ArticlePDF Available

Abstract and Figures

Recent morphological and molecular evidence has changed interpretations of arthropod phylogeny and evolution. Here we compare complete mitochondrial genomes to show that Collembola, a wingless group traditionally considered as basal to all insects, appears instead to constitute a separate evolutionary lineage that branched much earlier than the separation of many crustaceans and insects and independently adapted to life on land. Therefore, the taxon Hexapoda, as commonly defined to include all six-legged arthropods, is not monophyletic.
Content may be subject to copyright.
DOI: 10.1126/science.1078607
, 1887 (2003); 299Science
et al.Francesco Nardi,
Hexapod Origins: Monophyletic or Paraphyletic?
www.sciencemag.org (this information is current as of December 18, 2006 ):
The following resources related to this article are available online at
http://www.sciencemag.org/cgi/content/full/299/5614/1887
version of this article at:
including high-resolution figures, can be found in the onlineUpdated information and services,
http://www.sciencemag.org/cgi/content/full/299/5614/1887/DC1
can be found at: Supporting Online Material
found at:
can berelated to this articleA list of selected additional articles on the Science Web sites
http://www.sciencemag.org/cgi/content/full/299/5614/1887#related-content
http://www.sciencemag.org/cgi/content/full/299/5614/1887#otherarticles
, 5 of which can be accessed for free: cites 12 articlesThis article
82 article(s) on the ISI Web of Science. cited byThis article has been
http://www.sciencemag.org/cgi/content/full/299/5614/1887#otherarticles
8 articles hosted by HighWire Press; see: cited byThis article has been
http://www.sciencemag.org/cgi/collection/evolution
Evolution
: subject collectionsThis article appears in the following
http://www.sciencemag.org/help/about/permissions.dtl
in whole or in part can be found at: this article
permission to reproduce of this article or about obtaining reprintsInformation about obtaining
registered trademark of AAAS.
c 2005 by the American Association for the Advancement of Science; all rights reserved. The title SCIENCE is a
CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
Science (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by the
on December 18, 2006 www.sciencemag.orgDownloaded from
energetics models without acceleration-
based transport predicted the offshore mi-
gration (1, 2), they had limited skill pre-
dicting the total change to the beach over
45 days because they failed to predict on-
shore migration between storms (2). The
energetics model that was extended to in-
clude acceleration better predicted the
change in the sea-floor both onshore and
offshore of the bar crest (Fig. 4), and the
overall evolution of the cross-shore depth
profile (Fig. 5).
References and Notes
1. E. Thornton, R. Humiston, W. Birkemeier, J. Geophys.
Res. 101, 12097 (1996).
2. E. Gallagher, S. Elgar, R. Guza, J. Geophys. Res. 103,
3203 (1998).
3. D. Aubrey, J. Geophys. Res. 84, 6347 (1979).
4. L. Wright, A. Short, Mar. Geol. 56, 93 (1984).
5. J. Roelvink, M. Stive, J. Geophys. Res. 94, 4185
(1989).
6. L. Wright, J. Boon, S. Kim, J. List, Mar. Geol. 96,19
(1991).
7. A. Bowen, in The Coastline of Canada, S. McCann, Ed.
(Geol. Surv. Canada Pap. 10-80, Ottawa, Canada,
1980), pp. 1–11.
8. J. Bailard, J. Geophys. Res. 86, 10938 (1981).
9. J. Trowbridge, D. Young, J. Geophys. Res. 94, 10971
(1989).
10. S. Elgar, E. Gallagher, R. Guza, J. Geophys. Res. 106,
11623 (2001).
11. S. Elgar, R. Guza, M. Freilich, J. Geophys. Res. 93, 9261
(1988).
12. T. G. Drake, J. Calantoni, J. Geophys. Res. 106, 19859
(2001).
13. O. Madsen, in Proc. 14th Int. Conf. Coastal Eng.,
Copenhagen, Denmark (American Society of Civil
Engineers, Reston, VA, 1974), p. 776 –794.
14. P. Nielsen, Coastal Eng. 45, 53 (2002).
15. R. Hallermeier, Cont. Shelf Res. 1, 159 (1982).
16. D. Hanes, D. Huntley, Cont. Shelf Res. 6, 585 (1986).
17. E. Gallagher, S. Elgar, E. Thornton, Nature 394, 165
(1998).
18. R. Bagnold, U.S. Geol. Surv. Prof. Pap. 422-I (1966).
19. Support was provided by the Army Research Office,
the Office of Naval Research, NSF, and a fellowship
from Conselho Nacional de Desenvolvimento
Cientı´fico e Tecnolo´gico (CNPq), Brazil. E. Gal-
lagher, R. Guza, T. Herbers, and B. Raubenheimer
made valuable comments and helped obtain the
field observations. The staff of the Field Research
Facility and the Center for Coastal Studies provid-
ed excellent logistical support during arduous field
conditions.
12 December 2002; accepted 12 February 2003
Hexapod Origins: Monophyletic
or Paraphyletic?
Francesco Nardi,
1
* Giacomo Spinsanti,
1
Jeffrey L. Boore,
2
Antonio Carapelli,
1
Romano Dallai,
1
Francesco Frati
1
Recent morphological and molecular evidence has changed interpretations of ar-
thropod phylogeny and evolution. Here we compare complete mitochondrial ge-
nomes to show that Collembola, a wingless group traditionally considered as basal
to all insects, appears instead to constitute a separate evolutionary lineage that
branched much earlier than the separation of many crustaceans and insects and
independently adapted to life on land. Therefore, the taxon Hexapoda, as com-
monly defined to include all six-legged arthropods, is not monophyletic.
The phylum Arthropoda comprises the major
groups Hexapoda (insects and presumed allies),
Myriapoda (e.g., centipedes and millipedes),
Chelicerata (e.g., spiders and horseshoe crabs),
and Crustacea (e.g., crabs and lobsters). Many
studies have attempted to reconstruct the evolu-
tionary relationships among arthropods using
various approaches such as paleontology (1),
comparative morphology (2), comparative devel-
opmental biology (3, 4), and molecular phyloge-
netics (5, 6).
