12
Secondary Conformation of Short Lysine- and Leucine-Rich Peptides Assessed by Optical Spectroscopies: Effect of Chain Length, Concentration, Solvent, and Time Bele ´ n Herna ´ ndez 1,2 F.-Z. Boukhalfa-Heniche 1,2 Olivier Seksek 1,2 Yves-Marie Coı¨c 3 Mahmoud Ghomi 1,2 1 UMR CNRS 7033, BioMoCeTi, Universite ´ Pierre et Marie Curie,Case 138, 4 Place Jussieu, 75252 Paris cedex 05, France 2 UMR CNRS 7033, BioMoCeTi, UFR Sante ´ -Me ´ decine-Biologie Humaine, Universite ´ Paris 13, 74 rue Marcel Cachin, 93017 Bobigny cedex, France 3 Unite ´ de Chimie Organique, Institut Pasteur, 28 rue du Docteur Roux, 75724 Paris cedex 15, France Received 18 October 2004; revised 19 August 2005; accepted 22 August 2005 Published online 30 August 2005 in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/bip.20366 Abstract: Solution secondary structures of three synthetic cationic peptides, currently used in antisense oligonucleotide delivery into living cells, have been analyzed by means of circular dichroism (CD) and Raman scattering in different buffers as a function of concentration and time. All three peptides are of minimalist conception, i.e., formed by only two types of amino acids (leucine: L and lysine: K). Two of these peptides contain 15 aminoacids: N ter - KLLKLLLKLLLKLLK (L 10 K 5 ), N ter -KLKLKLKLKLKLKLK (L 7 K 8 ), and the third one has only 9 residues: N ter -KLKLKLKLK (L 4 K 5 ). The conformational behavior of the 15-mers in pure water differs considerably one from another. Although both of them are initially dis- ordered in the 50–350 lM range, L 10 K 5 gradually undergoes a disordered to a-helix transition for molecular concentrations above 100 lM. In all other solvents used, L 10 K 5 adopts a stable a-helical con- formation. In methanol and methanol/Tris mixture, nonnative a-helices can be induced in both KL-alter- nating peptides, i.e., L 7 K 8 and L 4 K 5 . However, in major cases and with a time delay depending on peptide concentration, b-like structures can be gradually formed in both solutions. In PBS and methanol/PBS mixture, the tendency for L 7 K 8 and L 4 K 5 is to form structures belonging to b-family. A discussion has been undertaken on the effect of counterions as well as their nature in the stabilization of ordered struc- tures in both KL-alternating peptides. # 2005 Wiley Periodicals, Inc. Biopolymers 81: 8–19, 2006 This article was originally published online as an accepted preprint. The ‘‘Published Online’’ date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at [email protected] Keywords: lysine; leucine; cationic peptides; secondary conformation; optical spectroscopy Correspondence to: M. Ghomi; e-mail: [email protected] Biopolymers, Vol. 81, 8–19 (2006) # 2005 Wiley Periodicals, Inc. 8

Secondary conformation of short lysine- and leucine-rich peptides assessed by optical spectroscopies: Effect of chain length, concentration, solvent, and time

Embed Size (px)

Citation preview

Secondary Conformation ofShort Lysine- and Leucine-RichPeptides Assessed by OpticalSpectroscopies: Effect of ChainLength, Concentration, Solvent,and Time

Belen Hernandez1,2

F.-Z. Boukhalfa-Heniche1,2

Olivier Seksek1,2

Yves-Marie Coıc3

Mahmoud Ghomi1,21 UMR CNRS 7033,

BioMoCeTi, Universite Pierre etMarie Curie,Case 138,

4 Place Jussieu,75252 Paris cedex 05, France

2 UMRCNRS7033, BioMoCeTi,UFR Sante-Medecine-BiologieHumaine, Universite Paris 13,74 rue Marcel Cachin, 93017

Bobigny cedex, France

3 Unite de Chimie Organique,Institut Pasteur,

28 rue du Docteur Roux,75724 Paris cedex 15, France

Received 18 October 2004;revised 19 August 2005;accepted 22 August 2005

Published online 30 August 2005 in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/bip.20366

Abstract: Solution secondary structures of three synthetic cationic peptides, currently used in antisenseoligonucleotide delivery into living cells, have been analyzed by means of circular dichroism (CD) andRaman scattering in different buffers as a function of concentration and time. All three peptides are ofminimalist conception, i.e., formed by only two types of amino acids (leucine: L and lysine: K). Two ofthese peptides contain 15 aminoacids: Nter- KLLKLLLKLLLKLLK (L10K5), N

ter-KLKLKLKLKLKLKLK(L7K8), and the third one has only 9 residues: N

ter-KLKLKLKLK (L4K5). The conformational behavior ofthe 15-mers in pure water differs considerably one from another. Although both of them are initially dis-ordered in the 50–350 lM range, L10K5 gradually undergoes a disordered to a-helix transition formolecular concentrations above 100 lM. In all other solvents used, L10K5 adopts a stable a-helical con-formation. In methanol and methanol/Tris mixture, nonnative a-helices can be induced in both KL-alter-nating peptides, i.e., L7K8 and L4K5. However, in major cases and with a time delay depending on peptideconcentration, b-like structures can be gradually formed in both solutions. In PBS and methanol/PBSmixture, the tendency for L7K8 and L4K5 is to form structures belonging to b-family. A discussion hasbeen undertaken on the effect of counterions as well as their nature in the stabilization of ordered struc-tures in both KL-alternating peptides. # 2005Wiley Periodicals, Inc. Biopolymers 81: 8–19, 2006

