRODUCTION

4.2 MATERIALS

4.21 Inorganic Solutions

4.22 Organic Chemicals

4.3 MAKING BILAYERS

4.31 BLM Formation

4.32 Membrane Stability; General Observations

4.4 CALIBRATION AND PERFORMANCE OF APPARATUS

4.41 Amplifier Matching and Corrections

4.42 Calibrating for Stray Capacitance

4.43 Calibrating of Impedance Standards

4.44 Overall Performance of the Apparatus

4.5 DATA ANALYSIS, REDUCTION AND PRESENTATION

4.51 Data Presentation

4.52 Data Reduction

4.53 Description of Theoretical Fitting Technique

4.54 Overall Performance of Fitting Technique

55

page

56

56

56

57

59

59

60

61

61

63

64

66

67

67

69

70

71

56

4.1 INTRODUCTION

The experimental work reported in this thesis had three main

facets:-

a) The formation of bimolecular lipid membranes

b) Measurements of the membrane impedance

c) Reduction of data, its analysis and hence extraction of

information concerning the membrane dielectric structure.

Various calibration and performance-testing measurements were

periodically carried out on the experimental apparatus. These tests

were vital in ensuring the validity of the impedance measurements and

section 4.4 has been devoted to a detailed description of some of these

procedures which are particularly relevant to the accurate determination

of membrane structure.

4.2 MATERIALS

4.21 Inorganic Solutions

All salts used in aqueous solutions were A.R. grade. Solutes were

weighed on a Mettler balance (H6T 160 g) with a precision of .lmg.

Water used for these solutions was three times distilled and was

deionized by passing through a ion exchange membrane. Electrolyte

57

solutions of concentrations less than lM were made up by the dilution of

aliquots taken from 1 M stock solutions using a microlitre pipette

(Pipettman P200).

4.22 Organic Chemicals

i) Egg-Lecithin

Lecithin is the generic name for 1-,2-diacylphosphatidylcholine; or

using the structural nomenclature, 3 sn-phosphatidylcholine. This is

one class of phospholipids. The general structural formula for

egg-lecithin is shown in figures 4.1 and 4.2. The relative composition

of egg-yolk lecithin is shown in table 4.1.

Lecithin, extracted from egg-yolk, was obtained from Sigma chemical

company. It was dissolved in chloroform (1 gram in 10 ml) and stored at

-4°C.

ii) Cholesterol

Cholesterol is a compound belonging to a group of crystalline

alcohols known as sterols and is the principal sterol in the tissue of

vertebrates.

Many of the bilayer membranes in this study were formed from

solutions containing oxidised cholesterol (ie. mixture of cholesterol

and its oxidation products). Various molecular models for cholesterol

and some of its main oxidation products are shown in figures 4.1 and

4.3. Oxidised cholesterol (95% pure cholesterol) was donated by

K.Thulborn, School of Biochemistry, University of Melbourne. It was

E

Figure 4.1. Space filled rrodels of dipalmidtoyl lecithin (left) and cholesterol (middle) FraTI Stockton and ~nith ( 1976). Glycerol rronooleate (right). Fran t~hite (1977).

A) a double bond on the acyl chain

B) gauche rotation of a carbon-carbon bond

C) lengt h of the polycyclic ring structure of cholesterol

D) length of the side chain of cholesterol

E) the hydroxyl group of .cholesterol

F) the carbonyl oxygen atan of egg-lecithin

0 CH 3

3 II I -O-P-0-CH -H C-N-CH

I 2 2 + I 3

O_ CH 3

HC 2 - O-C-R 4 2 0

H c'-o-c-R 2 ~ I

0

long chain hydrocarbons

Figure 4.2. The structural formula of 3-sn-phosphatidylcholine (egg-lecithin). Rl and R2 refer to the hydrocarbon chains of the fatty acids which form the hydrophobic pprtion of the molecules. The composition of Rl and R2 is given in table 4.1. The choline-phosphate segment of the molecule is frequently referred to in this thesis as the polar head group (from Ashcroft, 1979).

rro

A B C

Figure 4.3. The structural formula of cholesterol and some of its known oxidation products. The dark circles represent CH 2 groups. (Feiser and Feiser, 1959): 1

A) cholesterol

B) 7-ketocholesterol

C) 7-hydroxycholesterol

OH

TABLE 4. l

Composition and Unsaturation in Acyl Chains of Some Egg-Lecithins

%W/W %W/W (F.A.H.) CHAIN (L.S.)

14:0 0. l

26.2 16:0 36.0

2.0 16: l 1.4

15. l 18:0 14.2

31.9 18: l 35.3

12.2 18:2 9.9

18:3 0.4

5.4 20:4 1.2

4.4 22: 6} 1.4

22:5

r8:4 2.8 20:2

20:5

Table 4.1. The middle column refers to the number of carbon atoms and double bonds per fatty acid chain. The other columns list the percentage by weight of the total fatty acid as presented by,

L.S.: data of Laboratory Supply Pty. Ltd., 1975

F.A.H.: data from Fettiplace, Andrews and Haydon, 1971

From Ashcroft, 1979.

58

stored in a solution of chloroform and methanol (.5g in 10 ml) at -4°C.

Cholesterol (99.9% pure)

Biochemistry, University of

steroid.

iii) Glycerol Monooleate

was a gift from Dr.

New South Wales,

K. Barrow, School of

who synthesized the

Glycerol monooleate (GMO) is a monoglyceride consisting of a

glycerol polar region condensed onto an oleic acid residue. Though

rarely found in living membranes it is commonly used in bilayer

studies. The structural for~ula and space filling model of GMO are

shown in figures 4.1 and 4.4. Glycerol monooleate (>99% pure) was

obtained from both the Sigma Chemical Company and Nu Chek.

iv) Other Reagents

The n-alkanes (>99% pure) were purchased from BDH laboratory

reagents and Sigma Chemical Company.

Benzyl alcohol was donated by the School of Chemistry, University

of New South ~ales.

The p-aminobenzoic acid ester type local anaesthetics were obtained

from Sigma Chemical Company.

The n-alcanols (A.R. grade) were obtained from Sigma Chemical

Company.

CH2 - OH

I CH -OH

CH

Figure 4.4. The structural formula of Glycerol monooleate.

59

v) Preparation of Lipid Solutions

Fifty microlitres of a solution of egg-lecithin and chloroform (ie.

0.5mg of lecithin) were pipetted into a glass vial. The solvent was

evaporated off in a vacuum oven at room temperature for 3 hours. The

solute was then dispersed in .5 ml of the designated n-alkane forming a

solution of 15 m~ lecithin.

4.3 MAKING BILAYERS

4.31 SLM Formation

Bilayers were formed using a modification of the film drainage

method of Mueller et al. (1962). A solution of lecithin in alkane was

extruded from a syringe over a hole in a polycarbonate septum. The

solvent then slowly drained out of the thinning membrane. Interference

of the light reflecting from both membrane-solution interfaces ~roduced

coloured fringes. The spontaneous formation of "black" membrane usually

occurred in 1-15 minutes. However, this could often be initiated

earlier by carefully touching the lipid film with the tip of the syringe

needle. The formation of stable black membranes was found to be

dependent on many factors such as temperature, membrane composition, the

external electrolyte and the author's need to leave the experiment to go

home (invariably this would enhance bilayer stability !). These effects

are discussed in more detail in latter sections. Figure 4.5 shows the

interference pattern of the reflected light from a typical lipid film

during the formation of a lipid bilayer.

Fi gure 4.5. A photograph of a bimolecular egg-lecithin membrane (lower se~i circle) forming from a thick egg-lecithin / n-hexadecar.e film (upper semi circle). The aqueous phase contains lOOmM KCl. Note the presence of particulate matter (probably lipid aggregates) at the interface of the bilayer and the thick lipid film. The graticule used to determine the re lative bilayer area can be seen.

60

4.32 Membrane Stability; General Observations

The formation of bimolecular lipid membranes, that had stable

mechanical and electrical properties for long enough periods of time to

enable successful impedance measurements, was often an arduous task.

The factors found to be important in determining membrane stability were

the following:-

(a) Vibration and electric noise (in excess of 100-150 mV) caused

the rupture of bilayers.

(b) The optimum pH for membrane stability was between 6.5 and 8.

If the pH of the electrolyte was much outside this range the membranes

became less stable.

(c) The temperature at which the bilayers were formed was a

critical factor affecting the membrane stability. Stable bilayer

membranes could be formed best at higher temperatures. At lower

temperatures membranes ruptured during thinning or thinned unevenly,

resulting in membranes with large shunt conductances with an abundance

of "grey" areas and trapped 11 islands 11 of thick film.

(d) The presence of oxidised cholesterol in the membrane forming

solution was found to be a stablizing influence.

(e) The hole cut in the septum to support the membrane had to be a

planar circle with a s~ooth 11 knife edge 11 • This enabled accurate

estimates of the membrane area as well as the promotion of stable

membranes.

(f) The presence of surface active conta~inants (eg.

etc.) acted to reduce membrane stability.

61

detergents

(g) Bilayers formed from solutions with high li~id concentrations

were invariably more stable than those formed from weaker solutions.

4.4 CALIBRATION AND PERFORMANCE OF APPARATUS

It was important to carry out periodic checks on the experimental

apparatus to ensure that the impedance dispersions measured by BULFIS

were due to the membrane substructure and not an artifact of the

experi~1ental design. These checks entailed measuring differences in the

amplifiers' gain and phase response, measuring stray capacitance at the

amplifier inputs and calibrating the impedance standards in series with

the membrane.

4.41 Amplifier Matching and Corrections

All amplifiers have a limited frequency- and phase-response

especially at high frequencies. It was imperative that both amplifiers

be matched such that a common signal at the amplifier inputs produce an

amplitude ratio of one and a phase difference of zero at the outputs.

Any deviation from this ideal behavior was measured using the following

procedure:-

The amplifier inputs were connected in parallel across a common A.C.

voltage signal (see figure 4.6). Any phase, ~c, or amplitude, Ac,

difference between the two output signals could then have arisen solely

1 Krt..

R1 R2

Figure 4.6. The differential amplifiers are shown here connected for the amplifier calibration procedures. At low frequencies the upper configuration is used. However, at frequencies over 3 KHz the lower arrangement is adopted. The input capacitance of each amplifier is represented by the capacitors connected across the amplifier input terminals. Rl and R2 for calibration measurements are no more that Kohm

to ensure that the reactance of the input capacitance of each amplifier is negligible.

62

from small differences the phase and gain response in both amplifiers.

Measurements of the phase difference and amplitude ratio of the two

voltage signals were carried out over the experimental frequency range

using BULFIS operated in an impedance measuring mode. BULFIS then

calculated and filed the correction factors which were used whenever

measurements were done at each experimental frequency. The data at each

frequency were corrected using the following equations.

4.1

A'= R 4.2

Where •c and Ac are the correction factors for phase and amplitude,

•Rand AR respectively are the phase and amplitude parameters

representing the raw data and A~ and ·~ are the corrected raw data

parameters respectively.

The phase and gain response of each araplifier at high frequencies

( >3 Hz) was significantly affected by the presence of the other

amplifier connected in parallel with its input. The µrevious amplifier

configuration was clearly unsuitable for measuring differences in the

amplifiers' phase and gain response at these high frequencies. The

correction factors for phase and gain in the frequency range 3 - 10 KHz

were determined using the following procedure:-

The amplifiers were connected in series across two separate low

valued resistors (R=lK ohm) shown in figure 4.6. Low valued resistors

were used for this purpose because the stray input capacitance of the

amplifiers would not cause any significant phase difference at the

amplifier inputs. Phase angle corrections were then measured directly

63

from the signals at each amplifier output. Provided that the corrected

amplitude ratios at lou frequencies had been accurately measured and the

amplitude ratio of the voltage signals appearing across two series

resistors was independent of frequency indirect measurement of the gain

correction factors could be made at these high frequencies.

Trial-and-error estimates of the gain corrections were given to BULFIS

until a frequency-independent amplitude and phase response at high

frequencies was obtained.

4.42 Calibrating for Stray Capacitance

At the input of any amplifier exists a stray input capacitance

which arises from the electronic design of the amplifiers and the

dielectric properties of the material located between input terminals.

The reactance of this stray capacitance was included in the total

measured impedance of the membrane. It was important that its

contribution to the total impedance be known and appropriate corrections

made. Having matched the phase and gain responses of both amplifiers,

the following procedure was employed to calculate the stray capacitance

(Cl and C2) at the input of each amplifier from the phase difference

between the amplifier outputs.

The inputs of amplifiers l and 2 were each connected across one of

the two series resistors (Rl=lOK ohm and R2=50K ohm). Any phase

difference between the amplifier outputs must have been a consequence of I

the reactive impedance of the stray capacitance at the input of· each

amplifier. The corrected phase difference, •b, between the two output

signals was then measured at selected frequencies and recorded.

Resistor Rl was then replaced with Rl' (Rl '=SOK ohm) and the new phase

difference •o was measured. The parallel stray capacitance at the input

64

of each amplifier could then be derived by solving the following

simultaneous equations.

At a given angular frequency, w:-

$ 1 = w(ClRl - C2R2) D

$ 11 = w(ClRl 1 - C2R2) D

Subtracting equation 4.4 from 4.3 one obtains

$ I $11 Cl = D - D

w(Rl - Rl 1 )

4.43 Calibration of Irapedance Standards

4.3

4.4

Measurements of the phase difference and amplitude ratio of the

outputs of both amplifiers yielded incorrect values for the membrane

impedance only if the impedance standards in series with the membrane,

from which the current was calculated, was not accurately known.

Capacitance standards were ea l i brated ( at l KHz, 20° C) with accuracy

of one ?art in 10 5 on a General Radio capacitance measurement system

(Type 1621) incorporating a capacitance bridge (Type 1616) at the

C.S.I.R.0. Division of Applied Physics. The capacitors, used as

standards, were made from polystyrene and exhibited a

frequency-independent capacitance (within ±.01%).

Maintaining accurately known, high value resistance standards

(10-1000 M ohm) was difficult as their resistance was subject to drifts

65

with time (typically 1% over a few years) and therefore the resistors

had to be calibrated regularly.

When calibrating high resistances on a conventional resistance

raeasuring bridge high power supply voltages (100-500 V) were needed to

obtain adequate sensitivity. However, high valued resistors have a

voltage dependent resistance (varying typically by 10% over 100 V). In

these experiments the standard resistors were subject to very

voltages (approximately 10 mV). Therefore measurements on

resistance raeasuring bridge could only be ~ade at low voltages.

small

the

An alternative "self-calibrating" technique using BULFIS was

developed to measure the high valued resistors at very low voltages

(10 mV). The standard resistor to be calibrated was connected in

parallel with an accurately known standard capacitor. This parallel GC

network was then used as a standard by which the impedance of another

unknown parallel GC was measured. A series of impedance measurements

over the frequency range .01-10000 Hz was then coramenced using this

arrangewent which is shown in figure 4.7. A frequency dispersion in the

calculated value of the equivalent parallel capacitance and conductance

of the ''unknown" network would only occur if incorrect values of the

standard resistor were fed into the BULFIS software. Provided the time

constant of the standard and ''unknown" GC networks were sufficiently

different this method was very sensitive to small errors in estimates of

the standard resistor.

For exaraple, figure 4.7 shows the effect of trial-and-error

estimates of a nominal 10 9 ohm standard resistor. In this case the

correct value of the 10 9 ohra resistor was found to be 919±1 Mohm. In

this way the resistance of an unknown resistor was calibrated against

8-3

Unknown ReS1sfor

Ru

Cs Cu

ACCURATELY KNOWN UNKNOWN STANDARD CAPACITOR CAPACITOR

• 0

t

Rs 917 Mn. 919 Mfl

921 Mfl

\ t t'\. ~ +'-I- •••

--4-,1,,,11 ~ ~~ • I • • • • • • • • ---- • • • • • •·-=•'---? . . ..... . ;++

.01 .1 1 10 1000 FREQUENCY Hz

Figure 4.7. The main features of the resistance calibration technique (see text). The amplifiers were connected as shown above. If the wrong value of Rs was fed into the BULFIS software the measured capacitance of Cu will vary with frequency as shown below. In this case the correct value of Rs was 919 M. The method is very sensitive to errors in Rs when Rs.Ru/Cs.Cu is significantly different from unity. It should be noted that this technique required that the amplifiers have been calibrated and that the BULFIS software has taken account of the input capacitance and relative phase and gain responses of the differential amplifiers.

66

the reactance of an accurately known capacitor. Maintenance of standard

capacitors was not fraught with the same difficulties as that of

resistors. The values obtained by this method agreed (within the

experimental error of ±.1%) with those obtained using a 10 V power

supply on a resistance measuring bridge (Keithly 515A) at the C.S.I.R.O.

Division of Applied Physics.

4.44 Overall Performance of the Apparatus

The corrections for the different phase and gain responses over the

frequency range .001 - 10000 Hz were small. A computer print out of

these corrections at each frequency is shown in Appendix A. The stray

input capacitance of the amplifiers was found to less than 10 pF.

Though the effect of this on the impedance measurements was small (<. 1%)

the appropriate corrections were still made to the raw data. The

capacitance of the septum was less than 50 pF which is therefore also

negligible.

The signal distortion and random noise on the voltage signal

introduced by either the signal generator or differential amplifiers

could be measured using the sine wave least squares fitting routine in

BULFIS. The total noise/signal ratio and relative distortion was found

to be in the range .001-.1%. Any data collected from signals with

distortion or random noise greater than .1% of the signal amplitude were

rejected.

Measurements of impedance could be made over the frequency

.001 Hz to 10000 Hz with a maximum error of .03% in magnitude and

in phase angle.

range

.02°

67

4.5 DATA ANALYSIS, REDUCTION AND PRESENTATIO~

The analysis of the impedance data involved the following

procedures:-

a) The derivation of SLM impedance from the raw data recorded on

magnetic disc during the experiments.

b) Fitting the impedance data to a Maxwell-Wagner dispersion.

Software for this purpose was written in Focal and compiled with

machine language subroutines to run on a PDP ll/03 computer. This was

written jointly with Terry Chilcott at the University of New South

Wales.

4.51 Data Presentation

The frequency data file and the data file containing the raw data

at each experimental frequency was read into the computer memory from

magnetic disc. Once stored in the computer memory the raw data could be

plotted on a visual display unit as any one of the parameters defining

membrane impedance as a function of frequency.

?arameters are now considered.

a) Impedance ratio ZR and phase difference ~D

The nature of these

These are related directly to the corrected raw data. Z can be

calculated from the following:

Z - 1/A' R - R 4.6

68

ZR and •6 are essentially the corrected raw data parameters (albeit

with a few minor corrections) and thus had the advantage of being

closely related to the experimental raeasurements. However, ZR and •6 on

their mm had little physical significance as they are effectively a

comparison between the membranes electrical properties and that of an

arbitrary standard impedance. Thus direct comparison between

~easurements on different membranes using different standard impedances

would be difficult.

b) Total impedance Zand phase angle•·

These variables could be derived form the raw data using the

following equations:-

Z = Z /A' S R 4.7

\/here Z 5 and •s are the phase and magnitude of the standard

impedance respectively. These parameters were an absolute measure of

the electrical properties of the bilayer. Z(w) typically varied by five

orders of magnitude over the experimental frequency range. Most of this

variation was due to the frequency-dependent reactance of the membrane

capacitance with only a minor contribution from the membrane

substructure; that is, these parameters were insensitive to the presence

of dielectric substructure in lipid merabranes.

c) The inpedance expressed as the total equivalent parallel capacitance,

C and conductance, G. These were derived from the following equations:-

C(w)= sin t w z G(w) = cos t z 4.8

Data, when presented in the above form, had several advantages.

Firstly, any frequency dependence of C(w) and G(w) was due only to

59

inhomogeneity in the dielectric medium between the voltage measuring

electrodes. Thus if any dielectric substructure existed within the

bilayer or in the adjacent electrolyte it would have caused a dispersion

in these parameters.

Secondly, the physical significance of the membrane impedance could

be seen at a glance when expressed in these terms. For example the low

frequency capacitance (say 1 Hz) was simply related to the bilayer area

and thickness; and the relative size of the dispersion in capacitance

and conductance was indicative of the deviations from homogeneity in the

bilayer.

Small, time-dependent current shunts frequently occurred through or

around the membrane during the course of impedance measurements. These

introducea bogus frequency dispersions in the total measured

conductance. However, the frequency dispersion in membrane capacitance

remained undistorted by any time-dependent current''leaks~ Therefore to

obtain detail of the membrane substructure it was better to fit a

theoretical ~axwell-Wagner dispersion to the capacitance data than to

the raw impedance data. The theoretical fit to the conductance data was

merely a confirmation of the theoretical fit to the capacitance data.

4.52 Data Reduction

The following procedures were used to prepare the raw data for

theoretical fitting. The impedance data for each frequency scan,

usually expressed in terms of capacitance and conductance, were

displayed on a visual display unit. This was mainly done to chec~ that

the data were collected from bilayers that had stable electrical

properties during the course of data acquisition.

70

All the frequency scans where then normalised to a chosen frequency

scan to correct for small changes in the membrane area that might have

occurred between successive frequency scans. The average membrane

capacitance of each scan, over the experimental frequency range, was

used for calculating the normalization factors for the data. The

computer then automatically rejected those data points which did not

meet the S?ecified criteria (see section 3.42). The data were then

averaged and stored on magnetic disc. Averaged data were then ready for

fitting to a theoretical Maxwell-Wagner dispersion.

4.53 Description of Theoretical Fitting Technique

The theoretical fitting technique was capable of resolving six to

seven dielectrically distinct layers in the impedance data of

egg-lecithin bilayers over the experimental frequency range. The

fitting technique involved one of two alternative procedures.

a) A least squares fitting routine which fitted a Maxwell-Wagner

dispersion to any of the impedance related parameters ~entioned

previously.

b) A manual "trial and error'' method in which estimates of the

dielectric structural parameters were adjusted so that the theoretical

dispersion was best fitted to the data that were presented on the visual

display unit. This was used as a fine adjustment to the theoretical fit

after the least squares fitting routine had been used.

A good fit to the data was achieved when the x squared parameter

for the fit to the capacitance data was minimised. The number of

dielectrically distinct regions used to fit the data was ~rogressively

71

increased until no significant improvement in the fit could be obtained.~

In this way the highest possible resolution of bilayer structure was

obtained from the experimental data.

4.54 Overall Performance of fitting Technique

The data acquisition, reduction and theoretical fitting techniques

were tested for their accuracy and resolution for extracting structural

information of the bilayer. This was done in the following manner.

A series of impedance measurements were rnade over the frequency

range .01 - 10000 Hz on a hard-wire network of a number of parallel GC

elements in series. This simulated the type of substructural dielectric

properties of an egg-lecithin bilayer composed of six distinct layers.

The data were then averaged and fitted to a Maxwell-Uagner theoretical

dispersion in the same manner as was done with the bilayer impedance

data (see figure 4.8). The hard-wire network was dismantled and the

individual components measured on an impedance measuring bridge (\Jayne

Kerr Universal Bridge Type 8 224).

Table 4.1 the shows values of the individual resistors and

capacitors that gave the best theoretical fit to the measured impeaance

data obtained by BULFIS. The values of the individual co~ponents

r:ieasured on the impedance measuring bridge are also given. Figure 4.9

shows the frequency dispersion and Maxwell-Uagner theoretical fit to 1

hard-wire network simulating dielectrically distinct regions with

similar ti@e constants. The circuit elements were chosen so that the

Maxwell-Wagner dispersion of the whole network was similar to that

obtained from lipid bilayers.

lL C

• •• 8

10- 10- 1 10 10 1~ FREQUENCY Hz

1a5

~-m~-~-~10~-~--~;--~10--~h~2~-;-0~1--104 FREQUENCY Hz

Figure 4.8. The total equivalent parallel capacitance (above) and conductance (below) of a hard-wire network of parallel GC elements connected in series. The data presented here is the average of seven frequency scans. The error bars are too small to discern from this graph. The full curve is the Maxwell-Wagner theoretical fit to the data. The values for each of the circuit elements was calculated from the fit to the data. These values are shown in table 4.2.

TABLE 4.2

CG combination Value from (refer to fig 4. 10) fitting routine

2

3

4

5

6

2

3

4

5

Capacitance

8.21

436

530

850

1230

1000

Conductance

513

3100

26000

120000

9500000

nF

nS

Actual value

8.222

490

490

1050

1010

1070

450

3350

30000

95000

9300000

Table 4.2. The values of the individual circuit elements of the circuit shown in figure 4. 10, obtained from the theoretical fit to the data in figure 4.8 are listed in this table.· These are compared to the values determined from measurements made on individual circuit components using a Wayne-Kerr impedance measuring bridge.

72

The ability of the fitting technique to resolve regions of si~ilar

ti@e constant was tested by measuring the impedance dispersion of the

three element, GC, hard-wire circuit. Ins?ection of figure 4.9 shows

that electrical time constants differing by a factor of 3 can be

distinguished from a relative dispersion in capacitance of 2%.

The capacitance of the lowest time constant element was determined

to .1% accuracy. The impedance measurements and fitting technique were

capable of resolving time constants differing by a factor of three-five

from a relative dispersion in capacitance of 2%. It was found in these

tests that fitting a ~axwell-Wagner dispersion to the data yielded the

values of the various circuit elements to an accuracy of 20% for

capacitance and 30% for conductance.

ll. C

8·2

1 10 103

FREQUENCY Hz

Figure 4.9. The total capacitance of a hard-wire network as a function of frequency. The circuit consisted of a series combination of elements with similar electrical time-constant. The solid curve is that expected when the time constants are equal. The data represents the measured capacitance when the time-constants differed by a f1acto'r of 3 (0) and a factor of 1 0 (e).

0

Figure 4. 10. The circuit diagram of the impedance network from which the data in figures 4.8 and 4.9 was obtained.

0

CHAPTER 5

LIPID - n-ALKANE INTERACTIONS IN ARTIFICIAL BLM

5.1 INTRODUCTION

5.2 THEORETICAL CONSIDERATIONS

5.21 Calculation of n-Alkane Concentration in Egg-Lecithin Bilayers

5.22 Thermodynamic Considerations

5.23 The Bilayer Interior: Order and Its Effect on the Partitioning of n-Alkanes

5.3 MATERIAL AND METHODS

5.4 RESULTS

5.41 SLM Capacitance: Temperature Dependence

page

75

79

79

82

83

85

86

87

5.42 BLM Capacitance: Dependence on 87 Torus Alkane Concentration

5.43 The Effect of Cholesterol and Benzyl Alcohol 88 on Bilayer Capacitance

73

5.5 DISCUSSION

page

89

5.51 Interpretation of the Low Temperature Capacitance 89

5.52 Effect of Microlenses 90

5.53 n-Alkane Absorption: Chainlength Dependence 91

5.54 Bilayers Formed From Binary Mixtures of n-Alkanes 93

5.541 Bilayers Formed From Solutions 93 Containing n-Decane

5.542 The Assumption of Ideal Mixing 94 Between the n-Alkane- and Acyl Chains

5.55 Acyl Chain Order and its Effect 96 on n-Alkane Partitioning

5.56 Absorption of n-Alkanes: Interpretation 98 of Teraperature Dependence

5.57 Coraparison With GMO Bilayers 99

5.6 SUMMARY 101

74

75

5.1 INTRODUCTION

Single, planar lipid bilayers, in principle, should accurately

model the bilayer component of living membranes. The method co11111only

used in research on planar bilayers is the one described in this thesis;

namely the lipids are dispersed in a hydrophobic solvent such as one of

the n-alkanes and a film of this solution is established across an

aperture in a septum dividing two aqueous solutions. The bilayer

spontaneously forms as the solvent is expelled from the film into the

surrounding annular region (torus) between the bilayer and the septum

(see Chapter 2).

The bilayers so formed contain varying amounts of alkane solvent

which is presumably in thermodynamic equilibrium with the bulk

lipid-alkane solution of the torus and the lipid in the aqueous

solution.

Naturally occurring membranes probably do not contain any alkane

hydrophobic molecules. Further, the incorporation of solvents such as

the alkanes into living membranes significantly alters membrane

function; acting indiscriminately as l~cal anaesthetics (Haydon, Hendry,

Levinson and Requena, 1977). The presence of varying concentrations of

n-alkanes has also been reported to modulate the conduction properties

of reconstituted membranes containing passive ion pores such as

gramicidin (Hendry, Urb1an and Haydon, 1978).

In the past, differences in the response of multilayer lipid

preparations and single planar bilayers to the addition of molecules

such as benzyl alcohol (BZA) and cholesterol have been attributed in

76

part to the presence of n-alkane solvents in the latter (eg. Ebihara et

al. 1979). Hence the validity of artificial µlanar bilayers, which

contain solvent, as models for the cell membrane, might be questionable.

Therefore the interaction between hydrophobic solvents, such as the

n-alkanes, and lipid bilayers is important to our understanding of this

membrane model.

A systematic study of n-alkane absorption in egg-lecithin bilayers

and its dependence on variations in temperature and alkane chainlength

has been made. In previous studies it was assumed that ideal mixing of

the alkane chains and lipid acyl chains occurs (Haydon et al., 1977 and

Uhite, 1977). To test this assumption the µartition coefficient of the

n-alkanes between the bilayer and torus was measured for different

alkane concentrations in the torus. These measurements were made on

bilayer membranes containing mixtures of more than one alkane as this

effectively extended the µossible range of alkane concentrations in the

membrane torus.

The absorption of alkanes into the interior of artificial planar

bilayers is knmm to increase their thickness (Benz, Froh l i eh, Lauger

and Montal, 1975, ~Jhite, 1977 and Fettiplace, Andrews and Haydon, 1971),

alter their surface tension (White, 1975) and water permeability

(Fettiplace, 1978). By assuming the alkane concentration in the bilayer

is linearly related to the membrane thickness the alkane concentration

in the bilayer could be determined ( Fettiplace et al., 1971 and Uhite,

1377) (See also section 5.2). Studies of alkane absorption in bilayers

formed from monoglycerides such as glycerol monooleate (GMO) have shown

that the concentration of alkane solvent present depends on the

chainlength of both the ~onoglyceride and the alkane components of the

bilayer (Benz et al., 1975 and White, 1977). Uhite (1978) found that

77

bilayers could be formed with negligibly small (equilibrium) solvent

concentrations by the use of hydrophobic solvents that have such large

molecular diraensions (eg. squalene) that they are effectively too big to

fit into the bilayer structure (ie. they are similar in length to the

hydrocarbon chains of the monoglyceride). Independent Raman

spectroscopy studies of Simon, Lis, MacDonald and Kauffman (1977)

support White•s (1978) conclusion. Further it has been found that the

relative solubility of alkane in the bilayer is dependent on temperature

(Uhite, 1977). Decreasing the temperature causes the solvent in the

bilayer to collect into (perhaps frozen) microlenses, whereupon it is

removed from the bilayer proper. This effect has been exploited by

White (1974) as a method for producing essentially solventless GMO

bilayers.

Glycerol Monooleate is a neutral monoglyceride rarely found in

biological membranes. Much of the surface tension and alkane solubility

experiments have been made on artificial bilayers formed frori1 GMO-alkane

solutions. This is mainly because GMO is better defined chemically and

for~s bilayers more easily at physiological salt concentrations than do

natural lipid mixtures. Egg-lecithin bilayers have been studied in

these experiments because they more closely model biological membranes

than bilayers formed from GMO.

In order to account for the behaviour of lipids in bilayers a

statistical mechanical model of the acyl chain conformation of

dipalra.itoyl i)hosphatidylcholine in bilayer aggregates in the liquid

crystalline state was developed by Marcelja (1974). The partition

function was calculated for a single acyl chain located in a molecular

field which modelled the average behaviour of the neighbouring acyl

chains. Gruen (198Oa) refined the raean field model of Marcelja by

78

including explicit terms in the partition function which accounted the

polar group interactions and the energy at the oil-water interface.

Gruen (1980b,1980c) then extended this model to treat the partitioning

of alkanes between a saturated aqueous phase and the bilayer interior.

The results of the model are in good quantitative agreement with

experimental results.

BZA and cholesterol are known to alter the order parameter* of the

acyl chains in the bilayer (Turner and Oldfield (1979) for BZA and

Stockton and Smith (1976) for cholesterol). It was also suggested that

these r.1olecules may induce changes in the partitioning of n-alkanes into

the bilayer interior (Haydon et al. (1977) for Cholesterol and Ebihara

et al. (1979) for BZA). The effect of oxidised cholesterol and BZA on

the absorption of alkanes in the bilayers has been measured as part of

the present study. The results presented in the chapter will be

interpreted on the basis of Gruen's model of the bilayer interior.

The partitioning of n-alkanes (chainlength 10 to 16 carbon atoms)

between the bilayer and torus has been investigated in this study in

order to identify the factors which modulate the solubility of alkanes

in lipid bilayers.

* The molecular order parameter ni at the ith carbon atom of the acyl

chain is defined by the expression:

~- = <-3 cos 2 0 + l> 1

Uhere II e II is the angle spanned by the C-H bonds and the axis

~erpendicular to the plane of the bilayer. The parenthesis indicate

the thermodynamic average. The order parameter is a measure of the

internal entropy of the acyl chains.

79

In this way changes in the physical properties of lipid bilayers

due to changes in alkane absorption may be discerned from those due

other environmental changes which are of greater relevance to living

membranes.

5.2 THEORETICAL CONSIDERATIONS

5.21 Calculation of n-Alkane Concentration in Egg-Lecithin Bilayers

It has been shown that bilayer membranes containing alkane solvents

are thicker than those that are solvent free. Provided the difference

in thickness, ~o, between the solvent free and solvent containing

bilayer arises entirely from the partial molar volume of the alkane in

the bilayer (ie. provided the area density of the lipids is unchanged),

one can calculate the molar concentration of alkane per unit area in the

bilayer, ca, using the following expression:

5. l

l-Jhere "p" and "M" are the mass density and raolecular weight of the

solvent. The volume-averaged mole fraction of alkane with respect to

the acyl chains, Xa, is given by:

5.2

Where c 1 is the number of moles of acyl chains per unit of membrane

area.

80

Previous studies (eg. see Fettiplace, Andrews and Haydon, 1971,

White, 1974 and Fettiplace et al., 1975) have used measurements of

membrane capacitance to measure II C II

a and hence estimate the molar

concentration per unit area in the bilayer,

calculated using the following expression.

l t;

In this study IIC II

a was

5.3

Where "C "' and "C II are the capacitances of the al kane-1 i pi d bilayer m m and the same bilayer in its alkane free state respectively. 11 £ 11 and

0

"c" are the permittivity of free space and relative dielectric constant m

of the hydrophobic region which is about 2.1-2.l (Huang and Levitt,

1977). Equation 5.3 is valid provided that:

(a) The molecular volume of the alkane solvent contributes to the

volume of the hydrophobic region without contributing to the membrane

surface area as an increase in the membrane area per molecule is

energetically unfavorable. Thus the bilayer interfacial area occupied

by each lipid is considered independent of both temperature* and alkane

concentration (see Fettiplace et al., 1975).

(b) The dielectric constant of the alkanes used in these experiments

was in the range 2.02-2.06, which was considered to be approximately

equal (within error of ±2%) to that of the lipid acyl chains. The

temperature dependence of the dielectric constant is negligible; varying

less than ±2% in the temperature range of the experiments (Chemical

Rubber Company Handbook of Chemistry and Physics, 1976).

* Increasing the temperature, if anything, should on entropic grounds,

cause a (small) increase in area occupied per lipid molecule.

81

(cl The alkane volume density is temperature independent in the

temperature range of these experiments,

\~hite, 1975).

to within 3% (eg. see

(d) the error introduced by neglecting the effect of microlenses on the

measured area of these bilayers is sma 11 nJhite and Thompson, 1973 and

also see Discussion).

The thickness of the solvent free bilayer is largely determined by

the ratio of the partial molar volume of the acyl chains of the

molecules to the area of the lipid molecule in the ~lane of the bilayer

(Israelachvili, Mitchell and Ninham, 1976). The former quantity can be

considered constant.

Implicit in the calculation of the alkane mole fraction is the

assumption that the alkane is uniformly distributed throughout the

bilayer interior. However, considerable evidence has been accumulated

which shows that this is may not be the case, and that the alkane

preferentially occupies the region near the midplane of the bilayer; the

area near the bilayer-water interface being effectively inaccessible to

alkanes (eg. see Simon, Stone and Busto-Latorre, 1977, Simon, Stone and

Bennett, 1979, and White, King and Cain, 1981). Our estimates of the

alkane mole fraction do not take into account this heterogeneity and

therefore represent a weighted average of the position dependent alkane

distribution throughout the bilayer interior. Thus the present I

estimates of the alkane mole fraction in the bilayer are only

qualitative.

Previous measurements (Ashcroft et al., 1981 ) shm, that the

capacitance of egg-lecithin bilayers at frequencies of l Hz includes the

82

geometric capacitance of the central hydrophobic layer as well as the

series capacitance of the carboxyl-ester-oxygens on either side of the

membrane. Further, it is known that the capacitance of the Gouy-Chapman

ionic double layers on either side of the· membrane also contributes to

the measured capacitance at l Hz, although this should be significant

only at low ion concentrations in the aqueous phase (LaUger et al., 1967

and Ashcroft et al., 1981). Discussion of the effects of ionic double

layers on the measured bilayer capacitance is not presented here. For

further details the reader is referred to Chapter 6. However, on

inspection of equation 5.3 one can see that any errors introduced by

ignoring the effects of bilayer substructure or ion double layers cancel

exactly when calculating the alkane concentration within the bilayer.

5.22 Thermodynamic Considerations

Provided ideal mixing occurs between the lipid acyl chains and the

alkane one can apply the following thermodynamic analysis to treat the

absorption of n-alkanes in the bilayer interior. (The validity of this

assumption will be discussed latter.)

At thermodynamic equilibrium the difference in chemical potential

of alkane molecules between the torus and bilayer phases must be zero.

From this condition one can relate the standard chemical potential

difference, 6µ 0 , and the partition coefficient of n-alkanes, K, between

the two phases:

6µ 0 = -RT ln K 5.4

Where "T" is the absolute teraperature and "R" is the molar gas

constant. The standard chemical potential difference can be expressed

83

in terms of a difference in the internal entropy, 11 5°, and enthalpy, 11 H~

of the n-alkane. Thus:

5.5

The temperature dependence of 11 11µ 0 " is then given by:

5.6

Provided that 11 115° 11 and "11H 0 " are temperature independent, equation

5.5 and 5.6 reduce to the following well known expressions:

5.7

5.8

5.23 The Bilayer Interior; Order and Its Effect on the Partitioning of

Alkanes

The acyl chains in a lipid bilayer above its phase transition are

in a semi-ordered state. The lowest energy conformation of the lipid

acyl chains is a coil with random orientation like that of the alkanes

in bulk liquid. Order is imparted to the acyl chains by their

attachment at one end to polar head groups which are aligned at the

bilayer-water interface. It can be argued on theoretical grounds that

the acyl chains are straightened to minimize the free energy at the

hydrocarbon-water interface (see section 1.35). Deuterated NMR studies

of Seelig and Seelig (1974) and Stockton and Smith (1976) show that the

order parameter for the acyl chains is nearly constant (at .45) up to

84

the tenth carbon atom of the egg-lecithin molecule but decreases after

this towards the bilayer midplane. This indicates that the acyl chains

of the lipids are well ordered in the the outer .8 nm of the hydrophobic

region, behaving somewhat like a wax, while the central region of the

bilayer is much more disordered, more like an alkane liquid.