It has long been held that hexapods (7) con-
stitute a monophyletic taxon (8, 9) and that their
closest relatives are to be found in myriapods
(10). More recently, molecular and developmen-
tal studies have rejected this relationship (35,
11, 12) in favor of a closer affinity between
Hexapoda and Crustacea (Pancrustacea or Tetra-
conata). In this context, special attention must be
given to the apterygotes (springtails, silverfish,
and their allies), the wingless hexapods thought
to branch at the base of Hexapoda. The phyloge-
netic position of these groups is still unclear
(1316), casting doubt even on the monophyly of
the Hexapoda (17).
A potentially powerful technique for resolv-
ing deep relationships is to compare whole mito-
chondrial genomes (5, 17, 18). Phylogenetic
analysis of the only complete mitochondrial
sequence available for an apterygotan species
(17 ) suggested the possibility that Collem-
bola might not be included within Hexapoda,
contrasting with the classic view of a mono-
phyletic taxon that includes all six-legged
arthropods. Collembola have been clustered
within crustaceans in other molecular and/or
combined data sets (15, 16), but the possible
paraphyly of Hexapoda has not been given
specific attention and the deserved consider-
ation. We have now sequenced the complete
mitochondrial genomes of two additional
species (19) specifically chosen to address
this problem: Tricholepidion gertschi, repre-
senting one of the most basal lineages of the
Insecta (Zygentoma), and Gomphiocephalus
hodgsoni, another collembolan, to test sup-
port for the two competing hypotheses of a
monophyletic versus paraphyletic Hexapoda.
An initial phylogenetic analysis performed on
the 35-taxon data set (19) produced the tree
shown in Fig. 1. The tree has high support at most
nodes, with support decreasing toward deeper
relationships. This analysis strongly supports the
Pancrustacea hypothesis, with the exception of
the position of Apis and Heterodoxus. T. gertschi
is basal to all the pterygotan insects, supporting
the monophyly of the Insecta. The four crusta-
cean sequences are divided into two well-defined
groups (representing Malacostraca and Bran-
chiopoda), but their reciprocal relationships and
position relative to the Insecta are not resolved.
The Crustacea Insecta node is well supported,
and it excludes the two collembolans, which
cluster together as the basal lineage of the Pan-
crustacea. A second group unites the Cheli-
cerata Myriapoda [as in (20)] but also includes
the insects Apis and Heterodoxus, presumably as
an artefact.
Although this tree shows many interesting
outcomes, it also contains some evidently unten-
able relationships, which nevertheless have
strong statistical support. This indicates the pres-
ence of anomalies in the evolution of these se-
quences that introduce strong systematic errors
in the analysis. The most likely factors that can
cause these anomalies are unequal base compo-
sition [which can bias amino acid composition
(21)] and uneven rates of evolution among
different lineages. This problem might be
especially acute, because some taxa share an
extremely high AT bias—Apis (84.8%), Rh-
ipicephalus (78.0%), and Heterodoxus
(79.3%)—and different rates of evolution,
1
Department of Evolutionary Biology, University of
Siena, via Aldo Moro 2, 53100 Siena, Italy.
2
U.S.
Department of Energy Joint Genome Institute and
Lawrence Berkeley National Laboratory, 2800 Mitch-
ell Drive, Walnut Creek, CA 94598, USA.
*To whom correspondence should be addressed. E-
mail: nardifra@unisi.it
Fig. 5. Observed and predicted cross-shore bot-
tom elevation profiles spanning a 45-day period.
Sea-floor elevation relative to mean sea level
observed 1 September 1994, 1900 hours (solid
black curve), observed 15 October 1994, 2200
hours (dashed black), and predicted for 15 Octo-
ber 1994, 2200 hours by the energetics (blue) and
energetics plus acceleration (red) models versus
cross-shore position.
R EPORTS
www.sciencemag.org SCIENCE VOL 299 21 MARCH 2003 1887
on December 18, 2006 www.sciencemag.orgDownloaded from
which could potentially cause artefactual at-
traction (22) in this analysis. Such sequenc-
es are usually removed from phylogenetic
analyses owing to their evidently incorrect
placement and disturbance to the reconstruc-
tion. To recognize and exclude from the
analysis those sequences whose placement
in the phylogenetic tree could be influenced
by such anomalies in the mechanism of evo-
lution, rather than by the true historical pro-
cess, we performed a detailed statistical test
(19) to select a subset of sequences with
homogeneous modes of evolution and whose
rate of evolution is compatible with that of
Gomphiocephalus and Tricholepidion. The
placement of these two taxa is key to assess-
ing the monophyly of the Hexapoda, so it is
especially important that the taxa compared
are compatible with them. The methods of
analysis outlined above, applied to this re-
duced data set, produced the two trees
shown in Fig. 2, which differ only in the
placement of Ostrinia with respect to the
remaining Holometabola. Again, strong sup-
port is obtained for the Pancrustacea, with
Tricholepidion basal to the remaining ptery-
gotan insects, and the two collembolans
placed outside the Crustacea Insecta
clade. The trees also show monophyly of
Crustacea, although with a lower level of
support. Limulus is recovered as the sister
group of the Pancrustacea, in contrast with
the analysis based on the 35-taxon data set,
but again with very low support. The result-
ing trees do not seem to be sensitive to the
taxa included (fig. S2).
The most interesting result produced by
this study is certainly the nonmonophyly of
Hexapodathat is the position of the two
collembolans outside the Crustacea Insecta
clade, agreed upon by all analyses and with
high levels of support. To test the relative
positioning of Crustacea, Collembola, and
Insecta in more detail, we compared two
alternative topologies using analytical tests.
