ArticlePDF Available

A direct comparison of helix propensity in proteins and peptides

Authors:

Abstract and Figures

alpha-Helical secondary structure occurs widely in globular proteins and its formation is a key step in their folding. As a consequence, understanding the energetics of helix formation is crucial to understanding protein folding and stability. We have measured the helix propensities of the nonpolar amino acids for an alpha-helix in an intact protein, ribonuclease T1, and for a 17-residue peptide with a sequence identical to that of the alpha-helix in the protein. The helix propensities are in excellent agreement. This shows that when compared in the same sequence context, the helix propensities of the nonpolar amino acids are identical in helical peptides and intact proteins, and that conclusions based on studies of the helix-to-coil transitions of peptides may, in favorable cases, be directly applicable to proteins. Our helix propensities based on ribonuclease T1 are in good agreement with those from similar studies of barnase and T4 lysozyme. In contrast, our helix propensities differ substantially from those derived from studies of alanine-stabilized or salt bridge-stabilized model alpha-helical peptides.
Content may be subject to copyright.
Proc. Natl. Acad. Sci. USA
Vol. 94, pp. 2833–2837, April 1997
Biochemistry
A direct comparison of helix propensity in proteins and peptides
JEFFREY K. MYERS,C.NICK PACE
, AND J. MARTIN SCHOLTZ
Departments of Medical Biochemistry and Genetics, and Biochemistry and Biophysics, and Center for Macromolecular Design, Texas A&M University,
College Station, TX 77843-1114
Communicated by Robert L. Baldwin, Stanford Medical Center, Stanford, CA, January 3, 1997 (received for review December 5, 1996)
ABSTRACT
a
-Helical secondary structure occurs widely
in globular proteins and its formation is a key step in their
folding. As a consequence, understanding the energetics of
helix formation is crucial to understanding protein folding
and stability. We have measured the helix propensities of the
nonpolar amino acids for an
a
-helix in an intact protein,
ribonuclease T
1
, and for a 17-residue peptide with a sequence
identical to that of the
a
-helix in the protein. The helix
propensities are in excellent agreement. This shows that when
compared in the same sequence context, the helix propensities
of the nonpolar amino acids are identical in helical peptides
and intact proteins, and that conclusions based on studies of
the helix-to-coil transitions of peptides may, in favorable
cases, be directly applicable to proteins. Our helix propensi-
ties based on ribonuclease T
1
are in good agreement with those
from similar studies of barnase and T4 lysozyme. In contrast,
our helix propensities differ substantially from those derived
from studies of alanine-stabilized or salt bridge-stabilized
model
a
-helical peptides.
Predicting the three-dimensional structure of a protein from
its amino acid sequence and gaining a detailed understanding
of the mechanism of protein folding remain two of the most
difficult, unsolved problems in biochemistry. In both cases,
understanding the many forces that contribute to the confor-
mational stability of a protein and their interplay is a major
difficulty. One approach to this problem is to uncouple the
formation of secondary structure from overall protein folding
by studying the factors that influence secondary structure
formation in model peptides.
a
-Helices are of primary interest
because they occur widely in proteins and the isolated peptides
often form helical structures in solution so that they can be
used as convenient models for protein folding and stability
(1–5). Although model
a
-helical peptides have been studied in
detail, the relevance of these models to the folding of intact
proteins has not been carefully explored. Here we present a
direct comparison of the helix propensity of the nonpolar
amino acids measured in an
a
-helix in an intact protein, and
in an
a
-helical peptide with the identical sequence.
Ribonuclease T
1
(RNase T
1
) is a small (104 residue),
monomeric protein, which has proven to be a useful model for
the study of protein folding and stability (6). RNase T
1
is an
a
1
b
class protein with several strands of
b
-sheet packed
against a relatively long (17 residues and 4.5 turns)
a
-helix,
forming a hydrophobic core (7). The sequence of the single
a
-helix in wild-type RNase T
1
is: SSDVSTAQAAGYKLHED,
which corresponds to Ser-13 through Asp-29 in the intact
protein (Fig. 1). The helical portion of the RNase T1 protein
has a near ideal site at which to measure helix propensities:
alanine 21 is in the exact center of the helix, on the solvent
exposed face, and the side chains of residues (i, i13) and (i,
i14), which could interact with residues at position 21 are all
involved in other interactions. No residues outside of the
helical region of the protein appear to be close enough to
interact with side chains at position 21. Here we compare the
differences in helix propensity for the nonpolar amino acids in
the context of the intact RNase T1 protein and in a helical
peptide derived from RNase T1.
MATERIALS AND METHODS
Mutants were constructed by the polymerase chain reaction
single-mutagenic primer technique (9) and the proteins were
purified as described (10). Purity was confirmed by SDSy
PAGE. The peptides were synthesized using an Applied
Biosystems automated peptide synthesizer (model 431A) and
standard fluorenylmethoxycarbonyl chemistry. The N and C
termini of the peptides are blocked with acetyl and carbox-
amide, respectively. The peptides were purified using reversed-
phase liquid chromatography and the identity was confirmed
using matrix-assisted laser desorption spectroscopy–time-of-
flight mass spectrometry.
The stability of the mutant proteins was determined using
urea denaturation. Usually, 22 tubes were prepared for each
denaturation curve; each sample tube contained 0.01 mgyml
21
protein in 30 mM glycine (pH 2.5), along with various con-
centrations of urea (Sigma ultrapure). A urea stock solution
was prepared fresh for each curve; the urea concentration was
measured by refractive index using a relation given previously
(11). These tubes were incubated at 25.08C for at least 16 hr
The publication costs of this article were defrayed in part by page charge
payment. This article must therefore be hereby marked ‘‘advertisement’’ in
accordance with 18 U.S.C. §1734 solely to indicate this fact.
Copyright q 1997 by T
HE NATIONAL ACADEMY OF SCIENCES OF THE USA
0027-8424y97y942833-5$2.00y0
PNAS is available online at http:yywww.pnas.org.
Abbreviations: RNase T
1
, ribonuclease T
1
; CD, circular dichroism.
To whom reprint requests should be addressed at: Department of
Medical Biochemistry and Genetics. e-mail: pace@bioch.tamu.edu
or jm-scholtz@tamu.edu.
