throbber
THE JOURNAL OF GENE MEDICINE
`J Gene Med 2005; 7: 739–748.
`Published online 31 January 2005 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/jgm.717
`
`R E S E A R C H A R T I C L E
`
`New multivalent cationic lipids reveal bell curve for
`transfection efficiency versus membrane charge
`density: lipid–DNA complexes for gene delivery
`
`Ayesha Ahmad1
`Heather M. Evans1
`Kai Ewert1
`Cyril X. George2
`Charles E. Samuel2
`Cyrus R. Safinya1*
`
`1Departments of Materials, Physics,
`and Molecular, Cellular and
`Development Biology, University of
`California, Santa Barbara, Santa
`Barbara, CA 93106-5121, USA
`2Molecular, Cellular and
`Developmental Biology Department,
`Biomolecular Science and Engineering
`Program, University of California,
`Santa Barbara, Santa Barbara, CA
`93106, USA
`
`*Correspondence to:
`Cyrus R. Safinya, Materials Research
`Laboratory, UC Santa Barbara,
`Santa Barbara, CA 93106, USA.
`E-mail: safinya@mrl.ucsb.edu
`
`Received: 22 July 2004
`Revised: 9 September 2004
`Accepted: 20 September 2004
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`Abstract
`
`Background Gene carriers based on lipids or polymers – rather than on
`engineered viruses – constitute the latest technique for delivering genes into
`cells for gene therapy. Cationic liposome–DNA (CL-DNA) complexes have
`emerged as leading nonviral vectors in worldwide gene therapy clinical trials.
`To arrive at therapeutic dosages, however, their efficiency requires substantial
`further improvement.
`
`Methods Newly synthesized multivalent lipids (MVLs) enable control of
`headgroup charge and size. Complexes comprised of MVLs and DNA have
`been characterized by X-ray diffraction and ethidium bromide displacement
`assays. Their transfection efficiency (TE) in L-cells was measured with a
`luciferase assay.
`
`Results Plots of TE versus the membrane charge density (σM, average
`charge/unit area of membrane) for
`the MVLs and monovalent 2,3-
`dioleyloxypropyltrimethylammonium chloride (DOTAP) merge onto a
`universal, bell-shaped curve. This bell curve leads to the identification of three
`distinct regimes, related to interactions between complexes and cells: at low
`σM, TE increases with increasing σM; at intermediate σM, TE exhibits saturated
`behavior; and unexpectedly, at high σM, TE decreases with increasing σM.
`
`Conclusions Complexes with low σM remain trapped in the endosome.
`In the high σM regime, accessible for the first time with the new MVLs,
`complexes escape by overcoming a kinetic barrier to fusion with the
`endosomal membrane (activated fusion), yet they exhibit a reduced level
`of efficiency, presumably due to the inability of the DNA to dissociate from
`the highly charged membranes in the cytosol. The intermediate, optimal
`regime reflects a compromise between the opposing demands on σM for
`endosomal escape and dissociation in the cytosol. Copyright  2005 John
`Wiley & Sons, Ltd.
`
`Keywords gene therapy; cationic lipids; transfection efficiency; membrane
`charge density
`
`Introduction
`
`Cationic liposome–DNA (CL-DNA) complexes are attracting consider-
`able attention as gene vectors due to their safety and other inher-
`ent advantages over viral delivery methods [1,2]. These advantages
`
`Moderna Ex 1006-p. 1
`Moderna v Protiva
`
`

