Startseite Fliposomes: stimuli-triggered conformational flip of novel amphiphiles causes an instant cargo release from liposomes
Artikel Open Access

Fliposomes: stimuli-triggered conformational flip of novel amphiphiles causes an instant cargo release from liposomes

  • Vyacheslav V. Samoshin

    Vyacheslav V. Samoshin graduated with an Honorable Diploma (MS) from Moscow State University, USSR, in 1974. At the same University, he defended his PhD dissertation in Organic Chemistry under guidance from Academician Nikolay S. Zefirov in 1982, and defended his Doctor of Chemical Sciences dissertation in 1991. He worked as a researcher in the Department of Chemistry, MSU, and since 1992 as Professor and Head of the Division of Organic Chemistry at the Moscow State Academy of Fine Chemical Technology. He was awarded the title Honorary Professor by this Academy in 2012. He took his present position as Professor of Chemistry at the University of the Pacific, Stockton, California in 1999. His scientific interests include molecular switches, especially in application to liposome design, conformational analysis, carbohydrate mimetics and crown ethers.

    EMAIL logo
Veröffentlicht/Copyright: 10. März 2014

Abstract

This review presents a new strategy for the design of stimuli-responsive liposomes for targeted delivery – the construction of a liposome membrane (lipid bilayer) using amphiphiles able to perform a stimuli-triggered conformational flip (‘flipids’). When done simultaneously by a major or significant part of the bilayer molecules, this massive flip disrupts the liposome membrane and induces a rapid release of the liposome load specifically in response to the initial stimulus. The conformational switches incorporated into the amphiphilic molecules could potentially be controlled by various internal or external factors (pH, metal complexation, light, electric field, etc.). Using this concept, we designed a series of pH-triggerable flipids, and prepared and tested ‘fliposomes’ with extraordinary characteristics: high stability in storage and in serum combined with an instant release of their cargo in response to a weakly acidic medium.

Introduction

Liposomes are particulate vesicles made of one or several lipid bilayers enclosing an aqueous core and able to incorporate a variety of hydrophilic or lipophilic substances within the core or within the lipid membrane, respectively. Liposomes are mainly used as a potent drug delivery system (1–16), as they prolong the life of pharmaceuticals in the bloodstream. This allows the drugs to reach a destination without being decomposed. Therefore, lower doses could be used. For this and other reasons, toxic drugs often show a relevant improvement in their therapeutic index when they are administered as liposomal formulations. Currently, several liposome-based therapeutics are clinically approved (1–6). To provide a more efficient and targeted drug delivery, the liposomes should release their content exactly at the designated place in response to certain factors characteristic for a disease (pH, enzymatic activity, redox potential, temperature) or to a stimulus applied by external devices (temperature, light, radiation, ultrasound, and electric and magnetic field) (1–9, 15–20). These internal or external stimuli initiate a perturbation or even destruction of the liposome membrane (lipid bilayer), which triggers a liposome discharge. Thus, the leakage of the most studied thermosensitive liposomes is caused by the formation of local defects and phase transition in the lipid bilayer (1–4, 8, 12). Another classic approach is chemical or enzymatic cleavage of molecules constituting the liposome membrane (2–5, 9, 10, 16). Liposomes are often rendered stimuli sensitive by incorporation of additional stimuli-responsive components (peptides, proteins, polymers, metal nanoparticles, etc.) in relatively small amounts (1–7, 9–11, 15–24).

A novel promising strategy described here is based on a stimuli-triggered conformational flip performed simultaneously by a significant part of amphiphilic molecules composing the liposome membrane. This transformation does not require chemical cleavage of molecules or their relocation, and it is very fast. The amphiphiles have to be equipped with the stimuli-responsive conformational switches. Mechanical or conformational molecular switches are molecular systems that reversibly change the relative orientation of their parts under external influence. They play a central role in the design of molecular machinery, controllable compounds, and intelligent materials for possible use in many applications, including drug release, new sensor techniques, or information storage and transmission (25–31). Cyclohexane-based molecular devices have been designed as a new type of such switches (29–31).

pH-responsive systems

Acid-induced release is a promising approach to the design of liposomes as drug/gene delivery systems because increased acidity is characteristic for numerous physiological and pathological conditions, including endosome processing, inflammation, ischemia, and solid tumor growth (2–7, 10, 13–15). During the last three decades, this idea developed into many strategies for construction of pH-sensitive liposomes. One of the major approaches employs an acid-triggered change of interactions between molecules in a mixture of non-polar and anionic lipids that results in phase transition (5) and/or formation of domains with ‘leaky’ interfaces (32, 33). Variation of the pKa of anionic lipids and their molar ratio can slightly modify the pH sensitivity. Another popular concept is based on the acidic hydrolysis of ortho-esters, acetals, hydrazones, vinyl ethers, or other acid-labile linkers in lipidic amphiphiles transforming the latter into destabilizing detergents or conical lipids (2, 3, 5, 9, 10, 16, 34, 35). These reactions often require several hours to proceed, which may be a problem for timely drug release. Liposomes can also be made acid sensitive by incorporation into the lipid bilayer of additional minor components: peptides or polymers designed to have a pH-dependent conformational change and change of solubility, thus causing membrane perturbation, pore formation, phase separation, and fusion (2, 4, 10, 19–24). The relatively complicated syntheses of such components and their limited variability may be a disadvantage of this approach.

We recently suggested a novel strategy to render liposomes pH sensitive: a protonation-induced conformational switch of hydrocarbon chains in latent amphiphiles composing the liposome membrane (30, 36–42). Our liposome design is based on a drastic conformational flip performed simultaneously by a significant part of these structurally simple and synthetically accessible pH-sensitive molecules upon protonation. In our studies, the pH-responsive amphiphiles constituted from 25% to 90% (in most experiments, 50 mol%) of the whole lipid composition (36–42). This massive uprising disturbs the liposome membrane instantaneously and induces a rapid release of the liposome payload specifically in response to increased acidity of the medium.

To highlight the key role of the conformational flip in pH-triggered liposome leakage, we introduced the terms ‘flipids’ for amphiphiles equipped with a pH-sensitive conformational switch and ‘fliposomes’ for liposomes composed of this material (30, 36–42). This should not be confused with a lipid translocation (flip-flop) that was involved in many reported lipid phase changes (43–46). The lipid motion from one leaflet of the bilayer to the other is unlikely to significantly contribute to the pH-triggered release from the fliposomes because of its much slower kinetics: hours for flip-flop (44) versus seconds for fliposome leakage.

