throbber
Biophysical Journal Volume 84 May 2003 3307–3316
`
`3307
`
`Three-Dimensional Imaging of Lipid Gene-Carriers: Membrane
`Charge Density Controls Universal Transfection Behavior
`in Lamellar Cationic Liposome-DNA Complexes
`
`Alison J. Lin,* Nelle L. Slack,* Ayesha Ahmad,* Cyril X. George,
`and Cyrus R. Safinya*
`*Materials Department, Physics Department, and Biomolecular Science and Engineering Program, University of California,
`y
`Molecular, Cellular, and Developmental Biology Department, and Biomolecular Science
`Santa Barbara, California 93106; and
`and Engineering Program, University of California, Santa Barbara, California 93106
`
`y
`
`Charles E. Samuel,
`
`y
`
`ABSTRACT Cationic liposomes (CLs) are used worldwide as gene vectors (carriers) in nonviral clinical applications of gene
`delivery, albeit with unacceptably low transfection efficiencies (TE). We present three-dimensional laser scanning confocal
`
`C and inverted hexagonal HIIC
`microscopy studies revealing distinct interactions between CL-DNA complexes, for both lamellar La
`C complexes in cells identified two regimes. For low membrane
`nanostructures, and mouse fibroblast cells. Confocal images of La
`charge density (sM), DNA remained trapped in CL-vectors. By contrast, for high sM, released DNA was observed in the
`cytoplasm, indicative of escape from endosomes through fusion. Remarkably, firefly luciferase reporter gene studies in the
`C-mammalian cell system revealed an unexpected simplicity where, at a constant cationic to anionic charge
`highly complex La
`ratio, TE data for univalent and multivalent cationic lipids merged into a single curve as a function of sM, identifying it as a key
`C complexes climbs exponentially over  four decades with
`universal parameter. The universal curve for transfection by La
`increasing sM below an optimal charge density (sM* ), and saturates for sM [s
`M at a value rivaling the high transfection efficiency of
`
`
`HIIC complexes. In contrast, the transfection efficiency of HIIC complexes is independent of sM. The exponential dependence of TE
`
`on sM for LaC complexes, suggests the existence of a kinetic barrier against endosomal fusion, where an increase in sM lowers the
`C complexes and HIIC, confocal microscopy reveals the dissociation of lipid and DNA.
`
`barrier. In the saturated TE regime, for both La
`However, the lipid-released DNA is observed to be in a condensed state, most likely with oppositely charged macro-ion
`condensing agents from the cytoplasm, which remain to be identified. Much of the observed bulk of condensed DNA may be
`transcriptionally inactive and may determine the current limiting factor to transfection by cationic lipid gene vectors.
`
`INTRODUCTION
`
`The unrelenting research activity involving gene therapy with
`either synthetic vectors (carriers) or engineered viruses is
`currently unprecedented (Alper, 2002; Chesnoy and Huang,
`2000; Clark and Hersh, 1999; Ferber, 2001; Henry, 2001;
`Mahato and Kim, 2002; Miller, 1998). After the initial
`landmark studies (Felgner et al., 1987; Nabel et al., 1993;
`Singhal and Huang, 1994), cationic liposomes (CLs; closed
`bilayer membrane shells of lipid molecules) have emerged
`worldwide as the most prevalent synthetic vectors (carriers)
`(Ferber, 2001) whose mechanisms of action are investigated
`extensively in research laboratories, concurrently with on-
`going, mostly empirical, clinical trials to develop cancer vac-
`cines (Alper, 2002; Chesnoy and Huang, 2000; Clark and
`Hersh, 1999; Ferber, 2001; Henry, 2001; Mahato and Kim,
`2002; Miller, 1998). Primary among the advantages of CL
`over viral methods is the lack of immune response due to the
`absence of viral peptides and proteins. Moreover, while viral
`capsids have a maximum DNA-carrying capacity of ;40 kbp,
`CLs (which, when combined with DNA, form, in most
`
`Submitted October 3, 2002, and accepted for publication December 18,
`2002.
`
`Alison J. Lin and Nelle L. Slack contributed equally to this work.
`
`Address reprint requests to C. R. Safinya, MRL Rm. 2208, University of
`California, Santa Barbara, CA 93106. Tel.: 805-893-8635; Fax: 805-893-
`7221; E-mail: safinya@mrl.ucsb.edu.
