Skip to main content
  • Research article
  • Open access
  • Published:

New statistical potential for quality assessment of protein models and a survey of energy functions

Abstract

Background

Scoring functions, such as molecular mechanic forcefields and statistical potentials are fundamentally important tools in protein structure modeling and quality assessment.

Results

The performances of a number of publicly available scoring functions are compared with a statistical rigor, with an emphasis on knowledge-based potentials. We explored the effect on accuracy of alternative choices for representing interaction center types and other features of scoring functions, such as using information on solvent accessibility, on torsion angles, accounting for secondary structure preferences and side chain orientation. Partially based on the observations made, we present a novel residue based statistical potential, which employs a shuffled reference state definition and takes into account the mutual orientation of residue side chains. Atom- and residue-level statistical potentials and Linux executables to calculate the energy of a given protein proposed in this work can be downloaded from http://www.fiserlab.org/potentials.

Conclusions

Among the most influential terms we observed a critical role of a proper reference state definition and the benefits of including information about the microenvironment of interaction centers. Molecular mechanical potentials were also tested and found to be over-sensitive to small local imperfections in a structure, requiring unfeasible long energy relaxation before energy scores started to correlate with model quality.

Background

Statistical potentials are widely used tools for protein structure analysis, modeling and quality assessment. Many different aspects and properties of these potentials have been explored during the last few decades including the different theoretical foundations to derive them, the representation of interaction centers and types of interactions, and the various models for defining the reference state. Combinations of various types and flavors of potentials are often used together in order to boost their performance. Initially, statistical potentials were based on statistical mechanics [1–3], however knowledge-based potentials now employ many other ideas, including the use of conditional probabilities to observe particular atom or residue distributions in specific conditions [4], linear programming techniques [5, 6], linear and quadratic programming on various decoy sets [7], or information theory [8].

Despite the seemingly similar formalism to derive statistical potentials in general, the alternative definitions may result in very different performances. The majority of statistical potentials are pairwise potentials. In addition, single body potentials, like the ones accounting for solvent accessibility [9], were reported, as well as multibody potentials [10–12]. Although pairwise potentials are frequently used in combination with other types of potentials to improve their performance, multibody potentials are much less used, apparently due to the high computational cost to apply them. In the present work we focus on pairwise potentials.

The majority of statistical potentials employ the Boltzmann law to convert the observed frequencies of interactions into potentials. These potentials are obtained as the ratio of observed and expected frequencies, where the expected frequencies are derived from a hypothetical reference state when no interactions occur. While the observed interactions can be counted in experimentally solved structures, hypothetical protein models without interactions, which serve as reference states, are solely imaginary. Therefore depending on their actual design they are potential sources of great variability in the performance of statistical potentials.

Quasi-chemical approximation, a popular model for defining the reference state [2, 13–16], uses molar fractions of the corresponding interacting centers to calculate the expected frequency of their pairs in the system without interactions and otherwise does not provide any other assumptions regarding their spatial distribution. This approach implies a homogeneous, infinite system, which of course is not true for proteins. An interesting attempt to account for the finite size of proteins was to substitute the corresponding dissipation of the atom density with a reduced effective dimensionality of the space [17, 18]. "Isotropic" reference state, which is based on the occurrence of interacting pairs of any type at the given condition (distance, angle, etc.), appears natural and was also widely used [4, 19, 20]. However, it also approximates the system as infinite and homogeneous. A reference state that is free of these limitations was recently developed on a basis of shuffled systems [21] and a similar approach was suggested in the DOPE [22] potential. The reference state in the DOPE potential was defined as a homogeneous ensemble of non-interacting atoms in a sphere with the radius equal to the radius of gyration of a sample native structure, whereas our Shuffled Reference State model preserved spatial positions of the interacting centers in proteins, while their identities were shuffled. Further improvements to DOPE potential have been reported later [23]. Some other definitions of reference state, such as the use of decoys [5], were also suggested.

Different representations of interaction centers were explored in statistical potentials. Two major classes of explored representations are residue level or atomic. The residue level representations use Cα, Cβ atoms or side chain centroids and are usually based on the 20 naturally occurring amino acids [2, 15, 24, 25], although both reduced [7, 26, 27] and extended [28, 29] amino acid alphabets were explored, where the extended alphabet further classifies each residue according to the possible secondary structure types. Another representation of interaction centers utilizes profile-based representation of amino acids residues [30]. For each protein a PSI-BLAST [31] generated alignment is used to create a position specific scoring matrix, which is converted into a set of evolutionary allowed amino acid residues for each position in the protein. Then these sets are used to derive potentials in a similar way to methods published by Melo [20] and Sippl [19]. Side-chain-to-backbone and side-chain to side-chain residue level potentials were also described [32]. All-heavy-atom representations based on reduced [20] and detailed all-atom protein representations were suggested [4]. More elaborate modifications of atomic alphabets consist of reduced set of atom types grouped by their chemical types and substitution states [33]. Micro-environments of atoms were distinguished by their chemical nature and by the counts of surrounding atoms. A potential function based on two interaction centers per residue [34] was also reported (all above examples employ one interaction center per residue). These two centers were Cα atoms and the side chain center of masses (Cα atom in the case of Glycine).

