Skip to main content

Proteinortho: Detection of (Co-)orthologs in large-scale analysis

Abstract

Background

Orthology analysis is an important part of data analysis in many areas of bioinformatics such as comparative genomics and molecular phylogenetics. The ever-increasing flood of sequence data, and hence the rapidly increasing number of genomes that can be compared simultaneously, calls for efficient software tools as brute-force approaches with quadratic memory requirements become infeasible in practise. The rapid pace at which new data become available, furthermore, makes it desirable to compute genome-wide orthology relations for a given dataset rather than relying on relations listed in databases.

Results

The program Proteinortho described here is a stand-alone tool that is geared towards large datasets and makes use of distributed computing techniques when run on multi-core hardware. It implements an extended version of the reciprocal best alignment heuristic. We apply Proteinortho to compute orthologous proteins in the complete set of all 717 eubacterial genomes available at NCBI at the beginning of 2009. We identified thirty proteins present in 99% of all bacterial proteomes.

Conclusions

Proteinortho significantly reduces the required amount of memory for orthology analysis compared to existing tools, allowing such computations to be performed on off-the-shelf hardware.

Background

Genome annotation largely depends on the determination of sequence intervals that are homologous, and if possible, orthologous to sequences of known identity and function in related genomes. Orthologous genes (orthologs) are derived from a common ancestor by a speciation event [1]. Orthologs are of particular interest because they can be expected to have maintained at least part of their (ancestral) biological function. For protein-coding genes, several well-known databases, including InParanoid[2], OrthoMCL-DB[3], COG-database[4], Homogene[5], eggNOG[6], OMA Browser[7] and Ensembl Compara[8] compile such information. Their content is restricted to data previously published in comprehensive databases of protein sequences such as UniProt[9]. Updates with additional proteomic data thus are published relatively infrequently. Modern high-throughput technologies, however, produce huge amounts of protein data and even larger amounts of transcript data that are computationally translated to putative polypeptide sequences. Oftentimes, therefore, it would be desirable to generate the orthology relation for a particular dataset, so that the availability of orthology data does not limit the set of species or genes that can be included.

The computation of genome-wide orthology data, however, is a challenging and time consuming task with the currently available tools. In many cases, orthologs cannot be identified unambiguously by means of sequence comparison. The main difficulty arises from the presence of paralogs (homologous genes within the same genome) which can make it very difficult to recognize the correct ortholog among the other homologs. Gene duplications following the speciation, furthermore, create two or more genes in one lineage that are, collectively, orthologous to one or more genes in another lineage. Such genes are known as co-orthologs [10].

The most widely used approach to identify (putative) orthologs between two species is the reciprocal best alignment heuristic[1115]. This approach was more recently extended e.g. in OrthoMCL[16] and MultiParanoid[17] to detect (co-)orthologs within multiple species. All these tools, however, are limited to relatively small sets of species. In practise, analyzing the complete proteomes of more than about 50 prokaryote species goes beyond the capabilities of standard hardware and requires access to supercomputer resources. This limitation arises from both, technical issues such as insufficient parallelization and the algorithmic design that requires all reasonable alignments for each input protein to be held in the memory for efficient access in the clustering stage. Proteinortho is specifically designed to deal with hundreds of species together containing millions of proteins. It achieves this performance both by optimizing the implementation and by modifying the reciprocal best alignment method in a way that allows alignment processing on the fly.

Results and Discussion

Orthology prediction

Theory

As for other approaches to large-scale orthology detection, the starting point is a complete collection of pairwise comparisons, typically performed using blast. For simplicity of presentation, we assume that the individual sequences that are compared represent proteins, although algorithms and pipelines are applicable also to other sequence data such as non-coding RNA genes or conserved DNA regions. Typically, the results of pairwise comparisons are ranked by similarity, for instance based on blast statistics, evolutionary distances, or genome rearrangement analysis [7, 16, 18]. High-ranking alignments across multiple species then have to be combined in order to determine orthologous groups. However, these groups usually do not readily provide detailed insights since they can contain large numbers of related genes for each species. Hence, meaningful units have to be identified. For this purpose, a variety of clustering algorithms has been applied to determine Clusters of Orthologous Groups (COGs). The MCL-algorithm for instance uses a stochastic flow simulation to determine meaningful COGs [16, 19, 20]. In addition, MultiParanoid explicitly searches and tags in-paralogs, i.e., recent paralogs that represent species-specific gene expansions. This strategy requires to directly compare proteins within each species. Alternatively, data were curated by manual postprocessing [4].

