Stefano
Borocci
*ab,
Giuseppina
Bozzuto
c,
Cecilia
Bombelli
d,
Francesca
Ceccacci
d,
Giuseppe
Formisano
c,
Annarita
Stringaro
c,
Agnese
Molinari
*c and
Giovanna
Mancini
*b
aDipartimento per la Innovazione nei sistemi Biologici, Agroalimentari e Forestali (DIBAF), Università degli Studi della Tuscia, L.go dell'Università, s.n.c., 01100 Viterbo, Italy. E-mail: borocci@unitus.it
bCNR, Istituto per i Sistemi Biologici, Area della Ricerca di Roma 1, SP35d 9, 00010 Montelibretti, Roma, Italy. E-mail: giovanna.mancini@cnr.it
cCentro Nazionale per la Ricerca e la Valutazione preclinica e clinica dei Farmaci, Istituto Superiore di Sanità, Viale Regina Elena 299, 00161 Roma, Italy. E-mail: agnese.molinari@iss.it
dCNR, Istituto per i Sistemi Biologici, Sede Secondaria di Roma-Meccanismi di Reazione c/o Dipartimento di Chimica Università degli Studi di Roma “Sapienza”, P.le A. Moro 5, 00185 Roma, Italy
First published on 2nd July 2021
Though liposome-based drugs are in clinical use, the mechanism of cell internalization of liposomes is yet an object of controversy. The present experimental investigation, carried out on human glioblastoma cells, indicated different internalization routes for two diastereomeric liposomes. Molecular dynamics simulations of the lipid bilayers of the two formulations indicated that the different stereochemistry of a lipid component controls some parameters such as area per lipid molecule and fluidity of lipid membranes, surface potential and water organization at the lipid/water interface, all of which affect the interaction with biomolecules and cell components.
We previously reported that liposomes composed of dimyristoyl-sn-glycero-phosphocholine, DMPC, and one of the two diastereomeric cationic gemini amphiphiles, 1a or 1b (Fig. 1), showed different efficacy in the delivery of the photosensitizer meta-tetrahydroxyphenylchlorin (m-THPC) to malignant glioma cells, DMPC/1b liposomes being more efficient than DMPC/1a liposomes.12 Furthermore, it was found that the different stereochemistry of the gemini component also controls the intracellular distribution of m-THPC.12 These findings strongly suggest that the ‘diastereomeric’ liposomes might follow different pathways of internalization.
Herein we report on experimental and theoretical investigations aimed at elucidating, on the one hand, the pathways of internalization and the intracellular trafficking of the two liposomes (DMPC/1a, 6:
4, and DMPC/1b, 6
:
4) and, on the other hand, the parameters that might control cell internalization. Appropriate inhibitors along with flow cytometry, antibody labelling along with laser scanning confocal microscopy (LSCM), and transmission electron microscopy (TEM) observations were used to investigate the potential differences in internalization pathways and intracellular trafficking. Molecular dynamics (MD) simulations of the lipid bilayers of the two formulations allowed us to shed light on the control of gemini component stereochemistry on the organization of lipids and on parameters such as lipid fluidity, surface electrical features and distribution of water bound to the surface of liposomes.
In this study we investigated by flow cytometry analysis, in human glioblastoma LN229 cells, the effect of different endocytosis and trafficking inhibitors on the uptake of fluorescent m-THPC loaded in DMPC/1a and/or DMPC/1b liposomes. In particular, we used (i) chloropromazine, a cationic amphiphilic drug which interferes with clathrin-mediated endocytosis at multiple levels, by inhibiting the function of AP2, one of the key adaptor proteins in clathrin-mediated endocytosis, and by trapping receptors inside the endosomes, thus blocking their recycling; (ii) filipin that inhibits caveolae-mediated endocytosis by forming complexes with 3-β-hydroxysterols in the plasma membrane, thus inducing the disassembly of filamentous caveolin-1-coating; (iii) bafilomycin A1, a macrolide antibiotic that affects the clathrin pathway by specifically inhibiting vacuolar-type H(+)-ATPase, thus preventing the acidification of endosomes and lysosomes; and (iv) LY294002 that inhibits macropinocytosis by interacting with phosphatidylinositol 3-kinase whose activity controls the arrangement of actin filaments.16
The results of uptake experiments of liposome-included m-THPC, carried out after pre-treatment of cells with the different inhibitors, are reported in Table 1 and show that the uptake of m-THPC mediated by the two formulations was influenced to a different extent by the various endocytic inhibitors (Table 1), though a total inhibition never occurred because different processes can be functioning in parallel for cell internalization. The uptake of m-THPC delivered by DMPC/1a was mainly inhibited by filipin (∼75%) and, to a lesser extent, by chloropromazine (∼50%). On the other hand the uptake of m-THPC delivered by DMPC/1b liposomes was strongly inhibited by chlorpromazine (∼80%) and, to a lesser extent, by filipin (∼40%). Therefore, both formulations enter cells via endocytosis and cell internalization involves more than a single specific pathway. The extent of inhibition exerted by filipin and chlorpromazine indicates that caveolae are mainly involved in the internalization of DMPC/1a whereas the clathrin-mediated pathway is mainly involved in the internalization of DMPC/1b. Bafilomycin A1 and LY294002 inhibited to some extent (∼40% and ∼47%, respectively) the uptake of m-THPC loaded in DMPC/1b liposomes while they did not affect that mediated by DMPC/1a liposomes, showing on the one hand that the acidic compartments are involved only in the final destination of the DMPC/1b formulation and, on the other hand, that macropinocytosis is a concomitant route of its internalization.
