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. 21.1, pp. 502-503   | 1 | 2 |

Section 21.1.7.2.2. Torsion angles (dihedrals)

G. J. Kleywegta*

aDepartment of Cell and Molecular Biology, Uppsala University, Biomedical Centre, Box 596, SE-751 24 Uppsala, Sweden
Correspondence e-mail: gerard@xray.bmc.uu.se

21.1.7.2.2. Torsion angles (dihedrals)

| top | pdf |

The conformation of the backbone of every non-terminal amino-acid residue is determined by three torsion angles, traditionally called φ ([{\rm C}_{i -1}\hbox{---}{\rm N}_{i}\hbox{---}{\rm C}^{\alpha}_{i}\hbox{---}{\rm C}_{i}]), ψ ([{\rm N}_{i}\hbox{---}{\rm C}^{\alpha}_{i}\hbox{---}{\rm C}_{i}\hbox{---}{\rm N}_{i+1}]) and ω ([{\rm C}^{\alpha}_{i}\hbox{---}{\rm C}_{i}\hbox{---}{\rm N}_{i+1}\hbox{---}{\rm C}^{\alpha}_{i+1}]). Owing to the peptide bond's partial double-bond character, the ω angle is restrained to values near 0° (cis-peptide) and 180° (trans-peptide). Cis-peptides are relatively rare and usually (but not always) occur if the next residue is a proline (Ramachandran & Mitra, 1976[link]; Stewart et al., 1990[link]). The average ω-value for trans-peptides is slightly less than 180° (MacArthur & Thornton, 1996[link]), but surprisingly large deviations have been observed in atomic resolution structures (Sevcik et al., 1996[link]; Merritt et al., 1998[link]). The ω angle therefore offers little in the way of validation checks, although values in the range of ±20 to ±160° should be treated with caution in anything but very high resolution models. The φ and ψ torsion angles, on the other hand, are much less restricted, but it has been known for a long time that owing to steric hindrance there are several clearly preferred combinations of φ, ψ values (Ramakrishnan & Ramachandran, 1965[link]). This is true even for proline and glycine residues, although their distributions are atypical (Morris et al., 1992[link]). Also, an overwhelming majority of residues that are not in regular secondary-structure elements are found to have favourable φ, ψ torsion-angle combinations (Swindells et al., 1995[link]). For these reasons, the Ramachandran plot (essentially a φ, ψ scatter plot) is an extremely useful indicator of model quality (Weaver et al., 1990[link]; Laskowski, MacArthur et al., 1993[link]; MacArthur & Thornton, 1996[link]; Kleywegt & Jones, 1996b[link]; Kleywegt, 1996[link]; Hooft et al., 1997[link]). Residues that have unusual φ, ψ torsion-angle combinations should be scrutinized by the crystallographer. If they have convincing electron density, there is probably a good structural or functional reason for the protein to tolerate the energetic strain that is associated with the unusual conformation (Herzberg & Moult, 1991[link]). As a rule, the residue types that are most often found as outliers are serine, threonine, asparagine, aspartic acid and histidine (Gunasekaran et al., 1996[link]; Karplus, 1996[link]). The quality of a model's Ramachandran plot is most convincingly illustrated by a figure. Alternatively, the fraction of residues in certain predefined areas of the plot (e.g. core regions) can be quoted, but in that case it is important to indicate which definition of such areas was used. Sometimes, one may also encounter a Balasubramanian plot, which is a linear φ, ψ plot as a function of the residue number (Balasubramanian, 1977[link]).

In protein structures, the plane of the peptide bond can have two different orientations (approximately related by a 180° rotation around the virtual Cα—Cα bond) that are both compatible with a trans configuration of the peptide (Jones et al., 1991[link]). The correct orientation can usually be deduced from the density of the carbonyl O atom or from the geometric requirements of regular secondary-structure elements (in α-helices, all carbonyl O atoms point towards the C-terminus of the helix; in β-strands, carbonyl O atoms usually alternate their direction). In other cases, e.g. in loops with poor density, the correct orientation may be more difficult to determine and errors are easily made. By comparing the local Cα conformation to a database of well refined high-resolution structures, unusual peptide orientations can be identified and, if required, corrected (through a `peptide flip'; Jones et al., 1991[link]; Kleywegt & Jones, 1997[link], 1998[link]). Since flipping the peptide plane between residues i and i + 1 changes the ψ angle of residue i and the φ angle of residue i + 1 by ∼180°, erroneous peptide orientations may also lead to outliers in the Ramachandran plot (Kleywegt, 1996[link]; Kleywegt & Jones, 1998[link]).

