Open access peer-reviewed chapter

Determination of Nucleopolyhedrovirus’ Taxonomic Position

Written By

Yu-Shin Nai, Yu-Feng Huang, Tzu-Han Chen, Kuo-Ping Chiu and Chung-Hsiung Wang

Submitted: 19 May 2016 Reviewed: 28 October 2016 Published: 05 April 2017

DOI: 10.5772/66634

From the Edited Volume

Biological Control of Pest and Vector Insects

Edited by Vonnie D.C. Shields

Chapter metrics overview

1,741 Chapter Downloads

View Full Metrics

Abstract

To date , over 78 genomes of nucleopolyhedroviruses (NPVs) have been sequenced and deposited in NCBI. How to define a new virus from the infected larvae in the field is usually the first question. Two NPV strains, which were isolated from casuarina moth (L. xylina) and golden birdwing larvae (Troides aeacus), respectively, displayed the same question. Due to the identity of polyhedrin (polh) sequences of these two isolates to that of Lymantria dispar MNPV and Bombyx mori NPV, they are named LdMNPV-like virus and TraeNPV, provisionally. To further clarify the relationships of LdMNPV-like virus and TraeNPV to closely related NPVs, Kimura 2-parameter (K-2-P) analysis was performed. Apparently, the results of K-2-P analysis that showed LdMNPV-like virus is an LdMNPV isolate, while TraeNPV had an ambiguous relationship to BmNPV. Otherwise, MaviNPV, which is a mini-AcMNPV, also exhibited a different story by K-2-P analysis. Since K-2-P analysis could not cover all species determination issues, therefore, TraeNPV needs to be sequenced for defining its taxonomic position. For this purpose, different genomic sequencing technologies and bioinformatic analysis approaches will be discussed. We anticipated that these applications will help to exam nucleotide information of unknown species and give an insight and facilitate to this issue.

Keywords

  • nucleopolyhedroviruses
  • Kimura-2-parameter analysis
  • next-generation sequencing
  • bioinformatic analysis

1. Introduction

Baculoviruses are insect-specific viruses which have a large circular double-stranded DNA genome packaged in enveloped, rod-shaped nucleocapsid and occluded within a paracrystalline protein occlusion body (OB) [1, 2]. The family Baculoviridae has four genera, including Alphabaculovirus, Betabaculovirus, Gammabaculovirus and Deltabaculovirus. Nucleopolyhedrovirus (NPV) is a member of Alphabaculovirus (lepidopteran-specific NPV) [3]; NPV replicates in the nucleus of the infected host cell and causes a disease of nuclear polyhedrosis. Epidemic outbreak of NPV may play a role in regulation of the host nature population [4]. Thereby, it is a potential agent for biological control with a number of eco-friendly benefits including high virulence and specificity against target insects, environmental safety and sustainable existence with target insects. Several baculoviruses showing promising results have been commercialized as biopesticides for the control of insect pests around the world [5]. For biotechnological applications, baculoviruses have been constructed as a eukaryotic protein expression vectors (baculovirus expression vector system (BEVS)) over the last 30 years and used to gene therapy trials. So far, many recombinant proteins have been expressed in insect cells by BEVS and contribute to human life [6].

To date, baculoviruses are known to infect more than 660 insect species; most of them are belonging to the order of Lepidoptera, Diptera and Hymenoptera [7, 8]. Baculoviruses exhibit genetic variations among species and its isolates [9]. Although a large number of baculoviruses in the nature, only a few have been well studied. To the best of our knowledge, a total of 78 fully sequenced genomes have been deposited in GenBank [10] and also several baculoviruses of whole genomes may soon be sequenced and deposited ( Table 1 ). However, these published viral genomes represent only a small fraction and the genetic relationship among nucleopolyhedroviruses (NPVs) in the natural environment remains a puzzle.

