• Aucun résultat trouvé

The colloidal nature of phospholipid monolayers

N/A
N/A
Protected

Academic year: 2021

Partager "The colloidal nature of phospholipid monolayers"

Copied!
10
0
0

Texte intégral

(1)

HAL Id: jpa-00210487

https://hal.archives-ouvertes.fr/jpa-00210487

Submitted on 1 Jan 1987

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

The colloidal nature of phospholipid monolayers

A. Miller, C.A. Helm, H. Möhwald

To cite this version:

A. Miller, C.A. Helm, H. Möhwald. The colloidal nature of phospholipid monolayers. Journal de

Physique, 1987, 48 (4), pp.693-701. �10.1051/jphys:01987004804069300�. �jpa-00210487�

(2)

The colloidal nature of phospholipid monolayers

A. Miller, C. A. Helm and H. Möhwald

TU Munich, Physics Department E22 (Biophysics), D-8046 Garching, F.R.G.

(Requ le 1er septembre 1986, révisé le 11 décembre, accepti le 18 d6cembre 1986)

Résumé.

2014

On étudie des monocouches de phospholipides et d’un acide gras fluoré par des mesures de

potentiels et de pressions de surface, et aussi en observant par microscopie de fluorescence les mouvements de domaines sous l’ effet de champs électriques inhomogènes. Les résultats montrent systématiquement que les forces électrostatiques à longue portée sont produites par les dipôles situés sur les esters glycériques et non pas par les têtes polaires de la monocouche. Les domaines lipidiques solides observés en coexistence avec une

phase fluide montrent un excès de densité de dipôles qui varie entre 50 à 200 mD/nm 2 avec des signes différents

pour les phospholipides et les acides gras. Ceci explique les directions opposées des forces qui s’exercent sur les domaines lipidiques placés dans les champs électriques. Les énergies par domaine pour ces forces sont de l’ordre de 104 kT. Nous discutons l’origine moléculaire de ces moments dipolaires de surface.

Abstract.

2014

Monolayers of different phospholipids and of a fluorinated fatty acid are studied by surface

pressure and potential measurements and by fluorescence microscopic observation of domain movement in

inhomogeneous electric fields. The data consistently show that long range electrostatic forces result from molecular dipole moments of the glycerin ester and not from the head group region of the monolayer. In coexistence with fluid phase solid lipid domains exhibit excess dipole density between 50 mD/nm2 and 200 mD/nm2 with different signs for phospholipids and fluorinated fatty acids. The latter fact explains the opposite direction of forces on lipid domains in electric fields. Energies per domain corresponding to these

forces were calculated to about 104 kT. The molecular origin of the surface dipole moments is discussed.

Classification

Physics Abstracts

87.20

-

68.10

-

64.70M

1. Introduction.

Lipid monolayers at the air/water interface have

encountered increasing interest as models of biologi-

cal membranes [1, 2], but also as models and precursors of well-structured organic films promising

many technical applications [3]. Beyond that they

also exhibit very interesting physical properties as

two-dimensional and interfacial systems. Their struc-

ture is determined from the amphiphilic character of

the constituent molecules causing a preferential alignment in that the hydrophilic parts point towards

the water surface, the hydrophobic parts towards the air. As the molecules possess a dipole moment the monolayer can be considered as an array of dipoles

each with a component perpendicular to the water

surface [4]. This gives rise to long range electrostatic forces and thus to very peculiar condensation

phenomena and formation of periodic superstruc-

tures [5].

Whereas the above features are principally intellig-

ible there rises the central question inhowfar they

dominate other forces such as van-der-Waals attrac- tion and line tension between different phases. To

answer this, one has to determine quantitatively the dipole moments in different monolayer phases and

relate these to molecular and supramolecular struc-

tures. This is done within this work by combining

classical surface potential measurements with fluorescence microscopic observations of lipid

domain movement in inhomogeneous electric fields.

The data are interpreted in conjunction with simple

model calculations showing consistently that the

essential contributions to the surface dipole moment

result from a polarization close to the hydrophobic

membrane region for charged and uncharged phosp- holipids as well as for partly fluorinated fatty acids.