`

`740
`
`A. Ahmad et al.
`
`lack of
`include ease and variability of preparation,
`immunogenicity, and a capacity for DNA of unlimited size,
`allowing for delivery of human artificial chromosomes
`[3]. Recent setbacks in clinical trials with viral vectors, in
`particular a fatality induced by a severe inflammatory
`response [4] and insertional mutagenesis caused by
`retroviral vectors [5], have further promoted a diligent
`effort in developing efficient nonviral methods.
`Currently,
`lipofection is a prevalent nonviral gene
`transfer technology used in clinical
`trials worldwide
`[6]. CLs for transfection typically consist of a mixture
`of cationic and neutral (helper) lipid. Numerous lipids
`with varied chemical and physical properties have been
`synthesized [7–9] to improve the transfection efficiencies
`of CL-DNA complexes to the level of viral vectors. These
`include multivalent lipids, which have been described as
`superior to their monovalent counterparts [10,11].
`Despite this abundance of different cationic lipids,
`unifying themes and a comprehensive understanding
`of
`the interactions between CL-DNA complexes and
`mammalian cells are lacking. In order to rationally design
`and improve lipid-based delivery systems, however,
`such an understanding is essential. In particular, it is
`necessary to identify the interactions between the CL-DNA
`complexes and the cells along the transfection pathway
`to overcome the biological
`impediments to optimal
`transfection by directed alteration and optimization of
`CL-DNA complex formulations.
`In part, the lack of mechanistic understanding of gene
`delivery by CL-DNA complexes is due to the large num-
`ber of parameters involved. Few investigations to date
`include a complete examination of lipid performance as
`a function of lipid-bilayer composition and lipid/DNA
`charge ratio (ρchg). Even in comparative studies [12],
`typically only one or two data points per lipid are eval-
`uated, allowing the ideal lipid composition (the ratio of
`neutral to cationic lipid) or cationic lipid/DNA ratio to be
`overlooked [10,11].
`In previous experiments with commercially available
`lipids, Lin et al. identified the membrane charge density,
`σM, as a universal parameter for transfection by lamellar
`CL-DNA complexes, but the scope of these investigations
`was limited by the lipids used [13]. The membrane
`charge density is the average charge per unit area of
`the membrane. It is controlled by the ratio of neutral to
`cationic lipid in the liposome formulation. On the other
`hand, the lipid/DNA charge ratio, ρchg, is the number of
`charges on the cationic lipid divided by the number of
`charges on the DNA. In our experiments, we keep both
`the amount of DNA and ρchg (and thus the number of
`charges contributed by the cationic lipid) constant. Thus,
`σM is varied solely by changing the amount of neutral
`lipid per transfection assay, ‘‘diluting’’ the cationic lipid in
`the membrane.
`The work reported here presents both a confirmation
`and a significant extension of
`the earlier findings.
`We have synthesized a set of new multivalent lipids
`(MVLs) through methodical variation of headgroup size
`and charge and have examined the dependence of
`
`transfection efficiency (TE) on two key parameters,
`lipid composition and lipid/DNA charge ratio, ρchg. The
`MVLs [14], which form lamellar DNA complexes alone
`and when mixed with neutral 1,2-dioleoyl-sn-glycero-3-
`phosphatidylcholine (DOPC), enabled us to systematically
`probe very high membrane charge densities for the
`first time. For all DNA complexes of the MVLs as well
`as monovalent 2,3-dioleyloxypropyltrimethylammonium
`chloride (DOTAP), TE plotted versus σM fits the same, bell-
`shaped curve, confirming σM as a universal parameter.
`The curve shows three distinct regimes of TE and a
`clear maximum of TE at an optimal charge density,
`∗. Here, the TE rivals that of hexagonal CL-DNA
`σM
`∗ of the universal curve shifts
`complexes. The optimal σM
`systematically with ρchg, with an increase inρ chg resulting
`∗. Only the new, highly charged
`in higher values for σM
`lipids investigated here have permitted unambiguous
`identification of the universal maximum in TE and its shift
`with ρchg, as well as the discovery of a third regime of TE,
`where TE decreases (not saturates) with increasing σM.
`
`Materials and methods
`
`Materials
`
`The multivalent lipids (MVLs), MVL2 (molecular weight
`(MW) = 884.2 g/mol), MVL3
`(MW = 977.8 g/mol),
`MVL5 (MW = 1552.7 g/mol),
`and TMVL5 (MW =
`1253.0 g/mol) (Table 1), were synthesized according
`to the procedure previously described [14]. Nα,Nδ-
`Bis(Boc)-ornithine and Nδ-Boc-ornithine (Novabiochem)
`were used as the starting materials for MVL2 and
`MVL3, respectively. 2,2(cid:2)-(Ethylenedioxy)diethylamine for
`TMVL5 was purchased from Aldrich. For all lipids except
`MVL2, the final deprotection was performed by dissolving
`of the protected lipid in trifluoroacetic acid (Fisher),
`incubating at room temperature for 30 min and drying
`in vacuum,
`in deviation from the published protocol
`[14]. As described below, cationic liposomes were
`prepared containing these lipids as well as the cationic
`lipid 2,3-dioleyloxypropyltrimethylammonium chloride
`(DOTAP, MW = 698.55 g/mol), in combination with the
`lipids DOPC (MW = 786.13 g/mol) and 1,2-
`neutral
`dioleoyl-sn-glycero-3-phosphatidylethanolamine (DOPE,
`MW = 744.05 g/mol), all
`from Avanti Polar Lipids.
`CL-DNA complexes were formed from these cationic
`liposomes and the appropriate DNA. For X-ray samples,
`ethidium bromide (EtBr) experiments, and transfection
`assays, highly purified λ-phage DNA (New England
`Biolabs), highly polymerized calf thymus DNA (Amersham
`Life Sciences) and pGL3 plasmid DNA containing the
`luciferase gene (Promega Corp.) were used, respectively.
`
`Liposomes
`
`Lipid mixtures were prepared volumetrically by combin-
`ing chloroform/methanol (4 : 1) solutions of cationic and
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`J Gene Med 2005; 7: 739–748.
`
`Moderna Ex 1006-p. 2
`Moderna v Protiva
`
`

