Vol.2, No.1, 37-48 (2011) Journal of Biophysical Chemistry
doi:10.4236/jbpc.2011.21006
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
Peptoids with aliphatic sidechains as helical structures
without hydrogen bonds and collagen/ inverse-collagen
type structures
Fateh S. Nandel, Avneet Saini
Department of Biophysics, Panjab University, Chandigarh, India; fateh_nandel@yahoo.com
Received 29 September 2010; revised 30 October 2010; accepted 12 November 2010.
ABSTRACT
Aliphatic homo-polypeptoids of NAla, NVal, NIle
and NLeu both in the presence and absence of
protecting groups adopt helical structures
without hydrogen bonds with Φ, Ψ values of ~ 0,
± 90° with trans amide bonds. These structures
are stabilized by carbonyl-carbonyl interactions
and characterized by ~ 3.16 residues per turn
with a pitch of ~ 6.13 Å. It has been shown that
like polyvaline and polyleucine peptides, poly-
peptoids can also be exploited for the con-
struction of potential surfactant like molecules
by incorporating charged amino acid residues
at the N terminal. A single-handed template with
Φ, Ψ values of ~ 0, 90˚ can be attained by in-
corporating L-leu or L-val at the C-terminal of
poly-NIle. Analysis of the simulation results in
water as a function of time reveals that the
opening of helical structures without hydrogen
bonds takes place at sub-picosecond time scale
starting from the N-terminal. This leads to the
formation of collagen or inverse-collagen type
structures (Φ, Ψ ~ -60, 145˚ and 60, -145˚ re-
spectively) stabilized by interactions of water
molecules with the backbone carbonyl groups.
Keywords: Peptoids; Conformation; Hel i cal
Structure without Hydrogen Bonds; Collagen and
Inverse-Co llagen Type Structures
1. INTRODUCTION
Design of polymers and oligomers that mimic com-
plex structures and biological properties of proteins is an
important endeavor with both fundamental and practical
implications as polypeptides themselves are generally
poor drugs, due to their in vivo rapid degradation by
proteases and their immunogenic character [1,2]. To ad-
dress multiple design criteria for applications ranging
from medicinal chemistry to material science, focus is to
identify non-natural chemical scaffolds that recapitulate
the desirable attributes of polypeptides i.e. peptidomi-
metics. These should have good solubility in aqueous
solution, access to facile sequence specific assembly of
monomers containing chemically diverse side chains and
capacity to form stable structures.
‘Peptoids’ offer attractive peptidomimetics [2] as their
backbone is similar to that of peptides. Shifting of the
side chain to the main chain nitrogen atom, shown in
Figure 1, renders the alpha carbon achiral. Peptoid mo-
nomers are linked through polyimide bonds and lack the
amide hydrogen; precluding the formation of hydrogen
bond networks that stabilize peptide secondary structures.
Their conformational behavior can be related to both
proline and glycine residues. These modifications be-
stow protease resistance [3], while allowing suitable
mimicry of the spacing between the critical chemical
functionalities of bioactive peptides. Thus, it remains to
be an area of active research for conformational investi-
gations, analysis of interactions and hence, designing of
molecules. In addition, the efficient sub-monomer solid
phase synthesis gave a major breakthrough in peptoid
research, due to the access to a large number of diverse
primary amines that can be added and that too at low
costs [4].
Consequently, peptoids have found application as: i)
Figure 1. Primary structure of peptoid.
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
38
attractive targets for the drug discovery process [5], ii)
motif for combinatorial strategies [5-10], and iii) to de-
velop peptide mimics for biomedical applications
[11-17]. They have also been exploited for their cell pe-
netrating properties [18] and antifouling action on sur-
faces [19,20]. New approaches to direct the peptoid
backbone towards formation of specific secondary
structures such as helices (a structure in which residues
rotate and rise in a repeating manner along an axis, but
of what type?) are being explored for the discovery of
bioactive peptoid modules [21-26].
In spite of their various applications limited system-
atic structural/conformational data is available on pep-
toids [27-29]. Peptoids comprising of 100% achiral,
aromatic NPhe side chains displayed no net CD. It is not
clear whether the no net CD signal is a result of degen-
erate conformations of opposite handedness or other
structures adopted by the peptoid. It may be mentioned
that even systems having no chiral centers can adopt
well defined helical structures. Peptoids containing
chiral aromatic side chain i.e. NSpe is strongly reminis-
cent of the alpha-helical signal on the basis of CD signa-
tures that are almost similar to that of α-helices in pep-
tides [30,31]. On the other hand NMR and CD studies on
peptoids containing α-chiral aliphatic side chains have
shown features similar to those of polyproline type heli-
ces [28]. On the basis of CD spectroscopic results it is
even proposed that α-chiral aliphatic and α-chiral aro-
matic sidechains form helices of essentially the same
type.
Thus, the conformational behavior of peptoids and the
secondary structure adopted by them remains an active
area of research. In this study we report the conforma-
tional preferences of both homo- and hetero-polypepto-
ids of the type; Ac-(NXaa)n-NMe2 of varying chain
length where ‘NXaa’ is a peptoid residue with the amino
acid side chain attached to the amide nitrogen atom by
keeping the amide bond geometry both as trans or cis.
The conformational results thus obtained, shall aid in the
construction and design of peptoid based biomimetics.
2. METHODOLOGY
Knowledge about the global, local and low energy
minima for the peptoid models Ac-NXaa-NMe2 was
obtained from the Φ, Ψ maps and potential energy
curves that were constructed using standard bond lengths
and bond angles. Input for peptoids was given in terms
of internal co-ordinates i.e. bond lengths, torsion angles
and connectivity of the atoms. Energy calculations were
carried out using the semi-empirical quantum mechani-
cal method PCILO (Pertubative Configuration Interac-
tion using Localised molecular Orbitals) [32] and mini-
mization was done by the variation of torsion angles.
The various conformational states of all constructs were
generated based on the global, local and low energy mi-
nima in the Φ, Ψ maps and i, j curves/maps of model
peptoids and their energies computed. Minimization was
further refined by varying Φ, Ψ, ω and values in the
neighborhood of the minima in steps of 5/2 degrees.
Minima obtained by PCILO calculations are also the
minima at the ab initio level for the usual amino acids
[33] and for dehydroamino acids [34-36]. In addition the
PCILO results [37,38] for the peptides containing usual
and unusual amino acids are in conformity with ab initio
results [39,40] and knowledge based crystallographic
data [41,42]. The charges on various atoms in different
conformations of peptoid models were computed using
the GRINDOL method [43].
Molecular Dynamic (MD) simulations provide great
deal of information regarding the stability of peptoids in
water and to the mobility of the peptoid residues. The
results obtained by quantum mechanics calculations
were used as the starting geometries for simulation stud-
ies using GROMACS 3.3.1 MD software package [44].