FIG. 1. A ribbon drawing of RNase T
1
generated from the crystal
structure (Protein Data Base entry 9RNT) determined by Martinez-
Oyanedel et al. (7). The ribbon drawing was made with MOLSCRIPT (8).
The single
a
-helix in RNase T
1
spans residues 13–29. The site of
substitution, Ala-21, is shown.
2833
before measurements were taken to ensure that the tubes had
come to equilibrium. The intrinsic fluorescence of each sample
(kept at 25.0 6 0.18C by a circulating water bath and stirred
using a magnetic stirring apparatus) was measured by exciting
at 278 nm and monitoring emission at 320 nm in a SLM AB2
fluorescence spectrometer (Aminco). Analysis of the dena-
turation curves was performed using the two-state unfolding
model and the linear extrapolation method (11). These two
methods were combined into a single equation to describe the
shape of the denaturation curve (12, 13):
Y 5
Y
f
1 m
f
@urea# 1 ~Y
u
1 m
u
@urea#! 3
exp
@ 2 ~m~D
1y2
1 @urea#!yRT!#
1 1 exp@ 2 ~m~D
1y2
1 @urea#!yRT!#
,
where Y is the observed fluorescence (after subtracting out the
intrinsic fluorescence of buffer and urea), m
f
and Y
f
are the
slope and intercept, respectively, of the pretransition baseline,
m
u
and Y
u
are the slope and intercept, respectively, of the
posttransition baseline, m is the dependence of free energy of
unfolding on urea concentration, and D
1/2
is the midpoint of
the denaturation curve. The experimental curves were fit by
the above equation using standard data analysis software. The
free energy of unfolding in the absence of denaturant (defined
as the conformational stability of the protein) is the product of
m and D
1/2
. By calculating DDG from the difference in
midpoints times an average m value, a long extrapolation back
to 0 M urea is avoided. Denaturation curves were performed
at least twice for each mutant, and two mutants were measured
four times. The standard deviations in these sets of four give
an estimated error in D
1/2
of 0.04 M and an error in m of 50
kcal mol
21
M
21
. This gives a maximum error in DDG of 0.07
kcal mol
21
.
Helicity of the peptides was measured using circular dichro-
ism (CD) on an Aviv model 62DS CD spectropolarimeter.
Samples contained 30
m
M peptide in CD buffer (1 mM each
potassium phosphate, borate, and citrate) (pH 2.5) in a 0.5-cm
pathlength cuvette, maintained at 08C by a built-in tempera-
ture controlling unit. Peptide stock solutions were made in
water and the peptide concentration was determined by the
absorbance of the single tyrosine residue at 276 nm using an
extinction coefficient of 1390 M
21
cm
21
(14). To convert the
measured CD signal at 222 nm of the peptides into a free
energy scale it is necessary to use Lifson–Roig helix–coil
theory (15). Raw CD signal (in millidegrees) was converted to
mean residue ellipticity ([
u
]
obs
) and then to fraction helix using
F
helix
5
@
u
#
obs
2 @
u
#
coil
@
u
#
helix
2 @
u
#
coil
,
where [
u
]
helix
and [
u
]
coil
represent the mean residue ellipticity
of a complete helix (242,500z(1 2 (3yn)), where n is the
number of residues in the peptide and complete random coil
(1640), respectively (16). The units of mean residue ellipticity
are degycm
2
dmol
21
.
For analysis of the peptide data we define the wt* peptide
(see below) as the ‘‘host’’ peptide and all the other peptides
contained guest residues at position 21. The helix propensities
of the host and guest residues were calculated using a version
of the Lifson–Roig model for the helix–coil transition de-
scribed previously (17). The model employs the single-
sequence approximation, meaning that only one stretch of
helical residues was allowed to exist in any one peptide
molecule in the partition function (one nucleation site per
molecule). For peptides of this length, this model is equivalent
to the full treatment of Lifson–Roig theory (17). The nucle-
ation constant, v
2
, was taken to be 0.0023 (16). The model
treats the host peptide as a homopolymer, assigning only one
propagation parameter (w
host
) to the whole peptide based on
the measured helicity. The helix propagation parameters for
the guest residues (w
guest
) are determined from the changes in
measured helicity. Relative free energy changes were calcu-
lated using DG 52RT ln (wy(1 1 v)). We define DDG 5
DG
mut
2DG
wt*
, such that a positive DDG indicates destabili-
zation of the helix.
RESULTS AND DISCUSSION
The sequence of the single
a
-helix in wild-type RNase T
1
is:
SSDVSTAQAAGYKLHED (Ser-13 through Asp-29). A pep-
tide of this sequence shows a CD spectrum characteristic of a
random coil conformation, with at most only a few percent
helix. The G23A mutation (Gly-23 underlined in the sequence
above) increases the helicity of the peptide to 30% so that it
becomes a useful model. This variant, denoted wt*, serves as
the host for our helix propensity studies. The nonpolar amino
acids are substituted at position 21 (double underlined alanine
above) in the wt* peptide and wt* protein (also containing the
G23A mutation). We have substituted six nonpolar amino
acids at position 21 in the wt* RNase T
1
protein and measured
the resulting changes in conformational stability using urea
denaturation. The data and analysis are presented in Table 1.
All mutations resulted in stable, active ribonucleases. These
types of surface substitutions are not expected to cause large
changes in the structure of the protein (18, 19). The RNase T
1
with alanine at position 21 is the most stable and that with
glycine is the least stable. Alanine is typically found to be the
best helix former and glycine the worst (excluding proline) in
studies of other peptides and proteins.