The hypothesis of Crustacea external to a
monophyletic Hexapoda (here, Insecta
Collembola) is strongly rejected (Table 1) in
favor of the proposed nonmonophyly of
Hexapoda. We also applied the same tests to
the problem of the basal trichotomy between
Chelicerata, Myriapoda, and Pancrustacea. A
sister group relationship between Pancrusta-
cea and Myriapoda (Mandibulata) is
strongly rejected (Table 1), and no significant
difference in support was found for the other
two possible hypotheses. This accords with
the low levels of support found in all trees at
this node.
It has been generally accepted that the taxon
Hexapoda, including the basal apterygotan or-
ders, is monophyletic. This conclusion is
strongly supported by similarities in their body
organization (composed of head, thorax, and
abdomen), as well as other morphological char-
acters including eye and leg structure and the
absence of limbs in one of the cephalic seg-
ments (9). On the other hand, the interpretation
of such characters also depends on which is the
closest relative of the Hexapoda, and even on
the basal splitting of the latter taxon (9). Nev-
ertheless, apterygotan taxa, including Collem-
bola, show a number of peculiar features that at
least complicate the analysis of their affinities
with the Insecta sensu stricto (9, 23) and leave
some room to question these affinities altogeth-
er. The acceptance of nonmonophyly of
Hexapoda implies that the tripartite and six-
legged body plan typical of Hexapoda would be
a convergent acquisition of collembolans and
the true insects.
Fig. 1. Maximum-likelihood
[ProtML (24)] phylogenetic re-
construction, complete data set.
Numerals at each node show lo-
cal bootstrap probability values.
Branch lengths are drawn pro-
portionally to maximum-likeli-
hood estimates.
Fig. 2. Maximum-likeli-
hood [ProtML (24) and
MrBayes (25)] phyloge-
netic reconstructions, re-
duced data set. Al-
ternative placement of
Ostrinia follows MrBayes
reconstruction. Numerals
above each node show
local bootstrap probabili-
ty values (ProtML), and
numerals below each
node indicate posterior
probabilities (MrBayes).
Branch lengths are pro-
portionate to maximum-
likelihood estimates pro-
duced by ProtML.
Table 1. Tests of significance for competing hypotheses. Statistical tests of significance were conducted
for different competing phylogenetic hypotheses within Pancrustacea and within arthropod classes. au,
approximately unbiased test; kh, Kishino-Hasegawa test; sh, Shimodaira-Hasegawa test (26).
Tree ln L (ProtML) au kh sh
(Collembola, (Crustacea, Insecta)) 19723.73 0.991 0.979 0.979 Best
(Crustacea, (Collembola, Insecta)) 19744.96 0.009 0.021 0.021
(Myriapoda, (Chelicerata, Pancrustacea)) 19723.73 0.509 0.496 0.649 Best
((Myriapoda, Chelicerata), Pancrustacea) 19723.97 0.509 0.504 0.626
(Chelicerata, (Myriapoda, Pancrustacea)) 19739.90 0.006 0.032 0.084
R EPORTS
21 MARCH 2003 VOL 299 SCIENCE www.sciencemag.org1888
on December 18, 2006 www.sciencemag.orgDownloaded from
Our analysis, based on a large, specifical-
ly targeted data set and modern statistical
tools, strongly supports the view that
Hexapoda is not monophyletic, that at least
some apterygotes have adapted to life on land
independently from insects, and that those
features shared between some apterygotes
and insects might have originated indepen-
dently in these lineages.
References and Notes
1. D. E. G. Briggs, R. A. Fortey, Science 246, 241 (1989).
2. M. J. Emerson, F. R. Schram, in Arthropod Relation-
ships, R. A. Fortey, R. H. Thomas, Eds. (Chapman &
Hall, London, 1997), pp. 67– 86.
3. C. E. Cook, M. L. Smith, M. J. Telford, A. Bastianello, M.
Akam, Curr. Biol. 11, 759 (2001).
4. M. Duman-Scheel, N. H. Patel, Development 126,
2327 (1999).
5. J. L. Boore, D. V. Lavrov, W. M. Brown, Nature 392,
667 (1998).
6. J. W. Shultz, J. C. Regier, Proc. R. Soc. London Ser. B
267, 1011 (2000).
7. Hexapoda is defined as composed of the apterygotan
orders Protura, Collembola, Diplura, Microcoryphia,
Zygentoma, and all pterygotan orders (Pterygota).
Microcoryphia Zygentoma Pterygota constitute
the Insecta sensu stricto.
8. W. Hennig, Insect Phylogeny (Wiley, New York,
1981).
9. K.-D. Klass, N. P. Kristensen, Ann. Soc. Entomol. Fr.
37, 265 (2001).
10. O. Kraus, in Arthropod Relationships, R. A. Fortey,
R. H. Thomas, Eds. (Chapman & Hall, London, 1997),
pp. 295–303.
11. W. Dohle, Ann. Soc. Entomol. Fr. 37, 85 (2001).
12. M. Friedrich, D. Tautz, Nature 376, 165 (1995).
13. C. Bitsch, J. Bitsch, Zool. Scr. 29, 131 (2000).
14. A. Carapelli, F. Frati, F. Nardi, R. Dallai, C. Simon,
Pedobiologia 44, 361 (2000).
15. G. Giribet, C. Ribera, Cladistics 16, 204 (2000).
16. T. Spears, G. Abele, in Arthropod Relationships,R.A.
Fortey, R. H. Thomas, Eds. (Chapman & Hall, London,
1997), pp. 169 –188.
17. F. Nardi, A. Carapelli, P. P. Fanciulli, R. Dallai, F. Frati,
Mol. Biol. Evol. 18, 1293 (2001).
18. E. Garcia-Machado et al., J. Mol. Evol. 49, 142 (1999).
19. Materials and methods are available as supporting
online material on Science Online.
20. U. W. Hwang, M. Friedrich, D. Tautz, C. J. Park, W.
Kim, Nature 413, 154 (2001).
21. K. G. Helfenbein, W. M. Brown, J. L. Boore, Mol. Biol.
Evol. 18, 1734 (2001).
22.