The alkanes are chemically similar to the acyl chains. However,

the alkanes have no polar groups and therefore are not anchored to the

bilayer interface and are free to reside wholly within the bilayer.

Gruen (198Gb), when modelling the absorption of alkanes in lipid

bilayers, considered the energy cost of creating free space for the

alkanes and the free energy of mixing of acyl and alkane chains in the

plane of the bilayer.

Gruen's (1980a) modelling of this system predicted two important

factors which affect the alkane absorption and which are both closely

related to the order parameter of the acyl chains (Gruen, 1980b and

1980c). Firstly, in regions of high acyl chain order the alkane chains

are partially constrained to lie parallel to the acyl chains. Thus the

internal entropy of the alkane molecules is much lower in these regions

than in regions of low acyl chain order. Secondly, creating space for

alkane requires a straightening of the acyl chains and hence a reduction

in the area per acyl chain (as the area per lipid is constant). In

regions of high order a change in the area per acyl chain involves a

greater increase in free energy of the lipid molecules than in regions 1

of lower order. Therefore regions of higher order in the bilayer are

relatively hostile to the presence of alkanes.

In the light of this model, it should be expected that membrane

85

additives which alter the order of the acyl chains of the lipids should

also alter the concentrations of n-alkanes in the bilayer.

5.3 ~-IATERIALS AND METHODS

Black lipid membranes were generated from solutions of egg-lecithin

and oxidised cholesterol dissolved in n-alkanes ( 15 mM with respect to

lecithin) with chainlengths between that of ten carbon atoms (n-decane)

and sixteen carbon atoms (n-hexadecane). Membranes were generated at 0

temµeratures between 25-45 C (depending on the alkane solvent) with a

lml-1 KCl in the aqueous phase. Bilayer formation occurred spontaneously.

Hm'lever, quite often it was initiated pre;naturely by touching the thick

lipid film with the tip of a syringe.

In order to modulate the mole fraction of a particular n-alkane in

egg-lecithin bilayers (at constant temperature) it was necessary to form

bilayers from lipid solutions over a wide range of lipid concentrations.

However, using a single alkane solvent it is only possible to generate

stable membranes from lipid solutions over a narrow range of

solvent-liµid concentrations. Formation of bilayers from solutions

containing low solvent concentrations could not be achieved because of

the limited solubility of lipids in the solvent. This problem was

overcome by dissolving the lipids in a long chainlength alkane (eg.

hexadecane) which, at tem~eratures less than 35°C, is essentially

excluded from the bilayer phase. The effect of low concentratio~s of

shorter chainlength alkanes could be studied by adding small amounts to

the lipid-hexadecane mixture.

86

In these studies the bilayer and torus were assumed to be in

equilibrium when the capacitance of the membrane had attained a steady

value (varying less than 1% in 15 minutes). All measurements reported

here were on bilayers which had been allowed to come to thermodynamic

equilibrium with the torus and were bimolecular (ie. 'black') over the

entire aperture in the septum. The area, excluding the torus, was

determined (to an accuracy of ±2%) using a graticuled eyepiece mounted

on a 40x microscope.

Care was taken to ensure that the bilayer remained flat during

measurements of bilayer capacitance. The membrane was kept flat by

periodically adjusting the hydrostatic µressure across the film by

adding appropriate amounts of distilled water to the aqueous phase on

one side of the membrane. The progress of unbowing the membrane was

monitored visually under reflected white light.

5.4 RESULTS

Stable egg-lecithin bilayers could only be generated at

temperatures where significant amounts of alkane solvent would remain in

the bilayer after formation (at least l : 10 mole ratio alkane:

egg-lecithin). Only after the film had become bimolecular over the

entire aperture in the septum could the temperature be lowered. Upon

lowering the temperature solvent left the bilayer phase of the membrane

apparently aggregating into microlenses. In this case I am referring to

small droplets of liquid alkane visible as isolated ''pinpoints" of high

reflectance under the viewing microscope. The membrane would usually

attain a steady capacitance within 15-20 minutes of a change in

temperature.

87

5.41 BLM Capacitance; Temperature Dependence

The specific capacitance of egg-lecithin bilayers was found to

decrease with decreasing temperature. It can be seen from figure 5. 1

that the capacitance of bilayers generated with the longer chainlength

alkanes (C 1 ~-C 16 alkanes) appeared to have approximately the same upper

limiting value of capacitance at low temperatures.

The bilayer capacitance decreased with increasing temperature at

low solvent concentrations and then asymptotically approached a lower

li~it of approximately 3.5 ±.3 mF/m 2 • From the results shown in

figure 5.1 it is clear that the bilayer thickness (and hence alkane

solubility) was reduced with increasing chainlength of the alkane

present.

The upper limit of membrane capacitance at low temperatures was

repeatable within ±2%. However, the temperature dependence of membrane

thickness at higher temperatures showed greater variation between

different membranes. The magnitude of the experimentally measured

scatter in the results can be seen in figure 5.2.

5.42 BLM Capacitance: Dependence on Torus Alkane Concentration

Egg-lecithin bilayers were formed from solutions in which a given

alkane molecule could be varied over a wide range of concentrations (see

methods). I

Egg-lecithin bilayers formed with a mixture of n-decane and

n-hexadecane were found to have no well-defined, stable, capacitance.

The bilayer capacitance varied between 3.7 mF/~ 2 and 5 mF/m 2 over a

period of 2 hours. This long-term time dependence in bilayer

65

---------------.ou-.a._._ __ -•.&.-..-----------------------6

• • oooo ._ • • C 16 • 0 • • 0 •

"' D • 0 C 15

~ 0

D C 11. D • Li.. 55 • 0

E: D D C 12

Lu • 0

u <: 5 0 ~ D

u ~ • 0 D

4.5 • D

D

10 20 30 t.O 50 60

TEMPERATURE oc

Figure 5.1. The capacitance, measured at lHz, of representative bilayers in equilibrium with egg-lecithin solutions containing different chainlength alkanes at different temperatures. The horizontal dashed line is the low temperature upper limit to membrane capacitance which is common to all the egg-lecithin bilayers formed from C1 ~-C 16 alkane solutions in lmM KCl over the temperature range employed here.

6 t t ~ t t t t

5

('\j

~ • LEC I.!.. E

0 LEC: CHOL Lu 2: 1 u ~

~I. --u cl ~ u

20 25 30 35 1.0 1.5 50 TEMPERATURE o C

Figure 5.2. The capacitance of lipid bilayers in equilibrium with n-tetradecane solutions of egg-lecithin (•) and egg-lecithin-oxidised cholesterol (2:1 mole ratio) (0) in l mM KCl. The error bars refer to the variation in the measured capacitance of ten egg-lecithin bilayers and five egg-lecithin-oxidised cholesterol bilayers. Note that the scatter·increases with decreasing membrane capacitance.

88

capacitance was not present in bilayers formed from longer chainlength

alkanes such as n-dodecane.

The capacitance of bilayers in equilibrium with different mole

fractions of n-dodecane in the torus, as a function of temperature, is

shown in figure 5.3. It was found that a decrease in the mole fraction

of n-dodecane in the torus caused an increase in the temperature

dependent bilayer capacitance.

5.43 The Effect of Cholesterol and Benzyl Alcohol on Bilayer

Capacitance

The incorporation of oxidized cholesterol increased membrane

capacitance at all temperatures investigated. The presence of

cholesterol in the membrane forming solution increased the capacitance

of egg-lecithin - n-tetradecane bilayers over the entire temperature

range (see figure 5.2). The upper limit (ie. at lower temperatures) of

the capacitance of bilayers formed from n-hexadecane solutions increased

from 6.15 ±.1 mF/m2 (without cholesterol) to 6.4 ±.1 mF/ra 2 (50% mole

fraction of cholesterol) in lmM KCl.

The addition of 10-30 mM BZA to the external electrolyte induced a

dramatic decrease in the capacitance of bilayers formed from

n-tetradecane and n-hexadecane solutions. Inspection of figures 5.4a

and 5.4b reveals that the plot of the te~perature dependent membrane

capacitance is dispiaced to lower temperatures as the BZA concentration

in the electrolyte is increased. However, it was found that the upper

limiting value of the bilayer capacitance at low temperatures was

unaffected by the presence of BZA.

(\j

E

"

6.5

6.0

5.

u. 5.0 E

•---•---. ---. ---. -----o -----o----- • o_ o 0 -·---- ---::_ __ _

·-·-· 0 - 0 0·1 ---· -o __ _

0---00 ---..........

~o ·~ o

~ -~ 0 -~

~ . 0---------0

~ 0·25

• 0

0 -- oO·S

0 0·8

1·0

20 25 30 JS t.O t.5 TEMPERATURE °C

Figure 5.3. The capacitance of egg-lecithin bilayers in equilibrium with solutions containing various mixtures of n-hexadecane and n-dodecane in lmM KCl. The numbers at the right hand side of the graph indicate the mole fraction of n-dodecane in the membrane forming solution.

t t 6

5

20

t BARE

t"'

30 TEMPERATURE °C

t~f~

,,"'

lO

Figure 5.4a. The capacitance of representative egg-lecithin bilayers in equilibrium with n-hexadecane solutions in lmM KCl at different concentrations of benzyl alcohol in the aqueous phase. Th~ error bars shown here indicate the errors arising from the uncertainty (2%) in calculating the membrane area.

(O) bare (OmM BZA)

(e) lOmM BZA or 30mM BZA

\J

E

" lJ.. E

Lu l)

< ~ -l) ~ <:( l)

6 -t t ------------ --------------- t BA R E

--- t ~ 5

t~ "'f 30 rrM BZA

~t~ 4

t 20 30 40

TEMPERATURE oc

Figure 5.4b. The capacitance of representative egg-lecithin bilayers in equilibrium with n-tetradecane solutions at different concentrations of benzyl alcohol in the aqueous phase.

89

5.5 DISCUSSION

5.51 Interpretation of the Low Temperature Capacitance

The different chainlength alkanes had different, temperature-

dependent, effects on the hydrophobic region capacitance. At

sufficiently low temperatures these differences vanished so that the

capacitance (and hence thickness) of these egg-lecithin bilayers was

independent of the temperature and alkane chainlength for the longer

chainlength alkanes. This implies that the concentration of the longer

chainlength alkanes in the bilayer at these low temperatures must be

very small. It seems likely that the capacitance of bilayers containing

n-dodecane would continue to follow a similar dependence on temperature

as the longer chainlength alkanes and would have had the same limiting

low temperature values at temperatures below the lowest values that

could be employed in the current experiments (see figure 5.1).

Based on these data it is now assumed that bilayers containing

n-hexadecane in 1 mM KCl external electrolyte are essentially solvent 0

free* at temperatures up to 30 C. Thus solventless egg-lecithin

bilayers were produced by generating the bilayer films at elevated

temperatures and then reducing the temperature which lowered the alkane

concentration. This technique is analogous to the 11 freeze-out 11 method

of White (1974). The analogy, however, is only superficial as the

mechanisms for removal of solvent in egg-lecithin bilayers and GMO

bilayers may be different (see section 5.57).

* Solvent free only in that the bilayer thickness was unaffected

(within an experimental error of ±2%) by the possible presence of

trace amounts of solvent within the bilayer.

90

The capacitance of egg-lecithin bilayers formed by this method in

100 mM KCl was 6.8± .2mF/m 2 at 25°C (see results in Chapter 6) which

compared favourably with capacitance of bilayers formed by monolayer

apposition (7.21 ± .2~F/m 2). The latter are believed to have negligible

solvent concentrations as determined by the effect of D.C. voltage bias

on membrane capacitance (Benz et al., 1975).

Therefore the upper limiting value of bilayer capacitance at low

temperatures is assumed to be the capacitance of the solvent-free

bilayer which represents the variable, C , m

difference in the standard chemical potential,

in equation 5.3.

0 ~µ' in the bilayer

The

and

torus for different alkanes could then be calculated on the basis of

equation 5.4 and assuming that the alkane mole fraction in the torus is

unity*.

5.52 Effect of Microlenses

White (1974) observed that alkane which was displaced from the

bilayer upon changes in the bilayer-torus equilibrium, collected into

small lenses of bulk alkane called microlenses which were clearly

visible under a low power microscope. This phenomenon was also observed

in the present experiments. Being much thicker than the surrounding

bilayer, the lenses of trapped solvent contribute little to the

capacitance of the bilayer. If the microlenses do occupy a significant

fraction of the membrane surface area then this would introduce errors

into the estimates of bilayer thickness derived from the total membrane

capacitance.

* The lipid concentration in the torus was always very small ~ 30mM).

The total alkane mole fraction was then always greater than 99%.

91

A detailed study of the effect of microlenses on estimates of the

bilayer area has been carried out by White and Thompson (1973). It was

found that the total area occupied by microlenses depended on the amount

of solvent disproportioned from the bilayer and the size of the

microlenses formed. The calculations of Uhite and Thompson (1973)

sho\1ed that in extreme cases microlenses could occupy 10% of the bilayer

area.

Inspection of figures 5.2 and 5.5 shows that the scatter in the

~embrane capacitance between different membranes ( for a given solvent)

increased with increasing alkane concentration in the bilayer. The

scatter in the bilayer capacitance increased from 2%, when no solvent

was present, to a maximum of 20% which was similar to that reported by

White and Thompson (1973) for bilayers containing n-decane.

If variations in the area of raicrolenses between different

membranes were responsible for the experimentally observed scatter then

large variations in the upper limiting capacitance would also be

observed. Therefore it can be concluded that microlenses only had a

s@all effect on the total capacitance of the membrane ~2%). The

experimental scatter must then arise from variations in the alkane

concentrations within the bilayer interior.

5.53 n-Alkane Absorption: Chainlength Dependence

1 The scatter in the experimentally measured alkane mole fraction in

the egg-lecithin bilayers was approximately ±25%. The reason for the

considerable variability is not known. One possibility is that the

concentration of solvent in the torus is not well defined and that

variations in the torus concentration of different membranes can lead to

0·6

0·4

0·2

I I

I I

I I

I I

I

I

0·2

I

I I

I

I

I /

I

0·4

I

/ /

/

/ /

/ /

0·6

/ /

/ /

-/ /

/ /

/ /

20°c

0·8 1 ·0

Figure 5.5. The mole fraction of n-dodecane in egg-lecithin bilayers (Xb) plotted against the mo!e fractign of n-dodecane in the membrane forming solution (Xt) at 20 C and 40 C. The data presented in this figure was calculated from the capacitance data in Figure 5.3. The dashed line is the mole fraction of n-dodecane near the bilayer midplane, calculated from the data at 40°C, assuming that the dodecane was only distributed throughout 50% of the bilayer interior. Note the relation between Xb and Xt is essentially linear up to the values of Xb or approximately .4.

92

variations in the alkane mole fraction in the bilayer; an effect

reported previously in GMO bilayers by Waldbillig and Szabo (1978).

In spite of the significant experimental scatter, the chainlength

dependence of alkane absorption in egg-lecithin bilayers was found to be

quite significant and was consistent with results reported by Fettiplace

et al. (1971) in egg-lecithin bilayers and was also similar to that

reported by White (1977) and Benz et al. (1975).

Gruen (1980b) found that the free energy cost of putting an alkane

molecule into the disordered central part of the bilayer is quite low;

as it would be if it were mixing with an oil. However, if the same

solvent QOlecules were to transfer into the more ordered outer region of

the membrane they would have to loose much of their internal entropy as

~ell as creating energetically unfavourable conformations of the

hydrocarbon chains of the alkane and lipid molecules. This more ordered

region of the bilayer would be effectively inaccessible to any alkane

solvents. Therefore, short chainlength solvents such as n-hexane would

be able to partition into the hydrophobic region near the bilayer

midplane. Longer chainlength alkanes, being partly constrained to be

parallel to the acyl chains, would have a portion of their structure

located in the ordered outer part of the hydrophobic region of the

bilayer and so would not be able to partition into the bilayer as

easily. The more the alkane has to penetrate the ordered region in

order to be accommodated within the bilayer, the greater the standard

chemical potential difference, 6µ 0 , will be for that alkane between the

bilayer and torus.

From figure 5.6 it can be seen that the chainlength dependence of

11 6µ°'• is very pronounced; increasing by 2. 5 KJ/M for each add it i ona 1

10

30° C

8

I

6

2

0 '--------;:!;--------:--l;;::----------:+-----,--,!---------1--2 13 1l 15 16

n-ALKANE CHAIN LENGTH ( N'? CARBON A70MS)

Figure 5.6. The difference in the standard chemical potential (6µ 0)

between alkanes in the bilayer and torus phases of the membrane for alkanes of different chainlengths, at 30°C. 6µ0 was calculated from the data shown in Figure 5.1 using equations 5.3 and 5.4.

93

carbon atom in the alkane chain. The l O\'I values of "e-,·/" for n-dodecane

show that alkane near the midplane of the bilayer contributes very

little to the "e-,µ0 " of the longer alkanes. This suggests that the main

contribution to the standard chemical potential of the alkanes in the

bilayer arises from the terminal carbon atoms of the longer alkanes.

Thus it seems that the partitioning of n-alkanes into the bilayer would

be sensitive to the order parameters of the acyl chains near the

bilayer-water interface.

5.54 Bilayers Formed From Binary i4ixtures of n-Alkanes

5.541 Bilayers formed From Solutions Containing n-Decane

!~early all of the earlier bilayer work was carried out on

egg-lecithin bilayers formed from n-decane solutions. As found in many

previous studies, bilayers formed from n-decane solutions had no well

defined capacitance (eg. Andrews and Haydon, 1968 and White and

Thorapson, 1973). In these studies it was suggested that this was due to

a time varying disproportioning of n-decane into ~icrolenses. In the

present experiments it was apparent that n-decane was sufficiently

volatile to affect the time course of the capacitance of bilayers formed

from this solvent; an effect already known in bilayers containing

n-nonane (Haydon et. al., 1977). Thus the. bilayer thickness never

reached a stable value in the life time of the membrane (typically 2

hours). Therefore the bilayers containing n-decane must be treated as

~hree phase systems (ie. the bilayer, the torus and the atmosphere) and

precautions would need to be taken to ensure that the aqueous phase and

adjacent atmosphere are in equilibrium with the n-decane in the bilayer

(for the atmosphere this represents substantial technical problems).

The fact that such precautions were in general not ta~en may account for

94

the considerable variation in bilayer capacitances reported in the

literature for apparently identical bilayer systems using n-decane

solvents (eg the capacitance of egg-lecithin - cholesterol bilayers

containing n-decane; c.f. Hanai, Haydon and Taylor, 1965b and Haydon et.

a 1 . , 1977) .

5.542 The Assumption of Ideal Mixing Between the n-Alkane- and Acyl

Chains

The maximura thickness of egg-lecithin bilayers saturated with

solvent was estimated at about 4.7-5.Snm* which is approximately twice

the extended chainlength of a typical lipid in the bilayer each acyl

chain having an average chainlength of 17 carbon atoms). Indeed for a

bilayer to attain a greater thickness, a new, "bulk" alkane phase would

need to form between the two apposing monolayers. Formation of such a

phase is unlikely** as this would be not lead to any further increase in

entropy-due-to-mixing of the lipid acyl chains and the alkanes, since

any additional alkane would then partition into a separate phase.

For further details the reader is referred to the discussion in

the previous studies of Taylor and Haydon (1966) and White (1970).

* These values are slightly over estimated as the capacitance of ionic

double layers in series with the dielect;ic capacitance of the

bilayer was not taken into account.

** In any case the spontaneous formation of a bilayer from a thick film

shows that on the basis of free energy the formation of such a

central phase is unfavourable.

95

The fact that there is an upper limit to alkane absorption in

egg-lecithin bilayers indicates that the mixing of n-alkanes with the

lipid acyl chains is non-ideal. If this be so then the thermodynamic

analysis presented earlier in this chapter as well as in some previous

studies (eg. Haydon et al., 1977 and White, 1977) would not be

applicable to this system.

To check this the partition coefficient of n-dodecane was measured

as a function of its mole fraction in the bilayer by varying the alkane

mole fraction in the torus at constant temperature. Results for this

are shown in figure 5.5. The partition coefficient of n-dodecane

between the bilayer and torus was independent of n-dodecane mole

fractions in the bilayer up to 40%. After this the partition

coefficient decreased with increasing concentrations in the bilayer.

Hence for alkane mole fractions in the bilayer less than 40% the

assumption of ideal ~ixing of acyl chains and alkanes is valid.

The decrease in 11 K11 at high alkane mole fractions in the bilayer

can be accounted for if one recalls that the alkane molecules are not

distributed uniformly in the bilayer. Gruen (1980c) calculated that at

low alkane concentrations the alkane molecules are distributed fairly

uniformly throughout the hydrophobic interior since their mixing entropy

dominates the unfavourab~e terms mentioned earlier. At higher alkane

concentrations the n-alkane distribution is non-uniform. The alkane

mole fraction approaches unity near the bilayer mid~lane whereas that

near the bilayer-water interface is low. A further decrease in the

entropy-of-mixing of alkane and acyl chains could only occur if

additional alkane partitioned into the regions near the bilayer surface

(where "ti/"is higher). In this way the average partition coefficient

over the bilayer interior decreases.

96

5.55 Acyl Chain Order and its Effect on n-Alkane Partitioning

Gruen's @odel predicts that alkane absorption in lipid bilayers

should increase with increasing entropy of the lipid acyl chains.

C~olesterol and benzyl alcohol, which are known to alter the acyl chain

order ~arameter in multi-lamellar dispersions had a pronounced effect on

the partitioning of alkanes between the bilayer and torus. The

interpretation of these effects are now considered separately.

i) Cholesterol

The rigid ring structure of cholesterol hinders the random raovement

of the adjacent acyl chains thus creating a structure of increased

order. Measurements of the order parameters of the acyl chains of the

lipids in bilayers containing cholesterol confirm this (Stockton and

Smith, 1976). In these latter studies it was found that cholesterol

(30% cholesterol : lecithin mole ratio) increased the order parameters

of the first thirteen carbon atoms of the lipid acyl chains from .45 to

.8, with a decreasing effect on the order parameters further down the

acyl chains. According to Gruen (198Ob) this increased order would

effectively exclude alkanes from a greater proportion of the hydrophobic

region than vhat would be the case without cholesterol. This is

consistent with the reduction in the average ~ole fraction of alkane in

the bilayer (see figure 5.8). Similar effects have been reported I .

elsewhere for multilayer ( \Jhite et al., 1981) and vesicle preparations

of lipids (Simon et al., 1977).

One would expect that if cholesterol caused a straightening of the

acyl chains in the bilayer an increase in the bilayer thickness would

97

result; hence a decrease in the bilayer capacitance. The results

presented in this chapter show just the opposite. However, no account

of the impedance of the ionic double layer capacitance has been made.

In Chapter 6, where the effect of ionic double layers on membrane

capacitance is measured, the thickness of the bilayer was found to

marginally increase with the addition of cholesterol (see Table 6.1).

It should be noted here that the mole fraction of cholesterol in

the bilayer phase of egg-lecithin membranes is less than that of the

~embrane forming solution. The partition coefficient of cholesterol

between the bilayer and torus egg-lecithin bilayers formed from

equi-molar solutions of lecithin and cholesterol is in the range .2-.5

(Bunceand Hider, 1974). In fact, evidence suggests that the· maximum

mole fraction of cholesterol in egg-lecithin bilayers is 60% (Reiber,

1978). Therefore the dependence of alkane absorption on the cholesterol

mole fraction in the bilayer may be more sensitive than the results in

figure 5.7 would suggest if cholesterol partitioned equally between the

bilayer and torus.

ii) Benzyl Alcohol

The absorption of BZA into egg-lecithin bilayers was found to

decrease bilayer capacitance (ie. increase bilayer thickness). The

increased thickness was not due to BZA absorption into the bilayer

interior as the high dielectric constant of this molecule(£ =13) would , r

have significantly increased bilayer capacitance. Thus it can be

concluded that BZA caused an increase in the alkane absorption into the

bilayer interior.

O ·L.

0·2

0 0 0·2 0·4 0·6 0·8 1·0

MOLE FRACTION OF CHOLESTEROL

Figure 5.7. The mole fraction of n-dodecane in egg-lecithin bilayers in equilibrium with solutions containing different mole fractions of cholesterol (with respect to lecithin). The concentration of ' egg-lecithin was adjusted so that the cholesterol concentration in the alkane never exceeded 15mM. The error bars represent the variation over 3 to 6 different membranes.

98

Benzyl alcohol is an amphiphilic molecule and hence is surface

active. Absorption of benzyl alcohol at the bilayer interface is likely

to reduce the energy arising from the oil-water interface by reducing

the area of oil-water contact. This would then allow a slight increase

in the lipid head group area and consequently a reduction in the acyl

chain order parameter. This suggestion is supported by the results of

NMR experiments reported by Turner and Oldfield (1979).

The minimum low temperature limiting thickness of egg-lecithin

bilayers was unaffected by the addition of BZA. This is surprising

since one would expect a change in the order parameter of the acyl

chains in the bilayer to induce a change in bilayer thickness. Either

changes in the order parameter of the lipid acyl chains are insufficient

to produce detectable changes in bilayer thickness or alkane exists in

the bilayer which does not "freeze out" at the temperatures employed in

these experiments.

5.56 Absorption of n-Alkanes: Interpretation of Temperature Dependence

Figure 5.8 shows the temperature dependence of the difference in

the standard chemical potential, ~µ 0 , between the bilayer and torus for

different alkanes. The temperature dependence increased with increasing

chainlength of the alkane solvent. At higher temperatures the

temperature dependence of "~µ0" appeared to vanish ( at least for the

shorter chainlength alkanes).

Provided the entropy and enthalpy components of 11 ~µ011 are

temperature independent then one can use eGuations 5.7 and 5.8 to

evaluate the relative roles of entropy and enthalpy in the partitioning

of n-alkane between the bilayer and torus phases. The results presented

99

here suggest that "tiµ0 " is dependent on the ordering of the 1 i pi d acyl

chains in the bilayer. Deuterated lipid NMR studies (Stockton,

Polnaszek, Tullock, Hasan and Smith, 1976) on egg-lecithin multilayer

preparations showed that the order of the acyl chains is temperature

dependent; the order parameters of the first eight carbon atoms of the

chain decreasing from .46 ± .01 at 30°C to .37 ± .03 at 55°C.

As the ordering of the bilayer interior varies with temperature it atiH 0 0

unlikely " a tiS " t zero in equation 5.6 is is that setting 11 ..:::..:::.:.:.• and al 0 a aT

good approximation. Hence caution must be used when interpreting the

gradient and intercept of data presented in figure 5.8 on the basis of

equations 5.7 and 5.8.

5.57 Comparison With GMO ailayers

The temperature and alkane chainlength dependence of alkane

solubility was similar for GMO and egg-lecithin bilayers. However, the

alkane solubility in egg-lecithin bilayers was found to be lower than

that observed for GMO bilayers. The mole fractions of the

various n-alkanes in egg-lecithin bilayers reported in the present study

are compared to those found in GMO bilayers by Hhite (1977)

5. 1.

in Table

Though the extended chainlength of the acyl chains of GMO and

egg-lecithin are similar, capacitance measurements indicate that the

acyl chains of egg-lecithin in bilayer aggregates have a more extended

conformation (compare results in figures 6.1 and 6.2). This means that

the acyl chains of GMO are more disordered and thus provide an

environment more favourable to the presence of n-alkanes. This could

explain the relatively large alkane solubility of GMO membranes.

10

a., - 8 0 E ' -, ~ 6

~

C __,

I-4

0::: I

2

C16

C14

C 12

10 20 30 40 50

TEMPERATURE °C

Figure 5.8. The difference in the standard chemical potential between the bilayer and torus for different chainlength alkanes at different temperatures. The heavy central lines indicate the typical temperature dependence of 6µ 0 for a given bilayer. The shaded area represents the scatter over 5-10 different membranes. The main variation in the temperature dependence of 6µ 0 between different membranes was due to variation in the intercept rather than the slope of the temperature dependence.

60

TABLE 5.1

Alkane GMO Egg-lecithin chain length

16 .34

15 .44 • 11

14 .52 . 19

12 .56 .44

Table 5.1. The mole fraction on different chainlength n-alkanes in bilayers of GMO (100 mM KCl) and egg-lecithin (1 mM KCl) at 30°C. The GMO data was calculated from the data in White (1977).

100

The mechanism for the removal of solvent from the interior of

egg-lecithin bilayers at low temperatures may be different from that

proposed by White (1974) for Gi40 bilayers. ~Jhite interpreted the

removal of solvent from the bilayer as a freezing effect implying that

it is an intrinsic property of the alkane solvent. The temperature at

\lhich the solvent "condensed" out of egg-lecithin membranes was well

above the freezing point of the alkane present and was also found to be

dependent on the bilayer composition. While alkane in egg-lecithin

bilayers could also freeze at low temperatures, the effect could not be

observed because the partition coefficients of the alkanes in the

bilayer were very low at temperatures in the vicinity of their freezing

point.

101

5.6 SUMi>lARY

The partition coefficient of n-alkanes between the bilayer and

torus of egg-lecithin bilayers was measured for n-alkanes with ten to

sixteen carbon atoms using measurements of membrane capacitance.

The partition coefficient was found to decrease with increasing

alkane chainlength and increase with increasing temperature.

It was found that n-decane was unsuitable as a solvent in these

experiments as the partitioning of n-decane into the aqueous phase and

atmosphere could not be ignored and could not be controlled.

Egg-lecithin bilayers containing negligible amounts of solvent

could be produced using a method similar to the freeze out method of

White (1974). Bilayers formed using n-hexadecane were found to be 0 solvent free at temperatures below 30 C.

The partition coefficient of n-alkanes in the bilayer was found to

depend on the alkane mole fraction. Thus it was concluded that the

assumption of ideal raixing between acyl- and alkane chains in the

bilayer was not valid when the alkane mole fraction exceeded 40% (with

respect to the acyl chains of the lipid).

Membrane additives knmm to alter the order pararaeter of the acyl

chains in lipid bilayers had pronounced effects on the partitioning of

alkanes into egg-lecithin bilayers. Cholesterol, known to increase the

order para~eter, decreased the partition coefficient and benzyl alcohol,

known to decrease the order parameter, increased the partition

coefficient. This was found

statistical mechanical model of

crystalline state.

102

to be consistent with a well-known

lipid-alkane bilayers in the liquid

The variation of the standard che~ical potential with temperature

was Qeasured for alkanes of different chainlengths. From these results

it Has concluded that the enthalpy and entropy of the alkanes in the

bilayer are in themselves a function of temperature. This is indicative

of the different state of the hydrophobic interior of lipid bilayers at

different te~peratures.

Thus far, solvent retention in lipid bilayers has been considered a

major pitfall of the model. However, it seems that the partitioning of

n-alkanes into lipid bilayers may be a useful probe in detecting small

variations in the ordering of the acyl chains in the hydrophobic

interior of lipid bilayers.

CHAPTER 6

EFFECT OF EXTERNAL ELECTROLYTE ON THE

CAPACITANCE OF LIPID BILAYERS

6. l INTRODUCTION

6.2 PRELIMINARY THEORETICAL CONSIDERATIONS

6.3 METHODS

6.4 RESULTS

6.41 Glycerol Monooleate Bilayers

Page 104

107

111

111

111

6.42 Egg-Lecithin Bilayers 112

6.43 Egg-Lecithin Bilayers Containing Cholesterol 113

6.5 DISCUSSION 114

6.51 Effect of Ions on Bilayer Structure. 114

6.52 The Capacitance of Ionic Double Layers 115

6.53 The Nature of Bound Charge on Lipid Bilayers 118

6.54 The Effect of Cholesterol on Bilayer Capacitance 121

6.55 Comparison With Previous Work 122

6.6 SUMMARY 123

103

104

6.1 INTRODUCTION

Biological membranes are involved in many aspects of cellular

activity. It is now recognized that the function of these membranes is

sensitive to the composition of their aqueous environment; particularly

to the presence of monovalent and multivalent cations ( eg. see Cole,

1968, Hope and Walker, 1975). Consequently many studies have been made

on model membrane systems in an effort to understand the mechanism

whereby alkali metal and alkaline earth cations can modulate membrane

function (eg. see references sighted in Sacre and Tocanne, 1977).

Ion - lipid and water - lipid binding has been detected by

deuterated water and labeled sodium NMR techniques (eg. see review of

Pope and Cornell, 1978). Competitive binding of alkali metal ions to

the polar groups of egg-lecithin has been detected in the Na23 NMR study

of Persson, Lindblom and Lindman (1974). Furthermore, the chemical

nature of the polar groups of the lipid determines what effects ions in

the external aqueous phase have on the structure of lipids aggregates

(Sacre and Tocanne, 1977). Therefore it was of interest to see whether

binding of monovalent cations ions had any effects on the ordering of

lipid molecules in egg-lecithin bilayers.

Dielectric measurements of the effects of monovalent ions on the

capacitance of single planar lipid bilayers have been relatively few and

all reported measurements have been on bilayers containing undetermined

concentrations of n-alkane solvent ( eg. Hanai Haydon and Taylor, 1964,

Coster and Simons, 1970 and Ohki, 1970). Further, almost all the

reported measurements were carried out using two ter@inal impedance

measuring methods where it is difficult to separate the

105

electrode-solution impedance frora that of the membrane plus

electrode - solution interface.

those studies were raeasured at

Further,

frequencies

the membrane capacitances in

over 100 Hz. At these

frequencies the impedance of the external electrolyte was a significant

fraction of the measured total impedance; especially at the lower ion

concentrations eraployed in those studies.

Hanai et al. (1964) found no significant dependence of membrane

capacitance on the external ion concentration for bilayers for@ed from

egg-lecithin - n-decane solutions in electrolytes with concentrations

ranging from .001 - 4.18 M NaCl. White (1973) measured the effect of

varying external ion concentrations on the capacitance of GMO bilayers

formed from n-decane solutions. The capacitance of these membranes was

consistent with that predicted by the Gouy-Chapman theory for a

bilayer - electrolyte syste~ with a sraall concentration of charge fixed

at the membrane - aqueous interface. Coster and Smith (1974), using a

four terminal impedance measuring technique reported an increase in the

capacitance, measured at a frequency of lHz, of egg-lecithin -

n-tetradecane bilayers with increasing KCl concentrations in the aqueous

phase. However, the exact interpretation of these results was uncertain

as the n-tetradecane concentration in these bilayers was unknown.

In this study of the effects of ions on lipid bilayers, solventless

egg-lecithin and GMO bilayers have been employed as these eliminated the

need to account for varying solvent concentrations in the bilayer. GMO

bilayers in the present study were formed from squalene solutions as

these bilayers have been found to contain negligible squalene

concentrations (Simon et al., 1977 and White, 1978).

106

The alkane solubility in the acyl chain region of lipid bilayers is

sensitive the the entropy of the acyl chains (see Chapter 5). The

alkane solubility in egg-lecithin bilayers has been used as a means of

detecting possible changes in the structure of the hydrophobic region of

bilayers due to ion - lipid interactions at the choline phosphate polar

head groups.

Dielectric studies of egg-lecithin bilayers containing oxidised

cholesterol and n-tetradecane showed that oxidized cholesterol increased

the bilayer capacitance (Ashcroft, 1979). This was confir@ed in the

present study for solventless bilayers (see previous chapter). This

result i~plies that cholesterol decreased bilayer thickness. However,

X-ray diffraction studies on lipid multilayer preparations by ~lclntosh

( 1978) shm'led that cholesterol increases the thickness of bilayers in

the liquid crysta 11 i ne state. The apj)arent discrepancy between these

two findings will be riiscussed later in this chapter.

The aim was to measure the effects of different ion concentrations

and ion species on the dielectric and charge storage properties of

solventless egg-lecithin and GMO bilayers using low frequency impedance

measurements. The results have been interpreted in terms of the

predictions of the Gouy-Chapman theory apj)lied to the

membrane - electrolyte syste@.

Though GMO is rarely found in biological merabranes the use of GMO

in forming bilayers in this study conveys several advantages. For

example, GMO readily forras bilayers with stable electrical properties

over a wide range of electrolyte concentrations.

107

While the effect of ionic double layers have been discussed in

regard to their effects on the bilayer capacitance (eg. see Everitt and

Haydon, 1968 and Smith, 1977), as yet no complete theoretical treatment

of the impedance of ionic double layers near the electrostatic dipoles

of egg-lecithin has been made. Thus the effect of the electrostatic

dipoles of egg-lecithin on the capacitance of the ionic double layers is

largely unknown. GMO molecules are electrically neutral, with a

negligible electrostatic dipole field. This allows a simpler

interpretation to be made of the effects of ions on the capacitance of

GNO membranes on the basis of the Gouy-Chapman theory.

o.2 PREL114INARY THEORETICAL CONSIDERATIONS

The theoretical examination of the dielectric model of lipid

raembranes in section 2.42 showed that when an external potential

difference is applied across the membrane during impedance measurements

part of the potential difference will appear across the external aqueous

phase. As a consequence of this, ionic double layers exist at the

membrane - solution interface and these have a capacitance and

conductance that acts in series with the dielectric impedance of the

bilayer. The total membrane capacitance, Cm, is given by:

C - + { 1 2 }- I

m- "toi Si 6. 1

The dielectric capacitance, CD of the bilayer is given by:

E E o m

6.2

108

\·Jhere "E " is the dielectric constant of the membrane. The m

capacitance of the ionic double layers, 11.' on each side of the membrane

when there is an electrostatic potential at the membrane surface,t , (in 0

the absence of an externally applied field ie. t =0) is given by:-a,

6.3

Where "z" is the ion valency, "Ew", the dielectric constant of

water and ">. ", the Debye 1 ength in the aqueous phase. The membrane

surface potential can be related to the bound charge at the

membrane surface, as, using the following expression:

6.4

When there is no net bound surface charge equation 6.1 reduces to:-

At this point it is necessary to draw a distinction between bound

charges and those involved in electrostatic screening of membrane

surface potentials.

At equilibrium the electrochemical potential of both mobile and

bound ions is equal. The electro-chemical potential,µ, is given by:

6.6

\Jhere "v(x)"is the coulomb electrostatic potential and c is the ion C

concentration. An ion is considered as being at a bound-site when the

standard chemical potential , µ 0 , of that ion is less than that in the

109

bulk aqueous phase. The difference in standard chemical potential

between bound and mobile ions is the binding energy which could arise

from che~ical bonding or ion-specific electrostatic interactions (eg.

see Eisenman, 1961 and D'Arrigo 1978).

In this chapter three models describing charge binding to lipid

bilayers are considered; namely:

1) The concentration of bound charge at the membrane surface is

constant regardless of the electrolyte concentration.

2) The raain ion species in the electrolyte (~ajority ions) bind to the

membrane.

3) The binding-ion concentration in the electrolyte is small (minority

ions) and is independent of the electrolyte concentration (binding due

to im~urities in the electroly~ or the membrane Torus).