`Ó 2003 by the Biophysical Society
`0006-3495/03/05/3307/10 $2.00
`
`lamellar LC
`a or
`instances, self-assemblies with distinct
`inverted hexagonal HC
`II nanostructures; see Raedler et al.,
`1997; Koltover et al., 1998; Lasic et al., 1997), place no limit
`on the size of the DNA. Thus, if complexed, for example,
`with human artificial chromosomes (Wilard, 2000), optimally
`designed CL-carriers offer the potential of potent vectors
`comprised of multiple human genes and regulatory sequences
`extending over hundreds of thousands of DNA base pairs.
`Despite all the promise of CLs as gene vectors, their
`transfection efficiency (TE; ability to transfer DNA into cells
`followed by expression), compared to viral vectors, remains
`notoriously low, resulting in a flurry of research activity
`aimed at enhancing transfection (Alper, 2002; Chesnoy and
`Huang, 2000; Clark and Hersh, 1999; Ferber, 2001; Henry,
`2001; Mahato and Kim, 2002; Miller, 1998). A further sense
`of urgency for developing efficient synthetic carriers stems
`from the recent tragic events associated with the use of
`engineered adenovirus vectors leading to the death of a patient
`due to an unanticipated severe immune response (Marshall,
`2000). In addition, in the latest gene therapy trials using
`modified retrovirus vectors to treat children with severe
`combined immunodeficiency, a French gene therapy team
`has reported a major setback where one patient (out of eleven)
`developed a blood disorder similar to leukemia which is
`confirmed to have resulted from insertion of the modified
`retrovirus in the initial coding region of a gene related to the
`early development of blood cells (Marshall, 2002).
`
`Moderna Ex 1006-p. 1
`Moderna v Arbutus
`
`

`

`3308
`
`Lin et al.
`
`Our in vitro studies should apply to TE optimization in ex
`vivo cell transfection, where cells are removed and returned
`to patients after transfection. In particular, our studies, aimed
`at understanding the chemically and physically dependent
`mechanisms underlying TE in continuous (dividing) cell
`lines, should aid clinical efforts to develop efficient CL-vector
`cancer vaccines in ex vivo applications. The vaccines are
`intended to induce transient expression of genes coding for
`immunostimulatory proteins in dividing cells (Chesnoy and
`Huang, 2000; Nabel et al., 1993; Rinehart et al., 1997;
`Stopeck et al., 1998); thus, the nuclear membrane, which dis-
`solves during mitosis, is not considered a barrier to expression
`of DNA.
`A critical requirement for enhancing transfection via
`synthetic carriers is a full understanding of the different
`nanostructures of CL-DNA complexes and the physical
`and chemical basis of interactions between complexes and
`cellular components. Toward that end, we used a combination
`of three-dimensional laser scanning confocal microscopy
`(LSCM), which closely followed complexes across the
`plasma membrane and inside the cytoplasm, and reporter
`gene TE studies that gave a statistically meaningful measure
`of the total amount of protein synthesized by cells from deli-
`vered DNA. Furthermore, we examined the structure-de-
`pendent basis of transfection through imaging of CL-DNA
`complexes exhibiting either the LC
`a (cationic lipid DOTAP
`mixed with neutral lipid DOPC) or HC
`II (DOTAP mixed with
`neutral lipid DOPE) nanostructure (Koltover et al., 1998;
`Raedler et al., 1997). Previous to our report, TE studies had
`shown that in mixtures of DOTAP and neutral lipids, typically
`at a wt.:wt. ratio of between 1:1 and 1:3, DOPE aided, while
`DOPC severely suppressed,
`transfection (Farhood et al.,
`1995; Hui et al., 1996), hence suggesting that HC
`II complexes
`transfect more efficiently than LC
`a complexes.
`
`MATERIALS AND METHODS
`
`Materials
`
`Lipids included univalent cationic lipids DOTAP (1,2-dioleoyl-3-trimethyl-
`ammonium-propane) and DMRIE (n-(2-hydroxyethyl)-n,n-dimethyl-2,3-
`bis(tetradecyloxy)-1-propanaminium), multivalent cationic lipid DOSPA
`(2,3-Dioleyloxy-n-(2-(sperminecarboxamido)ethyl)-n,n-dimethyl-1-propan-
`iminium penta-hydrochloride), and neutral lipids DOPC (1,2-dioleoyl-sn-
`glycero-3-phosphocholine) and DOPE (1,2-dioleoyl-sn-glycero-3-phos-
`phoethanolamine). DOTAP, DOPC, DOPE were purchased from Avanti
`Polar Lipids, Inc., and DMRIE and DOSPA were gifts from Vical Inc. Plas-
`mid DNA containing the Luciferase gene and SV40 promoter/enhancer ele-
`ments was used (pGL3-control vector, Promega, Cat. E1741).