`

`Lipid–DNA Complexes for Gene Delivery
`
`741
`
`same manner as above, but, prior to transfer to cells, the
`complexes were mixed with 1.8 µl of a 6 mg/ml chloro-
`quine solution and incubated for 10 min. The normal
`transfection protocol was then resumed. To compensate
`for the variation in cell behavior over time, the data
`for DOTAP/DOPC complexes was normalized using data
`taken at the same time as TE data for MVL5.
`
`X-ray diffraction (XRD)
`
`CL-DNA complexes were prepared by mixing 75 µg
`of λ-phage DNA at 5 mg/ml with liposome solutions
`(20 mg/ml) in an Eppendorf centrifuge for approximately
`3 h. Samples were prepared at ρchg = 2.8. After storage
`for 3 days at 4 ◦C, allowing the samples to reach
`equilibrium, they were transferred to 1.5 mm diameter
`quartz X-ray capillaries (Hilgenberg, Germany). The
`high-resolution XRD experiments were carried out at
`the Stanford Synchrotron Radiation Laboratory. Two-
`dimensional powder diffraction images were obtained
`using an image plate detector (Mar Instruments).
`
`EtBr displacement assay
`
`Samples were prepared in a 96-well plate. Each well
`contained 2.4 µg of DNA, 0.28 µg of EtBr and the
`appropriate amount of cationic lipid. Water was added to
`each well, achieving a final volume of 200 µl per well.
`Fluorescence was measured on a Cary Eclipse fluorescence
`spectrophotometer.
`
`Results and discussion
`
`MVL design and structures
`
`Table 1 shows the chemical structures and maximum
`charges of the MVLs used in this study. These lipids
`were designed to achieve a systematic variation of the
`headgroup charge with minimal change in the chemical
`structure. Only a single aminopropyl unit per cationic
`charge was added to the headgroup, starting from the
`ornithine unit of MVL2. The oleyl chains provide strong
`anchoring in the membrane and miscibility with DOPC.
`TMVL5 has a slightly longer (triethylene glycol) spacer
`than the other MVLs.
`
`X-ray characterization of MVL-DNA
`complexes shows a lamellar phase
`
`We used XRD to determine the structure of MVL-
`DNA complexes. For all MVLs, at all
`investigated
`cationic/neutral
`lipid compositions of 0–90% DOPC,
`C) phase, the
`MVL-DNA complexes form the lamellar (Lα
`highly prevalent of the two known complex structures
`[15–17]. Moreover, XRD shows no evidence of phase
`separation, indicating that the complexes contain both
`
`Structure
`
`2 Cl-
`
`+
`NH3
`
`N
`H
`
`O
`
`O
`
`N
`H
`
`H2N+
`
`+
`NH3
`
`+
`NH3
`
`H
`N
`
`H
`N
`
`O
`
`O
`
`-
`3 Cl
`
`O O
`
`O O
`
`+
`NH3
`
`H
`N+
`
`+
`NH3
`
`+
`NH3
`
`+
`NH3
`
`O
`
`H2N+
`
`NH
`
`NH
`
`O
`
`-
`5 CF3CO2
`
`O O
`
`N+H
`
`+
`NH3
`
`+
`NH3
`
`+
`NH3
`
`O
`
`H2N+
`
`NH
`
`O
`
`2
`
`-
`
`5 Cl
`
`NH
`
`O
`
`OO
`
`Table 1.
`
`Name
`(max. chg.)
`
`MVL2
`(+2)
`
`MVL3
`(+3)
`
`MVL5
`(+5)
`
`TMVL5
`(+5)
`
`neutral lipids. The solvents were evaporated, first under a
`stream of nitrogen and subsequently in a high vacuum to
`ensure complete removal of the solvents. The dried lipid
`mixtures were hydrated at 37 ◦C for at least 6 h with the
`appropriate amount of deionized water of 18.2 M (final
`concentration of 20 mg/ml for X-ray samples; final con-
`centration of 0.5 mg/ml for transfection, and EtBr assay
`samples), sonicated to clarity with a VibraCell from Sonics
`and Materials Inc., and filtered through a 0.2 µm Teflon
`filter (Whatman). The obtained liposome solutions were
`stored at 4 ◦C.
`
`Cell Transfection
`
`Mouse fibroblast L-cells were cultured in Dulbecco’s mod-
`ified Eagle’s medium (DMEM, Gibco BRL) supplemented
`with 1% (v/v) penicillin-streptomycin (Gibco BRL) and
`5% (v/v) fetal bovine serum (Gibco BRL) at 37 ◦C in
`a humidified atmosphere with 5% CO2, reseeding the
`cells every 2–4 days to maintain subconfluency. The
`cells were transfected at 60–80% confluency in 24-well
`plates (7 mm diameter per well). Liposome (0.5 mg/ml)
`and DNA (1 mg/ml) stock solutions were diluted with
`DMEM to a final volume of 0.1 ml and complexes, con-
`taining 0.4 µg of pGL3-DNA per well, were prepared at
`the desired cationic-to-anionic charge ratio (ρchg). The
`cells were incubated with the complexes for 6 h, rinsed
`three times with phosphate-buffered saline (PBS, Gibco
`BRL), and incubated in supplemented DMEM for an
`additional 24 h (sufficient for a complete cell cycle) to
`allow expression of the luciferase gene. Luciferase gene
`expression was measured with the luciferase assay sys-
`tem from Promega Corp., and light output readings were
`performed on a Berthold AutoLumat luminometer. Trans-
`fection efficiency, measured as relative light units (RLU),
`was normalized to the weight of total cellular protein
`using the Bio-Rad protein assay dye reagent. For exper-
`iments with chloroquine, complexes were formed in the
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`J Gene Med 2005; 7: 739–748.
`
`Moderna Ex 1006-p. 3
`Moderna v Protiva
`
`