The promising preliminary results (36) encouraged us to study the physicochemical and biological properties of this novel type of liposomes in more detail. The compatibility of the pH-sensitive conformational trigger with commonly used lipids, ability to control the rate and extent of liposome content release, mechanism of lipid membrane destabilization, and viability of fliposomes for drug delivery have been studied (36–42). Of special interest was the ability of flipids to trigger the PEGylated liposomes due to their prolonged life in blood circulation and successful applications in drug/gene delivery (2, 4, 5, 10, 14, 34).

To address these issues, we designed, prepared, and characterized a series of PEGylated fliposomes (36–42). The pH-driven conformational interconversion of flipids was studied by 1H nuclear magnetic resonance (NMR) titration. pH-triggered release from the fliposomes was measured using the 8-aminonaphthalene-1,3,6 trisulfonic acid/p-xylene-bis-pyridinium bromide (ANTS/DPX) fluorescent assay. Freeze-fracture electron microscopy (FFEM) was used for probing the mechanism of the acid-triggered lipid membrane destabilization. Selected lipid compositions were used to construct liposomes encapsulating a widely used anticancer drug methotrexate (MTX), followed by the characterization of their pH-triggered drug release by equilibrium microdialysis and anticancer activity in HeLa cells (human cervical cancer). By incorporation of 50 mol% flipids into liposome membranes containing also phospholipids and mPEG2000-ceramide, we constructed pH-triggerable fliposomes with extraordinary characteristics: high stability in storage and in serum combined with an instant release of their cargo in response to a weakly acidic medium (36–42). The MTX-loaded liposomes demonstrated much higher cytotoxicity in HeLa cells than the free drug, indicating that they can serve as viable drug delivery systems (37).

Trans-2-aminocyclohexanol-based flipids

The first amphiphiles with the pH-triggerable conformational switch were derivatives of trans-2-aminocyclohexanol 1 that performed an acid-induced ring flip (36–41), spreading their lipophilic tails like peacocks (this conformational change can be dubbed a ‘peacock effect’; Figure 1). Previously, we used the trans-2-aminocyclohexanol moiety to construct the conformationally controlled crown ethers and podands (29, 30, 47, 48).

Figure 1 Protonation-induced conformational flip causes spreading and efficient shortening of the lipophilic tails (a ‘peacock effect’) along with a change of the charge, effective size, and shape of the polar head resulting in a quick perturbation of the lipid bilayer, lipid phase separation, and fast content leakage from the fliposomes (30, 36–41).
Figure 1

Protonation-induced conformational flip causes spreading and efficient shortening of the lipophilic tails (a ‘peacock effect’) along with a change of the charge, effective size, and shape of the polar head resulting in a quick perturbation of the lipid bilayer, lipid phase separation, and fast content leakage from the fliposomes (30, 36–41).

The driving force of this dramatic acid-triggered transition is a strong, protonation-generated intramolecular hydrogen bond of HO···H-N+ type and an electrostatic/dipole-dipole attraction stabilizing a conformer with the gauche form of O-C-C-N fragment (1BH+ in Figure 1). This impulse results in a conformational flip of the cycle that moves the ester groups COOR at the other end of the molecule away from each other into axial positions. The relocation of substituents changes their intra- and intermolecular interactions, for example, their ability to form complexes with cations or to pack into lipid bilayers (depending on the nature of substituents).

We examined the chair-chair flip of the cyclohexane ring in flipids 1 (Figure 1) in solution by the proton NMR, 1H NMR (36–41). The vicinal coupling constants 3JHH between several protons attached to the cyclohexane moiety and their chemical shifts are strongly conformation dependent, which allows an assignment of a predominant conformation and an estimation of the position of equilibrium. To characterize the acid-induced swing of the conformational equilibrium, the changes of 1H NMR spectra were monitored during titration of the diluted d4-methanolic solutions of 1 with d-trifluoroacetic acid (Figure 2). The observation of a single set of well-resolved multiplets in the course of acidification attested to high rates of both the acid-base and the conformational equilibria on the NMR time scale. During the incremental addition of acid, the signal parameters changed significantly, indicating a strong protonation-induced shift of the conformational equilibrium from A to BD+ (∼100% in excess acid). This shift did not occur gradually over the whole course of titration, but happened only within a narrow pH range, the value of which depended on the basicity of the amino group. Using the change of ΔGB-A values for the conformational equilibrium in methanol solution upon addition of acid, we estimated the power of this conformational pH trigger to be ≥11–12 kJ/mol. Remarkably, the intramolecular hydrogen bond HO···H-N+ wins a competition with hydrogen bonding to solvent molecules. Similar results were obtained for aqueous solutions of the model trans-2-aminocyclohexanols equipped with ethyl groups instead of long hydrocarbon chains and therefore slightly soluble in D2O (41).

Figure 2 Change of the signal width (W=ΣJHH) for the proton H4 in 1H NMR due to the conformational switch of the flipids 1a (NRR′=morpholinyl, n=m=1) and 1b (NRR′=CH3OCH2CH2NH, n=m=1) in CD3OD solution caused by titration with d-TFA (■) and then by backward titration of the resulting acidic solution with DBU (□) (39).
Figure 2

Change of the signal width (W=ΣJHH) for the proton H4 in 1H NMR due to the conformational switch of the flipids 1a (NRR′=morpholinyl, n=m=1) and 1b (NRR′=CH3OCH2CH2NH, n=m=1) in CD3OD solution caused by titration with d-TFA (■) and then by backward titration of the resulting acidic solution with DBU (□) (39).

PEGylated liposomes that encapsulated a fluorescent dye (ANTS) and a quencher (DPX) were prepared by the freeze-thawing method. Starting with any particular lipid composition, this method yielded liposomes of reproducible colloidal properties and thus allowed studies on the relationship between the lipid composition of the liposomes and their acid-triggered release of contents (36–41). Noteworthy, after storage at 4°C for >6 months, the optimized liposome formulations (1/POPC/PEG-ceramide, 50:45:5 mol%) did not show noticeable changes in hydrodynamic diameter or polydispersity index, and kept a near-zero ζ-potential. We measured the leakage of ANTS/DPX after injecting a small aliquot of the fliposome preparation into buffer solutions with pH varying from 9.5 to 2.8 as exemplified for the case of two flipids in Figure 3.

