Open Access Article
Karlo
Komorowski
a,
Jannis
Schaeper
a,
Michael
Sztucki
b,
Lewis
Sharpnack
b,
Gerrit
Brehm
a,
Sarah
Köster
a and
Tim
Salditt
*a
aInstitute for X-Ray Physics, University of Göttingen, Friedrich-Hund-Platz 1, 37077 Göttingen, Germany. E-mail: tsaldit@gwdg.de
bEuropean Synchrotron Radiation Facility, CS 40220, 38043 Grenoble Cedex 9, France
First published on 16th April 2020
We have used time-resolved small-angle X-ray scattering (SAXS) to study the adhesion of lipid vesicles in the electrostatic strong-coupling regime induced by divalent ions. The bilayer structure and the interbilayer distance dw between adhered vesicles was studied for different DOPC:DOPS mixtures varying the surface charge density of the membrane, as well as for different divalent ions, such as Ca2+, Sr2+, and Zn2+. The results are in good agreement with the strong coupling theory predicting the adhesion state and the corresponding like-charge attraction based on ion-correlations. Using SAXS combined with the stopped-flow rapid mixing technique, we find that in highly charged bilayers the adhesion state is only of transient nature, and that the adhering vesicles subsequently transform to a phase of multilamellar vesicles, again with an inter-bilayer distance according to the theory of strong binding. Aside from the stopped-flow SAXS instrumentations used primarily for these results, we also evaluate microfluidic sample environments for vesicle SAXS in view of future extension of this work.
At some point in its life, triggered by changes in environmental parameters or by strong interactions, a vesicle it is bound to transform. Structural observation and characterization of these shape transformations is of fundamental interest, but challenging in view of their transient nature and the small length scales involved, on the order of the bilayer thickness. In the biological context, membrane fusion is known as a well-controlled and important physiological process. Fusion of neurotransmitter-filled synaptic vesicles (SVs) with the presynaptic plasma membrane,4 for example, is an essential step in nerve conduction. This process is catalyzed and controlled by proteins, in particular by the soluble N-ethylmaleimide-sensitive factor attachment protein receptors (SNAREs).5 In general, the merger of two membranes involves a highly complex interplay on the molecular level among lipids, proteins, ions of the aqueous environment and water molecules. Important aspects of the membrane fusion pathway, concerning for example intermediate structures, such as the docking state,6,7 and the quantitative evaluation of underlying forces and energetics, are currently under investigation. To this end, numerical studies have provided interesting insight into possible structures and mechanisms,8,9 and now call for experimental verification.
Lipid vesicles without proteins have also been intensively studied. The equilibrium phase diagram of vesicle shapes was calculated long ago,10 and a quantitative understanding of inter-bilayer forces based on linear superposition of molecular interactions has entered textbooks.1,11 Self-consistent field-theoretical approaches including Helfrich-type steric interaction predict a phase transition from bound to unbound lipid membranes.12 More recently, effects from non-linear electrostatics, such as like-charge attraction, have been considered, and a regime of strong coupling induced by multi-valent ions has been identified.13–16 Binding of divalent ions, in particular Ca2+ and Mg2+ to lipid membranes has been studied by Molecular Dynamics (MD) simulations.17–19
We have recently studied the adhesion state of vesicles, as induced by the divalent ions Ca2+ and Mg2+ for different ion concentration, lipid composition, and surface charge density σs.20 Using small-angle X-ray scattering (SAXS), we could distinguish a strong adhesion state – probably caused by ion bridging – and a soft adhesion state in the presence of background monovalent salt concentration. While vesicle adhesion was observed when mixing anionic and zwitterionic or neutral lipids (binary and quaternary lipid mixtures), pure anionic membranes (DOPS) with correspondingly high σs were found to undergo a phase transition to a multilamellar state. The structure and interactions between charged MLVs in the presence of multivalent ions was also studied by SAXS in ref. 21. Recently, strong coupling has been studied between two highly charged solid-supported lipid bilayers in the presence of monovalent counter-ions by neutron and X-ray reflectivity.22 While interesting aspects of bilayer interactions beyond linear electrostatics were elucidated, studies of ion-induced transition, non-equilibrium effects or the kinetics of vesicle phase transitions still remain scarce, with the exception of experiments based on temperature changes (T-jump).23
In this work, we now extend our SAXS studies of the adhesion state, presenting a wider range of ions and first time-resolved studies monitoring the transition from SUVs to MLVs as illustrated in Fig. 1(a and b). We show that the divalent salts induced attractive interactions between bilayers, which are in quantitative agreement with the strong coupling theory put forward by R. Netz and coworkers.13–16 Furthermore, we have studied the adhesion transition by time-resolved SAXS using the rapid mixing technique as illustrated in Fig. 1(a) as well as with microfluidic devices. Importantly, we can identify a transient docking intermediate state in the SUV to MLV transition in DOPC:DOPS mixtures, induced by Ca2+.
