DNA delivery systems based on peptide-mimicking cationic lipids

Mar 6, 2019 - ... Elzbieta Malinowska , Christian Wölk , and Gerald Brezesinski. Langmuir , Just Accepted Manuscript. DOI: 10.1021/acs.langmuir.8b041...
0 downloads 0 Views 2MB Size
Subscriber access provided by ECU Libraries

Interface Components: Nanoparticles, Colloids, Emulsions, Surfactants, Proteins, Polymers

DNA delivery systems based on peptide-mimicking cationic lipids – the effect of the co-lipid on the structure and DNA binding capacity Stephanie Tassler, Bodo Dobner, Lisa Lampp, Robert Zió#kowski, Elzbieta Malinowska, Christian Wölk, and Gerald Brezesinski Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.8b04139 • Publication Date (Web): 06 Mar 2019 Downloaded from http://pubs.acs.org on March 10, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

DNA delivery systems based on peptide-mimicking cationic lipids – the effect of the co-lipid on the structure and DNA binding capacity

Stephanie Tasslera*, Bodo Dobnerb, Lisa Lamppb, Robert Ziółkowskic, Elżbieta Malinowskac, Christian Wölkb* and Gerald Brezesinskia a Max

Planck Institute of Colloids and Interfaces, Science Park Potsdam-Golm, Am Mühlenberg 1, 14476 Potsdam, Germany

b Martin-Luther-University

(MLU) Halle-Wittenberg, Institute of Pharmacy, Wolfgang-LangenbeckStraße 4, 06120 Halle (Saale), Germany

c

Warsaw University of Technology, Faculty of Chemistry, Department of Microbioanalytics, The Chair of Medical Biotechnology, ul. Noakowskiego 3, 00-664 Warszawa, Poland

Corresponding authors: S.T.

[email protected]

C.W.

[email protected]

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract: In continuation of previous work, we present a new promising DNA carrier: OO4, a highly effective peptide-mimicking lysine-based cationic lipid. The structural characteristics of the polynucleotide carrier system OO4 in mixtures with the commonly used co-lipid DOPE and the saturated phospholipid DPPE have been studied in 2D and 3D model systems in order to understand their influence on the physical-chemical properties. The phase behavior of pure OO4 and its mixtures with DOPE and DPPE were studied at the air-water interface using a Langmuir film balance combined with infrared reflection-absorption spectroscopy (IRRAS). In bulk, the self-assembling structures in presence and absence of DNA were determined by small-angle and wide-angle X-ray scattering (SAXS/WAXS). The amount of adsorbed DNA to cationic lipid bilayers was measured using a quartz crystal microbalance (QCM). The choice of the co-lipid has an enormous influence on the structure and the capability of binding DNA. DOPE promotes the formation of non-lamellar lipoplexes (cubic and hexagonal structures) whereas DPPE promotes the formation of lamellar lipoplexes. The correlation of the observed structures with the transfection efficiency and the serum stability indicates that OO4/DOPE 1:3 lipoplexes with a DNA containing cubic phase encapsulated in multilamellar structures seem to be most promising.

Keywords: cationic lipids; IRRAS; lipoplex structure; QCM; X-ray scattering

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Introduction

The concept of gene therapy was first proposed in 1972 by Friedman et al.1 to cure monogenetic diseases such as Cystic Fibrosis and Sickle-cell anaemia.2 This innovative therapy requires efficient nucleic acid transfer into cells. This challenge took 40 years until the first in-vivo therapy against lipoprotein lipase deficiency, Glybera®, was approved by the European Medicine Agency. In 2016, an ex-vivo therapy against ADA-SCID, called Strimvelis®, was made available on the European market. Both therapeutics rely on viral vectors which carry several disadvantages, like the immunogenic potential and the expensive as well as difficult manufacturing.3,

4

The costs for a treatment with

Glybera® and Strimvelis® are approximately 1 million $ and 665,000 $, respectively.5, 6 Taking into account the decrease of follow-up costs if the causal treatment of this deseases occurs justifies the price, but cheaper alternatives can push forward a broader clinical use of gene therapeutics for deseases with a higher prävalence. Non-viral vectors, mainly realized by cationic polymers or cationic lipids, are an alternative approach because they can be easier and cheaper produced. But despite of the many advantages of non-viral vectors based on cationic lipids such as high loading capacity, biodegradability, lower immunogenic potential and their comparatively simple, large-scale and low-cost manufacturing, their low efficiency is the main drawback. Especially in-vivo, these so called cytofectins show a weak performance compared to viral vectors.7, 8 The transfection efficiency depends on many parameters like lipoplex size and structure, temperature, pH value, cell type, ionic strength, N/P-ratio and resulting charge density. Up to now, there is still a great interest in uncovering the structure-activity-correlation of lipid-based non-viral gene carriers.9-11 One key parameter is the choice of the co-lipid and its influence on the resulting lipoplex formation.12,

13

In general, neutral or zwitterionic co-lipids are

added to cationic lipids in order to adjust liposome charge density and to tune the lipoplex stability. If the charge density is too low, the lipoplex remains trapped in the endosome.14 Therefore, the charge density has to be high enough to enable an escape from the endosome through active fusion. A too tight packing of DNA (due to very high charge density) reduces the lipoplex dissociation and only a low amount of DNA will be set free for effective gene expression. An optimum charged density would ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 36

be high enough to avoid endosomal entrapment and low enough to efficiently release the DNA cargo.15 Additionally, the loading degree of the vector, often termed as N/P-ratio of the lipoplex, is crucial. The lipoplex needs to remain overall positively charged to interact with the negatively charged cell membrane in order to induce endocytosis.16, 17 Furthermore, the complex structure can be adjusted by the choice of helper lipids. DOPE adopts the inverted hexagonal phase at room temperature due to its conical molecule shape. The gel to liquid-crystalline phase transition occurs around -8 °C, while the Lα to HII transition takes place around 10 °C.18,

19

Adding the zwitterionic

helper lipid DOPE promotes the formation of inverted hexagonal structures, which can rapidly fuse with the endosomal bilayer and allow cytoplasmic release of the DNA in high efficiency.20 In contrast, DPPE has a cylinder-like molecular architecture and favors the formation of stable lamellar structures. The gel to liquid-crystalline phase transition occurs at around 64 °C, while the Lα to HII transition takes place at around 123 °C.21 Also the miscibility of cationic lipid and helper lipid plays an important role for an efficient gene transfer.22 One key parameter to ensure miscibility is the adjustment of alkyl chain length of the cationic and the helper lipids. Similar chain length and phase state enhance miscibility. Up to now, non-lamellar phases (hexagonal and cubic) are believed to result in higher transfection rates due to their fusogenic properties with membranes (cell and endosome).20,

23, 24

Additionally, the curvature of liposomes can be tuned by co-lipids, since the

vesicle size is expected to depend on the composition. The size or curvature of the equilibrium vesicles is undoubtedly linked to their stability, what effects transfection efficiency.25, 26 In continuation of our structure-properties-activity study on lipid-based non-viral DNA carries for enhanced gene transfection, we investigate the interaction of calf thymus DNA and OO4 in mixtures with either DOPE or DPPE. The OO4/DOPE (1:3, n:n) mixture is more effective and has a higher stability in physiological conditions compared to the OO4/DPPE (1:3, n:n) mixture.27 OO4/DOPE (1:3, n:n) can efficiently loaded with DNA and has several beneficial properties in biological media which help to overcome the polycation dilemma.27 This work will enlighten the physical-chemical properties of both mixtures. OO4 belongs to our second generation of malonic acid based cationic lipids,28, 29

ACS Paragon Plus Environment

Page 5 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

and has been proven to be strongly internalized in NIH3T3 cells.30 This work focuses on the influence of the chain pattern of co-lipids on the physical-chemical properties of OO4 in 2D and 3D model systems. For the 2D model systems, experiments were performed with Langmuir monolayers (compression and adsorption isotherms, infrared reflection-absorption spectroscopy). The bulk (3D) systems are studied in small- and wide-angle X-ray scattering and quartz crystal microbalance experiments.