`
`Liposome preparation
`
`Neutral lipids (DOPE, DOPC) were dissolved in chloroform and cationic
`lipids (DOTAP, DOSPA, DMRIE) were dissolved in a chloroform/methanol
`mixture. The lipid solutions were mixed in required ratios and the solvent
`was evaporated, first under a stream of nitrogen and then in vacuum over
`night, leaving a lipid film behind. The appropriate amount of millipore water
`was added to the dried lipid film, resulting in the desired concentration (0.1
`mg/ml–25 mg/ml) and incubated at 378C for at least 6 h to allow formation
`of liposomes. To form small unilamellar vesicles, liposome solutions were
`vortexed for 1 min, tip-sonicated to clarity (for 5–10 min), and filtered with
`0.2 mm filters to remove metal particulates arising from the sonicator tip.
`Liposome solutions were then stored at 48C.
`
`Transfection
`
`L-cells were transfected at a confluency of 60–80%. Using liposome (0.5
`mg/ml) and DNA (1 mg/ml) stock solutions, CL-DNA complexes were
`prepared in DMEM, which contained 2 mg of pGL3-DNA at a cationic-to-
`anionic charge ratio of 2.8 and allowed to sit for 20 min for complex
`formation. The cells were then incubated with complexes for 6 h, the
`optimized time of transfer into cells before removal, rinsed three times with
`phosphate-buffered saline (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 System from Promega. Each transfection ex-
`periment was repeated between 3–6 times over a short period of a few days
`yielding the error bars (Slack, 2000; Lin et al., 2002). In addition, the
`experiments were repeated four times over a 12-month period using different
`cell batches. While the absolute value of the average of each transfection
`measurement between the different experiments (done with several months
`separating experiments) varied by as much as a decade (which is also
`commonly found by other groups; Boussif et al., 1995), the observed trend
`in the transfection data was completely reproducible. Transfection efficiency
`was normalized to mg of total cellular protein using the Bio-Rad Protein
`Assay Dye Reagent Concentration solution (Bio-Rad) and is expressed as
`relative light units per mg of total cellular protein 6 1 SD. The transfection
`protocol is commonly used by others (Boussif et al., 1995).
`
`Laser scanning confocal microscopy (LSCM)
`
`L-cells were seeded on coverslips in six-well plates and allowed to grow,
`reaching a confluency of 60–80%. DNA was labeled following the Mirus
`Label IT (PanVera Corporation) protocol, which is fluorescent at 492 nm.
`Lipids were labeled with 0.2% (wt) Texas Red DHPE (Molecular Probes),
`which is fluorescent at 583 nm. Using labeled liposome (0.5 mg/ml) and
`DNA (0.1 mg/ml) stock solutions, CL-DNA complexes were prepared in
`DMEM using 2 mg of pGL3-DNA (at a cationic to anionic charge ratio of
`2.8), allowed to sit for 20 min for complex formation, and incubated with
`cells for the optimal 6-h transfer time. Cells were rinsed three times with
`phosphate-buffered saline (Gibco BRL), fixed by soaking in a fixing solution
`(3.7% formaldehyde in PBS) for 20–30 min and mounted using SlowFade
`Light Antifade Kit (Molecular Probes) for microscopy. Confocal images
`were taken with a Leica DM IRBE confocal microscope. The images of Figs.
`3 and 5 (repeated more than 10 times) are representative of the typical
`behavior in a given field of view (Lin, 2001).
`
`Cell culture
`
`X-ray diffraction (XRD)
`
`Mouse fibroblast L-cell lines were subcultured in DMEM (Dulbecco’s
`modified Eagle’s medium, Gibco BRL) supplemented with 1% (vol/vol)
`penicillin-streptomycin (Gibco BRL) and 5% (vol/vol) fetal bovine serum
`(HyClone Lab) at 378C and 5% CO2 atmosphere every 2–4 days to maintain
`monolayer coverage.