Figure 3 pH dependence of the ANTS/DPX leakage from 1/POPC/PEG-ceramide fliposomes made with 50 mol% of flipid 1a, or 1b, or their 1:1 blend (1a+1b).The black curve (▴) was obtained for the 1:1 mixture of fliposomes 1a and 1b. The dashed line is an average of the curves 1a and 1b (39).
Figure 3

pH dependence of the ANTS/DPX leakage from 1/POPC/PEG-ceramide fliposomes made with 50 mol% of flipid 1a, or 1b, or their 1:1 blend (1a+1b).

The black curve (▴) was obtained for the 1:1 mixture of fliposomes 1a and 1b. The dashed line is an average of the curves 1a and 1b (39).

As the initial horizontal part of the titration curves shows, fliposome leakage does not occur in basic (for 1a and 1b) or neutral (for 1a) medium. However, the increase of acidity brings about a dramatic jump of leakage between pH 8.5 and 6.5 in the case of fliposomes containing 50 mol% 1b, and between pH 6.0 and 4.0 in the case of fliposomes with 50 mol% 1a. The relative position of these transitional areas meets the expectation of a higher basicity for the secondary amine 1b as compared with the tertiary amine 1a (in protic solvents). The resulting diagrams are very similar to the NMR titration curves for the conformational flip of amphiphiles 1a and 1b (Figure 2). We consider this similarity as an evidence for the intrinsic dependence of the pH-induced fliposome leakage on the pH-triggered conformational flip of the amphiphiles 1.

Further evidence for the predominant role of conformational flip in the liposome leakage was obtained from the comparison of diastereomeric amphiphiles 1a and 2 (Scheme 1) (36, 37). Compound 2 differs from flipid 1a only in the configuration of one lipid tail substituent; however, for this reason, it is unable to noticeably change its conformation after protonation. Thus, the results for the control 2/POPC/PEG-ceramide liposomes allowed observation of all possible effects of the protonation of morpholine group on the liposome permeability (change of headgroup charge and radius, change of hydrogen bondings, etc.), except for the effects caused by the change of conformation. Therefore, the much larger and faster acid-induced leakage of 1/POPC/PEG-ceramide liposomes than that of 2/POPC/PEG-ceramide liposomes could be attributed to loosening of the liposome membrane caused by the pH-driven conformational change of the lipid tails in 1a. As expected, control liposomes containing neither 1a nor 2 were not responsive to lowered pH (36, 37).

Scheme 1 Absence of protonation-induced conformational flip in amphiphile 2 (36, 37).
Scheme 1

Absence of protonation-induced conformational flip in amphiphile 2 (36, 37).

FFEM allowed us to gain more insight into the mechanism of triggered release from fliposomes. This powerful technique not only can image structures of lipid colloids at nanometer resolution but also takes snapshots of lipid phase transformation by rapid freezing. The unloaded 1/POPC/PEG-ceramide (50:45:5) fliposome formulation was studied both at pH 7.4 and after exposure to pH 5.5 (Figure 4) (37). At both pH values, the samples contained vesicles, although a large part of them had dissipated in the more acidic medium. The diameters of the fliposomes measured on the electron micrographs ranged from 20 to 230 nm, with an average diameter of about 100 nm. At pH 7.4, the fliposome formulation appeared to be composed mostly of unilamellar vesicles (Figure 4A). Upon decrease of pH from 7.4 to 5.5, the remaining fliposomes underwent drastic morphological transformations (Figure 4B,C), including division, fusion, and budding. It appears that a large number of buds pinched off the fliposomes as freely suspended small spherical particles. Defects of lipid packing that resembled long and deep cracks were also observed on some of the dividing fliposome membranes.

Figure 4 FFEM of the 1/POPC/PEG-ceramide fliposome formulation at pH 7.4 (A) and 5 min after adjusting the pH to 5.5 with diluted acetic acid (B, C).Examples of division, buds, and stripes are shown with white, black, and yellow arrows respectively. The bars represent 100 nm (37).
Figure 4

FFEM of the 1/POPC/PEG-ceramide fliposome formulation at pH 7.4 (A) and 5 min after adjusting the pH to 5.5 with diluted acetic acid (B, C).

Examples of division, buds, and stripes are shown with white, black, and yellow arrows respectively. The bars represent 100 nm (37).

On the basis of 1H NMR studies of the pH-induced conformational flip of compounds 1, the fluorometric and ultraviolet studies of the pH-induced leakage of fliposomes, and the FFEM images of the morphological changes of the 1/POPC/PEG-ceramide fliposomes, we proposed a novel mechanism of the pH-triggered content release from fliposomes (37). The protonation generates a strong intramolecular hydrogen bond between the amine and the neighboring hydroxy group, which flips the chair conformation of the cyclohexane ring and mechanically switches its two remote ester groups from equatorial to axial positions (Figure 1). Thereby, the spatial separation of the two lipid tails of 1 is increased, especially at their proximal end (peacock effect). At the distal end, the conformational freedom and the hydrophobic interactions may allow the chains to partially remain packed or to re-pack in the lipid bilayer, although with a decreased packable length of the hydrophobic moiety. Besides this conformational shortening and widening, as flipid 1 is protonated, its head group assumes a positive charge and increases its hydrodynamic size as a result of higher hydration. This leads to additional electrostatic and steric repulsion. All these changes are very fast, and they directly perturb the lipid bilayer and could induce phase separation of the fliposome membrane into thinner and thicker domains that are rich in 1 or in longer POPC molecules, respectively. The encapsulated contents of the fliposomes would then quickly leak through defects between the domains. Some domains of monolayers would subsequently bud from the fliposome membranes as small micelles, as observed in FFEM (Figure 4) (37).

Peacock effect: heads or tails?

A simple synthetic scheme used for preparation of flipids 1 allows substantial flexibility in the design of these molecular devices (30, 36–41). Their parameters can be conveniently modified by structural variation of all substituents, especially the hydrocarbon tails and the amino group in the relatively polar heads of the molecules.

Because the basicity of amines is determined by the nature of groups attached to the nitrogen atom, the flipid pH sensitivity can be tuned by appropriate modification of the amino group structure. This may allow tailoring of fliposomes for the triggered release at certain pH values. Moreover, by mixing such liposome preparations, each with different pH sensitivity, one could potentially obtain a system for simultaneous administration, but independent delivery, of different drugs that would be released each at a specific pH value. Furthermore, the reversibility of conformational flip upon addition of a base (Figure 2) suggests that fliposomes prepared in acidic medium could be used for targeted delivery to the more basic places, for instance, to the small intestine (49). We explored some of these possibilities (39–41).