Inhibitor | Percentage of inhibition | Inhibition route | |
---|---|---|---|
DMPC/1a | DMPC/1b | ||
Chloropromazine | 53 ± 5 | 80 ± 3 | Clathrin-mediated endocytosis |
Bafilomycin A | 0 | 40 ± 8 | Acidification of endosomes |
Filipin | 75 ± 4 | 38 ± 5 | Caveolae-mediated endocytosis |
LY294002 | 0 | 47 ± 4 | Macropinocytosis |
Fluorescent DMPC/1a liposomes (green) co-localize with caveolin-rich regions (red), as shown by the regions of yellow pixels in panel VI of Fig. 2A, and with organelles labeled with Rab5-antibody (blue signal), as shown by light blue regions (early endosomes, panel IX, Fig. 2A). Localization of the same liposomes is hardly observed in clathrin (panel III of Fig. 2A), Rab7-positive or Lamp-1-positive organelles (late endosomes and lysosomes, respectively; panels XII and XV of Fig. 2A). Conversely, merging of color-coded channels shows that DMPC/1b liposomes localize preferentially in cytoplasmic regions rich in clathrin (yellow regions in panel III of Fig. 2B), in early and late endosomes and in the lysosomes (light blue regions in panels IX, XII and XV of Fig. 2B), whereas colocalization of DMPC/1b liposomes with caveolin (panel VI of Fig. 2B) is sporadic.
These results nicely confirmed those obtained in the inhibition experiments, indicating that caveolae and clathrin-coated vesicles serve as preferential “gates of entry” for DMPC/1a and DMPC/1b liposomes, respectively. In fact, DMPC/1a liposomes were mainly internalized in caveolae and in early endosomes (Rab5+ organelles) whereas DMPC/1b liposomes in clathrin-positive vesicles, late endosomes (Rab7+ organelles) and lysosomes (Lamp-1+ organelles). These findings are in agreement with previous studies reporting that the mechanism of caveolae uptake does not transport the internalized material to late endosomes and lysosomes, whereas clathrin-mediated endocytosis causes the endocytosed material to end up in degradative lysosomes.6,12
Therefore, TEM observations confirm different endocytic routes, caveolae- and clathrin-mediated, for DMPC/1a and DMPC/1b, respectively, as traced by inhibition and immunofluorescence experiments. Actually, the different endocytic routes of the two formulations could explain why DMPC/1a liposomes were less efficient than DMPC/1b in delivering m-THPC in the same range of time.12 In fact, it is known that the uptake of caveolae-mediated endocytosis occurs at a much slower rate than that of clathrin-mediated endocytosis.6
We know from a previous investigation17 that DMPC/1a liposomes are characterized by a higher transition temperature and by a minor extent of lipid miscibility with respect to DMPC/1b liposomes. It was also found that DMPC/1a liposomes feature a surface potential, Ψs, higher than DMPC/1b liposomes.12 At 37 °C DMPC/1a and DMPC/1b liposomes feature a similar diameter (150–200 nm) after extrusion, and the ultrastructural investigation reported above did not suggest different sizes of the liposomes processed for internalization, which is/are the parameters that control the route of cell entry and how?
We carried out MD simulations of DMPC/1a, DMPC/1b (at 6:
4 molar ratio) and DMPC bilayers, the last one to both investigate the effect of the gemini components on the physicochemical features of the lipid bilayer and validate our calculations, since the DMPC bilayer has been largely investigated.18–31 In particular, we took into consideration molecular interactions and organization that might control parameters such as fluidity, electrical features, and surface interaction with biomolecules.
〈AL〉 was calculated by dividing the lateral (XY) dimensions of the simulated box by the number of lipid molecules in each leaflet. In the case of DMPC, 〈AL〉 was 0.638 ± 0.002 nm2 at 310 K, in agreement with experimental data and previous molecular dynamics simulation,18 whereas it was 0.716 ± 0.002 nm2 and 0.721 ± 0.003 nm2 for DMPC/1a and DMPC/1b, respectively. The increase of the area per lipid molecule in DMPC/1 bilayers is in agreement with literature data relative to other cationic mixed lipid bilayers. In particular, it is reported that while a low content of the cationic component causes a condensing effect and hence a decrease of area per lipid molecule,33,34 a high concentration of the cationic component, as in the case of 6:
4 DMPC/1 systems, involves electrostatic repulsion between charged headgroups and hence leads to an increase of the area per lipid with respect to pure phosphocholine.