Materials and Methods

Material If not stated otherwise, all materials were purchased from Sigma Aldrich. Milli-Q Millipore water with a specific resistance of 18.2 MΩ·cm was used for all measurements and sample preparations. Methods Sample preparation For the experiments, a 1 mM stock solution of OO4 (C51H101N7O3, 860.39 g/mol) was prepared in chloroform:methanol in a ratio of 8:2 (v:v) (CHCl3: J. T. Baker, Netherlands ; stabilized with 0.75 % of ethanol, CH3OH: Merck, Germany; purity > 99.9 %). The general synthesis of the malonic acid based lipids and the analytical data of OO4 have been already described.28, 30 The zwitterionic phospholipids DOPE (C41H78NO8P, 744.03 g/mol) and DPPE (C37H74NO8P, 691.96 g/mol) were purchased from Avanti (Avanti Polar Lipids, Inc., Alabaster, USA) and used without further purification. The deoxyribonucleic acid sodium salt from calf thymus (> 13 kbp) was purchased from Sigma Aldrich (Type 1, CAS: 73049-39-5). A solution of 1 mMnucleotides ct-DNA was freshly prepared in 1 mM NaCl solution by gently stirring at 5 °C overnight. The molar mass of DNA refers to a monomer containing one charge per phosphate moiety with 10 % of hydration (M ~ 370 g/mol).31 Due to the large amounts of DNA needed for this experiment, we choose calf thymus DNA instead of the biologic active plasmid used in other publications,27 because comparative studies in our lab show no difference in the structure (unpublished results). The bromide containing buffers had a constant concentration of bromide anions (2 mM). The pH 3 was adjusted with 1,4-diazabicyclo(2,2,2)octane (pKa1 = 4.2, pKa2 = 8.2) and pH 10 was obtained by using piperazine (pKa1 = 5.7, pKa2 = 9.8). These buffers were also used in the experiments performed ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

for the determination of the apparent pKa values.32 The 2-(N-Morpholino)ethansulfonic acid (MES) buffer (C6H13NO4S, 195.2 g/mol) had a constant concentration of either 10 mM (experiments except SAXS and WAXS) or 100 mM (SAXS and WAXS), and pH 6.5 was adjusted with hydrochloric acid (pKa = -3). This buffer was also used in the biological experiments.27 Liposomes for QCM or zeta-potential measurements were prepared with the film hydration procedure. Briefly, OO4 and the co-lipids were solved in chloroform/methanol (v:v 8:2) to obtain 1 mM solutions. Later on these solutions were mixed in the desired molar ratio of 1:3 (v:v OO4/colipid). The resulting solutions were dried under nitrogen flow for 2 hours and stayed in a desiccator under vacuum for 12 h. Afterwards, the lipid film was hydrated with sterile aqueous medium (bromide containing buffer (2 mM, pH 3) for QCM or pH adjusted water for zeta potential measurements). The liposome solution had a final concentration of 1 mg/mL. Since small uni-lamellar vesicles (SUV) were desired for the QCM experiments, the lipid dispersions were sonicated (37 kHz) for 3 minutes at room temperature. Langmuir Film Balance The lipid monolayers were examined on a computer-interfaced Langmuir trough equipped with a Wilhelmy balance to measure the surface tension with an accuracy of ±0.1 mN/m. The temperature was fixed with a precision of ± 0.1 °C by a thermostat. The lipid monolayer was compressed with a moveable barrier with a velocity of 5 Ų/molecule/min to the maximum surface pressure (approximately 45 mN/m). Afterwards, the lipid monolayer was expanded immediately to record the hysteresis. The 1 mM lipid solution was spread carefully onto the aqueous subphases (bromide containing buffers, 2 mM, different pH values) using a micro-syringe (Hamilton, Switzerland). Before compression, 10 min were given for evaporation of the organic solvent. In case of DNA adsorbing to the lipid monolayer, 60 min were given. All isotherms were measured at least twice for reproducibility. Infrared Reflection-Absorption Spectroscopy (IRRAS)

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

The infrared reflection-absorption spectra were collected with a Vertex 70 FT-IR spectrometer (Bruker Optics, Ettlingen, Germany). The set-up includes a film balance (R&K, Potsdam, Germany) inside a container (external air-water reflection unit XA-511, Bruker). A sample trough with two movable barriers and a reference trough (only water or buffer) allow the fast recording of the sample and reference spectra using a shuttle technique. The infrared beam is focused on the liquid surface by a set of mirrors. The IR-beam angle of incidence (ϕ) normal to the surface has been adjusted to 40°. A KRS-5 wire grid polarizer is used to polarize the infrared radiation either in the parallel (p) or perpendicular (s) direction. After reflection from the surface, the beam is directed to a narrow band Mercury-Cadmium-Telluride (MCT) detector cooled with liquid nitrogen. The final reflectance– absorbance spectra were obtained using −log(R/R0), with R being the reflectance of the film covered surface and R0 being the reflectance from the pure subphase. For each spectrum of s-polarized light 200 scans were performed with a scanning velocity of 20 kHz and a resolution of 8 cm−1, apodized using the Blackman–Harris 3-term function, and fast Fourier transformed after one level of zero filling.33 All spectra were corrected for atmospheric interference using the OPUS software and baseline corrected applying the spectra-subtraction software. The spectra are not smoothed.34 The maximum of the antisymmetric CH2-band (Lorentzian fit with ±0.2 cm−1) was taken to determine the lipid phase state, while the intensity of the antisymmetric PO2− band was used to estimate the amount of DNA coupled to the monolayers. ζ-potential measurements The electrophoretic mobility was measured using LASER Doppler electrophoresis technique with a ZetasizerNano ZS ZEN3600 (Malvern Instruments, Worcestershire, UK). Three measurements consisting of 30 runs with a voltage of 60 V were performed at 25 °C. For the calculations, a viscosity of η = 0.8872 mPa·s, a dielectric constant of ε = 78.5 F/m, and a refractive index of 1.33 were assumed. The mobility μ of the migrating particles was converted into the ζ-potential using the Smoluchowski relation (equation 1) (Zetasizer Software 6.34). Equation 1

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 36

The presented ζ potential titration curve contains three individual measurements on different days. Small- and Wide-Angle X-Ray Scattering The SAXS/WAXS data were recorded at the High Brilliance Beamline ID02 (ESRF, France) with an energy of the incident X-ray beam equal 12.5 keV (λ = 0.995 Å). The beam size was about 100x100 µm2 and the sample-to-detector distance was 1.2 m. The small-angle diffraction patterns were collected by a 4 FT-CCD detector (Rayonix MX-170HS). For SAXS, a q range from 0.006 Å−1 to 0.65 Å−1 with a detector resolution of 3·10−4 Å−1 (FWHM) was used. WAXS data were obtained in a q range from 0.72 Å−1 to 5.1 Å−1. To avoid sample damage by radiation, each sample was measured with 10 frames with an exposure time of 0.05 s per frame. For data analysis, the average of all 10 frames was used. The collected 2D powder diffraction spectra were reduced and background subtracted using SAXSutilities and analyzed in Origin. The SAXS/WAXS data were corrected for the empty sample holder with pure buffer solution at the corresponding temperature. The angular calibration of the SAXS detectors was performed using silver behenate powder as reference. For WAXS, para bromobenzoic acid was used. The temperature was adjusted with a Huber Unistat with an accuracy of ±0.1 °C. Experiments have been performed at 20 °C, 25 °C and 37 °C. The sealed glass capillaries containing the lipid dispersions were well-positioned in a Peltier controlled automatic sample changer. The real-space repeating distance d of the lattice planes was calculated from the position of the first diffraction peak by equation 2. Lorentzian-functions were fitted to the diffraction peaks. Equation 2 Dissipation-Enhanced Quartz Crystal Microbalance (QCM-D) Experiments The dissipation-enhanced quartz crystal microbalance experiments35 were carried out using a QCM-D E1 (Q-Sense, Gothenburg, Sweden) in order to quantify the adsorbed amount of DNA to a lipid bilayer. The liposomes were prepared as stated above to a final concentration of 1 mg/mL in sterile filtrated bromide containing buffer (2 mM, pH 3). The used DNA solution had a concentration of 0.1 mM in sterile filtrated bromide containing buffer. The AT-cut quartz crystals with a diameter of 14 mm and a frequency of (4.95 ± 0.05) MHz were coated with SiO2 (Q-Sense QSX 303). The flow rate ACS Paragon Plus Environment

Page 9 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

of sterile filtrated aqueous solutions was 20 µL/min. The QCM-chamber was immersed in a temperature-insulating box with a constant temperature of (20 ± 0.1) °C. The liposomes and the ctDNA solutions were injected with a flowrate of 20 µL/min. Each sample injection was stopped, when no change in the frequency occurred after 5 min. After each injection, the sample was flushed with buffer solution to remove dispensable molecules. In order to evaluate the real time recorded QCM data, the 3rd overtone has been used. The mass deposition has been analyzed according to the Sauerbrey equation (equation 3).36 Equation 3

Equation 4

With f0 being the natural frequency of the quartz crystal (equation 4), ρq being the density of the quartz material (ρq = 2.648 g/cm³) and µq being the shear modulus (µq = 2.947·1011 g/cm/s²), and dq being the thickness of the quartz crystal. The material constant C (

) refers to a mass

deposition of 17.7 ng onto an area of 1 cm² causing a frequency shift of 1 Hz for a 5 MHz quartz crystal. The calculation of the theoretical mass depositions is given in the Supporting Information chapter 1.

Results and Discussion

Structural Properties of the used lipids OO4 (Figure 1) is a lysine-based amino-functionalized peptide-mimicking lipid, which was synthesized for polynucleotide delivery. The term peptide-mimicking is based rather on the special aggregation behavior which includes the formation of hydrogen bond networks in a peptide-like manner, than on the fact that it contains L-lysine (see Figure 1).37,

38

The hydrophobic domain is unsaturated and

composed of two oleyl chains (C18:1), since they are known for high transfection efficiency. OO4 has a branched tris(2-aminoethyl)amine spacer and three primary amine groups leading to a pKa value of 6.32 The transfection capacity of OO4 alone is moderate, but the mixing with a co-lipid increases the transfection efficiency. As co-lipids, the unsaturated DOPE and the saturated DPPE were used. The ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 36

OO4/co-lipid ratio of 1:3 (n:n) was used in the following experiments because this was found to be the best ratio in transfection experiments.27 The pKa values of the used lipids as well as DNA are given in Table 1.

Figure 1: Chemical structures of N-{6-Amino-1-[N-(9Z)-octadec-9-enylamino]-1-oxohexan-(2S)-2-yl}-N`-{2-[N,Nbis(2-aminoethyl)amino]ethyl}-2-[(9Z)-octadec-9-enyl]propandiamide OO4, 1,2-dioleoyl-sn-glycero-3phosphoethanolamine DOPE, and 1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine DPPE. The structural features of OO4 are indicated. The primary amines are neglected in the H-bond donor/acceptor indication because the characteristic depends on the protonation state and they are also interacting with the DNA during lipoplex formation.