The variation of the ionic double layer capacitance with

electrolyte concentration for models 2 and 3 will now be calculated.

Provided the absorption of ions onto the surface of lipid bilayers

can be described by the Langmuir adsorption isotherm the bound charge

density on the membrane surface is given by:

ln 0 8 ~µo lj,

= ln cf - - q O + ln o ~ 7<T m 6.7

Provided 0 8 «oM

110

Where "°s' is the bound charge density and "oM" is the bound charge

density, when all the binding sites are occupied. "cf" is the 0

binding-ion concentration in the electrolyte and "Liµ" is the standard

chemical potential difference between ions in the bound and free states.

Substituting ~0 from equation 6.4 into equation 6.7 leads to the

following expression.

6.8

I:, 0

Kl = ---h- + ln oM 6.9

1ihere "Kl" is a constant. liowever if the bound charge does not

originate from the majority ions in the electrolyte and the

concentration of the impurity ions, either in the electrolyte or

~eLlbrane torus, does not vary with the electrolyte concentration then

one can write the following expression:

6.10

I:, 0 K2 = - kT + 1 n OM + 1 n cf 6.11

~~here "K2" is a constant. The surface charge density for cases 2

and 3 was determined from the solutions to the transcendental equations

(o.8 and 6.11). This was done using a graphical technique for different

electrolyte concentrations and different values of Kl and K2. The

membrane surface potential and double layer capacitance was then

calculated at each electrolyte concentration using equations 6.3 and

6.4. The solutions to these equations are presented and discussed in

section S.5.

111

6.3 METHODS

Glycerol Monooleate bilayer membranes were formed at 20°c from

squalene solutions using the film drainage technique described in

Chapter 4. Glycerol monooleate was obtained from two sources; from

Sigma chemical company (>99% pure) as well as from Nu-Chek (>99% pure).

Egg-lecithin bilayers were generated from n-hexadecane and

n-dodecane solutions at 40°C (10 ~M with respect to lecithin).

~easurements of capacitance reported in this chapter where made at a . 0

frequency of lHz and at temperatures in the range 20-30 C.

Bilayers were also for~ed from egg-lecithin and cholesterol

mixtures. Two different forms of cholesterol were used; oxidised

cholesterol (95% purity) and unoxidised cholesterol (>99% pure).

6.4 RESULTS

5.41 Glycerol ,,ionooleate i3ilayers

GMO bilayers formed rapidly fror,1 thick Gi;Q -squalene fili:ls; the

bilayers attaining equilibrium uith the torus within 5 minutes. The

appearance of the films during thinning was independent of the

concentration and ion type in the external electrolyte over

concentrations ranging from .l mM to 4 M. The life-time of these

bilayers was quite short; rarely exceeding 15 minutes.

112

The capacitance of solventless G~O bilayers was found to increase

with increasing electrolyte concentration; attaining an upper limit of

7.6 ±. 15mF/m2 at an electrolyte concentrations of .l to Molar. The

capacitance of the bilayer was measured as a function of salt

concentration for chloride salts of the alkali metals as well as Kr and

cuso ... It was found that membrane capacitance was independent of the

ion species for a variety of monovalent ions. However, the dependence

of membrane capacitance on electrolyte concentration in divalent

electrolytes was half that of bilayers in ~onovalent electrolytes (see

figure 6.1).

6.42 Egg-Lecithin Bilayers

The physical appearance of egg-lecithin - n-hexadecane films was

found to depend on the electrolyte concentration. At low ion

concentrations the films from which the bilayers were generated formed

readily at 40°C and were relatively fluid, compared to those formed in

higher electrolyte concentrations. The thinning of bilayers in

electrolyte concehtration in excess of 3 mM was slow and often higher 0 temperatures (45-50 C) had to be employed to ensure bilayer formation.

The thinning of the latter appeared to be hindered by clusters of lipid

aggregates forming a gel like region at the boundary of the bilayer and

the thick lipid film.

The capacitance of· solventless egg-lecithin bilayers increased with

increasing electrolyte concentration. The lHz capacitance of

egg-lecithin bilayers in .1 Molar KCl was 6.8 ±.2mF/m 2 • At lower

electrolyte concentrations the capacitance of egg-lecithin bilayers

decreased in a similar manner to that described for GMO bilayers (see

figure 6.2).

8 I

C'\,j

~ 7 ~ Ll.J (J

<'. ~ -(J

~ ~ (J 6

GMO/

t

t 10-

SQUALENE

t

t

t 'i

t

10-

ELECTROLYTE

I t

10-

CONCENTRATION

t

10-1

t 1 f

1 10

Mo! /m3

/

Figure 6.1. The membrane capacitance, measured at 1 Hz, for GMO bilayers in equilibrium with squalene solutions in different monovalent and divalent electrolytes, at 20°C.

(e) represents the average capacitance values obtained in LiCl, NaCl, KCl, CsCl and KF at each salt concentration.

( o) represents membrane capacitance va 1 ues obtained in CuS0 4 •

The error bars indicate the total experimental scatter. The effect of increasing the concentration of different monovalent ion species was the same ( within ±2%).

Lu (J <'. ~ ...... (J ~ "(

7

6

I

I

I I

I

I

(J4~----------~-------~------~ 10-4

ELECTROLYTE 10-3

CONCENTRATION

Figure 6.2. The capacitance, at lHz, of egg-lecithin bilayers as a function of KCl concentration in the external aqueous phase at 20 C ( pH=6) .

(!) represents bilayers formed from n-hexadecane solutions.

(0) represents bilayers formed from n-dodecane solutions.

The error bars at each datum point indicates the scatter in measured capacitance for 5 to 15 different membranes. The relatively large scatter in the capacitance of bilayers formed from n-dodecane has been discussed in Chapter 5.

113

The capacitance of egg-lecithin bilayers formed from n-dodecane

solutions of the lipid also showed a dependence on the ion concentration

(see figure 6.2). However, the dependence was less significant than

those of solventless bilayers (the scatter in the results greatly

increased when n-dodecane was present in the bilayer). This was due to

non-reproducibility in the measured mole fractions of n-dodecane in the

hydrophobic bilayer interior (see discussion in Chapter 5).

The mole fraction of the alkane in the bilayer phase was determined

from the meabrane capacitance (see Chapter 5). The mole fraction of

n-dodecane in egg-lecithin bilayers at 20°c was in the range 20-40%.

Figure 6.3 shows the measured mole fraction of n-dodecane at different

l t l t t t . t t t of 20°c. e ec ro ye concen ra ions a a empera ure The data in figure

6.3 indicates that the n-dodecane mole fraction in the bilayer at 100 mM

KCl was approximately 100% higher than that at low ion concentrations at

the same temperature.

6.43 Egg-lecithin Bilayers Containing Cholesterol

The capacitance of egg-lecithin bilayers containing cholesterol

(50% raole fraction) was measured for KCl concentrations in the range

. l m:~ to 10 mM. Oxidized cholesterol (50% mole fraction) significantly

increased the capacitance of egg-lecithin bilayers at low electrolyte

concentrations. The magnitude of the effect decreased with increasing

ion concentration until at 10 mN concentration the addition of oxidised

cholesterol had no measurable effect on membrane capacitance. However,

the presence of pure cholesterol in egg-lecithin membranes (50% mole

fraction) decreased the membrane capacitance at low electrolyte

concentrations and the magnitude of the effect Has not reduced at higher

electrolyte concentrations (see figure 6.4).

6

·4

< a --1--lJ ~ a: LL

UJ -.J

~-2

<: 0 L__----1...----:-----~~---___.L~2-----::10-1 1 o-t 10-3 10-

ELECTROLYTE CONCENTRATION Mo/ /rr,3

Fi~ure 6.3. The r,1ole fraction of n-dodecane in egg-lecithin bilayers at 20 Casa fuhction of external KCl concentration. The results were calculated from the data in Figure 6.2, using equations 5.2 and 5.3.

(\j

E:

" Lt. E:

Lu

~ ~ ..._ u ~ ~

7

X 6

X t X

5

4 10-l 10-3 10-2

ELECTROLYTE CONCENTRATION Mo! /m3

Figure 6.4. The membrane capacitance of solventless egg-lecithin bilayers as a function of KCl concentration in the aqueous phase.

( )() represents egg-1 ecithi n only,

X

10-1

(e) bilayers formed from solutions containing unoxidised cholesterol (~:l mole ratio lecithin:cholesterol) and

(0) bilayers formed from solutions containing oxidized cholesterol il :l mole ratio lecithin:oxidized cholesterol).

The error bars at each point indicate the scatter obtained from at least 3 different membranes. The error bars for bilayers formed from egg-lecithin solutions are shown in figure 6.3.

114

6.5 DISCUSSION

6.51 Effect of Ions on Bilayer Structure

The presence of ions in the aqueous phase had significant

observable effects on the for~ation of egg-lecithin bilayers. This

observation suggests that the configuration of lipids in the bilayer may

have been affected by interactions between the charged groups of the

egg-lecithin raolecules and the mobile ions in the electrolyte.

Inferring ion dependent structural changes in solventless

egg-lecithin bilayers from changes in the raembrane capacitance was

difficult because variations in the ionic double layer capacitance was

by far the most predominant effect (this will be shown in later sections

of this chapter). However, measurement of the alkane absorption into

lipid bilayers was not bedeviled by the effects of ionic double layer

capacitances. Changes in the ordering of lipids in the bilayer should

be reflected in changes in the partitioning of n-alkane raolecules

between the bilayer and torus. It was concluded (see Chapter 5) that

this should be a sensitive indicator of structural changes in the

hydrophobic region.

The experimental scatter rendered it difficult to detect relative

changes in the alkanes absorption much less than 50%. At. l M KCl the

alkane absorption showed an increase but it was difficult to say how

significant this increase was due to large experimental scatter. In any

case the effect that ions, at 100 raM concentration in the aqueous phase,

had on the absorption of alkane into egg-lecithin bilayers was small

compared to the large changes re~orted in Chapter 5. Thus it appears

that the presence of ions in the aqueous phase does affect lipid

115

ordering in bilayers. However, the effect on membrane ordering is

relatively small compared to that due to changes in temperature or

~embrane composition. Therefore it is not unreasonable to assume that

the effect of varying ion concentrations on membrane thickness is small

at ion concentrations less than 100 ml1.

Presented here is a tentative explanation of how it is possible

that ions can significantly effect the formation of egg-lecithin

bilayers without effecting the final bilayer st~ucture. Egg-lecithin

being sparingly soluble in n-alkane solutions, exists mainly in

mono-disperse aggregates which may be repelled by electrostatic forces.

At high ion concentrations the repulsive forces between neighbouring

aggregates are screened by the mobile ions in the electrolyte thus

allowing them to cluster. However, the distance between adjacent lipids

is much smaller than between aggregates so at these ion concentrations

the charge screening between individual lipids is negligible.

6.52 The Capacitance of the Ionic Double Layers

Considering the total capacitance of the membrane solution system,

as predicted by equation 6.1, it can be seen that at high electrolyte

concentrations the effect of ionic double layers on the total membrane

capacitance is small (since the double layer capacitance is large).

Thus at high electrolyte concentrations the membrane capacitance is

approximately equal to the dielectric capacitance of the bilayer. I

Inspection of figures 6.1 and 6.2 shows that the membrane capacitance

does approach an upper limit at high electrolyte concentrations as

expected from the Gouy-Chapman theory. At 100 mM KCl the membrane

capacitance approximates to within 2% the dielectric capacitance of the

bilayers.

116

At 1ow ion concentrations the ionic double layer capacitance is

small and hence this could significantly reduce the total membrane

capacitance. Provided the raembrane thickness and hence the dielectric

capacitances of solventless egg-lecithin and GMO bilayers are

indepenGent of electrolyte concentration then one can calculate the

capacitance of the ionic double layers by subtracting dielectric

capacitance from the total membrane capacitance as follows:

~/here "Cm" and "Cot are the membrane capacitance and ionic double

layer capacitance at an electrolyte concentration, c0 , and "Coo" is the

~embrane capacitance at high electrolyte concentrations (assuming c0~ C00 ).

The ionic double layer capacitances calculated on the basis of

equation 6.10 for egg-lecithin and GMO bilayers at different electrolyte

concentrations are shown in figure S.5. Comparing the measured double

layer capacitances to the predictions of the Gouy-Chapman theory (refer

to equations 6.2 and 6.3) in figure 6.5 shows that the results

significantly deviated from that expected for a neutral membrane or a

merabrane that possessed a constant bound charge. Therefore it appears

that the membrane possessed a bound charge that varied with the ion

concentration in the aqueous phase.

1 Alternatively it may be said that the deviation of the double layer

capacitance from that expected for a neutral membrane was due to an

underestimate of the dielectric constant of the aqueous phase adjacent

to the bilayer. This can be discounted as the dielectric constant of

the aqueous phase would need to be 320 at distances up to 30 nm from the

01)

C --

6 0 LEC

• GMO +

5

<Jb = 3x 10-3

4 \

3 L--__ _._ _______ ...__ ______ ___._ ______ ____.

10-4

ELECTROLYTE 10-1

CONCENTRATION

Figure 6.5. The capacitance (mF/m 2 ) of the ionic double layer external to egg-lecithin and GMO bilayers, plotted on a log scale, shown for different monovalent electrolyte concentrations. It was assumed that the membrane capacitance at high electrolyte concentrations was equal to the dielectric capacitance of the bilayer. The double layer capacitance was then calculated by subtracting the dielectric capacitance (assumed to be independent of electrolyte concentration) from the measured membrane capacitance using equation 6.12. The solid lines represent the double layer capacita_nce expected for bilayers cqntat-ning different amounts of bound surface charge (C/m 2 ), based on equations 6.3 and 6.4 (ie. model (1). Refer to text)

117

bilayer surface. The reorienting of water near the lipid polar moieties

in lipid bilayers should tend to reduce the dielectric constant of the

electrolyte in the immediate vicinity of the membrane surface.

Alternatively the presence of an electrostatic dipole field near

the membrane surface would produce a membrane surface potential higher

than that predicted by equation 6.4. Thus the ionic double layer

capacitance will be higher than that expected for a bilayer with zero

net charge. Considering the molecular structure of GMO and egg-lecithin

{see figure 7.6) it can be seen that the charge separation within the

electrostatic dipoles of egg-lecithin is about .4 nm, whereas GMO has no

charged groups. Thus it is clear that the electrostatic dipole

potential near the membrane surface in these two bilayer systems should

be vastly different. However, the results in figure 6.5 show that the

double layer capacitances in series with GMO and egg-lecithin bilayers

were about the same, indicating that the dipole potential of the choline

phosphate groups of egg-lecithin had little effect on the double layer

capacitance.

This is not surprising as the effect of electrostatic dipoles

should only be important when the Debye length in the electrolyte is

similar to the se?aration of the discrete charges (Cole, 1969). At low

concentrations where the ionic double layers contribute significantly to

the membrane capacitance the Debye length is about 10nm which is much

larger than the discrete charge separation. However, at concentrations '

where the Debye length is similar to the charge separation I in· the

choline phosphate groups (ie where the effect of electrostatic dipoles

cannot be ignored) the capacitance of the ionic double layers was large

and did not contribute significantly to the total membrane capacitance.

Thus if a deviation from the Gouy-Chapman theory did occur at high

118

electrolyte concentrations, as a result of electrostatic dipoles at the

membrane surface, it would not have been detectable in the present

study.

Therefore we are led back to conclude that the capacitance data

reported here can only be interpreted as being due to a net bound charge

on the bilayer which varies with the electrolyte concentration. The

partitioning of ions between the aqueous phase and bound sites on the

membrane are now considered in detail.

6.53 Nature of Bound Charge on Lipid Bilayers

The adsorption of ions to membrane-aqueous interfaces is not an

uncommon occurrence. Binding of monovalent ions has been detected in a

wide variety of lipid and soap films (eg. see Pope et al., 1978).

In section 6.2, three possible models for ion binding to bilayers

were described. The results obtained for the total membrane capacitance

have been compafed to that expected from these three ion binding models.

Model 1

The bound charge and membrane potential at the surface of GMO and

egg-lecithin bilayers, calculated on the basis of equations 6.2 and 6.3

is shown in figures 6.6a and 6.6b. Examination of these figures shows

that these data cannot be accounted for by a constant bound charge on

the bilayer membranes.

1 L,

12

10-4 ELECTROLYTE

, ,, ,,

, ,

,,,,,,'

J/,/

10-3 CONCENTRATION

10-2 Mo/ /m3

Figure 6.6a. The net surface charge density of egg-lecithin (O), and GMO (e) bilayers as a function of the monovalent electrolyte concentration, calculated from the ionic double layer capacitance data in Figure 6.5. (X) is the surface charge density on GMO bila;;ers ·in CuS04 solution.

The curves represent the bes~ theoretical fits to the data based on models 2 (full curve) and 3 (dashed curve). Kl and K2 (refer to equations 6.9 and 6. 11) for these two curves are 5.8 and -8.5 respectively.

70

60

0

( 30

20

10

' ' ", ' '

',,

10-4 ELECTROLYTE

' ' ' ' ' ' ' ' ' ', ',

',, ',

10-1 CONCENTRATION

',, .. , ',

' ', '

Figure 6.6b. This shows the membrane surface potential plotted against electrolyte concentration calculated from data in figure 6.6a using equation 6.4. ·

119

Model 2

The numerical solution to equation 6.7 was calculated using a

graphical technique for different values of the variable, Kl. The best

fit to the data using this model is shown in figures 6.7 and 6.8.

Provided that the number of available binding sites is known then the

ion binding energy can be calculated. The ion binding energy was

calculated by assuming that the number of available binding sites

corresponded to one per lipid. The values obtained for the different

lipid electrolyte systems in this study are given in Table 6. l.

Model 3

From solving equations 6.7 and 6.8 it was found that upon

increasing the electrolyte concentration the electrostatic potential

generated by the bound charges was partially screened by the mobile ions

in the electrolyte. Therefore even if the concentration of binding ions

in the bulk aqueous phase is constant, ie. when the majority ions in the

electrolyte were indifferent, the amount of bound charge at the membrane

surface is still a function of the indifferent ion concentration.

The solutions of these equations for models 2 and 3 were in good

agreement with the experimental results (see figures 6.6a, 6.6b and 6.7.

Therefore fro~ the results of these calculations both models 2 and 3

seem equally plausible. The values for the binding energies of the

absorbed ions were obtained by fitting the predictions of model 2 to the

capacitance data obtained from different membrane systems. These values

are compared in Table 6.1. The best fits to the data are shown in

figure 6.9. Obtaining quantitative values of ion binding energies for

co

I I I I I I I I I I I I I I I I I I I I I I I I I I I

...... . _............ :::>, ..

··---~:',..................... ·· ................ .... ...

-~

(Q

cW/.::JW

l/)

3JN

'v'1IJ'v'd

VJ ..... I a .....

Figure 6.7. The total membrane capacitance of egg-lecithin (0), and GMO (e) bilayers plotted against the monovalent electrolyte concentration. The curves correspond to the best fits to data obtained by the three ion binding models examined in this Chapter. For model l the bound charge density for GMO and egg-lecithin bilayers corresponded to one bound electronic charge per 50 nm 2 of membrane surface. The values or Kl and K2 for the dashed and dotted curves are given in figure 6.6a.

-- model l

----model 2

• ... -model 3.

C'\j

~ i ~

8

7

6

<'. 5 § l.)

~ 0

T

70-1. 10-.1 10-ELECTROLYTE Ca'I/CENTRATION Mol/m3

10- 1 10

Figure 6.8. A summary of the membrane capacitance data presented in this Chapter. This is compared to the predictions of model 2 (solid lines). The membrane surface potential, surface charge and bilayer dielectric capacitance derived from fitting this charge binding model to the membrane are given in Table 6.1.

(-&) GMO in Cu SO .. (•) GMO in monovalent electrolytes (e) Egg-lecithin in KCl (D) Egg-lecithin and unoxidised cholesterol (1:1 mole ratio) in KCl. (0) Egg-lecithin and oxidized cholesterol (1 :1 mole ratio) in KCl.

TABLE 6. l

membrane / ion "' mV tiµ 0/kl C O mF /m 2

- - - --- - - - - -- -- - ---- - - - - 0 - - - - - - - - - - - - - - - - - - - - - - - - - - -- - - - --

Lecithin I KCl 55 ±10 -11 ± • 2 6.9 ± • 15

GMO I KCl 60 ±10 -11 ± .2 7.6 ± • 15

GMO I CuSO 25 ± 10 -11 .2 ± .2 7.6 ± • 15

lec:chol (ox)/KCl 100 ±10 -10.5 ± .2 6.7 ± • 15

lec:chol (pure)/KCl < 55

Table 6.1. The membrane surface charge , ion binding potential, and the dielectric capacitance of the different bilayer electrolyte systems investigated in this study. The values were obtained from on the data in figure 6.8 on the basis of the predictions of model 2 (see text).

120

~odel 3 was difficult as the concentration of the binding ions, if any,

were unkno\'m.

If the surface charge on the lipid bilayers was due to the presence

of impurities in the lipid samples during isolation then one would

expect the quantity and nature of the impurities to vary significantly

between lipid samples from different supply companies. The results show

that there was no variation in the capacitance of bilayers formed from

GMO obtained from two different sources.

Alternatively binding of alkali metal cations from the electrolyte

may be occurring. Selective cation binding to lipid bilayers has been

detected in many studies (eg. see Persson et al., 1974). However, no

similar species dependence in ionic double layer capacitance was

apparent in G~lO bilayers nor was~i-~) any significant difference

between ion binding to egg-lecithin and GMO bilayers. Either the ion

binding due to alkali metal cations in this study is not species

specific or trace impurities, which may have been present in all the

experiments, were strongly absorbed onto the bilayer.

In any case it can be said that it is unlikely that the ion binding

was due to ion-ion interactions as this type of binding is strong and

highly selective for both the mobile ions and the lipid polar groups

(see Eisenman, 1961, and D'Arrigo, 1978). Therefore it is more likely

that weaker interactions such as ion dipole-dipole and induced I

dipole-dipole forces are responsible for ion adsorption onto these

bilayers.

12 l

6.54 The Effect of Cholesterol on Bilayer Ca~acitance

The substance which has become known as oxidised cholesterol

represents neither a single compound nor is it completely characterised.

However, it is extremely useful in bilayer studies as it acts as a

membrane stabilizer (see Chapter 1). Oxidised cholesterol and pure

cholesterol had opposite effects on the bilayer capacitance at low

electrolyte concentrations (see figure 6.4). At high electrolyte

concentrations the addition of oxidised cholesterol had very little

effect on the total capacitance of egg-lecithin bilayers. By comparing

the data to the numerical solutions of equation 6.7 it was found that

the effect of adding oxidised cholesterol was a marginal decrease in the

bilayer dielectric capacitance (see Table 6.1).

The surface charge density on egg-lecithin bilayers was calculated

from the ionic double layer capacitance as in previous sections.

However, as bilayers containing cholesterol are multicomponent systems

it is only possible to calculate the average ion binding energy over the

membrane surface (as the composition of the bilayer is not ~ell known).

Therefore the ion binding energies shown in Table 6.1 for bilayers

containing cholesterol only represent an average over all the bound

ions.

The presence of ~ure cholesterol (50% mole fraction) in

egg-lecithin bilayers decreased the bound surface charge density whereas

oxidised cholesterol doubled the surface charge density. Thus it

appears that either ionic impurities were present in oxidised

cholesterol that were not present in the pure sample or that ions could

bind more strongly to the some of the oxidation products of cholesterol.

122

The discrepancy between the conclusions of Ashcroft (1979) and

those of X-ray diffraction experiments (e.g. McIntosh, 1978) concerning

the effect of cholesterol on bilayer thickness was probably due to the

variations in the ionic double layer capacitance arising from the

presence of oxidised cholesterol which were unknown to Ashcroft (1979).

When the capacitance of the double layers was taken into consideration

the relative effects of cholesterol on bilayer thickness obtained from

dielectric measurements were similar to that obtained from X-ray

diffraction experiments.

6.55 Comparison With Previous Work

The values obtained for the upper limit of the capacitance of

GMO-squalene bilayers at high electrolyte concentrations were in good

agreement with those found by other workers (c.f. the present value of

7.6 ±. l5mF/m 2 \tith the value of 7.45 ± .24mF/m 2 (Benz et al., 1975, and

7.78 ± .005mF/m 2 (White, 1978)).

Previous measurements of capacitance as a function of electrolyte

concentration (White, 1973) of bilayers formed from GMO - n-decane

solution were consistent with the membrane having a bound surface charge

which was independent of ion concentration in the electrolyte. ~Jhi te

(1973) attributed this bound charge to the presence of ionic impurities

in the membrane forming solution or electrolyte. The results reported

here are different to those reported by White (1973) in that the I

variation of membrane capacitance with increasing ion concentration was

smaller and was consistent with a varying bound surface charge.

The capacitance of egg-lecithin bilayers was also found to depend

on the ion concentration; a similar effect to that found by Coster and

123

Coster and Smith (1974) but in disagreement with the findings of

Hanai et al. (1964). The increased scatter in the capacitance

measurements when n-dodecane was present in the bilayer tended to mask

the effect of varying ion concentrations. This raay account for why the

variation of bilayer capacitance with electrolyte concentration was

unnoticed by Hanai et al. (1964) in egg-lecithin bilayers containing

high n-decane concentrations.

6.6 SUMMARY

The capacitances of solventless egg-lecithin,

egg-lecithin - cholesterol and GMO bilayers were measured at different

ion concentrations in the external aqueous phase. The results were

interpreted in terms of the Gouy-Chapman theory applied to the

bilayer-water interface.

The capacitance of bilayers in electrolyte concentrations in excess

of .1 Molar was approximately equal to the dielectric capacitance of the

bilayer. At lower concentrations the capacitance of the ionic double

layers in series with the dielectric capacitance of the bilayer reduced

the total capacitance of the membrane.

The presence of a significant mole fraction of n-alkane in the

bilayer {in this case n-dodecane) introduced experimental scatter in the

capacitance which tended to raask the effect of ions on the membrane

capacitance. This might also account for the lack of effect reported by

Hanai et al. (1964) on egg-lecithin bilayers which also contained high

concentrations of solvent.

124

The results in figure 6.3 show that the n-dodecane mole fraction in

egg-lecithin bilayers increased with the external electrolyte

concentration. However, the ion concentration dependence of n-alkane

absorption was small compared that observed for other factors studied in

Chapter 5. It was concluded that, even though the ordering of the acyl

chains was slightly altered, the thickness of egg-lecithin bilayers was

largely independent of the electrolyte concentration.

The ion dependent capacitance of GMO bilayers measured in these

experiments was different to that obtained in previous measurements by

White (1973) on GMO - n-decane bilayers obtained using a two terminal

impedance measuring technique. The capacitance of the ionic double

layers obtained in this study deviated significantly from that expected

from Gouy-Chapman theory for a neutral membrane or one that had a

constant bound surface charge.

The Gouy-Chapman theory applied to ionic double layers external to

lipid bilayers by Everitt and Haydon (1968) was extended in this study

to include the possibility of an electrolyte concentration dependent ion

binding to the membrane surface. The results obtained is this study

were consistent with a weak, non-selective binding between either

majority ions in the electrolyte or trace impurities in the electrolyte

or membrane forming solution and bilayer surface.

Oxidised cholesterol had little effect on the bilayer thickness; if

anything there was a 2-3% increase. At low ion concentrations oxidised

cholesterol and pure cholesterol had opposite effects on me~brane

capacitance. The difference in the effects of these compounds on

bilayer capacitance at low electrolyte concentrations was chiefly

determined by variations in the capacitance of the ionic double layers.

125

It was concluded on the basis of the Gouy-Chapman analysis used in this

study that oxidised cholesterol caused a significant increase in the

bound charge density at the membrane surface whereas pure cholesterol

did not. Variation in the double layer capacitance upon the addition of

oxidised cholesterol to egg-lecithin bilayers at low electrolyte

concentrations seemed to be the main cause for its apparently opposite

effects on membrane thickness reported by Ashcroft (1979) and McIntosh

(1978).

CHAPTER 7

THE DIELECTRIC STRUCTURE OF THE HYDROPHOBIC - HYDROPHILIC INTERFACE OF EGG-LECITHIN AND GMO BILAYERS

7. l INTRODUCTION

7.2 METHODS

7.3 RESULTS 7.31 Frequency Dispersion in Bilayer Capacitance

and Conductance

7.32 Reliability of Fitted Parameters

7.33 Effect of Varying Ion Concentrations in the External Electrolyte

7 .4 DISCUSSION

7.41 Interpretation and Presentation of Structural Data

7.42 Dielectric Structure of GMO and Egg-Lecithin Bilayers Compared

7.43 Conductivity of the Hydrophobic - Hydrophilic Interface

7.44 The Conductance of the Ionic Double Layers

7.5 SUMMARY

126

Page

127

128

129 129

130

133

135

135

136

139

142

144

127

7. 1 INTRODUCTION

It is known that a slab of dielectric material consisting of a

series of discrete layers which have different electrical time-constants

exhibits a Maxwell-Wagner dispersion in total capacitance and

conductance. Coster and Smith (1974) were the first to measure this

dispersion in lipid bilayer membranes in the frequency range .1-100 Hz

which was associated with the dielectric inhomogeneity of the bilayer.

The results of Coster and Smith were successfully modelled with a

trilayer dielectric substructure consisting of a hydrophobic region

bounded by two relatively polar regions. The hydrophobic region was

associated with the acyl chains of the lipids and the polar regions with

the choline phosphate groups. Refinements to the measuring system by

Ashcroft, Coster and Smith (1981) revealed a third dielectrically

distinct region producing a frequency dispersion in bilayer capacitance

in the frequency range .01-.l Hz. This region had dielectric properties

midway between those of the polar and the hydrophobic regions. The

dielectric parameters \'/ere associated with the region of the bilayer

containing the carboxyl and ester-oxygen atoms of the lipids. Further

technical advances in the impedance measuring system now allow the

resolution of additional dielectric substructure with lipid membranes.

This chapter is concerned with the assignment of the dielectric j

parameters derived from fitting the theoretical Maxwell-Wagner impedance

dispersion to the impedance data and the interpretation of the

dielectric structure. To aid this, the dielectric structure of bilayer

membranes formed from egg-lecithin have been compared with that of

bilayers formed from glycerol monooleate. Although the dielectric

128

structure of bilayers formed from GMO and egg-lecithin should have some

similarities, in that both molecules have similar hydrophobic and acetyl

chemical moieties, there should also be differences in the dielectric

structure of these two bilayer systems as the polar head-group regions

of the former consists of the alignment of choline phosphate dipoles and

the latter contains mainly hydroxyl groups. Comparison of and

contrasting the dielectric dispersions of bilayers formed from these two

amphiphiles has allowed assignment of the dielectric parameters to

different parts of the bilayer structure. This has then provided a

basis for the study of the conductance mechanisms at the hydrophobic -

hydrophilic interface.

7.2 METHODS

Egg yolk lecithin bilayers were formed at 40°C from solutions

containing different chainlength n-alkanes. Solventless egg-lecithin

bilay~rs were generated using n-hexadecane solutions by the technique

described in Chapters 4 and 5. Egg-lecithin bilayers were formed in

aq~ous solutions containing l, 10, and 100 mM KCl.

-GMO bilayers were formed from 200mM solutions of glycerol

monooleate in n-hexadecane at 20°c. The ion concentrations in the

aqueous phase were in the range 10- 4 to 4.5 Molar.

Impedance dispersion measurements were conducted at 20-25°C. Uhen

membrane capacitance had sufficiently settled (varying by less than .2%

per 30 minutes) a series of frequency scans of membrane impedance was

commenced. Each scan involved measurements of membrane impedance at

129

35-45 frequencies, in increments of a factor of 1.4, varying in value

from .003 to 10000 Hz. Such a frequency scan took 30-45 minutes. All

impedance data reported here was an average of at least 3 such scans on

egg-lecithin bilayers. However, the relatively short life span of GMO

bilayers only allowed results of single frequency scans to be analysed.

7.3 RESULTS

7.31 Frequency Dispersion in Bilayer Capacitance and Conductance

The capacitance of egg-lecithin bilayers was found to reach

sufficiently steady values for frequency scans of the impedance in 20-40

minutes. The membrane conductance to a small extent remained time

dependent, usually increasing by 2-10% per hour. Hanai et al. {1965c)

demonstrated that a significant fraction of the membrane current was due

to border "leakage" {ie. current shunts through or around the torus).

Occasionally the conductance of the bilayer would jump to a new value,

presumably due to an abrupt change in the border "leak" current. This

"leak" conductance was estimated by plotting the linear relationship

between r.1er.1brane capacitance {proportional to bilayer area) and membrane

conductance and extrapolating to the conductance intercept. The

i@pedance of the current shunt does not reflect any intrinsic property

of the bilayer structure and therefore was subtracted from the data

before being fitted to a Maxwell-Wagner dispersion.

Bilayers formed from GMO thinned rapidly achieving a steady

capacitance in 3-10 minutes. However the bimolecular films were short

lived, lasting at most 60 minutes.

130

All experimental impedance data obtained from lipid bilayers showed

a frequency dependent capacitance and conductance which could be

accurately fitted to a Maxwell-Wagner dispersion expected from a

sandwich of substructural layers with 4-6 distinctly resolvable

electrical time-constants (ie,

bilayer is symmetric).

7-11 different layers assuming the

7.32 Reliability of Fitted Parameters

There were found to be two main sources of variability in the

fitted membrane parameters. These were:

a) the variations between successive membranes,

imperfect control of the bilayer environment

probably due to

b) uncertainties in the Maxwell-Wagner fit to the data, arising from

experimental error or the limited frequency range of the impedance

measurer.tents.

Here the various factors affecting the reliability and accuracy of

the dielectric parameters extracted from the least squares fitting

routine are considered.

i) Uncertainty in Accounting for the D.C. Conductance

j

Resolution of two time-constants within the bilayer was obtained

from the capacitance dispersion in the frequency range .003-.03 Hz. The

dielectric parameters giving the best Maxwell-Wagner fit to the

experimental data in this frequency range depended on what fraction of

the membrane current was attributed to the intrinsic conductance

131

(i.e. through the hydrophobic region of the membrane) and extrinsic

conductance (i.e. through aqueous channels shunting and hydrophobic

region) of the bilayer. Even accounting for the border "leak" current

using the method of Hana i et al . ( 1965c) there was st i 11 doubt as

to the nature of bilayer conduction (i.e. intrinsic or extrinsic;

see Chapter 9). The uncertainty in estimating the relative contributions

of these components of bilayer conductance introduced errors in the

dielectric parameters extracted from the impedance data. In the extrmee

case, it was found that by attributing the membrane conductance entirely

to extrinsic mechanisms and then subtracting the extrinsic conductance

from the data caused a 100% change in the ea lcul ated time-constant

of the hydrophobic region. However, the effect was much smaller for

regions with lower time-constants (i.e. the more polar regions).

The effect of this uncertainty could be ignored when:

7. l

where "GN " and "C N" are the conductance and capacitance of the

"Nth" polar region and "CM" and "GL" are the total bilayer capacitance

and "leak" conductance (G < 0.C. bilayer conductance) respectively.

For the bilayer membranes in this study subtracting the "leak" conduct­

ance from the membrane impedance had no effect on the fitted parameters

of polar regions with electrical time-constants less than about 10

seconds (see Table 7.1).

The unresolved question of intrinsic and extrinsic conductances

made the assignment of the electrical parameters derived from the

data in the frequency range . 003-. 03Hz uncertain. Thus if the membrane

conductance was intrinsic to the bilayer then the dielectric parameters

were consistent with there being two electrically distinct regions

TABLE 7. l

THE EFFECT OF THE SUBTRACTING "LEAK" CONDUCTANCE ON THE BILAYER FIT PARAMETERS

Leak Subtracted

------------------------------------------Element number Capacitance nF I -

n

l 7.8 7. 77

2 436 550

3 530 550

4 850 850

5 1230 1230

6 1000 1000

7 0.0 0.0

Conductance nS

l .96 0

2 513 600

3 3.1><10 3 3.2><10 3

4 2.6><10" 2.6><10"

5 l .2><10 5 l .2><10 5

6 9.5><10 5 9.5><10~

7 9. 5><10 5 9.5><10 5

-------------------------------------------

Table 7.1. The fit parameters to the impedance dispersion of a hard-wire network of resistors and capacitors designed to simulate the dielectric properties of of a six layered bilayer immersed in electrolyte. The parameters on the left were obtained by assuming that translocation of ions was through the hydrophobic region of the bilayer. The parameters on the right were obtained from the same impedance data after having subtracted .96 nS from the data at each frequency. The two methods gave results which differed significantly only when the ratio of C/G \'fas similar to C/GL where GL is the "leak" conductance subtracted fro~ the impedance data.

132

associated with the acyl chains of the lipids. On the other hand, if

the conductance was via an extrinsic mechanism, say transmembrane water

channels, then subtraction of the extrinsic component from the data

yielded dielectric parameters that were consistent with there being one

electrical time-constant associated with the hydrophobic region and

another associated with a more polar region.

ii) Errors due to Finite Frequency Range

At low frequencies the resistive component of the membrane

impedance is large compared with the reactive component (loosely

speaking the bilayer looks more like a resistor than a capacitor).

Accurate measurement of the me~brane capacitance at these low

frequencies was limited by the large relative errors in measuring the

small phase angles of the membrane impedance. Therefore as the

membrane conductance increased (ie. in the presence of concentrated

electrolyte solutions) the low frequency limit of data acquisition moved

to higher frequencies (.03-.l Hz). Thus at high salt concentrations the

structural information in the very low frequency range could not be

easily resolved as data at very low frequencies became progressively

inaccurate.

At high frequencies the bilayer impedance was much less than that

of the electrolyte between the membrane and the potential measuring

electrodes due to the relatively high capacitance of the bilayer. Thus I

the dielectric properties of the bilayer are not easily resolved at

these high frequencies. Therefore the high frequency limit of data

acquisition was determined by the relative impedances of the bilayer and

electrolyte. This explains why the polar substructure was more easily

133

resolved in high external KCl concentrations where the electrolyte

impedance was much smaller (see Table 7.2).

Tests on hard-wire impedance networks showed that the polar head

parameters could be determined to within an accuracy of 20% in

capacitance and 30% in conductance (see Table 7.1 and also section

4.54). The capacitance of the hydrophobic region could be determined

with a precision of ±0.2% and in the case of solventless bilayers was

repeatable to within ±2% between different membranes.

7.33 Effect of Varying Ion Concentrations in the External Electrolyte.

The capacitance of the egg-lecithin and GMO bilayers increased

with increasing KCl concentration in accord with the results of

Chapter 6. Increasing the KCl concentration in the external electrolyte

moved the large dispersion in capacitance and conductance arising from

the electrolyte impedance to higher frequencies. However, at low

frequencies the general shape of the dispersion remained unchanged (see

figures 7.la, 7.lb and 7.2a; also see Tables 7.2 and 7.3). It was also

found that the impedance dispersion of GMO bilayers was insensitive to

pH over the range l to 6.