To investigate the area per lipid of each component in the mixed lipid bilayer we applied the Voronoi tessellation using a set of selected key atoms as implemented in APL@voro,35 in particular, the phosphorus atom of DMPC and the mass center of the two stereogenic centers of the gemini headgroup. In the DMPC/1a system, the values of the area per lipid were 0.696 ± 0.003 nm2 and 0.719 ± 0.002 nm2 for DMPC and 1a, respectively, while in the DMPC/1b system the values were 0.706 ± 0.003 nm2 and 0.743 ± 0.006 nm2 for DMPC and 1b, respectively. Thus the two formulations feature a different averaged area per lipid (slightly larger in the case of DMPC/1b), and this difference is mostly due to the larger area per lipid of 1b with respect to the 1a component; the headgroup interactions involve also a slightly larger area of DMPC in DMPC/1b with respect to the DMPC/1a formulation (Fig. S1 in the ESI†). The larger area per lipid of 1b and DMPC, due to a higher extent of charge repulsion, suggests a higher charge exposure in gemini 1b with respect to 1a.
The area per lipid and its variation over simulation time were used to calculate the area compressibility modulus, KA, that provides a measure of the elastic properties of lipid bilayers, higher values of KA corresponding to higher rigidity of lipid bilayers36,37 (see the Material and methods section). The average KA calculated for the DMPC bilayer was 334 ± 22 mN m−1, whereas in the case of DMPC/1a and DMPC/1b it was 283 ± 19 mN m−1 and 253 ± 18 mN m−1, respectively. Therefore, the larger area per lipid, that characterizes the DMPC/1b bilayer with respect to the DMPC/1a one, involves an increase of the elasticity of the DMPC/1b membrane with respect to the DMPC/1a membrane.
Fig. 4 shows the density distribution of the most relevant atoms and functional groups of lipid components at the lipid/water interface, namely nitrogen (NDMPC), phosphorus (PDMPC) and carbonyl groups (COsn-1 and COsn-2) of DMPC, both nitrogen N1a and N1b and both oxygen, O1a and O1b, of 1a and 1b, respectively. The analysis of mass density shows that in DMPC/1a and DMPC/1b bilayers nitrogen atoms of the gemini component are located close to the phosphorus atom of DMPC and replace the nitrogen atom of choline in the formation of charge-pairs. The electrostatic repulsion between DMPC choline groups and gemini ammonium groups induces a reorientation of phosphocholine headgroups and a consequent exposure of their nitrogen atom (NDMPC) toward the water phase. These findings are in agreement with results obtained by molecular dynamics simulation of cationic lipid membranes reported previously.33,38 The presence of the gemini component also induces a decrease of the thickness of the lipid bilayer, calculated as the distance between the peaks of the phosphate groups from the density profile, which was 3.57 nm in the case of DMPC (Fig. S2†), 3.22 nm for DMPC/1a and 3.21 nm for DMPC/1b, this being in agreement with the increase of surface area per lipid and fluidity, described above.
On the other hand, a difference between DMPC/1a and DMPC/1b bilayers concerns the time-averaged position of methoxy group oxygen atoms of 1a and 1b (O1a and O1b) that suggests a different conformation of gemini headgroups and a different orientation of 1a and 1b methoxy groups.
Next, we evaluated O–C–C–O dihedral angle distribution of gemini headgroups and calculated the angular distribution of C → OCH3 vectors of gemini and P–N vectors with respect to the normal to the lipid bilayer. The evaluation of the first parameter gave us detailed information on the effect of the different configuration of stereogenic centers of 1a and 1b on the conformation of the O–C–C–O headgroup portion. In the case of 1a, the O–C–C–O segment adopts exclusively an anti conformation whereas in 1b it adopts exclusively a gauche conformation (g+ = 57.9%, g− = 42.1%) as shown in Fig. 5A.
The orientation of C → OCH3 vectors with respect to the normal to the lipid bilayer confirms the different conformation adopted by the O–C–C–O portion in the gemini components. In 1a, methoxy groups are oriented in opposite directions, one towards the lipid bilayer hydrophobic region and the other one toward the water phase ( – Fig. 5B), whereas in 1b, both methoxy groups are mainly oriented towards the hydrophobic region (
), as shown in Fig. 5C and in the snapshots of DMPC/1a and DMPC/1b bilayers reported in Fig. 6.
The orientation of the P–N vector is strongly affected by the presence of the gemini component as it switches from 84.5° in DMPC to ∼36.0° in the case of both mixed lipid bilayers (Fig. S3 in the ESI†); however, it is not affected by the gemini stereochemistry.
