Anda di halaman 1dari 12

The Complete Sequence of the Mitochondrial Genome of the Crustacean

Penaeus monodon: Are Malacostracan Crustaceans More Closely Related to


Insects than to Branchiopods?
Kate Wilson, Valma Cahill, Elizabeth Ballment, and John Benzie
Australian Institute of Marine Science, Townsville, Queensland, Australia

The complete sequence of the mitochondrial genome of the giant tiger prawn, Penaeus monodon (Arthropoda,
Crustacea, Malacostraca), is presented. The gene content and gene order are identical to those observed in Dro-
sophila yakuba. The overall AT composition is lower than that observed in the known insect mitochondrial genomes,
but higher than that observed in the other two crustaceans for which complete mitochondrial sequence is available.
Analysis of the effect of nucleotide bias on codon composition across the Arthropoda reveals a trend with the
crustaceans represented showing the lowest proportion of AT-rich codons in mitochondrial protein genes. Phylo-
genetic analysis among arthropods using concatenated protein-coding sequences provides further support for the
possibility that Crustacea are paraphyletic. Furthermore, in contrast to data from the nuclear gene EF1␣, the first
complete sequence of a malacostracan mitochondrial genome supports the possibility that Malacostraca are more
closely related to Insecta than to Branchiopoda.

Introduction
Mitochondrial genome sequence and structure is within the Arthropoda. The resulting analysis supported
widely used to provide information on phylogenetic re- a paraphyletic Crustacea, with Penaeus being more
lationships and on the genetic structure of populations closely related to insects than to the branchiopod crus-
and patterns of gene flow. This information may derive tacean, Artemia. However, this analysis was based on
from studies of gene order, the sequences of individual only partial mitochondrial sequence from P. notialis and
genes, restriction fragment length polymorphism a single branchiopod, A. franciscana, as the complete
(RFLP) analysis of mtDNA, or the sequences of com- mitochondrial sequence of D. pulex was not then avail-
plete genomes. able for inclusion in the analysis. Moreover, the mtDNA
The giant tiger prawn, Penaeus monodon, is the of Artemia appears to be evolving at an accelerated rate
major species of shrimp farmed in aquaculture world- (Crease 1999; see fig. 5), which could affect molecular
wide. The Penaeidae fall within the order Decapoda phylogenetic inference (Hillis, Moritz, and Mable 1996).
(class Malacostraca), which contains most commercially In this paper, we present the first complete mito-
important crustacean species. Analyses of penaeid mi- chondrial sequence for a malacostracan, P. monodon.
tochondrial genetics have been carried out using RFLP We discuss the composition of the P. monodon mito-
analysis and single-gene sequence analysis (Palumbi and chondrial genome and analyze the effect of nucleotide
Benzie 1991; Bouchon, Souty-Grosset, and Raimond bias on codon composition across the Arthropoda. Fi-
1994; Klinbunga et al. 1999). However, there is no com- nally, we describe the results of phylogenetic analyses
plete mitochondrial genome sequence available for any among members of the Arthropoda for which complete
decapod, or even for any malacostracan crustacean. To mitochondrial sequences are available, paying particular
date, the only complete crustacean mitochondrial ge- attention to the relationships of the different crustacean
nome sequences available are those of two branchio- groups and Insecta.
pods: the brine shrimp Artemia franciscana (Branchio-
poda, Anostraca) (Valverde et al. 1994) and the water Materials and Methods
flea Daphnia pulex (Branchiopoda, Cladocera) (Crease Determination of Sequence
1999). Thus, there is a dearth of mitochondrial sequence
data available for phylogenetic comparisons within the Mitochondrial DNA was extracted from pleopod
Crustacea and related groups. muscle from a single hatchery-grown P. monodon spec-
Recently, partial sequence (8,754 bp) was pub- imen derived from parents collected near Cairns, North
lished for the southern pink shrimp, Penaeus notialis Queensland, Australia, according to Dowling, Moritz,
(Garcı́a-Machado et al. 1999). This sequence encom- and Palmer (1990), using 10 mM EDTA in the homog-
passed seven protein- coding genes, and the aligned enization buffer. This mtDNA was cut with XbaI to pro-
amino acid sequences of this species, Artemia, and four duce three fragments (2.1, 3.8, and 10.0 kb), which were
insects were used to make inferences about phylogeny cloned into pUC19. The resulting three plasmids,
AIMS-P.mon74, AIMS-P.mon75, and AIMS-P.mon76,
were sequenced using fluorescent cycle-sequencing kits
Key words: Malacostraca, complete mitochondrial genome se- from the Applied Biosystems division of Perkin-Elmer.
quence, arthropod phylogeny, crustacean/insect relationships, penaeid
shrimp. Initial sequencing was carried out using vector primers,
and subsequent sequencing was performed by ‘‘primer
Address for correspondence and reprints: Kate Wilson, Australian
Institute of Marine Science, PMB 3, Townsville MC, Queensland 4810, walking’’; i.e., primers were synthesized corresponding
Australia. E-mail: k.wilson@aims.gov.au. to previously obtained P. monodon mitochondrial se-
Mol. Biol. Evol. 17(6):863–874. 2000 quence. Sequence was obtained for both strands for the
䉷 2000 by the Society for Molecular Biology and Evolution. ISSN: 0737-4038 complete genome. To determine the relative orientation

863
864 Wilson et al.

