Journal of Molecular Biology
Volume 353, Issue 2, 21 October 2005, Pages 232-241
Journal home page for Journal of Molecular Biology

How Large is an α-Helix? Studies of the Radii of Gyration of Helical Peptides by Small-angle X-ray Scattering and Molecular Dynamics

https://doi.org/10.1016/j.jmb.2005.08.053Get rights and content

Using synchrotron radiation and the small-angle X-ray scattering technique we have measured the radii of gyration of a series of alanine-based α-helix-forming peptides of the composition Ace-(AAKAA)n-GY-NH2, n=2–7, in aqueous solvent at 10(±1) °C. In contrast to other techniques typically used to study α-helices in isolation (such as nuclear magnetic resonance and circular dichroism), small-angle X-ray scattering reports on the global structure of a molecule and, as such, provides complementary information to these other, more sequence-local measuring techniques. The radii of gyration that we measure are, except for the 12-mer, lower than the radii of gyration of ideal α-helices or helices with frayed ends of the equivalent sequence-length. For example, the measured radius of gyration of the 37-mer is 14.2(±0.6) Å, which is to be compared with the radius of gyration of an ideal 37-mer α-helix of 17.6 Å. Attempts are made to analyze the origin of this discrepancy in terms of the analytical Zimm–Bragg–Nagai (ZBN) theory, as well as distributed computing explicit solvent molecular dynamics simulations using two variants of the AMBER force-field. The ZBN theory, which treats helices as cylinders connected by random walk segments, predicts markedly larger radii of gyration than those measured. This is true even when the persistence length of the random walk parts is taken to be extremely short (about one residue). Similarly, the molecular dynamics simulations, at the level of sampling available to us, give inaccurate values of the radii of gyration of the molecules (by overestimating them by around 25% for longer peptides) and/or their helical content. We conclude that even at the short sequences examined here (≤37 amino acid residues), these α-helical peptides behave as fluctuating semi-broken rods rather than straight cylinders with frayed ends.

Introduction

The α-helix is the most prevalent secondary structural element in proteins.1 However, in the absence of stabilizing tertiary interactions, α-helices rarely persist in isolation. One of the notable exceptions is a series of alanine-based, helix-forming peptides introduced by Robert Baldwin and his group.2, 3, 4 The study of these peptides has significantly furthered our understanding of the stability of α-helices,4, 5, 6 helical preference of individual amino acids,7, 8, 9 the role of helix-termination signals6, 10 and the kinetics of α-helix folding.11 In addition, these peptides have been studied extensively by computer simulations, which lead to a more detailed microscopic picture of α-helix folding and dynamics in general.12, 13, 14, 15, 16, 17, 18, 19, 20

The two most commonly used techniques to study α-helices in isolation, CD and NMR, are predominantly short distance-range techniques. They report on the local structure and dynamics of polypeptides while giving limited information on the long-range ordering and correlations. On the other hand, the small angle X-ray scattering technique (SAXS),21 with molecular radius of gyration as its principal readout, is a long-range technique. It reports on the global structure of a molecule and, as such, gives information complementary to that obtained by CD, NMR and other local techniques. This ability of SAXS to probe the long-range structure in polypeptides is shared by electron paramagnetic resonance22 and fluorescence resonance energy transfer23, 24 techniques, but with greater accuracy compared to these methods. In this study, we have used SAXS to measure the radii of gyration of six α-helix-forming peptides with composition Ace-(AAKAA)n-GY-NH2 (n=2–7) in aqueous solvent. The measured radii of gyration are smaller than those of the equivalent ideal α-helices or α-helices with frayed ends. This difference is analyzed in terms of the Zimm–Bragg–Nagai (ZBN) analytical theory25 and explicit solvent molecular dynamics (MD) simulations using world-wide distributed computing. The ZBN theory has been very successful in describing the behavior of long homopolymeric helices under various conditions,26 but fails to account for the low radii of gyration measured here. The ZBN theory describes helices as cylinders connected by random walk segments. It combines the standard Zimm–Bragg helix-coil theory27 with the statistics of ideal random walks. However, even if the persistence length of the random walk segments is taken to be very short (one amino acid), the ZBN theory predicts radii of gyration that are significantly larger than the values obtained by SAXS. The explicit solvent MD simulations described here were performed using distributed computing techniques and two variants of the AMBER force-field with a hope of finding a microscopic explanation for the low radii of gyration observed in the experiment. However, at the level of sampling used in this study (50 independent 50 ns long trajectories for each peptide and each force field), the radii of gyration of the simulated peptides deviate from the measured values, especially for the longest peptides where they are overestimated significantly. In addition, the secondary structure content from simulations deviates from the estimates based on CD. This suggests that capturing the behavior of flexible peptides is an area where modern atomistic force-fields could be improved. Our SAXS results are most consistent with alanine-based polypeptides adopting fluctuating semi-broken helix conformations, and are somewhat at odds with the picture of helices as straight cylinders with frayed ends.