The capacitance of GMO - n-hexadecane bilayers was less than that

measured for GMO-squalene bilayers reported in Chapter 6. This is

because n-hexadecane was slightly soluble in the hydrophobic interior of l

GMO bilayers (White, 1977 and Haydon et al., 1977). However, bilayers

for~ed using n-hexadecane solvent had longer life-times than those using

squalene and therefore where more suitable for impedance dispersion

measurements.

GMO

6.8

6.7

6.6

6.5

6.t.

C\j 6.3 E

.TM

~ E

6.2

10-1 1 10 FREQUENCY Hz

Figure 7.la. The capacitance spectrum of representative membranes formed from GMO/ n-hexadecane solutions at 20°c and in electrolytes of different concentration. The data shown for bilayers formed in .1 and lM KCl is an average of 2 runs. The error bars on the averaged data are too small to discern from this graph. The solid curves represent the Maxwe 11-Wagner dispersion expected for 4-6 layered dielectric · structures.

10

1

10-1 1 10 102 FREOUE NCY Hz

Figure 7.lb. The measured conductance spectrum of a GMO bilayer as a function of frequency in l Molar KCl at 20°C. The capacitance spectrum for the same membrane is shown in figure 7.la. The solid line is the Maxwell-Wagner theoretical curve. 1

7-0

6·8 N

E ....... lL E

w 6·6 u z <{ ..... u ~ <{ u 6-4

6·2

t I

10-1 1·0

FREQUENCY 101

HERTZ

• 1 mM KCI

a 10 mM KCI

• 100 mM KCI

10 2 103 104

Figure 7.2a. The measured capacitance spectrum (measured at 20°C) of egg-lecithin bilayers formed from n-hexadecane solutions. The results shown here represent the average of five frequency scans on single bilayers at 3 different KCl concentrations. The error bars indicate the stangard error on the mean at each frequency. The error bars are too smal1 to be seen on this graph. The full curves represent Maxwell-Wagner theoretical fits to the data.

• 1mMKCI

a 10 m M KCI ,a-1

" 100 m M KCI N

E '- ,a-2 Vl

w u z 10-J <( I-u :::> 0 z ,o-4 0 u

,a-s

10 2 10 3 10 4

FREQUENCY HERTZ

Figure 7.2b. The measured membrane conductance spectrum (corresponding to the capacitance spectrums in figure 7.2a) formed from egg-lecithin ( 20°c) .

TABLE 7.2

THE DIELECTRIC PARAMETERS OF EGG-LECITHIN BILAYERS

1 mM

( 10)

6.35 ±. 1

650 ±100

1200 ±200

1200 ±-200

1100 ±200

. 1-. 3

(4-20)x10

(3-10)xl0 2

(6-10)xl0 3

(5-9)x10"

10 mM

(5)

Capacitance mF/m 2

6.7 ±.15

600 ±-50

1250 ± 200

1750 ±300

2000 ±300

Conductance mS/m 2

.5-2.5

{4-5)xl0 2

(3-8)xl0 3

( 6-9) X 10"

(6-7)xl0 5

100 mM

(4)

7 .0 ±.2

540 ± 100

950 ± 100

1300 ± 300

1200 ± 200

1-3

(2-3)xl0 3

(4-7) X 10"

(6-10)xl0 5

(3-7)xl0 6

------------------------------------------------

Table 7.2. ~he dielectric parameters of egg-lecithin bilayers formed from n-hexadecane solutions in the temperature range 20-30°C. The errors are the standard deviation on the mean obtained for the number of membranes indicated at the top of each column.

l mM ( 3)

5.7 ±.2

500 ±100

860 ±200

1100 ±200

1000 ±200

.1-.3

(2-4)xl0 2

(l-3)xl0 3

(l-2)xl0"

(8-lO)xlO"

TABLE 7.3

THE DIELECTRIC PARAMETERS OF GMO BILAYERS

10 mM ( 2)

6.5 ±.2

700 ±100

650 ±200

850 ±200

-

. 1-. 3

(5-l0)xl0 3

(3-4)xl0"

(l-2)xl0 5

-

100 mM (3)

Capacitance mF /m 2

6.7 ±.2

700 ±100

850 ±100

1200 ± l 00

-

Conductance mS/m 2

. l -1

(l .5-3)xl0 3

(l-l .5)xl0"

(l-2)xl0 5

-

l M ( 2)

6.6 ±.2

l 200±WO

1600 ±400

2000 ±400

-

.01-.05

(4-5)xl0 3

(5-7)xl0"

(2-3)xl0 5

-

4 M (6)

6.8 ±.2

1300 ±200

1500 ±300

1500 ±300

1700 ±300

1-3

(l-2)xl0 3

(3-5)xl0"

(2-4)xl0 5

(1-2)xl0 6

Table 7.3. The dielectric parameters of GMO /n-hexadecane bilayers at 20°c in the presence of different KCl concentrations in the external electrolyte. The number at the top of each column refers to the number of bilayers from which the statistics were obtained.

134

The general shape of the dispersion of GMO and egg-lecithin

bilayers was distinctly different (see figure 7.3); the main difference

being that the capacitance of GMO bilayers was independent of frequency

at frequencies over 100 Hz whereas that for egg-lecithin continued to

disperse. At frequencies less than 100 Hz the shape of the impedance

dispersions for egg-lecithin and GMO bilayers was similar.

Inspection of Table 7.3 reveals that the electrical time-constant

and hence the relative conductivity, of the polar regions of GMO and

egg-lecithin bilayers was insensitive to large changes in the external

electrolyte concentration. Only the most conductive portions of the

polar region showed any dependence of time-constant on the electrolyte

concentration and then only at electrolyte concentrations in excess of

l Molar. The dependence of the conductivity of the most conductive

portion of the polar structure on the ion concentration in the external

electrolyte is shown in figure 7.4.

72

7. 0

""6.8

~ l( E

Lu

~ ~6.6 -u ~ 0

10-1 1 FREQUENCY Hz

10

o GMO • LEG

Figure 7.3. The capacitance dispersion of representative egg-lecithin and GMO bilayers at 100 mM KCl at 20°C. The capacitance of both membranes at low frequencies follows similar frequency dependence. However, at high frequencies the dispersions are distinctly different.

E

" l/) C:

J(X)

2(1]

100 >--I--:::::.. .::: l) :::> a :::: a l)

10-2 ELECTROLYTE

10- 1 CONCENTRATION

1 Mo! /m 3

I

10

Figure 7.4. The volume specific conductivity of the most conductive region in the polar structure of GMO bilayers at different external KCl concentrations. The values were calculated from the data presented in table 7.2. It was assumed that the most conductive region was .4 nm thick.

135

7.4 DISCUSSION

7.41 Interpretation and Presentation of Structural Data

The first measurements of membrane dielectric substructure by

Coster and Smith (1974) were able to identify a polar region which was

associated with the alignment of choline phosphate groups at the

membrane solution interface. Later, Ashcroft, Coster and Smith (1981)

using a refined measuring technique were able to identify another less

polar region which was associated with the acetyl and ester oxygen atoms

of the lecithin molecules. Each region was considered as a thin slab of

dielectric with a fairly well defined dielectric constant and thickness.

With the new, improved low frequency impedance spectrometer

(BULFIS) and the subsequent refinements in the measuring techniques, the

resolution of dielectric structure has much increased.

Tables 7.2 and 7.3 shows that six distinct electrical

have been detected in the polar regions of the bilayer.

Inspection of

time-constants

However, this

new wealth of information brings with it problems which have set new

limitations to the extended interpretation now possible. Firstly, the

assignment of dielectric parameters to different parts of the bilayer

and consequently the co~parison of different bilayer structures is more

difficult with a many layered dielectric model. Secondly, the

structural information inferred from the high frequency data is

truncated due to the limited frequency range of the data acquisition 1

system and the impedance of the electrolyte. Therefore there is some

uncertainty as to which part of the hydrophobic - hydrophilic interface

is manifest in the impedance dispersions. Thirdly, when one examines

the dielectric parameters in the previous Tables it can be shown that

they pertain to regions that are only .1-.4 nm thick. Over these short

136

distances the concept of a distinct region with a well defined

dielectric constant and conductivity is clearly unrealistic. Even if

step changes in dielectric constant do occur at the boundary of each

region the conductance and hence dielectric time constant would remain a

smoothly varying function of position. For a bilayer solution interface

the dielectric properties would exhibit a monotonic spatial variation

between those of the hydrophobic interior and those of the external

aqueous phase. The analysis of the data presented here models this

somewhat nebulous transition region between these two phases with series

of step changes.

In this chapter (and some subsequent chapters) the dielectric data

obtained from lipid bilayers will be presented in the form of electrical

time-constant diagra~s. These particular experimental plots readily

give insight into the possible spatial variation of dielectric

time-constant in the bilayer structure.

To assign a dielectric constant to each layer requires some

additional, a priori information. However, one can gain estimates of

the dielectric constant of different parts of the bilayer structure by

comparing the chemical and dielectric structures of egg-lecithin and GMO

bilayers. Assignment of the parameters in Tables 7.2 and 7.3 to

different parts of the bilayer structure will now be discussed.

7.42 Dielectric Structure of1 GMO and Egg-Lecithin Bilayers Compared

The dielectric structure of GMO was easier to interpret than

of egg-lecithin because the capacitance dispersion of the entire

region lay within the experimentally useful range of frequencies;

capacitance dispersion being complete before the dispersion due to

that

polar

the

the

137

presence of the aqueous phase began. The cessation of the irapedance

dispersion at about 100 Hz provided a convenient "raarker'' on the

experimental data which was associated with the most conductive part of

the G~O polar region (presumably the hydroxyl groups of the GMO

molecules).

The lowest electrical time constant detected in the GMO bilayer

structure was 10- 2 second and which we here associate with the hydroxyl

region which should be approxi~ately .3-.4 nm thick depending on the

relative orientation of the hydroxyl groups (see figure 7.5). The

dielectric constant of similar chemical moieties such as the short chain

alcohols and diols is in the range 25-40 (Chemical Rubber Company

Handbook of Che~istry and Physics, 1976). Providea the dielectric

constant of the hydroxyl region of GMO bilayers is also in the range

25-40 then the capacitance of the hydroxyl region would be

600-1200 mF/m 2• Comparing these values with those in Table 7.1 it seems

that the one or two regions with the lowest electrical time-constants

can be associated with the hydroxyl regions of the bilayer.

The other regions with higher electrical time-constant (>.l second)

must then be associated with the acetyl region of GMO bilayers. It is

expected that regions of the dielectric substructure of egg-lecithin

bilayers having similar time-constants (>.lsecond), can also be

associated with the the acetyl region of egg-lecithin bilayers.

Provided the acetyl region of egg-lecithin and GMO bilayers is

about .3 nm (see figure 7.5) then the total capacitance of the acetyl

region is consistent with it having a relative dielectric constant in

the range 6-12 which is akin to that expected from similar chemical

moieties such as acetic acid (see figure 7.6).

CH3

CH....._/ CHOLINE-3 + \'CH3 PHOSPHATE ·-· I P-o __ ,

OH P:::::::o I I

~ 0 HYDROXYL OH "

OH 0-• /• 0 ~ I ---- OH0 0 0 \ "'\ / ACETYL _ o '( 0,, a-"

ACETYL ~~ 0

Figure 7.5. A comparison of the molecular structures of glycerol monooleate and phosphatidylcholine. The dark circles represent CH groups. Note that the regions containing the acetyl and hydrocarbon moieties of these molecules are similar. The polar substructure of bilayers formed from these molecules would differ only in the choline phosphate and hydroxyl regions. The entire length of the acyl chains of these molecules does not appear in the figure. The full length of the acyl chains are not shown in this figure.

-I.

-5

-6

r..., -7

~ V')

t ~ ~ u ::, a

-8

~ -9~2 u .......

10

I. 5 6

11.

-.J )... 1-.

17

11

7

21

22-/

19 ~nULIIVC

18 I PHOSPHATE 21.

16 _:_J 15

12

8

/20

13 I_H_Y_D:.,_~-0-XYL-I

Q 0\ ~

t3 1/1 ~ -v 10 20 30 1.0 50

Er

2526

60

KEY 1 PENTANE 2 PROPIONIC ACID 3 ACETIC ACID I. BROMOFORM 5 CHLOROFORM 6 IDOMETHANE 7 PfRIDINE 8 PROPANOL 9 METHANOL 10 DIPIRID!NE 11 METHYL PROPIONATE 12 ACETONE 13 ETHANOL 11. METHYLAMINE 15 ACETY L CHLORIDE 16 ACETYL ACETONE 17 CHLOROACETIC ACID 18 PROP/ON ALDEHYDE 19 ACETALDEHYDE 20 NITROBENZENE 21 ALLYL I SOTHIOCYANATE 22 ALLYL ALCOHOL 23 GLYCEROL 21. FURFURAL 25 FORMIC ACID 26 ACETAMIDE

Figure 7.6. Presented here is a pictorial summary of the dielectric structural characterization of GMO and egg-lecithin bilayer systems described in this chapter. The possible range of dielectric constants of each main region within these two bilayer structures and their measured conductivities are compared with that reported for pure non-aqueous polar liquids (Washburn, 1929). The full line represents the assumed relationship between dielectric constant and conductivity within the bilayer. This relationship was used in generating the time-constant diagrams in figures 7.7 and 7.8) from the data in Tables 7.2 and 7.3.

138

Egg-lecithin bilayers exhibit a dispersion in capacitance and

conductance at frequencies much higher than that of GMO bilayers. The

dispersion data obtained from egg-lecithin bilayers at high frequencies

was indicative of regions with lower electrical time-constants than

those found in GMO bilayers. The additional dielectric parameters

(with tirae-constants in the range 10- 5 to 10- 3 second), obtained fro@ the

high frequency data are expected to be due to the presence of the region

containing the choline phosphate dipoles. From the molecular dimensions

of the choline phosphate group of egg-lecithin and the capacitance

values shown in Table 7.2 it is possible to estiraate the dielectric

consta~t of the region containing these chemical raoieties. However, it

is li~ely that only part of the dielectric structure of this region can

be detected (see section 7.32). Therefore only an upper limit to the

dielectric constant of this region can be estimated.

is calculated as approx. 40.

This upper li~it

Figure 7.6 shows volume specific conductivity of each main region

in the polar structure (viz. of the acetyl, hydroxyl and choline

phosphate) of egg-lecithin and GMO bilayers plotted against the

dielectric constant of each raain region.

The dielectric time-constant profiles for a number of egg-lecithin

and GMO bilayers are compared in figures 7.7 and 7.8. In order to

assign dielectric constants to each electrically distinct region in the

·membrane structure it was necessary to assurae a linear relationship l

between the conductivity and dielectric constant of the different parts

of the raembrane structure. This linear relationship was based on the

above estimates of the average dielectric constant of each region and

the conductances of each region, calculated from the parameters in

Tables 7.2 and 7.3 (see figure 7.6). It is important to note that the

(.'.)

"

2

0

ACETYL CHOLINE - PHOSPHATE

HYDROXYL

l) - 1 --2

-3

-lL....!:- --L----!-----1=----'-----..l,;:-- --~==::t====+== 0 ·1 ·2 ·3 · t. ·5 ·6 ·B

Relat,ve D,stonce nm

Figure 7.7. The possible spatial variation of dielectric time-constant in GMO and egg-lecithin bilayers calculated from the parameters in tables 7.2 and 7.3. The shaded area represents the total variation in the dielectric substructure of 8 GMO bilayers in l to 100 mM KCl. The unshaded enclosed area represents the total scatter on 19 egg-lecithin bilayers in l to 100 mM KCl. The thickness of each region was calculated from the membrane capacitance and the dielectric constant (see figure 7.6). The horizontal scale represents relative distance. The horizontal scale is monotonic but not necessarily linear (see text).

In the diagram, the different parts of the polar substructure of GMO . and egg - lec ithin bilayer s have been de lineated. The data shown here, pertaining t o egg-l ec i t hin i s t ru ncated on both the left and righ t as a resul t of the li mi ted freq uency r ange of the data acquisition. The GMO data i s only tr uncated to t he left (see text ).

--.. l'.)

" u '--

°' -2

2

(AQUEOUS HYDROXYL ACETYL ~

0

-1

-2

-3

-ll_ _____ ___J__ ___ .J_ __ ---4--__ _.__ ___ .._ __ ....._ __ _.

·O . 1 ·2 ·] ·5 ·6

Relat,ve Pos,flon nm

Figure 7.8. The time-constant profile of GMO bilayers at l to 100 mM KCl (unshaded) and 4M KCl (shaded). Note that the main difference between the two profiles shown here is in the hydroxyl region. The electrical time-constant of this region decreases at high electrolyte concentrations. The acetyl region is unaffected by variations in the ion concentrations in the aqueous phase.

139

scale of these time-constant diagrams, though monotonic, is certainly

not linear as the linearity of the relationship between conductivity and

dielectric constant used in deriving these plots is somewhat arbitrary.

However, from correlating the conductivity and dielectric constant of

~any other organic compounds (see figure 7.6) it appears that the

assumed relation between dielectric constant and conductivity may not be

a bad approximation.

7.43 Conductivity of the Hydrophobic - Hydrophilic Interface

The measured conductance of the polar regions of egg-lecithin and

GMO bilayers was found to be in the range_ . l to 104 Slm2 which

corresponds to a volu~e conductivity of 10- 10 to 10- 5 Sim. It should be

noted here that these values are small compared to that found in the

adjacent electrolyte (.l Sim for l mM KCl and 10 Sim for 100 mM KCl).

Furthermore the conductance of the acetyl region of both egg-lecithin

and GMO bilayers was considerably less than that of highly purified

water which has a conductance of 4xl0- 3 Sim (Bockris and Reddy, 1970).

These low conductances suggest that there is very little electrolyte

penetration into the acetyl region. On the other hand, the choline

phos~hate dipole region of the egg-lecithin bilayers _has a conductance

higher than that of pure water and it is likely that there is

significant ion penetration into this region.

i) Conductivity at Low Ion Concentrations

From the results shown in figure 7.4 it can be seen that the

conductivity of the polar regions of GMO and egg-lecithin is insensitive

to the ion concentration in the aqueous phase at concentrations less

than l Molar KCl. From these results two possible conclusions can be

140

drawn. Either the electrical conduction in the polar regions was due to

raechanisms other than ion migration or the ion concentrations in the

polar regions were bufferred against changes in the external electrolyte

concentration.

Consider the presence of a bound charge at the membrane surface

(see Chapter 6). When the electrolyte concentration in the bulk phase

is less than the surface charge density the total ion concentration near

the region of bound charge is equal to the bound charge concentration.

At higher electrolyte concentrations the bufferring effect of the bound

charge becomes negligible.

Theoretical calculations presented in Chapter 2 and Chapter 6

demonstrate that the capacitance of the ionic double layers is dependent

on the ion concentration near the membrane. The results presented in

Chapter 5 indicate that the ionic double layer capacitance varied with

salt concentration even at very low concentrations. This evidence

suggests that the ion concentrations near the polar regions of GMO and

egg-lecithin bilayers were not bufferred by regions of bound charge

(even at very low ion concentrations).

The fact that the polar head conductance of GMO cannot be

attributed to the bufferring of the ion concentrations in the polar

regions suggests that conduction mechanisms other than ion migration are

significant in deterraining bilayer conductivity. It was found that

large variations in the pH had no significant effect on the polar head

conductance of GMO bilayers. Therefore the possibility of conduction by

hydroxide and hydronium ion migration seems unli~ely.

141

The conductivity of the polar regions of GMO and egg-lecithin

bilayers was similar to that found for a wide variety of non-aqueous

polar liquids. Figure 7.6 co~pares the conductivities of the acetyl,

hydroxyl and choline-phosphate regions to that of pure liquids with

different dielectric constants. The results shown in figure 7.6 suggest

that when the electrolyte concentration is less than l Molar the

conductance of the hydrophobic - hydrophilic interface of egg-lecithin

and GMO bilayers is ~ainly due to the intrinsic conductivity of the

lipiri material.

ii) Conductivity at High Ion Concentrations

At high electrolyte concentrations sufficient concentrations of

ions partitioned into the polar region of GMO bilayers to contribute

significantly to the total electrical current. The time-constant

profiles for GMO at low and high concentrations are compared in figure

7.8. Provided the ion mobility in the polar regions was not very

different to that in the electrolyte then an order of magnitude estimate

of the partition coefficient between the bulk electrolyte and the polar

regions of GMO bilayers could then be made using the following equation:

conductivity of polar region conductivity of external electrolyte 7.2

The ~artition coefficient calculated from this expression was -8

approx. 10 which indicates that the energy barrier to ions in this

region is 40 KJ/Mole (loosely speaking 18 kT's).

Thus it appears that there are at least two conduction mechanisms

operating in the polar regions of egg-lecithin and GMO bilayers. That

is a non-ion migration type of conductance raechanism which is not

142

dependent on the external ion concentration, as well as an ion migration

type of conductance mechanism. The conduction due to ion migration is

negligible at low ion concentrations but increases linearly with ion

concentration until at high concentrations it

conductance mechanism.

7.44 The Conductance of the Ionic Double Layers

is the do~inant

In Chapter 2 the electrical properties of dielectrics in

equilibrium with electrolytes was calculated using the Nernst-Planck

equations. Provided the electric field is not a rapidly varying

function of position the ionic conduction in the membrane could be

approximated by the following expression:

7.3

As the ion concentrations in the ionic double layers are relatively

high one would expect these regions to have conductances similar to

that of the aqueous phase. From the values of the ionic double layer

capacitance obtained in the previous chapter the electrical

time-constant of the double layers should be many orders of magnitude

lower than that of the bilayer. In that case the dispersion in the

totai membrane impedance should have a contribution arising from the

different time-constants of the ionic double layers and the membrane.

However, it was found that no dispersion due to ionic double layers

was present at frequencies over .003 Hz, so it was impossible to

distinguish the time-constant of the ionic double layers fro~ that of

the bilayer. This indicated that the time-constant of the ionic double

layers was approximately equal to that of the membrane and also that the

143

conductance of the double layers was not proportional to the external

ion concentration. This result, as strange as it may seem, is in

agreement with the predictions of Smith (1977) which were based on the

solutions to the time-dependent Nernst-Planck equations. This

unexpected result is due to the fact that the electric field near the

membrane surface exhibits a large spatial variation (due to the high

values of space-charge in the double layers). Therefore the concepts

which are normally applied to macroscopic systems, in which the electric

field is relatively constant, do not apply in the electrolyte near the

membrane surface.

144

7.5 SUMMARY

Artificial BLM were produced from egg-lecithin - n-hexadecane and

GMO - n-hexadecane solutions. The impedance of artificial BLM was

measured using an updated version of the four-terminal digital technique

of Bell, Coster and Smith (1975) over the frequency range

.003 Hz to 10 KHz. ~odelling the impedance spectrura so obtained with a

Maxwell-Wagner dispersion enabled conclusions to be drawn concerning the

dielectric structure and conductance profile of the hydrophobic interior

and the hydrophobic - hydrophilic interface of lipid bilayers.

By comparing the dielectric structures of GMO and egg-lecithin

bilayers it was possible to assign values of resistance and capacitance

to the acetyl, hydroxyl and choline phosphate regions of these bilayers.

The dispersion in membrane capacitance in the frequency range

.003 Hz to 10 Hz was associated with the acetyl regions of the GMO and

egg-lecithin molecules. At higher frequencies the capacitance

dispersion was associated \'lith the hydroxyl and choline phosphate groups

of the Gl-10 and egg-lecithin molecules.

From the capacitance of each electrically distinct region within

the polar heads of the lipids it was possible to gain estimates of the

dielectric constant of these regions. It was found that the dielectric

constant of the polar head regions varied with position over the range

6-40 and was akin to that of chemically similar polar liquids.

The conductance of the regions containing the acetyl groups and

hydroxyl groups of the G~O and egg-lecithin molecules was much less than

that of purified water. Further, a 10- fold change in the external ion

145

concentrations had only marginal effects on the dielectric properties of

these regions. Thus it was concluded that electrolyte penetration into

the polar regions of these bilayers was negligible in egg-lecithin

bilayers and became significant in GMO bilayers only at ion

concentrations in excess of l Molar; with the possible exception of the

choline phosphate regions as these had relatively high conductivities.

The variation of polar head conductance with external ion

concentration was consistent with two conductance mechanisms in the

polar regions of egg-lecithin and GMO bilayers: a non ion-migration

mechanism which was dominant at low ion concentrations and an ion

migration type of conductance with was do~inant at high electrolyte

concentrations.

The partition coefficient of ions in the hydroxyl region of GMO

bilayers, calculated from the electrical conductivity at high

electrolyte concentrations, was found to be l~~

The time-constant of the ionic double layers was found to be equal

to that of the bilayer itself. This was in agreement with the

predictions of Smith (1977) based on the solutions to the time-dependent

Nernst-Planck equations.

CHAPTER 8

THE DIELECTRIC STRUCTURE OF ARTIFICIAL BLM:

I THE EFFECT OF CHOLESTEROL AND n-ALKANE INCLUSION

II THE EFFECT OF D20/H 2 0 REPLACEMENT IN THE AQUEOUS PHASE

8. l INTRODUCTION

8.2 METHODS

8.3 RESULTS

8.4

8.31 The Effect of Cholesterol on Bilayer Impedance

8.32 The Effect of D2 0/H 20 Replacement

8.33 The Effect of n-Alkane Absorption

DISCUSSION

8.41 The Location of Cholesterol in Egg-Lecithin Bilayers

8.42 The Effect of D2 0/H 20 Replacement on Bilayer Structure

8.43 The Location of n-Alkane Chains in Eg~-Lecithin Bilayers

8.5 SUMMARY

146

Page

147

150

150

150

151

151

152

152

154

155

156

147

8. l INTRODUCTION

In the previous chapter the dielectric structure of the

hydrophobic-hydrophilic interface of egg-lecithin and GMO bilayers was

characterized. This chapter is concerned with the effects on the

dielectric substructure of cholesterol and n-alkane incorporation into

egg-lecithin and GMO bilayers, as well as the effect of replacing the

aqueous phase with D2 0. The results will be interpreted on the basis of

the dielectric model of egg-lecithin bilayers developed in the previous

chapter.

Cholesterol is a major component of most biological membranes.

Though many studies on the effects of cholesterol have been made its

role in biological membranes is not well understood. In general it

seems that cholesterol acts as a moderator of membrane properties.

Cholesterol, when present in biological membranes, has been found to

reduce the thermal sensitivity in mammalian cells {Cress and Gerner,

1980) and increase ethanol tolerance in erythrocyte membranes in mice

(Chin, Parsons and Goldstein, 1978). Thus it seems that cell membranes

high cholesterol levels are much less susceptible to the action of

membrane soluble drugs and changes in the external environment.

However, cholesterol has also been found to induce local anaesthesia in

membranes which do not normally have cholesterol {Stephens and

Shinitzky, 1977).

X-ray and neutron diffraction studies { Franks, 1976, Worcester and

Franks, 1976 and McIntosh, 1978) indicate that cholesterol is aligned

perpendicular to the plane of the bilayer with its rigid ring structure

embedded in the acyl chain region, and its hydroxyl group located in the

148

vicinity of the glycerol region of the phospholipids. The exact location

of cholesterol in egg-lecithin bilayers has been the subject of some

debate. NMR studies by Darke, Finer ,Flook and Phillips (1972) on lipid

vesicles indicate interactions between the hydroxyl group of cholesterol

and the phosphate group of the lipids. This was later confirmed by

Phillips and Finer (1974). However, Huang (1976), on the basis of

previous reports concerning stereo specificity of phospholipid-steroid

interactions (eg. Brockerhoff, 1974), suggested that the hydroxyl group

of cholesterol interacts with the carbonyl groups of the phospholipids.

Subsequent NMR studies by Chatterjie and Brockerhoff (1978) support this

suggestion. A more recent study (Reiber, 1978) gave evidence indicating

that hydrogen bonding does not occur between cholesterol and lipid

molecules in lipid bilayers. Reiber went on to say that the stereo

specific interactions reported in earlier studies may have been due to

different degrees of hydration of the hydroxyl groups of the different

cholesterol isomers. However this evidence does not rule out the planar

alignment of the hydroxy groups of cholesterol and the carbonyl groups

of the lipids.

Low frequency impedance measurements similar to those described in

the previous chapter were also made on egg-lecithin bilayers containing

oxidised cholesterol. Oxidised cholesterol is a mixture of pure

cholesterol {95%) together with some of its oxidation products (see

Chapter 2). Some of the oxidation products of cholesterol have been

isolated from biological material {see Feiser and Feiser, 1959 and also

see Chapter 1). Hence, the composition and structure of egg-lecithin

bilayers, with added oxidised cholesterol, are presumably closer to that

of lipid bilayers present in cell ~embranes, than those of the pure

egg-lecithin bilayers described in the previous chapter. These

measurements also allow an investigation of whether the location of

149

cholesterol in single planar bilayers is similar to that determined from

other studies on different bilayer systems.

Neutron diffraction studies of lipid membrane preparations have

greatly enhanced our knowledge of the molecular organisation in these

systems. Such studies to a large extent have relied on the replacement

of H2 o by 02 0 both to obtain the relative phase of the various orders in

the diffraction patterns, and to determine the degree of water

penetration into the bilayer. In this method it is assumed that

replacement of H2 o by D2 0 in itself does not affect the membrane

structure although there is ample evidence suggesting that the presence

of D2 0 affects the function of living organisms. For example, Moore

(1975) found that the growth rate of bacteria is drastically inhibited

by the presence of D2 0 in the growth medium.

In order to validate this crucial assumption, experiments have been

carried out to determine the dielectric substructure of planar lipid

bilayers made in H2 o and D2 0 solutions of lmM KCl. The results of these

experiments will also be described in this chapter.

In Chapter 5 the alkane absorption properties of egg-lecithin

bilayers were investigated. An important assumption was that the alkane

chains did not penetrate the polar regions of the bilayers and thus the

bilayer area per lipid molecule could be considered constant, which can

be supported by sound thermodynaraic arguments (Gruen 1980b and 1980c).

In this study, this assumption was tested experimentally by determining

the effect of alkane on the dielectric substructure of the polar region.

If significant alkane penetration of the polar region occurs, one would

expect to detect significant changes in the dielectric structure of the

hydrophobic-hydrophilic interface.

150

8.2 METHODS

Bilayers were formed from solutions of egg-lecithin and oxidised

cholesterol (2:l mole ratio) in n-hexadecane (l5mM egg-lecithin in

n-hexadecane)

20-30°C.

0 at 40 C. Impedance measurements were carried out at

The effect of "heavy water" on egg-lecithin bilayers was

investigated by generating bilayers in an aqueous phase consisting of a

solution of lmM KCl in D2 0. In order to avoid significant exchange

between H2o in the atmosphere and D2 0 in the aqueous phase the impedance

measurements were carried out within 10 hours of exposing the D2 0 to the

atmosphere.

Bilayers containing relatively high concentrations of n-alkane

solv~nt were formed from n-dodecane solutions of egg-lecithin at 20°C.

8.3 RESULTS

8.31 The Effect of Cholesterol on Bilayer Impedance

The dispersion in bilayer capacitance and conductance for

egg-lecithin bilayers with and without cholesterol is show~ in figure l

8. l. Determination of the equivalent electrical substructural

parameters from the impedance data (see Table 8.1) revealed that the

pri~ary effects of the inclusion of cholesterol in egg-lecithin bilayers

formed from n-hexadecane was an increase in the total capacitance of the

7·0

6·8

N

E 6·6

' 1.1. E

UJ <.) z f,-l, <( I-u <( a.. <( <.)

6·2

10- 1

• LECITHIN / CHOLESTEROL

o LECITHIN

10 1 10 2

FREQUENCY HERTZ

10 3

Figure 8.1. The capacitance spectrum of representative egg-lecithin and egg-lecithin - cholesterol (2:1 mole ratio) bilayers formed from n-hexadecane solutions in lmM KCl electrolyte at 20°C. Note the increased dispersion at low frequencies for the bilayer containing cholesterol. The data represents an average of 5 frequency scans. The error bars are too small to be shown here.

TABLE 8.1

THE EFFECT OF CHOLESTEROL ON THE DIELECTRIC PARAMETERS OF EGG-LECITHIN BILAYERS

1 mM KCl 10 mM KCl 100 mM KCl Lee Lec:Chol Lee Lec:Chol Lee Lec:chol

Capacitance mF/m

6. 35 ± • l 6.6±.1 6.7±.15 6.8 ± .2 7.0±.2 7.0±.2

650 ± l 00 400 ± 50 600 ± 50 650 ± 50 540 ± 100 900 ± 200

1200 ± 200 600 ± 150 1250 ± 200 1500 ± l 00 950 ± 100 1200 ± 200

1200 ± 200 1000 ± 200 1750 ± 300 2000 ± 500 13QQ±3QO 1400±200

1100 ± 200 900 ± 150 2000 ± 300 - 1200 ± 200 1500 ± 500

Conductance mS/m 2

.5-.1.5 • 5-1 • 5 .5-2.5 1-2 1-3 2-3

(4-20) *10 ( 4-10) *10 (4-5) *10 2 (2-3) *10 2 (2-3) *1()3 (2-7) *10 3

(3-10) *10 2 (4-10) *10 2 (3-8) *10 3 (4-6) *10 3 (4-7) *10" ( 4-10) *10"

(6-10) *10 3 (7-10) *10 3 (6-9) *10" (6-7) *10" (6-10) *10 5 (6-10) *10 5

(5-9) *10" (5-10) *10" (6-7) *10 5 - (3-7) *10 6 (4-6) *10 6

Table 8.1. The capacitance and conductance of the dielectrically distinct regions of

151

hydrophobic region and a 40-50% reduction in the capacitance of the

acetyl region with a time constant greater than 0. l second.

These alterations in the substructural parameters showed that

cholesterol was indeed present in the bilayer phase and that the

previously reported absence of an effect of cholesterol on the total

measured capacitance of such bilayers (Benz and Lauger, 1977) could not

simply have been due to its exclusion from the bilayer phase.

8.32 The Effect of D20/H 20 Replacement

The dielectric dispersion of egg-lecithin - cholesterol (2:1 mole

ratio) was measured in lmM KCl solutions of 020 and H20. The dispersion

in capacitance for these different systems is compared in figure 8.2.

It is clear from these plots that the dielectric substructure of

egg-lecithin - cholesterol bilayers was insensitive to the two isotopes

of hydrogen in the water molecules. Table 8.2 shows the dielectric

parameters obtained from membrane impedance data.

8.33 The Effect of n-Alkane Absorption

The presence significant concentrations of alkanes in the bilayer

interior had very little effect on the dielectric parameters obtained

from the theoretical fitting of the impedance data to a Maxwell-Wagner

dispersion. Figure 8.3 compares the frequency dispersions in bilayer

"tl capac1 ance

solutions.

'

for bilayers formed from n-hexadecane and n-dodecane

The parameters obtained from fitting theoretical

Maxwell-Wagner curves to the data are shown in Table 8.3.

7 0

6·8

N

E 6·6

' lL E

w u z 6·4 <(

!:: u ~ <( u

6·2

o LECITHIN / CHOLESTEROL H20

• LECITHIN/ CHOLESTEROL 0 20

- THEORETICAL MAXWELL-WAGNER DISPERSION

10 1 10 3

FREQUENCY HERTZ

Figure 8.2. The capacitance spectrum of two representative egg-lecithin : cholesterol bilayers (2:1 mole ratio) formed in solutions of 0 20 and H2o containing lmM KCl at 20°c.

10 •

• HEXADECANE

o DODECANE

6.8

0

l"\j 6.6 58 ~ ~ 0

u.. E

IJJ l)

~ 56 ~ -l) ~ l)

5.4

5.2 ,.___-;n-=r-----:-------:;';;-----~r-------:'-,.------' 10- 1 FREQUENC?O Hz 1 10 101.

I

Figure 8.3. The capacitance spectrum of 2 egg-lecithin bilayers formed in lOOmM KCl. The data was obtained from single frequency scans on bilayers in equilibrium with n-hexadecane and n-dodecane solutions at 20°c.

TABLE 8.2

THE EFFECT OF D20/H 20 REPLACEMENT ON EGG-LECITHIN/ CHOLESTEROL BILAYERS

Capacitance mF /m 2

6.6 ±.1

450 ±50

600 ±lQQ

1000 ±200

850 ± 150

6.6±.1

400 ± 50

600 .tl 50

1000 ±200

900 ± 150

Conductance mS/m 2

.5-1.5 . 5-1 . 5

10-20 40-100

(4-7) *10 (4-10)

(6-12) *l O 2 (7-10)

( 4-10) *10' ( 5-10)

*10

*l O 2

*l O 1

----------------------------------------

Table 8.2. The capacitance and conductance of the dielectrically distinct regions within the substructure of egg-lecithin - cholest~rol bilayers in l mM KCl.

TABLE 8.3

THE EFFECT OF ALKANE ABSORPTION IN EGG-LECITHIN BILAYERS

C16 C12

Capacitance mF /m 2

6.35 ±.1

650 ±100

1200 t 200

1200 ± 200

3.8-5.7

750 ±50

1500 ± 400

1350 ± 300

Conductance mS/m 2

• 5-1 • 5

(4-20) *10

(3-10) *10 2

(6-10) *10'

.01-1.5

(10-20) *10

(1-6) *10 2

(1-3) *10'

Table 8.3. The effect of different solvent concentrations in the hydrophobic region on the dielectric parameters of egg-lecithin bilayers in lmM KCl.

152

8.4 DISCUSSION

8.41 The Location of Cholesterol in Egg-Lecithin Bilayers

On the basis of the studies described

concluded that the increase in the overall

in Chapter 6, it was

bilayer capacitance of

egg-lecithin bilayers when cholesterol was present, was not due to a

decrease in membrane thickness but rather to an increase in the

capacitance of the ionic double layers external to the membrane. The

apparent increase in the capacitance of the hydrophobic region

(extracted from the impedance dispersion data) also incorporated the

effects due to these same ionic double layers (see Chapter 7). By

taking the capacitance of the ionic double layers into account, the

dielectric capacitance of the bilayer was found to slightly decrease

(2-3%) with the addition of cholesterol.

The dielectric time-constant profiles of egg-lecithin bilayers with

and without cholesterol are compared in figure 8.4. The profiles were

generated in the same manner as those in Chapter 7. If either the

dielectric constant or thickness of the hydrophobic - hydrophilic

interface of egg-lecithin bilayers was altered by the presence of

cholesterol, then one would expect the time-constant profiles to be

different. The location of cholesterol within the bilayer could be

determined from the relative effects of its inclusion in the bilayer

structure upon the bilayer time-constant profile.

Examination of figure 8.4 reveals that the inclusion of cholesterol

in the bilayer structure had the effect of broadening (ie. decreasing

the capacitance) those regions with electrical time-constants in the

2 ACETYL CHOLINE - PHOSPHATE

0

-t.'.) "- -1 (_.) .,__

OI 0

-2

-3

-I. L..L ___ .J._ __ ...J.._ __ -4-__ _L __ ____.L ___ '::-__ --=-__ -:-__ ·2 ·] ·I. ·5 ·6 ·7 ·8 0 · 1 Relat,ve Distance nm

Figure 8.4. The dielectric time-constant profiles for bilayers formed from egg-lecithin (shaded) and egg-lecithin : cholesterol solutions (2:1 mole ratio) (unshaded). The profiles were generated in the same manner as those described in Chapter 7. The results presented here indicate the total variation in the time-constant profiles for 20 egg-lecithin bilayers and 23 bilayers containing cholesterol in aqueous solution~ of 1, 10 and 100 mM KCl. Note that cholesterol mainly altered the 1

capacitance of regions with electrical time-constants in the range . 1 to 10 seconds.

