• No results found

University of Groningen DNA nanotechnology as a tool to manipulate lipid bilayer membranes Meng, Zhuojun

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen DNA nanotechnology as a tool to manipulate lipid bilayer membranes Meng, Zhuojun"

Copied!
128
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

DNA nanotechnology as a tool to manipulate lipid bilayer membranes

Meng, Zhuojun

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2017

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Meng, Z. (2017). DNA nanotechnology as a tool to manipulate lipid bilayer membranes. University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

DNA nanotechnology as a tool to

manipulate lipid bilayer membranes

(3)

DNA nanotechnology as a tool to manipulate lipid bilayer membranes

Zhuojun Meng PhD thesis

University of Groningen October 2017

Zernike Institute PhD thesis series 2017-22 ISSN: 1570-1530

ISBN: 978-90-367-9976-8(printed version) ISBN: 978-90-367-9975-1(electronic version)

The research described in thesis was carried out in Polymer Chemistry and Bioengineering group at Zernike Institute for Advanced Materials, University of Groningen, The Netherlands. This work was financially supported by the Chinese Scholarship Council (CSC), the University of Groningen and the Netherlands Organization for Science Research (NWO).

Cover design by: Zhuojun Meng Printed by: Ridderprint BV

(4)

DNA nanotechnology as a tool to

manipulate lipid bilayer membranes

PhD thesis

to obtain the degree of PhD at the University of Groningen

on the authority of the Rector Magnificus Prof. E. Sterken

and in accordance with the decision by the College of Deans. This thesis will be defended in public on

Friday 13 October 2017 at 16.15 hours

by

Zhuojun Meng

born on 5 May 1987 in Henan, China

(5)

Supervisor

Prof. A. Herrmann

Assessment committee

Prof. S. Vogel

Prof. A. M. van Oijen Prof. D. J. Slotboom

(6)

Dedicated this book to myself

and my best friend

(7)
(8)

Contents

Chapter 1

Functionalization of LipidBilayer Membranes ... 9

1. 1 Lipid bilayer membranes ... 10

1.2 Classification and Preparation of Liposomes ... 12

1.3 Modification and Applications of liposomes ... 14

1.4 Motivation and Thesis Overview ... 24

References ... 26

Chapter 2 Stability Study of Lipid-DNA on the Liposomal Membrane ... 31

2.1 Introduction ... 32

2.2 Results and Discussion ... 35

2.3 Conclusion ... 42

2.4 Experimental Section ... 42

References ... 48

Chapter 3 Efficient Fusion of Liposomes by Nucleobase Quadruple-Anchored DNA .. 51

3.1 Introduction ... 52

3.2 Results and Discussion ... 54

3.3 Conclusion ... 63

3.4 Experimental Section ... 65

(9)

Chapter 4

DNA Replacement and Hybridization Chain Reaction on the Surface of

Liposome Membrane ... 73

4.1 Introduction ... 74

4.2 Results and Discussion ... 76

4.4 Experimental Section ... 84

References ... 88

Chapter 5 Performing DNA Nanotechnology Operations on a Zebrafish Surface ... 91

5.1 Introduction ... 92

5.2 Results and Discussion ... 94

5.3 Conclusion ... 101 5.4 Experiment Section ... 103 References ... 105 Summary ... 108 Samenvatting ... 114 Acknowledgements ... 119

(10)

Chapter 1

Functionalization of

Lipid Bilayer Membranes

(11)

10

1. 1 Lipid bilayer membranes

Lipids play an important role in the physiology and pathophysiology of living systems why they are produced, transported, and recognized by the concerted actions of numerous enzymes, binding proteins, and receptors.1

Micelles are formed by the aggregation of single-chain lipids in a polar solvent (such as water) beyond a particular concentration, known as Critical Micelle Concentration (CMC) (Fig. 1.1A). Therefore, the micelle formation and stability are highly dependent on the lipid concentration and solvent composition (Fig. 1.1B).

Two-chain lipids can hardly be packed into micelles due to the bulky hydrophobic part. They usually form a lipid bilayer membrane, which is a thin polar sheet made of two layers of lipid molecules and is characterized by hydrophobic tails facing inwards towards each other and hydrophilic head groups facing outwards to associate with aqueous solution.2 At this

moment, the hydrophobic parts of the molecules are still in contact with water, which leads to an energetically unfavorable state of the bilayer. This is overcome through folding of the bilayer membrane into a liposome with closed edges (Fig. 1.1C).3,4

(12)

Functionalization of Lipid Bilayer Membranes

11 Fig. 1.1 (A) Surface tension as a function of the surfactant concentration. Schematic structure of a micelle (B) and a liposome (C). (Fig. 1.1 C was adapted from reference 4)

(13)

12

1.2 Classification and Preparation of Liposomes

Depending on the number of bilayers, liposomes can be classified into two categories: unilamellar vesicles (ULV) and multilamellar vesicles (MLV). Unilamellar vesicles can also be classified into three categories on the basis of their sizes, which can vary from nanometer to micrometer range: small unilamellar vesicles (SUV), large unilamellar vesicles (LUV) and giant unilamellar vesicles (GUV). GUVs also include other morphologies such as multilamellar vesicles (MLV), which consist of SUVs or multiple concentric bilayers (Fig. 1.2).

Fig. 1.2 Schematic structure of unilamellar and multilamellar liposomes.

There are four classical methods to prepare liposomes, differing in the way how the lipids are dried from organic solvent and then redispersed in aqueous buffer.5 These steps can be performed individually or jointly.6

These four methods are:

1. Hydration of a Thin Lipid Film.7

2. Reverse-Phase Evaporation Technique.8

3. Solvent (Ether or Ethanol) Injection Technique.9,10

(14)

Functionalization of Lipid Bilayer Membranes

13 Since the “Hydration of a Thin Lipid Film” method, is widespread used and easy to handle, it is explained here in more details. Firstly, the lipids are dissolved and mixed in an organic solvent to assure a homogeneous mixture. Once the lipids are thoroughly dispersed in the organic solvent, the solvent is removed using a dry nitrogen stream in a fume hood to yield a lipid film. The lipid film is dried to remove residual organic solvent by using a vacuum desiccator overnight. Afterwards, hydration of the dry lipid film is accomplished by stirring in an aqueous buffer. The temperature of the hydrating buffer should be higher than the gel-liquid crystal transition temperature (Tc) of the lipid. Subsequently, several stirring (above the Tc) and freeze-thawing cycles of the swelling multilayer sample results in MLVs. Finally, the sample is extruded multiple times using an extruder and polycarbonate membranes to obtain unilamellar vesicles (LUVs or SUVs).

Fig. 1.3 shows the classical hydration method of liposome preparation.

Fig. 1.3 Schematic diagram of liposome preparation method. (Schematic obtained from www. avantilipids.com)

(15)

14

1.3 Modification and Applications of liposomes

The first discovery of liposomes in 1964 by A. D. Bangham12 was the

starting point for these self-assembled containers to become a multifunctional tool in biology, biochemistry and medicine today. Because of the structure, charge, chemical composition and colloidal size can be well controlled by preparation methods, liposomes can be useful in various applications. Vesicles can also be prepared from natural substances and are therefore in many cases nontoxic, biodegradable, biocompatible, targetable and non-immunogenic.13 Due to these properties, liposomes can be used as

drug14-16 protein, plasmid17 and gene18-21 delivery vehicles in medicine and

diagnosis.

1.3.1 Loading and surface modification

Molecular interactions between the cargo and the lipid bilayer membrane play an important role on liposome formation and cargo encapsulation.22

Liposomes consist of an aqueous core surrounded by a lipid bilayer, sectioning off two separate inner areas. They can carry hydrophobic molecules in their hydrocarbon tail region (between the phospholipid bilayer), or hydrophilic molecules in the core and direct the cargo to the required diseased site in the body with some targeting moieties on the surface.23 The thickness of the lipid bilayer is around 4 to 10 nm, which is a

natural barrier for many substances such as sugars and proteins.24 But

small hydrophilic substances such as water, gases, ammonia and glycerol can penetrate freely through the bilayer.25-27 Some large hydrophilic

substances can be encapsulated in the water core of the liposome during liposome preparation using the common thin layer hydration method. Cationic liposomes, which are made of positively charged lipids, appear to be better suited for DNA delivery due to the natural charge-charge interaction between the positively charged lipid head groups and the negatively charged phosphate groups of the DNA-backbone.28,29 Due to

their favorable interactions with negatively charged DNA and cell membranes,30-33 cationic liposome–DNA complexes are increasingly being

researched for their use in gene therapy and nucleic acid release.34,35 In

(16)

Functionalization of Lipid Bilayer Membranes

15 tissues, the use of targeted liposomes with surface modification has been suggested.

