Study of the Fusion Process between Solidand Soft-Supported

Aug 20, 1999 - In this context, the investigation of adhesion and fusion of vesicles with each other as well as with planar biomembranes (1,2) is of p...
1 downloads 0 Views 2MB Size
Chapter 15

Study of the Fusion Process between Solid­ -and Soft-Supported Phospholipid Bilayers with the Surface Forces Apparatus 1

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

Markus Seitz, Chad K. Park, Joyce Y. Wong , and Jacob N. Israelachvili Department of Chemical Engineering, University of California, Santa Barbara, C A 93106

Introduction. L i p i d vesicles and planar bilayers have been used extensively to model living cells i n the biophysical study of molecular mechanisms and fundamental interaction forces. In this context, the investigation of adhesion and fusion of vesicles with each other as w e l l as with planar biomembranes (1,2) is o f particular interest. F o r structural investigations, biomembranes generally have to be 'held i n place', such as by transfer onto solid substrates (3-5). Various preparation techniques have been suggested for mono- and multilayer systems, such as the Langmuir-Blodgett technique (4) as well as self assembly processes on planar surfaces (3,6,7) or on small glass beads (8). Several methods have been developed that allow the measurement of interaction profiles with a resolution on the sub-nanometer scale (9). One of them is the osmotic pressure (or osmotic stress) technique (10), by which the first quantitative data for the short-range repulsive interaction of lipid bilayers i n aqueous solution could be obtained (11,12). However, it is restricted to the repulsive contributions of the inter­ action profile. Another important method by which mechanical properties as well as the adhesion of vesicles in contact can be investigated is the "pipette aspiration tech­ nique" (13-15). In recent years, reflection microscopy techniques have been deve­ loped to monitor the distance between bilayer vesicles and modified surfaces which could also used to investigate interaction forces between bilayers and different types of substrates (16-18). The complete force-distance profile including adhesive interactions can be directly measured with the surface forces apparatus ( S F A , 19,20), by which most fundamental interactions between surfaces i n aqueous as w e l l as non-aqueous solvents have been identified and quantified such as van der Waals, electrostatic as well as repulsive hydration forces or attractive hydrophobic interactions (19,21-26). In a unique way, the S F A also allows the observation of the membrane fusion pro­ cess directly (in 'real time') with a resolution on the molecular level, at least i n one dimension. In this contribution we w i l l give a brief summary of earlier fundamental results of bilayer fusion studies employing the S F A technique as well as a brief overview of the most recent developments in this area. 1

Current address: Department of Biomedical Engineering, Boston University, Boston, MA 02215.

© 1999 American Chemical Society

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

215

216

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

Study of Normal Interbilayer Forces with the Surface Forces Apparatus. The S F A used i n the surface forces measurements has been described extensively (23,27,28). Briefly, the distance of two cylindrically curved, molecularly smooth m i c a surfaces, of radii R i and R2, is visualized using the optical interference technique described previously (26,29). Fringes of equal chromatic order ( F E C O ) are produced when white light is passed through the apposing sample surfaces, and reflect the shape of the surface contact region. Thus, the distance of the two mica surfaces from a separation of several micrometers (Figure L A ) to close contact (Figure l . B ) of the surfaces can be followed to within 1-2 A, while at the same time the topography of the contact region can be imaged. One of the surfaces is mounted on a spring with spring constant K . From the measured deflection A D of the spring, the force F between the surfaces can be determined. The corresponding energy per unit area E between two flat surfaces is simply related to the force between the curved surfaces by the so-called Derjaguin approximation, E = F/27CR, where R is the geometric mean radius R = (Ri-R2)" (21). For repulsive forces, the surfaces are simply pushed together and the spring deflection A D is measured optically, from which the force is obtained as F = K A D . In the case of attractive forces, the system becomes unstable i f the gradient of the force 3F/3D exceeds the spring constant K , and the surfaces jump to the next stable point of the interaction profile (30). Under very large compressive forces the two initially curved surfaces flatten elastically as the glue supporting the mica sheets becomes compressed. This deformation can be directly observed from the shapes of the fringes which represents a cross section of the contact zone. When one observes flattening, the Derjaguin approximation no longer applies. Instead of the energy per surface area, one can now determine the mean pressure from P a n = F / A . However, the maximum pressure occurs at the center of the contact area for repulsive interactions. If the surfaces deform elastically and do not mutually adhere, the "Hertz theory" (31,32) predicts that P = 0 at the edges and that it rises to a maximum at the center, where P x =1.5 Pmean- The elastic deformation of adhering surfaces is more complicated, and the resulting stress in the system is generally highest at the edges of the contact region (32,33). 2

