Chapter 15

Algorithmic Methodologies for Discovery of Nonsequential Protein Structure Similarities

BHASKAR DASGUPTA, JOSEPH DUNDAS, and JIE LIANG

15.1 Introduction

An increasing number of protein structures are becoming available that either have no known function or whose functional mechanism is unknown or incomplete. Using experimental methods alone to explore these proteins in order to determine their functional mechanism is unfeasible. For this reason, much research has been put into computational methods for predicting the function of proteins [5, 14, 31, 34, 44, 53]. One such computational method is functional inference by homology, where annotations from a protein with known function are transferred onto another protein on the basis of sequence or structural similarities.

Protein sequence comparisons have been used as a straightforward method for functional inheritance. If two proteins have a high level of sequence identity, frequently the two proteins have the same or related biological functions. This observation has been used as a basis for transferring annotations from a protein that is well characterized to a protein with unknown function when the two proteins have high sequence similarity [3, 4, 45]. Frequently, only the protein residues that are near the functional region of the protein are under evolutionary pressure for conservation. Therefore, the global sequence similarity may be relatively low while local regions within the two sequences maintain a higher level of sequence similarity. In this case, probabilistic models such as profiles have been constructed using only the local regions of high sequence similarity [3, 32, 28]. Sequence comparison methods have the advantage that large numbers of sequences are deposited into sequence databases such as Swiss-Prot [11], which provides adequate information for constructing probabilistic models. However, a relatively high level of sequence similarity is needed in order to accurately transfer protein function. In fact, problems begin to arise when the sequence identity between a pair of proteins is c15-math-0001% [57].

Because proteins often maintain structural similarities even when sequence identity falls as low as 30% [6], making protein structure more strongly correlated with protein function than protein sequence [24], many researchers have begun to compare the three-dimensional structure of proteins in an attempt to uncover more distant evolutionary relationships among proteins. The SCOP [40] and CATH [43] databases have organized protein structures hierarchically into different classes and folds on the basis of their overall similarity in topology and fold. Classification of protein structures relies heavily on the reliable protein structure comparison methods. Common structural comparison methods include DALI (see Section 14.3.4) [27] and CE [47]. However, structural alignment methods cannot guarantee optimal results and do not have an interpretability comparable to sequence alignment methods.

Several challenges arise when trying to compare protein structures:

1. When searching for global structural similarity, similar to sequence alignment methods, one can search for global similarity or similarity within local surface regions of interest. Unlike sequence alignment scoring methods, which are heavily based on models of protein evolution [13, 25], scoring systems for structural alignment must account for both the 3D positional deviations between the aligned residues or atoms, and other biologically important shared characteristics. Defining a robust quantitative measure of similarity is challenging. This difficulty is illustrated by the variety of structural alignment scoring methods that have been proposed [23].
2. Many alignment methods assume that the ordering of the residues follows that of the primary sequence [47, 51]. This sequence order dependence can lead to problems when comparing local surface regions that often contain residues and atoms from different locations on primary sequence fold together to form functional regions in 3D space. On the global backbone level, the existence of permuted proteins, such as the circular permutation [17, 37] also poses significant problems for sequence order–dependent alignment methods.
3. Proteins may undergo small sidechain structural fluctuations or larger backbone fluctuations in vivo that are not represented in a single static snapshot of a crystallized structure in the Protein Data Bank (PDB) [9]. Many structural alignment methods assume rigid bodies and cannot factor in structural changes.

In this chapter, we will discuss several issues of structural alignment and then discuss methods that we have implemented for sequence order–independent structural alignment at the global and local surface levels. We illustrate the utility of our methods by showing how our sequence order–independent global structural alignment method detects circular permuted proteins. We then show how our local surface sequence order–independent structural alignment method can be used to construct a basis set of signature pockets of binding surfaces for a specific biological function. The signature pocket representsstructurally conserved surface regions. A set of signature pockets can then be used to represent a functional family of proteins for protein function prediction.

15.2 Structural Alignment

Comparing the structure of two proteins is an important problem [23] that may detect evolutionary relationships between proteins even when sequence identity between two proteins is relatively low. A widely used method for measuring structural similarity is the root-mean-squared distance (RMSD) between the equivalent atoms or residues of two proteins. If the equivalence relationship is known, a rotation matrix c15-math-0002 and a translation vector c15-math-0003 that when applied to one of the protein structures will minimize the RMSD can be found by solving the minimization problem

15.1 c15-math-0004

where c15-math-0005 is the number of points in structure c15-math-0006 and c15-math-0007 is the number of points in structure c15-math-0008. If c15-math-0009, then the least-squares estimation of the parameters c15-math-0010 and c15-math-0011 in this equation can be found using singular value decomposition.

The equivalence relationship is rarely known a priori when aligning to protein structures. In this case, the structural alignment method consists in minimizing RMSD while maximizing the number of aligned points. Heuristics must be used to solve this multiobjective optimization problem.