For this study, we have used both a stopped-flow rapid mixer as well as microfluidic devices, and different beam conditions, instrumentation and optics have been evaluated for studies of time-resolved SAXS, in particular with respect to docking and fusion of vesicles. Therefore, it is adequate to briefly consider the state-of-the art in this field: SAXS combined with the stopped-flow rapid-mixing technique has been used previously to study osmotic shrinkage of sterically stabilized liposomes,24 as well as the role of calcium in membrane condensation and spontaneous curvature variations in model lipidic systems.25 A calcium triggered lamellar to hexagonal phase transition (Lα–H2) was studied by combined rapid mixing and time-resolved synchrotron SAXS.26 For SAXS in combinations with microfluidics, we refer to the review.27 Direct monitoring of calcium-triggered phase transitions in cubosomes using small-angle X-ray scattering combined with microfluidics has been reported in ref. 28. A microfluidic platform for the continuous production and characterization of multilamellar vesicles was presented in ref. 29. Finally, a microfluidic SAXS study of unilamellar and multilamellar surfactant vesicle phases was reported in ref. 30.
| lB = e2(4πεε0kBT)−1, | (1) |
| μ = (2πqlBσs)−1, | (2) |
![]() | (3) |
Together, these three length scales are sufficient to describe all regimes of linear electrostatics in solution. As mentioned above, the dominant attraction between like-charge membranes when adhesion is induced by divalent ions is described by the so-called strong coupling theory,13–16 which is characterized by ion bridging and/or ion correlation effects. In contrast, mean-field electrostatics which neglect ion–ion correlation effects is denoted as the weak coupling regime.2,3 The two regimes are delineated by the unitless coupling parameter Ξ =
B/μ, where
B = q2lB. The Poisson–Boltzmann approximation is valid for Ξ ≪ 1, while Ξ ≫ 1 is denoted as the strong coupling regime. In the strong coupling regime, an analytical expression for the interaction pressure is given as
![]() | (4) |
![]() | (5) |
In order to tightly connect the microfluidic device and the tubing in a leak-free manner, the device was integrated into a sample holder. The sample holder consists of two plates made of aluminium, which are designed with a gab to expose the channels. The two plates sandwich a PVC (polyvinylchloride) plate and the device itself. The front metal plate and the PVC plate have five small holes matching the holes of the device to attach the tubing to the device. For sealing, the tubing was further threaded through o-rings. A photograph of the sample holder integrated at the GINIX endstation at the P10 beamline is shown in Fig. 5.
To establish certain flow velocities on all four inlets and to be able to control the flow remotely the neMESYS pump system (Cetoni GmbH, Korbussen, Germany) is used in combination with Hamilton Gastight Syringes (Hamilton Bonaduz AG, Switzerland).
At ID02, the photon energy was set to 12.56 keV by a Si(111) mono crystal monochromator. The beam size at the sample plane was 30 μm × 30 μm for microfluidics experiments. The scattered X-rays were recorded using an Rayonix MX-170HS CCD pixel detector (Rayonix L.L.C., USA) with 3840 × 3840 pixels at a sample-to-detector distance of 1.5 m.