A detailed picture of the short-range order of headgroups at the lipid bilayer surface was obtained by calculating the 2D radial distribution function (2D-RDF) between the phosphate group of DMPC and significant atoms of the gemini component. The 2D-RDF (g2D(r)) between the phosphate group and nitrogen atoms of gemini, reported in Fig. 7A, shows an intense peak at a lateral distance of 0.47 nm due to the first shell of neighbor nitrogen atoms around the phosphate group, and two less intense peaks at 0.77–1.0 nm range distance, for both DMPC/1a and DMPC/1b bilayers. This analysis does not give information on the orientation of the whole gemini molecule with respect to phosphorus atoms, namely it does not indicate if both nitrogen atoms of the gemini face the phosphate group. Insight into this point was given by the analysis of 2D-RDF between the DMPC phosphate group and the center of mass of gemini stereogenic centers (Fig. 7B). In both systems the 2D-RDF shows a first peak at a lateral distance of r = 0.49 nm and a more intense peak at 0.76 nm. However, in the case of DMPC/1b the peak at 0.49 is less intense with respect to DMPC/1a indicating that this configuration is less abundant. Representative snapshots of DMPC/1 pairs selected from MD trajectories contributing to the peaks in the radial distribution (Fig. 7C and D) illustrate the different organization of headgroups in the two lipid bilayers. The first peak at a lateral distance of r = 0.49 nm in the RDF can be ascribed to interactions between DMPC and gemini molecules that orient both nitrogen atoms toward the phosphate group (Fig. 7C). In the case of 1a the head-group conformation of gemini allows a tighter interaction between the head groups of DMPC and gemini, with the phosphate group coordinating two nitrogen atoms of the same gemini molecule. The second and more intense peak at 0.76 nm is due to the interaction between the phosphate and only one nitrogen atom of the neighbor gemini, with the second nitrogen atom far from the phosphate group of DMPC (Fig. 7D). These findings are in agreement with the results relative to the area per lipids.
The order parameter, SCD, is defined as
Because we used a united atom force field, the order parameter of each methylene group was calculated based on the position of neighboring methylene and assuming a tetrahedral geometry.18,39Fig. 8A and B show the averaged order parameter, |SCD|, of methylene groups of sn-1 and sn-2 alkyl chains in the DMPC bilayer in the presence and in the absence of the gemini component. In all simulated systems, the value of |SCD| for the methylenes of DMPC tails is lower than 0.25, which indicates the occurrence of a fluid lipid bilayer with disordered alkyl chains.40 Actually the three bilayers were investigated at 310 K, i.e. above their transition temperature and are all in the liquid crystal phase. The presence of gemini involves a decrease of the order parameter of DMPC tails, indicating less ordered tails in the mixed bilayers with respect to the pure phospholipid, with very small differences due to the nature of the gemini. On the other hand, a relevant difference between the mixed systems concerns the averaged order parameter of methylene groups of gemini alkyl chains (Fig. 8C), the first four methylenes of 1a alkyl chains showing a higher order parameter (lower orientational mobility) with respect to 1b. The anti orientation of 1a methoxy groups involves a steric hindrance on the first methylenes of alkyl chains that reduces their orientational mobility. Conversely, in 1b the gauche conformation of the O–C–C–O torsional angle and the consequent syn orientation of methoxy groups allow a higher flexibility of the first methylenes of alkyl chains.
It is worth noting that these results are in good agreement with those concerning the surface area per lipid, as it is known that the SCD of alkyl chains is related to the surface area per lipid, a larger area per lipid corresponding to a lower order parameter, and, hence, to more disordered alkyl chains.41,42
Both mixed lipid bilayers feature a positive electrostatic potential with respect to bulk water; however, they feature a different boundary potential, Ψb, (defined as the difference between the electrostatic potential found at the center of the lipid bilayer and the electrostatic potential of the water phase), the Ψb of DMPC/1b being higher (1.175 V) than that of DMPC/1a (1.062 V). We analyzed also the surface charge density, σ(z), of the cationic lipid bilayers as a function of the distance from the center of the lipid bilayer, and the relative profiles are reported in Fig. 9B. In DMPC/1b the value of σ(z) close to the surface of the lipid bilayer, in the region of DMPC phosphate groups and of the headgroups of gemini (1.0 < z < 1.9 nm), is positive and slightly higher with respect to DMPC/1a. For z > 2.0 nm, in the region of nitrogen atoms of DMPC, surface charge density of DMPC/1a switches to a higher value with respect to DMPC/1b.
These results are not in contrast with those reported previously12 and obtained by experimental data that, as mentioned above, gave a higher surface potential in the case of DMPC/1a. In fact, the experimental determination of surface potential was carried out by an indirect method based on the effect of the microenvironment at the water/lipid interface of liposomes on the pKa of the umbelliferone fluorescent probe. The different headgroup organization and lipid packing of DMPC/1a and DMPC/1b bilayers could determine a different position of the probe at the lipid/water interface providing the value of the detected potential surface of slightly different zones in the case of DMPC/1a and DMPC/1b.
The analysis of water properties with respect to the membrane surface took into account the roughness of the surface. Fig. 10A shows the water density profiles relative to DMPC, DMPC/1a and DMPC/1b bilayers calculated as a function of the distance d from the surface of the bilayer (d = 0). All the systems do not show a smooth density profile and water density defines four regions: region I (−1.0 < d < −0.14 nm for DMPC and −1.0 < d < 0.05 nm for DMPC/1) includes water molecules close to carbonyl groups; region II (−0.14 < d < 0.40 nm for DMPC and 0.05 < d < 0.52 nm for DMPC/1) includes the first shell of water molecules close to phosphate and ammonium ions of DMPC and gemini; region III (0.40 < d < 0.80 nm for DMPC and 0.52 < d < 1.0 nm for DMPC/1) the second shell of water molecules close to ammonium ions of DMPC; region IV (d > 0.80 nm for DMPC and d > 1.0 nm for DMPC/1) bulk water.