We synthesized seven 17-residue peptides with sequences
identical to the
a
-helix in wt* RNase T
1
and the six nonpolar
mutants. Our peptide model, wt*, shows a CD spectrum
characteristic of an
a
-helix, with minima at 222 nm and 208 nm,
and a maximum around 190 nm (data not shown). This helical
Table 1. Protein urea denaturation and peptide helicity data
Residue
21
D
1y2
,
M
m,
kcal mol
21
M
21
D(DG)
kcal mol
21
2[U]
222
,
deg cm
2
dmol
21
F
helix
D(DG)
kcal mol
21
Ala 2.78 1.58 0 9,900 0.30 0
Leu 2.70 1.64 0.13 7,800 0.24 0.25
Met 2.69 1.60 0.14 8,300 0.25 0.18
Ile 2.51 1.57 0.44 6,600 0.20 0.38
Phe 2.43 1.69 0.56 5,100 0.16 0.61
Val 2.38 1.77 0.65 4,800 0.15 0.66
Gly 2.23 1.71 0.90 3,200 0.11 0.98
Change in conformational stability of the protein relative to wt* with alanine at position 21. Calculated
using D(DG) 5 m
avg
z(D
1y2
[wt*] 2 D
1y2
[mutant]), where m
avg
is the average m value for the seven proteins
(1.65 kcal mol
21
M
21
). Positive values of D(DG) indicate that the mutation is destabilizing. Errors in D
1y2
and m are approximately 0.04 M and 0.05 kcal mol
21
M
21
, respectively.
Change in free energy for helix formation of the peptides calculated from Lifson–Roig helix–coil theory
(see Materials and Methods).
2834 Biochemistry: Myers et al. Proc. Natl. Acad. Sci. USA 94 (1997)
structure shows no dependence on peptide concentration,
suggesting that the peptide exists as a monomer. The helical
contents of the seven peptides, as measured by CD at 222 nm,
are given in Table 1. The fractional helix contents ranged from
11 to 30%. We used Lifson–Roig helix–coil theory to translate
these fractional helicities into a relative scale of free energies
as described above. The range of free energies is about 1 kcal
mol
21
, just as in the mutant proteins.
Fig. 2 compares helix propensities measured in intact RNase
T
1
with those measured in peptides with sequences identical to
those of the
a
-helices of RNase T
1
. There is excellent agree-
ment between the two systems, with a slope of unity, a
y-intercept near zero, and a correlation coefficient of 0.98. This
is an important result. It shows that results from studies of the
helix-to-coil transition of a peptide may, in favorable cases, be
directly applicable to proteins. In earlier work, we also found
excellent agreement between studies of interactions at the
carboxyl terminus of the wt* peptide and similar studies of
intact RNase T
1
(20, 21).
Helix propensities were measured earlier in two other
proteins, barnase (22) and T4 lysozyme (18, 19). These results
are compared with our RNase T
1
results in Table 2 and Fig. 3A.
The agreement between results from three proteins and the
RNase T
1
peptide suggests that the helix propensities given in
Table 1 provide a good measure of helix propensity at an
exposed site near the center of an
a
-helix in a protein.
Intrinsic helical propensities of the amino acids were first
measured systematically by Scheraga and coworkers using a
host–guest random polymer system (24). These studies sug-
gested that the helical propensities of the amino acids, with the
exception of proline, did not vary greatly, and that short
peptides (,20 residues) would not exhibit significant helix
formation in aqueous solution. However, in 1968, Klee showed
that the S-peptide of ribonuclease A does form significant
amounts of
a
-helix in aqueous solution (27, 28). This triggered
studies, first by the Baldwin laboratory and then by several
other laboratories, to find simpler model
a
-helices that could
be used to probe the determinants of
a
-helix stability. It was
found that short, monomeric helices composed mostly of
alanine exhibit significant helix formation in water (29, 30).
This helix formation occurred even without favorable side-
chain interactions, showing that the helical propensity of
alanine is higher than indicated by host–guest studies. These
alanine-stabilized peptides have been widely used to model
protein folding (1–5, 16). An important question is whether
these and other model
a
-helices give results directly applicable
Table 2. Helix propensity measured in various systems
Model system
(sequence)
DG(Gly) 2DG(Ala),
kcal mol
21
Intercept Slope
Correlation
coefficient
Peptides
RNase T
1
peptide 0.98 (0.00)
(1.00)
(1.00)
(-STAQXAAYK-)
AK 1.97 0.00 1.84 0.97
(-AAKAXAAKA-)
E
4
K
4
0.74 0.06 0.70 0.97
(-KKKXXXEEE-)
EAK 1.95 20.04 1.79 0.90
(-AKEAXAKEA-)
Host–guest 0.35 20.11 0.35 0.80
(HBLP or HPLG)
AGADIR 1.10 20.04 1.02 0.96
(various)
Proteins
RNase T
1
protein-21 0.90 20.02 0.96 0.98
(-STAQXAAYK-)
Barnase-32 0.91 0.19 0.88 0.88
(-KSAQXLG-)
T4 lysozyme-44 0.96 20.13 0.92 0.93
(-QAAKXELDK-)
Coiled–coil 0.77 0.07 0.74 0.90
(-AALEXKLQA-)
The peptide or protein systems being compared to the RNase T
1
peptide are shown along with the
sequence near the substitution site (X), given as one-letter amino acid codes. For the peptide model
systems, the AK data are from Baldwin’s group (16), the E
4
K
4
data are from Kallenbach’s group (5),
the EAK data are from Stellwagen’s group (23) as analyzed by Chakrabartty and Baldwin (3), the
host–guest studies of Scheraga (24) and the algorithm AGADIR is from Mun˜oz and Serrano (25). HPLG
and HBLG refer to the host, hydroxypropyl- or hydroxybutyl-L-glutamine, respectively (24). The protein
models are from site 32 in barnase (22), site 44 in T4 lysozyme (18, 19) and a solvent-exposed site in
a model coiled–coil peptide (26).
The intercept, slope, and correlation coefficient are derived by plotting the data from the indicated model
system against the results for the RNase T
1
peptide.
FIG. 2. Comparison of measured helix propensity in the RNase T
1
peptide and protein systems. The differences in DG are expressed
relative to alanine for the other nonpolar amino acids. Data are from
Table 1.
Biochemistry: Myers et al. Proc. Natl. Acad. Sci. USA 94 (1997) 2835
to the
a
-helices found in proteins. With regard to the helix
propensities of nonpolar amino acids, our results suggest that
in some cases they do not.
A comparison of our peptide results with those from the
alanine-stabilized peptides studied by Baldwin and coworkers
(16) and with the salt bridge-stabilized peptides studied by
Kallenbach and coworkers (5) is shown in Fig. 3B. In both
cases, the correlation is excellent (Table 2), but there is a
sizable discrepancy in the range of propensities. The propen-
sities measured in alanine-stabilized helices are almost twice
those found with the ribonuclease T
1
helices, whereas those for
the salt bridge-stabilized helices are about 30% less (Table 2).