153

range 0.1 to 10 seconds. It can be concluded from the study described

in Chapter 7 that these regions were associated with the inner portion

of the acetyl region. The decrease in capacitance was presumably the

result of a 40% decrease in the dielectric constant in this part of the

bilayer.

Alteration of the effective dielectric constant of any region

should produce a corresponding change in the equivalent parallel

capacitance of that region. The dielectric constant of the cholesterol

molecule (with the exception of the hydroxyl group) is thought to be

2.27 (Fettiplace et al., 1971), whereas that of the lecithin acyl chains

is thought to be 2.1 - 2.2 (Huang and Levitt, 1977). Therefore the

presence of cholesterol should not significantly alter the average

dielectric constant of the hydrophobic region. However, the dielectric

constant of the acetyl region of the egg-lecithin bilayers appears to be

significantly higher (6-12) (see Chapter 7). Thus if the ring structure

of cholesterol, with its low dielectric constant, bridges the acetyl

region, an approximately 30% reduction in the average dielectric

constant of that region would result for a 2:1 lecithin-cholesterol mole

ratio in the bilayer. If the cholesterol ring structure extended into

the polar head region a similarly large decrease in capacitance of the

polar head region would be expected. The results in Table 8. l show a

reduction of 30% in the capacitance of the acetyl region without a large

alteration of the polar head capacitance. This indicates that the

cholesterol ring structure certainly extends into the glycerol bridge

region but not as far as the inner portion of the polar heads of the

lecithin molecules.

On the other hand, the hydroxyl group on the cholesterol molecule

should be reasonably polar, presumably having a significantly higher

CH CH / 3 3,

+l('CH3

·-· I o 0~11-

P::::::::o I o'\.

0-• /. \ "-• ~o

0~ "a--·"

Figure 8.5. The location of the cholesterol molecule in egg-lecithin bilayers as determined from its effects on the dielectric structure of the polar regions of egg-lecithin bilayers. The hydroxyl group of cholesterol molecule is located midway between the planes containing the carbonyl and phosphate groups of egg-lecithin.

154

effective dielectric constant than the rest of the molecule. Thus the

presence of the hydroxyl group within the acetyl region would not be

expected to produce the drawatic decrease always seen in the acetyl

region capacitance upon the inclusion of cholesterol. However the

dielectric constants of the hydroxyl group and that of the lecithin

polar head are likely to be similar and so no gross change of the

average dielectric constant (and hence capacitance) of the inner polar

head region would be expected.

The dielectric data is thus consistent with the cholesterol

hydroxyl group being located in between the phosphate group and glycerol

oxygens of the lecithin molecules (see figure 8.5). The dielectric data

presented here suggests a location for the cholesterol molecule midway

between those of Darke et al. (1972) and Chatterjie and Brockerhoff

(1978) and perhaps reflects the fact that this data refers to a

different lipid system (ie. a single bilayer in the presence of abundant

water).

8.42 The Effect of D20/H 20 Replacement on Bilayer Structure

The dielectric time constant profiles for egg-lecithin -

cholesterol bilayers formed in D20 and H20 are compared in figure 8.6.

Replacement of H20 by D20 had no detectable effect on the dielectric

structure of the hydrophobic - hydrophilic interface of these bilayers.

The total membrane capacitance in l and 100 m~1 KCl solutions was also

unaffected which indicates that both membrane thickness and ionic double

layer capacitance were insensitive to the presence of 020.

"' ('.)

\. lJ ......

°' 0 --

2

ACETYL CHOLINE -PHOSPHATE

1

0

-1

-2

-3

-4L....J. ___ _._ __ _L ___ --1-__ ----1. ___ J._ _____ ----j

·6 ·7 0 ·2 ·3 ·4 Distance

·5 nm

·1 Reio five

Figure 8.6 . The dielectric time-constant profiles for egg-lecithin bilayers containing cholesterol (2:1 mole ratio egg-lecithin to cholesterol) formed in aqueous solutions of H2 o (shaded) and 02 0 (unshaded) in l mM KCl at 20°C . The shaded and unshaded areas represent the total variation for 6 membranes in 02 0 and 10 membranes in H2 o.

155

8.43 The Location of n-Alkane Chains in Egg-Lecithin Bilayers

If n-alkanes were to partition into the polar regions of a lipid

bilayer, then the presence of a hydrophobic molecule (Er= 2. l) would

have the effect of significantly reducing the average dielectric

constant and hence capacitance of the polar regions. The presence of

n-alkanes in lipids bilayers was found to significantly alter the

structure and composition of the hydrophobic region (see Chapter 5).

However, from examining Table 8.3, it can be seen that the presence of

n-alkane in lipid bilayers had little effect on the polar dielectric

structure. From this it can be concluded that the n-alkane chains did

not penetrate into the acetyl or choline phosphate regions of

egg-lecithin bilayers. Hence the alkanes are mainly located deep within

the hydrophobic interior, in agreement with the results of neutron

diffraction experiments (White, King and Cain, 1981).

156

8.5 SUMMARY

In the present study the low frequency impedance dispersions of

bilayers formed from egg-lecithin and those from egg-lecithin - oxidised

cholesterol (2:1 mole ratio) were compared. The presence of oxidised

cholesterol in the bilayer structure caused a 5% increase in the bilayer

capacitance (in lmM KCl) over the frequency range .003-10000 Hz, as well

as a decrease in the capacitance of the acetyl region. The former

effect was due to changes in the ionic double layers external to the

membrane which has been discussed in Chapters 6 and 7.

The effect of oxidised cholesterol on the various dielectric

parameters of the bilayer hydrophobic - hydrophilic interface allowed

the location of cholesterol in the bilayer structure to be determined.

The data was consistent with the ring structure of cholesterol being

embedded in the acyl chain region and penetrating the acetyl region.

This suggested that the location of the hydroxyl group of the

cholesterol molecules was midway between the plane containing the

phosphate groups and the carbonyl groups of the lipids.

The presence of n-alkanes in egg-lecithin bilayers had a

significant effect on bilayer thickness. However, the dielectric

structure of the polar regions was insensitive to the presence of

n-alkanes. It was therefore concluded that n-alkanes are not present in '

the polar head regions df egg-lecithin bilayers and that they must be

located deep within the bilayer structure.

It was generally concluded in the present study that replacement of

H2o by 02 0 had essentially no effect on the dielectric structure of

157

egg-lecithin - cholesterol bilayers and therefore should not effect the

fidelity of the bilayer structure determined by neutron diffraction

methods.

CHAPTER 9

THE CONDUCTANCE OF LECITHIN BILAYERS

9.1 INTRODUCTION

9.2 MATERIALS AND METHODS

9.3 RESULTS

9.4

9.31 Variation of Bilayer Conductance With Area

9.32 Conductance Characteristics

DISCUSSION

9.41 Interpretation of the Area Dependent Membrane Conductance

9.42 The Effect of Different Electrolytes

9.43 The Energy Barrier to Ionic Conduction

9.44 Nature of Ionic Conduction

9.45 Possible Nature of Hydrophobic Conduction

9.5 SUMMARY

158

Page

159

161

162

162

164

167

167

168

170

174

177

179

159

9. 1 INTRODUCTION

Although the electrical properties of planar lipid bilayers have

often been studied, little is known about the nature and origin of their

intrinsic conductance. Several studies have reported that the

conductance is 'irreproducible' and quote only an order of magnitude for

the measured conductance. Even when reproducible conductance

measurements are obtained it is still difficult to distinguish between

the conductance of the bilayer component itself and the conductance of

the torus surrounding the bilayer ( Miyamoto and Thompson, 1967). Some

values reported for the area-specific conductance of egg-lecithin and

egg-lecithin - cholesterol bilayers are shown in table 9.1.

It has been proposed (Hanai, Haydon and Taylor, 1965c) that only

the very lowest measured values of bilayer conductance (approximately

.01 mS/m 2 ) reflect those intrinsic to the bilayer itself, and that any

higher values are a consequence of ''leaks" associated with the torus.

This conclusion was reached from the absence of any observed dependence

of the conductance upon the bilayer area, except for bilayers with very

low conductance.

In this chapter I describe experiments made specifically to deduce

the bilayer conductance. The dependence of bilayer conductance on area

was measured in order to distinguish between the conductance of the

bilayer and the "leak" conductance of the membrane border.

The origin of the membrane conductance was of some importance to

the study of the dielectric substructure of lipid bilayers undertaken in

this thesis. It was found (see Chapter 7) that a small portion of the

TABLE 9. l

SOME MEASURED VALUES OF THE CONDUCTIVITY OF EGG-LECITHIN BILAYER MEMBRANES

Conductivity mS/m 2 Reference

. 01 - 10

2.5 - 40

. 125 - . 19

.005 - .025

1.0

. 1 - 10

.05

1 - 10

3 - 30

. 3 - 10

1.0

3 - 15

.24 - 6

.76

.06

. 6 - . 7

.08-. 16

. 75 - 17

Hanai, Haydon and Taylor (1964)

Huang, Wheeldon and Thomson (1964)

Hanai, Haydon and Taylor (1965a)

Hanai, Haydon and Taylor (1965c)

Van den Berg (1965)

Lauger,Lesslauer, Marti and Richter (1967)

Tien and Diana (1967)

Rosen and Sutton (1968)

Simons (1968)

Ohki and Goldup (1968)

Rosenberg and Jendriasiak (1968)

Clowes, Cherry and Chapman (1971)

Coster and Smith (1974)

Ashcroft, Coster and Smith (1977)

Gutknecht and Walters (1980)

Gutknecht (1981)

Gutknecht and Walters (1981)

Ashcroft, Coster and Smith (1981)

Table 9.1. Reported values of the conductance of egg-lecithin bilayers formed from lipid solutions containing varying amounts of cholesterol. The bilayers were formed in electrolytes varying in concentration from lmM to lM.

160

dielectric substructure of lipid bilayers, determined from the membrane

impedance at very low frequencies (>.03 Hz), was dependent on whether

the measured conductance was attributed to the hydrophobic conductance

of the bilayer (i.e. the intrinsic conductance) or the presence of

aqueous channels traversing the membrane.

One way of obtaining useful information about the mechanisms

responsible for the conductance is to measure its temperature

dependence, and hence obtain the activation energy for the translocation

of charge across the bilayer. Previous attempts to measure the

temperature dependence of bilayer conductivity reported little or no

temperature dependence (eg. Hanai et al., 1965c and Simons, 1968). The

activation energies have indeed been measured for bilayers in the

presence of special additives such as various carriers and pore-inducers

(Ginsberg and Noble, 1974) and compounds thought to enhance electronic

conduction (Rosenberg and Bhowr11ik, 1969). These additives greatly

increased the conductance. However, the temperature dependence of the

intrinsic conductance of egg-lecithin bilayers themselves appears not to

have been studied in detail. Evaluation of the activation energy

permits some conclusions to be drawn concerning the concentration of

charge carriers in the bilayer, and their nature of entry and

translocation.

To investigate the mechanisms whereby charge translocation might

occur across the bilayer, the dependence of bilayer conductivity on

temperature and the type and concentration of ions in the aqueoues phase

has been measured.

Measurements of the A.C. conductance have been made at frequencies

low enough to accurately reflect that of the D.C. membrane conductance

161

and yet to allow sufficiently rapid simultaneous measurement of the

bilayer capacitance and conductance. The aim of this

identify possible mechanisms of charge translocation

bilayers.

9.2 MATERIALS AND METHODS

study is to

across lipid

Egg-lecithin, egg-lecithin - cholesterol (2:1 mole ratio) and GMO

bilayers were formed using the technique outlined in Chapter 4. GMO

bilayers were formed from squalene and n-hexadecane solutions. Bilayers

containing egg-lecithin were formed from n-hexadecane solutions of the

lipid.

The bilayer capacitance and conductance were measured

simultaneously to a precision of better than 0.1% using the digital

four-terminal impedance measuring technique described in Chapters 2 and

4. Measurements of A.C. membrane impedance were made at .1 and l Hz.

In an endeavour to determine the relative contributions that the

bilayer and torus components of the membrane make towards the total

membrane conductance, the bilayer conductance was measured as a function

of bilayer area using two methods. Both methods assumed that the low

frequency capacitance was proportional to bilayer area as has been

demonstrated previously by Hanai et al. (1965c) and Coster and Simons

(1968). The first method involved measuring me~brane capacitance and

conductance, while the planar bilayer increased in area from 20% to 100%

of its final planar area (which was approximately the area of the hole

in the septum), during the formation of the bilayer from the thick lipid

162

film. The second method involved measurements of membrane capacitance

and conductance whilst bowing the membrane under a net hydrostatic

pressure.

(1965c).

This second method was originally used by Hanai et al.

The former method was mainly used in this study to evaluate

the bilayer conductance.

Before the temperature was varied, the area dependence of the

conductance of each bilayer was checked to ensure that the area

independent contribution to bilayer conductance was not significant.

Measurements of the temperature dependent membrane conductance were

made during heating and cooling of the membrane to allow for the effects

of the ti~e-variation of bilayer conductance during the course of the

experiment.

9.3 RESULTS

9.31 Variation of Bilayer Conductance With Area

Figure 9.1 shows some results for the relationship between the

bilayer capacitance and conductance, measured at l Hz, as a function of

bilayer area (lecithin bilayers formed from n-hexadecane solutions in

mM electrolyte). It was apparent in each of these cases that the

relationship was reasonably linear and that the relation between bilayer j

capacitance and conductance was similar for planar and bowed bilayers.

This suggests that the contributions made to the measured electrical

properties by areas of the film which are not yet bimolecular, i.e. when

the bilayer area had not yet reached that of the hole in the septu~, are

negligible. Thus for the results shown in figure 9.1 it would be

0 10 20

CAPACITANCE nF

Figure 9.1. The relationship between the very low frequency capacitance and conductance of egg-lecithin bilayers formed from n-hexadecane solutions, measured at lHz. The data points shown are from two bilayers, one formed in lmM KCl and the other in lmM NH~Cl. The vertical arrow indicates the measured value of capacitance (11.4 nF) at which the bilayers were planar with an area equal to that of the hole in the septum ( area l. 7 r,1m 2 ). At this stage the area-speci fie conductances were in the range 2-2.5 mS/m 2 • The results to the left of the arrow are for different areas of the planar bilayer during formation from the thick lipid film while those to the right are for different areas of a "bowed" bilayer.

163

expected that the extrapolation of the relationship back to zero

capacitance (i.e. area=O) would give an unequivocal result for the

conductance of any element that was independent of bilayer area (i.e.

the torus). The full implications of this statement will be discussed

later. In these two examples the percentage contributions of the

conductance of the torus, relative to the conductance of the bilayer

(100% black), were 1.5% and 30% for lmM KCl and NH~Cl respectively.

Indeed for nearly all the bilayers studied in l mM electrolytes, using

fresh batches of egg-lecithin, it was found that the "leak" conductance

was less than 50% of the total membrane conductance.

Occasionally bilayers were generated from lipid solutions that had

been allowed to "go off". Bilayers formed from such solutions were less

stable and had markedly higher conductances than bilayers formed from

fresh lipid solutions (less than a few days old).

Thick lipid films, when formed in electrolytes of high

concentrations (. 10 mM), contained clusters of lipid aggregates, visible

under the binocular microscope, which appeared to be displaced into the

torus as the bilayer formed (see figure 4.5). These aggregates were

observed to remain at the boundary of the torus and the bilayer phases.

It was noticed that bilayers formed from films containing high

concentrations of these aggregates tended to possess higher "leak"

conductances. Further, bilayers formed from these films would undergo

abrupt changes in conductance after the bilayer was 100% black.

Mechanical movement (either by bowing or touching with a syringe needle)

could also induce abrupt changes in membrane conductance.

Nany times, attempts to distinguish the

conductances whilst bowing the membrane at high

bilayer and "leak"

ion concentrations

164

were frustrated by the occurrence of sudden changes in the ''leak"

current (see figure 9.2).

Therefore, instead of bowing the membrane to vary the bilayer area,

the membrane conductance was generally measured as a function of the

area of the planar bilayer during its formation from the thick lipid

film. The bilayers at this stage of their formation were certainly not

in equilibrium with the torus. However, there is evidence suggesting

that this did not affect the validity of this technique (see

discussion). Measurements of bilayer conductance obtained this way

where similar to those obtained by bowing the merabrane (see figure

9.1) but had the advantage of leaving the membrane relatively

undisturbed. Egg-lecithin bilayers formed in lmM KCl had relatively

stable electrical properties therefore most of the data was collected

from bilayers in aqueous solutions with low ion concentrations.

It was also found that the conductance generally would increase as

the membrane aged; it usually increased by 2-10% per hour.

The presence of cholesterol in egg-lecithin bilayers had no effect

on the bilayer conductance. Therefore the results obtained from bilayers

with and without cholesterol are presented together.

9.32 Conductance Characteristics

i) Effect of Different External Electrolytes

To investigate some of the possible mechanisms of charge

translocation, the conductance of bilayers in lmM electrolyte solutions

containing ions of different radii and charge was measured at lHz (see

100

-V) C -

++ +

++ +++

+

• • I •• • •• . .. -·· • •• •

+

0 L--------+----~-----~-----L----_.J_ ___ ___J

0 W 20 30 CAPACITANCE (nF)

Figure 9.2 The membrane conductance measured for a single membrane as a function of membrane capacitance (measured at 1 Hz) in 100 mM KCl. The vertical arrow indicated when the bilayer was planar with an area equal to that of the ho 1 e in the septum. The results show the re·sul ts of three consecutive runs (e) (•) (+). During the 1 ast run the bi 1 ayer formed a large "leak" during bowing. The fact that this was a "leak" can be easily seen by extrapolating the linear conductance versus capacitance relation back to the conductance intercept.

165

table 9.2). In the results presented in table 9.2 (also table 9.~,

the bilayer conductance was measured from the linear relationship

between bilayer capacitance and conductance. No correlation between

bilayer conductance and the ion radius and charge in the electrolyte was

found.

ii) Effect of Varying External pH

Table 9.3 shows the conductance of egg-lecithin - cholesterol

bilayers in lmM KCl solutions at different pH. It is apparent that the

bilayer conductance only varied slightly when the hydronium and hydroxyl

ion concentrations varied over six orders of magnitude.

iii) Effect of Increasing Electrolyte Concentration.

Egg-lecithin bilayers formed in aqueous solutions containing high

ion concentrations were found to possess relatively less stable

conductance properties than those formed at lower ion concentrations.

However, bilayers formed from GMO had stable mechanical and electrical

properties over the entire range of ion concentrations employed in this

study. The impedance measurements of GMO bilayers, at high electrolyte

concentrations, were not bedeviled by abrupt "jumps" in bilayer

conductance of the type shown in figure 9.2. The conductances of

egg-lecithin - cholesterol and GMO bilayers measured at .l Hz in

different external ion concentrations are shown in figure 9.3. It

should be noted here that the me~brane conductance at Hz was slightly

higher than that measured at .lHz (see dispersion data in Chapter 7). A

1O~ fold increase in external ion concentration had only a slight effect

on bilayer conductance. At high ion concentrations it can be seen that

GMO bilayers had lower conductances than egg-lecithin bilayers

10

• GMO

o LECITHIN

~ 10 I t E

" VJ

E

t ll.J u < q I---. 07

! u :)

l a < C) u

007 70-J 10-2 10-1 1 10

ELECTROLYTE CONCENTRATION Mol/m3

Figure 9.3. The effect of varying KCl concentrations in the external aqueous phase on the conductivity of egg-lecithin and GMO bilayers (measured at .1 Hz). The bilayer conductivity of egg-lecithin bilayers was determined from the variation of membrane conductance with area at 40°C. The values obtained for GMO bilayers are the tot~l m~mbrane conductance (measured at . l Hz) at 20°C.

TABLE 9.2

THE EFFECT OF DIFFERENT ELECTROLYTES ON BILAYER CONDUCTANCE

Salt Conductance (mS/m 2 )

Measured # Theoretical

KCl l. 7 ±. 4 ( 18) 3.10-21

NH.,Cl 5.0 ± l. 3 ( 5) 3 • lo- 2 I

TMA.Cl 2.7 ±.2 ( 5) 2. 10- 9

TEA. Cl l. 5 ±. l ( 5) 2. 10- 3

MgSO., 2.2 ±.4 ( 6) 6. 10-60

Kl 6.0 ±l.5 ( 5) 2 •lo- I 6

Table 9.2. The values of the measured conductivity (at l Hz) of egg-lecithin bilayer formed from n-hexadecane solutions at 40°C. # refers to the number of membranes measured. The conductance was determined from the linear relationship between bilayer conductance and capacitance. The theoretical values of bilayer conductance were calculated from the crystal radii if the ions using equations 9. 1,9.2 and 9.3 (see text).

TABLE 9.3

THE EFFECT OF pH UPON BILAYER CONDUCTANCE

pH Conductance (mS.m 2 )

#

4.3 l. l ±. 3 (3)

5.2 l. 7 +_. l ( 5)

6.0 l. 5 ±.4 ( 5)

7.0 1.6 -1:. 5 ( 9)

8.0 2.2 ±.4 ( 7)

9.0 3.0 ±. l ( 5)

9.8 3.0 ±.3 ( 4)

Table 9.3. Bilayers were formed in lmM KCl solutions of varying pH at 40°C. The conductance values were obtained from the measured impedance at l Hz.

166

iv) The Dependence Upon Temperature

The bilayer conductance was found to reversibly increase with

increasing temperature. The membrane conductance responded rapidly to

changes in the temperature. The rate of change of membrane conductance

seemed to be limited only by the thermal lag in the membrane

environment. The bilayer conductance,G, (normalised with respect to the

capacitance, C, measured simultaneously) was found to vary exponentially

with inverse temperature (see figure 9.4). The effect of a slight

time-dependence of bilayer conductance during heating and cooling of the

membrane can be seen in figure 9.4.

The Arrhenius plots of the inverse time constant (i.e. G/C)

obtained from impedance measurements at . 1 and 1 Hz are shown in figure

9.5 for egg-lecithin and egg-lecithin - cholesterol bilayers. It was

apparent that for those membranes which had reached equilibrium with the

torus, the Arrhenius plots were substantially linear. The slopes of the

Arrhenius plots were not dependent on the frequency at which the

measurements were made. The activation energy obtained from the slopes

of such plots, for nine different bilayers, was found to be independent

of their absolute conductance and had values of 35±2 KJ/mole.

Occasionally a bilayer would form that had an exceptionally low

conductance. The activation energy obtained from these membranes was

identical to that found. for membranes with higher conductances (eg. see

figure 9.5).

10

. 9

. 8

.7

.6

.5

- .I.

. 3

. 2

• • •

• • ••

•• •

3.0

.. • • • •• •• • ••

• • • ••

•• • • • • •

• • • •

• •

• •

• • • • • • • • • START

.. •• •

3.1

•• • •• .. ••

•• •• •• •

J.3

• • .. .. • • ••••

ll

r---~--~---~---~----.---~---r-- ----, ffJ 50 lO 30 20

T °C

Figure 9.4. An Arrhenius plot of the conductance (normalised with respect to the capacitance measured simultaneously) of a single lecithin/cholesterol bilayer formed with ~-hexadecane in lmM KCl (measured at lHz). Measurements shown on the figure were made when it was established that the "leak" conductance of the membrane was relatively small. The values of conductance shown here were obtained while the temperature of the electrolyte was increasing or decreasing. From the final slope EA was calculated to be 38 KJ/mole. The area specific conductance at 40'-t was 2.6 mS/m 2 •

(.'.)

" u

C -

0

-,

-2

0 0

-3

- t, ~-

-5 0

-6L_ ___ ~--'=-------='-:~---~~----J~-~----J,_....,.-103/ T(ti K)

so t.O 30 20 70

TEMPERATURE ( ° C)

Figure 9.5. The Arrhenius plots of bilayer conductance measured at . lHz (O), and lHz (e) for a number of representative bilayers. Note that the slope of the graphs are not dependent on the absolute value of bilayer, conductance. Also it can be seen that the activation energy determined from the slope of these plots was independent of the frequency at which the measurement was made.

167

9.4 DISCUSSION

9.41 Interpretation of the Area Dependent Membrane Conductance

Most of the measurements of bilayer conductance in this study were

made while the bilayer was forming from the thick lipid film. The fact

that the bilayers were not in equilibrium with the torus during

measurements of the bilayer conductance did not effect the validity of

the measurement. This assertion is based on two observations. Firstly,

the bilayer conductance responded much more rapidly to changes in

temperature than did the bilayer composition (as determined from bilayer

capacitance; see Chapter 5). Secondly, the bilayer conductance

determined from this technique was the same as that determined from

bowing the membrane, allowing the bilayer to come to equilibrium with

the torus at each stage of bowing. From this data it seems that the

approach to equilibrium between ions in the bilayer and the external

aqueous phase was much more rapid than the approach to equilibrium

between the lipid and alkane molecules in the bilayer and torus.

From the relation between membrane conductance and bilayer area one

could obtain an unequivocal value for the area-independent membrane

conductance. However it should be pointed out that the area-dependent

membrane conductance determined from this relation need not entirely

exclude border "leaks" .·

For example, one possible conductance mechanism giving rise to area

dependent "leaks" could be the following: if stable transmembrane

pores in the torus are in the form of non-disperse lipid aggregates

that reside ~ainly at the interface of the bilayer and torus (presumably

because the torus is thinnest at this point) and if the magnitude of

168

the "leak" is proportional to the concentrations of these aggregates

then the following could occur.

As the bilayer area increases (at the expense of the thick lipid

film) the lipid aggregates in the thick lipid film get trap~ed at the

bilayer-torus interface. The number of clusters at the interface will

increase proportionally to the bilayer area. This would cause the

"leak" current in the torus also to increase proportionally with bilayer

area.

Though the proposed mechanism is by no means substantiated,

particulate matter of some form (possibly lipid aggregates) was present

in thick lipid films (see figure 4.5) especially at high electrolyte

concentrations. Further these particles did get swept into the

bilayer-torus interface and significantly slowed the rate of bilayer

formation. It was also observed that bilayers with high concentrations

of these particles had relatively high ''leak" conductances.

Determination of the bilayer conductance from the linear relation

between membrane conductance and capacitance does not conclusively

exclude the effects of the torus "leak". However, it does give a closer

estimate of the bilayer conductance than that obtained from the total

membrane conductance.

9.42 The Effect of Different Electrolytes.

i) The Effect of Different Ion Species

To give some insights into the mechanism responsible for charge

translocation through the bilayer, we can now examine how the measured

169

conductance is dependent upon the radius and charge of the ionic species

present in the external aqueous solution. Table 9.2 indicates that a

three fold increase in cationic radius (from potassium to

tetramethylammonium (TEA)) has only a slight effect upon bilayer

conductance. Similarly replacing the monovalent electrolyte with one

that only contained divalent ions (MgSO~) also had little effect.

If the conductance was a consequence of ions passing through the

hydrophobic interior of the bilayer, then the area specific conductance

could be calculated from the Nernst-Planck equations. Provided that the

electric field in the hydrophobic region is reasonably constant, the

conductance, G, of the bilayer is given by the following equation:

9. l

where 11 q11 is the magnitude of the electronic charge, 11 k11 is the

Boltzmann constant and, 11 0 11 , is the thickness of the hydrophobic region.

11 c- 11 and 11 D- 11 are the concentration and diffusion constants of the ionic l l

species, i, respectively. A detailed treatment of the derivation of

equation 9. l is given in Chapter 2 (see equations 2.6-2. 14).

The ion concentration in the hydrophobic region can be calculated

using the following equations:

9.2

where t.U = z 2q2 {-1 - _l } 8ne: e: a e: e:

o r m w 9.3

where 11 t.U II here is the Born energy difference between the aqueous

phase and the bilayer interior. The activation energy for the

170

translocation of an ion through the hydrophobic interior will be at

least equal to that of the Born energy of partitioning.

In Table 9.5 the theoretical values of the bilayer conductance are

given. It is immediately apparent that the discrepancies between the

theory and experiment, both in magnitude of the conductance and the lack

of a strong variation in different electrolytes, are so large that

conduction due to "naked" ions through the hydrophobic region can be

discounted. Possible alternative forms of ion translocation mechanisms

across lipid membranes will be investigated later.

ii) The Effect of Different Ion Concentrations

It is difficult to distinguish a definite relation between bilayer

conductance and ion concentration from the data presented here.

However, the relative lack of effect of varying ion concentrations on

bilayer conductance is similar to that reported in some previous studies

(eg. Coster and Smith, 1974). The fact that GMO bilayers had -lower

conductances than egg-lecithin bilayers at high ion concentrations was

probably a consequence of the relative stability of the conduction

properties of GNO bilayers.

9.43 The Energy Barrier to Ionic Conduction

The value of activation energy, EA, for bilayer conduction deduced

from the present experiments is 35±2 KJ/mole I which is significantly

higher than that expected for ion diffusion in bulk water. This

suggests that the ionic conductance reported here is not associated with

macroscopic water channels as would be expected from "leak" paths in the

membrane torus. Similar values of EA were obtained from bilayers

171

possessing high (2m5/m 2 ) and low ( .01 mS/m 2 ) conductances. This

indicates that similar mechanisms were responsible for conduction in

both high and low conductance bilayers. This lends support to the

proposal that the conductance reported in this study is due to the

bilayer component of the membrane.

The measured activation energy for electrical conduction does not

itself reveal anything directly about the mechanism of charge

translocation through the membrane. However, it does allow us at least

to narrow down the range of possible mechanisms on the basis of

consistency of the predicted activation energy with that obtained

experimentally.

We consider here various mechanisms whereby ions can cross the

hydrophobic region of a lipid bilayer, and hence derive the minimum

activation energy expected for each. The mechanisms themselves were

originally discussed by Parsegian (1969) and MacDonald (1976).

i) Hydrated Ion Translocation.

The electrostatic energy difference is reduced as the radius of the

ion increases, and will thus be lower for hydrated ions than for 11 naked 11

ions. However, the partitioning of hydrated ions into the hydrophobic

interior of lipid bilayers involves the interfacial free energy of the

11 bubble 11 of water surrounding the ion in the hydrophobic phase as

suggested by MacDonald (1976) and Ashcroft and Coster (1978). The

energy difference creating additional oil-water interface, 6U5, plus the

electrostatic self energy, 6U, is given by:

172

= 4nyb 2 + z 2 q2 /8n£ E b {-1 - _l } 0 m E E m w

9.4

where "y" is the interfacial tension between the hydrophobic phase

and "b" is the radius of the water "bubble". "4U" increases with the s increasing radius of the hydration shell associated with each ion.

Another contribution to the total energy difference for an ion in

the aqueous phase and the hydrophobic region of the membrane, 4UH,

arises from the change in the free energy of hydration due to

differences in the hydration number (and hence radius) of the hydrated

ion in these two phases. As discussed by Ashcroft and Coster (1978)

this latter contribution probably does not play a very significant part

provided that the hydration number in the case of H+ does not drop below

3 or 4 (Ashcroft and Coster, 1978). For K+ and Cl- which have larger

radii, this will be even less significant as the hydration energies for

these ions are ~uch smaller than for H+. Therefore the total energy

difference will have a minimum value for some value of "b'' (MacDonald,

1976). This will occur when:

9.5

At a radius given by:

9.6

For y = 0.05 J/m and then b=0.34 nm and 4UT=l40 KJ/mole. Thus the

minimum possible energy difference for an ion enclosed in a "bubble" of

water in the membrane, relative to when it is in aqueous solution, is

still very much in excess of the measured activation energy.

173

ii) Formation of a Transmembrane Pore During Translocation.

The energy difference for translocation through the hydrophobic

region would be a great deal smaller if the ion could traverse through a

"pore" of higher dielectric constant. Thus for a cylindrical pore of

radius b, which contains water and spans a hydrophobic region of width

y, the electrostatic energy difference for an ion in the center of such

a pore and in the aqueous solution (see Parsegian, 1969) is given by:

9.7

where "a" is a geometrical constant depending on the form of the

pore. a= 0.175 for a cylindrical pore. For~ation of such a pore again

involves creating an interface between the aqueous pore and the

hydrophobic region. The interfacial free energy for this is given by:

9.8

Again, since "tiUs" increases with pore radius and "tiUE" decreases

with pore radius, a minimum energy exists.

point is given by:

The radius of this minimum

For a hydrophobic region, thickness 6=3nm, the optimum radius,

b , is equal to 0. 14nm. The minimum possible energy activation

energy for this process is then 160 KJ/mole.

This again is large compared with the experimental value. Further,

a pore of this radius is comparable in dimension to the ionic radii of

174

many ions, and for many ions therefore the pore would need to be larger

with a concomitant increase in II tiU 11•

T

iii) Ion Inside a Pre-Existing Transmembrane Pore.

The ion translocation itself may not involve the process of

creating a pore in the hydrophobic region, but proceeds through a

pre-existing pore formed by an independent process, which is part of the

intrinsic structure of the lipid bilayer. In this case the activation

energy for the translocation process will be far less. The Born energy

remains to be considered, and for a water filled pore is given by

equation 9.7. If the ionic diffusion in these pores is similar to that

in the bulk aqueous phase then the activation energy for diffusion

through the electrolyte in the pore (considered on a bulk phase) will be

"' l 7 KJ /mole. The remaining 18 KJ/mole required to match the

experimental value of 35 KJ/mole could be attributed in part to a

temperature dependent pore population as well as to the electrostatic

energy difference associated with the passage of an ion through the

hydrophobic interior via a narrow aqueous pore. The minimum radius of

the pores that would completely account for the remaining 18 KJ/Mole is

"'l nm.

9.44 Nature of Ionic Conduction

The vast difference between the calculated conductivity of ''naked"

ions in the bilayer and that measured experimentally was sufficient

evidence to completely discount the possibility of charge translocation

through the membrane by dehydrated ions. Further, the activation energy

of bilayer conduction was sufficiently low to also eliminate the

possibility that ions could pass directly through the hydrophobic

175

interior. It could also be concluded from the experimentaly measured

value of the activation energy that the charge translocation through

macroscopic aqueous channels was unlikely. Therefore it seems as though

the main charge translocation mechanism in these bilayers was ion

conduction via narrow aqueous channels. The lower limit to the radius

of these putative channels was estimated at l nm. The investigations

into the area-dependent membrane conductance did not conclusively locate

the aqueous channels as being in the bilayer or torus components of the

membrane.

Here, various consequences of the presence of these putative pores

will now be considered.

If the pores are evenly distributed across the surface of the

bilayer one can calculate the average number of pores which would be

consistent with the measured value of the area specific conductance

(approx. l lilS/m 2 ) • The conductance of each pore (G) can be estimated p

from its geometric dimensions and the average ion concentrations in the

pore (c ) via the following equation: p

9. 10

The equilibrium value of "c " is related to the external ionic p

concentration, c , via the relation: 0

For U =18KJ/mole and lml~ KCl external 0 concentration of ions in the pore will be lµM at 40 C.

9. 11

electrolyte the

The value of Gp

is then approx. 1.5x10- 1 ~ S/w 2 if b=lnm and 6=3nm. Approximately 6.10 10

176

pores/ra 2 would then be necessary to produce the observed conductance.

These pores would be separated by an average distance of 4µm, and would

only occupy 2.10~ % of the bilayer area. The area occupied by pores

will thus be so small that the pores would contribute insignificantly to

the total membrane capacitance.

Cass and Finkelstein (1967) found no evidence for the existence of

transmembrane pores on the basis of their study of the water

permeability of lipid membranes. However, the additional amount of

water present in the bilayer due to these putative pores represents an

average concentration of less than 10-~ % w/w which would be negligible

compared to that normally "dissolved'' in the bilayer. Hence

of variations in water channel population, and hence

conductance, would not be possible from measurements of the

conductivity of the bilayer.

detection

bilayer

hydraulic

It is of interest to note that the properties of membrane

conduction in this study and the hydraulic conductivities of bilayers

reported in other studies (eg. Fettiplace, 1978, and Finkelstein and

Cass, 1967) are distinctly different. The activation energy of

hydraulic conduction in other studies is significantly higher than that

found for charge translocation in this study. Further, the activation

energy and magnitude of the hydraulic conductivity in lipid bilayers was

sensitive to the bilayer composition, whereas the charge translocation

was insensitive to bilayer composition. Some measured values for the

activation energy of the hydraulic conductivity of bilayers are in the

range 55-61 KJ/raole for lecithin - cholesterol (Redwood and Haydon,

1969, Price and Thompson, 1969 and Graziani and Livne, 1972).

177

Pore formation in egg-lecithin bilayers may arise from the presence

of trace impurities such as lysolecithin which is known to be a pore

former (Israelachvili et al., 1980) and is also known to drastically

increase bilayer conductance (Van Zutphen and Van Oeenen, 1967).

Lysolecithin can be produced as a breakdown product of egg-lecithin.

That would account for the fact that bilayers formed from "old" lipid

solutions had markedly higher conductances. It is possible that trace

amounts of lysolecithin in fresh egg-lecithin solutions may be

responsible for transmembrane pores. However,this proposal is difficult

to reconcile with the lack of effect of membrane composition in the

aqueous channel population.

It should be noted that conduction through "pores" in the torus

still remains a possibility. These pores could be present as convoluted

aqueous channels through clusters of mechanically stable lipid

aggregates which are trapped in the torus due to surface tension

effects. The concentration of lipid in the membrane torus are far in

excess of the lipid critical micelle concentration. Therefore it is

reasonable to assume that there is an abundant supply of lipid

aggregates available for pore formation - the area dependence of the

total conductance of such pores has been discussed earlier (see section

9.41).

9.45 Possible Nature of Hydrophobic Conductance

I No definite conclusions as to the nature of the intrinsic

conduction properties (i.e. the conduction of the hydrophobic region) of

the bilayer can be ascertained from experiments described here-except

that it must be very small compared to the measured values of bilayer

conductivity. The intrinsic conductive properties of the

178

hydrophobic - hydrophilic interface were investigated in Chapter 7. It

was concluded that the conductance of the polar regions of GMO and

egg-lecithin bilayers formed in aqueous solutions of l to 100 mM KCl was

not due to an ion migration mechanism. The results were consistent with

the polar head conduction being an intrinsic property of the bilayer

material. If this is so then the intrinsic conduction in the

hydrophobic region might also be a result of the "intrinsic'' conductance

of the lipid material.

In order to obtain the dielectric substructure of the bilayers from

their impedance dispersion, the non-hydrophobic conductance of the

bilayer must be subtracted from the bilayer impedance (see Chapter 7).

From data presented in this chapter it appears that the hydrophobic

conductance of the bilayer is much smaller than the total bilayer

conductance, and therefore virtually 100% of the D.C. conductance should

be subtracted from the impedance data before extracting information

concerning the bilayer dielectric structure from the impedance data (see

Chapter 7).