`
`the Stanford Synchrotron
`The XRD experiments were carried out at
`Radiation Laboratory at 10 KeV. To prepare CL-DNA samples liposome
`solutions (25 mg/ml) and DNA solutions (5 mg/ml) were each diluted in
`DMEM at 1:1 (vol.:vol.), then mixed at the desired cationic to anionic
`charge ratio of 2.8 and centrifuged before loading into for 1.5-mm x-ray. We
`
`Biophysical Journal 84(5) 3307–3316
`
`Moderna Ex 1006-p. 2
`Moderna v Arbutus
`
`

`

`Imaging and Efficacy of Lipid-DNA Complex
`
`3309
`
`note that similar to previous findings (Koltover et al., 1998; Raedler et al.,
`1997) the self-assembled structures of CL-DNA complexes does not change
`in the concentration range between the x-ray samples and the samples for
`confocal microscopy and transfection.
`
`RESULTS AND DISCUSSION
`
`The initial electrostatic attraction between positively charged
`CL-DNA complexes and mammalian cells is known to be
`mediated by negatively charged cell
`surface sulfated
`proteoglycans (Fig. 1 a, expanded view; also see Mislick
`and Baldeschwieler, 1996). Consistent with previous studies,
`we found that LC
`a complexes gain entry into the cell through
`endocytosis (Fig. 1, b and c; see also Labat-Moleur et al.,
`1996; Zabner et al., 1995). LSCM of LC
`a CL-DNA particles
`in cells revealed two distinct types of behavior. At low
`membrane charge density, sM  0.005 e/A˚ 2 ¼ e/(200 A˚ 2),
`mostly intact LC
`inside cells
`a complexes were present
`implying that DNA was trapped by the lipid vector. Further
`TE experiments confirmed that the intact complexes were
`themselves trapped in endosomes (Fig. 1 c). At high sM 
`
`FIGURE 1 Model of cellular uptake of LC
`a complexes. (a) Cationic
`complexes adhere to cells due to electrostatic attraction between positively
`charged CL-DNA complexes and negatively charged cell-surface sulfated
`proteoglycans (shown in expanded views) of mammalian plasma mem-
`branes. (b and c) After attachment, complexes enter through endocytosis. (d )
`Only those complexes with a large enough membrane charge density (sM)
`escape the endosome through activated fusion with endosomal membranes.
`(e) Released DNA inside the cell is observed by comfocal microscopy to be
`present primarily in the form of aggregates. The DNA aggregates must reside
`in the cytoplasm because oppositely charged cellular biomolecules able to
`condense DNA are not present in the endosome. Arrows in the expanded
`view of c indicate the electrostatic attraction between the positively charged
`membranes of the complex and the negatively charged membranes of the
`endosome (comprised of sulfated proteoglycans and anionic lipids), which
`tends to enhance adhesion and fusion. Arrows in the expanded view in
`d indicate that the bending of the membranes hinders fusion.
`
`0.012 e/A˚ 2  e/(83 A˚ 2), we observed released DNA inside
`cells consistent with the escape of CL-DNA complexes into
`the cytoplasm through fusion with anionic endosomal
`membranes (Fig. 1 d ). LSCM further determined that the
`released DNA was condensed (Fig. 1 e), most likely, with
`oppositely charged cytoplasmic condensing agents absent in
`endosomes. Unlike the endosomal environment, the cyto-
`plasm contains many multivalent cationic biomolecules such
`as spermine and histones, which become available during the
`cell cycle in millimolar concentration levels, and are able to
`condense DNA (Bloomfield, 1991).
`Corresponding TE measurements exhibited two remark-
`able features. First and foremost, an unexpected simplicity
`emerged from this highly complex LC
`a-cell system, where sM
`was found to be a universal parameter controlling TE. The TE
`data for LC
`a complexes containing DOPC mixed with the
`multivalent cationic lipid DOSPA or the univalent cationic
`lipids DOTAP and DMRIE, at a constant cationic to anionic
`charge ratio, merged onto a universal curve when plotted
`versus sM. Second, this universal TE curve increased ex-
`ponentially, over four decades, with increasing sM. This be-
`havior is consistent with a model describing a kinetic barrier
`for fusion of CL-DNA complexes with the endosomal mem-
`brane (Fig. 1 d), where an increase in sM lowers the barrier
`height. This new understanding of the fundamental role of the
`lipid carrier sM has led to redesigned LC
`a DOPC-based carriers
`with efficacy competitive with the TE of HC
`II DOPE-based
`carriers.