The propensities measured by Stellwagen’s group (23) (see
also ref. 3) with peptides stabilized by both alanine residues
and salt bridges are very similar to those measured in the
alanine-stabilized peptides (Table 2). In contrast, our peptide
data are in excellent agreement with results from the program
AGADIR developed by Mun˜oz and Serrano (23) (Table 2).
The parameters used in AGADIR were obtained from an
analysis of the measured fractional helicities of over 400
peptides. It is also obvious from Table 2 that propensities from
host–guest studies are not applicable to either proteins or
other peptides.
The major discrepancy is that the D(DG) values from the
alanine-stabilized peptides are twice as large as the those
measured in RNase T
1
and in other proteins. When the amino
acid sequences of the various systems are compared, important
differences are found (Table 2). In the alanine-stabilized
peptides, the variable residue has adjacent alanines and has
alanines at three of the four (i, i13) and (i, i14) positions. The
RNase T
1
helix has bulkier residues at some of these positions,
as do the T4 lysozyme and barnase helices. The residues in the
salt bridge-stabilized peptides of Kallenbach are even bulkier,
with glutamic acid or lysine present at most positions. The
host–guest polymers consist mainly of very large host residues
(hydroxypropyl- or hydroxybutyl-
L-glutamine). It appears that
the range of propensities scales with the size of the residues
surrounding the variable position. This suggests that the
hydration of backbone amides and carbonyls might be an
important factor (for an example, see ref. 31). The local
sequence might also change the flexibility of the backbone or
side chain and thereby influence conformational entropy. It is
surprising that these differences in sequence exert such a large
effect on the measured helix propensities. It will be interesting
to see if the theoreticians who study the helix-to-coil transition
can explain this difference in behavior.
It has been suggested that the disagreement between ala-
nine-based peptides and other systems is due to oversimplifi-
cation in applying complex helix–coil theories (32). Our pep-
tide results confirm that the problem does not lie in helix–coil
theory. One recent attempt to explain the discrepancy between
propensities measured in peptides and proteins was made by
Qian and Chan (33). In their model, protein-based systems and
isolated peptides should only give identical measures of helix
propensity when helix formation and global protein folding are
tightly coupled. The excellent agreement between our peptide
and protein data suggests that secondary and tertiary structure
formation are indeed tightly coupled in RNase T
1
.
Several explanations for the different propensities exhibited
by the amino acids with nonpolar side chains have been
proposed. These include differences in conformational en-
tropy (34), in the hydrophobic effect (19), and in hydration of
the backbone (35). The helix propensities calculated by Her-
mans et al. (32) using molecular dynamics simulations are in
remarkably good agreement with the results in Table 1. They
predict a D(DG)of'0.2 kcal mol
21
for leucine and methionine
(based on their calculations with
a
-amino-n-butyric acid), of
'0.7 kcal mol
21
for valine, and '1.2 kcal mol
21
for glycine.
They attribute the difference between alanine and glycine
entirely to differences in backbone conformational entropy,
and those between alanine and the other side chains mainly to
differences in side-chain conformational entropy. These and
other calculations (19, 34, 36, 37) suggest that the differences
in helix propensity for the nonpolar amino acids are due mainly
to differences in conformational entropy.
Recently, a structure-based thermodynamic scale for
a
-helix
propensity has been developed (38). In this approach, the
differences in helical propensity between the amino acids
cannot solely be explained by differences in side-chain con-
formational entropy, but can be faithfully calculated when
backbone conformational entropy, solvation entropy, and en-
thalpic contributions are included. Using this approach, good
agreement is found between predicted and observed helix
propensities in T4 lysozyme, barnase, and coiled-coils (see
Table 2).
The conformational stability of proteins is remarkably low,
only 5–10 kcal mol
21
. The large conformational entropy ['1.7
kcal mol
21
per residue at 300 K (36, 37, 39)] that favors the
unfolded state is barely surmounted by a large number of weak,
stabilizing interactions: '1 kcal mol
21
per -CH
2
-group buried
(40) and '1 kcal mol
21
per intramolecular hydrogen bond
formed (41). Even though the helix propensities discussed here
are similarly weak, they are important because proteins typi-
cally contain 30% of their residues in
a
-helical conformations.
Intrinsic propensities are also very important for determining
the conformational preferences of peptides. However, the
contributions from helix propensity are generally destabilizing.
As pointed out (1–5, 34), only alanine residues contribute
favorably to the stability of
a
-helices, all other amino acids are
either neutral or destabilizing and make their contributions
mainly through unfavorable conformational entropy.
In summary, one important unresolved question has been
(1–5): Do helix propensities make an equivalent energetic
FIG.3. (A) Comparison of measured helix propensity in the wt*
peptide and T4 lysozyme (18, 19) (
F
) and barnase (22) (
M
).Values
given are the change in DG of folding relative to alanine. (B)
Comparison of measured helix propensity values for the nonpolar
amino acids in the RNase T
1
wt* peptide and alanine-based peptides
(16) (
F
) and salt bridge-stabilized peptides (5) (
M
). Values are the
change in DG of helix formation relative to alanine. Slopes, intercepts,
and correlation coefficients for best-fit linear regressions of the data
in A and B are given in Table 2.
2836 Biochemistry: Myers et al. Proc. Natl. Acad. Sci. USA 94 (1997)
contribution in peptides and proteins? Our results suggest that
the answer is YES.
We thank R. L. Baldwin and the Pace and Scholtz lab groups for
helpful discussion, Geoff Horn for DNA sequencing, and Kevin Shaw
for generating Fig. 1. We acknowledge the National Institutes of
Health for financial support (Grants GM52483 to J.M.S. and
GM37039 to C.N.P. and Predoctoral Training Grant T32 GM08523 to
J.K.M.) and the Robert A. Welch Foundation (Grants A-1281 to
J.M.S. and A-1060 to C.N.P.). C.N.P. is also supported by the Tom and
Jean McMullin Professorship and J.M.S. is an American Cancer
Society Junior Faculty Research Awardee (Grant JFRA-577).