The Dundee-PRODRG2 [45] server was used to obtain
the GROMACS topology and coordinate files. Interac-
tion parameters within the design sequence were taken
from GROMOS-96 force field G43a1 [46]. It is worth
mentioning that the simulation results obtained by
GROMOS force field are in good agreement with the
experimental results [29,47 and 48]. Energy of the sys-
tem was minimized by the steepest descent method, us-
ing the convergence criteria of 50 kJ mol-1 followed by
conjugate gradient method with a force constant of 20 kJ
mol-1. Next, the MD run was carried out in vacuum for
20 ns, with a time step of 2 fs using the Leap Frog Algo-
rithm. The temperature was controlled through weak
coupling to a bath of constant temperature [49], using a
coupling time; τp of 0.1ps and a reference temperature;
T0 of 300 K. LINCS algorithm [50] was used to restrict
all bonds to their equilibrium lengths and the center of
mass motion of the system was removed every step to
maintain the effective simulation temperature at 300 K.
For the evaluation of coulomb interactions and Van der
Waals interaction a cut off of 0.9 and 1.0 nm respec-
tively was applied. Long range forces were updated
every 10 fs during generation of the neighbor list. The
Long Range Electrostatic Interactions were calculated
using a Particle Mesh Ewald Summation. Initial veloci-
ties of all atoms were taken from a Maxwellian distribu-
tion at the desired initial temperature. After the vacuum
MD simulation a simple cubic periodic box was set up
using the Simple Point Charge (SPC) Water Model [51].
In order to allow equilibration of solvent around the
model sequence, position of all peptoid residues was
restrained for 20 ps. Finally, MD simulation for 1ns at
300 K, without any restrictions was carried out. The
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
39
Table 1. Conformational results* of the various dipeptoid models.
Φ, ψ, ω ΔE Φ, ψ, ω ΔE Φ, ψ, ω ΔE Φ, ψ, ω ΔE
χi, χj (deg) kcal/mol χi, χj (deg) kcal/mol χi, χj (deg) kcal/mol χi, χj (deg) kcal/mol
Ac-NAla-NMe2 Ac-NVal-NMe2
–2, 92, 180 0 –175, 85, 0 1.8 0, –85, 174 0 100, 135, –3 –1
175 75
–5, –80, 180 0 180, –95, 0 3.7 0, 90, 174 0.4 –70, 170, –10 0.7
175 120
180, –90, 180 1.8 –95, –170, 175 1.5
135
180, 90, 180 1.9 85, 175, 170 2.1
110
120, 180, 180 2
–120, 180, 180 2.4
Ac-NLeu-NMe2 Ac-NIle-NMe2
15, –105, 180 0 –125, –170, 03.2 5, 85, 180 0 –95, –155, 0 2.5
130, 175 –65, 175 145, 170 155, 180
15, 70, 178 1.5 120, –175, 03.5 0, –90, 180 0.3 90, 180, 0 5.2
130, 175 65, 70 145, 170 155, 180
120, 180, 180 2.6 –60, 175, –54 –90, 180, 180 2.9
–110, 55 115, 180 145, 170
–120, 180, 178 2.9 55, –160, 8 4.2 90, 180, 180 4
–5, –60 –115, 60 145, 170
* Φ, ψ values are in bold, ω in italics and χi, χj in normal text.
pressure was controlled using weak coupling with a time
constant of 0.5 ps and a reference pressure of 1 Bar.
3. RESULTS AND DISCUSSION
Conformational results in terms of Φ, Ψ, ω and χ val-
ues of the various model dipeptoids in both cis and trans
amide bond geometries are summarized in Table 1. In
general, there is not much difference in the energy of the
various conformers and hence, with a change in experi-
mental conditions different states may be populated.
Results in Table 1 also reveal that NAla can be popu-
lated in several conformations due to its simple and
small side chain with no branching. Thus, incorporation
of NAla in peptides may lead to disruption of helices as
the energy difference between the states is small and
different states may be populated depending on the en-
vironment. This observation is consistent with the ex-
perimental finding that incorporation of NAla in the oth-
erwise helical antimicrobial peptides decreases the heli-
cal content [52].
In higher mers of all peptoid models it is apparent
from the results in Table 2 that degenerate helical struc-
tures without hydrogen bonds with Φ, Ψ values of ~ 0, +
90˚ are most stable with trans amide bond geometry.
These structures characterized by n = 3.16, with a pitch
of 6.13 Å have also been reported in poly-dehydro ami-
no acids [36]. A molecular view of the models Ac-NVal7-
NMe2 and Ac-NIle7-NMe2 shown in Figure 2 with trans
amide bonds for the conformational states with Φ, Ψ
values ~ 0, -90˚ and 0, 90˚ respectively clearly shows
pore formation with an average diameter of 4.6 Å with
hydrophobic side chains protruding outwards. Such
structures are maximally stable in the hydrophobic en-
vironment masking the carbonyl oxygens that point in-
wards towards the central cavity. Negative charge on
carbonyl oxygen (~ -0.66) may provide a negative po-
tential gradient along the pore facilitating the cation to
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
40
Table 2. Conformational results* for homo-polypeptoid models.
ΔE Φ, ψ,
ω χi, χj (deg) kcal/mol˚
1 2 3 4 5 6 7
Ac-NAla6-NMe2
–15, 105, 178 –25, 120, 173 0, 90, 172 –25, 110, 170 –5, 95, 176 0, 95, 170 0
- - - - - -
5, –95, –179 15, –100, –178 10, –95, –179 15, –105, 180 5, –95, –174 10, –100, –179 0.17
- - - - - -
Ac-NVal6-NMe2
0, 90, 178 –10, 100, 176 –5, 95, 172 –5, 95, 174 –5, 95, 172 –5, 95, 176 0
90 95 95 90 95 90
–5, –85, –176 5, –95, –172 5, –95, –172 5, –95, –172 5, –95, –172 5, –95, –176 0.66
90 145 145 150 145 150
–90, 180, 178 –85, 175, 178 –95, –175, 178 –65, 170, 176 –80, 175, –178 –40, 135, 178 19.53
135 130 150 120 125 125
90, –175, –178 90, 175, 180 85, 180, –176 95, 170, –176 85, 180, –176 45, –145, –178 19.55
100 100 115 100 115 120
NVal6-NH2
–, 180, 176 0, 90, 172 –10, 100, 170–5, 95, 170 –10, 100, 172 0, 90, –176 0
–170 155 95 95 95 90
–, 175, –178 0, –90, 180 30, –120, –174–5, –90, –17630, –120, 1800, –90, 174 0.72
40 145 145 85 145 90
Ac-NLeu6-NMe2
5, 85, 176 5, 90, 170 5, 90, 168 10, 85, 172 0, 90, 176 –5, 95, 176 0
130, 175 120, 180 120, 175 115, 175 115, 175 115, 180
25, –115, –177 30, –120, 172 30, –120, 17735, –120, 174 30, –120, 170 15, –100, 173 3.84
130, 175 85, 75 90, 75 90, 75 90, 80 85, 85
Ac-Nile7-NMe2
5, 85, 174 –5, 95, 172 –5, 95, 172 –5, 95, 174 –5, 95, 172 –5, 95, 176 0, 90, 176 0
150, 165 95, 170 95, 175 90, 170 90, 175 90, 170 95, 170
0, –90, –178 5, –105, –174 5, –95, –174 10, –100, –174 –5, –90, –16630, –120, –174 5, –95, –178 0.25
110, 55 150, 170 150, 165 150, 170 145, 165 140, 170 85, 170
Ac-NLys-NLeu6-NMe2
–90, 160, 180 5, 90, 170 0, 90, 178 10, 85, 180 0, 90, 176 –5, 95, 176 0, 90, 180 0
180, 180, 160, –175 120, 180 120, 175 115, 175 115, 175 115, 180 120, 180
* Φ, ψ values are given in bold, ω in italics and χi, χj in normal text.