of the three clones and to ensure that no sequence was give a single multiple sequence alignment containing
missing due to a failure to clone very small XbaI frag- 3,327 characters (2,335 variable and 1,687 parsimony-
ments, primers that faced ‘‘out’’ of the clones were de- informative) for phylogenetic analysis. The alignments
signed and used to amplify junction fragments from in- generated are available from the corresponding author
tact mitochondrial DNA by the polymerase chain reac- on request.
tion (PCR). In each case, single bands of the predicted The aligned sequences were subjected to distance
size were obtained. These three PCR products were also analysis using MEGA 1.01 (Kumar, Tamura, and Nei
sequenced using ABI reagents. 1993) and PHYLIP 3.572c (Felsenstein 1993), parsi-
mony analysis using PAUP 3.1.1 (Swofford 1993), and
Analysis of Codon Bias maximum-likelihood analysis using MOLPHY 2.3b
(Adachi and Hasegawa 1996b). Distances between taxa
Potential amino acid bias due to underlying nucle-
were estimated in the program MEGA using the number
otide bias was analyzed using the ‘‘square plot’’ analysis
of differences, p-distance, the Poisson correction, and
of Foster, Jermiin, and Hickey (1997). In brief, this an-
gamma probabilities with gamma set to 1 or 2. Trees
alyzes the amino acid composition by looking at the
were then constructed using neighbor joining. Bootstrap
proportion of amino acids with codons that have (1) A
probabilities were estimated with 1,000 replications us-
or T in both the first and the second positions (2) A or
ing pairwise deletion for gaps and missing data. Dis-
T in either the first or the second position, and (3) only
tance analyses were also carried out with the PHYLIP
G or C in these two positions. The amino acid leucine
package using the Dayhoff PAM matrix for amino acid
was excluded from this analysis, as it can be encoded
substitutions and neighbor-joining tree building. Boot-
by codons from two different classes (TTR and CTR)
strap probabilities were estimated in PHYLIP with 100
and hence can accommodate changes in underlying base
replicates. For parsimony analysis, the heuristic search
composition without a change in encoded amino acids.
option in PAUP 3.1.1 was used. A simple addition se-
The data are presented here as a histogram rather than
quence was used and tree bisection-reconnection branch
as square plots to show underlying trends across the
swapping was performed. Bootstrap values were esti-
Arthropoda.
mated using 1,000 replications. For both distance and
parsimony analyses, L. terrestris was defined as the out-
Phylogenetic Analyses
group. For maximum-likelihood analysis, the mtREV-24
Protein-coding sequences were individually aligned model of amino acid substitution in mitochondrial pro-
using Pileup (Genetics Computer Group 1994) with the tein-coding genes was used (Adachi and Hasegawa
gap creation penalty set to 5.0 and the gap extension 1996a), and the data were analyzed using both the star
penalty set to 0.3. Sequences were from 10 of the 11 decomposition method and local rearrangement options.
arthropods for which complete mitochondrial genome For the latter, the 200 best trees using stepwise addition
sequences were available in GenBank: among the in- were first generated, and these tree topologies were then
sects (Hexapoda), the dipteran flies Drosophila yakuba subjected to local rearrangements to determine the max-
(X03240, Hexapoda, Insecta, Diptera), Drosophila me- imum-likelihood tree.
lanogaster (U37541, Hexapoda, Insecta, Diptera), Cer-
atitis capitata (CCA242872, Hexapoda, Insecta, Dip- Results
tera), Anopheles quadrimaculatus (L04272, Hexapoda, Genome Size
Insecta, Diptera), and Anopheles gambiae (L20934,
Three plasmids were constructed from the three
Hexapoda, Insecta, Diptera) and the orthopteran Locusta
major XbaI fragments of the P. monodon mitochondrial
migratoria (X80245, Hexapoda, Insecta, Orthoptera);
genome. As the total size of these fragments was esti-
the branchiopod crustaceans A. franciscana (X69067,
mated to be approximately 16 kb, it was assumed that
Crustacea, Branchiopoda, Anostraca) and Daphnia pu-
this comprised the majority of the mitochondrial ge-
lex (AF117817, Crustacea, Branchiopoda, Cladocera);
nome. PCR analysis of intact mtDNA using primers that
and from the Chelicerata, the prostriate tick Ixodes hex-
spanned the XbaI sites at the plasmid ends revealed that
agonus (AF081828, Chelicerata, Arachnida, Acari) and
one XbaI site had been missed in the cloning strategy.
the metastriate tick Rhipicephalus sanguineus
This was because there were two XbaI sites which were
(AF081829, Chelicerata, Arachnida, Acari). The hon-
almost adjacent within a region which was subsequently
eybee, Apis mellifera (L06178, Hexapoda, Insecta, Hy-
identified as containing the trnG gene. These sites are
menoptera), was not included in the analysis due to the
separated by only 6 bp and thus add 12 bp to the total
extreme nucleotide bias which confounds phylogenetic
sequence length obtained from the plasmids, making a
inference based on both nucleotide and protein sequence
total genome size of 15,984 bp. This sequence has been
analyses (Galtier and Gouy 1995; Foster, Jermiin, and
deposited in GenBank (accession number AF217843).
Hickey 1997). The earthworm Lumbricus terrestris
(U24570, Annelida, Oligochaeta) was used as an out-
Gene Content and Order
group. Internal sections of nad2 and nad4 were excluded
from the analysis due to severe ambiguities in align- The gene content was identified primarily by align-
ment. In addition, all sequences were trimmed at the N- ment with the known genes from D. yakuba, D. melan-
and C-terminal ends to give an exact and unambiguous ogaster, and A. gambiae. In addition, tRNA genes and
alignment. The alignments were then concatenated to their anticodons were detected using the program t-
Complete mtDNA Sequence of the Shrimp P. monodon 865

FIG. 1.—Gene order in the Penaeus monodon mitochondrial genome. tRNA genes are represented by the single-letter code for the cognate
amino acid. For Leu and Ser, for which there are two tRNAs, the codon recognized has been specified. The numbers are the numbers of bases
between genes, with negative numbers representing probable overlaps between genes (e.g., atp8 and atp6 are separated by ⫺7 bp, ATGTTAA).