1. Scholtz, J. M. & Baldwin, R. L. (1992) Annu. Rev. Biophys.
Biomol. Struct. 21, 95–118.
2. Scholtz, J. M. & Baldwin, R. L. (1995) in Peptides: Synthesis,
Structures, and Applications, ed. Gutte, B. (Academic, San Di-
ego), pp. 171–192.
3. Chakrabartty, A. & Baldwin, R. L. (1995) Adv. Protein Chem. 46,
141–176.
4. Bryson, J. W., Betz, S. F., Lu, H. S., Suich, D. J., Zhou, H. X.,
O’Neil, K. T. & DeGrado, W. F. (1995) Science 270, 935–941.
5. Kallenbach, N. R., Lyu, P. & Zhou, H. (1996) in Circular Di-
chroism and the Conformational Analysis of Biopolymers, ed.
Fasman, G. D. (Plenum, New York), pp. 201–259.
6. Pace, C. N., Heinemann, U., Hahn, U. & Saenger, W. (1991)
Angew. Chem. Int. Ed. Engl. 30, 343–360.
7. Martinez-Oyanedel, J., Choe, H. W., Heinemann, U. & Saenger,
W. (1991) J. Mol. Biol. 222, 335–352.
8. Kraulis, P. J. (1991) J. Appl. Crystallogr. 24, 946–950.
9. Landt, O., Grunert, H. P. & Hahn, U. (1990) Gene 96, 125–128.
10. Shirley, B. A. & Laurents, D. V. (1990) J. Biochem. Biophys.
Methods 20, 181–188.
11. Pace, C. N. (1986) Methods Enzymol. 131, 266–280.
12. Santoro, M. M. & Bolen, D. W. (1988) Biochemistry 27, 8063–
8068.
13. Scholtz, J. M. (1995) Protein Sci. 4, 35–43.
14. Pace, C. N., Vajdos, F., Fee, L., Grimsley, G. & Gray, T. (1995)
Protein Sci. 4, 2411–2423.
15. Lifson, S. & Roig, A. (1961) J. Chem. Phys. 34, 1963–1974.
16. Rohl, C. R., Chakrabartty, A. & Baldwin, R. L. (1996) Protein
Sci. 5, 2623–2637.
17. Qian, H. (1993) Biopolymers 33, 1605–1616.
18. Blaber, M., Zhang, X.-J. & Matthews, B. W. (1993) Science 260,
1637–1640.
19. Blaber, M., Zhang, X.-J., Lindstrom, J. D., Peipot, S. D., Baase,
W. A. & Matthews, B. W. (1994) J. Mol. Biol. 236, 600624.
20. Myers, J. K., Smith, J. S., Pace, C. N. & Scholtz, J. M. (1996) J.
Mol. Biol. 263, 390–395.
21. Walter, S., Hubner, B., Hahn, U. & Schmid, F. X. (1995) J. Mol.
Biol. 252, 133–143.
22. Horovitz, A., Matthews, J. M. & Fersht, A. R. (1992) J. Mol. Biol.
227, 560–568.
23. Park, S.-H., Shalongo, W. & Stellwagen, E. (1993) Biochemistry
32, 7048–7053.
24. Wo´jcik, J., Altman, K. H. & Scheraga, H. A. (1990) Biopolymers
30, 121–134.
25. Mun˜oz, V. & Serrano, L. (1995) J. Mol. Biol. 245, 275–296.
26. O’Neil, K. T. & DeGrado, W. F. (1990) Science 250, 646651.
27. Klee, W. A. (1968) Biochemistry 7, 2731–2736.
28. Brown, J. E. & Klee, W. A. (1971) Biochemistry 10, 470476.
29. Marqusee, S. & Baldwin, R. L. (1987) Proc. Natl. Acad. Sci. USA
84, 88988902.
30. Marqusee, S., Robbins, V. H. & Baldwin, R. L. (1989) Proc. Natl.
Acad. Sci. USA 86, 5286–5290.
31. Blaber, M., Baase, W. A., Gassner, N. & Matthews, B. W. (1995)
J. Mol. Biol. 246, 317–330.
32. Hermans, J., Anderson, A. G. & Yun, R. H. (1992) Biochemistry
31, 5646–5653.
33. Qian, H. & Chan, S. I. (1996) J. Mol. Biol. 261, 279–288.
34. Creamer, T. P. & Rose, G. D. (1992) Proc. Natl. Acad. Sci. USA
89, 5937–5941.
35. Avbelj, F. & Moult, J. (1995) Biochemistry 34, 755–764.
36. D’Aquino, J. A., Go´mez, J., Hilser, V. J., Lee, K. H., Amzel,
L. M. & Freire, E. (1996) Proteins Struct. Funct. Genet. 25,
143–156.
37. Lee, K. H., Xie, D., Freire, E. & Amzel, L. M. (1994) Proteins 20,
6884.
38. Luque, I., Mayorga, O. L. & Freire, E. (1996) Biochemistry 35,
13681–13688.
39. Spolar, R. S. & Record, M. T., Jr. (1994) Science 263, 777–784.
40. Pace, C. N. (1992) J. Mol. Biol. 226, 29–35.
41. Myers, J. K. & Pace, C. N. (1996) Biophys. J. 71, 2033–2039.
Biochemistry: Myers et al. Proc. Natl. Acad. Sci. USA 94 (1997) 2837
... where the variable θ is the measured difference in absorbed circularly polarized light in millidegrees, c is the millimolar concentration of the specimen, l is the path length of the cuvette in cm and b is the number of amide bonds in the polypeptide, for which the N-terminal acetyl bond was included but not the C-terminal amide. Peptide concentration was where MRE coil is calculated by 640-45*T; T is the temperature in °C; and n is the number of amide bonds in the sample (including the C-terminal amide) 73 . ...
Article
Full-text available
Recent advances in de novo protein design have delivered a diversity of discrete de novo protein structures and complexes. A new challenge for the field is to use these designs directly in cells to intervene in biological processes and augment natural systems. The bottom-up design of self-assembled objects such as microcompartments and membraneless organelles is one such challenge. Here we describe the design of genetically encoded polypeptides that form membraneless organelles in Escherichia coli . To do this, we combine de novo α-helical sequences, intrinsically disordered linkers and client proteins in single-polypeptide constructs. We tailor the properties of the helical regions to shift protein assembly from arrested assemblies to dynamic condensates. The designs are characterized in cells and in vitro using biophysical methods and soft-matter physics. Finally, we use the designed polypeptide to co-compartmentalize a functional enzyme pair in E. coli , improving product formation close to the theoretical limit.