`We used positively charged CL-DNA complexes prepared
`at r ¼ DOTAP/DNA (wt./wt.) ¼ 6 (r ¼ 2.2 is the isoelectric
`point) from mixtures of cationic and neutral lipids complexed
`with plasmid DNA (pGL3). The weight ratio r ¼ 6, which
`gave a cationic-to-anionic charge ratio of 2.8, was chosen as it
`corresponded to the middle of a typical plateau region
`observed for optimal transfection conditions as a function of
`increasing r above the isoelectric point. X-ray diffraction
`(XRD) results elucidated the structures of these complexes in
`water and in DMEM, a common environment for in vitro
`studies of cells. Synchrotron XRD of DOTAP/DOPC com-
`plexes at the mole fraction FDOPC ¼ 0.67 (Fig. 2, left)
`showed sharp peaks at q001 ¼ 0.083 A˚ ÿ1, q002 ¼ 0.166 A˚ ÿ1,
`with a shoulder peak at q003 ¼ 0.243 A˚ ÿ1 (due to the form
`factor), and q004 ¼ 0.335 A˚ ÿ1, resulting from the layered
`a phase (d ¼ interlayer spacing ¼ dm1 dw ¼
`structure of the LC
`2p/q001 ¼ 75.70 A˚ ) with DNA intercalated between cationic
`left,
`inset). For DOTAP/DOPE
`lipid bilayers (Fig. 2,
`complexes at FDOPE ¼ 0.69, XRD (Fig. 2, right) revealed
`four orders of Bragg peaks at q10 ¼ 0.103 A˚ ÿ1, q11 ¼ 0.178
`A˚ ÿ1, q20 ¼ 0.205 A˚ ÿ1, and q21 ¼ 0.270 A˚ ÿ1, denoting the HC
`phase (Fig. 2, right, inset) with a unit cell spacing of a ¼
`p
`q10Þ ¼ 70:44 ˚A: Except for a difference in lattice
`4p=ð
`3
`constants, the structures of CL-supercoiled DNA complexes
`are analogous to the ones reported recently for CL-linear
`DNA complexes (Koltover et al., 1998, 1999; Raedler et al.,
`1997; Salditt et al., 1997).
`
`II
`
`ffiffiffi
`
`Biophysical Journal 84(5) 3307–3316
`
`Moderna Ex 1006-p. 3
`Moderna v Arbutus
`
`

`

`3310
`
`Lin et al.
`
`FIGURE 2 Comparison of structure and transfection efficiency (TE). Left (mole fraction FDOPC ¼ 0.67) shows a typical XRD scan of lamellar (inset) LC
`a
`complexes. Right (mole fraction FDOPE ¼ 0.69) shows a typical XRD scan of inverted hexagonal (inset) HC
`II complexes. Middle displays the corresponding
`TE, as measured by luciferase enzyme assays of transfected mouse L-cells.
`
`Transfection experiments were done using plasmid DNA
`(pGL3), which contains the firefly luciferase reporter gene, to
`measure TE and its correlation to the solution structures of
`complexes. Mouse L-cells were transfected with CL-DNA
`complexes and incubated on average for at least a full cell-
`cycle, during which expression occurred. A standard luci-
`ferase assay allowed us to evaluate quantitatively the amount
`of synthesized protein by measuring the bioluminescence
`(number of emitted photons) in relative light units per mg of
`cell protein. Fig. 2 (middle) clearly demonstrates the higher
`TE, by more than two decades, attained by complexes in the
`II phase at FDOPE ¼ 0.69 compared to LC
`HC
`a complexes at
`FDOPC ¼ 0.67.
`To further understand the structure-function correlation,
`we examined the transfer process of CL-DNA complexes into
`cells and the mechanism of the subsequent DNA release
`using LSCM, which provides an optical resolution of ;0.3
`mm in the x and y, and ;3/4 mm in the z. The complexes were
`doubly tagged with fluorescent labels, a red one for lipid and
`a green covalent one for DNA. Fig. 3 shows LSCM pictures
`of mouse L-cells after 6 h of transfer time. By comparing
`images of the x-y, y-z, and x-z planes, we were able to
`determine the position of an object relative to a cell. Fig. 3 A
`shows a typical confocal image of a mouse cell transfected
`II complexes at FDOPE ¼ 0.69. The lipid fluorescence
`with HC
`clearly outlines the plasma membrane, indicating fusion of
`lipid with the plasma membrane before or after entry through
`the endocytic pathway (Wrobel and Collins, 1995). An
`aggregate of complexes (yellow) was seen inside a cell as well
`as a clump of DNA (green) in the cytoplasm. The image
`shows that the interaction between HC
`II complexes and cells
`leads to the dissociation and release of DNA from the CL-
`vector consistent with the measured high TE.