We synthesized a series of flipids 1 with a variety of ‘heads’ and ‘tails’, and studied the pH dependence of the ANTS/DPX leakage from the corresponding 1/POPC/PEG-ceramide fliposomes (36–41). The measured pH ranges for the leakage varied from pH 9 to pH 4 depending on the nature of the amino group (as shown in Figure 3 for flipids 1a and 1b with different basicity). Thus, the pH sensitivity of a fliposome preparation can be indeed custom-tailored. Very conveniently, the leakage range for each flipid can be accurately predicted from the results of 1H NMR titration (Figure 2).

Intriguing results were obtained when we used a 1:1 blend of flipids 1a and 1b for preparation of (1+2)/POPC/PEG-ceramide fliposomes (39, 40). Because the pH ranges for leakage of the fliposomes based on just one of these triggers practically do not overlap (Figure 3), one could expect the mixture of both flipids to produce an additive effect. That would extend the leakage over a pH range from 8.5 through 4.0 (dashed line in Figure 3 representing an average of the curves for 1a and 1b). Instead, in our experiments, the release was complete by pH 5.5. The first (more basic) part of the experimental curve, which can be credited to the conformational flip of 1b, practically coincides with the expectation (the dashed line). However, the second part of the curve, which was supposed to depend on switching of 1a, goes up and achieves the complete leakage much ‘earlier’ on the acidity scale than expected.

Very similar results have been obtained when we mixed two separate fliposome formulations, one based on 1a (50 mol% of lipid composition) and the second based on 1b (also 50 mol%). We supposed that fliposomes in this mixture of formulations were to respond to the change of pH independently – first those containing the more basic 1b and then those containing 1a. However, in this case, the second kind of fliposomes released its content also at less acidic conditions than expected, and the content release was complete by pH 5.0 (Figure 3).

Thus, when the flipids are blended before the preparation of fliposomes, or when the ready fliposome formulations are mixed, those flipids do not respond to acidification independently. There appears to be some complicated interaction between them. A plausible explanation of the observed effects may be as follows. When the more basic molecules 1b become protonated, the fliposomes that contain them are partially or completely destroyed. The ions 1bH+ thus released into the bulk aqueous media may play a role of a detergent, which permeates other fliposomes and releases their content. In other words, the acid-induced release of fliposome cargo may sometimes have an autocatalytic character. To verify this hypothesis, we performed the standard experiment by injection of the ANTS/DPX fliposome preparation based on 1a into buffer solutions with pH 7.5 and 5.5. Then, we added the dispersion of 1b in identical buffer. The molar amounts of 1a and 1b were equal. The addition of 1b produced no change when the medium had pH 7.5. However, the fluorescence released by the fliposome preparation of 1a in the pH 5.5 buffer increased upon injection of 1b. This observation seems to confirm the ability of protonated compound 1b to act as a detergent.

As the structure of lipid tails is critically important for the stability of lipid bilayers, the ability to release the liposome content could be tuned by appropriate design of these hydrophobic parts. We prepared and characterized the fliposomes containing new trans-2-morpholinocyclohexanol-based flipids equipped with either longer hydrocarbon chains, bent or branched hydrocarbon chains, and compared them with fliposomes based on flipid 1a (38, 40, 41). The pH-driven conformational flip was studied by 1H NMR titration. The 1H NMR and conformational parameters were practically identical between these flipids in all the studied conditions. Thus, the elongation, bending, and branching of the hydrocarbon chains did not produce any noticeable conformational effect in the polar heads of these structures and did not alter their pH sensitivity. (This observation suggests inter alia that in future studies, simpler model compounds with short alkyl tails can be used for a preliminary estimation of conformational properties of the designed flipids.) The pH dependence of the ANTS/DPX leakage from the fliposomes made of these flipids was similar to the NMR titration curve, which confirmed the intrinsic dependence of the pH-induced liposome leakage on the pH-triggered conformational flip.

However, the elongation and branching of the hydrocarbon chains reduced the content release from fliposomes (38, 40, 41). This observation may be considered as an additional confirmation to the proposed mechanism of pH sensitivity, wherein bilayer destabilization starts with an acid-triggered conformational flip of the trans-2-aminocyclohexanol moiety (the polar head group) that increases the spatial separation of the two lipid tails, which results in a set of membrane perturbations that cause the leakage (Figure 1). The tails should be separated the most in the vicinity of the polar head and may partially fold back at the distal end to re-pack in the lipid bilayer. However, the packable length of the tails would decrease (Figure 1), thus inducing the phase separation of the lipid bilayer and liposome leakage. Therefore, the effect of the conformational flip in the polar head on the packing of the tails should relatively decrease as the flipids carry longer tails. On the other hand, membrane perturbation by the conformational flip is generally possible because of a certain degree of order in the hydrophobic part of the lipid bilayer. When some additional disorder (fluidity) is introduced from the very beginning by the asymmetrically branched groups, the effect of the conformational flip on the membrane integrity becomes less significant. These results suggest that a better performance of the flipids may be achieved by making their hydrocarbon chains shorter and/or by introducing additional rigid fragments into their lipid tails. These hypotheses are currently under exploration.

Piperidinol-based flipids

We recently suggested the trans-3,4-bis(acyloxy)-piperidine structure 3 (Scheme 2) as a new platform for the pH-triggered conformational switches (42, 50, 51). The driving force in these devices is the same as in flipids 1: a strong, protonation-generated intramolecular hydrogen bond of O···H-N+ type and an electrostatic/dipole-dipole attraction stabilizing a conformer with the gauche form of the O-C-C-N fragment (3AH+ in Scheme 2).

Scheme 2 Protonation-induced conformational flip, shortening of the lipid tails, and change of the charge, effective size, and shape of the polar head cause a quick perturbation of the lipid bilayer, lipid phase separation, and fast content leakage from the fliposomes (42).
Scheme 2

Protonation-induced conformational flip, shortening of the lipid tails, and change of the charge, effective size, and shape of the polar head cause a quick perturbation of the lipid bilayer, lipid phase separation, and fast content leakage from the fliposomes (42).