Surface modification of liposomes with controlled propertied requires the chemical conjugation of peptides, DNA, antibodies or other targeting molecules. Moreover, some “smart” vesicle designs allow the release of the encapsulated cargo by incorporation of transport channels.36-39 Both

chemical attachment and physical interactions can be used to achieve surface modification (Fig. 1.4A).

Fig. 1.4 (A) Schematic representation of liposomes surface modifications. (B) Interaction of the particle with cell surface antigens and receptors.40 (C) Scheme of tetrac tagged liposome and

enhanced delivery by the ligand-mediated targeting strategy.41(Fig. 1.4 B was adapted from

reference 40. Fig. 1.4 C was adapted from reference 41)

To realize active targeting, the liposome surface can be coated with ligands or antibodies that will confer cell type-specificity to ensure that the liposomes are internalized and that their content is released, improving the efficacy and reducing side effects over non-targeted cells (Fig. 1.4B).40 For

(17)

16

to integrin αvβ3, was used for the surface modification of liposomes and successfully enhanced the tumor-targeting ability of PEGylated liposome (Fig. 1.4C).41 Although the physical properties of liposomes were not

significantly changed, tetrac-tagged liposomes showed significantly higher cancer cell localization than the unmodified PEGylated liposome, and tumor growth was effectively retarded. The ligand-mediated targeting strategy could provide better therapeutic effects with more accurate delivery of nanoparticles.

1.3.1.1 Membrane fusion

Surface modification of lipid bilayers can also be used for membrane fusion which is an essential process of life resulting in the highly regulated transport of bio-molecules both between and within cells.42-44 Membrane

fusion is an essential but not a spontaneous process as free energy is required to overcome the electrostatic and steric repulsions between two merging membrane surfaces and to break the hydration shell.45,46 A highly

conserved protein machinery, known as SNARE proteins (soluble N-ethylmaleimide sensitive factor attachment protein receptors), facilitates the communication within a cell.47-49 The SNAREs from synaptic vesicles

interact with the SNAREs from the target membrane to form a coiled-coil bundle of four helices, pulling the membranes tightly together and initiating fusion.

Design and construction of simplified artificial model systems mimicking natural systems are one of the most promising approaches for studying complex biological mechanisms.50 Several of these systems have been

reported for realizing membrane fusion, such as DNA51-53, peptides54, 55,

enzymes56 and polymers57. Yang et al. designed an artificial biorthogonal

targeting system that was able to target liposomes and other nanoparticles efficiently to the tissue of interest by using coiled coil forming peptides, E4[(EIAALEK)4] (E4) and K4[(KIAALKE)4] (K4) (Fig. 1.5C), which are

known to trigger liposomal membrane fusion when tethered to lipid vesicles in the form of lipopeptides.58 The same group proved that E4

peptide-modified liposomes could deliver far-red fluorescent dye TOPRO-3 iodide (E4-Lipo-TP3) and doxorubicin (E4-Lipo-DOX) into HeLa cells

(18)

Functionalization of Lipid Bilayer Membranes

17 expressing K4 peptide (HeLa-K) on the surface. Then, Lipo-TP3 and E4-Lipo-DOX were injected into zebrafish xenografts of HeLa-K (Fig. 1.5A, B). The results showed that E4-liposomes delivered TP3 to the implanted HeLa-K cells (Fig. 1.5D), and E4-Lipo-DOX could suppress cancer proliferation in the xenograft when compared to nontargeted conditions. These data demonstrated that coiled-coil formation enables drug selectivity and efficacy in vivo.

Fig. 1.5 Drug Delivery by E4/K4 Coiled-Coil Formation in Cells (A) and Zebrafish (B). (C) Schematic representation of coil structure between peptides E and K. (D) E4/K4 coiled-coil formation allows delivering the content in the liposome to cancer cells in the xenograft zebrafish.58 (This figure was reproduced with permission from reference58)

1.3.1.2 Controlled release

Conventional liposomes (Fig. 1.6A) are easily recognized by the mononuclear phagocyte system and are rapidly cleared from the blood stream.59 Many methods have been suggested to achieve long circulation of

(19)

18

polymers to delay the elimination process, such as coating the surface of liposomes with biocompatible polymers like poly(ethylene glycol) (PEG) linked phospholipids. These can be incorporated into the liposomal bilayer to form a hydrophilic polymer shield over the liposome surface, protecting the liposome from penetration or disintegration by plasma proteins60-65

(Fig. 1.6B). While many varieties have been synthesized by using chemically modified forms of PEG, in some cases it’s necessary to make the liposomes shed their cloak of modified PEG molecules when they reach their target (Fig. 1.6C). In this way they can interact with the target and release their payload. Using imaging technologies, visual evidence of the effect of PEGylation on the circulation kinetics of the liposomes was provided (Fig. 1.6D).66 The images clearly demonstrate that PEGylation

significantly enhances the persistence of liposomes in the blood stream. At the same time, the uptake of PEGylated liposomes in organs (liver and spleen) responsible for particle clearance decreased.

Fig. 1.6 Schematic representation of (A) conventional liposome, (B) PEG-liposome and (C) chemically modified PEG-liposome. (D) The effect of PEGylation on the circulation persistence of liposomes. The liposomes were labeled with Tc-99m, administered in rats, and the rats were imaged with a gamma camera over 24 h. As is evident from the heart (H) image signal, the PEG-liposomes remained in circulation even 24 h post-injection. The accumulation in the liver (L) and the spleen (S) was also lower in the case of PEG-liposomes, as compared to the plain liposomes.66 (Fig. 1.6 D was adapted from reference66)

(20)

Functionalization of Lipid Bilayer Membranes

19 Ligands conjugated with hydrophobic molecules form amphiphiles. The hydrophobic part can insert into the liposome bilayer, exposing the ligand outside of the liposomes for being recognized or for other interactions. For instance, DNA-b-polypropyleneoxide (DNA-ppo) has proven to be stably anchored into the lipid membrane for over at least 24 h. In this way, the containers are encoded with sequence information. The DNA-ppo present on the surface was used for anchoring a photosensitizer by hybridization. Upon light irradiation the PPO was oxidized leading to cargo release (Fig.

1.7).67

Fig. 1.7 Illustration of selective cargo release from DNA block copolymer (DBC) -decorated phospholipid vesicles. (1) DNA-ppo is stably anchored in unilamellar lipid vesicles; (2) DBC-decorated vesicles are functionalized with conjugated DNA-photosensitizers by hybridization; (3) singlet oxygen is generated by light irradiation; and (4) selective cargo release is induced by the oxidative effect of singlet oxygen.67 (This figure was reproduced with permission from

reference 67)

1.3.2 Stimuli-responsive liposomes

Liposomes can suspend cargos with their peculiar solubility properties and act as a sustained-release system for microencapsulated molecules. After modification, liposomes can be used as stimuli-responsive nanoparticles, which are visionary concepts to deliver and release a drug exactly where it is needed.68,69 There are several ways to trigger cargo release, such as

(21)

20

to be combined to appropriately improve the cargo release kinetics and distribution to reduce side effects.

1.3.2.1 Light responsive vesicle systems

Methods of sensitizing liposomes to light have progressed from the use of organic molecule moieties to the use of metallic plasmon resonant structures which can be broadly categorized as photochemical or photophysical release. Photochemical release can be achieved via photoisomerization, photocleavage and photopolymerization, which all lead to destabilization of the liposome bilayer and release of encapsulated contents (Fig. 1.8A-C).

Fig. 1.8 Release from liposomes mediated by photochemical responses: photoisomerization (A), photocleavage (B), or photopolymerization (C); and photophysical responses: molecular absorbers (D) and gold nanoparticles (E).

(22)

Functionalization of Lipid Bilayer Membranes

21 On the other hand, photophysical release from liposomes does not rely on any chemical changes of structures within or associated with the bilayer membrane. Examples of photophysical release discussed here take advantage of photothermal conversion of absorbed light with ensuing thermal and/or mechanical processes in the lipid membrane and the surrounding medium. The methods for achieving photophysical release are developed around molecular absorbers (Fig. 1.8D) or gold nanoparticles (Fig. 1.8E).70

1.3.2.2 Temperature responsive vesicle systems

Temperature-responsive liposomes are classified into two types: traditional temperature-responsive liposomes and liposomes modified with responsive polymers. Traditional temperature-responsive liposomes which are composed of temperature-temperature-responsive lipids show the greatest permeation of the lipid membrane at its gel-to-liquid crystalline phase transition temperature.