A number of heuristic methods have been developed [1, 22, 48, 49, 52, 56, 59] that can be divided into two main categories. Global methods are used to detect similarities between the overall fold of two proteins, and local alignment methods are used to detect similarities within local regions of interest within the two proteins. Most current methods are restricted to finding structural similarities only where the order of the structural elements within the alignment follows the order of the elements within the primary sequence. Sequence order–independent methods ignore the sequential ordering of the atoms or residues in primary sequence. These methods are better suited for finding more complex global similarities and can also be employed for finding all atoms local comparisons. We have implemented both sequence order–independent methods for both global and local structural alignments.

15.3 Global Sequence Order–Independent Structural Alignment

Looking for similarities between the overall fold can elucidate evolutionary or functional relationships between two proteins. However, most of the current methods for structural comparison are sequence order–dependent and are restricted to comparison of similar topologies between the two backbones. It has been discovered that throughout evolution, a genetic event can rearrange the topology of the backbone. One such example is regarded circular permutation. Conceptually, a circular permutation can be as a ligation of the c15-math-0012 and c15-math-0013 termini of a protein and cleavage somewhere else on the protein backbone. It has been observed that circular permutations often maintain a similar 3D spatial arrangement of secondary structures. In addition to circular permutations, research has shown that more complex topological rearrangements are possible [37]. Detection of these permuted proteins will be valuable for studies in homology modeling, protein folding, and protein design.

15.3.1 Sequence Order–Independent Global Structural Alignment

We have developed a sequence order–independent structural alignment algorithm for detecting structural similarities between two proteins that have undergone topological rearrangement of their backbone structures [17]. Our method is based on fragment assembly where the two proteins to be aligned are first exhaustively fragmented. Each fragment c15-math-0014 from protein structure c15-math-0015 is pairwise superimposed onto each fragment c15-math-0016 from protein structure c15-math-0017. The result is a set of fragment pairs c15-math-0018, where c15-math-0019 and c15-math-0020 are the indices in the primary sequence of the first residue of the two fragments. The variable c15-math-0021 is the length of the fragment. Each fragment pair is assigned a similarity score

15.2 c15-math-0022

where cRMSD is the measured RMSD value after optimal superposition, c15-math-0023 and c15-math-0024 are two constants, c15-math-0025 is a scaling factor to the measured RMSD values that depends on the secondary structure of the fragments, and SCS is a BLOSUM (blocks substitution matrix)-like measure of similarity in sequence of the matched fragments [25]. Details of the scoring method can be found in an earlier study [17].

The goal of the structural alignment is to find a consistent set of fragment pairs c15-math-0026 that minimizes the overall RMSD. Finding the optimal combination of fragment pairs is a special case of the well-known maximum-weight-independent set problem in graph theory. This problem is MAX-SNP-hard. We employ an approximation algorithm that was originally described for the scheduling of split-interval graphs [8] and is itself based on a fractional version of the local ratio approach.

To begin, a conflict graph c15-math-0027 is created, where a vertex is defined for each aligned fragment pair. Two vertices are connected by an edge if any of the fragments c15-math-0028 or c15-math-0029 from the fragment pair is not disjoint, that is, if both fragments from the same protein share one or more residues. For each vertex representing aligned fragment pairs, we assign three indicator variables c15-math-0030, and c15-math-0031, and a closed neighborhood c15-math-0032 indicates whether the fragment pair should be used (c15-math-0033 or not (c15-math-0034) in the final alignment. c15-math-0035 and c15-math-0036 are artificial indicator values for c15-math-0037 and c15-math-0038, which allow us to encode consistency in the selected fragments. The closed neighborhood of a vertex c15-math-0039 of c15-math-0040 is c15-math-0041, which is simply c15-math-0042 and all vertices that are connected to c15-math-0043 by an edge.

The sequence order–independent structural alignment algorithm can be described as follows. To begin, initialize the structural alignment c15-math-0044 equal to the entire set of aligned fragment pairs. We then

1. Solve a linear programming (LP) formulation of the problem:

15.3 c15-math-0045

15.4 c15-math-0046

15.5 c15-math-0046a

15.6 c15-math-0046b

15.7 c15-math-0046c

15.8 c15-math-0046d

2. For every vertex c15-math-0047 of c15-math-0048, compute its local conflict number c15-math-0049. Let c15-math-0050 be the vertex with the minimum local conflict number. Define a new similarity function c15-math-0051 from c15-math-0052 as follows:

equation

3. Create c15-math-0054 by removing from c15-math-0055 every substructure pair c15-math-0056 such that c15-math-0057. Push each removed substructure on to a stack in arbitrary order.
4. If c15-math-0058 then repeat from step 1, setting c15-math-0059 and c15-math-0060. Otherwise, continue to step 5.
5. Repeatedly pop the stack, adding the substructure pair to the alignment as long as the following conditions are met:
a. The substructure pair is consistent with all other substructure pairs that already exist in the selection.
b. The cRMSD of the alignment does not change beyond a threshold. This condition bridges the gap between optimizing a local similarity between substructures and optimizing the tertiary similarity of the alignment. It guarantees that each substructure from a substructure pair is in the same spatial arrangement in the global alignment.