RNAscan (Lowe and Eddy 1997). Like all of the other Coding Sequences
arthropod mtDNAs, the P. monodon mitochondrial ge-
nome contains genes for 13 proteins, 22 tRNAs, and two The P. monodon mitochondrial genome appears to
rRNAs. In addition, there is a 991-bp-long A⫹T-rich use the same genetic code as Drosophila. It is certainly
region that does not contain any known genes. As the the case that UGA specifies an amino acid rather than
origin of replication of the Drosophila mtDNA maps in a stop codon, since otherwise the majority of proteins
the corresponding region in the Drosophila mtDNA ge- would show premature termination. It is also reasonable
nome, this region is presumed to be the equivalent of to presume that the amino acids specified by each codon
the vertebrate control region (Wolstenholme 1992). The family are the same as in Drosophila, as use of this code
order and arrangement of the genes in P. monodon is gives rise to the expected open reading frames with a
illustrated in figure 1. It is identical to that observed in high degree of similarity to the same proteins identified
the insects D. melanogaster, D. yakuba, and C. capitata from Drosophila and other arthropods. However, this
and in the branchiopod D. pulex. could be formally proven only by purification and se-
The genes are densely packed, as in other mito- quencing of P. monodon mitochondrial proteins.
chondrial genomes. Excluding the 991-bp control re- The location and putative initiation and termination
gion, there are a total of only 139 bp found between the codons of the protein-coding sequences are shown in
different genes in the P. monodon mitochondrial ge- table 1. Twelve out of the 13 protein coding genes ap-
nome. There are three clear cases of overlapping genes, pear to initiate with the codon ATN. However, as is the
which all occur between adjacent genes encoded on the case with several other arthropod species, the initiation
same strand. There is a 5-bp overlap (TCTAA) between codon of cox1 is unclear. In P. monodon, it appears to
the last codon and the termination codon of cox1 and start with ACG based on alignment with other species.
trnL(taa) and a 7-bp overlap (ATGATAA) between the In Drosophila, Locusta, and Daphnia, there is a tetra-
termination of atp8 and the initiation of atp6 which is nucleotide sequence (ATAA or ATTA) immediately pre-
seen in most arthropod mitochondrial genomes (the hon- ceding the apparent first codon, and it has been sug-
eybee, A. mellifera, has an even greater overlap, 19 bp); gested that this tetranucleotide could serve as the initi-
likewise, the end of nad4L and nad4 overlap by 7 bp ation codon (Clary and Wolstenholme 1983; de Bruijn
(ATGTTAA) on the minority coding strand. In addition, 1983). However, no such sequence is present in P. mon-
there are a further three cases in which an overlap could odon, and we are not aware of any model indicating how
occur if the complete stop codon present in the mtDNA a tetranucleotide might function in this capacity. The
sequence is used. These are a 2-bp overlap (AA) be- absence of such a tetranucleotide has also been observed
tween the potential termination codon of nad2 (TAA) in other arthropods, e.g., A. gambiae (Beard, Mills, and
and the start of trnW, a 1-bp overlap (A) between the Collins 1993), leading to the conclusion that each of the
termination of nad6 (TAA) and the initiation of cob, and disparate array of first codons in aligned cox1 genes
a 2- bp overlap (AG) between the termination of cob most likely does serve as the initiation codon. An alter-
(TAG) and the start of trnS(uga). However, it is also native model to explain these atypical start codons is
possible that RNA processing clips the genes apart up- that they are converted to more conventional start co-
stream of these potential overlaps and that polyadenyl- dons in the mRNA; for example, in the tomato, the cox1
ation completes the stop codon on the mRNA (Anderson gene starts with an ACG codon, but this is converted to
et al. 1981). an AUG in the mRNA by RNA editing (Kadowaki et
866 Wilson et al.

Table 1
Protein-Coding Genes in the Penaeus monodon Mitochondrial Genome
Majority (J) or
Minority (N)
Gene Start Position End Positiona Start Codon End Codon Amino Acids Strand
nad2. . . . . . . . . 252 1250 ATT TAA 333 J
cox1 . . . . . . . . . 1465 3000 ACGb TAA 512 J
cox2 . . . . . . . . . 3071 3757 ATG T 229 J
atp8 . . . . . . . . . 3900 4055 ATT TAA 52 J
atp6 . . . . . . . . . 4052 4723 ATG TAA 224 J
cox3 . . . . . . . . . 4737 5525 ATG T 263 J
nad3. . . . . . . . . 5594 5944 ATG T 117 J
nad5. . . . . . . . . 8095 6374 ATA T or TA 574 N
nad4. . . . . . . . . 9512 8175 ATG TAA 446 N
nad4L . . . . . . . 9805 9509 ATG TAA 99 N
nad6. . . . . . . . . 9942 10460 ATT TAA 173 J
cob . . . . . . . . . . 10463 11596 ATG TAG 378 J
nada1. . . . . . . . 12627 11692 ATA TAA 312 N
a Designated as the last base of the codon encoding the final amino acid, i.e., excluding partial or complete stop codons.
b See discussion in the text and figure 2 regarding the initiation codon of cox1.

