You are on page 1of 167

University of Wollongong

Research Online
University of Wollongong Thesis Collection University of Wollongong Thesis Collections

2014

Mitochondrial genomes and their utility for the


recovery of phylogeny in the Hymenoptera
Meng Mao
University of Wollongong

Recommended Citation
Mao, Meng, Mitochondrial genomes and their utility for the recovery of phylogeny in the Hymenoptera, Doctor of Philosophy thesis,
School of Biological Sciences, University of Wollongong, 2014. http://ro.uow.edu.au/theses/4206

Research Online is the open access institutional repository for the


University of Wollongong. For further information contact the UOW
Library: research-pubs@uow.edu.au
Mitochondrial genomes and their utility for the recovery of
phylogeny in the Hymenoptera

A thesis submitted in fulfillment of the requirements for the award of the degree

Doctor of Philosophy

from

University of Wollongong

by

Meng Mao
(BSc, MAg)

School of Biological Sciences

2014
Thesis Certification

I, Meng Mao, declare that this thesis, submitted in fulfillment of the requirements for the award of

Doctor of Philosophy, in the School of Biological Sciences, University of Wollongong, is wholly my

own work unless otherwise referenced or acknowledged. The document has not been submitted for

qualifications at any other academic institution.

Meng Mao

18th August 2014

ii
Table of Contents

Thesis Certification ............................................................................................................................ ii

List of Tables .................................................................................................................................... viii

List of Figures .................................................................................................................................... ix

List of Abbreviations ......................................................................................................................... xi

Abstract ............................................................................................................................................ xiv

Acknowledgements.......................................................................................................................... xvi

List of publications ......................................................................................................................... xvii

Statement of Contribution ............................................................................................................ xviii

CHAPTER 1: General Introduction................................................................................................. 1

1.1 Preamble ............................................................................................................................. 1


1.2 The mitochondrial genome and phylogenetics................................................................. 2
1.3 Mitochondrial gene rearrangements ................................................................................ 4

1.3.1 Mechanisms of gene rearrangement .................................................................................... 5

1.3.2 Gene rearrangements as phylogenetic markers ................................................................... 8

1.4 Higher-level phylogenetic relationships within the Hymenoptera ................................. 9

1.4.1 Suborder Symphyta ............................................................................................................. 9

1.4.2 Suborder Apocrita ............................................................................................................. 10

1.5 Conclusions ....................................................................................................................... 13


1.6 Aims ................................................................................................................................... 14

CHAPTER 2: The first mitochondrial genome for the wasp superfamily Platygastroidea: the
egg parasitoid Trissolcus basalis ...................................................................................................... 17

2.1 Introduction ...................................................................................................................... 17


2.2 Materials and methods..................................................................................................... 18

DNA extraction, PCR amplification, and sequencing ................................................................ 19

Sequence analyses ...................................................................................................................... 20

Sequence alignment ................................................................................................................... 21

Phylogenetic analysis ................................................................................................................. 22

2.3 Results and discussion...................................................................................................... 23

Genome size and gene content ................................................................................................... 23


iii
Nucleotide content ..................................................................................................................... 25

Codon usage and amino acid composition ................................................................................. 25

Translation initiation and termination codons ............................................................................ 26

Transfer RNA genes ................................................................................................................... 27

Ribosomal RNA genes ............................................................................................................... 27

Non-coding regions .................................................................................................................... 28

Genome organization ................................................................................................................. 29

Phylogenetic analysis and considerations .................................................................................. 31

2.4 Conclusions ....................................................................................................................... 34

CHAPTER 3: Complete mitochondrial genomes of Ceratobaeus sp. and Idris sp. (Hymenoptera:
Scelionidae): shared gene rearrangements as potential phylogenetic markers at the tribal level
............................................................................................................................................................ 35

3.1 Introduction ...................................................................................................................... 35


3.2 Materials and methods..................................................................................................... 36

DNA extraction .......................................................................................................................... 36

Amplification, sequencing, and annotation of entire mt genomes ............................................. 37

Sequence alignment ................................................................................................................... 38

Genetic divergence and phylogenetic analysis ........................................................................... 39

3.3 Results and discussion...................................................................................................... 41

Genome size and gene content ................................................................................................... 41

Protein-coding genes .................................................................................................................. 42

Transfer RNA and Ribosomal RNA ........................................................................................... 43

Noncoding regions ..................................................................................................................... 44

Genome organizations ................................................................................................................ 46

Ka/Ks ratios among the three Scelionidae taxa .......................................................................... 48

iv
Phylogenetic analysis ................................................................................................................. 49

CHAPTER 4: Coexistence of minicircular and a highly rearranged mtDNA molecule suggests


that recombination shapes mitochondrial genome organization ................................................. 52

4.1 Introduction ...................................................................................................................... 52


4.2 Materials and methods..................................................................................................... 55

DNA extraction .......................................................................................................................... 55

Amplification, sequencing, and annotation of the entire mt genome ......................................... 55

Amplification and sequencing of minicircles and the maxicircle .............................................. 56

4.3 Results ............................................................................................................................... 57

The mt genome of Conostigmus sp. ........................................................................................... 57

Initial characterization of the putative maxicircle and minicircles of Conostigmus sp. ............. 58

Are the minicircles amplification artifacts? ............................................................................... 62

Elimination of the possibility of numts ...................................................................................... 64

4.4 Discussion .......................................................................................................................... 65

Importance of control reactions.................................................................................................. 65

Possible explanations for low copy number of the maxicircle ................................................... 66

Comparison with minicircles from other Metazoa ..................................................................... 66

Possible mechanism yielding the minicircle and maxicircle ...................................................... 68

Direct link between intra-mitochondrial recombination and mt gene rearrangement ................ 69

CHAPTER 5: Evolutionary dynamics of the mitochondrial genome in the Evaniomorpha


(Hymenoptera) – a group with an intermediate rate of gene rearrangement............................. 71

5.1 Introduction ...................................................................................................................... 71


5.2 Materials and Methods .................................................................................................... 74

DNA extraction .......................................................................................................................... 74

Mt genome amplification, sequencing, and annotation .............................................................. 75

Nucleotide composition, gene rearrangement, and repeat analyses ........................................... 76

v
Sequence alignment and phylogenetic analysis ......................................................................... 77

Rogue taxon identification ......................................................................................................... 78

5.3 Results and Discussion ..................................................................................................... 78

General features of mt genomes ................................................................................................. 78

Genome organizations ................................................................................................................ 80

The mt genome of Gasteruption sp. ........................................................................................... 83

The mt genomes of Megalyra sp. and O. pulchella ................................................................... 85

Evolutionary dynamics of evaniomorph mt genomes ................................................................ 87

The utility of gene rearrangements for assessing phylogeny in the Evaniomorpha ................... 89

Rates of gene rearrangement in the Evaniomorpha .................................................................... 90

Phylogenetic analysis ................................................................................................................. 90

CHAPTER 6: Higher-level phylogeny of the Hymenoptera inferred from mitochondrial


genomes ............................................................................................................................................. 97

6.1 Introduction ...................................................................................................................... 97


6.2 Materials and Methods .................................................................................................. 100

DNA extraction ........................................................................................................................ 100

Mt genome amplification, sequencing, annotation, and bioinformatic analysis ...................... 101

Sequence alignment and phylogenetic analysis ....................................................................... 102

6.3 Results and Discussion ................................................................................................... 103

General features of mt genomes ............................................................................................... 103

Genome organizations .............................................................................................................. 105

Phylogenetic analysis ............................................................................................................... 107

6.4 Conclusions ......................................................................................................................113

CHAPTER 7: General Conclusions...............................................................................................115

7.1 Main outcomes .................................................................................................................115


7.2 Recommendations and future research .........................................................................115

vi
7.2.1 Increasing taxon sampling of the Apocrita ................................................................115

7.2.2 Incorporating genome-based collection of markers ..................................................117

7.3 Conclusions ......................................................................................................................118

References ....................................................................................................................................... 120

vii
List of Tables

Table 1.1 Mt genomes (complete/near-complete) for the Hymenoptera available in GenBank

and unpublished in this study ..................................................................................................... 14

Table 2.1 List of taxa used in phylogenetic analyses. ................................................................ 22

Table 2.2 Locations and nucleotide sequence lengths of genes in the Trissolcus basalis mt

genome, the lengths of predicted amino acid sequences of protein-coding genes, and their

initiation and termination codons ............................................................................................... 24

Table 2.3 Nucleotide composition (%) of the Trissolcus basalis mt genome ............................ 25

Table 2.4 Codon usage for protein-coding genes of the Trissolcus basalis mt genome ............ 26

Table 3.1 Datasets used for phylogenetic analysis..................................................................... 39

Table 3.2 Estimation of Bayes Factors of different partition schemes. ..................................... 39

Table 3.3 Comparison of genetic divergences for all protein-coding genes among scelionds. . 48

Table 5.1 Species sequenced in this study, collection data and GeneBank accession numbers . 75

Table 5.2 The number of repeats (including direct and inverted repeats) within the A+T-rich

region and gene rearrangements identified for seven evaniomorph taxa ................................... 88

Table 5.3 Breakpoint numbers of the evaniomorph, hemipteroid and nematoceran mt genomes

relative to the ancestral mt genome organization. ...................................................................... 89

Table 6.1 Species sequenced in this study, collection data, and GeneBank accession numbers

.................................................................................................................................................. 101

Table 6.2 Characteristics of the three mitochondrial genomes sequenced in this study. ......... 104

Table 6.3 Summary of the major clades recovered by different dataset and analytical approach

.................................................................................................................................................. 107
viii
List of Figures

Figure 1.1 Map of the typical mitochondrial genome of animals ................................................ 2

Figure 1.2 Simple diagram illustrating the duplication/random loss model ................................ 6

Figure 1.3 The duplication/nonrandom loss model ..................................................................... 7

Figure 1.4 Model for the generation of mitochondrial gene inversions by recombination ......... 8

Figure 1.5 Major hymenopteran relationships proposed by Rasnitsyn (1988) .......................... 10

Figure 2.1 Mitochondrial genome organizations of four Proctotrupomorpha taxa ................... 30

Figure 2.2 Bayesian analysis of mitogenomic sequences based on the Muscle alignment dataset

.................................................................................................................................................... 31

Figure 3.1 Mitochondrial genome organization of three sceliond taxa, compared with the

ancestral pancrustacean mt genome organization ...................................................................... 42

Figure 3.2 Inferred secondary structures of tRNA genes .......................................................... 43

Figure 4.1 Organizations of the entire mt genome, maxicircle and minicircles that were initially

characterized by amplification ................................................................................................... 57

Figure 4.2 Mitochondrial genome organization of Conostigmus sp., compared with the

ancestral pancrustacean mt genome organization ...................................................................... 58

Figure 4.3 Size separations of amplicons on 1% agarose gels .................................................. 62

Figure 4.4 Model for the generation of minicircle and maxicircle mtDNA by recombination

during repair of DSBs. ............................................................................................................... 69

Figure 5.1 Simplified interpretation of the phylogenetic relationships among the major lineages

within the Apocrita from Rasnitsyn (1998) ................................................................................ 73

Figure 5.2 Structure of the A+T-rich region in the Ceraphron sp., Gasteruption sp., Megalyra
ix
sp. and Orthogonalys pulchella mt genomes ............................................................................. 81

Figure 5.3 Mitochondrial genome organization of two taxa from the Ceraphronoidea,

Ceraphron sp. (present study) and Conostigmus sp. (Mao et al. 2014) ..................................... 83

Figure 5.4 Mitochondrial genome organization of three taxa from the Evanioidea: Gasteruption

sp. (present study), Evania appendigaster (Wei et al., 2010b) and Pristaulacus compressus (Wei

et al., 2010b)............................................................................................................................... 85

Figure 5.5 Mitochondrial genome organization of Megalyra sp. (Megalyroidea) (present study)

and Orthogonalys pulchella (Trigonalyoidea) (present study) ................................................... 87

Figure 5.6 Bayesian analysis of the initial taxon dataset (without rogue taxon deletion) based

on mitogenomic sequences......................................................................................................... 93

Figure 5.7 Bayesian analysis of the reduced taxon dataset (with rogue taxon deletion) based on

mitogenomic sequences.............................................................................................................. 94

Figure 6.1 Mitochondrial genome organization of Ibalia leucospoides, Pelecinus polyturator,

and Monomachis antipodalis, compared with the ancestral pancrustacean/hymenopteran mt

genome organization ................................................................................................................ 105

Figure 6.2 Bayesian analysis of the Muscle alignment dataset ............................................... 109

x
List of Abbreviations

°C degrees Celsius

μl micro-litre

μm micro-metre

A adenine

a6 ATPase subunit 6

a8 ATPase subunit 8

BLAST Basic Local Alignment Search Tool

bp base pairs

C cytosine

cox1 cytochrome oxidase subunit I

cox2 cytochrome oxidase subunit II

cox3 cytochrome oxidase subunit III

cob cytochrome oxidase subunit b

DNA deoxyribonucleic acid

EDTA ethylendiaminetetraacetic acid

EST expressed sequenced tag

G guanine

GTR + G + I general time reversible model with rates invgamma

kb kilo base pair

Mb Mega base pairs

xi
MEGA Molecular Evolutionary Genetics Analysis

mg micro-gram

min minute

ml milli-litre

mM milli-molar

mt mitochondrial

nad1 NADH dehydrogenase subunit 1

nad2 NADH dehydrogenase subunit 2

nad3 NADH dehydrogenase subunit 3

nad4 NADH dehydrogenase subunit 4

nad4L NADH dehydrogenase subunit 4L

nad5 NADH dehydrogenase subunit 5

nad6 NADH dehydrogenase subunit 6

numts nuclear copies of mitochondrial DNA

ORF Open reading frame

PCR polymerase chain reaction

Poly T polythymidine

RNA ribonucleic acid

rRNA ribosomal RNA

rrnL gene for ribosomal RNA large-subunit

rrnS gene for ribosomal RNA small-subunit

s second

xii
T thymidine

TE Tris-EDTA

Tris tris (hydroxymethyl) aminomethane

tRNA transfer RNA

trnA transfer RNA gene for Alanine

trnC transfer RNA gene for cytosine

trnD transfer RNA gene for aspartic acid

trnE transfer RNA gene for glutamic acid

trnF transfer RNA gene for phenylalanine

trnH transfer RNA gene for histidine

trnK transfer RNA gene for lysine

trnL transfer RNA gene for leucine

trnM transfer RNA gene for metionine

trnN transfer RNA gene for asparagine

trnP transfer RNA gene for proline

trnR transfer RNA gene for arginine

trnS transfer RNA gene for serine

trnT transfer RNA gene for threonine

trnV transfer RNA gene for valine

trnW transfer RNA gene for triptophan

trnY transfer RNA gene for tyrosine

UOW University of Wollongong, Wollongong, NSW, Australia

xiii
Abstract

The Hymenoptera, comprising of sawflies, wasps, bees and ants, is the third largest order of insects.

It contains more than 155,000 described species, placed into two traditional suborders (Symphyta

and Apocrita). Many hymenopteran species play important roles in biological control, ecosystems

and production of commercial products. Unfortunately, higher-level relationships of the

Hymenoptera, particularly within the highly diverse parasitic lineages, remain unresolved in both

morphological and molecular studies.

The mitochondrial (mt) genome sequence is becoming a powerful tool for reconstructing

phylogenetic relationships. In addition, mt gene rearrangements also provide useful phylogenetic

information. However, deficient representation of a broad range of lineages restricts the evolutionary

utility of the mt genome in the Hymenoptera. For the four infraorders of the Apocrita, the Aculeata

and Ichneumonomorpha are reasonably sampled. By contrast, the Evaniomorpha and

Proctotrupomorpha are poorly represented. Therefore, this PhD research was focussed on increasing

our understanding of mt genomes of the Hymenoptera and their utility for the recovery of the

higher-level phylogeny by extending the taxonomic range of available mt genome sequences.

Eleven new mt genomes for representatives of four evaniomorph and four proctotrupomorph

superfamilies were determined. Several gene rearrangements were detected in each of the eleven mt

genomes when compared with the ancestral positions. In particular, protein-coding or rRNA gene

rearrangements were identified in six mt genomes. Mt genome organizations were also compared

within the Evaniomorpha and Proctotrupomorpha. The results showed that there were no shared,

derived gene rearrangements at the family level, indicating that the frequency of gene

xiv
rearrangements of these groups is too high to be useful for assessing relationships between families.

By contrast, comparison of mt genome organizations within the subfamily Scelioninae suggested

that gene rearrangements have much potential as phylogenetic markers at the subfamily level. The

large number of gene rearrangements identified in this study also facilitated the investigation of gene

rearrangement mechanisms. Direct and indirect evidence that supports the notion that recombination

is an important aspect of the gene rearrangement mechanism was identified. Most importantly, the

end products of recombination were detected (minicircles), in a megaspilid wasp Conostigmus sp.,

which provided support for the link between recombination and mt gene rearrangement. Furthermore,

a model of recombination which is important for our understanding of mtDNA evolution was

developed.

The first mt genome phylogeny of the Apocrita with a complete representation of superfamilies

(excluding Mymmaromatoidea) was reconstructed. The influence of inclusion/exclusion of 3rd codon

positions, alignment approaches, partition schemes and phylogenetic approaches on topology and

nodal support within the Hymenoptera was assessed. The results showed that the topologies were

sensitive to the variation of dataset and analytical approach. However, some robust and highly

supported relationships were recovered: the Ichneumonomorpha was monophyletic; the

Trigonalyoidea + Megalyroidea and the Diaprioidea + Chalcidoidea were consistently recovered; the

Cynipoidea was generally recovered as the sister group to the Diaprioidea + Chalcidoidea. In

addition, the monophyletic Aculeata and Proctotrupomorpha were recovered in some analyses.

Further phylogenetic analysis of the Hymenoptera will rely on increased sampling of both taxa and

sequence data.

xv
Acknowledgements

First of all, I would like to express my deepest appreciation to my principal supervisor, Professor

Mark Dowton. I have appreciated your guidance, great support, and encouragement throughout my

degree. Thank you for sharing your knowledge and laboratory techniques with me. I am also grateful

for your time and patience spent improving my writing skills in English as a second language.

Without your help, I would not have reached my dream of obtaining my PhD. I would also like to

thank my co-supervisor, Associate Professor James Wallman for his continual support and

encouragement. I feel so lucky and honoured to have had these two great supervisors. Gratitude also

goes to all the members of the Dowton lab, particularly Tracey Gibson, Angelique Hoolahan and

Kelly Meiklejohn, for their support, guidance and friendship.

I would like to express my sincere thanks to Andy Austin, Gary Gibson, Dave Smith, and John

Jennings for providing the biological specimens. A huge thank you also goes to Andy Austin,

Norman Johnson and Alejandro Valerio for providing the sequence data of Trissolcus basalis, and

Tracey Gibson for providing the sequence data of Ibalia leucospoides and Orthogonalys pulchella.

I would like to acknowledge the financial support from the Centre for Medical and Molecular

Bioscience, University of Wollongong.

Thanks to Xiushuai for his love, patience and support. It would not have been easy for me to

complete this thesis without his computing assistance. Finally, I would like to thank my wonderful

family, my mum, dad and younger sister. Thank you for your immense support and encouragement. I

appreciate everything you have done to help me to achieve my dream.


xvi
List of publications

1. Mao M, Valerio A, Austin AD, Dowton M, Johnson NF (2012) The first mitochondrial genome

for the wasp superfamily Platygastroidea: the egg parasitoid Trissolcus basalis. Genome

55:194-204.

2. Mao M, Austin AD, Johnson NF, Dowton M (2014) Coexistence of minicircular and a highly

rearranged mtDNA molecule suggests that recombination shapes mitochondrial genome

organization. Molecular Biology and Evolution 31:636-644.

3. Mao M, Dowton M (2014) Complete mitochondrial genomes of Ceratobaeus sp. and Idris sp.

(Hymenoptera: Scelionidae): shared gene rearrangements as potential phylogenetic markers at

the tribal level. Molecular Biology Reports doi: 10.1007/s11033-014-3522-x.

4. Mao M, Gibson T, Dowton M (2014) Evolutionary dynamics of the mitochondrial genome in

the Evaniomorpha (Hymenoptera) — a group with an intermediate rate of gene rearrangement.

Genome Biology and Evolution 6, 1862-1874.

5. Mao M, Gibson T, Dowton M Higher-level phylogeny of the Hymenoptera inferred from

mitochondrial genomes. Molecular Phylogenetics and Evolution (Under review).

xvii
xvi
ii
CHAPTER 1: General Introduction

1.1 Preamble

The Hymenoptera, comprising sawflies, ants, bees and wasps, is considered to be the third largest

order of insects, behind the Coleoptera and the Lepidoptera. There are more than 155,000 described

species in the Hymenoptera (Aguiar et al., 2013). The first fossil record of the Hymenoptera

appeared in the late Triassic, about 230 million years ago (Rasnitsyn, 2010). However, according to

recent divergence time estimation based on morphological and molecular data, the origin of the

Hymenoptera might date back to the Carboniferous, about 310 million years ago (Ronquist et al.,

2012b). Many hymenopterans play important ecological roles in the biological control of crop and

forest pests, the pollination of flowers and the production of commercial products, such as honey and

wax (LaSalle and Gauld, 1993).

Since most of this thesis has been published or submitted for publication, each chapter has been

written as a journal article. The chapters written as journal articles have been left in the general

submission format, except for minor changes for consistency. Consequently there is some repetition

between the General Introduction with the introduction of each chapter. To reduce the level of

repetition, the General Introduction chapter is not written as a comprehensive literature review, but

as a general introduction to the topics contained within the thesis. The literature is more intensely

reviewed in the introduction to each thesis chapter. Similarly, the General Conclusions chapter is not

written as a comprehensive discussion of the thesis chapters, as this would duplicate much of the

content in each chapter. Instead, the General Conclusions chapter outlines the main areas of future

research that the results of the thesis suggest. There is also some repetition among chapters of the

1
methodology used. The title page of each chapter details publication information. All references have

been compiled within a single list at the end of the thesis.

1.2 The mitochondrial genome and phylogenetics

The mitochondrion is an essential organelle present in most eukaryotic cells. It plays a central role in

energy metabolism, apoptosis, disease and aging (Boore, 1999). The mitochondrion contains its own

independent genome, known as the mitochondrial (mt) genome. In animals, the mt genome is a

double-stranded circular molecule (~16 kb in size), and usually consists of two ribosomal RNA

(rRNA) genes, 22 transfer RNA (tRNA) genes, along with 13 protein coding genes which code for

subunits of the enzymes required in electron transport or ATP synthesis (Moritz et al., 1987). There

is also generally a single large non-coding region (the A+T-rich region or control region) which is

known to contain regulatory elements for replication and transcription (Zhang and Hewitt 1997). The

37 genes distribute asymmetrically between the two strands, allowing the identification of a major (J)

and a minor (N) strand according to the number of encoded genes (Gissi et al., 2008) (Fig. 1.1).

Figure 1.1 Map of the typical mitochondrial genome of animals. The gene content on the major and minor DNA
strands are indicated, with the gene labels either external or internal to the molecule, respectively. tRNA genes are
indicated by single-letter amino acid codes, L1, L2, S1 and S2 denote trnLCUN, trnLUUR, trnSAGN and trnSUCN,
respectively. Adapted from Cook (2005).
2
Several features promote the mt genome as a good model for evolutionary genomic studies: (1) it is

easy to amplify due to the relatively small size and high copy number in the cell; (2) it is maternally

inherited in most species (Birky, 2001); (3) the nucleotide substitution rate of mtDNA is much higher

when compared with nuclear DNA (Lin and Danforth, 2004; Oliveira et al., 2008); (4) it lacks

recombination, so the whole molecule can be assumed to have the same genealogical history (Moritz

et al., 1987; Rokas et al., 2003). (5) In addition to the sequence data, some other mt genome

characters, such as gene rearrangements (Boore and Brown, 1998) and tRNA secondary structure

(Chen et al., 2014) also provide useful phylogenetic information.

However, the utility of mtDNA as a molecular marker is not consistently supported. Some

researchers have expressed their concerns about the usefulness of mtDNA. For example, mtDNA has

been suggested to be less informative than nuclear DNA to resolve deep relationships due to AT bias

and saturation (Burger et al., 2003). Furthermore, Galtier et al. (2009) argued that mtDNA is not

always clonal, that it does not evolve in a neutral manner and that it has a highly variable mutation

rate among lineages. These evolutionary properties of mtDNA might mislead phylogenetic

reconstructions (Galtier et al., 2009). Nevertheless, with the development of appropriate analytical

strategies and the rapid accumulation of sequenced mt genomes, most of these concerns can be

resolved. One such successful example of using mt genomes from a large taxon sampling for

phylogenetic reconstruction was the study of deuterostome phylogeny (Perseke et al., 2013). In this

study, more than 300 complete mt genomes were employed to deduce internal relationships within

the Deuterostomia. The results robustly recovered three chordata subgroups, Echinodermata and

their five subgroups, Enteropneusta and Ambulacraria. Similarly, a more recent phylogenetic study

of higher-level relationships within the Lepidoptera based on 106 mt genomes recovered topologies

3
that were highly congruent with nuclear and/or morphological studies (Timmermans et al., 2014). In

addition, several strategies have been developed to increase the performance of mitochondrial

phylogenetic analyses. For example, exclusion of 3rd codon positions or recoding them as purines

and pyrimidines (RY-coding) in the analysis can reduce the effect of saturation, and therefore recover

a more accurate topology (Cameron et al., 2012; Dowton et al., 2009a; Harrison et al., 2004).

Furthermore, appropriate data partitioning schemes and substitution models that account for

heterogeneity in evolutionary processes among sites are important to avoid artifactual relationships.

Recently, a program (named PartitionFinder) for simultaneously choosing partitioning schemes and

substitution models has been developed (Lanfear et al., 2012). This program can reduce systematic

error for large dataset as it allows optimal partitioning schemes to be efficiently and accurately

selected from predefined character sets (Lemmon and Lemmon, 2013). An example of using this

type of analytical strategy to improve the mitochondrial phylogeny was recently reported by Gibson

et al. (Gibson et al., 2005). In this study, Gibson et al. developed a customized three-state DNA

substitution model to reduce base composition bias in mammalian mt genomes. Phylogenetic trees

obtained using the new model showed increased congruence with studies of large sets of nuclear

genes. However, the performance of any particular analytical strategy can vary among different

lineages. For example, exclusion of 3rd codon positions has been found to improve the accuracy of

termite phylogenetic reconstructions (Cameron et al., 2012), while it did not appear to have any

effect on the phylogeny of the Orthoptera (Fenn et al., 2008). Therefore, extensive tests of the

effectiveness of different analytical strategies is crucial for accuracy when constructing a

mitochondrial phylogeny (Cameron, 2014).

1.3 Mitochondrial gene rearrangements


4
Mitochondrial gene rearrangements were generally considered to be rare evolutionary events in

animals (Boore, 1999). However, with more mt genomes from different groups available, the

diversity of mt genome organization has been shown to be much higher than expected. For example,

in insects, 13 orders have been found to possess gene rearrangements and the rate of gene

rearrangements are extremely high in the Hymenoptera (Dowton and Austin, 1999) and the

hemipteroid orders (Shao et al., 2001). The increasing number of gene rearrangements found in

different groups facilitates our understanding of molecular evolutionary processes and mechanisms

of gene rearrangement.

Gene rearrangements can be classified into four types: long-range translocation (across

protein-coding or rRNA genes), local inversion, remote inversion and local translocation (within a

tRNA gene cluster) (Dowton et al., 2003). Long-range translocation and inversion are rare in

vertebrates (but see Amer and Kumazawa, 2007; Inoue et al., 2001), while they are more common in

invertebrates (Dowton et al., 2003).

1.3.1 Mechanisms of gene rearrangement

Considerable effort has been devoted to explore the mechanisms of gene rearrangement. To date,

three main models have been proposed: the duplication/random loss model (Boore, 2000; Moritz et

al., 1987), the duplication/nonrandom loss model (Lavrov et al., 2002) and recombination (Dowton

and Campbell, 2001; Lunt and Hyman, 1997).

(1) The duplication/random loss model. During replication, a tandem duplication of part of the mt

genome may occur due to slipped-strand mispairing or imprecise termination. One copy of each

duplicated gene is likely to be rapidly eliminated or become nonfunctional due to the

accumulation of mutations. Therefore, the original gene arrangement is either restored or altered
5
depending on which copy of the genes is lost (Fig. 1.2). This model is supported by the

existence of pseudogenes and the position of intergenic spacers observed in some rearranged mt

genome organizations. It has long been considered as the principal mechanism of gene

rearrangement in vertebrates (San Mauro et al., 2006). Many translocations observed in

invertebrates are also consistent with this model. For example the rearrangement trnW-trnC →

trnC-trnW, which occurred in neuropteran mt genomes, can be well explained by this model

(Beckenbach and Stewart, 2009). However, this model cannot explain gene inversions and

long-range translocations.

Figure 1.2 Simple diagram illustrating the duplication/random loss model. A and B represent two adjacent genes.

(2) The duplication/nonrandom loss model. Duplication of the entire mt genome can result in a

dimeric molecule with the two monomers covalently linked head to tail. Such a dimeric

molecule would contain two copies of transcriptional promoters. Mutation to one copy of the

promoters would result in the genes under the control of the disabled promoter becoming

pseudogenes, and the sequences containing them would be free to be eliminated from the

genome. Therefore, the loss of genes would be predetermined by their transcriptional polarity.

After rearrangement, genes with identical transcriptional direction will be clustered in the mt

genome (Fig. 1.3). A typical example in insects that is consistent with this mechanism is the mt

genome of a winter crane fly, Paracladura trichoptera (Diptera) (Beckenbach, 2012). In this mt

genome, the genes are highly rearranged, and include protein-coding, rRNA and tRNA genes.

6
All these rearrangements fall into two main groups. Within each group, both the ancestral gene

order and transcriptional direction are maintained. The duplication/nonrandom loss model can

account for almost all gene rearrangements in Paracladura. However, as is the case with the

duplication/random loss model, this model cannot explain gene inversions and long-range

translocations.

Figure 1.3 The duplication/nonrandom loss model proposed by Lavrov et al. (2002) explains the gene
rearrangements (excepting trnC and trnT) observed in two millipedes. The proposed position for transcription
initiation and termination are marked by TI and TT. Genes are transcribed from left to right except when
underlined; underlining indicates the opposite transcriptional polarity. The copies of the genes that became
pseudogenes in a hypothetical intermediate arrangement are indicated by blue boxes. Adapted from Lavrov et al.
(2002).

(3) Recombination. A defining feature of this mechanism is the breakage and rejoining of mtDNA

strands (Fig. 1.4). Recombination had been considered to be absent in the animal mt genome

(Moritz et al., 1987; Satoh and Kuroiwa, 1991). However, with more effort devoted to mt

genome research in recent years, both direct and indirect evidence of animal mtDNA

recombination has been provided (Gantenbein et al., 2005; Kraytsberg et al., 2004; Ladoukakis

and Zouros, 2001; Lunt and Hyman, 1997; Tsaousis et al., 2005). Recombination can provide a

good explanation to the gene inversions and long-range translocations commonly observed in

7
invertebrate mt genomes. It is also supported by the non-tandem duplications present in some

vertebrate mt genomes. For example, the duplicated control regions in the mantellid frog mt

genomes were hypothesized to be produced by recombination (Kurabayashi et al., 2008).

Figure 1.4 Model for the generation of mitochondrial gene inversions by recombination. After double-strand
breakage at two neighbouring loci, rejoining of the broken ends can produce a genome with a short, inverted segment.
The linear fragment represents the minifragment that is deleted from the entire mt genome; the circle represents the
main fragment of the entire mt genome (excluding the minifragment). Adapted from Dowton et al. (1999).

1.3.2 Gene rearrangements as phylogenetic markers

Comparison of mitochondrial gene arrangements has been argued to be a powerful tool for deducing

evolutionary relationships mainly because (1) the enormous number of potential arrangements

makes convergence unlikely; (2) gene arrangements are selectively neutral (Boore and Brown, 1998).

It has been shown to be useful in resolving both shallow and deep evolutionary relationships across a

broad range of animal groups (Boore, 2006). For example, gene arrangement analysis of Sipuncula

taxa provided additional evidence for their close relationship with Annelida rather than Mollusca

(Boore and Staton, 2002; Mwinyi et al., 2009). Several analytical methods of gene rearrangement

have been developed in recent years, such as Crex (Bernt et al., 2007), TreeREx (Bernt et al., 2008),

BADGER (Larget et al., 2002) and UniMoG (Hilker et al., 2012). These methods facilitate the use of

gene rearrangements to infer phylogenetic relationships. However, gene rearrangement analysis is

8
still at a premature stage. Each of the available analytical methods has its limitations and has not

been extensively applied (Bernt et al., 2013). For example, gene duplications are commonly found in

rearranged mt genomes (Akasaki et al., 2005; Castro et al., 2006). However, most of the analytical

methods, such as Crex and TreeREx, cannot encode them. Furthermore, cases of convergent

evolution of mt gene organization have been occasionally reported (Dowton et al., 2009b; Flook et

al., 1995; Kilpert et al., 2012). Therefore, more taxa with gene rearrangements and more

sophisticated analytical methods are needed for detecting the potential of mt gene arrangements as

phylogenetic markers, as well as their performance and levels of homoplasy (Grande et al., 2008).

1.4 Higher-level phylogenetic relationships within the Hymenoptera

The Hymenoptera are traditionally divided into the Symphyta (without discernible waist) and

Apocrita (with wasp-waist) (Sharkey, 2007). The Symphyta are the most primitive members and

comprise about 7% of the Hymenoptera (Huber, 2009). Except for the parasitic Orussidae, all

symphytan families are herbivorous, with a fairly unspecialized, caterpillar-like lifestyle in larval

stage (Vilhelmsen, 2001). The Apocrita comprise about 93% of the Hymenoptera and are subdivided

into the Aculeata (including bees, ants and stinging wasps) and the Parastica (most of which are

parasitoids of insects and spiders) (Huber, 2009).