Fig. 10 shows that the hydrophobic region close to carbonyl groups of phosphocholine (region I) in the case of the DMPC bilayer is characterized by a deeper penetration of water molecules with respect to the DMPC/1 systems. Some differences between the cationic bilayers were observed in water density close to the lipid surface. In fact, DMPC/1a shows a higher density of water molecules in region I, whereas DMPC/1b shows it higher in the region of the second solvation shell of DMPC ammonium groups (region II).
Water molecule orientation with respect to the lipid bilayer was then evaluated by analyzing the average cosine of the angle formed by the dipole vector and the normal to the lipid bilayer (n). Note that for water randomly oriented the value of the averaged cosine is zero. On the other hand, positive and negative values of the averaged cosine correspond to an averaged orientation of water molecules with oxygen atoms pointing inward and outward the lipid bilayer, respectively. Hence, higher positive or negative values of averaged cosine correspond to a larger extent of orientation of water molecules. Fig. 10B shows water dipole orientation as a function of the distance d from the surface of DMPC/1 bilayers. In both systems, the oxygens of water molecules are oriented toward the lipid bilayer; however water molecules of the second hydration shell of DMPC ammonium ions (region III) bound to the DMPC/1a bilayer feature a higher grade of orientation with respect to those bound to DMPC/1b; only negligible differences were found in the other regions of hydration.
Lateral diffusion of water (Dlat) and the relaxation time (τ1) of dipolar moment of water molecules as a function of the distance d from the surface of the lipid bilayer were also calculated, because these parameters allow evaluating the translational mobility and the rotational diffusion of water close to the lipid bilayer.
For the model of water we used (SPC), at 310 K we obtained for the bulk D = (5.01 ± 0.02) × 10−5 cm2 s−1 and τ1 = 2.5 ± 0.2 ps. In all simulated systems we observed (Table 2) a slowdown of lateral mobility and rotational relaxation of water close to the lipid surface with respect to bulk water. In particular, in the region I the mobility and rotational relaxation of water are reduced, with respect to bulk water, by ∼10 and ∼80 times, respectively, due to deep penetration into the lipid bilayer. In the region II, the mobility and rotational relaxation of water are reduced, with respect to bulk water, by ∼4 and ∼16 times, respectively. Finally, in the region III and IV both lateral diffusion and rotational relaxation of water approach the value of bulk water. However, while in the case of the DMPC bilayer the values of both parameters relative to region IV are similar to those of bulk water, in the cases of DMPC/1 bilayers they are reduced with respect to bulk water, thus indicating that the perturbation effect of the lipid bilayer and chloride ions on the organization and dynamics of water extends beyond 1.5 nm from the surface of the lipid membrane. Therefore, this analysis indicates that the presence of the gemini component affects the mobility of water molecules at the surface of the lipid bilayer in regions II–IV; furthermore, it shows that its stereochemistry does not affect lateral mobility of water molecules (see also the ESI†), while slightly affecting water rotational relaxation in regions II and III. In particular water molecules bound to the DMPC/1b bilayer show a faster relaxation with respect to those bound to the DMPC/1a bilayer.
Region | DMPC | DMPC/1a | DMPC/1b | |||
---|---|---|---|---|---|---|
D lat | τ 1 | D lat | τ 1 | D lat | τ 1 | |
I | 0.47 ± 0.02 | 154.9 ± 6.0 | 0.46 ± 0.04 | 188.9 ± 7.0 | 0.48 ± 0.03 | 194.1 ± 7.0 |
II | 1.94 ± 0.03 | 31.3 ± 2.0 | 1.46 ± 0.03 | 42.3 ± 3.0 | 1.47 ± 0.02 | 38.9 ± 3.0 |
III | 3.98 ± 0.02 | 4.4 ± 0.2 | 3.31 ± 0.02 | 7.3 ± 0.2 | 3.32 ± 0.03 | 6.5 ± 0.3 |
IV | 4.90 ± 0.02 | 2.6 ± 0.2 | 4.47 ± 0.02 | 2.9 ± 0.2 | 4.45 ± 0.02 | 3.0 ± 0.2 |
Lateral and rotational diffusion of water are influenced by the capacity of forming and breaking hydrogen bonds with other molecules.48 Therefore, in order to better characterise the structure and dynamics of water molecules close to the surface of DMPC/1 bilayers we analysed hydrogen bonds between water molecules close to the lipid surface and those between water molecules and oxygen atoms of lipids.
Calculations indicated that in DMPC/1a and DMPC/1b bilayers each molecule of DMPC forms 5.8 and 5.7 hydrogen bonds with neighbour water molecules, respectively, a number slightly lower with respect to a mere DMPC bilayer where each molecule of DMPC forms 6.2 hydrogen bonds.