m e

m a

W h e n two lipid bilayer membranes have come into close, flattened contact, there is still one layer of hydration water separating the lipid head groups (Figure L C ) . A s a certain energy barrier is overcome (e.g. by applying a certain pressure), either the center or the outer edge, i.e. the regions of the highest effective pressure, of the ellipsoidal contact spot can "break" through, which can be followed by the F E C O technique. These regions then slowly spread out towards the side or towards the center of the contact until, eventually, the whole interference pattern is shifted by the same amount to smaller wavelengths, i.e. smaller distances (26). O n a molecular level, this relates to the hydration water layer as well as the two outer membrane leaflets being squeezed out (Figure l . D ) . Thus, the total change i n the mica-mica separation distance during the break­ through corresponds to the removal of one full bilayer including one layer of hydra­ tion water after which one bilayer membrane still remains i n the contact region (22,24,26). These observations have been related to a so-called "hemifusion" (Figure 1 .E) which is assumed to be an intermediate step in the process of full vesicle fusion (Figure l . F , 1,34). Forces and Fusion between Solid-Supported Phospholipid Bilayers. T o date, almost all S F A measurements on interacting lipid bilayer membranes have made use of solid (muscovite) mica substrates. This material provides an atomically flat surface so that the upper (distal) layer of the substrate-supported membrane can exist i n an almost undisturbed fluid state, and the adhesion and fusion of model membranes could be followed directly with these systems (22-24). In general, it was found that adhesion between two bilayers can be progressively increased by up to

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

217

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

SFA IMAGES

VESICLES

B "Contact"

D Breakthrough

Figure 1. Schematic representation of hemifusion and fusion between bilayer mem­ branes: Two mica-supported membranes in the S F A (left) represent the contact region during approach, contact and fusion of two free lipid vesicles (right).

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

218

two orders of magnitude i f they are stressed to expose more hydrophobic groups. The most important force leading to direct (hemi-)fusion is the hydrophobic attrac­ tion acting between the (exposed) hydrophobic interiors of the solid-supported bilayers (in most cases, the inner layer which is i n contact with the mica support consists of a crystalline dipalmitoylphosphatidylethanlolamine monolayer, D P P E ) . However, the occurrence of fusion is not simply related to the strength o f the attractive interbilayer forces but also to the internal bilayer stresses (intrabilayer forces). Fusion can be triggered spontaneously between two repelling bilayers when they are still at a finite distance from each other without their having to first overcome the short-range repulsive forces, e.g. the steric hydration barrier. L o c a ­ lized molecular rearrangements allow this to happen by a process o f locally bypassing these forces via a simple breakthrough mechanism. Hereby, fusion starts locally at the weakest point or most highly stressed point of two apposing bilayers (26). W i t h 'fully developed' bilayers no fusion is observed up to applied pressures of 150 atm. Fusion takes place between 'depleted bilayers', that is, when the solution is diluted below the C M C . Depletion also changes the force law between bilayers (26,34). In addition, hemifusion could be i n systems with packing defects, e.g. after cooling below the fluid-gel phase transition temperature, or for natural lipid mixtures such as "eggPC" (22). Nevertheless, high pressures of up to P = 50 100 atm were needed to achieve hemifusion (22,26). The second stage of the process depicted in Figure l . F , from hemifusion to full (or complete) fusion, involves the total removal of the central bilayer. This would be characterized i n S F A measurements by a second critical breakthrough step. O n mica-supported systems it could be only observed so far between hydrophilic monolayers i n humid air or i n water-wet organic solvents, when the monolayer binding to the mica surfaces becomes weak (34,35). The hydrophobic interaction between the inner layers o f apposing l i p i d membranes could also be established as the ultimate driving force during the calcium-induced bilayer fusion of mixed bilayers systems. D i m y r i s t o y l - p h o s phatidylcholine/dimyristoyl-phosphatidylglycerol ( D M P C / D M P G ) or dilauroylphosphatidylcholine/dilauroyl-phosphatidylglycerol ( D L P C / D L P G ) , i.e. mixtures of zwitterionic P C and anionic P G lipids were investigated (25). It was found that the critical role of calcium is in the induction of phase-separated domains of different lateral stresses followed by 'contraction' of the c a l c i u m / P G - r i c h domains and 'expansion' of the PC-rich domains, with subsequent attraction and fusion o f appo­ sed layers. This model was confirmed by the finding that pure P G (or P C ) bilayers do not fuse i n the presence of calcium as the entire lipid layers become uniformly stressed upon calcium binding and no additional area is exposed to the aqueous phase. In continuation of this earlier work, recently, the hemifusion of zwitterionic b i ­ layers on solid mica substrates was further studied (Park, C . K . ; Israelachvili, J. N . , to be published). The following parameters were found to have an effect on the fu­ sion event: deposition density, the lipid shape, and the phase state of the lipids with­ in the supported membrane. W h e n lipids are deposited at sufficiently low densities, there is not enough coverage of the innermost D P P E layer. The exposure of these hydrophobic regions results i n a larger adhesion than is accountable by van der Waals forces alone and frequently leads to hemifusion. Another way of reducing coverage can be accomp­ lished with conically-shaped lipids. These are molecules where the headgroup area is smaller than that of the side groups. When such a lipid, Dilinoleoyl-phosphatidylethanolamine, D L i n P E , was deposited at reasonable coverages, an increased adhesion and propensity for fusion was observed. This may be due to the outermost layers' inefficient packing which exposes its own tail groups to the aqueous environ­ ment. Finally, by depositing near a liquid-expanded ( L E ) to liquid-condensed ( L C ) phase transition we again see a large increase adhesion and frequent fusion events.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