al. 1995). The presumed initiation context of cox1 genes in P. monodon, in contrast to CUU in D. yakuba. The
from arthropods, for which there is complete mitochon- trnK anticodon is also UUU in the related shrimp P.
drial sequence, plus that of the penaeid shrimp P. no- notialis and in the honeybee A. mellifera (Crozier and
tialis, for which there is partial mitochondrial sequence Crozier 1993), but it is CUU in all other arthropods
(GenBank accession numbers X84350–X84357), is known to date. In fact, UUU is the anticodon which
shown in figure 2. would be predicted for lysine because, in most mtDNAs,
Nine of 13 genes have a complete termination co- two-codon families which end in a purine have a U in
don, either TAA (eight genes) or TAG (cytochrome b the wobble position of the anticodon. In the case of D.
only). The other genes all have incomplete termination yakuba and other arthropods, the CUU anticodon for
codons, T, or possibly TA for nad5 (table 1). These are lysine (codons AAA and AAG) requires C-A pairing
believed to be converted to complete TAA codons by (Clary and Wolstenholme 1985).
the addition of A residues during RNA processing (An-
derson et al. 1981). Base Composition
The A⫹T content of the P. monodon genome is
tRNAs
70.6% (A, 5,636; C, 2,669; G, 2,028; T, 5,651 on the
The predicted structures of the P. monodon tRNAs strand that contains the majority of open reading
are shown in figure 3. All tRNAs have the typical clo- frames). The 991-bp control region has an AT compo-
verleaf structures of other mitochondrial tRNAs except sition of 81.5%. These values are lower than those for
for trnS(ucg), which recognizes the codon AGN. This the published insect genomes (75.3%–84.9% for total
latter tRNA lacks the dihydrouridine loop, as observed A⫹T content and 86.0%–96.0% for the AT-rich control
for this tRNA in all other arthropod mitochondrial ge- region), but higher than those for the crustaceans A.
nomes (Wolstenholme 1992). franciscana (64.5% total AT content, 68.0% AT in the
The anticodons are identical to those observed in control region) and D. pulex (62.3% total AT, 67.1% AT
D. yakuba, except for the trnK anticodon, which is UUU in the control region). Among the chelicerates, the

FIG. 2.—Initiation context for cox1 genes from a variety of arthropods. The first five to seven codons are shown in uppercase letters. Where
the first codon is not a typical start codon for arthropod mitochondrial DNA (ATN), the upstream context is shown in lowercase letters. In the
case of the drosophilid flies, the preceding tetranucleotide is postulated to act as the initiation ‘‘codon.’’ However, while this theory might also
be applicable to the locust and to Daphnia, the upstream sequence for the Anopheles species, Ceratitis capitata, and Penaeus monodon do not
conform to this pattern. The presumed initial amino acids are shown on the right-hand side of the figure.
Complete mtDNA Sequence of the Shrimp P. monodon 867

FIG. 3.—The inferred secondary structure of the 22 tRNAs in the Penaeus monodon mitochondrial genome. The tRNAs for Leu and Ser
are identified in the figure by the codons recognized, rather than by the anticodon present in the tRNA itself.

horseshoe crab is 67.7% A⫹T in the 8,459 bp sequenced partial sequence available for P. notialis (for the protein-
(Staton, Daehler, and Brown 1997), and the ticks Ixodes coding genes atp6, atp8, cox1, cox2, cox3, nad2, nad3,
and Rhipicephalus are 72.6% and 77.9% AT-rich, and nad5) is 74.6%. Table 2 illustrates that these values
respectively. for the two penaeid shrimp, representing the class Ma-
There are a total of 3,716 codons in all 13 protein- lacostraca, are intermediate between those for the bran-
coding genes, excluding stop codons. The overall AT chiopod crustaceans and the insects. The percentage of
composition of protein-coding regions is 69.3% but at A⫹T in the third positions of fourfold-degenerate co-
third positions it is 83.7%. When the nucleotide com- dons for the two tick species is similar to that for the
position of the third positions of fourfold-degenerate co- penaeids.
dons is examined, it is 84.9% AT (see table 2). This The codon usage in P. monodon is shown in table
contrasts with much lower values of 68.7% in Artemia 3. A feature of most arthropod genomes sequenced to
and 62.9% in Daphnia. The corresponding value for the date is that the bias toward the nucleotides A and T also
868 Wilson et al.

leads to a bias in the amino acids used. This can be

Daphnia
analyzed by comparing the proportions of amino acids

pulex
17.3
26.7
36.2
19.8
62.9
37.1
with A or T versus C or G at the first and second codon
positions. Penaeus monodon appears to share the bias
of all arthropod genomes toward amino acids encoded

franciscana
by AT-rich codons (fig. 4). In P. monodon, the bias to-

Artemia

14.1
30.9
37.8
17.1
68.7
31.2
ward AT-rich codons appears to be very similar to that
observed in Artemia and Daphnia and is slightly less
than that observed in the insects. The gradation in the
proportion of codons with A or T in both of the first
Penaeus
natialis
10.6 two positions which is observed in the different taxo-
32.4
42.2
14.8
74.6
25.4
nomic groups indicates a lack of evolutionary station-
arity; i.e., there is heterogeneity in the evolutionary pro-
cess leading to differing average codon and, hence, ami-
monodon
Penaeus

