2
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: found
      Is Open Access

      Chromosome-Level Assembly of the Southern Rock Bream ( Oplegnathus fasciatus) Genome Using PacBio and Hi-C Technologies

      review-article

      Read this article at

      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          The Rock Bream ( Oplegnathus fasciatus) is an economically important rocky reef fish of the Northwest Pacific Ocean. In recent years, it has been cultivated as an important edible fish in coastal areas of China. Despite its economic importance, genome-wide adaptions of domesticated O. fasciatus are largely unknown. Here we report a chromosome-level reference genome of female O. fasciatus (from the southern population in the subtropical region) using the PacBio single molecule sequencing technique (SMRT) and High-through chromosome conformation capture (Hi-C) technologies. The genome was assembled into 120 contigs with a total length of 732.95 Mb and a contig N50 length of 27.33 Mb. After chromosome-level scaffolding, 24 chromosomes with a total length of 723.22 Mb were constructed. Moreover, a total of 27,015 protein-coding genes and 5,880 ncRNAs were annotated in the reference genome. This reference genome of O. fasciatus will provide an important resource not only for basic ecological and population genetic studies but also for dissect artificial selection mechanisms in marine aquaculture.

          Related collections

          Most cited references31

          • Record: found
          • Abstract: found
          • Article: found
          Is Open Access

          Canu: scalable and accurate long-read assembly via adaptive k-mer weighting and repeat separation

          Long-read single-molecule sequencing has revolutionized de novo genome assembly and enabled the automated reconstruction of reference-quality genomes. However, given the relatively high error rates of such technologies, efficient and accurate assembly of large repeats and closely related haplotypes remains challenging. We address these issues with Canu, a successor of Celera Assembler that is specifically designed for noisy single-molecule sequences. Canu introduces support for nanopore sequencing, halves depth-of-coverage requirements, and improves assembly continuity while simultaneously reducing runtime by an order of magnitude on large genomes versus Celera Assembler 8.2. These advances result from new overlapping and assembly algorithms, including an adaptive overlapping strategy based on tf-idf weighted MinHash and a sparse assembly graph construction that avoids collapsing diverged repeats and haplotypes. We demonstrate that Canu can reliably assemble complete microbial genomes and near-complete eukaryotic chromosomes using either Pacific Biosciences (PacBio) or Oxford Nanopore technologies and achieves a contig NG50 of >21 Mbp on both human and Drosophila melanogaster PacBio data sets. For assembly structures that cannot be linearly represented, Canu provides graph-based assembly outputs in graphical fragment assembly (GFA) format for analysis or integration with complementary phasing and scaffolding techniques. The combination of such highly resolved assembly graphs with long-range scaffolding information promises the complete and automated assembly of complex genomes.
            Bookmark
            • Record: found
            • Abstract: found
            • Article: found
            Is Open Access

            OrthoFinder: phylogenetic orthology inference for comparative genomics

            Here, we present a major advance of the OrthoFinder method. This extends OrthoFinder’s high accuracy orthogroup inference to provide phylogenetic inference of orthologs, rooted gene trees, gene duplication events, the rooted species tree, and comparative genomics statistics. Each output is benchmarked on appropriate real or simulated datasets, and where comparable methods exist, OrthoFinder is equivalent to or outperforms these methods. Furthermore, OrthoFinder is the most accurate ortholog inference method on the Quest for Orthologs benchmark test. Finally, OrthoFinder’s comprehensive phylogenetic analysis is achieved with equivalent speed and scalability to the fastest, score-based heuristic methods. OrthoFinder is available at https://github.com/davidemms/OrthoFinder.
              Bookmark
              • Record: found
              • Abstract: found
              • Article: not found

              Automated eukaryotic gene structure annotation using EVidenceModeler and the Program to Assemble Spliced Alignments