2. Materials and methods.

The lipids L-a-dilaurylphosphatidylethanolamine (DLPE), L-a-dimyristoylphosphatidylethanolamine (DMPE) and L-a-dimyristoylphosphatidic acid (DMPA) (Sigma) were used without further

purification. The ethoxylated perfluorinated car-

bonic acid (FCS), C6F7(CH2hO(CH2h-COOH, was

a gift of Drs. U. Scheunemann and W. Interthal, Hoechst, Frankfurt, BRD. It was chromatographi-

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01987004804069300

(3)

694

cally pure. The dye probe dipalmitoyl-nitrobenzo- oxadiazol-phosphatidylethanolamine (DP-NBD-PE) (Avanti) was previously shown to be insoluble in solid lipid domains, but is soluble in the fluid phase.

The water used was deionized and filtered using a

Milli-Q system. pH was established using NaOH (Fluka, p.a. grade). Contamination by divalent ions

was prevented by adding 10-5 M EDTA (ethylenediamine tetra-acetic acid, sodium salt, Sigma) to the subphase. All experiments were performed at 20 °C. The phospholipid/dye or fatty acid/dye mixtures were spread from a chlo-

roform/methanole (9 : 1) mixed solvent.

Surface potential measurements were performed

in the laboratory of Dr. D. Mobius, Gottingen with

the vibrating capacitance technique (6, 7). A con- ducting circular plate (diameter 1 cm) periodically changes its distance d from the surface (frequency

420 Hz, d ~ 0.2 mm, amplitude 0.1 mm). The cur-

rent towards the counter electrode in the subphase is compensated, and the voltage necessary for this

compared with that for the clean water surface is the surface potential. The accuracy of potential measure-

ments was better than 10 mV. In the experiments

the molecular area could be varied continuously

while simultaneously measuring potential and sur-

face pressure.

The film balance and fluorescence microscope

have been described previously [8]. Of special rele-

vance for this study is the fact that the surface texture was observed via an objective lens in the

bottom of the film balance. Thus manipulations

could be performed by arranging additional tools above the surface. We made use of this by moving an

electrode above the surface. The electrode consists of the metal tip of a syringe, having a tip radius of

about 5 iLm. It can be moved in three orthogonal

directions by a micromanipulator (Bachofer) to an

accuracy of 5 Rm. The electrode can be charged to yield potentials of up to ± 100 V with respect to the bulk water that is grounded via the objective lens in

the trough bottom. On charging the electrode we

observe a distortion of the surface geometry due to elevating the water level. However, the findings reported are not due to gravitational effects, as the

latter would be independent of the sign of the

electric field. Yet we observe a change in the sign of

force on changing the field direction.

3. Experimental results.

Figure 1 shows surface pressure and surface potential

as a function of molecular area for the uncharged phospholipids DLPE (Fig. la) and DMPE (Fig.1b).

Additionally given is also the molecular dipole

moment u derived- as described in chapter 4. Most pronounced in the pressure/area isotherms is the change into a nearly horizontal slope on increasing

the pressure above a critical value 7T C. This is known

Fig. 1.

-

Surface pressure 7T’, surface potential AV and

vertical dipole moment A

=

Eo . OV A as a function of molecular area A for DLPE (Fig. la) and DMPE (Fig. 1b) monolayers at the air/water interface. In the phase coexist-

ence region in figure 1b between Af

=

69 A2lmolecule and

At

=

42 A2/molecule the fraction f of fluid phase may be assumed to depend linearly on A as also sketched. The dotted non-horizontal line corresponds to an extrapolation

of thè JL (A) function if the coexistence range would extend towards f

=

0 without changing the nature of the

solid phase. (T = 21 °C, 10-5 M EDTA, no buffer, i.e.

pH 5.5).

to correspond to the onset of a fluid/gel phase

transition which, due to the shorter chain length

occurs at a much higher pressure for DLPE

(irc

=

33 mN/m) than for DMPE (7T c

=

5 mN/m).

Comparing surface pressure and surface potential as

a function of molecular area there is obviously a good correspondence :

For pressures below 1 mN/m and molecular areas

above 90 Å 2/molecule there are large potential fluc-

tuations due to extreme surface heterogeneities. As

observed in fluorescence microscopic studies [9]

there is a coexistence of large domains of high and

low lipid density (« gas/fluid» transition). These heterogeneities disappear on increasing the pressure to establish the fluid lipid phase. Then the potential

increases towards a defined and reproducible value.

For pressures between 1 mN/m and 7T c the potential

(4)

increases continuously with pressure as qualitatively expected for an increase in molecular density. The flattening of the AV versus area isotherm for

pressures slightly above 7T c will be explained as due

to the transition to a gel phase exhibiting a smaller

molecular dipole moment. Increasing the surface

pressure beyond the plateau region causes the poten-

tial to rise rather steeply with increasing density.

