throbber
THE JOURNAL OF BIOLOGICAL CHEMISTRY
`O 1994 by The American Society for Biochemistry and Molecular Biology, Inc.
`
`Vol. 269, No. 10, Issue of March 11, pp. 7185-7191, 1994
`Printed in U.SA.
`
`Interactions of Synthetic Peptide Analogs of the Class A
`Amphipathic Helix with Lipids
`EVIDENCE FOR THE SNORKEL HYP0THESIS*
`
`(Received for publication, October 12, 1993, and in revised form, December 6, 1993)
`
`Vinod K. Mishra, Mayakonda N. Palgunachari, Jere P. Segrest, and G. M. Anantharamaiaht
`From the Departments of Medicine and Biochemistry and the Atherosclerosis Research Unit, University of
`Alabama-Birmingham Medical Center, Birmingham, Alabama 35294
`
`Class A amphipathic helixes present in exchangeable
`plasma apolipoproteins are characterized by the loca-
`tion of positively charged amino acid residues at the
`non-polar-polar interface and negatively charged amino
`acid residues at the center of the polar face. The objec-
`tives of the present study were: (i) to investigate the role
`of hydrocarbon side chain length of the interfacial posi-
`tively charged amino acid residues in the lipid affinity
`of class A amphipathic helixes, and (ii) to investigate the
`importance of the nature of interfacial charge in the
`lipid affinity of class A amphipathic helixes. Toward this
`end, lipid interactions of the following two analogs of
`the class A amphipathic helix, Ac-18A-NH2 (acetyl-Asp-
`Trp-Leu-Lys-Ala-Phe-Tyr-Asp-Lys-Val-Ala-Glu-Lys-Leu-
`Lys-Glu-Ala-Phe-NH2), and Ac-18A(Lys>Haa)-NH2
`(acetyl-Asp-Trp-Leu-Haa-Ala-Phe-Tyr-Asp-Haa-Val-Ala-
`Glu-Haa-Leu-Haa-Glu-Ala-Phe-NH2) (Haa = homoamino-
`alanine), were studied. The side chain of Haa has two
`CH2 groups less than that of lysine. The lipid affinities of
`these two peptide analogs were compared with that of
`Ac-18R-NH2, an analog of Ac-18A-NH2 with positions of
`the charged amino acid residues reversed. The tech-
`niques used in these studies were circular dichroism,
`fluorescence spectroscopy, right-angle light scattering
`measurements, and differential scanning calorimetry.
`The results of these studies indicated the following rank
`order of lipid affinity: Ac-18A-NH2 > Ac-18A(Lys>Haa)-
`NH2 > Ac-18R-NH2. These results are in agreement with
`the "snorkel" model proposed earlier to explain the
`higher lipid affinity of class A amphipathic helixes (Seg-
`rest, J. P., Loof, H. D., Dohlman, J. G., Brouillette, C. G.,
`and Anantharamaiah, G. M. (1990) Proteins Struct.
`Funct. Genetics 8, 103-117). In addition, it was observed
`from the differential scanning calorimetry studies that
`Ac-18A-NH2 and Ac-18A(Lys>Haa)-NH2 interact more
`strongly than Ac-18R-NH2 with negatively charged di-
`myristoyl phosphatidylglycerol. The weaker interaction
`of Ac-18R-NH2 with dimyristoyl phosphatidylglycerol is
`suggested to be due to electrostatic repulsion between
`the negatively charged lipid and the interfacial negative
`charges of the peptide.
`
`Exchangeable plasma apolipoproteins are protein detergents
`and are responsible for carrying lipids in circulation in the form
`of lipoproteins. There is now considerable evidence that the
`structural motif that is responsible for their detergent property
`
`* This research was supported in part by National Institutes of
`Health Program Project HL34343. The costs of publication of this ar-
`ticle were defrayed in part by the payment of page charges. This article
`must therefore be hereby marked "advertisement" in accordance with 18
`U.S.C. Section 1734 solely to indicate this fact.
`To whom correspondence should be addressed.
`
`is the amphipathic a-helix (1). While many other proteins and
`biologically active peptides have been found to possess this
`secondary structural motif, the exchangeable apolipoproteins
`are unique in that they possess multiple copies of amphipathic
`a-helixes with positively charged amino acid residues at the
`non-polar-polar interface and negatively charged amino acid
`residues at the center of the polar face (2). The amphipathic
`a-helixes with this kind of distribution of charged amino acid
`residues in the polar face have been classified as class A (3-5).
`The lipid-associating properties of exchangeable apolipopro-
`teins are believed to reside predominantly in the amphipathic
`helical domains possessing the class A motif (3).