1.4.1 Suborder Symphyta

It has long been recognized that the Symphyta are paraphyletic with respect to the Apocrita (Sharkey,

2007) (Fig. 1.5). Both morphological and molecular evidence supports that the Orussoidea is sister

group to the Apocrita (Heraty et al., 2011; Rasnitsyn, 1988; Schulmeister, 2003b; Sharkey et al.,

2011; Vilhelmsen, 2001; Vilhelmsen et al., 2010). Several comprehensive phylogenetic studies based

9
on morphological and/or molecular characters have recovered many well-supported relationships:

the Xyelidae is sister group to all the remaining Hymenoptera; the clade Xiphydrioidea + Orussoidea

+ Apocrita is frequently recovered; and the Unicalcarida (all Hymenoptera except Xyelidae,

Tenthredinoidea and Pamphilioidea) is well supported (Ronquist et al., 1999; Schulmeister, 2003a, b;

Sharkey et al., 2011; Vilhelmsen, 2001).

Figure 1.5 Major hymenopteran relationships proposed by Rasnitsyn (1988). Adapted from Sharkey et al. (2007).

1.4.2 Suborder Apocrita

The Apocrita has long been recognized as a natural group (Sharkey, 2007). However, the

phylogenetic relationships within the Apocrita remain elusive. Neither morphological nor molecular

data have provided a convincingly resolved phylogeny. In 1988, Rasnitsyn presented the first

comprehensive, fully resolved phylogenetic hypothesis for the Apocrita. He suggested four major

infraorders based on morphological and fossil evidence: the Ichneumonomorpha, the Aculeata, the

Proctotrupomorpha, and the Evaniomorpha, with the Ichneumonomorpha sister to the Aculeata and

the Proctotrupomorpha sister to the Evaniomorpha (Rasnitsyn, 1988) (Fig 1.5). Although his analysis

was not performed in a strictly cladistic context, it sets a stage for subsequent hymenopteran

phylogenetic studies. Johnson’s (1988) analysis based on the midcoxal articulation characters

10
recovered the Stephanoidea as the most basal apocritan lineage and supported the remaining

Evaniomorpha as monophyletic (Johnson, 1988). The basal position of the Stephanoidea was

supported by most subsequent studies (Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al.,

2011; Vilhelmsen et al., 2010; Whitfield, 1992). Ronquist et al. (1999) recoded Rasnitsyn’s (1988)

data into a numerical cladistics matrix and conducted a parsimony analysis. However, only the

Ichneumonomorpha and Aculeata were recovered as monophyletic groups; the Evaniomorpha

appeared as a paraphyletic grade of basal apocritan lineages; the Proctotrupomorpha was also

paraphyletic due to the position of the Ceraphronoidea (Ronquist et al., 1999). A subsequent

reanalysis with a revised set of wing characters produced an even less resolved tree (Sharkey and

Roy, 2002). Recently, Vilhelmsen et al. (2010) conducted a comprehensive analysis based on 273

mesosomal characters scored for 89 taxa. The results recovered the Ichneumonomorpha and the

Aculeata as monophyletic and supported their sister relationship; the Evaniomorpha (excluding the

Stephanoidea) was also retrieved but placed as the basal apocritan lineage; the Proctotrupomorpha

failed to be recovered (Vilhelmsen et al., 2010). In a later study, Vilhelmsen analyzed the head

capsule characters and mapped these characters onto a phylogeny from Vilhelmsen et al. (2010).

Most of the characters provided further support at the superfamily or family level (Vilhelmsen,

2011).

Dowton and Austin (1994) provided the first comprehensive molecular phylogenetic analysis of the

Hymenoptera based on 16S rRNA data. The results supported the clade Ichneumonomorpha +

Aculeata and a monophyletic Proctotrupomorpha. A subsequent study with expanded taxon sampling

based on the same sequence data supported the above relationships and recovered a monophyletic

Evaniomorpha (excluding the Stephanoidea) (Dowton et al., 1997). Dowton and Austin (2001)

11
performed multiple analyses based on three genes (16S, COI and 28S) and morphological characters.

However, the relationships were sensitive to the analytical approaches and the inclusion of

morphological data (Dowton and Austin, 2001). More recently, Castro and Dowton (2006) added

18S rRNA sequences to the Dowton and Austin (2001) dataset and conducted Bayesian and

parsimony analyses. The results supported the Evaniomorpha as paraphyletic, with the Aculeata

nested within the Evaniomorpha and recovered a monophyletic Proctotrupomorpha (Castro and

Dowton, 2006). Two studies based on molecular data and one study based on molecular data and

morphological characters have been published recently (Heraty et al., 2011; Klopfstein et al., 2013;

Sharkey et al., 2011). These three studies consistently recovered the Aculeata nested within a

paraphyletic Evaniomorpha and generally supported a sister relationship between the

Ichneumonomorpha and the Proctotrupomorpha.

Several phylogenetic studies using mt genomes from the Hymenoptera have been published. In 2007,

Castro and Dowton performed a range of analyses to assess the ability of mt genome data to recover

a well-accepted phylogeny of five hymenopteran taxa. The results indicated that the outgroup

composition was important to recover the test phylogeny (Castro and Dowton, 2007). Dowton et al.

(2009) performed phylogenetic analyses of mt genomes from a much broader range of the

Hymenoptera using mutiple phylogenetic approaches. The analyses indicated that partitioned

Bayesian analysis of nucleotide data (excluding 3rd codon positions) recovered more

uncontroversial relationships. The results suggested that analysis of mt genomes holds promise for

the resolution of hymenopteran higher-level relationships (Dowton et al., 2009a). Wei and co-authors

published two papers to reconstruct the phylogeny of the Hymenoptera using mt genomes. The

monophyly of the Aculeata, Ichneumonomorpha and Proctotrupomorpha were robustly confirmed.

12
(Wei et al., 2014; Wei et al., 2010a). Some other studies also attempted to assess the utility of the mt

genome for phylogenetic analysis of the Hymenoptera (de Melo Rodovalho et al., 2014; Kaltenpoth

et al., 2012). However, a common drawback of the above studies is the limited taxon sampling,

especially the poorly represented Evaniomorpha and Proctotrupomorpha.

1.5 Conclusions

There has been much effort in generating a good taxon sampling and diversity of markers for

phylogenetic analysis of the Hymenoptera (Castro and Dowton, 2006; Heraty et al., 2011; Sharkey et

al., 2011). But with respect to the mitochondrial phylogeny, previous studies have not sampled the

diversity of the Hymenoptera well. So far, there are 38 mt genomes (this number excludes

congenerics) of the Hymenoptera available in GenBank. As shown in Table 1.1, the Aculeata (in blue

font) and Ichneumonomorpha (in green font) are reasonably sampled. By contrast, the

Evaniomorpha (in purple font – excludes highlighted taxa, which were sequenced in the present

study) and Proctotrupomorpha (red font) are poorly represented. To resolve higher-level

relationships within the Apocrita using mt genomes, there is a pressing need to sample entire mt

genomes from each apocritan superfamily. In addition, with an extended taxon sampling from the

Evaniomorpha and Proctotrupomorpha, more gene rearrangements are likely to be identified. This

will facilitate the investigation of the mechanisms of gene rearrangement and assessing the utility of

shared, derived gene rearrangements as phylogenetic markers in these two lineages. By

understanding the gene rearrangement mechanism, this will further lead to more realistic analyses of

gene rearrangement characters.

13
Table 1.1 Mt genomes (complete/near-complete) for the Hymenoptera available in GenBank and unpublished in this
study. The superfamilies from the same infraorder are shown in the same color. The mt genomes sequenced in this
study are highlighted by yellow.

Cephoidea Cephus
Orussoidea Orussus
Pamphilioidea
Siricoidea
Tenthredinoidea Monocellicampa Perga
Xiphydrioidea
Xyeloidea
Apoidea Apis Bombus Melipona philanthus
Chrysidoidea Cephalonomia Primeuchroeus
Vespoidea Atta Abispa Camponotus Leptomyrmex Polistes
Pristomyrmex Radoszkowskius Solenopsis Wallacidia Pristomyrmex
Ichneumonoidea Aphidius Cotesia Diachasmimorpha Diadegma Diadromus Enicospilus
Macrocentrus Meteorus Phanerotoma Spathius Venturia Macrocentrus
Ceraphronoidea Ceraphron Conostigmus
Evanioidea Evania Gasteruption Pristaulacus
Megalyroidea Megalyra
Stephanoidea Schlettererius
Trigonalyoidea Orthogonalys Taeniogonalys
Chalcidoidea Ceratosolen Nasonia Philotrypesis
Cynipoidea Ibalia
Diaprioidea Monomachis
Mymarommatoidea
Platygastroidea Ceratobaeus Idris Trissolcus
Proctotrupoidea Pelecinus Vanhornia

1.6 Aims

1. To more completely sample mt genomes within the Platygastroidea, and to

investigate mt gene rearrangements and mechanisms within this group (Chapters 2

and 3).

Mt gene positions were highly rearranged in the available Proctotrupomorpha mt genomes (Castro et

al., 2006; Oliveira et al., 2008; Xiao et al., 2011). To further investigate the degree of gene

rearrangements at different taxonomic levels and possible mechanism for these gene rearrangements,

14
I sequenced three sceliond taxa: Trissolcus basalis, Ceratobaeus sp. and Idris sp. These mt genomes

facilitated both closer (within subfamily) and more distant (between superfamilies) comparisons of

gene rearrangements.

2. To more completely sample mt genomes within the Evaniomorpha, and to

investigate phylogeny and mt gene rearrangements and mechanisms within this

group (Chapters 4 and 5).

Evolutionary dynamics of mt gene rearrangements are poorly studied mainly due to a lack of

suitable model systems. To investigate whether the Evaniomorpha might provide such a model

system, I sequenced five new mt genomes for representatives of four evaniomorph superfamilies,

Megalyra sp. (Megalyroidea), Orthogonalys pulchella (Trigonalyoidea), Ceraphron sp.

(Ceraphronoidea), Conostigmus sp. (Ceraphronoidea) and Gasteruption sp. (Evanioidea). A mt

genome phylogeny was also reconstructed. In addition, recombination has been considered as an

important mechanism for invertebrate mt gene rearrangements. The Conostigmus sp. mt genome was

examined for evidence of recombination.

3. To further sample Proctotrupomorpha mt genomes, in order to obtain

representative mt genomes from every apocritan superfamily (excluding

Mymmaromatoidea), and to conduct the first mt genome phylogeny of the

Apocrita with a complete representation of superfamilies (Chapter 6)

The analysis of mt genomes holds promise to resolve higher-level relationships within the Apocrita.

However, a comprehensive mt genome phylogeny of the Apocrita has not been previously reported

due to limited taxon sampling. In an attempt to increase our understanding of the apocritan

phylogeny, I further sequenced three Proctotrupomorpha mt genomes: Ibalia leucospoides

15
(Cynipoidea), Monomachis antipodalis (Diaprioidea) and Pelecinus polyturator (Proctotrupoidea).

With these new genomes, the taxon dataset represented every apocritan superfamily (excluding

Mymmaromattoidea). I analyzed this dataset using a range of phylogenetic analytical approaches.

16
CHAPTER 2: The first mitochondrial genome for the
wasp superfamily Platygastroidea: the egg parasitoid
Trissolcus basalis

This chapter is slightly modified from the paper:

Mao M, Valerio A, Austin AD, Dowton M, Johnson NF (2012) The first mitochondrial genome for

the wasp superfamily Platygastroidea: the egg parasitoid Trissolcus basalis. Genome 55:194-204.

doi: 10.1139/g2012-005.

2.1 Introduction

Mitochondrial (mt) genomes have been widely used for the reconstruction of phylogenetic

relationships at different taxonomic levels (Cameron et al., 2007b; Kim et al., 2009; Li et al., 2011;

Miya et al., 2003). In insects, mtDNA is typically a double-stranded circular molecule of about 16 kb

in size. It contains 13 protein-coding genes, 2 ribosomal RNA (rRNA) genes, and 22 transfer RNA

(tRNA) genes (Boore, 1999). Additionally, there is a major non-coding region, known as the

A+T-rich region or control region, which plays a role in initiation of transcription and replication

(Zhang and Hewitt, 1997).

The Hymenoptera is one of the most species-rich and biologically diverse insect orders (LaSalle and

Gauld, 1993). Although much progress has been made to investigate the phylogeny of the

Hymenoptera (Castro and Dowton, 2006; Dowton and Austin, 1994, 2001; Heraty et al., 2011;

Ronquist et al., 1999; Sharkey et al., 2011), there are also many relationships remaining unresolved.

For example, the monophyly of the Proctotrupomorpha (sensu Rasnitsyn 1988) is well supported by

recent studies, but the relationships among the proctotrupomorph families are still controversial

17
(Castro and Dowton, 2006; Dowton and Austin, 2001; Heraty et al., 2011; Rasnitsyn and Zhang,

2010). Several analyses have demonstrated that the whole mt genome is a useful tool to resolve

hymenopteran evolutionary relationships (Cameron et al., 2008; Castro and Dowton, 2007; Dowton

et al., 2009a; Wei et al., 2010a). To date, 18 complete mt genomes and 15 nearly complete mt

genomes have been successfully sequenced in the Hymenoptera, with 18 taxa added in the last 3

years. This large amount of data holds promise to improve phylogenetic analysis. Notwithstanding,

there remain shortcomings with this dataset, with only 11 superfamilies represented. For

approximately half of the hymenopteran superfamilies, there is no mt genome data available.

In this paper, we provide the mt genomic sequence of Trissolcus basalis (Wollaston) from the

superfamily Platygastroidea, the third largest of the parasitic superfamilies after the Ichneumonoidea

and Chalcidoidea (Austin et al., 2005). No mt genome sequences for the Platygastroidea have

previously been published. The characteristics of this mt genome, with respect to genome

composition, nucleotide content, and codon usage, were compared with other hymenopteran mt

genomes. We conducted phylogenetic analyses of mt genomes from 29 hymenopteran taxa and 7

other orders of holometabolous insects. Bayesian analyses were performed using nucleotide

sequences (including and excluding the third codon position). The performance of two alignment

methods (Clustal W and Muscle) was also compared. Another goal of this research was to further

explore the evolution of mt gene rearrangement. The hymenopteran mt genome has been reported to

rearrange more frequently than those from other insect orders (Crozier and Crozier, 1993; Dowton

and Austin, 1999; Dowton et al., 2009b). Here we compared the genome structures of four

Proctotrupomorpha taxa.

2.2 Materials and methods


18
DNA extraction, PCR amplification, and sequencing

454 sequencing. Genomic DNA was extracted from whole bodies of 25 adult male specimens of T.

basalis from a laboratory culture maintained by F. Bin at the Universitàdi Perugia. The extract was

sent to the DNA Sequencing Facility at the University of Pennsylvania Perelman School of Medicine.

Five full plates were run on a Roche/454 GS FLX sequencer (454 Life Sciences, Branfort, Conn.,

USA) using Titanium chemistry. This resulted in 5,080,113 reads of a total of 1,535,920,544 bases.

These reads were assembled using gsAssembler 2.5.3 with a minimum overlap of 30 bases and the

complex genome assembly option.

Illumina sequencing. To address the issue of possible homopolymer errors in 454 sequencing,

genomic DNA was extracted from an additional five female specimens from the same culture. The

extract was processed using the Nextera DNA sample preparation kit from Epicentre

Biotechnologies (Madison, Wis., USA) following the manufacturer’s instructions. The sample was

submitted to the Nucleic Acid Shared Resource of the College of Medicine at The Ohio State

University. A single lane of 51 base reads was run on an Illumina Genome Analyzer IIx (Illumina,

San Diego, Calif., USA) resulting in 29,780,645 reads and a total of 1,518,812,895 bases. These

reads were assembled using VCAKE 1.5 into repetitive and non-repetive contigs, and short contigs

(< 70 base) were then removed. The remaining contigs were then assembled together with the final

454 assembly using gsAssembler 2.5.3. Nesoni 0.42 software was used to correct errors in the hybrid

assembly.

Contigs of mt origin from the entire assembly were identified by a BLAST search using known

hymenopteran mt proteins. Two contigs emerged from the search, with lengths of 8,067 bases and

19
7,719 bases.

Sequence analyses

The two contigs were imported into Bioedit version 7.0.5 (Hall, 1999). Visual inspection revealed

that the same dinucleotide repeat, (TA)n was present at one end of each of the contigs. When the

contigs were oriented in the ancestral orientation (i.e. with protein-coding genes oriented on the

same strand and direction as the ancestral pancrustacean), the shorter contig had a (TA)13 repeat at

the 3’ end, while the longer contig had a (TA)9 repeat at the 5’ end. We interpret this as indicating

that a dinucleotide repeat that varies in length is situated between these two contigs. Joining these,

the ancestral orientation of all protein-coding and rRNA genes is maintained. The dinucleotide repeat

is situated in the intergenic space, between the trnH and nad4 genes.

Nineteen tRNA genes were identified by tRNA-Scan SE version 1.21

(lowelab.ucsc.edu/tRNAscan-SE/) (Lowe and Eddy, 1997), specifying mitochondrial/chloroplast

DNA as the source and choosing the invertebrate mitochondrial genetic code for tRNA isotype

prediction. The cove cutoff score was set to 5 in order to avoid missing tRNA genes with lower

cutoff scores. Two tRNA genes (trnN and trnS1) failed to be detected by tRNA-Scan SE, but they

were identified by visually comparing unassigned regions of the genome with previously determined

tRNA counterparts from the Hymenoptera. Thirteen protein-coding genes were identified using an

open reading frame (ORF) finder (www.ncbi.nlm.nih.gov/gorf/orfig.cgi), specifying the invertebrate

mt genetic code. The initiation and termination codons of some genes were identified according to

the boundaries of tRNA genes and comparison with other insect mt sequences. The precise

boundaries of rRNA genes are difficult to define. They were assumed to be bounded by the

20
neighboring tRNA genes (Beckenbach, 2011; Beckenbach and Stewart, 2009; Dowton et al., 2009a).

However, the boundary between rrnL and rrnS was difficult to define in T. basalis, as the trnV that is

usually between the two rRNA genes is not present here. Instead, we determined this boundary by

aligning the rRNA regions from T. basalis with other closely related insect rrnL and rrnS genes.

Sequence alignment

Mitogenomic sequences from 29 hymenopteran taxa and 7 taxa from other holometabolous insects

were obtained from GenBank (Table 2.1). Nucleotide sequences for each of the 13 protein-coding

genes and the 2 rRNA genes were imported into separate files using MEGA version 5.05 (Tamura et

al., 2011) and aligned using Clustal W (Thompson et al., 1994) or Muscle (Edgar, 2004) as

implemented within MEGA 5. For the protein-coding genes (excluding the stop codons), the

nucleotide sequences were translated into the amino acid sequences using the invertebrate mt genetic

code, and the amino acid sequences were aligned using Clustal W or Muscle. MEGA 5 then creates a

nucleotide alignment using the amino acid sequences as a guide. The alignment parameters were the

default settings for all genes. The Clustal W alignment parameters were as specified in Cameron et al.

(2008). The Muscle alignment parameters for the rRNA genes were gap open penalty = -400, gap

extend penalty = 0; and for the protein-coding genes were gap open penalty = -2.9, gap extend

penalty = 0. For both types of gene, the maximum memory allocated was 2,047 MB, with a

maximum of eight iterations. The clustering method used for all iterations was UPGMB, with the

length of the minimum diagonal set to 24. Following alignment, individual genes were concatenated

prior to phylogenetic analysis.

21
Table 2.1 List of taxa used in phylogenetic analyses.

Order Family Genus GenBank Accession Referance


Coleoptera Tenebrionidae Tribolium AJ312413 Friedrich and Muqim 2003
Diptera Ceratopogonidae Culicoides AB361004 Matsumoto et al. 2009
Mecoptera Bittacidae Bittacus HQ696578 Beckenbach 2011
Megaloptera Sialidae Sialis FJ859905 Cameron et al. 2009
Neuroptera Ascalaphidae Ascaloptynx FJ171324 Beckenbach and Stewart 2009
Lepidoptera Tortricidae Adoxophyes NC_008141 Lee et al. 2006
Raphidioptera Raphidiidae Mongoloraphidia FJ859902 Cameron et al. 2009
Hymenoptera
Suborder
Symphyta
Cephoidea Cephidae Cephus NC_012688 Dowton et al. 2009a, b
Orussoidea Orussidae Orussus NC_012689 Dowton et al. 2009a, b
Tenthredinoidea Pergidae Perga AY787816 Castro and Dowton 2005
Apocrita
Apoidea Apidae Apis NC_001566 Crozier and Crozier 1993
Apidae Bombus DQ870926 Cha et al. 2007
Apidae Melipona NC_004529 Silvestre et al. 2008
Chalcidoidea Agaonidae Ceratosolen AH016005 Chen et al. unpublished
Pteromalidae Nasonia NW_001815691 Oliveira et al. 2008
Chrysidoidea Bethylidae Cephalonomia FJ823227 Wei et al. unpublished
Chrysididae Primeuchroeus AH015389 Castro et al. 2006
Evanioidea Evaniidae Evania NC_013238 Wei et al. 2010b
Ichneumonoidea Braconidae Aphidius GU097658 Wei et al. 2010a
Braconidae Cotesia NC_014272 Wei et al. 2010a
Braconidae Diachasmimorpha GU097655 Wei et al. 2010a
Braconidae Macrocentrus GU097656 Wei et al. 2010a
Braconidae Meteorus GU097657 Wei et al. 2010a
Braconidae Phanerotoma GU097654 Wei et al. 2010a
Braconidae Spathius NC_014278 Wei et al. 2010a
Ichneumonidae Diadegma NC_012708 Wei et al. 2009
Ichneumonidae Enicospilus FJ478177 Dowton et al. 2009a, b
Ichneumonidae Venturia FJ478176 Dowton et al. 2009a, b
Platygastroidea Scelionidae Trissolcus JN903532 Present study
Proctotrupoidea Vanhorniidae Vanhornia NC_008323 Castro et al. 2006
Stephanoidea Stephanidae Schlettererius FJ478175 Dowton et al. 2009a, b
Vespoidea Formicidae Pristomyrmex NC_015075 Hasegawa et al. 2011
Formicidae Solenopsis NC_014669 Gotzek et al. 2010
Mutillidae Radoszkowskius NC_014485 Wei and Chen unpublished
Vespidae Abispa NC_011520 Cameron et al. 2008
Vespidae Polistes EU024653 Cameron et al. 2008

Phylogenetic analysis

Previous phylogenetic studies using hymenopteran mt genome sequences indicated that nucleotide

data analyses were superior to amino acid data analyses, and that Bayesian analyses were superior to

22
parsimony analyses (Castro and Dowton, 2007; Dowton et al., 2009a). For this reason, we employed

Bayesian analyses in this study. The analyses were conducted using MrBayes version 3.1.2

(Ronquist and Huelsenbeck, 2003) at the freely available Bioportal server (www.bioportal.uio.no).

The two datasets (one aligned using Clustal W, the other aligned using Muscle) were divided into

four partitions: the first, second, and third codon positions and rRNA genes. The GTR+G+I model

(nst = 6, rate = invgamma) was chosen as the best-fit model for all of the partitions (MrModelTest

2.3, Nylander et al., 2004). Two analyses (all codon positions included or the third codon positions

excluded) were performed for the two alignment datasets. The Markov chain Monte Carlo (MCMC)

process was set so that four chains (three heated and one cold) ran simultaneously. For each analysis,

we conducted four independent runs for 1,000,000 generations, with trees being sampled every 100

generations. Burn-in was discarded according to the plot of generation against the likelihood scores

and the sump command in MrBayes. The percentage of trees recovering a particular clade was used

as a measure of that clade’s posterior probability (Huelsenbeck and Ronquist, 2001).

2.3 Results and discussion

Genome size and gene content

The almost complete mt genome sequence of T. basalis is 15,768 bp in size (Fig. 2.1). We were not

able to obtain the sequence of part of the non-coding region. Although the sequencing strategy

should have sequenced the non-coding region just as frequently as the coding regions, the inability to

identify fragments that could be added to our mt contigs is possibly due to the presence of repeats.

We did attempt to amplify the noncoding region by conventional PCR, but we were unable to obtain

amplicons, as has been reported for a number of hymenopteran mt genomes (Castro and Dowton,

23
2005; Castro et al., 2006; Oliveira et al., 2008). We were able to identify 36 of the 37 genes usually

found in animal mt genomes (Table 2.2); 13 protein coding genes, 21 of the 22 tRNA genes, 2 rRNA

genes, and part of the A+T-rich region. Several short non-coding regions were identified. The longest

of these is 60 bp long, and it is located between the trnS1 and nad5 genes. There are five genes that

overlap: two between protein-coding genes (a8 and a6, and nad4 and nad4L) and three between

tRNA genes (trnD and trnK, trnN and trnE, and trnT and trnP).

Table 2.2 Locations and nucleotide sequence lengths of genes in the Trissolcus basalis mt genome, the lengths of
predicted amino acid sequences of protein-coding genes, and their initiation and termination codons.

Gene/region Nucleotide position Size Codon


No. of nucleotides No. of amino acids Initiation Termination
trnC 2-62 61
trnY 82-147 66
trnQ 178-249 73
trnI 253-318 66
trnA 322-387 66
nad2 447-1425 979 327 ATA T
trnW 1426-1489 64
cox1 1490-3025 1536 512 ATA TAA
trnL2 3040-3109 70
cox2 3110-3787 678 226 ATT TAA
trnD 3790-3858 69
trnK 3856-3927 72
a8 3928-4095 168 56 ATT TAA
a6 4079-4737 659 220 ATG TA
cox3 4738-5530 793 265 ATG T
trnG 5531-5598 68
nad3 5605-5949 345 115 ATA TAA
trnN 5952-6019 68
trnE 6017-6080 64
trnF 6084-6150 67
trnS1 6154-6212 59
nad5 6275-7984 1710 570 ATT TAA
trnH 7985-8049 65
nad4 8080-9429 1350 450 ATG TAA
nad4L 9423-9695 273 91 ATA TAG
trnT 9709-9774 66
trnP 9774-9837 64
nad6 9839-10428 590 197 ATA TA
cob 10429-11565 1137 379 ATG TAA
trnS2 11581-11650 70
nad1 11655-12566 912 304 ATA TAA
trnL1 12573-12638 66
rrnL 12639-13908 1270
rrnS 13909-14660 752
trnM 14661-14727 67
trnV 14729-14793 65

24
Nucleotide content

As has been widely published in other insect mt genomes (Simon et al., 1994), the nucleotide

content of T. basalis is heavily biased towards adenine and thymine (Table 2.3). The entire sequence

has 43.7% A, 9.8% C, 40.5% T, and 6.0% G. The total A+T content is 84.2%, which is typical of

other hymenopteran mt genomes. For example, it is almost as high as that of Apis mellifera (84.9%:

Crozier and Crozier, 1993) but higher than that of Perga condei (78%: Castro and Dowton, 2005)

and Vanhornia eucnemidarum (80.1%: Castro et al., 2006).

For the mt protein-coding genes of T. basalis, the A+T content is a little lower than that of the entire

mt genome sequence. However, the first and second codon positions have a lower A+T content

(78.9% and 75.4%, respectively), while the third codon position shows a much higher A+T content

(92.3%). Conversely, both the rRNA and tRNA genes have a higher A+T content (87.9% and 88.7%,

respectively) when compared with the average for the genome. The nucleotide content bias is

inferred to result from mutational pressure (Asakawa et al., 1991; Foster et al., 1997).

Table 2.3 Nucleotide composition (%) of the Trissolcus basalis mt genome.

Nucleotide Length (bp) A C T G A+T


Entire sequence 15768 43.7 9.8 40.5 6.0 84.2
Protein-coding sequence 11130 36.5 9.1 45.7 8.6 82.2
J-strand 6885 38.5 11.4 42.8 7.4 81.3
N-strand 4245 33.4 5.5 50.5 10.6 83.9
Codon position
1st 3710 40.9 8.7 38.0 12.4 78.9
2nd 3710 23.6 14.0 51.8 10.6 75.4
3rd 3710 45.0 4.8 47.3 2.9 92.3
Ribosomal RNA gene sequences 2022 42.7 4.0 45.1 8.1 87.9
Transfer RNA gene sequences 1396 46.4 4.7 42.3 6.5 88.7
A+T-rich
J-strand region 974
1002 42.8
46.6 5.1
5.3 47.0
42.1 5.0
5.9 89.8
88.7
N-strand 394 45.9 3.3 42.6 7.9 88.5

Codon usage and amino acid composition

25
The codon usage in the T. basalis mt genome is highly skewed towards codons that are high in A+T

content (Table 2.4). Codons such as CTG (Leu) and CGC (Arg), encoded by both C- and G-rich

codons, are rarely used. Conversely, the AT-rich codons ATT (Ile), TTA (Leu), TTT (Phe), and ATA

(Met) are the four most frequently used codons (465, 460, 407, and 341 times, respectively). Within

a particular synonymous codon family, AT-rich codons are predominantly used. For example, leucine

can be coded by six alternative codons, but TTA is used much more frequently than any other. The

relative synonymous codon usage (RSCU: Nei and Kumar, 2000) for TTA is 5.23. The RSCU

measures the relative proportion that a codon is used within a particular synonymous codon family –

the maximum RSCU value for the leucine codon family is 6. For each codon family, the sum of the

A- and T-rich codons is close to the maximum possible RSCU value. Similar codon usage bias has

been observed in other hymenopteran mt genomes (Castro et al., 2006; Cha et al., 2007; Wei et al.,

2009).

Table 2.4 Codon usage for protein-coding genes of the Trissolcus basalis mt genome.

codon aa nc RSCU codon aa nc RSCU codon aa nc RSCU codon aa nc RSCU


TTT Phe 407 1.92 TCT Ser 92 2.45 TAT Tyr 161 1.69 TGT Cys 25 1.85
TTC Phe 18 0.08 TCC Ser 9 0.24 TAC Tyr 30 0.31 TGC Cys 2 0.15
TTA Leu 460 5.23 TCA Ser 102 2.71 TAA * 12 1.85 TGA Trp 79 1.93
TTG Leu 15 0.17 TCG Ser 4 0.11 TAG * 1 0.15 TGG Trp 3 0.07
CTT Leu 28 0.32 CCT Pro 41 1.45 CAT His 50 1.64 CGT Arg 8 0.84
CTC Leu 1 0.01 CCC Pro 17 0.6 CAC His 11 0.36 CGC Arg 1 0.11
CTA Leu 23 0.26 CCA Pro 55 1.95 CAA Gln 54 1.93 CGA Arg 26 2.74
CTG Leu 1 0.01 CCG Pro 0 0.0 CAG Gln 2 0.07 CGG Arg 3 0.32
ATT Ile 465 1.88 ACT Thr 48 1.52 AAT Asn 252 1.82 AGT Ser 20 0.53
ATC Ile 30 0.12 ACC Thr 11 0.35 AAC Asn 25 0.18 AGC Ser 0 0.0
ATA Met 341 1.92 ACA Thr 66 2.1 AAA Lys 155 1.9 AGA Ser 70 1.86
ATG Met 15 0.08 ACG Thr 1 0.03 AAG Lys 8 0.1 AGG Ser 4 0.11
GTT Val 55 1.93 GCT Ala 34 1.92 GAT Asp 46 1.74 GGT Gly 17 0.45
GTC Val 4 0.14 GCC Ala 7 0.39 GAC Asp 7 0.26 GGC Gly 4 0.11
GTA Val 48 1.68 GCA Ala 27 1.52 GAA Glu 61 1.74 GGA Gly 99 2.61
GTG Val 7 0.25 GCG Ala 3 0.17 GAG Glu 9 0.26 GGG Gly 32 0.84
Note: aa, amino acid; nc, number of codons; RSCU, relative synonymous codon usage; *, stop codon.

Translation initiation and termination codons

26
Thirteen protein-coding genes were identified using an ORF finder. Table 2.2 shows the position of

these genes in the mt genome of T. basalis, their lengths, and translation initiation and termination

codons. Conventional ATA, ATT, or ATG initiation codons are assigned to all of the 13

protein-coding genes. Nine genes (a8, cob, cox1, cox2, nad1, nad3, nad4, nad4L, and nad5) are

predicted to use the complete termination codons TAA or TAG. Four other genes have incomplete

termination codons: TA for a6 and nad6, and T for nad2 and cox3. Incomplete termination codons

have been commonly reported for other insect species (Clary and Wolstenholme, 1985; Crozier and

Crozier, 1993). It has been proposed that the complete termination codon (TAA) is created by

polyadenylation after transcription (Ojala et al., 1981).

Transfer RNA genes

Twenty-one tRNA genes (59-73 bp in size) were identified in the T. basalis mt genome. Fourteen

tRNA genes are encoded on the J-strand (the strand encoding the majority of genes, as defined by

Simon et al 1994), with 7 encoded on the N-strand (the minority strand). Their predicted secondary

structures are typical cloverleaf structures, except for trnS1 (Supplementary Fig. 1,

http://nrcresearchpress.com/doi/suppl/10.1139/g2012-005). The D-stem in the dihydrouridine (DHU)

arm is absent in the trnS1 gene, which has been reported in other hymenopteran species (Castro et al.,

2006; Wei et al., 2010b). In addition, mismatched base pairs appear in the DHU stem and acceptor

stem, which is common in the insect tRNA genes (Shao and Barker, 2003; Zhou et al., 2009). The

anticodons are identical to their counterparts in most other published hymenopteran mt genomes.

Ribosomal RNA genes

The small and large ribosomal RNA genes (rrnS and rrnL) are located on the N-strand. Unlike other

27
hymenopteran mt genomes, they are directly adjacent to each other, without the trnV gene between

them. The rrnS gene is 752 bp in size with an A+T content of 86.2%. The rrnL gene is 1270 bp in

size with an A+T content of 88.8%. These sizes are similar when compared with other Hymenoptera.

Based on the range of hymenopteran taxa included in Table 2.1, rrnS ranges from 600 to 811

nucleotides, while rrnL ranges from 868 to 1541 nucleotides. The A+T content of rrnS is similar

when compared with other hymnopteran taxa (Table 2.1), in which the A+T content ranges from

74.1% to 91.2%. The A+T content of rrnL is at the higher end of the range; it is only lower than that

of Cotesia (89.72%) and Macrocentrus (89.55%).