Section snippets

Results

How helical are the six peptides that we studied? Figure 1 (a) shows the CD spectra of the molecules. All peptides other than AK12 give rise to standard α-helical profiles with double minima at around 208 nm and 222 nm. The 12-mer is known to be too short to form a sizable helix, and its CD spectrum is consistent with random-coil behavior with some small level of helix present (∼20%, see below). By measuring the mean molar ellipticity at 222 nm and using the standard length-dependent baseline,28

Discussion

The radii of gyration of the peptides studied here are significantly smaller than the radii of gyration of ideal α-helices with the same sequence length. The presence of 310 or π-helices, the occurrence of fraying at the ends and the potentially inaccurate ellipticity of the CD coil-state baseline do not provide a satisfactory explanation for this discrepancy. First, the 310-helix, which has been suggested as an important structural element in polyalanine peptides,29, 30 is more elongated

Peptide synthesis

The peptides (sequence composition Ace-(AAKAA)n-GY-NH2, with n=2–7) were synthesized at the Stanford University PAN facility using solid-phase synthesis using standard Fmoc synthetic chemistry. The final products of the synthesis and their purity were analyzed by reverse-phase HPLC and MALDI-TOF mass spectrometry. In the text, the peptides are referred to as AKxx, where xx corresponds to the number of amino acid residues in a given molecule.

Circular dichroism

Peptide stock solutions for CD measurements were made

Acknowledgements

We especially thank the thousands of Folding@Home contributors, without whom this work would not be possible. A complete list of contributors can be found at http://folding.stanford.edu. We thank Soenke Seifert and staff at APS for help with data collection. We thank Buzz Baldwin and Eric J. Sorin for useful discussions, and the anonymous referee for suggestions on how to improve the manuscript. This work was supported by grants from the NIH (R01GM62868), a postdoctoral fellowship by EMBO (to

References (56)

  • H.J.C. Berendsen et al.

    GROMACS: a message-passing parallel molecular dynamics implementation

    Comput. Phys. Commun.

    (1995)
  • T.E. Creighton
  • S. Marqusee et al.

    Helix stabilization by Glu–Lys+ salt bridges in short peptides of de novo design

    Proc. Natl Acad. Sci. USA

    (1987)
  • S. Marqusee et al.

    Unusually stable helix formation in short alanine-based peptides

    Proc. Natl Acad. Sci. USA

    (1989)
  • T.P. Creamer et al.

    Side-chain entropy opposes alpha-helix formation but rationalizes experimentally determined helix-forming propensities

    Proc. Natl Acad. Sci. USA

    (1992)
  • C.A. Rohl et al.

    Alanine is helix-stabilizing in both template-nucleated and standard peptide helices

    Proc. Natl Acad. Sci. USA

    (1999)
  • S. Padmanabhan et al.

    Relative helix-forming tendencies of nonpolar amino acids

    Nature

    (1990)
  • A. Chakrabartty et al.

    Helix propensities of the amino acids measured in alanine-based peptides without helix-stabilizing side-chain interactions

    Protein Sci.

    (1994)
  • T.P. Creamer et al.

    Alpha-helix-forming propensities in peptides and proteins

    Proteins: Struct. Funct. Genet.

    (1994)
  • W.A. Eaton et al.

    Fast kinetics and mechanisms in protein folding

    Annu. Rev. Biophys. Biomol. Struct.

    (2000)
  • V. Daggett et al.

    A molecular dynamics simulation of polyalanine: an analysis of equilibrium motions and helix-coil transitions

    Biopolymers

    (1991)
  • S. Huo et al.

    Direct computation of long time processes in peptides and proteins: reaction path study of the coil-to-helix transition in polyalanine

    Proteins: Struct. Funct. Genet.

    (1999)
  • A.E. Garcia et al.

    Alpha-helical stabilization by side chain shielding of backbone hydrogen bonds

    Proc. Natl Acad. Sci. USA

    (2002)
  • S. Chowdhury et al.

    Breaking non-native hydrophobic clusters is the rate-limiting step in the folding of an alanine-based peptide

    Biopolymers

    (2003)
  • H. Nymeyer et al.

    Simulation of the folding equilibrium of alpha-helical peptides: a comparison of the generalized Born approximation with explicit solvent

    Proc. Natl Acad. Sci. USA

    (2003)
  • A.E. van Giessen et al.

    Monte Carlo simulations of polyalanine using a reduced model and statistics-based interaction potentials

    J. Chem. Phys.

    (2005)
  • S. Doniach

    Changes in biomolecular conformation seen by small angle X-ray scattering

    Chem. Rev.

    (2001)
  • M.D. Rabenstein et al.

    Determination of the distance between two spin labels attached to a macromolecule

    Proc. Natl Acad. Sci. USA

    (1995)
  • Cited by (62)

    • Using small-angle X-ray scattering (SAXS) to study the structure of self-assembling biomaterials

      2018, Self-Assembling Biomaterials: Molecular Design, Characterization and Application in Biology and Medicine
    • Effect of external pulling forces on the length distribution of peptides

      2015, Biochimica et Biophysica Acta - General Subjects
      Citation Excerpt :

      The secondary structure of the peptide was assigned for each timeframe using the DSSPcont [42] criteria. This was then used to calculate the helicity (or helical fraction) of the peptide overall, with helicity defined as the fraction of 310, α, and π residues (i.e., G + H + I) [43]. It should be noted that the capping groups are included in the secondary structure assignment and are therefore a factor in the helicity fractions reported.

    • Harnessing the unique structural properties of isolated α-helices

      2014, Journal of Biological Chemistry
      Citation Excerpt :

      However, the breaks are unlikely to be spatially coordinated, resulting in far fewer conformations (α 1/L2) that bring the ends in close enough proximity to precipitate CaM-peptide interactions. Although this interpretation needs further testing, it is consistent with a previous SAXS study suggesting that certain α-helical peptides exhibit breaks along their length (38). The use of the ER/K linker to modulate protein-protein interactions was termed Systematic Protein Affinity Strength Modulation (SPASM).

    View all citing articles on Scopus
    View full text