`On the basis of the key structural features predicted for the
`lipid-associating properties of the amphipathic helix (1, 2),
`many laboratories, including ours, have studied structural and
`functional properties of de novo designed peptide analogs of the
`amphipathic helix (6-16). We have addressed the question
`whether or not the location of charged amino acid residues on
`the polar face of the class A amphipathic helical peptide play a
`role in determining its lipid affinity (17-19). A model peptide
`with 18 amino acid residues was designed to mimic the general
`features of the amphipathic helix described in the original
`model (2, 17). The model peptide 18A, with primary sequence
`Asp-Trp- Leu-Lys-Ala-Phe-Tyr-Asp- Lys-Val -Al a-Glu-Ly s- Leu-
`Lys-Glu-Ala-Phe, in the helical wheel representation has posi-
`tively charged Lys residues at the non-polar-polar interface and
`negatively charged Asp and Glu at the center of the polar face
`(17). To address the question of contribution of interfacial Lys
`residues to the lipid affinity of 18A, a peptide with the same
`amino acid composition, but the positions of charged amino acid
`residues reversed, was synthesized. The peptide reverse-18A
`(18R) had the following amino acid sequence: Lys-Trp-Leu-Asp-
`Ala-Phe-Tyr-Lys-Asp-Val-Ala-Lys-Glu-Leu-Glu-Lys-Ala-Phe. A
`helical wheel representation of this sequence shows negatively
`charged amino acid residues Asp and Glu at the non-polar-
`polar interface and positively charged Lys residues at the cen-
`ter of the polar face (17). Studies on these two peptides and
`their analogs showed that class A amphipathic peptides with
`positively charged residues at the non-polar-polar interface and
`negatively charged residues at the center of the polar face have
`higher lipid affinity compared with the corresponding charge-
`reversed peptide analogs (17-20).
`To explain the higher lipid affinity of class A amphipathic
`helixes, the "snorkel" model was proposed (3). The bulk of the
`van der Waals surface area of the positively charged lysine and
`arginine residues is hydrophobic. In the snorkel model, it was
`proposed that the interfacial positively charged amino acid
`residues of the peptide, when associated with lipid, extend
`toward the polar face of the helix to insert their charged moi-
`eties into the aqueous environment for solvation. Thus, in the
`snorkel orientation, the entire uncharged van der Weals sur-
`face of the class A amphipathic helix is buried within the hy-
`
`7185
`
`This is an Open Access article under the CC BY license.
`
`MYLAN EXHIBIT - 1030
`Mylan Pharmaceuticals, Inc. v. Bausch Health Ireland, Ltd. - IPR2022-00722
`
`

`

`7186
`
`Lipid Interactions of Amphipathic Helical Peptides
`
`drophobic interior of a lipid bilayer.
`Apolipoprotein A-I (apoA-I),' the major protein constituent of
`plasma high density lipoproteins, and class A amphipathic hel-
`ical peptide analogs have been shown to stabilize the bilayer
`structure of phospholipids, and these properties were corre-
`lated to their ability to protect against lytic peptide-induced
`erythrocyte lysis (21). These molecules also exhibit anti-viral
`and anti-inflammatory properties and protect fatty acid con-
`taining dye-entrapped phospholipid vesicles from albumin-in-
`duced leakage (22-24). These properties were attributed to the
`snorkeling effect of interfacial Lys residues in these molecules.
`This snorkeling effect was thought to create a "wedge" cross-
`sectional shape which is responsible for the stabilization of
`phospholipid vesicles (21, 24). It was, therefore, important to
`determine if indeed the longer hydrocarbon side chain of inter-
`facial Lys residues increases the lipid-associating ability, which
`is a key determinant of the properties of apoA-1 and the corre-
`sponding peptide analogs.
`Earlier, we reported that N- and C-terminal protection of 18A
`drastically increased its helicity as well as lipid affinity and the
`resulting peptide, Ac-18A-NH2 (acetyl-Asp-Trp-Leu-Lys-Ala-
`Phe-Tyr-Asp-Lys-Val-Ala-Glu-Lys-Leu-Lys-Glu-Ala-Phe-NH2),
`closely mimicked the properties of apoA-I (25). In the present
`report, we compare the lipid affinity of Ac-18A(Lys>Haa)-NH2
`(acetyl-Asp-Trp-Leu-Haa-Ala-Phe-Tyr-Asp-Haa-Val-Ala-Glu-
`Haa-Leu-Haa-Glu-Ala-Phe-NH2) with that of Ac-18A-NH2 and
`Ac-18R-NH2
`(acetyl-Lys-Trp-Leu-Asp-Ala-Phe-Tyr-Lys-Asp-
`Val-Ala-Lys-Glu-Leu-Glu-Lys-Ala-Phe-NH2). The side chain of
`Haa residue has two methylene groups less than that of Lys.