Table 1: pKa values

(apparent) pKa value Lysine (conjugated)

10.4 (ε-amino)39, 40, 6.1-9.1 (α-amino)40

OO4

632

DOPE

zwitterionic (PO4- and NH3+)

DPPE

zwitterionic (PO4- and NH3+)

DNA

9.6 (guanine)41, 10.5 (thymine)41

Monolayer Experiments Langmuir Isotherms Measuring compression isotherms of Langmuir monolayers gives first information about area requirements. OO4, DOPE, DPPE, OO4/DOPE (1:3), and OO4/DPPE (1:3) monolayers were spread on buffer (pH 3) and on the same buffer containing 0.1 mM calf thymus DNA at 20 °C. The π-A-isotherms are presented in Figures 2A and 2B. The buffer and the pH were chosen because the lipid OO4 is fully protonated at this pH value and hence we have no additional effects of different protonated lipid ACS Paragon Plus Environment

Page 11 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

species. Although pH 3 is far from biological pH values, it is very useful to understand interactions of OO4 with the co-lipids and/or DNA.

Figure 2: π-A-isotherms of A) OO4 (black), DOPE (red) and DPPE (green) on bromide containing buffer (pH 3) at 20 °C. B) OO4 (black), OO4/DOPE (1:3) (blue) and OO4/DPPE (1:3) (orange) on bromide containing buffer (straight lines) and the corresponding pressure/area isotherms on calf thymus DNA (1 mMnucleotides) in bromide containing buffer (pH 3) at 20 °C (dashed-dotted lines). C) Adsorption isotherms of ct-DNA to OO4 monolayers at pH 3, pH 7 and pH 10 (bromide-containing buffer, 2 mM) at 20 °C

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

OO4 is in a liquid-expanded state along the π-A-isotherm (Figure 2A) due to its unsaturated oleyl chains (C18:1). The results are in agreement with IRRAS (Figure 4A). DOPE contains also two oleyl chains and is therefore also in the liquid-expanded phase state. Since the head group of DOPE is smaller than the one of OO4, DOPE occupies a smaller in-plane area. Unlike the zwitterionic DOPE, the phase behavior of OO4 depends strongly on the pH value. At pH 3, the OO4 molecules have three charged amine groups leading to electrostatic repulsions,32 which result in large molecular areas. DPPE is in the liquid-condensed phase state already at close to zero pressures (re-sublimation). Due to strong van-der-Waals interactions of the saturated hexadecyl chains (C16:0), DPPE is tightly packed and occupies the smallest area per molecule (Figure 2A). With the area per molecule at 30 mN/m we can calculate the form factor of DPPE resulting in a value of 1.09 (using an A (30 mN/m) = 40.9 Å2 and the formulas presented in the SI of an already published article)65. This value proves the assumed cylindrical shape of DPPE. The binary mixture OO4/DOPE (1:3) is in the liquid-expanded phase state, but requires smaller molecular areas than OO4 alone. The isotherm of OO4/DOPE (1:3) looks similar to the isotherm of DOPE (compare Figures 2A and 2B). The addition of DPPE to OO4 changes the phase state. The stiff DPPE enhances the rigidity of the OO4/DPPE (1:3) monolayer (compare Figures 2A and 2B). OO4/DPPE (1:3) is in the liquid-condensed phase state with smaller molecular area than OO4 and OO4/DOPE (1:3). The addition of DNA to the subphase changes the behavior. Surface pressure and IRRAS experiment with DNA containing subphase in absence of cationic lipids show that the DNA will not appear at the air water interface within 12 h. But DNA adsorbs at lipid monolayers. The adsorbed DNA has different effects on the lipid monolayers. In case of OO4 + DNA (at pressures above 13 mN/m) and OO4/DOPE (1:3) + DNA, the area per molecule even decreases compared to the same monolayers in absence of DNA. Furthermore, the slope of the isotherms changes drastically, and a surface pressure of 25 mN/m is hardly reachable. This indicates that there is no stable Langmuir monolayer anymore and most likely some material dives into the subphase (hypothesis: lipoplex formation in bulk) upon compression. A comparable behavior was described for the assembling of lipoplexes at the air water ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

interface.42 In contrast, OO4/DPPE (1:3) + DNA and OO4 + DNA (at pressures below 13 mN/m) show an increase in the required area per molecule compared to OO4/DPPE (1:3) and OO4. At 30 mN/m (lateral pressure in biological membranes)43, the difference in area for OO4/DPPE (1:3) in presence and absence of DNA is about 29 Ų per molecule, what should be the area occupied by DNA segments adsorbing to the air-water interface and penetrating the OO4/DPPE monolayer. Furthermore, there is a kink in the π-A-isotherms of OO4 + DNA and OO4/DPPE (1:3) + DNA at 10.6 mN/m and 31.5 mN/m, respectively. IR-spectra clearly indicate that OO4 + DNA remains in the fluid phase state as characterized by high wavenumbers (asym-CH2 stretching vibrations ~ 2927 to 2926 cm-1). This hints that DNA and/or lipoplexes, respectively, are squeezed out from the monolayer at high surface pressures and dive into the subphase. pH dependence of OO4 at the air-water interface

Figure 3: pH dependence of OO4 on bromide containing buffer (2 mM, different pH values) at 20 °C. A) Molecular area and position of the symmetric and asymmetric methylene stretching bands at π = 30 mN/m and B) protonation state of OO4 at π = 5 mN/m (o) and π = 30 mN/m (▪). For experimental details see Tassler at al.32

As mentioned above, the cationic lipid OO4 is sensitive to the pH, therefore, its behavior in absence of the co-lipids was studied at different pH values. Figure 3A illustrates the isotherms of OO4 at three different pH values. The insert (Figure 3A) shows the methylene stretching modes for a pH titration in IRRAS experiments. Due to the two unsaturated oleyl chains (C18:1), OO4 remains in the fluid phase (gauche conformation indicated by the typical large wavenumbers) even at high pH values

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 36

(uncharged head group), whereas the head group protonation changes drastically with the pH values (Figure 3B). At pH 3, OO4 is fully protonated, and at pH 8, completely unprotonated. Since the protonation state is a key parameter for the interaction between DNA and cationic lipids, the pH value must have a huge influence on the adsorption of DNA to OO4 monolayers. The adsorption of ct-DNA to OO4 monolayers has been measured at pH 3, pH 7 and pH 10 (Figure 2C). In fact DNA is a polyelectrolyte and its protonation state also strongly depends on the pH value. At physiological pH 7.4, DNA is negatively charged due to the phosphate backbone (PO2-). In contrast, the cationic lipid is weakly protonated at pH 7.4 (only 59% of the molecules carry one positive charge). At the investigated pH of 7, 94% of the lipids have only a single protonated amine group (theoretically three amine groups can be protonated if the tertiary amine is excluded) available to interact electrostatically with DNA (see Figure 3B). At pH 10, the lipids are completely unprotonated, while DNA is close to strand separation. At higher pH values, guanine and thymine will be deprotonated and exist as negatively charged conjugated bases.44 On the contrary, DNA is weakly charged at pH 3, since its (PO2-) groups will be balanced by protons (H+). Lowering the pH value will induce hydrolysis and results in denaturation of DNA, since the hydrogen-bonds between the corresponding bases (A=T and G≡C) are going to break due to electrostatic repulsion of the positive charges. At pH 3, OO4 is almost fully protonated (see Figure 3B).

Table 2: Increase in surface pressure (Δπ) due to calf thymus DNA adsorption to OO4 monolayers on buffers with different pH values at 20 °C. The adsorption isotherms are given in the Supporting information Figure 2C.

ct-DNA containing bromide ion based buffer pH 3 pH 7 pH 10

πinitial [mN/m]

Δπ [mN/m]

8.1 9.2 9.1

1.7 11.8 11.2

The resulting change in the surface pressure (Δπ = π60min - πinitial) due to the adsorption of ct-DNA after 60 min is given in Table 2. At pH 7, the negatively charged DNA is attracted by the weakly protonated OO4 due to electrostatic interactions between the phosphate diester backbone (PO2-) of DNA and the amine groups (R-NH3+) in the lipid head group. The surface pressure increases strongly ACS Paragon Plus Environment