Genus Virus Virus Abbreviation GenBank accession Genome size (bp) A C G T GC content ORFs Sequencing Assembler Reference
Alphabaculovirus (Group I) Anticarsia gemmatalis MNPV AgMNPV-2D NC_008520 132,239 36,623 29,338 29,513 36,765 44.5% 158 Sanger PHRED/ALIGNER [11]
AgMNPV-26 KR815455 131,678 36,411 29,288 29,405 36,574 44.6% 157 Roche 454 GS FLX Geneious [12]
AgMNPV-27 KR815456 131,172 36,273 29,176 29,331 36,392 44.6% 157
AgMNPV-28 KR815457 130,745 36,185 29,018 29,242 36,300 44.6% 157
AgMNPV-29 KR815458 130,506 36,072 28,989 29,216 36,229 44.6% 157
AgMNPV-30 KR815459 130,741 36,195 29,011 29,173 36,362 44.5% 156
AgMNPV-31 KR815460 132,126 36,543 29,363 29,564 36,656 44.6% 158
AgMNPV-32 KR815461 131,494 36,341 29,234 29,384 36,535 44.6% 157
AgMNPV-33 KR815462 131,059 36,322 29,114 29,244 36,379 44.5% 157
AgMNPV-34 KR815463 131,543 36,435 29,233 29,383 36,492 44.6% 158
AgMNPV-35 KR815464 132,176 36,552 29,384 29,558 36,682 44.6% 159
AgMNPV-36 KR815465 131,216 36,293 29,127 29,270 36,526 44.5% 156
AgMNPV-37 KR815466 131,855 36,531 29,255 29,400 36,669 44.5% 156
AgMNPV-38 KR815467 130,740 36,194 29,012 29,172 36,362 44.5% 156
AgMNPV-39 KR815468 130,698 36,219 29,026 29,184 36,269 44.5% 157
AgMNPV-40 KR815469 132,180 36,542 29,409 29,583 36,646 44.6% 158
AgMNPV-42 KR815470 130,949 36,274 29,098 29,275 36,302 44.6% 157
AgMNPV-43 KR815471 132,077 36,539 29,369 29,528 36,641 44.6% 159
Antheraea pernyi NPV AnpeNPV NC_008035 126,629 29,513 34,041 33,664 29,406 53.5% 147 Sanger ContigExpress9.1.0 + SeqMan5.0/DNASTAR [13]
Autographa californica MNPV AcMNPV NC_001623 133,894 39,195 27,151 27,347 40,201 40.7% 156 Sanger GCG package [14]
Autographa californica MNPV-WP10 AcMNPV-WP10 KM609482 133,926 39,205 27,157 27,346 40,199 40.7% 151 Illumina HiSeq 2000 Newbler [15]
Bombyx mandarina NPV BomaNPV NC_012672 126,770 37,358 25,398 25,601 38,413 40.2% 141 Solexa GA GENETYX-win Software + DNASTAR [16]
Bombyx mori NPV BmNPV NC_001962 128,413 37,747 25,828 26,056 38,782 40.4% 143 Sanger DNASIS/PROSIS [17]
Catopsilia pomona NPV CapoNPV KU565883 128,058 38,938 25,348 25,444 38,328 39.7% 131 Roche 454 GS FLX+ GS de novo assembler [10]
Choristoneura fumiferana DEF MNPV CfDEFMNPV NC_005137 131,160 35,474 30,110 29,993 35,580 45.8% 149 Sanger MacVector + Lasergene/DNASTAR [18]
Choristoneura fumiferana MNPV CfMNPV NC_004778 129,593 32,224 32656 32,261 32,452 50.1% 146 Sanger Gene Runner [19]
Choristoneura murinana NPV ChmuNPV NC_023177 124,688 31,408 30,986 31,370 30,924 50.0% 147 Roche 454 CLC Genomics Workbench [20]
Choristoneura occidentalis NPV ChocNPV NC_021925 128,446 32,108 31,905 32,481 31,952 50.1% 148 Roche 454 GS FLX SeqMan Pro Lasergene/DNASTAR [21]
Choristoneura rosaceana NPV ChroNPV NC_021924 129,052 33,309 31,261 31,425 33,057 48.6% 149
Condylorrhiza vestigialis MNPV CoveMNPV NC_026430 125,767 35,904 26,937 27,038 35,886 42.9% 138 Roche 454 Geneious + MIRA [22]
Dasychira pudibunda NPV DapuNPV KP747440 136,761 31,022 37,008 37,454 31,277 54.4% 161 Illumina MiSeq Geneious [23]
Ectropis obliqua NPV EcobNPV NC_008586 131,204 40,683 24,676 24,708 41,137 37.6% 126 Sanger Genetyx-win [24]
Hyphantria cunea NPV HycuNPV NC_007767 132,959 36,031 30,039 30,465 36,424 45.5% 148 RISA-384 DNASIS [25]
Lonomia obliqua MNPV LoobMNPV KP763670 120,023 38,995 20,932 21,966 38,104 35.7% 134 Roche 454 GS FLX Geneious [26]
Maruca vitrata MNPV MaviMNPV NC_008725 111,953 34,041 21,669 21,563 34,680 38.6% 126 Sanger PHRED/PHRAP [27]
Orgyia pseudotsugata MNPV OpMNPV NC_001875 131,995 29,463 36,477 36,295 29,758 55.1% 152 Sanger GCG package [28]
Philosamia cynthia ricini NPV PhcyNPV JX404026 125,376 28,966 33,461 33,809 29,140 53.7% 138 Sanger N/A1 [29]
Plutella xylostella MNPV PlxyMNPV NC_008349 134,417 39,437 27,303 27,396 40,281 40.7% 152 Sanger Lasergene/DNASTAR [30]
Rachiplusia ou MNPV RoMNPV NC_004323 131,526 39,674 25,630 25,793 40,429 39.1% 149 Sanger Wisconsin package + Lasergene/DNASTAR [31]
Thysanoplusia orichalcea NPV ThorNPV NC_019945 132,978 40,022 26,388 26,142 40,426 39.5% 145 Solexa GA Edena [32]
Alphabaculovirus (Group II) Adoxophyes honmai NPV AdhoNPV NC_004690 113,220 36,505 20,025 20,328 36,362 35.6% 125 RISA-384 PHRED/PHRAP [33]
Adoxophyes orana NPV AdorNPV NC_011423 111,724 36,306 19,404 19,694 36,320 35.0% 121 Sanger SeqMan II Lasergene/DNASTAR [34]
Agrotis ipsilon MNPV AgipMNPV NC_011345 155,122 40,201 37,490 7,860 39,571 48.6% 163 Sanger Lasergene/DNASTAR [35]
Agrotis segetum NPV AgseNPV NC_007921 147,544 40,237 33,200 34,247 39,860 45.7% 153 Sanger Gap4 [36]
Agrotis segetum NPV B AgseNPV-B NC_025960 148,981 40,490 33,698 4,371 40,422 45.7% 150 Roche 454 DNASTAR [37]
Apocheima cinerarium NPV ApciNPV NC_018504 123,876 41,223 20,865 20,449 41,332 33.4% 117 Sanger SeqMan Pro Lasergene/DNASTAR unpublished
Buzura suppressaria NPV BusuNPV NC_023442 120,420 37,568 22,152 22,142 38,558 36.8% 127 Roche 454 GS FLX GS de novo assembler [38]
Chrysodeixis chalcites NPV ChchNPV NC_007151 149,622 45,151 29,304 29,060 46,107 39.0% 151 Sanger Gap4 [39]
Chrysodeixis chalcites SNPV ChchSNPV-TF1-A JX535500 149,684 45,090 29,324 29,133 46,137 39.1% 150 Roche 454 Newbler [40]
ChchSNPV-TF1-C JX560539 150,079 45,146 29,384 29,096 46,447 39.0% 150
ChchSNPV-TF1-B JX560540 149,080 44,989 29,152 28,987 45,952 39.0% 150
ChchSNPV-TF1-G JX560541 149,039 45,075 29,136 28,869 45,958 38.9% 151
ChchSNPV-TF1-H JX560542 149,624 45,162 29,285 29,034 46,143 39.0% 150
Clanis bilineata NPV ClbiNPV NC_008293 135,454 41,557 25,560 25,558 42,779 37.7% 129 Sanger N/A [41]
Epiphyas postvittana NPV EppoNPV NC_003083 118,584 35,221 24,287 23,956 35,120 40.7% 136 Sanger DNASTAR [42]
Euproctis pseudoconspersa NPV EupsNPV NC_012639 141,291 41,736 28,455 28,549 42,551 40.3% 139 Sanger Wisconsin package + GENETYX-win [43]
Helicoverpa armigera SNPV AC53 HaSNPV-AC53 NC_024688 130,442 39,121 25,389 25,606 40,326 39.1% 138 Ion Torrent PGM CLC Genomics Workbench [44]
Helicoverpa armigera MNPV HearMNPV NC_011615 154,196 46,371 30,731 31,060 46,031 40.1% 162 Sanger SeqMan 5.0/DNASTAR [45]
Helicoverpa armigera NPV HearNPV NC_003094 130,759 39,345 25,340 25,552 40,522 38.9% 137 Sanger Wisconsin package + Lasergene/DNASTAR [46,47]
Helicoverpa armigera NPV G4 HearNPV-G4 NC_002654 131,405 39,529 25,530 25,738 40,608 39.0% 135 Sanger PHRED/PHRAP [48]
Helicoverpa armigera NPV NNg1 HearNPV-NNg1 NC_011354 132,425 39,754 25,791 26,054 40,826 39.2% 143 RISA-384 DNASIS [49]
Helicoverpa zea SNPV HzSNPV NC_003349 130,869 39,273 25,471 25,675 40,450 39.1% 139 Sanger Wisconsin package + Lasergene/DNASTAR [50]
Hemileuca sp. NPV HespNPV NC_021923 140,633 42,827 26,977 26,595 44,234 38.1% 137 Sanger Wisconsin package + Lasergene/DNASTAR [51]
Lambdina fiscellaria NPV LafiNPV NC_026922 157,977 45,363 34,616 34,350 43,648 43.7% 137 Roche 454 CLC Genomics Workbench [52]
Leucania separata NPV LeseNPV NC_008348 168,041 42,546 40,683 40,927 43,885 48.6% 169 MegaBACE1000 DNASTAR [53]
Lymantria dispar MNPV LdMNPV NC_001973 161,046 34,229 46,226 46,331 34,260 57.5% 164 Sanger GCG package [54]
Lymantria dispar MNPV-27 LdMNPV-27 KP027546 164,158 35,020 47,133 47,118 34,887 57.4% 162 Illumina MiSeq CLC Genomics Workbench [55]
Lymantria dispar MNPV-BNP LdMNPV-BNP KU377538 157,270 38,788 39,579 39,567 39,336 50.3% 154 Illumina MiSeq Geneious [56]
Lymantria dispar MNPV-2161 LdMNPV-2161 KF695050 163,138 34,855 46,648 46,812 34,823 57.3% 174 Roche 454 GS Junior SeqMan NGEN Lasergene/DNASTAR [9]
Lymantria dispar MNPV-3029 LdMNPV-3029 KM386655 161,712 34,321 46,434 46,457 34,500 57.4% 163 Roche 454 Lasergene/DNASTAR [57]
Lymantria dispar MNPV-45 LdMNPV-45 KU862282 161,006 34,234 46,192 46,314 34,264 57.5% 155 Illumina CLC Genomics Workbench [58]
Lymantria dispar MNPV-3054 LdMNPV-3054 KT626570 164,478 35,151 47,119 47,140 35,068 57.3% 174 Roche 454 GS Junior LaserGene/DNASTAR [59]
Lymantria dispar MNPV-3041 LdMNPV-3041 KT626571 162,658 34,715 46,478 46,647 34,818 57.3% 178
Lymantria dispar MNPV-Ab-a624 LdMNPV-Ab-a624 KT626572 161,321 34,282 46,302 46,405 34,332 57.5% 176
Lymantria xylina MNPV LyxyMNPV NC_013953 156,344 36,207 41,674 41,933 36,530 53.5% 157 Sanger PHRED/PHRAP [60]
Mamestra brassicae MNPV MabrMNPV NC_023681 152,710 46,042 30,311 30,604 45,753 39.9% 159 Roche 454 GS de novo assembler [61]
Mamestra configurata NPV-A MacoNPV-A NC_003529 155,060 45,336 32,160 32,463 45,101 41.7% 169 Sanger Wisconsin package + Lasergene/DNASTAR [62]
Mamestra configurata NPV-B MacoNPV-B NC_004117 158,482 47,831 31,504 31,953 47,194 40.0% 168 Sanger Sequencher 4.0 [63]
Orgyia leucostigma NPV OrleNPV NC_010276 156,179 46,420 31,270 31,020 47,469 39.9% 135 Sanger Agencourt BioScience [64]
Peridroma NPV PespNPV NC_024625 151,109 35,060 40,593 39,822 35,633 53.2% 139 Roche 454 CLC Genomics Workbench [65]
Perigonia lusca single NPV PeluNPV NC_027923 132,831 39,968 26,167 26,362 40,256 39.6% 145 Roche 454 Geneious unpublished
Pseudoplusia includens SNPV PsinNPV NC_026268 139,132 41,843 27,452 27,210 42,609 39.3% 141 Roche 454 GS FLX MIRA [66]
Spodoptera exigua MNPV SeMNPV NC_002169 135,611 38,445 29,486 29,929 37,751 43.8% 139 Sanger Wisconsin package + Lasergene/DNASTAR [67]
Spodoptera frugiperda MNPV virus SfMNPV NC_009011 131,331 39,417 26,346 26,507 39,061 40.2% 143 Sanger Lasergene/DNASTAR [68]
Spodoptera litura MNPV SpliMNPV-AN1956 JX454574 137,998 37,469 30,803 30,846 38,880 44.7% 132 Roche 454 GS Junior LaserGene/DNASTAR [69]
Spodoptera litura NPV SpltNPV NC_003102 139,342 39,180 29,691 29904 40,567 42.8% 141 MegaBACE1000 DNASIS + DNASTAR [70]
Spodoptera litura NPV II SpltNPV-II NC_011616 148,634 40,998 33,210 33,671 40,755 45.0% 147 n/a N/A unpublished
Sucra jujuba NPV SujuNPV KJ676450 135,952 41,395 26,157 26,399 42,001 38.7% 131 Roche 454 GS de novo assembler [71]
Trichoplusia ni SNPV TnSNPV NC_007383 134,394 40,601 6,256 26,117 41,384 39.0% 145 Sanger PHRED/PHRAP [72]
Betabaculovirus Adoxophyes orana granulovirus AdorGV NC_005038 99,657 33,077 17,098 17,275 32,207 34.5% 119 Sanger SeqMan II Lasergene/DNASTAR [73]
Agrotis segetum granulovirus AgseGV NC_005839 131,680 41,892 25,179 23,953 40,656 37.3% 132 n/a n/a unpublished
Clostera anastomosis GV isolate Henan ClasGV-A NC_022646 101,818 27,115 23,832 23,739 27,132 46.7% 122 Illumina GA SOAPdenovo [74]
Clostera anastomosis granulovirus-B ClasGV-B KR091910 107,439 33,648 19,904 20,673 33,214 37.8% 123 Roche 454 GS FLX Newbler [75]
Cnaphalocrocis medinalis GV CnmeGV NC_029304 111,246 36,021 19,756 19,385 36,084 35.2% 118 Roche 454 GS FLX GS de novo assembler [76]
Cnaphalocrocis medinalis granulovirus CnmeGV KP658210 112,060 36,295 19,904 19,529 36,332 35.2% 133 PacBio RS II HGAP2.2.0 [77]
Choristoneura occidentalis GV ChocGV NC_008168 104,710 36,132 17,268 16,938 34,372 32.7% 116 Sanger PHRED/PHRAP [78]
Clostera anachoreta granulovirus ClanGV NC_015398 101,487 28,188 22,554 22,523 28,222 44.4% 123 Illumina GA SOAPdenovo [79]
Cydia pomonella granulovirus CpGV NC_002816 123,500 34,029 27,722 28,183 33,566 45.3% 143 Sanger Wisconsin package + Lasergene/DNASTAR [80]
Cryptophlebia leucotreta granulovirus CrleGV NC_005068 110,907 38,095 18,090 17,890 36,832 32.4% 128 Sanger Lasergene/DNASTAR [81]
Diatraea saccharalis granulovirus DisaGV NC_028491 98,392 32,133 17,032 17,337 31,880 34.9% 125 Roche 454 Geneious [82]
Epinotia aporema granulovirus EpapGV NC_018875 119,082 35,524 24,984 24,403 34,171 41.5% 132 Roche 454 GS FLX Newbler [83]
Erinnyis ello granulovirus ErelGV NC_025257 102,759 31,707 19,440 20,324 31,288 38.7% 130 Roche 454 GS FLX Geneious [84]
Helicoverpa armigera granulovirus HearGV NC_010240 169,794 50,336 34,518 34,810 50,130 40.8% 179 Sanger SeqMan Lasergene/DNASTAR [85]
Plodia interpunctella granulovirus PiGV KX1513952 112,536 n/a n/a n/a n/a n/a 123 Roche 454 GS Junior SeqMan NGEN Lasergene/DNASTAR [86]
Phthorimaea operculella granulovirus PhopGV NC_004062 119,217 38,306 21,127 21,431 38,353 35.7% 130 Sanger N/A [87]
Plutella xylostella granulovirus PlxyGV NC_002593 100,999 30,252 20,546 20,546 29,655 40.7% 120 DSQ-1000 L GENETYX-win [88]
Pieris rapae granulovirus PrGV NC_013797 108,592 36,619 17,863 18,168 35,942 33.2% 120 Sanger NN/A [89]
Pseudaletia unipuncta granulovirus PsunGV NC_013772 176,677 53,572 34,993 35,311 52,799 39.8% 183 n/a N/A unpublished
Spodoptera frugiperda GV isolateVG008 SpfrGV NC_026511 140,913 38,131 32,852 32,288 37,642 46.2% 146 Roche 454 GS FLX Newbler [90]
Spodoptera litura granulovirus SpliGV NC_009503 124,121 38,360 23,813 24,377 37,571 38.8% 136 Sanger N/A [91]
Xestia c-nigrum granulovirus XcGV NC_002331 178,733 53,166 36,079 36,627 52,861 40.7% 181 Sanger DNASIS/PROSIS [92]
Gammabaculovirus Neodiprion abietis NPV NeabNPV NC_008252 84,264 28,292 13,948 14,177 27,847 33.4% 93 Sanger PHRED/PHRAP [93]
Neodiprion lecontei NPV NeleNPV NC_005906 81,755 27,741 13,596 13,640 26,616 33.4% 89 Sanger SeqMan Lasergene/DNASTAR [94]
Neodiprion sertifer NPV NeseNPV NC_005905 86,462 29,158 14,444 14,745 28,115 33.8% 90 Sanger Sequencher 4.1 [95]
Deltabaculovirus Culex nigripalpus NPV CuniNPV NC_003084 108,252 26,623 27,228 27,839 26,562 50.9% 109 Sanger CAP3 [96]