We will argue here that orthology determination can be understood as the problem of finding nearly disjoint maximal nearly-complete multipartite subgraphs in an edge-weighted directed graph whose vertices are the proteins in the input set, and whose edges connect certain pairs of similar proteins of different species. The edge weights ωxyencode the similarity of x and y. In our implementation, the bit score of the blast alignment (xy) will serve as edge weight. An E-value cut-off is used beyond which blast alignments are not included into .

In order to motivate our point of view, we first consider an idealized dataset (see Figure 1) in which (1) each protein x (from species A) has at most one ortholog in any other species BA, (2) if y B is an ortholog of x A, then a blast search of x against B yields at least one alignment and (3) the unique best alignment of query x against B is the true ortholog y of x. In this case, the well-know "reciprocal best alignment heuristic" (RBAH), also known as "reciprocal best hits" (RBH) [11], can be used to retrieve the correct ortholog set. To see this, we construct a subgraph of as follows: For each protein x in species A and a given species BA we retain only the arc with maximal weight:

Figure 1
figure 1

Orthology relations. Idealized dataset for two species A and B. Proteins x A and y B are depicted by open boxes. Orthology relations between proteins x and y are represented by grey shadows. Arrows indicate alignments above a certain cut-off from the search of x against B. Solid lines refer to the best alignments. Cases (1), (2), and (3) cannot occur by definition in an idealized dataset, but of course do appear in real life applications.

(1)

The symmetric subgraph of , containing only reciprocal best alignments, can be regarded as an undirected graph ϒ RBAH . By construction, any two vertices are connected by edges in ϒ RBAH if and only if they are orthologs. A set of orthologs therefore corresponds to a complete multipartite subgraph of in which every species is represented at most once. Furthermore, we note that these subgraphs are disjoint, i.e., ortholog sets correspond to the connected components of .

When applied to real data, however, RBAH usually gives rise to several artifacts. In general, therefore, it does not produce correct and complete sets of orthologs. First, gene duplications produce co-orthologs, destroying the uniqueness of best blast alignments. The blast comparison of two species both containing two co-orthologs will in general produce slightly different scores among those genes, so that in general will not contain all arcs between them, see Figure 2a. RBAH now extracts the symmetric part of , i.e., it removes all non-reciprocal edges. Thus the undirected graph ϒ RBAH has an edge {x, y} if and only if both (x, y) and (y, x) are arcs in . A missing arc in the example of Figure 2a thus translates into a missing edge in ϒ RBAH . Ortholog sets, therefore, are still located in disjoint connected components, but they are only approximately multipartite and there is no guarantee that they remain connected. As a remedy, it could be proposed to use the k-best blast alignments instead of a single best one. This leads, however, to another type of artifact: If distant homologs or spurious blast alignments with scores above the blast threshold are present, see Figure 2b, the inclusion of up to k alignments can result in edges between paralogous but non-orthologous proteins.

Figure 2
figure 2

Adaptive RBAH. Reciprocal best alignment heuristic (RBAH). (a) If there is a pair of divergent co-orthologs x1, , and y1 and , resp., it is possible that there are no reciprocal best blast alignments. In this situation, RBAH will not identify any orthologs. (b) One possible remedy is to include the second best blast alignment (n = 2). However, in this case highly similar orthologs (x2 and y2 as well as x3 and y3), which in principle can clearly be divided, can get combined. (c) Proteinortho uses an adaptive approach that is (1) flexible with respect to the number of more diverged orthologs in absence of a reciprocal best blast alignment and (2) will not intermix orthologous groups that can be disentangled easily because of large differences in pairwise similarity.

In order to reduce such problems, we suggest to introduce a similarity cut-off value that itself depends on the quality of the matches, i.e., to consider all those blast alignments of each query against a given species whose E-values are only slightly smaller (by a factor f < 1) than the queries' best blast alignments. More precisely,

(2)

This has two advantages: it retains edges to likely co-orthologs while at the same time reducing the number of edges that are inserted in .

The symmetric part ϒ* of now retains more edges than ϒ RBAH . In particular, it includes all the edges connecting similar co-orthologs. On the other hand, the threshold at fairly high bit scores disconnects at least most of the more distant homologs. Sets of (co-)orthologs thus appear in ϒ* as nearly complete multipartite subgraphs. Typically they will contain more than one node from the same species, among them, in particular, all in-paralogs. Although this approach strongly reduces the problem with spurious edges, ϒ* may also contain additional edges connecting two or more sets of (co-)orthologs.