Various models of interactions were explored during the developments of statistical potentials. The most widely used ones are the distant-dependent potentials, which either treat all contacts uniformly within a cutoff distance [2, 15, 24, 25], or account for their radial distribution [9, 19–21, 35–37]. Similar to the distance-dependent potentials are the contact area [38] and packing density potentials proposed by Li and Liang, (unpublished but available for download from http://gila.bioengr.uic.edu/resources/geometric.html). Another frequently used interaction model is based on angular dependence. Distributions of backbone ϕ, ψ torsion angles [3, 32] as well as virtual κ, α angles [34, 36] were explored. Promising combination of these degrees of freedom depends on both distance and orientation, which became more widely used recently [39–42].

Comparative analysis of contact potentials demonstrated that majority of them can be approximated by simple sum of amino acid hydrophobicities, while the rest depends on the hydrophobicities as well as on electrostatic properties [43].

In addition to the variety of ways to derive potentials, some additional techniques to improve their accuracy have been proposed. A trivial source of errors in statistical potentials is sparse statistics. Two major workarounds were developed: the use of pseudo-counts [4] and a weighting scheme suggested by Sippl [19]. Pseudo-counts simply add a unity to every count to avoid a division by zero when calculating fractions and do not try to normalize potential values in the case of empty counts, which could result in arbitrarily high positive values in certain cases. The weighting scheme assigns the average of all interaction types to the potential in the case of an empty count.

Composite potentials combine various terms, which may include solvation, residue-level pairwise, atomic level pairwise, hydrogen bonding, steric, torsion or secondary structure packing. One such example is the Rosetta scoring function [44, 45]. Another, more recent example for a composite scoring function is QMEAN [28, 46], which consists of six different terms: a torsion angle potential, secondary structure-specific, distance-dependent residue and all-atom pairwise potentials, a solvation potential as well as terms accounting for agreement of predicted and calculated secondary structure and solvent accessibility. A combination of mean force potentials, which account for distributions of pseudo-bonds, pseudo-angles, pseudo-dihedrals and distances between centers of interactions was studied [34]. Another composite potential, utilizing both residue-level (Cα-based) [41] and its all-atom version [47] combines energy terms for distance-dependent pairwise interactions with orientation preference, hydrogen bonding, short-range interactions, packing, tri-peptide packing, three-body interactions, and solvation terms. Zhang and colleagues proposed a composite residue-level potential that consists of contact and local energy terms and employs a reduced alphabet of amino acids and a mapping of protein structures into a discrete state model [48]. The potential was generated by optimizing its components in order to guarantee a minimum energy gap between the native and decoy structures in a training set.

In the present work we perform a systematic comparison of many of the above listed scoring functions using a large and diverse decoy set that is based on models collected during various CASP experiments [49]. We analyze the differences in their performances of ranking protein models as a function of various flavors of scoring functions. Partially based on these results, we developed a novel residue level statistical potential that takes advantage of our earlier developed shuffled reference state definition [21] but utilizes orientation-dependent accounting for residue interactions. We demonstrate that this novel potential is highly competitive with other scoring functions.

Results and Discussion

Benchmarking potential functions

Evaluating the performance of various statistical potentials using protein-like decoys is not a trivial problem. Decoys must present a balanced range of difficulty or be specific for a particular task or property [50]. Some scoring functions identify the native structure easily among a set of decoys but perform very poorly when it comes to identifying the most accurate model from the rest of the decoys in the absence of the native structure. This can happen because of overtraining on native structures or because of significant structural differences between the decoys and the native structure. As a consequence, benchmarks that include the native structure in the decoy set may not be informative or challenging enough for most scoring functions. On the other hand, a decoy set without a native structure has its own limitations because it is not guaranteed that a decoy with the highest geometrical similarity to the native structure (e.g. lowest root mean square deviation) is also the one with the lowest energy. The model that is most similar to the native structure might have a higher energy due to some locally unfavorable features. Nevertheless, this approach seems more practical because scoring functions are typically used in scenarios when the native structure is not known and only a variety of possible alternative models are available.

Another problem arises when only one method or a limited number of methods is used to generate decoys, which is often the case for other available decoy sets [51–53]. In these cases a scoring function might be specific to implicit features of the decoy generation procedure but perform significantly worse if used to score decoys of different origin. These potential problems can be avoided with the use of a large number of targets in a decoy set and by a careful selection of decoy properties, such as using standardized similarity to the native structure and using a diversity of methods to generate decoy models.

In the present work, we tested scoring functions on decoys with and without the native structure, emphasizing on the latter set. "Global distance test - total score" (GDT_TS, which is (GDT_P1 + GDT_P2 + GDT_P4 + GDT_P8)/4, where GDT_Pn denotes percent of residues under distance cutoff ≤ nÅ) values [54, 55] were used to assess structural similarity of decoy models to the corresponding experimental solution structure of the target. Scores were binned in 2.5 GDT_TS units (i.e. models that are less than 2.5 GDT_TS units different from each other were considered indistinguishable), and bin numbers were used as rank values starting from the highest GDT_TS value. This scheme makes sets of decoys of different quality comparable to one another. Although the choice of 2.5 GDT_TS units for binning is subjective, any other value would be subjective to the same extent. Meanwhile this value provides enough granularity for a statistical survey, while groups together essentially indistinguishable models. However, when native structure is included in a decoy set, this approach may over-penalize mispredictions. The GDT_TS score of a native structure is 100 by definition and, according to the selection process, the closest model can be as low as 65 (see Methods for details). Therefore, if the native structure is included in the test set, it may be separated from the most accurate decoy model by a significant accuracy gap, up to 14 bins. Consequently, misrecognition of the native model, when it is included in the set, is heavily penalized. To overcome this effect, we always assign the rank of 2 to the first non-native structure, if the native one is present, regardless of the number of empty bins separating them.