Non-coding regions

The longest non-coding region (partial A+T-rich region) of T. basalis is 974 bp in size, which is

located between the rrnS and trnC genes (Fig. 2.1). The A+T content is 89.8%, which is not

markedly higher compared with the remainder of the genome. There are several other short

non-coding regions ranging in size from 2 to 60 bp, which are located throughout the genome.

Interestingly, the 60 bp non-coding region has the anticodon of trnR (the tRNA gene that could not

be found) and the length is similar to the tRNA genes, but it could not be folded into a clover-leaf

structure unless overlapping more than 30 bp with nad5. If this section is ascribed to the trnR gene,

the slightly shorter nad5 gene lacks a stop codon (either complete or incomplete). Recently, some

truncated tRNA genes were reported in the mt genomes of velvet worms, with tRNA editing

converting the primary transcripts to conventional tRNA structures (Segovia et al., 2011). It is

conceivable that the missing trnR gene is present in this non-coding region, and it is completed by a

similar mechanism. We also investigated whether trnR was present in other non-coding regions,

including the trnA-nad2 junction, but we could find no regions that could be folded into a trnR gene.
28
Genome organization

The mitochondrial genome organization of T. basalis is shown in Fig. 2.1. All of the protein-coding

and rRNA genes have identical positions and transcriptional directions when compared with the

organization of the ancestral pancrustacean (Cook, 2005; Crease, 1999) (Fig. 2.1), which (as far as is

known) is the same organization as the ancestral hymenopteran (Dowton et al., 2009b). However, the

positions of a number of tRNA genes differ from the ancestral organization. For example, the

ancestral positions of trnC, trnY, and trnW are between the nad2 and cox1 genes, while in T. basalis,

trnC and trnY are between nad2 and the non-sequenced region (presumably the non-coding region).

Similarly, trnA, trnD, trnI, trnM, trnQ, trnR, and trnS1 are not in the ancestral positions.

We then compared the organization of the T. basalis mt genome with three other hymenopteran taxa

(all from the Proctotrupomorpha), to investigate whether any of the tRNA gene rearrangements were

shared with close allies. The Proctotrupomorpha have been consistently recovered as a monophyletic

group in molecular phylogenetic studies (Castro and Dowton, 2006; Dowton and Austin, 2001;

Heraty et al., 2011). Shared gene rearrangements are considered reliable indicators of shared

ancestry (Boore et al., 1995), and they have been used to infer phylogeny in the Hymenoptera

(Dowton, 1999), although examples of homoplastic gene rearrangements have been reported

(Dowton and Austin, 1999; Flook et al., 1995). The mt genome organizations of these three

additional taxa from the Proctotrupomorpha are shown in Fig. 2.1 to facilitate comparisons. All of

the protein-coding and rRNA genes are positioned identically in T. basalis and Vanhornia. In the

available mitogenomic sequences of Ceratosolen and Nasonia, Ceratosolen has an inversion of at

least five protein-coding genes (Xiao et al., 2011), while Nasonia has an inversion of six

protein-coding genes and a translocation of one protein-coding gene (Oliveira et al., 2008). With
29
respect to the position of the larger protein-coding and rRNA genes, there are three blocks of genes

that are present in all four taxa: cox1-cox2-a8-a6-cox3, nad4-nad4L-nad6-cob, and

nad1-rrnL-rrnS. However, despite the conservation of the organization of the larger genes in these

blocks, the tRNA genes are not conserved, with only three tRNA genes remaining in their ancestral

locations in all four taxa (trnH, trnT, and trnP). The trnS2 gene is present in its ancestral locations in

three Proctotrupomorpha taxa, with its positions unknown in Nasonia. All remaining tRNA genes

have rearranged in one or more taxa. Interestingly, the trnA gene is in the same derived position

(next to the rrnS gene) in both Ceratosolen and Nasonia. These are the only two included

representatives from the Chalcidoidea – this gene rearrangement may thus be a synapomorphy for a

subset of the Chalcidoidea.

Figure 2.1 Mitochondrial genome organizations of four Proctotrupomorpha taxa, compared with the ancestral
pancrutacean mt genome organization (Crease 1999, Cook 2005). tRNA genes are indicated by single-letter amino
acid codes L1, L2, S1, and S2, and denote trnLCUU, trnLUUR, trnSAGN, and trnSUCN, respectively. Genes are transcribed
from left to right except those indicated by underlining. Gene movements, relative to the ancestral organization, are
indicated with arrows.
30
Phylogenetic analysis and considerations

Phylogenetic analyses were performed using 36 complete or partial mt genomes, with 29 of these

taxa from the Hymenoptera. This is more than double the number of hymenopteran taxa compared

with previous studies (14 taxa: Dowton et al., 2009a). The most comprehensive analyses of

holometabolan relationships place the Hymenoptera at the base of the Holometabola (Savard et al.,

2006; Wiegmann et al., 2009), which make it difficult to choose an outgroup. For this reason, we

included a range of homometabolous taxa in the analysis. MrBayes allows only a single taxon to be

the outgroup, so we chose Tribolium (Coleoptera) as the outgroup, but any of the other

holometabolous insects would have been appropriate. Two Bayesian analyses (all codon positions

included or the third codon positions excluded) were performed for each of the alignment datasets

(Clustal W and Muscle).

Figure 2.2 Bayesian analysis of mitogenomic sequences based on the Muscle alignment dataset, including first and
second codon positions from protein-coding genes, and the two rRNA genes. Tribolium (Coleoptera) was specified as
the outgroup. Posterior probabilities are shown at each node.

The intraordinal relationships within the Hymenoptera were broadly congruent across the four
31
analyses, although there were some differences in the topologies. The major difference occurred

between analyses that included all codon positons and those that excluded the third codon positions.

In the 29 hymenopteran taxa, there are 10 taxa from the Aculeata, which contains the superfamilies

Apoidea, Chrysidoidea and Vespoidea. The analyses with the third codon positions excluded

recovered the Aculeata as a monophyletic group, while the analyses with all codon positions

recovered a controversial clade, with Evania (a member of Evaniomorpha) disrupting the Aculeata.

This contrasts with previous studies (Dowton et al., 2009a), in which the Aculeata were recovered as

a clade in both analyses (i.e. including or excluding third codon positions). This suggests that the

exclusion of third codon positions improves phylogenetic accuracy when analyzing hymenopteran

relationships using the whole mt genome, which is consistent with previous analyses (Castro and

Dowton, 2007; Dowton et al., 2009a). By contrast, exclusion of third codon positions has not had a

positive effect on phylogenetic reconstruction in some insect orders (Diptera: Cameron et al., 2007b;

Orthoptera: Fenn et al., 2008).

Dataset alignment strategies may also affect the recovered topology (Cameron et al., 2004; Cameron

et al., 2009). Comparison of the topologies recovered from the datasets aligned by Clustal W and

Muscle (both with third codon positions excluded) indicated that the topology produced from the

Muscle alignment was more congruent with expected relationships than the Clustal W alignment.

Analysis of both datasets recovered a range of monophyletic superfamilies, such as Apoidea,

Chalcidoidea and Ichneumonoidea, but the relationships within the superfamilies varied between the

two strategies. For example, within the Braconidae, the Muscle analysis recovered relationships

entirely congruent with the most recent analysis of Sharanowski et al. (2011); the monophyly of the

noncyclostomes (Cotesia + Phanerotoma + Macrocentrus + Meteorus), the monophyly of the


32
cyclostomes (Spathius + Diachasmimorpha), the monophyly of the microgastroids (Cotesia +

Phanerotoma), and a sister relationship between the cyclostomes and the Aphidiinae (Aphidius). By

contrast, the Clustal W topology failed to support these expected relationships. This suggests that the

phylogenetic relationships at lower taxonomic levels are more sensitive to the alignment strategy.

For the reasons outlined above, we present the tree of the dataset aligned by Muscle, excluding the

third codon positons (Fig. 2.2). Relationships among the Symphyta are recovered as expected – there

are only three symphytans included, and they form a paraphyletic grade at the base of the

Hymenoptera; Tenthredinoidea + (Cephoidea + (Orussoidea + Apocrita)). As expected, a

monophyletic Aculeata is recovered, as is a monophyletic Apoidea. Evania + Aculeata form a clade,

consistent with the resolution of the Aculeata inside the Evaniomorpha in more taxon-rich analyses

(Castro and Dowton, 2006; Heraty et al., 2011; Sharkey et al., 2011). However, support for a range

of relationships within the Aculeata is low, with posterior probabilities below 0.95. Terminal branch

lengths are very long compared with internode lengths. Further, compared with previous analyses,

internal relationships within the Aculeata are less congruent with expectation, despite the increase in

taxonomic sampling. For example, the Chrysidoidea are not recovered as monophyletic; although

Heraty et al. (2011) also failed to recover a monophyletic Chrysidoidea. Similarly, the Vespoidea are

polyphyletic, but this is in agreement with recent, broader analyses of hymenopteran relationships

(Heraty et al., 2011; Sharkey et al., 2011). Alternatively, Pilgrim et al. (2008) recovered a

paraphyletic Vespoidea. Together, these observations suggest that the uncontroversial relationships

recovered by Dowton et al. (2009a) may have been due to the narrow taxon sampling – for example,

the two vespoids are closely related (both from the Vespidae), biasing the analysis toward a finding

of monophyly. It suggests that a higher taxonomic sampling is necessary to reliably recover

33
relationships within the aculeates.

With respect to the Proctotrupomorpha, four taxa from three superfamilies (Proctotrupoidea,

Platygastroidea, and Chalcidoidea) are included in our analysis. The four taxa failed to form a clade

as Vanhornia (Proctotrupoidea) was misplaced, while the Chalcidoidea and Platygastroidea were

recovered as a grade above the Ichneumonoidea. Within the Ichneumonoidea, the well-established

sister group relationship between the Ichneumonidae and Braconidae are well supported (Dowton et

al., 1997; Rasnitsyn, 1988). In addition, the two groups (cyclostomes and non-cyclostomes) of

Braconidae were well recovered.

2.4 Conclusions

We obtained the nearly complete mt genome of T. basalis, which is the first record of a mt genome

of the superfamily Platygastroidea. Comparison of the genome organization of T. basalis with other

Proctotrupomorpha taxa identifies a number of gene rearrangements, which might provide

informative characters for the resolution of relationships among the Proctotrupomorpha.

Phylogenetic analyses using the entire mt genome sequences indicate that, despite the large number

of characters provided by the mt genome, increased taxonomic sampling is likely necessary to

reliably recover evolutionary relationships at the family and superfamily levels in the Hymenoptera.

34
CHAPTER 3: Complete mitochondrial genomes of
Ceratobaeus sp. and Idris sp. (Hymenoptera:
Scelionidae): shared gene rearrangements as potential
phylogenetic markers at the tribal level

This chapter is slightly modified from the paper:

Mao M, Dowton M (2014) Complete mitochondrial genomes of Ceratobaeus sp. and Idris sp.

(Hymenoptera: Scelionidae): shared gene rearrangements as potential phylogenetic markers at the

tribal level. Mol Biol Rep. doi: 10.1007/s11033-014-3522-x.

3.1 Introduction

The Scelionidae (Apocrita: Platygastroidea) is a diverse and speciose family of parasitic

Hymenoptera, encompassing 3308 valid species worldwide (Johnson, 2013). They are generally

small (0.5-12 mm), with most species being morphologically simple compared with other parasitic

wasps (Austin et al., 2005; Masner, 1976). Scelionds are egg parasitoids of grasshoppers, locusts,

crickets, bugs and spiders, so they have great significance in the biological control of agricultural

pests (Galloway and Austin, 1984). The Scelionidae are considered a paraphyletic group based on

morphological and molecular data (Austin and Field, 1997; Murphy et al., 2007). The evolutionary

relationships within the family are poorly resolved, especially at the tribal level. According to the

most recent molecular phylogenetic analysis, none of the tribes was found to be monophyletic except

for the Scelionini, which is probably due to the problems of current classification (Murphy et al.,

2007). A solid phylogenetic scheme for the group is essential to systematic and biological

researchers.

35
Mitochondrial (mt) genome sequences have provided large and diverse datasets useful in improving

phylogenetic relationships (Cameron et al., 2007b; Fenn et al., 2008). Analysis of mt genome

organization is also particularly useful for retracing ancient evolutionary relationships (Boore and

Brown, 1998). In insects, mt genomes are typically double-stranded circular molecules of about 16

kb that contain 13 protein-coding genes, 2 rRNA genes and 22 tRNA genes (Boore, 1999). The mt

genome also contains a large non-coding region, known as the A+T-rich region (in invertebrates) or

control region (in vertebrates), which regulates transcription and replication (Zhang and Hewitt,

1997). Prior to the present study, only a single mt genome from the Scelionidae (Trissolcus basalis)

had been sequenced (Mao et al., 2012). When this sequence was included in a comparison of gene

organizations among the sequenced Proctotrupomorpha taxa, mt gene positions were highly

rearranged at the superfamily level. Thus it would be intriguing to explore mt genome organization

among more closely related taxa, in order to better understand the history of gene rearrangement. In

this paper, we present two complete mt genomes of two other Scelionidae taxa, Ceratobaeus sp. and

Idris sp. Both of these wasps are from the tribe Baeini (Scelionidae: Scelioninae). This provides the

capacity to make both closer (within subfamily) and more distant comparisons (between

superfamilies). We analyze their organizations together with Trissolcus and perform phylogenetic

analyses based on the three mt genome sequences and the Nasonia mt genome. We find that gene

organization is much more conserved at the tribal level and that the mt genome sequences recover

the expected topology with high confidence. These results indicate that the mt genome may provide

a very useful source of data for resolution of the phylogeny of the Scelionidae.

3.2 Materials and methods

DNA extraction
36
The specimens of Ceratobaeus sp. (UOW wasp alcohol library reference number M453) and Idris sp.

(UOW wasp alcohol library reference number M454) were kindly provided by Andy Austin, were

collected from Wistow, South Australia and preserved in 100% ethanol. Genomic DNA was

extracted as described (Aljanabi and Martinez, 1997). The DNA was resuspended in 100 μl of fresh

TE solution (1 mM Tris–HCl, 0.1 mM EDTA [pH 8]) and stored at 4°C.

Amplification, sequencing, and annotation of entire mt genomes

Internal gene fragments were amplified and sequenced using either published universal animal mt

primers (Simon et al., 1994; Simon et al., 2006), or using primers designed from consensus

hymenopteran mt sequences. Using the sequence information obtained, specific primers were

designed for long amplification of multigenic regions. Short PCRs were performed using

BIOTAQ™ DNA Polymerase (Bioline, Australia) with the following cycling conditions: 94oC for 2

min, 35 cycles of 94oC for 30 sec, 45oC-65oC for 1 min, 72oC for 2 min and a final elongation for 10

min at 72oC. Long PCRs were performed using Takara LA Taq (Takara Biomedical, Japan) with the

following conditions: 94oC for 2 min, 30 cycles of 94oC for 1 min, 45oC-65oC for 30 sec, 68oC for

12 min and a final elongation for 10 min at 68oC. The A+T-rich region of Ceratobaeus sp. was

amplified by nested PCR. All PCR products were visualised by agarose gel electrophoresis and

purified with ExoSAP-IT® (GE Healthcare, Bucks, UK). The primer-walking method was employed

for direct sequencing, using the ABIPRISM®BigDye™ Terminator v3.1 Cycle Sequencing Kit

(Applied Biosystems, Australia). Amplicons that were difficult to sequence directly were ligated into

the pGEM-T Easy Vector System (Promega, Madison, Wisconsin) and cloned into E. coli JM109

cells (Promega, Madison, Wisconsin). The sequences of both strands were determined for each of

the entire mt genomes.


37
The sequences were assembled into contigs in ChromasPro Ver 1.33 (Technelysium Ltd., Tewantin,

Australia). ORFinder (www.ncbi.nlm.nih.gov/gorf/gorf.html) was used to identify protein-coding

genes, specifying the invertebrate mt genetic code. The tRNA-scan SE 1.21

(lowelab.ucsc.edu/tRNAscan-SE/) (Lowe and Eddy, 1997) search server was used to annotate tRNA

genes using the following settings: source, Mito/Chloropast; genetic code for tRNA isotype

prediction, Invertebrate Mito; and cove score cutoff, 5. Ribosomal RNA genes were identified both

through sequence comparison with published hymenopteran mt ribosomal RNA sequences and

through comparison of the conserved secondary structure of the 3’ end of rrnS. The sequence data

have been deposited in GenBank, under accession numbers KF696669 and KF696670.

Sequence alignment

Two mt genomes (Ceratobaeus sp. and Idris sp.) sequenced in this study together with one mt

genome (T. basalis) sequenced in a previous study (Mao et al., 2012) were aligned, together with

Nasonia (which would form the outgroup for later phylogenetic analyses). Nucleotide sequences for

each of the 13 protein-coding genes and the 2 rRNA genes were imported into separate files using

MEGA5 (Tamura et al., 2011) and aligned using Clustal W (Thompson et al., 1994) or Muscle

(Edgar, 2004) as implemented within MEGA5. For the protein-coding genes (excluding the stop

codons), an amino acid alignment was generated first for each gene and a nucleotide alignment

inferred from the amino acid alignment. The alignment parameters were the default settings for all

genes, which have been specified in previous studies (Cameron et al., 2008; Mao et al., 2012).

Nucleotide alignments were concatenated prior to phylogenetic analysis.

38
Table 3.1 Datasets used for phylogenetic analysis.

Dataset Alignment method Partitions Number of partitions


All-DNA123 + rRNA Clustal W 1st, 2nd, 3rd codons, rrnL, rrnS 5
Muscle 1st, 2nd, 3rd codons, rrnL, rrnS 5
st st st
Clustal W (rrnS, rrnL) (a6_1 codon, cob_1 codon, cox3_1 10
codon, cox2_1st codon, cox1_1st codon) (a6_2nd
codon, cox3_2nd codon, cox2_2nd codon, nad1_2nd
codon, nad4l_2nd codon, nad4_2nd codon, nad5_2nd
codon) (a6_3rd codon, cob_3rd codon, cox3_3rd
codon, cox2_3rd codon, cox1_3rd codon, nad3_3rd
codon) (a8_1st codon, nad2_1st codon, nad3_1st
codon, nad6_1st codon) (a8_2nd codon, nad2_2nd
codon, nad3_2nd codon, nad6_2nd codon) (a8_3rd
codon, nad2_3rd codon, nad6_3rd codon) (cob_2nd
codon, cox1_2nd codon) (nad1_1st codon, nad4l_1st
codon, nad4_1st codon, nad5_1st codon) (nad1_3rd
codon, nad4l_3rd codon, nad4_3rd codon, nad5_3rd
codon)
Muscle (rrnS, rrnL) (a6_1st codon, cob_1st codon, cox3_1st 10
codon, cox2_1st codon, cox1_1st codon) (a6_2nd
codon, cob_2nd codon, cox3_2nd codon, cox2_2nd
codon, cox1_2nd codon, nad3_2nd codon) (a6_3rd
codon, cob_3rd codon, cox3_3rd codon, cox2_3rd
codon, cox1_3rd codon, nad3_3rd codon) (a8_1st
codon, nad2_1st codon, nad3_1st codon, nad6_1st
codon) (a8_2nd codon, nad2_2nd codon, nad6_2nd
codon) (a8_3rd codon, nad2_3rd codon, nad6_3rd
codon) (nad1_1st codon, nad4l_1st codon, nad4_1st
codon, nad5_1st codon) (nad1_2nd codon,
nad4l_2nd codon, nad4_2nd codon, nad5_2nd codon)
(nad1_3rd codon, nad4l_3rd codon, nad4_3rd codon,
nad5_3rd codon);
All-DNA12 + rRNA Clustal W 1st, 2nd codons, rrnL, rrnS 4
st nd
Muscle 1 , 2 codons, rrnL, rrnS 4

Table 3.2 Estimation of Bayes Factors of different partition schemes.

Partition* Bayes Factor


P4-Muscle/P4-ClustalW 343
P5-Muscle/P5-ClustalW 314
P10-Muscle/P10-ClustalW 807
P10-Muscle/P5-Muscle 1538
P10-ClustalW/P5-ClustalW 1045

* Partition schemes are outlined in Table 3.1. P4, P5, P10 correspond to the number of partitions given in Table 3.1

Genetic divergence and phylogenetic analysis

The ratios of Ka (the nonsynonymous substitution rate) and Ks (the synonymous substitution rate)

39
for protein coding genes between the Scelionidae species were calculated using the software Ka/Ks

calculator (v. 1.2) based on the model-averaging method (MA) (Zhang et al., 2006). For the

phylogenetic analyses, two partition schemes were used for the two datasets (one aligned using

Clustal W, the other aligned using Muscle). Previous investigations of partitioning approaches using

hymenopteran and termite mt genome sequences indicated that nucleotide data excluding third

codon positions better recovered traditional relationships (Cameron et al., 2012; Dowton et al.,

2009a; Mao et al., 2012). However, inclusion of third codon positions has been found to improve the

accuracy of phylogenetic reconstructions in other studies (Cameron et al., 2007b; Fenn et al., 2008;

Nelson et al., 2012). In addition, a more objective framework for choosing among competing

partition schemes was recently developed, PartitionFinder (Lanfear et al., 2012). For these reasons, a

range of partition schemes were compared (Tables 3.1, 3.2). A total of 41 data blocks were defined

(3 codon positions of 13 protein-coding genes + 2 rRNA genes). PartitionFinder distinguished

between competing partition schemes using the Bayesian Information Criterion (BIC) and a heuristic

search algorithm. In all cases, MrModelTest 2.3 (Nylander et al., 2004) was employed to define

models of nucleotide evolution for each partition.

Phylogenetic analysis was performed with MrBayes v. 3.1.2 (Ronquist and Huelsenbeck, 2003). The

Markov chain Monte Carlo (MCMC) process was set so that four chains (three heated and one cold)

ran simultaneously. For each analysis, we conducted four independent runs for 1,000,000

generations with sampling every 100 generations. Burn-in was discarded according to the plot of

generation against the likelihood scores. Stationarity was assessed by importing the parameter file

into Tracer v. 1.5 (Rambaut and Drummond, 2009), and assessing whether the ESS of all parameter

values were >100. The percentage of trees recovering a particular clade was used as a measure of

40
that clade’s posterior probability (Huelsenbeck and Ronquist, 2001).

3.3 Results and discussion

Genome size and gene content

Two complete mt genomes (Ceratobaeus sp. and Idris sp.) were sequenced. Each genome contained

13 protein-coding genes, 22 tRNA genes, 2 rRNA genes, and an A+T-rich region similar to most

other insects (Fig. 3.1). The Ceratobaeus sp. mt genome is 15,851 bp long. This size is well within

the range found in other completely sequenced hymenopteran insects (15,425 bp in Spathius agrili to

19,339 bp in Cephus cinctus) (Dowton et al., 2009a; Wei et al., 2010a). The A+T content of the

entire genome is 77.59% (A=40.87%; T=36.71%; C=14.03%; G=8.38%), which is similar to Perga

condei (78%) (Castro and Dowton, 2005). We detected five overlapping regions between genes

(trnD-trnK, a8-a6, trnN-trnF, trnF-trnS1, and trnE-trnR) for a total of 21 bp (supplementary Table 1,

http://dx.doi.org/10.1007/s11033-014-3522-x). The longest one is 8 bp, involving the trnE and trnR

genes. Twelve short noncoding regions ranging from 2 to 72 bp were identified.

At 15,137 bp, the mt genome of Idris sp. is the shortest hymenopteran complete mt genome to date.

The A+T content of the entire mt genome is 81.22%, which is a little higher than Ceratobaeus sp.

Six overlapping regions were detected. Four of these overlaps are in the same location as found in

Ceratobaeus sp. (trnD-trnK, a8-a6, trnN-trnF, and trnE-trnR), but there are another two located in

the trnQ-trnC and trnT-trnP regions (supplementary Table 1,

http://dx.doi.org/10.1007/s11033-014-3522-x). There are 14 short noncoding regions in Idris sp.,

ranging from 1 to 72 bp, which together total 155 bp.

41
Figure 3.1 Mitochondrial genome organization of three sceliond taxa, compared with the ancestral pancrustacean mt
genome organization (Cook 2005). tRNA genes are indicated by single-letter amino acid codes, L1, L2, S1, and S2
denote tRNA-Leu (CUN), tRNA-Leu (UUR), tRNA-Ser (AGN), and tRNA-Ser (UCN), respectively. Genes are
transcribed from left to right except those indicated by underlining. Gene movements, relative to the ancestral
organization, are indicated with arrows.

Protein-coding genes

The 13 protein-coding genes typical for insect mt genomes were identified using ORF finder. Two of

the conventional start codons ATA or ATG could be assigned to most of the protein-coding genes.

However, nad2, a8, and cox2 use ATC, ATC, and ATT in Ceratobaeus sp., respectively, and cox1

and nad5 use ATT in Idris sp. Nine genes end with complete termination codons (TAA) in

Ceratobaeus sp., while only six genes have complete termination codons in Idris sp. The remaining

genes appear to use incomplete termination codons (TA or T) that are presumably completed by

polyadenylation after transcription (Ojala et al., 1981).

42
The relative synonymous codon usage (RSCU) shows high base compositional biases for AT in the

mt genomes of Ceratobaeus sp. and Idris sp. (supplementary Table 2, 3,

http://dx.doi.org/10.1007/s11033-014-3522-x). Within a particular synonymous codon family,

codons with A or T in the third codon position are generally strongly overrepresented compared to

codons with G or C in the third position. However, the RSCU values of GCC (Ala), AGG (Ser) and

GGG (Gly) are higher than GCA (Ala), AGT (Ser) and GGT (Gly) in Ceratobaeus sp., respectively.

GCC (Ala) and AGG (Ser) have slightly higher RSCU values than GCA (Ala) and AGT (Ser),

respectively, while the RSCU value of GGG (Gly) is much higher than GGT (Gly) (1.31 to 0.35).

Anticodon analysis showed that both trnS-TCT and trnG-TCC contain wobble nucleotides and they

can decode AGG and GGG, respectively. This might be the reason for the higher RSCU value of

AGG (Ser) and GGG (Gly). However, it cannot explain the higher RSCU value of GCC (trnA-TGC).

Figure 3.2 Inferred secondary structures of tRNA genes: A-B. trnS1 identified in the mt genomes of Ceratobaeus sp.
and Idris sp., respectively. C. Pseudo-trnK identified in the mt genome of Idris sp. D-G, trnR identified in the three
sceliond mt genomes (D. Ceratobaeus sp. E. Idris sp. and F. Trissolcus basalis), compared with trnR reported in the
mt genome of Ceratosolen solmsi (Chalcidoidea). Watson-Crick pairs are indicated by lines and noncanonical pairs
are indicated by asterisks.

Transfer RNA and Ribosomal RNA

43
The length of tRNA genes ranges from 58 bp to 74 bp in the two species. Most tRNA genes have a

typical cloverleaf structure except for trnS1 and trnR, in which the D-stem pairings in the

dihydrouridine (DHU)-arm are absent (Fig. 3.2). The missing D-stem has been commonly noted for

the trnS1 gene in insects (Castro et al., 2006; Sheffield et al., 2008; Yang et al., 2013). It is

hypothesized that the atypical trnS1 occurred early in, or before the evolution of Metazoa (Garey and

Wolstenholme, 1989). Although the loss of the DHU arm of trnR has been reported in other

non-insect organisms (e.g. Taenia asiatica) (Jeon et al., 2005), the missing D-stem in trnR, to our

knowledge, has not been previously reported in insect mt genomes. This particular tRNA gene was

the most difficult to locate in all three sceliond mt genomes (Ceratobaeus sp., Idris sp. and T.

basalis); it overlaps with trnE in both Ceratobaeus sp. and Idris sp. In our previous study on the mt

genome of T. basalis, we were unable to find the trnR gene (Mao et al., 2012). However, using the

location and structure of the trnR gene in Ceratobaeus sp. and Idris sp. as a guide, we are now able

to locate a putative trnR gene in T. basalis; all three have similar secondary structures (Fig. 3.2). We

suggest that the atypical trnR might be a shared derived character for the Scelionidae.

Since the rrnL and rrnS genes are directly adjacent to each other in both species, we identified the

gene boundaries by comparing the nucleotide sequences with other hymenopteran mt genomes. The

rrnL and rrnS genes of Ceratobaeus sp. are 1,280 and 754 bp long, which are shorter in size by 9

and 21 bp than those genes of Idris sp.

Noncoding regions

The two mt genomes contain a large noncoding region (A+T-rich region) between trnM and trnQ. It

is 1,098 bp long in Ceratobaeus sp. In contrast to the A+T-rich region in most insect mt genomes,

44
the A+T content (75.16%) of this region is lower than that of the entire mt genome (77.59%). It

contains an array of tandem repeats with a total length of 1,066 bp present in eight copies (the last

one is a partial copy). The first, second, third, fifth, sixth and seventh copies are almost identical in

sequence, with only one to three site differences, while the fourth copy is 24 or 25 bp shorter than

other copies and with six to nine site differences. The A+T-rich region has been difficult to amplify

in other hymenopteran mt genomes (Castro et al., 2006; Oliveira et al., 2008). We could only

amplify this region in Ceratobaeus sp. by nested PCR. It suggests that the long tandem repeats might

cause amplification problems. In addition to the long tandem repeats, a microsatellite (AT)10 was

identified, which is a typical characteristic reported in other insect mitochondrial A+T-rich regions.

Several short non-coding regions with a total of 199 bp are scattered throughout the Ceratobaeus sp.

mt genome. The longest one (72 bp) is located between the nad4L and trnP genes, where the trnT

gene is missing compared with the ancestral organization (Fig. 3.1). Sequence analysis shows that

this noncoding region possesses the anticodon (TTG) of trnT but cannot be folded into a clover-leaf

structure. The 59.7% sequence identity compared with trnT indicates that it might be a remnant of an

historical intermediate generated by the tandem duplication/random loss model (Moritz et al., 1987)

(see below).

The A+T-rich region of Idris sp. is 414 bp with an A+T content of 84.3%, which is relatively higher

than that of the entire mt genome (81.22%). An 8 bp tandem repeat (TTTTACTA) is present in three

copies, as well as a microsatellite (AT)9. The second largest noncoding region of Idris sp. is 72 bp,

between the nad1 and trnL1 genes. This region contains a sequence that can be folded into a

secondary structure resembling an additional trnK with TTT as the anticodon, differing from the

designated trnK using CTT as the anticodon (Fig. 3.2). The cove score of trnKTTT (19.9) is relatively
45
lower than that for trnKCTT (28.8), as determined by tRNAscan-SE. The abnormal anticodon TTT is

more commonly used in the trnK gene of hymenopteran mt genomes (Castro et al., 2006; Gotzek et

al., 2010; Kaltenpoth et al., 2012; Wei et al., 2010a). However, trnK (CTT) is clearly the ancestral

form of this gene, as it is the form that is present in a range of holometabolous and other insects.

Although trnK (TTT) may also function as a bona fide trnK gene, the ancestral nature of the trnK

(CTT) gene, together with the higher cove score leads us to postulate that the trnK (CTT) gene is the

functional version of the trnK. The presence of tRNA-like sequences or extra tRNA genes has been

commonly reported in other insects (Beckenbach, 2011; Castro et al., 2006; Cha et al., 2007; Shao

and Barker, 2003).

Genome organizations

The two mt genome organizations are shown in Fig. 3.1. The protein-coding and rRNA genes are

conserved in positions and orientations relative to the organization of the ancestral pancrustacean

(Cook, 2005). However, the relative positions of tRNA genes are highly variable. As shown in Fig.

3.1, a total of 11 tRNA genes have rearranged in both Ceratobaeus sp. and Idris sp. when compared

with the ancestral positions. The rearrangement events mainly occur at three junctions (nad2-cox1,

cox2-a8 and nad3-nad5) and around the A+T-rich region. The cox2-a8 and nad3-nad5 junctions

have been reported as “hot spots” for gene rearrangements in the Hymenoptera (Dowton and Austin,

1999; Dowton et al., 2003). In addition, trnT and trnP have swapped positions in Ceratobaeus sp.

We then compared the genome organizations among the three Scelionidae taxa. The gene order of

Ceratobaeus sp. is nearly identical to that of Idris sp. Only two tRNA genes (trnT-trnP→trnP-trnT)

are arranged differently between the two mt genomes. The presence of a short noncoding region

46
between nad4L and trnP in Ceratobaeus sp. is highly consistent with the tandem duplication/random

loss model, which is the most plausible mechanism for local gene rearrangements in animal mt

genomes (Boore, 2000). The retention of the ancestral organization (trnT-trnP) in both Idris sp. and

T. basalis indicates the gene rearrangement occurred after the divergence of Ceratobaeus within the

Baeini. There are some derived gene arrangements shared among the three taxa: trnK and trnD have

swapped positions; trnM has inverted and moved into the junction between rrnS and the A+T-rich

region; trnC and trnY have moved out of the nad2-cox1 junction and trnC has moved to the junction

between nad2 and the A+T-rich region in all 3 scelionid taxa; trnV has moved into the junction

between rrnS and the A+T-rich region and trnY is inverted. Some of the rearrangements were also

characterized in other hymenopteran linages. For example, the derived position of trnM reported

here was identical with 7 known braconid (Ichneumonoidea) mt genomes (Wei et al., 2010a). But

these two superfamilies (Platygastroidea and Ichneumonoidea) are not closely related groups,

indicating this arrangement has likely evolved on two occasions independently. Furthermore, trnV

has also moved out of the rrnL-rrnS junction in three chalcidoid taxa (Oliveira et al., 2008; Xiao et

al., 2011). Although its derived position is unknown in these taxa, we speculate that the

rearrangement of trnV might be a shared derived character for Platygastroidea and Chalcidoidea,

which were recovered as sister groups in some phylogenetic analysis (Dowton and Austin, 2001;

Ronquist et al., 1999).

Shared gene rearrangements are considered valuable sources for deducing phylogenetic relationships

(Dowton et al., 2002). The two mt genomes reported here share almost all of the derived gene

positions. Both of them belong to the tribe Baeini, which was not resolved as monophyletic by

molecular data (Carey et al., 2006; Murphy et al., 2007). The morphological phylogeny for the

47
Baeini is also problematic as it is dominated by reductional synapomorphies (Iqbal and Austin,

2000). Idris Foerster and Ceratobaeus Ashmead are the two most speciose genera in Baeini and they

were found to be polyphyletic in previous studies (Carey et al., 2006; Murphy et al., 2007). Although

our data is not yet sufficient to comprehensively address these issues, the evidence of shared gene

positions between closely related groups suggests that these gene rearrangements have much

potential as phylogenetic markers at the tribal level in the subfamily Scelioninae. Further, the three

Scelionidae taxa also shared some derived gene arrangements. However, we consider that we require

more genome representatives before we can conclude that these shared rearrangements are derived

characters for the Scelionidae.