The problem of finding maximal complete multipartite subgraphs of a graph is NP complete [21]. Furthermore, we have seen above that ϒ* may lack a few edges which should connect orthologs ("false negatives"), while at the same time there are also some additional "false positive" edges. In a two-species comparison there is no information that could compensate missing and spurious edges, while in the multi-species case, the graph ϒ* is in a sense "self-correcting" since we can formulate orthology detection as an optimization problem. More precisely, we search for a decomposition of ϒ* into a disjoint collection of complete multipartite subgraphs so that the number of edge insertions and deletions is minimized.

Since no efficient approaches to this combinatorial optimization problem seem to be known, it appears fruitful to resort to a heuristic approach that employs a somewhat different point of view: nearly complete multipartite subgraphs are very dense subgraphs, which in our case either form connected components on their own, or which are connected to other dense clusters by a few additional edges. The problem thus is to determine for each connected component ϒ* of whether it is sufficiently densely connected, and if not, to partition it into its densely connected components by removing the spurious edges connecting them. Here, we approach this issue by means of spectral partitioning [22], see Additional File 1 for a detailed description.

We remark, finally, that one could efficiently add the explicit determination of in-paralogs after ϒ* has been constructed, although currently this is not implemented in Proteinortho. Following e.g. InParanoid, sets of in-paralogs are subsets of proteins from the same species within the same connected component ϒ* of that are more similar to each other than to any protein in another species. It is sufficient, thus, to determine alignment scores for pairs of nodes from the same species within connected components of ϒ*. As in MultiParanoid, in-paralogs could be collapsed to a single node.

Implementation

Blast searches

Proteinortho expects fasta files containing either nucleic acid or amino acid sequences as input. Proteinortho does not build a large database containing all proteins but rather keeps the protein complements of different species separate. This has multiple advantages: (1) the blast step can be partitioned into multiple runs. The available processor cores are used efficiently as each blast process can utilize one processor core to the full, which is usually not the case if threading is handled by blast itself. Furthermore, the blast jobs can be distributed to several computers in a network. (2) The computationally expensive blast search of species against themselves (which is implicit in searching a comprehensive protein database) can be avoided. (3) The E-values returned by blast depend only on the proteome of each input species, not on the database of all input sequences. Therefore, Proteinortho implicitly enforces higher similarity for related proteomes but allows lower levels of similarity for larger evolutionary distances. Conceptually, this replaces the normalization of the scores employed e.g. by OrthoMCL. (4) The scope of the analysis can be extended without the need to re-compute blast comparisons that have already been computed earlier. The E-values remain unchanged when species are included or excluded from the analysis. Proteinortho can use several PCs with a shared storage (such as an NFS file system). The implementation of distributed computing is illustrated in Additional File 2. Furthermore, cluster infrastructure (e.g. MPI or SGE) is supported. A benchmark illustrating the performance improvements is shown in Figure 3a. Proteinortho is faster than OrthoMCL even if only a single processor core is available and gains much of its practical advantage from parallelization. In applications to large numbers of species, for which Proteinortho is primarily intended, handling blast alignments dominates the spectral partitioning step by several orders of magnitude. As blast and the memory consumption of holding the graph structure for clustering are limiting, we do not investigate the complexity of the clustering step itself in detail.

Figure 3
figure 3

Benchmarks. CPU time and memory requirements of Proteinortho. (a) The speed benchmark was performed using an E. coli strain with 4132 proteins on an eight core Intel Xeon system using one thread (1) and eight threads (8) at 2.33 GHz. The encoded proteins were used multiple times to simulate multiple (identical) species. This is the worst case scenario for Proteinortho since in this case every protein has a link to at least one protein in every other species. Proteinortho is significantly faster than OrthoMCL. Using multiple threads we observe a substantial speed up. (b) The memory benchmark is performed using the same set as in (a). OrthoMCL quickly exhausts memory for larger sets. Proteinortho clearly performs more efficient, even though this artificial scenario is a more complex case than real world analysis. Both benchmarks outline that Proteinortho allows comprehensive studies which were not possible before.

Most importantly, the algorithm outlined in the previous section avoids the memory bottleneck that limits previous approaches. Suppose our input set comprises N species with, on average, m genes. The size of the input is thus n = N × m proteins. Instead of storing all n × n pairwise blast scores, Proteinortho processes the comparisons between any two species A and B immediately: first the blast alignments are filtered by two additional criteria: (1) The alignment must exhibit a minimum level of sequence identity. (2) The alignment must cover at least a minimum fraction of the query protein. This second rule ensures that fusion genes such as rice OsUK/UPRT1 [23] are eventually assigned as homologs of the dominating part of the protein. Then equ.(2) is evaluated for all x in A, so that Proteinortho directly constructs the sparse graph , while does not need to be stored at all. Proteinortho therefore uses chained arrays, requiring only n × k entries, where k is the average number of nearly optimal blast alignments per gene, and k = a × N, where a is the average number of (co-)orthologs of a gene in a single species. The value of a is independent of the size of the dataset. Empirically, we found a ≤ 1 in all datasets investigated so far. Thus Proteinortho saves a factor n2/N2ma = m/am of memory. Note that prokaryotes have m ≈ 103 ... 104 proteins.