`The corresponding confocal images with LC
`a complexes at
`FDOPC ¼ 0.67 are shown in Fig. 3 B. In striking contrast to
`transfection with HC
`II complexes, we observed no free DNA,
`
`Biophysical Journal 84(5) 3307–3316
`
`but rather many individual intact CL-DNA complexes inside
`cells. Fig. 3 B highlights one such typical complex. In the
`absence of fusion, complexes entered cells through endocy-
`tosis (Fig. 1). This was further substantiated in LSCM images
`of cells prepared at 48C, where endocytosis is inhibited,
`which showed complexes attached to the outside cell surface
`and none within the cell body (Lin et al., 2000; Lin, 2001;
`Safinya et al., 2002). At FDOPC ¼ 0.67, most of the DNA
`remained trapped by the CL-vector consistent with the
`measured low TE. As we discuss below, chloroquine-based
`experiments showed that the intact CL-DNA complexes were
`typically trapped within endosomes.
`As the concentration of DOPC decreased in the LC
`a CL-
`DNA complexes, we observed an unexpected enhancement
`in TE by two decades. Fig. 4 A (red diamonds) exhibits the
`nontrivial dependence of TE on FDOPC for DOPC/DOTAP-
`DNA complexes, which starts low for 0.5 \ FDOPC \ 0.7
`and increases dramatically to a value, at FDOPC ¼ 0.2,
`rivaling that achieved by DOPE/DOTAP-DNA complexes.
`Similar results were obtained for another univalent cationic
`lipid DMRIE (Fig. 4 A, black triangles). The key experi-
`ment, which led to a deeper understanding of TE, was
`a study done with the multivalent cationic lipid DOSPA (blue
`squares) replacing DOTAP. A qualitatively similar trend was
`observed with TE decreasing rapidly above a critical F
`DOPC,
`albeit with F
`DOPC shifted from ;0.2 (observed for DOTAP
`and DMRIE complexes) to 0.7 6 0.1 (DOSPA). The main
`difference between the cationic lipids is the notably larger
`charge density of DOSPA (Fig. 4 B, inset), with a larger head-
`group carrying potentially up to five cationic charges. Thus,
`at a given FDOPC, the membrane charge density (sM) is sig-
`nificantly larger in DOSPA compared to DOTAP or DMRIE
`containing complexes.
`We show in Fig. 4 B the same TE data of Fig. 4 A, now
`plotted versus the membrane charge density sM (i.e., the
`average charge per unit area of the cationic membrane):
`
`Moderna Ex 1006-p. 4
`Moderna v Arbutus
`
`

`

`Imaging and Efficacy of Lipid-DNA Complex
`
`3311
`
`area of the cationic to neutral lipid; scl ¼ eZ/Acl is the charge
`density of the cationic lipid with valence Z; and Fnl ¼ Nnl/
`(Nnl 1 Ncl) and Fcl ¼ Ncl/(Nnl 1 Ncl) are the mole fractions
`of the neutral and cationic lipids, respectively. For the plots in
`Fig. 4 B, we used Anl ¼ 70 A˚ (Langmuir trough data), rDOTAP
`¼ rDMRIE ¼ 1, rDOSPA ¼ 2, ZDOTAP ¼ ZDMRIE ¼ 1, and
`ZDOSPA¼ 3. We found that values of ZDOSPA between 3 and
`4 yield a good visual fit for the comparison between DOTAP
`and DOSPA while 2 and 5 clearly do not. For environments
`of neutral pH we expect ZDOSPA to be closer to 4. The value
`of ZDOSPA could be regarded as a fitting parameter in the
`range between 3 and 4. Remarkably, given the complexity of
`the CL-DNA-cell system, the data, spread out when plotted as
`a function of Fnl (Fig. 4 A), coalesce into a ‘‘universal’’ curve
`as a function of sM, with TE varying exponentially over
`nearly four decades as sM increases by a factor of  8 (Fig. 4
`B, sM between 0.0015 e/A˚ 2 and 0.012 e/A˚ 2), clearly demon-
`strating that sM is a key universal parameter for transfection
`with lamellar LC
`a CL-vectors. We now observe a single
`
`M  0:0104 6 0:0017 e= ˚A2  e=ð100 ˚A
`optimal s
`2
`) (Fig. 4
`B, arrow) where the universal TE curve saturates for
`sM [s
`M for both univalent and multivalent cationic lipid-
`containing CL-vectors. sM controls the average DNA
`interaxial spacing dDNA (Fig. 2, inset), which decreases as
`sM increases (Koltover et al., 1999; Raedler et al., 1997).