... Where MREcoil = 640-45T; T is the temperature in °C; and n is the number of amide bonds in the sample (including the C-terminal amide) [99]. ...
Preprint
Full-text available
Tardigrades are microscopic animals that survive desiccation by inducing biostasis. To survive drying tardigrades rely on intrinsically disordered CAHS proteins that form gels. However, the sequence features and mechanisms underlying gel formation and the necessity of gelation for protection have not been demonstrated. Here we report a mechanism of gelation for CAHS D similar to that of intermediate filaments. We show that gelation restricts molecular motion, immobilizing and protecting labile material from the harmful effects of drying. In vivo, we observe that CAHS D forms fiber-like condensates during osmotic stress. Condensation of CAHS D improves survival of osmotically shocked cells through at least two mechanisms: reduction of cell volume change and reduction of metabolic activity. Importantly, condensation of CAHS D is reversible and metabolic rates return to control levels after CAHS condensates are resolved. This work provides insights into how tardigrades induce biostasis through the self-assembly of CAHS gels.
... a-methyl Ala has been utilized for many years in peptide synthesis [14], but other type of a-methylated amino acids have only recently become commercially available. In general, compared to most other amino acids, Ala is already more helicogenic, because its short methyl side chain does not sterically interfere with helix formation [26,34]. Despite these two favorable factors, one of the major findings from this study is that other types of a-methylated amino acids may be better suited than a-methyl Ala in the design of ApoA-I mimetic peptides. ...
Article
Full-text available
Apolipoprotein A-I (ApoA-I) mimetic peptides are potential therapeutic agents for promoting the efflux of excess cellular cholesterol, which is dependent upon the presence of an amphipathic helix. Since α-methylated Ala enhances peptide helicity, we hypothesized that incorporating other types of α-methylated amino acids into ApoA-I mimetic peptides may also increase their helicity and cholesterol efflux potential. The last helix of apoA-I, peptide ‘A’ (VLESFKVSFLSALEEYTKKLNT), was used to design peptides containing a single type of α-methylated amino acid substitution (Ala/Aα, Glu/Dα, Lys/Kα, Leu/Lα), as well as a peptide containing both α-methylated Lys and Leu (6α). Depending on the specific residue, the α-helical content as measured by CD-spectroscopy and calculated hydrophobic moments were sometimes higher for peptides containing other types of α-methylated amino acids than those with α-methylated Ala. In ABCA1-transfected cells, cholesterol efflux to the peptides showed the following order of potency: 6α>Kα≈Lα≈Aα≫Dα≈A. In general, α-methylated peptides were resistant to proteolysis, but this varied depending on the type of protease and specific amino acid substitution. In summary, increased helicity and amphilicity due to α-methylated amino acid substitutions in ApoA-I mimetic peptides resulted in improved cholesterol efflux capacity and resistance to proteolysis, indicating that this modification may be useful in the future design of therapeutic ApoA-I mimetic peptides.
... It was derived by Khandogin and Brooks, and has been applied to model peptides of Aβ 37 . The helix content c was given by where N hel is the number of helix formed fragment in the peptide, n i is the number of helix formed residues in the i-th helix formed fragment, and k is a minimum number of helix formed residues required to produce the CD signal, which was set to the value of 3 54,56 . Here, the number of helix formed residues n i was obtained by the DSSP algorithm 43 . ...
Article
Full-text available
Aggregates of amyloid-β (Aβ) peptide are well known to be the causative substance of Alzheimer’s disease (AD). Recent studies showed that monosialotetrahexosylganglioside (GM1) clusters induce the pathological aggregation of Aβ peptide responsible for the onset and development of AD. However, the effect of GM1-glycan cluster on Aβ conformations has yet to be clarified. Interactions between Aβ peptide and GM1-glycan cluster is important for the earliest stage of the toxic aggregation on GM1 cluster. Here, we performed all-atom molecular dynamics (MD) simulations of Aβ40 on a recently developed artificial GM1-glycan cluster. The artificial GM1-glycan cluster facilitates the characterization of interactions between Aβ40 and multiple GM1-glycans. We succeeded in observing the binding of Aβ40 to the GM1-glycan cluster in all of our MD simulations. Results obtained from these MD simulations indicate the importance of HHQ (13-15) segment of Aβ40 for the GM1-glycan cluster recognition. This result is consistent with previous experimental studies regarding the glycan recognition of Aβ peptide. The recognition mechanism of HHQ (13-15) segment is mainly explained by non-specific stacking interactions between side-chains of histidine and rings of sugar residues, in which the HHQ regime forms coil and bend structures. Moreover, we found that Aβ40 exhibits helix structures at C-terminal side on the GM1-glycan cluster. The helix formation is the initial stage of the pathological aggregation at ceramide moieties of GM1 cluster. The binding of Lys28 to Neu triggers the helix formation at C-terminus side because the formation of a salt bridge between Lys28 and Neu leads to change of intrachain interactions of Aβ40. Our findings suggest that the pathological helix formation of Aβ40 is initiated at GM1-glycan moieties rather than lipid ceramide moieties.
Article
Full-text available
Tardigrades are microscopic animals that survive desiccation by inducing biostasis. To survive drying tardigrades rely on intrinsically disordered CAHS proteins, which also function to prevent perturbations induced by drying in vitro and in heterologous systems. CAHS proteins have been shown to form gels both in vitro and in vivo, which has been speculated to be linked to their protective capacity. However, the sequence features and mechanisms underlying gel formation and the necessity of gelation for protection have not been demonstrated. Here we report a mechanism of fibrillization and gelation for CAHS D similar to that of intermediate filament assembly. We show that in vitro, gelation restricts molecular motion, immobilizing and protecting labile material from the harmful effects of drying. In vivo, we observe that CAHS D forms fibrillar networks during osmotic stress. Fibrillar networking of CAHS D improves survival of osmotically shocked cells. We observe two emergent properties associated with fibrillization; (i) prevention of cell volume change and (ii) reduction of metabolic activity during osmotic shock. We find that there is no significant correlation between maintenance of cell volume and survival, while there is a significant correlation between reduced metabolism and survival. Importantly, CAHS D's fibrillar network formation is reversible and metabolic rates return to control levels after CAHS fibers are resolved. This work provides insights into how tardigrades induce reversible biostasis through the self‐assembly of labile CAHS gels.