Impact of different protein representations on performance

According the representation of interaction centers used scoring functions evaluated in this study can be classified in three major groups: (i) atom-based, i.e. all-heavy-atom or reduced set of atom types, namely "QMEAN-all_atom" [46], "OPUS_PSP" [47], "DOPE" [22], "dFIRE" [17], "Shortle2006" [35], VSCORE-pair [38], ANOLEA-like ("Melo-ANOLEA", "Melo-NL") [56], our "RF_HA", "RF_HA_SRS" [21], and "Liang-geometric" potentials, (ii) residue-based: "QMEAN-pairwise", "QMEAN-SSE_agree", "QMEAN-ACC_agree", "QMEAN-torsion", "QMEAN-solvation" [28], "Floudas-Ca" [5], "Floudas-CM" [6], "Dong-pair" [30] potentials, as well as potentials proposed in this work, "RF_CB_SRS_OD", "RF_CB_OD", "RF_CB_SRS", "RF_CB", and (iii) composite potentials: "PROSA-pair", "PROSA-combined" [57], "Rosetta" [44], "Shortle2005" [32], "QMEAN6" [28, 46], "OPUS_CA" [41], "VSCORE-combined" [38], and "PC2CA" [34]. Composite potential functions most often are defined as a linear combination of residue-based long-range potentials with different kinds of local potentials, which are in most cases residue-based as well. In addition to the knowledge based scoring functions, a molecular mechanics potential, CHARMM [58], as implemented in the NAMD [59] package, was also evaluated. In terms of protein representation CHARMM can be categorized as a composite all-atom potential. Models were evaluated after subjecting them to one or 1000 relaxation steps, indicated as "NAMD 1" and "NAMD 1000", respectively.

The results of the benchmarking survey of the scoring functions are shown in Table 1. Data are sorted by the average rank of the lowest energy decoy structure in the absence of the native structure. It is noticeable that the performance of different potentials varies significantly depending on the presence or absence of the native structure. In the presence of a native structure all-atom potentials are usually more sensitive (i.e., the RF_HA_SRS and the Shortle2006 potentials are the top two). Meanwhile, no interaction type preference is observed if the native structure is absent from the test set: residue or atom based or composite potentials all perform competitively. In addition, potentials with good performances in the presence of the native structure often exhibit rather mediocre performance if the native structure is removed from the decoy set. For instance RF_HA_SRS, our all-atom potential with shuffled reference state definition [21], is the best performing potential recognizing the native structure correctly in 137 out of 143 decoy sets but ranks only as the 6th best when tested on a set without the native structure (Table 1). Similarly the Shortle2006 potential, which is the second best recognizing the native structure among decoys ranks only 23rd among potentials when the native structure is removed. This may indicate that atomistic potentials are often over-trained to recognize native structures or, alternatively, it may indicate that side-chain placement by current modeling methods is not accurate enough. Indirect support for the former hypothesis is the observation that reduction in number of atom types by joining chemically equivalent but distinct by PDB nomenclature types like Phe-CD1 and Phe-CD2 atoms into one Phe-CD type results in the loss of potential performance (data not shown).

Table 1 Performance of various statistical potentials on models of CASP5-8 experiments.

Influence of different properties of scoring functions in test cases where the native structures are absent from the set of decoys is not as straightforward as it is in the case when the native structures are present. There is not a specific group of potentials that outperform others. The composite potential QMEAN6, with its individually evaluated all-atom term and components accounting for secondary structure and solvent accessibility agreement, is among the best performing potentials. The residue level RF_CB_SRS_OD potential proposed in the present work compares competitively in this test. However, QMEAN "agreement-based" terms perform rather modestly in the presence of the native structure, and all other functions discussed here (QMEAN6, QMEAN-all_atom and RF_CB_SRS_OD) underperform some other all-atom potentials (RF_HA_SRS and Shortle2006), as mentioned above.

Assessing statistical significance of performance differences

An important question in benchmarking various potentials is the assessment of the statistical significance of differences of their performances. We performed pairwise one-tailed Wilcoxon tests on results obtained in the absence of the native structure (Fig. 1). Potentials are sorted in the same order as in Table 1. Only p-values higher than 0.05 are shown, pointing out pairs of scoring functions that are not significantly different from one another. We employed the Wilcoxon test because the distributions of the calculated ranks of decoys that scored as best are highly different from normal. In this test the null hypothesis is that the ranks calculated by two methods under comparison share the same distribution and the one-sided alternative is that the ranks obtained with the method listed in the row of the Fig. 1 are lower than ones obtained with the method listed in the column.

Figure 1
figure 1

Statistical significance of rank distributions obtained with different methods. Only p-values obtained with one-tailed Wilcoxon test above 0.05 are shown. Alternative hypothesis in the test was that location of the distribution obtained with scoring function in a row is lower than the one in the column. Interaction centers are color coded in row titles as follows: black is used for residue-based potentials, blue is used for atom-based potentials, and red designates composite scoring functions. Certain interaction types are color coded in column titles, specifically, scoring functions making use of solvent accessibility (green), torsion angle dependence (blue), secondary structure dependence (red), and orientation-dependence (in italic).