Table 3.3 Comparison of genetic divergences for all protein-coding genes among scelionds.

Comparison Ka Ks Ka/Ks
Ceratobaeus/Idris 0.160421 6.48094 0.0247527
Ceratobaeus/Trissolcus 0.267965 9.04276 0.0296331
Idris/Trissolcus 0.265943 11.7398 0.0226532

Ka/Ks ratios among the three Scelionidae taxa

The calculation of nonsynonymous (Ka) and synonymous (Ks) substitution rate is informative for

understanding the evolutionary dynamics of coding regions across closely related species (Fay and

Wu, 2003). In this study, we measured the Ka and Ks of protein coding genes and determined the

Ka/Ks ratios between the three Scelionidae taxa, respectively (Table 3.3). The results show that Ks

(range 6.48-11.74) is about 34-44 times higher than Ka (range 0.16-0.27) in each comparison, which

indicates these protein coding genes are evolving under strong purifying selection (Roques et al.,

2006). The results also indicated that there was little difference between the Ka/Ks ratio when all

combinations of the various Scelionidae were compared. This indicates that there has not been a
48
substantial rate change during the evolution of the Scelionidae, in contrast to the rate change that is

evident when symphytan wasps are compared with apocritan wasps (Dowton et al., 2009a).

Phylogenetic analysis

Phylogenetic analyses were performed using four mt genomes, with three of them from the

Scelionidae. Nasonia (Chalcidoidea: Pteromalidae) was chosen as the outgroup. Bayesian analyses

were performed for each dataset, in which the alignment and partitioning approach were varied; the

analyses are outlined in Table 3.1. Each analysis recovered the same uncontroversial relationships

(supplementary Fig. 1, http://dx.doi.org/10.1007/s11033-014-3522-x). The two Baeini taxa

(Ceratobaeus sp. and Idris sp.) group together with strong support (pp = 1). In contrast to previous

studies using hymenopteran mt geomes, inclusion of third codon positions had no effect on topology

and nodal support in the current analysis, indicating that inclusion of third codon positions does not

appear to be problematic when constructing the phylogeny of closely related taxa, which is

consistent with a recent study in which intrafamily relationships within a family of Diptera were

examined (Nelson et al., 2012)

In order to test the effects of different alignment methods and partition schemes, we employed Bayes

Factors following the approach in a previous study (Dowton et al., 2009a). As shown in Table 3.2,

alignment of sequences using Muscle resulted in trees with improved likelihood scores when

compared with those aligned with Clustal W. This indicates that alignment with Muscle is preferable

to those with Clustal W in phylogenetic reconstruction of hymenopteran taxa, which is consistent

with our previous study (Mao et al., 2012). We then examined the effect of different partition

schemes. Choosing an appropriate partitioning scheme is important for accuracy of phylogenetic

49
reconstruction using mt genomes (Dowton et al., 2009a; Fenn et al., 2008). For the All-DNA123 +

rRNA dataset, we first divided the data into four partitions (1st, 2nd, 3rd codon positions and rRNA, as

reported in previous studies). In the second scheme, we employed PartitionFinder, which divided the

data into ten partitions in both the All-DNA123 + rRNA-Clustal W and the All-DNA123 +

rRNA-Muscle datasets.

The results of our PartitionFinder analyses suggest that more complex partitioning schemes can

result in vastly improved likelihood scores. The likelihood score of the tree found using the model

established by PartitionFinder was 769 log units higher than that found using our previous approach

(1st, 2nd, 3rd codon positions, and rRNA partitions). This increase in likelihood is not likely to simply

reflect an increased fit of the data due to the increased number of parameters in the model. This is

because PartitionFinder uses the BIC to distinguish between competing partition schemes, and this

approach penalizes over-parameterization (Sullivan and Joyce, 2005), particularly when the number

of sites is high (as is the case with mt genome based phylogenies). It is also of interest to examine

the details of the partitions found by PartitionFinder, as it may yield clues to the underlying

evolutionary processes operating on the hymenopteran mt genome. Genes or codon positions that are

placed into the same partition may be evolving under the same underlying influences. For example,

the third codon positions of protein coding genes were partitioned into three partitions: (3rd codon

positions of a6, cob, cox3, cox2, cox1, nad3), (3rd codon positions of a8, nad2, nad6), and (3rd codon

positions of nad1, nad4L, nad4, nad5). The third codon positions are the most likely to reveal any

underlying evolutionary directions, as they are expected to be least subject to selective constraints.

Interestingly, the four protein coding genes that are encoded on the minority strand (nad1, nad4L,

nad4, nad5) were partitioned together, while those encoded on the majority strand were also

50
partitioned together (although they were further divided into two separate partitions). This suggests

that the underlying mutational model is different for the two strands of the mitochondrial genome, as

has been reported in mammals (Reyes et al., 1998). Finally, of the two partitions that contained 3rd

codon positions of genes encoded on the majority strand, a8, nad2 and nad6 were grouped together,

and have been reported to have the highest evolutionary rates in mammals and other insects (Li et al.,

2012; Saccone et al., 1999). Together, these observations indicate that PartitionFinder is able to

discover superior partitioning schemes for the phylogenetic analysis of mt genome data – they

produce more likely trees and may be more biologically realistic. This suggests that it may be useful

to reinvestigate previously analysed datasets, ones that were produced prior to the availability of

PartitionFinder.

51
CHAPTER 4:Coexistence of minicircular and a highly
rearranged mtDNA molecule suggests that
recombination shapes mitochondrial genome
organization

This chapter is slightly modified from the paper:

Mao M, Austin AD, Johnson NF, Dowton M (2014) Coexistence of minicircular and a highly

rearranged mtDNA molecule suggests that recombination shapes mitochondrial genome

organization. Mol Biol Evol. 31:636-644. doi:10.1093/molbev/mst255.

4.1 Introduction

The mitochondrial (mt) genome is one of the most highly used sources of molecular markers for

reconstructing phylogeny, both through alignment of orthologous sequences (Simon et al., 1994) and

through comparison of gene position (Boore et al., 1998). Analysis of aligned sequences indicates

that there are inherent biases in the number and types of nucleotide substitutions that occur, and that

incorporating this information into phylogenetic analysis improves the power of the analysis (Hillis

et al., 1994). To appreciate whether these patterns are general or specific to the taxon being studied,

it is also important to understand the mechanism of nucleotide substitution (Bakker et al., 2000). For

the same reasons, it is critical to understand the mechanism of gene rearrangement. But knowledge

on the mechanism and biases of gene rearrangement has remained elusive, because they occur much

more sporadically. Information on the mechanism and frequency of different types of rearrangement

can be best inferred by characterizing lineages with high rates of gene rearrangement (Dowton and

Austin, 1999).

52
There are three main models proposed to explain gene rearrangement: duplication/random loss

(Moritz et al., 1987; San Mauro et al., 2006), duplication/nonrandom loss (Lavrov et al., 2002) and

recombination (Dowton and Campbell, 2001). In the duplication/random loss model, slipped-strand

mispairing or inaccurate termination during replication can cause duplication of part of the mt

genome. The supernumerary gene copies are subsequently eliminated due to selection, which will

either restore or alter the original gene order. The duplication/random loss model is consistent with

the observed gene rearrangements in the vertebrates (San Mauro et al., 2006). However, it cannot

explain gene inversions and long-range translocations, which are common in invertebrate mt

genomes (Dowton et al., 2002). A defining feature of the duplication/nonrandom loss model is the

duplication of the entire mt genome, resulting in a dimeric molecule linked head to tail. Such a

dimeric molecule would contain two copies of transcriptional promoters. Mutation to one copy of

the promoters would result in the genes under the control of the disabled promoter becoming

pseudogenes; eventually these would be eliminated. After rearrangement, genes with identical

transcriptional direction will be clustered in the mt genome. However, the duplication/nonrandom

loss model also cannot explain gene inversions and translocations with different transcriptional

direction to genes close to the translocation point. Recombination was considered to be absent in

animal mtDNA based on early studies. For example, recombinant haplotypes of mtDNA were not

detected in somatic cell hybrids; sequestration of mtDNA molecules into clusters prevents physical

contact of unrelated molecules and there is also evidence that excision repair activity and crossover

products are absent in mammalian mitochondria (Clayton et al., 1974; Satoh and Kuroiwa, 1991;

Zuckerman et al., 1984). However, both direct and indirect evidence of animal mtDNA

recombination has been demonstrated more recently. In 1996, the end products of homologous

53
recombination were detected in human mitochondria (Thyagarajan et al., 1996). Lunt and Hyman

(1997) provided the most convincing evidence by characterizing the end products of recombination

in the nematode Meloidogyne javanica. A number of subsequent studies also provided evidence of

recombination in other animal species (Awadalla et al., 1999; D'Aurelio et al., 2004; Hoarau et al.,

2002; Kraytsberg et al., 2004; Ladoukakis and Zouros, 2001; Ujvari et al., 2007). However, it is not

yet widely accepted that mt recombination occurs (e.g., Jorde and Bamshad, 2000; Kivisild and

Villems, 2000; Kumar et al., 2000; Parsons and Irwin, 2000), while there is very little information on

how frequently it occurs.

The occurrence of recombination in animal mtDNA has substantial impact on evolutionary studies

that utilize mt genes. First, the traditional methods for phylogenetic analysis are based on the

assumption that animal mtDNA does not recombine; ignoring recombination might mislead

phylogenetic reconstruction (Rokas et al., 2003; Slate and Gemmell, 2004). Second, it offers an

intriguing understanding of how mt genes might rearrange. As the duplication/random loss model

cannot provide a full explanation of gene rearrangements across the Metazoa, recombination might

explain the gene inversions and long-range translocations that are often observed in invertebrate mt

genomes (Dowton and Campbell, 2001). The key component of this mechanism is the introduction

of double-strand breaks (DSBs) in the mt circle, with subsequent rejoining of the broken ends. Two

recent studies provided support for this mechanism: DSBs were demonstrated to play an important

role in the generation of mtDNA rearrangements in mice (Bacman et al., 2009). A more recent study

speculated that the presence of mtMUTS in the octocoral mt genome might promote recombination

which accounts for the gene inversions observed in that genome (Brockman and McFadden, 2012).

Here, we report the discovery of a minicircular mtDNA molecule in the megaspilid wasp
54
Conostigmus sp., similar to that reported by Lunt and Hyman (1997) in the nematode M. javanica.

We used a similar approach to investigate the presence of the end products of recombination: the

maxicircle and minicircle. We detected such end products, although we caution that polymerase

errors can produce a range of apparent recombination products. Furthermore, the possibility that the

minicircle comes from nuclear copies of mt sequences (numts; nuclear pseudogenes) is eliminated in

our study. We then investigate the structure of the minicircles produced and develop a model of

recombination that is consistent with the sequence data. We suggest that the presence of minicircles

is associated with the most highly rearranged invertebrate mt genomes (the Hymenoptera and

Phthiraptera) (Shao et al., 2009), lending credibility to the notion that recombination is an important

component of the mt genome rearrangement mechanism.

4.2 Materials and methods

DNA extraction

Genomic DNA was extracted from 100% ethanol preserved specimens of Conostigmus sp. (collected

from Mount Barker, South Australia) using the “salting out” protocol (Aljanabi and Martinez 1997).

The DNA was resuspended in 100 μl of fresh TE solution (1 mM Tris–HCl, 0.1 mM EDTA [pH 8])

and stored at 4°C.

Amplification, sequencing, and annotation of the entire mt genome

Initially, small gene fragments were amplified and sequenced using either published universal

primers (Simon et al. 1994; Simon et al. 2006) or using primers designed from consensus

hymenopteran mt sequences. Using the sequence information obtained, specific primers were

designed for the amplification of the remaining regions. Short and long PCRs were performed using
55
BIOTAQ™ DNA Polymerase (Bioline, Australia) and Takara LA Taq (Takara Biomedical, Japan),

respectively, following the manufacturers’ instructions. For short PCRs, cycling condition consisted

of 94°C for 2 min, 35 cycles of 94°C for 30 sec, 45°C-65°C for 1 min, 72°C for 2 min, and a final

elongation for 10 min at 72°C. Long PCRs were performed with the following conditions: 94°C for

2 min, 30 cycles of 94°C for 1 min, 45°C-65°C for 30 sec, 68°C for 12 min, and a final elongation

for 10 min at 68°C. All PCR products were resolved by agarose gel electrophoresis and purified with

ExoSAP-IT® (GE Healthcare, Bucks, HP8 4SP, UK). The primer-walking method was used for

direct sequencing, using the ABIPRISM®BigDye™ Terminator v3.1 Cycle Sequencing Kit (Applied

Biosystems, Australia). Amplicons that were difficult to sequence directly were ligated into the

pGEM-T Easy Vector System (Promega, Madison, Wisconsin) and cloned into Escherichia coli

JM109 cells (Promega, Madison, Wisconsin). The sequences of both strands were determined for the

entire mt genome. Genes were identified using methods as previously described (Mao et al. 2012).

Amplification and sequencing of minicircles and the maxicircle

To detect the existence of minicircles, two specific, outward-facing primers (85-ATF and 85-ATR)

were designed, based on the A+T-rich region sequence, following the strategy of Lunt and Hyman

(Lunt and Hyman 1997). PCR amplification was carried out using Takara LA Taq. PCR products

were purified by agarose gel electrophoresis and cloned as described above. A total of 3-5 clones per

product were sequenced. The putative maxicircle was amplified using the primers 85-MR1 and

85-36R1 followed by a nested PCR with two internal primers 85-MR2 and 85-36R2. The PCR

product was purified and sequenced directly. The primers are indicated in Figure 4.1.

56
Figure 4.1 Organizations of the entire mt genome, maxicircle and minicircles that were initially characterized by
amplification. A. Organization of the entire mt A+T-rich region of Conostigmus sp. Different repeats are shown in
different colours. PCR primer sites and orientations are indicated by arrows. Breakpoint junctions are indicated by
black arrows. Regions with hatching indicate non-repeat regions. B. The real minicircles (L2) from different
individuals and the artifactual minicircles (L1, S1, and S2) generated by PCR errors. C. Organization of the
maxicircle A+T-rich region.

4.3 Results

The mt genome of Conostigmus sp.

The entire mt genome of Conostigmus sp. is 16,315 bp long. The overall A+T content is 82.9%. Like

other insect mt genomes, 13 protein-coding genes, 22 tRNA genes, 2 rRNA genes, and the A+T-rich

region were identified. A series of gene rearrangements are evident relative to the putative ancestral

pancrustacean mt genome (Cook, 2005). Although tRNA genes are highly rearranged among wasps

(Dowton et al., 2009b), the protein-coding genes are highly conserved (but see Oliveira et al., 2008;

Xiao et al., 2011). In Conostigmus sp., 11 tRNA and 2 protein-coding genes are rearranged (Fig. 4.2).

Cox1 and nad2 are translocated into two junctions (Cox3-nad3 and rrnL-rrnS), respectively.

Interestingly, the gene orientations are almost identical to that of the ancestral organization, except a

57
single tRNA inversion of trnW.

Figure 4.2 Mitochondrial genome organization of Conostigmus sp., compared with the ancestral pancrustacean mt
genome organization (Cook 2005). tRNA genes are indicated by single-letter amino acid codes, L1, L2, S1, and S2
denote trnLCUN, trnLUUR, trnSAGN, and trnSUCN. Genes are transcribed from left to right except those indicated by
underlining. Gene movements, relative to the ancestral organization, are indicated with arrows.

The A+T-rich region is 1,447 bp long with an A+T content of 86.8%. It can be divided into three

subsections (Sec I, Sec II and Sec III) (Fig. 4.1A). Sec I contains seven copies of tandem repeat A

(the last one is a partial copy), nine copies of tandem repeat B, and one copy of non-tandem repeat C

and D. In the second subsection, several conserved elements (poly T stretch, [TA (A) ]n-like stretch,

TATA motif and stem-loop structure) can be identified, which are putatively involved in the initiation

of replication and transcription (Zhang and Hewitt, 1997). Sec III possesses four copies of tandem

repeat B and the second copy of non-tandem repeat C and D. D overlaps with trnG by 12 bp.

Initial characterization of the putative maxicircle and minicircles of Conostigmus sp.

The presence of distinct repeats within the A+T-rich region was reminiscent of those described by

Lunt and Hyman (Lunt and Hyman, 1997) in the nematode M. javanica, in which they characterized

both minicircular and maxicircular products of recombination. For this reason, we designed outward

facing amplification primers (85-ATF and 85-ATR, Fig. 4.1A) within the A+T-rich region in an

58
attempt to amplify any minicircles that might be present in Conostigmus sp. A short extension time

should preclude the amplification of the entire mt genome, while any amplicons produced from the

entire mt genome would be easily distinguishable (based on their size of 16,100 bp) by agarose gel

electrophoresis. Thus, any amplicons smaller than 16,100 bp would be produced from minicircles. In

one reaction, four different sizes of minicircles (L1, L2, S1 and S2) were detected (Fig. 4.3A). To

test whether these putative minicircles could be amplified from a range of individuals, we amplified

them using the same primers in another two individuals caught at the same location (85B and 85C).

Four and three PCR products (L1 was lacking in 85C) were obtained respectively, with the sizes of

the various amplicons consistent with the minicircles from 85A (data for 85B-L1, 85B-L2 and

85C-L2, see supplementary Fig. 1A,

http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1; data not shown for 85B-S1,

85B-S2, 85C-S1 and 85C-S2). We obtained sequence data from the most strongly amplified product

(L2) from each individual, the two shorter products (S1 and S2) from 85A, and the longest product

(L1) from 85B (Fig. 4.1B). However, we were not able to purify and sequence all products from

each individual, due to the low levels of some amplicons in some reactions. For example, we failed

to purify and sequence L1 from 85A.

After amplification of these putative minicircles, we attempted to amplify the maxicircle that would

be formed as the product of reciprocal recombination. If reciprocal recombination was responsible

for production of the minicircle, we should be able to detect a maxicircle; together, the minicircle

and maxicircle sequences should precisely match the sequence of the entire mt genome. To detect

the maxicircle, we designed primers that would amplify the A+T-rich region of the entire mt genome

(85-MR1 and 85-36R1, see Fig. 4.1A). Any maxicircles would be evident by the presence of shorter

59
than expected amplicons (as they would be missing the section encompassed by the minicircle). The

PCR amplification was performed and the only amplicon that was produced was one that represented

the A+T-rich region of the entire mt genome. To investigate whether there were low levels of

maxicircles present, we designed two internal primers (85-MR2 and 85-36R2) and performed nested

PCR. In this experiment, two PCR products, representing the A+T-rich region of the entire mt

genome and the maxicircle, were obtained. Sequence analysis of the maxicircle amplicon indicated

that it did correspond to the A+T-rich region of the entire mt genome (but lacking the minicircle

fragment), although two sites differed with the corresponding sequence of the A+T-rich region of the

entire mt genome (Fig. 4.1C). These sites were not close to the breakpoint junction. The inability to

amplify the maxicircle from genomic DNA, despite its smaller size compared with the entire mt

genome, might suggest that the subgenomic circles are not formed by reciprocal recombination. We

attempted to design a primer that straddles the putative breakpoint junction of the maxicircle in order

to perform reactions that would only amplify the maxicircle. However, this was unsuccessful due to

the lack of any sequence differences between the maxicircle breakpoint junctions. These junctions

occurred at two perfect (i.e., without any mismatches) microhomologies (21 bp) of non-tandem

repeat D, with one complete copy of repeat D kept in the maxicircle after formation. The breakpoint

junction primer would therefore anneal at three places: to the two copies of repeat D in the entire mt

genome, and the repeat D of the maxicircle (Supplementary Fig. 2,

http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1).

We then carried out sequence analysis of the most strongly amplified PCR product L2 amplified

from 85A, representing the minicircle. The sequence of L2 corresponded to the fragment deleted

from the mt genome, based on the maxicircle sequence information. It comprised Sec II and part of

60
Sec I and Sec III, although some variation existed when compared with the A+T-rich region of the

entire mt genome. In Sec II, the main difference was the length of the “TA” repeats. In Sec III, the

first copy of repeat B contained one site substitution while the last three copies were smaller; repeat

B is 10 bp long in the entire mt genome, but the last three copies are 9, 7, and 7 bp in the putative

minicircle. In addition, repeat D in Sec I was shorter, only containing the last 14 bp of that repeat

and with one site substitution. The breakpoint junctions occurred at two imperfect microhomologies

(14 bp) of non-tandem repeat D (Supplementary Fig. 2,

http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1). This observation indicates

that the minicircle and maxicircle sequences do not precisely match the entire mt genome. This

suggests that the subgenomic circles are not formed by reciprocal recombination, unless the

recombination occurred some time ago, with the subgenomic circles accumulating mutations since

that time. Alternatively, the level of variation might be consistent with the notion that direct,

precisely matching repeat sequences are not necessary for the repair of DSBs (Bacman et al., 2009;

Mita et al., 1990; Samuels et al., 2004; Srivastava and Moraes, 2005).

We then compared the sequence of the fragments of L2 amplified from each individual. They were

mostly identical, with the main difference among them being the length of the TA repeats in Sec II.

The fragments S1 and S2 were 58% and 22% of the length of L2 in 85A. L1 was 50 bp longer than

L2 in 85B and most of the length difference was due to the number of TA repeats (21 repeats). All

the alignment data are shown in supplementary figure 1A and 1B

(http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1).

61
Figure 4.3 Size separations of amplicons on 1% agarose gels. A. Putative minicircular amplicons obtained using
outward facing primers and genomic DNA (extracted from individual 85A). Primers were 85-ATF and 85-ATR. B.
minicircular PCR artifact obtained from plasmid DNA containing the A+T-rich region of the entire mt genome using
primers 85-ATF and 85-ATR. Although the size of the amplicon is similar to L2, sequencing indicated that the two
amplicons were distinct. C. minicircular amplicon obtained from genomic DNA extracted from individual 85A using
primers 85-ATF2 and 85-ATR. M, molecular weight marker (lambda DNA digested with EcoRI and HindIII).

Are the minicircles amplification artifacts?

As the A+T-rich region contained a complex series of repeats, the possibility remained that some of

these putative minicircles were artifacts of amplification. For example, PCR jumping, in which the

newly synthesized DNA strand jumps from one repeat to another, could produce small amplicons

that resembled those that would be produced by minicircles (or maxicircles). To investigate whether

the putative minicircles were artifacts caused by PCR jumping, we used plasmid DNA containing

cloned fragments of the A+T-rich region of the entire mt genome. In this way, we could examine

whether amplification of a template known not to contain minicircles could produce smaller

amplicons (i.e., artifactual minicircles), while sequencing would indicate whether they coincided

precisely with the putative minicircle amplicons characterized above (L1, L2, S1, and S2). In the

first set of experiments, we used a plasmid containing the A+T-rich region of the entire mt genome

as template, together with the same primers used to produce the minicircle amplicons (L1, L2, S1,

62
and S2). This reaction produced a single amplicon of size ~600 bp (Fig. 4.3B), but sequencing of this

fragment indicated that it was 639 bp and not similar to any of the minicircle amplicons. It contained

one copy of repeat C and D and nine copies of repeat B – this organization was not evident in either

the entire mt genome or in any of the minicircle amplicons. The amplification of a unique product

was surprising as we expected that molecules amplified from the cloned A+T-rich region should also

be amplified from genomic DNA: both templates contain the same sequence. We then investigated

whether the cloned A+T-rich region had rearranged during bacterial growth, but the sequence

precisely resembled the full-length A+T-rich region. This does not rule out the possibility that the

cloned DNA contained a small proportion of rearranged molecules, but we think it is more likely that

PCR jumping occurred during the amplification of the cloned A+T-rich region to a greater degree

compared with the genomic DNA amplification. This may be attributable to the much higher

concentration of target DNA. In the amplification utilizing the cloned A+T-rich region, the target is

not diluted by nuclear DNA, whereas in the genomic DNA amplification, the target (mt DNA)

represents only a fraction of the total DNA.

To investigate whether the presence of minicircles might also lead to PCR jumping, we used another

plasmid DNA containing the strongest amplified fragment 85A-L2. If PCR jumping did not occur,

we would expect to produce a single amplicon using the same PCR primers and conditions as used

in the original amplification. However, this amplification produced four bands, which corresponded

(in size) to each of the putative minicircles. Two of these amplicons were subcloned and sequenced

and found to be identical to the sequences of 85A-L2 and 85A-S1. We consider this strong evidence

that S1 is an artifact, produced by PCR template switching between the repeats present on the

minicircle. Although we could not purify and sequence the band corresponding to L1 and S2, the

63
coincidence in size leads us to suspect that L1 and S2 are also artifacts of PCR jumping.

To further test whether the stongest amplicon (L2) was a PCR artifact, we performed two

experiments. First, we designed a primer (85-ATF2) that straddled the putative “breakpoint

junction” sequence and together with 85-ATR performed an amplification using genomic DNA (Fig.

4.1B). The breakpoint junction primer should only anneal to a minicircle, as this sequence is absent

in the entire mt genome. A single sharp band was obtained (Fig. 4.3C) and the sequence data showed

that it was identical to the corresponding part of L2. In the second experiment, we used the plasmid

containing the entire A+T-rich region as template, together with the same primer pair to perform

amplification. As expected, no PCR product was obtained, indicating that the breakpoint junction

primer does not misprime somewhere in the A+T-rich region. The two experiments showed that L2

is unlikely to be an artifact.

Elimination of the possibility of numts

numts commonly exist in metazoans (Bensasson et al., 2001). In our protocol, the amplification of

minicircles was carried out using genomic DNA. Therefore, the possibility that the minicircle came

from numts cannot be excluded. In order to investigate this possibility, we used a primer pair

(85-ATF2 and 85-MRF2) that would amplify the remainder of L2, if it was present on a minicircle.

These primers are inward facing on a minicircle but would be outward facing on a nuclear

chromosome. As expected, this reaction produced an amplicon with sequence identical to a section

of sec II of the A+T-rich region and, together with L2, conceptually comprises a circular molecule of

about 700 bp. The amplified fragment length varied slightly among different clones (see alignment

data in supplementary Fig. 1C,

64
http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1). These differences might be

caused by polymerase error or represent minor variations present on different minicircles. The use of

outward-facing primers reduces the risk of amplifying numts, but does not completely exclude it. If a

numt is present as a tandem copy, outward facing primers would produce an amplicon. Indeed, if a

tandemly repeated numt existed which had a junction that precisely matched the breakpoints

identified in the putative minicircle L2, this template would produce amplicons with both of the pairs

of outward-facing primers described earlier (85-ATF and 85-ATR and 85-ATF2 and 85-MRF2).

However, a tandemly repeated numt would not produce the heterogeneity observed between clones

(all copies would be the same) nor the variation in repeat B and D that was observed in the

85-ATF/85-ATR amplification. Although the existence of a tandemly repeated numt cannot be

absolutely excluded, we consider its existence unlikely.

4.4 Discussion

Importance of control reactions

In this study, we report convincing evidence of the product of recombination (the minicircle) in the

mt genome of the parasitic wasp Conostigmus sp. However, we found that it was critical to establish

appropriate control reactions to identify and characterize artifacts produced by PCR jumping. We

suggest that this is especially important when repeats exist within the DNA template. Our

experiments indicated that it was necessary to perform control amplifications using both the isolated

(cloned) A+T-rich region and the putative minicircle in order to identify all non-biological products

of amplification. PCR artifacts were especially prevalent when using the cloned minicircle fragment,

suggesting that this sequence is more prone to PCR jumping. The main difference between these two

65
control sequences is the presence of the breakpoint junction sequence in the minicircle fragment.

Finally, our results suggest that it is critical to use breakpoint junction primers – primers that straddle

the putative breakpoint junction – in order to firmly establish the biological nature of putative

minicircles. Once the biological orgin of the minicircle was confirmed, it was also necessary to

demonstrate that the minicircle was derived from a circular (mt) molecule, as the existence of numts

might interfere with our identification of real recombination end products.

Possible explanations for low copy number of the maxicircle

The maxicircle detected in this study appeared to be present at very low copy number, as it could

only be detected by nested PCR. A low copy number of the maxicircle was also observed in human

tissues (Kajander et al., 2000). However, we interpret the existence of the maxicircle in our study

with some caution. As it could only be detected after nested PCR, we cannot exclude the possibility

that it was caused by PCR error. If, as we suspect, the minicircle contains the origin of replication

(discussed later), the absence of an origin of replication in the maxicircle would explain its presence

at very low copy number.

Comparison with minicircles from other Metazoa

The discovery of minicircles in a hymenopteran expands the number of taxonomic groups in which

minicircles have been characterized. Previously they have been characterized in humans (Holt et al.,

1988), the Phthiraptera (Cameron et al., 2011; Shao et al., 2009), and in nematodes (Armstrong et al.,

2000; Gibson et al., 2007; Lunt and Hyman, 1997), although the characteristics of the minicircles

vary between these groups. In humans, minicircles are generated by partial mt genome deletion, and

these minicircles coexist with the entire mt genome. In the Phthiraptera, three types of minicircles

66
have been identified: the first type is similar to the minicircles in humans (Cameron et al., 2011), the

second type has multiple genes and a short non-coding region (Cameron et al., 2011), and the third

type has one to three genes and a large non-coding region (Cameron et al., 2011; Shao et al., 2009).

In the nematodes, the minicircle of M. javanica is excised from the entire mt genome and coexsits

with both its reciprocal partner (maxicircle) and the entire mt genome (Lunt and Hyman, 1997);

minicircles found in the potato cyst nematode range from 6 to 9 kb (Armstrong et al., 2000; Gibson

et al., 2007). Some of the potato cyst nematode mincircles are mosaics, comprising multigenetic

fragments derived from other minicircles. The minicircle reported in the present study is most

similar to those reported in the nematode M. javanica, and the discovery of minicircles in four

taxonomically divergent groups indicates that they are likely to occur much more commonly than is

currently appreciated.

We found that there were both similarities and differences between the minicircle of Conostigmus sp.

and the minicircle reported by Lunt and Hyman (Lunt and Hyman, 1997) in the nematode M.

javanica. The minicircles were similar in that they were 1) both excised from the A+T-rich region of

the entire mt genome and 2) comprised of different repeats flanked by unique sequences. The

differences were mainly in two aspects. First, the breakpoints in the mt genome of Conostigmus sp.

occurred in two copies of a non-tandem repeat, and imperfect microhomologies (14 bp) could be

found around the breakpoints (supplementary Fig. 2,

http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1). In contrast, only short (3 bp)

or no direct repeats could be found around the breakpoints in the entire mt genome of the nematode

M. javanica. Second, nucleotide deletion occurred upstream of the second breakpoint in the

minicircles of all individuals examined in this study, when compared with the entire mt genome. No

67
such deletions were evident in M. javanica. Two candidate explanations might be responsible for the

nucleotide variation within an individual. First, the minicircles were not recently formed but instead

were generated long ago and transmitted to subsequent generations. Another explanation for the

nucleotide variation is that mismatching might be tolerated during recombination as perfect

microhomologies are not necessary for the repair of DSBs. Both explanations might also be

responsible for the observation that the minicircles are highly similar in related individuals.

Unfortunately, it is very difficult to experimentally distinguish between these two scenarios.

We suspect that the minicircles contain origin of replication information, and that these smaller

molecules are likely to replicate faster than larger ones (Diaz et al., 2002). This could explain the

higher levels of minicircle compared with the maxicircle.

Possible mechanism yielding the minicircle and maxicircle

The mechanism that produces mtDNA minicircles is still unclear. The characterization of a suite of

minicircles, from evolutionarily divergent organisms, is perhaps the best approach for identifying the

key mechanistic components. Thus far, the production and repair of DSBs appears to be associated

with the existence of sublimons (ΔmtDNA) in humans (Krishnan et al., 2008). Recent research of

rodent mtDNA showed that DSBs promoted recombination, resulting in the formation of sublimons

(Bacman et al., 2009; Srivastava and Moraes, 2005).

Mitochondria have been shown to have an effective system of DSB repair in Drosophila (Morel et

al., 2008). Krishnan et al. (Krishnan et al., 2008) proposed that 3’→5’exonucleases could generate

single-stranded regions at DSBs, and these exposed single-strand regions could subsequently anneal

at regions containing microhomologous sequences, and that this would result in the formation of

68
sublimons. We speculate that the minicircle and maxicircle of Conostigmus sp. were probably

generated by a similar mechanism. However, in order to generate both a minicircle and a maxicircle,

we propose that two breakpoints must occur simultaneously near or within the direct repeats (14 bp)

(Fig. 4.4). After the minicircle sequence was excised, another pair of microhomologous sequences

(up to 21 bp, the sizes varies depending on the position of the breakpoints) could be detected around

thebreakpoints (supplementary Fig. 2,

http://mbe.oxfordjournals.org/content/suppl/2013/12/11/mst255.DC1), which probably facilitated the

formation of the maxicircle. Ligase III has been identified to play an important role in maintaining

mtDNA integrity (Gao et al., 2011); the ends of the linearized DNA around the breakpoints were

probably joined by ligase III.

Figure 4.4 Model for the generation of minicircle and maxicircle mtDNA by recombination during repair of DSBs.
Direct repeats which facilitate the formation of minicircle and maxicircle mtDNA are shown in red and purple regions,
respectively.