Briefly, we use the standard decomposition of the powder-averaged kinematic structure factor S(q) and the bilayer form factor F(q) = |f(q)|2 with the form factor amplitude f(q) to write the scattering intensity I(q) ∝ 〈F(
)S(
)〉, where 〈 … 〉 denotes the orientational average. q is given by the modulus of the momentum transfer vector q = |
| = (4π/λ)sinθ, where λ is the wavelength of the X-rays and θ the half of the scattering angle relative to the incident beam. The electron density profile (EDP) ρ(z) normal to the bilayer, which enters into the form factor by
, is parameterized by three Gaussian functions according to
![]() | (6) |
The final expression for the scattering intensity of the docking model used in the analysis is given by20
![]() | (7) |
| S(q) = 2 + 2cos(qd), | (8) |
![]() | (9) |
| Itot(q) = c1·Imod(q) + c2. | (10) |
![]() | (11) |
For the bilayer structure, a symmetric profile was enforced, reducing the number of free parameters. The Gaussian parameters representing the headgroups are σh = σh1 = σh2 and ρh = ρh1 = ρh2. While the width σc of the Gaussian representing the chain region is a free parameter, the amplitude and the position were fixed to ρc = −1 (arb. units) and zc = 0 (nm), respectively. The positions of the two outer Gaussians are denoted zh1,2 = ±zh. Estimation for standard deviation of the fit parameters was performed by calculation of the variance–covariance matrix approximated by cov = resnorm (JTJ)−1/N.
:
DOPS (1
:
1) vesicles, resulted in a stable adhesion state.20 The obtained interbilayer spacing was well reproduced by the strong-coupling theory. Moreover, dw was approximately the same for CaCl2 and for MgCl2, and was independent of the concentration of the ions. Contrarily, ν was increased when increasing the ion concentration. This indicates, that the concentration of the divalent ions has a minor effect on the equilibrium interbilayer spacing, but indeed affects the fraction of adhered bilayers.
The characteristic shape of the SAXS signal from adhered vesicles is discussed in detail in ref. 20. Briefly, in the case of vesicle adhesion, structure factor modulations can be observed in a most pronounced manner between the characteristic form factor minima of the lipid bilayer (depending on the bilayer thickness, approximately in the q-range of 0.4 and 2.5 nm−1). The structure factor modulations are superimposed by the bilayers form factor, so that the minima of the structure factor exhibit a smooth profile. The intermediate modulations of the structure factor vary in a characteristic manner with dw.
Here, we extend the previous study by using other divalent salts as well as by using different molar ratios in the DOPC:DOPS mixtures. Moreover, the results presented in Fig. 2 will be an important basis for the time-resolved SAXS experiments discussed later. Fig. 2(a and b) shows a series of SAXS data of lipid vesicles prepared in a 100 mM glucose solution, in which structural changes were induced by (a) ZnCl2 and SrCl2 for DOPC
:
DOPS (1
:
1), and by (b) CaCl2 for DOPC
:
DOPS (1
:
1, 1
:
2, and 1
:
4). Comparing the SAXS signal with the pure lipid vesicle suspension, e.g. no salt added, structure factor modulations characteristic for the adhesion state can be observed for almost all cases except for DOPC
:
DOPS (1
:
4), where two Bragg peaks are visible. This indicates, that increasing the surface charge density by increasing the content of DOPS results in a collapse of the vesicles and to an rearrangement of the vesicles to a multilamellar state at a certain point. Previously, this observation was also made with pure DOPS vesicles suspended in ultra-pure water.20
The quantitative analysis of the adhesion state is exemplified in (c) for DOPC
:
DOPS (1
:
1) with the addition of 12.5 mM SrCl2. From the docking model fit, the effective electron density profile (EDP) and the interbilayer spacing, here dw ≈ 1.6 nm, are obtained as indicated in the inset. The results are well in line with the previous studies20 for CaCl2 and MgCl2. The structural parameters are summarized in Table 1. The modeling of the interaction potentials by using the strong-coupling theory for the dominating attraction between the like-charge bilayers and an additional hydration potential for the repulsion is shown in (d) using the parameters described in the caption. The experimentally obtained interbilayer spacing of dw = 1.6 nm can be well reproduced by the superposition of the strong-coupling potential and the hydration potential. Thus, the small bilayer separations, e.g. strong like-charge attraction between the bilayers, are the result of ion-correlation effects associated with strong coupling according to theory. The simulation parameters are as summarized in the caption of Fig. 2.