              Background Accurate and comprehensive gene discovery in eukaryotic genome sequences requires multiple independent and complementary analysis methods including, at the very least, the application of ab initio gene prediction software and sequence alignment tools. The problem is technically challenging, and despite many years of research no single method has yet been able to solve it, although numerous tools have been developed to target specialized and diverse variations on the gene finding problem (for review [1,2]). Conventional gene finding software employs probabilistic techniques such as hidden Markov models (HMMs). These models are employed to find the most likely partitioning of a nucleotide sequence into introns, exons, and intergenic states according to a prior set of probabilities for the states in the model. Such gene finding programs, including GENSCAN [3], GlimmerHMM [4], Fgenesh [5], and GeneMark.hmm [6], are effective at identifying individual exons and regions that correspond to protein-coding genes, but nevertheless they are far from perfect at correctly predicting complete gene structures, differing from correct gene structures in exon content or position [7-10]. The correct gene structures, or individual components including introns and exons, are often apparent from spliced alignments of homologous transcript or protein sequences. Many software tools are available that perform these alignment tasks. Tools used to align expressed sequence tags (ESTs) and full-length cDNAs (FL-cDNAs) to genomic sequence include EST_GENOME [11], AAT [12], sim4 [13], geneseqer [14], BLAT [15], and GMAP [16], among numerous others. The list of programs that perform spliced alignments of protein sequences to DNA are much fewer, including the multifunctional AAT, exonerate [17], and PMAP (derived from GMAP). An extension of spliced protein alignment that includes a probabilistic model of eukaryotic gene structure is implemented in GeneWise [18], a popular homology-based gene predictor that serves a critical role in the Ensembl automated genome annotation pipeline [19]. In most cases, the spliced protein alignments and transcript alignments (derived from ESTs) provide evidence for only part of the gene structure, delineating introns, complete internal exons, and potential portions of other exons at their alignment termini. A comprehensive approach to eukaryotic gene structure annotation should utilize both the information intrinsic to the genome sequence itself, as is done by ab initio gene prediction software, and any extrinsic data in the form of homologies to other known sequences, including proteins, transcripts, or conserved regions revealed from cross-genome comparisons. Some of the most recent ab initio gene finding software is able to utilize such extrinsic data to improve upon gene finding accuracy. Examples of such software are numerous, and each falls within a certain niche based on the form of extrinsic data utilized. TWINSCAN [20], for example, uses an 'informant' genome to condition the probabilities of exons and introns in a closely related genome. Subsequently, TWINSCAN_EST [21] combined spliced transcript alignments with the intrinsic data, and finally N-SCAN [22] (also known as TWINSCAN 3.0) and N-SCAN_EST [21] utilized cross-genome homologies to multiple related genome sequences in the context of a phylogenetic framework. Other tools, including Augustus [23], Genie [24], and ExonHunter [25] include mechanisms to incorporate extrinsic data into the ab initio gene prediction framework to improve accuracy further. Each of these programs analyzes and predicts genes along a single target genome sequence, while using homologies detected to other sequences. A more specialized approach to gene-finding is employed by the tools SLAM [26] and TWAIN [27], which consider homologies between two related genome sequences and simultaneously predict gene structures within both genomes. Early large-scale genome projects relied heavily on the manual annotation of gene structures in order to ensure genome annotation of the highest quality [28-30]. Manual annotation involves scientists examining all of the evidence for gene structures as described above using a graphical genome viewer and annotation editor such as Apollo [31] or Artemis [32]. These manual efforts were, and continue to be, essential to providing the best community resources in the form of high quality and accurate genome annotations. Manual annotation is limited, though, because it is time consuming, expensive, and it cannot keep pace with the advances in high-throughput DNA sequencing technology that are producing increasing quantities of genome sequences. FL-cDNA projects have lessened the need for manual curation of every gene by providing accurate and complete gene structure annotations derived from high-quality spliced alignments. Software such as Program to Assemble Spliced Alignments (PASA) [33] has enabled high-throughput automated annotation of gene structures by exploiting ESTs and FL-cDNAs alone or within the context of pre-existing annotated gene structures. Other, more comprehensive computational strategies have been developed to play the role of the human annotator by combining precomputed diverse evidence into accurate gene structure annotations. These tools include Combiner [34], JIGSAW [35], GLEAN [36], and Exogean [37], among others. These algorithms employ statistical or rule-based methods to combine evidence into a most probable correct gene structure. We present a utility called EVidenceModeler (EVM), an extension of methods that led to the original Combiner development [34,38], using a nonstochastic weighted evidence combining technique that accounts for both the type and abundance of evidence to compute weighted consensus gene structures. EVM was heavily utilized for the genome analysis of the mosquito Aedes aegypti [39], and used partially or exclusively to generate the preliminary annotation for recently sequenced genomes of the blood fluke Schistosoma mansoni [40], the protozoan oyster parasite Perkinsus marinus, the human body louse Pediculus humanus, and another mosquito, Culex pipiens. The evidence utilized by EVM corresponds primarily to ab initio gene predictions and protein and transcript alignments, generated via any of the various methods described above. The intuitive framework provided by EVM is shown to be highly effective, exploiting high quality evidence where available and providing consensus gene structure prediction accuracy that approaches that of manual annotation. EVM source code and documentation are freely available from the EVM website [41]. Results and discussion In the subsequent sections, we demonstrate EVM as an automated gene structure annotation tool using rice and human genome sequences and related evidence. First, using the rice genome, we develop the concepts that underlie the algorithm of EVM as a tool that incorporates weighted evidence into consensus gene structure predictions. We then turn our attention to the human genome, in which we examine the role of EVM in concert with PASA to annotate protein-coding genes and alternatively spliced isoforms automatically. In each scenario, we include comparisons with alternative annotation methods. Evaluation of ab initio gene prediction in rice The prediction accuracy for each of the three programs