`The peptide Ac-18A(Lys>Haa)-N112 was designed to investigate
`the role of longer hydrocarbon side chain of the interfacial
`lysine residues in the lipid affinity of Ac-18A-NH2. The three
`peptides were studied for their ability to interact with dimy-
`ristoyl phosphatidylcholine (DMPC), dimyristoyl phosphatidyl-
`glycerol (DMPG), and egg yolk phosphatidylcholine (EYPC)
`multilamellar vesicles. The following criteria were used for the
`lipid affinity of these peptides (26-28): (i) increase in the a -hel-
`ical structure in the presence of the lipid as detected by far-UV
`CD spectra; (ii) blue-shift in the tryptophan emission maxi-
`mum and shielding from the aqueous-phase quenchers, iodide
`and acrylamide, in the presence of the lipid as monitored by
`steady-state fluorescence spectroscopy; (iii) clarification of the
`turbidity due to EYPC multilamellar vesicles as measured by
`right-angle light scattering; and (iv) reduction in the transition
`enthalpy and broadening in the gel to liquid-crystalline phase
`transition of the DMPC and DMPG multilamellar vesicles as
`measured by differential scanning calorimetry (DSC).
`
`EXPERIMENTAL PROCEDURES
`Materials—Synthesis and purification of the peptide Ac-18A-NH2
`has been described earlier (25). Peptide Ac-18R-NH2 was synthesized
`and purified following similar methodology (20). For the synthesis of the
`peptide Ac-18A(Lys>Haa)-NH2, Boc-Haa(Z), instead of Boc-Lys(2-C1Z),
`was used. The amino acid derivative was prepared using the procedure
`described elsewhere (29). The other steps in the synthesis and purifi-
`cation were similar to those described earlier for Ac-18A-NH2 (25).
`DMPC, DMPG, and EYPC, purity > 99%, were purchased from
`Avanti Polar Lipids, Inc. (Alabaster, AL) and used without further pu-
`rification. Potassium iodide (KO was obtained from Fisher Scientific
`Company, and acrylamide (>99.9%) was purchased from Bio-Rad. (1S )-
`(TFE,
`(99%), 2,2,2-trifluoroethanol
`(+)-10-Camphorsulfonic acid
`
`The abbreviations used are: apoA-I, apolipoprotein A-I; Boc, terti-
`ary-butyloxycarbonyl; Z, benzyloxycarbonyl; (2C1)Z, 2-chlorobenzyloxy-
`carbo-nyl; DMPC, 1,2-dimyristoyl-sn-glycero-3-phosphocholine; DMPG,
`1,2-dimyristoyl-sn-glycero-3-lphospho-rac-(1-glyceronlisodium salt);
`EYPC, egg yolk t-a-phosphatidylcholine; NATA, N-acetyltryptophan-
`amide; PBS, phosphate-buffered saline; TFE, 2,2,2-trifluoroethanol;
`DSC, differential scanning calorimetry; Haa, homoaminoalanine; Ac,
`acetyl; NH2, carboxamide.
`
`99.5+%, NMR grade), and guanidine hydrochloride (99%) were obtained
`from Aldrich. All other chemicals were of highest purity commercially
`available.
`Preparation of Peptide Solutions—Peptide solutions were prepared
`by dissolving the peptide in 4 M guanidine hydrochloride and dialyzing
`against phosphate-buffered saline (PBS, pH 7.4; KH2PO4 1.47 ms,
`Na2HPO4.7H2O 6.45 ms, NaCl 136.89 rnm, KCl 2.68 mm) extensively
`(overnight, with at least three buffer changes). All the studies described
`below were done in the same buffer. Peptide concentrations were deter-
`mined in 4 M guanidine hydrochloride by measuring absorbance at 280
`nm (€ 20. = 7300 m-1 cm-').
`Circular Dichroism—The CD spectra were recorded with an AVIV
`62DS spectropolarimeter interfaced to a personal computer. The instru-
`ment was calibrated with (1S )-(+)-10-camphorsulfonic acid (30, 31). The
`CD spectra were measured from 260 to 190 nm every nm with 1 s
`averaging per point, and a 2-nm bandwidth. An 0.01-cm path length cell
`was used for obtaining the spectra. All CD spectra were signal averaged
`by adding four scans, base-line corrected, and smoothed. All the CD
`spectra were recorded at 25 °C. Temperature was regulated with a
`Lauda RS2 circulating water bath. Final peptide concentrations of 100
`pi or less were used for obtaining the CD spectra. DMPC multilamellar
`vesicles suspension was prepared by dissolving a known amount of lipid
`(-10 mg) in ethanol in a test tube and evaporating the solvent slowly
`under a stream of dry nitrogen. The residual solvent was removed by
`storing the tube under high vacuum overnight in a vacuum oven at
`25 °C. To the dried lipid film appropriate volume of buffer was added to
`give a final lipid concentration of 14.7 mm. The lipid film was hydrated
`by vortexing the mixture for 30 min at room temperature. Peptide-
`DMPC complexes for CD studies were prepared as follows. Appropriate
`volume of peptide solution in buffer was added to the DMPC multila-
`mellar vesicles suspension to give lipid to peptide molar ratio of 20:1.