Article
De novo peptides and proteins that switch state in response to chemical and physical cues would advance protein design and synthetic biology. Here we report two designed systems that disassemble and reassemble upon site-specific phosphorylation and dephosphorylation, respectively. As starting points, we use hyperthermostable de novo antiparallel and parallel coiled-coil heterotetramers, i.e., A2B2 systems, to afford control in downstream applications. The switches are incorporated by adding protein kinase A phosphorylation sites, R-R-X-S, with the phosphoacceptor serine residues placed to maximize disruption of the coiled-coil interfaces. The unphosphorylated peptides assemble as designed and unfold reversibly when heated. Addition of kinase to the assembled states unfolds them with half-lives of ≤5 min. Phosphorylation is reversed by Lambda Protein Phosphatase resulting in tetramer reassembly. We envisage that the new de novo designed coiled-coil components, the switches, and a mechanistic model for them will be useful in synthetic biology, biomaterials, and biotechnology applications.
Thesis
Les α-hélices sont des éléments clés de la reconnaissance biomoléculaire, comme en témoigne le fait qu'une quantité significative de complexes protéine-protéine, dans la banque de données protéiques (PDB), présentent des interfaces inter-hélices. Cependant, de courtes hélices peptidiques isolées dans le but de bloquer ces interactions ne sont généralement que faiblement présentes en milieu aqueux et sont sensibles à la dégradation protéolytique, limitant ainsi leur potentiel thérapeutique. Diverses approches chimiques ont été proposées pour augmenter la propension au repliement en hélice des α-peptides. Une stratégie consiste à pré-organiser les premières liaisons amides à l'aide d'une « capping box » ou d'un substitut de liaison hydrogène. Récemment, nous nous sommes intéressés à la possibilité d'interfacer des peptides-α et des foldamères afin d’élaborer des « foldamères à blocs » générant ainsi un nouveau type d’architecture mime de l'hélice-α. Dans notre laboratoire, nous avons développé des foldamers à base d’urées qui s’organisent pour former des structures hélicoïdales. Les similitudes du sens d’enroulement, du pas et de la polarité entre l’hélice-α peptidique et l’hélice-2.5 d'oligourée suggéraient qu'il serait possible de combiner ces deux squelettes. Au cours de cette thèse, nous avons montré que les chimères : oligourée/α-peptide, forment des structures hélicoïdales bien définies dans les solvants organiques polaires avec la propagation d'un réseau de liaisons hydrogène intramoléculaires continu couvrant la totalité de la séquence. Ces études ont suggéré que le squelette de l'oligourée qui possède une forte propension à adopter une structuration en hélice pourrait conduire au développement d’un « cap » permettant la pré-organisation des quatre premiers NHs ainsi que des quatre derniers groupements carbonyles d’une hélice-α peptidique. Nous avons donc étudié l'influence de courts fragments oligourées sur la stabilisation de séquences peptidiques modèles solubles dans l'eau en hélices-α, conduisant au développement de la « foldamer capping box ». Grâce à cette nouvelle stratégie de stabilisation des hélices peptidiques, nous avons pu concevoir des inhibiteurs de l’interaction protéine/protéine p53/MDM2.
Article
Protein-protein interactions control a wide variety of natural biological processes. α-Helical coiled coils frequently mediate such protein-protein interactions. Due to the relative simplicity of their sequences and structures and the ease with which properties such as strength and specificity of interaction can be controlled, coiled coils can be designed de novo to deliver a variety of non-natural protein-protein-interaction domains. Herein, several de novo designed coiled coils are tested for their ability to mediate protein-protein interactions in Escherichia coli cells. The set includes a parallel homodimer, a parallel homotetramer, an antiparallel homotetramer and a newly designed heterotetramer, all of which have been characterized in vitro by biophysical and structural methods. Using a transcription repression assay based on reconstituting the Lac repressor, we find that the modules behave as designed in the cellular environment. Each design imparts a different property to the resulting Lac repressor-coiled coil complexes, resulting in the benefit of being able to reconfigure the system in multiple ways. Modification of the system also allows the interactions to be controlled: assembly can be tuned by controlling the expression of the constituent components; and complexes can be disrupted through helix sequestration. The small and straightforward de novo designed components that we deliver are highly versatile and have considerable potential as protein-protein-interaction domains in synthetic biology where proteins must be assembled in highly specific ways. The relative simplicity of the designs makes them amenable to future modifications to introduce finer control over their assembly and to adapt them for different contexts.
Article
Full-text available
The MOLSCRIPT program produces plots of protein structures using several different kinds of representations. Schematic drawings, simple wire models, ball-and-stick models, CPK models and text labels can be mixed freely. The schematic drawings are shaded to improve the illusion of three dimensionality. A number of parameters affecting various aspects of the objects drawn can be changed by the user. The output from the program is in PostScript format.
Chapter
This chapter elucidates the use of peptide helices to analyze the factors that determine protein stability: helix propensities and specific interactions between side chains. Most of the advances in the field of helix formation by peptides are due to the increased availability of solid-phase peptide synthesis methods. The initial studies in the field were directed toward an understanding of helix formation in water by the 13-residue C-peptide from the N terminus of ribonuclease A. The C-peptide and other model systems derived from proteins offer excellent systems for examining the qualitative roles of side-chain interactions as well as the effects of specific amino acid substitutions on helix formation. The C-peptide system was very complex as a host peptide for studying all the specific elements of helix formation in peptides. The major advancement in the peptide helix field came with the design of short peptide sequences that showed a good α-helix formation in water. With the availability of several de novo designed peptide systems, the interest has been revived in measuring the intrinsic helix-forming tendencies or helix propensities of all of the amino acids, and also in analyzing the side-chain interactions that can help stabilize helices in an aqueous solution.