pass through in a single file. Thus, like Grami-
cidin A, these peptoids may be exploited for the con-
struction and design of channels in membranes. It is also
obvious from the results in Table 2 that presence or ab-
sence of protecting groups hardly affected the nature of
the most stable conformation adopted by poly-NVal.
These helical structures without hydrogen bonds are
stabilized by carbonyl-carbonyl interactions between
carbonyl oxygen of ith residue and carbonyl carbon of
ith + 1 with dOi…Ci+1 being 2.18 Å. A careful look at the
data from penta peptoid onwards revealed that in the
conformation with Φ, Ψ values of ~ 0, 90˚ all ω values
lie between 168 ω 180˚ and in the conformation with
Φ, Ψ values of ~ 0, - 90˚ they lie between 180 ω
194˚.The deviation in ω values resulted in stronger car-
bonyl interactions relative to those in complete trans
amide bond geometries. In addition, C-H…O interac-
tions between the C = O of ith residue and HCα-N of the
side chain of ith + 3 residue (dO…H and dO…C being ~ 2.1
and 3.0 Å) leads to the formation of an eleven membered
ring (Figure 2). Based on a systematic study between
ketonic groups in the Cambridge structure database car-
bonyl-carbonyl interactions [53,54] have been modeled
by three main types of interaction motifs. Importance of
carbonyl interactions as a stabilizing factor in α-helices,
β-sheets and right-handed twist is well-documented [55,
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
41
Figure 2. Molecular view of (a) Ac-NVal7-NMe2 in the con-
formation state with Φ, Ψ, ω values of ~ 0, -90, 180˚depicting
carbonyl-carbonyl interactions and formation of an eleven
member ring due to C-H…O interactions between the ith and ith
+ 3 residue; (b) Ac-NIle7-NMe2 with Φ, Ψ , ω values of ~ 0, 90,
180˚showing the formation of a hydrophilic pore with the
masking of the carbonyl groups by the hydrophobic side
chains.
56]. They also stabilize the partially allowed Ramachandran
conformations of aspartic acid and asparagines [55] and
helical structures without hydrogen bonds in peptides con-
structed from achiral and unusual amino acids [36,58-59].
Degenerate helical structures without hydrogen bonds
with opposite handedness observed in polypeptoid mod-
els of NVal/NIle having branching at the N-Cα position
implies population of both states to the same extent and
hence, very well explains the very weak/no signal in CD
spectroscopy for aliphatic peptoids [28]. Interestingly,
for poly-NLeu only one state with Φ, Ψ values of ~ 0, 90
° is predicted; possibly due to the side chain branching at
the N-Cβ position. Thus, the difference in the branching
pattern changes the local environment and may be re-
sponsible for the difference in the conformational be-
havior of peptoid models.
Φ, Ψ values of 0, 90˚ appear to be unusual even in
peptides, but this region has been predicted a minima for
some amino acids [36]. Also, for the dipeptides of glycine
& alanine [60] a stationary point near Φ = 0˚ , Ψ = 90˚ at
the HF/3.21G and HF/6.31+G levels has been reported.
Ramachandran plots based on; i) PDB- 40 dataset [61]
corresponding to X-ray protein structures with the resolu-
tion of 2.5 Å or better for 470 proteins, 95778 total resi-
dues plotted, proline and glycine excluded and (ii) NMR
derived structure for 113 proteins, 84719 total residues
plotted, (proline and glycine excluded) show appreciable
density between the left-handed helical region and the
collagen-type structural region [62,63] and between the
right-handed helical region and the inverse-collagen type
structural region [60,61]. Though, the dataset for the
NMR derived structures is small compared to the X-ray
protein structures, yet in the mentioned regions the data
point density is more for NMR derived structures i.e. in
the solution phase. Thus, the conformational states with Φ,
Ψ values of ~ 0, 90˚ are not an over-estimation and may
be realized in solvents with low polarity.
4. CONSTRUCTION AND DESIGN
OF NANO-STRUCTURES
Single Handed Template: Alipihatic polypeptoids
were realized in degenerate helical structures without
hydrogen bonds with rise & rotation per residue of 1.94
Å and 114˚ respectively. To construct a template with a
given handedness from these peptoid residues, either
L/D-Leu or L/D-Val residue was incorporated either at
the N-terminal or the C-terminal of the poly Nile/Nval
peptoid sequences of required length. Incorporation of
such chiral amino acids has been shown to control the
screw sense of helical peptides constructed from achiral
residues [36,58,62-63]. Therefore, these amino acids
(L/D-Val or L/D-Leu having branching in the side chain)
were incorporated in achiral peptoids to control the screw
sense of helical structures. The conformational results
thus obtained are given in Table 3 and it is obvious that
in poly-Nile peptoid models the degeneracy was lifted
and only one conformation with Φ, Ψ values of ~ 0, 90˚
was realized by incorporating L-leu or L-val at the
C-terminal. Incorporation of L or D-leu either at the N or
C-terminal of poly-NVal sequence models did not lift the
degeneracy. This may be attributed to the local environ-
ment around N-Cα, which is asymmetrical in NIle resi-
dues and symmetrical in NVal residues. Thus, NIle with
its asymmetrical side chain plays a vital role in lifting
the degeneracy and hence, construction of a single
handed template.
Peptoids as Surfactants: The helical structures
adopted by homo-polypeptoid models (NVal, NLeu, NIle)
appear similar (although with different Φ, Ψ values) to
the corresponding helical structures in polypeptides of
leucine, valine, isoleucine and norleucine with the
non-polar side chains projecting outwards forming
well-defined hydrophobic structures with a hydrophilic
pore. Poly-leucine, poly-valine, poly-isoleucine and
poly-norleucine like peptides have shown surface activ-
ity that helped in accelerating the surface spreading at
the air-water interface and thus, exhibited improved dy-
namic activity. Poly-leucine peptides or peptides rich in
leucine have been exploited for their surfactant like
properties especially in lung surfactant proteins SP-B
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
42
Table 3. Construction of a single handed template for Ac-L-leu-(NIle)6-NMe2.