`The mixture was incubated overnight at room temperature. This
`yielded a clear solution of the peptide-DMPC complex.
`The mean residue ellipticity, lO1mRE(deg-cm2.dmol-1), was calculated
`using the following equation:
`
`[O1,
`
` = MRW.0/10.c.1
`
`where, MRW is mean residue weight of the peptide, O is the ob(sEergVeld)
`ellipticity in degrees, c is the concentration of the peptide in g/rril, and
`1 is the path length of the cell in centimeters. The percent helicity of the
`peptide was estimated from the following equation:
`
`(Eq. 2)
`
`% a helix = (l0 i222 + 3,000)/(36,000 + 3,000)
`where, [01222 is the mean residue ellipticity at 222 nm (32).
`fluorescence emission
`Fluorescence Measurements-Steady-state
`spectra were recorded in the ratiometric mode on an SLM 8000C photon
`counting spectrofluorometer (SLM Instruments, Inc., Urbana, IL) at
`25 °C. Excitation and emission slit widths were both 4 nm. An excitation
`wavelength of 280 nm was used for recording emission spectra. Wave-
`lengths at which maximum emission occurred (AOOO,) were determined
`by positioning the cursor manually at the peak maximum and reading
`the corresponding wavelength. The final peptide concentration used
`was 14 gm (A280 = 0.1).
`Fluorescence quenching experiments were performed using 295-nm
`excitation wavelength in order to minimize absorptive screening by the
`quenchers used. Aliquots (10 ul) of freshly prepared potassium iodide (4
`M) or acrylamide (4 M) stock solutions were added to a constantly stirred
`and thermostated (25 °C) 2-ml peptide solution. After every addition of
`the quencher, the emission spectra were recorded from 310 to 450 nm,
`and the emission intensity at Am. were determined as described above.
`Stock solution of potassium iodide contained 1 mm sodium thiosulfate
`(Na2S2O3) in order to prevent I3 formation (33). For acrylamide quench-
`ing, corrections for inner filter effects (€ 295 = 0.25 m--1 cm-1, for acryl-
`amide) were made (34). The fluorescence quenching data were analyzed
`according to the Stern-Volmer equation for the collisional quenching
`(35):
`
`FVF = 1 + Ksv[Q] = 70/r= 1 + iz,t0[Q]
`
`(Eq. 3)
`
`where, F. and F are fluorescence intensities in the absence and pres-
`ence of quencher, respectively, K. is Stern-Volmer constant for the
`collisional quenching process, fQJ is quencher concentration, +0 and r
`are fluorescence lifetimes of the flurophore in the absence and presence
`of the quencher, respectively, and kq is the rate constant for the bimo-
`lecular quenching process. The above equation predicts a linear plot of
`Fo/F (or +0/+) versus [Q] for a homogeneously emitting solution. The
`slope of this plot yields the value of K.„.
`
`

`

`Lipid Interactions of Amphipathic Helical Peptides
`
`7187
`
`35000
`
`30000
`
`2500D
`
`20000
`
`15000
`
`10000
`
`5000
`
`-5000
`
`- /0000
`
`-15000
`
`-200
`
`00
`
`E
`
`43
`47
`2
`
`B
`
`A
`
`70000
`
`60000
`
`50000
`
`40000
`
`J0000
`
`-
`
`71 2 71
`
`0
`
`-5 20000
`
`10000
`
`:
`
`- 10000
`
`00
`
`210
`
`220
`
`230
`
`2.0
`
`250
`
`26D
`
`- 3000 190
`
`00
`
`0
`
`220
`
`230
`
`240
`
`250
`
`260
`
`10000
`
`60000
`
`50000
`
`•0000
`
`30000
`
`ti
`
`m.7
`
`-0 20000
`
`'0000
`
`2
`
`- 10000
`
`-20000
`
`- 3000100
`
`200
`
`2/D
`
`220
`
`230
`
`2.0
`
`250
`
`260
`
`Wavelength (nm)
`Wavelength (nm)
`Flo 1. Far-UV CD spectra of peptides (---100 um) in buffer (PBS, pH 7.4) (A), 50% (v/v) TFE/PBS (B), and DMPC complex (lipid/
`peptide molar ratio 20:1) (C). Ac-18A-NH2 (O), Ac-18A(Lys>Haa)-NH2 (A), Ac-18R-NI-12 (0).