`

`742
`
`A. Ahmad et al.
`
`the lipid bilayers and corresponds to a DNA interhelical
`spacing dDNA = 2π/qDNA [17]. The experimental values
`are dDNA = 31 ˚A for MVL3 and dDNA = 27 ˚A for TMVL5.
`In Figures 1D and 1E, XRD patterns of complexes
`containing 100% MVL3 (D) and TMVL5 (E) under
`otherwise identical conditions are shown. The narrow
`peaks and multiple harmonics show that at the salt
`concentrations present in DMEM, i.e. under conditions
`as in the transfection experiments, stable and well-
`defined complexes form even from membranes containing
`exclusively the highly charged multivalent lipids. The
`salt present in DMEM screens the electrostatic repulsions
`between the headgroups and enables formation of stable
`complexes. However, when prepared in deionized water,
`complexes without neutral lipids exhibit broadening of
`the first lamellar peak (at q001) and do not show higher
`harmonics of the first lamellar peak, indicative of smaller
`size multilamellar assemblies (results not shown).
`
`Estimation of lipid headgroup charge:
`EtBr displacement assay
`
`Figure 2 shows data from an EtBr displacement assay
`[18–21], performed to examine the ability of the MVLs to
`condense DNA within the CL-DNA complexes. The data
`was acquired by collecting fluorescence measurements
`at various weight ratios of MVL to DNA (with a fixed
`weight of DNA and EtBr per point) and normalizing the
`intensity to the fluorescence of DNA and EtBr in solution.
`EtBr fluoresces when intercalated between the base pairs
`of DNA, but self-quenches in solution. As the MVL
`C
`liposomes are introduced and self-assembly into the Lα
`phase occurs, EtBr is displaced and overall fluorescence
`decreases until all DNA has been incorporated into the
`MVL-DNA complexes at the isoelectric point. Thus, this
`method allows a quick and efficient assessment of the
`effective charge on the headgroup of these lipids. The
`
`Figure 2. Normalized fluorescence data from the EtBr displace-
`ment assay for MVL2, MVL3 and MVL5. The dashed lines
`are drawn at the isoelectric points, determined as described
`in the text. They result in valencies of ZMVL2 = 2.0 ± 0.1,
`ZMVL3 = 2.5 ± 0.1, ZMVL5 = ZTMVL5 = 4.5 ± 0.1
`
`Figure 1. (A) Schematic of the lamellar phase indicating the
`characteristic dimensions. Reprinted with permission from [15].
`(B, C) Typical X-ray diffraction (XRD) scans from lamellar
`C) CL-DNA complexes, containing 40 mol% MVL ((B) MVL3;
`(Lα
`(C) TMVL5) and 60 mol% DOPC, at a lipid/DNA charge ratio
`ρchg = 2.8 in the presence of DMEM. (D, E) Typical XRD scans
`C CL-DNA complexes containing 100 mol% MVL ((D)
`from Lα
`MVL3; (E) TMVL5), prepared at ρchg = 2.8 in the presence of
`DMEM (Reproduced in part with permission from reference 15)
`
`MVL and DOPC, as intended. A schematic of the lamellar
`C) and its characteristic dimensions is shown
`phase (Lα
`in Figure 1A. Figures 1B and 1C shows typical XRD
`patterns of complexes containing 40 mol% MVL (MVL3
`(B); TMVL5 (C)) and 60 mol% DOPC, at a lipid/DNA
`charge ratio of 2.8, prepared in the presence of DMEM.
`The sharp peaks, labelled q001, q002, q003, respectively,
`give the lamellar repeat distance, d, which is the sum
`of the membrane thickness (δm) and the thickness of
`a water/DNA layer (δw): d = δm + δw = 2π/q001. The
`labeled qDNA, results from one-
`diffuse weaker peak,
`dimensional ordering of the DNA sandwiched between
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`J Gene Med 2005; 7: 739–748.
`
`Moderna Ex 1006-p. 4
`Moderna v Protiva
`
`