| Sample | ρ h | σ h, σc | d hh (nm) | d w (nm) | (1 − ν) | χ red 2 | Ref. |
|---|---|---|---|---|---|---|---|
DOPC : DOPS (1 : 1), 4 mM CaCl2 |
1.39 | 0.44, 0.95 | 3.71 | 1.56 | 0.89 | 1.14 | 20 |
DOPC : DOPS (1 : 1), 4 mM MgCl2 |
1.34 | 0.39, 0.84 | 3.72 | 1.72 | 0.98 | 0.89 | 20 |
DOPC : DOPS (1 : 1), 12.5 mM SrCl2 |
1.14 | 0.54, 0.98 | 3.63 | 1.59 | 0.83 | 1.44 | — |
DOPC : DOPS (1 : 1), 4 mM ZnCl2 |
1.13 | 0.48, 0.89 | 3.63 | 1.59 | 0.85 | 101.76 | — |
DOPC : DOPS (1 : 2), 4 mM CaCl2 |
1.31 | 0.46, 0.91 | 3.75 | 1.26 | 0.85 | 17.63 | — |
Note, that further flow-through SAXS experiments were performed to study the reaction between DOPC
:
DOPS (1
:
1) vesicles and the trivalent salts FeCl3, Al2(S04)3, and MgSO4 as shown in the ESI† (Fig. S2). In those cases, no signature was observed which could be attributed to the docking state of vesicles. Instead, phase transitions to different multilamellar states were observed.
:
DOPS vesicles with the a molar ratio of (a) 10
:
1, (b) 1
:
1 and (c) 1
:
4, either mixed with 15 mM CaCl2 (a and c) or 25 mM SrCl2 (b). After mixing, the first SAXS image is recorded approximately 0.05 after starting the measurement script, subsequently following the reaction over a time window of approximately 92 s (a and c) and 316 s (b). However, subtracting the dead time of the stopped flow mixing time, the true kinetic time can be smaller down to the millisecond range, accounting for an offset between the elapsed time in the data acquisition and the kinetic time. In all data sets, different states along the reaction pathway can be distinguished as a function time. In particular, the transition of unilamellar vesicles to adhered vesicles is visible, even if the final states are different. For DOPC
:
DOPS (10
:
1), notably, the final state is difficult to attribute. The docking modulation seems to be superimposed with several smaller (structure factor) peaks indicative of short range order between several adhering bilayers. Osmotic shrinkage of vesicles may also contribute to the signal.
For DOPC
:
DOPS (1
:
4), the final state is clearly characterized by the structure factor of multilamellar vesicles. This state is reached from a transient docking state. As revealed by the least-square fits, the collapse of vesicles occurs at ν ≈ 0.5. For the observation of the intermediate state of adhered vesicles in this system, the time-resolved SAXS combined with the stopped-flow technique was essential. SAXS experiments using the flow-through cell could only reveal the initial state of unilamellar vesicles without mixing with CaCl2, and the final state of multilamellar vesicles by measuring the equilibrated reaction upon mixing with CaCl2 (cf.Fig. 2). The quantitative analysis of the adhesion state is exemplified in (d) for representative SAXS curves of each data set, where the typical structure factor modulation associated with docking are observed.
Fig. 4 shows the main structural results combining the analysis of the stopped-flow SAXS data and of the flow-through SAXS data. The structural parameters d, dhh and dw are displayed in (a) as a function of the molar ratio of DOPC:DOPS (10
:
1, 1
:
1, 1
:
2, and 1
:
4). For (10
:
1), (1
:
1), and (1
:
4) the structural parameters are the mean values from the time-course analyzed by docking model fits to the stopped-flow SAXS data (corresponding to the analysis of at least ten SAXS signals, where a clear adhesion state could be identified). As shown in Fig. S2 (ESI†), the structural parameters dhh and dw remain nearly constant over time. The error is given as the standard deviation of the mean. For the molar ratio of (1
:
2), the results are from the analysis of the flow-through SAXS data shown in Fig. 2(b). The error is calculated by the approximated parameter co-variance matrix of the least-squares fit.