Moreover, liposomes modified with temperature-responsive polymers exhibit a lower critical solution temperature (LCST) behavior. These polymers are soluble in an aqueous solution below this temperature but dehydrate and aggregate if heated above the LCST. This behavior induces the release of a drug within a polymer-modified liposome. For instance, a temperature-responsive polymer, poly (N-isopropylacrylamide)-co-N,N'-dimethylaminopropylacrylamide (P(NIPAAm-co-DMAPAAm)) was synthesized and used for liposome modification. This research showed that the polymer underwent dehydration and aggregation above 40 °C and that temperature-responsive polymer-modified liposomes had faster cellular uptake and release compared to non-modified liposomes (Fig. 1.9).74

(23)

22

Fig. 1.9 Liposomes modified with temperature-responsive polymers are used for cellular uptake. The copolymer displayed a thermosensitive transition at a lower critical solution temperature (LCST) that is higher than body temperature. Above the LCST, the temperature-responsive liposomes started to aggregate and release their content. The liposomes showed a fixed aqueous layer thickness (FALT) at the surface below the LCST, and the FALT decreased with increasing temperature. Above 37°C, cytosolic release from the temperature-responsive liposomes was higher than that from the PEGylated liposomes, indicating intracellular uptake.74 (This figure was adapted from reference 74)

1.3.2.3 Magnetic responsive vesicle systems

Magnetoliposomes are composed of a lipid bilayer surrounding superparamagnetic iron oxide nanoparticles. Due to the biocompatibility, size, material-dependent physicochemical properties and potential applications as alternative contrast enhancing agents for magnetic resonance imaging, magnetoliposomes are ideal candidates to achieve a spatial and temporal control over drug release.75,76 Superparamagnetic iron

oxide nanoparticles (SPION) can be guided to their site of action using an externally applied magnetic field. The subsequent accumulation of SPION in the target site can be exploited for simultaneous drug delivery, MR imaging or hyperthermia therapy of cancer (Fig. 1.10).

(24)

Functionalization of Lipid Bilayer Membranes

23 Fig. 1.10 Superparamagnetic iron oxide nanoparticles can be guided to the site of action using an externally applied magnetic field.77 (This figure was adapted from reference 77)

In the beginning, liposomes were studied only for their physicochemical properties as models of membrane morphology. Today, they are used as delivery devices to encapsulate cosmetics, drugs, fluorescent detection reagents, and as vehicles to transport nucleic acids, peptides, and proteins to specific cellular sites in vivo. Advances in therapeutic applications of liposomes have been achieved through surface modifications. With these surface modifications, their biological stability could be increased, which includes reduced constituent exchange and leakage as well as reduced unwanted uptake by cells of the mononuclear phagocytic system.78

Targeting components such as antibodies can be attached to liposomal surfaces and were used to create large antigen-specific complexes. In this sense, liposomal derivatives are being used to target cancer cells in vivo, to enhance detectability in immunoassay systems.

(25)

24

1.4 Motivation and Thesis Overview

The overall goal of the work described in this thesis was to use DNA nanotechnology as a tool to manipulate lipid bilayer surfaces. Our group synthesized and characterized a new family of DNA amphiphiles containing modified nucleobases. The modification is introduced in uracil and consists of hydrophobic moieties. Through solid phase synthesis, the modified nucleotides can be incorporated in any desired position and several modifications per DNA strands can be introduced.79 The resulting DNA

sequences still undergo specific Watson-Crick base pairing. This property combined with the amphiphilic nature of this lipid-DNA qualifies the material as appealing candidate to interact with and manipulate biological membrane structures.

In chapter 2, a powerful new approach was introduced by modifying DNA with lipid chains at four nucleobases to tightly anchor the nucleotide to the lipid membrane. This strategy allows highly stable incorporation of DNA into the liposomal bilayer, thereby limiting dissociation. Several assays were employed proving the incorporation and stable anchoring in the phospholipid bilayer. These measurements involve small vesicles and fluorescence energy transfer. These experiments allow to measure how long the DNA amphiphiles remain in the bilayer.

In chapter 3, efficient fusion of liposomes was studied using lipid-DNA introduced in the chapter before. While the orientation of DNA hybridization played a significant role in the efficacy of full fusion of DNA-grafted vesicles, the number of anchoring units was found to be a crucial factor as well. As compared to vesicles functionalized with single-anchored or double-anchored DNA, liposomes containing quadruple-anchored oligonucleotides were found to be highly fusogenic, achieving considerable full fusion of up to 29% without notable leakage. This study demonstrates the importance of the DNA-anchoring strategy in hybridization-induced vesicle fusion, as not only the structural properties of the unit itself, but also the number of anchoring units determines its favorable fusion-inducing properties. Several fluorescence assays, dynamic light scattering and cryogenic transmission electron microscopy were utilized to prove these results.

(26)

Functionalization of Lipid Bilayer Membranes

25 In chapter 4, we expand the functionality of DNA encoded vesicles significantly. It was demonstrated that strand replacement can be carried out. In this chapter it will be outlined what sequences and what DNA amphiphiles are needed to reach this goal, i.e. changing the surface functionalities of liposomes by the simple addition of oligonucleotides. Moreover, it will be detailed how such a surface modification can be amplified by a simple DNA-triggered supramolecular polymerization. In chapter 5, we investigated whether it is possible to insert the lipid-modified DNA sequences into the membrane of live zebrafish to function as artificial receptor. We demonstrate that oligonucleotides functionalized with a membrane anchor can be immobilized on a zebrafish. Protruding single-stranded DNA atop the fish was functionalized by Watson-Crick base pairing employing complementary DNA sequences. In this way, small molecules and liposomes were guided and attached to the fish surface. The anchoring process can be designed to be reversible allowing exchange of surface functionalities by simple addition of DNA sequences. To achieve this on a fish surface, the strand exchange experiments established in

chapter 4 on simple vesicles as model were crucial. Finally, a DNA based

amplification process was performed atop of the zebrafish enabling the multiplication of surface functionalities from a single DNA anchoring unit.

(27)

26

References

1. Fahy, E.; Subramaniam, S.; Brown, H. A.; Glass, C. K.; Merrill, A. H. Jr.; Murphy, R. C.; Raetz, C. R.; Russell, D. W.; Seyama, Y.; Shaw, W.; Shimizu, T.; Spener, F.; Van Meer, G.; Van Nieuwenhze, M. S.; White, S, H.; Witztum, J. L.; Dennis, E. A.; A comprehensive classification system for lipids. Journal of Lipid Research 2005, 46, 839-861.

2. Margineanu, D. G.; Equilibrium and non-equilibrium approaches in biomembrane thermodynamics. Archives Internationales de Physiologie et de Biochimie 1987, 95, 381-422. 3. Svetina, S.; Zeks, B.; Shape behavior of lipid vesicles as the basis of some cellular processes. The Anatomical Record 2002, 268, 215-225.

4. Lasic, D. D.; Liposomes in Gene Delivery. 1997, March 13, by CRC Press.

5. Mozafari, M. R.; Liposomes: An Overview of Manufacturing Techniques. Cell Mol. Biol. Lett.

2005, 10, 711-719.

6. Laouini, A.; Jaafar-Maalej, C.; Limayem-Blouza, I.; Sfar, S.; Charcosset, C.; Fessi, H.; Preparation, Characterization and Applications of Liposomes: State of the Art. J. Colloid Sci. Biotechnol. 2012, 1, 147-168.

7. Bangham, A.; Gier, J. D.; Greville, G.; OSMOTIC PROPERTIES AND WATER PERMEABILITY OF PHOSPHOLIPID LIQUID CRYSTALS. Chem. Phys. Lipids 1967, 1, 225-246.

8. Szoka, F. J.; Papahadjopoulos, D.; Procedure for preparation of liposomes with large internal aqueous space and high capture by reverse-phase evaporation. Proc. Natl. Acad. Sci. U.S.A. 1978, 75, 4194-4198.

9. Szebeni, J.; Breuer, J. H.; Szelenyi, J. G.; Bathori, G.; Lelkes, G.; Hollan, S. R.; Oxidation and denaturation of hemoglobin encapsulated in liposomes. Biochim. Biophys. Acta 1984, 798, 60-67.

10. Batzri, S.; Korn, E. D.; Single bilayer liposomes prepared without sonication. Biochim. Biophys. Acta 1973, 298, 1015-1019.

