International
Tables for
Crystallography
Volume F
Crystallography of biological macromolecules
Edited by M. G. Rossmann and E. Arnold

International Tables for Crystallography (2006). Vol. F. ch. 22.2, pp. 548-549   | 1 | 2 |

Section 22.2.5.3. Secondary structures

E. N. Bakera*

aSchool of Biological Sciences, University of Auckland, Private Bag 92-109, Auckland, New Zealand
Correspondence e-mail: ted.baker@auckland.ac.nz

22.2.5.3. Secondary structures

| top | pdf |

Secondary structures provide the means whereby the polar C=O and NH groups of the polypeptide chain can remain effectively hydrogen bonded when they are buried within a folded globular protein. In doing so, they provide the framework of folding patterns and account for the majority of hydrogen bonds within protein structures. The three secondary-structure classes (helices, β-sheets and turns) are each characterized by specific hydrogen-bonding patterns, which can be used for objective identification of these structures (Stickle et al., 1992[link]).

22.2.5.3.1. Helices

| top | pdf |

Helices have traditionally been defined in terms of their N—H···O=C hydrogen-bonding patterns as α-helices [(i \rightarrow i - 4)], 310-helices [(i \rightarrow i - 3)], or π-helices [(i \rightarrow i - 5)]; in an α-helix, for example, the peptide NH of residue 5 hydrogen bonds to the C=O of residue 1. In fact, the vast majority of helices in proteins are α-helices; 310-helices are rarely more than two turns (six residues) in length, and discrete π-helices have not been seen so far.

The residues within helices have characteristic main-chain torsion angles, (φ, ψ), of around (−63°, −40°) that cause the C=O groups to tilt outwards by about 14° from the helix axis (Baker & Hubbard, 1984[link]). This results in somewhat less linear hydrogen bonding than in the original Pauling model (Pauling et al., 1951[link]), with a degree of distortion towards 310-helix geometry. Thus, weak [i \rightarrow i - 3] interactions are often made in addition to the more favourable [i \rightarrow i - 4] hydrogen bonds, giving hydrogen-bond networks that may enhance helix elasticity (Stickle et al., 1992[link]). Tilting outwards also makes the C=O groups more accessible for additional hydrogen bonds from side chains or water molecules. For the α-type, [i \rightarrow i - 4] interactions, the hydrogen-bond angles at both donor and acceptor atoms are quite tightly clustered (N—H···O ∼157° and C=O···H ∼147°). The hydrogen-bond lengths in helices average 2.06 (16) Å (O···H) or 2.99 (14) Å (O···N) (Baker & Hubbard, 1984[link]).

Few helices are regular throughout their length. Many are curved or kinked such that one side (often the outer, solvent-exposed side) of the helix is opened up a bit and has longer hydrogen bonds (Blundell et al., 1983[link]; Baker & Hubbard, 1984[link]). The bends are often associated with additional hydrogen bonds from water molecules or side chains to C=O groups that are tilted out more than usual. Curved helices are normal in coiled-coil structures and can enable long helices to pack more effectively in globular structures. Sometimes a kink can be functionally important, as in manganese superoxide dismutase, where a kink in a long helix, incorporating a π-type [(i \rightarrow i - 5)] hydrogen bond, enables the optimal positioning of active-site residues (Edwards et al., 1998[link]).

The beginnings and ends of helices are sites of hydrogen-bonding variations which can be seen as characteristic `termination motifs'. At helix N-termini, 310-type [i\rightarrow i - 3] (or bifurcated [i\rightarrow i - 3] and [i\rightarrow i - 4]) hydrogen bonds are often found. At C-termini, two common patterns occur. In one, labelled [\alpha_{\rm C1}] by Baker & Hubbard (1984[link]), there is a transition from α-type, [i\rightarrow i - 4] to 310-type, [{i\rightarrow i - 3}] hydrogen bonding, often with genuine bifurcated hydrogen bonds, as in Fig. 22.2.2.1(b)[link], at the transition point. The other, labelled [\alpha_{\rm C2}] (Baker & Hubbard, 1984[link]) or referred to as the `Schellman motif' (Schellman, 1980[link]), has a π-type, [i\rightarrow i - 5] hydrogen bond coupled with a [3_{10}]-type, [i - 1\rightarrow i - 4] hydrogen bond; residue [i - 1] has a left-handed α configuration and is often Gly. The beginnings and ends of helices are also the sites of specific side-chain hydrogen-bonding patterns, referred to as N-caps and C-caps (Presta & Rose, 1988[link]; Richardson & Richardson, 1988[link]); these are described below.

22.2.5.3.2. β-sheets

| top | pdf |

β-sheets consist of short strands of polypeptide (typically 5–7 residues) running parallel or antiparallel and cross-linked by N—H···O=C hydrogen bonds. Although the (φ, ψ) angles of residues within β-sheets can be quite variable, the hydrogen-bonding patterns within these segments tend to be quite regular, as in the original Pauling models (Pauling & Corey, 1951[link]). Occasional β-bulges in the middle of β-strands can interrupt the hydrogen-bonding pattern (Richardson et al., 1978[link]), but otherwise disruptions occur only at the ends of strands. The hydrogen bonds in β-sheets appear to be slightly shorter than those in helices, by ∼0.1 Å, and also more linear (N—H···O ∼ 160°, compared with ∼157° in helices) (Baker & Hubbard, 1984[link]). There also appears to be no difference between parallel and antiparallel β-sheets in the hydrogen-bond lengths and angles.