![]() | ||
Fig. 4 Analysis of the vesicle adhesion. (a) Structure of the adhesion state: The structural parameters d, dhh and dw for different molar ratio of DOPC : DOPS. The values for molar ratios (10 : 1), (1 : 1), and (1 : 4) are the mean of all least-square fit results for curves corresponding to the adhesion state in the stopped-flow SAXS series shown in Fig. 3. The corresponding standard deviation of the mean is taken as errorbar. For the molar ratio (1 : 2) the parameters are obtained from the flow-through SAXS data shown in Fig. 2. Here, the errorbar is given by the estimated standard deviation of the least-squares analysis. (b) Kinetics of adhesion: The surface fraction (1 − ν) of non-adhered bilayers as a function of time, obtained from least-squares analysis of the stopped-flow SAXS. (c) Analysis of the strong coupling regime: each data point represents a different system with varied charge density σs and corresponding coupling parameters, notably Ξ(10 : 1) = 3.97, Ξ(1 : 1) = 19.84, Ξ(1 : 2) = 26.19, and Ξ(1 : 4) = 29.76. The experimentally determined water layer thickness dw is plotted against the value predicted by the strong coupling theory. We can deduce that within the probed range of the four different coupling parameters, prediction and measurement are reasonably close, with a correlation coefficient r = 0.9948. The fact, however, that only 2 out of the 4 points are within error bars may suggest that the experimental error is slightly underestimated. | ||
In (a) it can be observed that with an increasing σs, i.e. increasing content of DOPS, the structural parameters d and dw decrease, while dhh shows a slight increase. While those parameters remain nearly constant over time, in (b) we can observe that (1 − ν) as a measure of the fraction of single (undocked) bilayers decreases significantly as a function of time for all stopped-flow data sets shown here. The increase of the fraction of adhered bilayers is most pronounced for DOPC
:
DOPS (1
:
4), indicating that the surface charge density σs is an important factor for triggering the adhesion-reaction. Prior to the collapse of the DOPC
:
DOPS (1
:
4) vesicles (cf.Fig. 3(c)), inferred from the emerging Bragg peak, the fraction of adhered bilayers is ν ≈ 0.5. Contrary, for DOPC
:
DOPS (10
:
1) and for DOPC
:
DOPS (1
:
1) ν ≈ 0.1 and ≈0.2, respectively.
In (c), the structural parameter dw obtained by the docking model analysis of the SAXS data (shown in (a)) is plotted as a function of dw,sim obtained from the simulations of the strong coupling theory. It can be observed, that the experimentally measured water spacings of the lipid bilayer, obtained for the different σs values, are well in line with the predictions from the strong coupling theory. For the simulated water distance predicted by strong coupling, the following parameters were chosen and kept constant for all systems: area per lipid headgroup dAH = 64 Å2, lB ≃ 7.11 × 10−10 m, λh = 2 × 10−10 m, Ph = 3.3 × 109 J m−3, and T = 294 K. In contrast, the thickness of the headgroup dh,FWHM is taken from the fit results of dh by calculating the full width half maximum (FWHM), and more importantly, the surface charge density σs is varied corresponding to the molar ratios of the DOPC
:
DOPS mixtures and the Gouy–Chapman length μ is calculated accordingly. Further, it is of interest to also discuss the multi-lamellar final state of the DOPC
:
DOPS 1
:
4 system in view of strong coupling. From the position of the first Bragg peak qn=1 ≈ 1.208 1/nm, we obatain d ≈ 5.2 nm, and by subtraction of the bilayer thickness (as obtained from the least-square fit of the adhesion state), the water layer can be determined to dw ≈ 5.2 − 3.935 = 1.27 nm, again in good agreement with the predicted value dw,sim ≈ 1.25 nm. All additional fit results, in particular for dhh and dw as a function of time are shown in the appendix (Fig. S1, ESI†).