`
`Wavelength (nm)
`
`Right-angle Light Scattering Measurements—The association of the
`peptides with EYPC multilamellar vesicles was followed by monitoring
`the rate of clarification of the turbidity due to the vesicles (36). The rate
`of turbidity clarification was measured by right-angle light scattering
`using SLM 8000C photon counting spectrofluorometer with both exci-
`tation and emission monochromators set at 400 nm. Data were acquired
`using slow time based acquisition. EYPC multilamellar vesicles sus-
`pension was prepared as described above for the DMPC. The sample
`containing 105 pm EYPC and an equimolar amount of the peptide was
`maintained at 25 °C and was continuously stirred. Complete dissolution
`of the EYPC vesicles was achieved by the addition of Triton X-100 to the
`vesicles suspension at a final concentration of 1 mm.
`Differential Scanning Calorimetry—The high sensitivity DSC stud-
`ies were performed in Microcal MC-2 scanning calorimeter (MicroCal,
`Inc., Amherst, MA) at the scan rate of 20`h- ' at an instrumental sen-
`sitivity of 1 and with a filtering constant of 10 s. The lipid multilamellar
`vesicles and the peptide-lipid mixtures for DSC were prepared as fol-
`lows. DMPC or DMPG (-2 mg) was dissolved in chloroform in a test
`tube and dried by slow evaporation under a stream of dry nitrogen. The
`residual solvent was removed under high vacuum in the vacuum oven
`as described above. To the dried lipid film either buffer alone or buffer
`containing the peptide was added to obtain lipid/peptide molar ratio
`100:1. The lipid was hydrated by vortexing at room temperature for 30
`min. The suspension was degassed for 30 min, and an aliquot of L2 ml
`of suspension was used and run against buffer in the reference cell.
`Four consecutive scans with 60 min equilibration time between each
`scan were run for the same sample. No significant changes in the
`thermograms were observed between the first scan and the fourth scan.
`The observed transitions were analyzed using software provided by
`MicroCal, Inc., Amherst, MA.
`
`RESULTS
`CD Studies—Secondary structures of the peptides were de-
`termined by recording their far-UV CD spectra in buffer (PBS,
`pH 7.4), in 50% (v/v) TFE/PBS, and in the presence of DMPC
`(lipid/peptide molar ratio 20:1). The CD spectra are shown in
`Fig. 1. The CD spectrum of an a-helix is characterized by two
`negative bands, one at 222 nm and another at 208 nm, and a
`positive band at 192 nm (37). The CD spectrum of Ac-18A-NH2
`in buffer (Fig. 1A) indicates that the peptide is largely helical.
`As estimated from mean residue ellipticity at 222 nm [O1222,
`the peptide is 55% a-helical. The CD spectrum of Ac-18R-NH2
`indicates that it is less helical (43%) than Ac-18A-NH2. The CD
`spectrum of Ac-18A(Lys>Haa)-NH2 in buffer is distinct from
`the CD spectra of the other two peptides because of its greatly
`reduced mean residue ellipticity values as well as a large shift
`in the zero cross-over point (wavelength at which mean residue
`ellipticity is zero). This suggests a largely non-helical structure
`of the peptide in buffer. The peptide was estimated to have 22%
`a-helical structure.
`TFE is a most commonly used structure-inducing, hydropho-
`bic, cosolvent (38). CD spectra of the three peptides in 50% (v/v)
`TFE/PBS were recorded to estimate their a-helical contents in
`a relatively hydrophobic environment. As is evident from Fig.
`
`O
`
`0
`
`325
`
`350
`375
`100
`Wavelength loin)
`
`• 5
`
`450
`
`FIG. 2. Fluorescence emission spectra of peptides in buffer
`(PBS, pH 7.4) and peptide-DMPC complexes (lipid/peptide mo-
`lar ratio 20:1). Peptide concentration = 14 pm, excitation wavelength =
`280 nm. Ac-18A-NH2 in buffer (O), Ac-18A-NH2-DMPC complex (6),
`Ac-18A(Lys>Haa)-NH2 in buffer (z), Ac-18A(Lys>Haa)-NH2-DMPC
`complex (A), Ac-18R-NH2 in buffer (0), Ac-18R-NIL-DMPC complex
`( • ).
`
`1B, among the three peptides, Ac-18A-NH2 has the maximum
`a-helical structure (67%). The a-helical contents ofAc-18R-NH2
`and Ac-18A(Lys>Haa)-NH2 are less than that of Ac-18A-NH2
`(61 and 59%, respectively). However, a closer inspection of the
`CD spectra of the peptides in TFE reveals that while Ac-
`18A(Lys>Haa)-NH2 has a zero cross-over point (202.7 nm)
`nearly identical with that of Ac-18A-NH2 (202.6 nm), the zero
`cross-over point of Ac-18R-NH2 is shifted toward a shorter
`wavelength (202 nm). This indicates that Ac-18A(Lys>Haa)-
`NH2 has a higher a-helical content than Ac-18R-NH2 in this
`solvent system (39).