22.2.5.3.3. Turns

| top | pdf |

By far the most common type of turn is the β-turn, a sequence of four residues that brings about a reversal in the polypeptide chain direction. Hydrogen bonding does not seem to be essential for turn formation, but a common feature is a hydrogen bond between the C=O group of residue 1 and the NH group of residue 4, a 310-type, [i\rightarrow i - 3] interaction. Turns are also often associated with characteristic side-chain–main-chain hydrogen-bond configurations (see below). The hydrogen bonds in turns tend to be longer and less linear than those in helices and β-sheets; in particular, the angle at the acceptor oxygen atom C—O···H is around 120° (Baker & Hubbard, 1984[link]).

In addition to β-turns, a small but significant number of γ-turns are found. In these three-residue turns, a hydrogen bond is formed between the C=O of residue 1 and the NH of residue 3, an [i\rightarrow i - 2] interaction. Although the approach to the acceptor oxygen atom is highly nonlinear (C—O···H ∼ 100°), the nonlinearity at the H atom is less pronounced (N—H···O ∼ 130–150°) (Baker & Hubbard, 1984[link]). γ-turns are again of several types, depending on the configuration of the central residue. The classic γ-turn, first recognised by Matthews (1972[link]) and Nemethy & Printz (1972[link]), has a central residue with (φ, ψ) angles around (70°, −60°), which puts it in the normally disallowed region of the Ramachandran plot. More common, however, are structures in which an [i\rightarrow i - 2] hydrogen bond is associated with a central residue with a configuration around (90°, −70°) (Baker & Hubbard, 1984[link]); these structures are not necessarily true turns in the sense of bringing about a sharp chain reversal, however.

22.2.5.3.4. Aspects of in-plane geometry

| top | pdf |

For hydrogen bonds involving sp2 donors and/or acceptors, optimal interaction is expected to occur when the donor D—H group and the acceptor lone-pair orbital are coplanar (Taylor et al., 1983[link]). Analysis of `in-plane' and `out-of-plane' components of N—H···O hydrogen bonds in proteins shows that these have characteristic values for different secondary structures (Artymiuk & Blake, 1981[link]; Baker & Hubbard, 1984[link]). The out-of-plane component is tightly clustered at ∼25° for helices and ∼60° for the most common β-turns (type I and type III), but is widely scattered around a mean of 0° for β-sheets. The latter reflects different twists or curvature of β-sheets. The large out-of-plane component for turns is consistent with a relatively weak interaction.

References

First citation Artymiuk, P. J. & Blake, C. C. F. (1981). Refinement of human lysozyme at 1.5 Å resolution. Analysis of non-bonded and hydrogen-bonded interactions. J. Mol. Biol. 152, 737–762.Google Scholar
First citation Baker, E. N. & Hubbard, R. E. (1984). Hydrogen bonding in globular proteins. Prog. Biophys. Mol. Biol. 44, 97–179.Google Scholar
First citation Blundell, T., Barlow, D., Borkakoti, N. & Thornton, J. (1983). Solvent-induced distortions and the curvature of α-helices. Nature (London), 306, 281–283.Google Scholar
First citation Edwards, R. A., Baker, H. M., Whittaker, M. M., Whittaker, J. W. & Baker, E. N. (1998). Crystal structure of Escherichia coli manganese superoxide dismutase at 2.1 Å resolution. J. Biol. Inorg. Chem. 3, 161–171.Google Scholar
First citation Matthews, B. W. (1972). The γ turn. Evidence for a new folded conformation in proteins. Macromolecules, 5, 818–819.Google Scholar
First citation Nemethy, G. & Printz, M. P. (1972). The γ turn, a possible folded conformation of the polypeptide chain. Comparison with the β turn. Macromolecules, 5, 755–758.Google Scholar
First citation Pauling, L. & Corey, R. B. (1951). Configurations of polypeptide chains with favoured orientations around single bonds: two new pleated sheets. Proc. Natl Acad. Sci. USA, 37, 729–740.Google Scholar
First citation Pauling, L., Corey, R. B. & Branson, H. R. (1951). The structure of proteins: two hydrogen-bonded helical configurations of the polypeptide chain. Proc. Natl Acad. Sci. USA, 37, 205–211.Google Scholar
First citation Presta, L. G. & Rose, G. D. (1988). Helix signals in proteins. Science, 240, 1632–1641.Google Scholar
First citation Richardson, J. S., Getzoff, E. D. & Richardson, D. C. (1978). The β-bulge: a common small unit of nonrepetitive protein structure. Proc. Natl Acad. Sci. USA, 75, 2574–2578.Google Scholar
First citation Richardson, J. S. & Richardson, D. C. (1988). Amino acid preferences for specific locations at the ends of α-helices. Science, 240, 1648–1652.Google Scholar
First citation Schellman, C. (1980). The alpha-L conformation at the ends of helices. In Protein folding, edited by R. Jaenicke, pp. 53–61. Amsterdam: Elsevier.Google Scholar
First citation Stickle, D. F., Presta, L. G., Dill, K. A. & Rose, G. D. (1992). Hydrogen bonding in globular proteins. J. Mol. Biol. 226, 1143–1159.Google Scholar
First citation Taylor, R., Kennard, O. & Versichel, W. (1983). Geometry of the N—H···O=C hydrogen bond. 1. Lone pair directionality. J. Am. Chem. Soc. 105, 5761–5766.Google Scholar








































to end of page
to top of page