179

9.5 SUMMARY

In this chapter the conductance characteristics of egg-lecithin,

egg-lecithin-cholesterol and GMO bilayers were measured. These

characteristics were then compared to those expected for various

mechanisms for the translocation of ions across the hydrophobic region

of lipid bilayers.

Any area independent component of membrane conductance (which was

presumably due to torus 11 leak 11 conductance) was subtracted from the

membrane conductance.

Varying the ion species, ion concentrations and pH in the external

aqueous phase had relatively little effect on the conductance over a

wide range of ion concentrations, ion crystal radii and charge. On the

basis of these results it was concluded that translocation of either

"bare'' or hydrated ions through the hydrophobic region could not account

for the measured bilayer conduction.

The temperature dependence of bilayer conduction was measured and

was found to vary exponentially with inverse temperature. The

activation energy of ion conduction ( obtained from Arrhenius plots of

bilayer conductivity) was found to be 35 ±2 KJ/mole. This value was

much lower than that predicted for mechanisms involving ion

translocation through a hydrophobic region. The data was consistent

with ion conduction via very small aqueous channels a few nanometres in

diameter. The activation energy of conduction was sufficiently high to

dismiss the possibility of ion conduction through macroscopic water

channels in the torus.

180

It was also concluded that the hydrophobic conductance of the

bilayer was negligible compared to the total bilayer conducance. In

view of the fact that the conductivity of the polar regions of

egg-lecithin and GMO bilayers were largely independent of the external

ion concentration it was postulated that the very small intrinsic

hydrophobic conductance of the bilayer may be a property of the bilayer

material rather than due to ion migration.

CHAPTER 10

ENERGY OF FORMATION OF LIPID BILAYERS

10. l INTRODUCTION

10.2 THEORETICAL CONSIDERATIONS

10.21 Thermodynamics of Lipid Partitioning Between the Bilayer and Torus

10.22 Free Energy of Lipid Bilayers

10.3 METHODS

10. 4 RESULTS

10.5 DISCUSSION

10.51 Rationale for the Method

10.52 Energy of Formation: Temperature Dependence

10.53 Bilayer Tension: the Effect of Electrolyte Concentration and Cholesterol

10.6 SUMMARY

181

Page

182

185

185

186

187

190

192

192

193

198

199

182

10. l INTRODUCTION

The plasma membranes of nearly all plant cells and most animal

cells are in a state of tension. The importance of this variable in the

functioning of cell membranes in vivo has not yet been demonstrated.

However, thermodynamic considerations suggest that the solubility and

activity of membrane-bound enzymes is dependent on the tension and

elasticity of the lipid bilayer (Bates and Wolfe, 1980, and Gruen and

Wolfe, 1982). Van Deenen et al. (1976) found that the partitioning of

phospholipase into lipid monolayers varied drastically with only small

(5mN/m) changes in the lateral pressure of the monolayer.

The mechanism for local anaesthesia due to alkanes has been

associated with variations in thickness of the bilayer component of the

plasma membrane of squid axon (Haydon et al., 1977). However, it is

likely that this mechanism does not apply to other groups of local

anaesthetics (eg. Aminobenzoic acid ester types - see Chapter 11). The

modulation of membrane function caused by changes in the bilayer tension

induced by local anaesthetics is an alternative mechanism for

consideration. Part of this chapter is devoted to characterizing the

lateral tension properties of egg-lecithin and egg-lecithin -

cholesterol bilayers.

The tension required to rupture most biological and artificial

membranes, studied to date, ~s about 5 mN/m. Assuming that the area

elastic modulus of the bilayer component of membranes is of the order of

100 mN/m (eg. see Wobschall, 1971) the maximum elastic deformation that

would occur in most biological membranes is about 2-4%. Variations in

bilayer area far in excess of this have been frequently observed in

183

artificial lipid bilayers (Hanai et al., lSS5c and Coster and Simons,

1968) and in plasma membranes of living cells (Wolfe and Steponkus

1981 and Curtain, unpublished results). Previous studies have found

that over sufficiently long times, the stress-strain relation in

artificial bimolecular lipid membranes (Coster and Simons, 1968) and the

plasma membranes of rye protoplasts (Wolfe and Steponkus, 1981) follows

a surface energy law.

Wolfe (1979) applied the Gibbs equation to the equilibrium that

exists between lipids in the torus and the bilayer of artificial lipid

membranes (I discussed the existence of this putative equilibrium in

Section 10.2). His calculations showed that the equilibrium tension in

lipid bilayers is to a large extent dependent on the condition of the

lipid in the torus. Whilst considerable progress has been made in

studying the bulk solution-monolayer equilibrium of different

amphiphiles at oil-water interfaces no detailed experimental

investigation has been made on the thermodynamics of lipid incorporation

into egg-lecithin bilayers. This chapter describes measurements of

bilayer tension of bilayers formed from GMO and egg-lecithin at

different temperatures. The results are interpreted in terms of Wolfe's

modelling of the bilayer-torus equilibrium.

The bilayer tension is derived from the variation of transmembrane

hydrostatic pressure with the inverse of the radius of curvature of a

bowed BLM. The radius of curvature is derived from the area of the

bowed BLM; the area itself being determined from the membrane

capacitance. This technique is an adaption of that of Coster and Simons

(1968); the improvements made to the method of Coster and Simons will be

described here also.

184

The energy of formation of lipid-water interfaces has been

measured using a variety of techniques. Tien and Diana (1967), Coster

and Simons (1968) and Wolfe (1979) deter~ined the bifacial surface

energy* of planar lipid bilayers by measuring the surface area increase

upon bowing the membrane. Wobschall (1971) used the periodic bowing

version of the technique of Coster and Simons (1968). At the low

frequencies of bowing, where there was sufficient time given for lipid

molecules to transfer into the bilayer, the bilayer stress-strain

relation followed a surface energy law. At high frequencies, where the

number of molecules in the bilayer remained essentially constant, the

bilayer stress-strain relation followed an elastic law. From this,

esti~ates of the elastic modulus of the bilayer could be obtained.

Haydon and Taylor (1968) calculated the surface tension of egg-lecithin

bilayers from measurements of the angle of contact between thick lipid

films and lipid bilayers. A novel method for measuring the resting

tension in flat bilayers was developed by Grabowski and Cowen (1977)

whereby the modes of thermal excitations in the bilayer were measured by

laser light scattering.

* The monofacial surface energy is that due to the lipid

monolayer-water interface. However, in these studies the surface energy

of two apposing monolayers was measured. Thus the bifacial surface

energy was double that of the monolayer.

185

10.2 THEORETICAL CONSIDERATIONS

10.21 Thermodynamics of Lipid Partitioning Between the Bilayer and

Torus

Single planar lipid bilayers, by virtue of their rigid support at

the septum, can maintain a net lateral tension. The analysis presented

here follows that of Wolfe (1979) who considered the equilibrium between

lipids in the torus and bilayer phases of the membrane at a constant

tension and applied the Gibbs equation to the equilibrium between lipids

in the bilayer and torus.

Bilayers are only stable when the lipid concentration in the torus

is far in excess of the critical micelle concentration. Therefore

essentially all the lipids are in aggregates of some form. The

treatment here considers the lipid molecules in the torus to exist in

monodisperse aggregates of size N (see figure 10.l).

The Gibbs free energy of the lipids in the torus and bilayers

phases at equilibrium are equal. Therefore one can write the following

expression:

kT XB kT XN 0 ( ) + ln ° + ln µBy --;;- --;;-=µN N N 10. l

Hhere "XN" is the lipid concentration that exists in aggregates

with aggregation number, N, and tiµ\(y) is the standard cherni ea l I

potential of the lipids in the torus and the bilayer at a tension "2y".

The translational entropy of the lipids in the effectively infinite

bilayer is much smaller than that of aggregates, therefore, this term

can be neglected in equation 10.l. Part of the standard chemical

potential of the lipids in the bilayer arises from its interfacial

e b

Figure 10.l. This figure, taken from Wolfe (1979), depicts the model used in this chapter to describe lipid partitioning between the bilayer and torus.

(a) Lipid Vesicles

( b) Inverted Micelle s

( C) Lipid Monomers

( d) Lipid Bilayer

( e) Alkane

186

energy. By considering this term separately and rearranging equation

10.2 we get:

X ( o) kT l n N

ya :: 6µ - r r 10.2

where "a" is the lipid head group area per molecule in one

interface and" 6µ~' is the tension independent component of the standard

chemical potential difference between lipids in the bilayer and torus.

The contribution of the entropy of mixing term in equation 10.2 to the

bilayer tension "2y" is shown in table 10.1.

10.22 Free Energy of Lipid Bilayers

An increase in the surface area of a lipid bilayer under a tension

"2y" requires the input of energy. The work done in increasing the

bifacial surface area against this tension goes into increasing the free

energy of the bilayer. The increase in free energy of this phase can

occur via an increase in the free energy per lipid (ie. arising from

elastic deformation) or due to the incorporation of new material into

the bilayer. The relative contributions of these two mechanisms to the

total free energy of the bilayer will now be considered.

i) Elastic Deformation

A lipid bilayer deforms elastically when its total number of

' molecules is conserved. Defining the area-elastic modulus of a bilayer

"KA" (see equation 10.3) one can calculate the work done to increase the

area of a bilayer by a relative amount "6A/A" (see equation 10.4).

y = K 6A A A 10. 3

10. 4

N

10

50

100

500

1000

TABLE 10. l

-2kT/aN ln{Xn/N)

Xn . 0 l

61

9.2

2.2

l.2

.29

. 15

{mN/m)

.001

92

12.2

2.8

l. 5

.35

. 18

.0001

122

15

3.5

l.8

. 41

. 21

--------------------------------------

Table 10. l. The contribution of the entropy of mixing of inverted micelles in the torus to the bilayer tension {2Y) for egg-lecithin shown for different values of micelle aggregation number and micelle concentration.

187

ii) Energy of Formation

Work done on lipid membranes can also go into incorporating extra

material into the bilayer; hence the formation of extra bilayer area.

If bilayers were to expand in this way then the bilayer tension at

equilibrium would be independent of surface area and the energy of

bilayer formation, Ef, would be given by the following equation:

10.5

The elastic modulus of some artifical bilayer membranes has been

calculated to be in the order of 100 mN/m (Wobschall, 1971). Assuming

that the maximum tension encountered in stable bilayers is about 5 mN/m

(Goldup et al, 1970), then the free energy contribution arising from

elastic deformation is about .1 mJ/m 2 , whereas that due to energy of

formation is about 5 mJ/m 2 • It is apparent from these considerations

that the elastic properties of lipid bilayers can be ignored in this

study.

10.3 METHODS

Egg-lecithin bilayers were formed by the method outlined in

previous chapters. In order to measure the bilayer tension it was

necessary to bow the membrane under a hydrostatic pressure difference.

Once the membrane had become bimolecular over most of the aperture the

membrane was bowed by withdrawing a small amount of electrolyte from the

outer compartment via a microlitre syringe.

188

When the bilayer area had increased sufficiently, part of the

hydrostatic pressure was then removed so that the membrane flattened

slightly. It was assumed that equilibrium between the bilayer and torus

was achieved when the membrane capacitance reached a steady value. When

equilibrium was attained the capacitance 11 C11 of the spherical bilayer

was measured and then the membrane was flattened by adding a known

quantity of electrolyte to the outer compartment. At this point two

variations of the protocol were tested. One involved unbowing the

bilayer in several stages, allowing the bilayer and torus to come to

equilibrium at each point and recording the capacitance. Then, having

completely flattened the bilayer, the capacitance, "C II of the planar 0 '

bilayer was measured. Alternatively, the bilayer was flattened in one

step, giving insufficient time for the torus to move before taking the

capacitance of the flat membrane. Achievement of the flat bilayer could

be easily detected under reflected light since the bilayer acts as a

spherical mirror whose focal point goes to infinity when it is flat.

This second variation of the method was found to be most suitable for

measuring bilayer tensions in this study (see discussion). The area of

the bowed membrane was determined from the membrane capacitance using

the following equation:

10.6

Hhere "A II is the area of the flat membrane. The radius of 0

curvature, R, of the membrane, just prior to being flattened, was then

determined from the area of the bowed membrane at equi l i br'i uril, A,

relative to its projected area normal to the septum (assumed to be equal

to "A") using the following equation: 0

_l R = A{4n(A - A ) } 2

0 l O. 7

189

The hydrostatic pressure difference across the membrane whilst

bowed could be calculated from the volume of the aqueous solution, v A

that was added to the outer compartment when flattening the membrane.

This was done using equation 10.8:

10.8

Where 10. 9

~Jhere 11 p 11 and "g" are the density of water and the acceleration due

to gravity. "S1" and 11 S2 11 are the areas of the air-water interface in

both water compartments and 11 v II is the volume of water displaced by the m

bowed membrane. The tension of the bowed bilayer with a radius of

curvature, R, supporting a pressure difference "tiP" is given by the

Laplace-Young equation:

2y = ~ 10. 10

The procedure was repeated 3-5 times for each membrane for

different relative increases in bilayer area during bowing in the range

10-60%. Measurements of bilayer tension at temperatures below that

necessary for rapid bilayer formation were made by bowing the membrane

at elevated temperatures. Thus after equilibrating the membrane at a

lower temperature, the capacitance was measured and then the membrane

was flattened. A period of 15-20 minutes was allowed for equilibration I

of the membrane at each temperature.

This method warrants some discussion.

is reserved for a later section.

However, such a discussion

190

10.4 RESULTS

Thick egg-lecithin films became bimolecular over most of the

supporting aperture in 5-15 minutes. Upon bowing the membrane, there

was no immediate increase in capacitance. Thus the initial area

increase was entirely due to an increase in the area of the torus. The

formation of the additional bilayer due to bowing proceeded at about the

same rate as that found for bilayer formation in flat membranes. At

different degrees of bowing the membrane torus was anchored at different

points on the septum (see figure 10.2). The capacitance would settle

within 10-20 minutes of bowing. However, during this time, different

evaporation rates in the aqueous phases in both water filled

compartments invariably produced additional, unknown, hydrostatic

pressure across the membrane. Sometimes, particularly at elevated

temperatures, the time variation in hydrostatic pressure due to

evaporation losses was sufficiently rapid to bow the bilayer into a

hemisphere over a period of 30 minutes. Capacitance changes upon

progressive flattening of an already-formed bilayer~re much more rapid.

For this reason it was fairly difficult to measure hysteresis curves of

membrane capacitance during bowing and flattening. A plot of the time

variation in bilayer capacitance during bowing and flattening is shown

in figure 10.3. Note that during unbowing the capacitance settled to a

steady value with two distinct "rates". The initial rapid decrease in

capacitance was associated with the decrease in bilayer area as a direct

result of a change in the curvature of the membrane. After the initial

rapid decrease a slower decrease in capacitance persisted for 5-10

minutes which was associated with a change in the shape of the torus.

a

E

Figure 10.2. A drawing of the typical changes in the geometry of the bilayer and torus during the course of a surface free energy measurement. The letters used in this key coincide with the time course in membrane capacitance during bowing shown in figure 10.3.

(A) The membrane had just become bimolecular and the bilayer was in equilibrium'with the torus.

(B) the membrane was bowed. The initial area increase was accommodated by an increase in the area of the torus.

(E) The bilayer was now in equilibrium with the torus and had moved to a new position with respect to the septum.

(F) Upon flattening the torus began to move slowly back to its original position. The area of the flat film "Ao" was approximately equal to the projected area of the film on the septum "Ap".

ll

13

12 lJ... C

Lu u ~

~ u r}_ <t: u 11

E

/

i D

~-F y- 1' A

l C

--------

..l ~--- ___.I_ ___ --- - -----'-- _____ _j__ _______ ___J

0 5 10 15 20 25

TIME MINUTES

Figure 10.3. The time course in membrane capacitance during bowing.

{A) The bilayer had just formed from the film. The capacitance was increasing because of the hydrostatic pressure arising from evaporation losses from the water compartments.

(B), (C) & (D) indicate times at which step changes in hydrostatic pressure (.5 pascal) were produced by withdrawing small volumes of aqueous solution from the outer compartment.

{E) The hydrostatic pressure difference was removed from the membrane (in this case the membrane had not come to equilibrium with the torus before flattening). The long term decrease in membrane capacitance was due the torus returning back to its original position {see figure 10.2).

Over the time scale presented here the hydrostatic due to evaporative losses was l pascal.

191

Two variations of the method described previously were tested.

When measurements were made by repeatedly flattening bowed membranes

over short times the measurements of inverse radius of curvature were

linearly proportional to the hydrostatic pressure. Consequently the

bifacial tension of the bilayer, calculated from the gradient of this

plot, was independent of membrane area which was similar to the findings

of Coster and Simons (1968). When the membrane was unbowed in

successive steps the calculated bilayer tension varied significantly

with membrane curvature. The comparison of results obtained by these

two variations is shown in figure 10.4. The significance of the

discrepancy in the values obtained by these two procedures will be

discussed later.

The temperature dependence of bilayer tension in bilayers formed

from egg-lecithin and cholesterol is shown in figure 10.5. The bilayer

tension in egg-lecithin membranes increased with decreasing temperature

whereas the tension in bilayers formed from GMO decreased with

decreasing temperature ( see Table 10.2). The values shown in Table

10.2 are in agreement with tension values obtained by Wolfe (1979) and

Andrews et al. (1970) for identical bilayer systems. Measurements of

bilayer tension on each membrane were repeatable to ±10%. The variation

in bilayer tension between different membranes was ±25%.

Inspection of figure 10.5 shows that the effect of using different

chainlength alkanes in the merabrane forming solution on the lateral

tension of egg-lecithin bilayers was not very significant; if anything

there was a slight reduction in bilayer tension for shorter chainlength

alkanes.

2

0

0+ 0

1

0 CJ

Q

LU 0 ~ ct

::)

~ Lu

[o 1 2 3 5 6 7

(1 / R)x10-2 m- 1

Figure 10.4. This shows the hydrostatic pressure difference plotted against the inverse radius of curvature of a single bilayer using two differing protocols (see text). (@) The membrane was flattened rapidly after being bowed different amounts. {-f-) The membrane was f1 attened slowly thus allowing time for measurements of membrane capacitance to be made at several stages during unbowing.

2

LEC CHOL C16 2 1 cu. •

• 15

E:

·········;1t1ii-"- 7 • I < E:

< Q ~ Lu I--

~ >--":t: -.J --(l:)

• I• • • I I II I •

5 • • I I

0 15 20 25 JO 35 1.0 1.5 50

TEMPERATURE oc

Figure 10.5. The bilayer tension of membranes formed from solutions of lecithin and oxidised cholesterol in n-alkanes of two different chainlengths. The error bars refer to the experimental scatter of the results obtained from repeated measurements on each membrane. The scatter of the values obtained from bilayers formed from both alkane solutions were similar.

TABLE 10. 2

Composition 2Y mN/m 20-25°( # 2Y mN/m 35-40°( # --------------------------------------------------------Lec:Chol 2 l

GMO

1.2

3.9

(KCl) Mol/m 3

• l

4.5

±.2 ( 4)

±.25 ( 5)

TABLE 10.3

GMO

2Y mN/m

3.9 ±.2

3. 7 ±. 2

.75

5.4

#

( 5)

( 2)

±.25 ( 12)

±.5 ( 3)

TABLE 10.4

Composition 2Y mN/m 35-45°( # Lee: Chol

0

2

0

.65 +.25

.75 +.25

( 5)

( 12)

1.0 +.25 (3)

2.5*-3.4**

* From Coster and Simons (1968) at 20°C ** From Grabowski and Cowen (1977) at l8°C

Table 10.2. The bilayer tension for GMO and egg-lecithin bilayers containing oxidised cholesterol over two different temperature ranges. The errors represent the standard deviations of the mean value for the number of different bilayers shown in each case (#).

Table 10.3. The bilayer tension of GMO bilayers formed in aqueous solutions containing different ion concentrations.

Table 10.4. The bilayer tension of egg-lecithin bilayers formed from solutions containing different mole ratios of oxidised cholesterol.

192

The bilayer tension of GMO and egg-lecithin bilayers was found to

be essentially unaffected by a 50 fold increase in the external

electrolyte concentration {see Table 10.3). Measurements of the bilayer

tension of egg-lecithin bilayers at high electrolyte concentrations were

exceedingly difficult as the formation of the bilayer was much slower

and the membrane was more fragile at these concentrations.

The inclusion of cholesterol in varying mole fractions in

egg-lecithin bilayer was found to increase the bilayer tension {see

Table 10.4). However, the temperature dependence of the bilayer tension

remained unaffected.

10.5 DISCUSSION

10.51 Rationale for the Method

Equilibrium between the bilayer and torus upon bowing the membrane

was difficult to detect as the approach to equilibrium was slow. This

was further complicated by the effect of differing rates of evaporation

from both electrolyte compartments; introducing unaccountable drifts in

the hydrostatic pressure over the equilibrating period of the membrane.

However, the approach to equilibrium during flattening was much more

rapid. Therefore it was desirable to conduct the tension measurements

whilst unbowing the membrane.

It was also found that the Plateau-Gibbs border was not rigidly

anchored to the septum and that the torus of lipid solution would move

away from the edge of the hole during bowing. Thus the area of the

bilayer, projected normal to the septum, varied with bowing and was

193

always greater than the area of the aperture. Examination of figure

10.3 shows that the movement of the torus over the 15 second time

interval immediately after flattening caused an underestimate of "A II

0

but not by more than 2%. However, over longer periods "A" could have 0

been underestimated by up to 10%. The effects of significant

underestimates of "A II in calculating the reciprocal radius of curvature 0

can be seen in figure 10.4 for the case where the bilayer was unbowed at

different rates. Therefore it was necessary to accurately determine the

area of the projection of the bowed film on the septum. Hence, after

measuring the capacitance of the membrane at equilibrium in the bowed

state, the membrane was flattened and the capacitance immediately

measured, not allowing sufficient time for the torus to move. From a

knowledge of the bilayer specific capacitance and the total capacitance

of the flattened bilayer the area could then be calculated.

The fact that the bilayer wasn't in mechanical equilibrium

immediately after unbowing has no bearing on the validity of this

approach as no intrinsic property of the bilayer was being investigated

after the membrane had been flattened. The capacitance of the bilayer

merely served as a measure of the effective projected area of the

membrane whilst it was bowed.

1-0.52 Energy of Formation: Temperature Dependence

In order to interpret the temperature dependence of bilayer tension

reported here we will calculate the partial derivative of "ya'' in

equation 10.2. The temperature-variation of the micelle concentration

"XN" is not likely to be important as the bilayer tension is logarithmic

in micelle concentration. Therefore terms arising from the temperature

194

variation of 11XN II can be neglected. The temperature dependence of the

bilayer surface energy is then given by:

where

0 aya _ a6µ + k O _ kT aN (D + l) ar-ar N war

X D=-ln;

10. 11

In these experiments 11 ~ 11 was in the range (2-4). 10- 3• Therefore,

as "N" is no less than unity, it is unlikely that 11 0 11 lies outside the

range 7-20. One expects that the variation of aggregation number with

increasing temperature to be negative both from the higher entropy of

smaller aggregates and from the relaxation of geometric packing

constraints (due to increased fluidity) at higher temperature. Hence:

aN > O 10.12 aT

As "D" lies in the range 7-20 then 1.):::0+1. Equation 10.11 can be

expressed as:

aya a6µ 0 kT + kT(D + l) (- aN) ar- =at+ N N2 aT 10. 13

From this equation it can be seen that the aggregation number, N,

of the lipids in the torus will effect the temperature dependence of the

bilayer tension. The smaller the aggregation number of the lipids in

the torus the more positive the temperature dependence of bilayer

tension. I

Differences of temperature dependence of bilayer tensions in

egg-lecithin and GMO bilayers have been assumed to be a result of

differences in the aggregation of the lipids in the torus. The

temperature dependence of bilayer tension in these two bilayer systems

will now be considered separately.

i) Egg-Lecithin

From the data in figure 10.3 the value of ayo aT

approximately -10-i kT/°C per molecule.

0

Thus a~T <O. 0

Then I ai\:r I > ~ + k~~ ( - ~)

195

was calculated as

10. 13

Thus it can be seen that the temperature dependence of the

standard chemical potential difference between lipids in the bilayers

and torus of egg-lecithin membranes is the dominant factor in the

temperature dependence of bilayer tension; the entropy terms being much 0

smaller. This implies that either .,a~µT " or "N" is large. It is not

certain which of these two alternatives is the case. Hm,ever, the

latter possibility is more likely as the internal conformation of the

lipid molecules in the bilayer is not expected to vary significantly with

temperature nor is it expected to be very different to that found in lipid

aggregates (inverted micelles) in the torus.

It can be concluded that the standard chemical potential difference

between lipids in the torus and bilayer decreases with increasing

temperature.

196

ii) GMO

The temperature variation of free energy per molecule in GMO

bilayers was found to be 3xl03 kT/°C (see Table 10.2). Hence:-

0

If atiµ <O al then l O. 14

Thus the positive temperature dependence of bilayer tension

indicates that the entropy terms play a much greater role in the

partitioning of lipid into the bilayer in this system than in

egg-lecithin bilayers. Hence the aggregation number of GMO molecules in

the torus must be much smaller. GMO is relatively soluble in n-alkane

solvent therefore one would expect the aggregation number of GMO in the

torus to be smaller than that of egg-lecithin. Further, from inspection

of equation 10.2 one would also expect with a smaller aggregation number

that the bilayers tension would be much higher than in egg-lecithin

bilayers which have much larger aggregation numbers.

Table 10.l shows that this was indeed the case.

Inspection of

From the results presented here an order of magnitude estimate of

the GMO aggregation number in the torus can be made by assuming that the

entropy terms in equations 10. l and 10.2 can be ignored for egg-lecithin

bilayers and that tiµ 0 for GMO and egg-1 eci thin membranes are equal at

all temperatures.

197

Firstly considering the relative surface energy of both bilayer

systems:

For egg-lecithin bilayers 0 = 6µ

kTD + N

If N is very large then the entropy of mixing can be neglected, hence

Hence for GMO bilayers:

Provided

0 = 6µ

kTD +N

6µ 0 = 60xl0 3 kT (as for egg-lecithin) Then

D N = -:iyg-"' 100

Having an estimate of the aggregation number in the torus GMO

membranes it is possible to get an estimate of the temperature

dependence of the aggregation number.

Now considering the temperature dependence of bilayer tension:

For egg-lecithin bilayers aya aT

If N for egg-lecithin is large then

Then for GMO bilayers:

If

a6 µ - l0- 3 kT/°C aT -

(as for lecithin)

198

Then N 1 N + lO aN = 0 ~ 1U aT 10. 15

And aN o when N ~ 100 then aT = -3.7( C)- 1

According to the above analysis the temperature dependence of N is

the dominant factor contributing to the temperature dependence of

bilayer tension in GMO bilayers. It should be noted that ~ estimated

here would only be for a limited temperature range. 0 bilayers could not form at temperatures over 60 C.

Otherwise GMO

10.53 Bilayer Tension: The Effect of Electrolyte Concentration and

Cholesterol

Bilayers formed from cholesterol have much higher tensions than do

those formed from egg-lecithin. However, the increase in bilayer

tension upon the addition of cholesterol was not as much as that

expected from a simple linearly proportional relationship between the

bilayer tension and the cholesterol mole fraction in the torus.

The exact interpretation of this effect is not known as the mixing

of cholesterol and egg-lecithin in the torus is unknown. A tentative

explanation is that cholesterol forms mixtures with relatively low

aggregation number because of its "concave"

(Israelachvili et al., 1980).

packing geometry

I

Widely varying ion concentrations in the external electrolyte had

very little effect on the bilayer tension of GMO bilayers. This was not

sur~rising as it was reported earlier in this thesis that the

partitioning of ions into the polar head region of GMO bilayers is

sma 11 ( about 1 o-e) •

199

10.6 SUMMARY

The bilayer tension of membranes formed from GMO and egg-lecithin

was measured using a variation of the technique of Coster and Simons

(1968). The method was adapted to overcome difficulties peculiar to the

present lipid bilayer membranes.

The stress-strain relation in lipid bilayers was that expected from

the surface energy law. Elastic contributions to the free energy of the

bilayer were negligible.

The temperature variation of bilayer tension in GMO and

egg-lecithin bilayers was distinctly different. GMO bilayers maintained

large tensions which increased with increasing temperature.

Egg-lecithin bilayers, on the other hand, had much smaller tensions

which decreased with increasing temperature. The difference in the

tension properties of these two systems was consistent with

egg-lecithin bilayers being in equilibrium with very large aggregates

and GMO bilayers being in equilibrium with relatively small -aggregates,

containing about 100 molecules. The increase in GMO bilayer tension

with increasing temperature was mainly due to a decrease in the

aggregation number of the molecules with increasing temperature.

From the data presented in this chapter it seems that entropy plays

a minor role in determining the energy of formation of egg-lecithin

bilayers. However, in GMO membranes entropy plays the dominant role in

bilayer-torus equilibrium.

200

The addition of cholesterol into egg-lecithin bilayers produced

bilayers which had higher tensions. The mechanism for this is not yet

known. However, it is possible that egg-lecithin - cholesterol

mixtures have smaller aggregation numbers than pure egg-lecithin.

A 50 fold increase in the ion concentration in the

had no measurable effect on GMO bilayers. This lack

aqueous phase

of effect is

consistent with a high energy barrier to ions in the polar regions of

GMO discussed in previous chapters.

CHAPTER 11

THE EFFECT OF SOME LOCAL ANAESTHETICS ON THE

PHYSICAL PROPERTIES OF EGG-LECITHIN BLM

11.l INTRODUCTION

11.2 MATERIALS AND METHODS

11.3 RESULTS

11.4

11. 31 Procaine

11.32 Other p-Aminobenzoic Acid Esters

11 . 33 Alcohols

DISCUSSION

11 . 41 p-Aminobenzoic Acid Esters

11 .411 Surface Charge

11.412 Dielectric Structure

11.413 Bilayer Conductance

Page

202

204

205

205

206

207

209

209

209

210

213

11 . 42 Alcohols: Their Effect on Dielectric Structure 214

11 . 43 Bilayer Tension: Effect of Local Anaesthetics 214

11 . 44 Comparison With Previous Work 215

11. 45 Possible Mechanisms For ~ocal Anaesthesia 216

11.5 SUMMARY 219

201

202

11. l INTRODUCTION

From the viewpoint of their practical application in medicine,

local anaesthetics are clinically important drugs.

their ability to inhibit the propagation of nerve

This is a result of

action potentials

without causing cell death.

is not well understood.

Their mode of action, at a molecular level,

An understanding of the mechanism of

anaesthesia by local anaesthetics will yield new insights, not only into

the action of anaesthetics per se, but also into the more general

problem of the function of biological membranes.

The inhibition of the action potential in squid axon by local

anaesthetics is mainly due to a reduction in the conductivity of the

sodium channel; the conductance of the potassium channel being much

less sensitive to the presence of local anaesthetics (see review by

Ritchie and Greengard, 1966).

Given the functional importance of the bilayer component of

biological membranes (see review in Chapter l) it was of interest in

this study to investigate the effects of local anaesthetics on the

structure of artificial lipid bilayers.

Some models for the action of local anaesthetics associate the

action of anaesthetic with specific proteins, acting at specific sites

on the protein (Hille, 1980). However, the wide variety of compounds

which act as anaesthetics, as well as their often additive effects

(Staiman and Seeman, 1975) on a wide variety of membranes and membrane

functions, suggests that the site of action of anaesthetics is a

203

non specific site with polar and non-polar properties (Franks and

Lieb, 1978).

Anaesthetics are known to increase membrane area and fluidity

(Seeman and Roth, 1972). It has been postulated that anaesthesia is a

direct consequence of this (Seeman, 1972). However, X-ray (Franks and

Lieb, 1978), and NMR experiments (Boggs, Yoong and Hsia, 1976, Boggs,

Roth, Yoong, Wong and Hsia, 1976 and Turner and Oldfield, 1979) found

that, at the relatively low concentrations required to induce

anaesthesia in living cells, anaesthetics had no detectable effects on

lipid organization in multilayer and vesicle preparations.

suggested that anaesthetics specifically act at sites on

It was then

the membrane

bound proteins (though this is difficult to reconcile with the notion of

non-specific sites referred to earlier) or at the surrounding annular

lipid region (Lee, 1976, Richards, 1976 and Franks and Lieb, 1978).

Ashcroft, Coster and Smith (1977) and Haydon et al. (1977) proposed

that the anaesthetic action of hydrophobic molecules, such as the

n-alkanes and amphiphilic molecules, such as benzyl alcohol, was due to

their ability to swell lipid bilayers, both in artificial systems as

well as in biological membranes. Ashcroft (1979) investigated this

possibility for amphiphilic molecules such as procaine and benzocaine by

measuring their effect on the dielectric structure of artifical

egg-lecithin bilayers generated from n-tetradecane lipid solutions. On

the basis of those studies it was postulated that anaesthetics generally

act by inducing changes in the thickness of the bilayer component of

biological membranes.

In this thesis artificial BLM have been re-characterised with

improved apparatus in the light of recent contributions to the

204

understanding of the physical properties of lipid bilayers. Namely,

alkane absorption (eg. White, 1977 and Gruen, 1980a and 1980b),

capacitance of ionic double layers (Smith, 1977) and bilayer tension

(eg. Wolfe, 1979). Given this re-characterisation, the effects of some

alcohols and p-aminobenzoic acid ester ty~es of local anaesthetic were

re-examined in an attempt to test the validity of conclusions of

Ashcroft (1979) in solventless bilayer systems. The results presented

in this chapter have been interpreted in terms of the physical and

electrical properties of lipid bilayers described in this thesis.

11.2 MATERIALS AND METHODS

Artificial bimolecular lipid membranes were formed from n-alkane

solutions containing egg-lecithin and egg-lecithin - cholesterol

(2:1 mole ratio).

The effects of some of the p-aminobenzoic acid ester types of local

anaesthetics on lipid bilayers were investigated by adding appropriate

amounts of these chemicals to the aqueous phase. Both procaine and

tetracaine were added to the membrane system in the form of a

hydrochloride salt. Alcohols were added to the membrane via either the

membrane forming solutions or the aqueous phase.

Control experiments were conduct~d at the beginning of each series

of measurements with a new anaesthetic. The physical and electrical

properties of the control membranes were compared to those obtained from

membranes known to be uncontaminated. This served as a check to ensure

that traces of anaesthetics employed in earlier experiments (adsorbed on

205

the walls of the membrane chambers, for instance) did not contaminate

bilayers in subsequent experiments.

When the effects of slightly volatile drugs on egg-lecithin

bilayers were being investigated, the impedance measurements performed

were made within 2 hours of exposing the drug to the atmosphere.

Neasurements of bilayer surface free energy were made using the

same procedure as that employed in similar measurements described in the

previous chapter.

The molecular models of the local anaesthetics used in this study

are shown in figure 11.1.

11.3 RESULTS

11.31 Procaine

The presence of procaine in the aqueous phase at concentrations in

the range 1-5 mM had no detectable effect on the capacitance (measured

at lHz) of solvent free egg-lecithin and egg-lecithin - cholesterol

bilayer (see Table 11.1).

The frequency dispersion in bilayer capacitance and conductance was

markedly effected by procaine at mM concentrations. At frequencies

above lHz the dispersion due to the polar regions of egg-lecithin

bilayers, with and without cholesterol, virtually disappeared when

procaine was present (see figure 11.2).

BENZOCAJNE

N2 N C 0~ II 0 CH3

PROCAINE

~CH3

H2 N C 0 N

II ~H3 0

TETRA CAINE

HN --c---o II 0

CH3

Figure 11. l. Molecular models for some of the local anaesthetics used in this study. The1dark circles represent the CH 2 groups of the molecules.

Lu l..)

~

~ G rf_ <::(

8

7

• 2 mM PROCAINE

o SAT. BENZOCAINE

• • X BARE

l..)6L_ ______ 1-_____ .J._ ______ _J_ ____ :-=-~.i_--~ 10- 70- 10- 1 10 100

FREQUENCY Hz

Figure 11.2. The capacitance spectra (at 20°C) of egg-lecithin bilayers formed from n-hexadecane solutions in lmM KCl in the presence of procaine and benzocaine. These are compared to the capacitance spectrum of a typical bare egg-lecithin bilayer. The solid curves represent the theoretical fits to the data.

TABLE 11. l

THE EFFECT OF PROCAINE HCl ON THE CAPACITANCE OF EGG-LECITHIN BLM

(MEASURED AT lHZ)

Chloride concentration

.2 mM

1.0 mM

KCl Procaine HCl

Capacitance mF/m 2

5.6 ±.15

6.1 ±.1

5.8±.15

6.15± .15

Table 11.1. The effect of different concentrations of procaine HCl in the aqueous phase on the membrane capacitance (measured at l Hz) at 25°C.

206

One membrane remained stable over sufficient time to allow

measurements of merabrane impedance at frequencies in the range

.003-.03 Hz. In this frequency range the dispersion was considerably

increased by the presence of procaine. The dielectric parameters

extracted from the impedance data obtained from raembranes in aqueous

solutions of 1-3 mM procaine are shown in Tables 11. 2 and 11. 3.

Procaine increased the capacitance of the choline phosphate region by

300-500% whilst reducing the capacitance associated with the acetyl

region.

Aqueous solutions containing Procaine (3 mM) reduced the surface

free energy of egg-lecithin bilayers containing cholesterol by 40% over

the temperature range 20-40°C (see figure 11.3).

11.32 Other p-Aminobenzoic Acid Esters

i) Benzocaine

The addition of benzocaine (at lmM concentrations) had very little

effect on the polar regions of egg-lecithin bilayers at frequencies over

.lHz. At very low frequencies (.003 -.03 Hz) the capacitance dispersion

was significantly increased in a manner similar to that found for

procaine (see figure 11.2, also see Tables 11.2 and 11.3). Only one

membrane remained stable over a long enough period to allow impedance

measurements at these low frequencies.

Benzocaine had only a marginal effect on the bilayer surface free

energy over the temperature range 20-30°C (see figure 11.3).

2.0 LEC CHOL BARE 2 7 • 1mM BENZOCAINE

~ 0 1mM PROCAINE

~,5 E

~ a --V) ~ • 11 I 1·11 ~10

1 .11 •

ct Lu )...

~ -J

---Q)

• • 01 • 0 • I I 0

0 0 0 • 0 • I I

0 I ·5 Oo I 0 0 I 0 0

0

Oo

0 15 20 25 30 35 l0 l5 50

TEMPERATURE oc

Figure ll .3. The bilayer tension of egg-lecithin membranes (lmM KCl) formed from n-hexadecane solutions at different temperatures in the presence of procaine and benzocaine. This is compared to that obtained for bare membranes.