Spectral partitioning

First we reduce the problem by determining the connected components ϒ* of since these can be treated separately. We use the well-known breath-first search approach [24] to this end. In order to check whether a connected component Ξ is sufficiently dense to represent a single set of co-orthologs we compute its normalized algebraic connectivity . Here n is the number of vertices of Ξ and αalpha2 is 2nd-smallest eigenvalue of the graph Laplacian L = D - A of Ξ [25]. Here A is the adjacency matrix of Ξ and D is the diagonal matrix of the vertex degrees. The eigenvalue αalpha2 can be computed iteratively, see Additional File 1. Values of indicate dense clusters that most likely correspond to coherent sets of (co-)orthologs. Small values , on the other hand indicate that Ξ has a low connectivity and either consists of two or more dense components or it has (nearly) tree-like protrusions. Very large components can arise when genes duplicate frequently and diverge quickly according to the duplication-degeneration-complementation (DDC) model [26].

The "Fiedler vector" x2, i.e., the eigenvector of L to eigenvalue α2 can be used to find a partition of Ξ into two connected components, one consisting of the vertices for which x2 has positive entries and one for which x2 has negative entries [27]. This decomposition is iterated until Ξ is partitioned into components with algebraic connectivity above a certain threshold value and tree-like pieces, which most likely correspond to false-positive edges of ϒ*. In order to speed up the computation, trees are therefore removed from the component Ξ before the algebraic connectivity and the Fiedler vector is computed. This is achieved by iteratively removing a vertex of degree 1 and its adjacent edge. This step is not performed if Proteinortho is used to compare only two species.

We remark that the memory and CPU consumption for the clustering step of OrthoMCL can be drastically reduced by using a novel algorithm [28], reaching a performance that is theoretically comparable to spectral partitioning as used by Proteinortho (see Additional File 1). Both require only the storage of edge or adjacency lists. The current implementation of spectral partitioning could be further optimized e.g. by employing the Lanczos algorithm [29] for computing the eigenvalues. Spectral partitioning on average scales as O(n2k). This leads to an expected runtime of O(N3) for Proteinortho which is comparable to the O(N3 log N) complexity bound achieved for COG clustering in [28].

Evaluation of Proteinortho

We compared Proteinortho with the COG-database [4] and OrthoMCL[3]. The latter is the main competitor in terms of speed and memory. The COG-database provides a manually curated dataset that can be regarded as more reliable than fully automated approaches. For benchmark analysis, a set of 16 randomly chosen bacteria from three different classes (six Gram-positive bacilli, six gamma- and four alpha-proteobacteria) are used. The input set comprises 53, 623 protein sequences.

Figure 4 summarizes the size distribution of co-ortholog sets for the three approaches. Using default settings (E-value cutoff 10-10, algebraic connectivity threshold , minimum 25% identity, 50% coverage of query sequence, similarity threshold f = 0.95) Both Proteinortho and OrthoMCL report fewer groups than listed in the COG-database. This is not surprising, however, since the COG-database relies on less restrictive criteria and thus tends to include multiple co-orthologs, while both stand-alone tools apply clustering algorithms that attempt to split the large components whenever possible.

Figure 4
figure 4

Coverage completion. Comparison of the results of Proteinortho with different thresholds of the normalized algebraic connectivity with the COG-database and OrthoMCL for a dataset consisting of 16 randomly chosen bacterial proteomes. The vertical dashed line marks the transition from clusters containing mainly a single ortholog from each species to sets including co-orthologs. The COG-database reports many large groups which often include co-orthologous proteins. OrthoMCL and Proteinortho focus on highly connected subsets in order to find orthologous sets and thus split those groups. Thereby, Proteinortho's clustering algorithm becomes more stringent with increasing values of in splitting in particular large groups. While these groups are left intact for , thresholds of 0.5 and higher drastically reduce the fraction of included co-orthologs.