`The CD spectra of peptide-DMPC complexes (lipid/peptide
`molar ratio 20:1) are shown in Fig. 1C. When complexed with
`the lipid, all the three peptides are largely a-helical. The three
`peptides were estimated to have the following a-helical con-
`tents: Ac-18A-NH2
`72%, Ac-18R-NH2 62%, and Ac-
`18A(Lys>Haa)-NH2 56%. We would like to point out that the
`a-helical content of Ac-18A-NH2 in DMPC complex was earlier
`reported to be 92% (25). This value was recently found to be an
`overestimation because of the improper calibration of the spec-
`tropolarimeter.
`Fluorescence Measurements—The fluorescence emission
`maximum of tryptophan is highly sensitive to its microenviron-
`ment (40). The microenvironments of tryptophan residues in
`the peptides were probed by recording their emission spectra in
`buffer and in the presence of DMPC (lipid/peptide molar ratio
`20:1). The emission spectra are shown in Fig. 2. All three pep-
`tides disrupt DMPC vesicles completely at lipid/peptide molar
`ratio 20:1 resulting in an optically clear solution. In buffer
`
`

`

`7188
`
`Lipid Interactions of Amphipathic Helical Peptides
`
`TABLE I
`Tryptophan emission maxima and fluorescence quenching parameters of the peptides and the peptide-DMPC complexes
`Peptide concentration was 14 um; lipid/peptide molar ratio was 20:1. Fluorescence emission maxima were determined using excitation wave-
`length of 280 nm; quenching studies were done using excitation wavelength of 295 nm. Ksv is Stern-Volmer quenching constant determined from
`the slopes of the lines for the plots of Fo/F-1 versus [Q]; slopes were determined by linear regression analysis of the fluorescence quenching data
`using the least squares method.
`
`Peptide
`
`Ac-18A-NH2
`Ac-18A(Lys>Haa)-NH2
`Ac-18R-NH2
`NATA
`* Not determined.
`
`(nm)
`
`DMPC
`complex
`
`329
`331
`336
`ND°
`
`Buffer
`
`343
`344
`344
`350
`
`Ksv (hi')
`
`Buffer,
`iodide
`quenching
`
`5.7
`7.2
`8.0
`11.0
`
`DMPC
`complex,
`iodide
`quenching
`1.4
`3.7
`5.4
`ND
`
`Buffer,
`acrylamide
`quenching
`
`13.5
`15.5
`17.4
`27.7
`
`DMPC
`complex,
`acrylamide
`quenching
`2.6
`6.2
`10.3
`ND
`
`2.0
`
`0.
`
`O 40 GO 00 100 120 140 ISO 100 200 220
`
`[Iodide] mil
`
`0
`
`20 40 SO SO l00 120 I40 16U
`
`100 200 220
`
`tAcrylemldel mil
`Fin. 3. Stern-Volmer plots of fluorescence quenching of the
`peptides in buffer (PBS, pH 7.4), and peptide-DMPC complexes
`(lipid/peptide molar ratio 20:1) by iodide (A) and acrylamide (B).
`Peptide concentration = 14 pm, excitation wavelength = 295 nm. Ac-
`(O), Ac-18A-NH2-DMPC complex (0), Ac-
`in buffer
`18A-NH2
`18A(Lys>Haa)-NH2 in buffer (1k), Ac-18A(Lys>Haa)-NH2-DMPC com-
`plex (A), Ac-18R-NH2 in buffer ( 0), Ac-18R-NH2-DMPC complex( • ),
`NATA in buffer (E).
`
`alone, while Ac-18A-NH2 has its emission maximum (Amex) at
`343 nm, Ac-18R-NH2 and Ac-18A(Lys>Haa)-NH2 both have
`their emission maxima at 344 nm. When complexed with
`DMPC, all the three peptides show an increase in the emission
`intensity as well as a blue-shift in their emission maxima com-
`pared with those in buffer, indicating partitioning of trypto-
`phan into a more hydrophobic environment. However, the ex-
`tent of the blue-shift in the tryptophan emission maximum,
`compared with that in buffer, is different for the three peptides.
`In the presence of DMPC, while Ac-18A-NH2 emission maxi-
`mum shows a blue-shift of 14 nm, Ac-18A(Lys>Haa)-NH2 and
`Ac-18R-NH2 show blue-shifts of 13 and 8 nm, respectively
`(Table I). This indicates that the tryptophan residue in Ac-18R-
`NH2 is in a less hydrophobic environment than the tryptophan
`residues in Ac-18A-NH2 and Ac-18A(Lys>Haa)-NH2.