TABLE 11 . 2

BARE PR0C (2mM) BENZ ( lmM) BUT ( l 00mM) BZA ( l 00mM) -------------------------------------------------------------------

Capacitance mF/m 2

6.6 ±.l 6.7 ±. l 6.7±.l 6. 35 ±. 2 6.0 ± .2

400 ±50 500 ±100 450 ± 20 375 ± 25 550 ± 50

600 ± 150 1000 ±200 550 ±40 700 ±150 500 ±100

1000 ±200 2200 ± 800 900 ± 40 l l 00 ± 300 900 ± 100

900 ±150 5000 ±2000 800 ±40 1400 ±300 900 ±200

Conductance mS/m 2

. 5- l. 5 . 4- l .0 .2-l.0 . 3- l. 5 .6-2

40-100 100 100 50 100-200

( 4-10) xl 02 (4-7)x10 2 (8-lO)xl0 2 300 (7-8)x10 2

(7-10)xl0 3 (1-2)x10" (1-2)xl0 3 (3-7lx10 3 (9-13)xl0 3

(5-lO)xlO" (5-20)x10 5 (7-8)x10" (2-3)x10" (5-8)x10"

Table 11.2. The effect of procaine (PR0C),benzocaine (BENZ~ butanol (BUT),and benzyl alcohol (BZA} on the dielectric parameters of egg-lecithin cholesterol membranes at 20-30°C in lmM KCl. The data in this table has also been presented in a graphical format in figures l l . 6b, and 11 . 9.

BARE

6.35 ±.1

650 ± 100

1200 ± 200

1200 ±200

1100 ±200

. 5-1. 5

40-200

(3-lO)xl0 2

(6-10)xl0 3

(5-9)xl0"

TABLE 11. 3

PR0C (lmM) BENZ (lmM)

Capacitance mF /m 2

6.7±.l 6.4±.l

160 250 ±50

1000 ±500 500 ± 50

10000 ± l 000 1700 ± 200

1600 ± 400

Conductance mS/m 2

. 3-1 . 0

20

(8-l3)xl0 2

(3-4)xl0 5

. 4-1.0

40

140-200

(4-9)xl0'

(7-lO)xlO"

PENTAN0L (sat.)

6.4 ±. l

600 ± 100

700 ±200

1500 ± 500

.l-.5

2000

(4-6)xl0"

Table 11.3. The effect of procaine (PR0C~ benzocaine (BENZ~ and pentanol on the dielectric parameters of pure egg-lecithin bilayers. The data in this table has also been presented in a graphical format in figures 11.6a, 11. 7, 11.8, and 11.10.

207

ii) Tetracaine

Attempts to measure the frequency dispersion in membrane

capacitance in the presence of tetracaine were frustrated by the

presence of large membrane conductances which increased rapidly as the

bilayer aged. With lmM tetracaine and lr.iM KCl in the aqueous phase the

r.iembrane conductance increased from l0mS/m 2 immediately after formation

of the bilayer to 200 mS/m 2 within 15 minutes of formation. This effect

was also present at r.iuch lower concentrations of the anaesthetic. When

the concentration of tetracaine in the aqueous phase was less than

.01 mM the conductance was stable enough to allow impedance measurements

at frequencies as low as .1 Hz. However, even at these low tetracaine

concentrations the bilayer conductance still increased by 5 mS/m 2 over

the course of an experimental run (approximately 20 minutes).

At .01 mM tetracaine in the aqueous phase there was no

change observed in the dielectric structure parameters

hydrophobic - hydrophilic interface of lipid bilayers.

11 . 33 Alcohols

i) n-Alkanols

detectable

of the

Ethanol at concentrations up to 100 mM in the aqueous phase and

150 mM in the r.ier.ibrane forming solution had no detectable effect on

either the ir.ipedance dispersion, the capacitance measured at lHz or the

surface free energy of egg-lecithin bilayers (see figure 11.4).

X ETHANOL 1.0 0 BUTANOL

t: • PENTANOL

~ t: ~ C) -l/)

~ Lu f--

·S

Q:: Lu 0 >- 0 "'{ -J -(l)

0 0 so 100

ALK ANOL CONCENTRATION mM/m 3

Figure 11 .4. The effect of different chainlength n-alkanols on the tension of egg-lecithin bilayers at 40°C in l mM KCl.

208

Longer chainlength alkanols such as butanol and pentanol at

concentrations in the range 10-50 mM in the aqueous phase significantly

reduced the surface free energy of egg-lecithin bilayers (see

figure 11.5).

The effect of these alkanols on the dielectric structure of the

bilayers was not detectable from these measurements (see Tables 11.2 and

11.3).

ii) Benzyl Alcohol

The effect of benzyl alcohol on the impedance dispersion of

egg-lecithin and egg-lecithin - cholesterol bilayers at frequencies over

.1 Hz was undetectable at BZA concentrations up to l00mM.

However the effect on the alkane absorption of lipid bilayers was

quite profound (see figures 5.4a and 5.4b).

70

6.8

16.6 ~ Lt. t:

UJ ~ 6.4 ~ -<..J rf_

• BARE o 100 mM BUTAN0L

X 10mM B Z A

~ . <..J 62L_.,.---___ _,1__~ ___ _.J_ ____ ---Jl---------:-10- 10-1 1 10 100

FREQUENCY Hz

Figure 11.5. The capacitance spectrum of egg-lecithin bilayers (lmM KCl) formed from n-hexadecane solutions in the presence of butanol and benzyl alcohol. The is compared to the capacitance spectrum obtained from bare egg-lecithin bilayers at 20°C.

209

ll.4 DISCUSSION

11.41 p-Aminobenzoic Acid Esters

11.411 Surface Charge

Procaine can exist in either a charged or an uncharged form

depending on the pH of the aqueous phase. The pKa of the amino group of

procaine is 9.0. Therefore in these experiments, where the pH of the

aqueous phase was approximately 6,

charged form.

procaine existed mainly in its

A study of the binding of charged drugs to lipid bilayers by

Lee, (1978) showed that the apparent pKa of charged diethylammonium

moieties of bound procaine is shifted from 9.0 in the aqueous phase to

8.0 at the membrane surface. This is a result of the fact that

additional work is needed to bring the charged form of procaine near to

the bilayer against the Born repulsive forces near the hydrophobic

region of the bilayer.

In Chapter 6 it was shown that at low external ion concentrations

the membrane capacitance was sensitive to changes in the ionic double

layer capacitance, which was in turn dependent on the bilayer bound

surface charge and electrolyte concentration. Therefore if the charged

form of procaine binds significantly ~o egg-lecithin bilayers one ~ould

expect to measure an increase in the total membrane capacitance. The

results presented here show that the addition of procaine to aqueous

solutions containing .l and l mM KCl had a negligible effect on the

total membrane capacitance. Thus it appears that the membrane surface

210

charge was unaffected by the absorption of procaine, and that the

charged form of procaine did not bind to egg-lecithin bilayers at a

pH of 6 in the aqueous phase. This indicates that the effects procaine

had on the properties of lipid bilayers in this study were probably due

to the presence of the neutral form of the anaesthetic.

11.412 Dielectric Structure

The addition of procaine {at l mM concentrations in the aqueous

phase) had a profound effect on the bilayer impedance dispersion of

egg-lecithin and egg-lecithin cholesterol bilayers. The dispersion in

bilayer capacitance over the frequency range l to 100 Hz almost vanished

and the low frequency dispersion was significantly increased. The

dielectric time constant diagrams, showing the effects of procaine and

benzocaine on the spatial variation of time-constant within egg-lecithin

and egg-lecithin cholesterol bilayers, are shown in figures

11 . 6a, 11 . 6b and 11 . 7

There are a number of different interpretations as to how the

dielectric structure of the hydrophobic-hydrophilic interface of

egg-lecithin bilayers is altered by the absorption of neutral procaine

molecules. These are set out below:

l) Procaine may significantly increase the dielectric capacitance of

the choline phosphate region. This can be discounted however, as the

thickness of the choline phosphate region would need to be less than

.1 nm to account for the high capacitance values. The relatively high

dielectric constant of the polar regions {see Chapter 7) suggests that

the possibility of an additional increase in the dielectric constant

upon the inclusion of procaine is unlikely.

2

0

- 7

-2

-J

-~!~~~it

~tlll '.Jj~;kiYit?J.W\~~cw+~~!~~;

i:;;M:

-l ~ ---0;1;--' ---~, ----;.2-:;----_-!:3----;_l~1 ------!:_s=-----.s!c---.-='-f---J.s'----RELATIVE DISTANCE nm

Figure 11 .6a. The time-constant profile for the hydrophobic-hydrophilic interface of egg-lecithin bilayers. The data represents the variation in the results obtained from 3 membranes in the presence of procaine at lmM concentrations (unshaded). This is compared to that obtained for 19 bare egg - lecithin bilayers (shaded)

2

1

0

-1

-2 C: --

-3

0 ·1 -2

RELATIVE

3 ·l ·5

DISTANCE nm

Figure 11 .6b. The effect of procaine on the time-constant profiles of egg-lecithin - cholesterol membranes (2:l mole ratio) formed from n-hexadecane solutions in lmM KCl. The shaded area represents the scatter in the results obtained from 23 bare membranes. The open are~ represents the scatter in the results obtained from 6 membranes in the presence of procaine at 2 mM concentrations.

2

1

0

-1

-c:: -

-]

-l 0 ·1 ·2 ·] ·l ·5

RELATIVE DISTANCE nm

Figure ll .7. The effect of benzocaine on the time-constant profile of the hydrophobic-hydrophilic interface of egg-lecithin bilayers. The shaded area represents the data obtained from 19 bare membranes. The open area represents the data from 3 membranes in the presence of benzocaine at lmM concentrations in the aqueous phase.

·6

211

2) Alternatively the conductivity of the choline phosphate region could

have been drastically decreased by the presence of procaine so that it

had a similar time-constant to that of the acetyl region. The large low

frequency dispersion could then have been due to the choline phosphate

region. Then the regions with dielectric time-constants in the range

.01 to 1 second (refer to figures 11.6a&b) represent the transition

region between the membrane polar groups and the aqueous phase. However

this interpretation does not account for the effects of benzocaine on

the dielectric structure of the lipid bilayers. The presence of

benzocaine also produced a similar effects to that of procaine on the

capacitance of regions with electrical time-constants in the range

10-100 seconds but had no effect on regions with higher time constant

(see figure 11.7).

3) Another possibility is that the conductivity of the choline

phosphate group was increased by the presence of the procaine molecules

to such an extent that the capacitance dispersion associated with this

polar region could not be detected from impedance measurements in the

frequency range currently employed in these experiments. Thus the

dielectric structure in the time-constant range .01-1 second (refer to

figures 11.6a&b) represents the interface between the choline phosphate

groups and the acetyl region. The large dispersion observed at low

frequencies could then possibly be due to the presence of a relatively

polar part of the procaine molecules embedded in the hydrophobic region;

thus creating slightly polar regions within the acyl chain region.

X-ray diffraction measurements by Coster, James, Berthet and

Hiller (1981) found the uncharged aromatic amino group of procaine to be

located 5 to 7 angstroms into the acyl chain region fro~ the acetyl

region in multilayer preparation of lecithin-cholesterol bilayers.

212

The large effect of procaine on regions in the bilayer with

time-constants in the range 10-100 seconds, observed in the present

dielectric experiments, could then be due to the presence of the

aromatic ring and the carbonyl groups of the anaesthetic molecules in

the acyl chain region. The presence of the aromatic amine and the

acetyl groups of the anaesthetic molecules could significantly increase

the dielectric constant of the outer part of the acyl chain region

endowing it with dielectric properties similar to those of the acetyl

region of egg-lecithin. This would effectively increase the thickness

of the acetyl region.

experiments.

Such was found to be the case in the present

NNR studies employing deuterated anaesthetic (Boulanger, Schreier,

Leitch and Smith, 1980) postulated data consistent with the notion that

some of the procaine molecules were intercalated in the acyl chains of

the lipids and other were weakly bound to the polar head groups of the

lipids. The increase in polar head conductivity observed in the current

study may then have been due to the binding of procaine to the polar

head groups.

Alternatively the large increase in the conductance of this region

may have been due to the presence of the diethylammonium groups of

procaine. Benzocaine does not have a diethylammonium group as part of

its molecular structure and does not affect the dielectric structure of

the choline phosphate region.. The X-ray studies of Coster et

al. (1981) indicate that the diethylammonium moiety of the procaine

molecule penetrates as far as the choline phosphate groups of the lipid

molecules. The diethylammonium groups, which are more polar than the

benzene ring, would be located in the polar head region of the lipid

bilayer; the benzene ring being e~bedded in the acyl chains.

213

Benzocaine, however, has a methyl group in place of the

diethylammonium group. The aromatic amine group is the more polar end

of the benzocaine molecule. Therefore benzocaine should be oriented

with the aromatic amine moiety located in the acetyl region or possibily

in the choline phosphate region of the bilayer. The reciprocal

orientations of procaine and benzocaine in the bilayer may account for

the distictly different effects that these otherwise similar molecules

have on the dielectric structure of egg-lecithin bilayers. Figure 11 .8

illustrates the relative orientations of benzocaine and procaine in

egg-lecithin bilayers inferred from these data.

11.413 Bilayer Conductance

The effects of the p-aminobenzoic acid ester types of local

anaesthetics on the conduction properties of artificial BLM have been

well characterised (eg. see Ohki, 1970 and McLaughlin, 1975). The

conductance of egg-lecithin bilayers in the presence of procaine and

tetracaine reported in these studies was similar to that obtained here.

McLaughlin's (1975) results indicate that the increased conductivity of

lipid bilayers in the presence of tetracaine was a result of these

molecules acting as carriers for electrical charge. Further, the data

was consistent with the carrier being a complex containing one neutral

and one charged tetracaine molecule.

The relative non-effect of procaine on bilayer conductance reflects

the fact that procaine is smalle, and more polar than tetracaine and

therefore its Born energy in the hydrophobic interior of egg-lecithin

bilayers would be much higher than that of tetracaine.

a

\ u=a

a="

\ 0

�) :i::

c...,

Figure 11 .8. This figure shows the orientation of procaine (left) and benzocaine (right) in egg-lecithin bilayers as determined from the dielectric measurements in this study.

214

11.42 Alcohols: Their Effect on Dielectric Structure

Though butanol and pentanol had pronounced effects on the bilayer

tension they had little effect on the dielectric substructure.* The

alcohols employed in this study were much smaller, simpler molecules

than the p-aminobenzoic acid ester types of local anaesthetics. If the

hydrocarbon chain of the alcohols was located within the acyl chain

region and the hydroxyl group was located in the polar regions, then the

alcohol molecules would have been dielectrically "invisible" in the

bilayer structure. However, their effect on bilayer tension bears

witness to their presence at the bilayer-solution interface. *(See figs.

11 . 9 and 11 . 10. )

11.43 Bilayer Tension: Effect of Local Anaesthetics

X-ray diffraction studies (Franks and Lieb, 1978) and NMR studies

(Boggs, Yoong et al., 1976 and Turner and Oldfield, 1979) found no

detectable changes in the lipid ordering and bilayer structure upon the

addition of local anaesthetics. It was therefore assumed that the site

of action of the local anaesthetics was not associated with the bilayer

but rather with the proteins or the surrounding lipid annulus.

However in this study it was found that procaine, butanol and

pentanol significantly reduced the bilayer tension of egg-lecithin­

cholesterol bilayers at clinical concentrations of the anaesthetics in

the aqueous phase.

In Chapter 10 the surface tension properties of lipid bilayers was

characterised. From the temperature dependence of bilayer tension in

egg-lecithin - cholesterol membranes it was concluded that the

contribution of the entropy-of-mixing of inverted raicelles in the torus

2

1

0

- 1

C -2

- 3

- lLL----L----+----:!;---_74r---~-St-------;:__6 0 ·1 ·2 3

RELATIVE DISTANCE nm

Figure ll .9. The dielectric time-constant profile of egg-lecithin -cholesterol bilayers (2:l mole ratio) formed from n-hexadecane solutions in l mM KCl. The shaded area represents the data obtained from 23 bare membranes. The open area represents that obtained from 3 membranes formed in solutions tontaining 100 mM butanol.

2

1

0

-1

-(.:) " -2 ~ '--

c:: --]

-l

0 . 1

RELATIVE

·2

DISTANCE

·]

nm

Figure 11. 10. The dielectric time-constant profile formed from n-hexadecane solutions in 1 mM KCl. The represents the data obtained from 19 bare membranes. represents that obtained from 2 membranes formed in of pentano 1.

·l ·5

of egg-lec'ithin shaded area The open area

saturated solutions

215

to the membrane tension was relatively small, and that the main

contribution arises from the difference in standard che~ical potential

between lipids in the torus and lipids in the bilayer. Therefore the

anaesthetics that reduce bilayer tension do so by reducing the standard

chemical potential of the lipids in the bilayer relative to that in the

torus. This might be achieved by a reduction in area of the

hydrocarbon - water interface due to 11 masking 11 effects of anaesthetic

molecules at the membrane surface. However, the interpretation

presented here is by no means conclusive.

The reason for the different effects of benzocaine and procaine on

the bilayer surface tension may be a result of the different

orientations of both molecules at the bilayer aqueous interface.

The relative effects of the different chainlength alcohols on the

bilayer tension correlated well with their relative solubilities in the

bilayer as well as their anaesthetic potency (see Table 11.4).

11.44 Comparison With Previous Work

The location of anaesthetic molecules within the membrane structure

deduced from dielectric impedance measurements in the present study

confirmed the conclusions of Ashcroft, (1979) concerning the location of

procaine and benzocaine in egg-lecithin bilayers. However, in the

present study it was found that the n-alkanols, benzyl alcohol

(see Chapter 5), procaine and benzocaine all had a negligible affect on

the thickness of solventless bilayers. It seems as though the thickness

changes reported by Ashcroft (1979) upon the addition of anaesthetics

were due to changes in the mole fraction of n-tetradecane in the bilayer

which were probably associated with an altered lipid ordering in the

TABLE 11. 4

Procaine l mM 2-20 mM

Ethanol .l50mM 1-2 M

Butanol 50 mM 50-l00mM

Pentanol 10 mM 20-40mM

Table 11.4. The concentration of the different anaesthetics, examined in this thesis, required to cause a 50% reduction in bilayer tension (left). This is compared to that required to block sodium conduction in nerve axon (right). The values on the right were obtained from Seeman (1972).

216

bilayer. This has been conclusively demonstrated in Chapter 5 for the

case of benzyl alcohol. Changes in the acyl chain ordering in lipid

bilayers have been observed upon the addition of relatively high

concentrations of benzyl alcohol, procaine and tetracaine to vesicle

preparations, by the use of deuterated NMR techniques (Turner and

Oldfield, 1979, Boulanger et al., 1980 and Boulanger et al., 1981).

Much smaller changes in the acyl chain ordering at clinical anaesthetic

concentrations, too small to be resolved by NMR methods, might also be

responsible for the variations in the alkane solubility in egg-lecithin

bilayers (c.f. also Chapter 5).

11.45 Possible Mechanisms For Local Anaesthesia

It is believed that the excitation in nerve axon results from

protein pores spanning the membrane which contain some form of voltage

dependent ion gating mechanism (Ehrenstein, 1976, Lee, 1976 and

Ashcroft, Coster and Smith, 1977).

From thermodynamic considerations of the fluid mosaic model

and

of

discussed in Chapter l, the function of membrane

proteins should be very sensitive to the structure

their immediate lipid environment.

bound enzymes

and composition

Ashcroft, Coster and Smith (1977) and Haydon et al. (1977)

postulated that changes in the thickness of the bilayer component of

biological membranes by anaesthetic molecules (such as n-alkanes) were

consistant with their anaesthetic action.

It was then postulated that anaesthesia could be induced by a

relative change in the dimensions of the hydrophobic region of the lipid

217

bilayer with respect to the sodium channels. A mismatch between the

dimensions of the hydrophobic regions of the protein and lipid

components of the membrane could lead to a distortion in the boundary

lipid region of the protein (as seen in figure 11. 10). It was proposed

(see Ashcroft, Coster and Smith, 1977) that the strain in the sodium

channel produced as a result of this mismatch mediated blocking of the

sodium channel.

In this study there was no indication that anaesthetics would cause

a general increase in the thickness of the bilayer component of

biological membranes; unless, of course, they were in equilibrium with a

saturated solution of relatively small hydrophobic molecules.

However, the fact that some anaesthetics in the present study

significantly reduced bilayer tension raises an alternative possibility.

If the lipid bilayer is norr.,ally distorted ("cusped" or "dimpled"), in

the vicinity of the sodium channel, then a change in the surface tension

of the bilayer could significantly alter the stresses on the protein.

However, it should be noted that the mechanisms that give rise to the

resting tension i~ the axon plasma membrane are likely to be different

to those responsible for tension artificially produced in lipid bilayers

in this study. It is thus not yet ce1~tain whether local anaesthetics do

in fact alter the surface tension properties of the lipid bilayer

component of nerve axon plasma membranes.

At this point it should be mentioned that general structural

changes in the bilayer component of biological membranes need not be

responsible for blockage of the sodium channels in nerve axon. This has

already been pointed out by (Lee, 1976) and hinted at in the concluding

remarks of Franks and Lieb (1978).

218

The fact that local anaesthetics in the present study could reduce

bilayer tension as a result of altering the equilibrium between lipids

in the bilayer and torus phases of the ~embrane raises the possibility

that anaesthetics can also alter the equilibrium between the boundary

lipids of the sodium channel and the lipid bilayer. Such a change in

the equilibrium between the boundary and bilayer phases of

multicomponent systems (like living membranes) could lead to significant

changes in the composition of the boundary lipid phase, which in turn

could give rise to changes in the fluidity, geometry, surface charge and

surface potential of the membrane in the immediate vicinity of the

sodium channel. In fact on this basis neutral anaesthetic molecules

such as benzocaine, by altering the lipid composition of the boundary

lipid region, could change the membrane surface potential near sodium

channels in nerves.

Unfortunately this hypothesis is difficult to test and it is

definitely beyond the scope of the simple me~brane model employed in the

present study. However, studies on artificial BLM reconstituted with

excitation inducing material may give additional insights into the mode

of action of local anaesthetics.

219

11.5 SUMMARY

The effects of some n-alkanol and p-aminobenzoic acid ester types

of local anaesthetics on egg-lecithin and egg-lecithin - cholesterol

artificial BLM were investigated. The results were interpreted in terms

of the dielectric and surface tension properties of the lipid bilayers

characterised in this thesis.

The presence of amphiphilic local anaesthetics in egg-lecithin

bilayer systems had little effect on the capacitance (measured at 1 Hz)

of solventless bilayers.

The presence of the protonated form of procaine in the aqueous

phase did not significantly alter the capacitance of the ionic double

layers external to the membrane. This suggested that the charged form

of procaine did not partition significantly into the membrane.

The presence of tetracaine in the aqueous phase drastically

increased the bilayer conductance. This increase in conductance was

consistent with previous measurements made by McLaughlin (1975). In

that study it was found that the increased bilayer conductance was due

to a charged tetracaine complex acting as a carrier of electrical

current in the bilayer interior.

I Procaine and benzocaine both had significant effects on the

frequency dependence of the bilayer capacitance. At low frequencies

(.003 Hz to .1 Hz) both procaine and benzocaine increased the relative

dispersion in bilayer capacitance. At higher frequencies procaine

virtually abolished the frequency dependence of the bilayer capacitance

220

due to the polar head regions whereas benzocaine had no detectable

effect on the dispersion.

From the relative effects of these two related compounds on the

dielectric structure it was possible to infer the location and

orientation of the anaesthetic molecules within the bilayer structure.

The data was consistent with the aromatic amine group of procaine being

embedded in the acyl chain region of the bilayer with the

diethylammonium moieties penetrating the region containing choline

phosphate groups of the lipid. The location of procaine within the

bilayer as determined from these dielectric measurements

with that previously found by X-ray (Coster et al.,

studies (Boulanger et al., 1980).

is consistent

1981) and NMR

Indications were that procaine and benzocaine probably had opposite

orientations within the bilayer, ie. the benzene ring of procaine was

intercalated among the acyl chains of the lipid whereas that of

benzocaine was penetrating the acetyl region.

The addition of benzyl alcohol, butanol and pentanol to artificial

BLM had no effect on the capacitance of solventless bilayers over the

entire experimental frequency range. This lack of effect could not be

interpreted as being due to the exclusion of these molecules from the

bilayer phase as their presence in the bilayer was reflected by changes

in the surface tension and alkane solubility properties of the bilayer.

Thus these molecules were dielectrically ''invisible" at their

binding sites within the bilayer.

221

Procaine, butanol and pentanol, at clinical concentrations in the

aqueous phase, significantly reduced the surface tension of

egg-lecithin - cholesterol bilayers. The magnitude of the reduction in

bilayer tension correlated well (with the exception of benzocaine) with

the anaesthetic potency of the anaesthetic molecules on nerve axon.

Previous studies of the dielectric properties of lipid bilayer

systems containing hydrophobic and amphiphilic anaesthetic molecules

(Ashcroft, Coster and Smith, 1977 and Haydon et al., 1977) proposed that

changes in thickness of the bilayer component of the plasma membrane,

induced by local anaesthetics, mediated the blockage of the sodium

channels in nerve axon. However, in this study there was no evidence

found which suggested that anaesthetics with amphiphilic molecular

structures could induce anaesthesia by altering the bilayer thickness.

The fact that some local anaesthetics lowered the surface tension

in egg-lecithin-cholesterol bilayers indicated that the bilayer-torus

equilibrium in these artificial bilayer systems was altered by the

incorporation of these molecules into the membrane structure. The fact

that the anaesthetics had this effect on artificial BLM raised the

possibility that the mechanism of anaesthesia (for amphiphilic molecules

at least) could involve changes in the equilibrium between the lipids in

the bilayer and those in the boundary lipid region of the sodium

channels in nerve. A change in the composition in the lipid

environment, caused by such a change in the bilayer/boundary-lipid

equilibrium could then mediate the blocking of the ion channels.

CHAPTER 12

SUMMARY

12.1 SUMMARY

12.2 SUGGESTIONS FOR FURTHER WORK

Page

223

233

222

223

12.l SUMMARY

i) Bilayer Formation

This thesis describes low frequency impedance measurements on

artificial bimolecular lipid membranes over the frequency range

.003-10000 Hz. Artificial bimolecular lipid membranes were formed from

n-alkane and squalene solutions of glycerol monooleate, egg-lecithin,

and mixtures of the latter with oxidised and unoxidised cholesterol.

The membranes were generated by using the film drainage method of

Mueller, Rudin, Tien, and Wescott (1962). Experiments were performed

with membranes generated from n-alkanes of various chainlengths from

that of n-decane to n-hexadecane.

ii) Electrical Measurements

A four terminal digital impedance measuring technique, based on

that of Bell, Coster and Smith (1975), was employed to measure bilayer

impedance. In this method the membrane impedance was calculated from

the relative phase and amplitude of sinusoidal voltage signals appearing

across the membrane and a known impedance standard when sinusoidal

current of accurately known frequency was passed through the membrane

and the series, known, impedance network. The recent technical

improvements to the impedance measuring system embodied in this system

(the Biophysics Ultra Low Impedance Spectrometer- or BULFIS) allowed

resolution of phase angles and impedances of .1% and .02° over the

frequency range .001 - l0O00Hz.

To accommodate the increased accuracy and extended frequency range

of the impedance measuring apparatus, several improvements to the

224

amplifier electronics were made. This thesis also describes the

improved procedures for calibrating and correcting the differences in

the phase and gain responses and input capacitances as well as the

calibration technique for the impedance standards.

iii) Characterising the n-alkane Solubility Properties of Egg-Lecithin

Bilayers

This thesis describes investigations into the dependence of alkane

solubility in egg-lecithin bilayers on the alkane chainlength,

temperature, its concentration in the torus and the acyl chain ordering

of the lipids.

It was found that the absorption of increasing amounts of n-alkane

into the bilayer interior caused a significant increase in bilayer

thickness. From the relative changes in the bilayer thickness it was

possible to estimate the total n-alkane concentration in the bilayer.

The partitioning of n-alkanes into the bilayer was found to

increase with increasing temperature and decrease with increasing alkane

chainlength. From the variation of thickness of solvent containing

bilayers with temperature and alkane chainlength it was possible to

conclude that for sufficiently long chainlength alkanes at sufficiently 0

low temperatures (ie n-hexadecane below 30 C) the concentration of the

alkane solvent in the bilayer was negligible. In this way egg-lecithin l

bilayers could be produced that were effectively solventless. The

essentially solvent-free nature of these membranes was verified by the

fact that the capacitance of these bilayers was similar to those formed

by monolayer apposition (which are generally accepted to be

solvent-free).

225

The assumption of ideal mixing between the acyl chain of the lipids

in the bilayer and the alkane chains was tested experimentally using

measurements of the n-alkane concentration in bilayers, when the latter

were in equilibrium with a torus containing more than one type of alkane

molecule. It was found that the assumption of ideal mixing was valid

only when the alkane mole fraction in the bilayer interior was less

than 40%.

The absorption of n-alkanes into lipid bilayers was found to be

very sensitive to the ordering of the acyl chains in the hydrophobic

interior. Benzyl alcohol, known to disorder the acyl chain region of

lipid bilayers at high concentrations, dramatically increased the

absorption of n-alkanes into egg-lecithin bilayers. On the other hand,

cholesterol, which is known to increase the order of the acyl chain

region, significantly reduced the absorption of n-alkanes into

egg-lecithin bilayers. This dependence on the acyl chain order was

found to be consistent with a very successful statistical mechanical

model of lipid-alkane bilayers in the liquid crystalline state

(Gruen, 1980b). That model has successfully accounted for the

chainlength dependence of n-alkane solubility in the bilayer.

The production of solventless bilayers allowed a more useful

comparison to be made between

artificial and living systems.

the properties of lipid bilayers

On the other hand the partitioning

in

of

n-alkane solvents between the bilayer and torus was found to be a useful

probe in detecting small variations in the ordering of the lipid acyl

chain. Indeed, the change in partitioning appeared to be more sensitive

to small changes in the ordering of the acyl chains than, for instance,

changes in the thickness of solventless bilayers.

226

iv) The Effect of Ions on Bilayer Capacitance.

The capacitance of all the membrane electrolyte systems

investigated in this thesis increased to an upper limit with increasing

ion concentration in the external aqueous phase.

From the relatively small effect of external ion concentration on

the alkane solubility in egg-lecithin bilayers it was concluded that the

variation in bilayer capacitance with ion concentration was mainly due

to the ion dependent capacitance of the ionic double layers external to

the membrane.

The capacitance of the ionic double layers in series with the

dielectric capacitance of the bilayer significantly reduced the total

measured capacitance of the membrane over the frequency range

.003-10000 Hz. From the analysis of the bilayer impedance dispersion

over this frequency range it was found that the time constant of the

ionic double layer was very close, if not equal (within the range of

experimental error of BULFIS), to that of the bilayer hydrophobic region

at all concentrations of the external electrolyte. This implied that

the conductance of the ionic double layer was very low and essentially

independent of the external ion concentration. This strange result had

been predicted from solutions to the time dependent Nernst-Planck

equations,

Smith (1977).

as applied to the bilayer electrolyte interface by

The capacitance of the ionic double layers external to the membrane

was calculated, assuming that the measured capacitance of the

at high electrolyte concentrations was approximately equal

dielectric capacitance of the membrane (which was assumed

membrane

to the

to be

227

independent of electrolyte concentration). The results were analysed in

terms of the Gouy-Chapman theory. They were consistent with an ion

concentration dependent net charge absorption at the membrane solution

interface, which could be described by the Langmuir absorption isotherm.

The inclusion of oxidised and non-oxidised cholesterol had

distinctly different effects on the measured capacitance of egg-lecithin

bilayers at low ion concentrations. However, the effect of these

compounds on bilayer thickness, although significant, was found to be

small. It was concluded, therefore, that one must be sure to take the

effects of ionic double layers into account when determining thickness

changes in lipid bilayers from measurements of membrane capacitance.

v) Dielectric Dispersion Measurements

a) Characterising the Dielectric Structure

It was found that over the frequency range .003 Hz to 10000 Hz the

total membrane capacitance and conductance of bilayers formed from

egg-lecithin and glycerol monooleate showed a dispersion with frequency.

This dispersion was fitted with a theoretical interfacial polarisation

model for a multilayer sandwich of dielectrics- a Maxwell-Wagner system.

The experimental data was fitted with a Maxwell-Wagner dispersion

modeling a sandwich structure containing 4-6 dielectrically distinct

layers.

By comparing and contrasting the dielectric structures of GMO and

egg-lecithin bilayers determined in this way, it was possible to

characterise the dielectric substructure of both these membrane systems.

228

The dielectric parameters obtained from the impedance data of both

bilayer systems for that part of the substructure with electrical

time-constants greater than .l-1 second were assigned to the acetyl

regions of both of the bilayer systems investigated. Parameters for

the regions with time-constants less than .l second were assigned to

either the hydroxyl region of the GMO bilayers or,

egg-lecithin bilayers, the choline phosphate region.

in the case of

The ion concentration in the aqueous phase had very little effect

on the polar head dielectric structure of any of the bilayer systems

investigated in this thesis.

The conductance of the hydroxyl region of GMO bilayers was

independent of the external electrolyte concentration up to l Molar

concentrations. After this the conductance increased with increasing

electrolyte concentration. This was consistent with there being two

conductance mechanisms operating within the polar region of GMO

bilayers: one being independent of the external ion concentration, and

the other varying more or less linearly with ion concentration. The

latter became significant only at high ion concentrations. This

suggested that a conduction mechanism, other than ion migration, was

operating in the polar head regions of GMO bilayers (perhaps associated

with the intrinsic conductivity of the lipid material). The fact that

the conductance of the polar regions of egg-lecithin was independent of

ion concentration in this study raises the possibility that a similar i

intrinsic conductance mechanism may also have been dominant in the

hydrophobic regions of lipid bilayers.

On the basis of the polar head conductivity at high ion

concentrations it was possible to calculate the partition coefficient

229

for ions between the aqueous phase and the polar head region. The

partition coefficient was of the order of 10- 8 which indicated that there

was no significant ion penetration of the membrane polar structure.

b) Cholesterol

From the relative effects of the inclusion of cholesterol in the

bilayer structure of egg-lecithin bilayers, it was possible to locate

this molecule in the bilayer. Cholesterol when present in egg-lecithin

bilayers significantly reduced the capacitance of the acetyl region.

This was interpreted as being due to the presence of the steroid ring

structure in the region containing the acetyl groups of the lipids. The

hydroxyl group of cholesterol would then have been located in a plane

midway between the carbonyl and phosphate groups of egg-lecithin. This

was also consistent with the effect of cholesterol on the dielectric

parameters determined for the region containing the choline phosphate

groups in the bilayers.

c) n-Alkanes

From the lack of effect of n-alkane absorption on the polar head

dielectric structure it was concluded that these molecules mainly reside

deep within the bilayer structure.

The replacement of H2o in the aqueous phase with D20 had neither a

detectable effect on the dielectric substructure nor an effect on the

ionic double layer capacitance in egg-lecithin-cholesterol bilayers.

This independently confirmed the validity of the use of D2 0/H 2 0

230

replacement in neutron diffraction experiments using lipid multilayers.

e) Anaesthetics

Procaine and benzocaine in the external aqueous phase at pH 6 and

at anaesthetising concentrations had significant effects on the

dielectric substructure of egg-lecithin bilayers, indicating that these

molecules were indeed absorbed into the bilayer structure. The data was

consistent with procaine and benzocaine having opposite orientations in

the bilayer. It was concluded that procaine was aligned perpendicular

to the plane of the bilayer with the aromatic group intercalated in the

lipid acyl chains and the diethylammonium group penetrating the choline

phosphate region of the bilayer. It was reasoned that benzocaine would

be oriented in such a way that its methyl group was embedded in the

acyl chain region and its aromatic ring structure was in the acetyl

region of the bilayer.

The incorporation of alcohols in the bilayer structure had no

effect on the dielectric substructure of the bilayer. Thus these

molecules were "invisible" to the impedance measuring techniques used in

this study.

vi) Membrane Electrical Conduction

The conductance of lipid bilayers in aqueous solutions of different

ion species and varying concentrations was measured. It was found that

the bilayer conductance was relatively insensitive to ion concentration,

radius or charge of the ions. On the basis of this is was possible to

discount the possibility of ion conduction directly through the

hydrophobic region of the bilayer.

231

The bilayer conductance increased with increasing temperature and

was found to vary exponentially with inverse absolute temperature. The

activation energy of bilayer conductance was 35 ± 2 KJ/mole which was

sufficiently low to confirm the previous conclusion that the conduction

did not involve the partitioning of the ions into the hydrophobic

region. On the other hand, the value was sufficiently high to eliminate

the possibility of bilayer conduction via water channels of macroscopic

dimensions.

The data obtained in this chapter was shown to be consistent with

ion conduction via aqueous channels with dimensions of the order of a

few nanometres. The location of these pores, whether evenly distributed

across the bilayer surface, or localised at the bilayer-torus interface

was not certain.

vii) Bilayer Surface Tension

a) Characterising Bilayer Tension

The tension of egg-lecithin and GMO bilayers was measured using the

membrane bowing technique of Coster and Simons (1968). The results were

interpreted in terms of the bilayer torus equilibrium as envisaged by

~~olfe (1979).

The bilayer tension of GMO bilayers were relatively high and I

increased with increasing temperature. On the other hand the tension of

egg-lecithin bilayers was low and decreased with increasing temperature.

This temperature dependence allowed certain insights into the bilayer

torus equilibrium.

232

The data were consistent with GMO molecules in the bilayer being in

equilibrium with small micellular aggregates in the torus (approximately

100 molecules per aggregate). The data indicated that egg-lecithin· in

the torus must be in relatively large aggregates, so that the entropy of

mixing of lipids in this system had little effect on the bilayer

tension.

b) Cholesterol

Cholesterol was found to increase the tension of egg-lecithin

bilayers but the mechanism for this effect was not certain.

c) Anaesthetics

Procaine, butanol and pentanol, at relatively low concentrations in

the aqueous phase, all significantly reduced the tension of egg-lecithin

bilayers. Therefore it could be concluded that these anaesthetics

significantly altered the lipid equilibrium between the bilayer and

torus components of the membrane. On the basis of this result it was

suggested that the mechanism of action of local anaesthetics on nerves

may involve alteration of the lipid equilibrium between the annulus of

lipids around the proteinaceous sodium channels and the lipids in the

rest of the bilayer.

233

12.2 SUGGESTIONS FOR FURTHER WORK

Artificial bimolecular membranes, used in this study to model the

bilayer component of living membranes, have physical properties peculiar

to the lipid equilibrium that exists between the bilayer and torus

components of the membrane. It is important to remember this when

applying the conclusions based on this artificial system to living

membrane systems, where the lipid equilibrium between the membrane and

the environment will be different.

An alternative model that could be employed in dielectric studies

would be large bimolecular vesicles. Vesicle systems would model the

bilayer component of living membranes that are not in equilibrium with a

lipid ''reservoir". Dielectric and surface tension studies on large

vesicles would be an interesting extension to the work reported in this

thesis.

The main advantage of using vesicles is that they contain

absolutely no hydrophobic solvents, as they can be readily formed

without the use of such solvents. The results obtained from vesicle

systems would be easier to interpret than those obtained from planar

bilayers which are in equilibrium with a solvent reservoir.