Fgenesh [5], GlimmerHMM [4], and GeneMark.hmm [6] was evaluated using a set of 1,058 cDNA-verified reference gene structures. All three were nearly equivalent in both their exon prediction accuracy (about 78% exon sensitivity [eSn] and 72% to 79% exon specificity [eSp]) and complete gene prediction accuracy (22% to 25% gene sensitivity [gSn] and 15% to 21% gene specificity [gSp]; Figure 1). The breakdown of prediction accuracy by each of the four exon types indicates that all gene predictors excel at predicting internal exons correctly (about 85% eSn) while predicting initial, terminal, and single exons less accurately (44% to 68% eSn; Figure 2). Figure 1 Rice Ab initio gene prediction accuracies. Gene prediction accuracies are shown for GeneMark.hmm, Fgenesh, and GlimmerHMM ab initio gene predictions based on an evaluation of 1058 cDNA-verified reference rice gene structures. The accuracy of EVidenceModeler (EVM) consensus predictions from combining all three ab initio predictions using equal weightings (weight = 1 for each) is also provided. Figure 2 Ab initio prediction sensitivity by exon type. Individual ab initio exon prediction sensitivities based on comparisons with 1,058 reference rice gene structures are shown for each of the four exon types: initial, internal, terminal, and single. Results are additionally shown for EVidenceModeler (EVM) consensus predictions where the ab initio predictions were combined using equal weights. Although each gene predictor exhibits a similar level of accuracy, they differ greatly in the individual gene structures they each predict correctly. The Venn diagrams provided in Figure 3 reveal the variability among genes and exons predicted correctly by the three programs. Although each program predicts up to 25% of the reference genes perfectly, only about a quarter of these (6.2%) were identified by all three programs simultaneously. It is also notable that more than half (54%) of the cDNA-verified genes are not predicted correctly by any of the gene predictors evaluated. At the individual exon level, there is much more agreement among predictions, with 60.5% of the exons correctly predicted by all three programs. Only 7.1% of exons are not predicted correctly by any of the three programs. The Venn diagrams indicate much greater overall consistency among internal exon predictions, correlated with the inherently high internal exon prediction accuracy, as compared with the greater variability and decreased prediction accuracy among other exon types. A relatively higher proportion of the single (22.1%), initial (14.4%), and terminal (13.9%) exon types found in our reference genes are completely absent from the set of predicted exons. Figure 3 Venn diagrams contrasting correctly predicted rice gene structure components by ab initio gene finders. Percentages are shown for the fraction of 1,058 cDNA verified rice genes and gene structure components that were predicted correctly by each ab initio gene predictor. The cDNA-verified gene structure components consist of 7,438 total exons: 86 single, 5408 internal, 972 initial, and 972 terminal. Consensus ab initio exon prediction accuracy Although there is considerable disagreement among exon calls between the various gene predictors, when multiple programs call exons identically they tend more frequently to be correct. Figure 4 shows that by restricting the analysis to only those exons that are predicted identically by two programs, exon prediction specificity jumps to 94% correct, regardless of the two programs chosen. Exon prediction specificity improves to 97% if we consider only those exons that are predicted identically by all three programs. Note that although the specificity improves to near-perfect accuracy, the prediction sensitivity drops from 78% to 60%. Although we cannot rely on shared exons to predict all genes correctly, we can in this circumstance trust those that are shared with greater confidence. EVM uses this increased specificity provided by consensus agreement among evidence for gene structure components and reports these specific components as part of larger complete gene structures; at the same time, EVM uses other lines of evidence to retain a high level of sensitivity. Figure 4 Exon prediction accuracy limited to consensus complete exon calls. Exon sensitivity (eSn) and exon specificity (eSp) were determined by comparing ab initio predicted exons. Exons were restricted to those perfectly agreed upon by either two or three different gene predictors. Only those predicted exons found within 500 base pairs flanking the 1,058 reference gene structures were considered for the specificity calculations. Consensus gene prediction by EVM Unlike conventional ab initio gene predictors that use only the composition of the genome sequence, EVM constructs gene structures by combining evidence derived from secondary sources, including multiple ab initio gene predictors and various forms of sequence homologies. In brief, EVM decomposes multiple gene predictions, and spliced protein and transcript alignments into a set of nonredundant gene structure components: exons and introns. Each exon and intron is scored based on the weight (associated numerical value) and abundance of the supporting evidence; genomic regions corresponding to predicted intergenic locations are also scored accordingly. The exon and introns are used to form a graph, and highest scoring path through the graph is used to create a set of gene structures and corresponding intergenic regions (Figure 5; see Materials and methods, below, for complete details). Because of the scoring system employed by EVM, gene structures with minor differences, such as small variations at intron boundaries, can yield vastly different scores. For example, a cDNA-supported intron that is only three nucleotides offset from an ab initio predicted intron could be scored extraordinarly high as compared with the predicted intron, although they differ only slightly in content. Likewise, an intron that is fully supported by multiple spliced protein alignments will be scored higher than an alternate intron of similar length yielded by only a single similarly weighted protein alignment. In this way, EVM uses the abundance and weight of the various evidence to score gene structure components appropriately to promote their selection within the resulting weighted consensus genome annotation. Figure 5 Consensus Gene Structure Prediction by EVM. The main aspects of the EVidenceModeler (EVM) weighted consensus prediction generating algorithm are depicted here, exemplified with a 7 kilobase region of the rice genome. The top view illustrates a genome browser-style view, showing the ab initio gene predictions GlimmerHMM, Fgenesh, and GeneMark.hmm, AAT-gap2 spliced alignments of other plant expressed sequence tags (ESTs), Program to Assemble Spliced Alignments (PASA) assemblies of rice EST and full-length cDNA (FL-cDNA) alignments, AAT-nap spliced alignments of nonrice proteins, and GeneWise protein homology-based predictions. Top strand and bottom strand evidence are separated by the sequence ticker. Evidence is dismantled into candidate introns and exons; candidate exons are shown in the context of the six possible reading frames at the figure bottom. A coding, intron, and intergenic score vector are shown; feature-specific scores (see Materials and methods) were