`

`Lipid–DNA Complexes for Gene Delivery
`
`743
`
`isoelectric point was determined as the intersection
`point of
`linear fits to the data at high and low
`lipid/DNA ratio [22,23]. This gives headgroup charges of
`ZMVL2 = 2.0 ± 0.1, ZMVL3 = 2.5 ± 0.1, ZMVL5 = 4.5 ± 0.1.
`The dashed lines in Figure 2 indicate the corresponding
`isoelectric MVL/DNA weight ratios.
`
`Transfection efficiency as a function of
`lipid composition
`
`Figure 3 shows TE results for complexes transfecting
`mouse fibroblast cells at various MVL/DOPC ratios.
`Also included is data for the monovalent lipid DOTAP
`mixed with DOPC, a well-investigated reference system
`which constituted the starting point of our studies [13].
`The complexes were prepared at ρchg = 2.8, which Lin
`et al. have found to be the optimum charge ratio for
`DOTAP/DOPC complexes [24]. The amount of DNA and
`cationic lipid per sample was kept constant. Thus, only the
`amount of neutral lipid varies between data points. All
`MVL-DNA complexes form globular particles of around
`0.2 µm diameter in water, as previously reported for
`DOTAP [25] and MVL5 [14]. In DMEM, these particles
`form much larger aggregates due to screening of their
`electrostatic repulsion by the high salt concentration, as
`we have shown for MVL5 by optical and epi-fluorescence
`microscopy [14]. Figure 3A shows the TE data as a
`function of the molar fraction of cationic lipid. For all
`cationic lipids, a maximum in TE as a function of lipid
`composition is observed: at 65 mol% for MVL2, 70 mol%
`for MVL3, 50 mol% for MVL5, 55 mol% for TMVL5, and
`90 mol% for DOTAP. The optimal molar ratio results in a
`TE that is over two decades higher than that of the lowest
`transfecting complexes in these systems, and each data
`set fits a skewed bell-shaped curve.
`
`Membrane charge density is a
`universal parameter: three regimes of
`transfection efficiency
`
`Figure 3B shows the data from Figure 3A plotted versus
`the membrane charge density, σM. A notable simplification
`occurs and all the data points merge onto a single curve.
`This identifies σM as a universal parameter for transfection
`by lamellar CL-DNA complexes.
`the average charge
`As mentioned above, σM is
`per unit area of the lipid membrane; therefore, the
`headgroup areas of
`the lipids,
`their charge, and
`the molar
`fractions of cationic and neutral
`lipid
`required to calculate σM. We
`are the parameters
`calculated σM as described by Lin et al. [13], with σM =
`total charge/total membrane area = eZNcl/(Ncl Acl +
`Nnl Anl) = [1 − nl/( nl + r cl)]σcl, where Ncl and Nnl
`are the number of cationic lipids and neutral lipids in
`the complexes, respectively; r = Acl/Anl is the ratio of the
`headgroup areas of the cationic and the neutral lipid;
`σcl = eZ/Acl is the charge density of the cationic lipid
`
`with valence Z; and nl and cl are the molar fractions of
`the neutral and cationic lipids, respectively. For our data,
`we used Anl = 72 ˚A2 [26,27], rDOTAP = 1, rMVL2 = 1.05 ±
`0.05, rMVL3 = 1.30 ± 0.05, rMVL5 = 2.3 ± 0.1, rTMVL5 =
`2.5 ± 0.1, ZDOTAP = 1, ZMVL2 = 2.0 ± 0.1, ZMVL3 = 2.5 ±
`0.1, ZMVL5 = ZTMVL5 = 4.5 ± 0.1. The values for Z were
`obtained by the EtBr displacement assay as described
`above. The values for r can be considered as fitting
`parameters, but they yield physically reasonable values
`that agree with chemical intuition. (The values of r were
`determined based on agreement with the Gaussian fit.
`For the monovalent lipid DOTAP, r was assumed to be
`equal to 1, consistent with previous findings [13].) It
`is interesting to note that the optimal TE for MVL3 is
`found at a larger molar fraction of cationic lipid than for
`MVL2 despite the fact that MVL3 has a higher headgroup
`charge. This can be attributed to the significantly larger
`headgroup size of MVL3 and shows the importance of the
`parameter r.
`The resulting curve of TE vs. σM can be described
`empirically by a simple Gaussian (solid line in Figure 3B):
`TE = TE0 + A exp−[(σM − σM
`∗
`)/w]2
`(1)
`where TE0 = −(1.9 ± 5.6) × 107 RLU/mg protein; A =
`(9.4 ± 0.6) × 108 RLU/mg protein; w = (5.8 ± 0.5) ×
`10−3 e/˚A2. For the optimal charge density, the fit gives
`∗ = (17.4 ± 0.2) × 10−3 e/˚A2.
`σM
`Remarkably, in extension of previous results which
`showed the increase and a subsequent levelling-off of TE
`with σM [13], we see an entire bell curve of efficiency,
`including a decrease in TE at higher charge densities.
`Previously, without the series of new MVLs, Lin et al.
`were not adequately prepared to measure this range of
`charge densities.
`The new universal TE curve of lamellar complexes
`exhibits three well-defined regimes. Regime I (dark
`gray), corresponding to low σM, features an exponential
`increase in efficiency over three orders of magnitude.
`Regime III (light gray), corresponding to high σM, is
`characterized by a decrease in efficiency with increasing
`σM, suggesting that there also is an obstacle of electrostatic
`nature to successful DNA delivery by lamellar CL-DNA
`complexes. The competition of the two effects that give
`rise to regimes I and III leads to the existence of the
`intermediate regime II (white) as the region of optimal
`charge density, corresponding to the highest TE. This
`clearly demonstrates the importance of including neutral
`lipid in the formulation of CL-DNA complexes, particularly
`those of newer, multivalent lipids. Due to the universality
`of the curve, it should also be possible to estimate the
`optimal composition for any given lipid by performing
`the simple EtBr displacement assay and estimating the
`headgroup size of the lipid from its chemical structure.
`We will address the implications of the universal curve
`for the mechanism of transfection in more detail below.
`In addition to data for
`the lamellar complexes,
`Figure 3B shows TE data for
`the commonly used
`C)
`DOTAP/DOPE lipid system. The inverted hexagonal (HII
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`J Gene Med 2005; 7: 739–748.
`
`Moderna Ex 1006-p. 5
`Moderna v Protiva
`
`