The importance of reference state definition

A large group of various potentials, specifically QMEAN residue-based pairwise and accessibility agreement terms (QMEAN-SSE_agree, QMEAN-ACC_agree, QMEAN-pairwise), our atomic level potentials with shuffled (RF_HA_SRS) and classic (RF_HA) reference states, our residue-level potentials with and without orientation dependence and shuffled reference state (RF_SRS_CB_OD, RF_CB_SRS, RF_CB_OD, RF_CB), Rosetta++ scoring function (Rosetta), both pairwise (VSCORE-pair) and composite (VSCORE-combined) versions of VSCORE potential, PROSA2003 composite scoring function (PROSA-combined), as well as profile based distance-dependent potential from Dong group (Dong-pair) and OPUS_PSP potentials do not demonstrate statistically significant difference to one another (Fig. 1). However, one can speculate that p-values obtained for the residue-level regular (RF_CB_SRS) and orientation dependent (RF_CB_SRS_OD) potentials, both of which utilize a recently introduced shuffled reference state definition [21], are superior to (RF_CB), which is a potential based on a classic reference state definition (p-values of differences are 0.06 and 0.007, respectively). Orientation dependence is another important factor, which contributes significantly to the potential performance, resulting in statistically significant superiority the (RF_CB_SRS_OD) potential over classic reference state potential (RF_CB). It is also interesting to mention that the distribution of ranks obtained with OPUS_CA scoring function is located significantly lower on the rank scale than most other potentials in this group, whereas the average rank value calculated with this potential is in the middle of this group. This fact can be explained by the observation that OPUS_CA is able to score decoys with the highest GDT_TS values as the best ones in many more cases than other potentials in this group (Table 1). However, the relatively low average rank for this potential is because it exhibits a drastically high error in cases in which it fails to find the best structure.

The performance of the molecular mechanics based CHARMM potential depends on the number of steps of structure relaxation. The performance of the CHARMM is close to random after one and even after 1000 steps of relaxation. A further 10-fold increase in the energy minimization steps brings CHARMM performance to the middle of the group of similarly performing potentials, discussed above (average rank in the absence of native structure is 5.27, data not shown in the Table 1 and Fig. 1). However, the exceedingly high computational cost makes the use of such long minimizations impractical.

Effect of accounting for microenvironments on the performance

It is interesting to survey the common features among the best performing potentials. As we noted above, the choice of the type of interaction center (either atomistic or residue level potential or a composite scoring function) does not correlate with the performance. Indeed, one can see, from the color coding of interaction center types in raw titles (Fig. 1), that potentials of every kind can be found over the entire range of performances. The very small number of residue level potentials that are based on interaction centers other than Cβ atoms (Floudas-Ca, Floudas-CM, PC2CA) does not allow us to draw a conclusion about their performance. Meanwhile, some conclusions can be drawn from the effect of certain other features of scoring functions, such as the use of solvent accessibility, torsion angle, accounting for secondary structure and consideration of orientation dependence. The aforementioned features are color-coded in the column titles of Fig. 1. The secondary structure dependent functions (red) perform better than average, whereas torsion angle dependent functions (blue) perform worse than average. Potentials using information on solvent accessibility (green) and orientation dependence (shown in italic) do not show a clear advantage.

It is interesting to see if the performance of various scoring functions varies with the quality of the best available model for a given target. This dependence is plotted in Fig. 2, panels (A) (B) and (C) display the accuracy dependence of composite, all-atom and residue-based potentials, respectively. One can observe a general trend in the case of composite (Fig. 2A) and especially of all-atom potentials (Fig. 2B), according to which the performance improves with the improvement of the quality of the best available model. Noticeable exceptions are composite VSCORE and PROSA potentials, which perform visibly worse for the highest accuracy groups, when the best model has GDT_TS 95.0 or higher. These two potentials include solvent accessibility term in addition to their distance dependent terms. Solvent accessibility term may have limited benefit at this high accuracy level, when solvent accessibility of alternative models is essentially identical. Another example of such "reversed" dependence is OPUS_PSP, which is the best in the group of targets in the bin of 80.0, but its performance decreases as higher quality models become available. The group of residue-based potentials (Fig. 2C) does not show the above trends. Instead, this group collectively shows inferior performance for targets in 72.5 bin as compared to the 65.0 bin, as well as for targets in 87.5 bin as compared to 80.0 bin. An interesting exception is the performance of QMEAN-SSE_agree (using secondary structure dependent term) and QMEAN-ACC_agree (using solvent accessibility dependent term) potentials. Both are among the best ones for sets of targets with lower quality (65 and especially 72.5 bins), QMEAN-SSE_agree keeps its leading position up to 87.5 GDT_TS target group but looses its sensitivity as nearly perfect models of GDT_TS 95.0 or higher become available. This observation together with the outstanding performance of the QMEAN-all_atom potential, which is also a secondary structure dependent one, confirms the previous observations about the general benefit of incorporating secondary structure information in the potential function. However, the QMEAN-ACC_agree potential with solvent accessibility term loses its sensitivity much earlier. This behavior of the QMEAN-ACC_agree potential is in agreement with earlier discussed behavior of composite VSCORE and PROSA potentials, which also dependent on solvent accessibility.

