research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL
BIOLOGY
ISSN: 2059-7983

The Buccaneer software for automated model building. 1. Tracing protein chains

CROSSMARK_Color_square_no_text.svg

aDepartment of Chemistry, University of York, Heslington, York YO10 5DD, England
*Correspondence e-mail: cowtan@ysbl.york.ac.uk

A new technique for the automated tracing of protein chains in experimental electron-density maps is described. The technique relies on the repeated application of an oriented electron-density likelihood target function to identify likely Cα positions. This function is applied both in the location of a few promising `seed' positions in the map and to grow those initial Cα positions into extended chain fragments. Techniques for assembling the chain fragments into an initial chain trace are discussed.

1. Introduction

Automated building of atomic models of protein structures from electron density is an important element of a high-throughput structure-solution environment and a useful tool in a non-automated environment. Current automated building tools incorporate a range of ideas, some of which have been purpose-designed for automation and others that have been adopted from techniques which already exist in graphical model-building programs.

The approach to automated model building described here incorporates one new technique, the use of an oriented electron-density likelihood target function to identify likely Cα positions, along with a range of methods adapted from existing approaches. The resulting combination of methods is very simple and yet shows significant promise as the basis for a new automated model-building system. Before the new developments are described in detail, some other approaches which have been influential in this work will be discussed.

1.1. Graphical model-building tools

Much of the fundamental work on which current automated model building depends is drawn from the work of Jones and coworkers (e.g. Jones, 2004[Jones, T. A. (2004). Acta Cryst. D60, 2115-2125.]) on graphical tools for model building and in particular the O software. Two approaches are used. The first involves the calculation of a `skeleton' of ridge-lines connecting peaks in the electron density. The skeleton is then interpreted in terms of Cα positions, which commonly occur near branch points in the skeleton (Jones et al., 1991[Jones, T. A., Zou, J.-Y., Cowan, S. W. & Kjeldgaard, M. (1991). Acta Cryst. A47, 110-119.]). The second involves the location of secondary-structure features, in particular helices and strands, by performing a six-dimensional rotation and translation search with an idealized fragment and evaluating the electron density at the atomic centres (Kleywegt & Jones, 1997[Kleywegt, G. J. & Jones, T. A. (1997). Acta Cryst. D53, 179-185.]). These features provide a starting point from which the rest of the protein chain may be traced.

Another important contribution by Jones and coworkers is the docking of the protein sequence to the main-chain trace (Zou & Jones, 1996[Zou, J.-Y. & Jones, T. A. (1996). Acta Cryst. D52, 833-841.]). This step provides validation of the chain direction and is often necessary before completion of the main-chain trace because flexible loops connecting the more easily interpreted core regions of the protein may not be visible in the electron density. Zou & Jones (1996[Zou, J.-Y. & Jones, T. A. (1996). Acta Cryst. D52, 833-841.]) score possible side-chain types by a combination of rotomer fitting and a real-space residual and then `slide' the known sequence against the residue scores to find the most likely match.

Oldfield went on to develop graphical chain-tracing tools to perform assisted and automated building (Oldfield, 2002[Oldfield, T. J. (2002). Acta Cryst. D58, 487-493.]). Secondary-structure features are identified by geometrical analysis of the skeleton ridge-lines and these features are then automatically grown to model the loop regions of the molecule by automatic identification of branch points in the skeleton which extend the chain fragments. The resulting method stands out from subsequent automatic procedures because of its speed. Oldfield suggests that the procedure is limited in most cases to data at resolutions better than 4.0 Å resolution (Oldfield, 2003[Oldfield, T. J. (2003). Acta Cryst. D59, 483-491.]).

1.2. Non-graphical model-building tools

Automated electron-density interpretation by the identification of atoms with electron-density peaks has a long history in the field of small-molecule direct methods and has also been applied to macromolecules at high resolution (see, for example, Sheldrick et al., 2001[Sheldrick, G., Hauptman, H., Weeks, C., Miller, R. & Usón, I. (2001). International Tables for Crystallography, Vol. F, edited by M. G. Rossmann & E. Arnold, pp. 333-351. Dordrecht: Kluwer Academic Publishers.]). The ARP/wARP package has extended this approach to work at successively lower resolutions (Morris et al., 2002[Morris, R. J., Perrakis, A. & Lamzin, V. S. (2002). Acta Cryst. D58, 968-975.]). At lower resolutions, atoms are not resolved and therefore individual atomic peaks disappear; however, it is still possible to construct (under-determined) atomic models which account for the observed data. Morris et al. (2002[Morris, R. J., Perrakis, A. & Lamzin, V. S. (2002). Acta Cryst. D58, 968-975.]) apply information about protein geometry to select plausible Cα atoms from these redundant models and then conduct an exhaustive search of possible routes through the resulting list of candidate Cα atoms to identify a best trace. This approach works reliably when data is available to 2.5 Å resolution and in some cases to worse resolutions (Cohen et al., 2004[Cohen, S. X., Morris, R. J., Fernandez, F. J., Ben Jelloul, M., Kakaris, M., Parthasarathy, V., Lamzin, V. S., Kleywegt, G. J. & Perrakis, A. (2004). Acta Cryst. D60, 2222-2229.]). Automated sequence docking and refinement lead to a near-complete model in many cases.