D. L. Swofford, G. J. Olsen, P. J. Waddell, D. M. Hillis, in
Molecular Systematics, D. M. Hillis, C. Moritz, B. K. Mable,
Eds. (Sinauer, Sunderland, MA, 1996), pp. 407–514.
23. R. Dallai, B. A. Afzelius, in The Male Gamete, from
Basic Science to Clinical Applications, C. Gagnon, Ed.
(Cache River, Vienna, 1999), pp. 333–350.
24. J. Adachi, M. Hasegawa, Comput. Sci. Monogr. 28,1
(1996).
25. J. P. Huelsenbeck, F. Ronquist, Bioinformatics 17, 754
(2001).
26. H. Shimodaira, Syst. Biol. 51, 492 (2002).
27. This study was supported by grants from the Minis-
tero dell’Istruzione, dell’Universita` e della Ricerca,
from the Progetto Nazionale di Ricerca in Antartide,
and from the University of Siena. Part of this work
was performed under the auspices of the U.S. Depart-
ment of Energy, Office of Biological and Environmen-
tal Research, and by the University of California,
Lawrence Berkeley National Laboratory, under con-
tract DE-AC03-76SF00098.
Supporting Online Material
www.sciencemag.org/cgi/content/full/299/5614/1887/
DC1
Materials and Methods
Figs. S1 and S2
References
20 September 2002; accepted 7 February 2003
Role of EphA4 and EphrinB3 in
Local Neuronal Circuits That
Control Walking
Klas Kullander,
1,2
* Simon J. B. Butt,
3
James M. Lebret,
3
Line Lundfald,
3
Carlos E. Restrepo,
3
Anna Rydstro¨m,
2
Ru¨diger Klein,
4
Ole Kiehn
3
*
Local circuits in the spinal cord that generate locomotion are termed central
pattern generators (CPGs). These provide coordinated bilateral control over the
normal limb alternation that underlies walking. The molecules that organize the
mammalian CPG are unknown. Isolated spinal cords from mice lacking either
the EphA4 receptor or its ligand ephrinB3 have lost left-right limb alternation
and instead exhibit synchrony. We identified EphA4-positive neurons as an
excitatory component of the locomotor CPG. Our study shows that dramatic
locomotor changes can occur as a consequence of local genetic rewiring and
identifies genes required for the development of normal locomotor behavior.
Rhythmic movements such as locomotion and
swimming require that muscles contract and
relax in a complex repetitive pattern. Central
pattern generators, or CPGs, are local spinal
neuronal networks that generate and coordinate
these rhythmic muscle activities (1, 2). In the
fruit fly, it was recently shown that the CPG for
peristaltic crawling develops in the complete
absence of sensory input (3). In two nonmam-
malian vertebrate species, the lamprey and the
Xenopus tadpole, the critical neuronal compo-
nents of the locomotor CPG have been identi-
fied (4, 5). In mammals, the CPGs controlling
limb movements are located in the ventromedial
part of the spinal cord (6). However, the neuro-
nal organization is still poorly understood (2),
and no molecules that contribute to CPG devel-
opment have been identified. Because CPGs are
important for spinal control of walking in hu-
mans (7), understanding their neuronal organi-
zation and molecular determination is essential
in the ongoing effort to reestablish locomotor
1
Department of Medical Biochemistry, Gothenburg
University, Medicinaregatan 9 A, 405 30 Gothenburg,
Sweden.
2
AstraZeneca Transgenics and Comparative
Genomics, AstraZeneca, 431 83 Mo¨lndal, Sweden.
3
Mammalian Locomotor Laboratory, Department of
Neuroscience, The Karolinska Institute, Retzius vag 8,
171 77 Stockholm, Sweden.
4
Max-Planck Institute of
Neurobiology, Am Klopferspitz 18A, D-82152 Martin-
sried, Germany.
*To whom correspondence should be addressed. E-
mail: klas.kullander@medkem.gu.se (K.K.); ole.kiehn@
neuro.ki.se (O.K.)
These authors contributed equally to this work.
Fig. 1. The lumbar seg-
ments of the spinal
cords of neonatal
ephA4- and ephrinB3-
null mice exhibit syn-
chronous left-right
ventral root activity.
(A to C) Images of WT
mice displaying nor-
mal locomotor activity
(A) or ephA4-null mice
(B) and ephrinB3-null
mice (C) displaying a
rabbitlike gait. (D to F)
Recorded activity after
application of NMDA
and serotonin to the
isolated spinal cord
(a 4-M solution of
each drug) of WT mice
(D), ephA4-null mice
(E), and ephrinB3-null
mice (F) in flexor (L2)
and extensor ventral
(L5) roots. r, right; l,
left. (G to I) Circular
phase diagrams derived from 20 locomotor cycles for the WT (G), ephA4-null (H), and ephrinB3-null (I)
mice shown in (A) to (C), respectively, (J to L) Plots show the vector points of L2 pairings for all
experiments conducted on WT mice (n 5) ( J, green squares); ephA4 heterozygotes (n 13) (K, black
triangles) and homozygotes (n 14) (K, blue circles); and ephrinB3 homozygotes (n 9) (L).
R EPORTS
www.sciencemag.org SCIENCE VOL 299 21 MARCH 2003 1889
on December 18, 2006 www.sciencemag.orgDownloaded from
... Mitochondrial genomes (mt-genomes) provide two primary data types for phylogenetic inference: DNA sequence and gene arrangements. Mt-sequence data (often translated and concatenated coding sequences) have been used extensively in molecular phylogenetics [30][31][32][33][34] and are particularly well suited for the analysis of demosponge relationships because of the low rate of evolution and relatively homogeneous composition of mtDNA sequences in this group [26,[35][36][37][38]. Mt-gene arrangement data have also been used both for the reconstruction of global animal relationships [26,35] and for testing specific phylogenetic hypotheses (e.g., [39]). However, this dataset provides fewer characters for phylogenetic inference and its analysis is computationally more challenging, so the latter application is more common. ...