Page 15 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

during adsorption until an adsorption-desorption equilibrium is established. Interestingly, even at pH 10, at which the OO4 is unprotonated, a Δπ comparable to that at pH 7 was observed (Table 2), indicating an interaction with DNA. The increasing surface pressure can be explained by two scenarios: i) the DNA partially penetrates into the lipid monolayer,37 or ii) the DNA triggers the protonation of the lipids at the interface resulting in increased space requirements for the lipid head groups.45 Surprisingly, at pH 3 only a small increase in the surface pressure is detected (Δπ = 1.7 mN/m). Therefore, it seems that DNA does not occupy additional space at the surface or trigger conformational changes in the lipid leading to increased area requirement per molecule. A change in the protonation state of the lipids can also be excluded, because the lipid is at pH 3 nearly completely positively charged. Most-likely, DNA segments are aligned underneath the monolayer by the protonated amine groups of OO4. The position of the methylene stretching vibrational modes gives information about the phase state of the lipids. OO4 is in the fluid phase with most oleyl chain segments in gauche conformation on the bromide containing buffers at pH 3, pH 7 and pH 10 as can be seen in Figure 4A (asymmetric methylene stretching vibration above 2924 cm-1). Selected IRRA-spectra can be found in the Supporting Information, Figure S1A. Upon compression the wavenumbers decrease only slightly, meaning that OO4 remains in the fluid phase state even at high surface pressures. The adsorption of DNA does not influence the phase state of OO4 strongly. At pH 3, the wavenumbers increase by about 1 cm-1, indicating a slight increase in fluidity of the OO4 monolayer. At pH 7 and pH 10, the wavenumbers are marginally lower than for OO4 without adsorbed DNA but within the error of ± 0.2 cm-1. We also get information about the amount of adsorbed DNA from IRRAS evaluating the phosphate modes. For OO4 on all DNA containing buffers, the νasym PO2- band is at 1220 cm-1 which implies strong hydration of the DNA phosphate groups and the presence of hydrogen bonds (see Supporting Information, Figure S1B).46 The relative amount of attached calf thymus DNA to the lipid monolayers of OO4 on the bromide ion based buffers at pH 3, pH 7 and pH 10 has been estimated by the integration of the antisymmetric PO2- stretching vibration band after subtraction of the spectra of DNA solutions at the corresponding pH values. The intensity plots of the νasym PO2- band versus the ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

area per molecule of the corresponding lipids are given in Figure 4B. The relative amount of attached DNA to the OO4 monolayer strongly depends on the pH values of the subphase. As expected, the largest amount of DNA is attached to OO4 at pH 3, at which OO4 has the highest protonation state. The increase of the pH value (presults in a lower protonation degree of OO4 and therewith the amount of attracted DNA decreases. At pH 10, the relative amount of attached DNA is very small. At pH 3 and pH 7, the amount of adsorbed DNA increases with decreasing molecular areas (increasing charge density) upon compression.

Figure 4: A) Phase state of OO4 on bromide containing buffer (filled symbols) and 0.1 mM calf thymus DNA containing bromide buffer (2 mM) (empty symbols) at different pH values (pH 3 – black, pH 7 – red, and pH 10 blue) - Lorentzian fit maximum as position of the νasym CH2 band (20 °C, s-polarized light, incidence angle of 40°) versus lateral pressure and B) attached amount of calf thymus DNA to a Langmuir monolayer of OO4 – Integral of νasym PO2- band versus area per molecule from Langmuir isotherms. Dashed lines are linear fits.

In summary, the IRRAS experiments demonstrate that OO4 can bind DNA efficiently in a wide pH range from pH 3 to pH 7. The adsorption of DNA at monolayers composed of the lipid mixtures was not investigated, because the phospholipids (co-lipids) contain also phosphate in the head group what makes it more difficult to quantify the amount of DNA. Therefore, quartz microbalance experiments have been performed to answer this open question in bulk experiments. Bulk Experiments Quantification of adsorbed DNA by Quartz Crystal Microbalance (QCM) experiments The QCM-D technique has been used in order to quantify the amount of DNA, which attaches to mixed lipid bilayers. The experiments have been performed at pH 3. The obtained results and

ACS Paragon Plus Environment

Page 16 of 36

Page 17 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

calculated theoretical mass depositions are summarized in Table 3. The zwitterionic lipid 1,2dimyristoyl-sn-glycero-3-phosphocholine (DMPC, C14:0 alkyl chains, Tm = 24 °C) was used as reference. The obtained frequency shift of DMPC amounts to 26 Hz (see Figure S2, Supporting Information), which clearly indicates the formation of a supported lipid bilayer (SLB).47, 48 As already described in literature, the formation of supported DMPC bilayers at pH 3 takes place without the observation of a critical coverage.48 The resulting mass deposition is higher than the expected theoretical mass. The difference (19 ng/cm²) is most-likely based on 3 to 4 adsorbed water molecules (strongly hydrated phosphocholine head group). From NMR measurements it is known, that one DMPC molecule binds 4 to 5 water molecules in the Lß’ phase state.49

Table 3: Frequency shift (∆f) and resulting mass deposition (∆m) compared to theoretical mass.

sample

remarks

∆f [Hz]

∆m [ng/cm²]

theoretical mass [ng/cm²]

DMPC

reference

26

460

441

ct-DNA

1 base pair

25.7

OO4/DOPE (1:3)

28

496

OO4/DOPE (1:3) + ct-DNA

16

283

OO4/DPPE (1:3)

16.5

292

OO4/DPPE (1:3) + ct-DNA

15.5

274

469

601

The supported OO4/DOPE 1:3 (n:n) bilayer is also formed without observing a critical packing density of vesicles on the surface (Θc) (see Figure 5A). Its theoretical mass amounts to 469 ng/cm2. However, the measured frequency shift (Δf = 28 Hz) results in a mass deposition of 496 ng/cm2. The mass difference of 27 ng/cm2 is due to 5 water molecules attached to the SLB. Specular X-ray reflectivity data of OO4 monolayers support the assumption of hydrated amino head groups (see XRR curves Figure S3 and Table S1, Supporting Information). Also DOPE can bind up to 12 water molecules in the liquid-crystalline phase state.50 Adding DNA to the OO4/DOPE SLBs causes a frequency shift of 16 Hz.

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The mass of adsorbed DNA amounts to 283 ng/cm² (~2.58x1011 kbp DNA per cm2 OO4/DOPE bilayer equivalent to 2.5 base pairs per 100 Å2). The Langmuir isotherm of the OO4/DOPE mixture at 30 mN/m indicates that one molecule needs 60 Å2 (isotherm on the pure buffer without DNA). This value translates into 1.5 base pairs per lipid molecule. A significantly lower mass deposition was observed for OO4/DPPE (see Figure 5B) with only a small shift in frequency of 16.5 Hz. The resulting mass deposition amounts to 292 ng/cm². Compared to the theoretical value (mtheoretical = 601 ng/cm²), the small Δm-value implies a lipid monolayer or rather lipid bilayer patches of OO4/DPPE on the quartz crystal instead of a SLB. If we assume the latter case, roughly half of the area is covered with lipid bilayer patches. The adsorption of calf thymus DNA to this system caused a frequency shift of 15.5 Hz. The measured mass deposition of DNA to the OO4/DPPE bilayer patches (Δm = 274 ng/cm²) is marginally smaller than that observed for OO4/DOPE (Δm = 283 ng/cm²). Assuming, that the cationic charged lipid patches bind the highest proportion of the DNA, ~5.16 x1011 kbp DNA per cm² are attached to OO4/DPPE bilayers and therewith roughly 2 base pairs per lipid molecule (using a molecular area of 40 Å2 at 30 mN/m). This assumption is justified by the work of Vandeventer at al., they demonstrate that only small amounts of DNA attach to the untreated silica covered quartz crystal.66 Consequently, ct-DNA more likely attaches to positively charged OO4/DPPE patches than to negatively charged silica.

Figure 5: Δf(t) of A) OO4/DOPE 1:3 and B) OO4/DPPE 1:3 (1 mg/mL) in bromide containing buffer (pH 3) and calf thymus DNA (0.1 mM) containing bromide buffer (pH 3) at 20 °C. The 3rd overtone is shown. ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Although we investigated the adsorption of DNA at pH3 and the OO4/DPPE bilayer did not cover the whole area of the crystal we can assume that OO4/DPPE bilayers can bind a higher amount of DNA compared with OO4/DOPE bilayer. If we assume that OO4 carries the same charge in both mixtures, the charge density is higher for the OO4/DPPE 1:3 mixture because the lipids need less space compared to the OO4/DOPE 1:3 mixture. A surface with a higher charge density should attract more DNA. Protonation state of OO4 in bulk The protonation state of the cationic lipid OO4 in vesicles (bulk system) has been estimated by pHdependent ζ-potential measurements. The obtained titration curve is presented in Figure 6. The zeta potential is the potential at the shear plane of the ion cloud surrounding a particle, consequently the real surface charge (Nernst potential) is not accessible. Furthermore, the zeta potential is used to estimate the colloidal stability of a system. For liposomes, a ζ-potential above ± 30 mV has been reported for colloidal stable systems.51 OO4 has a positive ζ-potential (>30 mV) below pH 10. The isoelectric point at pH 11 leads to the first assumption that the lipids in the vesicle are partially charged up to this pH-value. The ζ-potential is constant between pH 3 and pH 7, whereas the protonation state in the monolayer system decreases from pH 3 to pH 7 (compare Figure 3B and Figure 6). An increased ζ-potential is observed between pH 8 and pH 9. However, it is very unlikely that the protonation state (positive charge) increases. Above pH 9, the ζ-potential decreases continuously. Consequently, the monolayer experiments (especially TRXF)32 give exact information about the protonation state of OO4 in contrast to the ζ-potential. Nevertheless, the zeta potential curve is needed to better understand the structural investigations of the 3D systems.

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6: ζ-potential titration curve of OO4 in 20 mM NaCl at different pH values. The pH was adjusted by adding NaOH or HCl. Dashed lines are for guiding the eyes only.

Small- and Wide-Angle X-Ray Scattering in non-physiological HBr buffer

Figure 7: A) SAXS and B) WAXS of OO4 in bromide containing buffer pH 3 (black line) and pH 10 (red line) at 25 °C.