Fig. 5 presents an overview of the setup and illustrates the micro-focus SAXS experiment at GINIX. In Fig. 5(a) a photo of the microfluidic device is shown, mounted on the sample tower, for alignment and subsequent scanning in x, y, and z. Coarse alignment and inspection of the flow was enabled by an on-axis video camera (OAV). A microfluidic layout with four-inlet/one-outlet channels was chosen for all devices. Flow properties were simulated by finite elements (COMSOL), see Fig. 5(b). Lipid vesicles were simulated as particles with 60 nm in diameter with the diffusion constant of D ≈ 8 × 10−12 m2 s−1 and injected into the upper diagonal inlet. The flow rates are 100 μL h−1 and 25 μL h−1 for the diagonal inlets and for the side inlets, respectively. In this configuration, the side inlets focus the flow from the diagonal inlets. Along the outlet, diffusion of the particles outside of the focused stream can be observed. Finally, Fig. 5(c, d) show examples of integrated scattering intensity maps (darkfield maps), as recorded by scanning-SAXS with 1 s accumulation time for each frame. As indicated, 10 mg ml−1 extruded DOPS vesicles and 10 mM CaCl2 suspended in ultra-pure water were injected through the diagonal inlets at the flow rates of 25 μl h−1 to study the reaction upon mixing along the outlet. Ultra-pure water was injected through the side-inlets at the flow rates of 100 μl h−1. Each pixel represents the integrated photon counts from a single 2d-diffraction pattern. The darkfield images show an increase of the intensity in the vesicle inlet as well as in the mixing region, corresponding to the scattering from vesicles. In particular, a large increase of the intensity can be observed after the mixing of DOPS vesicles with CaCl2. The one-dimensional curves I(q) vs. q, obtained by azimuthal integration of the 2d diffraction patterns, are exemplified in Fig. 5(e) for selected points along the reaction line. The corresponding locations are indicated by numbers in Fig. 5(d). Note that the raw data are shown before background subtraction. Qualitatively, the curves 1 and 2 show an aggregated state of vesicles with an emerging multilamellar Bragg peak modulated by the bilayer form factor. Along the outlet, a phase coexistence regime of different multilamellar states can be observed, see Bragg peaks of curves 3 and 4. From previous experiments it is known that DOPS vesicles mixed with CaCl2 show a transition to a monophasic multilamellar state.20 Accordingly, the multilamellar state obtained from the microfluidics SAXS experiments shown here has not reached its equilibrium state.
The darkfield maps and the SAXS curves reveal the limitations of the configuration: high background and limited q-range. Aside from background generated by the chamber itself, the micro-focusing by compound refractive lenses (CRL) and the relatively large beam path in air both compromise the SAXS quality. While the strong multilamellar signals can still be extracted under these conditions, the much weaker signal of unilamellar vesicles (see the vesicle inlet in the darkfield map) does not reveal much structural information. For this reason, the second experiment at ID02/ESRF used a much more optimized configuration for SAXS, and keeping air paths at minimum, albeit sacrificing spatial and hence also temporal resolution.
Fig. 6 and 7 presents examples of SAXS data recorded in this configuration. In this setting, it became possible to study vesicle adhesion of DOPC:DOPS mixtures at varied molar ratios mixing with CaCl2. Fig. 6 shows the results obtained by microfluidic SAXS for DOPC
:
DOPS (1
:
1) vesicles mixed with a 4 mM CaCl2 solution suspended in ultra-pure water. The samples were injected through the diagonal inlets as indicated in (a) by the darkfield obtained by scanning SAXS of the microfluidic device. The flow rates were 100 μl h−1 and 0 μl h−1 for the diagonal and the vertical inlets, respectively. For a point-wise background subtraction, the device was previously measured with only ultra-pure water being injected in the channels. (b) shows an example of a background corrected SAXS signal corresponding to the position [1] highlighted in the darkfield. The background corrected SAXS signal were analyzed by fitting the docking model to the SAXS data as exemplified in (c) for three SAXS signals, corresponding to the positions [1], [2] and [3] highlighted in the darkfield. For the positions [1] and [3] no reaction with CaCl2 is expected, while position [2] corresponds to the mixing region. The fraction of adhered bilayers quantified by ν is slightly increased for the SAXS signal from position [2] (ν = 0.04 at position [2], compared to ν = 0.01 at position [1], and [3]), indicating that a small adhesion-reaction can be observed. The water spacing dw ≈ 1.7 nm is well in line with the results obtained by flow-through and stopped-flow SAXS discussed above.