added to corresponding vectors here for illustration purposes only, and note that all introns have feature-specific scores. The selection of exons, introns, and intergenic regions that define the highest scoring path is shown by the connections between exon features within the six-frame feature partition. This highest scoring path yields two complete gene structures, shown as an EVM tier at top, corresponding to the known rice genes (left) LOC_Os03g15860 (peroxisomal membrane carrier protein) and (right) LOC_Os03g15870 (50S ribosomal protein L4, chloroplast precursor). To demonstrate the simplest application of EVM, we combine only the three ab initio gene predictions and weight each prediction type equally. Figures 1 and 2 display the results in comparison with the ab initio prediction accuracies; we demonstrate that, by incorporating shared exons and introns into consensus gene structures, complete gene prediction accuracy is improved by at least 10%. Exon prediction accuracy is increased by about 6%, and exon prediction accuracies for each exon type are mostly improved, with the exception of the initial exon type, for which GeneMark.hmm alone is slightly superior. Consensus gene prediction accuracy using varied evidence types and associated weights A gene structure consensus as computed by EVM is based on the types of evidence available and their corresponding weight values. In the example above, each evidence type provided in the form of ab initio gene predictions was weighted identically. In the case where each prediction type is equivalent in accuracy, this may be sufficient, but when an evidence type(s) is more accurate, a higher weight(s) applied to that evidence is expected to drive the consensus toward higher prediction accuracy. Figure 6 illustrates the impact of varied weight combinations and sources of evidence on exon and complete gene structure prediction sensitivity. In the first set (iterations 1 to 10), only the three ab initio gene predictions are combined using random weightings. Prediction accuracy ranges from 22% to 38% gSn and 77% to 84% eSn. In the second set (iterations 11 to 20), sequence homologies are additionally included in the form of spliced protein alignments (using nap of AAT), spliced alignments of ESTs derived from other plants (using gap2 of AAT), and GeneWise protein-homology-based gene predictions. There, complete prediction accuracy ranges from 44% to 62% gSn and 88% to 92% eSn. In the third and final set (iterations 21 to 30), PASA alignment assemblies derived from rice transcript alignments were included, from which a subset define the correct gene structure. In the presence of our best evidence and randomly set weights, prediction accuracy ranges from 75% to 96% gSn and 95% to 99% eSn. Figure 6 Response of EVM prediction accuracy to varied evidence types and weights. Iterations (30) of randomly weighted evidence types were evaluated by EVidenceModeler (EVM). Iterations 1 to 10 included only the ab initio predictors GlimmerHMM, Fgenesh, and GeneMark.hmm. Iterations 11 to 20 additionally included AAT-nap alignments of nonrice proteins, GeneWise predictions based on nonrice protein homologies, and AAT-gap2 alignments of other plant expressed sequence tags. Iterations 21 to 30 included Program to Assemble Spliced Alignments (PASA) alignment assemblies and corresponding supplement of PASA long-open reading frame (ORF)-based terminal exons. Exon and complete gene prediction sensitivity values resulting from EVM using the corresponding weight combinations are plotted below. Although this represents just a minute number of possible random weight combinations, it demonstrates the effect of the weight settings and the inclusion of different evidence types on our consensus prediction accuracy. By including evidence based on sequence homology, our prediction accuracy improves greatly, doubling to tripling complete gene prediction accuracy of ab initio programs alone or in combination. Also, very different weight settings can still lead to similar levels of performance, particularly in the presence of sequence homology data. EVM consensus prediction accuracy using trained evidence weights Given the variability in consensus gene prediction accuracy observed using different combinations of weight values, finding the single combination of weights that provides the best consensus prediction accuracy is an important goal. Searching all possible weight combinations to find the single best scoring combination is not tractable, given the computational effort needed to explore such a vast search space. To estimate a set of high scoring weights, we employed a set of heuristics that use random weight combinations followed by gradient ascent (see Materials and methods, below). For the purpose of choosing high performing weights and evaluating their accuracy, we selected 1,000 of our cDNA-verified gene structures and used half for estimating weights and the other half for evaluating accuracy using these weights (henceforth termed 'trained weights'). In both the training and evaluation process, accuracy statistics were limited to each reference gene and flanking 500 base pairs (bp). However, EVM was applied to regions of the rice genome including the 30 kilobase (kb) region flanking each reference gene, to emulate gene prediction by EVM in a larger genomic context. Because the training of EVM is not deterministic, and each attempt at training can result in a different set of high-scoring weights, we performed the process of training and evaluating EVM on the rice datasets three times separately. The trained weight values computed by each training process are provided in Additional data file 2 (Table S1), and the consensus gene prediction accuracy yielded during each evaluation is provided in Additional data file 2 (Table S2). The average gene prediction accuracy is provided in Figure 7. On this set of 500 reference genes, the average exon and complete gene prediction accuracies for the ab initio predictors are similar to those computed earlier for the larger complete set of 1,058 cDNA-verified genes. EVM applied to the ab initio predictions alone using optimized weights yielded 38% gSn and 34% gSp, approximately 10% better than the best corresponding ab initio accuracy. By including the additional evidence types in the form of protein or EST homologies independently, complete gene prediction sensitivity increases to 49% to 56% gSn and 44% to 50% gSp. Using all evidence minus the PASA data, complete gene sensitivity reaches 62% gSn and 56% gSp. Note that each gain in sensitivity is accompanied by a gain in specificity, indicating overall improvements in gene prediction accuracy. Figure 7 Rice consensus gene prediction accuracy using optimized evidence weights. Gene prediction accuracy for EVidenceModeler (EVM) was calculated at the nucleotide, exon, and complete gene level using trained weights and specific sets of evidence, applied to 500 of the reference rice gene structures. The evidence evaluated is described as follows: EVM:GF includes ab initio gene predictions (GF) alone; EVM:GF+gap2 includes GF plus the AAT-gap2 alignments of other plant expressed sequence tags (gap2); EVM:GF+nap includes GF plus AAT-nap alignments of nonrice proteins (nap); EVM:GF+GeneWise includes GF plus the GeneWise predictions based on nonrice protein homologies (GeneWise); EVM:ALL(-PASA) includes GF, nap, gap2, and GeneWise; EVM:ALL(+PASA) additionally includes the Program to Assemble Spliced Alignments (PASA) alignment assemblies and PASA long-open reading frame (ORF)-based terminal exon supplement. Sn, sensitivity; Sp, specificity. Intuitive versus trained weights Although we can computationally address the problem of finding a set of weights that yield optimal performance, it is clear from our analysis of randomly selected weights that there could be numerous weight combinations that provide reasonable accuracy. In general, we find that combinations of assigned weightings in the following form provides adequate consensus prediction accuracy: (ab initio predictions) ≤ (protein alignments, EST alignments) 1 megabase), the data are partitioned into overlapping segments, and the EVM predictions from the separate partitions are subsequently joined into a single nonredundant set of predictions. Dismantling predictions and alignments into exons and introns Exons of eukaryotic gene structures are commonly treated as four distinct types: initial exon, including the start codon to a donor splice junction; internal exon, including an acceptor splice junction to a donor splice junction; terminal exon, including the acceptor splice junction to the stop codon; and the single exon, which corresponds to an intronless gene from start codon to stop codon. These are the four types of exons considered by EVM. The ab initio gene predictions provided as inputs to EVM are dismantled into their component exons and introns and added to a nonredundant corresponding exon or intron feature set. Each exon of a given type is stored by EVM with its coordinates, the codon position of its leading base, and a list of all evidence types that perfectly support it. Introns are likewise stored as discrete features based on unique coordinate pairs and their supporting evidence. Only the consensus GT or GC donor and AG acceptor dinucleotide splice sites are treated as valid by EVM; the more rare AT-AC consensus introns, although accepted by PASA, are currently disallowed by EVM. No maximum intron length is enforced by EVM, but a minimum intron length of 20 bp is set and can be tuned as required. Protein and transcript spliced alignment inputs to EVM, by default, are only capable of contributing internal exons and introns to EVM's feature set. Spliced alignments contribute internal exons to the feature set for those internal alignment segments that have consensus splice sites and encode an ORF in at least one of the three reading frames. An internal exon is added to the feature set for each incident codon position that provides an ORF on that strand. A final way for alignment data to contribute initial, terminal, or single exons to the feature set is by explicitly providing such candidate exons to EVM a priori. This is one mechanism that allows EVM to better exploit gene structures provided by PASA. PASA includes functions to provide the longest ORF within each PASA assembly, and EVM includes a utility that extracts initial, terminal, and single exons from gene structures corresponding to the longest ORF within each PASA assembly. This list of PASA-based exon candidates can be provided directly to EVM. Internal exons provided by PASA alignment assemblies are included in the feature set exactly as other forms of spliced alignment data described above. Experiments performed on the rice genome utilizing PASA evidence as input instead included the structure of the longest ORF (minimum length of 50 amino acids) within each PASA alignment assembly in place of the alignment assemblies themselves supplemented with the terminal exon candidates, as described above. These PASA longest ORF structures were provided to EVM as an OTHER_PREDICTION evidence class. Utilization of the PASA data in this way was necessary to allow provision of identical PASA-based evidence to the alternative annotation tools Glean and JIGSAW as part of the rice combiner accuracy comparison. Scoring genome features The candidate unique exon, intron, and intergenic region feature types derive their score from either a feature-specific score and/or a corresponding feature type scoring vector, as described below. Each type of evidence provided to EVM is specified as having a numerical weight value and belonging to one of the four allowable classes: PROTEIN, TRANSCRIPT, ABINITIO_PREDICTION, or OTHER_PREDICTION. Table 1 indicates the scoring mechanism for each feature type and classification. Primary differences between these four classes of evidence are that the PROTEIN and TRANSCRIPT classes are not expected to encode complete gene structures from start to stop codon, but instead contribute components of gene structures such as internal exons and, in the case of the PROTEIN class, an indication of coding nucleotides. Complete gene predictions are partitioned into the classes ABINITIO_PREDICTION and OTHER_PREDICTION, where the ABINITIO_PREDICTION class predicts noncoding intergenic regions (GeneMark.hmm) and OTHER_PREDICTION allows for the inclusion of high-specificity forms of complete predictions that are not intended to delineate the noncoding intergenic regions (KnownGene). Table 1 EVM scoring mechanism based on feature class and type Class Type Scoring vector Feature-specific score ABINITIO_PREDICTION Exon X ABINITIO_PREDICTION Intron X ABINITIO_PREDICTION Intergenic X TRANSCRIPT Exon X TRANSCRIPT Intron X PROTEIN Exon X PROTEIN Intron X OTHER_PREDICTION Exon X OTHER_PREDICTION Intron X EVM, EVidenceModeler. A feature type scoring vector contains a numerical value for each nucleotide across the genome sequence. Evidence that contributes to a feature type scoring vector contributes its corresponding weight value to each nucleotide within the span of its feature coordinates. Evidence that contributes a feature-specific score instead contributes a value of its (weight × feature_length) to that unique feature that it supports, in this case either that complete intron or exon. Exons derive their scores from a combination of feature-specific scores and a corresponding scoring vector. In this case, the feature-specific scores are summed with the values in the corresponding scoring vector for each nucleotide position within its span. For example, a complete feature with coordinates a to b would be scored like so: S c o r e ( a , b ) = ∑ a < = i < = b S c o r i n g V e c t o r [ i ] ​ ​ + ∑ e v i d e n c e _ e n d 5 ' = a e v i d e n c e _ e n d 3 ' = b f e a t u r e L e n g t h * w e i g h t ( e v i d e n c e ) As each gene prediction or spliced alignment is dismantled into its component parts, the parts contribute the weight of that evidence to the scoring scheme. For example, a single spliced protein alignment is dismantled into the protein alignment segments and intervening gaps, possibly contributing to feature types exon and intron of feature class PROTEIN. Those 'perfect' complete introns and exons yielded by dismantling of this protein alignment chain are added to the candidate exon and intron feature set if those features do not already exist. Each protein alignment segment contributes its corresponding evidence weight to each overlapping nucleotide position in the exon feature type scoring vector. Those protein alignment gaps that correspond to complete introns in our feature set contribute a value of (weight × length) to the feature-specific score of each corresponding intron. The abundance of evidence is reflected in both the feature-specific and vectored scores. For example, often many protein homologies will exist at a given locus. Each protein database match (accession) at a given locus is scored separately, and so exon and introns supported by vast quantities of evidence will have scores that reflect both the weight and abundance of that evidence. For the purpose of scoring exons and introns and minimizing the memory requirements required for storing the scoring vectors, each strand and associated set of evidence is initially examined separately; note that our final gene prediction examines both strands simultaneously. During the initial strand-based analysis, distinct exons and introns are collected from the evidence restricted to the strand being analyzed and scored accordingly. After collecting properly scored gene structure components from each strand, they are grouped together as a single collection of features from both DNA strands. Dynamic programming is used to find the highest scoring set of connected exons, introns, and intergenic regions across the entire genome sequence (see Figure 5). Unlike exon and intron features, the intergenic features are not precomputed and are instead scored during the dynamic programming stage; scores for intergenic regions are computed when attempting to connect candidate gene termini while building the directed acyclic graph of connectable feature components (also referred to as the feature trellis). The highest scoring path of connected features is extracted from the feature trellis and separated into the individual gene predictions. A primary restriction within our feature trellis is that the introns connecting exons must exist as explicit components of our feature set; EVM will not connect two otherwise compatible exons unless the required intron exists within the inputted evidence, such as provided by a gene prediction, or spliced protein or transcript alignment. Note that, by default, EVM will re-examine long introns to identify candidate nested genes. Although we find this functionality extraordinarily useful for automated annotation, especially for insect genomes, this function was not employed in any analysis described here. Although improvements in sensitivity can result from the nested gene search, there are associated costs in specificity (data not shown). Augmenting intergenic scores from approximate beginnings and ends of genes Because the ABINITIO_PREDICTION class of evidence is the only class that contributes explicitly to the prediction of intergenic regions, coping with cases in which the consensus of ab initio predictions merges multiple adjacent genes into a single gene structure is particularly problematic. To split the merged consensus into separate individual predictions, the true intergenic region would need a score that is suitable to offset the alternative, typically involving a predicted intron that joins what should be distinct loci. To encourage the selection of separate complete gene structures supported by protein homologies instead of the merged gene, EVM augments the scores of intergenic regions supported indirectly by protein evidence, as elaborated below. The approximate boundaries of candidate intergenic regions supported by protein homologies are localized by examining the boundaries of protein alignment chains. The beginnings and ends of all PROTEIN evidence structures (the far bounds of all spliced alignment chains, not the individual segments) are tallied. A sliding window of 300 nucleotides is applied to each strand, and all peaks of beginnings and ends are separately tallied. In addition to the protein alignment chains, the terminal exons provided by the extraction of long ORFs from PASA alignment assemblies also contribute to the tally of candidate beginnings and ends of genes. From each begin peak, a corresponding initial exon is located from the feature set. The intergenic score for each nucleotide from the candidate initial exon upstream to the preceding gene is set to the maximal intergenic score, corresponding to the sum of the weights for ABINITIO_PREDICTION evidence classes. Likewise, from each candidate gene end, a terminal exon is located from the feature set, and the genome region downstream to the next gene is set to the maximal intergenic score. Note that single exon genes are also treated similarly as initial or terminal exons in the search for the next possible adjacent gene structure. Although this search for gene boundaries is not very precise, the heuristic employed here tends to work acceptably well in practice. Choosing the proper boundaries of a gene structure is critical for predicting the entire gene correctly, as demonstrated by the greater variability in initial and terminal exon prediction among the various ab initio gene prediction programs. Filtering EVM predictions with low support Instead of reporting the single best scoring gene structure at each locus, EVM reports the set of gene structures that, when connected together with the intervening intergenic regions, provides an optimal cumulative score. There are sometimes cases in which low scoring adventitious genes are included in the preliminary EVM gene set, largely as a consequence of ABINITIO_PREDICTION introns called on either strand in what are really intergenic regions. To remove these adventitious genes from the EVM gene set, the score of each EVM prediction is re-examined in the context of ab initio predicted introns being scored as if they were intergenic regions. An alternative noncoding score is computed for each EVM gene prediction by summing the predicted intergenic regions with the ab initio predicted intron regions. This noncoding score is then compared with the initial EVM prediction score, and those EVM predictions with a coding/noncoding score ratio below 0.75 are eliminated. An example of a low scoring EVM prediction removed during this postprocessing stage is illustrated in Additional data file 1 (Figure S5). An option is available in the EVM software to report these eliminated genes. In those cases in which all predictions agree, predictions lack introns, and the corresponding intergenic score is zero, the score ratio is set to an arbitrary high value and reported accordingly. Evaluating prediction accuracy Gene prediction accuracy (sensitivity and specificity) was computed at the level of nucleotides, exons, transcripts, and complete genes, as described previously [10] but with slight modifications. Although some gene structures include untranslated region annotations, only the protein-coding portions of each exon were considered when we computed accuracy. In our evaluation of the reference gene structures in rice, alternative splicing was ignored, and no attempt was made to generate a reference gene set for rice that included alternatively spliced transcripts. Therefore, given the one transcript per gene in the rice dataset, gene prediction accuracy calculations would necessarily be identical to the transcript accuracy calculations, and so only the gene prediction accuracy was reported. Although each reference gene region was provided as input to EVM in the context of the flanking 30 kb of genome sequence and corresponding evidence, all accuracy calculations were based on the gene predictions isolated from reference gene region including a flanking 500 bp. In our comparison of the accuracy of EVM to the annotation tools Glean and JIGSAW, we obtained the most current versions of the software available from their respective sites, namely version 3.2.9 for JIGSAW [55] and version 1.0.1 for GLEAN [56], downloaded directly from the subversion source repository. Accuracy calculations on the human ENCODE genome regions included these regions and corresponding predictions in their entirety. Given that the GENCODE annotations included alternatively spliced transcripts, the prediction of alternatively spliced genes was a major component of our analysis, and so transcript prediction accuracy calculations were reported along with complete gene, exon, and nucleotide prediction accuracies. Estimating optimal evidence weights The EVM training process is divided into three phases described below: Initially optimized PREDICTION weights In the first stage, optimal weights are explored for the ABINITIO_PREDICTION class in isolation from evidence of the other classes. The proper balance between the evidence weights applied to exons, introns, and intergenic regions is explored to optimize gene prediction accuracy. Weights are randomly chosen for each ab initio gene prediction type and normalized so that they sum to one. EVM is applied to each reference gene and specified length of flanking region included. EVM prediction accuracy is measured, and a conglomerate accuracy score is computed as follows: AccuracyScore = F + gSn + eSn Where F = (2 × nSn × nSp)/(nSn + nSp), Sn = TP/(TP + FN), and Sp = TP/(TP + FP). (TP, FP, FN correspond to true positives, false positives, and false negatives, respectively. The nSn and nSp indicate nucleotide sensitivity and specificity, respectively.) Twenty random trials are performed. The weight combination that yielded the greatest AccuracyScore is chosen. These weight values are gradually adjusted while applying gradient ascent to find weight values that improve performance. Initially optimized best individual evidence weights Using the combination of weights now temporarily fixed for the ABINITIO_PREDICTION evidence, each other evidence type is introduced separately to find the minimum corresponding weight that provides the greatest AccuracyScore in the context of the ABINITIO_PREDICTION types. The weight for the other evidence type is first set to zero and evaluated. Next, the weight is set to the average weight value of the ABINITIO_PREDICTION types and evaluated. Gradient ascent is performed to explore adjusted weight values and a higher scoring weight. The minimum weight value that yielded the highest AccuracyScore is initially assigned to the other evidence type. Simultaneous application of all evidence and relative weight refinements The weight values for all evidence types are adjusted to find weight combinations that demonstrate improved prediction accuracies when all evidence is examined simultaneously. Evidence types are examined in descending order of their initially set weight values computed from phase 1 (ABINITIO_PREDICTION) or phase 2 (other) above. Weight values are gradually adjusted and gradient ascent is applied to explore better performing weight value in the context of the other evidence types. Cycling through the evidence types in this manner occurs until no appreciable improvement in performance is observed, in which case the training process ceases and the final weight values are reported. Evidence weights and EVM prediction accuracies encountered during the training process using the rice data are illustrated in Additional data file 1 (Figure S6). Manual annotation of gene structures The genome sequence, ab initio gene predictions, protein alignments, GeneWise predictions, and other plant EST alignments were examined using the Neomorphic/Affymetrix Annotation Station software (described by Haas and coworkers [28]). No rice transcript alignments either alone or in the context of PASA assemblies were made available to users so that we could reasonably estimate optimal gene structure annotation accuracy in the context of ab initio gene predictions and homologies to sequences derived from other organisms. A group of annotators were provided with the same data sets evaluated by EVM, only in graphical form. Annotators were instructed to model a gene structure in the targeted region that best reflected the available evidence using the Annotation Station software. Annotators were not allowed to examine the data deeper than the visual display provided. The sequence alignments themselves were not available except in the context of the glyphs highlighting their end points, and no additional sequence analyses such as running blast was allowed. The focus of this effort was not to measure the maximal accuracy of manual gene annotation accuracy in general, but only to measure the maximal possible accuracy of an automated annotation such as EVM given the restricted inputs. Abbreviations bp = base pairs; EGASP = ENCODE Genome Annotation Assessment Project; ENCODE = ENCyclopedia of DNA Elements; eSn = exon sensitivity; eSp = exon specificity; EST = expressed sequence tag; EVM = EVidenceModeler; FL-cDNA = full-length cDNA; FN = false negatives; gSn = gene sensitivity; gSp = gene specificity; HMM = hidden Markov model; kb = kilobase; nSn = nucleotide sensitivity; nSp = nucleotide specificity; ORF = open reading frame; PASA = Program to Assemble Spliced Alignments; TN = true negatives; TP = true positives. Authors' contributions BJH carried out all analyses, software development, and wrote the initial version of the manuscript while under the guidance of JW, OW, and SLS. SLS, MP, and JA helped to develop many of the underlying concepts of EVM. Analyses using the rice genome data were assisted by WZ and CRB. JO was responsible for generating all evidence for the rice and human genome sequences. All authors contributed to and approved the final version of the manuscript. Additional data files The following additional data are available with the online version of this paper. Additional data file 1 includes the supplementary figures described throughout the report. Additional data file 2 contains supplementary data tables. Supplementary Material Additional file 1 Supplementary figures. (Figure S1 shows the difference in rice gene prediction accuracy between using trained and intuitively set evidence weights. Figure S2 shows the change in human gene prediction accuracy due to application of PASA. Figure S3 shows the comparison of 1,058 reference gene structure exon distribution to all rice gene annotations. Figure S4 shows the gene prediction accuracies for EGASP gene sets. Figure S5 shows filtering EVM predictions with low support. Figure S6 shows optimization of evidence weights by exploring weight and evidence combinations. Click here for file Additional file 2 Supplementary data tables. Table S1 provides trained weights for evidence based on evaluating 500 rice gene structures. Table S2 shows the gene prediction accuracy for EVM measured using 500 reference rice gene structures. Table S3 provides trained EVM weights including PASA. Table S4 provides trained EVM evidence weights for the ENCODE regions. Table S5 shows the EVM prediction accuracy using trained evidence weights for ENCODE regions. Click here for file
                Bookmark