This will be discussed as due to the onset of another

phase transition of the gel phase.

The above features are qualitatively identical for

the two lipids, the diference being that due to the

extensive fluid phase region of DLPE this lipid is

more suitable for studying this phase whereas DMPE, showing the pronounced plateau region best

serves to study the coexistence region.

Qualitatively similar behaviour of surface pressure and potential is obtained for monolayers of the charged lipid DMPA, displayed in figure 2 for

different subphases. Abrupt and not well defined

potential changes may appear for very low pressures

( 1 mN/m, Fig. 2a) followed by a gradual potential

increase with density in the fluid phase. In the

Fig. 2.

-

Surface pressure iT, surface potential AV and

vertical dipole moment u calculated as described in

chapter 4.1 for a DMPA monolayer at pH 5.5,

1 mM NaCI (Fig. 2a), at pH 11, 1 mM NaCI (Fig. 2b) and

at pH 11, 10 mM NaCI (Fig. 2c). (T

=

20 °C, 10-5 M EDTA).

coexistence region above iTc the potential is nearly

flat and at the end of the coexistence phase the potential exhibits another steep rise (Figs. 2b, c).

The fluorinated fatty acid FCS studied exhibits a

break in the slope of the pressure/area isotherm

analogously to that corresponding to the fluid/gel phase transition of phospholipids. In accordance

with this are also similar features of the OV versus area isotherm. The main difference between FCS and phospholipids resides in the fact that changes in

the absolute values of the potential are about a

factor of two larger for FCS and, what is most important, have opposite signs for the two classes of lipids (Fig. 3).

Fig. 3.

-

Surface pressure 7r, surface potential AV and

effective vertical dipole moment u

=

Eo AV. A for the fluorinated fatty acid FCS. (T

=

20 °C,10-5 M EDTA, no buffer, i.e. pH 5.5).

The difference in sign of the surface potential is

also reflected in the response to an electric field as

visualized in figure 4. Increasing the surface pressure above iTc gel phase lipid domains are formed. These

domains are observed as dark areas in the fluor-

escence micrographs because the dye probes used

are more soluble in fluid compared to gel phase lipid [9, 10]. Placing a tip electrode above the water

surface one expects a field geometry as sketched at

the bottom of figure 4. Upon charging the electrode gel phase phospholipid domains are pulled under the

electrode for negative potential (Fig. 4a) and repel-

led for a positive potential (Fig. 4b). This holds for

the neutral phospholipids DMPE and DLPE as well

as for the charged lipid DMPA irrespective of its degree of dissociation (measured for 0.2 a 1.5). The electrostatic force is directed oppositely

for gel phase domains of FCS (Fig. 4c).

4. Analysis and discussion.

4.1 EVALUATION OF SURFACE POTENTIAL DATA.

4.1.1 Homogeneous monolayer.

-

In case of an

uncharged monolayer the surface potential AV can

(5)

696

Fig. 4.

-

Movement of crystalline (dark) lipid domains

under a tip electrode about 20 um above the water surface. Figure 4a : DMPA, pH 11,

w

= 18 mN/m, A

=

63 A2lmolecule, electrode potential - 100 V ; Figure 4b : DMPA, pH 11,

7T =

20.5 mN/m, A

=

56 A2lmolecule,

electrode potential + 100 V ; Figure 4c : FCS, pH 5.5,

7T =

29 mN/m, A

=

33 A2lmolecule, electrode potential

-100 V ; Figure 4d : Sketch of geometry assumed for calculation of forces : Circular discs at the water surface

(Z

=

0) move horizontally under the influence of a charge

q smeared over the surface of a- sphere with radius rK. (T

=

20 °C, 10-5 M EDTA).

be related to the density p of dipoles oriented perpendicular to the surface according to

( Eo

=

8.9 x 10-12 C2 /m2 N).

As p is proportional to the molecular density one can

convert it into a dipole moment per molecule u obtaining [6, 11]

ii determined from equation (2) is given in figure 1.

One realizes u

=

const. for surface pressures

betvyeen 1 mN/m and 7Tc where the monolayer is in a single homogeneous (fluid) phase. This also means

that the molecular orientation does not change while varying the surface pressure within this range. One obtains almost identical values of A for the two lipids showing that u does not depend on chain length.