Fig. 11 shows the average probability of hydrogen bond formation between water molecules and oxygen atoms of DMPC. The different probability to form hydrogen bonds reflects the different exposure of DMPC oxygen atoms to the solvent and hence their different environment.21,23
The analysis, shown in Fig. 11, indicates that ester oxygens OA, OB, OE, and OG (OA and OB belonging to C–O–R bonds and OE and OG to P–O–R bonds) have a lower tendency to form hydrogen bonds with respect to carbonylic oxygen atoms, OF and OH (CO) and non-ester oxygen of the phosphate group OC and OD. In fact, in the DMPC bilayer the ester oxygens OA, OB, OE, and OG give rise exclusively to a single hydrogen bond with a probability of 30% for OA and of ∼15% for OB, OG and OE. This tendency slightly increases in the presence of the gemini component, except for OE, which shows a modest reduction with respect to the bilayer of pure DMPC.
In the case of OC and OD oxygen atoms the probability of forming one hydrogen bond increases in the presence of the gemini (56% in DMPC/1 bilayers versus 48% in DMPC, Fig. 11G and H), whereas the probability to form two hydrogen bonds for each oxygen is reduced (30% in DMPC/1 bilayers versus 37% in DMPC). The same trend was observed also in the case of OF oxygen atoms (sn-2 carbonyl oxygen), the probability of forming one hydrogen bond being 54% in DMPC/1 bilayers versus 50% in DMPC bilayer, and for the formation of two hydrogen bonds 25% in DMPC/1versus 32% in DMPC. The other carbonylic oxygen atom, sn-1 OH, shows a lower tendency to bind water molecules with respect to sn-2 OF, (Fig. 11C and D), this difference reflecting the different orientation of the sn-1 and sn-2 carbonyl groups with respect to the lipid surface and a different hydration of the two ester groups, as evidenced by the analysis of the radial distribution function (Fig. S4 in the ESI†).
The stereochemistry of the gemini component does not affect the capability of DMPC to bind water molecules by hydrogen bonds.
Differently from what observed in the case of DMPC oxygen atoms, the oxygen atoms of gemini headgroups form hydrogen bonds with water molecules very rarely. In fact, an averaged value of 1.2 hydrogen bonds was calculated over all 52 molecules of 1. The absence in the gemini components of functional groups able to form stable intermolecular hydrogen bonds with water and the hydrophobic nature of the charged ammonium groups lead to the formation of “clathrate-like” structures of water around the headgroups of cationic gemini molecules (Fig. 11J) that play a key role in the structure and dynamics of water close to the lipid bilayer.20,49,50
The dynamical properties of water molecules close to the lipid bilayer surface were then investigated by analysing their ability to break and reform hydrogen bonds between themselves and with DMPC. The analysis of lifetime of hydrogen bonds gave in the case of relaxation time (τ) of water–water hydrogen bonds τw–wHB = 3.2 ps in the case of DMPC, τw–wHB = 4.7 ps in DMPC/1a and τw–wHB = 4.4 ps in DMPC/1b.
The hydrogen bonds between water and DMPC have a longer lifetime with respect to water–water hydrogen bonds, namely τhg-wHB = 41.5 ps, τhg-wHB = 50.0 ps and τhg-wHB = 49.4 ps in DMPC, DMPC/1a and DMPC/1b bilayers, respectively.
Therefore, in the interfacial region of DMPC/1 bilayers we observed a stronger interaction between water molecules and between water and lipid molecules with respect to the DMPC bilayer due to longer lasting hydrogen bonds. This could be due to lower density of water in the interfacial region of DMPC/1 with respect to DMPC (Fig. 10) that reduces hydrogen bond switching between water molecules.51 In addition, the orientation of water molecules around gemini headgroups and the formation of “clathrate-like” structures might play a key role in the increase of the lifetime of water–water hydrogen bonds. In fact, in “clathrate-like” structures formed around hydrophobic surfaces, water molecules are ordered and translational and rotational diffusion is reduced.52–57
The two liposomes, DMPC/1a and DMPC/1b, differ exclusively for the stereochemistry of the gemini component, 1a being chiral because of the S configuration of both its stereogenic centres and 1b being a meso form, i.e. achiral. This difference, that might seem subtle, was found to control some of the physicochemical features12,17 of the considered lipid membranes such as the transition temperatures of the two liposomes and the lipid miscibility. In this work it was shown that the stereochemical information also controls the biological fate of DMPC/1a and DMPC/1b liposomes, the first formulation entering cells mainly by caveolae mediated endocytosis and the second one by clathrin mediated endocytosis and partly by macropinicytosis. Hence the stereochemistry of the gemini component controls the intracellular target of the DMPC/1 carriers and of their payloads. In the case of DMPC/1a the target is early endosomes whereas in the case of DMPC/1b it is lysosomes; this means that, if a plasma membrane were the target of a given drug, DMPC/1a would be the formulation of choice, whereas if the nucleus were the target, as in the case of DNA, then DMPC/1b would be the carrier for reaching it. This is confirmed by our previous results that indicated DMPC/1b as a better formulation than DMPC/1a for the delivery of genetic materials such as a DNA plasmid58 or siRNA.59,60
Finding a connection between the stereochemical structure, the physicochemical properties and finally the biological fate of liposomes is a complicated task, and one should keep in mind that physicochemical properties are interrelated and contribute altogether in defining liposome biological features.