11. Zumbuehl, O.; Weder, H. G.; Liposomes of controllable size in the range of 40 to 180 nm by defined dialysis of lipid/detergent mixed micelles. Biochim. Biophys. Acta 1981, 640, 252-262.

12. Bangham, A. D.; Horne, R.W.; Negative staining of phospholipids and their structural modification by surface active agents as observed in the electron microscope. J Mol Biol

1964, 8, 660-668.

13. Lasic, D. D.; Applications of Liposomes. Handbook of Biological Physics 2005, chapter 10. 14. Chen, Y.; Sen, J.; Bathula, S. R.; Yang, Q.; Fittipaldi, R.; Huang, L.; molecular pharmaceutics

2009, 6, 696-705.

15. Dutta, D.; Pulsipher, A.; Luo, W.; Yousaf, M. N.; Synthetic Chemoselective Rewiring of Cell Surfaces: Generation of Three-Dimensional Tissue Structures J. Am. Chem. Soc. 2011, 133, 8704-8713.

(28)

Functionalization of Lipid Bilayer Membranes

27 16. Chen, H.; Kim, S.; Li, L.; Wang, S.; Park, K.; Cheng, J.; Release of hydrophobic molecules from polymer micelles into cell membranes revealed by Förster resonance energy transfer imaging. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 6596-6601.

17. Samad, A.; Sultana, Y.; Aqil, M.; Liposomal Drug Delivery Systems: An Update Review. Current Drug Delivery 2007, 4, 297-305.

18. Li, S.; Huang, L.; In vivo gene transfer via intravenous administration of cationic lipid-protamine-DNA (LPD) complexes. Gene Therapy 1997, 4, 891-900.

19. Torchilin, V. P.; Levchenko, T. S.; Rammohan, R.; Volodina, N.; Sternberg, B. P.; D’Souza, G. G. M.; Cell transfection in vitro and in vivo with nontoxic TAT peptide-liposome–DNA complexes. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 1972-1977.

20. Dzau, V. J.; Mann, M. J.; Morishta, R.; Kaneda, Y.; Fusigenic viral liposome for gene therapy in cardiovascular diseases. Proc. Natl. Acad. Sci. U.S.A. 1996, 93, 11421-11425.

21. Shi, N.; Pardridge, W. M.; Noninvasive gene targeting to the brain. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 7567-7572.

22. Villasmil-Sánchez, S.; Rabasco, A. M.; González-Rodríguez, M. L.; Thermal and 31P-NMR studies to elucidate sumatriptan succinate entrapment behavior in Phosphatidylcholine/Cholesterol liposomes. Comparative 31P-NMR analysis on negatively and positively-charged liposomes. Colloids and Surfaces B: Biointerfaces 2013, 105, 14-23. 23. Mozafari, M. R.; Liposomes: an overview of manufacturing techniques. Cell Mol. Biol. Lett.

2005, 10, 711-719.

24. Gubernator, J.; Active methods of drug loading into liposomes: recent strategies for stable drug entrapment and increased in vivo activity. Expert Opin. Drug Deliv. 2011, 567-582.

25. Cohen, B. E.; Bangham, A. D.; Diffusion of small nonelectrolytes across liposome membranes. Nature 1972, 236, 173-174.

26. Cullis, P. R, Hope, M. J.; Bally, M. B.; Influence of pH gradients on the transbilayer transport of drugs, lipids, peptides and metal ions into large unilamellar vesicles. Biochim. Biophys. Acta 1997, 1331, 187-211.

27. Lasic, D. D.; Liposomes: from physics to aplications. Biophysical Jounal 1994, 67, 1358-1362.

28. Chonn, A.; Cullis, P. R.; Devine, D. V.; THE ROLE OF SURFACE CHARGE IN THE ACTIVATION AND ALTERNATIVE PATHWAYS OF COMPLEMENT BY LIPOSOME. J Immunol

1991, 46, 4234-4241.

29. Eastman, S. J.; Siegel, C.; Tousignant, J.; Smith, A.E.; Cheng, S.H.; Scheule, R.K.; Biophysical characterization of cationic lipid: DNA complexes. Biochim. Biophys. Acta 1997, 1325, 41-62. 30. Leventis, R.; Silvius, J. R.; Interactions of mammalian cells with lipid dispersions containing novel metabolizable cationic amphiphiles. Biochim. Biophys. Acta 1990, 1023, 124-132.

31. Felgner, P. L.; Gadek, T. R.; Holm, M.; Roman, R.; Chan, H. W.; Wenz, M.; Northrop, J. P.; Ringold, G. M.; Danielson, M.; Lipofection: a highly efficient, lipid-mediated DNA-transfection procedure. Proc. Natl. Acad. Sci. U.S.A. 1987, 84, 7413-7417.

32. Templeton, N. S.; Cationic liposome-mediated gene delivery in vivo. Biosci. Rep. 2002, 22, 283-295.

(29)

28

33. Wrobel, I.; Collins, D.; Fusion of cationic liposomes with mammalian cells occurs after endocytosis. Biochim. Biophys. Acta 1995, 1235, 296-304.

34. Shim, G.; Kim, M.; Park, J. Y.; Oh, Y.; Application of cationic liposomes for delivery of nucleic acids. ASIAN JOURNAL OF PHARMACEUTICAL SCIENCES 2013, 8, 72-80.

35. Kang, S. H.; Cho, H.; Shim, G.; Lee, S.; Kim, S.; Choi, H.; Kim, C.; Oh, Y.; Cationic Liposomal Co-delivery of Small Interfering RNA and a MEK Inhibitor for Enhanced Anticancer Efficacy. Pharm Res. 2011, 28, 3069-3078.

36. Dudia, A.; Koҫer, A.; Subramaniam, V.; Kanger, J. S.; Biofunctionalized Lipid-Polymer Hybrid Nanocontainers with Controlled Permeability. Nano Lett. 2008, 8, 1105-1110. 37. Cisse, I.; Okumus, B.; Joo, C.; Ha, T.; Fueling protein–DNA interactions inside porous nanocontainers. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 12646-12650.

38. Birkner, J. P.; Poolman, B.; Koҫer, A.; Hydrophobic gating of mechanosensitive channel of large conductance evidenced by single-subunit resolution. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 12944-12949.

39. Louhivuori, M.; Risselada, H. J.; Giessen, van der E.; Marrink, S. J.; Release of content through mechano-sensitive gates in pressurized liposomes. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 19856-19860.

40. Yhee, J. Y.; Lee, S.; Kim, K.; Advances in targeting strategies for nanoparticles in cancer imaging and therapy. Nanoscale 2014, 6, 13383.

41. Lee, S.; Kim, J.; Shim, G.; Kim, S.; Han, S. E.; Kim, K.; Kwon, I. C.; Choi, Y.; Kim, Y. B.; Kim, C.; Oh, Y.; Tetraiodothyroacetic acid-tagged liposomes for enhanced delivery of anticancer drug to tumor tissue via integrin receptor. J Control Release 2012, 164, 213-220.

42. Ma, M.; Bong, D.; Controlled Fusion of Synthetic Lipid Membrane Vesicles. Acc Chem Res.

2013, 46, 2988-2997.

43. Kumar, P.; Guha, S.; Diederichsen, U.; SNARE protein analog-mediated membrane fusion. J. Pept. Sci. 2015, 21, 621-629.

44. Kong, L.; Askes, S. H. C.; Bonnet, S.; Kros, A.; Campbell, F.; Temporal Control of Membrane Fusion through Photolabile PEGylation of Liposome Membranes. Angew. Chem. Int. Ed. 2016, 55, 1396-1400.

45. Chernomordik, L. V.; Kozlov, M. M.; Protein-lipid interplay in fusion and fission of biological membranes. Annu. Rev. Biochem. 2003, 72, 175-207.

46. Cohen, F.S.; Melikyan, G. B.; The energetics of membrane fusion from binding, through hemifusion, pore formation and pore enlargement. J. Membr. Biol. 2004, 199, 1-14.

47. Weber, T.; Zemelman, B. V.; Mcnew, J. A.; Westermann, B.; Gmachl, M.; Parlati, F.; Söllner, T. H.; Rothman, J. E.; SNAREpins: minimal machinery for membrane fusion. Cell 1998, 92, 759-772.

48. Jahn, R.; Scheller, R. H.; SNAREs–engines for membrane fusion. Nat. Rev. Mol. Cell Biol.

2006, 7, 631-643.

49. Hong, W. J.; Lev, S.; Tethering the assembly of SNARE complexes. Trends Cell Biol. 2014, 24, 35-43.

50. Kumar, P.; Guha, S.; Diederichsen, U.; SNARE protein analog-mediated membrane fusion. J. Pept. Sci. 2015, 21, 621-629.