15.3.2 Detecting Permuted Proteins

This algorithm was implemented in a large-scale study to search for permuted proteins in the Protein Data Bank (PDB) [9]. A subset of 3336 protein structures taken from the PDBSELECT90 dataset [26] are structurally aligned in a pairwise fashion. From the subset of 3336 proteins, we aligned two proteins if they met the following conditions (see Ref. 17 for details):

1. The difference in their lengths was no more than 75 residues.
2. The two proteins shared approximately the same secondary structure content.

Within the approximately 200,000 structural alignments performed, we found many known circular permutations and three novel circular permutations, as well as a more complex pair of noncyclic permuted proteins. Here we describe the details of the circular permutation that we found between a neucleoplasmin core and an auxin binding protein, as well as details of the more complex noncyclic permutation.

15.3.2.1 Nucleoplasmin Core and Auxin Binding Protein

We found a novel circular permutation between the nucleoplasmin core protein in Xenopu laevis (PDB ID 1k5j, chain E) [19] and the auxin binding protein in maize (PDB ID 11rh, chain A, residues 37–127) [58]. The structural alignment between 1k5jE (Fig. 15.1, top) and 11rhA (Fig. 15.1, bottom) consisted of 68 equivalent residues superimposed with a RMSD of 1.36 Å. This alignment is statistically significant with a c15-math-0061 after Bonferroni correction. Details of the c15-math-0062 value calculation can be found in our earlier study [17]. The short loop connecting two antiparallel strans in nucleoplasmin core protein (in circle, top of Fig. 15.1b) becomes disconnected in auxin binding protein 1 (in circle, bottom of Fig. 15.1b), and the N and C termini of the nucleoplasmin core protein (in square, top of Fig. 15.1b) are connected in auxin binding protein 1 (square, bottom of Fig. 15.1b). For details of other circular permutations we found, including permutations between microphage migration inhibition factor and the c15-math-0063-terminal domain of arginine repressor, please see our earlier study [17].

Figure 15.1 A newly discovered circular permutation between nucleoplasmin core [1k5j, chain E, top panel), and a fragment of auxin binding protein 1 (residues 37–127) (11rh, chain A, bottom panel). (a) These two proteins align well with a RMSD value of 1.36 Å over 68 residues, with a significant c15-math-0064 after Bonferroni correction. (b) The loop connecting strands 4 and 5 of nucleoplasmin core (in rectangle, top) becomes disconnected in auxin binding protein 1. The c15-math-0065 and c15-math-0066 termini of nucleoplasmin core (in rectangle, top) become connected in auxin binding protein 1 (in rectangle, bottom). To facilitate visualization of the circular permutation, residues in the c15-math-0067-to-c15-math-0068 direction before the cut in the nucleoplasmin core protein are colored by the lighter shade, and residues after the cut are colored by the darker shade. (c) The topology diagram of these two proteins. In the original structure of nucleoplasmin core, the electron density of the loop connecting strands 4 and 5 is missing in the PDB structure file. (This figure is modified from Hruscedila_lowercaseka et al. [17].)

c15f001

15.3.2.2 Complex Protein Permutations

Because of their relevance in understanding the functional and folding mechanism of proteins, circular permutations have received much attention [37, 55]. However, the possibility of more complex backbone rearrangements were experimentally verified by artificially rearranging the topology of the ARC repressor and were found to be thermodynamically stable [50]. Very little is known about this class of permuted proteins, and the detection of noncyclic permutations is a challenging task [2, 15, 29, 46].

Our database search uncovered a naturally occurring noncyclic permutation between chain F of AML1/core binding factor (AML1/CBF, PDB ID 1e50, Fig. 15.2a, top) and chain A of riboflavin synthase (PDB ID 1pkv, Fig. 15.2a, bottom). The two structures align well with an RMSD of 1.23 Å at an alignment length of 42 residues, with a significant c15-math-0080 after Bonferroni correction.

Figure 15.2 A noncyclic permutation discovered between AML1/core binding factor (AML1/CBF PDB ID 1e50, chain F, top) and riboflavin synthase (PDB ID 1pkv, chain A, bottom). (a) These two proteins structurally align with an RMSD of 1.23 Å over 42 residues and have a significant c15-math-0069 after Bonferroni correction. The residues that were assigned equivalences from the structural alignment are colored by the darker shade. (b) These proteins are structurally related by a complex permutation. The steps to transform the topology of AML1/CBF (top) to riboflavin (bottom) are as follows: (c) Remove the loops connecting strand 1 to helix 2, strand 4 to strand 5, and strand 5 to helix 6; (d) Connect the c15-math-0070-terminal end of strand 4 to the original c15-math-0071 termini; (e) connect the c15-math-0072-terminal end of strand 5 to the c15-math-0073-terminal end of helix 2; (f) connect the original c15-math-0074-termini to the c15-math-0075-terminal end of strand 5. The c15-math-0076-terminal end of strand 6 becomes the new c15-math-0077 termini, and the c15-math-0078-terminal end of strand 1 becomes the new c15-math-0079 termini. We now have the topology diagram of riboflavin synthase. (This figure is modified from Hruska et al. [17]).