Another approach to the problem of limited resolution is to search for structures larger than atoms. The template-convolution method (Kleywegt & Jones, 1997[Kleywegt, G. J. & Jones, T. A. (1997). Acta Cryst. D53, 179-185.]) is an example of this and inspired the Fourier-based FFFear method (Cowtan, 1998[Cowtan, K. D. (1998). Acta Cryst. D54, 750-756.]), later applied as an electron-density-based likelihood function (Cowtan, 2001[Cowtan, K. D. (2001). Acta Cryst. D57, 1435-1444.]) for locating secondary-structure features and larger domains. A similar approach was later adopted by Terwilliger (2001[Terwilliger, T. C. (2001). Acta Cryst. D57, 1755-1762.]) for the location of secondary-structure elements, implemented in the RESOLVE phase-improvement and model-building software. The secondary-structure elements may then be grown and joined to complete the structure by adding residues in conformations consistent with geometrical constraints (Terwilliger, 2003[Terwilliger, T. C. (2003). Acta Cryst. D59, 38-44.]). One particularly powerful technique employed in RESOLVE is the building of two additional residues at a time, with the best combined electron-density fit for the pair of residues determining the final position for the first of the two residues. This `look-ahead' approach is more reliable than building a single residue on the basis of density alone.

The CAPRA software of Ioerger & Sacchettini (2002[Ioerger, T. R. & Sacchettini, J. C. (2002). Acta Cryst. D58, 2043-2054.]) has some significant parallels with the current work in that it uses pattern-recognition techniques to identify likely Cα positions in the electron-density map. An electron-density skeleton is calculated and orientation-invariant features of the electron density in a 4 Å sphere about a candidate point are processed using a neural network to identify which points on the skeleton are most likely to represent Cα positions. The chain is then traced by selecting connected candidate positions using the scores and geometrical constraints. This approach is effective at 2.8 Å resolution or better (Ioerger & Sacchettini, 2002[Ioerger, T. R. & Sacchettini, J. C. (2002). Acta Cryst. D58, 2043-2054.]).

2. Method

2.1. Overview

The approach to chain tracing described here is built on the idea of locating likely Cα positions and extending these into a chain. The first step resembles the CAPRA approach, but with one very significant difference: CAPRA locates likely Cα positions on the basis of orientation-independent density features, whereas Buccaneer uses an orientation-dependent measure. This has two benefits.

  • (i) The result of the search is a list of oriented amino-acid groups, rather than just positions. This provides additional directional information to assist the process of assembling the amino acids into chains.

  • (ii) Since orientation-dependent information is not being excluded from the identification of the Cα positions, the target function may be more sensitive. However, this is offset by limitations imposed upon the target function by the search algorithm.

One other difference between this and some previous implementations is that the whole calculation takes place in `crystal' space, in which the space-group symmetry and cell repeat are implicit. As a result, there is no need to `locate' the molecule in the cell before building is attempted, since all symmetry copies of any atom are by definition built simultaneously. The implementation in crystal space is a benefit of the use of the `Clipper' crystallographic libraries (Cowtan, 2003[Cowtan, K. D. (2003). IUCr Comput. Commun. Newsl. 2, 4-9.]).

The discussion here is given in terms of locating Cα groups, where a Cα group is considered to include the Cα atom, the bonded N, C and H atoms and the Cβ atom when present, these atoms forming a rigid group. However, exactly the same techniques are equally applicable to the location of planar peptide groups (Cα, C, O, N, Cα) or of nucleotides for the tracing of DNA and RNA.

Likely Cα positions will be located using a density-likelihood function, which will score possible positions and orientations in the electron-density map in a six-dimensional search. Each possible configuration will be scored according to how well the density features reproduce the density features of real Cα groups in a simulated electron-density map for a known structure.

A vital element of the calculation is the preparation of this simulated electron-density map. For the likelihood target function to be valid, the simulated electron density must be on the same scale, represent broadly similar thermal motion and have the same size and type of noise features as the electron-density map to be interpreted. This simulation process is a complex calculation in itself and is the basis for both the Pirate statistical phase-improvement software and the Buccaneer chain-tracing software and will be described in another paper (Cowtan, 2006[Cowtan, K. D. (2006). In preparation.]).

The whole calculation can therefore be described in terms of four steps.

  • (i) Finding initial Cα `seed' positions.

  • (ii) Growing `seed' positions into chain fragments.

  • (iii) Joining chain fragments into chains.

  • (iv) Pruning of clashing chains.

Each of these will be discussed in more detail in the following sections.

2.2. Finding initial Cα `seed' positions

The aim of the finding step is to locate a few very probable Cα positions in the electron-density map for use as seed points from which longer chains will be grown. This process is related to the location of α-helices in the RESOLVE model-building software. The location of the Cα `seed' positions requires a six-dimensional search in both position and orientation. The accomplishment of this search in reasonable time places some constraints on the type of target function which can be used. The approach adopted here is to use a target function for which the translational search may be achieved with a few fast Fourier transforms (FFTs) and to perform an FFT-based translation search for every possible orientation of the density target.