All amino-acid residues whose side chain extends beyond the Cβ atom contain one or more conformational side-chain torsion angles, termed χ1 (N—Cα—CβX γ, where X may be carbon, sulfur or oxygen, depending on the residue type; if there are two γ atoms, the χ1 torsion is calculated with reference to the atom with the lowest numerical identifier, e.g. Oγ1 for threonine residues), χ2 (Cα—CβX γX δ) etc. Early on, it was found that the values that these torsion angles assume in proteins are similar to those expected on the basis of simple energy calculations and that in addition certain combinations of χ1, χ2 values are clearly preferred (so-called rotamer conformations; Janin et al., 1978[link]; James & Sielecki, 1983[link]; Ponder & Richards, 1987[link]). Analogous to Ramachandran plots, χ1, χ2 scatter plots can be produced that show how well a protein's side-chain conformations conform to known preferences (Laskowski, MacArthur et al., 1993[link]; Carson et al., 1994[link]). Alternatively, a score can be computed for each residue that shows how similar its side-chain conformation is to that of the most similar rotamer for that residue type. This score can be calculated as an r.m.s. distance between corresponding side-chain atoms (Jones et al., 1991[link]; Zou & Mowbray, 1994[link]; Kleywegt & Jones, 1998[link]) or it can be expressed as an r.m.s. deviation of side-chain torsion-angle values from those of the most similar rotamer (Noble et al., 1993[link]).

Other torsion angles that have been used for validation purposes include the proline φ torsion (restricted to values near −65° owing to the geometry of the pyrrolidine ring; Morris et al., 1992[link]) and the χ3 torsion in disulfide bridges (defined by the atoms Cβ—S—S′—Cβ′ and restricted to values near +95 and −85°; Morris et al., 1992[link]). In addition to the torsion-angle values of individual residues, pooled standard deviations of χ1 and/or χ2 torsions have been used for validation purposes (Morris et al., 1992[link]; Laskowski, MacArthur et al., 1993[link]).

To assess the `geometric strain' in a model on a per-residue basis, the refinement program X-PLOR (Brünger, 1992b[link]) can produce geometric pseudo-energy plots. In such a plot, the ratio of Egeom(i)/r.m.s.(Egeom) is calculated as a function of the residue number i. The pseudo-energy term Egeom consists of the sums of the geometric and stereochemical pseudo-energy terms of the force field (Egeom = Ebonds + Eangles + Edihedrals + Eimpropers), involving only the atoms of each residue.

It has been observed that the more high-resolution protein structures become available, the more `well behaved' proteins turn out to be, i.e. the distributions of conformational torsion angles and torsion-angle combinations become even tighter than observed previously and the numerical averages tend to shift somewhat (Ponder & Richards, 1987[link]; Kleywegt & Jones, 1998[link]; EU 3-D Validation Network, 1998[link]; MacArthur & Thornton, 1999[link]; Walther & Cohen, 1999[link]).