(30)

Functionalization of Lipid Bilayer Membranes

29 51. Stengel, G.; Zahn, R.; Höök, F.; DNA-induced programmablefusionof phospholipidvesicles. J. Am. Chem. Soc. 2007, 129, 9584-9585.

52. Chan, Y. H. M.; van Lengerich, B.; Boxer, S. G.; Lipid-anchored DNAmediates vesicle fusion as observed by lipid content mixing. Biointerphases 2008, 3,17-21.

53. Chan, Y. H. M.; van Lengerich, B.; Boxer, S. G.; Effects of linker sequences on vesicle fusion mediated by lipid-anchored DNA oligonucleotide. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 979-984.

54. Zheng, T.; Voskuhl, J.; Versluis, F.; Zope, H. R.; Tomatsu, I.; Marsden, H. R.; Kros, A. Controlling The Rate of Coiled Coil Driven Membrane Fusion. Chem. Commun. 2013, 49, 3649-3651.

55. Kong, L.; Askes, S. H.; Bonnet, S.; Kros, A.; Campbell, F. Temporal Control of Membrane Fusion through Photolabile PEGylation of Liposome Membranes. Angew. Chem. Int. Ed. 2016, 55, 1396-1400.

56. Mukai, M.; Sasaki, Y.; Kikuchi, J.; Fusion-Triggered Switching of Enzymatic Activity on an Artificial Cell Membrane. Sensors 2012, 12, 5966-5977.

57. Su, W.; Luo, Y.; Yan, Q.; Wu, S.; Han, K.; Zhang, Q.; Gu, Y.; Li, Y.; Photoinduced Fusion of Micro-Vesicles Self-Assembled from Azobenzene-Containing Amphiphilic Diblock Copolymers. Macromol. Rapid Commun. 2007, 28, 1251-1256.

58. Jian Yang, Yasuhito Shimada, René C. L. Olsthoorn, B. Ewa Snaar-Jagalska, Herman P. Spaink, and Alexander Kros, Application of Coiled Coil Peptides in Liposomal Anticancer Drug Delivery Using a Zebrafish Xenograft Model. ACS Nano. 2016, 10, 7428-7435.

59. Nag, O. K.; Awasthi, V.; Surface Engineering of Liposomes for Stealth Behavior. Pharmaceutics 2013, 5, 542-569.

60. Papahadjopoulos, D.; Allen, T. M.; Gabizon, A.; Sterically stabilized liposomes-improvements in pharmacokinetics and antitumor therapeutic efficacy. Proc. Natl. Acad. Sci. U.S.A. 1991, 88, 11460-11464.

61. Pashin, Y. V.; Bakhitova, L. M.; Bentkhen, T. I.; Antimutagenic activity of simple phenols and its dependence on the number of hydroxyl groups. Bull Exp Biol Med. 1986, 102, 1121-1123.

62. Woodle, M. C.; Lasic, D. D.; Sterically stabilized liposomes. Biochim. Biophys. Acta 1992, 1113, 171-199.

63. Woodle, M. C.; Newman, M. S.; Cohen, J. A.; Sterically stabilized liposomes: physical and biological properties. J Drug Target 1994, 2, 397-403.

64. Woodle, M. C.; Newman, M. S.; Collins, L. R.; Efficient evaluation of long circulating or stealth liposomes by studies of in vivo blood-circulation kinetics and final organ distribution in rats. Biophys J. 1990, 57, A261.

65. Çağdaş, M.; Sezer, A.D.; Bucak, S.; Liposomes as Potential Drug Carrier Systems for Drug Delivery. Application of Nanotechnology in Drug Delivery. 2014 Chapter 1.

66. Nag, O. K.; Yadav, V. R.; Hedrick, A.; Awasthi, V.; Post-modification of preformed liposomes with novel non-phospholipid poly(ethylene glycol)-conjugated hexadecylcarbam -oylmethyl hexadecanoic acid for enhanced circulation persistence in vivo. Int. J. Pharm.

(31)

30

67. Rodríguez-Pulido, A.; Kondrachuk, A. I.; Prusty, D. K.; Gao, J.; Loi, M. A.; Herrmann, A.; Light-Triggered Sequence-Specific Cargo Release from DNA Block Copolymer–Lipid Vesicles. Angew. Chem. Int. Ed. 2013, 52, 1008-1012.

68. Chiu, H. C.; Lin, Y. W.; Huang, Y. F.; Chuang, C. K.; Chern, C. S.; Polymer Vesicles Containing Small Vesicles within Interior Aqueous Compartments and pH-Responsive Transmembrane Channels. Angew. Chem. Int. Ed. 2008, 47, 1875-1878.

69. Volodkin, D. V.; Skirtach, A. G.; Möhwald, H.; Near-IR Remote Release from Assemblies of Liposomes and Nanoparticles. Angew. Chem. Int. Ed. 2009, 48, 1807-1809.

70. Leung, S. J.; Romanowski, M.; Light-Activated Content Release from Liposomes. Theranostics 2012, 2, 1020-1036.

71. Li, L.; Hagen, ten T. L.M.; Hossann, M.; Süss, R.; Rhoon, van G. C.; Eggermont, A. M.M.; Haemmerich, D.; Koning, G. A.; Mild hyperthermia triggered doxorubicin release from optimized stealth thermosensitive liposomes improves intratumoral drug delivery and efficacy. J Control Release 2013, 168, 142-150.

72. Tai, L.; Tsai, P.; Wang, Y.; Wang, Y.; Lo, L.; Yang, C.; Thermosensitive liposomes entrapping iron oxide nanoparticles for controllable drugrelease. Nanotechnology 2009, 20, 1-9.

73. Nappini, S.; Bombelli, F. B.; Bonini, M.; Nord, B.; Baglioni. P.; Magnetoliposomes for controlled drug release in the presence of low-frequency magnetic field. Soft Matter 2010, 6, 154-162.

74. Wang, J.; Ayano, E.; Maitani, Y.; Kanazawa, H.; Tunable Surface Properties of Temperature-Responsive Polymer-Modified Liposomes Induce Faster Cellular Uptake, ACS Omega 2017, 2, 316-325.

75. Monnier, C. A.; Burnand, D.; Rutishauser, B. R.; Lattuada, M.; Petri-Fink, A.; Magnetoliposomes: opportunities and challenges. Eur J Nanomed. 2014, 6, 201-215. 76. Amstad, E.; Kohlbrecher, J.; Müller, E.; Schweizer, T.; Textor, M.; Reimhult, E.; Triggered Release from Liposomes through Magnetic Actuation of Iron Oxide Nanoparticle Containing Membranes. Nano Lett. 2011, 11, 1664-167.

77. Laurent, S.; Saei, A.A.; Behzadi, S.; Panahifar, A.; Mahmoudi, M.; Superparamagnetic iron oxide nanoparticles for delivery of therapeutic agents: opportunities and challenges. Expert Opin Drug Deliv. 2014, 11, 1449-1470.

78. Woodle, M. C.; Surface-modified liposomes: assessment and characterization for increased stability and prolonged blood circulation. Chem. Phys. Lipids 1993, 64, 249-262. 79. Anaya, M.; Kwak, M.; Musser, A. J.; Müllen, K.; Herrmann, A.; Tunable Hydrophobicity in DNA Micelles: Design, Synthesis, and Characterization of a New Family of DNA Amphiphiles. Chem. Eur. J. 2010, 16, 12852-12859.