A suitable target function for this type of calculation is the weighted density agreement function described by Cowtan (1998[Cowtan, K. D. (1998). Acta Cryst. D54, 750-756.]), which described how a localized density agreement function could be efficiently calculated using FFTs. This work was extended (Cowtan, 2001[Cowtan, K. D. (2001). Acta Cryst. D57, 1435-1444.]) by the use of the same function to calculate a density likelihood function in the presence of noise. The analysis here follows the same approach, with the exception that the term introduced there to account for the noise in the target map is replaced by the use of a simulated noisy map in the construction of the likelihood target function.

The search function is constructed using Bayes theorem,

[P({\rm model}|{\rm data}) = {{P({\rm data}|{\rm model})P({\rm model})} \over {P({\rm data})}}. \eqno (1)]

In this case the data are the electron-density map and the model is a specific placement of the search density for the Cα group. Let F represent the case that the electron density arises from a correctly positioned and oriented Cα group and [{\overline F}] represent the case that the electron density arises from any other source (i.e. an incorrectly positioned Cα group or density arising from a completely different source). Then, the probability of a correctly positioned Cα group given an individual density value from the map is given by

[P[F|\rho({\bf x})] = {{P[\rho({\bf x})|F]P(F)}\over {P[\rho({\bf x})]}}. \eqno (2)]

P[ρ(x)] is the probability of the `observed' map density at x. It may be calculated as a marginal distribution of P[ρ(x), C], C ∈ (F, [{\overline F}]), i.e.

[\eqalignno {P[\rho({\bf x})] & = P[\rho({\bf x}), F] + P[\rho({\bf x}), {\overline F}] \cr & = P[\rho({\bf x})|F]P(F) + P[\rho({\bf x})|\overline {F}]P({\overline F}). & (3)}]

It is more likely that an electron-density value will arise from any other source than from a correctly positioned and oriented Cα group, therefore [P({\overline F})] will dominate over P(F). Neglecting both this first term and also the prior probabilities P(F) and [P({\overline F})], which will be assumed to be uniform, (2)[link] becomes

[P[F|\rho({\bf x})] \simeq {{P[\rho({\bf x})|F]}\over {P[\rho({\bf x})|\overline {F}]}}. \eqno (4)]

There are a number of Cα groups in the reference map, each represented by a different pattern of electron-density values in the region around it. For each position in the region of a standard Cα group placed at the origin, a distribution of electron densities will be calculated based on the different values appearing in that position relative to the Cα atom when considering all the Cα groups in the reference structure. The probability of a particular electron-density value given a particular correctly positioned Cα group will be approximated by a Gaussian whose mean is the expected electron density and whose variance is given by the variance of the distribution of densities at that position when calculated over all the Cα atoms in the simulated reference map. These will be termed ρfrag(x) and σfrag(x)2.

The probability of an observed density value arising from a correctly positioned Cα group is then

[P[\rho({\bf x})|F] \propto \exp \left \{-{{[\rho({\bf x})-\rho_{\rm frag}({\bf x}')]^{2}}\over {2 \sigma_{\rm frag}({\bf x}')^{2}}}\right \}, \eqno (5)]

where x′ is the coordinate relative to the Cα which maps to the point x in the map under the current translation and orientation of the Cα group.

The probability of an observed density arising from some other source than a correctly positioned Cα group is estimated from the simulated reference density map by examining the density in regions not correlated with Cα features (but avoiding solvent). If the mean and variance of such uncorrelated density are given by ρrand and σrand, then

[P[\rho ({\bf x})|{\overline F}] \propto \exp \left \{ -{{[\rho({\bf x})-\rho_{\rm rand}]^{2}}\over {2 \sigma_{\rm rand}^{2}}}\right \}. \eqno (6)]

Substituting these expressions in (4)[link] and discarding the constant terms gives

[\eqalignno {P[F|\rho({\bf x})] &\propto {{\exp {\displaystyle \left \{- {{ [\rho({\bf x}) - \rho_{\rm frag}({\bf x}')]^{2}}\over {2\sigma_{\rm frag}({\bf x}')^{2} }} \right \}}\over {\exp - {\displaystyle \left \{ {{[\rho ({\bf x}) - \rho_{\rm rand}]^{2}}\over {2 \sigma^{2}_{\rm rand}}}\right \}}}}} \cr & \propto \exp \left \{ - {{[\rho ({\bf x}) - \rho''({\bf x}')]^{2}}\over {2 \sigma''({\bf x}')^{2}}} \right \}, &(7)}]

where

[\rho''({\bf x}') = {{ \sigma_{\rm rand}^2\rho_{\rm frag}({\bf x}') - \sigma_{\rm frag}({\bf x}')^2 \rho_{\rm rand} }\over {\sigma_{\rm rand}^2 - \sigma_{\rm frag}({\bf x}')^2 }}]

and

[\sigma''({\bf x}')^2 = {{\sigma_{\rm frag}({\bf x}')^2\sigma_{\rm rand}^2 }\over { \sigma_{\rm rand}^2 - \sigma_{\rm frag}({\bf x}')^2 }}.]