`

`744
`
`A. Ahmad et al.
`
`Figure 3. (A) TE in RLU per mg total cellular protein plotted as a function of mol% DOPC for DNA complexes prepared with MVL2
`(green diamonds), MVL3 (red squares), MVL5 (blue triangles) and TMVL5 (purple inverted triangles) as well as DOTAP (gray
`circles). All data was taken at ρchg = 2.8, using the same amount of DNA for each data point. (B) The same TE data plotted against
`the membrane charge density, σM. Also included are data for DOTAP/DOPE complexes (gray open circles) and a Gaussian fit to the
`DOPC systems. The three regimes of transfection efficiency are indicated by different shading in the plot
`
`Figure 4. TE data taken in the presence of chloroquine (chlq) to
`assess the relevance of endosomal escape in the three regimes
`(regime I, dark gray; regime II, white; regime III, light gray).
`The relative increase in TE, TEchlq/TE-1, is plotted for MVL3
`(black bars) and MVL5 (white bars) in the three regimes. TE is
`enhanced by about a factor of three in regime I (10 mol% MVL),
`but not in regime II (60 mol% MVL) and regime III (100% MVL).
`This indicates that endosomal release is not a limiting factor for
`regimes II and III
`
`[15] DOTAP/DOPE complexes exhibit high TE even at
`low σM, due to their distinct mechanism of cellular entry,
`which relies on the fusogenic properties of DOPE [2,13].
`This has made DOPE a popular choice as a co-lipid, and
`complexes with DOPE are abundant in the literature.
`However, the vast majority of neutral
`lipids lead to
`lamellar CL-DNA complexes, and, even with PE-based
`lipids, the hexagonal phase is only observed in a small
`window of composition [15]. Furthermore, the in vivo
`performance of PE-based complexes is disappointingly
`poor and cholesterol, which leads to lamellar complexes, is
`increasingly used as a neutral lipid for in vivo applications
`[2,7]. Therefore, our result that the TE of optimized
`
`Figure 5. (A) Three-dimensional transfection phase diagram of
`lamellar complexes, combining data for MVL2, MVL3, and MVL5.
`Membrane charge density, σM, lipid/DNA charge ratio, ρchg, and
`TE are plotted along the x-axis, y-axis, and z-axis, respectively.
`∗ (from the data shown in (A)) is plotted
`(B) The optimal σM
`∗)
`against the charge ratio, ρchg. (C) The maximum TE (at σM
`plotted against ρchg. This TE essentially remains constant
`
`C
`C complexes rivals that of the highly transfecting HII
`Lα
`DOTAP/DOPE complexes is of great significance. It adds
`a compositional degree of freedom, because efficiently
`transfecting complexes can be prepared from a broad
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`J Gene Med 2005; 7: 739–748.
`
`Moderna Ex 1006-p. 6
`Moderna v Protiva
`
`