Φ, ψ, ω
χi, χj (deg)
ΔE
kcal/mol
1 2 3 4 5 6 7
Ac-L-leu-(NIle)6-NMe2
–100, 130, –92 10, 80, 176 0, 95, 170 –5, 95, 172 –5, 100, 172 –5, 95, 172 0, 90, 174 0
–55, 170 155, 170 90, 175 95, 175 95, 175 95, 170 90, 175
–155, 150, 90 0, –90, –178 –5, –90, –172 5, –95, –176 25, –120, –174 0, –90, –176 0, –90, –178 1.99
55, 150 150, 170 155, 160 150, 170 150, 170 90, 170 150, 170
Ac-(NIle)6-L-leu-NMe2
15, 90, 90 0, 95, 170 0, 95, 170 –15, 100, 1760, 90, 172 15, 70, 178 –20, 115, 174 0.68
140, 170 100, 170 95, 170 95, 165 90, 170 90, 175 –65, 170
–10, –80, –86 10, –100, –178 0, –95, –170 25, –120, –1760, –90, –176 –5, –75, 176 180, 160, –178 7.00
130, 50 145, 175 150, 165 140, 175 90, 170 155, 170 60, 140
Ac-(NIle)6-L-val- NMe2
5, 85, 178 –5, 100, 170 5, 90, 166 –15, 100, 176 0, 90, 172 5, 75, –176 –10, 105, 174 0
145, 170 90, 170 100, 165 90, 160 90, 170 90, 170 180
–5, –85, –178 15, –105, –176 0, –95, –166 0, –95, –170 0, –90, –172 –5, –75, 176 –30, 115, –170 3.56
115, 60 145, 170 150, 170 150, 160 150, 170 155, 170 50
and SP-C [13,14]. On similar lines the corresponding
peptoid mimics are constructed to design surfactant like
molecules by incorporation of NLys residue at the
N-terminal in poly-NLeu and the conformational results
are summarized in Table 2. A graphical view of the
peptoid in the most stable state as shown in Figure 3
clearly depicts that NLys residue resembles the polar
head group and the NLeu residues formed the hydro-
phobic tail like structure similar to that of lipids.
5. SIMULATION STUDIES
Simulation studies on aliphatic peptoids reveals almost
Figure 3. A graphical view of the surfactant peptoid design
‘Ac-NLys-(NLeu)6-NMe2’.
similar conformational results irrespective of the initial
geometry taken for simulations (i.e. with Φ, Ψ = 0,
90˚ ; 60, 180˚ ; or 120, 180˚ ) and the side chain.
This means shifting of the side chain from the carbon
alpha to the amide nitrogen makes the peptoid backbone
less flexible and this observation is consistent with the
experimental and computational results [29,64-65].
A molecular view of the model Ac-NVal7-NMe2 after
1ns simulation in water (with starting geometry having
Φ, Ψ values of ~ 0, 90˚ and ~ 0, - 90˚ ) is shown in Fig-
ure 4. It is apparent from the results in Table 4 and the
molecular graphics shown in Figures 4 & 5; that the
interactions of water molecules with the carbonyl oxy-
gen of the backbone lead to the population of structures
with average Φ, Ψ values around -60, 145˚ and 60, -145˚.
The conformation with Φ, Ψ values of ~ -60, 145˚ is
called collagen type structure [66] and the structure with
inverse Φ, Ψ values of collagen type structure i.e. 60,
-145˚ is named as inverse-collagen type structure. Col-
lagen type structures are also well documented in globu-
lar proteins [67,41]. The distance between the carbonyl
oxygen of the peptoid backbone and hydrogen atoms of
water molecules are found to lie between 1.5 to 1.9 Å.
The angle/OHO(C) are observed to lie in the range from
150 to 180˚. This observation is consistent with the ex-
perimental finding that the hydrogen bond angle in bio-
logical systems is never 180˚ but less [68]. In QM cal-
culations both collagen and inverse-collagen type struc-
tures are predicted to be higher in energy (Table 4).
Helical structures with Φ = ± 79.2˚, Ψ = ± 174.6˚ with
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
43
Table 4. Torsion angles for the peptoids in trans amide bond geometry after 1ns simulation with different starting geometries having
1) Φ, Ψ values of ~ 0˚, - 90˚and 2) Φ, Ψ values of ~ 0˚, 90˚.
Φ ψ ω χi, χj (deg) Φ ψ ω χi, χj (deg)
Ac-NAla7-NMe2 Ac-NVal7-NMe2
1 72.3 –159.2 178.2
1 56.3
131.9–163.0 108.1
107.7 –153.3 175.2
61.0
143.0–170.8 121.6
70.5 –113.8 157.7
79.6
137.7 158.4 138.0
55.9 –120.1 175.3
63.4
122.6–174.7 126.7
57.7 –128.4 177.2
64.9
125.6 177.1 122.3
92.4 –137.1 159.9
77.5
151.2–175.1 122.7
72.9 –138.5 179.3
61.4 99.0 –172.5 113.3
2 –63.8 148.8 –162.6
2 –49.7 113.9 178.3 –104.2
–79.4 134.9 –174.4
–59.4 127.7 –174.0 –118.8
–117.7 148.1 –173.6
–66.4 153.8 166.6 –123.4
–83.5 –146.0 163.5
–58.5 128.7 168.8 –116.0
–90.5 –164.7 163.8
–75.4 125.9 –179.3 –120.9
–104.3 –156.7 160.2
–44.9 124.1 –171.8 –121.4
70.4 –117.5 157.2
–76.1 145.4 175.1 –104.6
Ac-NLeu7-NMe2 Ac-NIle7-NMe2
1 79.8 166.6 –169.6 95.5, –88.4
1 65.3 –154.6
–176.1 103.6, 69.4
46.8 –117.7 –176.2 –90.5, 148.2
57.8 –127.4
–162.7 112.6, 71.6
65.6 –162.9 –165.7 –145.6, –157.4
78.6 –155.4
–175.3 103.0, 110.6
81.1 –107.2 168.0 105.1, –79.0
64.3 –142.8
–189.6 124.8, 87.6
61.4 –108.2 179.6 –110.7, –162.9
65.9 –135.4 177.8 97.3, 87.4
55.3 –136.2 172.4 –97.3, 78.7
89.4 169.3
–168.7 102.9, 143.5
94.1 –152.8 163.2 131.6, –84.2
66.8 –120.6
–175.0 108.0, 76.5
2 –68.2 133.9 –170.5 113.5, –62.1
2 –87.8 –169.7 167.9 109.8, 87.9
–67.9 145.6 179.5 118.1, –83.6
–61.6 110.4 177.6 131.1, 152.9
–57.3 123.4 –176.0 119.7, –85.9
–75.1 146.8 –176.1 85.9, 153.6
–72.4 129.6 –160.9 85.1, –117.3
–61.8 152.9 174.1 123.7, –174.6
–64.6 142.5 –170.3 107.9, –70.7
–64.5 145.7 174.3 109.4, 95.6
–76.3 157.2 179.4 –118.9, –98.8
–50.8 135.3 –174.1 110.6, 76.9
–52.5 152.8 –174.9 115.5, –84.1
–67.0 137.1 176.5 104.9, 95.7
Table 5. Torsion angles for the model Ac-NVal7-NMe2 at different simulation time in water with starting conformation of Φ, Ψ, ω ~
0˚, -90˚, 180˚.