`
`O
`
`C
`
`SS
`
`O
`
`O
`
`0
`
`o
`
`400
`
`SOO
`
`1200
`
`'GOO
`
`2000
`
`,•••••.•
`
`Time (tlee)
`Fin. 4. Rate of decrease of scattered-light intensity measured
`at 90° from a 105 pm suspension of EYPC vesicles. Equimolar
`amounts of the peptides were added to the lipid suspension. Complete
`dissolution was achieved by adding Triton X-100 solution to the lipid
`suspension at a final concentration of 1 inm. Excitation and emission
`wavelengths both were 400 nm. EYPC alone (M), Ac-18A-NH2 (0), Ac-
`18A(Lys>Haa)-NH2 (A), Ac-18R-NH2 ( • ), Triton X-100 (El).
`
`To further probe the location of the tryptophan residues of
`the three peptides in the lipid bilayer, fluorescence quenching
`experiments were carried out. Iodide and acrylamide were used
`as aqueous-phase quenchers of the tryptophan fluorescence
`(35). While iodide is an anionic quencher, acrylamide is a neu-
`tral but polar quencher. Iodide is considered to have access only
`to surface tryptophans, whereas acrylamide has good access to
`all but the most highly buried tryptophan residues (41). The
`Stern-Volmer plots of the quenching of tryptophan fluorescence
`by iodide and acrylamide are shown in Fig. 3. For comparison,
`Stern-Volmer plots of fluorescence quenching of NATA (N-acet-
`yltryptophanamide), a model compound for free tryptophan, in
`buffer by iodide and acrylamide are also included. In buffer,
`compared with NATA, tryptophans in all the three peptides are
`less exposed to the quenchers. In the presence of DMPC, com-
`pared with buffer, tryptophans in all the three peptides become
`less accessible to the quenchers, suggesting shielding by the
`lipid bilayer. However, the extent of shielding from the quench-
`ers is different for the three peptides. In the presence of DMPC,
`compared with buffer, the following order was observed in the
`shielding of the tryptophan residue of the three peptides from
`the quenchers: Ac-18A-NH2 > Ac-18A(Lys>Haa)-NH2 > Ac-18R-
`NH2. Thus, among the three peptides, tryptophan in Ac-18R-
`NH2 experiences least shielding in going from buffer to the lipid
`bilayer. The Stern-Volmer quenching constants for a bimolecu-
`lar collisional quenching process were calculated from the ap-
`parent slopes of the plots of Fo/F-1 versus [Q] (slopes were
`calculated by linear regression analysis using the "least
`squares" method; 0.989 < r < 0.999) and are given in Table I.
`The Stern-Volmer plots of the fluorescence quenching by acryl-
`amide for the three peptides as well as that of NATA in solution
`
`

`

`Lipid Interactions of Amphipathic Helical Peptides
`
`7189
`
`absence and in the presence of the peptides by DSC. The heat-
`ing endotherms of the pure lipid vesicles and the peptide-lipid
`mixtures (lipid/peptide molar ratio 100:1) are shown in Fig. 5.
`At neutral pH, while DMPC is zwitterionic, DMPG is nega-
`tively charged. Interactions of the peptides with DMPG vesicles
`were studied to investigate possible electrostatic interactions
`between the positively charged amino acid residues of the pep-
`tides and the negative charges of the lipid. DMPC vesicles
`alone exhibited endothermic transitions at 13.7 and 23.1 °C,
`which correspond to the pretransition (lamellar to periodic gel)
`and the gel to liquid-crystalline (order-disorder) phase transi-
`tion of the lipid, respectively (43). Addition of peptides to the
`DMPC vesicles resulted in lowering of the transition enthalpy
`and broadening of the gel to liquid-crystalline phase transition
`of DMPC vesicles (Table II). Among the three peptides, Ac-18A-
`NH2 caused the maximal reduction in the enthalpy of the main
`phase transition, while Ac-18A(Lys>Haa)-NH2 is intermediate
`to Ac-18A-NH2 and Ac-18R-NH2 (Table II). None of the peptides
`changed the chain-melting temperature (T„) of DMPC vesicles
`by more than 0.2 °C. While the pretransition in the DMPC
`vesicles is observed with Ac-18A(Lys>Haa)-NH2 and Ac-18R-
`NH2, it is not detected in the presence of Ac-18A-NH2. Similar
`to DMPC vesicles, DMPG vesicles exhibited endothermic tran-
`sitions at 11.5 and 22.7 °C, corresponding to the pretransition
`and the main phase transition of the lipid, respectively (44).