Article
Full-text available
Class Demospongiae is the largest in the phylum Porifera (Sponges) and encompasses nearly 8,000 accepted species in three subclasses: Keratosa, Verongimorpha, and Heteroscleromorpha. Subclass Heteroscleromorpha contains ∼90% of demosponge species and is subdivided into 17 orders. The higher level classification of demosponges underwent major revision as the result of nearly three decades of molecular studies. However, because most of the previous molecular work only utilized partial data from a small number of nuclear and mitochondrial (mt) genes, this classification scheme needs to be tested by larger datasets. Here we compiled a mt dataset for 136 demosponge species—including 64 complete or nearly complete and six partial mt-genome sequences determined or assembled for this study—and used it to test phylogenetic relationships among Demospongiae in general and Heteroscleromorpha in particular. We also investigated the phylogenetic position of Myceliospongia araneosa , a highly unusual demosponge without spicules and spongin fibers, currently classified as Demospongiae incertae sedis , for which molecular data were not available. Our results support the previously inferred sister-group relationship between Heteroscleromorpha and Keratosa + Verongimorpha and suggest five main clades within Heteroscleromorpha: Clade C0 composed of order Haplosclerida; Clade C1 composed of Scopalinida, Sphaerocladina, and Spongillida; Clade C2 composed of Axinellida, Biemnida, Bubarida; Clade C3 composed of Tetractinellida; and Clade C4 composed of Agelasida, Clionaida, Desmacellida, Merliida, Suberitida, Poecilosclerida, Polymastiida, and Tethyida. The inferred relationships among these clades were (C0(C1(C2(C3+C4)))). Analysis of molecular data from M. araneosa placed it in the C3 clade as a sister taxon to the highly skeletonized tetractinellids Microscleroderma sp. and Leiodermatium sp. Molecular clock analysis dated divergences among the major clades in Heteroscleromorpha from the Cambrian to the Early Silurian, the origins of most heteroscleromorph orders in the middle Paleozoic, and the most basal splits within these orders around the Paleozoic to Mesozoic transition. Overall, the results of this study are mostly congruent with the accepted classification of Heteroscleromorpha, but add temporal perspective and new resolution to phylogenetic relationships within this subclass.
... Previously, people have tried to use morphological data or sequence data of a few genes to estimate the phylogenetic relationships between leafhopper groups, but little research has been performed on Typhlocybinae [10][11][12]. Nowadays, the emergence of a new generation of sequencing technology has brought a breakthrough to solve this problem so that the mitochondrial genome data can verify the existing classification of Typhlocybinae [13][14][15]. ...
Article
Full-text available
Leafhoppers (Hemiptera, Auchenorrhyncha, Cicadellidae) are distributed worldwide and include around 2550 genera, more than 21,000 species, including almost 2000 species in China. Typhlocybinae is the second largest subfamily in Cicadellidae after Deltocephalinae. Previously, morphological characteristics were the diagnostic basis of taxonomy, but they were not combined with molecular biology. The genus Ziczacella Anufryev, 1970 has only six known species worldwide. The mitogenomes of Ziczacella steggerdai Ross, 1965, Ziczacella dworakowskae Anufriev, 1969 and Ziczacella heptapotamica Kusnezov, 1928 were sequenced and identified here for the first time. They all contained 13 PCGs, 22 tRNA genes, 2 rRNA genes, and a control region, and the complete mitochondrial genomes were 15,231 bp, 15,137 bp, and 15,334 bp, respectively. The results show heavy AT nucleotide bias. Phylogenetic analysis yielded the following topology: (Empoascini + Alebrini) + ((Erythroneurini + Dikraneurini) + (Zyginellini + Typhlocybini)). In this study, three newly sequenced species were closely related to Mitjaevia dworakowskae and M. shibingensis. We confirmed the monophyly of the four tribes within Typhlocybinae again, and Zyginellini should be combined with Typhlocybini, which supports Chris’s points.
... S4, Supplementary Material online) albeit with low support. We suspect this is an artifact, but hexapod paraphyly has been suggested before (Nardi et al. 2003). In the full ASTRAL analysis, the node leading to Remipedia and Protura + Diplura + Insecta has 0.96 PP, but support for this node decreased when nodes with low support in gene trees were collapsed to polytomies; when gene tree nodes with <10% and <30% BS support were collapsed prior to ASTRAL, the support values for the Remipedia + Protura + Diplura + Insecta node decreased to 0.91 PP and 0.45 PP, respectively, demonstrating that this node in ASTRAL was supported by gene trees with low support at this node (supplementary fig. ...
Article
Full-text available
The clade Pancrustacea, comprising crustaceans and hexapods, is the most diverse group of animals on earth, containing over 80% of animal species and half of animal biomass. It has been the subject of several recent phylogenomic analyses, yet relationships within Pancrustacea show a notable lack of stability. Here, the phylogeny is estimated with expanded taxon sampling, particularly of malacostracans. We show small changes in taxon sampling have large impacts on phylogenetic estimation. By analyzing identical orthologs between two slightly different taxon sets, we show that the differences in the resulting topologies are due primarily to the effects of taxon sampling on the phylogenetic reconstruction method. We compare trees resulting from our phylogenomic analyses with those from the literature to explore the large tree space of pancrustacean phylogenetic hypotheses and find that statistical topology tests reject the previously published trees in favor of the Maximum Likelihood trees produced here. Our results reject several clades including Caridoida, Eucarida, Multicrustacea, Vericrustacea, and Syncarida. Notably, we find Copepoda nested within Allotriocarida with high support and recover a novel relationship between decapods, euphausiids, and syncarids that we refer to as the Syneucarida. With denser taxon sampling, we find Stomatopoda sister to this latter clade, which we collectively name Stomatocarida, dividing Malacostraca into three clades: Leptostraca, Peracarida, and Stomatocarida. A new Bayesian divergence time estimation is conducted using 13 vetted fossils. We review our results in the context of other pancrustacean phylogenetic hypotheses and highlight 15 key taxa to sample in future studies.