Φ, Ψ
Residue number
Time/ ps
1 2 3 4 5 6 7
0.5 57.7, –149.1 68.8, –134.7 23.2, –88.7 12.8, –100.5 36.6, –112.7 17.4, –110.4 –20.9, –67.1
1.0 70.0, –152.6 62.5, –141.5 53.5, –128.9 43.2, –103.7 20.2, –96.2 68.4, –134.6 –27.7, –70.9
5.0 51.6, –137.2 70.8, 178.9 67.6, –125.4 66.5, –119.5 37.3, –100.6 62.9, –126.4 –30.2, –73.7
10.0 49.9, –123.8 73.2, 178.0 63.1, –140.2 46.8, –120.1 48.1, –133.7 67.0, –129.7 18.7, –100.6
1000.0 58.3, –131.9 61.0, –143.0 79.6, –137.7 63.4, –122.6 64.9, –125.6 77.5, –151.2 –61.4, 99.0
trans amide bond geometry for sarcosine have also been
reported by ab initio calculations using dielectric con-
stants to replicate an aqueous environment [69] but not
in explicit solvent. Our results are also consistent with
the peptoid backbone conformational landscape de-
scribed by Butterfoss et al. using ab initio Gaussian03
package [29].
Opening of Helical Structure without hydrogen
bonds: To gain insight into the opening of the helical
structures, whether opening starts from the N or C-terminal
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
44
(a)
(b)
Figure 4. Molecular view of the (a) Collagen and (b) in-
verse-collagen type structures observed in the model
‘Ac-NVal7-NMe2 after 1ns simulation in water with starting
geometry having Φ, Ψ values of ~ 0, 90˚and 0, -90˚ respec-
tively. As evident, such structures are stabilized by the interac-
tions of water molecules with carbonyl moieties of the peptoid
backbone. For clarity purposes, water molecules within 3 Å of
the peptoid surface are shown.
a pictorial view of the peptoid Ac-(NVal)7-NMe2 at dif-
ferent simulation times with water molecules within 3 Å
of the peptoid surface is shown in Figure 6. It is appar-
ent from the graphics that the opening of the helical
structure into the collagen or inverse-collagen type
structure begins at the N-terminal due to initial interac-
tions of water molecules with the acetyl carbonyl oxygen
that lead to the change of torsion angles. Analysis of the
simulation results at different time intervals given in
Table 5 reveal that opening of the helical structure with-
out hydrogen bonds takes place at the sub picoseconds
time scale. The number of hydrogen bonds formed be-
tween water molecules and the carbonyl moieties attain
the maximum val ues within 10 ps and thereafter remain
Figure 5. Plot of total energy (E) as a function of simula-
tion time for the model Ac-(NIle)7-NMe2 (black) and
Ac-(NLeu)7-NMe2 (red) for the collagen type conforma-
tions.
(
a
)
(
c
)
(b)
Figure 6. Snapshots of the peptoid ‘Ac-(NVal)7-NMe2 at dif-
ferent simulation intervals (a) 1ps, (b) 5ps and (c) 10ps with
water molecules within 3 Å of the peptoid surface indicates
that the opening of the helical structure without hydrogen
bonds starts at the N-terminal.
constant throughout the simulation period (Figure 7).
6. CONCLUSIONS
Aliphatic polypeptoid models adopt degenerate helical
structures without hydrogen bonds with Φ, Ψ values of ~
0, 90˚ in trans amide bond geometry that are stabilized
(a)
(b)
Figure 7. Number of hydrogen bonds formed between car-
bonyl groups of the backbone and water molecules for the
model ‘Ac-(NVal)7-NMe2’ (a) increases within first 10 ps and
thereafter, (b) remains constant on simulation in water.
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
45
by carbonyl interactions. The population of poly-NLeu
in only one state with Φ, Ψ values of ~ 0, 90˚ is attrib-
uted to the difference in the side chain branching pat-
terns of these residues. A single handed template was
realized by incorporating L-leu or L-val at the C-termi-
nal of poly-NIle sequences. Such templates provide pat-
terns and find utility in generating a comple- mentary
molecule. In peptoids, helical structures without hydro-
gen bonds can also be exploited for the design of surfac-
tant molecules by incorporating charged residues at the
terminal positions. Simulation studies reveal that inter-
actions of water molecules with the carbonyl moieties of
peptoid backbone act as the primary driving force behind
the opening of helical structures without hydrogen bonds
resulting in the population of collagen or in-
verse-collagen type structures.
7. ACKNOWLEDGMENTS
We gratefully acknowledge the funding for this project by the De-
partment of Biotechnology, India.
REFERENCES
[1] Latham, P.W. (1999) Therapeutic Peptides Revisited.
Nature Boitechnology, 17, 755-757.
doi:10.1038/11686
[2] Patch, J.A. and Barron, A.E. (2002) Mimicry of bioactive
peptides via non-natural, sequence-specific peptidomi-
metic oligomers. Current Opinion in Chemical Biology,
6(6), 872-877.
doi:10.1016/S1367-5931(02)00385-X
[3] Miller, S.M., Simon, R.J., Ng, S., Zuckermann, R.N.,
Kerr, J.M. and Moos, W.H. (1995) Comparison of the
proteolytic susceptibilities of homologous L-amino acid,
D-amino acid and N-substituted glycine peptides and
peptoid oligomers. Drug Developement Research, 35(1),
20-32.
doi:10.1002/ddr.430350105
[4] Zuckermann, R.N., Kerr, J.M., Kent, S.B.H. and Moos,
W.H. (1992) Efficient method for the preparation of pep-
toids [oligo (N-substituted) glycines] by submonomer
solid-phase synthesis. Journal of American Chemical So-
ciety, 114(26), 10646-10647.
doi:10.1021/ja00052a076
[5] Zuckermann, R.N., Martin, E.J., Spellmeyer, D.C., Stau-
ber, G.B., Shoemaker, K.R., Kerr, J.M., Flgliozzi, G.M.,
Goff, D.A., Siani, M.A., Simon, R.J., Banville, S.C.,
Brown, E.G., Wang, L., Richter, L.S. and Moos, W.H.
(1994) Discovery of nanomolar ligands for
7-transmembrane G-Protein coupled receptors from a
diverse (N-substituted) glycine peptoid library. Journal
of Medicinal Chemistry, 37(17), 2678-2685.
doi:10.1021/jm00043a007
[6] Gibbons, J.A., Hancock, A.A., Vitt, C.R., Knepper, S.,
Buckner, S.A., Brune, M.E., Milicic, I., Kerwin, J.F.,
Richter, L.S., Taylor, E.W., Spear, K.L., Zuckermann,
R.N., Spellmeyer, D.C., Braeckman, R.A. and Moos,
W.H. (1996) Pharmacologic characterization of CHIR
2279, an N-substituted glycine peptoid with high-affinity
binding for alpha 1-adrenoceptors. Journal of Pharma-
cology and Experimental Therapeutics, 277(2), 885-899.