`The effect of the three peptides on the phase transition prop-
`erties of the DMPG vesicles is similar to that observed with
`DMPC (Table II). The maximum change in T, of the lipid
`vesicles observed in the presence of the peptides is less than
`0.3 °C. The width at half of the peak height of the main phase
`transition in DMPG is more than that in DMPC (Table ID. This
`is presumably due to the poor packing of the lipid molecules in
`the bilayer because of electrostatic repulsion between the nega-
`tive charges in the head group of the phospholipid molecules
`(36).
`
`DISCUSSION
`A comparison of the CD spectra of the three peptides in the
`DMPC complex (Fig. 1C) shows the following rank order of
`helicity: Ac-18A-NH2 > Ac-18R-NH2 > Ac-18A(Lys>Haa)-NH2.
`While the same order is observed in the helicity of the three
`peptides in buffer (Fig. 1A), the helicity of Ac-18A(Lys>Haa)-
`NH2 in the presence of TFE is more than that of Ac-18R-NH2
`(Fig. 1B). In the presence of TFE, Ac-18A(Lys>Haa)-NH2 was
`estimated to have less helical structure than Ac-18R-NH2
`based on the mean residue ellipticity at 222 nm of the two
`peptides. However, comparison of the zero cross-over point of
`the two peptides revealed that Ac-18A(Lys>Haa)-NH2 has more
`helical structure than Ac-18R-NH2 (39). It has been pointed out
`that since the shape and zero cross-over of a CD spectrum
`directly reflect the proportional contribution of different sec-
`ondary structures, any genuine change in secondary structure
`would lead to a change in either the shape or the zero cross-
`over, or both, of that spectrum (39). Therefore, it is important to
`note that a comparison of helicity should not be based on ellip-
`ticity values alone. The results of the CD studies indicate that
`the lipid affinities of Ac-18R-NH2 and Ac-18A(Lys>Haa)-NH2
`are less than the lipid affinity of Ac-18A-NH2. The lipid affinity
`of Ac-18A(Lys>Haa)-NH2, based on the results of the CD stud-
`ies alone, is lower than that of Ac-18R-NI-12.
`A comparison of the fluorescence emission maxima of the
`peptides, however, reveals that in the DMPC complex, while
`tryptophan in Ac-18A(Lys>Haa)-NH2 is in a less hydrophobic
`environment than the tryptophan in Ac-18A-NH2, it is in a
`more hydrophobic environment than the tryptophan in Ac-18R-
`NH2 (Table I). These conclusions are further supported by the
`results from the fluorescence quenching studies (Fig. 3). The
`
`A
`
`(I)
`
`(a)
`
`Endothermic
`
`ID
`
`16
`
`20
`
`26
`
`50
`
`Temperature (cC)
`
`B
`
`(I)
`
`Ili)
`
`wn
`
`(JO
`
`Endothermic
`
`15
`
`20
`
`25
`
`30
`
`Temperature (CC)
`FIG. 5. DSC heating thermograms of lipids and lipid-peptide
`mixtures (lipid/peptide molar ratio 100:1). A, DMPC alone (1.5 ma)
`(i), DMPC + Ac-18R-NH2 (ii), DMPC + Ac-18A(Lys>Haa)-NH2 (iii),
`DMPC + Ac-18A-NH, (iv). B, DMPG alone (1.4 mM) (i), DMPG + Ac-
`18R-NH2 (ii), DMPG + Ac-18A(Lys>Haa)-NH2 (iii), DMPG -4- Ac-18A-
`N112 (it)).
`
`TABLE II
`Phase transition parameters for lipids and peptide-lipid mixtures as
`determined by DSC
`DMPC concentration 1.5 ma, DMPG concentration 1.4 ma, lipid/
`peptide molar ratio 100:1. T,,, is the temperature at which excess heat
`capacity is maximum; Alto is calorimetric enthalpy; AlivH is Van't Hoff
`enthalpy; AT1/2 is width of the transition at half of the peak heat capac-
`ity.
`
`Sample
`
`DMPC
`DMPC/Ac-18A-NH2
`DMPC/Ac-18A(Lys>Haa)-NHz
`DMPC/Ac-18R-NH2
`DMPG
`DMPG/Ac-18A-NI-12
`DMPG/Ac-18A(Lys>Haa)-NH2
`DMPG/Ac-18R-NH2
`
`T,,,
`oc
`23.1
`23.0
`23.3
`23.1
`22.3
`22.2
`22.0
`22.1
`
`AHvii
`
`ATv.
`
`kcal I mol
`5.4
`4284.4
`3.3
`1389.6
`4.1
`1566.8
`4.2
`2537.5
`5.5
`1171.1
`3.0
`480.4
`3.8
`568.4
`4.8
`674.3
`
`°C
`0.2
`0.6
`0.6
`0.3
`0.6
`0.8
`0.7
`0.4
`
`(Fig. 3B) curve upward, indicating the occurrence of static
`quenching (42).
`Right-angle Light Scattering Measurements—Interactions

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