Finally, the probability indications for the presence of a Cα group on the basis of each individual density value in the map are combined to give an overall indication of the probability of a the Cα group being present with the given translation and orientation,

[P(F|\rho) = \textstyle \prod \limits_{x} P[F|\rho({\bf x})]. \eqno (8)]

It is more convenient to calculate the logarithm of this expression,

[\eqalignno {\log P(F|\rho) & = {\textstyle \sum \limits_{x}} \log P[F|\rho({\bf x})] \cr & = {\textstyle \sum \limits_{x}} - \left \{ {{[\rho ({\bf x})-\rho''({\bf x}')]^{2}}\over {2 \sigma''({\bf x}')^{2}}} \right \} + c. & (9)}]

The resulting function may be efficiently calculated for a single orientation as a function of position in the cell using an FFT approach. Let the translation search function, which gives the agreement between the Cα group density (in the current orientation) and the electron density as a function of translation, be called t(x). As a simplification, let μ′′(x) = 1/[2σ′′(x′)2]. The search function may then be written as

[\eqalignno {t(x) & = \textstyle \sum \limits_{x'} \mu''(x') [\rho''(x') - \rho(x'-x)]^2 & (10) \cr & = \textstyle \sum \limits_{x'} \mu''(x')\rho''(x')^2 - 2\mu''(x')\rho''(x')\rho(x'-x) \cr &\ \quad +\ \mu''(x')\rho(x'-x)^2.}]

Note that in the expansion the first term is independent of x and so is only calculated once, whereas the second two terms are convolutions and may therefore be efficiently calculated in reciprocal space as follows,

[\eqalignno {t(x) & = \textstyle \sum \limits_y \mu''(y)\rho''(y)^2 + (1/V) {\cal F}\{{\cal F}^{-1} [\mu''(x)] {\cal F}^{-1} [\rho(x)^2] ^* \cr &\ \quad {-}\ 2{\cal F}^{-1} [\mu''(x)\rho''(x)] {\cal F}^{-1} [\rho(x)] ^*\}, & (11)}]

where [{\cal F}] represents the Fourier transform, [{\cal F}^{-1}] the inverse Fourier transform and * complex conjugation. If the Fourier coefficients of the density and squared density are pre-calculated, then the translation function may be calculated by three Fast Fourier Transforms (FFTs) per orientation. Since the Cα group has no symmetry, the FFTs must be performed in P1.

The electron-density target function for the location of a Cα group is determined by considering the electron density within a 4 Å sphere around each Cα in the simulated reference map. This radius was initially inspired by CAPRA (Ioerger & Sacchettini, 2002[Ioerger, T. R. & Sacchettini, J. C. (2002). Acta Cryst. D58, 2043-2054.]) and subsequent testing proved it to be a good choice for this method too. The density means and variances are calculated on a fine (0.5 Å) orthogonal grid.

Fig. 1[link] shows the mean and variance density for a typical search model. The mean density shows the expected pattern of density around the atoms of the Cα group, with weaker Cβ density and bulges in likely Cγ directions. However, the variance density shows more interesting features, in particular that the most conserved density is concentrated not only at the main-chain atomic sites, but also at low-density positions between the atoms. This highlights the power of the FFFear search function to select both high- and low-density positions. Note also the hollows around the Cβ at common Cγ sites.

[Figure 1]
Figure 1
Representation of Buccaneer target function for a Cα group showing regions of (a) high mean density and (b) low variance (i.e. strongly conserved) density. Figures generated using CCP4MG (Potterton et al., 2002[Potterton, E., McNicholas, S., Krissinel, E., Cowtan, K. & Noble, M. (2002). Acta Cryst. D58, 1955-1957.]).

The six-dimensional search is performed over every possible translation and orientation of the Cα group and the highest scoring matches are assumed to be correct. Each position and orientation is then refined by a simplex algorithm search and then stored as a `seed' position for chain growth. As a default, one seed position is stored for every five residues expected in the final model, although this parameter is not very critical.

2.3. Growing `seed' positions into chain fragments

The `seed' Cα groups are grown into chains by adding additional Cα groups both before and after the seed group in positions which optimize the log-likelihood fit to density for the new group while not disobeying the constraints of the Ramachandran plot. The same log-likelihood function is used for evaluating Cα positions added by growth as for the initial finding stage; however, it is now evaluated in real space for each candidate position and orientation instead of using the FFT approach.

For the purposes of this calculation, the Ramachandran plot classified by residue type and contoured at two levels: a frequency of >0.0005 rad−2 describing an `allowed' region and a frequency of >0.01 rad−2 describing a `favoured' region, using the imprecise but commonplace terminology.