Applying the approaches and methods developed for the cyclohexane derivatives 1, we studied the conformational behavior of the piperidine-based flipids 3 and their properties as the lipidic components of fliposomes. The 1H NMR data showed that flipids 3 changed conformation when acidity increased, and this flip occurred between apparent pD 5.5 and 3.5 in d4-methanol solution. The fliposomes comprising flipid 3, POPC and PEG-ceramide, were tested using the ANTS/DPX fluorescent assay and demonstrated a leakage that started at pH 5.6 and became substantial at pH 4.0. The release results achieved thus far with the flipids 3 are below the efficiency of the fliposomes containing trans-2-aminocyclohexanol-based flipids 1. Perhaps, the perturbation of the lipid bilayer is stronger in the case of flipids 1 because the latter have more substituents, which drastically changed their position during the acid-triggered conformational flip (Figure 1).

Targeted drug and gene delivery

To confirm that fliposomes are capable of releasing encapsulated drugs in response to lowered pH, we constructed the 1a/POPC/PEG-ceramide (50:45:5) and the 1a/POPC/PEG-DPPE fliposomes (50:45:5) containing the anionic anticancer drug MTX (37). Both fliposome formulations were subjected to equilibrium microdialysis as a more versatile method to quantify drug release regardless of its fluorescence. The percentage of MTX release was determined by measuring the UV absorbance (λ=306 nm) of the free drug in the solution outside the dialysis bag after the microdialysis had reached equilibrium. None of the fliposome formulations released a significant amount of MTX after dialysis at pH 7.4 for 4 h. Upon dialysis at pH 5.5, both fliposome formulations containing flipid 1a released most of the encapsulated MTX. The concentration of MTX generated in the buffer chamber by far exceeded the typical levels of MTX in patients’ plasma and serum. For comparison, liposomes containing the diastereomeric amphiphile 2 released much less MTX after microdialysis at pH 5.5 (37). Such observations confirm that pH-triggered release of contents from fliposomes (ANTS, MTX, etc.) results mainly from the conformational change of the lipid tails, which takes place in 1a but not in its diastereomer 2.

Finally, to test whether fliposomes can serve as viable drug delivery systems, selected liposome formulations encapsulating the anticancer drug MTX were applied to HeLa cells followed by an assessment of the cell viability using the MTS assay (37). HeLa cells treated with 1 μm (final concentration) of MTX solution retained 65.0±4.3% viability, consistent with prior reports on the anticancer activity of the free drug. In comparison, HeLa cells treated with MTX-loaded fliposomes built of 1a/POPC/PEG-ceramide (50:45:5) showed a significantly lower viability (39.4±4.3%), indicating the superior anticancer cytotoxicity of such liposomal formulations of MTX compared with the free drug at the same dosage. The HeLa cells treated with MTX-loaded liposomes containing the diastereomeric analogue 2 instead of 1a retained most of the viability, thus highlighting the importance of the conformational flip in enhancing the cytotoxicity of the payload MTX. The cell culture media were buffered at ∼pH 7.4; therefore, discharge of the 1a/POPC/PEG-ceramide/MTX fliposomes most likely took place in the acidic endosomal compartment of the HeLa cells after their cellular uptake. Because MTX is too hydrophilic to passively diffuse across biomembranes, it is likely that the fliposomes not only released MTX in response to the lowered pH in the endosome but also facilitated the diffusion of the released MTX from the endosome to the cytosol. One possible mechanism of such facilitated diffusion could be a destabilization of the endosomal membranes by the micelles generated from the destruction of fliposomes.

It is known that pH sensitivity improves the efficiency of a number of gene delivery systems, including viral vectors, polyplexes, and lipoplexes (10, 52). Inspired by the drug delivery experiments, we started exploring the possibility to use flipids 1 as helper lipids in cationic lipoplexes for gene transfection. The preliminary tests using luciferase assay produced very promising results (41, 53), and this project is currently in progress.

Metal-responsive systems

Another group of researchers suggested 3,7-diazabicyclo[3.3.1]nonan-9-one 4 (Scheme 3) as a conformational switch for the lipid-like amphiphiles (54, 55). In neutral and weakly basic media, these compounds adopt a chair-boat conformation (4B) that flips into a chair-chair form (4C) in the presence of acid or upon complexation with metal cations.

Scheme 3 Protonation- or complexation-induced conformational flip causes a perturbation of the lipid bilayer and the content leakage from fliposomes (54, 55).
Scheme 3

Protonation- or complexation-induced conformational flip causes a perturbation of the lipid bilayer and the content leakage from fliposomes (54, 55).

The liposomes with compounds 4 incorporated into membrane released their fluorescent cargo in response to the addition of aqueous CuSO4. This effect was attributed to the complexation-induced change of conformation, which affected the package of lipids in the bilayer and sharply increased the permeability of the liposomal membrane. The liposomal containers with such flipids can be used for the encapsulation and following release of drugs that control the copper level in patients with various pathologies, e.g., hepato-cerebral dystrophy (Wilson disease) (54, 55). Interestingly, the acid- or metal-triggered flip brings the hydrocarbon tails in 4 closer to each other – the change in the direction opposite to the peacock effect observed for flipids 1 (see above).

Phototriggerable flipids

In terms of the dramatic change of molecular shape and size, the conformational switch of flipids is similar to the cis-trans photoisomerization of azobenzene derivatives (56). Mechanistically, the isomerization of the latter can be described as a conformational flip proceeding through inversion at the nitrogen atom (Scheme 4) (56). Therefore, the lipidic amphiphiles equipped with the azobenzene photo-switch (1, 57, 58) match the definition of flipids presented above.

Scheme 4 Photoisomerization of the lipidic azobenzene derivatives causes a perturbation of the lipid bilayer and the content leakage from liposomes (1, 57, 58).
Scheme 4

Photoisomerization of the lipidic azobenzene derivatives causes a perturbation of the lipid bilayer and the content leakage from liposomes (1, 57, 58).

The artificial lipids containing the azobenzene moiety were synthesized and demonstrated the ability to cause a disruption of the bilayer packing and the release of entrapped solutes from liposomes upon light irradiation either in vitro or in cell cultures (1, 57, 58). The phototriggerable lipids synthesized thus far are triggerable by light in the UV or visible range, which has a limited ability to penetrate into biological tissues and deliver sufficient photon energy. This poses certain limitations on the potential in vivo application of such delivery systems (1). However, a new approach that is currently gaining momentum may help overcome this problem: the use in photochemotherapy of compounds and materials designed to upconvert near-infrared light into higher-energy visible photons. It has been applied to activation of ruthenium complexes suggested as prodrugs for photoactivatable anticancer therapy (59, 60). Perhaps, the same or similar compounds could help trigger the azobenzene photoisomerization in lipid amphiphiles by infrared light.