Table 1.

List of sequenced baculoviruses genomes

N/A: no information is available either in the paper or GenBank file.

The GenBank file with accession number KX1513952 is not available in GenBank website.

Previously, Sanger sequencing was employed to sequence the viral genomic sequences cloned in plasmids. With the advances of sequencing technologies, next-generation sequencing (NGS) is becoming an important technology for large-scale viral genomic sequencing. The high cost of NGS and requirement of intensive bioinformatic analysis remain a hurdle for this application. In a word, NGS is an available tool to facilitate on the study of the genetic relationship of baculoviruses.

Advertisement

2. Identification of NPVs

Biochemical and biotechnology-based methods are the most common approaches employed to identify the NPVs. In most cases, more than one method is employed to compensate the pros and cons for each other. For example, restriction enzyme profiling of viral genomic DNA was used to reveal genetic variations among different isolates [9799] and to distinguish one species from another between closely related viruses such as Rachiplusia ou (RoMNPV), AcMNPV, Trichoplusia ni (TnMNPV), Galleria mellonella (GmMNPV) [100, 101] and the MNPVs of Spodoptera frugiperda [102].

Polymerase chain reaction (PCR)-based methods were then established. These methods have been shown not only to be more sensitive and faster but also more reliable than restriction enzyme analysis for classifying baculoviral species [4, 103105]. Multiple genetic markers (e.g., egt, ac17, lef-2, polh, p35, pif-2) could be used for the identification of baculoviruses [7, 106109]. The late expression factor 8 (lef-8), late expression factor 9 (lef-9) and polyhedrin(polh) were found in a highly conserved genes among baculoviruses [110], therefore, used as targets for degenerating PCR to characterize lepidopteran NPVs through the amplification of the conserved regions from a variety range of baculoviruses [111113]. The Kimura 2-parameter (K-2-P) distances between the aligned polh/gran, lef-8 and lef-9 nucleotide sequences were described by Jehle et al. for baculoviruses identification and species classification [3]. The K-2-P nucleotide substitution model from aligned nucleotide sequences were determined by using the pairwise distance calculation of MEGA version 3.0 applying the Kimura 2-parameter model [114].

Due to the higher cost of NGS for viral genome sequencing, it is frequently required to combine various approaches to cut down the cost but still ensure precision, e.g., PCR-based K-2-P analysis and NGS approach for identifying the potential new NPV species. Two NPVs were isolated from casuarina moth (Lymantria xylina) and golden birdwing larvae (Troides aeacus) collected from the fields, respectively, will be as representative cases for explanation in the following sections. We will focus on the characterization of these two potential new NPVs first and then the use of the sequences of three genes, lef-8, lef-9 and polyhedrin of two NPV candidates was used to examine their taxonomic position by K-2-P analysis. Finally, we will focus on the genome sequencing technology and bioinformatic analysis on NPVs.

Advertisement

3. The identification of ambiguous NPVs

In this section, the discussion of molecular identification of NPV species based on K-2-P distance [3] is presented. Two new NPVs were used as examples in this study to reveal different issues regarding the classification of NPVs.

3.1. LdMNPV-like virus

The K-2-P distances, based on the sequences of three genes, between different viruses could mostly evaluate the ambiguous relationship among the NPVs. It was defined that distances less than 0.015 indicates that the two isolates are the same baculovirus species. On the other hand, the difference between two viruses is more than 0.05 should be considered as different virus species. For the distances between 0.015 and 0.05, complementary information is needed to determine whether these two viruses are of the same or different species [3, 9, 115].

A new multiple nucleopolyhedrovirus strain was isolated from casuarina moth, L. xylina Swinhoe, (Lepidoptera: Lymantriidae) in Taiwan. Since the polyhedrin sequence of this virus had high identity to L. dispar MNPV (98%), it was named LdMNPV-like virus [116]. To precisely clarify the relationship of three Lymantriidae-derived NPVs (LdMNPV-like virus, LdMNPV and LyxyMNPV [60]), the K-2-P of polh, lef-8 and -9 was performed. The distances between LdMNPV-like virus and LyxyMNPV exceeded 0.05 for each gene, polh, lef-8, or lef-9 and also for concatenated polh/lef-8/lef-9 ( Figure 1 ). For LdMNPV-like virus and LdMNPV, not only the single lef-8 and lef-9 sequences but also concatenated polh/lef-8/lef-9, the distances were generally lower than 0.015, but only the polh sequence distance (0.016) exceeded slightly 0.015 ( Figure 1 ). These results strongly suggested that LdMNPV-like virus is an isolate of LdMNPV. However, as indicated by our previous report, the genome of LdMNPV-like virus is approximately 139 Kb, due to large deletions compared to that of LdMNPV [116]. To further investigate the LdMNPV-like virus, a HindIII-PstI fragment (7,054 nucleotides) was cloned, sequenced and compared to the corresponding region of LdMNPV. Nine putative ORFs (including seven with full lengths and two with partial lengths) and two homologous regions (hrs) were identified in this fragment ( Figure 2 ) and those genes, in order from the 5′ to 3′ end, encoded part of rr1, ctl-1, Ange-bro-c, LdOrf151, LdOrf-152-like peptides, Ld-bro-n, two Ld-bro-o and part of LdOrf155-like peptides ( Table 2 ). The physical map of HindIII-PstI fragment of LdMNPV-like virus showed that the gene organization was highly conserved compared to the corresponding region of LdMNPV, although several restriction enzyme recognition sites were different. Additionally, the ld-bro-o gene in the LdMNPV-like virus was split into two ORF7 and ORF8, due to a point deletion in the downstream (+669) of ORF7 and this deletion causes a frameshift that results in the formation of a stop codon (TGA) after 73 bp. Afterward, ORF8 was overlapped with the last four base pairs (ATGA) in ORF7. The nucleotide identities of these genes were 96–100% homologous to those of LdMNPV, except ORF3 which was 68% homologous to Ange-bro-c and ORF7 and ORF8 showing low identities to Ld-bro-o (73% and 26%, respectively). The deduced amino acid sequences of these genes were similar to those of LdMNPV, with identities of 81–100%, except the similarity of ORF3 to Ange-bro-c was 70% and ORF7 and ORF8 also showed low similarity to Ld-bro-o (67% and 26%, respectively). These results imply that the LdMNPV-like and LdMNPV viruses are closely related but not totally identical.