Direct link between intra-mitochondrial recombination and mt gene rearrangement

The Hymenoptera is now the second insect order after the Phthiraptera in which minicircles have

been detected (Shao et al., 2009). Coincidently, both groups possess the most rearranged mt genomes

among insects. We suggest that there might be a direct link between these two phenomena. Although
69
intra-mitochondrial recombination has been considered as a possible mechanism for gene

rearrangement in animal mt genomes (Dowton and Campbell, 2001), there has been very little direct

evidence to support this. One reason for this is that the evidence, in the form of excised partial mt

genomes, direct repeats, or inverted repeats, may be erased soon after (in evolutionary terms) the

recombination event (Brockman and McFadden, 2012). The presence of minicircles, combined with

the highly rearranged mt genome reported in this study, supports the notion that recombination is an

important component of the mt gene rearrangement mechanism. Our working hypothesis is that the

generation of minicircles is a crucial first step in the rearrangement mechanism, but that these

minicircles may have to be long-lived (i.e. contain an origin of replication). The integration of these

minicircles back into the full-length mt genome, via homologous recombination, then generates gene

rearrangements. This hypothesis not only gives a good explanation to the existence of minicircles

but also accommodates the existence of the non-tandem repeat fragments that are found in some taxa,

such as those whose mt genome contains two or three copies of the control regions (Kumazawa et al.,

1996; Kurabayashi et al., 2008; Mueller and Boore, 2005). These duplicated control regions have

been postulated to be generated and maintained by recombination (Kurabayashi et al., 2008; Mueller

and Boore, 2005).

70
CHAPTER 5: Evolutionary dynamics of the
mitochondrial genome in the Evaniomorpha
(Hymenoptera) – a group with an intermediate rate of
gene rearrangement

This chapter is slightly modified from the paper:

Mao M, Gibson T, Dowton M (2014) Evolutionary dynamics of the mitochondrial genome in the

Evaniomorpha (Hymenoptera) - a group with an intermediate rate of gene rearrangement. Genome

Biol Evol. 6, 1862-1874. doi:10.1093/gbe/evu145.

5.1 Introduction

Knowledge of the forces that shape the organization of the animal mitochondrial (mt) genome has

suffered from a lack of suitable model systems. This is because the rate of gene rearrangement in

most animal mt genomes is generally extremely slow; for example most vertebrates have identically

arranged mt genomes (Boore, 1999). The ideal model system would provide multiple gene

rearrangements for comparison (in order to identify common themes), but in a lineage that is not

rearranging so frequently that hidden and convergent rearrangements obscure the interpretation of

evolutionary events. In the insects, a number of higher level lineages have been identified with

accelerated rates of mt gene rearrangement – the Hymenoptera (Dowton and Austin, 1999) and the

hemipteroid orders (Shao et al., 2001) are two examples. Within these rapidly rearranging lineages,

taxonomic sampling at lower levels may be all that is required to identify good model groups; that is,

ones in which the evolutionary trajectory of genome reorganization is straightforward to interpret. In

the present study, we investigated whether one group of the Hymenoptera (the Evaniomorpha) might

71
provide such a model system.

The Hymenoptera (sawflies, wasps, ants and bees) is one of the most important components of insect

diversity. As the third largest insect order, it is estimated to contain more than 140,000 extant species,

placed into two traditional suborders (Symphyta and Apocrita) (Huber, 2009). The Apocrita has long

been considered as a natural group and comprises more than 90% of the Hymenoptera (Huber, 2009;

Sharkey, 2007). Many apocritan species play valuable roles in biological control, ecosystem and

production of commercial products (LaSalle and Gauld, 1993).

Traditionally, the Apocrita is subdivided into two groups, the Aculeata (stinging wasps) and the

Parasitica (parasitoid wasps), with the Aculeata now widely accepted as being derived from within

the Parasitica (Sharkey, 2007). The reconstruction of a robust phylogeny for the Apocrita has long

been of great interest to hymenopteran systematists (reviewed by Sharkey, 2007). In 1988, Rasnitsyn

proposed a fully resolved phylogenetic hypothesis of higher level hymenopteran relationships based

on morphological and fossil evidence (Rasnitsyn, 1988). Despite the use of non-cladistic

methodology, Rasnitsyn’s research remains influential and sets a stage for current research of

hymenopteran phylogeny. In his hypothesis, the Apocrita was divided into four lineages, the

Ichneumonomorpha, the Vespomorpha (Aculeata), the Proctotrupomorpha and the Evaniomorpha

(Fig. 5.1). The Ichneumonomorpha and Aculeata have long been recovered as natural groups (Castro

and Dowton, 2006; Dowton and Austin, 2001; Dowton et al., 1997; Heraty et al., 2011; Klopfstein et

al., 2013; Rasnitsyn and Zhang, 2010; Sharkey et al., 2011; Vilhelmsen et al., 2010) and the

monophyly of the Proctotrupomorpha has been supported with several comprehensive studies

(Castro and Dowton, 2006; Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al., 2011).

However, the monophyly of the Evaniomorpha (including the Ceraphronoidea, Evanioidea,


72
Megalyroidea, Stephanoidea and Trigonalyoidea) remains unclear from both morphological and

molecular analyses. In morphological analyses, a subsequent numerical cladistic analysis of

Rasnitsyn’s (1988) data failed to retrieve the Evaniomorpha (Ronquist et al. 1999). Gibson (1999)

suspected that the mesocoxal articulatory structure, which supported the monophyly of the

Evaniomorpha according to Rasnitsyn (1988), was a retained symplesiomorphy rather than a

synapomorphy (Gibson 1999). Moreover, Rasnitsyn himself proposed that the Evaniomorpha was

not monophyletic and divided it into three lineages (Rasnitsyn and Zhang, 2010). Most molecular

analyses recovered the Evaniomorpha as paraphyletic, with the Aculeata nested within the

Evaniomorpha. Further, the sister group of the Aculeata varied among different analyses (Castro and

Dowton, 2006; Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al., 2011). Within the

Evaniomorpha, there is little consensus on the phylogenetic relationships among the superfamilies

due to a lack of reliable morphological characters and the limited number of molecular markers.

Figure 5.1 Simplified interpretation of the phylogenetic relationships among the major lineages within the Apocrita
from Rasnitsyn (1998).

Prior to this study, there were only four evaniomorph mt genomes available in GenBank,

representing three superfamilies [Evania appendigaster (Evanioidea: Evaniidae), Pristaulacus

compressus (Evanioidea: Aulacidae), Schlettererius cinctipes (Stephanoidea: Stephanidae) and

Conostigmus sp. (Ceraphronoidea: Megaspilidae)] (Dowton et al., 2009a; Mao et al., 2014a; Wei et

al., 2010b; Wei et al., 2013). Each of these mt genomes displayed a relatively small number of

73
changes to their mt genome organization, when compared with the ancestral

pancrustacean/hymenopteran organization, and these rearrangements involved both tRNA genes and

(to a lesser extent) protein-coding genes. Schlettererius had evidence of two tRNA gene

rearrangements (Dowton et al., 2009b), Pristaulacus had evidence of one protein-coding gene (nad1)

rearrangement and two tRNA gene rearrangements (Wei et al., 2013), Evania had evidence of four

tRNA gene rearrangements (Wei et al., 2010b), whereas Conostigmus had 2 protein-coding genes

rearranged (cox1, nad2), and 10 tRNA genes rearranged (Mao et al., 2014a). Thus this group

represents a good candidate model system in which to investigate the dynamics of mt genome

rearrangement, as individual rearrangements might be inferred through denser taxonomic sampling.

For this reason, in this study we present four new mt genomes for representatives of four

evaniomorph superfamilies, Megalyra sp. (Megalyroidea: Megalyridae), Orthogonalys pulchella

(Trigonalyoidea: Trigonalyidae), Ceraphron sp. (Ceraphronoidea: Ceraphronidae) and Gasteruption

sp. (Evanioidea: Gasteruptiidae). We also use these entire mt genomes to further investigate

relationships among the Evaniomorpha and Aculeata. Because of the likely paraphyletic nature of

the Evaniomorpha, we include the Aculeata in order to examine the phylogeny within a natural

group.

5.2 Materials and Methods

DNA extraction

The collection details for each study species are listed in Table 5.1. Genomic DNA was extracted

from 100% ethanol preserved specimens using the ‘salting out’ protocol (Aljanabi and Martinez,

1997). The DNA was resuspended in 100 μl of fresh TE solution (1 mM Tris–HCl, 0.1 mM EDTA

74
[pH 8]) and stored at 4°C.

Table 5.1 Species sequenced in this study, collection data and GeneBank accession numbers.

Species Library reference Family Superfamily Collection locality Accession number


Ceraphron sp. M247 Ceraphronidae Ceraphronoidea California, USA KJ570858
Gasteruption sp. M19 Gasteruptionidae Evanioidea Auburn, South Australia KJ619460
Megalyra sp. M272 Megalyridae Megalyroidea Wistow, South Australia KJ577600
Orthogonalys pulchella M16 Trigonalyidae Trigonalyoidea Clarke County, Washington, USA KJ619461

Mt genome amplification, sequencing, and annotation

Short gene fragments were amplified and sequenced using a range of universal insect mt primers

(Simon et al., 1994; Simon et al., 2006) and primers that had been previously designed from

consensus hymenopteran mt sequences. Using the sequence information obtained, taxon-specific

primers were designed for each sample to amplify the remaining regions. PCR and sequencing

reactions were conducted as previously described (Mao et al., 2014a). The sequences of both strands

were determined for the entire mt genomes of Ceraphron sp., Gasteruption sp., and O. pulchella, In

addition, we sequenced both strands for all coding regions and most of the A+T-rich region of

Megalyra sp., but failed to get the full double stranded sequence for the entire A+T-rich region due

to an array of 50 bp tandem repeats. This array contained 20 repeats, and as a result, no unique

internal sequencing primers could be designed.

Raw sequences were assembled into contigs in ChromasPro Ver 1.33 (Technelysium Ltd., Tewantin,

Australia). tRNA genes were identified using tRNA-scan SE 1.21 (lowelab.ucsc.edu/tRNAscan-SE/)

(Lowe and Eddy, 1997) and ARWEN 1.2 (http://130.235.46.10/ARWEN/) (Laslett and Canbäck,

2008). ORFinder (www.ncbi.nlm.nih.gov/gorf/gorf.html) was used to identify protein-coding genes,

specifying the invertebrate mt genetic code. The start and stop codons of some genes were corrected

75
according to the boundaries of tRNA genes and through alignment with other hymenopteran mt

sequences. rRNA genes were identified using BlastN and the ends of genes were assumed to extend

to the boundaries of the neighboring tRNA or protein-coding genes. However, the boundary between

rrnL and rrnS was difficult to define in Megalyra sp., as there is no gene between them. In this case,

we determined this boundary through sequence comparison with published hymenopteran mt rRNA

sequences.

Nucleotide composition, gene rearrangement, and repeat analyses

The A+T content was determined for the major strand of each of the four mt genomes by MEGA5

(Tamura et al., 2011). In order to investigate whether gene inversions influenced the nucleotide

compositional biases, we also measured the AT and GC skews for some protein-coding genes (nad2,

cox1, nad6, and cob) and two rRNA genes of 8 evaniomorph taxa and 12 aculeate taxa. These genes

were inverted in at least one of the four new mt genomes (see below). The formulae used were

AT-skew = (A - T) / (A + T) and GC-skew = (G - C) / (G + C) (Perna and Kocher, 1995).

Gene rearrangement analyses were conducted with CREx via the freely available CREx web server

(http://pacosy.informatik.uni-leipzig.de/crex) (Bernt et al. 2007). We compared gene orders using a

dissimilarity measurement – number of breakpoints and visually inspected the output diagram to

identify shared, derived gene rearrangements.

Direct and inverted repeats longer than 12 bp were identified for the A+T-rich region and the entire

mt genome of each evaniomorph taxa using UGENE v.1.13.1 (Okonechnikov et al., 2012). We

calculated the number of different size repeats around the breakpoints and in the A+T-rich region to

investigate whether any particular size of repeat facilitate gene rearrangement.

76
Sequence alignment and phylogenetic analysis

A total of 24 taxa were analyzed in this study, including 8 evaniomorph taxa, 12 aculeate taxa, and 4

symphytan taxa. Nucleotide sequences for each of the 13 protein-coding genes and the 2 rRNA

genes were imported into separate files using MEGA5 and aligned using Muscle (Edgar, 2004) or

MAFFT (Katoh et al., 2005). For the protein-coding genes (excluding the stop codons), an amino

acid alignment was generated first for each gene in Muscle as implemented within MEGA5, or

MAFFT at the freely available TranslatorX server (http://translatorx.co.uk/) (Abascal et al., 2010). A

nucleotide alignment was then inferred from the amino acid alignment. The alignment parameters

for all genes in Muscle were the default settings, which have been specified in a previous study (Mao

et al., 2012). For the MAFFT alignment of the rRNA genes, we used the G-INS-i algorithm as

implemented in the MAFFT web server (http://mafft.cbrc.jp/alignment/server/) (Katoh et al., 2005).

MAFFT (G-INS-i) has been shown to be more accurate than other programs (Golubchik et al., 2007).

No regions were excluded from the rRNA alignments. Individual gene alignments were

concatenated prior to phylogenetic analysis.

The best partitioning schemes and corresponding nucleotide substitution models were determined

with PartitionFinder version 1.0.1 (Lanfear et al., 2012) using the Bayesian Information Criterion

(BIC) and a heuristic search algorithm. A total of 41 data blocks were predefined (3 codon positions

of 13 protein-coding genes + 2 rRNA genes). Maximum Likelihood and Bayesian approaches were

employed to infer phylogenetic trees. Maximum Likelihood analyses were conducted with RAxML

via the available RAxML BlackBox server (http://embnet.vital-it.ch/raxml-bb/) (Stamatakis et al.,

2008). The GAMMA model of rate heterogeneity was employed for all partitions. The “Maximum

likelihood search” and “Estimate proportion of invariable sites” boxes were selected, with a total of
77
100 bootstrap replicates performed. Bayesian analyses were conducted with MrBayes v. 3.2.2

(Ronquist et al., 2012a) via the online CIPRES Science gateway portal (Miller et al., 2010). The

Markov chain Monte Carlo (MCMC) process was set so that four chains (three heated and one cold)

ran simultaneously. Four separate runs using unlinked partitions (unlink statefreq = all; unlink

revmat = all; unlink shape = all; unlink pinvar = all; prset applyto = all; ratepr = variable) for a total

of 1,000,000 generations were performed, with sampling every 100 generations. Stationarity for each

run was assessed by importing the parameter files into Tracer v. 1.5 (Rambaut and Drummond,

2009).

Rogue taxon identification

‘Rogue’ taxa, defined as those that have an unstable position in topological trees (Wilkinson, 1996),

can obscure relationships that are consistently recovered during bootstrap and bayesian analyses

(Aberer et al., 2013). In order to reveal those relationships that were consistently recovered by our

analyses, we conducted rogue taxon filtering using the approach implemented in RogueNaRok

(http://rnr.h-its.org/rnr) (Aberer et al., 2013). The RAxML bootstrap tree sets and Bayesian tree sets

(excluding burn-in) were employed to perform these analyses.

5.3 Results and Discussion

General features of mt genomes

Three complete and one nearly complete mt genomes were sequenced for this study: Ceraphron sp.

(14,947 bp), Gasteruption sp. (17,884 bp), O. pulchella (17,277 bp) and Megalyra sp. (18,996 bp).

For Megalyra sp., we were able to estimate the size of the genome with a reasonable degree of

precision, because imperfect copies exist in the array of 50 bp tandem repeats, for which we failed to

78
get the full double stranded sequence. Each genome contains all of the 37 genes commonly found in

animal mt genomes (Boore, 1999). In addition, a second copy of trnE was identified in the O.

pulchella mt genome. As the shortest hymenopteran complete mt genome to date, the Ceraphron sp.

mt genome is extremely compact with a total of only 105 bp of short non-coding regions (excluding

the A+T-rich region). The short non-coding regions in Megalyra sp., Gasteruption sp. and O.

pulchella are 345 bp, 909 bp and 919 bp long, respectively. The four mt genomes have a similar A+T

content, ranging from 80% (Ceraphron sp.) to 83.8% (O. pulchella). The AT and GC skew analysis

showed that gene inversions had no influence on nucleotide compositional biases (data not shown).

Three of the conventional start codons (ATA, ATG or ATT) could be assigned to all of the

protein-coding genes. Most tRNA genes have a typical cloverleaf structure. The exceptions are trnS1

in all four species, trnS2 in Ceraphron sp. and trnR in Ceraphron sp. and Gasteruption sp. All of

these tRNA genes lack the D-stem pairings in the dihydrouridine (DHU)-arm. The missing D-stem

has been commonly noted for the trnS1 gene in insects (Mao et al., 2012; Sheffield et al., 2008; Yang

et al., 2013), while to our knowledge, it has not been previously reported in trnS2 in insect mt

genomes. The missing D-stem in trnR is likely a common feature of hymenopteran mt genomes. In

addition to the two mt genomes reported here, it was also found in another two evaniomorph taxa

(Conostigmus sp. and S. cinctipes) (Dowton et al., 2009a; Mao et al., 2014a) and in two sceliond mt

genomes (Ceratobaeus sp. and Idris sp.) sequenced in our laboratory (Mao and Dowton, 2014).

Furthermore, the 3’ end of the aminoacyl acceptor stem of trnR overlaps with the downstream gene

in Ceraphron sp., Megalyra sp. and O. pulchella. The truncated aminoacyl acceptor stem is probably

restored to conventional structure by extensive tRNA editing (Dowton and Austin, 1999; Lavrov et

al., 2000; Masta and Boore, 2004; Segovia et al., 2011). The considerable difference in genome size

79
among the four evaniomorph taxa is mainly due to variation in the size of the A+T-rich region,

which ranges from 692 bp (Ceraphron sp.) to 3,871 bp (Megalyra sp.). One or more series of tandem

repeats could be found in the Megalyra sp., O. pulchella, and Gasteruption sp. mt genomes, whereas

only three short non-tandem repeats are present in the Ceraphron sp. mt genome (Fig. 5.2 and

Supplementary Fig. 1A, http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1).

Most notably, two copies of the A+T-rich region exist in the Gasteruption sp. mt genome, which are

separated by trnL2. The second copy is 77.7% of the length of the first copy and the length difference

is mainly due to the variation in repeat number of an 11 bp microsatellite sequence (Supplementary

Fig. 1B, http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1). Similarly, the

major portion of the A+T-rich region in the O. pulchella mt genome is also duplicated

(Supplementary Fig. 1C, http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1).

The presence of duplicated A+T-rich regions (control regions) has been reported in a diverse range

of taxa, such as mantellid frogs (Kurabayashi et al., 2008), Amazona parrots (Eberhard et al., 2001)

and Australasian Ixodes ticks (Shao et al., 2005). It is proposed that the presence of two control

regions may be advantageous and be maintained either through stabilizing selection or through gene

conversion (Eberhard et al., 2001).

Genome organizations

The four mt genome organizations are shown in Figures 5.3-5.5. All of them possess dramatic gene

rearrangements when compared with the ancestral pancrustacean mt genome organization (Cook,

2005), which is also thought to represent the ancestral organization of the hymenopteran mt genome

(Mao et al., 2012). Remarkably, each of the newly sequenced mt genomes has protein-coding or

rRNA gene rearrangement(s), a feature which has not been commonly reported in the Hymenoptera
80
[only 8 of the 38 available mt genomes (this number excludes congenerics) have protein-coding or

rRNA gene rearrangements] (Castro et al., 2006; Dowton et al., 2009b; Mao et al., 2014a; Oliveira et

al., 2008; Wei et al., 2010a; Wei et al., 2013; Xiao et al., 2011).

Figure 5.2 Structure of the A+T-rich region in the Ceraphron sp., Gasteruption sp., Megalyra sp. and Orthogonalys
pulchella mt genomes. TR1-N and NTR1-N indicate the distinct tandem repeats and non-tandem repeats in each
genome, respectively. The 80 bp repeat in the O. pulchella mt genome, which was difficult to classify as either
tandem or non-tandem, is labeled as R1. The length and copy number of each repeat are indicated above each repeat.
The small repeats that nest within the large repeats are not shown. The width of different boxes is not proportional but
indicative of the size of a repeat within a particular genome. Different repeats in each mt genome are indicated with
different colors. Boxes with asterisks underneath represent a partial copy of a repeat.

The mt genome of Ceraphron sp.

In Ceraphron sp., two protein coding genes (nad2 and nad6) and 10 tRNA genes have rearranged

relative to their ancestral positions. The tRNA gene rearrangements mainly occur at the nad3-nad5

junction and around the A+T-rich region. Four tRNA genes (trnA, trnR, trnN, and trnS1) have moved

from the nad3-nad5 junction to the tRNA gene cluster downstream of the A+T-rich region. trnE has

also moved out of the nad3-nad5 junction, and we propose that it might be involved in two gene
81
rearrangement events. First, it moved to the tRNA gene cluster downstream of the A+T-rich region

together with the other four tRNA genes, then it inverted and moved to the nad1-trnL1 junction

together with nad2. This is a more parsimonious scenario than independent movement with

inversion of both genes (trnE and nad2) to the same gene junction. We also compared the mt

genome organization of Ceraphron sp. with another ceraphronoid taxon, Conostigmus sp. (Fig. 5.3),

the sequence of which was reported in a previous study (Mao et al., 2014a). In both ceraphronoids,

the trnQ, trnN, trnS1, and trnV genes have moved to the tRNA gene cluster upstream of the cox1

gene, whereas the trnW gene has inverted. The nad2 gene has rearranged in both taxa but into

different positions. The rearrangement of nad2 has also been reported in two chalcidoid taxa,

Nasonia and Philotrypesis, in which nad2 has also moved into different positions (Oliveira et al.,

2008; Xiao et al., 2011). The most striking gene rearrangement in both ceraphronoids is that the two

rRNA genes are separated by protein-coding genes (nad6 in Ceraphron sp. and nad2 in Conostigmus

sp.) instead of trnV lying between them, which has not been observed in other Hymenoptera. We

consider that the existence of different protein-coding genes between the two rRNA genes might be

useful markers to resolve the internal relationships of the Ceraphronoidea.

82
Figure 5.3 Mitochondrial genome organization of two taxa from the Ceraphronoidea, Ceraphron sp. (present study)
and Conostigmus sp. (Mao et al. 2014), compared with the ancestral pancrustacean/hymenopteran mt genome
organization. tRNA genes are indicated by single-letter amino acid codes, L1, L2, S1 and S2 denote trnLCUN, trnLUUR,
trnSAGN and trnSUCN, respectively. Genes are transcribed from left to right except those indicated by underlining.
Different types of gene movements, relative to the ancestral organization, are indicated with different colors: blue
represents long-distance translocation; red represents long-distance translocation and inversion; purple represents
local translocation; green represents local inversion.

The mt genome of Gasteruption sp.

The Gasteruption sp. mt genome possesses 11 gene rearrangements (two protein-coding genes and

nine tRNA genes) and a duplication of the A+T-rich region (Fig. 5.4) relative to the ancestral

hymenopteran. The cox1 gene has inverted and moved into a position upstream of the A+T-rich

region. The nad2 gene has also inverted, but remains in its ancestral location. Among the rearranged

tRNA genes, most of them have inverted and moved to positions remote from their ancestral

locations, although trnY has inverted, but remains in its ancestral location (downstream of nad2). In

addition, trnQ has moved to a position downstream of nad2. The existence of duplicate A+T-rich

regions separated by trnL2 is consistent with the duplication/random loss model (Moritz et al., 1987).

A parsimonious explanation is the cox1 and trnL2 genes inverted and moved to the position upstream

83
of the A+T-rich region simultaneously (cox1-trnL2→trnL2-cox1), then a duplication of cox1, trnL2

and the A+T-rich region followed by random deletions of the redundant copy of the cox1 and trnL2

genes occurred. An important evidence for this explanation is that two copies of a 10 bp

complementary repeat exist at the beginning of rrnS and at the end of trnL2 (AAAAGTATTT in rrnS

and TTTTCATAAA in trnL2). We consider this short repeat may promote the cox1 and trnL2

inversion and long-range movement into the region between rrnS and A+T-rich region via

recombination (Dowton and Campbell, 2001; Mao et al., 2014a). We then compared the mt genome

organization of Gasteruption sp. with another two evanioid taxa, E. appendigaster and P.

compressus (Fig. 5.4). The gene positions are relatively conserved in E. appendigaster and P.

compressus with only four and three gene rearrangements, respectively, when compared with the

ancestral pancrustacean/hymenopteran. Although two tRNA genes (trnS2 and trnW) have inverted in

both Gasteruption sp. and Evania appendigaster, no mt gene rearrangements are shared among the

three taxa. Similarly, trnS2 is no longer adjacent to cob in all three taxa, but it has different, derived

positions. Thus, among the three sequenced evanioid mt genomes, we identified 18 independent

gene rearrangements, indicating that these could be extremely useful characters for assessing

phylogeny in this superfamily.

84
Figure 5.4 Mitochondrial genome organization of three taxa from the Evanioidea: Gasteruption sp. (present study),
Evania appendigaster (Wei et al., 2010b) and Pristaulacus compressus (Wei et al., 2010b). Genes are transcribed
from left to right except those indicated by underlining. Different types of gene movements, relative to the ancestral
organization, are indicated with different colors: blue represents long-distance translocation; red represents
long-distance translocation and inversion; purple represents local translocation; green represents local inversion.

The mt genomes of Megalyra sp. and O. pulchella

Figure 5.5 shows the mt genomes of Megalyra sp. and O. pulchella sequenced in this study.

Megalyra sp. is from the superfamily Megalyroidea, whereas O. pulchella is from the superfamily

Trigonalyoidea. The sister relationship of these two superfamilies has been recovered by some

molecular analyses (including this study, see below) (Castro and Dowton, 2006; Dowton and Austin,

2001; Dowton et al., 1997; Heraty et al., 2011; Klopfstein et al., 2013), so we illustrate them in the

one figure to facilitate comparisons. Both of them possess dramatic gene rearrangements when

compared with the ancestral pancrustacean/hymenopteran. In Megalyra sp., there are eight gene

rearrangements. The most striking rearrangement is the one in which the two rRNA genes have

inverted and moved into the nad2-cox1 junction. Inverted rRNA genes have been reported in other

insect orders, the Phthiraptera and Thysanoptera (Cameron et al., 2007a; Cameron et al., 2011;

Covacin et al., 2006; Shao and Barker, 2003). Interestingly, these four studies reported that both
85
rRNA genes were inverted, similar to our observations in the Megalyra sp. mt genome. We suspect

that there might be an advantage to having the two rRNA genes encoded on the same strand in

insects. One candidate mechanism responsible for this phenomenon is the transcriptional feedback

system regulating mtDNA replication. In this mechanism, the initiation of mtDNA replication is

hypothesized to be positively correlated with the transcription rate, with the latter negatively

regulated by the accumulation of gene products. Therefore, mtDNA replication is indirectly

regulated by the amount of synthesized products (Axel and Thomas, 2014). The rate of rRNA

synthesis was found to be up to 60-fold higher than mRNA synthesis (i.e. of protein-coding genes) in

humans due to the transcription termination factor binding mtDNA immediately downstream of the

two rRNA genes (Martin et al., 2005). A significant difference between the transcription levels of

rRNA and mRNA was also found in Drosophila (Torres et al., 2009). We suspect that the

hymenopteran mt genome might also have a higher rate of rRNA synthesis. Therefore, the

mechanism of transcriptional feedback system regulating mtDNA replication would confer an

advantage to maintaining both rRNA genes on the same strand, as regulation of the synthesis of both

rRNA genes would occur concomitantly. The retention of the rRNA genes on the minor strand in

other available evaniomoph mt genomes indicates that the inversion occurred after the divergence of

the Megalyroidea. In O. pulchella, 4 protein coding genes (nad2, cox1, nad6, and cob) and 10 tRNA

genes have rearranged relative to the ancestral positions. nad2 and cox1 have moved into the

position upstream of the A+T-rich region and the nad3-nad5 junction, respectively. nad6 and cob

have inverted and swapped positions (nad6-cob→cob-nad6). The gene boundary cob-nad6 was also

characterized in Venturia canescens (Ichneumonoidea) and Cotesia vestalis (Ichneumonoidea)

(Dowton et al., 2009b; Wei et al., 2010a), similar to our observations in O. pulchella. However,

86
these two superfamilies (Trigonalyoidea and Ichneumonoidea) are not closely related, indicating that

this arrangement has likely evolved on two or three occasions independently. Therefore, it is

probably one of the few examples of convergent evolution of gene order involving protein-coding

gene rearrangements. Two tRNA genes (trnQ and trnY) have inverted in both Megalyra sp. and O.

pulchella, but they are in different, derived positions. In Megalyra sp., both trnQ and trnY remain in

their ancestral locations, whereas in O. pulchella, they have moved to the position downstream of

rrnS. Furthermore, trnH is no longer between nad5 and nad4, but the derived positions are different

in the two taxa.

Figure 5.5 Mitochondrial genome organization of Megalyra sp. (Megalyroidea) (present study) and Orthogonalys
pulchella (Trigonalyoidea) (present study). The sister relationship of these two superfamilies is supported by present
study and some other molecular studies. Genes are transcribed from left to right except those indicated by underlining.
Different types of gene movements, relative to the ancestral organization, are indicated with different colors: blue
represents long-distance translocation; red represents long-distance translocation and inversion; purple represents
local translocation; green represents local inversion.

Evolutionary dynamics of evaniomorph mt genomes

A total of 64 gene rearrangements were identified in the seven evaniomorph mt genomes, including

4 local translocations (within a tRNA gene cluster and without inversion), 32 gene inversions and 49

long-range movements (some genes were involved in two rearrangement events). The local

translocations are readily explained by the duplication/random loss model (Moritz et al., 1987),

87
whereas the dominant rearrangement events (inversion and long-range movement) are more

consistent with the intramolecular recombination mechanism (Dowton and Campbell, 2001; Mao et

al., 2014a). In this mechanism, the repeat fragments located around the breakpoints may play an

important role in facilitating the recombination (Mao et al., 2014a). The A+T-rich region (control

region) has been suggested as a “hot spot” of recombination (Kurabayashi et al., 2008). Among the

64 gene rearrangements, there are 43 rearrangements that occurred close to the A+T-rich region. A

series of repeat sequences detected in each evaniomorph mt genome A+T-rich region may have

played a role in the gene rearrangements, via recombination (Table 5.2).

Table 5.2 The number of repeats (including direct and inverted repeats) within the A+T-rich region and gene
rearrangements identified for seven evaniomorph taxa.

Species 12-20 bp 21-50 bp 51-100 bp >100 bp Close* Not close* Inversion Long range Local
Ceraphron sp. 10 8 0 0 11 1 5 10 0
Conostigmus sp. 23 0 9 2 5 7 1 11 1
Evania appendigaster 35 31 0 0 4 0 3 3 1
Gasteruption sp. 50 0 0 2 7 4 10 7 0
Megalyra sp. 84 31 0 4 5 3 6 6 0
Pristaulacus 37 0 0 0 3 0 0 2 1
Orthogonalys pulchella 21
compressus 2 5 2 8 6 7 10 1

Note: *close/not close refers to whether the gene rearrangements occurred close to or far from the A+T-rich region.
The definition of ‘close to the A+T-rich region’ is: (1) when a tRNA gene moves to or moves out from a position
between the rrnS-nad2 gene junction, or (2) when a tRNA moves together with rrnS or nad2.

There have been few studies on the association of repeats with rearrangements in the mt genome. In

a recent study on the plastid genome (Weng et al., 2014), repeats were strongly associated with the

breakpoints in the rearranged genomes. In particular, large repeats (>20 bp and >60 bp) were

significantly correlated with the degree of genome rearrangement. In order to better understand the

relationship between repeats and gene rearrangements, we conducted a similar analysis for the

A+T-rich region and the entire mt genome among the seven evaniomorph taxa. However, no

significant correlations between the size of repeats and gene rearrangements could be detected (the

88
number of different size of repeats in the A+T-rich region is listed in table 5.2). We suspect that this

might be due to the high degree of divergence between the taxa studied here, with the number and

type of repeats that are present at the time of rearrangement being obscured by subsequent evolution.

The utility of gene rearrangements for assessing phylogeny in the Evaniomorpha

Gene rearrangement characters have been suggested as reliable markers for deducing phylogenetic

relationships (Boore and Brown, 1998; Dowton et al., 2002). To explore the potential phylogenetic

signals in the evaniomorph mt gene orders, such as shared, derived gene rearrangements, we visually

inspected each pairwise comparison of gene orders in the output diagrams from CREx. A large

number of individual gene rearrangements were detected between the evaniomorph mt genomes.

However, no shared, derived gene rearrangements could be identified. This result indicates that the

frequency of gene rearrangements might still be too high to be useful for assessing higher-level

relationships. We investigate this point further in the following section.

Table 5.3 Breakpoint numbers of the evaniomorph, hemipteroid and nematoceran mt genomes relative to the ancestral
mt genome organization.

Species Number of breakpoints Accession number Reference


Hemipteroids
Triatoma 0 NC_002609 Dotson and Beard 2001
Heterodoxus 35 NC_002651 Shao et al. 2001
Thrips 30 AF335993 Shao and Barker 2003
Lepidoscopid 15 AF335994 Shao et al. 2003
Diptera: Nematocera
Anopheles 5 NC_000875 Mitchell et al. 1993
Culicoides 0 NC_009809 Matsumoto et al, unpublished
Cramptonomyia 0 JN861747 Beckenbach 2012
Tipula 0 JN861743 Beckenbach 2012
Sylvicola 0 JN861752 Beckenbach 2012
Paracladura 21 JN861751 Beckenbach 2012
Arachnocampa 2 JN861748 Beckenbach 2012
Mayetiola 12 GQ387648 Beckenbach and Joy 2009
Hymenoptera: Evaniomorpha
Megalyra 16 KJ577600 Present study
Gasteruption 14 KJ619460 Present study
Evania 10 FJ593187 Wei et al. 2010b
Pristaulacus 5 KF500406 Wei et al. 2013
Conostigmus 16 KF015227 Mao et al. 2014
Ceraphron 18 KJ570858 Present study
Orthogonalys 21 KJ619461 Present study

89
Rates of gene rearrangement in the Evaniomorpha

Do the Evaniomorpha have an intermediate rate of gene rearrangement? To answer this question, we

calculated the breakpoint numbers of the evaniomorph, hemipteroid (hemipteran orders; Hemiptera,

Phthiraptera, Psocodea, Thysanoptera) and nematoceran (Diptera: Nematocera; crane flies, gnats,

and mosquitoes) mt genomes relative to the ancestral organization of insects. The hemipteroids

(excluding the Hemiptera) have been identified as a group with a high rate of mt gene rearrangement

(e.g. Shao et al., 2001, 2003), whereas most Nematocera have a low rate of mt gene rearrangement

(e.g., Beckenbach, 2012). As presented in Table 5.3, the breakpoint numbers of the seven

evaniomorph taxa range from 5 to 21, whereas the numbers of the three hemipteroid taxa (i.e.,

excluding Triatoma from the Hemiptera) are high (15-35), with most nematocerans low (0-5),

although Paracladura is an obvious exception. These numbers confirm the intermediate rate of gene

rearrangement of the Evaniomorpha. However, the inability to identify shared, derived gene

rearrangements indicates that the rate of the gene rearrangement is still too high to be reliably used

for reconstructing phylogenetic relationships. Therefore, increasing taxonomic sampling, particularly

in those evaniomorph groups with a relatively lower rate of gene rearrangement might be more

useful to establish phylogenetic relationships. Within the Evaniomorpha, the three evanioid taxa

yield the lowest breakpoint numbers, which suggests the Evanioidea is the most promising candidate

for further analyses. Although the monophyly of the Evanioidea was not recovered by some previous

research (Castro and Dowton, 2006; Dowton and Austin, 2001), two most recently comprehensive

studies robustly supported its monophyly (Heraty et al., 2011; Klopfstein et al., 2013).