The MD simulation allowed understanding what happens at the molecular level, providing information that gathered together give us a clear picture of the two systems, DMPC/1a and DMPC/1b. Its results are summarized in Table 3 together with those obtained in the biological evaluation. The stereochemistry dictates the conformation of the head group of the gemini component embedded in the lipid bilayer. The different conformation assumed by the two stereoisomers results in a different exposure of charges. From this difference derive the differences between DMPC/1a and DMPC/1b in terms of extent of charge repulsion between gemini molecules and interaction with the phosphate group of DMPC, and hence the difference in area per lipid (higher in the case of DMPC/1b and 1b itself with respect to DMPC/1a and 1a, respectively) with a consequent effect on the elasticity of lipid bilayers. The different conformation assumed by 1a and 1b also implies a different order parameter of the respective alkyl tails and hence a different fluidity of DMPC/1a and DMPC/1b, DMPC/1b being less ordered and more fluid than DMPC/1a. The different reciprocal organization of lipids components due to the different stereochemistry of gemini components is also reflected in the surface electrical features and in the penetration and organization of water at the lipid/water interface. In fact, the DMPC/1b bilayer features a higher boundary potential than DMPC/1a. Penetration of water is different in the regions defined with respect to the surface of the lipid bilayer, in particular, in the case of DMPC/1b water density is higher than in DMPC/1a, in the region that includes the first shell of water molecules close to phosphate groups and ammonium ions of DMPC and gemini (region II), at the boundary with the region of the second shell of water molecules close to ammonium ions of DMPC (region III). This implies a higher probability of hydrogen bond switching between water molecules and a shorter hydrogen bond lifetime with respect to water at the surface of the DMPC/1a bilayer. Consistently, water molecules in regions II and III of the DMPC/1a bilayer were more oriented and characterized by a longer rotational dipolar relaxation time with respect to those in the DMPC/1b bilayer.
Liposome | Biological features | Lipid organization and physicochemical features | Water at surface | H-Bond | |||
---|---|---|---|---|---|---|---|
Cell internalization | Intracellular trafficking | 〈AL〉 (nm2) | K A (mN m−1) | Ψ b (V) | τ 1 region III (ps) | τ w-wHB (ps) | |
DMPC/1a | Caveolae | Early endosomes | 0.716 ± 0.002 | 287 ± 20 | 1.062 | 7.3 ± 0.2 | 4.7 ± 0.2 |
DMPC/1b | Clathrin and macropinocytosis | Late endosomes and lysosomes | 0.722 ± 0.002 | 257 ± 18 | 1.175 | 6.5 ± 0.3 | 4.4 ± 0.2 |
All these differences, some marked and other subtle, account for the different biological behaviour of the formulations, because they all affect the interactions with biomolecules and the cell membrane. In particular, the organization of water and the surface potential control the interaction with protein in serum and thus the nature of biomolecular corona that in turn control the interaction with cell components and the intracellular fate.6,7
The message is that ascribing the control of biological features exclusively to membrane fluidity or particle charge or size or biological water might be too generic and restrictive since, as shown here, physicochemical parameters are interrelated and it is their whole to control the biological outcome.
Gemini surfactants (S,S)-2,3-dimethoxy-1,4-bis(N-hexadecyl-N,N-dimethylammonium)butane bromide, 1a, and (S,R)-2,3-dimethoxy-1,4-bis(N-hexadecyl-N,N-dimethylammonium)butane bromide, 1b, were synthesized as previously described.61,62
m-THPC containing liposomes were prepared by adding an appropriate volume of a m-THPC(5 × 10−4 M, EtOH abs) stock solution to the chloroform solution of the lipids to obtain, after hydration, a 50 μM m-THPC final concentration.
At the end of each treatment, cells were washed with ice-cold Hank's balanced salt solution (HBSS), detached with EDTA and 0.25% trypsin, resuspended in ice-cold PBS and immediately analyzed for the photosensitizer content. Fluorescence signals were analyzed with a FACScan™ flow cytometer (Becton Dickinson, Mountain View, CA) equipped with a 15 mW, 488 nm and air-cooled argon ion laser. The fluorescence emission was collected through a 670 nm band-pass filter and acquired in “log” mode. At least 10000 events were analyzed. The m-THPC content was evaluated as fluorescence intensity, expressed as the mean fluorescence channel (MFC). The analysis was performed using CellQuest™ software (Becton Dickinson). Results analyzed were reported as mean percent of inhibition obtained from 3 independent experiments.
Cells grown for 24 h on glass coverslips were incubated with NBD tagged liposomes for 18 h at 37 °C. At the end of the treatment LN229 cells were fixed with 3.7% paraformaldehyde in PBS for 15 min at room temperature, washed twice in PBS, and permeabilized with 0.04% Triton in PBS + 1% BSA. Cells were washed twice, incubated with 1:
100 primary antibodies in PBS + 1% BSA for 45 min at 37 °C, washed twice and incubated with 1
:
50 secondary antibody 633 nm Alexa Fluor® Dyes in PBS + 1% BSA for 30 min at 37 °C. Finally, samples were washed and observed under a Leica TCS SP2 confocal microscope (Leica Microsystems, Mannheim, Germany).