Fig. 7 shows much stronger evidence for the adhesion state for DOPC
:
DOPS (10
:
1) vesicles and 10 mM CaCl2, again along with the corresponding least-square fits. However, this signal is observed only at the stagnation point (almost zero flow rate), and not at other points in the channel with higher flow rates, see the darkfield map in (a) with the point of stagnation highlighted by a red box. The scattering curves I(q) vs. q taken from the point of stagnation show indeed the characteristic structure factor modulations corresponding to adhesion of the vesicles, which is exemplified in (b) by the background corrected SAXS signal corresponding to the position marked in (a). The docking model analysis reveals an interbilayer spacing of dw = 2.4 nm, which is in good agreement with the results obtained by the stopped-flow SAXS data and the theory of strong coupling, as discussed above.
The current study has also solidified and extended the experimental basis of this very interesting effect of like-charge attraction, predicted by the theory of strong coupling. For the first time, the surface charge density of the bilayers and hence the coupling parameter was varied to study the corresponding variation in water layer distance. To this end, different molar ratios of the DOPC
:
DOPS mixtures were probed. Our previous study of vesicle adhesion in a DOPC
:
DOPS (1
:
1) mixture20 was also extended concerning the type of divalent ions used. As we have shown, adhesion is induced not only by CaCl2 and MgCl2, but also by SrCl2 and ZnCl2. Importantly, variation of ions results only in small changes of the water layer thickness, as expected from strong coupling theory. Whether ion radius and polarizability can account for these differences remains to be investigated. The interbilayer spacings for different surface charge densities, i.e. for different coupling parameters ranging from Ξ ≈ 4… 30, were in good agreement with the strong coupling theory. Deviations between the experiments and the simulations of the strong coupling theory could eventually be attributed to a change of the effective dielectric constant of water, which was kept constant (ε = 80) for all coupling parameters. Recent water-explicit numerical simulations of nanometer-separated charged surfaces revealed a surface charge-induced reorientation of hydration water, which modifies the dielectric constant of water as well as the hydration repulsion.39
Concerning experimental settings of the microfluidics experiments, we conclude that the signal quality of scanning SAXS with focused beams, while desirable in terms of temporal and spatial resolution, strongly compromises data quality at the present configurations. Future improvements will address more options for cleaning of the beam as well as a configuration with tightly evacuated flight paths. Further reduction of the chamber background scattering for example by thinner windows (presently 120 μm total thickness), and possibly also by the processing protocol are also important. Finally, a layout with 5 inlets would be well suited to avoid a stagnation point, which often led to unwanted formation of large lipid aggregates which in turn impeded proper microfluidic flow.
With all of these improvements in place, we anticipate that both microfluidics and stopped flow experiments will enable kinetic studies of vesicle transformation and reactions in a complementary manner. Already in the present proof-of-concept study and under current parameters, the time-resolved SAXS data has allowed us to probe the structure of a transient adhesion state of highly charged vesicles induced by calcium injection. Importantly, the binding of like-charged vesicles in this state results in a spacing which is in very good agreement with strong binding theory, predicted more than fifteen years ago. This example demonstrates, that transient vesicle states can not only be evidenced but that the corresponding structures can be described in quantitative terms. This opens up an interesting perspective in applications of this approach for biological vesicles undergoing shape transformations and functionally important transitions. Understanding the docking and fusion reaction of synaptic vesicles (SVs), for example, is of significant interest in view of a quantitative understanding of chemical synapses. To this end, we include a first test exposure of SVs in a microfluidic device, comparing the present data quality with respect to static SAXS (Fig. S3, ESI†).
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/d0sm00259c |
| This journal is © The Royal Society of Chemistry 2020 |