References

First citation EU 3-D Validation Network (1998). Who checks the checkers? Four validation tools applied to eight atomic resolution structures. J. Mol. Biol. 276, 417–436.Google Scholar
First citation Balasubramanian, R. (1977). New type of representation for mapping chain-folding in protein molecules. Nature (London), 266, 856–857.Google Scholar
First citation Brünger, A. T. (1992b). X-PLOR. A system for crystallography and NMR. Yale University, New Haven, Connecticut, USA.Google Scholar
First citation Carson, M., Buckner, T. W., Yang, Z., Narayana, S. V. L. & Bugg, C. E. (1994). Error detection in crystallographic models. Acta Cryst. D50, 900–909.Google Scholar
First citation Gunasekaran, K., Ramakrishnan, C. & Balaram, P. (1996). Disallowed Ramachandran conformations of amino acid residues in protein structures. J. Mol. Biol. 264, 191–198.Google Scholar
First citation Herzberg, O. & Moult, J. (1991). Analysis of the steric strain in the polypeptide backbone of protein molecules. Proteins Struct. Funct. Genet. 11, 223–229.Google Scholar
First citation Hooft, R. W. W., Sander, C. & Vriend, G. (1997). Objectively judging the quality of a protein structure from a Ramachandran plot. Comput. Appl. Biosci. 13, 425–430.Google Scholar
First citation James, M. N. G. & Sielecki, A. R. (1983). Structure and refinement of penicillopepsin at 1.8 Å resolution. J. Mol. Biol. 163, 299–361.Google Scholar
First citation Janin, J., Wodak, S., Levitt, M. & Maigret, B. (1978). Conformation of amino acid side-chains in proteins. J. Mol. Biol. 125, 357–386.Google Scholar
First citation Jones, T. A., Zou, J.-Y., Cowan, S. W. & Kjeldgaard, M. (1991). Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Cryst. A47, 110–119.Google Scholar
First citation Karplus, P. A. (1996). Experimentally observed conformation-dependent geometry and hidden strain in proteins. Protein Sci. 5, 1406–1420.Google Scholar
First citation Kleywegt, G. J. (1996). Use of non-crystallographic symmetry in protein structure refinement. Acta Cryst. D52, 842–857.Google Scholar
First citation Kleywegt, G. J. & Jones, T. A. (1996b). Phi/Psi-chology: Ramachandran revisited. Structure, 4, 1395–1400.Google Scholar
First citation Kleywegt, G. J. & Jones, T. A. (1997). Model-building and refinement practice. Methods Enzymol. 277, 208–230.Google Scholar
First citation Kleywegt, G. J. & Jones, T. A. (1998). Databases in protein crystallography. Acta Cryst. D54, 1119–1131.Google Scholar
First citation Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993). PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Cryst. 26, 283–291.Google Scholar
First citation MacArthur, M. W. & Thornton, J. M. (1996). Deviations from planarity of the peptide bond in peptides and proteins. J. Mol. Biol. 264, 1180–1195.Google Scholar
First citation MacArthur, M. W. & Thornton, J. M. (1999). Protein side-chain conformation: a systematic variation of χ1 mean values with resolution – a consequence of multiple rotameric states? Acta Cryst. D55, 994–1004.Google Scholar
First citation Merritt, E. A., Kuhn, P., Sarfaty, S., Erbe, J. L., Holmes, R. K. & Hol, W. G. J. (1998). The 1.25 Å resolution refinement of the cholera toxin B-pentamer: evidence of peptide backbone strain at the receptor-binding site. J. Mol. Biol. 282, 1043–1059.Google Scholar
First citation Morris, A. L., MacArthur, M. W., Hutchinson, E. G. & Thornton, J. M. (1992). Stereochemical quality of protein structure coordinates. Proteins Struct. Funct. Genet. 12, 345–364.Google Scholar
First citation Noble, M. E. M., Zeelen, J. P., Wierenga, R. K., Mainfroid, V., Goraj, K., Gohimont, A. C. & Martial, J. A. (1993). Structure of triosephosphate isomerase from Escherichia coli determined at 2.6 Å resolution. Acta Cryst. D49, 403–417.Google Scholar
First citation Ponder, J. W. & Richards, F. M. (1987). Tertiary templates for proteins. Use of packing criteria in the enumeration of allowed sequences for different structural classes. J. Mol. Biol. 193, 775–791.Google Scholar
First citation Ramachandran, G. N. & Mitra, A. K. (1976). An explanation for the rare occurrence of cis peptide units in proteins and polypeptides. J. Mol. Biol. 107, 85–92.Google Scholar
First citation Ramakrishnan, C. & Ramachandran, G. N. (1965). Stereochemical criteria for polypeptide and protein chain conformations. II. Allowed conformations for a pair of peptide units. Biophys. J. 5, 909–933.Google Scholar
First citation Sevcik, J., Dauter, Z., Lamzin, V. S. & Wilson, K. S. (1996). Ribonuclease from Streptomyces aureofaciens at atomic resolution. Acta Cryst. D52, 327–344.Google Scholar
First citation Stewart, D. E., Sarkar, A. & Wampler, J. E. (1990). Occurrence and role of cis peptide bonds in protein structures. J. Mol. Biol. 214, 253–260.Google Scholar
First citation Swindells, M. B., MacArthur, M. W. & Thornton, J. M. (1995). Intrinsic φ, ψ propensities of amino acids, derived from the coil regions of known structures. Nature Struct. Biol. 2, 596–603.Google Scholar
First citation Walther, D. & Cohen, F. E. (1999). Conformational attractors on the Ramachandran map. Acta Cryst. D55, 506–517.Google Scholar
First citation Weaver, L. H., Tronrud, D. E., Nicholson, H. & Matthews, B. W. (1990). Some uses of the Ramachandran (φ, ψ) diagram in the structural analysis of lysozymes. Curr. Sci. 59, 833–837.Google Scholar
First citation Zou, J. Y. & Mowbray, S. L. (1994). An evaluation of the use of databases in protein structure refinement. Acta Cryst. D50, 237–249.Google Scholar








































to end of page
to top of page