`

`Lipid–DNA Complexes for Gene Delivery
`
`745
`
`range of lipids without diminishing TE if σM is optimized.
`This includes lipids with specialized functions in the
`delivery process, such as peptide-lipids.
`
`Effect of chloroquine on transfection
`efficiency in the three regimes
`
`Endocytosis is the dominant mechanism of entry of
`lamellar CL-DNA complexes, as evident
`from recent
`confocal microscopy and transfection data [13] and work
`from other laboratories [28,29]. After cellular uptake
`via endocytosis, CL-DNA complexes must escape from
`endosomes in order for the DNA to progress toward the
`cell nucleus. Only a finite amount of time is available for
`this, since the endosomal pathway involves degradation of
`the contents of the endosome: initially through a lowering
`of the pH within the endosome and then through fusion
`with low-pH lysosomes.
`The importance of endosomal escape as a barrier to
`transfection can be assessed by performing transfection
`experiments in the presence of chloroquine (chlq),
`a well-established bio-assay known to enhance the
`release of material
`from endosomes by osmotically
`bursting the vesicle [30]. We have repeated efficiency
`experiments for MVL3 and MVL5 in the presence of
`chloroquine for data points characteristic of the three
`transfection regimes. The resulting data is shown in
`Figure 4, plotted as the relative increase in transfection
`efficiency, (TEchlq-TE)/TE = TEchlq/TE − 1. TE in regimes
`II (60 mol% MVL) and III (100% MVL) is not enhanced
`by the addition of chloroquine. However, an increase
`in TE is seen in regime I, at low σM (10 mol% MVL),
`where transfection efficiency is enhanced approximately
`threefold. These results suggest that complexes with lower
`σM remain trapped within endosomes, while complexes
`with higher σM are able to overcome this barrier.
`Therefore, the decrease in TE at highest σM, in regime
`III, is not related to endosomal entrapment but must be
`due to other effects.
`In regime I, a straight line fits the data of Figure 3B
`well, particularly for small values of σM. This regime was
`previously investigated by Lin et al. [13], who proposed
`that endosomal escape limits TE in this regime, consistent
`with the enhancement of TE by chloroquine. The
`escape from the endosome likely occurs via an activated
`fusion process of the oppositely charged membranes of
`endosome and complex [13]. The activation energy for
`this can be written as δE = aκ − bσM, where a and b are
`constants >0. The parameter κ is the bending rigidity of
`the membrane, which is mainly determined by the lipid
`tails and the area per lipid chain and therefore constant in
`our experiments. Bending of membranes, as required for
`fusion, results in an energy cost proportional to κ. Since
`the interacting membranes are oppositely charged, the
`activation energy decreases with increasing σM, making
`fusion with the endosomal membrane more likely. For
`endosomal entrapment being the main impediment to
`transfection as proposed by Lin et al., the activation
`
`energy for fusion directly relates σM to the transfection
`efficiency via an Arrhenius-type equation [13]:
`TE ∝ rate of fusion = 1/τ e
`−δE/kT
`
`(2)
`
`Here, 1/τ is the collision rate between the trapped CL-
`DNA particle and the endosomal membrane.
`At higher charge densities, in regimes II and III, TE
`is no longer limited by endosomal escape, as shown
`by the negligible effect of chloroquine. Here, the data
`of Figure 3B bends strongly away from a straight line
`but is well approximated by the empirical bell curve of
`Equation (1). Except for the lowest values of σM, this bell