(32)

Chapter 2

Stability Study of Lipid-DNA on

the Liposomal Membrane

(33)

32

2.1 Introduction

Deoxyribonucleic acid (DNA) is a macro molecule that carries hereditary information of all known living organisms and many viruses. Its double-stranded helix structure was discovered by Watson & Crick in 1953,1 which

has greatly fueled many technologies dealing with DNA and hence revolutionized modern science. In recent years DNA has become a valuable functional building block and tool in nanotechnology and material science due to the unique nature and properties of DNA and DNA hybrid materials. A wide variety of products and applications have been realized using DNA technologies among which is incorporating DNA with a functional group and utilizing its information-carrying capability to develop DNA detection systems. For instance, fluorescent dye-labeled DNA was used as probe monitor in PCR2 or for sequence analysis.3,4 Additionally, coupling DNA

strands with moieties like polymers or nanoparticles changes the morphological structure and introduces new functionalities, which are

(34)

Stability Study of Lipid-DNA on the Liposomal Membrane

33 different from conventional polymers. For instance, DNA conjugated gold nanoparticles were used in DNA microarray technology.5-7 Another

functional moiety chemically conjugated with DNA consists of hydrophobic molecules, such as long alkyl chains, cholesterol, or fatty acids, resulting in amphiphiles, which spontaneously form nanoparticles in solution and enhance the pharmacokinetic behavior and trans-membrane delivery. Their amphiphilic nature arises from the hydrophilic DNA backbone containing charged phosphodiester bonds and the hydrophobicity of attached alkyl chains.8 These nanoparticles can be further functionalized

through hybridization of a modified complimentary DNA or internalization of payloads in the hydrophobic core.9

Our group reported the synthesis and characterization of a family of DNA amphiphiles containing hydrophobically modified nucleobases.10,11

Specifically, 1-dodecyne (C12H22) was attached to a uracil base which was

further attached to the 5’ or 3’ position of a DNA sequence (Fig. 2.1A). In aqueous environment, due to their amphiphilic nature, lipid-DNA self-assembles into micelles whereby the hydrophilic DNA strands shield the hydrophobic lipid core. These DNA micelles can be loaded with cargo by hydrophobic interactions or hybridization with functionalized complementary DNA (Fig. 2.1B). The aggregation properties of lipid-DNA can be relatively easy manipulated by changing the length of the lipid part or the number and position of the modified uracil bases within the DNA sequence. Fig. 2.1C shows three different lipid-DNAs. U2M and U2T are lipid DNA with two modified uracil bases either in the middle or at the terminus and U4T represents lipid DNA with four modified uracil bases at the 5’ end.10

Because of the amphiphilic and sequence specific properties, lipid-DNA can be used for liposome surface modification by insertion of the hydrophobic part into the membrane while the hydrophilic DNA is exposed to the aqueous medium. Compared with existing terminal modifications, our design allows the precise and easy introduction of hydrophobic units at arbitrary positions and numbers in a DNA sequence through conventional solid-phase synthesis. In this chapter, DNA was modified with four lipid

(35)

34

chain modified nucleobases at both terminals and it was used to anchor it to phospholipid membranes.

Fig. 2.1 Structure of lipid modified nucleotide and representation of lipid–DNAs. (A) Chemical structure of the lipid-modified uracil nucleobase. (B) Lipid-DNAs self-assemble to form DNA micelles due to their amphiphilic nature. These self-assembled structures can carry cargo by hydrophobic interaction (1) or by hybridization with a functionalized complementary DNA (2). (C) Schematic representation of the ss and ds lipid–DNA amphiphiles (U2M, U2T, and U4T) and their propensity to undergo Watson-Crick base pairing.10

(36)

Stability Study of Lipid-DNA on the Liposomal Membrane

35

2.2 Results and Discussion

2.2.1 Lipid-DNA design and characteristics

To obtain stable incorporation of DNA into the liposomal bilayer, we use lipid-DNA (U4T-18), which has been designed to contain four modified uracil nucleobases at the 5’ position of a 18-mer oligonucleotide (including the 4 lipid modified uracil bases). CrU4T-18 is complementary to U4T-18 with the lipid anchor at the opposite terminus (i.e. the 3’ position). Cr-ATTO488 is a 14-mer DNA complementary to U4T-18 and was covalently attached an ATTO488 dye to the 3’ end (Table 2.1).

Table 2.1 Sequences of modified DNA.

Name Sequence (5’→ 3’)*

U4T-18 UUUUGCGGATTCGTCTGC

CrU4T-18 GCAGACGAATCCGCUUUU

14mer GCGGATTCGTCTGC

Cr-ATTO488 GCAGACGAATCCGC-ATTO488

*: U represents the lipid-modified uracil base.

U4T-18 can be attached to the liposome surface by insertion of four lipid-modified nucleobases into the lipid membrane while the remaining 14mer DNA part is protruding into the aqueous medium. This DNA unit can hybridize with the DNA part from CrU4T-18 or Cr-ATTO488 (Fig. 2.2A). According to the results from polyacrylamide gel electrophoresis (PAGE), a lower electrophoretic mobility of hybridized lipid-DNA (lane 2) is observed compared to ssDNA controls (lane 1 and lane 3), indicating successful Watson-Crick base pairing (Fig. 2.2B).

After confirming hybridization, the melting temperature (Tm) of the

ds-lipid-DNA (U4T-18+Cr-ATTO488) was determined. The ds-ds-lipid-DNA and ds14mer (14mer+Cr-ATTO488) were heated at 0.5 °C/min while measuring the absorption at 260 nm. Afterwards the first derivative of the curve was calculated and Tm of the ds DNA was taken at maximum slope.

(37)

36

The Tm value of lipid-DNA (62.5 °C) is very close to that of 14mer (63.6 °C)

(Fig 2.2C, D). The result indicates that lipid chains have little influence on the melting temperature.

Fig. 2.2 (A) Schematic representation of U4T-18 hybridization with Cr-ATTO488 on the surface of liposomes. (B) Native PAGE characterization of lipid-DNA (20% TBE gel, 100V, 80min). Lane 1: U4T-18, lane 2: U4T-18 + Cr-ATTO488, lane 3: Cr-ATTO488. (C) Melting curve of dsDNA, U4T-18 + Cr-ATTO488. (D) Melting curve of dsDNA, 14mer + Cr-ATTO488. Melting curve (black squares, left Y-axis) and calculated

derivative

for corresponding sample (red circle, right Y-axis).

2.2.2 Characterization of the incorporation of lipid-DNA in liposomal

bilayer.

After synthesis of the nucleobase-modified DNA hybrids and testing their ability for Waston-Crick base pairing, the lipid DNAs were stably anchored into the membrane of DOPC:DOPE:cholesterol lipid vesicles, while the oligonucleotides remained available for hybridization, as demonstrated by a Fluorescence Resonance Energy Transfer (FRET) assay.12

(38)

Stability Study of Lipid-DNA on the Liposomal Membrane

37 Since ATTO488 and rhodamine dyes show energy transfer when there is a sufficiently short distance between them,13 ATTO488 was covalently

attached to the 3’ end of a 14-mer DNA complementary to U4T-18 (Cr-ATTO488) to act as a donor. In parallel, rhodamine-functionalized phospholipid (Rh-DHPE) was incorporated in the liposomal bilayer to function as an acceptor (Fig. 2.3A). To observe the dynamic emission changes of donor and acceptor after adding Cr-ATTO488, fluorescence emission spectra with excitation at 470 nm of Cr-ATTO488 (donor, emission maximum 520 nm) and Rh-DHPE (acceptor, emission maximum 592 nm) were recorded over 30 min (Fig. 2.3B). The fluorescence of donor significantly decreased by adding Cr-ATTO488 and the fluorescence of acceptor slightly increased at the same time, illustrating that FRET is induced by DNA hybridization.

Fig. 2.3 (A) Schematic of FRET assay demonstrating that oligonucleotides anchored into liposomal bilayers via lipid-DNA remain available for hybridization. (B) Fluorescence emission of Cr-ATTO488 (donor, emission maximum 520 nm) and Rh-DHPE (acceptor, emission maximum 592 nm) followed over 30 min after adding Cr-ATTO488.

Meanwhile, as demonstrated by the increase in the maximum intensity ratio I592/I520 (acceptor/donor peak) (Fig. 2.4D), hybridization only

occurred upon mixing of Cr-ATTO488 with U4T-18-grafted Rh-DHPE-containing vesicles, positioning both dyes sufficiently close to each other to achieve FRET (Fig. 2.4A), whereas for vesicles containing non-complementary lipid-DNA, CrU4-18, (Fig. 2.4B) or no lipid-DNA at all (Fig.

(39)

38

Fig. 2.4 Anchoring of lipid-DNA in the membrane and hybridization on the vesicle surface leads to Fluorescence Resonance Energy Transfer (FRET) upon hybridization of donor-modified complementary DNA with DNA-functionalized, acceptor-containing vesicles. (A) FRET is achieved when complementary Cr-ATTO488 DNA hybridizes with U4T-18 and brings the donor close to the acceptor, rhodamine, positioned in the membrane. If hybridization is not possible, either due to mismatch of the two DNA strands (B) or the absence of membrane-grafted DNA (C) FRET does not occur. (D) Fluorescence spectra of systems capable of FRET (red) and non-FRET controls, either due to DNA mismatch (blue) or absence of membrane-grafted DNA (green).