This article was originally published online as an accepted preprint. The ‘‘Published Online’’ datecorresponds to the preprint version. You can request a copy of the preprint by emailing theBiopolymers editorial office at [email protected]

Keywords: lysine; leucine; cationic peptides; secondary conformation; optical spectroscopy

Correspondence to:M. Ghomi; e-mail: [email protected]

Biopolymers, Vol. 81, 8–19 (2006)

# 2005 Wiley Periodicals, Inc.

8

INTRODUCTION

We have recently evidenced the ability of a cationic

peptide, i.e., the 15-mer Nter-KLLKLLLKLLLKLLK,

to translocate a nonspecific phosphorothioate oligo-

deoxynucleotide (PS-ODN) into glioma cells with

low cytotoxicity.1 Spectroscopic investigations (CD

and Raman scattering) could emphasize that this pep-

tide adopts an �-helix structure in phosphate buffer

and in methanol. The capability of cationic vectors,2–5

particularly minimalist cationic peptides,6,7 to con-

dense DNA and to facilitate its in vitro cellular trans-

fection were shown in the framework of antisense ther-

apy. Among the minimalist peptides, those formed

only by lysine (K: cationic and hydrophilic monomer)

and leucine (L: neutral and hydrophobic monomer),

were found to be highly efficient.1,8–15 This is mainly

due to their ability to mimic the behavior of natural

peptides as regards their strong lytic activity and their

aptitude to perturb cell membranes.9–15 On the other

hand, it is now well established that the secondary

structure of peptides and proteins does influence their

interactions with membranes and model membranes.8–

12,15–18 Structural properties of a few KL-minimalist

peptides were subject to previous investigations by

means of CD and vibrational spectroscopy,8–12,19,20

with the consideration of chain length, molecular con-

centration, and ionic force.20 We can recall here the

results obtained from some (LK)n and (LKKL)n pepti-

des,20 those from the peptides belonging to LiKj (i ¼2j)1,8–10,12 and LnKnþ1

11 families. Among the LiKj

(i ¼ 2j) peptides, the 9-mer Nter-KLLLKLLLK (L6K3)

revealed a very low level of structuring, whereas the

15-mer Nter-KLLKLLLKLLLKLLK (L10K5) was

shown to be capable of forming an �-helical secondarystructure.8 Furthermore, it was shown that, because of

their amphipathic feature, the hemolytic activity of

LiKj (i ¼ 2j) peptides increases proportionally with

their chain length. The activity of L10K5 is even higher

than that of mellitin taken as a standard.10 As far as

LnKnþ1 peptides (L4K5, L5K6, L6K7, L7K8) are con-

cerned, a previous work based on FT-ir spectroscopic

measurements performed in bulk solid state (samples

obtained after evaporation of methanolic solutions) is

to be mentioned.11 It turned out that all of these pepti-

des adopt preferentially �-sheet or �-turn secondary

structures and that the amount of �-sheet conformation

increases with the chain length in going from 9- to 15-

mer. On the other hand, it has been shown that, in

aqueous solution, KL-alternating peptides tend to form

�-type family conformations, whereas –LKKL– pepti-

des adopt preferentially an �-helical structure.20 In

both cases, the appearance of structured species

depends on the peptide concentration and/or on the

amount of added salt. In diluted solutions without

added salt, all of these peptides are unstructured.

In some previous reports it has been mentioned that

CD spectra alone cannot elucidate unambiguously sec-

ondary conformation of peptides and proteins.21–24 A

nonnegligible uncertainty remains in structural interpre-

tations mainly in the case of �-family structures, due to

the presence of numerous subfamilies23 and to the fact

that CD spectroscopy may assign extended loops and

strands to unordered elements.21 Therefore, to circum-

vent eventual misinterpretations, CD and vibrational

studies can be combined together. The simultaneous

use of CD25,26 and vibrational markers27–32 leads to a

more precise conformational determination in solution.

As far as the solvents are concerned, we have mainly

considered those routinely used in cell transfection

experiments, i.e., Tris, PBS, methanol (but in a lesser

extent), as well as their bicomponent mixtures. We

have also considered the temporal evolution of the sec-

ondary conformations of peptides in different solvents.

This study can be considered as the first step prior to

the analysis of the interactions of cationic peptides with

antisense oligonucleotides.

MATERIALS AND METHODS

Three different peptides, containing either 15 or 9 amino

acids, have been synthesized with the following primary se-

quences: Nter-KLLKLLLKLLLKLLK, Nter-KLKLKL-

KLK, and Nter-KLKLKLKLKLKLKLK, hereafter abbre-

viated and referred to as L10K5, L4K5, and L7K8, respec-

tively (Table I). They were synthesized at the Institut

Pasteur (Paris) by solid phase methodology following the

Fmoc chemistry protocol.33 Chemical synthesis of L10K5

and L4K5 was made from a Fmoc-PAL-PEG-PS resin on a

Pioneer continuous-flow peptide synthesizer using 4-fold

molar excess of Fmoc amino acid derivatives with double

coupling cycles. L7K8 was synthesized on a Model 433A

batch synthesizer from Fmoc Amid Resin using single cou-

pling cycles with 10-fold molar excess of Fmoc amino acid

derivatives. HATU [0-(7-azabenzotriazol-1-yl)-1,1,3,3,tetra-

methyluronium hexafluorophosphate] and DIPEA (N,N,diiso-

propylethylamine) were used as coupling reagents. Both ap-

paratus as well as chemical reagents were obtained from

Applied Biosystems (Foster City, CA, USA). All peptides

were cleaved from the resin using trifluroacetic acid (TFA)

with 5% water. After resin filtration, the cleavage mixtures

were poured into cold diethyl ether. The fully deprotected

peptides were then recovered by filtration of the precipitate

(L4K5 and L7K8) or extraction with water (L10K5) and

lyophilized. Crude peptides were dissolved with a mixture

of 0.08% aqueous TFA/acetonitrile and directly purified by

reverse-phase MPLC using a Nucleoprep 20 �m C18 100 A

preparative column and a linear gradient (1%/min) of aceto-

nitrile in 0.08% aqueous TFA (pH 2) over 60 min at a

Optical Spectroscopy Assessment of LK Peptides 9

TableI

TemporalEvolutionoftheSecondaryConformationsasaFunctionoftheMolecularConcentrationandSolventNatureasSuggestedbyCDSpectra