If the monolayer is charged one has additionally to

take into account the potential drop qio in going from charged interface to bulk solution yielding

qf 0 can be calculated from Gouy-Chapman theory [12, 13] taking into account that the degree of

dissociation of lipids in turn depends on qio. The

surface charge density a is related to the density n of

monovalent ions in the subphase according to

(k

=

Boltzmann constant, T

=

absolute tempera- ture, e

=

elementary charge).

Neglecting binding of ions other than protons the

mass action law for the dissociation of phospholipid

head groups yields

([H+ ]

=

bulk proton concentration).

The intrinsic dissociation constants Kl and K2 for

dissociation of two protons from the phosphatidic

acid head groups were assumed to be 60 M-1 [13, 14] and 108 M-1, respectively. The latter value is not known for phospholipids but deviates by less

than an order of magnitude from the corresponding

value known for phosphatidic acid. Calculations

varying this value did not show a significant in-

fluence.

Connecting equations (4) and (5) we obtain a non- analytical equation that can be solved numerically

for t/10 and then also obtain the surface charge density and the degree of dissociation [14]. As a

result of the calculation figure 5 gives the degree of

dissociation and the surface potential as a function of molecular area A for three different ionic conditions where experiments with DMPA were performed.

Most important for the data evaluation is the poten- tial qio that amounted to between - 175 mV and

-

275 mV. As the measured AV is some hundred mV (Fig. 2) but positive, one may conclude that the

dipolar contribution to AV exceeds t/10 and is of

comparable magnitude.

One also realizes that the area dependence of t/10 can be linearized over a limited range, e.g.

between 60 and 120 A2/molecule, 4io = a + bA, with parameters a, b given in the first columns of table I.

Experimentally these parameters can be determined

by assuming also a linear relation 4/0 = a’ + b’A

and adjusting a’ and b’ in such a way that u obtained

from equation (3) is independent of molecular area

for the monolayer in the fluid phase. The latter condition is justified from the observation for the

uncharged monolayer where u does not depend on

pressure while being in the fluid phase. Comparison

of parameters derived from the experiment with theoretically calculated values in table I shows a

reasonable agreement. This also justifies the assump- tions used in the calculations. Among these the most

critical ones may be the neglect of sodium ion

binding and assuming a value of K2 =108 M-1 1 for

dissociation of the second proton of DMPA. Varying

(6)

Fig. 5.

-

Degree of dissociation and potential qio calcu-

lated as described in chapter 4.1.1 for the DMPA mono-

layers of figure 2. A : pH 5.5, 1 mM NaCI ; B : pH 11, 1 mM NaCI; C : pH 11, 10 mM NaCI. Parameters : T

=

20 °C, K,

=

60 M-I, K2 =108 M-I.

Table I.

-

Parameters a, b derived from the numerical calculation of g/o

=

a + bA and parameters a’, b’

obtained from surface potential data as described in the

text.

the corresponding parameters would yield an even

better agreement between theory and experiment,

but this would result in too many adjustable par- ameters without getting a better physical insight.

The latter statement also holds concerning an improvement of the constancy in 4 while the mono- layer is in the fluid phase (see Fig. 2). Nevertheless,

one realizes that the analysis yielded an almost

constant u and that values determined for the fluid

phase of phospholipids are of comparable mag- nitude. They range between 0.64 D and 0.77 D per molecule for DMPA in different ionization states and are 0.52 D and 0.54 D for DLPE and DMPE, respectively. This again indicates the consistency of

the analysis and also that the type of head group does not grossly alter the dipole moment.

The fluorinated fatty acid was studied as an example of an amphiphilic compound with surface

potential sign opposite to that of phospholipids and

of fatty acids containing exclusively hydrocarbon

chains. However, because there is presently much

less information available on this class of compounds

an analysis as extensive as above is not performed.

Hence we do not discriminate between molecular

charge and dipole moment and in figure 3 give merely an effective dipole moment JLeff defined as

This value describes the behaviour of the monolayer

in electric fields and also allows conclusions on

pressure induced orientations of molecules in the fluid lipid phase. ueff as a function of molecular area

in figure 3 is nonmonotonous even for the fluid

phase. The potential contribution gio due to partial charging of the carboxylic acid head group is negative

and increasing in absolute value with molecular

density. This causes u I

=

I geff I - C (A) with

C (A ) being positive and larger for large areas A due

to the .po (A) relationship which is less steep than

inversely proportional. Thus g increases on increas- ing the pressure towards iT indicating a change of

molecular orientation with surface pressure. This

can be verified, e.g. by subtracting an estimated

value of .po

= -

200 mV from the measured value AV and then determining g from equation (1). For a

more refined analysis, however, more refined data

at different subphase conditions would have to be collected.