c15f002

The topology diagram of AML1/CBF (Fig. 15.2b) can be transformed into that of riboflavin synthase (Fig. 15.2f) by the following steps. Remove the loops connecting strand 1 to helix 2, strand 4 to strand 5, and strand 5 to strand 6 (Fig. 15.2c). Connect the c15-math-0081-terminal end of strand 4 to the original c15-math-0082 terminal (Fig. 15.2d). Connect the c15-math-0083-terminal end of strand 5 to the c15-math-0084-terminal end of helix 2 (Fig. 15.2e). Connect the original c15-math-0085 termini to the c15-math-0086-terminal end of strand 5. The c15-math-0087-terminal end of strand 6 becomes the new c15-math-0088 termini, and the c15-math-0089-terminal end of strand 1 becomes the new c15-math-0090 termini (Fig. 15.2f).

15.4 Local Sequence Order–Independent Structural Alignment

Comparison of the global backbone can lead to discovery of distant evolutionary relationships between proteins. However, when attempting to detect similar functions or functional mechanisms between two proteins, global backbone similarity is not a robust indicator [20, 36, 41]. It can be assumed that the physicochemical properties of the local region where function takes place (i.e., substrate binding) is under more evolutionary pressure to be conserved. This assumption has been backed up by several studies [21, 30, 38, 42, 53, 54].

A typical protein contains many concave surface regions, commonly referred to as surface pockets. However, only a few of the surface pockets supply a unique physiochemical environment that is conducive to the protein carrying out its function. The protein must maintain this surface pocket throughout evolution in order to conserve its biological function. For this reason, shared structural similarities between functional surfaces among proteins may be a strong indicator of shared biological function. This has led to a number of promising studies, in which protein functions can be inferred by similarity comparison of local binding surfaces [7, 10, 21, 35, 39]

The inherent flexibility of the protein structure makes the problem of structural comparison of protein surface pockets challenging. A protein is not a static structure as represented by a PDB [9] entry. The whole protein as well as the local functionalsurface may undergo various degrees of structural fluctuations. The use of a single surface pocket structure as a representative template for a specific protein function can lead to many false negatives. This is due to the inability of a single representative to capture the full functional characteristics across all conformations of a protein.

We have addressed this problem by developing an algorithm that can identify the atoms of a surface pocket that are structurally preserved across a family of protein structures that have similar functions. Using a sequence order–independent local surface alignment method to pairwise-align the functional pockets across a family of protein structure, we automatically find the structurally conserved atoms and measure their fluctuations. We call these structurally conserved atoms the signature pocket. More than one signature pocket may result for a single functional class. In this case, our method can automatically create a basis set of signature pockets for that functional family. We can then use these signature pockets as representatives for scanning a structure database for functional inference by structural similarity.

15.4.1 Bipartite Graph Matching Algorithm for Local Surface Comparison

We have modified and implemented a sequence order–independent local structural alignment algorithm based on the maximum-weight bipartite graph matching formulation developed by Chen et al. [12].

As mentioned earlier, the structural alignment problem boils down to a problem of finding an equivalence relation between residues of a reference protein c15-math-0091 and a query protein c15-math-0092 that when applied will optimize the superposition of the two structures. The formulation here does this in an iterative two-step process: (1) an optimal set of equivalent atoms are determined under the current superposition using a bipartite graph representation and (2) the new equivalence relation is used to determine a new optimal superposition. The two steps are then repeated until a stopping condition is met.

The equivalence relationship is found between the two atoms of the functional pocket surfaces by representing the atoms the atoms of c15-math-0093 and c15-math-0094 as nodes in a graph. This graph is bipartite, meaning that edges exist only between atoms of c15-math-0095 and atoms of c15-math-0096. A directed edge is drawn between two nodes if a similarity threshold is met. In our implementation, the measure of similarity takes into account both spatial distances and the chemical property similarities between the two corresponding atoms.

Each edge c15-math-0097 connecting nodes c15-math-0098 and c15-math-0099 is assigned a weight c15-math-0100 equal to the similarity score between the two corresponding atoms (see [12] for details). The optimal equivalence relationship given the current superposition is a subset of the edges within this bipartite graph that have maximum combined weight, where at most one edge can beselected per atom, making this a maximum-weight bipartite graph matching problem. The solution to this problem can be found using the Hungarian algorithm [33].

The Hungarian method is as follows. Initially, an overall score c15-math-0101 is set. Additionally, an artificial source node c15-math-0102 and an artificial destination node c15-math-0103 are added to the bipartite graph. A directed edge c15-math-0104 with zero weight is added for each of the atom nodes c15-math-0105 from c15-math-0106 and similarly, directed edges c15-math-0107 with zero weight are drawn from each of the atoms nodes of c15-math-0108. The algorithm then proceeds as follows:

1. Find the shortest distance c15-math-0109 from the source node c15-math-0110 to every other node c15-math-0111 using the Bellman–Ford algorithm.
2. Assign a new weight c15-math-0112 to each edge that does not originate from the source node c15-math-0113 as follows:

15.9 c15-math-0114

3. Update c15-math-0115 as c15-math-0116.
4. Reverse the direction of the edges along the shortest path from c15-math-0117 to c15-math-0118.
5. If c15-math-0119 and a path exists between c15-math-0120 and c15-math-0121, then repeat step 1.