No* LdMNPV-like virus LdMNPV§
Position Length Name Identity (%)
nt aa nt aa
1 1 → 654 654 217 rr1 96 81
2 1063 → 1224 162 53 Ctl-1 100 100
3 1397 → 2473 1077 358 Ange-bro-c 68 70
4 2590 → 3596 504 168 LdOrf-151 99 98
5 3200 → 3952 753 251 LdOrf-152 99 99
6 4019 → 5026 1005 335 Ld-bro-n 93 91
7 5645 → 6391 744 248 Ld-bro-o 73 67
8 6388 → 6654 264 88 Ld-bro-o 26 26
9 6758 → 7054 297 99 LdOrf-155 100 100

Table 2.

Comparison of the nucleotide (nt) and deduced amino acid (aa) sequences for putative ORFs in LdMNPV-like virus genomic fragment and their corresponding LdMNPV homologues.

The directions of the transcripts are indicated by arrows.

§Reference from the genome of LdMNPV (Kuzio et al. [63])

*The nine potentially expressed ORFs are numbered in the order in which they occur in the LdMNPV-like virus genomic fragment from the 5′ to 3′ end. Two ORFs extend past this cloning site are printed in bold; only the N-terminus which contains 217 amino acids (654 nucleotides) and 99 amino acids (297 nucleotides) was examined.

Figure 1.

Pairwise K-2-P distances of the nucleotide sequences of polh, lef-8 and lef-9 and concatenated polh/lef-8/lef-9 fragments of LdMNPV-like virus, LyxyMNPV and LdMNPV. Modified data reproduced with permission of the Elsevier [116].

Figure 2.

Comparison of relative restriction sites and gene locations in the LdMNPV-like virus HindIII-PstI fragment with those of the corresponding LdMNPV fragment. Arrows denote ORFs and their direction of transcription. Gray boxes represent the homologous repeat regions (hrs). ORF homologues in the corresponding regions are drawn with the same patterns. Numbers below the arrows indicate the nine putative ORFs listed in Table 2 .

Based on these results, LdMNPV-like virus has a genomic size significantly smaller than that of LdMNPV and LyxyMNPV and appears to be an NPV isolate distinct from LdMNPV or LyxyMNPV. Moreover, a gene, ange-bro-c of LdMNPV-like virus, was truncated into two ORF7 and ORF8 and the sequence showed relatively low identity to that of LdMNPV ( Table 2 ). Taken together, these results indicate that LdMNPV-like virus is a distinct LdMNPV strain with several novel features. Otherwise, LdMNPV-like virus and LdMNPV have distinct geographical locations (from subtropical and cold temperate zones, respectively) and are distinct in genotypic and phenotypic characteristics and it also showed broad genetic variation among LdMNPV isolates [9].

3.2. An NPV isolate from T. aeacus larvae

A nucleopolyhedrosis disease of the rearing of the golden birdwing butterfly (T. aeacus) larvae was found and the polyhedral inclusion bodies (PIBs) were observed under light microscopy ( Figure 3 ). PCR was performed to amply the polh gene by 35/36 primer set ( Figure 3 ) to further confirm NPV infection [117, 118]. Therefore, this NPV was named provisionally TraeNPV. The three genes, polh, lef-8 and lef-9 of TraeNPV, were cloned and sequenced and then the K-2-P distances between the aligned single and concatenated polh, lef-8 and lef-9 nucleotide sequences were analyzed. The results indicated that TraeNPV belonged to the group I baculoviruses and closely related to BmNPV group. Figure 4 showed that most of the distances between TraeNPV and other NPVs were between 0.015 and 0.050, whereas the distances for polh between TraeNPV, PlxyNPV, RoNPV and AcMNPV group exceeded 0.05. It should be noted that for all the concatenated polh/lef-8/lef-9 sequences, the distances were apparently much more than 0.015 and even to 0.05. These results left an ambiguous situation of this NPV isolate; so far, we could conclude that TraeNPV neither belongs to BmNPV group nor AcMNPV group. More complementary information is needed to determine the viral species of TraeNPV.

Figure 3.

Identification of unknown NPV. (A) Light microscopy observation of liquefaction from the cadavers of T. aeacus larvae, scale bar = 20 μm. Black arrows indicated the polyhedral inclusion bodies (PIBs). (B) PCR detection of partial polyhedrin gene, M = 100 bp marker, (+) = positive control and (−) = negative control.

Figure 4.

Pairwise Kimura-2-parameter distances of the nucleotide sequences of lef-8, lef-9 and polh and concatenated polh/lef-8/lef-9 fragments of TraeNPV and 12 viruses.

In summary, K-2-P distances were employed to further clarify the relationship between closely related NPVs. We discussed two different cases analyzed by K-2-P. From the sequence data of LdMNPV-like virus, results strongly supported that LdMNPV-like virus is an isolate of LdMNPV. Since the RFLP profiles of the LdMNPV-like virus showed the genome of this isolate was deleted tremendously, this deletion also showed coordinately in our partial sequences of genomic DNA fragments and the results of K-2-P. The K-2-P distances between TraeNPV and BmNPV or AcMNPV were among 0.05 and 0.015. Anyway, we cannot define that this virus is a new species with the evidences of RFLP, part gene sequences and K-2-P results; therefore, it is necessary to get more data, especially the whole genome sequence of TraeNPV.

Advertisement

4. The importance of whole genome sequencing on baculoviruses

The rapidly growing mass of genomic data shifts the taxonomic approaches from traditional to genomically based issues. The K-2-P distance supported LyxyMNPV as a different viral species (K-2-P values = 0.067–0.088), even though they were still a closely relative species phylogenetically. But, “how different did LyxyMNPV and LdMNPV?” become another question. Thus, the whole genome sequence could provide deep information of this virus. For example, as the genomic data revealed, the most part of the ORF (151 ORFs) between LyxyMNPV and LdMNPV was quite similar while still have several different ORF exhibits or absent in LyxyMNPV, e.g., two ORFs were homologous to other baculoviruses and four unique ORFs were identified in the LyxyMNPV genome and LdMNPV contains 23 ORFs that are absent in LyxyMNPV [60]. Besides, there is a huge genomic inversion in LyxyMNPV compared to LdMNPV [60]. Another example is Maruca vitrata NPV (MaviNPV). All of the K-2-P distance-supported MaviNPV is quite different from other NPVs (K-2-P values = 0.092–0.237) ( Figure 6 ). While the gene content and gene order of MaviNPV were highly similar to that of AcMNPV and BmNPV, through the genomic sequencing, it showed the 100% collinear to AcMNPV [27] and MaviNPV shared 125 ORFs with AcMNPV and 123 with BmNPV. The detailed information could only be captured after whole genome sequencing rather than partial gene sequences or other phylogenetic analyses. Sometimes, usage of K-2-P data may raise other problems, which we mentioned above; it seems LdMNPV-like virus and LdMNPV were the same viral species. While through the restriction enzyme profile and partial genomic data, we could identify that there are some deletion fragments and different gene contents within the LdMNPV-like virus genome. For the TraeNPV, most of the K-2-P values were ranged from 0.015 to 0.05; thus, whole genome sequencing could be one of the best ways to figure out this ambiguous state. The more detailed information we can get, the more deep aspect we can evaluate, e.g., the taxonomic problems and further evolutionary studies.

Figure 5.

Pairwise Kimura-2-parameter distances of the nucleotide sequences of lef-8, lef-9 and polh and concatenated polh/lef-8/lef-9 fragments of MaviNPV and 12 viruses.

Figure 6.

Common bioinformatic workflow for genome assembly and analysis.

Advertisement

5. Genomic sequences of NPVs

5.1. Genome sequencing technology

Previous NPV genome sequencing employed three types of approaches: plasmid clone (or template) enrichment, NGS, or a combination of the two methods. Initially, the most common approach used restriction enzymes to fragmentize the viral genome into smaller pieces. Plasmid-based clone amplification was then employed to enrich templates for sequencing. Later, conventional Sanger sequencing and/or next-generation sequencing was employed for genome assembly. In addition, purely high-throughput sequencing-based approach from isolated viral genome was also employed [9, 15]. To date, next-generation sequencing technology plays an increasingly important role on viral genome assembly. Previous researches showed that Illumina HiSeq has superior performance in yield than 454 FLX [119121]. Baculoviruses usually contain a novel homologous region (hr) feature, which comprises a palindrome that is usually flanked by short direct repeats located elsewhere in the genome [122]. Thereby, the shorter single-read length of Illumina sequencers might lead the difficulty during genome assembly. Further application of paired-end read sequencing method could certainly provide alternative for sequencing overlap the hrs in baculoviral genomes.

5.2. Bioinformatic analysis

Construction of a complete genome map is essential for future genomic investigations. Besides sequencing, bioinformatic approaches are also required for determining the order and content of the nucleotide sequence information for the viral genome of interest. In general, bioinformatic approaches can be separated into three consecutive steps: genome assembly, genome annotation and phylogenetic relationship inference ( Figure 5 ).

5.2.1. Genome assembly

Sequence reads are the building blocks for genome sequencing and assembly. Thus, quality control of sequence reads plays a key role in determining the fidelity of a genome assembly. The procedure of read quality checking includes, but not limited to, the removal of unrelated sequences such as control sequences, adaptors, vectors, potential contaminants, etc., trimming of low-quality bases and selection of high-quality reads. The control sequences (e.g., PhiX control reads in Illumina sequencers, control DNA beads in Roche 454 sequencer) are routinely used by sequencer manufacturers to evaluate the quality of each sequencing run. There are software applications made available to be utilized to identify and remove control sequences and low-quality bases. For NGS, sequencing adapters could be identified in reads if the fragment size is shorter than read length. Cutadapt [123] was implemented to trim the adapter sequences. Ambiguous bases or bases with lower-quality values can be removed by PRINSEQ [124] from either 5′ or 3′ end. NGS QC Toolkit [125] has programmed module to select high-quality reads. If paired-end technology was applied, paired-end reads could be joined by PANDAseq [126], PEAR [127], FLASH [128] and COPE [129], if a fragment size is shorter than read length.