4.1.2 Inhomogeneous monolayer.

-

As monolayers undergo several first order phase transitions one

generally expects coexistence of phases with different

surface potential for certain pressure regimes. To

take this into account in the analysis one may

envisage two extreme situations :

(i) The lateral dimensions of coexisting lipid

domains are of comparable size or larger than the air

electrode area. The surface potential may then

assume any value between those of the homogeneous coexisting phases depending on the fraction of these

phases under the electrode. This condition seems to hold .for pressures below 1 mN/m where strong potential fluctuations are encountered (e.g. Fig. 1, Fig. 2a). These fluctuations cannot be assigned more specifically to definite molecular areas and will

therefore not be considered further.

(ii) The more interesting case is that where

domain dimensions are much smaller than the elec- trode size. This holds for coexistence between fluid and gel phase domains for pressures above 1T, where

gel phase domains with sizes below 100 f.Lm are

(7)

698

observed [9, 10]. For this case the data analysis is

refined the following way :

Concerning the measurement, the two phases can

be represented by two parallel capacitors C, and C2.with magnitudes proportional to the correspond- ing areas Fl, F2 and by parallel voltage sources V1 and V2 according to their surface potentials (see Fig. 6). The external voltage AV necessary to com- pensate the current I

=

Il + 12 is measured as

surface potential. For I

=

0 one obtains

and

Hence

Fig. 6.

-

Equivalent electrical circuit to describe the surface potential AV of a monolayer with coexisting regions of potentials V1 and V2 and areas proportional to C1 and C2.

According to equation (8) AV is the area weighted

average of the surface potential of the coexisting phases. The potential can in turn be converted into

an average dipole moment tt that relates to the

molecular dipole moments ILl’ IL2 in the correspond- ing phases according to

In equation (9) f

=

N, + 2 N2 N1 + N2 N 2 is the fractional number (N2 ) of total molecules (N1 + N2) in phase 2, which we henceforth will consider the fluid

area.

The validity of equation (9) is most clearly demon-

strated by inspection of the coexistence region (above 7Tc) of the DMPE monolayer (Fig. la).

Assuming a coexistence between a

«

solid » phase of

molecular area 42 A2 and a fluid phase of area Ac corresponding to w,, f depends linearly on

molecular area as sketched also in figure la. Conse- quently according to equation (9) A depends linearly

on molecular area. This is experimentally observed

with DMPE for 0.4 f 1 and, although less pronounced, for the other monolayer systems for high enough liquid area fraction. Extrapolation of

the linear V versus A plot towards f

=

0 then yields

the dipole moments A in the

«

solid » phases given in

table II.

There are obviously strong deviations from lineari- ty on reducing the fraction f of fluid area below 0.4.

These can be ascribed to a structural change in the

«

solid » phase, as was deduced from Synchrotron X-

ray diffraction [15] and fluorescence microscopic

data [16]. Then it was also shown that the

«

solid »

phase coexisting with the fluid one for pressures only slightly above 7rc is one of slightly higher molecular

area than expected for the solid phase. Taking this

into account in the analysis would yield a correction

of data on dipole density given below by less than

2 % and is thus negligible.

4.2 DIPOLE DENSITY IN FLUID AND SOLID PHASES AND FORCES IN ELECTRIC FIELDS.

-

The last column in table II also gives an effective difference in dipole density determined with equation (6) and with the

Table II.

-

Dipole density p and differences in dipole densities for systems studied in this work.

(8)

model derived to analyse the heterogeneous mono- layer. This value is given since it is basically deter- mining solid phase domain movement in in- homogeneous electric fields discussed below.

Inspection of table II shows that dipole density

differences Ap amount to between 5 % and 20 % of

the absolute values of g. This means that errors in the determination of p can lead to errors in Ap of up to 50 %. On the other hand the signs of Ap are correct as they are given by the sign of the slope in the coexistence range of the AV versus area

diagram.

Values of p do not differ much for phospholipids

under various ionic conditions indicating again that they possess the same groups determining the dipole

moment.

The situation is entirely different for FCS where

the sign of p and Ap as well as the absolute magnitude of the dipole potential are distinguished

from that of phospholipids as well as of other fatty

acids [6] containing exclusively hydrocarbon chains.

This proves the dominant influence of the fluorocar- bon moiety.