Peptide(PrimarySequence)

Abbreviation

SolventNature

MolecularConcentration

InitialPeaksb

,c(nm)

Initial

Secondary

Structure

FinalPeaksc

(nm)

Final

Secondary

Structure

Nter -KLLKLLLKLLLKLLKa

L10K5

Pure

water

100–350�M

202(�

),222(�

w)

R208(�

)224(�

)�

Pure

water

50�M

200(�

),223(�

w)

R200(�

),223(�

w)

R

Allother

solventsused

100�M–5mM

207(�

),220(�

)�

207(�

),220(�

)�

Nter -KLKLKLKLKLKLKLK

L7K8

Pure

water

<350�M

197(�

),229(�

w)

R197(�

),229(�

w)

R

100%

methanol

100�M

208(�

),220(�

)�

199(þ

),219(�

)�

100%

methanol

5mM

207(�

),220(�

)�

195(þ

),219(�

)�

100%

PBS

100�M–5mM

199(þ

),219(�

)�

199(þ

),220(�

)�

50%

PBSþ

50%

methanol

100�M–5mM

195(þ

),219(�

)�

195(þ

),219(�

)�

50%

Trisþ

50%

methanol

100�M

205(þ

),220(�

)�

205(�

),219(�

)�

3mM–5mM

199(þ

),205(�

),217(�

)�þ

�199(þ

),219(�

)�

Nter -KLKLKLKLK

L4K5

Pure

water

<100�M

198(�

),228(�

w)

R197(�

),228(�

w)

R

100%

methanol

100�M

195(þ

),217(�

)�

196(þ

),217(�

)�

100%

methanol

5mM

207(�

),220(�

)�

199(þ

),219(�

)�

100%

PBS

100�M

��

100%

PBS

5mM

197(þ

),206(�

),224(�

)�þ

�þ

Rd

199(þ

),214(�

)�

50%

PBSþ

50%

methanol

100�M–5mM

196(þ

),219(�

)�

196(þ

),219(�

)�

50%

Trisþ

50%

methanol

100�M

199(�

),225(�

)R

198(�

),225(�

)R

3mM

206(�

),219(�

)�

198(þ

),219(�

)�þ

�5mM

196(þ

),219(�

)�

196(þ

),219(�

)�

aNter ,N-terminalofthepeptide;K,lysine;L,leucine;PBS,phosphatebuffered

saline;�,�-helix;�,�-likestructure;Rrandom

structure.

bThesign(þ

or�)

oftheobserved

CDbandisindicated

foreach

wavelength.

cInitialpeaksareobserved

inCDspectrajustaftersamplepreparation.Finalpeaksarethose

observed

afteratimedelay

necessary

fortheevolutionofthesecondarystructure

(see

text).

dSee

ResultsandDiscussion.

10 Hernandez et al.

18 mL/min flow rate. The purity of the peptides was

checked by HPLC on a nucleosil 5 �m C18 300 A analyti-

cal column, using a linear gradient of acetonitrile (1%/min)

in 0.08% aqueous TFA (pH 2) over 20 min at a 1 mL/min

flow rate. Purified peptides were quantified by amino acid

analysis and finally characterized by using positive ion

electrospray ionization mass spectrometry (ESþ). PBS sol-

utions as well as methanol and Tris were purchased from

PROLABO (Paris, France).

CD and Raman setups were described extensively in a

recent paper.34 Peptides were dissolved in i) pure water, ii)

methanol, iii) phosphate buffer containing 10 mM of Kþ

and Naþ cations (pH 6.8, for details see reference 35), iv)

PBS containing 10 mM phosphate buffer, 2.7 mM KCl,

137 mM NaCl (pH 7.4), v) Tris buffer containing 20 mM

Tris and 130 mM NaCl (pH 7.4), vi) a mixture of 50%

methanol þ 50% Tris or 50% PBS þ 50% methanol. CD

spectra were recorded in suprasil quartz cells with two dif-

ferent path lengths to demonstrate the effect of sample con-

centration either in the 50 to100 �M (2 mm path-length

cell) or in the 100 �M to 5 mM (0.01 mm path-length cell)

ranges. Spectra were analyzed at room temperature in

the 180- to 300-nm range. Each spectrum corresponds to

5 min of accumulation. The measured ellipticity for each

sample, [�]observed, was normalized to obtain the so-called

mean residue ellipticity, [�], by using the expression: [�]¼ [�]observed/10ncl, where n, c, and l are the number of resi-

dues in the peptide, the molar concentration, and the optical

path length of the sample, respectively.36 The normalized

ellipticity was expressed in deg cm2 dmol–1. The theoretical

estimates of [q]100% �-helix at 222 nm as calculated by the

expression given in reference 36, are �30,000 and �23,400

deg cm2 dmol–1, for a 15- and a 9-mer, respectively. These

values allow us to have a rough estimation of �-helix con-

tent in each sample when the general shape of its CD spec-

trum resembles that generally obtained for an �-helix.Raman spectra were recorded in the 5 to 8 mM concentra-

tion range. Samples were placed in a microcell of 13 �L inner

volume and excited with the 488 nm line of an argon laser.