Disruption of vesicles by addition of Triton X-100 to a final concentration of 0.3% (v/v) resulted in a drop in FRET in the U4T-18 vesicles hybridized with Cr-ATTO488 (Fig. 2.5A vs Fig 2.4D), confirming that FRET was indeed caused by bringing the donor in close vicinity to the acceptor dye located in the liposomal membrane. As expected, in two control non-FRET systems in which DNA hybridization could not occur, either due to absence of DNA in the membrane (Fig. 2.5B) or the presence of non-complementary DNA (Fig. 2.5C) energy transfer from donor to acceptor

(40)

Stability Study of Lipid-DNA on the Liposomal Membrane

39 was prevented. Therefore, similar spectra were observed before and after liposomal disruption.

Fig. 2.5 To further investigate the engraftment of the lipid-DNA hybrids into the membrane, FRET liposomes were disrupted with Triton X-100 at a final concentration of 0.3 % (v/v). Fluorescence spectra of FRET liposomes before and after adding Triton X-100 (A). Similar spectra were observed in control experiments before and after liposomal disruption, either due to the absence of DNA on the membrane (B) or the presence of non-complementary DNA (C).

2.2.3 Temporal stability of lipid-DNA in the liposomal membrane.

To study whether the incorporation of U4T-18 in the membrane is stable overtime, FRET (U4T-18/Cr-ATTO488/Rh-DHPE) liposomes were incubated with non-FRET (NF) liposomes (Fig. 2.6A) at different ratios (1:1, 1:5 and 1:10).

(41)

40

Fig. 2.6 Measurement of stability of lipid-DNA in liposomes over time. FRET (U4T-18/Cr-ATTO488/Rh-DHPE) liposomes were incubated with non-FRET (NF) liposomes (A) at different ratios (1:1, 1:5 and 1:10), and the relative Rh-DHPE/ATTO488 (IA/ID) emission intensity ratio

was monitored over 24 h after mixing (B). Fluorescence spectra of Cr-ATTO488/Rh-DHPE pair in FRET liposomes mixed with NF liposomes at different ratio (v/v): 1:1(red line), 1:5(blue line), 1:10(green line) (C). Solid and dashed lines represent the spectra of the mixed systems before and after adding Triton X-100, respectively.

The relative Rh-DHPE/ATTO488 (IA/ID) emission intensity ratio of the

three systems was monitored over 24 h after mixing (Fig. 2.6B). If lipid-DNA redistributes from FRET liposomes to NF liposomes, a decrease in relative fluorescence of acceptor peak would be observed. After 24 h, some of the acceptor intensity had dropped, but the relative fluorescence IA/ID of

the mixture remained at a similar value as that during the initial measurement before non-FRET liposomes were added. The results demonstrate that the lipid–DNA is stably anchored in the liposomes over at last 24 hours. Fig. 2.6C shows the fluorescence spectra of Cr-ATTO488/Rh-DHPE pair in FRET liposomes mixed with NF liposomes at different ratio before and after liposomal disruption.

(42)

Stability Study of Lipid-DNA on the Liposomal Membrane

41 Moreover, lipids were mixed with U4T-18 at different molar ratios (5000, 1000, 100, 62.5). The final concentration of Cr-ATTO488 and lipid-mixture (DOPC+DOPE) were kept at 7.32 µM and 0.45 mM, respectively, in all FRET experiments. The results show the I592/I520 ratio increased markedly with

higher U4T-18 densities in the membrane (Fig. 2.7, Table 2.2). These results demonstrate that when more lipid DNA is incorporated into the membrane more DNA strands can be attached to this vesicle surface by hybridization (Table 2.2).

Fig. 2.7 U4T-18/Rh-DHPE fluorescence spectra of FRET liposomes mixed with Cr-ATTO488 at different lipid/U4T-18 ratios. The inset shows a zoom-in of the acceptor Rh-DHPE peak. Solid lines and dashed lines represent the spectra of the FRET system before and after adding Triton X-100, respectively. Lipids were mixed with U4T-18 at different molar ratios (5000, 1000, 100, 62.5).

Table 2.2 The acceptor/donor fluorescence intensity ratios (I592/I520) at different lipid/U4T-18

ratios.

Lipid : U4T-18 ratio

U4T-18:liposome

ratio I592/I520 FRET system

5000 8 0.22

1000 38 0.24

100 380 0.31

(43)

42

2.3 Conclusion

In conclusion, we proposed a powerful new approach employing lipid-DNA which contains four lipid chains modified nucleobases to tightly anchor the nucleotide to the lipid membrane. The incorporation and stability of lipid-DNA on the liposomal membrane were proved by FRET. FRET was achieved when the hybridization occurred between Cr-ATTO488 and U4T-18, which brought the donor (Cr-ATTO488) close to the acceptor (rhodamine) that was positioned in a U4T-18 functionalized membrane. Meanwhile, the I592/I520 (acceptor/donor peak) ratio increased markedly

with higher U4T-18 densities in the membrane, and disruption of vesicles by addition of Triton X-100 resulted in a drop of FRET vesicles system, confirming that FRET was indeed caused by bringing the donor in close vicinity to the acceptor dye located in the liposomal membrane. Finally, the lipid–DNA remained stably anchored in the liposomes for at least 24 hours.

2.4 Experimental Section

2.4.1 Materials

Cholesterol (Chol), 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE) and 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) were purchased from Avanti Polar Lipids (Alabaster, USA) (Fig 2.8A-C, purity >99%) and used without further purification. Headgroup-labeled phospholipid, Lissamine rhodamine B 1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine (triethylammonium salt) (Rh-DHPE) was purchased from Invitrogen (Amsterdam, Netherlands), and used as received (Fig.

2.8D). The DNA-dye conjugate Cr-ATTO488 was purchased from

Biomers.net GmbH (Ulm, Germany). Trition X-100 (10% in water), and Tris/HCl buffer were purchased from Sigma-Aldrich (St. Louis, United States). Anhydrous CHCl3 was purchased from Acros Organics (Geel,

Belgium) and stored over molecular sieves. For all experiments, ultrapure water (specific resistance > 18.4 MΩ cm) was obtained by a Milli-Q water purification system (Sartorius).

(44)

Stability Study of Lipid-DNA on the Liposomal Membrane

43 Fig. 2.8 Structures of lipids: (A) DOPC, (B) DOPE, (C) Cholesterol; fluorescent lipids: (D) Rh-DHPE.

2.4.2 Synthesis and characterization of amphiphilic oligonucleotides

The synthesis of 5-(dode-1-cynyl) deoxyuracil and 5-(dode-1-cynyl) deoxyuracil phosphoramidite were reported previously (Fig. 2.9).10,11 In

short, the modified uracil phosphoramidite was dissolved in CH3CN to a

final concentration of 0.15 M in the presence of 3 Å molecular sieves and the prepared solution was directly connected to a DNA synthesizer (ÄKTA oligopilot plus, GE Healthcare (Uppsala, Sweden)). Oligonucleotides were synthesized on a 10 μmol scale using standard β-cyanoethylphosphoramidi -te coupling chemistry. Deprotection and cleavage from the PS support was carried out by incubation in concentrated aqueous ammonium hydroxide solution for 5 h at 55 °C. Following deprotection, the oligonucleotides were purified by using reverse-phase chromatography, using a C15 RESOURCE RPCTM 3 mL reverse phase column (GE Healthcare) through a custom gradient elution (A: 100 mM triethylammonium acetate (TEAAc) and 2.5% acetonitrile, B: 100 mM TEAAc and 65% acetonitrile). Fractions were

(45)

44

desalted using centrifugal dialysis membranes (MWCO 3000, Sartorius Stedim). Oligonucleotide concentrations were determined by UV absorbance using extinction coefficients. Finally, the identity and purity of the oligonucleotides was confirmed by RPC-HPLC (Fig. 2.10) and MALDI-TOF mass spectrometry (Fig. 2.11).

Fig. 2.9 Synthesis of 5-(dode-1-cynyl) deoxyuracil 2 and 5-(dode-1-cynyl) deoxyuracil phosphoramidite 3.

(46)

Stability Study of Lipid-DNA on the Liposomal Membrane

45 Fig. 2.11 RPC HPLC analysis of purified lipid-DNAs: (A) U4T-18, (B) CU4T-18 and (C) CrU4T-18. Numbers beside the elution peaks represent the buffer B contents when lipid-DNAs were eluted.