Phylogenetic analysis

90
Previous phylogenetic analyses of hymenopteran mt genome data by our group has established that

uncontroversial relationships are only recovered with partitioned, model-based (Bayesian) analyses

of nucleotide data (Dowton et al., 2009a). Analysis of amino acid sequences using either maximum

parsimony or Bayesian approaches failed to recover these uncontroversial relationships, as did

analysis of nucleotide sequence data using the maximum parsimony criterion. Although our earlier

studies also found that the exclusion of third codon positions markedly improved the recovery of

these relationships, more recently we have found that the use of PartitionFinder (Lanfear et al., 2012)

now facilitates the inclusion of third codon positions (Mao and Dowton, 2014). PartitionFinder tends

to group proteins into the strand on which they are encoded, an approach that we did not use in our

earlier studies. For this reason, we analyzed the present dataset using model-based analyses of

nucleotide data which was partitioned by PartitionFinder. Nevertheless, in order to assess how

sensitive our phylogeny was to the analytical approach, we analyzed our dataset using eight

analytical approaches, which are outlined below.

We analyzed two taxon datasets (with and without rogue taxon deletion) using two alignment

approaches (Muscle and MAFFT) and two phylogenetic approaches (Maximum likelihood and

Bayesian inference). Our initial dataset contained all of the available evaniomorph and aculeate

species. The trees yielded a number of expected relationships with high support (posterior

probabilities 0.98-1.00 in Bayesian analyses and bootstrap proportions 57-100 in maximum

likelihood analyses). For example, the Ceraphronoidea, Apoidea, Formicidae and Vespidae were

consistently recovered as monophyletic (Fig. 5.6 and Supplementary Fig. 2,

http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1). However, some

traditionally monophyletic groups failed to be recovered in all of the analyses. For example, the two

91
chrysidoid genera (Cephalonomia and Primeuchroeus) failed to group together; Evania was placed

within the Aculeata and grouped as a sister of Radoszkowskius. This latter relationship was also

reported in a recent mt phylogeny of the Hymenoptera (Kaltenpoth et al., 2012). We then

investigated whether some taxa were being placed in variable positions during the bootstrap and

Bayesian analyses, and as a result were obscuring other, more stable relationships (Wilkinson, 1996).

We prefer this objective approach to taxon pruning, rather than basing taxon exclusion upon, for

example, inconsistency with previous phylogenetic hypotheses. Therefore, we conducted rogue

taxon identification on all of the tree sets using RogueNaRok. For the RAxML bootstrap tree sets,

Evania appendigaster and Primeuchroeus spp. were consistently identified as rogue taxa, whereas

only Primeuchroeus spp. was identified in Bayesian tree sets. For consistency between analyses, we

removed the concatenated sequences corresponding to the two rogue taxa (Evania and

Primeuchroeus) and repeated the sequence alignment (as removal of rogue taxa might impact on the

alignment) and phylogenetic analyses. The phylogenetic resolution was improved compared with the

initial analyses. For example, a monophyletic Aculeata was consistently recovered in all analyses.

Therefore, our discussion below focuses on the trees generated by the dataset with these two taxa

removed.

The general topologies were broadly congruent across the four analyses with similar levels of

support for major clades (Fig. 5.7 and Supplementary Fig. 3,

http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1). The two alignment

approaches had no effect on topology and limited effect on nodal support (Fig. 5.7 and

Supplementary Fig. 3A, http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1).

When the two phylogenetic inference approaches are compared, there is only one topological

92
difference, which was caused by the variable positions of Radoszkowskius (Vespoidea) and

Cephalonomia (Chrysidoidea). In Bayesian analyses, the two taxa grouped together and this clade

was sister to Apoidea + the remaining Vespoidea (Fig. 5.7 and Supplementary Fig. 3A,

http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1). However, in maximum

likelihood analyses, Radoszkowskius was recovered as the sister group to the remaining Aculeata and

Cephalonomia formed a clade with three formicid taxa (Camponotus, Pristomyrmex and Solenopsis)

(Supplementary Fig. 3B and C,

http://gbe.oxfordjournals.org/lookup/suppl/doi:10.1093/gbe/evu145/-/DC1). Nodal support was

generally stronger for the Bayesian than the maximum likelihood analyses as has been commonly

noted (Cameron et al., 2012; Nelson et al., 2012). The data matrix and the two Bayesian trees in Fig.

5.6 and 5.7 have been deposited in TreeBASE (Accession number S16039, http://treebase.org/).

Figure 5.6 Bayesian analysis of the initial taxon dataset (without rogue taxon deletion) based on mitogenomic
sequences including 13 protein-coding genes, and two rRNA genes. All the Symphyta, Evaniomorpha and Aculeata
were recovered as paraphyletic grades. Posterior probabilities are shown at each node.

93
Figure 5.7 Bayesian analysis of the reduced taxon dataset (with rogue taxon deletion) based on mitogenomic
sequences including 13 protein-coding genes, and two rRNA genes. Both the Symphyta and Evaniomorpha were
recovered as paraphyletic grades. Posterior probabilities are shown at each node.

In the pruned dataset, there are 22 hymenopteran taxa, with four taxa coming from the Symphyta.

Our analyses strongly supported the Cephoidea (Cephus) as sister to Orussoidea + Apocrita, which is

congruent with previously morphological studies (Vilhelmsen, 2001; Vilhelmsen et al., 2010). The

Orussoidea (Orussus) was recovered as sister to the Apocrita with high nodal support.

Within the Apocrita, the Aculeata was consistently recovered as monophyletic. Ignoring the unstable

positions of Radoszkowskius and Cephalonomia, the internal relationships are highly congruent with

our previous study (Mao et al., 2012). Compared with the previous study, there are two new

representatives: Camponotus from the Vespoidea and Philanthus from the Apoidea. As expected,

Camponotus formed a clade with another two vespoid taxa: Pristomyrmex and Solenopsis. All these

three taxa are from the same family Formicidae. Philanthus was recovered at the base of the

Apoidea.

The Evaniomorpha was recovered as paraphyletic, with the Aculeata nested within them. Of the

representatives from the Evaniomorpha, the Stephanoidea was placed as the sister group to the
94
remaining Evaniomorpha+Aculeata. The sister relationship of the Ceraphronoidea and Evanioidea

was also highly supported in each analysis. The Trigonalyoidea was recovered as sister to the

Megalyroidea, and this clade was sister to the Aculeata. A paraphyletic Evaniomorpha with respect

to the Aculeata has commonly been recovered in previous molecular studies (Castro and Dowton,

2006; Dowton and Austin, 2001; Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al., 2011).

However, the likely sister group to the Aculeata remains controversial. In our previous studies, the

Aculeata were generally recovered among a clade with the Stephanoidea, Trigonalyoidea and

Megalyroidea, but with weak support (Castro and Dowton, 2006; Dowton and Austin, 2001). A

recent analysis of combined morphological and molecular data supported a sister group of Aculeata

+ Evanioidea, but this clade was not supported by the morphology-only tree or individual gene trees

(Sharkey et al., 2011). Additionally, some recent molecular analyses recovered the same sister group

to the Aculeata as in this study, but the results were sensitive to the alignment approach (Heraty et al.,

2011; Klopfstein et al., 2013). In contrast, in this study the Aculeata + (Trigonalyoidea +

Megalyroidea) clade was insensitive to variations in alignment approach and phylogenetic inference

approach. We consider this clade to be the best supported hypothesis, based on current evidence. The

phylogenetic relationships among evaniomorph superfamilies are also far from settled. Since

Rasnitsyn (1988) placed the Stephanoidea within the Evaniomorpha, there has been no robust

morphological or molecular evidence to support this affiliation (Rasnitsyn, 1988). Instead, most

subsequent analyses placed the Stephanoidea as the most basal apocritan lineage (Dowton and

Austin, 1994; Heraty et al., 2011; Huber, 2009; Peters et al., 2011; Sharkey et al., 2011; Vilhelmsen

et al., 2010; Whitfield, 1992). Our results supported the basal position of the Stephanoidea, but failed

to give a clear indication to the affiliation of this superfamily. The sister relationship between the

95
Trigonalyoidea and Megalyroidea has been consistently proposed in molecular analyses (Castro and

Dowton, 2006; Dowton and Austin, 2001; Dowton et al., 1997; Heraty et al., 2011; Klopfstein et al.,

2013). However, recent morphological analyses favor the Ceraphronoidea as the sister to the

Megalyroidea based on mesosomal characters (Vilhelmsen et al., 2010). Our results confirmed the

Trigonalyoidea + Megalyroidea clade with robust molecular support and yielded a novel sister

relationship of the Ceraphronoidea and Evanioidea. The inclusion of other taxa is required to further

test this novel relationship.

To date, most molecular phylogenies on the internal relationships of the Apocrita have relied on

relatively short mt and nuclear gene fragments. The few studies that have used the much larger mt

genome to assess this group have done so with relatively few taxa – usually one representative from

each superfamily. Our study is the first attempt to use the entire mt genome, and a denser taxonomic

sampling, to reconstruct phylogeny. Nevertheless, we do so for just two of the four apocritan groups,

the Evaniomorpha and Aculeata. The molecular phylogenetic hypothesis presented here provides a

reliable basis for future analyses.

96
CHAPTER 6: Higher-level phylogeny of the
Hymenoptera inferred from mitochondrial genomes
This chapter is slightly modified from a paper that was submitted for publication in Molecular

Phylogenetics and Evolution and is currently in review:

Mao M, Gibson T, Dowton M. Higher-level phylogeny of the Hymenoptera inferred from

mitochondrial genomes.

6.1 Introduction

The Hymenoptera, comprising of sawflies, wasps, bees and ants, is one of the most speciose insect

orders. They live in all terrestrial and some aquatic habitats (Huber, 2009). The Hymenoptera

conventionally comprises two suborders, the Symphyta and the Apocrita, with 27 superfamilies (9

superfamilies in the Symphyta and 18 in the Apocrita) (Aguiar et al., 2013). Members of the

Symphyta are more primitive, with phytophagous lifestyle. The Orussidae is the only parasitoid

symphytan family (Sharkey, 2007). More than 90% of the described hymenopteran species belong to

the Apocrita (Huber, 2009). As one of the most biologically diverse suborder of insects, the Apocrita

exhibits a rich spectrum of lifestyles including parasitoidism, predation, omnivory, mycophagy and

secondary reversals to phytophagy (Huber, 2009; New, 2012).

Robust phylogenetic relationships could provide a framework to understand the evolution of the

various biologies exhibited by the Apocrita (Dowton and Austin, 2001). However, the higher-level

relationships within the Apocrita remain contentious (Sharkey, 2007). Based on morphological and

fossil evidence, Rasnitsyn (1988) proposed a new infraorder system, which formed a framework for

subsequent research of hymenopteran phylogeny. In this system, the Apocrita was divided into four

97
infraorders, the Ichneumonomorpha, the Vespomorpha (Aculeata), the Proctotrupomorpha and the

Evaniomorpha, with the Ichneumonomorpha sister to the Aculeata and the Proctotrupomorpha sister

to the Evaniomorpha. The monophyly of the Ichneumonomorpha and Aculeata was consistently

supported by subsequent morphological and molecular analyses (Castro and Dowton, 2006; Dowton

and Austin, 1994, 2001; Dowton et al., 1997; Heraty et al., 2011; Klopfstein et al., 2013; Ronquist et

al., 1999; Sharkey et al., 2011; Vilhelmsen et al., 2010). However, the sister relationship between the

two infraorders was only recovered by some morphological and early molecular analyses (Dowton

and Austin, 1994; Dowton et al., 1997; Rasnitsyn and Zhang, 2010; Vilhelmsen et al., 2010). Instead,

most of the later molecular analyses supported the Aculeata nested within the paraphyletic

Evaniomorpha and a sister relationship between the Ichneumonomorpha and the Proctotrupomorpha

(Dowton and Austin, 2001; Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al., 2011).

The Proctotrupomorpha (including the Proctotrupoidea, Cynipoidea, Diaprioidea,

Mymarommatoidea, Platygastroidea and Chalcidoidea) and Evaniomorpha (including the

Ceraphronoidea, Evanioidea, Megalyroidea, Stephanoidea and Trigonalyoidea) were two novel

concepts in Rasnitsyn’s infraorder system. The clade Evaniomorpha has not been consistently

recovered in both morphological and molecular analyses (Castro and Dowton, 2006; Gibson, 1999;

Heraty et al., 2011; Klopfstein et al., 2013; Rasnitsyn and Zhang, 2010; Ronquist et al., 1999;

Sharkey et al., 2011). For example, most recent analyses supported the Stephanoidea as the most

basal apocritan lineage (Heraty et al., 2011; Peters et al., 2011; Sharkey et al., 2011; Vilhelmsen et al.,

2010). Furthermore, recent molecular analyses recovered the Evaniomorpha as paraphyletic relative

to the Aculeata (Castro and Dowton, 2006; Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al.,

2011). Similarly, the monophyly of the Proctotrupomorpha was not supported by early

98
morphological or molecular analyses. In morphological analyses, Ronquist et al. (1999) reanalyzed

Rasnitsyn’s (1988) data using numerical cladistic method and the trees showed that the

Ceraphronoidea fell among the Proctotrupomorpha, close to the Chalcidoidea and Platygastroidea

(Ronquist et al., 1999). An early molecular analysis conducted by Dowton et al. (1997) using 37 taxa

based on 16S rRNA gene also failed to retrieve a monophyletic Proctotrupomorpha, as the

Cynipoidea was placed as a relatively basal apocritan lineage. But the reason for this might be the

limited taxon samplings (only two cynipoid representatives) (Dowton et al., 1997). Later, Dowton

and Austin (2001) performed multiple analyses with a dramatically increased taxon dataset and more

molecular markers (16S rRNA, COI and 28S rRNA). However, the Heloridae (Proctotrupoidea) was

recovered outside the Proctotrupomorpha in most of these analyses. Interestingly, when the

morphological characters were included, the Ceraphronoidea was placed within the

Proctotrupomorpha (Dowton and Austin, 2001). Although later and more comprehensive molecular

analyses consistently recovered the Proctotrupomorpha as monophyletic, the internal relationships

remain elusive (Castro and Dowton, 2006; Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al.,

2011). For example, Sharkey (2007) separated the Diaprioidea from the Proctotrupoidea as a new

superfamily based on previous molecular evidence (Castro and Dowton, 2006; Dowton and Austin,

2001), but the monophyly of the Diaprioidea was not well supported by later molecular analyses

(Heraty et al., 2011; Klopfstein et al., 2013).

The entire mitochondrial (mt) genome is a popular molecular marker used in phylogenetic studies

mainly because of its maternal inheritance and higher rate of nucleotide substitution compared with

the nuclear DNA (Moritz et al., 1987). In addition, some other mt genome characters, such as gene

rearrangements, also provide useful phylogenetic information (Boore and Brown, 1998). However,

99
poor representation of a broad range of lineages restricts the evolutionary utility of the mt genome.

Among the insect orders, hymenopteran mt genomes are not well represented (Cameron, 2014).

Therefore, the main purpose of the present study is to increase our understanding of the phylogeny

of the Hymenoptera by extending the taxonomic range of available mt genome sequences. Here, we

present three new mt genomes for representatives of three hymenopteran superfamilies: Ibalia

leucospoides (Cynipoidea), Monomachis antipodalis (Diaprioidea) and Pelecinus polyturator

(Proctotrupoidea). The mt genomes from the superfamilies Cynipoidea and Diaprioidea are

sequenced for the first time. We conducted phylogenetic analyses using these new mt genomes

together with the previously published mt genomes of the Hymenoptera. For the first time, we

present analyses of mt genome data from representatives of every extant apocritan superfamily, with

the exception of Mymarommatoidea. A range of analytical approaches were employed. In addition,

we compared the organization of the newly sequenced mt genomes with other Proctotrupomorpha

taxa and with the putative ancestral organization for the Hymenoptera, in order to identify any

shared, derived gene rearrangements.

6.2 Materials and Methods

DNA extraction

The collection details for each species sequenced for this study are listed in Table 6.1. All specimens

were stored at 4°C in the UOW wasp alcohol library before extraction. Genomic DNA was extracted

from 100% ethanol preserved specimens using the ‘salting out’ protocol (Aljanabi and Martinez,

1997). The DNA was resuspended in 100 μl of fresh TE solution (1 mM Tris–HCl, 0.1 mM EDTA

[pH 8]) and stored at 4°C.

100
Table 6.1 Species sequenced in this study, collection data, and GeneBank accession numbers.

Species Library reference Family Superfamily Collection locality Accession number


Ibalia leucospoides M4 Ibaliidae Cynipoidea Mt Gambier, South Australia KJ814197
Monomachis antipodalis M241 Monomachidae Diaprioidea ACT, Australia KM104168
Pelecinus polyturator M22 Pelecinidae Proctotrupoidea Ottawa, Canada KM104167

Mt genome amplification, sequencing, annotation, and bioinformatic analysis

PCR amplifications and sequencing reactions were conducted as previously described (Mao et al.,

2014a). Briefly, short fragments were amplified and sequenced using a range of universal insect

mitochondrial primers (Simon et al., 1994; Simon et al., 2006) and primers that had been previously

designed from consensus hymenopteran mt sequences. Using the sequence information obtained,

taxon-specific primers were designed for each sample to amplify the remaining regions by long PCR.

The primer walking method was employed to determine the complete, double stranded sequence for

each long PCR product.

Raw sequences were assembled into contigs in ChromasPro Ver 1.33 (Technelysium Ltd., Tewantin,

Australia). tRNA genes were identified using tRNA-scan SE 1.21 with a cove cutoff score of 5

(lowelab.ucsc.edu/tRNAscan-SE/) (Lowe and Eddy, 1997) and ARWEN 1.2

(http://130.235.46.10/ARWEN/) (Laslett and Canbäck, 2008). ORFinder

(www.ncbi.nlm.nih.gov/gorf/gorf.html) was used to identify protein-coding genes, specifying the

invertebrate mt genetic code. The start and stop codons of some genes were corrected according to

the boundaries of tRNA genes and through alignment with other hymenopteran mt sequences.

Ribosomal RNA genes were identified by sequence comparison with published hymenopteran mt

ribosomal RNA sequences.

Nucleotide compositions were determined using MEGA5 (Tamura et al., 2011). The AT and GC

101
skews were measured for the major (J) strand of each genome. The formulae used were AT-skew =

(A - T) / (A + T) and GC-skew = (G - C) / (G + C) (Perna and Kocher, 1995).

Sequence alignment and phylogenetic analysis

Nucleotide sequences for each of the 13 protein-coding genes and the 2 rRNA genes were imported

into separate files using MEGA5 (Tamura et al., 2011) and aligned using Muscle (Edgar, 2004) or

MAFFT (Katoh et al., 2005). For the protein-coding genes (excluding the stop codons), an amino

acid alignment was generated first for each gene in Muscle as implemented within MEGA5, or

MAFFT at the freely available TranslatorX server (http://translatorx.co.uk/) (Abascal et al., 2010). A

nucleotide alignment was then inferred from the amino acid alignment. The alignment parameters

for all genes in Muscle were the default settings, which have been specified in a previous study (Mao

et al., 2012). For the MAFFT alignment of the rRNA genes, we used the G-INS-i algorithm as

implemented in the MAFFT web server (http://mafft.cbrc.jp/alignment/server/) (Katoh et al., 2005).

Individual gene alignments were concatenated prior to phylogenetic analysis.

The best partitioning schemes and corresponding nucleotide substitution models were determined

with PartitionFinder version 1.0.1 (Lanfear et al., 2012) using the Bayesian Information Criterion

(BIC) and a heuristic search algorithm with branch lengths linked and unlinked, respectively. Two

data blocks were predetermined for analyses: 41 data blocks [2 rRNA genes + 3 codon positions of

13 protein-coding genes (rrnS + rrnL + PCG123)] and 28 data blocks [2 rRNA genes + 1st and 2nd

codon positions of 13 protein-coding genes (rrnS + rrnL + PCG12)]. Maximum Likelihood and

Bayesian approaches were employed to infer phylogenetic trees. Analyses were performed using

MrBayes v. 3.2.2 (Ronquist et al., 2012a) and RAxML v. 8.0.9 (Stamatakis, 2014) via the online

102
CIPRES Science gateway portal (Miller et al., 2010). For Maximum likelihood analyses, a total of

1000 bootstrap replicates were performed with the GTRGAMMA substitution model applied to all

partitions. For Bayesian analyses, the Markov chain Monte Carlo (MCMC) process was set so that

four chains (three heated and one cold) ran simultaneously. Four separate runs using unlinked

partitions for a total of 1,000,000 generations were performed, with sampling every 100 generations.

Stationarity for each run was assessed by importing the parameter files into Tracer v. 1.5 (Rambaut

and Drummond, 2009).

6.3 Results and Discussion

General features of mt genomes

One complete mt genome of I. leucospoides (17,212 bp) and two nearly complete mt genomes of P.

polyturator (14,896 bp) and M. antipodalis (14,066 bp) were sequenced and have been deposited in

GenBank (Table 6.1). We were unable to amplify a portion of the A+T-rich region for P. polyturator

and the entire A+T-rich region, three tRNA genes (trnI, trnM, and trnQ) and part of nad2 and rrnS

for M. antipodalis. All 37 genes typical of animal mt genomes were identified in both I. leucospoides

and P. polyturator mt genomes (Boore, 1999). However, unlike most animal mt genomes, the I.

leucospoides mt genome contains two extra trnM genes and inverted, duplicated A+T-rich regions.

The three trnM genes are highly similar, with 92% – 97% sequence identities. They are separated by

35 bp and 41 bp short non-coding regions. The two inverted copies of the A+T-rich region are 457

bp and 848 bp long, respectively. The first copy is a perfect inverted repeat of the major portion of

the second copy. The presence of duplicated A+T-rich regions (control regions) has been reported in

a diverse range of lineages, such as mantellid frogs (Kurabayashi et al., 2008), Amazona parrots

103
(Eberhard et al., 2001), plague thrips (Shao and Barker, 2003) and Australasian Ixodes ticks (Shao et

al., 2005). However, to our knowledge, the inverted, duplicated A+T-rich regions have not been

previously reported in other animals. The 100% sequence identity indicates that (1) the inverted

duplication occurred very recently; or (2) the two inverted A+T-rich regions might evolve in concert

as has been reported for several duplicated control regions (Eberhard et al., 2001; Kurabayashi et al.,

2008).

Table 6.2 Characteristics of the three mitochondrial genomes sequenced in this study.

Species Length (bp) A+T content (%) AT-skew GC-skew


Ibalia leucospoides 17217 86.3 -0.032 -0.012
Monomachis antipodalis 14066 83.8 -0.045 0.138
Pelecinus polyturator 14896 81.2 -0.031 -0.341

Some general characteristics of the three genomes are given in Table 6.2. As widely reported in other

hymenopteran mt genome, the nucleotide base composition is similar for the three mt genomes with

a strong AT bias (81.2% – 86.3%). Each of the three mt genomes exhibits negative AT-skews for the

J strand, ranging from -0.045 to -0.031. The GC-skews of the J strand in the I. leucospoides and P.

polyturator mt genomes are negative (-0.012 and -0.341, respectively) while the M. antipodalis mt

genome shows a positive GC-skew (0.138). It has been proposed that the reversal of strand

asymmetry might be caused by inversion of the replication origin (Hassanin et al., 2005). All of the

protein-coding genes start with the typical ATA, ATG or ATT codon and stop with the completed

(TAA or TAG) or truncated (TA or T) termination codons. Several short non-coding regions were

identified in each mt genome. The P. polyturator mt genome contains a total of 219 bp of short

non-coding regions with the largest one of 59 bp between nad4L and trnT genes. The sizes of the

short non-coding regions within the M. antipodalis mt genome are highly variable, ranging from 1

bp to 298 bp. A total of 982 bp of short non-coding regions was identified in the I. leucospoides mt
104
genome; the largest one (256 bp) contains an array of 9 bp tandem repeats (TTATTAATC or

TTATTAATT) present in 29 copies (the last one is a partial copy).

Figure 6.1 Mitochondrial genome organization of Ibalia leucospoides, Pelecinus polyturator, and Monomachis
antipodalis, compared with the ancestral pancrustacean/hymenopteran mt genome organization. tRNA genes are
indicated by single-letter amino acid codes, L1, L2, S1 and S2 denote trnLCUN, trnLUUR, trnSAGN and trnSUCN,
respectively. Genes are transcribed from left to right except those indicated by underlining. Gene movements, relative
to the ancestral organization, are indicated with arrows.

Genome organizations

The three mt genome organizations are shown in Figure 6.1. Several gene rearrangements were

revealed in each mt genome when compared with the organization of the ancestral

pancrustacean/hymenopteran (Cook, 2005; Mao et al., 2012). In M. antipodalis, three tRNA genes

(trnS2, trnY, and trnV) have rearranged relative to their ancestral positions. All of them have inverted

and moved to the cox1 – cox2 gene junction. The derived gene rearrangements were not found in

other proctotrupomorph taxa, indicating they occurred during the divergence of Monomachis. There

are four tRNA gene rearrangements in the P. polyturator mt genome relative to the ancestral

105
pancrustacean/hymenopteran: trnQ has swapped position with trnI; trnM has moved to the cob –

nad1 gene junction; trnE and trnF have inverted and swapped positions. We then compared the

organization of P. polyturator with another species of the Proctotrupoidea, Vanhornia eucnemidarum

(Castro et al., 2006), to investigate whether any of the tRNA gene rearrangements were shared

within this superfamily. However, no shared gene rearrangements could be identified in the two taxa.

Although trnM is no longer in the tRNA gene cluster upstream of nad2 in both taxa, it has different,

derived positions. In V. eucnemidarum, it has inverted and moved to the nad1 – rrnL gene junction.

We could not compare the position of trnI and trnQ as they failed to be amplified in the V.

eucnemidarum mt genome. The I. leucospoides mt genome has extensive rearrangements involving

15 tRNA genes as well as 7 protein-coding genes. The most striking one is a large inversion of a

group of genes from trnE to trnS2, genes from this group were found in three separate, derived

positions. trnE and trnF remain in their ancestral positions; genes from nad5 to cob have moved to a

position downstream of rrnS; trnS2 has moved to a position downstream of the A+T-rich region. In

addition, the inverted copy of the A+T-rich region lies next to trnE. We propose that the A+T-rich

region might have first duplicated and moved to the position downstream of trnS2 via a

duplication/random loss model (Moritz et al., 1987), then become inverted with the trnE – trnS2 gene

block simultaneously via recombination (Dowton and Campbell, 2001; Mao et al., 2014a). The

breakpoints occurred in the nad3-nad5 and cob-nad1 gene junctions. This explanation is not only a

more parsimonious scenario than several independent inversion events, but also consistent with the

hypothesis that the A+T-rich region is important for the formation of long-lived minicircles during

recombination (Mao et al., 2014a). The inverted trnE – trnS2 gene block is unique when compared

with other proctotrupomorph mt genomes. However, this is not the first report of large gene block

106
inversion in the Proctotrupomorpha. In previous studies, large inversions of cox1-cox3 or cox1-nad3

gene blocks have been reported in three chalcidoid wasps (Oliveira et al., 2008; Xiao et al., 2011).

Interestingly, the breakpoints in these mt genomes also occurred in the nad3-nad5 gene junction. We

suspect that there might be some structural characteristics in this region that facilitate the occurrence

of double strand breaks, and that repair of these breaks might lead to gene rearrangement (Mao et al.,

2014a).

Table 6.3 Summary of the major clades recovered by different dataset and analytical approach. Support values are
shown for each recovered relationship (Posterior probability for Bayesian analyses and bootstrap proportion for
maximum likelihood analyses). Abbreviations: -, not recovered; PCG123, 1st, 2nd and 3rd codon positions of 13
protein-coding genes; PCG12, 1st and 2nd codon positions of 13 protein-coding genes; BI, Bayesian analysis; ML,
maximum likelihood analysis; Link, branch lengths = linked in PartitionFinder; Unlink, branch lengths = unlinked in
PartitionFinder; rEV, the remaining Evaniomorpha; PR, Proctotrupomorpha; IC, Ichneumonomorpha; Cy, Cynipoidea;
Di, Diaprioidea; Ch, Chalcidoidea.

rrnS+rrnL+PCG123 rrnS+rrnL+PCG12
BI_Muscle BI_MAFFT ML_Muscle ML-MAFFT BI_Muscle BI_MAFFT ML_Muscle ML-MAFFT
Clade
Link Unlink Link Unlink Link Unlink Link Unlink Link Unlink Link Unlink Link Unlink Link Unlink
Orussoidea + Apocrita 0.91 0.74 1 1 - - 78 86 1 1 1 1 61 64 86 90
Stephanoidea + the remaining Apocrita 1 1 0.96 0.93 86 89 77 79 1 1 0.99 0.98 76 75 70 76
(Evaniidae + Aculeata) + (rEV+PR+IC) 0.92 1 0.98 - - - - - 0.72 0.8 - 0.82 34 34 - -
Aculeata - 0.48 - - 51 47 - - 0.7 - - - - - - -
Evanioidea + Aculeata - - - - - - - - - - 0.88 - - - 17 -
Apoidea 1 1 1 1 100 99 100 99 1 1 1 1 100 100 99 100
Chrysidoidea 0.34 - 0.63 0.42 - - - - - - - - - - - -
Trigonalyoidea + Megalyroidea 1 1 1 1 100 100 100 100 1 1 1 1 100 100 100 100
Ichneumonomorpha 1 1 1 1 100 100 100 100 1 1 1 1 100 100 100 100
(Trigonalyoidea + Megalyroidea) + IC 1 0.98 1 1 85 84 85 81 1 1 1 1 60 57 78 76
Ceraphronoidea + Evanioidea 0.62 0.79 0.53 - - - - - 0.89 0.64 - 0.63 - 22 - -
Ceraphronoidea + Platygastroidea - - - 0.63 48 46 53 56 - - - - 50 - - 47
Proctotrupomorpha - - - - - - - - 0.9 0.63 0.69 0.65 - 33 43 -
Ceraphronoidea + Proctotrupomorpha - - - - - - - - - - 0.88 - - - 32 -
(Ceraphronoidea + Evanioidea) + PR - - - - - - - - 0.78 0.58 - - - 17 - -
Cynipoidea + Proctotrupoidea - - - - 17 20 37 36 - - - - - - - 26
Diaprioidea + Chalcidoidea 0.99 0.85 1 0.98 74 70 92 87 0.9 0.69 0.98 0.74 66 57 78 72
Cynipoidea + (Diaprioidea + Chalcidoidea) 1 1 1 1 - - - - 0.99 1 1 1 50 50 42 -
Proctotrupoidea + (Cy + (Di + Ch)) - - - - - - - - 1 1 1 0.98 68 67 63 -
Proctotrupoidea + Platygastroidea 0.66 0.94 0.66 - - - - - - - - - - - - -

Phylogenetic analysis

107
Phylogenetic analyses were performed using 48 complete or partial mt genomes, representing 14

apocritan and 3 symphytan superfamilies. We analyzed two datasets (rrnS + rrnL + PCG123 and rrnS

+ rrnL + PCG12) using two alignment approaches (Muscle and MAFFT), two partition schemes

(branch lengths = linked and unlinked) and two phylogenetic approaches (Maximum likelihood and

Bayesian inference). Previous phylogenetic studies of mt genome data have indicated that Bayesian

analyses were superior to parsimony analyses (Castro and Dowton, 2007; Dowton et al., 2009a). For

this reason, we do not conduct parsimony analyses in the present study. A total of 11 topologies of

the relationships among the apocritan superfamilies were produced by the 16 separate analyses. As

shown in Table 6.3, some clades were consistently recovered by all analyses. However, the

relationships between the four infraorders and within the Proctotrupomorpha and Evaniomorpha

vary widely between datasets and across analytical approaches. Our recent study found that the use

of PartitionFinder facilitated the inclusion of 3rd codon positions when investigating the phylogeny

of more closely related taxa (Mao and Dowton, 2014). However, the taxa included here are more

remotely related, and excluding 3rd codon positions produced topologies more consistent with other

analyses of higher-level relationships. For example, a monophyletic Proctotrupomorpha was only

recovered by the rrnS + rrnL + PCG12 dataset (Fig. 6.2). The analytical approaches also influenced

the topology and the level of nodal support. We discuss this in more detail below.

Outgroup and basal relationships of the Apocrita

Our analyses generally support the Orussoidea as the sister group to the Apocrita, which is consistent

with previous morphological and molecular analyses (Klopfstein et al., 2013; Rasnitsyn, 1988;

Vilhelmsen, 2001; Vilhelmsen et al., 2010). Two maximum likelihood analyses of the rrnS + rrnL +

PCG123 dataset aligned by Muscle recovered the clade (Orussoidea + Cephoidea) + Apocrita, but
108
the nodal supports for the sister group of Orussoidea + Cephoidea were very low (BP = 32 / 42).

Within the Apocrita, the Stephanoidea was consistently placed as the most basal apocritan lineage

with high nodal support, a placement that conflicts with Rasnitsyn’s infraorder system (1988), but

consistent with a number of subsequent studies (Heraty et al., 2011; Peters et al., 2011; Sharkey et al.,

2011; Vilhelmsen et al., 2010).