219

In this case, it is believed that the solid-like nature of the substrate induces the outer lipids to pack i n a more solid fashion. In doing so, the L E - L C phase boundaries may provide defects which facilitate hemifusion. T o determine the effect of lateral phase separation on the adhesion and fusion of D L P E bilayers was investigated (Park, C . K . ; Israelachvili, J. N . , to be published). Deposition densities corresponding to fluid phases and solid phases (42 and 55 A per molecule respectively) gave rise to stable systems. Forces measured here were comparable to that expected from D L V O theory. Further, the F E C O fringes d i d not show any deformations indicating hemifusion. If the lipid was deposited at very low densities, the resulting hemifusion was attributed to insufficient coverage of the pro­ ximal D P P E layer. This effect was slightly larger for the inverted phase forming lipid D L i n P E . When the deposition of the D L P E was done at temperatures near the fluidlike-solidlike phase transition, the layers were highly susceptible to hemifusion. Interestingly, fluorescence microscopy on this system showed the presence o f do­ mains. It is likely that the domains observed are providing defects or exposed edges which lowers the hemifusion energetic barrier. Essentially these effects are methods of (slightly) exposing the proximal D P P E layer to the bulk water. The resulting hemifusion is thought to be driven by strong attraction these hydrophobic groups have for each other.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

2

Soft-supported bilayers for the investigation of normal interaction forces. In recent years, the S F A was also used for the investigation o f the interactions i n more complex biological assemblies, such as the ligand-receptor system biotinstreptavidin (36-39). Whereas i n most of the cases discussed so far, the fluidity of the outer lipid layer was sufficient for the study of the interaction between simple as well as functionalized lipid membranes, the direct contact between the proximal l i ­ pid monolayer with the substrate poses a serious constraint i n many desired systems. The restriction of membrane mobility by the supporting mica substrate provides a significant difference compared to studies on bilayer vesicles or liquid-crystalline multilamellar lipid structures. Bilayer systems which are allowed to fluctuate freely around an average distance exhibit repulsive undulation forces. This effect could not be studied with the S F A technique so far. Although i n general there is good agree­ ment i n the measured forces between lecithin bilayers using the S F A and the osmo­ tic pressure technique, the interaction forces measured by the latter are usually of longer range (40,41). W i t h respect to probing the fusion process o f supported biomembranes, the major drawback resulting from the direct proximity of the solid substrate is the change in the viscoelastic properties of the overall bilayer which are important i n governing the initial breakthrough. A l s o , full fusion is hindered by the substrate which does not allow the study of all aspects of the fusion process. For the investigation of such dynamic processes involving membrane deformations and rearrangements of lipids or membrane-incorporated proteins within lipid bilayers, the realization o f substrate-supported, but fluid model membranes is desired. One promising approach is the use of a water-swellable cationic polyelectrolyte, bran­ ched polyethyleneimine (PEI), which electrostatically binds onto the negatively charged surface of the mica substrate (Figure 2 . A ) . This polymer gel can act as a deformable and mobile substrate and can build up an aqueous compartment between the membrane and the mica substrate. Recently, such "soft-supported membranes" have been used for the investigation of biomembrane structure and function (42). It is feasible that alternating polyelectrolyte multilayers can be formed stepwise by the use of a polyanion "counterlayer" which would allow control of the thickness of the polymeric cushion between the membrane and the solid support (43-45). Apart from their use i n the study o f fundamental principles governing biomembrane inter­ actions, technical and medical interest arises from their potential use i n biosensor systems (5,42,46).

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

220

Figure 2. Polymer-supported bilayers; A . physisorbed lipid membrane on support­ ing water swollen polyelectrolyte; B . partial covalent attachment of lower (proximal) lipid monolayer to the supporting polymer as indicated by the arrow.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

221

The chemical nature of the supporting polyethyleneimine layer also allows for the binding of the lipid headgroups onto the gel support by electrostatic interactions or v i a the formation of chemical bonds between amino reactive membrane inserting lipids ("tethered supported membranes", Figure 2.B). The covalent fixation can pro­ vide an additional stabilization factor which aids in slowing down the membrane de­ sorption from the polymeric substrate. In recent years, this approach has been discussed frequently in the literature as a concept of modelling the effect of the cytoskeleton i n living cells (42,47-53).