Each spectrum was accumulated during 20 min at room tem-

perature and analyzed in the 1750–700 cm–1 region. Tempo-

ral evolution of CD and Raman spectra has been analyzed by

recording them as a function of time after sample preparation.

Especially at 5 mM in PBS, CD and Raman data were col-

lected simultaneously on the samples taken from the same

stock solution. To estimate the contribution of TFA anions to

Raman spectra of peptides, we have prepared reference solu-

tions as follows: solutions containing 1 mL of pure TFA

diluted in 9 mL of H2O (millipore) were neutralized by add-

ing NaOH (8 M) to reach pH 7; they then were lyophilized

and dissolved in water for preparing TFA anion solutions at

25 or 40 mM. These concentrations correspond well to those

existing in peptide solutions. Raman spectra were collected

from buffer, TFA, and peptides. Subtraction of buffer from

peptide spectrum was performed by normalizing each of them

to the water bending band at ca.1645 cm–1. In the same man-

ner, the contribution of water in TFA spectrum was corrected.

Finally, the TFA contribution was subtracted from the water-

corrected peptide spectrum after normalization to the intense

Raman band at 1436 cm–1 (TFA marker).

Postprocessing (subtraction of buffer contribution, base-

line correction, smoothing, and curve fitting) of all spectral

data were performed using GRAMS/32 software (Galactic

Industries). To get some more information about the secon-

dary structure of these peptides, curve fitting of their amide

I Raman profile observed in the 1750–1600 cm–1 range was

attempted. This procedure has been previously applied to

vibrational spectra to probe the secondary conformations

existing in the whole structure of proteins.16,37,38 To

decompose Raman bands, mixed Gaussian þ Lorentzian

functions, with Lorentzian contribution kept equal to or

greater than 50%, were employed. Initial guesses for their

maximum wavenumbers were based on the second derivative

analysis in the amide I region. Figures shown in this paper

have been drawn using the SIGMAPLOT (Systat Software

Inc., Point Richmond, CA) package.

RESULTS AND DISCUSSION

CD Spectra

L10K5, the 15-mer with (KLLL) Repeat Unit. A

selection of CD spectra of this peptide at the highest

concentration (5 mM) in PBS and methanol is dis-

played in Figure 1. The presence of �-helix as the

major conformer in solution is proven by two nega-

tive bands at 207 and 220 nm. For lower concentra-

tions (50–350 �M), the behavior of this peptide in

pure water (data not shown, see Table I) is worth

mentioning here: it remains unstructured at 50 �M(lowest concentration); showing initially a predomi-

nant random structure in agreement with previous

FIGURE 1 Temporal evolution of the room temperature

CD spectra of the 15-mer Nter-KLLKLLLKLLLKLLK (L10K5)

at 5 mM, recorded in PBS and methanol.

Optical Spectroscopy Assessment of LK Peptides 11

results,20 this peptide undergoes a gradual transition

toward an �-helical structure (complete transition

observed after �3 h). In all other cases (solvent, con-

centration), L10K5 adopts a stable �-helical structure.

L7K8 and L4K5, the 15- and 9-mer with (KL) RepeatUnit. In contrast to L10K5, these two peptides present

a random structure in pure water, remaining unchanged

with time in the 50 to 350 �M concentration range

(data not shown, see Table I).

Figures 2 (15-mer) and 3 (9-mer) display the CD

spectra recorded in methanol at 100 �M and 5 mM.

At both concentrations, the structural shape of the ini-

tial CD spectrum of 15-mer (Figure 2) reveals the

presence of �-helical conformation with a gradual

temporal transition toward �-form family (as evi-

denced by a unique negative band at ca. 219 nm)

within 120 min at 100 �M (Figure 2a) and a shorter

time (90 min) at 5 mM (Figure 2b). The 9-mer L4K5

shows a CD spectrum belonging to �-form family at

low concentration (100 �M, Figure 3a). At 5 mM

Figure 3b), the initial spectrum, confirming the pres-

ence of �-helical conformation, undergoes within

FIGURE 2 Temporal evolution of the room temperature

CD spectra of the 15-mer Nter-KLKLKLKLKLKLKLK (L7K8)

recorded in methanol. (a) 100 �M, (b) 5 mM.

FIGURE 3 Temporal evolution of the room temperature

CD spectra of the 9-mer Nter-KLKLKLKLK (L4K5) recorded

in methanol. (a) 100 �M, (b) 5 mM.

12 Hernandez et al.

180 min a substantial change in the spectral shape, in

accordance with an �-helix toward �-structure con-

version.