At 25 °C, OO4 arranges in multi-lamellar bilayers in the bromide containing buffer pH 10 (d = 55.9 Å, Figure 7A). The alkyl chains are in the liquid-crystalline phase (Figure 7B). The lamellar phase has a high correlation length and the reflexes are visible up to the 4th order. In contrast, the SAXS pattern of OO4 in bromide containing buffer pH 3 is poorly resolved. A broad peak is observed indicating non-correlated lamellar bilayers (d = 55.1 Å), an effect of the electrostatic repulsion (ζpotential ~ 43 mV, see Figure 6). The correlation length is extremely small. Additionally, the small

ACS Paragon Plus Environment

Page 20 of 36

Page 21 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

peaks at low q values indicate the co-existence of a cubic mesophase. The Bragg peaks are in the ratio √3:√4:√8:√11:√12:√16:√19:√20:√24:√27 and characterize a micellar cubic phase with Fm3m symmetry (Qα225, see Supporting Information Table S2). The cubic Fm3m lattice parameter amounts to 283.3 Å. The characteristic cubic lattice parameter a is the slope of the linear function passing through the origin (0, 0). Here, h, k and l are Miller indices of the cubic lattice and the slope is equal to 1/a.52 The appearance of the micellar cubic phase indicates that OO4 has a conical shape, meaning that the highly charged head group needs more space than the fluid alkyl chains. OO4 was mixed with either DOPE or DPPE in order to adjust the liposome properties like charge density and phase behavior. DOPE forms inverted hexagonal cylinders due to the inverse conical molecular shape (the fluid chains require more space than the head group). DPPE forms multilamellar vesicles. For the mixtures with calf thymus DNA, the N/P-ratio of 4 was chosen because it showed the best performance in the transfection experiments.27 At pH 3 and 25 °C, DOPE has a hexagonal lattice parameter a equal to 74.4 Å (Supporting Information Figure S4A). The mixing of DOPE and OO4 in the ratio 1:3 results in phase separation as can be seen in Figure 8A. The lamellar phase LOO4 (d = 53.4 Å) corresponds most likely to phase separated OO4, since pure OO4 arranges in multi-lamellar bilayers with a similar d-value (55.1 Å). Additionally, there are Bragg peaks in the SAXS pattern in a ratio of √6:√8:√14:√16:√20:√22:√24:√34:√41 indicating an inverted cubic Ia3d phase (Qα230, so-called Gyroid (G) minimal surface, see Table S3).53, 54 The cubic Ia3d lattice parameter amounts to 187.6 Å. The cubic phase contains most likely DOPE and OO4 for three reasons. First, because the Ia3d phase is usually located between lamellar bilayers with chains in the liquid-crystalline state (Lα like OO4) and inverted hexagonal phases (HII like DOPE) in the phase diagram.55 Second, no phase-separated DOPE can be detected. Third, for OO4 without co-lipid a cubic phase of the Fm3m symmetry coexisting with the lamellar phase was found. Adding DNA to the system OO4/DOPE 1:3 (n:n) leads to further coexisting phases (Figure 8B). The cubic phase disappears, while the lamellar phase remains, and two hexagonal phases appear (peak ratio

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

√1:√3:√4:√7:√9:√12:√13:√16:√21). The lamellar phase LOO4 (d = 51.5 Å) is most probably formed by phase-separated OO4 (d = 55.1 Å) or it is an OO4-rich phase, while the hexagonal phase HDOPE (a = 68.2 Å) contains most likely phase-separated DOPE (a = 74.4 Å) or is a DOPE-rich phase. Therefore, an inverse hexagonal structure is very likely. Nevertheless, DNA, with a double helix diameter of 20 Å,56 seems not to be incorporated in these two phases. The DNA is most probably complexed inside the tubes of a hexagonal lipoplex indexed as HDNA. Since the lattice parameter of HDNA is 138.1 Å and therewith much larger as the a-value determined for HDOPE, it is reasonable to assume that the DNA is completely incorporated in the HDNA phase. If the HDNA phase is an inverted hexagonal or a honey comb arranged lipoplex structure cannot be determined by the pattern. A DOPE-rich lipid mixture could lead to the first described structure due to the inverse conical shape of DOPE,20 while an OO4-rich mixture would lead to the honeycomb hexagonal lipoplex due to the conical shape of OO4.57 The alkyl chains in all three phases are in the liquid-crystalline phase state.

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 8: SAXS pattern of A) OO4/DOPE (black line), B) OO4/DOPE + DNA (red line) in comparison with OO4/DOPE (black line), C) OO4/DPPE (black line) and D) OO4/DPPE + DNA (red line) in comparison with OO4/DPPE (black line) at 25 °C in bromide containing buffer pH 3. All lipids were mixed in the molar ratio OO4/co-lipid 1:3. The samples with DNA were complexed in a N/P-ratio of 4.

The mixtures with DPPE show a different behavior. At pH 3 and 25 °C, the multi-lamellar DPPE is in the gel phase state with chains in all-trans conformation (d = 54.9 Å) (Supporting Information Figure S4). The 1:3 mixture of DPPE and OO4 arranges in multi-lamellar bilayers (d = 54.8 Å, see Figure 8C). The shoulder at the first reflex might indicate a phase separation into two lamellar phases (OO4-rich and DPPE-rich), but the higher order peaks are missing to support this assumption. Adding DNA to the OO4/DPPE mixture leads to the appearance of two lamellar phases which can be evaluated separately as shown in Figure 8D. At least one phase has tilted chains in gel state (orthorhombic packing of the chains indicated by two peaks in the WAXS region). One lamellar phase (named as LOO4/DPPE, d = 54.8 Å) is comparable with the structure found for the OO4/DPPE mixture. The second phase (named as LDNA) has a higher d-value (d = 64.6 Å) and incorporates most probably the DNA. The broad peak at q = 2.47 nm-1 corresponds to a DNA-DNA in-plane distance of 25.5 Å. In the LDNA lipoplex the DNA is complexed sandwich-like between the lipid bilayers and the DNA rods align in a 1D lattice.58 With the assumption that the DNA rod diameter is 20 Å,56 the observed DNADNA in-plane distance of 25.5 Å indicates a tight packing. Obviously the high positive charge density of the lipid composition compensates the repulsive forces between the DNA strands. This tight packing means also that the OO4/DPPE mixture has a high DNA loading capacity, even with the coexisting empty lamellar phase. This observation is in good agreement with the QCM experiments presented above. Small- and Wide-Angle X-Ray Scattering in MES buffer The X-ray experiments were additionally performed in MES buffer to allow the comparison with the biological experiments.27 The SAXS and WAXS patterns of OO4, DOPE and DPPE at 25 °C and pH 6.5 are presented in Figures 9A and 9B, respectively. The cationic lipid OO4 is most probably in the lamellar mesophase with its oleyl chains in the liquid-crystalline phase state (Lα). The Bragg peaks are very broad indicating a structure with many defects. Most likely the electrostatic repulsions between

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the charged head groups are responsible for the small correlation length (ζ-potential ~ 43 mV, Figure 6). The estimated d-value amounts to 53.1 Å. Comparable to the results at pH 3, DOPE arranges in inverted hexagonal cylinders with a lattice parameter of 74.2 Å. DPPE forms multilamellar vesicles with tilted chains in the gel state (Lß’, d = 56 Å). The Bragg peak at qH = 13 nm-1 can be associated to intermolecular hydrogen bonds.

ACS Paragon Plus Environment

Page 24 of 36

Page 25 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 9: A) SAXS and B) WAXS of the single lipids OO4 (black line), DOPE (red line) and DPPE (green line), C) SAXS and D) WAXS of OO4/DOPE (black line) and OO4/DOPE + DNA (red line), E) SAXS and F) WAXS of OO4/DPPE (black line) and OO4/DPPE + DNA (red line) in MES buffer (pH 6.5) at 25 °C. The samples with DNA were complexed in a N/P-ratio of 4. At 20 °C and 37 °C, comparable diffraction patterns were measured (Supporting Information Figure S5-S7). F) The observed peaks ca be indexed as (10), (01) and (1-1).

The 1:3 mixture of OO4 and DOPE gives a broad signal in the SAXS region (Figure 9C) and a halo in the WAXS region (Figure 9D). Since no distinct peaks corresponding to DOPE can be seen in the SAXS pattern, OO4 and DOPE seemed to mix well arranging in a Lα mesophase. Adding DNA to the OO4/DOPE mixture leads to the formation of two phases: a micellar cubic phase (Qα223) and a lamellar phase (Lα). Due to the large Pm3n lattice parameter (a = 353.4 Å, reflections at √13 and √17 are missing), DNA is most likely complexed within the cubic mesophase. The lamellar phase consists of phase-separated OO4/DOPE (LαOO4/DOPE, d = 67.3 Å), but we cannot safely exclude that it does not contain DNA. Lipids with a comparable backbone to OO4 can form cubic lipoplex structures. In recently published study, TT10 and OT10 were found to form cubic Im3m lipoplexes with DOPE in the ratio 1:4 complexed with calf thymus DNA (N/P = 4).38 Furthermore lipid 6, a malonic acid diamide of the first generation forms cubic Im3m lipoplexes after mixing with DOPE.64 At 25 °C and pH 6.5, the cationic lipid OO4 and the zwitterionic lipid DPPE form multi-lamellar phases in 1:3 mixtures. The repeating distance equals 56.1 Å (Figure 9E). The three Bragg peaks at q10 = 15 nm-1, q01 = 15.4 nm-1 and q1-1 = 16.5 nm-1 in the WAXS region (Figure 9F) characterize an oblique chain lattice. The additional peak at qH = 13 nm-1 can be assigned to intermolecular hydrogen bonds. Adding DNA to the OO4/DPPE liposomes leads to the appearance of two phases. The Bragg peaks in the SAXS region imply a co-existence of two lamellar mesophases, namely LOO4/DPPE (d = 56.1 Å) and LDNA (d = 75.7 Å). The latter one describes the lamellar lipoplex with DNA entrapped between lipid bilayers. The difference in the spacing amounts to ~20 Å,56 which is the diameter of the DNA double helix. The WAXS region does not change. The three Bragg peaks indicating an oblique unit cell and the peak characterizing H-bonds are still present. The signals seem not to be superimposed by a halo of lipids in the liquid-crystalline phase, what implies that the mixture