Figure 2
figure 2

Performance of different potentials as a function of the quality of the best available model. Average rank calculated by (A) composite functions, (B) heavy-atom based functions, and (C) residue-based potential functions for targets having best model with GDT_TS better than 95.0 (11 targets), between 87.5 and 95.0 (25 targets), between 80.0 and 87.5 (32 targets), between 72.5 and 80.0 (47 targets) and between 65.0 and 72.5 (28 targets).

We also reviewed the performance of the scoring functions as a function of various structural classes. Because only 56 out of 143 targets are currently classified in the SCOP database [60, 61], the significance of such analysis is limited. We could not find a significant correlation between particular scoring function features and the fold classes (Additional file 1, Fig. A1). In general, all scoring functions show a better performance in case of α/β proteins, an average performance can be observed for all-α proteins and α+β proteins, while the worst performance is detected for all-β proteins.

Conclusions

In summary, the correct definition of the reference state used in statistical potentials is critical. In addition, there seems to be a benefit of including information on various protein microenvironments. An effective reference state definition should be free of systematic errors, as it is in our SRS model, and actual interactions should be a function of amino acid frequency variations caused by local microenvironments such as different secondary structure preferences, and other deviations of local characteristics from the average.

Methods

Set of proteins and parameters used to derive residue-level potentials

A novel distance dependent residue-level potential (RF_CB_SRS_OD), utilizing shuffled reference state [21] and featuring orientation dependence, was derived from a representative set of 375 globular proteins selected from the Protein Data Bank[62]. General procedure and details for the protein set selection are described previously for all-heavy-atom potential [21]. Briefly, the set was composed of X-ray solved structures of proteins of at least 50 residues long, which crystallographic resolution and R-value were better than 2.1 Ã… and 0.2, respectively; all PDB structures with incomplete, missing, modified, or nonstandard residues were excluded except structures that had missing residues in the terminal positions only; structures co-crystallized with ions were also discarded; additionally, the pairwise sequence identity between any two proteins in the set was required to be less than 40%.

Three additional potentials were generated in order to evaluate its improvement over the "isotropic" reference state and over unidirectional accounting for interacting pairs. Two of these additional potentials employed averaging over all residue types [19], where one was built as orientation-dependent (RF_CB and RF_CB_OD, respectively). A third potential (RF_SRS) was based on shuffled reference state, but lacked the orientation dependence. For all of the potentials the first bin for spatial separation spanned the distance between 0-4 Å and every next bin spanned a 1 Å increment thereafter. Cβ atoms were used for system representation. A virtual Cβ atom was built for Glycine residue. No minimal sequence separation between interacting residues was required. For sparse data treatment the scheme introduced by Sippl [19] was used.

In order to generate a shuffled reference state, randomized model sets were obtained by shuffling residue identities within each protein. Shuffling procedure was repeated 1000 times using different seed values for the random number generator. Potentials were derived as described previously [21]

For orientation-dependent potentials residue pairs were classified into three groups (Fig. 3): pairs with "parallel" Cα-Cβ vectors, pairs with "antiparallel" Cα-Cβ vectors facing each other, and ones with "antiparallel" Cα-Cβ vectors pointing away from each other. This definition of the orientation dependence is substantially simplified in comparison to ones suggested earlier [39, 42]. However, this simplification results in more representative statistics of contacts.

Figure 3
figure 3

Definition of residue orientation used to derive potentials. (A) interaction i → j is considered "parallel" if the scalar product of Cα-Cβ vectors for residues i (vector a) and j (vector b) is positive; (B) interaction i → j is considered "antiparallel i facing j" if scalar product of a and b is non-positive and scalar product of vector a and vector from Cα atom of residue i to Cα atom of residue j (vector c) is positive; (C) interaction i → j is considered as "antiparallel i pointing away from j" if both a·b and a·c scalar products are non-positive.

Potentials developed in the present study are labeled as "RF_CB", with "_SRS" suffix for shuffled reference state and "_OD" suffix for the orientation dependence.

Set of decoys

Predicted models for 143 targets collected from the CASP5-CASP8 experiments [49] were used as decoys, including a total of 2628 models produced by a large variety of groups and methods. These models were selected using the following procedure: (i) only all-atom models were used; (ii) the set for a given target was required to include at least one model with GDT_TS score upon superposition to the experimental solution structure 65.0 or better; (iii) all models for each target were clustered by their lengths, and models from the most populated cluster were used; (iv) models were binned by their GDT_TS scores with increments of 2.5 and one random representative was kept from each bin. Only targets for which the experimental solution is publicly available were kept. As a result, a ranked list of representative models was selected for each of the 143 targets. Fig. 4 gives an example of superimposition of models of different quality to the experimental structure. Lists of selected targets and their models along with corresponding GDT_TS values can be downloaded from our website http://www.fiserlab.org/potentials/casp_decoys

Figure 4
figure 4

Superimposition of models of different quality with the experimental solution structure. Experimental structure of the CASP8 target T0502 (violet) and its models, (A) METATASSER_TS5, GDT_TS = 80.357, (B) 3Dpro_TS4, GDT_TS = 60.204, and (C) panther_server_TS2, GDT_TS = 44.643 are shown as Cα traces. Those parts of the models, where the experimental positions are not known are colored white. Parts of models deviating from experimentally determined positions less than 4 Å are colored green, and the rest is colored bronze. Cα-Cα pseudobonds longer than 3.9 Å are shown thin. This plot has been generated using MOLSCRIPT software [63].