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

Formation of soft-supported D M P C bilayers on mica. Physisorbed Bilayers. The most straightforward method to build up a soft-suppor­ ted, physisorbed D M P C bilayer membrane on polyethyleneimine would be the un­ rolling of small unilamellar lipid vesicles onto the polyelectrolyte layer, which was shown to result in polymer-supported bilayers for negatively charged lipid vesicles on negatively charged polymer surfaces in the presence of C a ions (45). However, S F A measurements on zwitterionic vesicles adsorbed onto cationic polyethylene­ imine indicated a rather undefined multilayer structure which was confirmed in neutron scattering investigations on quartz supports (Majewski, J.; et al. Biophys. J., in press). A stepwise preparation was thus chosen for this particular system, in which the inner, polymer-supported D M P C monolayer was transferred onto the mica substrate by the Langmuir-Blodgett method from a subphase containing the polyelectrolyte. A s reported in the literature for the formation of other bilayer systems supported on solid substrates (6,7,48,54), the second layer was created by spreading unilamellar vesicles onto these preformed polymer-supported monolayers (Wong, J. Y . ; et al. Biophys. J., submitted). This approach provides two advantages: first, the sample preparation can easily be done within the S F A on preinstalled surfaces. Second, this method would allow the incorporation of membrane proteins directly from proteoliposomes. Structural investigations by neutron scattering confirmed the successful formation of the desired structures (the detailed results w i l l be published elsewhere). The preservation of lateral lipid mobility within the fluid membrane was studied by fluorescence microscopy in similar fashion as it w i l l be discussed below i n the context of partially attached membranes. 2 +

Covalently attached membranes. Most recently, reactive membrane-inserting amphiphiles have been used for the formation of tethered lipid bilayers on glass substrates (55). Adopting this approach for the purpose of studying membrane inter­ actions employing the S F A technique, a new isothiocyanate-functionalized l i p i d D M P C - N C S was synthesized from dimyristoylphosphatidylethanolamine ( D M P E ) and carbon disulfide. Its reactive isothiocyanate head group can selectively react with amino groups to form a N-N'-disubstituted thiourea structure i n the presence of water which allows the partial attachment of an interfacial reactive lipid monolayer to the amino functions of branched polyethyleneimine (PEI) dissolved i n the water subphase (compare Figure 3). The details of the lipid synthesis, as well as the mono­ layer characterization at the air-water interface have been described recently (Seitz, M . ; et al. Thin Solid Films, in press). The covalent fixation of the lipid monolayer onto the polymeric substrate was shown by IR measurements. A l s o , their increased stability against water could be observed i n studies of the contact angle of a water droplet i n contact with the monolayer surface. The results are summarized in Table I. It can be seen that the observed contact angles of water droplets on pure D M P C monolayers as well as on mixed monolayers of D M P C containing approximately 10 m o l % D M P E - N C S on polyethyleneimine were rather low. The advancing angle for the mixture was 0 A = 43° and the receding angle was 0 R < 15°. Moreover, the

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

222

Figure 3. Formation of tethered D M P C monolayer and bilayers partially attached to supporting polyethylenimine (PEI). Bilayers are formed by adsorption of small unilamellar vesicles on the monolayers.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

223 static contact angle of a water drop left on the monolayer decreased with time i n both cases, starting at a value of approximately 20°, until within a few minutes the water spread on the surface and wetted the support such that the contact angle was no longer measurable ( 0 « 0). Obviously, as long as the monolayers consist mainly of physisorbed D M P C they are not stable against water and are being "washed o f f the substrate or rearrange, presumably into bilayer patches. In contrast, the mono­ layers of pure reactive lipid on PEI/mica was much more hydrophobic and stable with their quality depending on the transfer conditions: the measured dynamic values on a monolayer transferred at room temperature from the liquid-expanded ("fluid") state were 0 A = 79° and 0 R = 16°. Despite this rather high hysteresis, the static contact angle of initially 64° dropped to only 47° after one hour. W h i l e these observations point to an increased stability of the PEI-supported monolayers at higher covalent binding densities of the l i p i d headgroups, the high hysteresis observed between the advancing and receding angles points to a rather heteroge­ neous surface structure. The quality of the layers can be improved when the films are transferred i n a more condensed state in which the monolayer exists at lower temperature (note that isotherm measurements on the reactive lipid D M P E - N C S on pure water as well as on polyethyleneimine subphase have been described recently, Seitz, M . ; et al. Thin Solid Films, in press).

Table I: Contact angle measurements on PEI-supported monolayers OA

OR

®stat.

Quality and Stability

21°C, 30 mN/m, "fluid" phase

31°

0 ° )

not stable against water

DMPE-NCS/ DMPC = 10:1

21°C, 35 mN/m, "fluid" phase

43°

0°)

not stable against water

DMPE-NCS

21°C, 35 mN/m, "fluid" phase

79°

16°

64° (-> 47°)

more hydrophobic and more stable under water, heterogeneous surface

13°C, 40 mN/m, "solid" phase

92°

52°

73° (stable)

further improvement of monolayer quality

Lipids

Transfer Conditions

DMPC

However, at the same time as the stability of polymer-supported monolayers increases when a higher amount of the lipid head groups is attached to the polymer substrate, the lateral lipid mobility within the monolayers is slowed down signifi­ cantly. Fluctuations along the surface normal should still be allowed in such tethered systems, and although this provides interesting perspectives with respect to studying the coupling of viscoelastic properties of the underlying polymer layer with the bound lipid layer, a complete fixation of the bilayer is not desired for S F A studies of membrane fusion processes. Thus, in the following a tethered supported membrane with only partial fixation of the proximal lipid layer w i l l be discussed.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