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

OO4/DPPE forms rarely described lamellar lipoplexes with lipids in the gel state and not the typical Lαc phase.58

Discussion

The presented work helps to understand the characteristics of both lipid mixtures in biological systems. OO4/DOPE 1:3 is the more effective nucleic acid carrier, which is even stable in presence of biological relevant substances (serum proteins, salt, glucosaminoglycans).27 The structural approach made in this study leads to the assumption, that the structures formed by the lipid formulations after complexing DNA determine the properties rather than charge density and capacity to bind DNA. If we assume that the charge state of OO4 in the OO4/DOPE 1:3 and the OO4/DPPE 1:3 mixtures is the same, the space requirement of the second mixture is smaller (see monolayer experiments) and therewith the charge density is higher. Therefore, the DNA binding capacity of the OO4/DPPE 1:3 mixture must be higher compared to the OO4/DOPE 1:3 mixture. A higher charge density would also imply a more stable complexation of DNA and therewith a higher stability against disassembling promoted by salts, anionic proteins and anionic polycarbonates. In reality, we have observed the opposite behavior.27 The reason seems to be the different structures of both lipid mixtures complexed with DNA. OO4/DOPE 1:3 forms cubic Pm3n structures at pH 6.5 which complex DNA, a lipoplex structure known to be fusiogenic. Such structures promote cellular entry and endosomal escape.59-61 However, there is a higher risk for disassembling by anionic molecules compared with the lamellar lipoplex structures. The SAXS pattern also proves a co-existing lamellar phase in OO4/DOPE 1:3 lipoplexes at pH 6.5. If the lamellar phase surrounds the cubic lipoplex structures, a highly protective effect against the anion-induced lipoplex disassembling would occur. This effect was observed in biological experiments supporting the assumption. Also freeze-fracture TEM images show only spherical structures,27 and dried samples of negatively stained TEM exhibit only aggregated spherical structures with a certain substructure (Figure S8, Supporting information). Therefore, we propose a cubic complexing DNA mesophase which is surrounded and therewith protected by multi-lamellar structures (Figure 10). Lowering the pH to 3 results in hexagonal lipoplexes and additional lamellar as well as hexagonal structures. Hexagonal mesophases are highly ACS Paragon Plus Environment

Page 26 of 36

Page 27 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

fusiogenic if they adopt the inverse hexagonal arrangement due to the high content of DOPE.11 The stable lipoplex structure, pre-formed at pH 6.5, becomes unstable during the endosomal maturation to the lysosome (pH 6.5 to pH 4.5)62. This model can explain the high serum stability and the high DNA transfer efficiency of OO4/DOPE 1:3, and further how this lipoplex formulation is able to overcome the polycationic dilemma.27, 63 The OO4/DPPE 1:3 mixture forms lamellar complexes with DNA both at pH 6.5 and pH 3. Lamellar lipoplexes are known to be stable in presence of serum and other polyanionic biological molecules, but due to the low interaction potential with cellular membranes the efficiency of DNA transfer is often lower compared to non-lamellar phases.11 Our investigations demonstrate that both the efficiency and the stability in presence of salts and polyanions are lower compared to the OO4/DOPE 1:3 mixture.27 Consequently, the interaction between the lipid mixture and DNA is weak although the DNA binding capacity is higher in the DPPE-containing mixture compared with the DOPE-containing mixture.

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 36

Figure 10: Proposed structures and uptake of DNA complexed with the two investigated lipid mixtures. Table 4: Structures of liposomes and lipoplexes in HBr-containing buffers (pH 3 and pH 10) and MES buffer (pH 6.5) at 25 °C.

OO4 OO4 DOPE OO4/DOPE OO4/DOPE/DNA DPPE OO4/DPPE OO4/DPPE/DNA OO4 DOPE OO4/DOPE

pH 3 10

3

6.5

mesophase cubic Fm3m (Qα225)/lamellar (Lα) lamellar (Lα) hexagonal (Hα) cubic Ia3d (Qα230) hexagonal (Hα) lamellar (Lß’) lamellar lamellar with 1D alignment of DNA lamellar (Lα) hexagonal (Hα) lamellar (Lα)

ACS Paragon Plus Environment

d/a[Å] 283.3 /55.1 55.9 74.4 187.6 138.1 54.9 54.8 64.6 53.1 74.2

additional phase

LOO4 HDOPE, LOO4

LOO4/DPPE

Page 29 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

OO4/DOPE/DNA DPPE OO4/DPPE OO4/DPPE/DNA

cubic Pm3n (Qα223) lamellar (Lß’) lamellar lamellar

353.4 56 56.1 75.7

LOO4/DOPE

LOO4/DPPE

Conclusion

In continuation of our previous work, the peptide-mimicking lysine-based amino-functionalized lipid OO4 was investigated in mixtures with the commonly used unsaturated co-lipid DOPE and the saturated DPPE in 2D and 3D model systems. The choice of the zwitterionic co-lipid has an enormous influence on the cationic lipid OO4 in terms of phase behavior, interaction with calf thymus DNA as well as biological behavior.27 At the air-liquid interface, OO4 is in the liquid-expanded phase state due to its two oleyl chains independent of the pH value. Since OO4 contains three primary amino groups in the head group, its protonation state is pH-sensitive. As a consequence, OO4 is fully protonated at pH 3 and completely uncharged above pH 8. Therefore, different scenarios of DNA coupling to OO4 monolayers have been proposed for different pH values. The saturated co-lipid DPPE (C16:0) strongly increases the packing density in the OO4/DPPE monolayer in comparison to the OO4/DOPE monolayer containing two unsaturated lipids. In bulk, the structures of different lipid mixtures in presence and absence of DNA and at different pH values have been investigated (Table 4). OO4 arranges in multi-lamellar bilayers with fluid chains at pH 10. Interestingly, OO4 mixes better with DPPE, since both lipids form lamellar mesophases, than with DOPE, since the phase structures are different due to drastically different molecule shapes even with the same chain composition. OO4/DPPE liposomes and OO4/DPPE/DNA lipoplexes are multilamellar. The tightly complexed DNA aligns in a 1D pattern between the sandwich-like OO4/DPPE bilayers. Furthermore, OO4/DPPE bilayers have a higher capacity to bind DNA compared to OO4/DOPE. The mixing of DOPE and OO4 results in phase separation and the formation of non-lamellar mesophases. At pH 3, OO4/DOPE forms bicontinuous cubosomes (Qα230), while phase-separated OO4

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

remains in the Lα mesophase. DNA is complexed by OO4/DOPE in an inverted hexagonal lipoplex at pH 3. At pH 6.5, it is assumed that OO4/DOPE arranges in lamellar bilayers. However, the adding of DNA causes phase separation and the formation of a cubic lipoplex (Qα223). Phase-separated OO4/DOPE arranges in the lamellar mesophase with molten chains. Cubic lipoplexes encapsulated in lamellar structures have been proposed. This is the most reasonable explanation for the high stability and good efficiency of this lipoplex formulation. The exchange of the co-lipid (DOPE → DPPE) has two major consequences for the gene transfection because of changed lipoplex structure and DNA binding capacity. The lipoplexes with DOPE are hexagonal (pH 3) and cubic (pH 6.5), while the lipoplexes with DPPE remain lamellar at both pH values. Non-lamellar lipoplexes are supposed to rapidly fuse with the endosomal bilayer and allow the release of DNA with high efficiency. In the lamellar lipoplex, DNA is complexed between the OO4/DPPE bilayers. DPPE increases the charge density in the mixed monolayer. However, the fusiogenic potential of the lamellar lipoplex is low what results in poor DNA release performance.