References

  1. Levitt M, Warshel A: Computer simulation of protein folding. Nature 1975, 253(5494):694. 10.1038/253694a0

    CAS  PubMed  Google Scholar 

  2. Miyazawa S, Jernigan RL: Estimation of effective interresidue contact energies from protein crystal structures: quasi-chemical approximation. Macromolecules 1985, 18: 534. 10.1021/ma00145a039

    CAS  Google Scholar 

  3. Pohl FM: Empirical protein energy maps. NatNew Biol 1971, 234(52):277.

    CAS  Google Scholar 

  4. Samudrala R, Moult J: An all-atom distance-dependent conditional probability discriminatory function for protein structure prediction. JMolBiol 1998, 275(5):895.

    CAS  Google Scholar 

  5. Rajgaria R, McAllister SR, Floudas CA: A novel high resolution Calpha--Calpha distance dependent force field based on a high quality decoy set. Proteins 2006, 65(3):726–741. 10.1002/prot.21149

    CAS  PubMed  Google Scholar 

  6. Rajgaria R, McAllister SR, Floudas CA: Distance dependent centroid to centroid force fields using high resolution decoys. Proteins 2008, 70(3):950–970. 10.1002/prot.21561

    CAS  PubMed  Google Scholar 

  7. Qiu J, Elber R: Atomically detailed potentials to recognize native and approximate protein structures. Proteins: Structure, Function, and Bioinformatics 2005, 61(1):44–55. 10.1002/prot.20585

    CAS  Google Scholar 

  8. Solis AD, Rackovsky S: Improvement of statistical potentials and threading score functions using information maximization. Proteins 2006, 62(4):892. 10.1002/prot.20501

    CAS  PubMed  Google Scholar 

  9. Sippl MJ: Boltzmann's principle, knowledge-based mean fields and protein folding. An approach to the computational determination of protein structures. JComputAided MolDes 1993, 7(4):473.

    CAS  Google Scholar 

  10. Ngan SC, Inouye MT, Samudrala R: A knowledge-based scoring function based on residue triplets for protein structure prediction. Protein Engineering Design and Selection 2006, 19(5):187. 10.1093/protein/gzj018

    CAS  Google Scholar 

  11. Krishnamoorthy B, Tropsha A: Development of a four-body statistical pseudo-potential to discriminate native from non-native protein conformations. Bioinformatics 2003, 19(12):1540–1548. 10.1093/bioinformatics/btg186

    CAS  PubMed  Google Scholar 

  12. Masso M, Vaisman II: Accurate prediction of enzyme mutant activity based on a multibody statistical potential. Bioinformatics 2007, 23(23):3155–3161. 10.1093/bioinformatics/btm509

    CAS  PubMed  Google Scholar 

  13. Miyazawa S, Jernigan RL: Self-consistent estimation of inter-residue protein contact energies based on an equilibrium mixture approximation of residues. Proteins 1999, 34(1):49. 10.1002/(SICI)1097-0134(19990101)34:1<49::AID-PROT5>3.0.CO;2-L

    CAS  PubMed  Google Scholar 

  14. Lu H, Skolnick J: A distance-dependent atomic knowledge-based potential for improved protein structure selection. Proteins 2001, 44(3):223. 10.1002/prot.1087

    CAS  PubMed  Google Scholar 

  15. Bahar I, Jernigan RL: Inter-residue potentials in globular proteins and the dominance of highly specific hydrophilic interactions at close separation. JMolBiol 1997, 266(1):195.

    CAS  Google Scholar 

  16. Skolnick J, Jaroszewski L, Kolinski A, Godzik A: Derivation and testing of pair potentials for protein folding. When is the quasichemical approximation correct? Protein Science 1997, 6(3):676.

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Zhou H, Zhou Y: Distance-scaled, finite ideal-gas reference state improves structure-derived potentials of mean force for structure selection and stability prediction. Protein Science 2002, 11(11):2714. 10.1110/ps.0217002

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Zhang C, Liu S, Zhou H, Zhou Y: An accurate, residue-level, pair potential of mean force for folding and binding based on the distance-scaled, ideal-gas reference state. Protein Science 2004, 13(2):400. 10.1110/ps.03348304

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Sippl MJ: Calculation of conformational ensembles from potentials of mean force. An approach to the knowledge-based prediction of local structures in globular proteins. JMolBiol 1990, 213(4):859.

    CAS  Google Scholar 

  20. Melo F, Feytmans E: Novel knowledge-based mean force potential at atomic level. JMolBiol 1997, 267(1):207.

    CAS  Google Scholar 

  21. Rykunov D, Fiser A: Effects of amino acid composition, finite size of proteins, and sparse statistics on distance-dependent statistical pair potentials. Proteins: Structure, Function, and Bioinformatics 2007, 67(3):559–568. 10.1002/prot.21279

    CAS  Google Scholar 

  22. Shen MY, Sali A: Statistical potential for assessment and prediction of protein structures. Protein Sci 2006, 15(11):2507–2524. 10.1110/ps.062416606

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Fitzgerald JE, Jha AK, Colubri A, Sosnick TR, Freed KF: Reduced Cbeta statistical potentials can outperform all-atom potentials in decoy identification. Protein Sci 2007, 16(10):2123–2139. 10.1110/ps.072939707