Vesicles, provided they are of

aspirated into small glass pipettes.

sufficient diameter could be

The pipette would act as an

insulating border (ie. functioning as a septum) as well as providing a

means of mechanically clamping the vesicle for the purposes of

electrical and mechanical measurements. The aspiration of single cells

234

into glass pipettes is now a commonly used technique in membranes

studies (eg. Wolfe and Steponkus, 1981).

Further work could also be undertaken in characterising the

electrical and surface tension properties of artificial bilayers

containing proteinaceous excitability inducing material (EIM). The

effect of changes in bilayer tension and thickness on the electrical

activity of these membrane components could then be investigated. This

would allow additional insight into the dependence of sodium channel

conductance in nerve axon on the state of the surrounding lipid matrix.

The effect of membrane thickness on the electrical activity of

gramicidin (a passive, transmembrane ion pore) in lipid bilayers has

been investigated by Hendry, Urban and Haydon (1978). The electrical

properties of EIM isolated from Enterobacter cloacae, reconstituted with

lipid bilayers has been well characterised in previous studies

(Kushnir,1968 and Bean, 1973). However, no study of the effects of

bilayer tension and thickness or addition of local anaesthetics on

membrane excitability in these model membranes has been made.

235

APPENDIX A

DATA STORAGE AND PRESENTATION

The BULFIS system required information necessary for programing the

signal generator and transient recorder boards as well as correcting for

the relative phase and gain response of the differential amplifiers on

each input channel of BULFIS. This information was stored on magnetic

disc in the "frequency file". A print-out of this file is shown below.

19 10 1 0 0 .-,c-c:- 0 .-,C' -. 44 -.44 .00 .994700 "-·-'·J -'·-' 20 10 1 0 0 .-,C'C' 0 .-,c- .-.c-

-·· .39 00 _9·;,4700 ..::.._,._, .::.,._, - . . ,:_,._, . 21 10 1 0 0 .-.c-c- 0 50 - 11--, - .-.. -. .00 .994700 ..::,._,._, ..... ,;- .. ,:.. . .::-.-.. -. 10 1 0 0 .-,c-c- 0 100 • :~:0 -.20 00 .994700 L..:.:. ~--'·-' . -.. -. 10 1 0 0 .-,c-c:- 0 150 - .-.. -. - 1 ·-:• .00 . •;,94 70 0 ~-.;, ..::,._,._, . . _;..::_ . .:..

24 10 .-. 0 0 .-,cc:- 0 200 -·. 27 -.77E-1 00 . ·;,·~4730 ..:. ,.::.._,._, . .-,e ..:.:.-J 10 1 0 0 .-.c-c:-

L..·-'·-· 0 .-.c:-c-.::.,.._,._, -.72E-2 - .-,c- .00 .994700 . .::.,.._, 26 10 .-. 0 0 .-,c-c:- 1 .-,c- - • :31 - ,-,,-, 00 .994700 . .:,, ,.::.._,._, .:.,._, .v•::• . 27 10 C" 0

,. ·--C"C' 1 .-.c- - . -.-, - .-,c- 00 .994700 ·-' '(/ ..::,._,.J L·-' • 4'. I .. ::,._, . .-.. -. 10 l, ,. 0 .-,c::-c- 1 . -,c _ ... 21.; . - .-.. -. 00 .994700 ..:.:.·~ '(/ ,.:;.._,._, ..::,._, • v•=- . --.. -. 10 .-. 0

,. .-,c-c- 1 .-,c- - .-.-, - .-,c- .00 .-•• -. " -, "' /1. .::..~ ,:;:, 1(/ ..:.:.,._,._, £,_, • 4/ .. ,::.._, . ?)"TI 'f/'(J

:30 . ,. 15 ,. 0 215 1 .-,C' -· 1 ·-=· - "C" 00 . 'i'l4700 J. '(/ '(/ ,£._I . ·-· . ..,.._, .

31 10 • C" 0 0 .-,c-c- • .-,c- - .-. , - .-.. -. .00 . ·;:•i4 70 0 J.._, ,.::._,._. J. ..::,._, • . ,,:.,1:, • -:.-•J -.. -, 10 • C" 0 0 .-.c-c.~ 1 .-,c- .-. ...,. --- .-,c- 00 • 9'?'4700 ;,..::.. J.-• ..::,._,._, .:,..._, - •._-.I . .,:,._, . ,·,.-. 10 15 0 0 .-.c:c:- 1 .-,c- - .-.-,. -- . :3·;, .00 . 9'?4700 ._:,.,;. L·-'·-' .a::,._, . -~ / .·-, 11 10 • C" 0 0 . -,c-c- 1 .-,c .. .-.. -. .-,c- ,. ,. .994700 ._:,.,. J.--· ,.:,,._,._, .::,.._, • -.:.-·=· .. :;,_ .. _. . 1(/'(/

·,c::- 10 • C" ,. ,. .-,c-c- 1 . -,c .-,.-. " . .00 • ·:;,·~4700 .:..•--• L ._, '(/ '(/ ..::.-J·-' ..::,._, - . .,;;_.,,:._ • .,. J. -,, 10 • C" ,. 0 .-,c-c- 1 .-,c- -. 40 -· .-. " 00 • '?'7'4 70 0 ,;.,:._. J. ._, '(/ ..::_._,._, ..,:;,.._, . ..,:..-., . ·V""7 10 15 0 0 .-,c-c 1 --.c- /I ,. -- .-. , .00 .994700 j/ L•.J•J ,.:::..._, . .,. '(/ .. .:;.,,~ -.. -. 10 15 0 0 .-,C"C" • .-.c- ·- /I • .. .-,c- 00 . '7"~14 70 0 _-.o L·-'·-' J. .::,,.._, • .,. J. . ~--· . ·-,.-. 10 • C" 0 0 .-,C'C' • . -,c- " . ·- .-, I .00 • '7"7'4700 J7 J.•..J ..:.:,._,._, L ,._, • .,. J. .. :._,,.:.,

40 • ,. 15 ,. 0 .-,c-c:.- • .-,c -- .41 ~-. 34 00 . ·;,·~J4700 J. V V ..::,._,._, J. L·-' . ., . 10 • C" r • 0 .-,C"C' • C",. .-,-, -- .-... 12E • . •7•;•4700 ·t J. L ._, 1(/ ..:,:.._,._. L ·-''fl • • .:;,1 / • • ,:.. J. . -L

11,-. • ,. • C" 0 ,. .-,c-c- • .-,c- /I • .-. " ... -,,- .. .·;:·~4700 ·-t"- J. l(J

.l ·-· V .:::,.._,._, J. .;.._. • .,. J. .. ~ .. ..., . J..::.c:;.··-J. ,,.-. • ,. • C" ,. ,. .-,c-c- • . -,c- ,..-. .-,c- .. . -,r- '4 • '?"~4700 't.:;. J."' J.._, "' V ..:.:,._,.J J. .::_._, • -t-L . -~·--· . J. .::.c.·-- J. . " • ,. • C" ,. 0 .-.. -.. -. 1 .-,c - /I • ,., " • 20E'-· 1 .994700 "T"t J. l(J

.l ·-· V ..::..::.~ .:,.._, •• .,. .L . ..:., .....

,1 a= • ,. • C" 0 0 .-.. -.. -. • ,. ,,-, ".-. .-•• -,,- <f • ·~·:;-4700 t,_I J. '(/ J. ·-·

..:,;. . .;., . .:, L V • "t I ....... ~ • ,.:,.::..c,-·~ L II I 10 • C" ,.

0 24:~:i • 0 "-, - " . /I.,- 1 • •;1•;J4 79•;1 .. .,.,.::, L ._, '(/ J. • "T / • .,. J. • .,. J. t:. -· ., .-, • C" ,. 0 .-,c:c- • 0 /I-, 4·-::· • f,0E··-· 1 . 'l';-'5b'i'i t I ::r L ._, 1(/ .:;.._ .. _, J. • ""t I . ~

"·-· .-. • C" ,. ,. .-.c-c· 1 10 -· /IC" -·· .-.. -. r,Ar- • .:~·;15:.:~00 •t•:.• / J. ·-' V 1(/ .:;,._,._, • 6"t·-' • -;t•.:• . ."'(JC. ·J.· 11,-. .-, 15 0 0 .-.c:-c- • 50 ·-=··~~ . -.. -. • • • '7"/55'7''? ·t / ·=·

..:.:,._,._, J. . ·-· .. · • L 7 . J. J. ~-, ,-, • C" - ,. .-,c-c: 1 100 • !' .-.. -. 1 l, . ,7 .. ;:55·;·;· ·t I ;,

J. ·-' ·, V .:.:,._,._, . J. I ·- • .;.1.::\ . -,,.-, .-, 15 .-, ,. . -,c-c- 1 200 .-. n. • I t .-. . ·;··;:54·-;:··;-t•.:• .:• - V ..:,;,._,._, . . .:_, '(/ . J.I.:,;., . J. .:•

,, .-. .-. • c::- -: ,. .-.c-c- • .-,--,c .-, n. • • .-,11 ,-, ,-, C' II ,-, ,-, 1 .- •.:..• J. ._, ·- V ..:.:,._,._, J. ..:.:.~·-' . . .:_, '(/ . J. J. • L6t . :-· ;;• ._,.., .-=' .:• .,--, .. -. • c::- -- " .-,c-c- • . -,c.-~ '4 .-• • ,. .-,--, .-.. -, " .-, .-.. -.

j / :• J. ._, V .&:.:.,_,._. J. _.:_._,._, . L ,_. . J. '(/ • .:.:, I . 7l'...-L?:· ... 11,-, c:- .-. ,. .-,c-c:- • .-,cc- .-.11 -,,-., 1 .-,c- .-. ,-•• -. --, II. (1.

·t·~· -' V .:_._,._, J. ..:.:,._,._, • ..:.. V . I •:.:•c.. .. ,,:;,.,._. . / .:-1.,:_., vv ., .-, ,·-, i:::- ·- ,. -.. -., --, 0 .-, / .-.. -. ,. ,. ,-, ,-, .-. &:: .-•• -, 't.·· ,..:., ··-' V .:_.,_,J ..:: .. .:._,,:.:, . . ::,, . .,;. .... \:J . :r,::,, ;.•._• _; .. 7-

~--, - c; .-,-, 0 . -,c·c.- ,., ,.. .-. --y .-, I c:-,-• ·~-·:::l, ·?·~,-~ .. t I - ,;__ I .:.:,._,._, .::. V - • . .;_, I . ..:,,.:;.. •• _, :.v .

The significance of the columns of data (from left to right)

described:

236

are now

(F)- programs the rate at which the function in the signal

generator is clocked out. The clocking rate is calculated

using the following formula:

clocking rate= lMHz x 2 (F)/2 1024

(S)- sets the number of analog steps for the sinusoidal output of

the signal generator. The number of steps is given by:

steps per period= 2 (S)

(V)- sets the number of periods over which the transient recorder

boards sample and average.

(G)- The first argument sets the number of periods of the sine

function loaded into the signal generator and the second tells

BULFIS where to find the function (either in EPROM (0) or on

di SC ( l ) ) .

(A)- determines the amplitude of the output signal (O=OV to 255=5V).

(L)- both arguments select the RC filters on the output of the

signal generator.

{D)- sets the D.C. offset on the output signal (volts).

{M)- are the correction factors for the relative phase and gain of

the differential amplifiers at each frequency.

237

When BULFIS was in an impedance measuring mode the impedance data

was stored on magnetic disc an simultaneously displayed on a visual

display unit (VDU). The format of the data appearing on the VDU is

given below:

5.0:.1E-3.5E-4 .745E-2 -7.4 -~. 0 : . ·;:E -4. 5E -4 • 1 0l,E -1 1 0 7.0:.5E-4.6E-4 .149E-1 -14 ::::.0: .4E--4.bE--4 .212E-1 -20 ;, . 0: . :3E--·4. 8E-4 10:. 2E-4. 'iE·-··4 11 : • 2E-4. 1 E-3 12: .2E-4.1E-:3 1 :3: • 2E -4 • 1 E-:3 14:. lE-4. lE·-3 15: . 2E -4 • l E -:3 U.~ •• 1 E-4. 0E--:3 1 7 : • :3E -5 • l ... E -4 1:3:. 7E-5.5E-4 1'?:. lE-4. lE-3 20:. 5E-5. :3E-4 2l:.4E-5.2E-4 22:.4E-5.2E-4 2:3: .4E-5.2E-4 24: .4E--5.2E-4 25: . 3E-·5. 2E-4 26: . :3E·-5. 2E-4 27:. :3E-·5. 2E-4 .-.,-, • .-, r- C' .-. ,- II ,..::,:-, • •.,;.c.-._,• .L.C -61'

2'i:. :3E-5. 2E-4 -~:0:. :3E-5. lE-4 :31:. 2E-5. lE-4 :::2:. 2E-5.1E-4 . 3:3: • 2E -5 • 1 E -4 :34 : • ::::E -5 • 1 E -4 :35: • :3E -5. 1 E -4 3b:. :]E-5. lE-4 :37 : • 2E -·5 • 7E -5 :3:3: • 2E -5. 7E -5 .3·;-:. 4E-5. 6E-5 40: • :~:E-5. 7E-·-5 41:.7E-5.9E-5

.425E-1 .596E-1 • :351 E·-1 .119 . 170

.-.. -.. -. • ,.:;_.~,:::,

.:340

. 476

.6:31 .-,c---, • 7._, . .,

1. :36 1. "i0 .-. ,.-. ..::. • I..::.,

C" II II ._ .. ~.., 7.62 10. :3 15.2 21. 7 :30 .5 4:3.5 61.0 :37. 1 .. .-.. -. J.£~

174 244 :34:::: 4:3:3 6'i7 ·;.,71:.. 14l:,E 1 20'iE1

42:.lE-4.lE-4 292El 43:.2E-4.2E-4 439El 44:.bE 4.bE-4 627E1 45:.1E--:3.1E-:~: 87:3E1 46-.3E 3.2E 3 131E2

-·27 -36 -45

c-c:--._,._,

-7/~

.-.. -. -.,:-. . .,;.

.-,c­·-·=-·-' -:36

,-,-, -,,:-., .-,.-, -·=-·=-.-.. -. -·=-·=--:3•;, -:3'?' .-..-. -,,:..7 .-.. -. -·=-·"'

-:3•;1 -:3•-;, -:3'"i' -:3•-;, -:3•-;, --:3•-;i

-:3·/ -:3•-;, ---:3·-;: -:3•-;, -:3·;) -:3•-;, -:3•-;, -11El -11E1

331E6

:32:3El:1

:31:3El:, 2'7'7E6 269E6 23:3E6

14bE6 107E6

5l,1E5 403E5 .-•• -,,-,r-c:-..:.:.•=•·:;ac..,_,

202E5 142E5 101E5 710E4 508E4 :355E4 254E4 177E4 127E4 .-•• - •• -. r- .-, ·=-·=- ,':'C,,.;t I.-, 11r-,-, ,_ .... .:- ..... c.-~ /I II II,-.-, "1""1' "'tC.-:>

317E3 222E:3 .. c-,-• .--.-, J. ·-··=-C.·-' 4 <t 4 ,-.-. J. J. .L c.-.:>

C'C"C'r-•-, ._,._,._,c,, :3'?6E2 .-,-,-,,.-. £ I I C:.L.

1 ·;,::::E2 1:32E2 'i24E1 l,b0El 441E1 :309E1 221El 14::::E 1

. 2·:;,9L0E-2 • 29:389E-2 • 29•:;,03E-2 • 2·~·7'05E-2 .29867E-2 .29888E-2 • 2·i8•;J:3E-2 • 29886E-2 .29980E-2 • :30055E-2 .2'7'837E-2 .30069E-2 .30131E-2 .29814E-2 .29606E-2 .30652E-2 . 302:33E-2 .:30622E-2 .30870E-2 • :33331E-2 .29219E-2 .34';>03E-2 .30246E-2 .27445E-2 .30275E-2 . 36620E-2 .44062E-2 .50119E-2 .35796E-2 .14644E-1 . 10465E-1 .54022E-2 • 77181E--2 .29329E-1 • :300'7'5E-1 .42990E-1

• :3:3077E-2 . :3:3128E-2 • 8281:3E-2 .:32459E-2 .02400E-2 • :322:34E-2 .:32233E-2 .82130E-2 • 8219'7'E-2 • :=:2212E-2 • :32144E-2 . :32058E-2 • :3211 'IE-2 .82139E-2 . :32120E-2 .82137E-2 • :3213:3E-2 • :32158E-2 .82085E-2 .82133E-2 . :32 0'i7E-2 • :320:3·;,E-2 .:32116E-2 . :3207bE-2 • :32130E-2 .82l56E-2 • :321 ';12E-2 .:32205E-2 .:32202E-2 • :321 ·~bE-2 .82173E-2 .:32207E-2 .82275E-2 • :322:33E-2 .:32283E-2 • :=:226 7E-2 .:32262E-2

.32397E-1 .82235E-2

.39481 .82046E-2

.75156 .81995E~2 -.24818 .81b63E-2 -3.3537 .81239E-2

238

The data in the columns (from left to right) are:

#)- The entry number of the data stored on the disc

COEFF'S (V,C)- is the normalised fit parameters of a sine-wave function

to the raw voltage signals appearing at the voltage and current

channels of BULFIS.

HERTZ-

DEGREES}

OHMS

US/CMA2

US/CMA2

The frequency of the A.C.

generator.

signal output from the signal

The phase and magnitude of the total measured impedance

The impedance expressed in terms of the equivalent

parallel

capacitance (uF/m2 ) and conductance (uS/m2 ).

The data stored on disc was stored in the "data fi 1 e" whi eh had. the

following format:

·-· . ,.

I ,-. • <i • ...

L ...;_ "

.l ·-' •

t ,, '

' , .

,-·., • .._.. .-··,r·- ,-.,-,r-r- r- , ,- r-, 1 l A,-.,- , r, A r, T ,._ ",,-. , ·-• 1- 11 I '-'t 1-· 1-'LI I ·-• I I 11-1--•L \ I \t··tL1 .I. f··1t•1._, /

,t C.-,,-·, r ,-·, i r· .·-, • J. ·-1 .:... •-• I •. • l L

,. c:- .. / 11 .-,c:- r- .-, • .i. ._, J. ·--· \I .:.. ·-· '- ·-·

.. -4. ,-, .... , .. .-, .-, .-- .·-, • J. J. •• · ...: .. .L ,._._,L,

i .. ,-•• - ... /),.; r-- J. L ,,..;, .. L 'V ,:.,;,L -•

,. .-,.-1 r .. c:.-, c-r- · . • J. .:. .. • ;.) \,- ·-' ,_, ._1 L ..

i ,-,r_- ,••, l I .-.,··· • .l ..;.-.-•..:..,_,,_.;t'L-

-1 /t C:- r.-· '" .-, . r-• J . .,.._1._1 '<I ,._,_IL

.. , .. , ,···, ,, /I 'l ,--• .i. •-• ·'T .·· '"'t --,- '1 L

•rr-• J. I ,._ --1· •-••-•·-'L _, . ·-, - .... --,.-,,-,.-,,--

··, ., -'• ,., r· J •- ' '-, ~ .. : L.

-4 I 'l,-, -, /lc-r-.__ l ·-· i ...

" .-. ,::-- ,. ,:-:- .-.. -. • ·-r ._, ._,-,- ._, ,_, ._,

C"C 4 ,-. II ,-,r:-• ._1 ._, J. •-• ,.T •• · •.-•

I II, ... ,,-,-.-,,-·, • •-• ·•·1· •-•~• I ._, '-·•

••• , ... - •• -. !'1 .-, .·-,

• I .l ~--•"'T~·--• ....,. ., .-, I\,-,,,--;

• , ... ·-· i,i . ··, , ' r·- .-, ; 1' I

• ·-··-······-··-· ...... J c:·· - _.., ,-,c- ,-, .·-. .. _,' '·-··-' .· ..;...

" r.::-- ,-, .-, .·-, c- ,, . ;· ·-· .· ·- ·-' -· . -, r·- -·-.. -, ~ r·· r-

• ·-' ._., ~ ·-· ... ·-· ··-· ·-, r- ,-, ,-, r~ ,-, .-.

·- ·- -··- ·.-' -' ,-,,-·, .·-,.-. ·"' '

I._,,_,-·

r-1 A ,- T , -. i I '

f\t·, ' J. ·-· ' ... , "' ., .-.r··., _. r'.'",.

• •T.:..·-•· • ,_,._,-·1 ., .-..... ~ ,-, .-, _.·t . ..,. ·-· \, ·,· ·-· ~ .. , C' I I • --,,-, 11

•• _ .. _. ,_, . __ . -· • 1

..,. .,. .-, .-.. -, .-.. -. • I ~ ._, ··-· _._ . ~

,-.~--:, .. - 4 .-,r.::-. ... _,' r.,"·,- ..

' -.~· ··-· ·-· -I .-, r

J. • ·-· '-'' I .I. , . .... • j ·-

... • ·-· J. ...

..... ·-·

239

The left column contains the sum of the normalised fit parameters

of the voltage and current channels of BULFIS at each frequency. The

second and third columns contain the phase difference and amplitude

ratio of the voltage signals arriving at both input channels of BULFIS.

240

REFERENCES

Andrews,D.M. and Haydon,D.A. (1968) J. Mol. Biol 32,149-150

Andrews,D.M.,Manev,E.D. and Haydon,D.A. (1970) Spee.Disc. Far. Soc. 1,46-56

Ashcroft,R.G. (1979) PhD. thesis, University of New South Wales

Ashcroft,R.G., Coster,H.G.L. and Smith,J.R. (1977) Biochim. Biophys. Acta 469, 13-22

Ashcroft,R.G., and Coster,H.G.L. (1978) Bioelectrochem. Bioenrg. 5, 37-42

Ashcroft,R.G., Coster,H.G.L. and Smith,J.R. (1981) Biochim. Biophys. Acta 643, 191-204

Bean, R.C. (1973) Membranes 2, 409-477

Bates,E.H. and Wolfe,J. (1980) in Advances in Liquid Crystals Research and Applications (Bata,L., ed.) pp 739-749 Akademiai Kiado, Budapest

Bell,D.J., Coster,H.G.L. and Smith,J.R. (1975) J. Physics E. 8, 66-70

Benz,R., Frohlich,0., LaUger,P. and Montal,M. (1975) Biochim. Biophys. Acta 394, 323-334

Benz,R. and LaUger,P. (1977) Biochim. Biophys. Acta 468, 245-258

Birrell,G.B. and Griffith,0.H. (1976) Biochem. 15, 2925-2929

Bockris,J. and Reddy,A.K.N. (1970) Modern Electrochemistry, Plenum press, N.Y.

Boggs,J.M., Roth,S.H., Yoong,T., Wong,E. and Hsia,J.C. (1976) Mol. Pharmacol. 12, 136-143

Boggs,J.M., Yoong,T. and Hsia,J.C. (1976) ~Mol. Pharmacol. 12, 127-135

Born,M. (1920) Z. Phys. 1, 45-48

Boulanger,Y., Schreier,S.,Leitch,L.C. and Smith,I.C.P. (1980) Can. J. Biochem. 58, 968-995

Boulanger,Y., Schreier,S. and Smith,I.C.P. (1981) Biochem. 20, 6824-6830

Brockerhoff,H. (1974) Lipids 9, 645-650

BUldt,G., Gally,H.U., Seelig,A., Seelig,J. and 1 Zaccai,G. (1979) J. Mol. Biol. 134, 673-691

Bunce,A.S. and Hider,R.C. (1974) Biochim. Biophys. Acta 363, 423-427

241 Caffrey,M. and Feigenson,G.W. (1981) Biochem. 20, 1949-1961

Carnie,S. Israelachvili,J.N. and Pailthorpe,B.A. (1979) Biochim. Biophys. Acta 554, 340-357

Cass,A. and Finkelstein,A. (1967) J. Gen. Physiol. 50 1765- 1784

Chatterjie,N. and Brockerhoff,H. (1978) Biochim. biophys. Acta 511, 116-119

Chemical Rubber Company Handbook of Chemistry and Physics (1976) 57th edition., West,R.C. ed. Chemical Rubber Company Press.

Chin,J.H., Parsons,L.M. and Goldstein,D.B. (1978) Biochim. Biophys. Acta 513, 358-363

Ciani,S., Gambale,F., Gliozzi,A. and Rolandi,R. (1975) J. Memb. Biol. 24, 1-34

Clowes,A.W., Cherry,R.J. and Chapman,D. (1971) Biochim. Biophys. Acta 249, 301-317

Cole,K.S. (1968) "Membranes, Ions and Impulses; a chapter of classical biophysics" University of California Press pp 458-463

Cole,K.S. (1969) Biophys. J. 9, 465-469

Coleman,R. (1973) Biochim. Biophys. Acta 300, 1-30

Coster,H.G.L. (1973) Biophys. J. 13, 1119-1123

Coster,H.G.L., James,V.J., Berthet,C. and Miller,A. (1981) Biochim. Biophys. Acta 641, 281-285

Coster, H.G.L. and Simons, R. (1968) Biochim. Biophys. Acta 163, 234-239

Coster, H.G.L. and Simons, R. (1970) Biochim. Biophys. Acta 203, 17-27

Coster,H.G.L. and Smith,J.R. (1974) Biochim. Biophys. Acta 373, 151-164

Cress,A.E. and Gerner,E.W. (1980) Nature 283, 677-679

Curatolo,W., Sakura,J.D., Small,D.M. and Shipley,G.G. (1977) Biochem. 16, 2313-2319

Danielli,J.F. (1966) J. Theoret. Biol. 12, 439-441

Darke,A., Finer,E.G., Flook,A.G. and Phillips,M.C. (1972) J. Mol. Biol. 63, 265-279

D'Arrigo,J.S. (1978) Am. J. Physiol. 235, Cl09-Cll7

Ebihara,L., Hall,J.E., MacDonald,R.C., Mcintosh,T.J. and Simon,S.A. (1979) Biophys. J. 28,185-196

Edidin,M. (1974) Ann. Rev. Biophys. Bioenerg. 3, 179-201

Ehrenstein,G. (1976) Physics Today (october) pp 33-39

242

Everitt,C.T. and Haydon, 0.A. (1968) J. Theoret. Biol. 18, 371-379

Feiser,L.F. and Feiser,M. (1959) Steroids, pp 233-235 (Reinhold)

Fettiplace,R., Andrews,D.M. and Haydon,D.A. (1971) J. Membrane Biol. 5, 277-296

Fettiplace,R., Gordon,L.G.M., Hladky,S.B., Requena,J., Zingsheim,H.P. and Haydon, D.A. Methods of Membrane Biol. (1975) (Korn,E.0., ed.) Vol 4. pp 1-75. Plenum press New York

Fettiplace,R. (1978) Biochim. Biophys. Acta 513, 1-10

Finkelstein,A. and Cass,A. (1967) Nature 216, 717-718

Franks,N.P. (1976) J. Mol. Biol. 100, 345-358

Franks,N.P. and Lieb,W.R. (1978) Nature 274, 339-342

Ginsberg,S. and Noble,D. (1974) J. Memb. Biology. 18:163-76

Goldman,D.E. (1943) J. Gen. Physiol. 27, 37-60

Goldup,A., 0hki,S. and Danielli,J.F. (1970) Ree. Proc. in Surf. Sci. 3, 193-260

Grabowski,E.F. and Cowen,J.A. (1977) Biophys. J. 18, 23-28

Graziani,Y. and Livne,A. (1972) J. Memb. Biol. 7, 275-284

Gruen,0.W.R. (1980a) Biochim. Biophys. Acta 595, 161-183

Gruen,0.W.R. (1980b) Biophys. J. 33,149-166

Gruen,D.W.R. (1980c) Biophys. J. 33,167-187

Gruen,D.W.R. and Wolfe,J. (1982) Biochim. Biophys. Acta 688, 572-580

Gutknecht,J. (1981) J. Memb. Biol. 61, 61-66

Gutknecht,J. and Walters,A. (1980) J. Memb. Biol. 56,65-72

Gutknecht,J. and Walters,A. (1981) Biochim. Biophys. Acta 645, 161-162

Hanai,T.,Haydon,D.A., and Taylor,J. (1964) Proc.Royal.Soc 281,377-391

Hanai,T., Haydon,D.A. and Taylor,J. (1965a) J. Theoret. Biol. 9, 278-296

Hanai,T.,Haydon,D.A. and Taylor,J. (1965b) J. Theoret. Biol. 9,422-432

Hanai,T., Haydon,D.A. and Taylor,J. (1965c) J. Theoret. Biol. 9, 433-443

Haydon,0.A.,Hendry,B.M.,Levinson,S.R., and Requena,J. (1977) Biochim. Biophys. Acta 470,17-34

Haydon,D.A. and Taylor,J. (1968) Nature 217, 739-740

Hendry,B.M., Urban,B.W. and Haydon, D.A. (1978) Biochim. Biophys. Acta 513, 106-116

Hille,B. (1980) "Progress in Anesthesiology" V2 ed. Fink,B.R. Raven Press pp 1-5

Huang,C.H. (1976) Nature 259, 242-244

Huang,C.H., Wheeldon,L., and Thompson,T.E. (1964) J. Mol. Biol. 8, 148-160

Huang,W.T. and Levitt,D.G. (1977) Biophys. J. 17, 111-128

243

Hope,A.B. and Walker,N.A. (1975). "The Physiology of Giant Algal Cells" Cambridge Press.

Jacobson,K. and Papahadjopolous,D. (1975) Biochem. 14, 152-161

Israelachvili,J.N. (1977) Biochim. Biophys. Acta 469, 221-225

Israelachvili,J.N. (1978) in "Light Transducing Membranes Structure, Function and Evolution", Ed. Deamer,D.W., Acced. Press, pp 91-107

Israelachvili,J.N.,Marcelja,S. and Horn,R.G. (1980) Quart. Rev. Biophys. 13, 121-200

Israelachvili,J.N., Mitchell,0:J. and Ninham,B.W. (1976) J. Chem. Soc. Faraday Trans. II 72, 1525-1568

Kushnir,L.D. (1968) Biochim. Biophys. Acta 150, 285-299

Langmuir,I. and Waugh,D.F. (1938) J. Gen. Physiol. 21, 745-755

LaUger,P.,Lesslauer,W.,Marti,E. and Richter,J. (1967) Biochim. Biophys. Acta 135,20-32

Lee,A.G. (1976) Nature 262,545-548

Lee,A.G. (1978) Biochim. Biophys. Acta 514, 95-105

Mac0onald,R.C. (1976) Biochim. Biophys. Acta 448:193-198

Mac0onald,R.C., Simon,S.A. and Baer,E. (1976) Biochem. 15, 885-891

Lorrain,P. and Corson,D. (1970) Electromagnetic Feilds and Waves, 2nd. ed., W.H.Freedman and Co.

Marcelja,S. (1974) Biochim. Biophys. Acta 367, 165-176

Marcelja,S. (1976) Biochim. Biophys. Acta 367, 165-176

Mauro,A. (1961) Biophys. J. l, 353-372

Mauro,A. (1962) Biophys. J. 2, 179-198

Mcintosh,T.J. (1978) Biochim. Biophys. Acta 513,43-58

Mclaughlin,S. (1975) "Progress in Anesthesiology" Vl ed. Fink,B.R. Raven Press pp 193-220

Miyamoto,V.K. and Thompson,T.E. (1967) J. Col. Int. Sci. 25, 16-25

244

Montal,M. and Mueller,P. (1972) Proc. Nat. Accad. Sci. USA 69, 3561-3566

Moore,P.B. (1975) in Brookhaven Symp. (Shoenborn,B.P. ed.), Vol. 27, VII 67-74

Mueller,P., Rudin,0.0., Tien,H.T. and Wescott,W.C. (1962) Nature 194, 979-981

Neumcke,B. and Lauger,P. (1969) Biophys. J. 9, 1160-1170

Neumcke,B. Walz, and Lauger (1970) Boiphys. J. 10, 172-182

Nicolson,G.L. (1976) Biochim. Biophys. Acta 457, 57-108

Ninham,B.W. and Parsegian,V.A. (1970) J. Chem. Phys. 53, 3398-3402

0hki ,S. (1968) J. Theoret. Biol. 19, 97-115

0hki ,S. (1970) Biochim. Biophys. Acta 219, 18-27

0hki,S. and Goldup,A. (1968) Nature 217, 458-459

Pagano,R.E.,Rysschaert,J.M. and Miller,I.R., J. Memb. Biol. (1972) 10, 11-30

Pagano,R. and Thompsom,T.E. (1967) Biochim. Biophys. Acta 144, 666-669

Parsegian,A. (1969) Nature 221, 844-846

Parsegian,V.A. and Ninham,B.W. (1969) Nature 217, 1197-1198

Pang, K.Y. and Miller, K.W. (1978) Biochim. Biophys. Acta 511, 1-9

Persson,N.0., Lindblom,G. and Lindman,B. (1974) Chem. Phys. Lipids 12, 261-270

Pillips,M.C. and Finer,E.G. (1974) 356, 199-206

Pope,J.M. and Cornell ,B.A. (1978) Prog. in Surf. Memb. Sci. 12, 183-243

Price,H.D. and Thompson,T.E. (1969) J. Mol. Biol. 41, 443-457

Redwood,W.R. and Haydon,D.A. (1969) J. Theoret. Biol. 22, 1-8

Reiber,H. (1978) Biochim. Biophys. Acta 512, 72-83

Ritchie,J.M. and Greengard,P. (1966) An. Rev. Pharmacol. 6, 405-430

Richards,C.D. (1976) Nature 262, 534

Rosen,D. and Sutton,A.M. (1968) Biochim. Biophys. Acta 163, 226-233

Rosenberg,B. and Bhowmik,B.B. (1969) Chem.Phys.Lipids. 3:109-24

Rosenberg,B. and Jenriasiak,G.L. (1968) Chem. Phys. Lipids 2, 47-54

Sacre,M.M. and Tocanne,J.F. (1977) Chem. Phys. Lip. 18, 334-354

Sandermann,H. (1978) Biochim. Biophys. Acta 515, 209-237

245

Seelig,J. and Seelig,A. (1974) Biochem. 13, 4839-4845

Seelig,J., Gally,H. and Wohlgemuth,R. (1977) Biochim. Biophys. Acta 467, 109-119

Seeman,P. (1972) Pharmacol. Rev. 24, 583-655

Seeman,P. and Roth,S. (1972) Biochim. Biophys. Acta 255, 171-177

Simon,S.A.,Lis,L.J.,MacDonald,R.C. and Kauffman,J.W. (1977) Biophys. J. 19,83-90

Simon,S.A., Stone,W.L. and Busto-Latorre,P. (1977) Biochim. Biophys. Acta 468, 378-388

Simon, S.A., Stone,W.L. and Bennett,P.B. (1979) Biochim. Biophys. Acta 550,38-47

Simons,R. (1968) J. Mol. Biol. 36, 287-288

Singer,S.J. and Nicolson,G.L. (1972) Science 175, 720-731

Smith,J.R. (1977) PhD thesis, university of New South Wales.

Staehelin,L.A. and Arntzen,C.J. (1979) in "Chlorophyll Organization and Energy Transfer in Photosynthesis" Ciba Found. Symp. 61, 147-175

Staiman,A.L. and Seeman,P. (1975) Can. J. Physiol. Pharmacol. 53, 513-524

Stephens,C.L. and Shinitzky,M. (1977) Nature 270, 267-268

Stier,A. and Sackmann,E. (1973) Biochim. Biophys. Acta 311, 400-408

Stockton,G.W. and Smith,I.C.P. (1976) Chem. Phys. Lipids 17, 251-263

Stockton,G.W., Polnaszek,C.F., Tullock,A.P., Hasan,F. and Smith,I.C.P. (1976) Biochem. 15, 954-966

Takagi,M., Azuma,K. and Kishimoto,U. (1965) Ann. Rep. Biol. Works. Fae. Sci. Osaka University 13, 107-110

Taylor,J. and Haydon,D.A. (1966) Disc. Far. Soc 42,51-59

Tien,H.T., Carbone,S. and Dawidowicz,E.A. (1966) Nature 212, 718-719

Tien,H.T. and Dawidowicz,E.A. (1966) J. Col. lnt. Sci 22, 438-453

Tien,H.T. and Diana,A.L. (1967) Nature 215, 1199-1200

Turner,G.L. and Oldfield,E. (1979) Nature 277, 669-670

246

Van Deenen,L.L.M., Demel,R.A., Geurts van Kessel,W.S.M., Kamp,H.H., Reolofsen,B., Verkleij,A.J., Wirtz,K.W.A. and Zwaal,R.F.A. (1976) in The Structural Basis of Membrane Function (Hatefi,Y. and Djadavi-0haniance,L., eds.), pp 21-38 Accedemic Press, New York

Van den Berg,H.J. (1965) J. Mol. Biol. 12, 290-291

Van Zutphen,H. and Van Deenen,L.L.M. (1967) Chem. Phys. Lipids 1, 389-

Waldbillig,R. and Szabo,G. (1978) Nature 272, 839-840

Warren,G.B., Houslay,M.D., Metcalfe,J.C. and Birdsall,N.J.M. (1975) Nature 255, 684-687

Washburn,E.W. ed. (1929) International Critical Tables V6 pp 142-144. (McGraw Hill Book Company).

White,S.H. ( 1970) Biophys. J . l 0 , l 1 2 7 - 11 48

White, S. H. ( 1973) Biochim. Biophys. Acta 323, 343-350

White, S. H. (1974) Biochim. Biophys.Acta 356, 8-16

~Jhite,S.H. (1975) Biophys. J. 15, 95-117

White,S.H. ( 1977) annals N.Y. Acad. Sci. 303, 243-265

White, S.H. (1978) Biophys. J. 23, 337-347

White,S.H., King,G.I. and Cain,J.E. (1981) Nature 290, 161-163

White,S.H.,Petersen,D.C.,Simon,S. and Yafuso,M. (1976) Biophys. J. 16, 481-481

White,S.H. and Thompson,T.E. (1973) Biochim. Biophys. Acta 223, 7-22

Wobschall ,D. (1971) J. Col. Int. Sci. 36, 385-396

Wolfe,J. (1979) PhD thesis, Australian National University

Wolfe,J. and Steponkus, (1981) Biochim. Biophys. Acta 643, 663-668

Worcester,D.L. and Franks,N.P. (1976) J. Mol. Biol. 100, 359-378

247

PUBLICATIONS

During the course of this project some of the material presented in this

thesis has been published in the scientific literature. These were as

follows:

'The Molecular Organisation of Bimolecular Lipid Membranes: The effect

of Cholesterol Inclusion' by R.G. Ashcroft, H.G.L. Coster, D.R. Laver

and J.R. Smith (submitted for publication to Biochimica Biophysica

Acta.)

'The Molecular Basis of Anaesthesia' by H.G.L.

J.R. Smith. In Bioelectrochemistry. Edited by H.

Plenum Press. New York, N.Y. 1980.

Coster, D.R. Laver, and

Keyzer and F. Gutmann.

'Effect of D20/H20 Replacement on the Dielectric Structure of Lipid

Bilayer Membranes' by H.G.L. Coster, D.R. Laver and B.P. Schoenborn.

Biochimica Biophysica Acta (1982) 686, 141-143