no acid compositions.
6.3
38.5
46.4
8.8
84.9
15.1

Phylogenetic Analysis
quadrimaculatus

The P. monodon mitochondrial sequence represents


Aopheles

the 12th complete arthropod mitochondrial genome se-


6.0
46.5
43.0
4.5
89.5
10.5

quenced to date; there are now complete mitochondrial


sequences available for seven insects, two chelicerates,
and three crustaceans. Hence, it was of interest to see
whether mitochondrial sequence comparisons could
Aopheles
gambiae

yield any evidence on phylogenetic relationships be-


3.2
47.8
46.0
2.9
93.8
6.1

tween the different subphyla represented, particularly on


crustacean/insect relationships. As the phylogenetic
comparisons being made are between widely divergent
Ceratitis

phylogenetic groups, most sequence signal will be ob-


capitata
3.1
44.7
48.8
3.4
93.5
6.5

scured by saturation. This was indeed found to be the


case when nucleotide sequences of individual genes
(small- and large-subunit rRNA genes and cox1) were
Base Composition at the Third Codon Positions of Fourfold-Degenerate Codons

melanogaster
Drosophila Drosophila

used in phylogenetic analyses: the phylogenetic trees


produced from these data were largely unresolved
3.3
45.7
48.8
2.2
94.5
5.5

(nodes supported by bootstrap values of ⬍70% were


considered insufficiently supported; Hillis and Bull
1993). However, by aligning the amino acid sequences
yakuba

of all mitochondrially encoded proteins and using the


3.5
46.2
48.3
2.0
94.5
5.5

concatenated set of aligned sequences in phylogenetic


analysis, more robust phylogenetic trees were obtained.
The phylogenetic analysis used data from 11 of the
migratoria

12 complete arthropod mitochondrial genomes available


Locusta

2.3
54.4
40.2
3.1
94.6
5.4

(10 from the database plus P. monodon). Sequence from


the honeybee A. mellifera was omitted as the extreme
AT bias in this species makes phylogenetic analyses
even of protein-coding regions unreliable (Foster, Jer-
mellifera

miin, and Hickey 1997), as evidenced by the fact that


Apis

0.8
62.7
34.1
2.4
96.8
3.2

A. mellifera clustered together with the ticks with a


bootstrap value of 99–100 when included in a similar
phylogenetic analysis (Black and Roehrdanz 1998). The
sanguineus hexagonus
Rhipicephalus Ixodes

earthworm L. terrestris (Boore and Brown 1995) was


10.1
38.5
36.6
14.9
75.1
25.0

used as the outgroup.


Three different tree structures resulted from appli-
cation of different phylogenetic algorithms (fig. 5A–C).
In all cases, the two chelicerate species form a clade
3.5
48.7
40.2
7.5
88.9
11.0

which is quite distinct from a Hexapoda/Crustacea


clade. In the first tree structure (fig. 5A), Artemia and
Daphnia form a clade separate from Hexapoda (repre-
% AT . . . . .
% GC . . . .
G ........
A ........
T ........
C ........
Table 2

sented by the class Insecta) and P. monodon; in the sec-


%

ond (fig. 5B), all Crustacea and Hexapoda form a single


clade, with Penaeus being the crustacean most closely
Complete mtDNA Sequence of the Shrimp P. monodon 869

Table 3
Codon Usage in 13 Protein-Coding Genes of Penaeus monodon
Phe . . . . . . . .TTT 234 Ser . . . . . . . .TCT 109 Tyr . . . . . . . .TAT 113 Cys . . . . . . . .TGT 32
TTC 71 TCC 26 TAC 30 TGC 9
Leu . . . . . . . .TTA 340 TCA 89 End. . . . . . . .TAA 8 Trp . . . . . . . .TGA 88
TTG 40 TCG 13 TAG 1 TGG 14
Leu . . . . . . . .CTT 91 Pro . . . . . . . .CCT 89 His . . . . . . . .CAT 58 Arg . . . . . . . .CGT 22
CTC 21 CCC 13 CAC 27 CGC 2
CTA 74 CCA 36 Gln . . . . . . . .CAA 69 CGA 34
CTG 19 CCG 6 CAG 7 CGG 3
Ile . . . . . . . . .ATTa 250 Thr . . . . . . . .ACT 92 Asn. . . . . . . .AAT 106 Ser . . . . . . . .AGT 43
ATC 48 ACC 19 AAC 35 AGC 8
Met. . . . . . . .ATA 192 ACA 68 Lys . . . . . . . .AAA 70 AGA 71
ATG 26 ACG 16 AAG 10 AGG 0
Val . . . . . . . .GTT 100 Ala . . . . . . . .GCT 129 Asp. . . . . . . .GAT 58 Gly . . . . . . . .GGT 109
GTC 20 GCC 35 GAC 17 GGC 5
GTA 131 GCA 60 Glu . . . . . . . .GAA 65 GGA 87
GTG 11 GCG 6 GAG 17 GGG 33
a Three of the ATT codons are likely to specify methionine (Fearnley and Walker 1987), as they appear to act as the initiation codons of nad2, atp8, and nad6.