Outlook

The first attempts to design stimuli-triggerable, especially pH-triggerable, amphiphiles equipped with conformational switches (flipids) and to use them as construction material for stimuli-responsive liposomes (fliposomes) produced encouraging results. Considering a broad variety of known and potential conformational switches, this strategy promises to be the subject of intense investigation in the following years.


Corresponding author: Vyacheslav V. Samoshin, Department of Chemistry, College of the Pacific, University of the Pacific, Stockton, CA 95211, USA, e-mail:

About the author

Vyacheslav V. Samoshin

Vyacheslav V. Samoshin graduated with an Honorable Diploma (MS) from Moscow State University, USSR, in 1974. At the same University, he defended his PhD dissertation in Organic Chemistry under guidance from Academician Nikolay S. Zefirov in 1982, and defended his Doctor of Chemical Sciences dissertation in 1991. He worked as a researcher in the Department of Chemistry, MSU, and since 1992 as Professor and Head of the Division of Organic Chemistry at the Moscow State Academy of Fine Chemical Technology. He was awarded the title Honorary Professor by this Academy in 2012. He took his present position as Professor of Chemistry at the University of the Pacific, Stockton, California in 1999. His scientific interests include molecular switches, especially in application to liposome design, conformational analysis, carbohydrate mimetics and crown ethers.

Acknowledgments

The author is deeply grateful to his coworkers, Drs. Nataliya M. Samoshina, Xin Guo, Andreas H. Franz, Vyacheslav A. Chertkov, and Galina V. Grishina, and to his former graduate students, Drs. Yu Zheng, Xin Liu, and Barbora Brazdova. This work was supported in part by Teva US Scholar Grant (2009–2012) administered by the American Chemical Society, by the National Science Foundation Major Research Instrumentation grant MRI-CHE-0722654 (funding of JEOL ECA-600 NMR-spectrometer), and by the Department of Chemistry, University of the Pacific.

References

1. Puri A. Phototriggerable liposomes: current research and future perspectives. Pharmaceutics 2014; 6: 1–25.10.3390/pharmaceutics6010001Suche in Google Scholar PubMed PubMed Central

2. Perche F, Torchilin VP. Recent trends in multifunctional liposomal nanocarriers for enhanced tumor targeting. J Drug Deliv 2013; 2013: 705265, 32 pp.10.1155/2013/705265Suche in Google Scholar PubMed PubMed Central

3. Ganta S, Devalapally H, Shahiwala A, Amiji M. A review of stimuli-responsive nanocarriers for drug and gene delivery. J Control Release 2008; 126: 187–204.10.1016/j.jconrel.2007.12.017Suche in Google Scholar PubMed

4. Bibi S, Lattmann E, Mohammed AR, Perrie Y. Trigger release liposome systems: local and remote controlled delivery? J Microencapsulation 2012; 29: 262–76.10.3109/02652048.2011.646330Suche in Google Scholar PubMed

5. Morilla MJ, Romero EL. Liposomal pH-sensitive nanomedicines in preclinical development. In: Reisner DE, editor. Bionanotechnology II. Boca Raton: CRC Press, 2012: 383–413.Suche in Google Scholar

6. Lehner R, Wang X, Marsch S, Hunziker P. Intelligent nanomaterials for medicine: carrier platforms and targeting strategies in the context of clinical application. Nanomedicine 2013; 9: 742–57.10.1016/j.nano.2013.01.012Suche in Google Scholar PubMed

7. Lehner R, Wang X, Wolf M, Hunziker P. Designing switchable nanosystems for medical application. J Control Release 2012; 161: 307–16.10.1016/j.jconrel.2012.04.040Suche in Google Scholar PubMed

8. Oude Blenke E, Mastrobattista E, Schiffelers RM. Strategies for triggered drug release from tumor targeted liposomes. Expert Opin Drug Deliv 2013; 10: 1399–410.10.1517/17425247.2013.805742Suche in Google Scholar PubMed

9. Huang Z, Szoka FC Jr. Bioresponsive liposomes and their use for macromolecular delivery. In: Gregoriadis G, editor. Liposome technology, 3rd ed., vol. 1. New York: Informa Healthcare, 2007: 165–96.Suche in Google Scholar

10. Hatakeyama H, Akita H, Harashima H. The polyethyleneglycol dilemma: advantage and disadvantage of PEGylation of liposomes for systemic genes and nucleic acids delivery to tumors. Biol Pharm Bull 2013; 36: 892–9.10.1248/bpb.b13-00059Suche in Google Scholar PubMed

11. Yue X, Dai Z. Recent advances in liposomal nanohybrid cerasomes as promising drug nanocarriers. Adv Colloid Interface Sci 2014. Epub ahead of print; doi: 10.1016/ j.cis.2013.11.014.Suche in Google Scholar

12. Ta T, Porter TM. Thermosensitive liposomes for localized delivery and triggered release of chemotherapy. J Control Release 2013; 169: 112–25.10.1016/j.jconrel.2013.03.036Suche in Google Scholar PubMed PubMed Central

13. Kristl J. Implication of nanotechnology on development of medicines. Farmacevtski Vestnik 2012; 63: 67–73.Suche in Google Scholar

14. Patel S, Bhirde AA, Rusling JF, Chen X, Gutkind JS, Patel V. Nano delivers big: designing molecular missiles for cancer therapeutics. Pharmaceutics 2011; 3: 34–52.10.3390/pharmaceutics3010034Suche in Google Scholar PubMed PubMed Central

15. Alaouie AM, Sofou S. Liposomes with triggered content release for cancer therapy. J Biomed Nanotechnol 2008; 4: 234–44.10.1166/jbn.2008.335Suche in Google Scholar

16. Guo X, Szoka FC Jr. Chemical approaches to triggerable lipid vesicles for drug and gene delivery. Acc Chem Res 2003; 36: 335–41.10.1021/ar9703241Suche in Google Scholar PubMed

17. Hamano N, Negishi Y, Omata D, Takahashi Y, Manandhar M, Suzuki R, Maruyama K, Nomizu M, Aramaki Y. Bubble liposomes and ultrasound enhance the antitumor effects of AG73 liposomes encapsulating antitumor agents. Mol Pharmaceutics 2013; 10: 774–9.10.1021/mp300463hSuche in Google Scholar PubMed