The growing process proceeds as follows. To grow a single residue in the forward direction, a search is conducted over the `allowed' values of the Ramachandran angle ψ for the current residue and φ for the next residue. The angles are searched with a uniform angle step of 20°, rejecting any ψ values forbidden by the Ramachandran plot. (When building the first new residue in a chain, no information is available concerning the first φ.) Next, a second residue is built, using a coarser angle search of 30°, but again applying Ramachandran constraints. The best combined log-likelihood score for the two residues is used to select the position of the first residue. The second residue is discarded, having served its sole purpose in validating the position of the first.

This two-residue look-ahead approach is similar to that of Terwilliger (2003[Terwilliger, T. C. (2003). Acta Cryst. D59, 38-44.]). The Ramachandran data used here is from the `Top 500' structures database of Lovell et al. (2003[Lovell, S., Davis, I., Adrendall, W., de Bakker, P., Word, J., Prisant, M., Richardson, J. & Richardson, D. (2003). Proteins, 50, 437-450.]). For the first residue any `allowed' conformation for any residue type is allowed, whereas for the second only `favoured' conformations for non-Gly residues are allowed.

Building in the reverse direction occurs in exactly the same manner, except for the reversal of the Ramachandran angles.

A cutoff threshold for the log-likelihood function is required to determine when to stop growing the chain in either direction. This cutoff is established through an effective ad hoc procedure: for each of the initial seed points, three residues are grown in a forward direction. It is then assumed that 90% of the resulting terminal Cα atoms will be correct. The scores for the terminal Cα atoms are sorted and the value separating the worst 10% from the remainder is used as the cutoff. This is a crude ad hoc criterion which provides only a rudimentary coupling to the quality of the map; however, in practice it is effective in providing useful fragments for processing by the subsequent stages.

Several optimizations are used to improve the performance of this approach. For the full angle search, the log-likelihood function is approximated by only using a subset of the grid points in the calculation. Since each calculation requires a density interpolation from the target map, this saves a significant amount of time. The best 50 conformations of the first residue are used to build the second residue and the best 30 combined scores are then rescored using all the points in the log-likelihood function. Finally, the Ramachandran angles for the best solution are refined using a simplex algorithm search.

2.4. Joining chain fragments

At this stage of the model-building process the model consists of many overlapped chain segments which may or may not be consistent with one another. From these, a single consistent model must be constructed either for visual assessment or for use in conventional refinement programs. This is achieved in two steps: joining of consistent fragments, followed by pruning of inconsistent fragments.

The joining stage merges overlapped fragments wherever this is possible and makes some initial selections between fragments when multiple possible merges are possible. The calculation proceeds as follows.

Firstly, every chain segment is split into a series of overlapping fragments, each containing three residues; i.e. a chain of n residues is split into n − 2 fragments of three residues, with each tri-residue fragment overlapping its neighbours by two residues.

Next, multiple traces of the same chain segment are merged by combining any pair of tri-residues for which all three Cα atoms match to within 2.0 Å. The combination is achieved by averaging all the coordinates of each main-chain atom of each tri-residue. This leads to a model in which multiple consistent traces of the same chain segment have been removed.

Next, the tri-residues are examined to see how they can be reassembled into chains. A search is conducted over every pair of tri-residues to identify any pair for which the second and third Cα atoms of the first tri-residue match the first and second Cα atoms of the second tri-residue, to within 2.0 Å. Every such pair is marked as a potential join.

A problem arises when a single tri-residue joins to several possible precursors or successors. At this point a decision must be made about the correct routing of the chain. Following the example of Cohen et al. (2004[Cohen, S. X., Morris, R. J., Fernandez, F. J., Ben Jelloul, M., Kakaris, M., Parthasarathy, V., Lamzin, V. S., Kleywegt, G. J. & Perrakis, A. (2004). Acta Cryst. D60, 2222-2229.]), the different possible routings of the chain are considered and that which yields the longest non-looped chain is assumed to be correct. An assumption here is that tracing is more likely to skip residues than to insert extra residues.

Identification of the longest possible trace through a list of multiply linked tri-residue fragments is a problem of finding the longest path through a directed graph. This is a simple computational problem which is conventionally solved by a dynamic programming technique called `critical path analysis'. However, the conventional implementations must be adjusted to deal with the possibility of looped chains in the hypothetical trace. The implementation is therefore as follows.

Each remaining tri-residue is considered to be a numbered node in a directed graph which may have zero or more predecessors and zero or more successors.[link]

[Scheme 1]

The calculation is very fast and leads to the longest possible chain trace through the given fragments in the case where there are no loops. In the case of looped chains, the results are not guaranteed to be optimal; however, they are usually optimal or near-optimal.

Once a set of chains have been traced through the tri-residues, the final atomic coordinates are assembled from the coordinates of the successive tri-residues in the chain. A weighted combination of all the overlapped atoms is used in order to achieve a smooth transition from one tri-residue to the next and thus maintain connectivity in the merged model. The weighting of each tri-residue decreases linearly from the central Cα towards its extremities.

2.5. Pruning of clashing chain fragments

The previous step will have merged all consistent chain fragments, selecting a single path where fragments branch in different directions. There still remains the problem of inconsistent chain fragments, including two common cases: firstly the case where two fragments trace the same chain in opposite directions and secondly the case where chains cross or clash without any commonality.