Acknowledgement This work was supported by the Max Planck Society, the German Academic Exchange Service (Project-ID 57064041) and the Deutsche Forschungsgemeinschaft (DFG) project-ID 396823779. We thank the ESRF for beam time and for providing excellent facilities and support, especially Dr. Gudrun Lotze and Dr. Alessandro Mariani. Further we acknowledge Dr. Mariusz Pietrzak, Dr. Tomasz Kobiela for the support with the QCM-D E1 device in Warsaw, Poland. We thank PD Dr. Annette Meister and PD Dr. Simon Drescher (Martin Luther University Halle Wittenberg) for the TEM image. Supporting Information  Calculation of the theoretical mass for QCM-D  IRRA spectra of OO4 on bromide containing buffer and ct-DNA containing bromide buffer (CH2 and PO2- stretching vibration region)  QCM-D of DMPC  Specular X-Ray Reflectivity curves of OO4 on bromide containing buffer  Cubic Mesophases  SAXS/WAXS of DOPE and DPPE in bromide containing buffer  Additional SAXS/WAXS data of OO4, DOPE and DPPE in MES buffer at 20 °C and 37 °C  TEM image

ACS Paragon Plus Environment

Page 30 of 36

Page 31 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Abbreviation: a = lattice parameter, arb. units = arbitrary units, π-A-Isotherm = surface pressureArea-Isotherm, ct- DNA = calf thymus Deoxyribonucleic acid, d = spacing repeating distance (including one lipid bilayer and one neighbouring water layer), DMPC = 1,2-dimyristoyl-sn-glycero-3phosphocholine,

DOPE

=

1,2-dioleoyl-sn-glycero-3-phosphoethanolamine,

DPPE

=

1,2-

dihexadecanoyl-sn-glycero-3-phosphoethanolamine, Hα = hexagonal liquid-crystalline phase state, IRRAS = Infrared Reflection Absorption Spectroscopy, Lα = lamellar liquid-crystalline phase state in bulk, Lβ = gel state with untilted alkyl chains in bulk, Lβ’ = gel state with tilted alkyl chains in bulk, N/P ratio = ratio of moles of amine groups of cationic lipids and DNA phosphate groups, OO4 = N-{6Amino-1-[N-(9Z)-octadec-9-enylamino]-1-oxohexan-(2S)-2-yl}-N`-{2-[N,N-bis(2aminoethyl)amino]ethyl}-2-[(9Z)-octadec-9-enyl]propandiamide, Qα223 = micellar cubic Pm3n lattice liquid-crystalline phase state, Qα225 = micellar cubic Fm3m lattice liquid-crystalline phase state, Qα230 = bicontinuous cubic Ia3d lattice liquid-crystalline phase state, QCM = Quartz Crystal Microbalance, SAXS = Small-Angle X-Ray Scattering, TRXF = Total Reflection X-Ray Fluorescence, WAXS = Wide-Angle X-Ray Scattering, XRR = Specular X-Ray Reflectivity

References 1. Friedmann, T.; Roblin, R., Gene Therapy for Human Genetic Disease? Science 1972, 175, (4025), 949-955. 2. O'Connor, T. P.; Crystal, R. G., Genetic medicines: treatment strategies for hereditary disorders. Nature Reviews Genetics 2006, 7, 261. 3. Adhikari, B.; Nanda, J.; Banerjee, A., Multicomponent hydrogels from enantiomeric amino acid derivatives: helical nanofibers, handedness and self-sorting. Soft Matter 2011, 7, (19), 89138922. 4. Maria, S.; Hildegard, B., Improving the Quality of Adeno-Associated Viral Vector Preparations: The Challenge of Product-Related Impurities. Human Gene Therapy Methods 2017, 28, (3), 101-108. 5. Touchot, N.; Flume, M., The payers' perspective on gene therapies. Nature Biotechnology 2015, 33, 902. 6. Senior, M., After Glybera's withdrawal, what's next for gene therapy? Nature Biotechnology 2017, 35, 491. 7. Yin, H.; Kanasty, R. L.; Eltoukhy, A. A.; Vegas, A. J.; Dorkin, J. R.; Anderson, D. G., Non-viral vectors for gene-based therapy. Nature Reviews Genetics 2014, 15, 541. 8. Glover, D. J.; Lipps, H. J.; Jans, D. A., Towards safe, non-viral therapeutic gene expression in humans. Nature Reviews Genetics 2005, 6, 299. 9. Dan, N.-D.; James, H.; Caroline, J. S., Structure-Activity Relationship in Cationic Lipid Mediated Gene Transfection. Current Medicinal Chemistry 2003, 10, (14), 1233-1261. 10. Hart, S. L., Multifunctional nanocomplexes for gene transfer and gene therapy. Cell Biology and Toxicology 2010, 26, (1), 69-81.

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

11. Ma, B.; Zhang, S.; Jiang, H.; Zhao, B.; Lv, H., Lipoplex morphologies and their influences on transfection efficiency in gene delivery. Journal of Controlled Release 2007, 123, (3), 184-194. 12. Du, Z.; Munye, M. M.; Tagalakis, A. D.; Manunta, M. D. I.; Hart, S. L., The Role of the Helper Lipid on the DNA Transfection Efficiency of Lipopolyplex Formulations. Scientific Reports 2014, 4, 7107. 13. Dabkowska, A. P.; Barlow, D. J.; Campbell, R. A.; Hughes, A. V.; Quinn, P. J.; Lawrence, M. J., Effect of Helper Lipids on the Interaction of DNA with Cationic Lipid Monolayers Studied by Specular Neutron Reflection. Biomacromolecules 2012, 13, (8), 2391-2401. 14. Lin, A. J.; Slack, N. L.; Ahmad, A.; George, C. X.; Samuel, C. E.; Safinya, C. R., ThreeDimensional Imaging of Lipid Gene-Carriers: Membrane Charge Density Controls Universal Transfection Behavior in Lamellar Cationic Liposome-DNA Complexes. Biophysical Journal 2003, 84, (5), 3307-3316. 15. Ayesha, A.; M., E. H.; Kai, E.; X., G. C.; E., S. C.; R., S. C., New multivalent cationic lipids reveal bell curve for transfection efficiency versus membrane charge density: lipid–DNA complexes for gene delivery. The Journal of Gene Medicine 2005, 7, (6), 739-748. 16. Düzgüneş, N.; Felgner, P. L., [23] Intracellular delivery of nucleic acids and transcription factors by cationic liposomes. In Methods in Enzymology, Academic Press: 1993; Vol. 221, pp 303306. 17. Wasungu, L.; Hoekstra, D., Cationic lipids, lipoplexes and intracellular delivery of genes. Journal of Controlled Release 2006, 116, (2), 255-264. 18. Van Dijck, P. W. M., Negatively charged phospholipids and their position in the cholesterol affinity sequence. Biochimica et Biophysica Acta (BBA) - Biomembranes 1979, 555, (1), 89-101. 19. Cullis, P. R.; Van Dijck, P. W. M.; De Kruijff, B.; De Gier, J., Effects of cholesterol on the properties of equimolar mixtures of synthetic phosphatidylethanolamine and phosphatidylcholine. A 31P NMR and differential scanning calorimetry study. Biochimica et Biophysica Acta (BBA) Biomembranes 1978, 513, (1), 21-30. 20. Koltover, I.; Salditt, T.; Rädler, J. O.; Safinya, C. R., An Inverted Hexagonal Phase of Cationic Liposome-DNA Complexes Related to DNA Release and Delivery. Science 1998, 281, (5373), 78-81. 21. Seddon, J. M.; Cevc, G.; Marsh, D., Calorimetric studies of the gel-fluid (L.beta.-L.alpha.) and lamellar-inverted hexagonal (L.alpha.-HII) phase transitions in dialkyl- and diacylphosphatidylethanolamines. Biochemistry 1983, 22, (5), 1280-1289. 22. Janich, C.; Hädicke, A.; Bakowsky, U.; Brezesinski, G.; Wölk, C., Interaction of DNA with Cationic Lipid Mixtures—Investigation at Langmuir Lipid Monolayers. Langmuir 2017, 33, (39), 1017210183. 23. Zuhorn, I. S.; Bakowsky, U.; Polushkin, E.; Visser, W. H.; Stuart, M. C. A.; Engberts, J. B. F. N.; Hoekstra, D., Nonbilayer phase of lipoplex–membrane mixture determines endosomal escape of genetic cargo and transfection efficiency. Molecular Therapy 2005, 11, (5), 801-810. 24. Koynova, R., Lipid Phases Eye View to Lipofection. Cationic Phosphatidylcholine Derivatives as Efficient DNA Carriers for Gene Delivery. Lipid Insights 2008, 2, LPI.S864. 25. MacKintosh, F. C.; Safran, S. A., Phase separation and curvature of bilayer membranes. Physical Review E 1993, 47, (2), 1180-1183. 26. Safran, S. A.; Pincus, P. A.; Andelman, D.; MacKintosh, F. C., Stability and phase behavior of mixed surfactant vesicles. Physical Review A 1991, 43, (2), 1071-1078. 27. Giselbrecht, J.; Janich, C.; Pinnapireddy, S. R.; Hause, G.; Bakowsky, U.; Wölk, C.; Langner, A., Overcoming the polycation dilemma – Explorative studies to characterise the efficiency and biocompatibility of newly designed lipofection reagents. International Journal of Pharmaceutics 2018, 541, (1), 81-92. 28. Wölk, C.; Drescher, S.; Meister, A.; Blume, A.; Langner, A.; Dobner, B., General Synthesis and Physicochemical Characterisation of a Series of Peptide-Mimic Lysine-Based Amino-Functionalised Lipids. Chemistry – A European Journal 2013, 19, (38), 12824-12838. 29. Wölk, C.; Janich, C.; Bakowsky, U.; Langner, A.; Brezesinski, G., Malonic acid based cationic lipids – The way to highly efficient DNA-carriers. Advances in Colloid and Interface Science 2017, 248, 20-34. ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