Figure 6.2 Bayesian analysis of the Muscle alignment dataset (branch lengths = linked in PartitionFinder) based on
mitogenomic sequences including first and second codon positions from 13 protein-coding genes, and the two rRNA
genes. Posterior probabilities are shown at each node.

Aculeata and Ichneumonomorpha

The Aculeata was recovered as monophyletic in four Muscle-aligned analyses. It failed to be

109
retrieved in other analyses due to the placement of the non-aculeate genus Evania, which grouped as

a sister of Radoszkowskius (Aculeata: Mutillidae). This relationship has been recovered in some

analyses in our previous study and a recent mitochondrial phylogeny of the Hymenoptera

(Kaltenpoth et al., 2012; Mao et al., 2012). In our previous study, a monophyletic Aculeata was

recovered when excluding 3rd codon positions. By contrast, excluding 3rd codon positions did not

improve the resolution of a monophyletic Aculeata here. Within the Aculeata, the Apoidea was

strongly supported as monophyletic and placed within a paraphyletic Vespoidea, which is in

agreement with previous morphological and molecular analyses (Debevec et al., 2012; Heraty et al.,

2011; Pilgrim et al., 2008; Sharkey et al., 2011; Vilhelmsen et al., 2010). The two chrysidoid taxa

(Cephalonomia and Primeuchroeus) were also nested within the Vespoidea, but they failed to group

together in most of the analyses. Apparently, a better taxon sampling is needed to provide more

stable resolution. Two vespoid families, Vespidae and Formicidae, were monophyletic across all

analyses.

The Ichneumonomorpha was consistently and strongly supported in our analyses. The internal

relationships were robust to the variation of datasets and analytical approaches and highly congruent

with our mt genome previous study (Mao et al., 2012).

Rasnitsyn (1988) proposed a sister relationship between the Aculeata and the Ichneumonomorpha

based on the presence of ovipositor valvilli and distinct articulation involving a pair of propodeal

condyles. However, this sister relationship was not supported by a subsequent cladistic analysis of

his data (Ronquist et al., 1999). Although early molecular studies using 16S rRNA data retrieved the

clade Aculeata + Ichneumonomorpha (Dowton and Austin, 1994; Dowton et al., 1997), most of the

later molecular analyses recovered the Aculeata nested within the Evaniomorpha and grouped the
110
Ichneumonomorpha and the Proctotrupomorpha together (Dowton and Austin, 2001; Heraty et al.,

2011; Klopfstein et al., 2013; Sharkey et al., 2011). Our analyses consistently recovered a novel

clade with high nodal support – (Trigonalyoidea + Megalyroidea) + Ichneumonomorpha, while the

position of the Aculeata within the Apocrita remains variable. Most of the Bayesian analyses (except

rrnS + rrnL + PCG123_MAFFT_Unlink and rrnS + rrnL + PCG12_MAFFT_Link analyses)

supported the Evaniidae + Aculeata as sister to the remaining Evaniomorpha + Ichneumonomorpha

+ Proctotrupomorpha. Two analyses of the rrnS + rrnL + PCG12 dataset aligned by MAFFT

recovered a close relationship between the Evanioidea and the Aculeata, but the nodal support was

weak (PP = 0.88 and BP = 17).

Evaniomorpha

The Evaniomorpha is the most controversial concept in Rasnitsyn’s infraorder system (1988). The

monophyly and internal relationships remain problematic in both morphological and molecular

analyses (summarized in Mao et al., 2014b). Our recent mitochondrial phylogeny based on 8

evaniomorph and 12 aculeate taxa robustly recovered the Trigonalyoidea + Megalyroidea as the

sister group to the Aculeata and a sister relationship between the Ceraphronoidea and the Evanioidea

(Mao et al., 2014b). The sister group relationship of Trigonalyoidea + Megalyroidea was strongly

confirmed in our present analyses, but they were recovered as the sister group to the

Ichneumonomorpha rather than the Aculeata. The clade Ceraphronoidea + Evanioidea was generally

supported by Bayesian analyses, indicating this relationship is sensitive to the phylogenetic inference

approach.

Proctotrupomorpha

111
The Proctotrupomorpha was generally recovered as monophyletic by the rrnS + rrnL + PCG12

dataset (except BI_Muscle_Unlink and ML_MAFFT_Unlink analyses). The Ceraphronoidea +

Evanioidea clade or the Ceraphronoidea was placed as the sister group to the Proctotrupomorpha.

These two sister relationships are at odds with most previous molecular studies that found a sister

relationship between the Ichneumonomorpha and the Proctotrupomorpha (Dowton and Austin, 2001;

Heraty et al., 2011; Klopfstein et al., 2013; Sharkey et al., 2011). In the remaining analyses, the

Platygastroidea fell outside the Proctotrupomorpha and grouped with the Ceraphronoidea in six

maximum likelihood and one Bayesian analyses. This is not unprecedented, as the sister group of

Platygastroidea + Ceraphronoidea has been proposed by Ronquist et al. (1999) based on

morphological characters and supported by some analyses based on molecular and morphological

data in Dowton and Austin (2001). But the difference is that the Ceraphronoidea fell within the

Proctotrupomorpha in these two studies (Dowton and Austin, 2001; Ronquist et al., 1999). Within

the Proctotrupomorpha, the Diaprioidea was consistently recovered as the sister group to the

Chalcidoidea. The Diaprioidea is a relative new superfamily separated from the Proctotrupoidea by

Sharkey (2007) based on molecular evidence (Castro and Dowton, 2006; Dowton and Austin, 2001).

The placement of this superfamily was unstable in early molecular analyses (Dowton and Austin,

2001). However, later analyses supported a sister relationship with the Chalcidoidea (Castro and

Dowton, 2006; Heraty et al., 2011). Our results supported this sister relationship with moderate to

high nodal support. The support values were higher in the analyses using the rrnS + rrnL + PCG123

dataset, indicating the 3rd codon positions contributed phylogenetic signal to support this relationship.

There is little consensus concerning the placement of the Cynipoidea in previous analyses. Rasnitsyn

(1988) placed the Cynipoidea as a sister group to the Diapriidae + Monomachidae + Austroniidae

112
clade (Rasnitsyn, 1988). Early molecular analyses favored the Cynipoidea as a relatively basal

apocritan lineage (Dowton et al., 1997). A later molecular study based on expanded taxon sampling

and molecular data recovered the Cynipoidea as a part of the Proctotrupomorpha, but its placement

was highly variable (Dowton and Austin, 2001). Although recent molecular analyses generally

recovered a sister relationship of Cynipoidea + Platygastroidea (Castro and Dowton, 2006; Dowton

and Austin, 2001; Heraty et al., 2011; Klopfstein et al., 2013), a recent morphological study

supported a close relationship between the Cynipoidea and the Diapriidae (Vilhelmsen et al., 2010).

Contrary to previous studies, our results recovered the Cynipoidea as a sister group to the

Proctotrupoidea (5 maximum likelihood analyses) or to the clade Diaprioidea + Chalcidoidea (all

Bayesian analyses + 3 maximum likelihood analyses). The support values for the clade Cynipoidea +

(Diaprioidea + Chalcidoidea) were much higher than the clade Cynipoidea + Proctotrupoidea. The

placement of the Proctotrupoidea was also unstable in our analyses. Except for the clade Cynipoidea

+ Proctotrupoidea, the remaining analyses recovered it as a sister group to the Cynipoidea +

(Diaprioidea + Chalcidoidea) (7 analyses based on the rrnS+rrnL+ PCG12 dataset) or

Platygastroidea (3 Bayesian analyses based on the rrnS+rrnL+ PCG123 dataset).

6.4 Conclusions

Although our study provides the most comprehensive mitochondrial phylogeny of the higher-level

relationships within the Hymenoptera, many relationships are far from resolved, especially within

the Proctotrupomorpha and the Evaniomorpha. Considering the limited representatives of mt

genomes in each hymenopteran superfamily, a denser taxon sampling is needed for future studies.

For example, only one cynipoid taxon–I. leucospoides from a small family Ibaliidae–was present

in our analyses. Further taxon sampling from other cynipoid families, such as the two largest
113
families Figitidae and Cynipidae, should be included to clarify the placement of the Cynipoidea

within the Proctotrupomorpha. Further taxon sampling might also prove useful to provide

phylogenetic signal in regions of the topology where internodes are relatively short. As shown in Fig.

6.2, the internode of the clade Proctotrupoidea + (Cynipoidea + (Diaprioidea + Chalcidoidea))

recovered in our analyses was very short at the base when compared with external branches. In

recent comprehensive studies, a close relationship among the Proctotrupoidea, Chalcidoidea,

Diaprioidea and Mymarommatoidea was strongly recovered (Heraty et al., 2011; Klopfstein et al.,

2013; Sharkey et al., 2011). Therefore, inclusion of taxa from the Mymarommatoidea might be a

good choice to provide signal in these problematic areas.

114
CHAPTER 7: General Conclusions

7.1 Main outcomes

This study provides fresh insight into the mt genomes of the Hymenoptera and their utility for the

recovery of higher-level phylogeny. Eleven complete or nearly complete mt genomes representing

eight apocritan superfamilies were sequenced, which dramatically expanded the cladistic diversity of

the mt genome dataset of the Hymenoptera. With these new mt genomes, a more accurate estimate of

the degree of gene rearrangements was obtained for two major apocritan lineages: the Evaniomorpha

and Proctotrupomorpha. Work was also undertaken to investigate the possible mechanisms for gene

rearrangement. The most important finding was the coexistence of minicircular and a highly

rearranged mtDNA in the megaspilid wasp Conostigmus sp., which provided evidence to support the

notion that recombination is an important aspect of the gene rearrangement mechanism. The first mt

genome phylogeny of the Apocrita with a complete representation of superfamilies (excluding

Mymmaromatoidea) was conducted with a range of phylogenetic analytical approaches. Some robust

and highly supported relationships were recovered, indicating that the mt genome is a useful marker

to resolve higher-level relationships within the Hymenoptera.

7.2 Recommendations and future research

7.2.1 Increasing taxon sampling of the Apocrita

The expanded taxon sampling facilitated the investigation of mt gene rearrangements and their

utility to resolve phylogenetic relationships within the Hymenoptera. However, no shared, derived

gene rearrangements could be identified between and within the superfamilies of the Evaniomorpha

and Proctotrupomorpha, indicating the frequency of gene rearrangements of these groups is too high

115
to be useful for assessing relationships between families. By contrast, comparison of mt genome

organizations within the subfamily Scelioninae suggested that gene rearrangements have much

potential as phylogenetic markers at the subfamily level. Therefore, additional work should be

focused on obtaining denser taxonomic sampling at the subfamily level for the investigation of gene

rearrangements. The groups with a relatively lower rate of gene rearrangement would be a good

starting point, such as the evanioid families in the Evaniomorpha and the diaprioid families in the

Proctotrupomorpha.

This study provided the first comprehensive mitochondrial phylogeny of the higher-level

relationships within the Apocrita. However, the taxon sampling used in this study was still limited,

with no representatives from the Mymmaromatoidea and no more than three representatives from

each of the evaniomorph superfamilies and the remaining proctotrupomorph superfamilies. The

limited taxon sampling resulted in the position and monophyly of some superfamilies remaining

unresolved, such as the Evanioidea, Cynipoidea and Platygastroidea. Future taxonomic sampling

would need to be included in a more comprehensive study. Representative taxa from the

Mymmaromatoidea would be the first option. Additionally, increasing taxon sampling in those

superfamilies, whose position was variable across analytical approaches, is also essential.

The long PCR amplification of complete mt genomes is the most widely used tool in insects. In this

study, taxon-specific primers were designed for each sample to improve the efficiency of the large

fragment amplification. However, large gene (protein-coding or rRNA genes) rearrangements in

some mt genomes brought some difficulties to the amplification. Thus, different primer

combinations were tested, which was a time-consuming process. Furthermore, the area surrounding

the A+T-rich region has proven difficult to amplify in the Hymenoptera (Castro and Dowton, 2006;
116
Dowton et al., 2009a). Although most of the mt genomes in this study were amplified entirely

through optimizing PCR conditions, designing more specific primers for the hymenopteran mt

genomes or using nested PCR, the entire mt genome sequence for Monomachis antipodalis

(Diaprioidea) and Pelecinus polyturator (Proctotrupoidea) failed to be amplified. The missing data

in some mt genomes may not have a significant impact on phylogenetic analysis (Dowton et al.,

2009a), but it has been shown to hinder the identification of shared, derived gene rearrangements in

some groups in this study, such as the Chalcidoidea (Chapter 2) and Proctotrupoidea (Chapter 6). So

it is also important to obtain the entire mt genome (at least the complete coding region) for further

taxon sampling. The recently developed "Next-Generation Sequencing" (NGS) technologies, such as

Roche and Illumina, provide a robust high-throughput, low-cost platform for de novo sequencing of

mt genomes, either from specific amplicons or directly from total genomic DNA. These technologies

can rapidly obtain large amounts of mt genomes from multiple individuals in one reaction and

substantially improve coverage and sequencing depth. Furthermore, a mapping assembly approach,

MITObim, for reconstructing mt genomes directly from raw NGS data has also been developed

recently. This approach is applicable for specific assembly of mitochondrial genomes from mixed

samples containing different species (Hahn et al., 2013). NGS technologies have already been used

for generating densely sampled trees from whole mt genomes in two large insect orders, the

Coleoptera and Lepidoptera (Gillett et al., 2014; Timmermans et al., 2014). Future studies could

therefore consider using NGS technologies to construct large dataset of hymenopteran mt genomes.

7.2.2 Incorporating genome-based collection of markers

In Chapter 5, topologies reconstructed with taxa from the Evaniomorpha and Aculeata robustly

supported the Trigonalyoidea + Megalyroidea as the sister group to the Aculeata. However, this
117
sister relationship collapsed when the Proctotrupomorpha and Ichneumonomorpha were included to

reconstruct the apocritan phylogeny in Chapter 6. The conflicting results demonstrate the limits of

the present dataset to resolve higher-level relationships of the larger and more diverse Hymenoptera

(or Apocrita). In addition to a broader taxon sampling mentioned above, incorporation of other

phylogenetic markers would also be beneficial to improve the resolution in further studies. A recent

phylogenetic analysis of relationships among families and higher groupings of the Diptera provides a

good example. This study reconstructed the evolution of a major branch of the fly tree of life based

on 14 nuclear loci, complete mt genomes and 371 morphological characters(Wiegmann et al., 2011).

So far, a very limited number of nuclear markers, such as 18S, 28S and EF-1α, have been used for

the reconstruction of hymenopteran phylogeny (Castro and Dowton, 2006; Heraty et al., 2011).

Sharanowski et al.’s (2010) research based on EST data was a good attempt to explore a large

number of molecular loci (24 ESTs) to resolve relationships among hymenopteran superfamilies

(Sharanowski et al., 2010). However, the taxon sampling was limited in their analyses and the

topologies were sensitive to analytical methods. Given the advantage of NGS technologies, further

studies could consider incorporating genome-based collection of markers from large taxon sampling.

In particular, nuclear protein-coding genes would be promising candidates that could extend our

knowledge of higher-level relationships of the Hymenoptera.

7.3 Conclusions

This PhD research has made a major progress in understanding the evolution of hymenopteran mt

genomes and their utility as phylogenetic markers. The large number of gene rearrangements

detected here has facilitated the investigation of the mechanisms of gene rearrangement, which will

further promote more realistic analyses of gene rearrangement characters in the Hymenoptera and
118
other invertebrate groups. This work has also provided a reliable basis for future phylogenetic

studies on the Hymenoptera, which will rely on increased taxon sampling and genomic sampling.

119
References

Abascal, F., Zardoya, R., Telford, M.J., 2010. TranslatorX: multiple alignment of nucleotide

sequences guided by amino acid translations. Nucleic Acids Res. 38, 7-13.

Aberer, A.J., Krompass, D., Stamatakis, A., 2013. Pruning rogue taxa improves phylogenetic

accuracy: an efficient algorithm and webservice. Syst. Biol. 62, 162-166.

Aguiar, A.P., Lohrmann, V., Mikó, I., Ohl, M., Rasmussen, C., Taeger, A., Yu, D.S.K., Deans, A.R.,

Engel, M.S., Forshage, M., Huber, J.T., Jennings, J.T., Johnson, N.F., Lelej, A.S., Longino, J.T., 2013.

Order Hymenoptera. Zootaxa 3703, 51-62.

Akasaki, T., Nikaido, M., Tsuchiya, K., Segawa, S., Hasegawa, M., Okada, N., 2005. Extensive

mitochondrial gene arrangements in coleoid Cephalopoda and their phylogenetic implications. Mol.

Phylogenet. Evol. 38, 648-658.

Aljanabi, S.M., Martinez, I., 1997. Universal and rapid salt-extraction of high quality genomic DNA

for PCR-based techniques. Nucleic Acids Res. 25, 4692-4693.

Amer, S.A.M., Kumazawa, Y., 2007. The mitochondrial genome of the lizard Calotes versicolor and

a novel gene inversion in South Asian draconine agamids. Mol. Biol. Evol. 24, 1330-1339.

Armstrong, M.R., Blok, V.C., Phillips, M.S., 2000. A multipartite mitochondrial genome in the

potato cyst nematode Globodera pallida. Genetics 154, 181-192.

Asakawa, S., Kumazawa, Y., Araki, T., Himeno, H., Miura, K., Watanabe, K., 1991. Strand-specific

nucleotide composition bias in echinoderm and vertebrate mitochondrial genomes. J. Mol. Evol. 32,

511-520.

120
Austin, A., Field, S., 1997. The ovipositor system of scelionid and platygastrid wasps (Hymenoptera:

Platygastroidea): comparative morphology and phylogenetic implications. Invertebr. Syst. 11, 1-87.

Austin, A.D., Johnson, N.F., Dowton, M., 2005. Systematics, evolution, and biology of scelionid and

platygastrid wasps. Annu. Rev. Entomol. 50, 553-582.

Awadalla, P., Eyre-Walker, A., Smith, J.M., 1999. Linkage disequilibrium and recombination in

hominid mitochondrial DNA. Science 286, 2524-2525.

Axel, K., Thomas, B.L.K., 2014. Transcription could be the key to the selection advantage of

mitochondrial deletion mutants in aging. Proc. Natl. Acad. Sci. USA 111, 2972-2977.

Bacman, S.R., Williams, S.L., Moraes, C.T., 2009. Intra- and inter-molecular recombination of

mitochondrial DNA after in vivo induction of multiple double-strand breaks. Nucleic Acids Res. 37,

4218-4226.

Bakker, F.T., Culham, A., Gomez-Martinez, R., Carvalho, J., Compton, J., Dawtrey, R., Gibby, M.,

2000. Patterns of nucleotide substitution in angiosperm cpDNA trnL (UAA)-trnF (GAA) regions.

Mol. Biol. Evol. 17, 1146-1155.

Beckenbach, A.T., 2011. Mitochondrial genome sequences of representatives of three families of

scorpionflies (Order Mecoptera) and evolution in a major duplication of coding sequence. Genome

54, 368-376.

Beckenbach, A.T., 2012. Mitochondrial genome sequences of Nematocera (lower Diptera): evidence

of rearrangement following a complete genome duplication in a winter crane fly. Genome Biol. Evol.

4, 89-101.

121
Beckenbach, A.T., Joy, J.B., 2009. Evolution of the mitochondrial genomes of gall midges (Diptera:

Cecidomyiidae): rearrangement and severe truncation of tRNA genes. Genome Biol. Evol. 1,

278-287.

Beckenbach, A.T., Stewart, J.B., 2009. Insect mitochondrial genomics 3: the complete mitochondrial

genome sequences of representatives from two neuropteroid orders: a dobsonfly (order Megaloptera)

and a giant lacewing and an owlfly (order Neuroptera). Genome 52, 31-38.

Bensasson, D., Zhang, D.X., Hartl, D.L., Hewitt, G.M., 2001. Mitochondrial pseudogenes:

evolution's misplaced witnesses. Trends Ecol. Evol. 16, 314-321.

Bernt, M., Braband, A., Middendorf, M., Misof, B., Rota-Stabelli, O., Stadler, P.F., 2013.

Bioinformatics methods for the comparative analysis of metazoan mitochondrial genome sequences.

Mol. Phylogenet. Evol. 69, 320-327.

Bernt, M., Merkle, D., Middendorf, M., 2008. An algorithm for inferring mitogenome

rearrangements in a phylogenetic tree. Comparative Genomics International Workshop,

RECOMB-CG 2008, Lecture Notes in Bioinformatics (LNBI). Springer, pp. 143-157.

Bernt, M., Merkle, D., Ramsch, K., Fritzsch, G., Perseke, M., Bernhard, D., Schlegel, M., Stadler,

P.F., Middendorf, M., 2007. CREx: inferring genomic rearrangements based on common intervals.

Bioinformatics 23, 2957-2958.

Birky, J.C.W., 2001. The inheritance of genes in mitochondria and chloroplasts: laws, mechanisms,

and models. Annu. Rev. Genet. 35, 125-148.

Boore, J.L., 1999. Animal mitochondrial genomes. Nucleic Acids Res. 27, 1767-1780.

122
Boore, J.L., 2000. The duplication/random loss model for gene rearrangement exemplified by

mitochondrial genomes of deuterostome animals. Comparative genomics. Kluwer Academic

Publishers, pp. 133-147.

Boore, J.L., 2006. The use of genome-level characters for phylogenetic reconstruction. Trends Ecol.

Evol. 21, 439-446.

Boore, J.L., Brown, W.M., 1998. Big trees from little genomes: mitochondrial gene order as a

phylogenetic tool. Curr. Opin. Genet. Dev. 8, 668-674.

Boore, J.L., Collins, T.M., Stanton, D., Daehler, L.L., Brown, W.M., 1995. Deducing the pattern of

arthropod phylogeny from mitochondrial DNA rearrangements. Nature 376, 163-165.

Boore, J.L., Lavrov, D.V., Brown, W.M., 1998. Gene translocation links insects and crustaceans.

Nature 392, 667-668.

Boore, J.L., Staton, J.L., 2002. The mitochondrial genome of the Sipunculid Phascolopsis gouldii

supports its association with Annelida rather than Mollusca. Mol. Biol. Evol. 19, 127-137.

Brockman, S.A., McFadden, C.S., 2012. The mitochondrial genome of Paraminabea aldersladei

(cnidaria: anthozoa: Octocorallia) supports intramolecular recombination as the primary mechanism

of gene rearrangement in octocoral mitochondrial genomes. Genome Biol. Evol. 4, 994-1006.

Burger, G., Gray, M.W., Franz Lang, B., 2003. Mitochondrial genomes: anything goes. Trends Genet.

19, 709-716.

Cameron, S.L., 2014. Insect mitochondrial genomics: implications for evolution and phylogeny.

Annu. Rev. Entomol. 59, 95-107.

123
Cameron, S.L., Dowton, M., Castro, L.R., Ruberu, K., Whiting, M.F., Austin, A.D., Diement, K.,

Stevens, J., 2008. Mitochondrial genome organization and phylogeny of two vespid wasps. Genome

51, 800-808.

Cameron, S.L., Johnson, K.P., Whiting, M.F., 2007a. The mitochondrial genome of the screamer

louse Bothriometopus (Phthiraptera: Ischnocera): effects of extensive gene rearrangements on the

evolution of the genome. J. Mol. Evol. 65, 589-604.

Cameron, S.L., Lambkin, C.L., Barker, S.C., Whiting, M.F., 2007b. A mitochondrial genome

phylogeny of Diptera: whole genome sequence data accurately resolve relationships over broad

timescales with high precision. Syst. Entomol. 32, 40-59.

Cameron, S.L., Lo, N., Bourguignon, T., Svenson, G.J., Evans, T.A., 2012. A mitochondrial genome

phylogeny of termites (Blattodea: Termitoidae): robust support for interfamilial relationships and

molecular synapomorphies define major clades. Mol. Phylogenet. Evol. 65, 163-173.

Cameron, S.L., Miller, K.B., D'Haese, C.A., Whiting, M.F., Barker, S.C., 2004. Mitochondrial

genome data alone are not enough to unambiguously resolve the relationships of Entognatha, Insecta

and Crustacea sensu lato (Arthropoda). Cladistics 20, 534-557.

Cameron, S.L., Sullivan, J., Song, H.J., Miller, K.B., Whiting, M.F., 2009. A mitochondrial genome

phylogeny of the Neuropterida (lace-wings, alderflies and snakeflies) and their relationship to the

other holometabolous insect orders. Zool. Scr. 38, 575-590.

Cameron, S.L., Yoshizawa, K., Mizukoshi, A., Whiting, M.F., Johnson, K.P., 2011. Mitochondrial

genome deletions and minicircles are common in lice. BMC Genomics 12, 394.

124
Carey, D., Murphy, N.P., Austin, A.D., 2006. Molecular phylogenetics and the evolution of wing

reduction in the Baeini (Hymenoptera: Scelionidae): parasitoids of spider eggs. Invertebr. Syst. 20,

489-501.

Castro, L.R., Dowton, M., 2005. The position of the Hymenoptera within the Holometabola as

inferred from the mitochondrial genome of Perga condei (Hymenoptera: Symphyta: Pergidae). Mol.

Phylogenet. Evol. 34, 469-479.

Castro, L.R., Dowton, M., 2006. Molecular analyses of the Apocrita (Insecta: Hymenoptera) suggest

that the Chalcidoidea are sister to the diaprioid complex. Invertebr. Syst. 20, 603-614.

Castro, L.R., Dowton, M., 2007. Mitochondrial genomes in the Hymenoptera and their utility as

phylogenetic markers. Syst. Entomol. 32, 60-69.

Castro, L.R., Ruberu, K., Dowton, M., 2006. Mitochondrial genomes of Vanhornia eucnemidarum

(Apocrita: Vanhorniidae) and Primeuchroeus spp. (Aculeata: Chrysididae): evidence of rearranged

mitochondrial genomes within the Apocrita (Insecta: Hymenoptera). Genome 49, 752-766.

Cha, S.Y., Yoon, H.J., Lee, E.M., Yoon, M.H., Hwang, J.S., Jin, B.R., Han, Y.S., Kim, I., 2007. The

complete nucleotide sequence and gene organization of the mitochondrial genome of the bumblebee,

Bombus ignitus (Hymenoptera: Apidae). Gene 392, 206-220.

Chen, W.J., Koch, M., Mallatt, J.M., Luan, Y.X., 2014. Comparative analysis of mitochondrial

genomes in Diplura (Hexapoda, Arthropoda): taxon sampling is crucial for phylogenetic inferences.

Genome Biol. Evol., 6, 105-120.

Clary, D.C., Wolstenholme, D.R., 1985. The Mitochondrial DNA Molecule of Drosophila yakuba:

125
Nucleotide Sequence, Gene Organization, and Genetic Code. J. Mol. Evol. 22, 252-271.

Clayton, D.A., Doda, J.N., Friedberg, E.C., 1974. The absence of a pyrimidine dimer repair

mechanism in mammalian mitochondria. Proc. Natl. Acad. Sci. USA 71, 2777-2781.

Cook, C.E., 2005. The complete mitochondrial genome of the stomatopod crustacean Squilla mantis.

BMC Genomics 6, 105.

Covacin, C., Shao, R., Cameron, S., Barker, S.C., 2006. Extraordinary number of gene

rearrangements in the mitochondrial genomes of lice (Phthiraptera: Insecta). Insect Mol. Biol. 15,

63-68.

Crease, T.J., 1999. The complete sequence of the mitochondrial genome of Daphnia pulex

(Cladocera: Crustacea). Gene 233, 89-99.

Crozier, R.H., Crozier, Y.C., 1993. The Mitochondrial Genome of the Honeybee Apis melliferu:

Complete Sequence and Genome Organization. Genetics 133, 97-117.

D'Aurelio, M., Gajewski, C.D., Lin, M.T., Mauck, W.M., Shao, L.Z., Lenaz, G., Moraes, C.T.,

Manfredi, G., 2004. Heterologous mitochondrial DNA recombination in human cells. Hum. Mol.

Genet. 13, 3171-3179.

de Melo Rodovalho, C., Lyra, M.L., Ferro, M., Bacci Jr, M., 2014. The mitochondrial genome of the

leaf-cutter ant Atta laevigata: a mitogenome with a large number of intergenic spacers. PLoS One 9,

e97117.

Debevec, A.H., Cardinal, S., Danforth, B.N., 2012. Identifying the sister group to the bees: a

molecular phylogeny of Aculeata with an emphasis on the superfamily Apoidea. Zool. Scr. 41,

126
527-535.

Diaz, F., Bayona-Bafaluy, M.P., Rana, M., Mora, M., Hao, H., Moraes, C.T., 2002. Human

mitochondrial DNA with large deletions repopulates organelles faster than full-length genomes

under relaxed copy number control. Nucleic Acids Res. 30, 4626-4633.

Dotson, E.M., Beard, C.B., 2001. Sequence and organization of the mitochondrial genome of the

Chagas disease vector, Triatoma dimidiata. Insect Mol. Biol. 10, 205-215.

Dowton, M., 1999. Relationships among the cyclostome braconid (Hymenoptera: Braconidae)

subfamilies inferred from a mitochondrial tRNA gene rearrangement. Mol. Phylogenet. Evol. 11,

283-287.

Dowton, M., Austin, A.D., 1994. Molecular phylogeny of the insect order Hymenoptera: apocritan

relationships. Proc. Natl. Acad. Sci. USA 91, 9911-9915.

Dowton, M., Austin, A.D., 1999. Evolutionary dynamics of a mitochondrial rearrangement " hot

spot" in the Hymenoptera. Mol. Biol. Evol. 16, 298-309.

Dowton, M., Austin, A.D., 2001. Simultaneous analysis of 16S, 28S, COI and morphology in the

Hymenoptera: Apocrita – evolutionary transitions among parasitic wasps. Biol. J. Linn. Soc. 74,

87-111.

Dowton, M., Austin, A.D., Dillon, N., Bartowsky, E., 1997. Molecular phylogeny of the apocritan

wasps: the Proctotrupomorpha and Evaniomorpha. Syst. Entomol. 22, 245-255.

Dowton, M., Cameron, S.L., Austin, A.D., Whiting, M.F., 2009a. Phylogenetic approaches for the

analysis of mitochondrial genome sequence datain the Hymenoptera – A lineage with both rapidly

127
and slowly evolving mitochondrial genomes. Mol. Phylogenet. Evol. 52, 512-519.

Dowton, M., Cameron, S.L., Dowavic, J.I., Austin, A.D., Whiting, M.F., 2009b. Characterization of

67 mitochondrial tRNA gene rearrangements in the Hymenoptera suggests that mitochondrial tRNA

gene position is selectively neutral. Mol. Biol. Evol. 26, 1607-1617.

Dowton, M., Campbell, N.J.H., 2001. Intramitochondrial recombination – is it why some

mitochondrial genes sleep around? Trends Ecol. Evol. 16, 269-271.

Dowton, M., Castro, L.R., Austin, A.D., 2002. Mitochondrial gene rearrangements as phylogenetic

characters in the invertebrates: the examination of genome 'morphology'. Invertebr. Syst. 16,

345-356.

Dowton, M., Castro, L.R., Campbell, S.L., Bargon, S.D., Austin, A.D., 2003. Frequent mitochondrial

gene rearrangements at the Hymenopteran nad3–nad5 junction. J. Mol. Evol. 56, 517-526.

Eberhard, J.R., Wright, T.F., Bermingham, E., 2001. Duplication and concerted evolution of the

mitochondrial control region in the parrot genus Amazona. Mol. Biol. Evol. 18, 1330-1342.

Edgar, R.C., 2004. MUSCLE: multiple sequence alignment with high accuracy and high throughput.

Nucleic Acids Res. 32, 1792-1797.

Fay, J.C., Wu, C.I., 2003. Sequence divergence, functional constraint, and selection in protein

evolution. Annu. Rev. Genomics Hum. 4, 213-235.

Fenn, J.D., Song, H., Cameron, S.L., Whiting, M.F., 2008. A preliminary mitochondrial genome

phylogeny of Orthoptera (Insecta) and approaches to maximizing phylogenetic signal found within

mitochondrial genome data. Mol. Phylogenet. Evol. 49, 59-68.

128
Flook, P., Rowell, H., Gellissen, G., 1995. Homoplastic rearrangements of insect mitochondrial

tRNA genes. Naturwissenschaften 82, 336-337.

Foster, P.G., Jermiin, L.S., Hickey, D.A., 1997. Nucleotide composition bias affects amino acid

content in proteins coded by animal mitochondria. J. Mol. Evol. 44, 282-288.

Friedrich, M., Muqim, N., 2003. Sequence and phylogenetic analysis of the complete mitochondrial

genome of the flour beetle Tribolium castanaeum. Mol. Phylogenet. Evol. 26, 502-512.

Galloway, I.D., Austin, A.D., 1984. Revision of the Scelioninae (Hymenoptera: Scelionidae) in

Australia. Aust. J. Zool. 32, 1-138.

Galtier, N., Nabholz, B., Glemin, S., Hurst, G., 2009. Mitochondrial DNA as a marker of molecular

diversity: a reappraisal. Mol. Ecol. 18, 4541-4550.

Gantenbein, B., Fet, V., Gantenbein-Ritter, I.A., Balloux, F., 2005. Evidence for recombination in

scorpion mitochondrial DNA (Scorpiones: Buthidae). Proc. Biol. Sci. 272, 697-704.

Gao, Y., Katyal, S., Lee, Y., Zhao, J., Rehg, J.E., Russell, H.R., 2011. DNA ligase III is critical for

mtDNA integrity but not Xrcc1-mediated nuclear DNA repair. Nature 471, 240-244.

Garey, J.R., Wolstenholme, D.R., 1989. Platyhelminth mitochondrial DNA: evidence for early

evolutionary origin of a tRNA(serAGN) that contains a dihydrouridine arm replacement loop, and of

serine-specifying AGA and AGG codons. J. Mol. Evol. 28, 374-387.

Gibson, A., Gowri-Shankar, V., Higgs, P.G., Rattray, M., 2005. A comprehensive analysis of

mammalian mitochondrial genome base composition and improved phylogenetic methods. Mol. Biol.