... 42 Mitochondrial genomes (mt-genomes) provide two primary data types for 43 phylogenetic inference: DNA sequence and gene arrangements. Mt-sequence data (often 44 translated and concatenated coding sequences) have been used extensively in molecular 45 phylogenetics [30][31][32][33][34] and are particularly well suited for the analysis of demosponge 46 relationships because of the low rate of evolution and relatively homogeneous 47 composition of mtDNA sequences in this group [26,[35][36][37][38]. Mt-gene arrangement data 48 have also been used both for the reconstruction of global animal relationships [26,35] 49 and for testing specific phylogenetic hypotheses (e.g., [39]). However, this dataset 50 provides fewer characters for phylogenetic inference and its analysis is computationally 51 more challenging, so the latter application is more common. ...
Preprint
Full-text available
Class Demospongiae is the largest in the phylum Porifera (Sponges) and encompasses nearly 8,000 accepted species in three subclasses: Keratosa, Verongimorpha, and Heteroscleromorpha. Subclass Heteroscleromorpha contains 90% of demosponge species subdivided into 17 orders. The higher level classification of demosponges underwent major revision as the result of nearly three decades of molecular studies. However, because most of the previous molecular work only utilized partial data from a small number of nuclear and mitochondrial (mt) genes, this classification scheme needs to be tested by larger datasets. Here we compiled a mt dataset for 136 demosponge species --including 64 complete or nearly complete and six partial mt-genome sequences determined or assembled for this study -- and used it to test phylogenetic relationships among Demospongiae in general and Heteroscleromorpha in particular. We also investigated the phylogenetic position of Myceliospongia araneosa, a highly unusual demosponge without spicules and spongin fibers, currently classified as Demospongiae incertae sedis, for which molecular data were not available. Our results support the previously inferred sister-group relationship between Heteroscleromorpha and Keratosa + Verongimorpha and suggest five main clades within Heteroscleromorpha: Clade C0 composed of order Haplosclerida; Clade C1 composed of Scopalinida, Sphaerocladina, and Spongillida; Clade C2 composed of Axinellida, Biemnida, Bubarida; Clade C3 composed of Tetractinellida; and Clade C4 composed of Agelasida, Clionaida, Desmacellida, Merliida, Suberitida, Poecilosclerida, Polymastiida, and Tethyida. The inferred relationships among these clades were (C0(C1(C2(C3+C4)))). Analysis of molecular data fromM. araneosa placed it in the C3 clade as a sister taxon to the highly skeletonized tetractinellids Microscleroderma sp. and Leiodermatium sp. Molecular clock analysis dated divergences among the major clades in Heteroscleromorpha from the Cambrian to the Early Silurian, the origins of most heteroscleromorph orders in the middle Paleozoic, and the most basal splits within these orders around the Paleozoic to Mesozoic transition. Overall, the results of this study are mostly congruent with the accepted classification of Heteroscleromorpha, but add temporal perspective and new resolution to phylogenetic relationships within this subclass.
... There are diverging views as to the evolutionary relationships between wingless and pterygote hexapods. On the basis of mitochondrial genome sequences, Nardi et al. (2003) suggested that Hexapoda is paraphyletic, and that the hexapod body plan originated twice: once in the insects and once in Collembola. However, nuclear markers strongly suggest that Hexapoda is monophyletic and Collembola are related to either Protura or Diplura (Timmermans et al., 2008). ...
... These findings could be explained by the decay of the phylogenetic signal or a limited signal in the mitogenomic sequences. The limitations of mitogenomes applied in deep phylogeny of Arthropod have already been pointed out [63] and emphasized [1,[4][5][6]64]. When the non-phylogenetic signal was higher than the phylogenetic signal due to mutational saturation, high AT-content, parasitic life-styles or explosive radiation events, considerable systematically erroneous relationships were recovered [6]. ...
Article
Full-text available
Conflicting phylogenetic signals are pervasive across genomes. The potential impact of such systematic biases may be reduced by phylogenetic approaches accommodating for heterogeneity or by the exclusive use of homoplastic sites in the datasets. Here, we present the complete mitogenome of Lynceus grossipedia as the first representative of the suborder Laevicaudata. We employed a phylogenomic approach on the mitogenomic datasets representing all major branchiopod groups to identify the presence of conflicts and concordance across the phylogeny. We found pervasive phylogenetic conflicts at the base of Diplostraca. The homogeneity of the substitution pattern tests and posterior predictive tests revealed a high degree of compositional heterogeneity among branchiopod mitogenomes at both the nucleotide and amino acid levels, which biased the phylogenetic inference. Our results suggest that Laevicaudata as the basal clade of Phyllopoda was most likely an artifact caused by compositional heterogeneity and conflicting phylogenetic signal. We demonstrated that the exclusive use of homoplastic site methods combining the application of site-heterogeneous models produced correct phylogenetic estimates of the higher-level relationships among branchiopods.
... Specific mitochondrial genes, such as the 12S ribosomal RNA (12S rRNA), 16S ribosomal RNA (16S rRNA), NADH dehydrogenase subunit 4 (ND4), cytochrome c oxidase subunit I (COI) and cytochrome b (Cytb) have been often used to compare phylogenetic relationships between vertebrates (Malhotra and Thorpe 2004;Rivera et al. 2018;Sidharthan et al. 2021). Although phylogenetic studies using mitochondrial sequences have both advantages and disadvantages (Rubinoff and Holland 2005), using the complete mitochondrial genome may provide higher phylogenetic resolution between species (Inoue et al. 2003;Nardi et al. 2003;Qian et al. 2018). In addition, a comparison of gene organisations, using complete mitochondrial genomes, could give a more precise understanding of the evolutionary histories between species (Boore 1999). ...