Figure 5 presents the outcome in more detail. Proteinortho and OrthoMCL report comparable results. Proteinortho is more stringent. The amount of completely new groups which have to be regarded as false positives is considerably lower. OrthoMCL reports slightly more groups. Even though both tools were applied with the same E-value threshold the results vary due to the different blast strategies: OrthoMCL employs one large database containing all proteins, while Proteinortho uses species-specific databases. Most pairwise E-values derived by Proteinortho thus tend to be somewhat larger than those derived by OrthoMCL for the same set of proteins. In turn, OrthoMCL reports some more and larger groups. However, the overall outcome can be regarded as similar while the runtime and memory requirements of Proteinortho are substantially reduced in comparison to OrthoMCL.

Figure 5
figure 5

Comparison of results. Comparison of OrthoMCL and Proteinortho to the COG-database. The following assignments were defined: identity: the group equals a COG-group; subset: the group is subset of a COG-group, at least two proteins are equal; superset: the group is a superset of a COG-group, at least two proteins are equal; new: none of the above-noted criteria matched. Both tools reveal comparable results with respect to the manually curated COG-database. OrthoMCL covers more identical and differently composed groups while Proteinortho is more restrictive and reports fewer new groups which are not present in the COG-database. All groups with less than six species were omitted from the OrthoMCL and Proteinortho data. See Additional File 2 for comparisons with different minimal coverage.

Domain-wide commons

In order to demonstrate that Proteinortho is suitable for large-scale analysis we asked Which proteins can be found in all bacterial species? Proteins that are conserved domain-wide are likely to be useful for the construction of a phylogeny of eubacteria as an alternative to the prevalent usage of 16S rRNA sequences [30]. They can also serve as protein-based markers for identifying novel bacterial species as members of an established phylogenetic group. In addition, they can give insight into basic protein equipment of bacterial life. Hence, we applied Proteinortho to the set of all eubacterial proteomes available at NCBI at the beginning of 2009 (Additional File 3).

The input dataset comprises 2, 155, 620 proteins annotated in 717 bacterial genomes. The Proteinortho run took less than two weeks using 50 processor cores (Intel Xeon at 2.00-2.33 GHz) distributed over multiple PCs. Only 2 GB memory were required. OrthoMCL could not be employed for this task on the hardware available in our lab. Extrapolating from the benchmarks in Figure 3, we estimate that hundreds of gigabytes of memory and years of runtime would have been required.

Proteinortho identified 152 proteins as core of the bacterial protein complement, occurring in at least 90% of all 717 free-living and endosymbiotic bacteria. Of these, 32 are ribosomal subunits. The 30 apparently most indispensable proteins, occurring in 99% of all bacteria, are:

Elongation factor Tu (often co-orthologous to elongation factor 1-alpha)

Elongation factor G

Translation initiation factor IF-2

RNA polymerase subunits β and β'

ATP-dependent metalloprotease FtsH

O-sialoglycoprotein endopeptidase

Methionine aminopeptidase

F0F1 ATP synthase subunits α and β

Dimethyladenosine transferase

  1. 1

    ribosomal protein of the 30S rRNA subunit

  2. 3

    ribosomal proteins of the 50S rRNA subunit

  3. 3

    GTP-binding proteins

  4. 12

    tRNA synthetases

A more detailed list, including unique identifiers for groups of proteins can be found in the supplemental material. The sensitivity of Proteinortho in this survey is limited by incomplete annotation in many species, which we complemented here by tblastn searches with fairly restrictive cut-off values, see Methods for details. Figure 6 summarizes the results quantitatively, emphasizing the effect of incomplete annotation. We remark that this example also shows that Proteinortho can be used to complement existing annotations in an automatic fashion.

Figure 6
figure 6

Orthologous groups. Number of orthologous groups present in nearly all bacterial species. The dashed line represents Proteinortho results based on the NCBI annotation. Using tblastn the annotation was complemented with high scoring genomic matches (solid curve). Note that this is a cumulative plot, i.e., each group of co-orthologs present in x species is also included in the count of groups contained in x' <x species.

Nevertheless, about one third of the 30 most conserved proteins could not be recovered in the genomes of the two species with the smallest proteomes in our dataset: Candidatus Carsonella ruddii PV and Candidatus Sulcia muelleri GWSS. Both are endosymbionts that are considered as organelle-like [31, 32]. Numerous genes that are otherwise considered to be essential for life have been reported as missing in both species. A more detailed and larger list of domain-wide common proteins can be downloaded at http://bioinf.pharmazie.uni-marburg.de/supplements/proteinortho/.

Conclusions

Proteinortho implements a blast-based approach to determine sets of (co-)orthologous proteins or nucleic acid sequences that generalizes the reciprocal best alignment heuristic. The software is optimized for large datasets, and in particular provides a drastic reduction of the memory requirements compared to earlier tools. It can therefore be run on off-the-shelf PC hardware for large datasets. Our implementation scales very well with the number of available processor cores. The blast searches can be trivially parallelized and distributed easily to multiple PCs without the need for a cluster management system, while deployment to existing cluster infrastructure is also supported.

Proteinortho views orthology detection as a variant of graph clustering since co-orthologous sets correspond to maximal complete multipartite subgraphs, which at the same time are well separated from each other. Due to the unavoidable noise in the real data, however, co-orthologous sets appear as dense subgraphs without clearly recognizable low-weight cuts. This property is measured quite well by the algebraic connectivity. At the same time, low-weight cuts between dense regions are identified very well by the corresponding Fiedler vector. We therefore employ spectral partitioning instead of a direct graph clustering approach. The quality of the co-orthologous sets proposed by Proteinortho is comparable to the performance of OrthoMCL.

Both time and memory requirements are significantly reduced compared to earlier approaches, enabling applications that were infeasible before. For instance, we applied Proteinortho to the complete set of 2.1 million proteins from the 717 bacterial genomes available at NCBI at the beginning of 2009. We found 30 proteins that are present in more than 99% of the investigated sequences.

Methods

All analysis with Proteinortho and OrthoMCL were applied using default values unless described otherwise. These are E - value < 10-10, 25% percent identity, Markov Inflation Index of 1.5 for OrthoMCL and E-value < 10-10, 25% percent identity, adaptive best alignments similarity of f = 0.95, algebraic connectivity > 0.1 for Proteinortho. OrthoMCL version 1.4 was downloaded from http://OrthoMCL.org/common/downloads/.

Speed and memory benchmark were performed multiple times using the proteome of Escherichia coli K12 substr. MG1655 data from the NCBI. The protein ids were renamed systematically to prevent duplicated ids for benchmarking purposes which cannot be handled by Proteinortho. A script continuously observed the memory consumption and reported the maximum peak for each run, Figure 3.

For the domain-wide commons we applied Proteinortho with default values of the parameters. Bacterial proteomes and genomes were downloaded from NCBI ftp://ftp.ncbi.nih.gov/genomes/Bacteria/ in March 2009. A detailed list can be found in Additional File 3. In order to recover missing annotation, we selected all orthologous groups covering at least 75% of all species that are good candidates of domain-wide commons. The unique set of sequences of each orthologous group was blasted against all genomes that lack an annotated ortholog using tblastn (E-value < 10-20). The sequence of the best alignment was then added to the orthologous group.

For evaluation we used the proteome data from the COG-database ftp://ftp.ncbi.nih.gov/pub/COG/COG/ downloaded in November 2009. We have chosen Bacillus halodurans, Bacillus subtilis, Lactococcus lactis, Listeria innocua, Streptococcus pneumoniae TIGR4, Streptococcus pyogenes M1 GAS from the Gram-positive bacilli class, Buchnera sp. APS, Escherichia coli K12, Pasteurella multocida, Salmonella typhimurium LT2, Vibrio cholerae, Yersinia pestis from the gamma proteobacteria class and Brucella melitensis, Caulobacter vibrioides, Mesorhizobium loti, Rickettsia prowazekii from the alpha proteobacteria class. Both, Proteinortho and OrthoMCL were applied to this set. All groups with proteins covering at least 6 species were compared to the COG-database, illustrated in Figure 4 and Figure 5.

Availability

The source code of Proteinortho can be obtained under the GPLv2 (or later) from http://www.bioinf.uni-leipzig.de/Software/proteinortho/

References

  1. Fitch WM: Distinguishing homologous from analogous proteins. Syst Zool 1970, 19(2):99–113. 10.2307/2412448

    Article  CAS  PubMed  Google Scholar 

  2. Berglund AC, Sjölund E, Ostlund G, Sonnhammer EL: InParanoid 6: eukaryotic ortholog clusters with inparalogs. Nucleic Acids Res 2008, (36 Database):D263–266.

  3. Chen F, Mackey AJ, Stoeckert CJ Jr, Roos DS: OrthoMCL-DB: querying a comprehensive multi-species collection of ortholog groups. Nucleic Acids Res 2006, (34 Database):D363–368.

  4. Tatusov RL, Galperin MY, Natale DA, Koonin EV: The COG database: a tool for genome-scale analysis of protein functions and evolution. Nucleic Acids Res 2000, 28: 33–36. 10.1093/nar/28.1.33

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  5. Wheeler DL, Barrett T, Benson DA, Bryant SH, Canese K, Chetvernin V, Church DM, Dicuccio M, Edgar R, Federhen S, Feolo M, Geer LY, Helmberg W, Kapustin Y, Khovayko O, Landsman D, Lipman DJ, Madden TL, Maglott DR, Miller V, Ostell J, Pruitt KD, Schuler GD, Shumway M, Sequeira E, Sherry ST, Sirotkin K, Souvorov A, Starchenko G, Tatusov RL, Tatusova TA, Wagner L, Yaschenko E: Database resources of the National Center for Biotechnology Information. Nucleic Acids Res 2008, (36 Database):D13–21.

  6. Jensen LJ, Julien P, Kuhn M, von Mering C, Muller J, Doerks T, Bork P: eggNOG: automated construction and annotation of orthologous groups of genes. Nucleic Acids Res 2008, (36 Database):D250–254.

  7. Schneider A, Dessimoz C, Gonnet GH: OMA Browser-exploring orthologous relations across 352 complete genomes. Bioinformatics 2007, 23(16):2180–2182. 10.1093/bioinformatics/btm295

    Article  CAS  PubMed  Google Scholar 

  8. Hubbard TJ, Aken BL, Beal K, Ballester B, Caccamo M, Chen Y, Clarke L, Coates G, Cunningham F, Cutts T, Down T, Dyer SC, Fitzgerald S, Fernandez-Banet J, Graf S, Haider S, Hammond M, Herrero J, Holland R, Howe K, Howe K, Johnson N, Kahari A, Keefe D, Kokocinski F, Kulesha E, Lawson D, Longden I, Melsopp C, Megy K, Meidl P, Ouverdin B, Parker A, Prlic A, Rice S, Rios D, Schuster M, Sealy I, Severin J, Slater G, Smedley D, Spudich G, Trevanion S, Vilella A, Vogel J, White S, Wood M, Cox T, Curwen V, Durbin R, Fernandez-Suarez XM, Flicek P, Kasprzyk A, Proctor G, Searle S, Smith J, Ureta-Vidal A, Birney E: Ensembl 2007. Nucleic Acids Res 2007, (35 Database):D610–617.

    Google Scholar 

  9. UniProt Consortium: The Universal Protein Resource (UniProt) in 2010. Nucleic Acids Res 2010, (38 Database):D142–8.

  10. Koonin EV: Orthologs, paralogs, and evolutionary genomics. Annu Rev Genet 2005, 39: 309–38. 10.1146/annurev.genet.39.073003.114725

    Article  CAS  PubMed  Google Scholar 

  11. Bork P, Dandekar T, Diaz-Lazcoz Y, Eisenhaber F, Huynen M, Yuan Y: Predicting function: from genes to genomes and back. J Mol Biol 1998, 283(4):707–725. 10.1006/jmbi.1998.2144

    Article  CAS  PubMed  Google Scholar 

  12. Rivera MC, Jain R, Moore JE, Lake JA: Genomic evidence for two functionally distinct gene classes. Proc Natl Acad Sci USA 1998, 95(11):6239–6244. 10.1073/pnas.95.11.6239

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  13. Hirsh AE, Fraser HB: Protein dispensability and rate of evolution. Nature 2001, 411(6841):1046–1049. 10.1038/35082561

    Article  CAS  PubMed  Google Scholar 

  14. Remm M, Storm CE, Sonnhammer EL: Automatic clustering of orthologs and in-paralogs from pairwise species comparisons. J Mol Biol 2001, 314(5):1041–1052. 10.1006/jmbi.2000.5197

    Article  CAS  PubMed  Google Scholar 

  15. Jordan IK, Rogozin IB, Wolf YI, Koonin EV: Essential genes are more evolutionarily conserved than are nonessential genes in bacteria. Genome Res 2002, 12(6):962–968.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Li L, Stoeckert CJ Jr, Roos DS: OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res 2003, 13(9):2178–2189. 10.1101/gr.1224503

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  17. Alexeyenko A, Tamas I, Liu G, Sonnhammer EL: Automatic clustering of orthologs and inparalogs shared by multiple proteomes. Bioinformatics 2006, 22(14):e9–15. 10.1093/bioinformatics/btl213

    Article  CAS  PubMed  Google Scholar 

  18. Fu Z, Chen X, Vacic V, Nan P, Zhong Y, Jiang T: MSOAR: a high-throughput ortholog assignment system based on genome rearrangement. J Comput Biol 2007, 14(9):1160–1175. 10.1089/cmb.2007.0048

    Article  CAS  PubMed  Google Scholar 

  19. van Dongen SM: Graph Clustering by Flow Simulation. PhD thesis. University of Utrecht, The Netherlands; 2000.

    Google Scholar 

  20. Enright AJ, Van Dongen S, Ouzounis CA: An efficient algorithm for large-scale detection of protein families. Nucleic Acids Res 2002, 30(7):1575–1584. 10.1093/nar/30.7.1575

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  21. Cornaz D: A linear programming formulation for the maximum complete multipartite subgraph problem. Math Program, Ser B 2006, 105: 329–344. 10.1007/s10107-005-0656-6

    Article  Google Scholar 

  22. Guattery S, Miller GL: On the performance of spectral graph partitioning methods. In Proceedings of the sixth annual ACM-SIAM Symposium on Discrete Algorithms. Philadelphia, PA: Society for Industrial and Applied Mathematics; 1995:233–242.

    Google Scholar 

  23. Sikdar M, Kim JS: Expression of a natural fusion gene for uracil phosphoribosyltransferase and uridine kinase from rice shows growth retardation by 5-fluorouridine or 5-fluorouracil in Escherichia coli. African Journal of Biotechnology 2010, 9(9):1295–1303.

    CAS  Google Scholar 

  24. Hopcroft J, Tarjan R: Efficient algorithms for graph manipulation. Commun ACM 1973, 16: 372–378. 10.1145/362248.362272

    Article  Google Scholar 

  25. Fiedler M: Algebraic Connectivity of Graphs. Czechoslovak Math J 1973, 23: 298–305.

    Google Scholar 

  26. Force A, Lynch M, Pickett FB, Amores A, Yan YL, Postlethwait J: Preservation of duplicate genes by complementary, degenerative mutations. Genetics 1999, 151(4):1531–1545.

    PubMed Central  CAS  PubMed  Google Scholar 

  27. Fiedler M: A property of eigenvectors of nonnegative symmetric matrices and its application to graph theory. Czechoslovak Math J 1975, 25: 619–633.

    Article  Google Scholar 

  28. Kristensen DM, Kannan L, Coleman MK, Wolf YI, Sorokin A, Koonin EV, Mushegian A: A low-polynomial algorithm for assembling clusters of orthologous groups from intergenomic symmetric best matches. Bioinformatics 2010, 26: 1481–1487. 10.1093/bioinformatics/btq229

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  29. Lanczos C: An Iteration Method for the Solution of the Eigenvalue Problem of Linear Differential and Integral Operators. J Res Natl Bureau Standards 1950, 45: 255–281.

    Article  Google Scholar 

  30. Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, Glöckner FO: SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res 2007, 35(21):7188–7196. 10.1093/nar/gkm864

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  31. Nakabachi A, Yamashita A, Toh H, Ishikawa H, Dunbar HE, Moran NA, Hattori M: The 160-kilobase genome of the bacterial endosymbiont Carsonella. Science 2006, 314(5797):267. 10.1126/science.1134196

    Article  CAS  PubMed  Google Scholar 

  32. McCutcheon JP, Moran NA: Parallel genomic evolution and metabolic interdependence in an ancient symbiosis. Proc Natl Acad Sci USA 2007, 104(49):19392–19397. 10.1073/pnas.0708855104

    Article  PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported in part by DFG STA-850/6, STA-850/2, EU-project "Quantomics" and VW-project "Complex networks" and SPP 1258: "Sensory and regulatory RNAs in Prokaryotes".

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcus Lechner.

Additional information

Authors' contributions

SP implemented an initial RBAH prototype. ML implemented, optimized, and evaluated the later versions of the program. All authors contributed to the details of implementation, testing and interpretation of the data, and collaborated in writing and approved the final manuscript.

Electronic supplementary material

12859_2010_4535_MOESM1_ESM.PDF

Additional File 1:Algebraic connectivity and Fiedler vector. Iterative approximation of the Algebraic Connectivity using the Fiedler Vector. (PDF 73 KB)

12859_2010_4535_MOESM2_ESM.PDF

Additional File 2:Supplemental figures. Supplemental Figures showing how multiple instances of Proteinortho can cooperate and how the comparison to OrthoMCL looks using smaller cutoffs. (PDF 129 KB)

12859_2010_4535_MOESM3_ESM.PDF

Additional File 3:Species list for domain wide commons. Table of species and accession numbers used in this analysis. (PDF 65 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

Lechner, M., Findeiß, S., Steiner, L. et al. Proteinortho: Detection of (Co-)orthologs in large-scale analysis. BMC Bioinformatics 12, 124 (2011). https://doi.org/10.1186/1471-2105-12-124

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2105-12-124

Keywords