                Author and article information

                Contributors
                Journal
                Front Genet
                Front Genet
                Front. Genet.
                Frontiers in Genetics
                Frontiers Media S.A.
                1664-8021
                21 December 2021
                2021
                : 12
                : 811798
                Affiliations
                [1] 1 State Key Laboratory of Marine Environmental Science, College of Ocean and Earth Sciences, Xiamen University , Xiamen, China
                [2] 2 State Key Laboratory of Large Yellow Croaker Breeding, Ningde Fufa Fisheries Company Limited , Ningde, China
                [3] 3 Fujian Key Laboratory of Genetics and Breeding of Marine Organisms, College of Ocean and Earth Sciences, Xiamen University , Xiamen, China
                Author notes

                Edited by: Roger Huerlimann, Okinawa Institute of Science and Technology Graduate University, Japan

                Reviewed by: Tianxiang Gao, Zhejiang Ocean University, China

                Syed Farhan Ahmad, Kasetsart University, Thailand

                *Correspondence: Peng Xu, xupeng77@ 123456xmu.edu.cn
                [ † ]

                These authors have contributed equally to this work

                This article was submitted to Livestock Genomics, a section of the journal Frontiers in Genetics

                Article
                811798
                10.3389/fgene.2021.811798
                8724560
                2da81fd4-1320-4a55-a209-eaf773591002
                Copyright © 2021 Bai, Gong, Zhou, Li, Zhao, Ke, Zou, Pu, Wu, Zheng, Zhou and Xu.

                This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

                History
                : 09 November 2021
                : 29 November 2021
                Categories
                Genetics
                Data Report

                Genetics
                oplegnathus fasciatus,genomic resources,pacbio,hi-c,phylogenetic analysis
                Genetics
                oplegnathus fasciatus, genomic resources, pacbio, hi-c, phylogenetic analysis

                Comments

                Comment on this article