Irrespective of molecular interpretation one may

use the data of table II to calculate forces on solid

lipid domains. Considering the tip electrode geomet- ry (Fig. 4 bottom) we calculated the electric field

assuming the water surface conducting [17, 18]. At

the interface the field has only a perpendicular

component E, and the force F on a dipole with

component P. is

F points into a direction r parallel to the surface with

r

=

0 being the position of maximum field strength

under the electrode. For a domain of radius ro the effective dipole moment Pz is calculated from the difference in dipole density according to

If h, the electrode distance from the water surface is much larger than ro, the maximum electrostatic energy gain (or loss) Gmax of a domain pulled under (or repelled from) the electrode is given by

The charge q of the electrode is related to the tip radius rt and the potential U by

To calculate Gmax we use the following parameters :

-

U = 100 V as applied in the experiment of figure 4

- rt

=

5 um determined from microscopic inspec-

tion of the tip

-

ro = 5 um as measured for the domains of

figure 4

-

åPeff

=

50 mDlnm2 as given for DMPE in

table II

-

h

=

20 um. This value was estimated from a

vertical movement of the electrode after touching

the surface. It is the uncertainty in this value that determines the error in Gmax which may be an overestimate by up to a factor of 3. With these parameters we obtain Gmax = 140 eV - 6 000 kT.

This clearly shows that it is possible to move solid

domains against thermal motion, and in fact it is

generally not thermal but convective motion which

frequently disturbs the experiment.

The above value of Gmax is about a factor of

30 smaller than that estimated previously [17]. This

is due to the fact that we now could use measured values of the dipole moment density which turned out to be considerably smaller than anticipated. On

the other hand one realizes that APeff differs only by

a factor of up to 10 for different lipids and enters linearly in the energy calculation. This means that the orders of magnitudes of the forces will be similar for different systems.

The above calculation showed that due to the colloidal nature of solid lipid domains, essentially being dipolar discs, reasonably high forces result in the electric fields applied. This also implies that we

do not have to discuss any other nonquantifiable

effects like trapping of charges in head group or

hydrocarbon tails to explain the observed forces. An

especially strong argument in favour of the above picture also results from a comparison of findings

with phospholipids and FCS: surface potential

measurements have revealed a different sign of the dipole moment difference between fluid and solid

phase for the two classes of compounds. As expected

within the above model also the forces on solid domains are of opposite orientation.

4.3 MOLECULAR INTERPRETATION OF SURFACE PO- TENTIAL DATA.

4.3.1 Homogeneous fluid phase.

-

Figure 7

sketches a presumed arrangement of the polar

groups of the three types of molecules compared in

this work at the air/water interface. Included by

arrows (from

« - »

to

«

+ ») are also local dipole

moments due to polarization or due to charge

accumulation at the interface. Since we want to discuss the experimental findings in view of the molecular structure we have to exclude the influence of the water structure at the interface that is changed by the presence of surfactants. This is imaginable,

because the surface potential of the free water

surface amounted to about - 500 mV (for Millipore water) in our-experiments and we basically measured

the difference from this potential.

(9)

700

Fig. 7.

-

Probable arrangement of DMPE and FCS at the air/water interface. Included by arrows are positions of largest dipole moment contributions. DMPA is oriented similar to DMPE but lacks the ethanolamine head group and the corresponding dipole.

The strongest argument against the dominance of the water structure results from a comparison of the

surface potential signs which are negative for the

fluorinated fatty acid and positive for fatty acids with hydrocarbon chains [6] as well as phospholipids.

Hence even for the same hydrophilic group opposite signs of the potential are measured excluding that

the water structure determines the potential.

Now considering the positive sign of the potential

for charged and uncharged phospholipids we deduce

from figure 7 that the charged phosphatidic acid and

the uncharged phosphatidylethanolamine head

groups would give a negative contribution, and the only positive contributions result from a polarization

of the carbonyl groups. This means that contribu- tions of the former groups are of minor relevance.

This conclusion is also valid for phosphatidylcholines

where the same sign of the surface potential was

measured [19] and is obvious comparing data and

molecular structure. Yet although there existed

surface potential data for phospholipids in the fluid phase qualitatively identical to ours the head group

region was usually assumed to be responsible for the potential and the origin of long range electrostatic forces [4, 5, 20]. That this is not the case has the following reasons :

-

Polar groups protruding into the water phase

are highly screened due to high conductivity and

dielectric constant of water.

-

Due to the large gradient of the dielectric

constant e across the interface dipoles in the sub-

phase are screened by image forces above the water level [21]. These image forces reduce long range interactions by more than 3 orders of magnitude but

are less effective considering short range interac- tions. This fact seems to explain why the influence of the head group charge on lipid phase transitions of vesicles [13] and -monolayers [14] is strong and

theoretically well describable.