Article
Full-text available
Studies using complete mitochondrial genome data have the potential to increase our understanding on gene organisations and evolutionary species relationships. In this study, we compared complete mitochondrial genomes between all 22 squamate species listed in South Korea. In addition, we constructed Maximum Parsimony (MP), Maximum Likelihood (ML) and Bayesian Inference (BI) phylogenetic trees using 13 mitochondrial protein-coding genes. The mitochondrial genes for all six species in the suborder Sauria followed the same organisation as the sequenced Testudines (turtle) outgroup. In contrast, 16 snake species in the suborder Serpentes contained some gene organisational variations. For example, all snake species contained a second control region ( CR2 ), while three species in the family Viperidae had a translocated tRNA-Pro gene region. In addition, the snake species, Elaphe schrenckii , carried a tRNA-Pro pseudogene. We were also able to identify a translocation of a tRNA-Asn gene within the five tRNA ( WANCY gene region) gene clusters for two true sea snake species in the subfamily Hydrophiinae. Our BI phylogenetic tree was also well fitted against currently known Korean squamate phylogenetic trees, where each family and genus unit forms monophyletic clades and the suborder Sauria is paraphyletic to the suborder Serpentes. Our results may form the basis for future northeast Asian squamate phylogenetic studies.
Article
Full-text available
The coffee leaf miner (Leucoptera coffeella) is one of the major pests of coffee crops in the neotropical regions, and causes major economic losses. Few molecular data are available to identify this pest and advances in the knowledge of the genome of L. coffeella will contribute to improving pest identification and also clarify taxonomy of this microlepidoptera. L. coffeella DNA was extracted and sequenced using PacBio HiFi technology. Here we report the complete L. coffeella circular mitochondrial genome (16,407 bp) assembled using Aladin software. We found a total of 37 genes, including 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs), 2 ribosomal RNA genes (rRNAs) and an A + T rich-region and a D-loop. The L. coffeella mitochondrial gene organization is highly conserved with similarities to lepidopteran mitochondrial gene rearrangements (trnM-trnI-trnQ). We concatenated the 13 PCG to construct a phylogenetic tree and inferred the relationship between L. coffeella and other lepidopteran species. L. coffeella is found in the Lyonetiidae clade together with L. malifoliella and Lyonetia clerkella, both leaf miners. Interestingly, this clade is assigned in the Yponomeutoidea superfamily together with Gracillariidae, and both superfamilies displayed species with leaf-mining feeding habits.
Article
Larval feeding by the moth genus Ogmograptis (Bucculatricidae: Lepidoptera) creates one of the most iconic features of the Australian bush—the ‘scribbles’ found on smooth‐barked Eucalyptus . The taxonomic history of Ogmograptis has been challenging, with members of the genus being initially described in four different genera representing three different superfamilies. While prior phylogenetic analysis has placed Ogmograptis within the Bucculatricidae, these findings were not strongly supported and there was poor resolution of the early diverging, non‐Apoditrysia superfamilies that Ogmograptis has been assigned to by different authors. As a consequence, the unique larval biology of scribbly moths cannot yet be interpreted in an evolutionary context. Phylogenomic analysis of whole mitochondrial (mt) genome data for Ogmograptis , related non‐Apoditrysia and taxa representing the superfamily‐level diversity of the order strongly supports its placement within the Bucculatricidae, a monophyletic Gracillarioidea and a clade of Gracillarioidea + Yponomeutoidea that was sister to the Apoditrysia. The hypermetamorphic larval development in Ogmograptis can thus be interpreted as an elaboration of the ancestral pattern of the clade Gracillarioidea + Yponomeutoidea that has specialised for phellogen/callus feeding within the bark. The utility of mt genomes for deep‐level phylogenetic study of the Lepidoptera is reviewed against prior multi‐locus and nuclear phylogenomic datasets. Mt phylogenomic analyses are sensitive to analytical methods and the inclusion versus exclusion of high‐variability data partitions for deep‐level relationships, already shown to be uncertain by multi‐locus or nuclear phylogenomic analyses, in particular relationships between apoditrysian and obtectomeran superfamilies. While mt genomes are ideal for examining the relationships of rare, physically small or difficult to collect taxa such as Ogmograptis , due to the low technical hurdles to collecting whole genomes, continued attention to the analytical sensitivities of phylogenies that use this data source is needed to reliably advance our understanding of deep lepidopteran evolution.
Chapter
Evolutionary developmental biology, or 'evo-devo', is the study of the relationship between evolution and development. Dealing specifically with the generative mechanisms of organismal form, evo-devo goes straight to the core of the developmental origin of variation, the raw material on which natural selection (and random drift) can work. Evolving Pathways brings together contributions that represent a diversity of approaches. Topics range from developmental genetics to comparative morphology of animals and plants alike, and also include botany and palaeontology, two disciplines for which the potential to be examined from an evo-devo perspective has largely been ignored until now. Researchers and graduate students will find this book a valuable overview of current research as we begin to fill a major gap in our perception of evolutionary change.
Article
Full-text available
We sequenced the complete mitochondrial DNA (mtDNA) of the articulate brachiopod Terebratalia transversa. The circular genome is 14,291 bp in size, relatively small compared with other published metazoan mtDNAs. The 37 genes commonly found in animal mtDNA are present; the size decrease is due to the truncation of several tRNA, rRNA, and protein genes, to some nucleotide overlaps, and to a paucity of noncoding nucleotides. Although the gene arrangement differs radically from those reported for other metazoans, some gene junctions are shared with two other articulate brachiopods, Laqueus rubellus and Terebratulina retusa. All genes in the T. transversa mtDNA, unlike those in most metazoan mtDNAs reported, are encoded by the same strand. The A+T content (59.1%) is low for a metazoan mtDNA, and there is a high propensity for homopolymer runs and a strong base-com positional strand bias. The coding strand is quite G+T-rich, a skew that is shared by the confamilial (laqueid) species L. rubellus but is the opposite of that found in T. retusa, a cancellothyridid. These compositional skews are strongly reflected in the codon usage patterns and the amino acid compositions of the mitochondrial proteins, with markedly different usages being observed between T. retusa and the two laqueids. This observation, plus the similarity of the laqueid noncoding regions to the reverse complement of the noncoding region of the cancellothyridid, suggests that an inversion that resulted in a reversal in the direction of first-strand replication has occurred in one of the two lineages. In addition to the presence of one noncoding region in T. transversa that is comparable with those in the other brachiopod mtDNAs, there are two others with the potential to form secondary structures; one or both of these may be involved in the process of transcript cleavage.