Genome can be assembled from quality paired-end or single-end reads with de novo or reference-guided approaches. There are two standard methods known as the de Bruijn graph (DBG) approach and the overlap/layout/consensus (OLC) approach for de novo genome assembly. The idea of de Bruijn graph is to decompose a read into kmer-sized fragments with sliding window screening. Each kmer-sized fragment will be used to construct graph for longer path (e.g., contigs). Then, long-range paired reads can be utilized to build scaffolds from contigs with given insert size and read orientation. SOAPdenovo [130] is one of the DBG assembler that has an extreme speed by utilizing threads parallelization [131]. The OLC assembler starts by identifying all pairs of reads with higher overlap region to construct an overlap graph. The contig candidates are identified by pruning nodes to simplify the overlap graph. The final contigs are then output based on consensus regions. Additionally, Newbler [132] is a widely used OLC assembler distributed by 454 Life Sciences.

Reference-guided genome assembly is another solution for genome assembly if the genome of a closely related species is already available. For viral genome assembly, closely related species can be identified by mapping quality reads against sequenced viral genomes deposited in GenBank (http://www.ncbi.nlm.nih.gov/genome/viruses/) and select top-ranked species as the reference genome(s) to facilitate the assembly of the genome of interest. Reference-guided assembler is also called mapping assembler that the complete genome is generated by mapping quality reads with variant (single nucleotide polymorphism (SNP), insertion and deletion) identification. For example, MIRA (a computer program) [133] can create a reference-based assembly by detecting the difference between references.

During the assembly process, gap filling (or gap elimination) is conducted to resolve the undetermined bases either by bioinformatics or other approaches such as PCR and additional sequencing. Bioinformatic approaches normally use paired-end reads to eliminate gaps. PCR coupled with Sanger sequencing is a common approach to finalize the undetermined regions [134]. In addition, Sanger sequencing can also be used for genome validation and homologous region (hr) checking.

5.2.2. Genome annotation

Annotation determines the locations of protein-coding and noncoding genes as well as the functional elements in the genome. Glimmer [135], N-SCAN [136], NCBI ORF Finder (https://www.ncbi.nlm.nih.gov/orffinder/), GeneMark [137] and VIGOR [138] are gene prediction tools for identifying protein-codivng genes in the genome. Repetitive sequence regions were detected by RepeatMasker (http://www.repeatmasker.org/). Viral microRNA candidate hairpins can be predicted by Vir-Mir [139]. The circular map of the viral genome was generated by CGView [140].

5.2.3. Phylogenetic analysis

Phylogenetic relationship inference reveals the evolutionary distances of various, especially closely related, species. MEGA [141] was the most widely used software suite that provides the sophisticated and integrated user interface for studying DNA and protein sequence data from species and populations. Alternatively, phylogenetic relationships among species based on the complete viral genomes or functional regions could also be estimated with Clustal Omega [142]. Clustal Omega was employed for multiple sequence alignment on the complete genomes and DNA fragments, respectively. ClustalW [143] was employed to do file format conversion of multiple sequence alignment. Ambiguously aligned positions were removed by using Gblocks version 0.91b [144, 145] under default settings. Phylogenetic tree inference could be constructed by hierarchical Bayesian method (e.g., MrBayes [146]) or maximum likelihood method (e.g., RAxML [147]) to estimate phylogeny [148]. Tree was depicted with FigTree version 1.4.2 (http://tree.bio.ed.ac.uk/software/figtree/). The divergence times of different species were estimated using BEAST version 1.8 or version 2.3.2 [149]. In addition, pairwise sequence identity was determined by BLASTN (NCBI BLAST Package) [150] to analyze sequence-level variation. Also, whole genome pairwise alignment can be done by LAGAN [151]. CGView comparison tool (CCT) [152] was used to represent the block similarity among different species. Mauve [153], one of the multiple genome alignment tools, can help us to visualize the consensus sequence blocks among distant-related species.

Up to 78 baculoviruses have been reported; most of baculoviruses have a narrow host range, only infect their homogenous hosts, such as BmNPV, SpltNPV, SpeiNPV, MaviNPV and so on; LyxyNPV can infect LD and LY cell lines, while AcMNPV has a wide host range; at least 40 hosts in vitro have be found. Therefore, a new baculovirus isolate needs to define its taxonomic position and to analyze its phylogenetic relationship with a known baculovirus member.

Advertisement

6. Conclusion

With the accomplishment of the sequencing technologies, more NPV genomes were sequenced. So far, more than 78 baculoviruses have been fully sequenced and based on the sequencing methods, we can divide into two parts, one is sequencing by Sanger method and another is sequencing by NGS method ( Table 1 ). Among these sequenced genomes, 35 genomes were sequenced by Sanger method and 43 genomes were sequenced by NGS methods. It could be expected that whole genome sequencing by NGS method would get much common in this field; however, the upcoming metagenomic era is imperative that one remains aware of and careful about the shortcomings of the information presented about the organisms that are being sequenced and that these databases can oversee neither the correctness of the organismal identifications nor of the sequences entered into the databases.

The natural environment harbors a large number of baculoviruses. However, only a few of them have been sequenced and studied. A lot more information related to the genetic relationship of NPVs in the natural environment is needed to facilitate our understanding of these creatures. Though NGS technology has become an important technology for viral genomic sequencing, high cost of NGS for whole viral genome sequencing remains a barrier. To reduce the cost, it is necessary to evaluate whether the newly collected NPVs are suitable for whole genome sequencing or not. Alternatively, biochemical approaches and biological tools, such as PCR-based K-2-P analysis, can be good options to facilitate the process. As expected, all these applications are anticipated to help us reveal the genetic information of unknown species, so that more detailed insights of their genetic makeup and functional composition can be obtained to help us better understand the nature of these viruses. By using the powerful sequencing technique, the metagenomic progress (e.g., transcriptome analysis of insect host), new pathogen species in the natural environment would be easier to be found in the future. With the increase of new baculoviral genomic data, improvement of bioinformatic analysis methods and further validation of biological information would generate a group of genes, which connect to the viral host range and solve the contradiction situation in the baculoviral genomics.

Advertisement

Acknowledgments

This research was supported by Grant 105AS-13.2.3-BQ-B1 from the Bureau of Animal and Plant Health Inspection and Quarantine, the Council of Agriculture, Executive Yuan and Grant 103-2313-B-197-002-MY3 from the Ministry of Science and Technology (MOST).