30. Zhang, X.; Dai, L.; Wang, A.; Wölk, C.; Dobner, B.; Brezesinski, G.; Tang, Y.; Wang, X.; Li, J., The Directional Observation of Highly Dynamic Membrane Tubule Formation Induced by Engulfed Liposomes. Scientific Reports 2015, 5, 16559. 31. Schindler, T.; Nordmeier, E., The stability of polyelectrolyte complexes of Calf-Thymus DNA and synthetic polycations: Theoretical and experimental investigations. Macromolecular Chemistry and Physics 1997, 198, (6), 1943-1972. 32. Tassler, S.; Wölk, C.; Janich, C.; Dobner, B.; Brezesinski, G., Lysine-based amino-functionalized lipids for gene transfection: the protonation state in monolayers at the air-liquid interface. Physical Chemistry Chemical Physics 2017, 19, (30), 20271-20280. 33. Maltseva, E.; Brezesinski, G., Adsorption of Amyloid Beta (1-40) Peptide to Phosphatidylethanolamine Monolayers. ChemPhysChem 2004, 5, (8), 1185-1190. 34. Kauppinen, J. K.; Moffatt, D. J.; Mantsch, H. H.; Cameron, D. G., Smoothing of spectral data in the Fourier domain. Appl. Opt. 1982, 21, (10), 1866-1872. 35. Rodahl, M.; Höök, F.; Krozer, A.; Brzezinski, P.; Kasemo, B., Quartz crystal microbalance setup for frequency and Q-factor measurements in gaseous and liquid environments. Review of Scientific Instruments 1995, 66, (7), 3924-3930. 36. Sauerbrey, G., Verwendung von Schwingquarzen zur Wägung dünner Schichten und zur Mikrowägung. Zeitschrift für Physik 1959, 155, (2), 206-222. 37. Tassler, S.; Pawlowska, D.; Janich, C.; Dobner, B.; Wölk, C.; Brezesinski, G., Lysine-based amino-functionalized lipids for gene transfection: the influence of the chain composition on 2D properties. Physical Chemistry Chemical Physics 2018, 20, (10), 6936-6944. 38. Tassler, S.; Pawlowska, D.; Janich, C.; Giselbrecht, J.; Drescher, S.; Langner, A.; Wölk, C.; Brezesinski, G., Lysine-based amino-functionalized lipids for gene transfection: 3D phase behaviour and transfection performance. Physical Chemistry Chemical Physics 2018, 20, 17393-17405. 39. Isom, D. G.; Castañeda, C. A.; Cannon, B. R.; García-Moreno E., B., Large shifts in pKa values of lysine residues buried inside a protein. Proceedings of the National Academy of Sciences 2011, 108, (13), 5260-5265. 40. Oregioni, A.; Stieglitz, B.; Kelly, G.; Rittinger, K.; Frenkiel, T., Determination of the pKa of the N-terminal amino group of ubiquitin by NMR. Scientific Reports 2017, 7, 43748. 41. Verdolino, V.; Cammi, R.; Munk, B. H.; Schlegel, H. B., Calculation of pKa Values of Nucleobases and the Guanine Oxidation Products Guanidinohydantoin and Spiroiminodihydantoin using Density Functional Theory and a Polarizable Continuum Model. The Journal of Physical Chemistry B 2008, 112, (51), 16860-16873. 42. Oberle, V.; Bakowsky, U.; Hoekstra, D., Lipoplex Assembly Visualized by Atomic Force Microscopy. In Methods in Enzymology, Academic Press: 2003; Vol. 373, pp 281-297. 43. Marsh, D., Lateral pressure in membranes. Biochimica et Biophysica Acta (BBA) - Reviews on Biomembranes 1996, 1286, (3), 183-223. 44. Wood, J. L., pH-controlled hydrogen-bonding (Short Communication). Biochemical Journal 1974, 143, (3), 775-777. 45. Antipina, M. N.; Schulze, I.; Heinze, M.; Dobner, B.; Langner, A.; Brezesinski, G., Physical– Chemical Properties and Transfection Activity of Cationic Lipid/DNA Complexes. ChemPhysChem 2009, 10, (14), 2471-2479. 46. Falk, M.; Hartman, K. A.; Lord, R. C., Hydration of Deoxyribonucleic Acid. II. An Infrared Study. Journal of the American Chemical Society 1963, 85, (4), 387-391. 47. Keller, C. A.; Kasemo, B., Surface Specific Kinetics of Lipid Vesicle Adsorption Measured with a Quartz Crystal Microbalance. Biophysical Journal 1998, 75, (3), 1397-1402. 48. Seantier, B.; Breffa, C.; Félix, O.; Decher, G., Dissipation-Enhanced Quartz Crystal Microbalance Studies on the Experimental Parameters Controlling the Formation of Supported Lipid Bilayers. The Journal of Physical Chemistry B 2005, 109, (46), 21755-21765. 49. Chrystel, F.; Lida, B.; J, D. E., Determination of DMPC hydration in the Lα and Lβ′ phases by 2H solid state NMR of D2O. FEBS Letters 1997, 405, (3), 263-266.

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

50. Phillips, M. C.; Finer, E. G.; Hauser, H., Differences between conformations of lecithin and phosphatidylethanolamine polar groups and their effects on interactions of phospholipid bilayer membranes. Biochimica et Biophysica Acta (BBA) - Biomembranes 1972, 290, 397-402. 51. Crommelin, D. J. A., Influence of Lipid Composition and Ionic Strength on the Physical Stability of Liposomes. Journal of Pharmaceutical Sciences 1984, 73, (11), 1559-1563. 52. Alexandridis, P.; Olsson, U.; Lindman, B., A Record Nine Different Phases (Four Cubic, Two Hexagonal, and One Lamellar Lyotropic Liquid Crystalline and Two Micellar Solutions) in a Ternary Isothermal System of an Amphiphilic Block Copolymer and Selective Solvents (Water and Oil). Langmuir 1998, 14, (10), 2627-2638. 53. Lindblom, G.; Rilfors, L., Cubic phases and isotropic structures formed by membrane lipids — possible biological relevance. Biochimica et Biophysica Acta (BBA) - Reviews on Biomembranes 1989, 988, (2), 221-256. 54. Fontell, K., Cubic phases in surfactant and surfactant-like lipid systems. Colloid and Polymer Science 1990, 268, (3), 264-285. 55. Holmberg, K.; Jönsson, B.; Kronberg, B.; Lindman, B., Surfactants and Polymers in Aqueous Solution. 2 ed.; 2002. 56. Podgornik, R.; Rau, D. C.; Parsegian, V. A., The action of interhelical forces on the organization of DNA double helixes: fluctuation-enhanced decay of electrostatic double-layer and hydration forces. Macromolecules 1989, 22, (4), 1780-1786. 57. Ewert, K. K.; Evans, H. M.; Zidovska, A.; Bouxsein, N. F.; Ahmad, A.; Safinya, C. R., A Columnar Phase of Dendritic Lipid−Based Cationic Liposome−DNA Complexes for Gene Delivery:  Hexagonally Ordered Cylindrical Micelles Embedded in a DNA Honeycomb Lattice. Journal of the American Chemical Society 2006, 128, (12), 3998-4006. 58. Rädler, J. O.; Koltover, I.; Salditt, T.; Safinya, C. R., Structure of DNA-Cationic Liposome Complexes: DNA Intercalation in Multilamellar Membranes in Distinct Interhelical Packing Regimes. Science 1997, 275, (5301), 810-814. 59. Angelov, B.; Angelova, A.; Filippov, S. K.; Narayanan, T.; Drechsler, M.; Štěpánek, P.; Couvreur, P.; Lesieur, S., DNA/Fusogenic Lipid Nanocarrier Assembly: Millisecond Structural Dynamics. The Journal of Physical Chemistry Letters 2013, 4, (11), 1959-1964. 60. Martínez-Negro, M.; Kumar, K.; Barrán-Berdón, A. L.; Datta, S.; Kondaiah, P.; Junquera, E.; Bhattacharya, S.; Aicart, E., Efficient Cellular Knockdown Mediated by siRNA Nanovectors of Gemini Cationic Lipids Having Delocalizable Headgroups and Oligo-Oxyethylene Spacers. ACS Applied Materials & Interfaces 2016, 8, (34), 22113-22126. 61. Leal, C.; Bouxsein, N. F.; Ewert, K. K.; Safinya, C. R., Highly Efficient Gene Silencing Activity of siRNA Embedded in a Nanostructured Gyroid Cubic Lipid Matrix. Journal of the American Chemical Society 2010, 132, (47), 16841-16847. 62. Mindell, J. A., Lysosomal Acidification Mechanisms. Annual Review of Physiology 2012, 74, (1), 69-86. 63. Wang, T.; Upponi, J. R.; Torchilin, V. P., Design of multifunctional non-viral gene vectors to overcome physiological barriers: Dilemmas and strategies. International Journal of Pharmaceutics 2012, 427, (1), 3-20. 64. Dittrich, M., Brauer, C., Funari, S.S., Dobner, B., Brezesinski, G., Wölk, C., Interactions of Cationic Lipids with DNA: A Structural Approach. Langmuir 2018, 34, (49), 14858-14868. 65. Janich, C., Wölk, C., Taßler, S., Drescher, S., Meister, A., Brezesinski, G., Dobner, B., Langner, A., Composites of malonic acid diamides and phospholipids – Structural parameters for optimal transfection efficiency in A549 cells. European Journal of Lipid Science and Technology 2014, 116, (9), 1184-1194. 66. Vandeventer, P.E., Mejia, J., Nadim, A., Johal, M.S., Niemz, A., DNA Adsorption to and Elution from Silica Surfaces: Influence of Amino Acid Buffers. The journal of Physical Chemistry B 2013, 117, (37), 10742-10749.

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

TOC

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

294x173mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 36 of 36