224 In a similar way as physisorbed membranes, tethered supported bilayers can be formed, e.g., by adsorption of D M P C vesicles onto polymer-supported mixed mono­ layers of D M P C and the synthetic reactive lipid D M P E - N C S (Figure 3). This process was followed by fluorescence microscopy. Figure 4 . A shows a mixed lipid monolayer of D M P C containing 10 m o l % D M P E - N C S and 0.5 m o l % T R - D H P E (Texas Red-labelled Dihexadecanoylphosphatidylethanolamine) on polyethyleneimine/mica in air which was prepared by L B transfer as described above. Note that two octagonal spots of the homogeneous fluorescing monolayer were bleached with the fluorescent light beam of the microscope, neither of w h i c h changed its shape during an hour after exposure of the sample proving that the fluorescent-labelled lipid is rather immobile within the polymer-supported mono­ layer in air. After the addition of a D M P C vesicle solution at 28 °C, the fluorescent-dye stayed on the substrate and the fluorescence of the sample remained homogenous. Moreover, the fluorescence of the sample recovered with time as T R - D H P E , now laterally mobile, diffused into a previously bleached spot. Figures 4.B and 4.C show the micrographs taken of the same spot 10 seconds and 10 minutes, respectively, after the initial bleaching with the fluorescent beam. Note that the radius o f the darker region decreased with time as a result of the diffusion of the fluorescent lipid. After one hour, the dark spot had almost completely disappeared. Note that these observations are only qualitatively comparable with F R A P experiments (fluore­ scence recovery after photobleaching, 56,57). However, the observed time-depen­ dent fluorescence recovery agrees well with an increased fluidity of the soft-suppor­ ted bilayers in aqueous solution as compared to the soft-supported monolayer system in air. A s expected, literature reported F R A P measurements for similar tethered bilayer systems show a lowered diffusion constant D = 1.5 ju,m -s o f the dyelabelled lipid, while within a polymer-supported D M P C bilayer D = 3 u r n ^ s (49). A n additional observation that is in agreement with the aforementioned contact angle measurements should be mentioned. After the addition of a pure water droplet to a PEI-supported monolayer of the lipid mixture fluorescence recovery is observed around the wetted region, but where the monolayer is still exposed to air. This points to a swelling of the underlying polymer gel which induces lateral lipid mobility in the supported monolayer. However, once the droplet covers the monolayer a rearrangement i n the monolayer takes place resulting in domain formation on the support as well as the detection of a strong fluorescence signal from the water surface. Considering the homogeneous fluorescence in the monolayer systems exposed to a D M P C vesicle solution directly, it appears that a second lipid layer is adsorbed on top of the transferred monolayer on initial contact with the solution which yields a stable, tethered supported bilayer as depicted in Figure 2.B. It appears that the successful formation of bilayer membranes on polymer supports by the described vesicle adsorption technique depends both on the quality as well as on the lipid mobility in the preformed polymer-supported monolayers, while the long-term stability is improved by the attachment to the membrane sup­ port. Whereas a polymer-supported monolayer can be created from a pure layer of reactive lipid on polyethyleneimine which is perfectly stable in water, the bilayer formation by vesicle adsorption is highly critical for these fully attached systems. A high extent of covalent binding between lipid headgroups and polymer support does no longer allow for lateral mobility within the proximal lipid layer which i n most cases would be crucial for the system in order to reorganize into a perfectly packed bilayer during the vesicle adsorption process. 2

_1

-1

Normal interaction forces of soft-supported D M P C bilayers. The phase transition temperature between gel and liquid crystalline phase of T = 24 °C for D M P C bilayers allows to study the effect of membrane fluidity in S F A measurements close to room temperature. The temperature dependence of the interc

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

225

Figure 4. Fluorescence micrographs of a mixed monolayer ( D M P C / D M P E N C S , 9.6:1 containing 0.5 mol% T R - D H P E on PEI/mica; A . in air; B . mono-layer in D M P C vesicle solution, 10 seconds after bleaching; C. same spot, 10 minutes after bleaching.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