Evol. 22, 251-264.

129
Gibson, G.A.P., 1999. Sister-group relationships of the Platygastroidea and Chalcidoidea

(Hymenoptera) — an alternate hypothesis to Rasnitsyn (1988). Zool. Scr. 28, 125-138.

Gibson, T., Blok, V.C., Phillips, M.S., Hong, G., Kumarasinghe, D., Riley, I.T., Dowton, M., 2007.

The mitochondrial subgenomes of the nematode Globodera pallida are mosaics: evidence of

recombination in an animal mitochondrial genome. J. Mol. Evol. 64, 463-471.

Gillett, C.P.D.T., Crampton-Platt, A., Timmermans, M.J.T.N., Jordal, B.H., Emerson, B.C., Vogler,

A.P., 2014. Bulk de novo mitogenome assembly from pooled total DNA elucidates the phylogeny of

weevils (coleoptera: curculionoidea). Mol. Biol. Evol. 31, 2223-2237.

Gissi, C., Iannelli, F., Pesole, G., 2008. Evolution of the mitochondrial genome of Metazoa as

exemplified by comparison of congeneric species. Heredity 101, 301-320.

Golubchik, T., Wise, M.J., Easteal, S., Jermiin, L.S., 2007. Mind the gaps: evidence of bias in

estimates of multiple sequence alignments. Mol. Biol. Evol. 24, 2433-2442.

Gotzek, D., Clarke, J., Shoemaker, D., 2010. Mitochondrial genome evolution in fire ants

(Hymenoptera: Formicidae). BMC Evol. Biol. 10, 300.

Grande, C., Templado, J., Zardoya, R., 2008. Evolution of gastropod mitochondrial genome

arrangements. BMC Evol. Biol. 8, 61.

Hahn, C., Bachmann, L., Chevreux, B., 2013. Reconstructing mitochondrial genomes directly from

genomic next-generation sequencing reads—a baiting and iterative mapping approach. Nucleic

Acids Res., e129.

Hall, T.A., 1999. BioEdit: a user-friendly biological sequence alignment editor and analysis program

130
for Windows 95/98/NT. pp. 95-98.

Harrison, G.L.A., McLenachan, P.A., Phillips, M.J., Slack, K.E., Cooper, A., Penny, D., 2004. Four

new avian mitochondrial genomes help get to basic evolutionary questions in the late cretaceous.

Mol. Biol. Evol. 21, 974-983.

Hasegawa, E., Kobayashi, K., Yagi, N., Tsuji, K., 2011. Complete mitochondrial genomes of normal

and cheater morphs in the parthenogenetic ant Pristomyrmex punctatus (Hymenoptera: Formicidae).

Myrmecological News 15, 85-90.

Hassanin, A., Léger, N., Deutsch, J., Collins, T., 2005. Evidence for multiple reversals of asymmetric

mutational constraints during the evolution of the mitochondrial genome of metazoa, and

consequences for phylogenetic inferences. Syst. Biol. 54, 277-298.

Heraty, J., Ronquist, F., Carpenter, J.M., Hawks, D., Schulmeister, S., Dowling, A.P., Murray, D.,

Munro, J., Wheeler, W.C., Schiff, N., Sharkey, M., 2011. Evolution of the hymenopteran

megaradiation. Mol. Phylogenet. Evol. 60, 73-88.

Hilker, R., Sickinger, C., Pedersen, C.N., Stoye, J., 2012. UniMoG—a unifying framework for

genomic distance calculation and sorting based on DCJ. Bioinformatics 28, 2509-2511.

Hillis, D.M., Huelsenbeck, J.P., Cunningham, C.W., 1994. Application and accuracy of molecular

phylogenies. Science 264, 671-677.

Hoarau, G., Holla, S., Lescasse, R., Stam, W.T., Olsen, J.L., 2002. Heteroplasmy and evidence for

recombination in the mitochondrial control region of the flatfish Platichthys flesus. Mol. Biol. Evol.

19, 2261-2264.

131
Holt, I.J., Harding, A.E., Morgan-Hughes, J.A., 1988. Deletions of muscle mitochondrial DNA in

patients with mitochondrial myopathies. Nature 331, 717-719.

Huber, J.T., 2009. Biodiversity of Hymenoptera. In: Foottit RG and Adler PH, editors. Insect

biodiversity: science and society Wiley-Blackwell, Oxford, UK. p. 303-323.

Huelsenbeck, J.P., Ronquist, F., 2001. MRBAYES: Bayesian inference of phylogenetic trees.

Bioinformatics (Oxford, England) 17, 754-755.

Inoue, J.G., Miya, M., Tsukamoto, K., Nishida, M., 2001. Complete mitochondrial DNA sequence of

Conger myriaster (Teleostei: Anguilliformes): novel gene order for vertebrate mitochondrial

genomes and the phylogenetic implications for anguilliform families. J. Mol. Evol. 52, 311-320.

Iqbal, M., Austin, A.D., 2000. A preliminary phylogeny for the Baeini (Hymenoptera: Scelionidae):

endoparasitoids of spider eggs. Hymenoptera: evolution, biodiversity and biological control. CSIRO

Publishing, Collingwood, VIC, Australia, 178-191.

Jeon, H., Lee, K., Kim, K., Hwang, U., Eom, K., 2005. Complete sequence and structure of the

mitochondrial genome of the human tapeworm, Taenia asiatica (Platyhelminthes; Cestoda).

Parasitology 130, 717-726.

Johnson, N.F., 1988. Midcoxal articulations and the phylogeny of the order Hymenoptera. Ann.

Entomol. Soc. Am. 81, 870-881.

Johnson, N.F., 2013. Fauna Europaea: Scelionidae. In: Noyes, J. (ed.). Fauna Europaea:

Hymenoptera. Fauna Europaea Version 2.6. http://www.faunaeur.org (accessed 9 April 2013).

Jorde, L.B., Bamshad, M., 2000. Questioning evidence for recombination in human mitochondrial

132
DNA. Science 288, 1931.

Kajander, O.A., Rovio, A.T., Majamaa, K., Poulton, J., Spelbrink, J.N., Holt, I.J., Karhunen, P.J.,

Jacobs, H.T., 2000. Human mtDNA sublimons resemble rearranged mitochondrial genoms found in

pathological states. Hum. Mol. Genet. 9, 2821-2835.

Kaltenpoth, M., Corneli, P.S., Dunn, D.M., Weiss, R.B., Strohm, E., Seger, J., 2012. Accelerated

evolution of mitochondrial but not nuclear genomes of Hymenoptera: new evidence from crabronid

wasps. PLoS One 7, e32826.

Katoh, K., Kuma, K.-i., Toh, H., Miyata, T., 2005. MAFFT version 5: improvement in accuracy of

multiple sequence alignment. Nucleic Acids Res. 33, 511-518.

Kilpert, F., Held, C., Podsiadlowski, L., 2012. Multiple rearrangements in mitochondrial genomes of

Isopoda and phylogenetic implications. Mol. Phylogenet. Evol. 64, 106-117.

Kim, K.G., Hong, M.Y., Kim, M.J., Im, H.H., Kim, M.I., Bae, C.H., Seo, S.J., Lee, S.H., Kim, I.,

2009. Complete mitochondrial genome sequence of the yellow-spotted long-horned beetle

Psacothea hilaris (Coleoptera: Cerambycidae) and phylogenetic analysis among coleopteran insects.

Mol. Cells 27, 429-441.

Kivisild, T., Villems, R., 2000. Questioning evidence for recombination in human mitochondrial

DNA. Science 288, 1931.

Klopfstein, S., Vilhelmsen, L., Heraty, J.M., Sharkey, M., Ronquist, F., 2013. The hymenopteran tree

of life: evidence from protein-coding genes and objectively aligned ribosomal data. PLoS One 8,

e69344.

133
Kraytsberg, Y., Schwartz, M., Brown, T.A., Ebralidse, K., Kunz, W.S., Clayton, D.A., Vissing, J.,

Khrapko, K., 2004. Recombination of human mitochondrial DNA. Science 304, 981.

Krishnan, K.J., Turnbull, D.M., Reeve, A.K., Samuels, D.C., Chinnery, P.F., Blackwood, J.K., Taylor,

R.W., Wanrooij, S., Spelbrink, J.N., Lightowlers, R.N., 2008. What causes mitochondrial DNA

deletions in human cells? Nat. Genet. 40, 275-279.

Kumar, S., Hedrick, P., Dowling, T., Stoneking, M., 2000. Questioning evidence for recombination

in human mitochondrial DNA. Science 288, 1931.

Kumazawa, Y., Ota, H., Nishida, M., Ozawa, T., 1996. Gene rearrangements in snake mitochondrial

genomes: highly concerted evolution of control-region-like sequences duplicated and inserted into a

tRNA gene cluster. Mol. Biol. Evol. 13, 1242-1254.

Kurabayashi, A., Sumida, M., Yonekawa, H., Glaw, F., Vences, M., Hasegawa, M., 2008. Phylogeny,

recombination, and mechanisms of stepwise mitochondrial genome reorganization in mantellid frogs

from Madagascar. Mol. Biol. Evol. 25, 874-891.

Ladoukakis, E.D., Zouros, E., 2001. Direct evidence for homologous recombination in mussel

(Mytilus galloprovincialis) mitochondrial DNA. Mol. Biol. Evol. 18, 1168-1175.

Lanfear, R., Calcott, B., Ho, S.Y.W., Guindon, S., 2012. PartitionFinder: combined selection of

partitioning schemes and substitution models for phylogenetic analyses. Mol. Biol. Evol. 29,

1695-1701.

Larget, B., Simon, D.L., Kadane, J.B., 2002. Bayesian phylogenetic inference from animal

mitochondrial genome arrangements. J. Roy. Stat. Soc. B 64, 681-693.

134
LaSalle, J., Gauld, I.D., 1993. Hymenoptera: their diversity and their impact on the diversity of other

organisms. In LaSalle, J. and Gauld, I.D. editors. Hymenoptera and Biodiversity. C. A. B.

International, Wallingford., 1-26.

Laslett, D., Canbäck, B., 2008. ARWEN: a program to detect tRNA genes in metazoan

mitochondrial nucleotide sequences. Bioinformatics 24, 172-175.

Lavrov, D.V., Boore, J.L., Brown, W.M., 2002. Complete mtDNA sequences of two millipedes

suggest a new model for mitochondrial gene rearrangements: duplication and nonrandom loss. Mol.

Biol. Evol. 19, 163-169.

Lavrov D.V., Brown W.M., Boore J.L., 2000. A novel type of RNA editing occurs in the

mitochondrial tRNAs of the centipede Lithobius forficatus. Proc. Natl. Acad. Sci. U S A.

97,13738–13742.

Lee, E.S., Shin, K.S., Kim, M.S., Park, H., Cho, S., Kim, C.B., 2006. The mitochondrial genome of

the smaller tea tortrix Adoxophyes honmai (Lepidoptera: Tortricidae). Gene 373, 52-57.

Lemmon, E.M., Lemmon, A.R., 2013. High-throughput genomic data in systematics and

phylogenetics. Annu. Rev. Ecol. Evol. Syst., 44, 99-121.

Li, H., Gao, J., Liu, H., Liu, H., Liang, A., Zhou, X., Cai, W., 2011. The architecture and complete

sequence of mitochondrial genome of an assassin bug Agriosphodrus dohrni (Hemiptera:

Reduviidae). Int. J. Biol. Sci. 7, 792-804.

Li, H., Liu, H., Shi, A., Stys, P., Zhou, X., Cai, W., 2012. The complete mitochondrial genome and

novel gene arrangement of the unique-headed bug Stenopirates sp. (Hemiptera: Enicocephalidae).

135
PLoS One 7, e29419.

Lin, C.P., Danforth, B.N., 2004. How do insect nuclear and mitochondrial gene substitution patterns

differ? Insights from Bayesian analyses of combined datasets. Mol. Phylogenet. Evol. 30, 686-702.

Lowe, T.M., Eddy, S.R., 1997. tRNAscan-SE: a program for improved detection of transfer RNA

genes in genomic sequence. Nucleic Acids Res. 25, 955-964.

Lunt, D.H., Hyman, B.C., 1997. Animal mitochondrial DNA recombination. Nature 387, 247.

Mao, M., Austin, A.D., Johnson, N.F., Dowton, M., 2014a. Coexistence of minicircular and a highly

rearranged mtDNA molecule suggests that recombination shapes mitochondrial genome organization.

Mol. Biol. Evol. 31, 636-644.

Mao, M., Dowton, M., 2014. Complete mitochondrial genomes of Ceratobaeus sp. and Idris sp.

(Hymenoptera: Scelionidae): shared gene rearrangements as potential phylogenetic markers at the

tribal level. Mol. Biol. Rep. doi: 10.1007/s11033-014-3522-x.

Mao, M., Gibson, T., Dowton, M., 2014b. Evolutionary dynamics of the mitochondrial genome in

the Evaniomorpha (Hymenoptera) — a group with an intermediate rate of gene rearrangement.

Genome Biol. Evol. 6, 1862-1874.

Mao, M., Valerio, A., Austin, A.D., Dowton, M., Johnson, N.F., 2012. The first mitochondrial

genome for the wasp superfamily Platygastroidea: the egg parasitoid Trissolcus basalis. Genome 55,

194-204.

Martin, M., Cho, J., Cesare, A.J., Griffith, J.D., Attardi, G., 2005. Termination factor-mediated DNA

loop between termination and initiation sites drives mitochondrial rRNA synthesis. Cell 123,

136
1227–1240.

Masner, L., 1976. Revisionary notes and keys to world genera of Scelionidae (Hymenoptera:

Proctotrupoidea). Mem. Entomol. Soc. Can. 108, 1-87.

Masta, S.E., Boore, J.L., 2004. The complete mitochondrial genome sequence of the spider

Habronattus oregonensis reveals rearranged and extremely truncated tRNAs. Mol. Biol. Evol. 21,

893–902.

Matsumoto, Y., Yanase, T., Tsuda, T., Noda, H., 2009. Species-specific mitochondrial gene

rearrangements in biting midges and vector species identification. Med. Vet. Entomol. 23, 47-55.

Miller, M.A., Miller, M.A., Pfeiffer, W., Pfeiffer, W., Schwartz, T., Schwartz, T., 2010. Creating the

CIPRES Science Gateway for inference of large phylogenetic trees. Proceedings of the Gateway

Computing Environments Workshop (GCE). New Orleans, LA., pp. 1-8.

Mita, S., Rizzuto, R., Moraes, C.T., Shanske, S., Arnaudo, E., Fabrizi, G.M., Koga, Y., DiMauro, S.,

Schon, E.A., 1990. Recombination via flanking direct repeats is a major cause of large-scale

deletions of human mitochondrial DNA. Nucleic Acids Res. 18, 561-567.

Mitchell, S.E., Cockburn, A.F., Seawright, J.A., 1993. The mitochondrial genome of Anopheles

quadrimaculatus species A: complete nucleotide sequence and gene organization. Genome 36,

1058-1073.

Miya, M., Takeshima, H., Endo, H., Ishiguro, N.B., Inoue, J.G., Mukai, T., Satoh, T.P., Yamaguchi,

M., Kawaguchi, A., Mabuchi, K., Shirai, S.M., Nishida, M., 2003. Major patterns of higher

teleostean phylogenies: a new perspective based on 100 complete mitochondrial DNA sequences.

137
Mol. Phylogenet. Evol. 26, 121-138.

Morel, F., Renoux, M., Lachaurne, P., Alziari, S., 2008. Bleomycin-induced double-strand breaks in

mitochondrial DNA of Drosophila cells are repaired. Mutat. Res. 637, 111-117.

Moritz, C., Dowling, T.E., Brown, W.M., 1987. Evolution of animal mitochondrial DNA: relevance

for population biology and systematics. Annu. Rev. Ecol. Syst. 18, 269-292.

Mueller, R.L., Boore, J.L., 2005. Molecular mechanisms of extensive mitochondrial gene

rearrangement in plethodontid salamanders. Mol. Biol. Evol. 22, 2104-2112.

Murphy, N., Carey, D., Castro, L., Dowton, M., Austin, A., 2007. Phylogeny of the platygastroid

wasps (Hymenoptera) based on sequences from the 18S rRNA, 28S rRNA and cytochrome oxidase I

genes: implications for the evolution of the ovipositor system and host relationships. Biol. J. Linnean

91, 653-669.

Mwinyi, A., Meyer, A., Bleidorn, C., Lieb, B., Bartolomaeus, T., Podsiadlowski, L., 2009.

Mitochondrial genome sequence and gene order of Sipunculus nudus give additional support for an

inclusion of Sipuncula into Annelida. BMC Genomics 10, 27.

Nei, M., Kumar, S., 2000. Molecular evolution and phylogenetics. Oxford University Press, Oxford.

Nelson, L.A., Lambkin, C.L., Batterham, P., Wallman, J.F., Dowton, M., Whiting, M.F., Yeates, D.K.,

Cameron, S.L., 2012. Beyond barcoding: A mitochondrial genomics approach to molecular

phylogenetics and diagnostics of blowflies (Diptera: Calliphoridae). Gene 511, 131-142.

New, T.R., 2012. Hymenoptera and Conservation. Wiley-Blackwell,Oxford, pp. 1-230.

Nylander, J., A,A,, Ronquist, F., Huelsenbeck, J.P., Nieves-Aldrey, J.L., 2004. Bayesian phylogenetic

138
analysis of combined data. Syst. Biol. 53, 47-67.

Ojala, D., Montoya, J., Attardi, G., 1981. tRNA punctuation model of RNA processing in human

mitochondria. Nature 290, 470-474.

Okonechnikov, K., Syabro, M., Tleukenov, T., Golosova, O., Fursov, M., Varlamov, A., Vaskin, Y.,

Efremov, I., German Grehov, O.G., Kandrov, D., Rasputin, K., Team, U., 2012. Unipro UGENE: A

unified bioinformatics toolkit. Bioinformatics 28, 1166-1167.

Oliveira, D.C.S.G., Raychoudhury, R., Lavrov, D.V., Werren, J.H., 2008. Rapidly evolving

mitochondrial genome and directional selection in mitochondrial genes in the parasitic wasp Nasonia

(Hymenoptera: Pteromalidae). Mol. Biol. Evol. 25, 2167-2180.

Parsons, T.J., Irwin, J.A., 2000. Questioning evidence for recombination in human mitochondrial

DNA. Science 288, 1931.

Perna, N.T., Kocher, T.D., 1995. Patterns of nucleotide composition at fourfold degenerate sites of

animal mitochondrial genomes. J. Mol. Evol. 41, 353-358.

Perseke, M., Golombek, A., Schlegel, M., Struck, T.H., 2013. The impact of mitochondrial genome

analyses on the understanding of deuterostome phylogeny. Mol. Phylogenet. Evol. 66, 898-905.

Peters, R.S., Meyer, B., Krogmann, L., Borner, J., Meusemann, K., Schütte, K., Niehuis, O., Misof,

B., 2011. The taming of an impossible child: a standardized all-in approach to the phylogeny of

Hymenoptera using public database sequences. BMC Biol. 9, 55.

Pilgrim, E.M., Von Dohlen, C.D., Pitts, J.P., 2008. Molecular phylogenetics of Vespoidea indicate

paraphyly of the superfamily and novel relationships of its component families and subfamilies. Zool.

139
Scr. 37, 539-560.

Rambaut, A., Drummond, A.J., 2009. Tracer version 1.5. http://treebioedacuk.

Rasnitsyn, A.P., 1988. An outline of the evolution of hymenopterous insects (order Vespida). Orient.

Insects 22, 115-145.

Rasnitsyn, A.P., 2010. Molecular phylogenetics, morphological cladistics, and fossil record. Entomol.

Rev. 90, 263-298.

Rasnitsyn, A.P., Zhang, H.C., 2010. Early evolution of Apocrita (Insecta, Hymenoptera) as indicated

by new findings in the middle Jurassic of Daohugou, Northeast China. Acta. Geol. Sin-Engl. 84,

834-873.

Reyes, A., Gissi, C., Pesole, G., Saccone, C., 1998. Asymmetrical directional mutation pressure in

the mitochondrial genome of mammals. Mol. Biol. Evol. 15, 957 - 966.

Rokas, A., Ladoukakis, E., Zouros, E., 2003. Animal mitochondrial DNA recombination revisited.

Trends Ecol. Evol. 18, 411-417.

Ronquist, F., Huelsenbeck, J.P., 2003. MrBayes 3: Bayesian phylogenetic inference under mixed

models. Bioinformatics 19, 1572-1574.

Ronquist, F., Huelsenbeck, J.P., Teslenko, M., van der Mark, P., Ayres, D.L., Darling, A., Höhna, S.,

Larget, B., Liu, L., Suchard, M.A., Stockholms, u., Matematiska, i., Naturvetenskapliga, f., 2012a.

MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model

space. Syst. Biol. 61, 539-542.

Ronquist, F., Klopfstein, S., Vilhelmsen, L., Schulmeister, S., Murray, D.L., Rasnitsyn, A.P., 2012b.

140
A total-evidence approach to dating with fossils, applied to the early radiation of the hymenoptera.

Syst. Biol. 61, 973-999.

Ronquist, F., Rasnitsyn, A.P., Roy, A., Eriksson, K., Lindgren, M., 1999. Phylogeny of the

Hymenoptera: A cladistic reanalysis of Rasnitsyn’s (1988) data. Zool. Scr. 28, 13-50.

Roques, S., Fox, C.J., Villasana, M.I., Rico, C., 2006. The complete mitochondrial genome of the

whiting, Merlangius merlangus and the haddock, Melanogrammus aeglefinus: A detailed genomic

comparison among closely related species of the Gadidae family. Gene 383, 12-23.

Saccone, C., De Giorgi, C., Gissi, C., Pesole, G., Reyes, A., 1999. Evolutionary genomics in Metazoa:

the mitochondrial DNA as a model system. Gene 238, 195-209.

Samuels, D.C., Schon, E.A., Chinnery, P.F., 2004. Two direct repeats cause most human mtDNA

deletions. Trends Genet. 20, 393-398.

San Mauro, D., Gower, D.J., Zardoya, R., Wilkinson, M., 2006. A hotspot of gene order

rearrangement by tandem duplication and random loss in the vertebrate mitochondrial genome. Mol.

Biol. Evol. 23, 227-234.

Satoh, M., Kuroiwa, T., 1991. Organization of multiple nucleoids and DNA molecules in

mitochondria of a human cell. Exp. Cell Res. 196, 137-140.

Savard, J., Tautz, D., Richards, S., Weinstock, G.M., Gibbs, R.A., Werren, J.H., Tettelin, H., Lercher,

M.J., 2006. Phylogenomic analysis reveals bees and wasps (Hymenoptera) at the base of the

radiation of Holometabolous insects. Genome Res. 16, 1334-1338.

Schulmeister, S., 2003a. Review of morphological evidence on the phylogeny of basal Hymenoptera

141
(Insecta), with a discussion of the ordering of characters. Biol. J. Linn. Soc. 79, 209-243.

Schulmeister, S., 2003b. Simultaneous analysis of basal Hymenoptera (Insecta): introducing robust

‐choice sensitivity analysis. Biol. J. Linn. Soc. 79, 245-275.

Segovia, R., Pett, W., Trewick, S., Lavrov, D.V., 2011. Extensive and evolutionarily persistent

mitochondrial tRNA editing in velvet worms (phylum Onychophora). Mol. Biol. Evol. 28,

2873-2881.

Shao, R., Barker, S.C., 2003. The highly rearranged mitochondrial genome of the plague thrips,

Thrips imaginis (Insecta: Thysanoptera): convergence of two novel gene boundaries and an

extraordinary arrangement of rRNA genes. Mol. Biol. Evol. 20, 362-370.

Shao, R., Barker, S.C., Mitani, H., Aoki, Y., Fukunaga, M., 2005. Evolution of duplicate control

regions in the mitochondrial genomes of metazoa: a case study with Australasian Ixodes ticks. Mol.

Biol. Evol. 22, 620-629.

Shao, R., Campbell, N.J., Barker, S.C., 2001. Numerous gene rearrangements in the mitochondrial

genome of the wallaby louse, Heterodoxus macropus (Phthiraptera). Mol. Biol. Evol. 18, 858-865.

Shao, R., Dowton, M., Murrell, A., Barker, S.C., 2003. Rates of gene rearrangement and nucleotide

substitution are correlated in the mitochondrial genomes of insects. Mol. Biol. Evol. 20, 1612-1619.

Shao, R., Kirkness, E.F., Barker, S.C., 2009. The single mitochondrial chromosome typical of

animals has evolved into 18 minichromosomes in the human body louse, Pediculus humanus.

Genome Res. 19, 904-912.

Sharanowski, B.J., Robbertse, B., Walker, J., Voss, S.R., Yoder, R., Spatafora, J., Sharkey, M.J., 2010.

142
Expressed sequence tags reveal Proctotrupomorpha (minus Chalcidoidea) as sister to Aculeata

(Hymenoptera: Insecta). Mol. Phylogenet. Evol. 57, 101-112.

Sharkey, M.J., 2007. Phylogeny and classification of Hymenoptera. Zootaxa, 521-548.

Sharkey, M.J., Carpenter, J.M., Vilhelmsen, L., Heraty, J., Liljeblad, J., Dowling, A.P.G.,

Schulmeister, S., Murray, D., Deans, A.R., Ronquist, F., 2011. Phylogenetic relationships among

superfamilies of Hymenoptera. Cladistics 27, 1-33.

Sharkey, M.J., Roy, A., 2002. Phylogeny of the Hymenoptera: a reanalysis of the Ronquist et al.

(1999) reanalysis, emphasizing wing venation and apocritan relationships. Zool. Scr. 31, 57-66.

Sheffield, N.C., Song, H., Cameron, L., Whiting, M.F., 2008. A comparative analysis of

mitochondrial genomes in Coleoptera (Arthropoda: Insecta) and genome descriptions of six new

beetles. Mol. Biol. Evol. 25, 2499-2509.

Silvestre, D., Dowton, M., Arias, M.C., 2008. The mitochondrial genome of the stingless bee

Melipona bicolor (Hymenoptera, Apidae, Meliponini): Sequence, gene organization and a unique

tRNA translocation event conserved across the tribe Meliponini. Genet. Mol. Biol. 31, 451-460.

Simon, C., Buckley, T.R., Frati, F., Stewart, J.B., Beckenbach, A.T., 2006. Incorporating molecular

evolution into phylogenetic analysis, and a new compilation of conserved polymerase chain reaction

primers for animal mitochondrial DNA. Annu. Rev. Ecol. Evol. Syst. 37, 545-579.

Simon, C., Frati, F., Beckenbach, A., Crespi, B., Liu, H., Flook, P., 1994. Evolution, weighting, and

phylogenetic utility of mitochondrial gene sequences and a compilation of conserved polymerase

chain reaction primers. Ann. Entomol. Soc. Am. 87, 651-701.

143
Slate, J., Gemmell, N.J., 2004. Eve ‘n’ Steve: recombination of human mitochondrial DNA. Trends

Ecol. Evol. 19, 561-563.

Srivastava, S., Moraes, C.T., 2005. Double-strand breaks of mouse muscle mtDNA promote large

deletions similar to multiple mtDNA deletions in humans. Hum. Mol. Genet. 14, 893-902.

Stamatakis, A., 2014. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large

phylogenies. Bioinformatics 30, 1312-1313.

Stamatakis, A., Hoover, P., Rougemont, J., 2008. A rapid bootstrap algorithm for the RAxML web

servers. Syst. Biol. 57, 758-771.

Sullivan, J., Joyce, P., 2005. Model selection in phylogenetics. Annu. Rev. Ecol. Evol. Syst. 36,

445-466.

Tamura, K., Peterson, D., Peterson, N., Stecher, G., Nei, M., Kumar, S., 2011. MEGA5: molecular

evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum

parsimony methods. Mol. Biol. Evol., 1-18.

Thompson, J.D., Higgins, D.G., Gibson, T.J., 1994. CLUSTAL W: improving the sensitivity of

progressive multiple sequence alignment through sequence weighting, position-specific gap

penalties and weight matrix choice. Nucleic Acids Res. 22, 4673-4680.

Thyagarajan, B., Padua, R.A., Campbell, C., 1996. Mammalian mitochondria possess homologous

DNA recombination activity. J. Biol. Chem. 271, 27536-27543.

Timmermans, M.J.T.N., Lees, D.C., Simonsen, T.J., 2014. Towards a mitogenomic phylogeny of

Lepidoptera. Mol. Phylogenet. Evol. 79, 169-178.

144
Torres, T.T., Dolezal, M., Schlotterer, C., Ottenwalder, B., 2009. Expression profiling of Drosophila

mitochondrial genes via deep mRNA sequencing. Nucleic Acids Res. 37, 7509–7518.

Tsaousis, A.D., Martin, D.P., Ladoukakis, E.D., Posada, D., Zouros, E., 2005. Widespread

recombination in published animal mtDNA sequences. Mol. Biol. Evol. 22, 925-933.

Ujvari, B., Dowton, M., Madsen, T., 2007. Mitochondrial DNA recombination in a free-ranging

Australian lizard. Biol. Lett. 3, 189-192.

Vilhelmsen, L., 2001. Phylogeny and classification of the extant basal lineages of the Hymenoptera

(Insecta). Zool. J. Linn. Soc. 131, 393-442.

Vilhelmsen, L., 2011. Head capsule characters in the Hymenoptera and their phylogenetic

implications. ZooKeys 130, 343-361.

Vilhelmsen, L., Miko, I., Krogmann, L., 2010. Beyond the wasp-waist: structural diversity and

phylogenetic significance of the mesosoma in apocritan wasps (Insecta: Hymenoptera). Zool. J. Linn.

Soc. 159, 22-194.

Wei, S.J., Li, Q., van Achterberg, K., Chen, X.X., 2014. Two mitochondrial genomes from the

families Bethylidae and Mutillidae: Independent rearrangement of protein-coding genes and

higher-level phylogeny of the Hymenoptera. Mol. Phylogenet. Evol. 77, 1-10.

Wei, S.J., Shi, M., He, J.H., Sharkey, M., Chen, X.X., 2009. The complete mitochondrial genome of

Diadegma semiclausum (Hymenoptera: ichneumonidae) indicates extensive independent

evolutionary events. Genome 52, 308-319.

Wei, S.J., Shi, M., Sharkey, M.J., van Achterberg, C., Chen, X.X., 2010a. Comparative

145
mitogenomics of Braconidae (Insecta: Hymenoptera) and the phylogenetic utility of mitochondrial

genomes with special reference to Holometabolous insects. BMC Genomics 11, 371.

Wei, S.J., Tang, P., Zheng, L.H., Shi, M., Chen, X.X., 2010b. The complete mitochondrial genome of

Evania appendigaster (Hymenoptera: Evaniidae) has low A+T content and a long intergenic spacer

between atp8 and atp6. Mol. Biol. Rep. 37, 1931-1942.

Wei, S.J., Wu, Q.L., van Achterberg, K., Chen, X.X., 2013. Rearrangement of the nad1 gene in

Pristaulacus compressus (Spinola) (Hymenoptera: Evanioidea: Aulacidae) mitochondrial genome.

Mitochondrial DNA. doi: 10.3109/19401736.2013.834436..

Weng, M. L., Blazier, J.C., Govindu, M., Jansen, R.K., 2014. Reconstruction of the ancestral plastid

genome in Geraniaceae reveals a correlation between genome rearrangements, repeats, and

nucleotide substitution rates. Mol. Biol. Evol. 31, 645-659.

Whitfield, J.B., 1992. Phylogeny of the non-aculeate Apocrita and the evolution of parasitism in the

Hymenoptera. J. Hymenoptera Res. 1, 3-14.

Wiegmann, B.M., Heimberg, A.M., Wheeler, B.M., Peterson, K.J., Pape, T., Sinclair, B.J.,

Skevington, J.H., Blagoderov, V., Caravas, J., Kutty, S.N., Schmidt-Ott, U., Trautwein, M.D.,

Kampmeier, G.E., Thompson, F.C., Grimaldi, D.A., Beckenbach, A.T., Courtney, G.W., Friedrich, M.,

Meier, R., Yeates, D.K., Winkler, I.S., Barr, N.B., Kim, J.W., Lambkin, C., Bertone, M.A., Cassel,

B.K., Bayless, K.M., 2011. Episodic radiations in the fly tree of life. Proc. Natl. Acad. Sci. USA 108,

5690-5695.

Wiegmann, B.M., Trautwein, M.D., Kim, J.W., Cassel, B.K., Bertone, M.A., Winterton, S.L., Yeates,

D.K., 2009. Single-copy nuclear genes resolve the phylogeny of the holometabolous insects. BMC
146
Biol. 7, 34.

Wilkinson, M., 1996. Majority-rule reduced consensus trees and their use in bootstrapping. Mol.

Biol. Evol. 13, 437-444.

Xiao, J.H., Jia, J.G., Murphy, R.W., Huang, D.W., 2011. Rapid evolution of the mitochondrial

genome in chalcidoid wasps (Hymenoptera: Chalcidoidea) driven by parasitic lifestyles. PLoS One 6,

e26645.

Yang, X., Xue, D., Han, H., 2013. The complete mitochondrial genome of Biston panterinaria

(Lepidoptera: Geometridae), with phylogenetic utility of mitochondrial genome in the Lepidoptera.

Gene 515, 349-358.

Zhang, D.X., Hewitt, G.M., 1997. Insect mitochondrial control region: A review of its structure,

evolution and usefulness in evolutionary studies. Biochem. Syst. Ecol. 25, 99-120.

Zhang, Z., Li, J., Zhao, X.-Q., Wang, J., Wong, G.K.-S., Yu, J., 2006. KaKs_Calculator: calculating

Ka and Ks through model selection and model averaging. Genom. Proteom. Bioinf. 4, 259-263.

Zhou, Z., Huang, Y., Shi, F., Ye, H., 2009. The complete mitochondrial genome of Deracantha onos

(Orthoptera: Bradyporidae). Mol. Biol. Rep. 36, 7-12.

Zuckerman, S.H., Solus, J.F., Gillespie, F.P., Eisenstadt, J.M., 1984. Retention of both parental

mitochondrial DNA species in mouse-Chinese hamster somatic cell hybrids. Somatic. Cell Mol.

Genet. 10, 85-91.

147

You might also like