related to the Hexapoda and Artemia being the most the insects D. yakuba (16,019 bp) (Clary and Wolsten-
distantly related. In the third tree, Penaeus and Hexap- holme 1985), D. melanogaster (19,517 bp) (Lewis, Farr,
oda form one clade, and the relationship of Daphnia and and Kaguni 1995), A. gambiae (15,363 bp) (Beard,
Artemia to each other and to this clade remains unre- Mills, and Collins 1993), A. quadrimaculatus (15,455
solved. The most striking feature held in common by all bp) (Cockburn, Mitchell, and Seawright 1990), C. cap-
three trees is the paraphyly of the Crustacea, with Pen- itata (15,980 bp) (Spanos et al. 2000), A. mellifera
aeus being more closely related to the insects than to (16,343 bp) (Crozier and Crozier 1993), and L. migra-
the branchiopods. This relationship of Penaeus to In- toria (15,722 bp) (Flook, Rowell, and Gellissen 1995),
secta is supported by bootstrap values of ⬎90% with all the crustaceans A. franciscana (15,822 bp) (Valverde et
tree-building methods. al. 1994) and D. pulex (15,333 bp) (Crease 1999), and
the chelicerates I. hexagonus (14,539 bp) and R. san-
Discussion guineus (14,710 bp) (Black and Roehrdanz 1998).
The gene order in the P. monodon mitochondrial
The length of the P. monodon mitochondrial ge-
genome is identical to that of the three dipteran species
nome falls within the range observed for most animal
D. yakuba, D. melanogaster, and C. capitata, as well as
mitochondrial genomes, including those of the other ar-
to the branchiopod D. pulex. All other arthropod
thropods for which complete sequences are available:
mtDNAs for which complete mitochondrial sequences
are available show distinct differences in the placement
of certain genes. The two Anopheles species differ in
the placement of two tRNAs, Locusta in the placement
of one tRNA, Apis in the placement of eight tRNAs,
Artemia in the placement of the block of trnI to trnQ,
and Ixodes in the placement of one tRNA, and Rhipi-
cephalus shows major rearrangements. In the latter spe-
cies, the block of genes from nad1 to trnQ appears to
have transposed, and there are two further tRNA
rearrangements.
Boore et al. (1995), and Boore and Brown (1998),
and Boore, Lavrov, and Brown (1998) argue that rear-
rangements of metazoan mtDNA are highly unlikely to
be duplicated by convergences and hence can be used
to infer phylogenetic lineages. These authors therefore
conclude that the Crustacea and Insecta form a sister
group, with the Chelicerata and Myriapoda occupying
FIG. 4.—Amino acid content of protein-coding genes in arthropod
separate arthropod lineages (Boore, Lavrov, and Brown
mitochondrial genomes is affected by codon bias. The histogram rep- 1998). The arrangement of genes in the P. monodon
resents the proportion of amino acids for codons that have (1) A or T genome is in accord with this hypothesis. This arrange-
in both the first and the second positions (AT/AT), (2) A or T in either ment is identical to that observed in the crustacean D.
the first or the second position (AT/CG or CG/AT), or (3) only G or
C in these two positions (CG/CG). Leucine and termination codons
pulex and appears to be also conserved in the partially
are excluded. All arthropod taxa for which complete mitochondrial characterized genomes of the malacostracan crustaceans
genome sequences are available are included. Homarus americanus (Boore et al. 1995) and P. notialis
870 Wilson et al.

clusions regarding arthropod phylogeny, although it


should be noted that these trees are necessarily based on
the restricted set of arthropod taxa for which complete
mitochondrial sequences are currently available. In par-
ticular, the available sequences only represent three of
the four extant arthropod subphyla, with no complete
mitochondrial sequence from the Myriapoda being
available in the database.
The first conclusion is that the Chelicerata form a
separate clade from Crustacea/Hexapoda. This appears
to be widely supported by molecular evidence on gene
sequence (Friedrich and Tautz 1995; Black and Roehr-
danz 1998), on mitochondrial gene order (Boore, Lav-
rov, and Brown 1998), and on the molecular basis of
development (Averof and Akam 1995). It is also a view
held by most morphological taxonomists (e.g., Snod-
grass 1938; Kukalová-Peck 1992; Wheeler, Cartwright,
and Hayashi 1993), although there have been proposals
that have placed crustaceans closer to chelicerates
(Schram and Emerson 1991; Briggs, Fortey, and Willis
1992).
The second conclusion is that Crustacea are a par-
aphyletic group. This contradicts many phylogenetic
trees proposed on the basis of morphological data
(Snodgrass 1938; Wheeler, Cartwright, and Hayashi
1993). However, the concept of crustacean paraphyly
has been proposed previously based on both morpho-
logical and molecular characters (Briggs and Fortey
1989; Averof and Akam 1995), although the details of
the paraphyly may vary (see discussion below). The
conclusion that Crustacea is paraphyletic with respect to
FIG. 5.—Phylogenetic trees based on analysis of concatenated, Hexapoda was also drawn by Garcı́a-Machado et al.
aligned protein-coding sequences from 13 mitochondrial coding genes. (1999) in their analysis of partial mitochondrial se-
For all parsimony and distance trees, only nodes with ⬎70% bootstrap
support have been retained (Hillis and Bull 1993). All trees are rooted
quence of P. notialis.
with L. terristris as outgroup. A, Tree derived from maximum-likeli- The third conclusion relates to the relationship of
hood analysis; the same tree was obtained using both star decompo- the two classes of crustaceans that are now represented
sition and local rearrangement options. Bootstrap probabilities shown by complete mitochondrial genome sequences, the Ma-
are local bootstrap probabilities for the subsequent node estimated by lacostraca and the Branchiopoda, to each other and to
the RELL method with 1,000 replications (Adachi and Hasegawa
1996b). The total branch length of the tree is 349.73, and the log the Hexapoda, represented here by the class Insecta. A
likelihood of the data given this tree is ⫺48,236.81. B, Tree obtained corollary of several theories supporting the idea that the
using parsimony analysis. A single tree of 8,282 steps was found with Crustacea are paraphyletic is that Hexapoda derive from
a consistency index excluding uninformative characters of 0.761, a a crustacean ancestor. If this is the case, then it is an
homoplasy index excluding uninformative characters of 0.239, and a
retention index of 0.561. Similar trees were obtained from the neigh-
inescapable conclusion that some Crustacea will be
bor-joining algorithm in the program MEGA using distance matrices more closely related to Hexapoda than to other crusta-
generated using either number of differences or p-distances. C, Tree cean groups. This theory is consistent with the trees gen-
obtained using distance analysis, as implemented in the PHYLIP pro- erated in the present analysis, which indicate that P.
gram. Bootstrap values (100 replicates) are shown. Similar trees were monodon is related to the Hexapoda (Insecta), with
obtained from the neighbor-joining algorithm in the program MEGA
using distance matrices generated using either the Poisson correction bootstrap support ranging from 92% to 100%. The exact
or the gamma parameter (␥ ⫽ 1 or ␥ ⫽ 2). relationship of the two branchiopods both to each other
and to the Penaeus/insect clade is unclear, but in all
cases our analysis indicates that the branchiopods are
(Garcı́a-Machado et al. 1999). Taken together, these data less closely related to insects than is the malacostracan
provide further support for the proposed relationship be- P. monodon. This result was also obtained in the anal-
tween Crustacea and Insecta (Boore, Lavrov, and Brown ysis of Garcı́a-Machado et al. (1999), although their
1998), and the conclusion that the drosophilid mito- analysis included only one branchiopod, A. franciscana,
chondrial gene order represents the ancestral crustacean/ which appears to be subject to accelerated evolution
insect mitochondrial gene order (Crease 1999). (Crease 1999; see long branch lengths for Artemia in
Phylogenetic relationships among the major arthro- fig. 5).
pod groups were also investigated by sequence compar- This conclusion on the relative relationships of Ma-
isons using concatenated protein-coding sequences. The lacostraca/Branchiopoda/Insecta contradicts the results
resulting trees all gave strong support for several con- of two other molecular studies that included represen-
Complete mtDNA Sequence of the Shrimp P. monodon 871