18. Yi J, Barrow AJ, Yu N, O’Neill BE. Efficient electroporation of liposomes doped with pore stabilizing nisin. J Liposome Res 2013; 23: 197–202.10.3109/08982104.2013.788024Suche in Google Scholar PubMed

19. Kaiden T, Yuba E, Harada A, Sakanishi Y, Kono K. Dual signal-responsive liposomes for temperature-controlled cytoplasmic delivery. Bioconjugate Chem 2011; 22: 1909–15.10.1021/bc2000353Suche in Google Scholar PubMed

20. Yuba E, Kono Y, Harada A, Yokoyama S, Arai M, Kubo K, Kono K. The application of pH-sensitive polymer-lipids to antigen delivery for cancer immunotherapy. Biomaterials 2013; 34: 5711–21.10.1016/j.biomaterials.2013.04.007Suche in Google Scholar PubMed

21. Li W, Nicol F, Szoka FC. GALA: a designed synthetic pH-responsive amphipathic peptide with applications in drug and gene delivery. Adv Drug Delivery Rev 2004; 56: 967–85.10.1016/j.addr.2003.10.041Suche in Google Scholar PubMed

22. Lee HM, Chmielewski J. Liposomal cargo unloading induced by pH-sensitive peptides. J Pept Res 2005; 65: 355–63.10.1111/j.1399-3011.2005.00229.xSuche in Google Scholar PubMed

23. Zhou W, An X, Wang J, Shen W, Chen Z, Wang X. Characteristics, phase behavior and control release for copolymer-liposome with both pH and temperature sensitivities. Colloids Surf A 2012; 395: 225–32.10.1016/j.colsurfa.2011.12.034Suche in Google Scholar

24. Cho EC, Lim HJ, Kim HJ, Son ED, Choi HJ, Park JH, Kim J-W, Kim J. Role of pH-sensitive polymer-liposome complex in enhancing cellular uptake of biologically active drugs. Mater Sci Eng C 2009; 29: 774–8.10.1016/j.msec.2008.07.014Suche in Google Scholar

25. Schneider H-J, editor. Applications of supramolecular chemistry. Boca Raton: CRC Press, 2012.Suche in Google Scholar

26. Shahinpoor M, Schneider H-J, editors. Intelligent materials. Cambridge, UK: RSC Publishing, 2008.10.1039/9781847558008Suche in Google Scholar

27. Kay ER, Leigh DA, Zerbetto F. Synthetic molecular motors and mechanical machines. Angew Chem Int Ed 2007; 46: 72–191.10.1002/anie.200504313Suche in Google Scholar PubMed

28. Feringa BL, editor. Molecular switches. Weinheim, Germany: Wiley-VCH, 2001.10.1002/3527600329Suche in Google Scholar

29. Samoshin VV. Cyclohexane-based conformationally controlled crowns and podands. Mini-Rev Org Chem 2005; 2: 225–35.10.2174/1570193054368882Suche in Google Scholar

30. Samoshin VV. Conformational control of cyclohexane derivatives by external stimuli. Rev J Chem 2011; 1: 250–74.10.1134/S2079978011020026Suche in Google Scholar

31. Costero AM, Parra M, Gil S, Andreu MR. Multichannel sensors based on biphenyl and cyclohexane conformational changes. Springer Ser Chem Sens Biosens 2013; 12: 1–32.Suche in Google Scholar

32. Mamasheva E, O’Donnell C, Bandekar A, Sofou S. Heterogeneous liposome membranes with pH-triggered permeability enhance the in vitro antitumor activity of folate-receptor targeted liposomal doxorubicin. Mol Pharmaceutics 2011; 8: 2224–32.10.1021/mp200079ySuche in Google Scholar PubMed

33. Bandekar A, Zhu C, Gomez A, Menzenski MZ, Sempkowski M, Sofou S. Masking and triggered unmasking of targeting ligands on liposomal chemotherapy selectively suppress tumor growth in vivo. Mol Pharmaceutics 2013; 10: 152–60.10.1021/mp3002717Suche in Google Scholar PubMed

34. Kim H-K, Van dBJ, Hyun S-H, Thompson DH. Acid-triggered release via dePEGylation of fusogenic liposomes mediated by heterobifunctional phenyl-substituted vinyl ethers with tunable pH-sensitivity. Bioconjugate Chem 2012; 23: 2071–7.10.1021/bc300266ySuche in Google Scholar PubMed PubMed Central

35. Wehunt MP, Winschel CA, Khan AK, Guo TL, Abdrakhmanova GR, Sidorov V. Controlled drug-release system based on pH-sensitive chloride-triggerable liposomes. J Liposome Res 2013; 23: 37–46.10.3109/08982104.2012.727423Suche in Google Scholar PubMed

36. Brazdova B, Zhang N, Samoshin VV, Guo X. trans-2-Aminocyclohexanol as a pH-sensitive conformational switch in lipid amphiphiles. Chem Commun 2008; 4774–6.10.1039/b807704eSuche in Google Scholar PubMed

37. Samoshina NM, Liu X, Brazdova B, Franz AH, Samoshin VV, Guo X. Fliposomes: pH-sensitive liposomes containing a trans-2-morpholinocyclohexanol-based lipid that performs a conformational flip and triggers an instant cargo release in acidic medium. Pharmaceutics 2011; 3: 379–405.10.3390/pharmaceutics3030379Suche in Google Scholar

38. Zheng Y, Liu X, Samoshina NM, Chertkov VA, Franz AH, Guo X, Samoshin VV. Fliposomes: pH-controlled release from liposomes containing new trans-2-morpholinocyclohexanol-based amphiphiles that perform a conformational flip and trigger an instant cargo release upon acidification. Nat Prod Commun 2012; 7: 353–8.10.1177/1934578X1200700317Suche in Google Scholar

39. Liu X, Zheng Y, Samoshina NM, Franz AH, Guo X, Samoshin VV. Fliposomes: pH-triggered conformational flip of new trans-2-aminocyclohexanol-based amphiphiles causes instant cargo release in liposomes. J Liposome Res 2012; 22: 319–28.10.3109/08982104.2012.698420Suche in Google Scholar

40. Liu X. Fliposomes: pH-sensitive liposomes comprising novel trans-2-aminocyclohexanol-based amphiphiles as conformational switches for the liposome membrane. PhD Dissertation, University of the Pacific, Stockton, CA, 2013.Suche in Google Scholar