`Future designs of CL-vectors, which further enhance the
`packing of DNA based on the recent theoretical understand-
`ing of
`intermolecular
`interactions within the complex
`(Bruinsma, 1998; Harries et al., 1998; O’Hern and Lubensky,
`1998), may be expected to enhance TE.
`The TE data suggest vastly diverse behaviors of LC
`a CL-
`DNA complexes between low and high sM. As we discussed
`earlier, for low sM ¼ e/(200 A˚ 2) where TE is low, confocal
`images show DNA locked within complexes after endocy-
`tosis (Fig. 3 B). To test the idea that it is the endosomal vesicle
`that traps the complex, we carried out transfection experi-
`ments in the presence of chloroquine, a well-established bio-
`assay known to enhance the release of trapped material within
`endosomes by osmotically bursting the vesicle. The endo-
`cytic pathway involves the fusion of endosomes with lyso-
`somes (vesicles containing enzymes for degradation of
`material within endosomes) leading to late-stage endosomes,
`limiting the time available for CL-DNA complexes to escape.
`Chloroquine, a weak base, penetrates the lysosome and accu-
`mulates in a charged state; thus, lysosomes and late-stage
`endosomes tend to rupture due to increased osmotic pressure
`caused by counterions rushing in (Voet and Voet, 1995).
`The fractional increase (TEchloroquine/TE) for the DOSPA/
`DOPC and DOTAP/DOPC systems with added chloroquine
`as a function of sM (Fig. 4 C) shows the large increase in TE
`by as much as a factor of 60 as sM decreases and indicates that
`at low sM lamellar LC
`a complexes are trapped within endo-
`somes, consistent both with the confocal images (Fig. 3 B)
`and the measured low TE without chloroquine. At high sM,
`chloroquine has a much smaller effect on TE with the frac-
`
`Biophysical Journal 84(5) 3307–3316
`
`FIGURE 3 Laser scanning confocal microscopy (LSCM) images of
`transfected mouse L-cells. Red denotes lipid; green, DNA; and yellow, the
`overlap of the two denotes CL-DNA complexes. For each set, middle is the
`x-y top view at a given z; right is the y-z side view along the vertical dotted
`line; bottom is the x-z side view along the horizontal dotted line. Objects in
`circles are indicated by arrows in the x-z and y-z plane side views. (A) Cells
`II complexes (FDOPE ¼ 0.69), show fusion of lipid (red )
`transfected with HC
`with the cell plasma membrane and the release of DNA (green in the circle)
`within the cell. Thus, HC
`II complexes display clear evidence of separation of
`lipid and DNA, which is consistent with the high transfection efficiency of
`such complexes. The released exogenous DNA is in an aggregated state,
`which implies that
`it has been condensed with oppositely charged
`biomolecules of the cell, which remain to be identified. (B) Cells transfected
`a complexes at FDOPC ¼ 0.67, which results in a low membrane
`with LC
`charge density sM  0.005 e/A˚ 2 and low transfection efficiency (as plotted
`in Fig. 4 B). First, no fusion is observed. Second, intact CL-DNA complexes
`are observed inside cells (one such yellow complex is shown in the circle).
`The intact complex implies that DNA remains trapped within the complex,
`which is consistent with the observed low transfection efficiency of such low
`sM=LC
`a complexes. Because of the lack of fusion (which aided observation
`of the cell outline in A), we achieved cell outline by observation in reflection
`mode, which appears as blue. Bars ¼ 5 mm applies to all planes.
`sM ¼ eZNcl=ðNnlAnl1NclAclÞ
`¼ ½1 ÿ Fnl=ðFnl1rFclފscl:
`Here, Nnl and Ncl are the number of neutral and cationic
`lipids, respectively; r ¼ Acl/Anl is the ratio of the headgroup
`
`(1)
`
`Moderna Ex 1006-p. 5
`Moderna v Arbutus
`
`

`

`3312
`
`Lin et al.