tatives of both crustacean classes. In the first, a com- siderable horizontal gene transfer in the evolutionary
bined analysis of the 18S and 28S nuclear ribosomal process.
RNA genes gave strong support for a single crustacean/ If the phylogeny suggested by this analysis of mi-
hexapod clade and weak support for a possible closer tochondrial sequences is correct, this would have strong
relationship of the branchiopod Artemia salina to Hex- implications for the polarity of evolution of morpholog-
apoda than the malacostracan Procambarus clarkii ical characters used in the taxonomy of Crustacea and
(Friedrich and Tautz 1995). In a more wide-ranging Hexapoda. In particular, it would imply that characters
analysis using protein-coding sequence from the nuclear which are used to distinguish crustaceans from hexapods
gene encoding EF1␣, Regier and Schultz (1997, 1998) have been secondarily lost in hexapods, rather than be-
propose that Crustacea is polyphyletic, with Branchio- ing autapomorphies of Crustacea. However, as the Crus-
poda forming a single clade with Hexapoda and Cheli- tacea are notoriously diverse in body form (Schram
cerata being more closely related to this clade than 1986) and the characters that serve to uniquely define
Malacostraca. them are few, it is possible to develop plausible scenar-
The reasons for the disparity in phylogenetic trees ios by which this may have happened. In fact, evolu-
obtained with the analysis of different molecules are un- tionary scenarios which would conform with the pro-
clear. One possibility is that the heterogeneity in the posed phylogeny have been already developed on both
evolutionary processes occurring in the different arthro- morphological and molecular-developmental grounds
pod lineages leads to a misleading result. This was cer- (Averof and Akam 1995; Telford and Thomas 1995).
tainly the case in the results of Black and Roehrdanz For example, one character which has been used to
(1998), who carried out a similar analysis using concat- define Crustacea is the structure of the head. It has been
enated protein sequences which placed A. mellifera in described as having five segments, two preoral segments
the same clade as the hard ticks with 99% bootstrap bearing the first and second antennae and three postoral
support. The reason for the latter result is the extreme segments bearing, respectively, the mandibles and the
AT bias in the Apis mitochondrial genome, which has a first and second maxillae (Cisne 1982; Schram 1986).
significant influence on the resulting amino acid com- Insects, in contrast, have been described as having six
position of the encoded proteins (see fig. 4). However, segments in the head, three preoral and three postoral
as figure 4 illustrates, although this bias is evident in (Lawrence, Nielsen, and Mackerras 1991). However, the
the P. monodon mitochondrial genome, it is not as ex- nature of the most anterior portion of the crustacean
treme as in the case of Apis or, indeed, any of the other head was never clear (see Averof and Akam 1995), and
recent work indicates that both crustaceans and insects
insects. In fact, the degree to which amino acid com-
have six cephalic segments, as revealed by looking at
position is affected by codon bias appears to be no dif-
the expression of genes that specify segmental identity
ferent for P. monodon than for A. franciscana, and only
(Popadic et al. 1998; Scholtz 1998). Another commonly
slightly higher than for D. pulex. Hence, it is unlikely
cited character defining Crustacea is the presence of a
to affect the relative positions of these three species on
second antenna on the second appendage-bearing seg-
a phylogenetic tree with respect to each other or to In- ment of the head. However, it is easy to envisage sec-
secta. In addition, the proposed phylogenetic tree struc- ondary loss of such a structure in Insecta, and, indeed,
ture is also strongly supported by maximum-likelihood some insects develop small appendages on this segment
analysis. This method, although still susceptible to pro- during the embryonic stage which could be vestiges on
ducing erroneous phylogenetic trees in cases of extreme an earlier adult appendage (Tamarelle 1984).
bias in nucleotide or amino acid composition, is far less A third set of characters often deemed to be unique
sensitive to small compositional differences than are dis- to Crustacea relate to embryology and larval develop-
tance and parsimony analyses (Galtier and Gouy 1995). ment, namely, the nature of the embryonic fate map and
A more likely explanation for the different trees is the formation of a nauplius larva (larvae with three sets
that each of these molecular phylogenies is based on of appendages used for propulsion, corresponding to the
analysis of only one or two genes or, in the case of the antennules, antennae, and mandibles) or passage through
present paper, a set of genetically linked genes. There is an egg-nauplius phase (Anderson 1982; Schram 1986).
always a danger that such analyses may give misleading However, more recent studies have contested Anderson’s
results due to factors such as insufficient informative (1982) claim that crustacean egg cleavage is spiral
characters or convergent evolution. (Hertzler and Clark 1992), and the wisdom of using a
A more far-fetched, but not impossible, scenario is character as variable as fate maps in phylogenetic anal-
that some horizontal transfer of either nuclear or mito- ysis has also been questioned (Averof and Akam 1995).
chondrial sequences has occurred from insects to crus- Moreover, as larval development has clearly undergone
taceans or vice versa. As increasing amounts of data considerable diversification within Hexapoda, ranging
emerge from complete genome sequences, evidence from ametabolous insects that show no marked meta-
suggests that the evolutionary origins of different groups morphosis at all, to hemimetabolous insects with a grad-
of genes may differ. For example, Rivera et al. (1998) ual larval-to-adult transition, to the holometabolous in-
recently demonstrated that some groups of genes in eu- sects which show complete metamorphosis from larval
karyotes are derived from archaebacteria, whereas other to adult forms (Chapman 1991), it is certainly possible
groups appear to derive primarily from proteobacteria, to envisage that the nauplius larval stage has been sec-
leading to the conclusion that there may have been con- ondarily lost.
872 Wilson et al.