Both of these cases are handled by a simple pruning step. Each chain is compared against every other chain, noting any cases where any pair of Cα atoms approach to within 2.0 Å. Any clashing Cα atoms are removed from the shorter chain. Any segments of the shorter chain which are less than five residues in length are then deleted.

One aim of this approach is to encourage chain tracing in the right direction, under the assumption that reversed chain traces will tend to be shorter than forward traces owing to the use of the Ramachandran constraint in the chain-growing step. (This is of course only true in loop regions since helices and strands may be traced in either direction without violating Ramachandran constraints; however, in practice the approach is effective.)

2.6. Results

The procedure has been implemented in a software package called Buccaneer using the Clipper crystallographic libraries (Cowtan, 2003[Cowtan, K. D. (2003). IUCr Comput. Commun. Newsl. 2, 4-9.]). The implementation is extremely simple, involving about 2000 lines of C++ code. The software as outlined here is incomplete in comparison to other software in the field: no refinement of the model or recycling takes place to complete the model and no sequence docking or side-chain building is performed. As a result, the software is not comparable to competing methods at this point. However, some initial results obtained using real data can provide some indications to the capabilities of the method.

The procedure was tested using 58 structures from the Joint Center for Structural Genomics (JCSG) data archive (Joint Center for Structural Genomics, 2006[Joint Center for Structural Genomics (2006). JCSG Data Archive. https://www.jcsg.org/datasets-info.shtml.]). This is a database of structures solved by largely automated methods. The chosen structures were solved using experimental phasing. For each structure, the JCSG software pursued multiple phasing paths using different software and parameters. A single initial phasing set was chosen for each structure by automatically selecting a structure on the basis of the statistics of the electron-density map. The selection criteria were crude, however, and so in some cases poor, low-resolution or even wrong phasing has been selected; all of these were kept as a means to test the behaviour of the software.

The selected set of experimental phases for each structure was then subjected to three cycles of phase improvement using the Pirate software (Cowtan, 2000[Cowtan, K. D. (2000). Jnt CCP4/ESF-EACBM Newsl. Protein Crystallogr. 38, 7.]). The resulting phases were used as a starting point for the Buccaneer chain-tracing calculation. The calculation for a single structure took between 2 and 30 min on a 2.4 GHz PC, depending on the volume of the asymmetric unit and resolution.

The quality of the starting data is described in terms of the data resolution and of the E-map correlation with the map from the final refined structure, the latter being a measure of phase error weighted by E value and figure of merit.

As an initial test, the performance of the `Cα-finding' step was examined. The six-dimensional FFFear search was used to identify the most probable Cα positions in both a good 2.0 Å map (PDB code 1z82) and a poorer low-resolution map (PDB code 1vrb). The likelihood score for the best matches was compared against the distance in angstroms from the candidate position to the nearest Cα in the solved structure. The results are shown in Fig. 2[link]. Note that with a good map the likelihood function accurately identifies Cα positions. In the poorer map there are a number of wrong positions identified along with the correct ones, although 75% of the candidate positions are still within 1.5 Å of a true Cα position. At lower resolutions, later stages in the chain-tracing calculation will have to remove fragments traced from incorrect candidates.

[Figure 2]
Figure 2
Distance to the nearest true Cα as a function of Buccaneer likelihood score (on an arbitrary scale; larger is better) for (a) a good high-resolution map (PDB code 1z92) and (b) a poorer low-resolution map (PDB code 1vrb). Results are plotted for the best Cαs found by the six-dimensional FFFear search in each case. The likelihood score is a reliable indicator of Cα position when density is good, but picks a mixture of good and bad positions when density is poor.

Next, the whole procedure was used to trace connected chains. The quality of the Buccaneer model is described in terms of the proportion of the known structure which was correctly built (i.e. completeness) and the proportion of the built model which was correct (i.e. accuracy). These were calculated by counting the proportion of real Cα atoms correctly built and the proportion of built Cα which were correct. For the purposes of this analysis, a correctly built Cα is one which is within 1.9 Å of a true Cα position in the known structure and has a neighbour which is in turn within 1.9 Å of a neighbouring Cα in the known structure. (1.9 Å was chosen as half the distance separating two Cα atoms, a registration error of up to half a residue.)

Table 1[link] describes the results for the 58 test structures in terms of number of residues, the quality measures of the starting data and the quality measures for the Buccaneer chain trace. The same data are visualized in Figs. 3[link], 4[link], 5[link] and 6[link].

Table 1
Results of applying Buccaneer chain tracing to density modified phases for 58 JCSG data sets

The columns give the deposition code, number of residues in the deposited model, resolution and E-map correlation for the starting density-modified phases and the completeness and accuracy of the resulting chain trace.