In PBS, the 15-mer L7K8, gives rise to CD spectra

characteristic of stable �-structures, whatever the

molecular concentration, as does the 9-mer L4K5 at

100 �M in PBS (data not shown, Table I). At 5 mM,

the 9-mer shows a slow conformational transition

toward �-structures with a complete achievement

after 240 min (Figure 4). It should be mentioned

that the shape of the final CD spectrum (taken after

240 min, Figure 4) resembles that previously reported

and assigned to �-hairpins (characterized by an in-

tense positive band at 199 nm).39

Considering the structuring effect of methanol on

both KL peptides (see above), we have studied their

structural behavior in two bicomponent mixtures,

such as methanol/PBS and methanol/Tris. Irrespec-

tive of the peptide concentration, L4K5 and L7K8

present stable �-forms in methanol/PBS (data not

shown, Table I). However, some surprising effects

have been observed in methanol/Tris, as shown in

Figure 5. CD spectrum of the 15-mer L7K8, reveals

the existence of a stable �-helical structure at 100 �M(Figure 5a). At 3 and 5 mM (Figures 5b and 5c,

respectively), complex CD spectra are obtained, lead-

ing us to conclude that several secondary conforma-

tions are present in solution. After 1 h these spectra

show a clear conformational transition toward �-struc-tures. The behavior of the 9-mer L4K5, is completely

different in this bicomponent solvent (Figure 5). At

the lowest concentration (100 �M) this peptide is un-

structured (Figure 5d), as characterized by a negative

band at 198 nm. Structured conformers appear at higher

concentrations. At 3 mM, CD spectra (Figure 5e) of

the 9-mer prove initially the existence of �-helix struc-

ture, followed by a very slow transition (within 24 h)

toward �-structures. Conversely, at 5 mM, CD spec-

tra (Figure 5f) are consistent with the formation of

stable �-structures.

Raman Spectra

Despite the clear structuring effect of methanol

described in the above section devoted to CD spectro-

scopy, we do not present here the Raman spectra

obtained in methanol, because of the difficulties

encountered to control the alcohol content in the sam-

ples used for Raman measurements. In fact, we have

noticed in the course of the accumulation procedure a

gradual evaporation of methanol in the capillaries

used as the Raman sample holder. Here, we only

comment on the main Raman spectral features of

the three peptides at 5 mM concentration recorded in

phosphate buffer (L10K5) and in PBS (L4K5 and

L7K8).

Correction of Raman Spectra for TFA Anions. Un-

der Materials and Methods, we have described the

manner in which to correct Raman spectra of peptides

for the contribution of TFA anions. We illustrate in

Figure 6 the treatment procedure through a typical

example (L10K5). As shown, TFA anions barely con-

tribute to the Raman intensity in the 1700–1600 cm–1

spectral region (amide I region, see below). Normal-

ization to the most intense TFA band at 1436 cm–1

affords the intensity of all other TFA modes located

at 1204, 843, and 726 cm–1, thus leading to a perfect

subtraction.

Analysis of the Amide I Region in Raman Spectraof Peptides. The amide I characteristic vibrations

(mentioned above) yield appreciable information on

the secondary structure of peptides.27–32 Decomposi-

tion of the amide I region for the three peptides is dis-

played in Figure 7. Table II reports the wavenumber,

band width at half height (BWHH), and normalized

area (in %) of each decomposed component. The

main features of these spectra can be described as fol-

lows:

� In agreement with the CD spectra (Figure 1,

Table I), the �-helix conformation of the 15-

mer L10K5 was confirmed by the presence of an

intense and narrow amide I Raman band located

FIGURE 4 Temporal evolution of the room temperature

CD spectra of the 9-mer Nter-KLKLKLKLK (L4K5) at 5 mM,

recorded in PBS.

Optical Spectroscopy Assessment of LK Peptides 13

FIGURE 5 Temporal evolution of the room temperature CD spectra of the 15-mer Nter-

KLKLKLKLKLKLKLK (L7K8) and the 9-mer Nter-KLKLKLKLK (L4K5), recorded in 50% Tris þ50% methanol. (a) L7K8 at 100 �M, (b) L7K8 at 3 mM, (c) L7K8 at 5 mM. (e) L4K5 at 100 �M,

(b) L4K5 at 3 mM, (c) L4K5 at 5 mM.

14 Hernandez et al.

at 1653 cm–1 (Figure 7a). Moreover, the obser-

vation of a medium amide III band at 1293 cm–1

and two intense Raman markers at 1345 and

935 cm–1, all known as �-helix markers,29,38,40–

43 reinforce the conformational assignment of

this peptide (Figure 6).

� As far as the conformation of L7K8 is con-

cerned, the existence of the �-marker Raman

band at 1666 cm-1 Figure 7b) as the major com-

ponent observed in this region and assigned to

H-bonded �-strands16,44 is to be emphasized.

We also mention the existence of an amide III

�-marker16,43 at 1241 cm–1 (data not shown).

� To confirm the temporal evolution of L4K5 ob-

served in CD spectra in PBS at 5 mM (see above,

Figure 4), we show in Figure 7 three Raman

spectra in the amide I region, recorded 30 min

(Figure 7c), 3 h (Figure 7d), and 5 h (Figure 7e)

after sample preparation. Changes in the secon-

dary conformations can be first confirmed by the

global spectral shape. Several secondary confor-

mations belonging to �-form structures can be

evidenced by band decomposition: �-strand res-

idues implying the presence of interstrand hy-

drogen bonds (marker at ca.1665 cm–1),16,44 free

�-strands and �-turns (marker at ca. 1676 cm–1,

see reference 16 and references therein), and

antiparallel �-sheet (a doublet in the frequency

intervals at ca. 1690–1680 cm–1 and at ca. 1640–

1615 cm–1).21,45 The persistence of an unordered

chain marker at ca. 1648 cm–1, even at times lon-

ger than 3 h, should be emphasized.46,47 Another

interesting effect is the presence of the character-

istic �-helix marker at 1653 cm–1 in the Raman

spectrum recorded after 30 min, becoming negli-

gible after 3 h and vanishing completely after 5 h.

In other words, �-helical conformers existing ini-

tially as a minor population in this peptide, are

transformed to �-type conformers with time (see

also CD spectra, Figure 4).

CONCLUSION

Beyond the consideration of chain length and molec-

ular concentration, we have emphasized the temporal

evolution of secondary conformations in solution.