[7] Ng, S., Goodson, B., Ehrhardt, A., Moos, W.H., Siani, M.
and Winter, J. (1999) Combinatorial discovery process
yields antimicrobial peptoid. Bioorganic and Medicinal
Chemistry, 7(9), 1781-1785.
doi:10.1016/S0968-0896(99)00132-7
[8] Garcia-Martinez, C., Humet, M., Planells-Cases, R.,
Gomis, A., Caprini, M., Viana, F., Pena, E.D, San-
chez-Baeza, F., Carbonell, T., Felipe, C.D., Perez-Paya,
E., Belmonte, C., Messeguer, A. and Ferre-Montiel, A.
(2002) Attenuation of thermal nociception and hyperal-
gesia by VR1 blockers. Proceedings of the National
Academy of Sciences USA., 99(4), 2374-2379.
doi:10.1073/pnas.022285899
[9] Heizmann, G., Hildebrand, P., Tanner, H., Ketterer, S.,
Panksy, A., Froidevaux, S., Beglinger, C. and Eberle,
A.N. (1999) A combinatorial peptoid library for the iden-
tification of novel MSH and GRP/Bombesin receptor li-
gands. Journal of Receptors and Signal Transduction
Research, 19(1-4), 449-466.
doi:10.3109/10799899909036664
[10] Thompson, D.A., Chai, B.X., Rood, H.L.E., Siani, M.A.,
Douglas, N.R., Grantiz, I. and Millhauser, G.L. (2003)
Peptoid mimetics of agouti related protein. Bioorganic
and Medicinal Chemistry Letters, 13(8), 1409-1413.
doi:10.1016/S0960-894X(03)00164-1
[11] Statz, A.R., Park, J.P., Chongsiriwatana, N.P. , Barron,
A.E. and Messersmith, P.B. (2008) Surface-immobilized
antimicrobial peptoids. Biofouling, 24(6), 439-448.
doi:10.1080/08927010802331829
[12] Brown, N.J., Wu, C.W., Seurynck-Servoss, S.L. and
Barron, A.E. (2008) Effects of hydrophobic helix length
and side chain chemistry on biomimicry in peptoid ana-
logues of SP-C. Biochemistry, 47(6), 1808-1818.
doi:10.1021/bi7021975
[13] Seurynck-Servoss, S.L., Brown, N.J., Dohm, M.T., Wu,
C.W. and Barron, A.E. (2007) Lipid composition greatly
affects the in vitro surface activity of lung surfactant
protein mimics. Colloids & Surfaces B: Biointerfaces,
57(1), 37-55.
doi:10.1016/j.colsurfb.2007.01.001
[14] Seurynck-Servoss, S.L., Dohm, M.T. and Barron, A.E.
(2006) Effects of including an N-terminal insertion re-
gion and arginine-mimetic side chains in helical peptoid
analogues of lung surfactant protein B. Biochemistry,
45(39), 11809-11811.
doi:10.1021/bi060617e
[15] Rouhi, A.M. (2001) A peptoid promise: Synthetic lung
surfactant mimics may widen access to treatment of res-
piratory disease. Chemical and Engineering News, 79,
50-51.
doi:10.1021/cen-v079n038.p050
[16] Seurynck, S.L., Patch, J.A. and Barron, A.E. (2004) Sim-
ple, helical peptoid analogs of lung surfactant protein B.
Chemistry and Biology, 12(1), 77-88.
doi:10.1016/j.chembiol.2004.10.014
[17] Wu, C.W., Seurynck, S.L., Lee, K.Y.C. and Barron, A.E.
(2003) Helical peptoid mimics of lung surfactant protein
C. Chemistry and Biology, 10(11), 1057-1063.
doi:10.1016/j.chembiol.2003.10.008
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
46
[18] T.M. Rossa, R.N. Zuckermann, C.R. William and H. Frey,
“Intranasal administration delivers peptoids to the rat
central nervous system”, Neuroscience Letters, Vol. 439,
No. 1, 2008, pp. 30-33.
doi:10.1016/j.neulet.2008.04.097
[19] A.R. Statz, R.J. Meagher, A.E. Barron and P.B. Mess-
ersmith, “New peptidomimetic polymers for antifouling
surfaces”, Journal of the American Chemical Society, Vol.
127, No. 22, 2005, pp. 7972-7973.
doi:10.1021/ja0522534
[20] R. Statz, A.E. Barron and P.B. Messersmith, “Protein,
cell and bacterial fouling resistance of polypep-
toid-modified surfaces: Effect of side-chain chemistry”,
Soft Matter, Vol. 4, 2008, pp. 131-139.
doi:10.1039/b711944e
[21] B. Yoo and K. Kirshenbaum, “Peptoid architectures:
Elaboration, actuation and application”, Current Opinion
in Chemical Biology, Vol. 12, No. 6, 2008, pp. 714-21.
doi:10.1016/j.cbpa.2008.08.015
[22] T.S. Ryge, N. Frimodt-Møller and P.R. Hansen, ” Antim-
icrobial Activities of Twenty Lysine-Peptoid Hybrids
against Clinically Relevant Bacteria and Fungi”, Che-
motherapy, Vol 54, No. 2, 2008, pp. 152-156.
doi:10.1159/000119707
[23] D. Daelemans, D. Schols, M. Witvrouw, C. Pannecouque,
S. Hatse, S.V. Dooren, F. Hamy, T. Klimkait, E. Clercq
and A.M. Vandamme, “A second target for the peptoid
Tat/transactivation response element inhibitor CGP64222:
Inhibition of human immunodeficiency virus replication
by blocking CXC-chemokine receptor 4-mediated virus
entry”, Molecular Pharmacology, Vol. 57, No. 1, 2000,
pp. 116-124.
[24] F. Hamy, E.R. Felder, G. Heizmann, J. Lazdins, F. Ab-
oul-Ela, G. Varani, J. Karn and J. Klimkait, “An inhibitor
of the Tat/TAR interaction that effectively suppresses
HIV-1 replication”, Proceedings of the National Academy
of Sciences USA, Vol. 94, No. 8, 1997, pp. 3548-3553.
doi:10.1073/pnas.94.8.3548
[25] J.A. Patch and A.E. Barron, “Helical peptoid Mimics of
Magainin-2 amide”, Journal of American Chemical So-
ciety, Vol. 125, No. 40, 2003, pp. 12092-12093.
doi:10.1021/ja037320d
[26] N.P. Chongsiriwatana, J.A. Patch, A.M. Czyzewski, M.T.
Dohm, A. Ivankin, D. Gidalevitz, R.N. Zuckermann and
A.E. Barron, “Peptoids that mimic the structure, function,
and mechanism of helical antimicrobial peptides”, Pro-
ceedings of the National Academy of Sciences USA, Vol.