The strong influence of the carbonyl groups has

already been pointed out by Paltauf et al. [19]

comparing phospholipids with and without these

groups. Using the interfacial dipole moment of

0.36 D for the C

=

0 bond [22] one may estimate that they are responsible for the total potential if they are tilted by an angle of 20° with respect to the surface. On the other hand one realizes that there

are also large in-plane moments which may give rise

to ferroelectric transitions [23].

Looking at the structure of FCS we see that the potential determining groups are at the interface

between hydro- and fluorocarbons. The electronega-

tive character of fluorine giving rise to this polariza-

tion has already been recognized previously [6, 24].

The fact that the potential determining group resides in the hydrophobic chain region also explains the

strong dependence of u on surface pressure of the fluid monolayer. On compression the chains are aligned to be more vertically oriented with respect to

the surface thus effecting a more negative potential.

On the other hand groups closer to the polar head

group region are less sensitive to any pressure

change as long as the lipid phase is preserved. This is supposedly valid for the carbonyl groups and ex-

plains the pressure independence of tt for fluid phospholipid monolayers.

At the end of this section we should also remark that the conclusion that dipolar contributions of the head group region to the surface potential are of

minor relevance is in accordance with the fact that u

measured for phospholipids with different head

groups does not differ drastically.

4.3.2 Fluidlsolid phase coexistence region.

-

A major contribution of this work results from the fact that experiments and analysis also considered the fluid/solid phase coexistence region. A molecular interpretation of corresponding dipole moment changes will, however, remain speculative, because

there effects of less than 10 % of the overall values

are discussed. This means that we can no longer

discard the influence of changes in the head group

region that may result from reorientations or from a

change in water structure affecting the screening of

forces.

We can still exclude the direct influence of the water polarization that may be different under a

fluid and a solid lipid phase. This may change the potential into one direction irrespective if the lipid is phospholipid or FCS. However, on solidifying the monolayer u changed in different directions for these classes of compounds.

On the other hand changes in the absolute values

of JL on increasing the solid area ratio occurred in a

similar way suggesting analogous mechanisms : In-

creasing the pressure above irc one first observes a more or less clearly pronounced linear decrease of JL

with area A and then a steep increase in 1L for areas

below which also the pressure increases. Yet one has

(10)

to be careful in drawing analogies too far, because

the two classes of compounds are vastly different.

Considering first the linear decrease in g with A

for phospholipids the most appealing idea is that the carbonyl group orientation changes in the solid

phase where the hydrocarbon chains are condensed

and almost vertical to the surface. Assuming a

decrease of the angle between bond orientation and surface from 20° to 3° would explain the observed values. The fluorocarbons in their condensed phase

are distinguished from phospholipids in that the limiting area per chain is larger (

~

30 A2 compared

to 20 A2 for hydrocarbon chains) and therefore the

carboxylic head group has a higher degree of

freedom. Assuming that in the solid phase the carboxylic head group, due to coulombic repulsion

with neighbouring groups would protrude further

into the water phase would cause a potential change

into a direction as measured for pressures slightly

above ’TT C.

At first glance one might expect that the conden- sation of fluorocarbons involving a chain alignment

also effects the chain contribution to the dipole

moment. This, however, would correspond to an

increase of u, in contrast to the findings. The

absence of this influence may be due to the fact that the chains are almost vertically oriented already in

the fluid phase near irc.

Another explanation of the linear decrease of tL

would be that screening of head group forces varies

due to a different water structure or water level in different phases. We cannot exclude this possibility

because it is also consistent with the findings : the

head group contributions would be changed in

absolute values not in sign. Nevertheless, this picture

is highly speculative and synchrotron X-ray experi-

ments are under way to conclude on this.

There is presently no consistent picture to con-

clude on the increase in A with increasing surface

pressure starting roughly at the end of the linear, nearly horizontal, region of the pressure/area

isotherm. From recent synchrotron X-ray diffraction

experiments [15] we know that this corresponds to

the onset of another phase transition where the

positional coherence length increases from about 20 to 60 lattice spacings at high pressures. One specula-

tion on the transition would be that the monolayer in

the condensed phase is further removed from the water surface thus increasing the influence of chain

polarization in the interfacial area. This would give a positive contribution to it for phospholipids and a negative one for FCS as observed.