The iterative process of the Hungarian algorithm stops when either there is no path from c15-math-0122 to c15-math-0123 or the shortest distance from the source node to the destination node c15-math-0124 is greater than the current overall score c15-math-0125. At the end of the process, the graph will consist of a set of directed edges that have been reversed (they now point from nodes of c15-math-0126 to nodes from c15-math-0127. These reversed edges represent the new equivalence relationships between the atoms of c15-math-0128 and the atoms of c15-math-0129.

The equivalence relationship found by the bipartite matching algorithm can now be used to superimpose the two proteins using the singular value decomposition. After superpositioning the new equivalent atoms, a new bipartite graph is createdand the process is iterated until the change in RMSD on superposition falls below a threshold.

15.4.2 A Basis Set of Binding Surface Signature Pockets

The ability to compare structural similarities between to protein surface regions can provide insight into shared biological functions. As mentioned earlier, when dealing with local surface regions, one has to be careful when choosing a functional representative pocket because of the inherent flexibility of the binding surfaces. We have developed a method that automatically generates a set of functional pocket templates, called signature pockets of local surface regions that can be used as a representative a functional surface for structural comparison. These signature pockets contain broad structural information and have discriminating ability.

A signature pocket is derived from sequence order–independent structural alignments of precomputed surface pockets. Our signature pocket method does not require the atoms of the signature pocket to be present in all member structures. Instead, signature pockets can be created at varying degrees of partial structural similarity and can be hierarchically organized according to their structural similarity.

The input of our signature pocket algorithm is a set of functional pockets from the CASTp database [18]. All versus all pairwise sequence order–independent local surface alignment is performed on the input functional surface pockets. A distance is calculated on the basis of the RMSD and the chemistry of the paired atoms of the structural alignment [16]. The resulting distance matrix is used by an agglomerative clustering method. The signature of the functional pocket is then derived using a recursive process following the hierarchical tree.

The recursive process begins by finding the two closest siblings (pockets c15-math-0130 and c15-math-0131), and combining them into a single structure c15-math-0132. During the recursive process, c15-math-0133 or c15-math-0134 may themselves already be a combination of several structures. When combining two structures, we follow these criteria:

1. If two atoms were considered equivalent in a structural alignment, a single coordinate is created in the new structure to represent both atoms. The new coordinate is calculated as the average of the two underlying atom coordinates.
2. If no equivalence was found for an atom during the structural alignment, the coordinates of that atom are transferred directly into the new pocket structure.

A count of the number of times that an atom at the position c15-math-0135 was present in the underlying set of pockets (c15-math-0136) is recorded during each step in the recursive process. A preservation ratio c15-math-0137 is calculated for each atom of the signature pocket by dividing c15-math-0138 by the total number of constituent pockets. In addition, the mean distance of the coordinates of the aligned atoms to their geometric center is recorded as the location variation c15-math-0139. At the end of each step, the new structure c15-math-0140 replaces the two structures c15-math-0141 and c15-math-0142 in the hierarchical tree and the process is repeated on the updated hierarchical tree.

The recursive process can be stopped at any point during its traversal of the hierarchical tree by selecting a c15-math-0143 threshold. Depending on the choice of the c15-math-0144 threshold, a single or multiple signature pockets can be created. Figure 15.3a shows a low c15-math-0145 threshold that results in a set of three signature pockets. As the threshold is raised, fewer signature pockets are created (Fig. 15.3b). A single signature pocket representing all surface pockets in the dataset can be generated by raising the threshold even further (Fig. 15.3). The set of signature pockets from different clusters in the hierarchical tree form a basis set that represents an ensemble of differently sampled conformations of the surface pockets in the PDB. The basis set of signature pockets can be used to accurately classify and predict enzymatic function.

Figure 15.3 Different basis sets of signature pockets can be produced at different levels of structural similarity by raising or lowering the similarity threshold (vertical dashed line): (a) a low threshold will produce more signature pockets; (b) as the threshold is raised, fewer signature pockets will be created; (c) a single signature pocket can, in principle, be created to represent the full surface pocket dataset by raising the threshold.

c15f003

15.4.2.1 Signature Pockets of NAD Binding Proteins

To illustrate how signature pockets and the basis set help identify structural elements that are important for binding and to show their accuracy in functional inference, we discuss a study performed on the nicotinamide adenine dinucleotide (NAD) binding proteins. NAD plays essential roles in metabolisms where it acts as a coenzyme in redox reactions, including glycolysis and the citric acid cycle.