105, No. 8, 2008, pp. 2794-2799.
doi:10.1073/pnas.0708254105
[27] S.A. Fowler, R. Luechapanichkul and H.E. Blackwell,
“Structure-function relationships in peptoids: Recent ad-
vances toward deciphering the structural requirements
for biological function”, Journal of Organic Chemistry,
Vol. 74, No. 4, 2009, pp. 1440-1449.
doi:10.1021/jo8023363
[28] C.W. Wu, K. Kirshenbaum, T.J. Sanborn, K. Huang, R.N.
Zuckermann, A.E. Barron, J.A. Patch and K.A. Dill,
“Structural and spectroscopic studies of peptoid oli-
gomers with -chiral, aliphatic side chains”, Journal of
American Chemical Society, Vol. 125, No. 44, 2003, pp.
13525-13530.
doi:10.1021/ja037540r
[29] G.L. Butterfoss, P.D. Renfren, B. Kuhlman and K.A.
Kirshenbaum, “Preliminary survey of the peptoid folding
landscape”, Journal of American Chemical Society, Vol.
131, No. 46, 2009, pp. 16798-16807.
doi:10.1021/ja905267k
[30] C.W. Wu, T.J. Sanborn, R.N. Zuckermann and A.E. Bar-
ron, “Peptoid oligomers with α-chiral, aromatic side
chains: effect of chain length on secondary structure”,
Journal of American Chemical Society, Vol. 123, No.13,
2001, pp. 2958-2963.
doi:10.1021/ja003153v
[31] W. Wu, T.J. Sanborn, K. Huang, R.N. Zuckermann and
A.E. Barron, “Peptoid oligomers with -chiral, aromatic
side chains: Sequence requirements for the formation of
stable peptoid helices”, Journal of American Chemical
Society, Vol. 123, No. 28, 2001, pp. 6778-6784.
doi:10.1021/ja003154n
[32] B. Pullman and A. Pullman, “Molecular orbital calcula-
tions on the conformations of amino acid residues of
proteins”, Advances in Protein Chemistry, Vol. 28, 1974,
pp. 347-526.
doi:10.1016/S0065-3233(08)60233-8
[33] R.P. Lawrence and C. Thompson, “The boron analogue
of glycine: A theoretical investigation of structure and
properties”, Journal of Molecular Structure: Theochem,
Vol. 88, No. 1-2, 1982, pp. 37-43.
doi:10.1016/0166-1280(82)80105-X
[34] C. Aleman and J. Casanovas, “Ab initio SCF and
force-field calculations on low-energy conformers of
2-acetylamino-2,N-dimethylpropanamide”, Journal of
the Chemical Society of Perkin Transactions, Vol. 2,
1994, pp. 563-568.
doi:10.1039/p29940000563
[35] C. Aleman and J. Casanovas, “Molecular conformational
analyses of dehydroalanine analogues”, Biopolymers, Vol.
36, No. 1, 1995, pp. 71-82.
doi:10.1002/bip.360360107
[36] F.S. Nandel and B. Khare, “Conformation of peptides
constructed from achiral amino acid residues Aib and
ZPhe: Computational study of the effect of L/D- Leu at
terminal positions”, Biopolymers, Vol. 77, No. 1, 2005,
pp. 63-73.
doi:10.1002/bip.20128
[37] F.S. Nandel, N. Malik, B. Singh and D.V.S. Jain, “Con-
formational structure of peptides containing dehydroa-
lanine. Formation of beta bend ribbon structure”, Inter-
national Journal of Quantum Chemistry, Vol. 72, No. 1,
1999, pp. 15-23.
doi:10.1002/(SICI)1097-461X(1999)72:1<15::AID-QUA
2>3.0.CO;2-2
[38] F.S. Nandel, N. Malik, B. Singh and M. Virdi, “Design-
ing of peptides with left handed helical structure by in-
corporating the unusual amino acids”, Indian Journal of
Biochemistry and Biophysics, Vol. 36, No. 3, 1999, pp.
195-203.
[39] S.J. Weiner, P.A. Kollman, D.T. Nguyen and D.A. Case,
“An all atom force field for simulations of proteins and
nucleic acids”, Journal of Computational Chemistry, Vol.
7, No. 2, 1986, pp. 230-252.
doi:10.1002/jcc.540070216
[40] K. Mohle, H.J. Hoffman, “Secondary structure formation
in N-substituted peptides”, J. Peptide Research, Vol. 51,
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
47
No. 1, 1998, pp. 19-28.
doi:10.1111/j.1399-3011.1998.tb00412.x
[41] A.A. Adzhubei, F. Eisenmenger, V.G. Tumanyan, M.
Zinke and N. G. Esipova, “Third type of secondary
structure: Noncooperative mobile conformation. Protein
Data Bank analysis”, Biochemical and Biophysical Re-
search Communications, Vol. 146, No. 3, 1987, pp.
934-938.
doi:10.1016/0006-291X(87)90736-4
[42] A.A. Aduzbei and J.E. Sternberg, “Left-handed poly-
proline II helices commonly occur in globular proteins”,
Journal of Molecular Biology, Vol. 229, No. 2, 1993, pp.
472-493.
doi:10.1006/jmbi.1993.1047
[43] J. Lipinski, “Modified all-valence INDO/spd method for
ground and excited state properties of isolated molecules
and molecular complexes”, International Journal of
Quantum Chemistry, Vol. 34, No. 5, 1988, pp. 423-435.
doi:10.1002/qua.560340504
[44] D. van der Spoel, E. Lindahl, B. Hess, A.R. van Buuren,
E. Apol, P.J. Meulenhoff, D.P. Tieleman, A.L.T.M. Si-
jbers, K.A. Feenstra, R. van Drunen and H.J.C. Berend-
sen, Gromacs User Manual version 3.3, 2005.
[45] A.W. Schuettelkopf and D.M.F. van Aalten, “PRODRG
-a tool for high-throughput crystallography of pro-
tein-ligand complexes” Acta Crystallographica, Vol. D60,
No. 8, 2004, pp. 1355-1363.
[46] W.F. van Gunsteren, S.R. Billeter, A.A. Eising, P.H.
Hünenberger, P. Krüger, A.E. Mark, W.R.P. Scott and
I.G. Tironi, “Biomolecular Simulation: The GROMOS96
manual and user guide. Z¨urich, Switzerland: Hoch-
schulverlag AG an der ETH Zürich, 1996.
[47] N.H. Shah, G.L. Butterfoss, K. Nguyen, B. Yoo, R. Bon-
neau, D.L. Rabenstein and K. Kirshenbaum, “Oligo
(N-aryl glycines): A new twist on structural peptoids”,
Journal of American Chemical Society, Vol. 130, No. 49,
2008, pp. 16622-16632.
doi:10.1021/ja804580n
[48] Q. Sui, D. Borchardt and D.L. Rabenstein, “Kinetics and
equilibria of cis/trans isomerization of backbone amide
bonds in peptoids”, Journal of American Chemical Soci-
ety, Vol. 129, No. 39, 2007, pp. 12042-12048.
doi:10.1021/ja0740925
[49] H.J.C. Berendsen, J.P.M. Postma, A. DiNola and J.R.