226 action profile of two PEI-supported, pure D M P C membranes is shown i n Figure 5. W h i l e the details of these experiments have been discussed i n a recent publication (Wong, J. Y . ; et al. Biophys. J., submitted), a qualitative summary of the forcedistance profile at short distances shown i n Figure 5 and its implications with respect to the fusion process of biological membranes w i l l be given here. A b o v e the chain melting temperature of the lipid membrane, at 28 °C, a hemi­ fusion of the apposing bilayers can be observed discontinuity i n the force profile from 10.5 to 5.5 nm at F / R « 20 m N / m corresponding to pressures (Pmean = 7-10 atm) which are considerably smaller than for bilayers directly supported on mica (see above). The fluidity of the polymer supported bilayers is also indicated by their recovery after a previous hemifusion event. It was seen that after appropriate healing times (< 1-2 hrs) an essentially identical force profile was obtained i n subsequent measurements on the same contact spot. A l s o , a complete fusion of two lipid membranes in the liquid crystalline state can now be observed with the S F A as a second discontinuous decrease of the mica-mica separation by A D = 3 . 5 n m at higher pressures (Figure 5). In contrast, this could not be observed with the S F A for lipid bilayers deposited directly onto mica where a full removal of both lipid layers w o u l d result i n a direct mica-mica contact (22,24,26). Once fully fused and completely removed from the polymer substrate, even the fluid PEI-supported D M P C bilayers do not recover within the time of the experiment (16 hours after fusion). A t lower temperatures, when the D M P C bilayers are i n the gel state, very high pressures are needed for hemifusion. In measurements at 21 °C, hemifusion was observed at very high F / R values corresponding to pressures above P e a n ~ 50 atm (Figure 5). In addition, the healing process after hemifusion was very slow, and a second step corresponding to a full fusion of the two apposing membranes could not be achieved under pressures up to 150 atm (within observation times of up to two hours). The temperature dependence of the systems dimensions can be extracted from the interaction profile between PEI-supported D M P C bilayers at short distances (Figure 5). They correlate well with reported X-ray diffraction data on multilamellar D M P C aggregates in an excess of water; e.g. the observed shift of the steric barrier ("hard wall" ) by 1.4 nm below the phase transition temperature can be explained by these literature reported numbers (20 °C, Pp'-phase, bilayer thickness: d s = 4.45 nm, thickness of hydration water layer: d = 2.04 nm; 37 °C, L - p h a s e , a© = 3.55 nm, d = 2.45 nm (58); thickness of hard wall, Dhw = 2dfi + d ) . m

w

w

a

w

Conclusions and implications: factors governing the fusion of lipid bilayer membranes. The fusion barrier of two apposing bilayers (i.e. the pressure and energy required for their hemifusion) is directly related to the measured, temperature-dependent interbilayer forces during the breakthrough process. In comparison with the results obtained for solid-supported l i p i d bilayers on mica, the observations on softsupported membranes clearly reflect the influence of the underlying substrate. Bilayers supported on mica alone are restricted in the mobility of their proximal, solid-like D P P E layer which is i n direct contact with the solid substrate. This restricted mobility of the lower layer can result i n packing stresses and hole formation i n the distal D M P C layers when undergoing phase transitions, since the solid and laterally immobile D P P E layers do not accommodate for changes i n the mean molecular area of the outer lipid layer. In contrast, polymer supported D M P C membranes can stay intact while undergoing the phase change from the fluid L to the Pp' gel phase. Thus, the data obtained on soft-supported systems more closely relate to the properties of the whole bilayer membrane, whereas measurements on solid-supported bilayers merely probe effects related to structural changes o f the outer membrane leaflets. a

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

227

Figure 5. Force distance profile measured during compression of two physisorbed PEI-supported D M P C bilayers in the liquid crystalline state (at 28 °C, full symbols) and in the gel phase (at 21 °C, open symbols; adapted from Wong, J. Y . ; et al. Biophys. / . , submitted).

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

228 It is interesting to compare the S F A results on soft-supported bilayers with fusion studies on spherical bilayer vesicles. A s mentioned above, the fusion barrier, i.e. the pressure required to achieve hemifusion of two apposing membranes i n the S F A , drastically increases in the gel state. When comparing to unilamellar vesicles of small size (R < 50 nm), this observation seems to contradict findings on free unilamellar vesicles i n solution for which fusion of the bilayer membranes was found to be favored below the fluid-gel transition, T (27). However, this difference can be explained considering that compared to nearly planar bilayers these small, curved vesicles i n the gel state are highly stressed which they can overcome by fusion to larger vesicles (7,2,27). This is supported by the finding that phospha­ tidylcholine vesicles in contact that have a larger radius of curvature (R ~ 2 mm) show a similar trend as the nearly planar bilayer systems (R ~ 1 cm) investigated in S F A studies, as they also fuse more easily with increasing temperature (59,60). The same general driving force has been discussed both for the hemifusion of solid-supported bilayer membranes in the S F A (24,26) as well as for free vesicles in solution (60). In the first case, it has been argued that the exposure of hydrophobic chains on thinning of the outer membrane layer (which itself can be caused i n diffe­ rent ways as described above) has been argued to give rise to an attraction between the inner D P P E layers of the two membranes which ultimately results in the removal of the two outer lipid layers. In the latter case, it was assumed that by increasing the fluidity of the membrane, the freedom of motion of the hydrocarbon chains of lipid molecules can be increased so as to allow the hydrocarbon chains to be exposed at the membrane surface. In both scenarios, the actual fusion process w o u l d be governed by an attractive, hydrophobic interaction. However, i n addition to this general hydrophobic driving force several other factors can play a crucial role in the fusion of lipid bilayer membranes, and their relative influence on the process can strongly depend on the particular system under investigation. A s discussed before, for solid-supported lipid bilayers, most of these additional factors such as packing defects and phase separation are related to the properties of the outer lipid monolayer which can be considerably different than those of the supporting D P P E layer. In soft-supported bilayers as well as large uni­ lamellar vesicles, the fusion of planar bilayers in contact is facilitated with increa­ sing fluidity of the entire membrane. The additional influence of bilayer curvature is seen i n small vesicles for which the fusion of is promoted by their high curvature. A l s o , it is important to consider how the bilayers are brought into contact. Thus, i n studies on free vesicles in solution their enhanced aggregation below T contributes to the overall number of fusion events (27).