2.4.3 Preparation and characterization lipid-DNA liposomes

An appropriate amount of freeze-dried lipid-DNA was mixed with DOPC:DOPE:Cholesterol (50:25:25 mol% in chloroform), to obtain the required lipid:lipid-DNA ratio. Afterwards, chloroform was removed by evaporation under an air stream and then under vacuum overnight. An aqueous buffer (100 mM NaCl, 20 mM Tris, pH 7.5) was added to the flask and the solution was vortexed and freeze-thawed 5 times. Subsequently, the dispersion was extruded 21 times, using an extruder and 100 nm polycarbonate membranes (Whatman), to obtain unilamellar vesicles. After extrusion, external buffers of each sample were removed by size exclusion chromatography. The column was filled with Sephadex G-75 (GE Healthcare Life Sciences) and equilibrated with buffer (100 mM NaCl, 20 mM Tris, pH 7.5). Lipid-DNA liposomes were used within one day. All liposomal formulations had an average diameter of around 130 nm as determined by DLS (ALV/CGS-3 ALV-Laser Vertriebsgesellschaft mbH, Langen, Germany). The ratio between lipid and U4T-18 was 500:1, unless stated otherwise.

(47)

46

2.4.4 Calculation of lipid-DNA/liposome ratio.

The amount of lipid-DNAs per liposome was calculated using the equation:

where Φ is the number of lipids per liposome which can be calculated from geometrical considerations:

where Souter and Sinner are the outer and inner surface area of the spherical

liposomes. Assuming the thickness of the lipid bilayer is 5 nm.14,15 α is the

average cross-sectional area of the lipid headgroups, which is assumed to be (2*80+65)/3=75 Å for DOPC:DOPE(2:1 molar ratio).16 Router is the

averaged radius of spherical liposomes, which was determined by DLS.

2.4.5 Characterization of lipid-DNA incorporation in liposomes

measured by Fluorescence Resonance Energy Transfer (FRET) assay

Fluorescence emission spectra of Cr-ATTO488 (donor) and Rh-DHPE (acceptor) in the 500–700 nm region were recorded with excitation at 470 nm using a SPECTRAMAX M2 (Molecular Devices) fluorescence spectrophotometer. Measurements were carried out at constant temperature of 25.0 °C, using a 100 mM NaCl, 20 mM Tris, pH 7.5 buffer.

2.4.6 FRET assay via DNA hybridization

U4T-18 was incorporated in Rh-DHPE/(DOPC+DOPE) (3:97 molar ratio) liposomes to obtain U4T-18 liposomes with a lipid to U4T-18 ratio of 500:1. Subsequently, an aliquot of these liposomes was mixed with a small amount of Cr-ATTO488 such that [U4T-18] = [Cr-ATTO488] = 0.906 μM and with a final lipid (DOPC+DOPE) concentration of 0.45 mM. Then, U4T-18 and Cr-ATTO488 were hybridized using an Eppendorf Mastercycler (Germany). The protocol consisted of heating the mixture 15 min to 40 °C

(48)

Stability Study of Lipid-DNA on the Liposomal Membrane

47 and slowly cooling to 4 °C over a period of 140 min. Afterwards, the emission spectra of Cr-ATTO488/Rh-DHPE pair were measured.

Author contributions

Meng Z designed and conducted the experiments, performed data analysis and wrote the manuscript. Liu Q and de Vries JW synthesized lipid-DNA. Herrmann A supervised the project.

(49)

48

References

1. Watson, J. D.; Crick, F. H.; Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature 1953, 171, 737-738.

2.Kapandis, A. N.; Weiss, S.; Fluorescent probes and bioconjugation chemistries for single-molecule fluorescence analysis of biosingle-molecules. J. Chem. Phys. 2002, 117, 10953-10964. 3. Demidov, V. V.; PNA and LNA throw light on DNA. TRENDS in Biotechnology 2003, 4-7. 4. Liu, Z.; Liu, B.; Ding, J.; Liu. J.; Fluorescent sensors using DNA-functionalized graphene oxide. Anal Bioanal Chem 2014, 406, 6885-6902.

5. Cho, H.; Jung, J.; Chung, B. H.; Scanometric analysis of DNA microarrays using DNA intercalator-conjugated gold nanoparticle. Chem. Commun. 2012, 48, 7601-7603.

6. Niemeyer, C. M.; Ceyhan, B.; Noyong, M.; Simon, U.; Bifunctional DNA–gold nanoparticle conjugates as building blocks for the self-assembly of cross-linked particle layers. Biochem Biophys Res Commun. 2003, 311, 995-999.

7. Park, S. Y.; Lytton-Jean, A. K. R.; Lee, B.; Weigand, S.; Schatz, G. C.; Mirkin, C. A.; DNA-programmable nanoparticle crystallization. Nature 2008, 451, 553-556.

8. Kwak M.; Herrmann, A.; Nucleic acid amphiphiles: synthesis and self-assembled nanostructures. Chem. Soc. Rev. 2011,40, 5745-5755.

9. Kwak, M.; Musser, A. J.; Lee, J.; Herrmann, A.; DNA-functionalised blend micelles: mix and fix polymeric hybrid nanostructures. Chem. Commun. 2008, 29, 326.

10. Anaya, M.; Kwak, M.; Musser, A. J.; Müllen, K.; Herrmann, A.; Tunable Hydrophobicity in DNA Micelles: Design, Synthesis, and Characterization of a New Family of DNA Amphiphiles. Chem. Eur. J. 2010, 16, 12852-12859.

11. Kwak, M.; Minten, I. J.; Anaya, D. M.; Musser, A. J.; Brasch, M.; Nolte, R. J. M.; Müllen, K.; Cornelissen, J. J. L. M.; Herrmann, A.;Virus-like Particles Templated by DNA Micelles: A General Method for Loading Virus Nanocarriers. J. Am. Chem. Soc. 2010, 132, 7834-7835. 12. Rodríguez-Pulido, A.; Kondrachuk, A. I.; Prusty, D. K.; Gao, J.; Loi, M. A.; Herrmann, A.; Light-Triggered Sequence-Specific Cargo Release from DNA Block Copolymer–Lipid Vesicles. Angew. Chem. Int. Ed. 2013, 52, 1008-1012.

13. Alfonta, L.; Singh, A. K.; Willner, I.; Liposomes Labeled with Biotin and Horseradish Peroxidase: A Probe for the Enhanced Amplification of Antigen-Antibody or Oligonucleotide-DNA Sensing Processes by the Precipitation of an Insoluble Product on Electrodes. Anal. Chem. 2001, 73, 91-102.

14. Leonenko, Z. V.; Finot, E.; Ma, H.; Dahms, T. E. S.; Cramb, D. T.; Investigation of Temperature-Induced Phase Transitions in DOPC and DPPC Phospholipid Bilayers Using Temperature-Controlled Scanning Force Microscopy. Biophys J. 2004, 86, 3783-3793.

(50)

Stability Study of Lipid-DNA on the Liposomal Membrane

49 15. Gramse, G.; Perez, A. D.; Edwards, M. A.; Fumagalli, L.; Gomila, G.; Nanoscale Measurement of the Dielectric Constant of Supported Lipid Bilayers in Aqueous Solutions with Electrostatic Force Microscopy. Biophys J. 2013, 104, 1257-1262.

16. Wiethoff, C. M.;Gill, M. L.; Koe, G. S.; Koe, J. G.; Middaugh, C. R.; The Structural Organization of Cationic Lipid-DNA Complexes. J. Biol. Chem. 2002, 277, 44980-44987.

(51)

Referenties

GERELATEERDE DOCUMENTEN

This work was financially supported by the Chinese Scholarship Council (CSC), the University of Groningen and the Netherlands Organization for Science Research

designed an artificial biorthogonal targeting system that was able to target liposomes and other nanoparticles efficiently to the tissue of interest by using coiled coil

2.4.5 Characterization of lipid-DNA incorporation in liposomes measured by Fluorescence Resonance Energy Transfer (FRET) assay Fluorescence emission spectra of Cr-ATTO488

bivalent cholesterol-anchored DNA was similarly effective as that of their bivalent double-stranded counterparts. A second DNA-mediated vesicle fusion strategy, reported by

Disruption of liposomes by addition of Triton X-100 resulted in an increased donor peak (C488, 520 nm) and slightly decreased acceptor peak (Rh-DHPE, 592 nm) (Fig 4.2C),

Studies performed in GUVs and GPMVs show different values for lateral diffusion, partitioning of raft associated proteins and differences in membrane order between the L o and L d

Single monosaccharides (glucose and fructose) and reducing disaccharides (including palatinose, maltose, and gentiobiose) do not affect coexisting L o and L d phases,

the acyl chains to α-helical membrane peptides or proteins would shift their partitioning from L d to L o , but this is not what we observe in phase-separating GUVs composed of DPPC,