Article
Full-text available
Peer Reviewed http://deepblue.lib.umich.edu/bitstream/2027.42/62560/1/392667a0.pdf
Article
Full-text available
The program MRBAYES performs Bayesian inference of phylogeny using a variant of Markov chain Monte Carlo. Availability: MRBAYES, including the source code, documentation, sample data files, and an executable, is available at http://brahms.biology.rochester.edu/software.html. Contact: johnh{at}brahms.biology.rochester.edu
Chapter
The monophyletic origin of the Tracheata (synonyms: Antennata; Atelocerata) has never been questioned seriously by previous authors (Snodgrass, 1938; Hennig, 1966, 1982; Manton, 1977; Hennig, 1981; Ax, 1987; Boudreaux, 1987; Kraus and Kraus, 1994, 1996). Apparently, the Tracheata form the sister taxon of the Crustacea (or a part of them). Recent work by developmental biologists also points in this direction: for example, this is shown by details of the segmental composition of the crustacean head as compared with insects (Scholtz, 1995).
Chapter
Resolution concerning issues of higher-order crustacean phylogeny remains elusive even after years of thorough morphological and palaeontological scrutiny. Surprisingly, there is as yet no consensus regarding even the number of constituent crustacean classes. One view (Schram, 1986) based on a cladistic analysis of morphological characters suggests there are four classes: Remipedia, Phyllocarida, Maxillopoda and Malacostraca (Figure 14.1(a)); an alternative cladistic analysis (Brusca and Brusca, 1990) suggests there are five: Remipedia, Branchiopoda, Cephalocarida, Maxillopoda and Malacostraca (Figure 14.1(b)). Still another view (Bowman and Abele, 1982), presented as a classification rather than a phylogeny, divides crustaceans into six classes: the five aforementioned groups and the Ostracoda (Figure 14.1(c)).
Article
In a series of papers some years ago (Emerson and Schram, 1990a,b; Schram and Emerson, 1991) we proposed Arthropod Pattern Theory (APT). Before Michael Emerson’s death (22 March, 1990), he and I had planned to perform a cladistic analysis of the fossil and recent arthropods using APT features. He prepared a preliminary data set, and we actually ran some analyses. A series of circumstances, however, intervened after the preparation of Schram and Emerson (1991), and time and opportunity only now have presented themselves to allow one of us (F.R.S.) to take this project up again. This still remains, nevertheless, very much a joint effort, since Michael had prepared before his death extensive notes and observations concerning arthropod body plans, and these have formed the foundation of this analysis.
Article
Basal hexapodan orders (Apterygota) are traditionally divided in two well defined taxonomic groups, Entognatha and Ectognatha. Entognathy (the enclosed mouthparts condition) is considered the most distinctive character joining the Ellipura (Protura+ Collembola) with the order Diplura, whereas the presence of exposed mouthparts (ectognathy) is the feature shared by Microcoryphia and Zygentoma with the pterygote insects. In spite of the growing interest for the evolutionary history of the Apterygota, there is no complete agreement among the general phylogenetic hypotheses based on the study of morphological characters. In this study we analyzed the DNA sequence of segments of the nuclear Elongation Factor-1α(EF-1α) and of the mitochondrial 12S rRNA genes, and used different methods of phylogenetic reconstruction. The 12S gene seems to be more suited than the EF-1α to resolve some of the most outstanding systematic disputes, whereas the lack of resolution at the deeper nodes does not allow to assess the phylogenetic relationships within Microcoryphia and between ectognathan orders. We have obtained a fairly high support for the monophyly of the orders Diplura and Zygentoma. In the 12S analysis, the Ellipura and the Entognatha form monophyletic assemblages. In addition, the study of the distribution of introns in the EF-1α suggests a relationship between Collembola and Diplura.
Article
The phylogeny of the basal hexapods, the so-called apterygote insects, was studied using parsimony analysis procedures. Most analyses took into account 47 characters mainly based on external morphology, and 19 taxa including 14 apterygote representatives, 3 pterygotes and also 2 distantly related myriapods were used as outgroups. The binary and multistate characters are discussed in detail and treated as unordered and equally weighted. Other analyses were performed using a second data set in which 28 characters, based on internal anatomy and already used in a previous work (Bitsch & Bitsch 1998), were added to the first data set. This second matrix was restricted to 12 terminal taxa, the same as those of our previous work. The results of the different analyses are generally congruent. They strongly support the monophyly of several orders (Protura, Collembola, Archaeognatha) and of two groupings (Ectognatha, Dicondylia). Three other assemblages (Ellipura, Diplura, Entognatha) appear as parsimonious phylogenetic hypotheses, but they are never supported by the cladistical analyses and are based on a very small number of autapomorphies; so, the monophyly of each of them is not firmly established. Archaeognatha appears as the sister group of the Dicondylia. The three unresolved representatives of the Zygentoma are found as the sister group of the Pterygota. The results are discussed in the light of current concepts in hexapod phylogeny.
Article
The evolutionary relationships among arthropods are of particular interest because the best-studied model system for ontogenetic pattern formation, the insect Drosophila, is a member of this phylum. Evolutionary inferences about the developmental mechanisms that have led to the various designs of the arthropod body plan depend on a knowledge of the phylogenetic framework of arthropod evolution. Based on morphological evidence, but also on palaeontological consideration, the sister group of the insects is believed to be found among the myriapods. Using nuclear ribosomal gene sequences for constructing a molecular phylogeny, we provide strong evidence that the crustaceans and not the myriapods should be considered to be the sister group of the insects. Moreover, the degree of sequence divergence suggests that the diversification of the myriapods occurred during the Cambrian. Our findings have general implications for the course of land colonization by the different arthropod groups, as well as for the interpretation of primitive and derived features of arthropod morphology.