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

m

m

Summary. Phospholipid bilayers supported by a polymer cushion which can swell and act as a deformable and mobile substrate can preserve their membrane fluidity since their direct interaction with the solid substrate is prevented. For S F A measurements it is convenient to use branched polyethyleneimine as a water-swellable polymeric subs­ trate, as it can be easily adsorbed electrostatically onto atomically flat mica substra­ tes. A lipid bilayer can be formed by adsorption of small unilamellar vesicles onto polymer-supported monolayers which can be partially covalently attached to the polymer layer using reactive lipids that insert into the membrane in order to provide additional stability to the system. In comparison to lipid bilayer systems deposited on m i c a directly, the interaction profile - particularly the hemifusion - o f two polymer-supported membranes is less affected by the proximity of the supporting substrate and thus more closely resembles the properties of free bilayer vesicles. This also allows the observation of a complete fusion o f two apposing l i p i d membranes with the S F A . M a n y molecular processes on the outer membrane surface or within the outer membrane layers can be conveniently studied employing solid-supported bilayers,

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

229

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

such as phase separation, phase changes in the lipid layer, or ligand-receptor-interactions, and the influence o f adsorbed polymers on the force profile between biomembranes. In addition to this broad variety o f effects that can be investigated and understood with the S F A to date, the described soft-supported bilayers open up new possibilities. They w i l l allow to further study dynamic interactions o f more complex, and biologically more relevant model membranes, such as their adhesion and fusion under controllable and near in-vivo conditions. Acknowledgments. This work was supported by the National Institutes of Health (NIH) under grant G M 47334, by a N R S A individual postdoctoral fellowship G M 1 7 8 7 6 ( J . Y . W . ) . M . S . acknowledges partial funding by the Deutsche Forschungsgemeinschaft.

References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10.

11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26.

Chernomordik, L . V.; Melikyan, G. B.; Chizmadzhev, Y. A., Biochim. Biophys. Acta 1987, 906, 309-352. Cell Fusion; Sowers, A. E., Ed.; Plenum Press: New York, 1987. Brian, A . A.; McConnell, H . M . , Proc. Nat. Acad. Sci. USA 1984, 81, 61596163. Tamm, L . K.; McConnell, H . M . , Biophys. J. 1985, 47, 105-113. McConnell, H . M . ; Watts, T. H.; Weis, R. M . ; Brian, A . A., Biochim. Biophys. Acta 1986, 864, 95-106. Kalb, E.; Frey, S.; Tamm, L . K., Biochim. Biophys. Acta 1992, 1103, 307-316. Lang, H . ; Duschl, C.; Gratzel, M . ; Vogel, H . , Thin Solid Films 1992, 210/211, 818-821. Bayerl, T.; Bloom, M . , Biophys. J. 1990, 58, 357-362. Israelachvili, J. N., Chemtracts-Analyt. Phys. Chem. 1989, 1, 1-12. Parsegian, V . A . ; Rand, P. R.; Fuller, N . L . ; Rau, D. C . In Methods in Enzymology; L . Packer, Ed.; Academic Press: New York, 1986; Vol. 127; pp 400-416. LeNeveu, D. M . ; Rand, R. P.; Parsegian, V . A., Nature 1976, 1976, 601. Lis, L . J.; McAlister, M . ; Fuller, N.; Rand, R. P.; Parsegian, V . A., Biophys. J. 1982, 37, 657-665. Evans, E . A., Biophys. J. 1980, 31, 425. Evans, E.; Metcalfe, M . , Biophys. J. 1984, 46, 423-426. Evans, E., Langmuir 1991, 7, 1900-1908. Prieve, D. C.; Alexander, B. A., Science 1986, 231, 1269-1270. Prieve, D. C.; Frej, N. A., Langmuir 1990, 6, 396-403. Rädler, J.; Sackmann, E., J. Phys.II1993, 3, 727-748. Israelachvili, J. N.; Adams, G. E . , J. Chem. Soc., Faraday Trans I 1978, 9751001. Israelachvili, J., Acc. Chem. Res. 1987, 20, 415-421. Israelachvili, J. Intermolecular and Surface Forces; Academic Press Ltd.: London, 1991. Horn, R. G., Biochim. Biophys. Acta 1984, 778, 224-228. Marra, J.; Israelachvili, J., Biochemistry 1985, 24, 4608-4618. Helm, C. A.; Israelachvili, J. N.; McGuiggan, P. M . , Science 1989, 246, 919922. Leckband, D. E . ; Helm, C. A.; Israelachvili, J., Biochemistry 1993, 32, 11271140. Helm, C. A.; Israelachvili, J. N.; McGuiggan, P. M . , Biochemistry 1992, 31, 1794-1805.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.