41. Zheng Y. Synthesis and conformational study of trans-2-aminocyclohexanol-based pH-triggered molecular switches and their application in gene delivery. PhD Dissertation, University of the Pacific, Stockton, CA, 2013.Suche in Google Scholar

42. Samoshin AV, Veselov IS, Chertkov VA, Yaroslavov AA, Grishina GV, Samoshina NM, Samoshin VV. Fliposomes: new amphiphiles based on trans-3,4-bis(acyloxy)-piperidine able to perform a pH-triggered conformational flip and cause an instant cargo release from liposomes. Tetrahedron Lett 2013; 54: 5600–4.10.1016/j.tetlet.2013.07.156Suche in Google Scholar

43. Contreras FX, Sanchez-Magraner L, Alonso A, Goni FM. Transbilayer (flip-flop) lipid motion and lipid scrambling in membranes. FEBS Lett 2010; 584: 1779–86.10.1016/j.febslet.2009.12.049Suche in Google Scholar

44. Demina T, Grozdova I, Krylova O, Zhirnov A, Istratov V, Frey H, Kautz H, Melik-Nubarov N. Relationship between the structure of amphiphilic copolymers and their ability to disturb lipid bilayers. Biochemistry 2005; 44: 4042–54.10.1021/bi048373qSuche in Google Scholar

45. Schreier S, Malheiros SVP, de Paula E. Surface active drugs: self-association and interaction with membranes and surfactants. Physicochemical and biological aspects. Biochim Biophys Acta Biomembr 2000; 1508: 210–34.10.1016/S0304-4157(00)00012-5Suche in Google Scholar

46. Yaroslavov AA, Melik-Nubarov NS, Menger FM. Polymer-induced flip-flop in biomembranes. Acc Chem Res 2006; 39: 702–10.10.1021/ar050078qSuche in Google Scholar PubMed

47. Samoshin VV, Chertkov VA, Gremyachinskiy DE, Dobretsova EK, Shestakova AK, Vatlina LP. trans-2-Aminocyclohexanols as pH-triggers for conformationally controlled crowns and podands. Tetrahedron Lett 2004; 45: 7823–6.10.1016/j.tetlet.2004.09.004Suche in Google Scholar

48. Samoshin VV, Brazdova B, Chertkov VA, Gremyachinskiy DE, Shestakova AK, Dobretsova EK, Vatlina LP, Yuan J, Schneider H-J. trans-2-Aminocyclohexanols as pH-triggered molecular switches. ARKIVOC 2005; 129–41.10.3998/ark.5550190.0006.410Suche in Google Scholar

49. Chourasia MK, Jain SK. Pharmaceutical approaches to colon targeted drug delivery systems. J Pharm Pharm Sci 2003; 6: 33–66.Suche in Google Scholar

50. Samoshin AV, Huynh L, Tran C, Samoshin VV. trans-3,4-Diacetoxypiperidine as a prototype of novel pH-triggered molecular switches. J Undergrad Chem Res 2011; 10: 50–5.Suche in Google Scholar

51. Samoshin AV, Veselov IS, Huynh L, Shestakova AK, Chertkov VA, Grishina GV, Samoshin VV. trans-3,4-Diacetoxypiperidine as a model for novel pH-triggered conformational switches. Tetrahedron Lett 2011; 52: 5375–8.10.1016/j.tetlet.2011.08.038Suche in Google Scholar

52. Jones CH, Chen C-K, Ravikrishnan A, Rane S, Pfeifer BA. Overcoming nonviral gene delivery barriers: perspective and future. Mol Pharmaceutics 2013; 10: 4082–98.10.1021/mp400467xSuche in Google Scholar PubMed PubMed Central

53. Liu X, Zheng Y, Samoshina NM, Franz AH, Samoshin VV, Guo X. Luciferase gene transfection mediated by cationic liposomes comprising novel trans-2-aminocyclohexanol-based amphiphiles. 243rd ACS National Meeting & Exposition. San Diego, CA: American Chemical Society, 2012, MEDI-414.Suche in Google Scholar

54. Veremeeva PN, Lapteva VL, Palyulin VA, Davydov DA, Yaroslavov AA, Zefirov NS. Novel amphiphilic compounds for the design of stimulus-sensitive liposomal containers. Dokl Chem 2012; 447: 275–7.10.1134/S0012500812120051Suche in Google Scholar

55. Veremeeva PN, Lapteva VL, Palyulin VA, Sybachin AV, Yaroslavov AA, Zefirov NS. Bispidinone-based molecular switches for construction of stimulus-sensitive liposomal containers. Tetrahedron 2014; 70: 1408–11.10.1016/j.tet.2014.01.012Suche in Google Scholar

56. Merino E, Ribagorda M. Control over molecular motion using the cis-trans photoisomerization of the azo group. Beilstein J Org Chem 2012; 8: 1071–90.10.3762/bjoc.8.119Suche in Google Scholar PubMed PubMed Central

57. Liu X-M, Yang B, Wang Y-L, Wang J-Y. Photoisomerizable cholesterol derivatives as photo-trigger of liposomes: effect of lipid polarity, temperature, incorporation ratio, and cholesterol. Biochim Biophys Acta Biomembr 2005; 1720: 28–34.10.1016/j.bbamem.2005.10.016Suche in Google Scholar PubMed

58. Kauscher U, Samanta A, Ravoo BJ. Photoresponsive vesicle permeability based on intramolecular host-guest inclusion. Org Biomol Chem 2014; 12: 600–6.10.1039/C3OB41893FSuche in Google Scholar

59. Ruggiero E, Habtemariam A, Yate L, Mareque-Rivas JC, Salassa L. Near infrared photolysis of a Ru polypyridyl complex by upconverting nanoparticles. Chem Commun 2014; 50: 1715–8.10.1039/c3cc47601dSuche in Google Scholar PubMed

60. Askes SHC, Bahreman A, Bonnet S. Activation of a photodissociative ruthenium complex by triplet-triplet annihilation upconversion in liposomes. Angew Chem Int Ed 2014; 53: 1029–33.10.1002/anie.201309389Suche in Google Scholar PubMed

Received: 2014-1-16
Accepted: 2014-2-3
Published Online: 2014-3-10
Published in Print: 2014-5-31

©2014 by Walter de Gruyter Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Heruntergeladen am 19.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/bmc-2014-0002/html
Button zum nach oben scrollen