`
`D. At high FDOPE [ 0.56, DOTAP/DOPE-DNA complexes
`are in the HC
`II phase (green open squares) and exhibit high
`TE. We further see that in contrast to DOPC containing
`complexes, which show a strong dependence on the mole
`fraction of neutral lipid (or equivalently sM), TE of DOPE
`containing complexes is independent of sM. The data show
`unambiguously that sM is a critical parameter for TE by LC
`a
`complexes and not so for HC
`II complexes. The mechanism
`of transfection by DOPE containing HC
`II complexes is in-
`dependent of sM and dominated by other effects; possibly,
`for example, the known fusogenic properties of inverted
`hexagonal phases. In contrast, for LC
`a complexes, high TE
`requires sM [s
`to produce a high TE of LC
`M: Thus,
`a
`complexes with large mole fraction of neutral lipid ;0.70
`(i.e., similar to those mole fractions where DOPE containing
`HC
`II complexes show high TE) requires the incorporation of
`multivalent cationic lipids (e.g., DOSPA) to ensure that
`sM [s
`M.
`LSCM images of cells transfected with LC
`a complexes
`at high sM displayed a path of complex uptake and DNA
`release distinct from transfections done with both LC
`a com-
`plexes at low sM (Fig. 3 B) and HC
`II complexes (Fig. 3 A).
`a complexes at FDOPC ¼
`A typical confocal image with LC
`0.18 (sM  0.012 e/A˚ 2) is shown in Fig. 5. Intact complexes
`were found inside the cell (Fig. 5, label 2 in x-y plane; box 2
`shows the equal green (DNA) and red (lipid) fluorescence
`intensity along the dotted line in x-y; see inset), but more
`interestingly, a mass of exogenous DNA successfully trans-
`ferred into the cytoplasm was also clearly evident (Fig. 5,
`label 1 in x-y plane; box 1 shows the much larger green
`fluorescence (DNA) intensity along the x-y direction). In the
`absence of fusion, high sM complexes enter cells through
`endocytosis. The integrated fluorescence intensity of the
`observed DNA (Fig. 5, box 1) is comparable to that of
`DNA complexed with lipids (Fig. 5, box 2), indicating that
`the released DNA is in the form of aggregates. Because
`endosomes contain no known DNA-condensing agent, these
`aggregates must reside in the cytoplasm (Fig. 1 e). The
`presence of lipid-released DNA in the cytoplasm after
`endocytic uptake of complexes agrees with the measured
`high TE, and moreover, implies fusion between CL-DNA
`lipids and endosomal membranes (Fig. 1 d), enabling escape
`from the endosomes. This is consistent with our finding that
`chloroquine has a small effect at high sM because endosomal
`escape is no longer a major obstacle.
`The confocal image also captured an aggregate of comp-
`lexes in contact with a large area of the cell surface,
`displaying no tendency toward fusion with the plasma mem-
`brane (Fig. 5, label 3 in x-y plane). Comparing the changes
`in fluorescence intensity along the x-y (Fig. 5, box 3) and z
`(Fig. 5, box 4) directions, from the outside toward the inside
`of the cell, we clearly see an aggregate of complexes caught in
`the process of dissociation after endocytosis, with released
`DNA toward the inside of the cell.
`
`(A) Transfection efficiency (TE) expressed as relative light
`FIGURE 4
`units per mg total cellular protein plotted as a function of varying mole fraction
`DOPC with cationic lipids DOSPA, DOTAP, and DMRIE. (B) TE plotted
`versus the membrane charge density sM, demonstrating universal behavior of
`CLs containing cationic lipids with different molecular charges and head-
`group areas (inset). For all three systems, TE increases with sM until an
`M (arrow at  0.0104 e/A˚ 2; determined by the intersection of two
`optimal s
`straight lines fit to the data (dashed lines) above and below the ‘‘knee’’), at
`which TE plateaus. (C ) The fractional increase TEchloroquine/TE for the
`DOSPA/DOPC and DOTAP/DOPC systems with added chloroquine as
`a function of sM. (D) TE plotted as a function of mole fraction of neutral lipid
`for the DOTAP/DOPC and the DOTAP/DOPE systems. The DOTAP/DOPC
`system is LC
`a . The DOTAP/DOPE system goes through two phase transitions:
`
`
`LCa ( filled green square) to coexisting LCa 1 HC
`II (green square with cross) to
`HC
`II (open green square). All transfection measurements were done with 2 mg
`of plasmid DNA at a constant cationic to anionic charge ratio of 2.8. (2.8 was
`chosen as it corresponded to the middle of a typical plateau region observed
`for optimal transfection conditions as a function of increasing cationic to
`anionic charge ratio above the isoelectric point of the complex.) Thus, every
`TE data point of A and B (which is found to vary by nearly four orders of
`magnitude) used the same amount of total charged species (anionic charge
`from DNA, cationic charge from cationic lipid) and t

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