PDB code No. of residues Resolution (Å) E-map correlation Completeness (%) Accuracy (%)
1vr8 135 1.75 0.912 52 100
1vqs 551 1.80 0.807 60 100
1vp8 190 1.53 0.916 57 100
1vmg 83 1.46 0.808 85 100
1vlo 364 1.70 0.798 68 100
1vku 85 1.94 0.718 47 100
1vk4 283 1.91 0.882 63 100
1vr5 1072 1.73 0.805 81 99
1vme 802 1.80 0.923 75 99
1z82 624 2.00 0.786 90 98
1vlc 362 2.46 0.774 75 98
1vqz 330 1.99 0.686 78 97
1vpm 460 1.66 0.828 86 97
1vp4 836 1.82 0.671 74 97
1vkm 1752 2.60 0.800 79 97
1vkh 513 2.30 0.790 88 97
1vk2 190 2.10 0.786 73 97
1vmf 407 1.73 0.695 86 96
1vr0 708 2.49 0.807 72 95
1vmi 329 2.32 0.646 83 95
1vpb 437 1.80 0.883 78 94
1vp7 417 3.00 0.741 90 94
1vk8 373 2.00 0.705 58 93
1vjf 167 2.60 0.892 88 93
1vpy 251 2.40 0.681 85 92
1vlm 414 2.20 0.638 78 92
1vl5 906 1.85 0.724 78 92
1vjx 149 2.30 0.620 92 92
1vqy 847 2.40 0.687 74 91
1vkz 782 2.90 0.799 71 91
1vkn 1351 2.45 0.871 83 91
1z85 428 2.12 0.621 84 90
1vjz 325 2.50 0.792 85 89
1o6a 168 1.90 0.791 90 88
1vli 358 2.38 0.700 81 87
1vl6 1486 2.80 0.794 75 87
1vpz 113 2.05 0.701 69 85
1vk3 586 2.80 0.630 71 84
1vjv 367 2.65 0.900 70 84
1vr3 179 2.06 0.789 92 82
1vlu 792 3.00 0.644 76 81
1vrb 1224 3.20 0.628 59 80
1vl4 856 2.32 0.735 67 80
1vll 642 3.00 0.779 66 76
1vkd 1956 2.60 0.705 62 76
1vky 563 3.00 0.664 71 72
1vk9 147 2.70 0.713 89 67
1vjn 383 3.00 0.857 63 66
1vkb 147 1.90 0.704 69 64
1zej 282 2.00 0.751 84 53
1vjr 261 2.50 0.592 86 43
1vqr 1101 3.00 0.528 27 35
1vpg 301 2.10 0.450 3 23
1vpj 356 1.74 0.401 20 16
1vl0 842 2.50 0.222 14 15
1vkw 217 2.00 0.074 2 4
1vr9 242 2.00 0.042 0 1
1vjo 377 2.00 0.018 2 1
[Figure 3]
Figure 3
Completeness of the Buccaneer models for 58 JCSG test structures as a function of resolution.
[Figure 4]
Figure 4
Accuracy of the Buccaneer models for 58 JCSG test structures as a function of resolution.
[Figure 5]
Figure 5
Completeness of the Buccaneer models for 58 JCSG test structures as a function of the quality of the initial phases (E-map correlation).
[Figure 6]
Figure 6
Accuracy of the Buccaneer models for 58 JCSG test structures as a function of the quality of the initial phases (E-map correlation).

Completeness varies between 0 and 92% and accuracy between 0 and 100%. Note that neither completeness or accuracy vary strongly as a function of the resolution of the starting data. There is a slight drop in accuracy for the lowest resolution models, but completeness remains consistent. The method appears to be usable at least to the 3.2 Å low-resolution limit of the data available for these tests.

However, completeness and accuracy are strongly related to the quality of the initial phases. An initial E-map correlation of less than 0.6 leads to a poor model. From these results, it can be concluded that the method is not strongly sensitive to the data resolution, but is sensitive to the quality of the phases. Thus, the method appears to be complementary to ARP/wARP (Cohen et al., 2004[Cohen, S. X., Morris, R. J., Fernandez, F. J., Ben Jelloul, M., Kakaris, M., Parthasarathy, V., Lamzin, V. S., Kleywegt, G. J. & Perrakis, A. (2004). Acta Cryst. D60, 2222-2229.]), which is more sensitive to the data resolution but can give results with quite poor phases.

For the purposes of automated model building, it is common for a first model to be incomplete and to be extended in a recycling process with refinement and map calculation. The accuracy of the initial model is therefore probably more important than its completeness.

The model for the lowest resolution data set in the test, 1vrb (Joint Center For Structural Genomics, unpublished work), shows some interesting features. One subunit of the Buccaneer model from the 3.2 Å experimental phasing data set is shown in Fig. 7[link]. Note that helical regions of the molecule have been well traced; even at low resolution the precision of atom placement is high, often within 0.3 Å. This is to be expected since helical conformations are more common and more uniform than other conformations and so contribute more strongly to the likelihood density target. Non-helical regions are much more variable and the precision of the chain traces is accordingly much lower. This suggests a future approach involving the use of different density targets for growth in different regions of the Ramachandran plot.

[Figure 7]
Figure 7
Buccaneer trace of the A subunit of 1vrb at 3.2 Å. The true structure is shown in thin blue lines and the Buccaneer trace in thicker red lines.