We obtained a set of 457 NAD binding proteins of diverse fold and diverse evolutionary origin. We extracted the NAD binding surfaces from the CASTp database of protein pockets [18]. We obtained the hierarchical tree using the results of our sequence order–independent surface alignments. The resulting nine signature pockets of the NAD binding pocket form a basis set, shown in Figure 15.4.

Figure 15.4 Topology of the hierarchical tree and signature pockets of the NAD binding pockets: (a) the resulting hierarchical tree topology; (b–j) the resulting signature pockets of the NAD binding proteins, along with the superimposed NAD molecules that were bound in the pockets of the member proteins of the respective clusters. The NAD coenzymes have two distinct conformations. Those in an extended conformation are marked with an X, and those in a compact conformation are marked with a C.

c15f004

The signature pockets of NAD contain biological information. The signature pocket show in Figure 15.4j is based on a cluster of NAD binding proteins that act on the aldehyde group of donors, the signature pockets in (Fig. 15.4f,g) are for oxioreductases that act on the CH–CH group of donors, and the signature pockets of Figure 15.4e, h, and i are for clusters of alcohol oxioreductases that act on the CH–OH group of donors. The NAD–binding lyase family is represented in two signature pockets. The first represents lyases that cleave both C–O and P–O (Fig. 15.4d) and the second, containing lyases that cleave both C–O and CC bonds (Fig. 15.4b). These two signature pockets from two clusters of lyase conformations have a different class of conformations of the bound NAD cofactor (extended and compact).

In addition to the structural fold, the signature pockets are also determined by the conformation of the bound NAD cofactor (Fig. 15.4a). It can be seen in Fig. 15.4b–j that there are two general conformations of the NAD coenzyme. The coenzymes labeled C (Fig. 15.4b,c,f,g,h,j) have a closed conformation, while the conenzymes labeled X (Fig. 15.4d,e,i) have an extended conformation. This indicates that the binding pocket may take multiple conformations yet bind the same substrate in the same general structure. For example, the two structurally distinct signature pockets shown in Fig. 15.4f,g are derived from proteins that have the same biological function and SCOP fold. All of these proteins bind to the same NAD conformation.

We further evaluated the effectiveness of the NAD basis set by determining its accuracy at correctly classifying enzymes as either NAD binding or non-NAD-binding. We constructed a testing dataset of 576 surface pockets from the CASTp database [18]. This dataset is independent of the 457 NAD binding proteins that we used to create the signature pockets. We collected the 576 surface pockets by selecting the top three largest pockets by volume from 142 randomly chosen proteins and 50 proteins that have NAD bound in the PDB structure. We then structurally aligned each signature pocket against each of the 576 testing pockets. The testing pocket was assigned to be an NAD binding pocket if it structurally aligned to one of the nine NAD signature pockets with a distance under a predefined threshold. Otherwise it was classified as non-NAD-binding. The results show that the basis set of nine signature pockets can classify the correct NAD binding pocket with sensitivity and specificity of 0.91 and 0.89, respectively. We performed further testing to determine whether a single representative NAD binding pocket, as opposed to a basis set, issufficient for identifying NAD binding enzymes. We chose a single pocket representative from one of the nine clusters at random and attempted to classify our testing dataset by structural alignment. We used the same predefined threshold used in the basis set study. This was repeated 9 times using a representative from each of the nine clusters. We found that the results deteriorated significantly with an average sensitivity and specificity of 0.36 and 0.23, respectively. This strongly indicates that the construction of a basis set of signature pockets to be used as a structural template provides significant improvement for functional inference of a set of evolutionarily diverse proteins.

15.5 Conclusion

We have discussed methods that provide solutions to the problems that arise during functional inference by structural similarity at both the global and at local surface levels. Both of our methods disregard the ordering of residues in the protein's primary sequence, making them sequence order–independent. The global method can be used to address the challenging problem of detecting structural similarities even after topological rearrangements of the proteins backbone. The fragment assembly approach based on the formulation of a relaxed integer programming problem and an algorithm based on scheduling split-interval graphs is guaranteed by an approximation ratio. We showed that this method is capable of discovering circularly permuted proteins and other more complex topological rearrangements.

We also described a method for sequence order–independent alignment of local surfaces on proteins. This method is based on a bipartite graph matching problem. We further show that the surface alignments can be used to automatically construct a basis set of signature pockets representing structurally preserved atoms across a family of proteins with similar biological functions.

Acknowledgments

DasGupta and Liang was partially supported by NSF grant DBI-1062328.

References

1. Aghili AS, Agrawal D, El Abbadi A, PADS: Protein structure alignment using directional shape signatures, Proc. Database Systems for Advanced Applications (DASFAA) Conf., 2004.

2. Alexandrov NN, Fischer D, Analysis of topological and nontopological structural similarities in the PDB: New examples from old structures, Proteins 25:354–365 (1996).

3. 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. 25:3389–3402 (1997).

4. Altschul SF, Warren G, Miller W, Myers EW, Lipman DJ, Basic local alignment search tool, J. Mol. Biolo., 215:403–410 (1990).

5. Andrade MA, Brown NP, Leroy C, Hoersch S, de Daruvar A, Reich C, Franchini J, Tamames A, Valencia A, Ouzounis C, Sander C, Automated genome sequence analysis and annotation, Bioinformatics, 15:391–412 (1999).

6. Rost B, Twilight zone of protein sequence alignments, Protein Eng. 12:85–94 (1999).

7. Bandyopadhyay D, Huan J, Liu J, Prins J, Snoeyink J, Wang W, Tropsha A, Functional neighbors: Inferring relationships between nonhomologous protein families using family-specific packing motifs, IEEE Trans. Inf. Technol. Biomed. 14(5):1137–1143 (2010).

8. Bar-Yehuda R, Halldorsson MM, Naor J, Shacknai H, Shapira I, Scheduling split intervals, Proc. 14th ACM-SIAM Symp. Discrete Algorithms, 2002, pp. 732–741.

9. Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE, The protein data bank, Nucleic Acids Res. 28:235–242 (2000).

10. Binkowski TA, Joachimiak A, Protein functional surfaces: Global shape matching and local spatial alignments of ligand binding sites, BMC Struct. Biol. 8:45 (2008).

11. Boeckmann B, Bairoch A, Apweiler R, Blatter MC, Estreicher A, Gasteiger E, Martin MJ, Michoud K, O'Donovan C, Phan I, Pilbout S, Schneider M, The SWISS-PROT protein knowledgebase and its supplement TrEMBL in 2003, Nucleic Acids Res. 31:365–370 (2003).

12. Chen L, Wu LY, Wang R, Wang Y, Zhang S, Zhang XS, Comparison of protein structures by multi-objective optimization, Genome Inform. 16(2):114–124 (2005).

13. Dayhoff MO, Schwartz RM, Orcutt BC, A model of evolutionary change in proteins, Atlas Protein Seq. Struct. 5(3):345–352 (1978).

14. Deng M, Zhang K, Mehta S, Chen T, Sun F, Prediction of protein function using protein-protein interaction data, J. Comput. Biol. 10(6):947–960 (2009).

15. Dror O, Benyamini H, Nussinov R, Wolfson HJ, MASS: Multiple structural alignment by secondary structures, Bioinformatics 19:i95–i104 (2003).

16. Dundas J, Adamian L, Liang J, Structural Signatures of Enzyme Binding Pockets from Order-Independent Surface Alignment: A Study of Metalloendopeptidase and NAD Binding Proteins, J. Mol. Biol. 406(5):713–729 (2011).

17. Dundas J, Binkowski TA, DasGupta B, Liang J, Topology independent protein structural alignment, BMC Bioinformatics 8:388 (2008).

18. Dundas J, Ouyang Z, Tseng J, Binkowski TA, Turpaz Y, Liang J, CASTp: Computed atlas of surface topography of proteins with structural and topographical mapping of functionally annotated residues, Nucleic Acids Res., 34:W116–W118 (2006).

19. Dutta S, Akey IV, Dingwall C, Hartman KL, Laue T, Nolte RT, Head JF, Akey CW, The crystal structure of nucleoplasmin-core implications for histone binding and neucleosome assemply, Mol. Cell 8:841–853 (2001).

20. Fischer D, Norel R, Wolfson H, Nussinov R, Surface motifs by a computer vision technique: Searches, detection, and implications for protein-ligand recognition, Proteins 16:278–292 (1993).

21. Binkowski TA, Adamian L, Liang J, Inferring functional relationship of proteins from local sequence and spatial surface patterns, J. Mol. Biol. 332:505–526 (2003).

22. Gold ND, Jackson RM, Fold independent structural comparisons of protein-ligand binding sites for exploring functional relationships, J. Mol. Biol. 355:1112–1124 (2006).

23. Hasegawa H, Holm L, Advances and pitfalls of protein structural alignment, Curr. Opin. Struct. Biol. 19:341–348 (2009).

24. Hegyi H, Gerstein M, The relationship between protein structure and function: A comprehensive survey with application to the yeast genome, J. Mol. Biol., 288:147–164 (1999).

25. Henikoff S, Henikoff JG, Amino acid substitution matrices from protein blocks, Proc. Natl. Acad. Sci. USA 89(22):10 915–10 919 (1992).

26. Hobohm U, Sander C, Enlarged representative set of protein structures, Protein Sci. 3:522 (1994).

27. Holm L, Sander C, Protein structure comparison by alignment of distance matrices, J. Mol. Biol. 233:123–138 (1993).

28. Hulo N, Sigrist CJA, Le Saux V, Recent improvements to the PROSITE database, Nucleic Acids Res. 32:D134–D137 (2004).

29. Ilyin VA, Abyzov A, Leslin CM, Structural alignment of proteins by a novel TOPOFIT method, as a superimposition of common volumes at a topomax point, Protein Sci. 13:1865–1874 (2004).

30. Jeffery C, Molecular mechanisms for multi-tasking; recent crystal structures of moon-lighting proteins, Curr. Opin. Struct. Biol. 14:663–668 (2004).