Haak, “Molecular dynamics with coupling to an external
bath”, Journal of Chemical Physics, Vol. 81, No. 8, 1984,
pp. 3684-3690.
doi:10.1063/1.448118
[50] B. Hess, H. Bekker, H.J.C. Berendsen and J.G.E.M.
Fraaije, “LINCS: A linear constraint solver for molecular
simulations”, Journal of Computational Chemistry, Vol.
18, No. 12, 1997, pp. 1463-1472.
doi:10.1002/(SICI)1096-987X(199709)18:12<1463::AID
-JCC4>3.3.CO;2-L
[51] H.J.C. Berendsen, J.P.M. Postma, W.F. van Gunsteren
and J. Hermans, “Interaction models for water in relation
to protein hydration” In: Intermolecular Forces. Dor-
drecht: D Reidel Publishing Company, 1981, pp.
331-342.
[52] Y.M. Song, Y.K. Park, S.S. Lim, S.T. Yang, E.R. Woo,
I.S. Park, J.S. Lee, J.I. Kim, K.S. Hahm, Y. Kim and S.U.
Shin, “Cell selectivity and Mechanism of action of an-
timicrobial model peptides containing peptoid residues”,
Biochemistry, Vol. 44, No. 36, 2005, pp. 12094-106.
doi:10.1021/bi050765p
[53] F.H. Allen, C.A. Baalham, J.P.M. Lommerse and P.R.
Raithby, “Carbonyl-carbonyl interactions can be com-
petitive with hydrogen bonds”, Acta Crystallographica,
Vol. B54, No. 3, 1998, pp. 320-329.
[54] F.H. Allen, J.E. Dacies, J.J. Galloy, O. Johnson, O. Ken-
nard, C.F. Macrae, E.M. Mitchell, G.F. Mitchell, J.M.
Smith and D.G. Watson, “The development of version 3
and 4 of the Cambridge Structural Database System”,
Journal of Chemical Information Computer Science, Vol.
31, No. 2, 1991, pp. 187-204.
doi:10.1021/ci00002a004
[55] P.H. Maccallum, R. Poet and E.J. Milner-White,
“Coulombic interactions between partially charged
main-chain atoms not hydrogen-bonded to each other in-
fluence the conformations of alpha-helices and an-
ti-parallel beta-sheet. A new method for analyzing the
forces between hydrogen bonding groups in proteins in-
cludes all the Coulombic interactions”, Journal of Mo-
lecular Biology, Vol. 248, No. 2, 1995, pp. 361-373.
doi:10.1016/S0022-2836(95)80056-5
[56] P.H. Maccallum, R. Poet and E.J. Milner-White,
“Coulombic attractions between partially charged
main-chain atoms stabilize the right-handed twist found
in most beta-strands”, Journal of Molecular Biology, Vol.
248, No. 2, 1995, pp. 374-384.
doi:10.1016/S0022-2836(95)80057-3
[57] C.M. Deane, F.H. Allen, R. Taylor and T.L. Blundell,
“Carbonyl-carbonyl interactions stabilize the partially
allowed Ramachandran conformations of asparagine and
aspartic acid”, Protein Engineering, Vol. 12, No. 12,
1999, pp. 1025-1028.
doi:10.1093/protein/12.12.1025
[58] F.S. Nandel and H. Kaur, “Effect of terminal achiral and
chiral residues on the conformational behaviour of poly
zPhe and analysis of various interactions”, Indian
Journal of Biochemistry and Biophysics, Vol. 40, No. 4,
2003, pp. 265-273.
[59] F.S. Nandel, H. Kaur, N. Malik, N. Shankar and D.V.S.
Jain, “Conformational study of peptides containing de-
hydrophenylalanine: Helical structures without hydrogen
bonds”, Indian Journal of Biochemistry and Biophysics,
Vol. 38, No. 6, 2001, pp. 417-425.
[60] T.M. Head-Gordon, M.J. Frisch, C.L. Brooks III and J.A.
Pople, “Theoretical study of blocked glycine and alanine
peptide analogs”, Journal of American Chemical Society,
Vol. 113, No. 16, 1991, pp. 5989-5997.
doi:10.1021/ja00016a010
[61] J. Park, S.A. Teichmann, T. Hubbard and C. Chothia,
“Intermediate sequences increase the detection of ho-
mology between sequences”, Journal of Molecular Bi-
ology, Vol. 273, No. 1, 1997, pp. 349-354.
doi:10.1006/jmbi.1997.1288
[62] F.S. Nandel and R. Jaswal, “New type of helix and 27
ribbon structure formation in poly Leu peptides: Con-
struction of a single-handed template”, Biomacro-
molecules, Vol. 8, No. 10, 2007, pp. 3093-3101.
doi:10.1021/bm700504h
[63] Y. Demizu, N. Yamagata, Y. Sato, M. Doi, M. Tanaka, H.
Okuda and M. Kurihara, “Controlling the helical screw
F. S. Nandel et al. / Journal of Biophys ical Chemistry 2 (2011) 37-48
Copyright © 2011 SciRes. Openly accessible at http://www.scirp.org/journal/JBPC/
48
sense of peptides with C-terminal L-valine”, Journal of
Peptide Science, Vol. 16, No. 3, 2010, pp. 153-158.
[64] Z. Liu, K. Chen, A. Ng, Z. Shi, R.W. Woody and N.R.
Kallenbach, “Solvent dependence of PII conformation in
model alanine peptides”, Journal of American Chemical
Society, Vol. 126, No. 46, 2004, pp. 15141-15150.
doi:10.1021/ja047594g
[65] B.W. Chellgren and T.P. Creamer, “Effects of H2O and
D2O on polyproline II helical structure”, Journal of
American Chemical Society, Vol. 126, No. 45, 2004, pp.
14734-14735.
doi:10.1021/ja045425q
[66] D. Voet, J.G. Voet and C.W. Pratt, “Fundamentals of Bi-
ochemistry”, John Wiley and Sons, Inc., 2008.
[67] F.S. Nandel, A. Ahluwalia and H. Kaur, “A conforma-
tional structure of the amphipathic peptide insecticide
L-KALA”, International Journal of Quantum Chemistry,
Vol. 55, No. 1, 1995, pp. 61-69.
doi:10.1002/qua.560550109
[68] G.R. Desiraju and T. Steiner, “The Weak Hydrogen
Bond”, In: Structural Chemistry and Biology, Oxford
Science Publications, 1999.
[69] C. Baldauf, R. Günther and H.J. Hofmann, “Helices in
peptoids of α- and β-peptides”, Physical Biology, Vol. 3,
No. 1, 2006, pp. S1-S9.
doi:10.1088/1478-3975/3/1/S01