However, it was found that all of these parameters

(environment, chain length, and time) are undoubt-

edly interdependent. From the present investigations,

it can be concluded that

� The primary sequence of the 15-mer L10K5,

responsible for its amphipathic character, and

its length should be considered as the main ele-

ments favoring intramolecular H-bonds, neces-

sary for an �-helix stabilization. In contrast, the

formation of �-sheets needs intermolecular H-

bonds. This situation is particularly favored in

the other two peptides (L7K8 and L4K5).

� The nature of the solvent can undoubtedly alter

the balance between the intra- and intermolecu-

lar H-bonds (see reference 48 and references

cited therein). For example, in methanol, unsta-

ble helical conformations can be induced ini-

tially in L7K8 and L4K5, whereas L4K5 at low

concentrations (below 100 �M) shows a natural

tendency to adopt a �-structure. However the

initially induced �-helical conformation shows

a gradual transition toward �-type family struc-

tures. A similar effect has been described in tri-

fluoroethanol (TFE) for �-sheet proteins.49 Theeffect of solvents with a low dielectric constant

such as TFE, methanol, or ethanol, capable of

inducing nonnative �-helical conformation in

proteins and peptides, has been extensively

studied.49,50 Most of these investigations have

been performed in a TFE/H2O mixture, because

of its ability to mimic the physiological environ-

ment.51 It has been emphasized that all these

alcohols might induce a common mechanism of

activity. Particularly, a direct mechanism in

which TFE molecules bind to residues in helical

conformation,52,53 or an indirect mechanism

mainly based on the effect of low dielectric

FIGURE 6 Top: (continuous trace) Raman spectrum of

the 15-mer Nter-KLLKLLLKLLLKLLK (L10K5) with TFA

anions, recorded in a phosphate buffer at room temperature.

(dotted trace) Raman spectrum of TFA at 25 mM (concen-

tration of TFA anions in the peptide solution). Bottom:

Raman spectrum of L10K5 after subtraction of TFA contri-

bution.

Optical Spectroscopy Assessment of LK Peptides 15

media on the intermolecular/intramolecular

hydrogen bonding balance, have both been pro-

posed to interpret the experimental observa-

tions. Cammers-Goodwin and coworkers54 have

suggested that lowering the dielectric constant

penalizes conformations with solvent-exposed

CO and NH groups, thus reinforcing a possible

�-helix formation. Furthermore, it has been

concluded that, generally in the presence of

TFE (reference 55 and other references cited

therein), most alcohols,50 and other organic sol-

vents,56 the hydrogen bonding between the

amide protons and water molecules decreases,

thus leading to a strengthening of intramolecular

hydrogen bonding, which stabilizes helices.57

This phenomenon has been observed even in

the case of charged peptides (containing lys and

glu residues).55 Recent NMR-based molecular

dynamics (MD) studies suggest that TFE or

alcohol molecules lead to the formation of clus-

ters around peptides in solution, reducing the

number of water molecules in direct interaction

with peptides.58,59 However, one can also

expect that random structures capable of mini-

mizing the electrostatic repulsion between

charged side chains are more suitable compared

to helical conformations, keeping these charged

moieties closer to each other. But, on the basis

of our results, it appears that, in L7K8 and in

L4K5 in methanol (at high molecular concentra-

tions), the increase of ‘‘attractive forces’’

between NH. . .CO in helices might be able to

compensate the side chain repulsive forces

between lysine groups. This assumption is in

agreement with the results obtained in the case

of poly(lys) and poly(glu).55 We should con-

sider that a minimal chain length seems to be

necessary for the LnKnþ1 peptides at low con-

centrations to adopt an �-helical conformation

in methanol.

FIGURE 7 Decomposition of the Raman spectra observed in the amide I spectral region at room

temperature. (a) Raman spectrum of the 15-mer Nter-KLLKLLLKLLLKLLK (L10K5), recorded in

phosphate buffer. (b) Raman spectrum of the 15-mer Nter-KLKLKLKLKLKLKLK (L7K8), recorded in

PBS. Traces d, e, and f refer to Raman spectra of the 9-mer Nter-KLKLKLKLK (L4K5) recorded in

PBS, 30 min, 3 h, and 5 h after sample preparation, respectively. See also Table II.

16 Hernandez et al.

TableII

DecompositionoftheRamanSpectraPerformedin

theAmideISpectralRegion

15-m

er(L

10K5)

15-m

er(L

7K8)in

PBSa

9-m

er(L

4K5)at5mM

inPBS(30min)a

9-m

er(L

4K5)at5mM

inPBS(3

h)

9-m

er(L

4K5)at5mM

inPBS(5

h)

ProposedSecondary

Structure

Wavenumber

(cm

�1)

Area

(%)

BWHH

(cm

�1)

Wave-

number

(cm

�1)

Area

(%)

BWHH

(cm

�1)

Wave-

number

(cm

�1)

Area

(%)

BWHH

(cm

�1)

Wave-

number

(cm

�1)

Area

(%)

BWHH

(cm

�1)

Wave-

number

(cm

�1)

Area

(%)

BWHH

(cm

�1)

1653

100

23

1681

18

1686

913

1689

12

14

1687

12

13

Antiparaleel�-sheetc

1670

67

1678

40

22

1676

41

22

1674

46

23

�-Turn

and/ornotH-bonded

1666

84

13

1665

23

21

1662

25

20

�-Stand,random

1660

16

17

H-bonded

�-strands(�-sheet)

1653

16

1652

418

�-helix

1645

26

1645

819

1648

10

15

Random

1640

711

1635

51637

915

�-sheetd

1625

720

1628

415

1618

523

1619

720

1612

412

aPBS,phosphatebuffered

saline;BWHH,bandwidth

athalfheight.

bThetimedelay

afterthesamplepreparationforrecordingRam

anspectraisshownin

parentheses.

cHighwavenumber

componentofthedoubletcorrespondingto

�-sheet(see

ResultsandDiscussion).

dLowwavenumber

componentofthedoubletcorrespondingto

�-sheet(see

ResultsandDiscussion).

Optical Spectroscopy Assessment of LK Peptides 17

� The increase in peptide concentration, espe-

cially in the case of a short peptide such as

L4K5 dissolved in methanol or in methanol/Tris,

favors the formation of �-helix, by an entropic

effect arising from the helix–helix interactions.

Previous results were consistent with the forma-

tion of short helices through multihelix bundle

effect.60 These nonnative �-helices are not sta-

ble and undergo a conformational transition

toward �-type structures with a temporal evolu-

tion faster in pure methanol.

� Solutions containing methanol þ Tris (and con-

sequently NaCl) bring Cl– anions in the vicinity

of peptides and the shielding effect of these

counterions reduces the repulsive interactions of

adjacent lysines, thus explaining a higher stabil-

ity of helical conformation. Moreover, a Tris

molecule (containing 3 OH and 1 NH2 groups)

can also contribute to the formation of H-bonds

with the peptides and consequently stabilize the

helical structures. In contrast, in PBS as well as

in methanol/PBS, �-type structures are favored

in L4K5 and L7K8. Indeed, ‘‘bulky’’ phosphate

anions might have the capability of forming

bridges between positively charged side chains,

favoring strand–strand interactions.

F.-Z. Boukhalfa-Heniche thanks the Fondation pour la

Recherche Medicale (FRM) and the Association pour la

Recherche contre le Cancer (ARC) for doctoral fellowships.

B. Hernandez acknowledges the Spanish Ministry of Edu-

cation, Culture, and Sport for a postdoctoral fellowship

EX2001 12751081. The authors thank Catherine Herve du

Penhoat and Vladimir Baumruk for their careful reading of

the manuscript and helpful suggestions.

REFERENCES

1. Boukhalfa-Heniche, F. Z.; Hernandez, B.; Gaillard, S.;

Coıc, Y. M.; Huynh-Dinh, T.; Lecouvey, M.; Seksek,

O.; Ghomi, M. Biopolymers 2004, 73, 727–734.

2. Wyman, T. B.; Nicol, F.; Zelphati, O.; Scaria, P. V.;

Plank, C.; Szoka F.C., Jr. Biochemistry 1997, 36,

3008–3017.

3. Li, W.; Nicol, F.; Szoka F.C., Jr.; Adv Drug Deliv Rev

2004, 56, 967–985.

4. Morris, M. C.; Chaloin, L.; Mery, J.; Heitz, F.; Divita,

G. Nucleic Acids Res 1999, 27, 3510–3517.

5. Derossi, D.; Chassaing, G.; Prochiantz, A. Trends Cell

Biol 1998, 8, 84–87.

6. Niidome, T.; Takaji, K.; Urakawa, M.; Ohmori, N.;

Wada, A.; Hirayama, T.; Aoyagi, H.; Bioconjug Chem

1999, 10, 773–780.

7. Kubo, T.; Yokoyama, K.; Ueki, R.; Abe, S.; Goto, K.;

Niidome, T.; Aoyagi, H.; Iwakuma, K.; Ando, S.;

Ono, S.; Fujii, M. Nucleic Acids Symp Ser 2000, 44,

49–50.

8. Dufourcq, J.; Neri, W.; Henry-Toulme, N. FEBS Lett

1998, 421, 7–11.

9. Castano, S.; Desbat, B.; Laguerre, M.; Dufourcq, J.

Biochim Biophys Acta 1999, 1416, 176–194.

10. Castano, S.; Cornut, I.; Buttner, K.; Dasseux, J. L.;

Dufourcq, J. Biochim Biophys Acta 1999, 1416, 161–175.

11. Castano, S.; Desbat, B.; Dufourcq, J. Biochim Biophys

Acta 2000, 1463, 65–80.

12. Beven, L.; Castano, S.; Dufourcq, J.; Wieslander, A.;

Wroblewski, H. Eur J Biochem 2003, 270, 2207–2217.

13. Albiol Matanic, V. C.; Castilla, V. Int J Antimicrob

Agents 2004, 23, 382–389.

14. Kagan, B. L.; Selsted, M. E.; Ganz, T.; Lehrer, R. I.

Proc Natl Acad Sci U S A 1990, 87, 210–214.

15. Blondelle, S. E.; Houghten, R. A. Biochemistry 1992,

31, 12688–12694.

16. Maiti, N. C.; Apetri, M. M.; Zagorski, M. G.; Carey, P.

R.; Anderson, V. E. J Am Chem Soc 2004, 126, 2399–

2408.

17. Maget-Dana, R.; Lelievre, D. Biopolymers 2001, 59,

1–10.

18. Kiyota, T.; Lee, S.; Sugihara G.Biochemistry 1996, 35,

13196–13204.

19. Cornut, I.; Buttner, K.; Dasseux, J. L.; Dufourcq, J.

FEBS Lett 1994, 349, 29–33.

20. Baumruk, V.; Huo, D.; Dukor, R. K.; Keiderling, T. A.;

Lelievre, D.; Brack, A.; Biopolymers 1994, 34, 1115–

1121.

21. Dong, A.; Kendrich, B.; Kreilgard, L.; Matsuura, J.;

Manning, M. C.; Carpenter, J. F. Arch Biochem Bio-

phys 1997, 347, 213–220.

22. Glattli, A.; Daura, X.; Seebach, D.; van Gunsteren, W.

F. J Am Chem Soc 2002, 124, 12972–12978.

23. Pelton, J. T.; McLean, L. R. Anal Biochem 2000, 277,

167–176.

24. Yada, R. Y.; Jackman, R. L.; Nakai, S. Int J Peptide

Protein Res 1988, 98–108.

25. Greenfield, N. J. Anal Biochem 1996, 235, 1–10.

26. Kelly, S. M.; Price, N.C. Biochim Biophys Acta 1997,

1338, 161–185.

27. Jackson, M.; Mantsch, H. H. Crit Rev Biochem Mol

Biol 1995, 30, 95–120.

28. Krimm, S.; Bandekar, J. Adv Protein Chem 1986, 38,

181–364.

29. Miura, T.; Thomas, G. J., Jr. Subcell Biochem 1995,

24, 55–99.

30. Peticolas, W. L. Methods Enzymol 1995, 226, 389–

416.

31. Sane, S. U.; Cramer, S. M.; Przybycien, T. M. Anal

Biochem 1999, 269, 255–272.

32. Marx, J.; Berjot, M.; Alix, A. J. P. J Raman Spectrosc

1987, 18, 289–300.

33. Chan, W. C.; White, P. D. In Fmoc Solid Phase Peptide

Synthesis: A Practical Approach. Oxford University

Press: Oxford, UK, 2000.

34. Hernandez, B.; Baumruk, V.; Gouyette, C.; Ghomi, M.

Biopolymers 2005, 78, 21–34.

18 Hernandez et al.

35. Baumruk, V.; Gouyette, C.; Huynh-Dinh, T.; Sun, J. S.;

Ghomi, M. Nucleic Acids Res 2001, 29, 4089–4096.

36. Lazo, N. D.; Downing, D. T. J Peptide Res 2001, 58,

457–463.

37. Munishkina, L. A.; Phelan, C.; Uversky, V. N.; Fink,

A. L. Biochemistry 2003, 42, 2720–2730.

38. Debelle, L.; Alix, A. J. P.; Wei, S. M.; Jacob, M. P.;

Huvenne, J. P.; Berjot, M.; Legrand, P. Eur J Biochem

1998, 258, 533–539.

39. Hilario, J.; Kubelka, J.; Keiderling, T. A. J Am Chem

Soc 2003, 125, 7562–7574.

40. Overman, S. A.; Thomas, G. J., Jr.; Biochemistry 1998,

37, 5654–5665.

41. Tsuboi, M.; Suzuki, M.; Overman, S. A.; Thomas, G.

J., Jr. Biochemistry 2000, 39, 2677–2684.

42. Overman, S. A.; Thomas, G. J., Jr. Biochemistry 1999,

38, 4018–4027.

43. Williams, R. W.; Dunker, A. K. J Mol Biol 1981, 152,

783–813.

44. Shao, Z.; Vollrath, F.; Sirichaisit, J.; Young, R. J. Poly-

mer 1999, 40, 2493–2500.

45. Kubelka, J.; Keiderling, T. A. J Am Chem Soc 2001,

12, 12048–12058.

46. Dong, J.; Wan, Z.; Popov, M.; Carey, P. R.; Weiss, M.

A. J Mol Biol 2003, 330, 431–442.

47. Thomas, P. J.; Ko, Y. H.; Pedersen P. L. FEBS Lett

1992, 312, 7–9.

48. Awasthi, S. K.; Shankaramma, S. C.; Raghothama, S.;

Balaram, P. Biopolymers 2001, 58, 465–476.

49. Dong, A.; Matsuura, J.; Manning M. C.; Carpenter J. F.

Arch Biochem Biophys 1998, 355, 275–281.

50. Arunkumar, A. I.; Kumar, T. K. S.; Yu, C. Int J Biol

Macromol 1997, 21, 223–230.

51. Buck, M. Q Rev Biophys 1998, 31, 297–355.

52. Jasanoff, A.; Fesht A. R. Biochemistry 1994, 33, 2129–

2135.

53. Storrs, R. W.; Truckses, D.; Wemmer, D. E. Biopoly-

mers 1992, 32, 1695–1702.

54. Cammers-Goodwin, A.; Allen, T. J.; Oslik, S. L.;

McClure, K. F.; Lee, J. H.; Kemp, D. S. J Am Chem

Soc 1996, 118, 3082–3090.

55. Luo, P.; Baldwin, R. L. Proc Natl Acad Sci U S A.,

1992, 96, 4930–4935.

56. Arunkumar, A. I.; Kumar, T. K. S.; Sivaraman, T.; Yu,

C. Int J Biol Macromol 1997, 21, 299–305.

57. Nelson, J. W.; Kallenbach, N. R. Proteins Struct Funct

Genet 1986, 1, 211–217.

58. Roccatano, D.; Colombo, G.; Fiorini, M.; Mark, A. E.

Proc Natl Acad Sci U S A 2002, 99, 12179–12184.

59. Fiorini, M.; Diaz, M. D.; Buger, K.; Berger, S. J Am

Chem Soc 2002, 124, 7737–7744.

60. Eisenberg, D.; Wilcox, W.; Eshita, S. M.; Pryciak, P.

M.; Ho, S. P.; DeGrado, W. F. Proteins 1986, 1, 16–22.

Reviewing Editor: Nancy Stellwagen

Optical Spectroscopy Assessment of LK Peptides 19