31. Jensen LJ, Gupta R, Blom N, Devos D, Tamames J, Kesmir C, Nielsen H, Staerfeldt HH, Rapacki K, Workman C, Andersen CAF, Knudsen S, Krogh A, Valencia A, Brunak S, Prediction of human protein function from post-translational modifications and localization features, J. Mol. Biol. 319:1257–1265 (2002).

32. Karplus K, Barret C, Hughey R, Hidden Markov models for detecting remote protein homologues, Bioinformatics 14:846–856 (1998).

33. Kuh HW, The Hungarian method for the assignment problem, Nav. Res. Logist. Q. 2:83–97 (1995).

34. Laskowski RA, Watson JD, Thornton JM, ProFunc: A server for predicting protein function from 3D structure, Nucleic Acids Res. 33:W89–W93 (2005).

35. Lee S, Li B, La D, Fang Y, Ramani K, Rustamov R, Kihara D, Fast protein tertiary structure retrieval based on global surface shape similarity, Proteins 72:1259–1273 (2008).

36. Lichtarge O, Bourne HR, Cohen FE, An evolutionary trace method defines binding surfaces common to protein families, J. Mol. Biol. 7:39–46 (1994).

37. Lindqvist Y, Schneider G, Circular permutations of natural protein sequences: Structural evidence, Curr. Opin. Struct. Biol. 7:422–477 (1997).

38. Meng E, Polacco B, Babbitt P, Superfamily active site templates, Proteins, 55:962–967 (2004).

39. Moll M, Kavraki L, LabelHash, A Flexible and Extensible Method for Matching Structural Motifs, available from Nature Precedings (http://dx.doi.org/10.1038/npre.2008.2199.1) (2008).

40. 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. 247:536–540 (1995).

41. Norel R, Fischer H, Wolfson H, Nussinov R, Molecular surface recognition by computer vision-based technique, Protein Eng. 7(1):39–46 (1994).

42. Orengo C, Todd A, Thornton J, From protein structure to function, Curr. Opin. Struct. Biol. 9:374–382 (1999).

43. Orengo CA, Michie AD, Jones DT, Swindells MB, Thornton JM, CATH: A hierarchical classification of protein domains structures, Structure 5:1093–1108 (1997).

44. Pal D, Eisenberg D, Inference of protein function from protein structure, Structure 13:121–130 (2005).

45. Shah I, Hunterm L, Predicting enzyme function from sequence: A systematic appraisal, Intell. Syst. Mol. Biol. 5:276–283 (1997).

46. Shih ES, Hwang MJ, Alternative alignments from comparison of protein structures, Proteins 56:519–527 (2004).

47. Shindyalov IN, Bourne PE, Protein structure alignment by incremental combinatorial extension (CE) of the optimal path, Protein Eng. 11(9):739–747 (1998).

48. Standley DM, Toh H, Nakamura H, Detecting local structural similarity in proteins by maximizing the number of equivalent residues, Proteins: Struct., Funct., Genet. 57:381–391 (2004).

49. Szustakowski JD, Weng Z, Protein structure alignment using a genetic algorithm, Proteins: Struct., Func., Genet. 38:428–440 (2000).

50. Tabtiang RK, Cezairliyan BO, Grant RA, Chochrane JC, Sauer RT, Consolidating critical binding determinants by noncyclic rearrangement of protein secondary structure, Proc. Natl. Acad. Sci. USA 7:2305–2309 (2004).

51. Teichert F, Bastolla U, Porto M, SABERTOOTH: Protein structure comparison basedon vectorial structure representation, BMC Bioinformatics 8:425 (2007).

52. Teyra J, Paszkowski-Rogacz M, Anders G, Pisabarro MT, SCOWLP classification: structural comparison and analysis of protein bindign regions, BMC Bioinformatics 9:9 (2008).

53. Tseng YY, Dundas J, Liang J, Predicting protein function and binding profile via matching of local evolutionary and geometric surface patterns, J. Mol. Biol. 387(2):451–464 (2009).

54. Tseng YY, Liang J, Estimation of amino acid residue substitution rates at local spatial regions and application in protein function inference: A Bayesian Monte Carlo approach, Mol. Biol. Evol. 23:421–436 (2006).

55. Uliel S, Fliess A, Amir A, Unger R, A simple algorithm for detecting circular permutations in proteins, Bioinformatics 15(11):930–936 (1999).

56. Veeramalai M, Gilbert D, A novel method for comparing topological models of protein structures enhanced with ligand information, Bioinformatics 24(23):2698–2705 (2008).

57. Weidong T, Skolnick J, How well is enzyme function conserved as a function of pairwise sequence identity, J. Mol. Biol. 333:863–882 (2003).

58. Woo EJ, Marshall J, Bauly J, Chen JG, Venis M, Napier RM, Pickersgill RW, Crystal structure of the auxin-binding protein 1 in complex with auxin, EMBO J. 21:2877–2885 (2001).

59. Zhu J, Weng Z, A novel protein structure alignment algorithm, Proteins: Struct., Funct., Bioinform. 58:618–627 (2005).

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.147.85.175