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Miyazawa S, Jernigan RL: Residue-residue potentials with a favorable contact pair term and an unfavorable high packing density term, for simulation and threading. J Mol Biol 1996, 256(3):623–644. 10.1006/jmbi.1996.0114

    CAS  PubMed  Google Scholar 

  25. Miyazawa S, Jernigan RL: An empirical energy potential with a reference state for protein fold and sequence recognition. Proteins 1999, 36(3):357. 10.1002/(SICI)1097-0134(19990815)36:3<357::AID-PROT10>3.0.CO;2-U

    CAS  PubMed  Google Scholar 

  26. Thomas PD, Dill KA: An iterative method for extracting energy-like quantities from protein structures. Proceedings of the National Academy of Sciences 1996, 93(21):11628. 10.1073/pnas.93.21.11628

    CAS  Google Scholar 

  27. Melo F, Marti-Renom MA: Accuracy of sequence alignment and fold assessment using reduced amino acid alphabets. Proteins 2006, 63(4):986–995. 10.1002/prot.20881

    CAS  PubMed  Google Scholar 

  28. Benkert P, Tosatto SC, Schomburg D: QMEAN: A comprehensive scoring function for model quality assessment. Proteins 2008, 71(1):261–277. 10.1002/prot.21715

    CAS  PubMed  Google Scholar 

  29. Zhang C, Kim SH: Environment-dependent residue contact energies for proteins. Proceedings of the National Academy of Sciences 2000, 97(6):2550. 10.1073/pnas.040573597

    CAS  Google Scholar 

  30. Dong Q, Wang X, Lin L: Novel knowledge-based mean force potential at the profile level. BMC Bioinformatics 2006, 7: 324. 10.1186/1471-2105-7-324

    PubMed  PubMed Central  Google Scholar 

  31. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ: Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res 1997, 25(17):3389–3402. 10.1093/nar/25.17.3389

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Fang Q, Shortle D: A consistent set of statistical potentials for quantifying local side-chain and backbone interactions. Proteins 2005, 60(1):90. 10.1002/prot.20482

    CAS  PubMed  Google Scholar 

  33. Summa CM, Levitt M, Degrado WF: An atomic environment potential for use in protein structure prediction. JMolBiol 2005, 352(4):986.

    CAS  Google Scholar 

  34. Fogolari F, Pieri L, Dovier A, Bortolussi L, Giugliarelli G, Corazza A, Esposito G, Viglino P: Scoring predictive models using a reduced representation of proteins: model and energy definition. BMC Struct Biol 2007, 7: 15. 10.1186/1472-6807-7-15

    PubMed  PubMed Central  Google Scholar 

  35. Fang Q, Shortle D: Protein refolding in silico with atom-based statistical potentials and conformational search using a simple genetic algorithm. JMolBiol 2006, 359(5):1456.

    CAS  Google Scholar 

  36. Reva BA, Finkelstein AV, Sanner MF, Olson AJ: Residue-residue mean-force potentials for protein structure recognition. Protein Eng 1997, 10(8):865. 10.1093/protein/10.8.865

    CAS  PubMed  Google Scholar 

  37. Hendlich M, Lackner P, Weitckus S, Floeckner H, Froschauer R, Gottsbacher K, Casari G, Sippl MJ: Identification of native protein folds amongst a large number of incorrect models. The calculation of low energy conformations from potentials of mean force. JMolBiol 1990, 216(1):167.

    CAS  Google Scholar 

  38. McConkey BJ, Sobolev V, Edelman M: Discrimination of native protein structures using atom-atom contact scoring. ProcNatlAcadSciUSA 2003, 100(6):3215.

    CAS  Google Scholar 

  39. Buchete NV, Straub JE, Thirumalai D: Orientational potentials extracted from protein structures improve native fold recognition. Protein Science 2004, 13(4):862. 10.1110/ps.03488704

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Buchete NV, Straub JE, Thirumalai D: Continuous anisotropic representation of coarse-grained potentials for proteins by spherical harmonics synthesis. JMolGraphModel 2004, 22(5):441.

    CAS  Google Scholar 

  41. Wu Y, Lu M, Chen M, Li J, Ma J: OPUS-Ca: a knowledge-based potential function requiring only Calpha positions. Protein Sci 2007, 16(7):1449–1463. 10.1110/ps.072796107

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Bagci Z, Kloczkowski A, Jernigan RL, Bahar I: The origin and extent of coarse-grained regularities in protein internal packing. Proteins: Structure, Function, and Bioinformatics 2003, 53(1):56–67. 10.1002/prot.10435

    CAS  Google Scholar 

  43. Pokarowski P, Kloczkowski A, Jernigan RL, Kothari NS, Pokarowska M, Kolinski A: Inferring ideal amino acid interaction forms from statistical protein contact potentials. Proteins 2005, 59(1):49. 10.1002/prot.20380

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Simons KT, Ruczinski I, Kooperberg C, Fox BA, Bystroff C, Baker D: Improved recognition of native-like protein structures using a combination of sequence-dependent and sequence-independent features of proteins. Proteins 1999, 34(1):82. 10.1002/(SICI)1097-0134(19990101)34:1<82::AID-PROT7>3.0.CO;2-A

    CAS  PubMed  Google Scholar 

  45. Rohl CA, Strauss CE, Misura KM, Baker D: Protein structure prediction using Rosetta. Methods Enzymol 2004, 383: 66–93. full_text