References

  1. 1. Takatsuka, J., Lymantria mathura nucleopolyhedrovirus: identification, occurrence and genetic diversity in Iwate Prefecture, Japan. J Invertebr Pathol, 2016. 138: pp. 1-9.
  2. 2. Boucias, D. and Pendland, J.C., Principles of insect pathology. 1998, Boston: Kluwer Aca demic Publishers. 537p.
  3. 3. Jehle, J.A., et al., Molecular identification and phylogenetic analysis of baculoviruses from Lepidoptera. Virology, 2006. 346(1): pp. 180-93.
  4. 4. Herniou, E.A., et al., The genome sequence and evolution of baculoviruses. Annu Rev Entomol, 2003. 48: pp. 211-34.
  5. 5. Moscardi, F., Assessment of the application of baculoviruses for control of Lepidoptera. Annu Rev Entomol, 1999. 44: pp. 257-89.
  6. 6. Smith, G.E., Summers, M.D. and Fraser, M.J., Production of human beta interferon in insect cells infected with a baculovirus expression vector. Mol Cell Biol, 1983. 3(12): pp. 2156-65.
  7. 7. Mehrvar, A., R.R.J., Veenakumari, K., Narabenchi, G.B., Molecular and biological characteristics of some geographic isolates of nucleopolyhedrovirus of Helicoverpa armigera (Lep.: Noctuidae). J Entomol Soc Iran, 2008. 28(1): pp. 39-60.
  8. 8. Murhammer, D.W., Useful tips, widely used techniques and quantifying cell metabolic behavior. Methods Mol Biol, 2007. 388: pp. 3-22.
  9. 9. Harrison, R.L., Keena, M.A. and Rowley, D.L., Classification, genetic variation and pathogenicity of Lymantria dispar nucleopolyhedrovirus isolates from Asia, Europe and North America. J Invertebr Pathol, 2014. 116: pp. 27-35.
  10. 10. Wang, J., et al., Genome sequencing and analysis of Catopsilia pomona nucleopolyhedrovirus: a distinct species in group I Alphabaculovirus. PLoS One, 2016. 11(5): p. e0155134.
  11. 11. Oliveira, J.V., et al., Genome of the most widely used viral biopesticide: Anticarsia gemmatalis multiple nucleopolyhedrovirus. J Gen Virol, 2006. 87(Pt 11): pp. 3233-50.
  12. 12. Brito, A.F., et al., The pangenome of the Anticarsia gemmatalis multiple nucleopolyhedrovirus (AgMNPV). Genome Biol Evol, 2016. 8(1): pp. 94-108.
  13. 13. Nie, Z.M., et al., Complete sequence and organization of Antheraea pernyi nucleopolyhedrovirus, a dr-rich baculovirus. BMC Genomics, 2007. 8: pp. 248.
  14. 14. Ayres, M.D., et al., The complete DNA sequence of Autographa californica nuclear polyhedrosis virus. Virology, 1994. 202(2): pp. 586-605.
  15. 15. Chateigner, A., et al., Ultra deep sequencing of a baculovirus population reveals widespread genomic variations. Viruses, 2015. 7(7): pp. 3625-46.
  16. 16. Xu, Y.P., et al., Comparative analysis of the genomes of Bombyx mandarina and Bombyx mori nucleopolyhedroviruses. J Microbiol, 2010. 48(1): pp. 102-10.
  17. 17. Gomi, S., Majima, K. and Maeda, S., Sequence analysis of the genome of Bombyx mori nucleopolyhedrovirus. J Gen Virol, 1999. 80 (Pt 5): pp. 1323-37.
  18. 18. Lauzon, H.A., et al., Gene organization and sequencing of the Choristoneura fumiferana defective nucleopolyhedrovirus genome. J Gen Virol, 2005. 86(Pt 4): pp. 945-61.
  19. 19. de Jong, J.G., et al., Analysis of the Choristoneura fumiferana nucleopolyhedrovirus genome. J Gen Virol, 2005. 86(Pt 4): pp. 929-43.
  20. 20. Rohrmann, G.F., Erlandson, M.A. and Theilmann, D.A., Genome sequence of an alphabaculovirus isolated from Choristoneura murinana. Genome Announc, 2014. 2(1): e01135-13.
  21. 21. Thumbi, D.K., et al., Comparative genome sequence analysis of Choristoneura occidentalis Freeman and C. rosaceana Harris (Lepidoptera: Tortricidae) alphabaculoviruses. PLoS One, 2013. 8(7): p. e68968.
  22. 22. Castro, M.E., et al., Identification of a new nucleopolyhedrovirus from naturally-infected Condylorrhiza vestigialis (Guenée) (Lepidoptera: Crambidae) larvae on poplar plantations in South Brazil. J Invertebr Pathol, 2009. 102(2): pp. 149-54.
  23. 23. Krejmer, M., et al., The genome of Dasychira pudibunda nucleopolyhedrovirus (DapuNPV) reveals novel genetic connection between baculoviruses infecting moths of the Lyman-triidae family. BMC Genomics, 2015. 16: p. 759.
  24. 24. Ma, X.-C., et al., Genome sequence and organization of a nucleopolyhedrovirus that infects the tea looper caterpillar, Ectropis obliqua. Virology, 2007. 360(1): pp. 235-46.
  25. 25. Ikeda, M., et al., Gene organization and complete sequence of the Hyphantria cunea nucleopolyhedrovirus genome. J Gen Virol, 2006. 87(Pt 9): pp. 2549-62.
  26. 26. Aragao-Silva, C.W., et al., The complete genome of a baculovirus isolated from an insect of medical interest: Lonomia obliqua (Lepidoptera: Saturniidae). Sci Rep, 2016. 6: p. 23127.
  27. 27. Chen, Y.R., et al., Genomic and host range studies of Maruca vitrata nucleopolyhedrovirus. J Gen Virol, 2008. 89(Pt 9): pp. 2315-30.
  28. 28. Ahrens, C.H., et al., The sequence of the Orgyia pseudotsugata multinucleocapsid nuclear polyhedrosis virus genome. Virology, 1997. 229(2): pp. 381-99.
  29. 29. Qian, H., et al., Analysis of the genomic sequence of Philosamia cynthia nucleopolyhedrin virus and comparison with Antheraea pernyi nucleopolyhedrin virus. BMC Genomics, 2013. 14: p. 115.
  30. 30. Harrison, R.L. and Lynn, D.E., Genomic sequence analysis of a nucleopolyhedrovirus isolated from the diamondback moth, Plutella xylostella. Virus Genes, 2007. 35(3): pp. 857-73.
  31. 31. Harrison, R.L. and Bonning, B.C., Comparative analysis of the genomes of Rachiplusia ou and Autographa californica multiple nucleopolyhedroviruses. J Gen Virol, 2003. 84(Pt 7): pp. 1827-42.
  32. 32. Wang, Y.S., et al., Genome of Thysanoplusia orichalcea multiple nucleopolyhedrovirus lacks the superoxide dismutase gene. J Virol, 2012. 86(21): pp. 11948-9.
  33. 33. Nakai, M., et al., Genome sequence and organization of a nucleopolyhedrovirus isolated from the smaller tea tortrix, Adoxophyes honmai. Virology, 2003. 316(1): pp. 171-83.
  34. 34. Hilton, S. and Winstanley, D., Genomic sequence and biological characterization of a nucleopolyhedrovirus isolated from the summer fruit tortrix, Adoxophyes orana. J Gen Virol, 2008. 89(Pt 11): pp. 2898-908.
  35. 35. Harrison, R.L., Genomic sequence analysis of the Illinois strain of the Agrotis ipsilon multiple nucleopolyhedrovirus. Virus Genes, 2009. 38(1): pp. 155-70.
  36. 36. Jakubowska, A.K., et al., Genome sequence of an enhancin gene-rich nucleopolyhedrovirus (NPV) from Agrotis segetum: collinearity with Spodoptera exigua multiple NPV. J Gen Virol, 2006. 87(Pt 3): pp. 537-51.
  37. 37. Wennmann, J.T., Gueli Alletti, G. and Jehle, J.A., The genome sequence of Agrotis segetum nucleopolyhedrovirus B (AgseNPV-B) reveals a new baculovirus species within the Agrotis baculovirus complex. Virus Genes, 2015. 50(2): pp. 260-76.
  38. 38. Zhu, Z., et al., Genome sequence and analysis of Buzura suppressaria nucleopolyhedrovirus: a group II Alphabaculovirus. PLoS One, 2014. 9(1): p. e86450.
  39. 39. van Oers, M.M., et al., Genome sequence of Chrysodeixis chalcites nucleopolyhedrovirus, a baculovirus with two DNA photolyase genes. J Gen Virol, 2005. 86(Pt 7): pp. 2069-80.
  40. 40. Bernal, A., et al., Complete genome sequences of five Chrysodeixis chalcites nucleopolyhedrovirus genotypes from a Canary Islands isolate. Genome Announc, 2013. 1(5).
  41. 41. Zhu, S.Y., et al., Genomic sequence, organization and characteristics of a new nucleopolyhedrovirus isolated from Clanis bilineata larva. BMC Genomics, 2009. 10: p. 91.
  42. 42. Hyink, O., et al., Whole genome analysis of the Epiphyas postvittana nucleopolyhedrovirus. J Gen Virol, 2002. 83(Pt 4): pp. 957-71.
  43. 43. Tang, X.D., et al., Morphology and genome of Euproctis pseudoconspersa nucleopolyhedrovirus. Virus Genes, 2009. 38(3): pp. 495-506.
  44. 44. Noune, C. and Hauxwell, C., Complete genome sequences of Helicoverpa armigera single nucleopolyhedrovirus strains AC53 and H25EA1 from Australia. Genome Announc, 2015. 3(5):e01083-15..
  45. 45. Tang, P., et al., Genomic sequencing and analyses of HearMNPV--a new Multinucleocapsid nucleopolyhedrovirus isolated from Helicoverpa armigera. Virol J, 2012. 9: p. 168.
  46. 46. Zhang, C.X., Ma, X.C. and Guo, Z.J., Comparison of the complete genome sequence between C1 and G4 isolates of the Helicoverpa armigera single nucleocapsid nucleopolyhedrovirus. Virology, 2005. 333(1): pp. 190-9.
  47. 47. Zhang, C.X. and Wu, J.C., Genome structure and the p10 gene of the Helicoverpa armigera nucleopolyhedrovirus. Acta Biochim Biophys Sinica, 2001. 33(2): pp. 179-84.
  48. 48. Chen, X., et al., The sequence of the Helicoverpa armigera single nucleocapsid nucleopolyhedrovirus genome. J Gen Virol, 2001. 82(Pt 1): pp. 241-57.
  49. 49. Ogembo, J.G., et al., Comparative genomic sequence analysis of novel Helicoverpa armigera nucleopolyhedrovirus (NPV) isolated from Kenya and three other previously sequenced Helicoverpa spp. NPVs. Virus Genes, 2009. 39(2): pp. 261-72.