230 27. 28. 29. 30. 31. 32. 33.

Downloaded by NORTH CAROLINA STATE UNIV on September 7, 2012 | http://pubs.acs.org Publication Date: August 20, 1999 | doi: 10.1021/bk-1999-0736.ch015

34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61.

Israelachvili, J. N.; Adams, G. E., Nature 1976, 262, 774-776. Israelachvili, J. N.; McGuiggan, P. M . , J. Mater. Res. 1990, 5, 2223-2231. Israelachvili, J., J. Colloid Interface Sci. 1973, 44, 259-272. Chan, D. Y. C.; Horn, R. G., J. Chem. Phys. 1985, 83, 5311. Hertz, H. J., Reine Angew. Math. 1881, 92, 156. Horn, R. G.; Israelachvili, J. N.; Pribac, F., J. Colloid Interface Sci. 1987, 115, 480. Johnson, K. L.; Kendall, K.; Roberts, A . D., Proc. R. Soc. London A 1971, 324, 301. Helm, C. A.; Israelachvili, J. N., Methods in Enzymology 1993, 220, 130-131. Chen, Y . L. E.; Gee, M . L.; Helm, C. A.; Israelachvili, J. N.; McGuiggan, P. M . , J. Phys. Chem. 1989, 93, 7057-7059. Leckband, D. E.; Schmitt, F. J.; Israelachvili, J. N.; Knoll, W., Biochemistry 1994, 33, 4611-4624. Helm, C. A.; Knoll, W.; Israelachvili, J. N., Proc. Natl. Acad. Sci. USA 1991, 88, 8169-8173. Leckband, D. E.; Israelachvili, J. N.; Schmitt, F. J.; Knoll, W., Science 1992, 255, 1419-1421. Wong, J. Y.; Kuhl, T. L.; Israelachvili, J. N.; Mullah, N.; Zalipsky, S., Science 1997, 275, 820-822. Horn, R. G.; Israelachvili, J. N.; Marra, J.; Parsegian, V . A.; Rand, R. P., Biophys. J. 1988, 54, 1185-1187. Pezron, I.; Pezron, E.; Bergenstahl, B. A.; Claesson, P. M . , J. Phys. Chem. 1990, 94, 8255-8261. Sackmann, E., Science 1996, 271, 43-48. Decher, G.; Hong, J. D., Ber. Bunsenges. Phys. Chem. 1991, 95, 1430-1434. Decher, G., Science 1997, 277, 1232-1237. Lindholm-Sethson, B., Langmuir 1996, 12, 3305-3314. Stelzle, M . ; Weissmüller, G.; Sackmann, E., J. Phys. Chem. 1993, 97, 2974-2981. Häussling, L.; Knoll, W.; Ringsdorf, H.; Schmitt, F. J.; Yang, J. L . , Makromol. Chem., Makromol. Symp. 1991, 46, 145-155. Spinke, J.; Yang, J.; Wolf, H.; Liley, M . ; Ringsdorf, H . ; Knoll, W., Biophys. J. 1992, 63, 1667-1671. Beyer, D.; Knoll, W.; Ringsdorf, H.; Elender, G.; Sackmann, E., Thin Solid Films 1996, 285, 825-828. Jacobson, K.; Sheets, E. D.; Simson, R., Science 1995, 268, 1441-1442. Janmey, P. In Handbook of Biological Physics; R. Lipowsky and E . Sackmann, Ed.; Elsevier: 1995; Vol. 1; pp 805-849. Heysel, S.; Vogel, H.; Sanger, M . ; Sigrist, H., Protein Sci. 1995, 4, 2532-2544. Cornell, B. A.; Braach-Maksvytis, V . L. B.; King, L. B.; Osman, P. D. J.; Raguse, B.; Wieczorek, L.; Pace, R. J., Nature 1997, 387, 580-583. Kühner, M . ; Tampe, R.; Sackmann, E., Biophys. J. 1994, 67, 217-226. Beyer, D.; Elender, G.; Knoll, W.; Kühner, M . ; Maus, S.; Ringsdorf, H . ; Sackmann, E., Angew. Chem. Int. Ed. Engl. 1996, 35, 1682-1685. Axelrod, O.; Kappel, D. E.; Schlessinger, J.; Elson, E.; Webb, W. W., Biophys. J. 1976, 16, 1055. Merkel, R.; Sackmann, E.; Evans, E., J. Phys. France 1989, 50, 1535. Janiak, M . J.; Small, D. M . ; Shipley, G. G., Biochemistry 1976, 15, 4575-4580. Breisblatt, W.; Ohki, S., J. Membr. Biol. 1975, 23, 385-401. Ohki, S. In Membrane Fusion Techniques, Part A; N . Düzgünes, Ed.; Academic Press: San Diego, 1993; Vol. 220; pp 79-89. Wong, J.Y., et al. "Interaction Forces and Fusion Between Lipid Bilayers Softly Supported' on Thin Polyelectrolyte Films." Biophys. J., in press.

In Supramolecular Structure in Confined Geometries; Manne, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1999.