    CAS  PubMed  Google Scholar 

  46. Benkert P, Kunzli M, Schwede T: QMEAN server for protein model quality estimation. Nucleic Acids Research 2009. gkp322 gkp322

    Google Scholar 

  47. Lu M, Dousis AD, Ma J: OPUS-PSP: An Orientation-dependent Statistical All-atom Potential Derived from Side-chain Packing. Journal of Molecular Biology 2008, 376(1):288–301. 10.1016/j.jmb.2007.11.033

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Zhang J, Chen R, Liang J: Empirical potential function for simplified protein models: Combining contact and local sequence-structure descriptors. Proteins: Structure, Function, and Bioinformatics 2006, 63(4):949–960. 10.1002/prot.20809

    CAS  Google Scholar 

  49. Moult J, Fidelis K, Rost B, Hubbard T, Tramontano A: Critical assessment of methods of protein structure prediction (CASP)--round 6. Proteins 2005, 61(Suppl 7):3. 10.1002/prot.20716

    CAS  PubMed  Google Scholar 

  50. Handl J, Knowles J, Lovell SC: Artefacts and biases affecting the evaluation of scoring functions on decoy sets for protein structure prediction. Bioinformatics 2009, 25(10):1271–1279. 10.1093/bioinformatics/btp150

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Samudrala R, Levitt M: Decoys 'R' Us: a database of incorrect conformations to improve protein structure prediction. Protein Science 2000, 9(7):1399. 10.1110/ps.9.7.1399

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Lee MR, Tsai J, Baker D, Kollman PA: Molecular dynamics in the endgame of protein structure prediction. J Mol Biol 2001, 313(2):417–430. 10.1006/jmbi.2001.5032

    CAS  PubMed  Google Scholar 

  53. Tsai J, Bonneau R, Morozov AV, Kuhlman B, Rohl CA, Baker D: An improved protein decoy set for testing energy functions for protein structure prediction. Proteins 2003, 53(1):76–87. 10.1002/prot.10454

    CAS  PubMed  Google Scholar 

  54. Zemla A: LGA: A method for finding 3D similarities in protein structures. Nucleic Acids Res 2003, 31(13):3370–3374. 10.1093/nar/gkg571

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Zemla A, Venclovas C, Moult J, Fidelis K: Processing and evaluation of predictions in CASP4. Proteins: Structure, Function, and Genetics 2001, 45(S5):13–21. 10.1002/prot.10052

    Google Scholar 

  56. Ferrada E, Vergara IA, Melo F: A knowledge-based potential with an accurate description of local interactions improves discrimination between native and near-native protein conformations. Cell Biochem Biophys 2007, 49(2):111–124. 10.1007/s12013-007-0050-5

    CAS  PubMed  Google Scholar 

  57. Sippl MJ: Recognition of errors in three-dimensional structures of proteins. Proteins 1993, 17(4):355. 10.1002/prot.340170404

    CAS  PubMed  Google Scholar 

  58. Brooks BR, Bruccoleri RE, Olafson BD, States DJ, Swaminathan S, Karplus M: CHARMM: A program for macromolecular energy, minimization, and dynamics calculations. Journal of Computational Chemistry 1983, 4(2):187–217. 10.1002/jcc.540040211

    CAS  Google Scholar 

  59. Phillips JC, Braun R, Wang W, Gumbart J, Tajkhorshid E, Villa E, Chipot C, Skeel RD, Kale L, Schulten K: Scalable molecular dynamics with NAMD. Journal of Computational Chemistry 2005, 26(16):1781–1802. 10.1002/jcc.20289

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Murzin AG, Brenner SE, Hubbard T, Chothia C: SCOP: a structural classification of proteins database for the investigation of sequences and structures. J Mol Biol 1995, 247(4):536–540.

    CAS  PubMed  Google Scholar 

  61. Andreeva A, Howorth D, Chandonia JM, Brenner SE, Hubbard TJ, Chothia C, Murzin AG: Data growth and its impact on the SCOP database: new developments. Nucleic Acids Res 2008, (36 Database):D419–425.

    Google Scholar 

  62. Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE: The Protein Data Bank. Nucleic Acids Res 2000, 28(1):235. 10.1093/nar/28.1.235

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Kraulis PJ: MOLSCRIPT: A Program to Produce Both Detailed and Schematic Plots of Protein Structures. JApplCryst 1991, 24: 946–950.

    Google Scholar 

Download references

Acknowledgements

The Authors thank all colleagues who provided their software for evaluation. The authors also thank Joseph Dybas for reading and commenting on the manuscript.

This work was supported by NIH grant GM62519-04.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andras Fiser.

Additional information

Authors' contributions

DR and AF conceived and designed the study and wrote the manuscript. DR carried out the calculations. All authors read and approved the final version.

Electronic supplementary material

12859_2009_3585_MOESM1_ESM.PDF

Additional file 1: Figure A1. Performance of different potentials as a function of SCOP class definitions. Average ranks were obtained for target structures of specific SCOP classes using various scoring functions. Connecting lines facilitate visual tracking of results for a given scoring function. (PDF 20 KB)

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Rykunov, D., Fiser, A. New statistical potential for quality assessment of protein models and a survey of energy functions. BMC Bioinformatics 11, 128 (2010). https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2105-11-128

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2105-11-128

Keywords