  50. 50. Chen, X., et al., Comparative analysis of the complete genome sequences of Helicoverpa zea and Helicoverpa armigera single-nucleocapsid nucleopolyhedroviruses. J Gen Virol, 2002. 83(Pt 3): pp. 673-84.
  51. 51. Rohrmann, G.F., Erlandson, M.A. and Theilmann, D.A., The genome of a baculovirus isolated from Hemileuca sp. encodes a serpin ortholog. Virus Genes, 2013. 47(2): pp. 357-64.
  52. 52. Rohrmann, G.F., Erlandson, M.A. and Theilmann, D.A., Genome sequence of an alphabaculovirus isolated from the Oak Looper, Lambdina fiscellaria, contains a putative 2-kilobase-pair transposable element encoding a transposase and a FLYWCH domain-containing protein. Genome Announc, 2015. 3(3): e00186-15.
  53. 53. Xiao, H. and Qi, Y., Genome sequence of Leucania seperata nucleopolyhedrovirus. Virus Genes, 2007. 35(3): pp. 845-56.
  54. 54. Kuzio, J., et al., Sequence and analysis of the genome of a baculovirus pathogenic for Lymantria dispar. Virology, 1999. 253(1): pp. 17-34.
  55. 55. Kabilov, M.R., et al., Complete genome sequence of a Western Siberian Lymantria dispar multiple nucleopolyhedrovirus isolate. Genome Announc, 2015. 3(2).
  56. 56. Rabalski, L., et al., An alphabaculovirus isolated from dead Lymantria dispar larvae shows high genetic similarity to baculovirus previously isolated from Lymantria monacha – An example of adaptation to a new host. J Invertebr Pathol, 2016. 139: pp. 56-66.
  57. 57. Harrison, R.L. and Rowley, D.L., Complete genome sequence of the strain of Lymantria dispar multiple nucleopolyhedrovirus found in the gypsy moth biopesticide Virin-ENSh. Genome Announc, 2015. 3(1):e01407-14.
  58. 58. Martemyanov, V.V., et al., The enhancin gene: one of the genetic determinants of population variation in baculoviral virulence. Dokl Biochem Biophys, 2015. 465: pp. 351-3.
  59. 59. Harrison, R.L., Rowley, D.L. and Keena, M.A., Geographic isolates of Lymantria dispar multiple nucleopolyhedrovirus: Genome sequence analysis and pathogenicity against European and Asian gypsy moth strains. J Invertebr Pathol, 2016. 137: pp. 10-22.
  60. 60. Nai, Y.S., et al., Genomic sequencing and analyses of Lymantria xylina multiple nucleopolyhedrovirus. BMC Genomics, 2010. 11: pp. 116.
  61. 61. Choi, J.B., et al., Complete genomic sequences and comparative analysis of Mamestra brassicae nucleopolyhedrovirus isolated in Korea. Virus Genes, 2013. 47(1): pp. 133-51.
  62. 62. Li, Q., et al., Sequence and organization of the Mamestra configurata nucleopolyhedrovirus genome. Virology, 2002. 294(1): pp. 106-21.
  63. 63. Li, L., et al., Complete comparative genomic analysis of two field isolates of Mamestra configurata nucleopolyhedrovirus-A. J Gen Virol, 2005. 86(Pt 1): pp. 91-105.
  64. 64. Thumbi, D.K., et al., Complete sequence, analysis and organization of the Orgyia leucostigma nucleopolyhedrovirus genome. Viruses, 2011. 3(11): pp. 2301-27.
  65. 65. Rohrmann, G.F., Erlandson, M.A. and Theilmann, D.A., A distinct group II alphabaculovirus isolated from a Peridroma species. Genome Announc, 2015. 3(2):e00185-15.
  66. 66. Craveiro, S.R., et al., The genome sequence of Pseudoplusia includens single nucleopolyhedrovirus and an analysis of p26 gene evolution in the baculoviruses. BMC Genomics, 2015. 16: p. 127.
  67. 67. WF, I.J., et al., Sequence and organization of the Spodoptera exigua multicapsid nucleopolyhedrovirus genome. J Gen Virol, 1999. 80 (Pt 12): pp. 3289-304.
  68. 68. Harrison, R.L., Puttler, B. and Popham, H.J., Genomic sequence analysis of a fast-killing isolate of Spodoptera frugiperda multiple nucleopolyhedrovirus. J Gen Virol, 2008. 89(Pt 3): pp. 775-90.
  69. 69. Breitenbach, J.E., et al., Determination and analysis of the genome sequence of Spodoptera littoralis multiple nucleopolyhedrovirus. Virus Res, 2013. 171(1): pp. 194-208.
  70. 70. Pang, Y., et al., Sequence analysis of the Spodoptera litura multicapsid nucleopolyhedrovirus genome. Virology, 2001. 287(2): pp. 391-404.
  71. 71. Liu, X., et al., Genomic sequencing and analysis of Sucra jujuba nucleopolyhedrovirus. PLoS One, 2014. 9(10): p. e110023.
  72. 72. Willis, L.G., et al., Sequence analysis of the complete genome of Trichoplusia ni single nucleopolyhedrovirus and the identification of a baculoviral photolyase gene. Virology, 2005. 338(2): pp. 209-26.
  73. 73. Wormleaton, S., Kuzio, J. and Winstanley, D., The complete sequence of the Adoxophyes orana granulovirus genome. Virology, 2003. 311(2): pp. 350-65.
  74. 74. Liang, Z., et al., Genomic sequencing and analysis of Clostera anachoreta granulovirus. Arch Virol, 2011. 156(7): pp. 1185-98.
  75. 75. Yin, F., et al., The complete genome of a New Betabaculovirus from Clostera anastomosis. PLoS One, 2015. 10(7): p. e0132792.
  76. 76. Zhang, S., et al., Genome sequencing and analysis of a granulovirus isolated from the Asiatic rice leafroller, Cnaphalocrocis medinalis. Virol Sin, 2015. 30(6): pp. 417-24.
  77. 77. Han, G., et al., Genome of Cnaphalocrocis medinalis granulovirus, the first Crambidae-infecting betabaculovirus isolated from rice leaffolder to sequenced. PLoS One, 2016. 11(2): p. e0147882.
  78. 78. Escasa, S.R., et al., Sequence analysis of the Choristoneura occidentalis granulovirus genome. J Gen Virol, 2006. 87(Pt 7): pp. 1917-33.
  79. 79. Liang, Z., et al., Comparative analysis of the genomes of Clostera anastomosis (L.) granulovirus and Clostera anachoreta granulovirus. Arch Virol, 2013. 158(10): pp. 2109-14.
  80. 80. Luque, T., et al., The complete sequence of the Cydia pomonella granulovirus genome. J Gen Virol, 2001. 82(Pt 10): pp. 2531-47.
  81. 81. Lange, M. and Jehle, J.A., The genome of the Cryptophlebia leucotreta granulovirus. Virology, 2003. 317(2): pp. 220-36.
  82. 82. Ardisson-Araujo, D.M., et al., A betabaculovirus-encoded gp64 homolog codes for a functional envelope fusion protein. J Virol, 2016. 90(3): pp. 1668-72.
  83. 83. Ferrelli, M.L., et al., Genome of Epinotia aporema granulovirus (EpapGV), a polyorganotropic fast killing betabaculovirus with a novel thymidylate kinase gene. BMC Genomics, 2012. 13: p. 548.
  84. 84. Ardisson-Araujo, D.M., et al., Genome sequence of Erinnyis ello granulovirus (ErelGV), a natural cassava hornworm pesticide and the first sequenced sphingid-infecting betabaculovirus. BMC Genomics, 2014. 15: p. 856.
  85. 85. Harrison, R.L. and Popham, H.J., Genomic sequence analysis of a granulovirus isolated from the Old World bollworm, Helicoverpa armigera. Virus Genes, 2008. 36(3): pp. 565-81.
  86. 86. Harrison, R.L., Rowley, D.L. and Funk, C.J., The complete genome sequence of Plodia interpunctella granulovirus: evidence for horizontal gene transfer and discovery of an unusual inhibitor-of-apoptosis gene. PLoS One, 2016. 11(7): p. e0160389.
  87. 87. Taha, A., et al., Comparative analysis of the granulin regions of the Phthorimaea operculella and Spodoptera littoralis granuloviruses. Virus Genes, 2000. 21(3): pp. 147-55.
  88. 88. Hashimoto, Y., et al., Sequence analysis of the Plutella xylostella granulovirus genome. Virology, 2000. 275(2): pp. 358-72.
  89. 89. Zhang, B.Q., et al., The genome of Pieris rapae granulovirus. J Virol, 2012. 86(17): p. 9544.
  90. 90. Cuartas, P.E., et al., The complete sequence of the first Spodoptera frugiperda Betabaculovirus genome: a natural multiple recombinant virus. Viruses, 2015. 7(1): pp. 394-421.
  91. 91. Wang, Y., et al., Genomic sequence analysis of granulovirus isolated from the tobacco cutworm, Spodoptera litura. PLoS One, 2011. 6(11): p. e28163.
  92. 92. Hayakawa, T., et al., Sequence analysis of the Xestia c-nigrum granulovirus genome. Virology, 1999. 262(2): pp. 277-97.
  93. 93. Duffy, S.P., et al., Sequence analysis and organization of the Neodiprion abietis nucleopolyhedrovirus genome. J Virol, 2006. 80(14): pp. 6952-63.
  94. 94. Lauzon, H.A., et al., Sequence and organization of the Neodiprion lecontei nucleopolyhedrovirus genome. J Virol, 2004. 78(13): pp. 7023-35.
  95. 95. Garcia-Maruniak, A., et al., Sequence analysis of the genome of the Neodiprion sertifer nucleopolyhedrovirus. J Virol, 2004. 78(13): pp. 7036-51.
  96. 96. Afonso, C.L., et al., Genome sequence of a baculovirus pathogenic for Culex nigripalpus. J Virol, 2001. 75(22): pp. 11157-65.
  97. 97. Miller, L.K. and Dawes, K.P., Restriction endonuclease analysis for the identification of baculovirus pesticides. Appl Environ Microbiol, 1978. 35(2): pp. 411-21.
  98. 98. Smith, G.E. and Summers, M.D., Analysis of baculovirus genomes with restriction endonucleases. Virology, 1978. 89(2): pp. 517-27.
  99. 99. Lee, H.H. and Miller, L.K., Isolation of genotypic variants of Autographa californica nuclear polyhedrosis virus. J Virol, 1978. 27(3): pp. 754-67.
  100. 100. Miller, L.K. and Dawes, K.P., Restriction endonuclease analysis to distinguish two closely related nuclear polyhedrosis viruses: Autographa californica MNPV and Trichoplusia ni MNPV. Appl Environ Microbiol, 1978. 35(6): pp. 1206-10.