Other differences between the Buccaneer and final models are worth noting. Fig. 8[link] shows a typical auto-tracing error where the chain trace has jumped between strands by means of side-chain density. Note this also illustrates how Buccaneer building can be counter-intuitive in comparison to programs which seek high density.

[Figure 8]
Figure 8
Buccaneer tracing error in the A subunit of 1vrb at 3.2 Å. The true structure is shown in thin black lines and the Buccaneer trace in thicker red lines.

Fig. 9[link] shows a place where Buccaneer has built a loop which was missing from the original model, along with the electron density. Comparison of the number of residues inserted against the sequence of the final model suggests that the trace is correct. The loop density is present, but the connectivity is only evident when the contour level is lowered. This case highlights another feature of Buccaneer. Since the likelihood target function keys on expected low-density features as well as high density features, Buccaneer is capable of building regions where the electron density is low. This feature has a cost: Buccaneer can also overinterpret solvent in terms of protein features in some cases. This can be seen in Table 1[link] in the cases where the completeness of the model is high but the accuracy is low (e.g. 1vjr), i.e. in addition to correctly tracing the protein region, Buccaneer has built protein chain in the solvent region. These cases can be trivially identified using the `Density fit analysis' feature of the Coot model-building program (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. D. (2004). Acta Cryst. D60, 2126-2132.]) and will be implemented in future developments of Buccaneer.

[Figure 9]
Figure 9
Buccaneer trace for a loop missing in the deposited structure. The true structure is shown in thin black lines and the Buccaneer trace in thicker red lines.

3. Conclusions

The chain-tracing approach described here is extremely simple, relying on the application of a single likelihood function in several different ways to trace protein main chains in experimentally phased electron-density maps. The method is reasonably fast, taking minutes to an hour, and can give a partial trace even at low resolutions (i.e. worse than 3.0 Å). However, the method is dependent on the quality of the initial experimental phasing and phase improvement.

The method as presented here is incomplete, lacking implementations for sequence docking, removal of incorrectly traced features, refinement of the resulting mode or recycling to model completion. However, the initial results suggest that the approach described here provides a suitable basis for future development.

Acknowledgements

The author would like to thank P. Emsley and E. Dodson for their helpful suggestions and the JCSG data archive for providing a source of well curated test data. This work was supported by The Royal Society under a University Research Fellowship.

References

First citationCohen, S. X., Morris, R. J., Fernandez, F. J., Ben Jelloul, M., Kakaris, M., Parthasarathy, V., Lamzin, V. S., Kleywegt, G. J. & Perrakis, A. (2004). Acta Cryst. D60, 2222–2229.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCowtan, K. D. (1998). Acta Cryst. D54, 750–756.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCowtan, K. D. (2000). Jnt CCP4/ESF–EACBM Newsl. Protein Crystallogr. 38, 7.  Google Scholar
First citationCowtan, K. D. (2001). Acta Cryst. D57, 1435–1444.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCowtan, K. D. (2003). IUCr Comput. Commun. Newsl. 2, 4–9.  Google Scholar
First citationCowtan, K. D. (2006). In preparation.  Google Scholar
First citationEmsley, P. & Cowtan, K. D. (2004). Acta Cryst. D60, 2126–2132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationIoerger, T. R. & Sacchettini, J. C. (2002). Acta Cryst. D58, 2043–2054.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationJoint Center for Structural Genomics (2006). JCSG Data Archive. https://www.jcsg.org/datasets-info.shtmlGoogle Scholar
First citationJones, T. A. (2004). Acta Cryst. D60, 2115–2125.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationJones, T. A., Zou, J.-Y., Cowan, S. W. & Kjeldgaard, M. (1991). Acta Cryst. A47, 110–119.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationKleywegt, G. J. & Jones, T. A. (1997). Acta Cryst. D53, 179–185.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationLovell, S., Davis, I., Adrendall, W., de Bakker, P., Word, J., Prisant, M., Richardson, J. & Richardson, D. (2003). Proteins, 50, 437–450.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMorris, R. J., Perrakis, A. & Lamzin, V. S. (2002). Acta Cryst. D58, 968–975.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationOldfield, T. J. (2002). Acta Cryst. D58, 487–493.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationOldfield, T. J. (2003). Acta Cryst. D59, 483–491.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPotterton, E., McNicholas, S., Krissinel, E., Cowtan, K. & Noble, M. (2002). Acta Cryst. D58, 1955–1957.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G., Hauptman, H., Weeks, C., Miller, R. & Usón, I. (2001). International Tables for Crystallography, Vol. F, edited by M. G. Rossmann & E. Arnold, pp. 333–351. Dordrecht: Kluwer Academic Publishers.  Google Scholar
First citationTerwilliger, T. C. (2001). Acta Cryst. D57, 1755–1762.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationTerwilliger, T. C. (2003). Acta Cryst. D59, 38–44.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZou, J.-Y. & Jones, T. A. (1996). Acta Cryst. D52, 833–841.  CrossRef CAS Web of Science IUCr Journals Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoSTRUCTURAL
BIOLOGY
ISSN: 2059-7983
Follow Acta Cryst. D
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds