PLOS/Inferring horizontal gene transfer

From Wikiversity
Jump to navigation Jump to search

PLOS Topic Pages
PLOS Computational Biology • PLOS Genetics • PLOS ONE

OPEN ACCESS (CC BY 4.0)

This article has been published as a PLOS Topic Page

Published in PLOS Computational Biology
Adapted for English Wikipedia
Public peer review comments can be seen here
Published under Creative Commons License CC BY 4.0
10.1371/journal.pcbi.1004095


Authors
Matt Ravenhall#, Nives Škunca# , Florent Lassalle , Christophe Dessimoz
About the Authors 

Matt Ravenhall#
AFFILIATION: University College London, Gower St, London WC1E 6BT, United Kingdom,

Nives Škunca#
AFFILIATION: ETH Zurich, Universitaetstrasse 19, 8092 Zurich, Switzerland;
Swiss Institute of Bioinformatics, Universitaetstrasse 19, 8092 Zurich, Switzerland,

Florent Lassalle
AFFILIATION: University College London, Gower St, London WC1E 6BT, United Kingdom,

Christophe Dessimoz
AFFILIATION: University College London, Gower St, London WC1E 6BT, United Kingdom;
European Molecular Biology Laboratory, European Bioinformatics Institute, Wellcome Trust Genome Campus, Hinxton, Cambridge CB10 1SD, United Kingdom,


#equal contribution

Horizontal or Lateral Gene Transfer (HGT or LGT) is the transmission of portions of genomic DNA between organisms through a process decoupled from vertical inheritance. In the presence of HGT events, different fragments of the genome are the result of different evolutionary histories. This can therefore complicate the investigations of evolutionary relatedness of lineages and species. Also, as HGT can bring into genomes radically different genotypes from distant lineages, or even new genes bearing new functions, it is a major source of phenotypic innovation and a mechanism of niche adaptation. For example, of particular relevance to human health is the lateral transfer of antibiotic resistance and pathogenicity determinants, leading to the emergence of pathogenic lineages.[1]

Computational identification of HGT events relies upon the investigation of sequence composition or evolutionary history of genes. Sequence composition-based ("parametric") methods search for deviations from the genomic average whereas evolutionary history-based ("phylogenetic") approaches identify genes whose evolutionary history significantly differs from that of the host species. The evaluation and benchmarking of HGT inference methods typically rely upon simulated genomes, for which the true history is known. On real data, different methods tend to infer different HGT events, and as a result it can be difficult to ascertain all but simple and clear-cut HGT events.

Introduction[edit | edit source]

Horizontal gene transfer was first observed in 1928, in Frederick Griffith's experiment: showing that virulence was able to pass from virulent to non-virulent strains of Streptococcus pneumoniae, Griffith demonstrated that genetic information can be horizontally transferred between bacteria via a mechanism known as transformation.[2] Similar observations in the 1940s[3] and 1950s[4] showed evidence that conjugation and transduction are additional mechanisms of horizontal gene transfer.[5]

To infer HGT events, which may not necessarily result in phenotypic changes, most contemporary methods are based on analyses of genomic sequence data. These methods can be broadly separated into two groups: parametric and phylogenetic methods (see figure 1). Parametric methods search for sections of a genome that significantly differ from the genomic average, such as GC content or codon usage.[6] Phylogenetic methods examine evolutionary histories of genes involved and identify conflicting phylogenies. Phylogenetic methods can be further divided into those that reconstruct and compare phylogenetic trees explicitly, and those that use surrogate measures in place of the phylogenetic trees.[7]

The main feature of parametric methods is that they only rely on the genome under study to infer HGT events that may have occurred on its lineage. It has been a considerable advantage at the early times of the sequencing era, when few closely related genomes were available for comparative methods. However, because they rely on the uniformity of the host's signature to infer HGT events, not accounting for the host's intra-genomic variability will result in overpredictions — flagging native segments as possible HGT events.[8] Similarly, the transferred segments need to exhibit the donor's signature and to be significantly different from the recipient's.[6] Furthermore, genomic segments of foreign origin are subject to the same mutational processes as the rest of the host genome, and so the difference between the two tends to vanish over time, a process referred to as amelioration.[9] This limits the ability of parametric methods to detect ancient HGTs.

Phylogenetic methods benefit from the recent availability of many sequenced genomes. Indeed, as for all comparative methods, phylogenetic methods can integrate information from multiple genomes, and in particular integrate them using a model of evolution. This lends them the ability to better characterize the HGT events they infer — notably by designating the donor species and time of the transfer. However, models have limits and need to be used cautiously. For instance, the conflicting phylogenies can be the result of events not accounted for by the model, such as unrecognized paralogy due to duplication followed by gene losses. Also, many approaches rely on a reference species tree that is supposed to be known, when in many instances it can be difficult to obtain a reliable tree. Finally, the computational costs of reconstructing many gene/species trees can be prohibitively expensive. Phylogenetic methods tend to be applied to genes or protein sequences as basic evolutionary units, which limits their ability to detect HGT in regions outside or across gene boundaries.

Conceptual overview of HGT inference methods. (1) Parametric methods infer HGT by computing a statistic, here GC content, for a sliding window and comparing it to the typical range over the entire genome, here indicated between the two red horizontal lines. Regions with atypical values are inferred as having been horizontally transferred. (2) Phylogenetic approaches rely on the differences between genes and species tree evolution that result from HGT. Explicit phylogenetic methods reconstruct gene trees and infer the HGT events likely to have resulted into that particular gene tree. Implicit phylogenetic methods bypass gene tree reconstruction, e.g., by looking at discrepancies between pairwise distances between genes and their corresponding species.

Because of their complementary approaches — and often non-overlapping sets of HGT candidates — combining predictions from parametric and phylogenetic methods can yield a more comprehensive set of HGT candidate genes. Indeed, combining different parametric methods has been reported to significantly improve the quality of predictions.[10][11] Moreover, in the absence of a comprehensive set of true horizontally transferred genes, discrepancies between different methods[12][13] might be resolved through combining parametric and phylogenetic methods. However, combining inferences from multiple methods also entails a risk of an increased false-positive rate.[14]

Parametric methods[edit | edit source]

Parametric methods to infer HGT use characteristics of the genome sequence specific to particular species or clades, also called genomic signatures. If a fragment of the genome strongly deviates from the genomic signature, this is a sign of a potential horizontal transfer. For example, because bacterial GC content falls within a wide range (see figure 2), GC content of a genome segment is a simple genomic signature. Commonly used genomic signatures include nucleotide composition,[15] oligonucleotide frequencies,[16] or structural features of the genome.[17]

To detect HGT using parametric methods, the host's genomic signature needs to be clearly recognizable. However, the host's genome is not always uniform with respect to the genome signature: for example, GC content of the third codon position is lower close to the replication terminus [18] and GC content tends to be higher in highly-expressed genes.[19] Not accounting for such intra-genomic variability in the host can result in over-predictions, flagging native segments as HGT candidates.[8] Larger sliding windows can account for this variability at the cost of a reduced ability to detect smaller HGT regions.[12]

Just as importantly, horizontally transferred segments need to exhibit the donor's genomic signature. This might not be the case for ancient transfers where transferred sequences are subjected to the same mutational processes as the rest of the host genome, potentially causing their distinct signatures to "ameliorate"[9] and become undetectable through parametric methods. For example, Bdellovibrio bacteriovorus, a predatory δ-Proteobacterium, has homogeneous GC content, and it might be concluded that its genome is resistant to HGT.[20] However, subsequent analysis using phylogenetic methods identified a number of ancient HGT events in the genome of B. bacteriovorus.[21] Similarly, if the inserted segment was previously ameliorated to the host's genome, as is the case for prophage insertions[22], parametric methods might miss predicting these HGT events. Also, the donor's composition must significantly differ from the recipient's to be identified as abnormal, a condition that might be missed in the case of short- to medium-distance HGT, which are the most prevalent. Furthermore, it has been reported that recently acquired genes tend to be AT-richer than the recipient's average[15], which indicates that differences in GC-content signature may result from unknown post-acquisition mutational processes rather than from the donor's genome.

Nucleotide composition[edit | edit source]

Average GC content of coding regions compared to the genome size for selected Bacteria. There is considerable variation in average GC content across species, which makes it relevant as a genomic signature.

Bacterial GC content falls within a wide range, with Ca. Zinderia insecticola having a GC content of 13.5%[23] and Anaeromyxobacter dehalogenans having a GC content of 75%[24] (see figure 2). Even within a closely related group of α-Proteobacteria, values range from approximately 30% to 65%.[25] These differences can be exploited when detecting HGT events as a significantly different GC content for a genome segment can be an indication of foreign origin[15] (see figure 2).

Oligonucleotide spectrum[edit | edit source]

The oligonucleotide spectrum (or k-mer frequencies) measures the frequency of all possible nucleotide sequences of a particular length in the genome. It tends to vary less within genomes than between genomes and therefore can also be used as a genomic signature.[26] A deviation from this signature suggests that a genomic segment might have arrived through horizontal transfer.

The oligonucleotide spectrum owes much of its discriminatory power to the number of possible oligonucleotides: if n is the size of the vocabulary and w is oligonucleotide size, the number of possible distinct oligonucleotides is nw; for example, there are 45=1024 possible pentanucleotides. Some methods can capture the signal recorded in motifs of variable size[27], thus capturing both rare and discriminative motifs along with frequent, but more common ones.

Codon usage bias, a measure related to codon frequencies, was one of the first detection methods used in methodical assessments of HGT.[16] This approach requires a host genome which contains a bias towards certain synonymous codons (different codons which code for the same amino acid) which is clearly distinct from the bias found within the donor genome. The simplest oligonucleotide used as a genomic signature is the dinucleotide, for example the third nucleotide in a codon and the first nucleotide in the following codon represent the dinucleotide least restricted by amino acid preference and codon usage.[28]

It is important to optimise the size of the sliding window in which to count the oligonucleotide frequency: a larger sliding window will better buffer variability in the host genome at the cost of being worse at detecting smaller HGT regions.[29] A good compromise has been reported using tetranucleotide frequencies in a sliding window of 5 kb with a step of 0.5kb.[30]

A convenient method of modelling oligonucleotide genomic signatures is to use Markov chains. The transition probability matrix can be derived for endogenous vs. acquired genes,[31] from which Bayesian posterior probabilities for particular stretches of DNA can be obtained.[32]

Structural features[edit | edit source]

Just as the nucleotide composition of a DNA molecule can be represented by a sequence of letters, its structural features can be encoded in a numerical sequence. The structural features include interaction energies between neighbouring base pairs,[33] the angle of twist that makes two bases of a pair non-coplanar,[34] or DNA deformability induced by the proteins shaping the chromatin.[35]

The autocorrelation analysis of some of these numerical sequences show characteristic periodicities in complete genomes.[36] In fact, after detecting archaea-like regions in the thermophilic bacteria Thermotoga maritima,[37] periodicity spectra of these regions were compared to the periodicity spectra of the homologous regions in the archaea Pyrococcus horikoshii.[17] The revealed similarities in the periodicity were strong supporting evidence for a case of massive HGT between the bacteria and the archaea kingdoms.[17]

Genomic context[edit | edit source]

The existence of genomic islands, short (typically 10-200kb long) regions of a genome which have been acquired horizontally, lends support to the ability to identify non-native genes by their location in a genome.[38] For example, a gene of ambiguous origin which forms part of a non-native operon could be considered to be non-native. Alternatively, flanking repeat sequences or the presence of nearby integrases or transposases can indicate a non-native region.[39] A machine-learning approach combining oligonucleotide frequency scans with context information was reported to be effective at identifying genomic islands[40]. In another study, the context was used as a secondary indicator, after removal of genes which are strongly thought to be native or non-native through the use of other parametric methods.[10]

Phylogenetic methods[edit | edit source]

The use of phylogenetic analysis in the detection of HGT was advanced by the availability of many newly sequenced genomes. Phylogenetic methods detect inconsistencies in gene and species evolutionary history in two ways: explicitly, by reconstructing the gene tree and reconciling it with the reference species tree, or implicitly, by examining aspects that correlate with the evolutionary history of the genes in question, e.g., patterns of presence/absence across species, or unexpectedly short or distant pairwise evolutionary distances.

Explicit phylogenetic methods[edit | edit source]

The aim of explicit phylogenetic methods is to compare gene trees with their associated species trees. While weakly-supported differences between gene and species trees can be due to inference uncertainty, statistically significant differences can be suggestive of HGT events (see figure 1a). For example, if two genes from different species share the most recent ancestral connecting node in the gene tree, but the respective species are spaced apart in the species tree, an HGT event can be invoked. Such an approach can produce more detailed results than parametric approaches because the involved species, time and direction of transfer can potentially be identified.

As discussed in more details below, phylogenetic methods range from simple methods merely identifying discordance between gene and species trees to mechanistic models inferring probable sequences of HGT events. An intermediate strategy entails deconstructing the gene tree into smaller parts until each matches the species tree (genome spectral approaches).

Explicit phylogenetic methods rely upon the accuracy of the input rooted gene and species trees, yet these can be challenging to build.[41] Even when there is no doubt in the input trees, the conflicting phylogenies can be the result of evolutionary processes other than HGT, such as duplications and losses, causing these methods to erroneously infer HGT events when paralogy is the correct explanation. Similarly, in the presence of incomplete lineage sorting, explicit phylogeny methods can erroneously infer HGT events.[42] That is why some explicit model-based methods test multiple evolutionary scenarios involving different kinds of events, and compare their fit to the data given parsimonious or probabilistic criteria.

Tests of topologies[edit | edit source]

To detect sets of genes that fit poorly to the reference tree, one can use statistical tests of topology, such as the Kishino-Hasegawa (KH),[43] Shimodaira-Hasegawa (SH),[44] and Approximately Unbiased (AU)[45] tests. These tests assess the likelihood of the gene sequence alignment when the reference topology is given as the null hypothesis.

The rejection of the reference topology is an indication that the evolutionary history for that gene family is inconsistent with the reference tree. When these inconsistencies cannot be explained using a small number of non-horizontal events, such as gene loss and duplication, an HGT event is inferred.

One such analysis checked for HGT in groups of homologs of the γ-Proteobacterial lineage.[46] Six reference trees were reconstructed using either the highly conserved small subunit ribosomal RNA sequences, a consensus of the available gene trees or concatenated alignments of orthologs. The failure to reject the six evaluated topologies, and the rejection of seven alternative topologies, was interpreted as evidence for a small number of HGT events in the selected groups.

Tests of topology identify differences in tree topology taking into account the uncertainty in tree inference but they make no attempt at inferring how the differences came about. To infer the specifics of particular events, genome spectral or subtree pruning and regraft methods are required.

Genome spectral approaches[edit | edit source]

In order to identify the location of HGT events, genome spectral approaches decompose a gene tree into substructures (such as bipartitions or quartets) and identify those that are consistent or inconsistent with the species tree.

Bipartitions Removing one edge from a reference tree produces two unconnected sub-trees, each a disjoint set of nodes — a bipartition. If a bipartition is present in both the gene and the species trees, it is compatible; otherwise, it is conflicting. These conflicts can indicate an HGT event or may be the result of uncertainty in gene tree inference. To reduce uncertainty, bipartition analyses typically focus on strongly supported bipartitions such as those associated with branches with bootstrap values or posterior probabilities above certain thresholds. Any gene family found to have one or several conflicting, but strongly supported, bipartitions is considered as an HGT candidate.[47][48]

Quartet decomposition Quartets are trees consisting of four leaves. In bifurcating (fully resolved) trees, each internal branch induces a quartet whose leaves are either subtrees of the original tree or actual leaves of the original tree. If the topology of a quartet extracted from the reference species tree is embedded in the gene tree, the quartet is compatible with the gene tree. Conversely, incompatible strongly supported quartets indicate potential HGT events.[49] Quartet mapping methods are much more computationally efficient and naturally handle heterogeneous representation of taxa among gene families, making them a good basis for developing large-scale scans for HGT, looking for highways of gene sharing in databases of hundreds of complete genomes[50][51].

Subtree pruning and regrafting[edit | edit source]

A mechanistic way of modelling an HGT event on the reference tree is to first cut an internal branch—i.e., prune the tree—and then regraft it onto another edge, an operation referred to as subtree pruning and regrafting (SPR).[52] If the gene tree was topologically consistent with the original reference tree, the editing results in an inconsistency. Similarly, when the original gene tree is inconsistent with the reference tree, it is possible to obtain a consistent topology by a series of one or more prune and regraft operations applied to the reference tree. By interpreting the edit path of pruning and regrafting, HGT candidate nodes can be flagged and the host and donor genomes inferred.[48][53] To avoid reporting false positive HGT events due to uncertain gene tree topologies, the optimal "path" of SPR operations can be chosen among multiple possible combinations by considering the branch support in the gene tree. Weakly supported gene tree edges can be ignored a priori[54] or the support can be used to compute an optimality criterion[55][56].

Because conversion of one tree to another by a minimum number of SPR operations is NP-Hard,[57] solving the problem becomes considerably more difficult as more nodes are considered. The computational challenge lies in finding the optimal edit path, i.e., the one that requires the fewest steps,[58][59] and different strategies are used in solving the problem. For example, the HorizStory algorithm reduces the problem by first eliminating the consistent nodes[60]; recursive pruning and regrafting reconciles the reference tree with the gene tree and optimal edits are interpreted as HGT events. The SPR methods included in the supertree reconstruction package SPRSupertrees substantially decrease the time of the search for the optimal set of SPR operations by considering multiple localised sub-problems in large trees through a clustering approach.[61]

Model-based reconciliation methods[edit | edit source]

Reconciliation of gene and species trees entails mapping evolutionary events onto gene trees in a way that makes them concordant with the species tree. Different reconciliation models exist, differing in the types of event they consider to explain the incongruences between gene and species tree topologies. Early methods exclusively modelled horizontal transfers (T).[52][55] More recent ones also account for duplication (D), loss (L), incomplete lineage sorting (ILS) or homologous recombination (HR) events. The difficulty is that by allowing for multiple types of events, the number of possible reconciliations increases rapidly. For instance, a conflicting gene tree topologies might be explained in terms of a single HGT event or multiple duplication and loss events. Both alternatives can be considered plausible reconciliation depending on the frequency of these respective events along the species tree.

Reconciliation methods can rely on a parsimonious or a probabilistic framework to infer the most likely scenario(s), where the relative cost/probability of D, T, L events can be fixed a priori or estimated from the data.[62] The space of DTL reconciliations and their parsimony costs—which can be extremely vast for large multi-copy gene family trees—can be efficiently explored through dynamic programming algorithms.[63][64][65] In some programs, the gene tree topology can be refined where it was uncertain to fit a better evolutionary scenario as well as the initial sequence alignment.[63][66][67] More refined models account for the biased frequency of HGT between closely related lineages,[68] reflecting the loss of efficiency of HR with phylogenetic distance,[69] for ILS,[70] or for the fact that the actual donor of most HGT belong to extinct or unsampled lineages.[71] Further extensions of DTL models are being developed towards an integrated description of the genome evolution processes. In particular, some of them consider horizontal at multiple scales — modelling independent evolution of gene fragments[72] or recognising co-evolution of several genes (e.g., due to co-transfer) within and across genomes.[73]

Implicit phylogenetic methods[edit | edit source]

In contrast to explicit phylogenetic methods, which compare the agreement between gene and species trees, implicit phylogenetic methods compare evolutionary distances or sequence similarity. Here, an unexpectedly short or long distance from a given reference compared to the average can be suggestive of an HGT event (see figure 1). Because tree construction is not required, implicit approaches tend to be simpler and faster than explicit methods.

However, implicit methods can be limited by disparities between the underlying correct phylogeny and the evolutionary distances considered. For instance, the most similar sequence as obtained by the highest-scoring BLAST hit is not always the evolutionarily closest one.[74]

Top sequence match in a distant species[edit | edit source]

A simple way of identifying HGT events is by looking for high-scoring sequence matches in distantly related species. For example, an analysis of the top BLAST hits of protein sequences in the bacteria Thermotoga maritima revealed that most hits were in archaea rather than closely-related bacteria, suggesting extensive HGT between the two;[37] these predictions were later supported by an analysis of the structural features of the DNA molecule.[17]

However, this method is limited to detecting relatively recent HGT events. Indeed, if the HGT occurred in the common ancestor of two or more species included in the database, the closest hit will reside within that clade and therefore the HGT will not be detected by the method. Thus, the threshold of the minimum number of foreign top BLAST hits to observe to decide a gene was transferred is highly dependent on the taxonomic coverage of sequence databases. Therefore, experimental settings may need to be defined in an ad-hoc way.[75]

Discrepancy between gene and species distances[edit | edit source]

The molecular clock hypothesis posits that homologous genes evolve at an approximately constant rate across different species.[76] If one only considers homologous genes related through speciation events (referred to as “orthologous" genes), their underlying tree should by definition correspond to the species tree. Therefore, assuming a molecular clock, the evolutionary distance between orthologous genes should be approximately proportional to the evolutionary distances between their respective species. If a putative group of orthologs contains xenologs (pairs of genes related through an HGT), the proportionality of evolutionary distances may only hold among the orthologs, not the xenologs.[77]

Simple approaches compare the distribution of similarity scores of particular sequences and their orthologous counterparts in other species; HGT are inferred from outliers.[78][79] The more sophisticated DLIGHT ('Distance Likelihood-based Inference of Genes Horizontally Transferred') method considers simultaneously the effect of HGT on all sequences within groups of putative orthologs[7]: if a likelihood-ratio test of the HGT hypothesis versus a hypothesis of no HGT is significant, a putative HGT event is inferred. In addition, the method allows inference of potential donor and recipient species and provides an estimation of the time since the HGT event.

Phylogenetic profiles[edit | edit source]

A group of orthologous or homologous genes can be analysed in terms of the presence or absence of group members in the reference genomes; such patterns are called phylogenetic profiles.[80] To find HGT events, phylogenetic profiles are scanned for an unusual distribution of genes. Absence of a homolog in some members of a group of closely related species is an indication that the examined gene might have arrived via an HGT event. For example, the three facultatively symbiotic Frankia sp. strains are of strikingly different sizes: 5.43 Mbp, 7.50 Mbp and 9.04 Mbp, depending on their range of hosts.[81] Marked portions of strain-specific genes were found to have no significant hit in the reference database, and were possibly acquired by HGT transfers from other bacteria. Similarly, the three phenotypically diverse Escherichia coli strains (uropathogenic, enterohemorrhagic and benign) share about 40% of the total combined gene pool, with the other 60% being strain-specific genes and consequently HGT candidates.[82] Further evidence for these genes resulting from HGT was their strikingly different codon usage patterns from the core genes and a lack of gene order conservation (order conservation is typical of vertically-evolved genes).[82] The presence/absence of homologs (or their effective count) can thus be used by programs to reconstruct the most likely evolutionary scenario along the species tree. Just as with reconciliation methods, this can be achieved through parsimonious[83] or probabilistic estimation of the number of gain and loss events.[84][85] Models can be complexified by adding processes, like the truncation of genes [86], but also by modelling the heterogeneity of rates of gain and loss across lineages[87] and/or gene families[88][85].

Clusters of polymorphic sites[edit | edit source]

Genes are commonly regarded as the basic units transferred through an HGT event. However it is also possible for HGT to occur within genes. For example, it has been shown that horizontal transfer between closely related species results in more exchange of ORF fragments[89][90], a type a transfer called gene conversion, mediated by homologous recombination. The analysis of a group of four Escherichia coli and two Shigella flexneri strains revealed that the sequence stretches common to all six strains contain polymorphic sites, consequences of homologous recombination.[91]. Clusters of excess of polymorphic sites can thus be used to detect tracks of DNA recombined with a distant relative[92]. This method of detection is, however, restricted to the sites in common to all analysed sequences, limiting the analysis to a group of closely related organisms.

Evaluation[edit | edit source]

The existence of the numerous and varied methods to infer HGT raises the question of how to validate individual inferences and of how to compare the different methods.

A main problem is that, as with other types of phylogenetic inferences, the actual evolutionary history cannot be established with certainty. As a result, it is difficult to obtain a representative test set of HGT events. Furthermore, HGT inference methods vary considerably in the information they consider and often identify inconsistent groups of HGT candidates:[93][6] it is not clear to what extent taking the intersection, the union, or some other combination of the individual methods affects the false positive and false negative rates[14].

Parametric and phylogenetic methods draw on different sources of information; it is therefore difficult to make general statements about their relative performance. Conceptual arguments can however be invoked. While parametric methods are limited to the analysis of single or pairs of genomes, phylogenetic methods provide a natural framework to take advantage of the information contained in multiple genomes. In many cases, segments of genomes inferred as HGT based on their anomalous composition can also be recognised as such on the basis of phylogenetic analyses or through their mere absence in genomes of related organisms. In addition, phylogenetic methods rely on explicit models of sequence evolution, which provide a well-understood framework for parameter inference, hypothesis testing, and model selection. This is reflected in the literature, which tends to favour phylogenetic methods as the standard of proof for HGT.[94][95][96][97] The use of phylogenetic methods thus appears to be the preferred standard, especially given that the increase in computational power coupled with algorithmic improvements has made them more tractable,[71][61] and that the ever denser sampling of genomes lends more power to these tests.

Considering phylogenetic methods, several approaches to validating individual HGT inferences and benchmarking methods have been adopted, typically relying on various forms of simulation. Because the truth is known in simulation, the number of false positives and the number of false negatives are straightforward to compute. However, simulating data do not trivially resolve the problem because the true extent of HGT in nature remains largely unknown, and specifying rates of HGT in the simulated model is always hasardous. Nonetheless, studies involving the comparison of several phylogenetic methods in a simulation framework could provide quantitative assessment of their respective performances, and thus help the biologist in choosing objectively proper tools.[56]

Standard tools to simulate sequence evolution along trees such as INDELible[98] or PhyloSim[99] can be adapted to simulate HGT. HGT events cause the relevant gene trees to conflict with the species tree. Such HGT events can be simulated through subtree pruning and regrafting rearrangements of the species tree.[54] However, it is important to simulate data that are realistic enough to be representative of the challenge provided by real datasets, and simulation under complex models are thus preferable. A model was developed to simulate gene trees with heterogeneous substitution processes in addition to the occurence of transfer, and accounting for the fact that transfer can come from now extinct donor lineages.[100] Alternatively, the genome evolution simulator ALF[101] directly generates gene families subject to HGT, by accounting for a whole range of evolutionary forces at the base level, but in the context of a complete genome. Given simulated sequences which have HGT, analysis of those sequences using the methods of interest and comparison of their results with the known truth permits study of their performance. Similarly, testing the methods on sequence known not to have HGT enables the study of false positive rates.

Simulation of HGT events can also be performed by manipulating the biological sequences themselves. Artificial chimeric genomes can be obtained by inserting known foreign genes into random positions of a host genome.[102][103][104][12] The donor sequences are inserted into the host unchanged or can be further evolved by simulation,[7] e.g., using the tools described above.

One important caveat to simulation as a way to assess different methods is that simulation is based on strong simplifying assumptions which may favour particular methods.[105]

Acknowledgements[edit | edit source]

The authors thank Daniel A. Dalquen, Nick Goldman, Kevin Gori, Jelena Repar, Hannes Röst, Greg Slodkowicz, Fran Supek, and Stefan Zoller for helpful comments and suggestions, as well as Tom Williams and Robert Beiko for constructive peer-reviews. We also thank Daniel Mietchen for his engaged and valuable editorial work. This article started as an assignment for the graduate course "Reviews in Computational Biology" (263-5151-00L) at ETH Zurich.

References[edit | edit source]

  1. Hiramatsu, K.; Cui, L.; Kuroda, M.; Ito, T. (2001). "The emergence and evolution of methicillin-resistant Staphylococcus aureus". Trends in Microbiology 9 (10): 486–493. doi:10.1016/s0966-842x(01)02175-8. PMID 11597450. 
  2. Griffith, F. (1928). "The Significance of Pneumococcal Types". The Journal of Hygiene 27 (2): 113–159. doi:10.1017/s0022172400031879. PMID 20474956. PMC 2167760. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2167760/. 
  3. Tatum, E. L.; Lederberg, J. (1947). "Gene Recombination in the Bacterium Escherichia coli". Journal of Bacteriology 53 (6): 673–684. doi:10.1128/jb.53.6.673-684.1947. PMID 16561324. PMC 518375. //www.ncbi.nlm.nih.gov/pmc/articles/PMC518375/. 
  4. ZINDER ND; LEDERBERG J (1952). "Genetic exchange in Salmonella". Journal of Bacteriology 64 (5): 679–699. doi:10.1128/jb.64.5.679-699.1952. PMID 12999698. PMC 169409. //www.ncbi.nlm.nih.gov/pmc/articles/PMC169409/. 
  5. Jones, D.; Sneath, P. H. (1970). "Genetic transfer and bacterial taxonomy". Bacteriological Reviews 34 (1): 40–81. doi:10.1128/br.34.1.40-81.1970. PMID 4909647. PMC 378348. //www.ncbi.nlm.nih.gov/pmc/articles/PMC378348/. 
  6. 6.0 6.1 6.2 Lawrence, J. G.; Ochman, H. (2002). "Reconciling the many faces of lateral gene transfer". Trends in Microbiology 10 (1): 1–4. doi:10.1016/s0966-842x(01)02282-x. PMID 11755071. 
  7. 7.0 7.1 7.2 Dessimoz, Christophe, Daniel Margadant, and Gaston H Gonnet. 2008. DLIGHT - Lateral Gene Transfer Detection Using Pairwise Evolutionary Distances in a Statistical Framework. Springer, 4955:315-330.
  8. 8.0 8.1 Guindon, S.; Perrière, G. (2001). "Intragenomic base content variation is a potential source of biases when searching for horizontally transferred genes". Molecular Biology and Evolution 18 (9): 1838–1840. doi:10.1093/oxfordjournals.molbev.a003972. PMID 11504864. 
  9. 9.0 9.1 Lawrence, J. G.; Ochman, H. (1997). "Amelioration of bacterial genomes: Rates of change and exchange". Journal of Molecular Evolution 44 (4): 383–397. doi:10.1007/pl00006158. PMID 9089078. 
  10. 10.0 10.1 Azad, R. K.; Lawrence, J. G. (2011). "Towards more robust methods of alien gene detection". Nucleic Acids Research 39 (9): e56. doi:10.1093/nar/gkr059. PMID 21297116. PMC 3089488. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3089488/. 
  11. Xiong, D.; Xiao, F.; Liu, L.; Hu, K.; Tan, Y.; He, S.; Gao, X. (2012). "Towards a better detection of horizontally transferred genes by combining unusual properties effectively". PLOS ONE 7 (8): e43126. doi:10.1371/journal.pone.0043126. PMID 22905214. PMC 3419211. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3419211/. 
  12. 12.0 12.1 12.2 Becq, J.; Churlaud, C.; Deschavanne, P. (2010). "A benchmark of parametric methods for horizontal transfers detection". PLOS ONE 5 (4): e9989. doi:10.1371/journal.pone.0009989. PMID 20376325. PMC 2848678. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2848678/. 
  13. Poptsova, M. (2009). "Testing phylogenetic methods to identify horizontal gene transfer". Horizontal Gene Transfer. Methods in Molecular Biology. 532. pp. 227–240. doi:10.1007/978-1-60327-853-9_13. ISBN 978-1-60327-852-2. 
  14. 14.0 14.1 Poptsova, M. S.; Gogarten, J. P. (2007). "The power of phylogenetic approaches to detect horizontally transferred genes". BMC Evolutionary Biology 7: 45. doi:10.1186/1471-2148-7-45. PMID 17376230. PMC 1847511. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1847511/. 
  15. 15.0 15.1 15.2 Daubin, V.; Lerat, E.; Perrière, G. (2003). "The source of laterally transferred genes in bacterial genomes". Genome Biology 4 (9): R57. doi:10.1186/gb-2003-4-9-r57. PMID 12952536. PMC 193657. //www.ncbi.nlm.nih.gov/pmc/articles/PMC193657/. 
  16. 16.0 16.1 Lawrence, J. G.; Ochman, H. (1998). "Molecular archaeology of the Escherichia coli genome". Proceedings of the National Academy of Sciences of the United States of America 95 (16): 9413–9417. doi:10.1073/pnas.95.16.9413. PMID 9689094. PMC 21352. //www.ncbi.nlm.nih.gov/pmc/articles/PMC21352/. 
  17. 17.0 17.1 17.2 17.3 Worning, P.; Jensen, L. J.; Nelson, K. E.; Brunak, S.; Ussery, D. W. (2000). "Structural analysis of DNA sequence: Evidence for lateral gene transfer in Thermotoga maritima". Nucleic Acids Research 28 (3): 706–709. doi:10.1093/nar/28.3.706. PMID 10637321. PMC 102551. //www.ncbi.nlm.nih.gov/pmc/articles/PMC102551/. 
  18. Deschavanne, P.; Filipski, J. (1995). "Correlation of GC content with replication timing and repair mechanisms in weakly expressed E.coli genes". Nucleic Acids Research 23 (8): 1350–1353. doi:10.1093/nar/23.8.1350. PMID 7753625. PMC 306860. //www.ncbi.nlm.nih.gov/pmc/articles/PMC306860/. 
  19. Wuitschick, J. D.; Karrer, K. M. (1999). "Analysis of genomic G + C content, codon usage, initiator codon context and translation termination sites in Tetrahymena thermophila". The Journal of Eukaryotic Microbiology 46 (3): 239–247. doi:10.1111/j.1550-7408.1999.tb05120.x. PMID 10377985. 
  20. Rendulic, S.; Jagtap, P.; Rosinus, A.; Eppinger, M.; Baar, C.; Lanz, C.; Keller, H.; Lambert, C. et al. (2004). "A predator unmasked: Life cycle of Bdellovibrio bacteriovorus from a genomic perspective". Science 303 (5658): 689–692. doi:10.1126/science.1093027. PMID 14752164. 
  21. Gophna, U.; Charlebois, R. L.; Doolittle, W. F. (2006). "Ancient lateral gene transfer in the evolution of Bdellovibrio bacteriovorus". Trends in Microbiology 14 (2): 64–69. doi:10.1016/j.tim.2005.12.008. PMID 16413191. 
  22. Vernikos, G. S.; Thomson, N. R.; Parkhill, J. (2007). "Genetic flux over time in the Salmonella lineage". Genome Biology 8 (6): R100. doi:10.1186/gb-2007-8-6-r100. PMID 17547764. PMC 2394748. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2394748/. 
  23. McCutcheon, J. P.; Moran, N. A. (2010). "Functional convergence in reduced genomes of bacterial symbionts spanning 200 My of evolution". Genome Biology and Evolution 2: 708–718. doi:10.1093/gbe/evq055. PMID 20829280. PMC 2953269. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2953269/. 
  24. Liu, Z.; Venkatesh, S. S.; Maley, C. C. (2008). "Sequence space coverage, entropy of genomes and the potential to detect non-human DNA in human samples". BMC Genomics 9: 509. doi:10.1186/1471-2164-9-509. PMID 18973670. PMC 2628393. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2628393/. 
  25. Bentley, S. D.; Parkhill, J. (2004). "Comparative genomic structure of prokaryotes". Annual Review of Genetics 38: 771–792. doi:10.1146/annurev.genet.38.072902.094318. PMID 15568993. 
  26. Karlin, S.; Burge, C. (1995). "Dinucleotide relative abundance extremes: A genomic signature". Trends in Genetics : Tig 11 (7): 283–290. doi:10.1016/s0168-9525(00)89076-9. PMID 7482779. 
  27. Vernikos, G. S.; Parkhill, J. (2006). "Interpolated variable order motifs for identification of horizontally acquired DNA: Revisiting the Salmonella pathogenicity islands". Bioinformatics (Oxford, England) 22 (18): 2196–2303. doi:10.1093/bioinformatics/btl369. PMID 16837528. 
  28. Hooper, S. D.; Berg, O. G. (2002). "Detection of genes with atypical nucleotide sequence in microbial genomes". Journal of Molecular Evolution 54 (3): 365–375. doi:10.1007/s00239-001-0051-8. PMID 11847562. 
  29. Deschavanne, P. J.; Giron, A.; Vilain, J.; Fagot, G.; Fertil, B. (1999). "Genomic signature: Characterization and classification of species assessed by chaos game representation of sequences". Molecular Biology and Evolution 16 (10): 1391–1399. doi:10.1093/oxfordjournals.molbev.a026048. PMID 10563018. 
  30. Dufraigne, C.; Fertil, B.; Lespinats, S.; Giron, A.; Deschavanne, P. (2005). "Detection and characterization of horizontal transfers in prokaryotes using genomic signature". Nucleic Acids Research 33 (1): e6. doi:10.1093/nar/gni004. PMID 15653627. PMC 546175. //www.ncbi.nlm.nih.gov/pmc/articles/PMC546175/. 
  31. Cortez, D.; Forterre, P.; Gribaldo, S. (2009). "A hidden reservoir of integrative elements is the major source of recently acquired foreign genes and ORFans in archaeal and bacterial genomes". Genome Biology 10 (6): R65. doi:10.1186/gb-2009-10-6-r65. PMID 19531232. PMC 2718499. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2718499/. 
  32. Nakamura, Y.; Itoh, T.; Matsuda, H.; Gojobori, T. (2004). "Biased biological functions of horizontally transferred genes in prokaryotic genomes". Nature Genetics 36 (7): 760–766. doi:10.1038/ng1381. PMID 15208628. 
  33. Ornstein, R. L.; Rein, R. (1978). "An optimized potential function for the calculation of nucleic acid interaction energies I. Base stacking". Biopolymers 17 (10): 2341–2360. doi:10.1002/bip.1978.360171005. PMID 24624489. 
  34. El Hassan, M. A.; Calladine, C. R. (1996). "Propeller-twisting of base-pairs and the conformational mobility of dinucleotide steps in DNA". Journal of Molecular Biology 259 (1): 95–103. doi:10.1006/jmbi.1996.0304. PMID 8648652. 
  35. Olson, W. K.; Gorin, A. A.; Lu, X. J.; Hock, L. M.; Zhurkin, V. B. (1998). "DNA sequence-dependent deformability deduced from protein-DNA crystal complexes". Proceedings of the National Academy of Sciences of the United States of America 95 (19): 11163–11168. doi:10.1073/pnas.95.19.11163. PMID 9736707. PMC 21613. //www.ncbi.nlm.nih.gov/pmc/articles/PMC21613/. 
  36. Herzel, H.; Weiss, O.; Trifonov, E. N. (1999). "10-11 bp periodicities in complete genomes reflect protein structure and DNA folding". Bioinformatics (Oxford, England) 15 (3): 187–193. doi:10.1093/bioinformatics/15.3.187. PMID 10222405. 
  37. 37.0 37.1 Nelson, K. E.; Clayton, R. A.; Gill, S. R.; Gwinn, M. L.; Dodson, R. J.; Haft, D. H.; Hickey, E. K.; Peterson, J. D. et al. (1999). "Evidence for lateral gene transfer between Archaea and bacteria from genome sequence of Thermotoga maritima". Nature 399 (6734): 323–329. doi:10.1038/20601. PMID 10360571. 
  38. Langille, M. G.; Hsiao, W. W.; Brinkman, F. S. (2010). "Detecting genomic islands using bioinformatics approaches". Nature Reviews. Microbiology 8 (5): 373–382. doi:10.1038/nrmicro2350. PMID 20395967. 
  39. Hacker, J.; Blum-Oehler, G.; Mühldorfer, I.; Tschäpe, H. (1997). "Pathogenicity islands of virulent bacteria: Structure, function and impact on microbial evolution". Molecular Microbiology 23 (6): 1089–1097. doi:10.1046/j.1365-2958.1997.3101672.x. PMID 9106201. 
  40. Vernikos, G. S.; Parkhill, J. (2008). "Resolving the structural features of genomic islands: A machine learning approach". Genome Research 18 (2): 331–342. doi:10.1101/gr.7004508. PMID 18071028. PMC 2203631. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2203631/. 
  41. Altenhoff, A. M.; Dessimoz, C. (2012). "Inferring orthology and paralogy". Evolutionary Genomics. Methods in Molecular Biology. 855. pp. 259–279. doi:10.1007/978-1-61779-582-4_9. ISBN 978-1-61779-581-7. 
  42. Than, C.; Ruths, D.; Innan, H.; Nakhleh, L. (2007). "Confounding factors in HGT detection: Statistical error, coalescent effects, and multiple solutions". Journal of Computational Biology : A Journal of Computational Molecular Cell Biology 14 (4): 517–535. doi:10.1089/cmb.2007.A010. PMID 17572027. 
  43. Goldman, N.; Anderson, J. P.; Rodrigo, A. G. (2000). "Likelihood-based tests of topologies in phylogenetics". Systematic Biology 49 (4): 652–670. doi:10.1080/106351500750049752. PMID 12116432. 
  44. Shimodaira, H, and M Hasegawa. 1999. Multiple Comparisons of Log-Likelihoods with Applications to Phylogenetic Inference. Molecular Biology and Evolution 16 (8): 1114.
  45. Shimodaira, H. (2002). "An approximately unbiased test of phylogenetic tree selection". Systematic Biology 51 (3): 492–508. doi:10.1080/10635150290069913. PMID 12079646. 
  46. Lerat, E.; Daubin, V.; Moran, N. A. (2003). "From gene trees to organismal phylogeny in prokaryotes: The case of the gamma-Proteobacteria". PLOS Biology 1 (1): E19. doi:10.1371/journal.pbio.0000019. PMID 12975657. PMC 193605. //www.ncbi.nlm.nih.gov/pmc/articles/PMC193605/. 
  47. Zhaxybayeva, O.; Hamel, L.; Raymond, J.; Gogarten, J. P. (2004). "Visualization of the phylogenetic content of five genomes using dekapentagonal maps". Genome Biology 5 (3): R20. doi:10.1186/gb-2004-5-3-r20. PMID 15003123. PMC 395770. //www.ncbi.nlm.nih.gov/pmc/articles/PMC395770/. 
  48. 48.0 48.1 Beiko, R. G.; Harlow, T. J.; Ragan, M. A. (2005). "Highways of gene sharing in prokaryotes". Proceedings of the National Academy of Sciences of the United States of America 102 (40): 14332–14337. doi:10.1073/pnas.0504068102. PMID 16176988. PMC 1242295. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1242295/. 
  49. Zhaxybayeva, O.; Gogarten, J. P.; Charlebois, R. L.; Doolittle, W. F.; Papke, R. T. (2006). "Phylogenetic analyses of cyanobacterial genomes: Quantification of horizontal gene transfer events". Genome Research 16 (9): 1099–1108. doi:10.1101/gr.5322306. PMID 16899658. PMC 1557764. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1557764/. 
  50. Bansal, M. S.; Banay, G.; Gogarten, J. P.; Shamir, R. (2011). "Detecting highways of horizontal gene transfer". Journal of Computational Biology : A Journal of Computational Molecular Cell Biology 18 (9): 1087–1114. doi:10.1089/cmb.2011.0066. PMID 21899418. 
  51. Bansal, M. S.; Banay, G.; Harlow, T. J.; Gogarten, J. P.; Shamir, R. (2013). "Systematic inference of highways of horizontal gene transfer in prokaryotes". Bioinformatics (Oxford, England) 29 (5): 571–579. doi:10.1093/bioinformatics/btt021. PMID 23335015. 
  52. 52.0 52.1 Hallett MT, Lagergren J. RECOMB 2001. Montreal: ACM; 2001. Efficient Algorithms for Lateral Gene Transfer Problems; pp. 149–156.
  53. Baroni, M.; Grünewald, S.; Moulton, V.; Semple, C. (2005). "Bounding the number of hybridisation events for a consistent evolutionary history". Journal of Mathematical Biology 51 (2): 171–182. doi:10.1007/s00285-005-0315-9. PMID 15868201. 
  54. 54.0 54.1 Beiko, R. G.; Hamilton, N. (2006). "Phylogenetic identification of lateral genetic transfer events". BMC Evolutionary Biology 6: 15. doi:10.1186/1471-2148-6-15. PMID 16472400. PMC 1431587. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1431587/. 
  55. 55.0 55.1 Nakhleh L, Ruths DA, Wang L: RIATA-HGT: A Fast and Accurate Heuristic for Reconstructing Horizontal Gene Transfer. COCOON, August 16-29, 2005; Kunming 2005.
  56. 56.0 56.1 Abby, S. S.; Tannier, E.; Gouy, M.; Daubin, V. (2010). "Detecting lateral gene transfers by statistical reconciliation of phylogenetic forests". BMC Bioinformatics 11: 324. doi:10.1186/1471-2105-11-324. PMID 20550700. PMC 2905365. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2905365/. 
  57. Hickey, G.; Dehne, F.; Rau-Chaplin, A.; Blouin, C. (2008). "SPR distance computation for unrooted trees". Evolutionary Bioinformatics Online 4: 17–27. doi:10.4137/ebo.s419. PMID 19204804. PMC 2614206. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2614206/. 
  58. Hein, Jotun, Tao Jiang, Lusheng Wang, and Kaizhong Zhang. 1995. On the Complexity of Comparing Evolutionary Trees. http://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.28.861.
  59. Allen, Benjamin L., and Mike Steel. 2001. Subtree Transfer Operations and Their Induced Metrics on Evolutionary Trees. Annals of Combinatorics 5 (1) (June): 1-15. doi:10.1007/s00026-001-8006-8.
  60. MacLeod, D.; Charlebois, R. L.; Doolittle, F.; Bapteste, E. (2005). "Deduction of probable events of lateral gene transfer through comparison of phylogenetic trees by recursive consolidation and rearrangement". BMC Evolutionary Biology 5: 27. doi:10.1186/1471-2148-5-27. PMID 15819979. PMC 1087482. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1087482/. 
  61. 61.0 61.1 Whidden, C.; Zeh, N.; Beiko, R. G. (2014). "Supertrees Based on the Subtree Prune-and-Regraft Distance". Systematic Biology 63 (4): 566–581. doi:10.1093/sysbio/syu023. PMID 24695589. PMC 4055872. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4055872/. 
  62. Doyon, J. P.; Hamel, S.; Chauve, C. (2012). "An efficient method for exploring the space of gene tree/Species tree reconciliations in a probabilistic framework". IEEE/ACM Transactions on Computational Biology and Bioinformatics 9 (1): 26–39. doi:10.1109/TCBB.2011.64. PMID 21464510. https://hal-lirmm.ccsd.cnrs.fr/lirmm-00448486/file/RR-10002.pdf. 
  63. 63.0 63.1 David, L. A.; Alm, E. J. (2011). "Rapid evolutionary innovation during an Archaean genetic expansion". Nature 469 (7328): 93–96. doi:10.1038/nature09649. PMID 21170026. 
  64. Doyon, J. P.; Hamel, S.; Chauve, C. (2012). "An efficient method for exploring the space of gene tree/Species tree reconciliations in a probabilistic framework". IEEE/ACM Transactions on Computational Biology and Bioinformatics 9 (1): 26–39. doi:10.1109/TCBB.2011.64. PMID 21464510. https://hal-lirmm.ccsd.cnrs.fr/lirmm-00448486/file/RR-10002.pdf. 
  65. Szöllosi, G. J.; Boussau, B.; Abby, S. S.; Tannier, E.; Daubin, V. (2012). "Phylogenetic modeling of lateral gene transfer reconstructs the pattern and relative timing of speciations". Proceedings of the National Academy of Sciences of the United States of America 109 (43): 17513–17518. doi:10.1073/pnas.1202997109. PMID 23043116. PMC 3491530. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3491530/. 
  66. Nguyen, T. H.; Ranwez, V.; Pointet, S.; Chifolleau, A. M.; Doyon, J. P.; Berry, V. (2013). "Reconciliation and local gene tree rearrangement can be of mutual profit". Algorithms for Molecular Biology : Amb 8 (1): 12. doi:10.1186/1748-7188-8-12. PMID 23566548. PMC 3871789. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3871789/. 
  67. Szöllosi, G. J.; Tannier, E.; Lartillot, N.; Daubin, V. (2013). "Lateral gene transfer from the dead". Systematic Biology 62 (3): 386–397. doi:10.1093/sysbio/syt003. PMID 23355531. PMC 3622898. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3622898/. 
  68. Bansal, M. S.; Alm, E. J.; Kellis, M. (2012). "Efficient algorithms for the reconciliation problem with gene duplication, horizontal transfer and loss". Bioinformatics (Oxford, England) 28 (12): i283-91. doi:10.1093/bioinformatics/bts225. PMID 22689773. PMC 3371857. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3371857/. 
  69. Majewski, J.; Zawadzki, P.; Pickerill, P.; Cohan, F. M.; Dowson, C. G. (2000). "Barriers to genetic exchange between bacterial species: Streptococcus pneumoniae transformation". Journal of Bacteriology 182 (4): 1016–1023. doi:10.1128/JB.182.4.1016-1023.2000. PMID 10648528. PMC 94378. //www.ncbi.nlm.nih.gov/pmc/articles/PMC94378/. 
  70. Sjöstrand, J.; Tofigh, A.; Daubin, V.; Arvestad, L.; Sennblad, B.; Lagergren, J. (2014). "A Bayesian method for analyzing lateral gene transfer". Systematic Biology 63 (3): 409–420. doi:10.1093/sysbio/syu007. PMID 24562812. 
  71. 71.0 71.1 Szöllõsi, G. J.; Rosikiewicz, W.; Boussau, B.; Tannier, E.; Daubin, V. (2013). "Efficient exploration of the space of reconciled gene trees". Systematic Biology 62 (6): 901–912. doi:10.1093/sysbio/syt054. PMID 23925510. PMC 3797637. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3797637/. 
  72. Haggerty, L. S.; Jachiet, P. A.; Hanage, W. P.; Fitzpatrick, D. A.; Lopez, P.; O'Connell, M. J.; Pisani, D.; Wilkinson, M. et al. (2014). "A pluralistic account of homology: Adapting the models to the data". Molecular Biology and Evolution 31 (3): 501–516. doi:10.1093/molbev/mst228. PMID 24273322. PMC 3935183. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3935183/. 
  73. Szöllősi, G. J.; Tannier, E.; Daubin, V.; Boussau, B. (2015). "The inference of gene trees with species trees". Systematic Biology 64 (1): e42-62. doi:10.1093/sysbio/syu048. PMID 25070970. PMC 4265139. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4265139/. 
  74. Koski, L. B.; Golding, G. B. (2001). "The closest BLAST hit is often not the nearest neighbor". Journal of Molecular Evolution 52 (6): 540–542. doi:10.1007/s002390010184. PMID 11443357. 
  75. Wisniewski-Dyé, F.; Borziak, K.; Khalsa-Moyers, G.; Alexandre, G.; Sukharnikov, L. O.; Wuichet, K.; Hurst, G. B.; McDonald, W. H. et al. (2011). "Azospirillum genomes reveal transition of bacteria from aquatic to terrestrial environments". PLOS Genetics 7 (12): e1002430. doi:10.1371/journal.pgen.1002430. PMID 22216014. PMC 3245306. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3245306/. 
  76. Zuckerkandl, E. and Pauling, L.B. 1965. Evolutionary divergence and convergence in proteins. In Bryson, V.and Vogel, H.J. (editors). Evolving Genes and Proteins. Academic Press, New York. pp. 97–166.
  77. Novichkov, P. S.; Omelchenko, M. V.; Gelfand, M. S.; Mironov, A. A.; Wolf, Y. I.; Koonin, E. V. (2004). "Genome-wide molecular clock and horizontal gene transfer in bacterial evolution". Journal of Bacteriology 186 (19): 6575–6585. doi:10.1128/JB.186.19.6575-6585.2004. PMID 15375139. PMC 516599. //www.ncbi.nlm.nih.gov/pmc/articles/PMC516599/. 
  78. Lawrence, J. G.; Hartl, D. L. (1992). "Inference of horizontal genetic transfer from molecular data: An approach using the bootstrap". Genetics 131 (3): 753–760. doi:10.1093/genetics/131.3.753. PMID 1628816. PMC 1205046. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1205046/. 
  79. Clarke, G. D.; Beiko, R. G.; Ragan, M. A.; Charlebois, R. L. (2002). "Inferring genome trees by using a filter to eliminate phylogenetically discordant sequences and a distance matrix based on mean normalized BLASTP scores". Journal of Bacteriology 184 (8): 2072–2080. doi:10.1128/JB.184.8.2072-2080.2002. PMID 11914337. PMC 134965. //www.ncbi.nlm.nih.gov/pmc/articles/PMC134965/. 
  80. Pellegrini, M.; Marcotte, E. M.; Thompson, M. J.; Eisenberg, D.; Yeates, T. O. (1999). "Assigning protein functions by comparative genome analysis: Protein phylogenetic profiles". Proceedings of the National Academy of Sciences of the United States of America 96 (8): 4285–4288. doi:10.1073/pnas.96.8.4285. PMID 10200254. PMC 16324. //www.ncbi.nlm.nih.gov/pmc/articles/PMC16324/. 
  81. Normand, P. et al. (2007). "Genome characteristics of facultatively symbiotic Frankia sp. Strains reflect host range and host plant biogeography". Genome Research 17 (1): 7–15. doi:10.1101/gr.5798407. PMID 17151343. PMC 1716269. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1716269/. 
  82. 82.0 82.1 Welch, R. A.; Burland, V.; Plunkett g, 3rd; Redford, P.; Roesch, P.; Rasko, D.; Buckles, E. L.; Liou, S. R. et al. (2002). "Extensive mosaic structure revealed by the complete genome sequence of uropathogenic Escherichia coli". Proceedings of the National Academy of Sciences of the United States of America 99 (26): 17020–17024. doi:10.1073/pnas.252529799. PMID 12471157. PMC 139262. //www.ncbi.nlm.nih.gov/pmc/articles/PMC139262/. 
  83. Csűrös M. 2008. Ancestral Reconstruction by Asymmetric Wagner Parsimony over Continuous Characters and Squared Parsimony over Distributions. In Comparative Genomics (eds. C.E. Nelson and S. Vialette), Lecture Notes in Computer Science, pp. 72–86, Springer Berlin Heidelberg http://link.springer.com/chapter/10.1007/978-3-540-87989-3_6
  84. Pagel, M. (1999). "Inferring the historical patterns of biological evolution". Nature 401 (6756): 877–884. doi:10.1038/44766. PMID 10553904. 
  85. 85.0 85.1 Csurös, M.; Miklós, I. (2009). "Streamlining and large ancestral genomes in Archaea inferred with a phylogenetic birth-and-death model". Molecular Biology and Evolution 26 (9): 2087–2095. doi:10.1093/molbev/msp123. PMID 19570746. PMC 2726834. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2726834/. 
  86. Hao, W.; Golding, G. B. (2010). "Inferring bacterial genome flux while considering truncated genes". Genetics 186 (1): 411–426. doi:10.1534/genetics.110.118448. PMID 20551435. PMC 2940306. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2940306/. 
  87. Hao, W.; Golding, G. B. (2006). "The fate of laterally transferred genes: Life in the fast lane to adaptation or death". Genome Research 16 (5): 636–643. doi:10.1101/gr.4746406. PMID 16651664. PMC 1457040. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1457040/. 
  88. Hao, W.; Golding, G. B. (2008). "Uncovering rate variation of lateral gene transfer during bacterial genome evolution". BMC Genomics 9: 235. doi:10.1186/1471-2164-9-235. PMID 18492275. PMC 2426709. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2426709/. 
  89. Ochman, H.; Lawrence, J. G.; Groisman, E. A. (2000). "Lateral gene transfer and the nature of bacterial innovation". Nature 405 (6784): 299–304. doi:10.1038/35012500. PMID 10830951. 
  90. Papke, R. T.; Koenig, J. E.; Rodríguez-Valera, F.; Doolittle, W. F. (2004). "Frequent recombination in a saltern population of Halorubrum". Science 306 (5703): 1928–1929. doi:10.1126/science.1103289. PMID 15591201. 
  91. Mau, B.; Glasner, J. D.; Darling, A. E.; Perna, N. T. (2006). "Genome-wide detection and analysis of homologous recombination among sequenced strains of Escherichia coli". Genome Biology 7 (5): R44. doi:10.1186/gb-2006-7-5-r44. PMID 16737554. PMC 1779527. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1779527/. 
  92. Didelot, X.; Falush, D. (2007). "Inference of bacterial microevolution using multilocus sequence data". Genetics 175 (3): 1251–1266. doi:10.1534/genetics.106.063305. PMID 17151252. PMC 1840087. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1840087/. 
  93. Ragan, M. A. (2001). "On surrogate methods for detecting lateral gene transfer". FEMS Microbiology Letters 201 (2): 187–191. doi:10.1111/j.1574-6968.2001.tb10755.x. PMID 11470360. 
  94. Ragan, M. A.; Harlow, T. J.; Beiko, R. G. (2006). "Do different surrogate methods detect lateral genetic transfer events of different relative ages?". Trends in Microbiology 14 (1): 4–8. doi:10.1016/j.tim.2005.11.004. PMID 16356716. 
  95. Kechris, K. J.; Lin, J. C.; Bickel, P. J.; Glazer, A. N. (2006). "Quantitative exploration of the occurrence of lateral gene transfer by using nitrogen fixation genes as a case study". Proceedings of the National Academy of Sciences of the United States of America 103 (25): 9584–9589. doi:10.1073/pnas.0603534103. PMID 16769896. PMC 1480450. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1480450/. 
  96. Moran, N. A.; Jarvik, T. (2010). "Lateral transfer of genes from fungi underlies carotenoid production in aphids". Science 328 (5978): 624–627. doi:10.1126/science.1187113. PMID 20431015. 
  97. Danchin, E. G.; Rosso, M. N.; Vieira, P.; De Almeida-Engler, J.; Coutinho, P. M.; Henrissat, B.; Abad, P. (2010). "Multiple lateral gene transfers and duplications have promoted plant parasitism ability in nematodes". Proceedings of the National Academy of Sciences of the United States of America 107 (41): 17651–17656. doi:10.1073/pnas.1008486107. PMID 20876108. PMC 2955110. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2955110/. 
  98. Fletcher, W.; Yang, Z. (2009). "INDELible: A flexible simulator of biological sequence evolution". Molecular Biology and Evolution 26 (8): 1879–1888. doi:10.1093/molbev/msp098. PMID 19423664. PMC 2712615. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2712615/. 
  99. Sipos, B.; Massingham, T.; Jordan, G. E.; Goldman, N. (2011). "PhyloSim - Monte Carlo simulation of sequence evolution in the R statistical computing environment". BMC Bioinformatics 12: 104. doi:10.1186/1471-2105-12-104. PMID 21504561. PMC 3102636. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3102636/. 
  100. Galtier, N. (2007). "A model of horizontal gene transfer and the bacterial phylogeny problem". Systematic Biology 56 (4): 633–642. doi:10.1080/10635150701546231. PMID 17661231. 
  101. Dalquen, D. A.; Anisimova, M.; Gonnet, G. H.; Dessimoz, C. (2012). "ALF--a simulation framework for genome evolution". Molecular Biology and Evolution 29 (4): 1115–1123. doi:10.1093/molbev/msr268. PMID 22160766. PMC 3341827. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3341827/. 
  102. Cortez, D. Q.; Lazcano, A.; Becerra, A. (2005). "Comparative analysis of methodologies for the detection of horizontally transferred genes: A reassessment of first-order Markov models". In Silico Biology 5 (5–6): 581–592. PMID 16610135. 
  103. Tsirigos, A.; Rigoutsos, I. (2005). "A new computational method for the detection of horizontal gene transfer events". Nucleic Acids Research 33 (3): 922–933. doi:10.1093/nar/gki187. PMID 15716310. PMC 549390. //www.ncbi.nlm.nih.gov/pmc/articles/PMC549390/. 
  104. Azad, R. K.; Lawrence, J. G. (2005). "Use of artificial genomes in assessing methods for atypical gene detection". PLOS Computational Biology 1 (6): e56. doi:10.1371/journal.pcbi.0010056. PMID 16292353. PMC 1282332. //www.ncbi.nlm.nih.gov/pmc/articles/PMC1282332/. 
  105. Iantorno, S.; Gori, K.; Goldman, N.; Gil, M.; Dessimoz, C. (2014). "Who watches the watchmen? An appraisal of benchmarks for multiple sequence alignment". Multiple Sequence Alignment Methods. Methods in Molecular Biology. 1079. pp. 59–73. doi:10.1007/978-1-62703-646-7_4. ISBN 978-1-62703-645-0.