Chapter
The proposal by Pauling and his coworkers (1951) of an atomic model for the structure of the alpha helix stimulated research in several areas of protein chemistry. It excited chemists as few discoveries have before or since, giving impetus to structural modeling efforts that resulted in the structure of DNA 2 years later, and in a whole new field of structural biology within two decades. Pauling’s feat pointed out the importance of understanding the conformation of the peptide group itself, rather than building models based on idealized helical structures. Working on the same problem, Bragg et al. (1950) failed to produce a structure of comparable elegance because they were unaware the peptide bond was planar (Crick, 1988). The alpha helix could be identified in the diffraction patterns from crystals of the globular proteins myoglobin and hemoglobin, as well as in the classical “α” patterns from fibrous proteins like keratin and synthetic polypeptides, poly(γ-l-glutamate) being the first to show the α pattern (Elliott, 1967).
Article
The evaluation of the configurational partition function of a polypeptide molecule, with the internal rotation angles as variables, leads to an improved treatment of the phenomenon of helix-coil transition in polypeptide molecules. The conditional probabilities of occurrence of helical and coiled states of the peptide units are obtained in the form of a 3×3 matrix. The order of this matrix is the lowest possible for the model employed, and is derived by a logical procedure which serves to eliminate redundancies in the enumeration of states. The eigenvalues of this matrix yield the various molecular averages as functions of the degree of polymerization, temperature, and molecular constants. Explicit formulas are given for the degree of intramolecular hydrogen bonding, average number of helical sequences, and the distribution of their lengths, as well as the number average and the weight average of these lengths.
Article
This chapter focuses on the mechanism of helix formation in an isolated peptide and the factors that determine the stability of a peptide helix. Helix propensities are considered together with N-cap and C-cap propensities, because measurement of helix propensities requires knowing values of the N-cap and C-cap propensities, and vice versa. The chapter considers side-chain interactions: these include both the interaction of a charged side chain with the helix macrodipole and specific interactions between a particular pair of side chains, such as ion pair and H-bond interactions. Measurement of these interactions is of interest for two reasons: their values are needed to relate the stability of a peptide helix to its amino acid composition and sequence; and peptide helices provide one of the best systems, and probably the most sensitive system, for quantifying the energetics of side-chain interactions. It also considers briefly the present status of the Chou-Fasman hypothesis and the relation between the mechanism of α-helix formation in peptides and proteins. It is necessary to use helix-coil transition theory to understand the populated intermediates and to analyze the energetics of helix formation. The two closely related theories of α-helix formation are the Zimm-Bragg theory and the Lifson-Roig theory.
Article
Proteins carry out the most important and difficult tasks in all living organisms. To do so, they must often interact specifically with other small and large molecules. This requires that they fold to a globular conformation with a unique active site that is used for the specific interaction. Consequently, protein folding can be regarded as the ''secret of life''. Biochemists and chemists have a great interest in elucidating the mechanism by which proteins fold and in predicting the folded conformation and its stability given just the amino acid sequence. This challenge is sometimes called the ''protein folding problem''. The ability to construct proteins differing in sequence by one or more amino acids and to analyze their three-dimensional structures by X-ray crystallography and NMR spectroscopy is a powerful tool for investigating the conformational stability and folding of proteins. Several proteins are now under intensive study by this approach. One of these is ribonuclease T1.
Article
Ribonuclease S-peptide is the amino-terminal eicosapeptide segment of bovine pancreatic ribonuclease A (i.e., residues 1-20). In the intact ribonuclease molecule a large portion of this segment is found to be helical (Kartha, G., Bello, J., and Harker, D. (1967), Nature 213, 862; Wyckoff, H. W., Hardman, K. D., Allewell, N. M., Inagami, T., Johnson, L. N., and Richards, J. M. (1967), J. Biol. Chem 242, 3984). The circular dichroism of S-peptide in dilute aqueous solution has been measured over a broad range of pH (1-11.5) and temperature (4-79°). These measurements show negative circular dichroism peaks near 200 and 225 mμ which indicate the presence of both helical and random coil structures in S-peptide throughout this range of conditions. In 5 M guanidine hydrochloride the peptide behaves as a random coil whereas 2-chloroethanol induces a high degree of helicity. These results are taken to indicate that, in the course of its conformational fluctuations, S-peptide often assumes a structure which is similar to that of the amino-terminal segment of intact ribonuclease.
Article
The free form of ribonuclease T1 (RNase T1) has been crystallized at neutral pH, and the three-dimensional structure of the enzyme has been determined at 1.5 Å nominal resolution. Restrained least-squares refinement yielded an R value of 14.3% for 12,623 structure amplitudes. The high resolution of the structure analysis permits a detailed description of the solvent structure around RNase T1, the reliable rotational setting of several side-chain amide and imidazole groups and the identification of seven disordered residues. Among these, the disordered and completely internal Val78 residue is noteworthy. In the RNase T1 crystal structures determined thus far it is always disordered in the absence of bound guanosine, but not in its presence. A systematic analysis of hydrogen bonding reveals the presence in RNase T1 of 40 three-center and an additional seven four-center hydrogen bonds. Three-center hydrogen bonds occur predominantly in the α-helix, where their minor components close 310-type turns, and in β-sheets, where their minor components connect the peptide nitrogen and carbonyl functions of the same residue. The structure of the free form is compared with complexes of RNase T1 with filled base recognition site and/or catalytic site. Several structural rearrangements occurring upon inhibitor or substrate binding are clearly apparent. In conjunction with the available biochemical knowledge, they are used to describe probable steps occurring early during RNase T1-catalyzed phosphate transesterification.
Article
The solvent-exposed residue Ala32 in the second alpha-helix of barnase was replaced by all other naturally occurring amino acids and the concomitant effects on the protein stability were determined. The results are assumed to reflect both the distinct conformational preferences of the different amino acids and also possible intrahelical interactions. The conformational preferences may be fully rationalized by invoking only a few physical principles. The results agree well with recently experimentally determined rank-order of helix-forming tendencies determined on a model peptide. There is very weak correlation between the results and the experimental host-guest values. There is a weak correlation between our results and the statistical helix propensities and a slightly better correlation with the positional-dependent statistical parameters of J. S. Richardson, and D. C. Richardson.