  101. 101. Smith, G.E. and Summers, M.D., Restriction Maps of Five Autographa californica MNPV Variants, Trichoplusia ni MNPV and Galleria mellonella MNPV DNAs with Endonucleases SmaI, KpnI, BamHI, SacI, XhoI and EcoRI. J Virol, 1979. 30(3): pp. 828-38.
  102. 102. Loh, L.C., et al., Analysis of the Spodoptera frugiperda nuclear polyhedrosis virus genome by restriction endonucleases and electron microscopy. J Virol, 1982. 44(2): pp. 747-51.
  103. 103. de Moraes, R.R. and Maruniak, J.E., Detection and identification of multiple baculoviruses using the polymerase chain reaction (PCR) and restriction endonuclease analysis. J Virol Methods, 1997. 63(1-2): pp. 209-17.
  104. 104. Ernoult-Lange, M., et al., Characterization of the simian virus 40 late promoter: relative importance of sequences within the 72-base-pair repeats differs before and after viral DNA replication. J Virol, 1987. 61(1): pp. 167-76.
  105. 105. Woo, S.D., Rapid detection of multiple nucleopolyhedroviruses using polymerase chain reaction. Mol Cells, 2001. 11(3): pp. 334-40.
  106. 106. Wang, L.H., et al., Sequence analysis of the Bam HI-J fragment of the Spodoptera litura multicapsid nucleopolyhedrovirus. Acta Biochim Biophy Sinica, 2001. 33(6): pp. 615-20.
  107. 107. Pijlman, G.P., A.J. Pruijssers and Vlak, J.M., Identification of pif-2, a third conserved baculovirus gene required for per os infection of insects. J Gen Virol, 2003. 84(Pt 8): pp. 2041-9.
  108. 108. Herniou, E.A., et al., Use of whole genome sequence data to infer baculovirus phylogeny. J Virol, 2001. 75(17): pp. 8117-26.
  109. 109. Somasekar, S., Jayapragasam, M., Rabindra, R. J., Characterization of five Indian isolates of the nuclear polyhedrosis virus of Helicoverpa armigera. Phytoparasitica, 1993. 21(4): pp. 333-7.
  110. 110. Lange, M., et al., Towards a molecular identification and classification system of lepidopteran-specific baculoviruses. Virology, 2004. 325(1): pp. 36-47.
  111. 111. Acharya, A. and Gopinathan, K.P., Characterization of late gene expression factors lef-9 and lef-8 from Bombyx mori nucleopolyhedrovirus. J Gen Virol, 2002. 83(Pt 8): pp. 2015-23.
  112. 112. Crouch, E.A., et al., Inter-subunit interactions of the Autographa californica M nucleopolyhedrovirus RNA polymerase. Virology, 2007. 367(2): pp. 265-74.
  113. 113. Toprak, U., et al., Preoperative evaluation of renal anatomy and renal masses with helical CT, 3D-CT and 3D-CT angiography. Diagn Interv Radiol, 2005. 11(1): pp. 35-40.
  114. 114. Kumar, S., K. Tamura and Nei, M., MEGA3: integrated software for molecular evolutionary genetics analysis and sequence alignment. Brief Bioinform, 2004. 5(2): pp. 150-63.
  115. 115. Jose, J., et al., Molecular characterization of nucleopolyhedrovirus of three lepidopteran pests using late expression factor-8 gene. Indian J Virol, 2013. 24(1): pp. 59-65.
  116. 116. Nai, Y.S., et al., A new nucleopolyhedrovirus strain (LdMNPV-like virus) with a defective fp25 gene from Lymantria xylina (Lepidoptera: Lymantriidae) in Taiwan. J Invertebr Pathol, 2009. 102(2): pp. 110-9.
  117. 117. Chou, C.M., et al., Characterization of Perina nuda nucleopolyhedrovirus (PenuNPV) polyhedrin gene. J Invertebr Pathol, 1996. 67(3): pp. 259-66.
  118. 118. Wang, C.H., et al., Continuous cell line from pupal ovary of Perina nuda (Lepidoptera: Lymantriidae) that is permissive to nuclear polyhedrosis virus from P. nuda. J Invertebr Pathol, 1996. 67(3): pp. 199-204.
  119. 119. Sims, D., et al., Sequencing depth and coverage: key considerations in genomic analyses. Nat Rev Genet, 2014. 15(2): pp. 121-32.
  120. 120. Goodwin, S., McPherson, J.D. and McCombie, W.R., Coming of age: ten years of next-generation sequencing technologies. Nat Rev Genet, 2016. 17(6): pp. 333-51.
  121. 121. Luo, C., et al., Direct comparisons of Illumina vs. Roche 454 sequencing technologies on the same microbial community DNA sample. PLoS One, 2012. 7(2): p. e30087.
  122. 122. Garcia-Maruniak A.et al., A variable region of Anticarsia gemmatalis nuclear polyhedrosis virus contains tandemly repeated DNA sequences. Virus Res, 1996. 41:123-132.
  123. 123. Martin, M., Cutadapt removes adapter sequences from high-throughput sequencing Reads. EMBnet.journal, 2011. 17(1): pp. 10-12.
  124. 124. Schmieder, R. and Edwards, R., Quality control and preprocessing of metagenomic datasets. Bioinformatics, 2011. 27(6): pp. 863-4.
  125. 125. Patel, R.K. and Jain, M., NGS QC toolkit: a toolkit for quality control of next generation sequencing data. PLoS One, 2012. 7(2): p. e30619.
  126. 126. Masella, A.P., et al., PANDAseq: paired-end assembler for illumina sequences. BMC Bioinformatics, 2012. 13: pp. 31.
  127. 127. Zhang, J., et al., PEAR: a fast and accurate illumina paired-end reAd mergeR. Bioinformatics, 2014. 30(5): pp. 614-20.
  128. 128. Magoc, T. and Salzberg, S.L., FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics, 2011. 27(21): pp. 2957-63.
  129. 129. Liu, B., et al., COPE: an accurate k-mer-based pair-end reads connection tool to facilitate genome assembly. Bioinformatics, 2012. 28(22): pp. 2870-4.
  130. 130. Luo, R., et al., SOAPdenovo2: an empirically improved memory-efficient short-read de novo assembler. Gigascience, 2012. 1(1): p. 18.
  131. 131. Zhang, W., et al., A practical comparison of de novo genome assembly software tools for next-generation sequencing technologies. PLoS One, 2011. 6(3): p. e17915.
  132. 132. Margulies, M., et al., Genome sequencing in microfabricated high-density picolitre reactors. Nature, 2005. 437(7057): pp. 376-80.
  133. 133. Chevreux, B., et al., Using the miraEST assembler for reliable and automated mRNA transcript assembly and SNP detection in sequenced ESTs. Genome Res, 2004. 14(6): pp. 1147-59.
  134. 134. Untergasser, A., et al., Primer3--new capabilities and interfaces. Nucleic Acids Res, 2012. 40(15): p. e115.
  135. 135. Salzberg, S.L., et al., Microbial gene identification using interpolated Markov models. Nucleic Acids Res, 1998. 26(2): pp. 544-8.
  136. 136. van Baren, M.J., Koebbe, B.C. and Brent, M.R., Using N-SCAN or TWINSCAN to predict gene structures in genomic DNA sequences. Curr Protoc Bioinformatics, 2007. 4: p. Unit 4 8.
  137. 137. Lukashin, A.V. and Borodovsky, M., GeneMark.hmm: new solutions for gene finding. Nucleic Acids Res, 1998. 26(4): pp. 1107-15.
  138. 138. Wang, S., Sundaram, J.P. and Spiro, D., VIGOR, an annotation program for small viral genomes. BMC Bioinformatics, 2010. 11: pp. 451.
  139. 139. Li, S.C., Shiau, C.K. and Lin, W.C., Vir-Mir db: prediction of viral microRNA candidate hairpins. Nucleic Acids Res, 2008. 36(Database issue): pp. D184-9.
  140. 140. Stothard, P. and Wishart, D.S., Circular genome visualization and exploration using CGView. Bioinformatics, 2005. 21(4): p. 537-9.
  141. 141. Kumar, S., et al., MEGA: a biologist-centric software for evolutionary analysis of DNA and protein sequences. Brief Bioinform, 2008. 9(4): pp. 299-306.
  142. 142. Sievers, F., et al., Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol Syst Biol, 2011. 7: pp. 539.
  143. 143. Thompson, J.D., Gibson, T.J. and Higgins, D.G., Multiple sequence alignment using ClustalW and ClustalX. Curr Protoc Bioinformatics, 2002. 2: p. Unit 2 3.
  144. 144. Talavera, G. and Castresana, J., Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Syst Biol, 2007. 56(4): pp. 564-77.
  145. 145. Castresana, J., Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol, 2000. 17(4): pp. 540-52.
  146. 146. Ronquist, F., et al., MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol, 2012. 61(3): pp. 539-42.
  147. 147. Stamatakis, A., RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics, 2014. 30(9): pp. 1312-3.
  148. 148. Douady, C.J., et al., Comparison of Bayesian and maximum likelihood bootstrap measures of phylogenetic reliability. Mol Biol Evol, 2003. 20(2): pp. 248-54.
  149. 149. Drummond, A.J., et al., Bayesian phylogenetics with BEAUti and the BEAST 1.7. Mol Biol Evol, 2012. 29(8): pp. 1969-73.
  150. 150. Camacho, C., et al., BLAST+: architecture and applications. BMC Bioinformatics, 2009. 10: p. 421.
  151. 151. Brudno, M., et al., LAGAN and Multi-LAGAN: efficient tools for large-scale multiple alignment of genomic DNA. Genome Res, 2003. 13(4): pp. 721-31.
  152. 152. Grant, J.R., Arantes, A.S. and Stothard, P., Comparing thousands of circular genomes using the CGView comparison tool. BMC Genomics, 2012. 13: pp. 202.
  153. 153. Darling, A.C., et al., Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res, 2004. 14(7): pp. 1394-403.

Written By

Yu-Shin Nai, Yu-Feng Huang, Tzu-Han Chen, Kuo-Ping Chiu and Chung-Hsiung Wang

Submitted: 19 May 2016 Reviewed: 28 October 2016 Published: 05 April 2017