`curve also fits the data in regime I, implying that the
`activation energy δE most likely contains contributions
`both linear and quadratic in σM.
`
`Effect of lipid/DNA charge ratio:
`transfection phase diagram
`
`All data shown in Figure 3 were taken at a fixed lipid/DNA
`charge ratio, ρchg, of 2.8. For other values of ρchg, we have
`observed similar, universal behavior. An efficiency phase
`diagram, shown in Figure 5A, summarizes these results.
`In this graph, σM varies along the x-axis, ρchg along the
`y-axis, and TE along the z-axis. At very low lipid/DNA
`charge ratios (i.e. below and near the isoelectric point,
`ρchg ≤ 1, where complexes are negatively charged or
`neutral), TE is low for all values of σM. This is to
`be expected, as an overall positive CL-DNA charge is
`required to promote initial electrostatic interactions with
`cell membranes [31–33]. As ρchg is increased to above
`unity, a maximum in TE, defining the optimal membrane
`∗, emerges and a bell curve of efficiency
`charge density σM
`∗ shifting to higher values
`is observed, with the optimal σM
`with increasing ρchg. This means that much more cationic
`lipid is required to achieve optimal TE at large lipid/DNA
`charge ratios. For clarity, this surprising trend is shown
`∗ against ρchg.
`in Figure 5B, which plots the optimal σM
`The descending part of the bell curve beyond the optimal
`∗ for ρchg >1 (i.e. regime III) cannot be seen for
`σM
`the highest lipid/DNA charge ratios. However, it seems
`likely that, given liposomes with an even higher charge
`density than the 27.17 × 10−3 e/˚A2 attainable with 100%
`MVL5, the decrease in the efficiency predicted by the bell
`curve shape would also be seen for the highest ρchg. As
`shown in Figure 5C, the maximum TE (TE at the optimal
`∗) does not change appreciably with ρchg. A relatively
`σM
`low lipid/DNA charge ratio, therefore, can be considered
`optimal since it allows for achievement of maximum TE
`with the least amount of cationic lipid. This is due to
`∗ with ρchg. Minimizing the
`the unexpected increase of σM
`amount of cationic lipid is desirable to reduce cost as well
`as potential toxic effects of the cationic lipid. In addition,
`achieving a given σM with fewer, more highly charged
`molecules should mean a smaller metabolic effort for the
`elimination of the lipids from the cell. This reasoning
`would favor multivalent over monovalent lipids. In this
`
`Copyright  2005 John Wiley & Sons, Ltd.
`
`J Gene Med 2005; 7: 739–748.
`
`Moderna Ex 1006-p. 7
`Moderna v Protiva
`
`

`

`746
`
`A. Ahmad et al.
`
`context, it is important to note that with the amounts
`of cationic lipid employed in our in vitro experiments,
`we find no toxic effects on the cells as judged by cell
`morphology and the amount of total cellular protein.
`
`A model of the intracellular CL-DNA
`complex pathway as a function of σM
`
`The schematic shown in Figure 6 summarizes our current
`C complexes in the
`understanding of the cellular fate of Lα
`three regimes. Initial attachment mediated by electrostatic
`attractions between CL-DNA complexes and negative
`charges at the cell surface (Figure 6a) is followed by
`endocytosis of the complex particle (Figure 6b), resulting
`in endosomal entrapment (Figure 6c) [13,31,34].
`If
`cellular attachment and uptake were limiting TE via a
`σM-dependent mechanism, a linear increase of TE with
`σM would be predicted. Thus, our data, which shows an
`exponential increase, excludes this possibility.
`∗ (regime I), trans-
`For complexes with low σM < σM
`fection is limit

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket