ArticlePDF Available

C-9/12 Ribbon-Like Structures in Hybrid Peptides Alternating - and Thiazole-Based -Amino Acids

Wiley
Chemistry - A European Journal
Authors:

Abstract

According to their restricted conformational freedom, heterocyclic γ-amino acids are usually considered related to Z-vinylogous γ-amino acids. In such a context, oligomers alternating α- and thiazole-based γ-amino acids, named ATCs were expected to fold into a canonical 12-helical shape as described for α/γ-hybrid peptides composed of cis α/β-unsaturated γ-amino acids. However, we herein demonstrate by combining X-ray crystallography, NMR and FT-IR experiments and DFT calculations that the folding behavior of ATC-containing hybrid peptides is much more complex. While the homochiral α/(S)-ATC sequences were not able to adopt a stable conformation, heterochiral α/(R)-ATC peptides displayed original ribbon structures stabilized by unusual C9, C12-bifurcated hydrogen bonds. These ribbon structures that could be considered as a succession of pre-organized γ/α dipeptides may provide the basis for designing original α-helix mimics.
Supported by
A Journal of
Accepted Article
Title: C9/12-ribbon-like structure in hybrid peptides alternating α- and
thiazole-based γ-amino acids
Authors: Clément Bonnel, Baptiste Legrand, Matthieu Simon,
Jean Martinez, Jean-Louis Bantignies, Young Kee Kang,
Emmanuel Wenger, François Hoh, Nicolas Masurier, and
Ludovic Thierry Maillard
This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.
To be cited as: Chem. Eur. J. 10.1002/chem.201704001
Link to VoR: http://dx.doi.org/10.1002/chem.201704001
COMMUNICATION
C9/12-ribbon-like structure in hybrid peptides alternating α- and
thiazole-based γ-amino acids
Clément Bonnel,[a] Baptiste Legrand,[a] Matthieu Simon,[a] Jean Martinez,[a] Jean-Louis Bantignies,[b]
Young Kee Kang,[c] Emmanuel Wenger,[d] Francois Hoh,[e] Nicolas Masurier,[a] Ludovic T. Maillard*[a]
Abstract: According to their restricted conformational freedom,
heterocyclic γ-amino acids are usually considered related to Z-
vinylogous γ-amino acids. In such a context, oligomers alternating α-
and thiazole-based γ-amino acids, named ATCs were expected to
fold into a canonical 12-helical shape as described for α/γ-hybrid
peptides composed of cis α/β-unsaturated γ-amino acids. However,
we herein demonstrate by combining X-ray crystallography, NMR
and FT-IR experiments and DFT calculations that the folding
behavior of ATC-containing hybrid peptides is much more complex.
While the homochiral α/(S)-ATC sequences were not able to adopt a
stable conformation, heterochiral α/(R)-ATC peptides displayed
original ribbon structures stabilized by unusual C9, C12-bifurcated
hydrogen bonds. These ribbon structures that could be considered
as a succession of pre-organized γ/α dipeptides may provide the
basis for designing original α-helix mimics.
Introduction
Substantial progress has been made over the last decades
in establishing abiotic architectures, named foldamers,
displaying helices, sheets and ribbon shapes.[1] Initial works
have focused on pseudopeptide oligomers[2] comprising of a
single type of monomer subunits that mainly consist in β-, γ-,
and δ-amino acids (selected examples:[3]) and on backbones
closely related to peptides such as peptoids or oligourea
(selected examples:[4]). In addition to homogeneous oligomers,
hybrid skeletons composed of different types of monomer units,
typically a combination of α, β or γ amino acids permit to explore
unusual folding patterns offering distinct ways to project side
chains in the three dimensional space.[5] Following the
development of α/β-hybrid peptides[2d, 3b, 6] and stimulated by a
systematic conformational analysis,[7] the 12-[8] and the mixed
12/10 helix[9] have been recognized as the most stable
conformations for (α/γ)n peptides.[10] More lately, Gopi et al.
described that α/γ-hybrid peptides composed of Aib and cis α/β-
Scheme 1. Scheme Caption. Chart 1. Nomenclature used for ATCs and
torsion angles description. Dipeptides 1 and 2, and oligomers 3-9.
unsaturated γ-amino acids also accommodate a 12-helical
shape. Surprisingly, the cis double bonds and the carbonyl
groups in the Z-vinylogous residues are not π-conjugated and
largely deviate from the planarity (Table 1).[11]
We have recently described a class of constrained heterocyclic
γ-amino acids named ATCs built around a thiazole ring that were
used to template C9-helical γ-peptide foldamers.[12] Because the
ζ torsion angle is locked around 0°, ATCs are related to Z-
vinylogous γ-amino acids (Chart 1).[13] In line with these studies,
we herein explored the structural feature of sequences
alternating α-amino acids and (S)- or (R)-ATCs using XRD,
NMR and FT-IR spectroscopies, and DFT calculations. Our
results demonstrate that, compared to other γ-amino acids
including constrained vinylogous building-blocks, ATC residues
show unique conformational properties resulting from its
aromatic ring.
Results and discussion
We first explored the intrinsic conformational behavior of the
(γ/α) dipeptide subunit. The two diastereomers Z(2-Br)-(S/R)-
ATC-Phe-NH-iPr 1 and 2 differing by the absolute
stereochemistry of the γ-amino acid (Chart 1) were prepared
NS
R2
O
R1
N
H
O
R1
N
Hα
γβ
φζψθ α
γβ
NS
O
N
H
N
HO
Ph
NS
O
HN H
N
H
N
O
O
O
Br
Ph
n
(S-)Phe-(S-)ATC
3: n = 2
*
NS
O
N
H
N
HO
NS
O
HN H
N
O
n
Ph
*
Ph
O
O
O
(R-)ATC-(S-)Phe
7: n = 1
8: n = 2
9: n = 3
(αγ)n alternating sequences
(S-)Phe-(R-)ATC
4: n = 1
5: n = 2
6: n = 3
(γα)n alternating sequences
N
HO
H
N
N
H
O
NS
O
O
Br
Ph
*
1: (S-)ATC
2: (R-)ATC
ATC building-block
Z-vinylogous
γ-amino acid
[a] Dr. C. Bonnel, Dr. B. Legrand, Dr. M. Simon, Prof. J. Martinez, Prof.
N. Masurier, Dr. L. T. Maillard
Institut des Biomolé cules Max Mousseron, UMR CNRS-UM-ENSCM
5247, UFR des Sciences Pharmaceutiques et Biologiques
15 Avenue Charles Flahault, 34093 Montpellier Cedex 5, France
E-mail: ludovic.maill ard@umontpel lier.fr
[b] Prof. J.L. Bantignies
LC2 - UMR 5221 CNRS-UM, Montpellier, France
[c] Prof. Y. K. Kang
Department of Chemistry, Chungbuk National University, Cheongju,
Chungbuk 28644, Republic of Korea
[d] E. Wenger
Laboratoire de Cristallographie, Résonance Magnétique et
Modélisation, Université de Lorraine, CNRS, UMR 7036, Nancy,
FRANCE
[e] Dr. F. Hoh
Centre de Biochimie Structurale, CNRS UMR 5048-INSERM 1054
University of Montpellier, Montpellier, France
Supportin g information for this article is given via a link at the end of
the document.
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
Figure 1. A/ Superimposition of the 15 lowest energy NMR structures of 1 (blue) and 2 (green) in CDCl3. The OBn and NH-iPr moieties were omitted for clarity.
RMSD are 0.448 and 0.208 Å respectively. B/ Crystal structures of 1 (blue) and 2 (green). C/ C9/7-, C9- and C9/12-hydrogen bonds accessible to 1 and 2. D/
Variation of the amide proton NMR resonances of 1 and 2 upon the addition of 40% CD3OH in CDCl3 and E/ FT-IR spectra (Black) and deconvolution (Red) of
oligomers 1 and 2 in the amide I region (3 mM in CHCl3).
according to formerly reported procedure[14]. Their structures
were solved by NMR in CDCl3 and CD3OH (Figures 1A, S14 and
S16) and X-ray crystallography (Figures 1B). 1 and 2 displayed
turn-conformations stabilized by a C9-hydrogen bond between
the urethane C=O and the HN-Phe surrounding the ATC
residues. The γ−residues shared similar torsion angle values at
the solution and solid states, in accordance with those
previously reported for ATC oligomers (Table 1). It is noteworthy
that the ATC ψ dihedral angles significantly differed from those
reported in Z-vinylogous-γ-amino acid-containing hybrid peptides
by Gopi et al.[11] Nevertheless, the C-terminal moieties remained
ambiguous. The sets of NOEs were compatible with the
formation of subsequent C7 or C12 hydrogen bonds associated to
a single rotation of the ϕ torsion angle of the α-residue. At the
solid-state, the α-amino acids exhibited extended conformations
(Figure 1B and Table S42) since the ATC C=O and C-terminus
N-H groups participated in the cohesion of the crystals via strong
intermolecular H-bonds. Thus, such crystal packings could
prevent the establishment of intramolecular C7- and C12-
Hydrogen bonds observed in solution. Consequently, based on
NMR and XRD data, 1 and 2 could adopt three different
conformations, named C9/7-, C9/12- and C9-models depending on
the hydrogen-bonding pattern (Figure 1C).
We compared their relative stabilities by calculating the free
energies using DFT methods at the ωB97X-D/def2-TZVP//SMD
M06-2X/6-31G(d) level of theory in chloroform (Table S47). The
strength of hydrogen bonds in the optimized conformers were
then evaluated in terms of the second perturbation energy (ΔE2)
of the lone pair orbitals of the carbonyl oxygen with the
corresponding NH antibonding orbital using natural bond orbital
(NBO) analysis[15] at the ωB97X-D/def2-TZVP level of theory
(see Supporting information). We found that the C9/12-conformer
folded around a highly strong C9-hydrogen bond (ΔE2 = 10.70
kcal mol1) was the most preferred for 1 followed by the C9-
conformer (ΔG = 0.89 kcal mol1 and ΔE2 = 13.56 kcal mol1).
The C9/7-folding was much less favored with ΔG = 3.78 kcal
mol1. Conversely, 2 displayed a different conformational
preference with the most preferred conformers being the C9/12
Table 1. Average
φ
, θ and ψ dihedral angles (°) for the γ-amino acids in ATC-
and Z-vinylogous γ-amino acid-containing sequences.
Cpd
φ
θ
ζ
ψ
(S,S)-1
XRD
-69 ± 3
123 ± 2
10 ± 1
-44 ± 1
NMR (CDCl3)
-85 ± 2
115 ± 7
-8 ± 3
-24 ± 5
NMR (CD3OH)
-103 ± 25
110 ± 25
-6 ± 2
-19 ± 7
(R,S)-2
XRD
83
-115
0
29
NMR (CDCl3)
71 ± 6
-118 ± 2
7 ± 2
24 ± 2
NMR (CD3OH)
64 ± 2
-118 ± 1
11 ± 1
26 ± 1
(R,S)-5
NMR (CDCl3)
76 ± 20
-109 ± 12
4 ± 5
17 ± 9
NMR (CD3OH)
91 ± 39
-112 ± 23
6 ± 6
15 ± 13
poly-(S)-
ATC[12]-
[16]
XRD
-78 ± 3
127 ± 14
0 ± 3
-41 ± 4
NMR (CDCl3)
-76 ± 16
128 ± 12
-6 ±5
-28 ± 6
Gopi's[11]a
XRD
-119 ± 4
100 ± 5
0 ± 3
-78 ± 4
a The oligomers reported by Gopi et al. alternate S-vinylogous γ-amino acids
and Aib.
C9#C9#
(S)-ATC'1"
Phe#2#
(R)-ATC'1"
Phe#2#
C9#
C12'
C7'
Phe#2#
(S)-ATC'1"
C9#
C12'
C7'
(R)-ATC'1"
C/#
A/#
B/#
Compound 1
Δδ#(ppm)#
0'
0.2'
0.4'
0.6'
0.8'
1'
1.2'
1.4'
1.6'
ATC1# F2# iPr# ATC1# F2# iPr#
Free#NH#
C12#
C9#
Compound 2
CD3OH#
D/#
Phe#2#
9
7
N
H
(S)(S)
O
H
N
N
H
O
O
Br
Ph
NS
O
9
N
H
(S)(S)
O
H
N
N
H
O
O
Br
Ph
NS
O
9
N
H
(S)(S)
O
H
N
N
H
O
O
Br
Ph
NS
O
12
10
8
6
4
2
0
x10-3
1800 1700 1600 1500 1400
10
8
6
4
2
0
x10-3
1800 1700 1600 1500 1400
1702#cm-1#
Free#CO#(Phe#2)#
1670#cm-1#
Bound#CO#carbamate#
1641#cm-1#
Free#CO#(ATC#1)#
1706#cm-1#
Free#CO#(Phe#2)#
1673#cm-1#
Bound#CO#carbamate#
1643#cm-1#
Free#CO#(ATC#1)#
Wavenumber#(cm-1)# Wavenumber#(cm-1)#
Absorbance#
#
X#10-3#
#x#10-3#
E/#
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
Table 2. Relative conformational free energies (ΔG) and hyperconjugation
energies (ΔE2) of conformers for 1 and 2.a
Cpd
Conf
ΔGb
ΔE2c
C9
C12
C7
1
C9
0.89
13.56
C9/12
0.00
10.70
7.70
C9/7
3.78
3.84
8.95
2
C9/7
0.01
17.73
6.12
C9/12
0.00
18.03
0.28
a All energies are in kcal mol-1. b Gibbs free energies were calculated at the
ωB97X-D/def2-TZVP//SMD M06-2X/6-31G(d) level of theory in chloroform
(see computational details in the Supporting Information). c The
hyperconjugation due to charge transfer for each hydrogen bond was
calculated in terms of the second perturbation energy (ΔE2) of the lone pair
orbitals of the carbonyl oxygen with the corresponding NH antibonding
orbital[15] at the ωB97X-D/def2-TZVP level of theory.
and C9/7 (ΔG = 0.00 and 0.01 kcal mol1, respectively) although
the latter had more favorable energy (ΔE = 1.65 kcal mol1) and
enthalpy (ΔH = 1.46 kcal mol1) than the former. The C9-
conformer was not a local minimum and converged into the C9/7-
conformer after geometry optimization. Whatever the model
considered, the C9-hydrogen bonds were rather strong (ΔE2 =
17.73 and 18.03 kcal mol-1 respectively) compared to the C7 or
C12 (ΔE2 = 6.12 and 0.28 kcal mol-1 respectively) and remained
the main stabilizing feature of the ATC-based structures.
To gather experimental evidences on the intramolecular
hydrogen bonding of 1 and 2, we performed CDCl3/CD3OH NMR
titration. In contrast to CDCl3, CD3OH can form H-bonds with the
compounds that may compete with intramolecular H-bonds. We
could dissociate two different NH behaviors from 1 (Figure 1D
Table S34). The NH-Phe exhibited small solvent chemical shifts
dependency along the titration (Δδ = 0.39 ppm), indicating its
involvement in the C9-hydrogen bond. By contrast both HN-ATC
and HN-iPr were accessible to the solvent according to their
significant resonance variation upon CD3OH addition (Δδ = 1.34
and 1.44 ppm, respectively). These data strongly support the C9-
model compared to the C9/7-, C9/12-ones. Conversely, 2 displayed
three different NH behaviors (Figure 1D Table S35). As in 1,
NH-Phe had a small chemical shift variation (Δδ = 0.21 ppm)
while HN-ATC was highly sensitive to CD3OH addition (Δδ =
1.33 ppm). Interestingly, HN-iPr showed an intermediate
comportment (Δδ = 0.78 ppm) denoting a partial protection from
the solvent in line with the C9/7- or C9/12-bifurcated hydrogen
bond models. To definitively ascertain the H-bonding network of
1 and 2 we achieved FT-IR experiments in CHCl3. The spectra
were almost similar for both compounds. The band at 1706-1702
cm-1 was attributed to the free Phe C=O vibration while the high
frequency amide I bands at 1673-1670 cm-1 were assigned to
the urethane (Figure 1E). We showed in a recent study that the
amide I frequencies provide unambiguous structural markers of
the H-bond network for ATC-containing oligomers.[17] According
to this previous work and based on calculation of the vibrational
frequencies at the SMD M06-2X/6-31G(d) level of theory in
chloroform, free ATC carbonyls were expected to give amide I
absorption at ν(CO)>1640 cm-1 (calculated at 1661 and 1677 cm-
1 for C9- and C9/12-conformers of 1, respectively, and at 1660 cm-
1 for C9/12-conformer of 2) while the bounded ATC C=O should
be highly redshifted around 1626 cm- 1 (calculated at 1624 and
1630 cm-1 for C9/7-conformers of 1 and 2, respectively). Bands
were measured at 1643-1641 cm-1 but no vibration was observed
around 1620-1630 cm-1 for both compounds 1 and 2, excluding
that the ATC carbonyl was engaged into a C7-hydrogen bond
with the NH-iPr. Consequently, taking our results all together, we
could conclude that the homochiral dipeptide 1 may adopt a C9-
fold while the heterochiral analog 2 may prefer the C9/12-
conformation. These data are consistent with our previous report
demonstrating that when incorporated in short peptide
sequences, ATC acts as a strong turn inducer stabilized either
by C9- or C9/12 bidentate hydrogen-bond depending on the γ-
amino acid absolute configuration.[16] Interestingly, such a
hydrogen bonding system has been observed by Grison et al.
for peptides incorporating a Z-vinylogous fragment.[13]
After shedding light on the pre-organization of small dipeptides,
we were interested in the folding capacity of homo- and
heterochiral (α/ATC)n oligomers. Hexapeptides 3 and 5 (Chart 1)
were prepared by alternating (S-)phenylalanine and (R)- or (S)-
ATC respectively (See the Supporting Information). Despite
many efforts, crystallization assays were not successful.
Nevertheless, the NMR signals in CDCl3 were well dispersed
and nearly all 1H, 13C and 15N resonances could be assigned for
both compounds, combining homonuclear COSY, TOCSY,
ROESY experiments and heteronuclear 15N-HSQC, 13C-HSQC
at 15N and 13C natural abundance (Tables S6 and S10). Any
chemical shifts variation was observed from 1 to 10 mM
suggesting that no peptide self-association occurred under
analysis conditions. The amide proton signals following the γ-
amino acids were strongly downfield, i.e. from 9.02 to 9.53 ppm
for 3 and from 9.10 to 9.75 ppm for 5. Such a strong NH
deshielding (> 9 ppm) was formerly related to the formation of
the typical C9-hydrogen bonds surrounding the ATC residues.[17]
Also, we detected strong ATC-δCH(i)/PheNH(i+1) and weak
ATC- γCH(i)/PheNH(i+1), ATC-γCH(i)/Phe-αCH(i+1) and ATC-
γCH(i)/Phe-βCH(i+1) correlations which were characteristics of
the C9-turn (Figure 2A).[16-17] In addition, long range
Phe(i)/ATC(i+3) NOEs were observed along the backbone of 3
and 5 (Figure 2B). These data were strong evidences of well-
organized systems in CDCl3, nevertheless the observed strong
ATC-δCH(i)/PheNH(i+1) correlations were not compatible with
the α/γ 12-helix proposed by Gopi et al.[11] In CD3OH, we
detected a similar set of i/i+1 and i/i+3 correlations for the
heterochiral sequence 5, supporting the hypothesis that the
structure remained in methanol. Inversely for the homochiral
hexamer 3, these characteristic long range Phe(i)/ATC(i+3)
crosspeaks disappeared. NOE correlations (Tables S21 and
S24) were then used as distance restraints for NMR structure
calculations using a typical simulated annealing protocol with
AMBER 11.[18] Peptide 3 displayed a non-canonical helical
shape in CDCl3 while calculations did not converged in CD3OH
demonstrating that the homochiral (α/ATC)n did not provide
stable platforms (see Figures S11 and S18). In contrast, the
solution structures of the heterochiral sequence 5 converged into
a well-defined ribbon shape in both CDCl3 (Figure 2D) and
CD3OH (Figure S15). The fold was stabilized by a regular C9/12-
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
Figure 2. A/ Typical short range NOE correlations consistent with the formation of a C9-hydrogen bond and characteristic i/i+1 and Phe(i)/ATC(i+3) NOEs
observed along the backbone of 4-6. B/ Hydrogen-bond patterns for heterochiral (α/ATC)n oligomers C/ FT-IR spectra of 4, 5 and 6 in the amide A and amide I
regions (3 mM in CHCl3). D/ Superimposition of the 15 lowest energy NMR structures of oligomers 3-6 in CDCl3 (lateral chains are omitted for clarity). E/
Superimposition of the lowest energy NMR structure (green) and of the DFT-optimized geometries (purple) of oligomer 5 (lateral chains were omitted for clarity).
H-bonding pattern involving the C=O of the α residue (i) and both
the amide protons of the ATC (i+1) and the α amino acid (i+2).
We then optimized the lowest-energy NMR structures of 3 and 5
by DFT calculations at the M06-2X/6-31G(d) level theory and
calculating the solvation free energies at the SMD M06-2X/6-
31G(d) level of theory in chloroform (Figure 2E). RMSD of the
backbone heavy atoms between NMR and DFT structures were
1.00 and 1.06 Å. According to DFT calculations, the heterochiral
hexamer 5 was 4.80 kcal mol1 more stable in Gibbs free energy
than the homochiral hexamer 3 in chloroform. Conversely, 3 had
a favorable entropic contribution (TΔS) of 1.57 kcal mol1 than
5. This indicates that 5 is more rigid than 3. The stabilities of the
C9- and C12-hydrogen bonds were investigated by calculating the
second perturbation energies (ΔE2) of the lone pair orbitals of
the carbonyl oxygen with the corresponding NH antibonding
orbital at the ωB97X-D/def2-TZVP level of theory (Table S48).
The sum of ΔE2 values of two C9/12-bifurcated and one C9-
hydrogen bonds for 5 was 14.68 kcal mol1 greater than 3, to
which the largest contribution was due to the C9-hydrogen bond
between Phe(3)-CO with HN-Phe(5). Thus, the ribbon structures
could be regarded as a succession of highly pre-organized α/γ
dipeptides comprising strong C9- and C12-hydrogen bonds
(Figure 2D). Similar NMR data sets were obtained for other
heterochiral α/γ (Chart 1, compounds 4-6) and γ/α sequences
(Chart 1, compounds 7-9). Whatever the lengths, they shared
similar C9/12-ribbon structures (Figures 2B-C and S12-13 and
S15-16) with average backbone torsion angle values
comparable to those measured on the C9/12-model of the (R)-γ/α
dipeptide 2 (Table S44). The lateral chains are distributed along
four projection axes over the γ-peptide backbone. The thiazole
rings belonging to the ATCs (i, i+4) project perpendicularly to the
ribbon plane and are spaced by around 9 Å (Figure 3).
Interestingly, as shown in Figure 3, the substituents (in green) at
the γ-position of ATCs (i, i+4) and the lateral chains of the α-
amino acids (i-1, i+3) are almost aligned on the same face. They
are separated by approximately 5.1 Å, which reminds the α-helix
pitch.
The structural behavior of the α/(R)-ATC 4-6 and (R)-ATC/α
oligomers 7-9 peptides was confirmed by circular dichroism. The
CD signatures of oligomers 4-9 in methanol (Figures S20 and
S21) were similar to those of other ATC-containing peptides
which displayed turn conformations.[16] They exhibited two
positive maxima at 203 and 228 nm and a large negative band
centred at 255 nm. The α/γ or γ/α alternations of similar lengths
have closed CD signal intensities indicated that they were
almost stable. When the temperature was increased up to 55°C,
A/# B/#
NS
O
N
H
N
HO
NS
O
HN
H
N
O
* *
D/#
4#
5#
6#
C9#
C9#
C12#
C12#
C9#
E/#
5#
C#
N#
N#
N#
C/#
4"
5"
6"
40x10-3
30
20
10
0
Absorbance
1800 1700 1600 1500 1400
Wavenumber (cm-1)
1715$cm-1$
1721$cm-1$
1721$cm-1$
1643$cm-1$
1640$cm-1$
1640$cm-1$
Amide$II$
8x10-3
6
4
2
0
Absorbance
3500 3400 3300 3200 3100
Wavenumber (cm-1)
3227$cm-1$
3225$cm-1$
3223$cm-1$
3417$cm-1$
3420$cm-1$
3420$cm-1$
3387$cm-1$
3382$cm-1$
3378$cm-1$
NS
O
N
H
N
HO
*
> 9.0 ppm
NS
R3
O
N
H
R2
N
H
R1
O
NS
R3
O
HN
R2
H
N
O
R1
*
C9
C12
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
Figure 3. A/ Side chain projections along the α/(R)-ATC backbone. B/
Superimposition of the α/(R)-ATC backbone and a canonical α-helix (ribbon).
Substituents of residues (i, i+3) of the α-helix are fleshed-colored.
we observed a slight decrease of the molar ellipticity without
variation in the shape of the spectra. A total recovery of the CD
signal was observed when the temperature was returned to
20°C, suggesting that the limited thermal loss of structure
induced by higher temperatures was fully reversible. Importantly,
the per-residue molar ellipticity values of the extrema did not
vary with the number of α/γ dipeptides, as previously described
for other ribbon structures.[19] Once again, such result supported
the idea that the ribbon structure was governed by each pre-
organized α/(R)-ATC C9/12-turns. The regular C9/12 H-bond
pattern was finally established by CDCl3/CD3OH NMR titration
experiments and FTIR spectroscopy. As observed on dimer 2
upon CDCl3/CD3OH titration experiments, each peptide showed
three different NH behaviors attributed to the free N-terminus
amides (Δδ > 1.1 ppm), Phe-NH (Δδ 0.40 ppm) and ATC-NH
(0.40 < Δδ < 0.80 ppm) engaged in C9- and C12-hydrogen
bonding patterns respectively (Tables S36-41 and Figure S10).
The three NH populations were also evidenced by FT-IR
analyses (Figure 3C and Table S52) in the ν(NH) region (3600-
3100 cm-1). The band around 3420 cm-1 was assigned to the
free NH while the lower vibration frequency (~3220 cm-1 )
corresponded to the NH-Phe engaged in strong C9-hydrogen
bonds. The intermediate band at about 3380 cm-1 was assigned
to the NH-ATC forming C12 pseudocycles. Analysis of the amide
I region (1800-1600 cm-1) was also consistent with the C9/12-
ribbon structure (Figure 3C). The high frequency band observed
for oligomers 4, 5 and 6 at around 1720 cm-1 was attributed to
the free urethane C=O group (Table S53). In sequences 7, 8
and 9, the frequency shifted below 1700 cm-1 probably at about
1685 cm-1 due to H-bonding of the urethane C=O (Table S53).
Consequently, the unsymmetrical broad band between 1640-
1670 cm-1 was attributed to a combination of two vibrations
corresponding to bound Phe C=O and free ATC C=O.
Importantly, estimations of positions and contributions of
discrete subcomponent absorptions in the amide I region have
ultimately been achieved using curve-fitting approaches with
Gaussian functions. As reported on Tables S54-S59,
quantitative analysis revealed that contributions of each C=O
perfectly settled the proposed structure. For both series, we
observed a concomitant downshift of the ν(NH) and ν(CO) band
position to lower frequencies along with the increase of the
oligomer size suggesting stronger hydrogen bonds. Interestingly,
for oligomers of identical sizes, the γ/α series seemed to exhibit
a higher conformational stability than the α/γ oligomers. Indeed,
the amide proton resonances were less sensitive upon methanol
addition for the γ/α series in the CDCl3/CD3OH NMR titration,
and their bound NH and CO displayed significantly downshifted
ν(NH) and ν(CO) absorption bands compared to the
corresponding bands in the α/γ series (Figure S10). This could
be explained because, in the α/γ sequences, the first α-amino
acids is not implicated in the hydrogen-bonding network of the
ribbon structure, thus some fraying at the N-terminus may
occurred which slightly affect the overall stability of the edifice.
Conclusions
We showed in this study that (S/R)-ATC/α dipeptides
adopt C9 and C9/12-turns, respectively. While the homochiral
α/(S)-ATC oligomers do not adopt a stable structure, we pointed
out the ability of the pre-organized C9/12-turn to propagate along
repetitive α/(R)-ATC sequences forming unusual ribbon
structure stabilized by C9/12-bifurcated hydrogen bonds.
Moreover, the (R)-ATC/α sequences were more stable than the
α/(R)-ATC ones. This identified structure is different from those
recently described by Gopi et al. for α/γ-hybrid peptides
alternating achiral Aib and Z-vinylogous γ-amino acids which
adopt a 12-helix. Nevertheless, instead of the highly helicogenic
disubstituted Aib residue, we used proteogenic α-amino acids in
order to maximize the diversification points. Thus, at this stage,
it is not clear if the conformational preference of such sequences
is mainly driven by the α-moiety or is specific to the ATC building
block since 1,4-O,S interactions may occurred which strongly
reduced its conformational freedom (ψ ~ 20°).[20] Such question
is currently under study.
Experimental Section
Commercially available reagents and solvents were used without any
further purification. Reactions were monitored by Thin-Layer
Chromatography (TLC) and analytical HPLC. Products were purified by
column chromatography on a Merck Kieselgel 60 (230-400 mesh) or by
flash column chromatography of silica gel using a Biotage Isolera One
system. Analytical TLCs were run on Merck Kieselgel 60F254 plates.
Visualization was accomplished by UV light (254 nm) and by heating at
300°C after spraying with a commercial phosphomolybdic acid solution
(20% in EtOH). Analytical HPLC analyses were run on an Alliance HT
Waters 2795 separations module equipped with a Chromolith Speed Rod
RP-C18 185 Pm column (50 x 4.6 mm, 5 µm) with a gradient from 100%
(H2O/TFA 0.1%) to 100% (CH3CN/TFA 0.1%) in 3 min; flow rate 5
mL/min; UV detection was done at 214 and 254 nm with a Waters 996
Photodiode Array Detector. Retention times are reported as follows: Rt =
N
S
O
NH
NH
ON
S
O
H
N
HN
O
NH
O
N
S
ONH
O
NH2
N"
C"
N"
C"
5.3$Å$
4.4$Å$
5.7$Å$
A/$ B/$
90°$
90°$
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
(min). LC/MS analyses were recorded on a Quattro micro ESI triple
quadrupole mass spectrometer (Micromass, Manchester, UK) equipped
with a Chromolith Speed Rod RP-C18 185 Pm column (50 x 4.6 mm, 5
µm) and an Alliance HPLC System (Waters, Milford, USA); gradient from
100% (H2O/HCO2H 0,1%) to 100% (CH3CN/HCO2H 0,1%) in 3 min;
flowrate 3 mL/min; UV detection at 214 nm. High-Resolution Mass
Spectrometric analyses were performed with a time-of-flight (TOF) mass
spectrometer fitted with an Electrospray Ionisation source (ESI) in
positive ion mode. 1H NMR and 13C NMR spectra were recorded at room
temperature (r.t.) on a Bruker 400 spectrometer at 400.13 MHz and
100.62 MHz respectively; chemical shifts (δ) are reported in parts per
million (ppm) relative to the solvent [1H: δ (CDCl3) = 7.26 ppm; 13C: δ
(CDCl3) = 77.16 ppm; 1H: δ (DMSO-d6) = 2.50 ppm; 13C: δ (DMSO-d6) =
39.52 ppm]. NMR spectra are reported as follows: δ (signal multiplicity,
coupling constant(s), number of protons, numbered protons). 1H NMR
signal multiplicities are designated as broad (br), singlet (s), doublet (d),
doublet of doublet (dd), triplet (t), triplet of doublet (td), quadruplet (q),
quintet (quin), multiplet (m), broad multiplet (brm), sextet (sext), octuplet
(oct), nonet (non).
Synthesis of compounds 1-9: Fmoc-(S)- and (R)-ATC-OH were
prepared according to reported procedures.[14] Syntheses of peptides 1-9
are described in the Supporting Information.
NMR experiments and simulated annealing protocol: The NMR
samples contained 2 mM of (1)-(9) dissolved in CDCl3 and CD3OH.
Spectra were recorded on a Bruker Avance 600 AVANCE III
spectrometer equipped with a 5 mm quadruple-resonance probe (1H, 13C,
15N, 31P). Homonuclear 2D spectra DQF-COSY, TOCSY (DIPSI2) and
ROESY were typically recorded in the phase-sensitive mode using the
States-TPPI method as data matrices of 256-512 real (t1) × 2048 (t2)
complex data points; 8-48 scans per t1 increment with 1.0 s recovery
delay and spectral width of 7210 Hz in both dimensions were used. The
mixing times were 80 ms for TOCSY and 350 ms for the ROESY
experiments. In addition, 2D heteronuclear spectra 13C and 15N-HSQC
were acquired to assign 15N and 13C resonances (32-96 scans, 64-256
real (t1) × 2048 (t2) complex data points). Spectra were processed with
Topspin (Bruker Biospin) and visualized with Topspin or NMRView on a
Linux station. The matrices were zero-filled to 1024 (t1) x 2048 (t2) points
after apodization by shifted sine-square multiplication and linear
prediction in the F1 domain. Chemical shifts were referenced to the
tetramethylsilane (TMS). 1H chemical shifts were assigned according to
classical procedures. NOE cross-peaks were integrated and assigned
within the NMRView software. The volume of a ROE between methylene
pair protons was used as a reference of 1.8 Å. The lower bound for all
restraints was fixed at 1.8 Å and upper bounds at 2.7, 3.3 and 5.0 Å, for
strong, medium and weak correlations, respectively. Pseudo-atoms
corrections of the upper bounds were applied for unresolved aromatic,
methylene and methyl protons signals as described previously. Structure
calculations were performed with AMBER 11 in two stages: cooking and
simulated annealing (SA). The cooking stage was performed at 1000 K to
generate 100 initial random structures. SA calculations were carried
during 20 ps (20000 steps, 1 fs long). First, the temperature was risen
quickly and was maintained at 1000 K for the first 5000 steps, then the
system was cooled gradually from 1000 K to 100 K from step 5001 to
18000 and finally the temperature was brought to 0 K during the 2000
remaining steps. For the 3000 first steps, the force constant of the
distance restraints was increased gradually from 2.0 kcal mol-1to 20
kcal.mol-1.Å. For the rest of the simulation (step 3001 to 20000), the force
constant is kept at 20 kcal.mol-1.Å. The 15 lowest energy structures with
no violations > 0.3 Å were considered as representative of the peptide
structure. The representation and quantitative analysis were carried out
using MOLMOL and PyMOL (Delano Scientific). Ptraj were used to
measure the dihedrals angles.
DFT calculations: All DFT calculations have been performed using the
hybrid-meta-GGA M06-2X functional[21] implemented in the Gaussian 09
program.[22] For (1) and (2), the lowest-energy NMR structures with C9/7-
and C9/12-hydrogen bonds and crystal structures with the C9-hydrogen
bond were used as initial structures for optimization. For 2, the gauche+
and gauche- torsions of the Phe2 residue were also considered for the
initial structures to obtain the feasible local minima. In addition, lowest-
energy NMR structures with two C9/12- and one C9-hydrogen bonds of (3)
and (5) were used as initial structures for optimization. All Br atoms of N-
terminal Cbz groups were replaced by H atoms. First, all initial structures
of (1), (2), (3) and (5) were optimized at the M06-2X/6-31G(d) level of
theory. For (1) and (2), further optimizations were performed in
chloroform using the solvation model based on the density (SMD)
method[23] at the M06-2X/6-31G(d) level of theory. The vibrational
frequencies of local minima of (1) and (2) in chloroform were calculated
at the SMD M06-2X/6-31G(d) level of theory at 25 °C and 1 atm.
However, the vibrational frequencies of local minima of (3) and (5) were
calculated at the M06-2X/6-31G(d) level of theory at 25 °C and 1 atm and
single-point energies in chloroform were calculated at the SMD M06-
2X/6-31G(d) level of theory. Each local minimum was confirmed by
verifying the absence of imaginary frequencies after the frequency
calculations. The scale factor used was 0.9323 at the M06-2X/6-31G(d)
level of theory; this value was chosen to reproduce the experimental
frequency at 1653 cm1 for the amide I band of the free carbonyl group of
the C-terminal end of Ac-(S)-ATC-NHiPr in chloroform.[17] To obtain the
new scale factor, the C9-hydrogen bonded structure of Ac-(S)-ATC-NHiPr
was optimized in chloroform at the SMD M06-2X/6-31G(d) level of theory
and followed by the frequency calculation. In addition, single-point
energies were calculated using the wB97X-D functional[24] with the larger
def2-TZVP basis set[25] to improve the conformational energies. Zero-
point energy and thermal energy corrections were employed to calculate
the Gibbs free energy of each conformation. Here, the ideal gas, rigid
rotor, and harmonic oscillator approximations were used for the
translational, rotational, and vibrational contributions to the Gibbs free
energy, respectively.[26] The strengths of hydrogen bonds in local minima
of (1), (2), (3) and (5) were evaluated in terms of the second perturbation
energy (ΔE2) of the lone pair orbitals of the carbonyl oxygen with the
corresponding NH antibonding orbital using natural bond orbital (NBO)
analysis[15] at the wB97X-D/def2-TZVP level of theory.
FT-IR experiments: Middle-infrared experiments (400-5000 cm-1) were
recorded in the transmission mode. The measurements were carried out
on a Bruker Tensor 27 spectrometer equipped with a deuterated (L)-
alanine doped triglycene sulphate (DLATGS) pyroelectric detector, a
Globar source, and potassium bromide (KBr) beam splitter. The spectral
resolution was 4 cm-1, and 128 scans were co-added for each spectrum.
The compounds were dissolved at 1 mM or 3 mM concentration in CHCl3
transferred in a liquid cell equipped with CaF2 windows separated by a
teflon spacer (thickness: 50 µM). A spacer thickness of 50 µM provide
exploitable signal in term of signal-to-noise ratio S/N (S/N (Peak to peak
= 4x10-5), S/N (quadratic mean (RMS) = 1x10-5 between 2000 and 2100
cm-1). Bellow this value, significant compensation problems between
solvent spectral bands used as reference and sample spectrum appear.
The FT-IR spectra were not smoothed. Baseline substraction and
deconvolution were performed using IGOR Pro 6.0 software
(WaveMetrics).
Circular dichroism (CD): CD experiments were carried out using a
Jasco J815 spectropolarimeter. The spectra were obtained in MeOH
using a 1 mm path length CD cuvette, at 20°C, over a wavelength range
of 190-300 nm. Continuous scanning mode was used, with a response of
1.0 s with 0.2 nm steps and a bandwidth of 2 nm. The signal to noise
ratio was improved by acquiring each spectrum over an average of two
scans. Baseline was corrected by subtracting the background from the
sample spectrum. The samples were dissolved in aspectrophotometric
grade MeOH at 100-200 µM.
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
Acknowledgements
Thanks to funding provided by the ANR (French National
Research Agency) CatFOLD Project. The authors also thanks
the Pôle Chimie Balard and the labex CheMISyst which financed
collaborative exchanges between Montpellier University
(France) and Chungbuk National University (Republic of Korea).
Keywords: keyword 1 keyword 2 keyword 3 keyword 4
keyword 5
[1] a/ D. J. Hill, M. J. Mio, R. B. Prince, T. S. Hughes, J. S. Moore, Chem.
Rev. 2001, 101, 3893-4012; b/ S. H. Gellman, Acc. Chem. Res. 1998,
31, 173180; c/ G. Guichard, I. Huc, Chem. Commun. 2011, 47, 5933-
5941; d/ C. M. Goodman, S. Choi, S. Shandler, W. F. DeGrado, Nat.
Chem. Biol. 2007, 3, 252-262.
[2] a/ D. Seebach, D. F. Hook, A. Glattli, Biopolymers 2006, 84, 23-37; b/ D.
Seebach, J. Gardiner, Acc. Chem. Res. 2008, 41, 1366-1375; c/ R. M.
Liskamp, D. T. Rijkers, J. A. Kruijtzer, J. Kemmink, ChemBioChem.
2011, 12, 1626-1653; d/ P. G. Vasudev, S. Chatterjee, N. Shamala, P.
Balaram, Chem. Rev. 2011, 111, 657-687.
[3] a/ R. P. Cheng, S. H. Gellman, W. F. DeGrado, Chem. Rev. 2001, 101,
3219-3232; b/ D. Seebach, A. K. Beck, D. J. Bierbaum, Chem.
Biodivers. 2004, 1, 1111-1239; c/ S. Hanessian, X. Luo, R. Schaum, S.
Michnick, J. Am. Chem. Soc. 1998, 120, 85698570; d/ T. Hintermann,
K. Gademann, B. Jaun, D. Seebach, Helv. Chim. Acta 1998, 81, 983-
1002; e/ D. Seebach, M. Brenner, M. Rueping, B. Schweizer, B. Jaun,
Chem. Commun. 2001, 207-208; f/ M. Brenner, D. Seebach, Helv.
Chim. Acta 2001, 84, 1181-1189; g/ D. Seebach, M. Brenner, M.
Rueping, B. Jaun, Chem. Eur. J. 2002, 3, 573-584; h/ K. Basuroy, B.
Dinesh, M. B. M. Reddy, S. Chandrappa, S. Raghothama, N. Shamala,
P. Balaram, Org. Lett. 2013, 15; i/ F. Bouillere, S. Thetiot-Laurent, C.
Kouklovsky, V. Alezra, Amino Acids 2011, 41, 687-707.
[4] a/ R. N. Zuckermann, Biopolymers 2011, 96, 545-555; b/ A. Violette, M.
C. Averlant-Petit, V. Semetey, C. Hemmerlin, R. Casimir, R. Graff, M.
Marraud, J.-P. Briand, D. Rognan, G. Guichard, J. Am. Chem. Soc.
2005, 127, 2156-2164; c/ C. Caumes, O. Roy, S. Faure, C. Taillefumier,
J. Am. Chem. Soc. 2012, 134, 9553-9556; dR. Wechsel, J. Maury, J.
Fremaux, S. P. France, G. Guichard, J. Clayden, Chem. Commun.
2014, 50, 15006-15009.
[5] W. S. Horne, S. H. Gellman, Acc. Chem. Res. 2008, 41, 1399-1408.
[6] a/ S. Chatterjee, R. S. Roy, P. Balaram, J. R. Soc. Interface 2007, 4,
587-606; b/ B. Legrand, C. André, L. Moulat, E. Wenger, C. Didierjean,
E. Aubert, M. C. Averlant-Petit, J. Martinez, M. Calmes, M. Amblard,
Angew. Chem. Int. Ed. Engl. 2014, 53, 1313113135.
[7] C. Baldauf, R. Günther, H.-J. Hofmann, J. Org. Chem. 2006, 71, 1200-
1208.
[8] a/ A. Bandyopadhyay, H. N. Gopi, Org. Lett. 2012, 14, 2770-2773; b/ A.
Bandyopadhyay, S. V. Jadhav, H. N. Gopi, Chem. Commun. 2012, 48,
7170-7172; c/ S. V. Jadhav, A. Bandyopadhyay, H. N. Gopi, Org.
Biomol. Chem. 2013, 11, 509-514; d/ S. V. Jadhav, R. Misra, S. K.
Singh, H. N. Gopi, Chem. Eur. J. 2013, 19, 16256-16262; e/ S. V.
Jadhav, R. Misra, H. N. Gopi, Chem. Eur. J. 2014, 20, 16523-16528; f/
K. Basuroy, B. Dinesh, N. Shamala, P. Balaram, Angew. Chem. Int. Ed.
Engl. 2012, 51, 8736-8739; g/ R. Sonti, B. Dinesh, K. Basuroy, S.
Raghothama, N. Shamala, P. Balaram, Org. Lett. 2014, 16, 1656-1659.
[9] G. V. Sharma, V. B. Jadhav, K. V. Ramakrishna, P. Jayaprakash, K.
Narsimulu, V. Subash, A. C. Kunwar, J. Am. Chem. Soc. 2006, 128,
14657-14668.
[10] a/ L. Guo, Y. Chi, A. M. Almeida, I. A. Guzei, B. K. Parker, S. H.
Gellman, J. Am. Chem. Soc. 2009, 131, 16018-16020; b/ L. Guo, W.
Zhang, I. A. Guzei, L. C. Spencer, S. H. Gellman, Org. Lett. 2012, 14,
2582-2585; c/ M. W. Giuliano, S. J. Maynard, A. M. Almeida, L. Guo, I.
A. Guzei, L. C. Spencer, S. H. Gellman, J. Am. Chem. Soc. 2014, 136,
15046-15053; d/ B. F. Fisher, L. Guo, B. S. Dolinar, I. A. Guzei, S. H.
Gellman, J. Am. Chem. Soc. 2015, 137, 6484-6487; e/ B. F. Fisher, S.
H. Gellman, J. Am. Chem. Soc. 2016, 138, 10766-10769.
[11] M. Ganesh Kumar, V. J. Thombare, M. M. Katariya, K. Veeresh, K. M.
Raja, H. N. Gopi, Angew. Chem. Int. Ed. Engl. 2016, 55, 7847-7851.
[12] L. Mathieu, B. Legrand, C. Deng, L. Vezenkov, E. Wenger, C.
Didierjean, M. Amblard, M. C. Averlant-Petit, N. Masurier, V. Lisowski, J.
Martinez, L. T. Maillard, Angew. Chem. Int. Ed. Engl. 2013, 52, 6006-
6010.
[13] a/ C. Baldauf, R. Gunther, H. J. Hofmann, J. Org. Chem. 2005, 70,
5351-5361; b/ C. Grison, P. Coutrot, S. Geneve, C. Didierjean, M.
Marraud, J. Org. Chem. 2005, 70, 10753-10764.
[14] L. Mathieu, C. Bonnel, N. Masurier, L. T. Maillard, J. Martinez, V.
Lisowski, Eur. J. Org. Chem. 2015, 2015, 2262-2270.
[15] F. Weinhold, in Encyclopedia of Computational Chemistry, P. v.R.
Schleyer, N. L. Allinger, T. Clark, J. Gasteiger, P. A. Kollman, H. F.
Schaefer III, P. R. Schreiner (Eds.), (John Wiley & Sons, Chichester,
UK, 1998), Vol. 3, pp. 1792-1811.
[16] B. Legrand, L. Mathieu, A. Lebrun, S. Andriamanarivo, V. Lisowski, N.
Masurier, S. Zirah, Y. K. Kang, J. Martinez, L. T. Maillard, Chem. Eur. J.
2014, 20, 6713-6720.
[17] C. Bonnel, B. Legrand, J. L. Bantignies, H. Petitjean, J. Martinez, N.
Masurier, L. T. Maillard, Org. Biomol. Chem. 2016, 14, 8664-8669.
[18] D. A. Case, T. A. Darden, T. E. Cheatham, C. L. Simmerling, R. E. D. J.
Wang, R. Luo, R. C. Walker, W. Zhang, K. M. Merz, B. Roberts, B.
Wang, S. Hayik, A. Roitberg, G. Seabra, I. Kolossváry, K. F. Wong, F.
Paesani, J. Vanicek, J. Liu, X. Wu, S. R. Brozell, T. Steinbrecher, H.
Gohlke, Q. Cai, X. Ye, J. Wang, M.-J. Hsieh, G. Cui, D. R. Roe, D. H.
Mathews, M. G. Seetin, C. Sagui, V. Babin, T. Luchko, S. Gusarov, A.
Kovalenko, P. A. Kollman, University of California, San Francisco, 2010.
[19] V. Martin, B. Legrand, L. L. Vezenkov, M. Berthet, G. Subra, M. Calmès,
J.-L. Bantignies, J. Martinez, M. Amblard, Angew. Chem. Int. Ed. Engl.
2015, 1396613970.
[20] B. R. Beno, K. S. Yeung, M. D. Bartberger, L. D. Pennington, N. A.
Meanwell, J. Med. Chem. 2015, 58, 4383-4438.
[21] Y. Zhao, D. G. Truhlar, Theor. Chem. Acc. 2008, 120, 215.
[22] M. J. T. Frisch, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;
Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson,
G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A.
F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.;
Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda,
Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J.
E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.;
Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.;
Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega,
N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.;
Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.;
Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;
Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.;
Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman,
J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J., in Gaussian 03, Revision
D.01; Gaussian, Inc.: Wallingford, CT, 2009.
[23] A. V. C. Marenich, C. J.; Truhlar, D. G. J. , Phys. Chem. B . 2009, 113,
6378.
[24] J.-D. Chai, M. Head-Gordon, Phys. Chem. Chem. Phys. 2008, 10,
6615-6620.
[25] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297-3305.
[26] W. J. R. Hehre, L.; Schleyer, P. v. R.; Pople, J. A., Ab Initio Molecular
Orbital Theory; John Wiley & Sons: New York, 1986; Chapter 6.
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
COMMUNICATION
Entry for the Table of Contents (Please choose one layout)
Layout 1:
COMMUNICATION
A ribbon structure in hybrid peptides
alternating α- and thiazole-based γ-
amino acids
Author(s), Corresponding Author(s)*
Clément Bonnel, Baptiste Legrand,
Matthieu Simon, Jean Martinez, Jean-
Louis Bantignies, Young Kee Kang,
Emmanuel Wenger, Francois Hoh,
Nicolas Masurier, Ludovic T. Maillard*
Page No. Page No.
C9/12-ribbon-like structure in hybrid
peptides alternating α- and thiazole-
based γ-amino acids
N
S
O
NH
NH
ON
S
O
H
N
HN
O
NH
O
N
S
ONH
O
NH2
N"
C"
N"
C"
90°$
90°$
10.1002/chem.201704001
Accepted Manuscript
Chemistry - A European Journal
... While the latter adopt a 12-helical shape, [106] The ATC-containing oligomers fold into C9/12-ribbons. [107] In ATC-based molecules, the carbonyls and double bonds are coplanar while for the 12-helix of (Z)-vinylogous γ-amino acids they are not. Whatever the sequences studied, the ATC carbonyl group point towards the sulfur atom of the thiazole ring suggesting a dipole-dipole interaction between the C=O and the heterocycle. ...
... previously recognized as a structural marker related to the formation of the poly-ATC C9-helix pattern. [104,107] The less deshielded proton was attributed to the N-terminal ATC NH, an indication that this hydrogen was not engaged in such hydrogen bond network. Similar NH behaviours were observed for the trimer ATC3-2-NHiPr. ...
... It was decided to conduct FTIR experiments on ATC-oligomers in isopropanol at room temperature. Considering that the Amide I frequencies (1600-1800 cm -1 ) were formerly recognized as structural markers of the C9 H-bond network for ATC-containing oligomers in a previous work performed in the laboratory, [104,107] thus, we focused our efforts on the interpretation of signals in this area. ...
Thesis
The work described in this manuscript is devoted to the synthesis of heterocyclic constrained γ-amino acids, named ATCs (4-Amino-(methyl)-1,3-Thiazole-5-Carboxylic acids), their application in enamine catalysis and their structural study. ATC monomers are built around a thiazole ring providing a conformational limitation around the Cα and Cβ at 0°. The presence of two diversification points both on the γ asymmetric carbon and on the position 2 of the aromatic ring, allows a large structural diversification of the ATCs. Therefore, several oligomers were synthesized using solid phase peptide synthesis. A structural study of these oligomers, employing NMR, FTIR, circular dichroism and crystallography RX, demonstrated that they adopt a C9-right-handed helix stabilized by a hydrogen bond pattern between COi---NHi+2 along the helix. The objective of the project presented in this manuscript was the design and the structural characterization of molecular edifices with predictable folding properties and the systematic study of structure-function relationships in the nitro-Michael addition reaction, for three different substrates. Eventually, the last part of this work focused on the development of a new methodology, specific to ATC-oligomers, to perform 3D-modelling studies using NMR refinement.
... Such a helical structure was however somewhat unexpected since among the various helices predicated by Hoffmann for α/γ hybrid peptides, none displayed ζ dihedral angle around 0°. [60] Additionally, despite the structural similarity between Z-and Het-γAAs, the homochiral sequences alternating αAAs and (S)-ATC Het-γAAs also failed to fold. [61] As a consequence it is likely that Aib residues have a key impact on the folding of both 1 : 1 Aib/γ 4 AA and 1 : 1 Aib/Z-dγAA oligomers. [62] More recently, another 12-helix was described for an oligomer alternating Aibs and β,γ-unsaturated Z-γAAs. ...
... Finally, a last structure was obtained for heterochiral oligomers, alternating αAAs and (R)-ATC. [61] Analysis from tetrato octamers showed that the local conformation of tripeptide R-62 (Figure 6c) was retained all along the backbones of longer oligomers. The folding of hexamer 66 is depicted in Figure 7 (right). ...
Article
Full-text available
Despite their concomitant emergence in the 1990s, γ‐peptide foldamers have not developed as fast as β‐peptide foldamers and to date, only a few γ‐oligomer structures have been reported, and with sparse applications. Among these examples, sequences containing α,β‐unsaturated γ‐amino acids have recently drawn attention since the Z/E configurations of the double bond provide opposite planar restrictions leading to divergent conformational behaviors, from helix to extended structures. In this Review, we give a comprehensive overview of the developments of γ‐peptide foldamers containing α,β‐unsaturated γ‐amino acids with examples of applications for health and catalysis, as well as materials science.
... The authors have cited additional references within the Supporting Information. [20][21][22][23] ...
Article
Full-text available
Amidations employing mixed (carbonic) anhydrides have long been favoured in peptide synthesis because of their cost‐effectiveness and less waste generation. Despite their long history, no study has compared the effects of additives on the activation of mixed anhydrides and carbonic anhydrides. In this study, we investigated the amidation of mixed (carbonic) anhydride in the presence of a base and/or Brønsted acids. The use of NMI⋅HCl significantly improved the conversion of the mixed carbonic anhydride, while expediting nucleophilic attacks on the desired carbonyl group. In contrast, in the case of mixed anhydrides, neither the conversion nor the desired nucleophilic attack improved significantly. We developed a C‐terminus‐free N‐methylated peptide synthesis method using mixed carbonic anhydrides in a micro‐flow reactor. Fourteen N‐alkylated peptides were synthesized in moderate to high yields (55–99 %) without severe racemization (<1 %). Additionally, a significant enhancement in the amidation between mixed carbonic anhydrides and bis‐TMS‐protected N‐methyl amino acids with the inclusion of NMI⋅HCl was observed for the first time. In addition, we observed unexpected C‐terminal epimerization of the C‐terminus‐free N‐methyl peptides.
... [1][2][3][4][5][6][7][8][9] In particular, peptide foldamers can stabilize various helical structures, of which the type, handedness, and macrodipole direction of helices can be controlled by the substitutions and/or stereochemistry of the residues. [1][2][3][4][5][6][7][8][9][10][11][12][13][14][15][16][17][18][19][20] Helical peptide foldamers have been used to design (a) antimicrobial peptides (AMPs) with cationic groups [3,4,[21][22][23][24] and (b) catalysts for various organic reactions by incorporating catalytic functional groups. [25][26][27][28][29][30][31] It is well known that the polymer nylon 6 of the ɛ-amino caproic acid (6-aminohexanoic acid, Ahx; Figure 1a) forms fibrils composed of β-sheet-like chain structures. ...
Article
Full-text available
The conformational preferences of oligopeptides of an ϵ-amino acid (2-((1R,3S)-3-(aminomethyl)cyclopentyl)acetic acid, Amc5 a) with a cyclopentane substituent in the Cβ -Cγ -Cδ sequence of the backbone were investigated using DFT methods in chloroform and water. The most preferred conformation of Amc5 a oligomers (dimer to hexamer) was the H16 helical structure both in chloroform and water. Four residues were found to be sufficient to induce a substantial H16 helix population in solution. The Amc5 a hexamer adopted a stable left-handed (M)-2.316 helical conformation with a rise of 4.8 Å per turn. The hexamer of Ampa (an analogue of Amc5 a with replacing cyclopentane by pyrrolidine) adopted the right-handed mixed (P)-2.918/16 helical conformation in chloroform and the (M)-2.416 helical conformation in water. Therefore, hexamers of ϵ-amino acid residues exhibited different preferences of helical structures depending on the substituents in peptide backbone and the solvent polarity as well as the chain length.
... γ-Thiazole based AAs are able to induce pseudo βor a fused β/α-turn depending on the γ-AA absolute configuration. [33] In particular, a series of sequencies bearing a dipeptide formed from L-Phe and S-or R-γ-thiazoles 33 ( Figure 6) were synthesized. ...
Article
Full-text available
The main goal in developing peptidomimetics is the stabilization of the bioactive conformation of peptides. Among all the secondary structures, turns are of paramount importance. This review highlights the recent advances in the design and synthesis of cyclic turn mimics. Both amino acids, carbo‐ and hetero‐cyclic scaffolds have been considered, focusing on their use in the preparation of peptidomimetics and on their ability to induce γ‐, β‐, and α‐turns.
... Over the last years, our group described a class of constrained heterocyclic γ-amino acids built around a thiazole ring, named 4-amino(methyl)-1,3-thiazole-5-carboxylic acids or ATCs (Figure 1a), 35 which were initially used to template an amphipathic hairpin peptide with antibacterial activities analogous to gramicidin S. 36 Furthermore, ATC-containing oligomers showed high propensity to adopt defined structures in both organic solvents and water. 35,37,38 The facial anisotropy of the edifice enables a precise 3D positioning of the appended functionalities, making ATC γ-peptides suitable for material 39 and biomedical applications. 40,41 In line with our research agenda, we herein addressed the prospects of ATC oligomers as a tunable scaffold to design polycationic AMP mimetics. ...
Article
Antimicrobial peptides (AMPs) are amphipathic molecules displaying broad-spectrum bactericidal activity providing opportunities to develop new generation of antibiotics. However their use is limited either by poor metabolic stability or high hemolytic activity. We herein addressed the potential of thiazole-based γ-peptide oligomers named ATCs as tunable scaffolds to design poly-cationic AMPs mimetics. Knowing the side chain distribution along the backbone, we rationally designed facially amphiphilic sequences with bactericidal effect in the micromolar range. Since no hemolytic activity was detected up to 100 µM, this class of compounds has shown the potential for therapeutic development.
Article
Peptides that are composed of an alternating pattern of α- and γ-amino acids are potentially valuable as metabolism-resistant bioactive agents. For optimal function, some kind of conformational restriction is usually...
Article
Full-text available
The Cover Feature shows the conformational preferences of oligomers of δ‐amino acid (δAc5a) with a cyclopentyl constraint in the Cβ−Cγ bond of the backbone investigated by using DFT methods in the gas phase and in solution. In particular, the hexameric Ac‐(δAc5a)6‐NHMe preferentially adopted the mixed H16/14 helical structure in chloroform and the H14 helical structure in water. Pyrrolidine‐substituted analogues also adopted helical structures. More information can be found in the Full Paper by H. S. Park and Y. K. Kang.
Article
Full-text available
The conformational preferences of oligomers of δ‐amino acid (δAc5a) with a cyclopentyl constraint in the Cβ−Cγ bond of the backbone were investigated by using DFT methods in the gas phase and in solution. The folded structures with C10 H‐bonded pseudocycles were most preferred for dimer and tetramer of δAc5a residues both in chloroform and water. However, for the hexameric Ac‐(δAc5a)6‐NHMe, the mixed H16/14 helical structure was found to be most preferred in chloroform (populated at 68 %), whereas the H14 helical structure was the most dominant conformation in water (populated at 60 %). The stability of the former was ascribed to the intrinsic conformational energy, whereas the solvation free energy was crucial to stabilize the latter. Pyrrolidine‐substituted analogues of the hexameric Ac‐(δAc5a)6‐NHMe, with adjacent amine diads that are almost exactly one turn apart with two nitrogen atoms separated by ca. 5.5 Å, adopted helical structures. They are potential catalysts in nonpolar and polar solvents as they have similar structures to a helical 1 : 2 α:β‐heptapeptide that exhibited good catalytic performance in the crossed aldol condensation.
Article
Full-text available
Molecular chirality is ubiquitous in nature. The natural biopolymers, proteins and DNA, preferred a right‐handed helical bias due to the inherent stereochemistry of the monomer building blocks. Here, we are reporting a rare co‐existence of left‐ and right‐handed helical conformations and helix‐terminating property at the C‐terminus within a single molecule of α,γ‐hybrid peptide foldamers composed of achiral Aib (α‐aminoisobutyric acid) and 3,3‐dimethyl‐substituted γ‐amino acid (Adb; 4‐amino‐3,3‐dimethylbutanoic acid). At the molecular level, the left‐ and right‐handed helical screw sense of α,γ‐hybrid peptides are representing a macroscopic tendril perversion. The pronounced helix‐terminating behaviour of C‐terminal Adb residues was further explored to design helix–Schellman loop mimetics and to study their conformations in solution and single crystals. The stereochemical constraints of dialkyl substitutions on γ‐amino acids showed a marked impact on the folding behaviour of α,γ‐hybrid peptides.
Article
Full-text available
Nuclear magnetic resonance (NMR) spectroscopy has been established as a potent method for the determination of foldamer structures in solution. However, the NMR techniques could be limited by averaging, so additional experimental techniques are often needed to fully endorse the folding properties of a sequence. We have recently demonstrated that oligo-γ-peptides composed of 4-amino(methyl)-1,3-thiazole-5-carboxylic acids (ATCs) adopt an original helical fold stabilized by hydrogen bonds forming C9 pseudocycles. The main objective of the present work is to reinvestigate the folding of ATC oligomer 1 in order to identify reliable FT-IR and NMR structural markers that are of value for tracking the degree of organization of ATC-based peptides.
Article
Full-text available
The impact of geometrically constrained cis α,β-unsaturated γ-amino acids on the folding of α,γ-hybrid peptides was investigated. Structure analysis in single crystals and in solution revealed that the cis carbon–carbon double bonds can be accommodated into the 12-helix without deviation from the overall helical conformation. The helical structures are stabilized by 41 hydrogen bonding in a similar manner to the 12-helices of β-peptides and the 310 helices of α-peptides. These results show that functional cis carbon–carbon double bonds can be accommodated into the backbone of helical peptides.
Article
Full-text available
The impact of geometrically constrained cis a,b-unsaturated g-amino acids on the folding of a,g-hybrid peptides was investigated. Structure analysis in single crystals and in solution revealed that the cis carbon–carbon double bonds can be accommodated into the 12-helix without deviation from the overall helical conformation. The helical structures are stabilized by 4!1 hydrogen bonding in a similar manner to the 12-helices of b-peptides and the 3 10 helices of a-peptides. These results show that functional cis carbon–carbon double bonds can be accommodated into the backbone of helical peptides.
Article
α/γ-Peptide foldamers containing either γ(4)-amino acid residues or ring-constrained γ-amino acid residues have been reported to adopt 12-helical secondary structure in nonpolar solvents and in the solid state. These observations have engendered speculation that the seemingly flexible γ(4) residues have a high intrinsic helical propensity and that residue-based preorganization may not significantly stabilize the 12-helical conformation. However, the prior studies were conducted in environments that favor intramolecular H-bond formation. Here, we use 2D-NMR to compare the ability of γ(4) residues and cyclic γ residues to support 12-helix formation in more challenging environments, methanol and water. Both γ residue types support 12-helical folding in methanol, but only the cyclically constrained γ residues promote helicity in water. These results demonstrate the importance of residue-based preorganization strategies for achieving stable folding among short foldamers in aqueous solution.
Article
The conformational control of molecular scaffolds allows the display of functional groups in defined spatial arrangement. This is of considerable interest for developing fundamental and applied systems in both the fields of biology and material sciences. Peptides afford a large diversity of functional groups, and peptide synthetic routes are very attractive and accessible. However, most short peptides do not possess well-defined secondary structures. Herein, we developed a simple strategy for converting peptide sequences into structured γ-lactam-containing oligomers while keeping the amino acids side chain diversity. We showed the propensity of these molecules to adopt ribbon-like secondary structures. The periodic distribution of the functional groups on both sides of the ribbon plane is encoded by the initial peptide sequence.
Article
Structural characterization of new α/γ-peptide foldamers containing the cyclically constrained γ-amino acid I is described. Crystallographic and 2D NMR analysis shows that γ residue I promotes the formation of a 12/10-helical secondary structure in α/γ-peptides. This helix contains two different types of internal H-bond, and the data show that the 12-atom C═O(i) → H-N(i+3) H-bond is more favorable than the 10-atom C═O(i) → H-N(i-1) H-bond. Several foldamer helices featuring topologically distinct H-bonds have been discovered, but our findings are the first to show that such H-bonds may differ in their favorability.
Article
Electron deficient, bivalent sulfur atoms have two areas of positive electrostatic potential, a consequence of the low-lying σ* orbitals of the C-S bond that are available for interaction with electron donors including oxygen and nitrogen atoms and, possibly, π-systems. Intramolecular interactions are by far the most common manifestation of this effect, which offers a means of modulating the conformational preferences of a molecule. Although a well-documented phenomenon, a priori applications in drug design are relatively sparse and this interaction, which is often isosteric with an intramolecular hydrogen-bonding interaction, appears to be underappreciated by the medicinal chemistry community. In this Perspective, we discuss the theoretical basis for sulfur σ* orbital interactions and illustrate their importance in the context of drug design and organic synthesis. The role of sulfur interactions in protein structure and function is discussed and although relatively rare, intermolecular interactions between ligand C-S σ* orbitals and proteins are illustrated.
Article
4-Amino(methyl)-1,3-thiazole-5-carboxylic acids (ATCs) are a new class of constrained heterocyclic γ-amino acids built around a thiazole ring; these compounds are valuable as design mimics of the secondary structures of proteins such as helices, β-sheets, turns, and β-hairpins. We report herein a short and versatile chemical route to orthogonally protected ATCs. The synthesis is centered on cross-Claisen condensations between N-Fmoc-amino acids and sterically hindered 1,1-dimethylallyl acetate. The optimized conditions are compatible with aliphatic, aromatic, acidic, and basic amino acids. The resulting N-Fmoc-β-keto ester intermediates were engaged in a two-step process to give ATCs in 45–90 % yields. The synthetic protocol provides a highly flexible method for the introduction of a wide variety of lateral chains either on the γ-carbon atom or on the thiazole core of the γ-amino acids.
Article
Supramolecular assembly of various artificially folded 12-helical architectures composed of γ4-Val, γ4-Leu and γ4-Phe residues is investigated. In contrast to the 12-helices composed of γ4-Val and γ4-Leu residues, the helices with γ4-Phe residues displayed unique elongated nanotubular architectures. The elongated nanotube assembly was further explored as a template for biomineralization of silver ions to silver nanowires. A comparative study using an analogous α-peptide helix reveals the importance of the spatial arrangement of aromatic side chains along the helical cylinder in a 12-helix. These results suggested that the proteolytically and structurally stable α,γ4-hybrid peptide 12-helices may serve as a new generation of potential templates in the design of functional biomaterials.
Article
The ability of urea-linked oligomers of achiral diamines (achiral analogues of the well-established chiral oligourea foldamers) to adopt helical conformations was explored spectroscopically. Up to four achiral units were ligated either to a well-formed helical trimer or to a single chiral diamine, and the extent to which they adopted a screw-sense preference was determined by NMR and CD. In the best performing cases, a trimeric chiral oligourea and even a single cis-cyclohexanediamine monomer induced folding into a helical conformation.