For freeze-fracturing cells were washed twice in 0.1 M cacodylate buffer, resuspended in the same buffer containing 30% glycerol, and incubated overnight at 4 °C. Samples were then put on carriers and quickly frozen in Freon 22, partially solidified at the liquid nitrogen temperature. The mounted carriers were then transferred into a Bal-Tec BAF 060 freeze-etch unit (BAL-TEC, Balzers, Liechtenstein), cleaved at −100 °C at a pressure of 2–4 × 10−7 mbar, shadowed with 2 nm of platinum–carbon and replicated with a 20 nm carbon film. Cells were digested for 2 h from the replicas by chlorox. The replicas were mounted on grids, and examined with a Philips 208S transmission electron microscope (FEI Company, Eindhoven, The Netherlands).
Starting coordinates of the lipid bilayer, composed of 128 molecules of DMPC, were taken from the website of Biocomputing laboratory at the University of Calgary (http://wcm.ucalgary.ca/tieleman/downloads) and solvated with 5227 water molecules in a rectangular box of 6.5 × 6.5 × 7.0 nm3. The mixed bilayers of DMPC and gemini components were obtained by a random selection of 52 molecules of DMPC replaced by the same number of gemini molecules. The positive charges of the mixture lipid bilayers were neutralised by replacing 104 water molecules with chloride ions.
The partial atomic charges of the headgroups of gemini molecules were evaluated using the RESP fit method72 of the electrostatic potential obtained from the HF/6-31G(d) wave function using the Merz–Singh–Kollman scheme. The electrostatic potential was calculated on the B3LYP/6-31 G(d,p) optimized geometry of the gemini headgroups by using a high point density around the molecule (17 points per Å2 and 10 layers around the van der Waals molecular surface). The RESP fit of the electrostatic potential was performed by using the antechamber module of AmberTools16.73 The quantum mechanical calculations were performed by using the GAUSSIAN03 program package.74 The partial atomic charges and the atom type of the headgroups of gemini molecules are reported in the ESI.†
The torsional parameters for O-CH2-CH2-O and CH3-O-CH2-CH2 dihedrals of the gemini molecules were parameterized as described in the ESI.†
Lennard–Jones and electrostatic interactions were calculated using a cut-off of 1.0 nm and the long-range electrostatic interactions were accounted by using the particle mesh Ewald method (PME).75
All bonds were constrained using the P-LINCS algorithm76,77 whereas the geometry of water molecules was fixed with the SETTLE algorithm.78 The simulations were carried out with a time step of 2 fs in the NPT ensemble. After the initial energy minimization, the simulated systems were warmed up by five consecutive unrestrained MD from 100 K to 323 K in 500 ps. After 120 ns of equilibration at 323 K the temperature of each system was cooled down from 323 K to 310 K in 1.5 ns by four consecutive MD and finally simulated for 500 ns at 310 K.
The lipids, gemini surfactants, water and ions were coupled separately to a temperature bath using the velocity rescale method79 with a time constant of 0.1 ps. The pressure was kept at 1 bar by weakly coupling to a pressure bath,80 using a coupling constant of 1.0 ps and an isothermal compressibility of 4.5 × 10−5 bar−1 and in the last 400 ns the Parrinello–Rahman barostat81 (P = 1 bar, τP = 4.0 ps) was used. Pressure coupling was applied semi-isotropically: the z and the x–y (isotropic) box dimensions were allowed to vary independently. Periodic boundary conditions were applied in all three dimensions.
For each simulated system three simulations with different random initial velocity were performed.
The surface charge density as a function of z was calculated using the relationship:
The lateral diffusion coefficient Dlat of water molecules as a function of the distance d from the surface of the lipid bilayer was obtained by dividing the d dimension into 0.2 nm slabs.85 The lateral diffusion coefficient Dlat was calculated from the mean square displacement (MSD) of water molecules using the Einstein equation:
The MSD of water in each zone was calculated over 20 ps intervals by considering only the water molecules located in the same zone at time t0 and t.
The rotational dynamics of water molecules in the different regions of lipid bilayers was investigated by the analysis of the dipole autocorrelation function
The relaxation time of dipole autocorrelation function was determined according to
The mean and standard error of diffusion coefficients were obtained by splitting the trajectories into pieces of 20 ns and using the block averaging method.86
Hydrogen bonds formed by water with other water molecules or with oxygen atoms of lipid molecules were identified by using the geometric criteria; we considered two water molecules or a water molecule and an oxygen atom linked by hydrogen bonds when the oxygen–oxygen distance was less than 0.35 nm and the angle formed by the O–H bond of the donor molecule with the O–O direction was smaller than 30°.87
The number of hydrogen bonds formed by a DMPC lipid or gemini molecule was obtained by counting the number of water molecules bound to oxygen atoms of the bilayer component (according the geometric criteria) and by averaging over time (300 ns) and over all DMPC or gemini molecules.
The time correlation functions of hydrogen bonds were calculated by using the following equation88
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1nr02175c |
This journal is © The Royal Society of Chemistry 2021 |