5. Acknowledgments.

We thank Dr. D. Mobius for allowing us to use his equipment for surface potential measurements and V. Vogel for performing the potential measurements with FCS. This work was supported by the Deutsche

Forschungsgemeinschaft.

References

[1] PHILLIPS, M. C. and CHAPMAN, D., Biochim. Bioph-

ys. Acta 163 (1968) 301.

[2] SACKMAN, E., Ber. Bunsenges. Phys. Chem. 82 (1978) 891.

[3] ROBERTS, G. G., Adv. Phys. 34 (1985) 475.

[4] ANDELMAN, D., BROCHARD, F., DE GENNES, P. G.

and JOANNY, J. F., Comptes Rendus 301 (1985)

675.

[5] FISCHER, A., LÖSCHE, M., MÖHWALD, H. and SACKMANN, E., J. Physique Lett. 45 (1984) L-

785.

[6] GAINES, G. L., Insoluble Monolayers at Liquid-Gas Interfaces, (Interscience, N.Y.) 1966.

[7] KUHN, H., MÖBIUS, D. and BÜCHER, H. in Physical

Methods of Chemistry eds. A. Weisberger and

B. Rossiter, Vol.1 (3B) (Wiley, N.Y.) 1972.

[3] LÖSCHE, M. and MÖHWALD, H., Rev. Sci., Instrum..

5 (1984) 1968.

[9] LÖSCHE, M., SACKMANN, E. and MÖHWALD, H., Ber. Bunsenges. Phys. Chem. 87 (1983) 848.

[10] WEISS, R. M. and McCONNELL, H. M., Nature 310 (1984) 5972.

[11] DAVIES, J. T., Roc. Roy. Soc. A 208 (1951) 224.

[12] McLAUGHLIN, S. A., in Current Topics in Membrane

Transport, Vol. 9 (1977) p. 71-144.

[13] TRÄUBLE, H., TEUBNER, M., WOOLLEY, P. and EIBL, H.-J., Biophys. Chem. 4 (1976) 319.

[14] HELM, C. A., LAXHUBER, L. A., LÖSCHE, M. and MÖHWALD, H., Coll. Polym. Sci., 264 (1986) 46.

[15] KJAER, K., ALS-NIELSEN, J., HELM, C. A., .

LAXHUBER, L. A. and MÖHWALD, H., subm. to Phys. Rev. Lett.

[16] LÖSCHE, M., THESIS, TU Munich 1986.

[17] MILLER, A. and MÖHWALD, H., Europhys. Lett. 2 (1986) 67.

[18] MILLER, A., THESIS, TU Munich, 1986.

[19] PALTAUF, F., HAUSER, H. and Phillips, M. C.,

Biochim. Biophys. Acta 249 (1971) 539.

[20] KELLER, D. J., McCONNELL, H. M. and MOY, V. T., J. Phys. Chem. 90 (1986) 2311.

[21] FLEWELLING, R. F. and HUBBELL, W. L. Biophys. J.

49 (1986) 541.

[22] FORT, T. and ALEXANDER, A. E., J. Coll. Science 14

(1959) 190.

[23] HECKL, W. M. and MÖHWALD, H., Ber. Bunsenges., Phys. Chem. 90 (1986) 1159.

[24] BERNETT, M. K. and ZISMAN, W. A., J. Chem.

Phys. 7 (1963) 1534.

Références

Documents relatifs

(« Novartis ») pour la commercialisation canadienne par Valeo de deux thérapies contre l’asthme : Enerzair MD Breezhaler MD (indacatérol [sous forme

The second order projection methods for the INS system with variable fluid density have been studied in [22], which used a combination of midpoint method and Crank–Nicolson method

Le terme français de « réadaptation » en est la traduction et convient donc pour désigner ce processus d’entraînement d’habiletés pour que la personne souffrant d’une

countries should contribute to and support the national strategy to achieve the MDGs. National efforts to increase outreach and access to health services should aim at

Prédiction : construire un modèle à partir des individus complètement renseignés et l’utiliser pour prédire les données correspondant aux données manquantes. •

Abstract: We experimentally demonstrated that the lifetime of quantum emitters with strongly mixed electric dipole ED and magnetic dipole MD transitions can directly probe the

Atomic force microscopy imaging and force spectroscopy were used to directly observe the adsorption of casein particles on supported phospholipid bilayers with controlled

Dissolvez la(les) pastille(s) en mettant le tube plusieurs fois à l'envers.. Placez la cuvette réservée à l’échantillon dans la chambre