Other features which in the past have been used to genome of two hard ticks, which lie within the same
argue that Crustacea and Hexapoda occupy distinct evo- chelicerate family, has revealed major mitochondrial re-
lutionary lineages, namely, the origin of the mandibles arrangements in four out of five subfamilies within that
(gnathobasic vs. whole-limb) and the nature of the limbs group (Black and Roehrdanz 1998; Campbell and Bark-
(uniramous vs. bi- or polyramous) have already effec- er 1998). Thus, the various tRNA rearrangements pre-
tively been discounted as being only apparent, rather sent in the mitochondrial genomes of Anopheles, Lo-
than real, differences by a number of authors (Kukalová- custa, Apis, and Artemia all appear to be characters that
Peck 1992; Averof and Akam 1995; Telford and Thom- were derived after the initial radiation of hexapods and
as 1995; Scholtz, Mittmann, and Gerberding 1998). crustaceans.
The proposed phylogeny also has implications for In summary, the complete sequence of the mala-
the relative affinities of different crustacean classes to costracan crustacean P. monodon gives further weight
each other and to Insecta. The divergence between the to the conclusion that the mtDNA gene order observed
branchiopod clade and the malacostracan/insect clade is in D. yakuba is the ancestral crustacean/hexapod ar-
consistent with the fossil record. Although their exact rangement. Phylogenetic analysis comparing available
affinities may be disputed, fossils with substantial sim- arthropod mitochondrial genomes indicates a sister
ilarities to both Branchiopoda and Malacostraca are group relationship between Malacostraca and Insecta
known from the Cambrian (Dahl 1983; Briggs 1983), and hence throws further fat on the controversial fire of
and these classes were certainly established as indepen- arthropod phylogeny. However, a more in-depth under-
dent lineages well before the appearance of the first fos- standing of arthropod phylogeny will require the com-
sil insects in the Devonian (Kukalová-Peck 1991). pletion both of the sequence of a greater variety of ar-
Moreover, there are certain features held in common be- thropod mitochondrial genomes, most notably with the
tween Malacostracan crustaceans and insects that are not inclusion of some myriapod genomes, as well as a better
shared with ‘‘simpler crustaceans,’’ most notably the de- understanding of the processes of mitochondrial
tailed structures of the compound eye and of the seg- evolution.
mental nervous systems (Averof and Akam 1995; Tel-
ford and Thomas 1995; Osorio, Averof, and Bacon Acknowledgments
1995). The former has been alluded to as ‘‘one of the
most disconcerting problems of arthropod phylogeny’’ The authors thank Lars Jermiin for his assistance
by Tiegs and Manton (1958), who argued that Arthrop- with the analysis of codon bias in arthropod genomes
oda were polyphyletic and hence that the malacostracan and with the maximum- likelihood analysis, and Ger-
and insect eyes had evolved by convergence. hard Scholtz for discussions of crustacean phylogeny.
A question that the present paper cannot address is We also thank two anonymous reviewers for their help-
the relative position of the myriapods on the proposed ful comments on the manuscript. This is contribution
phylogenetic tree, as no complete myriapod mitochon- number 994 of the Australian Institute of Marine
drial gene sequences were publicly available at the time Science.
of writing. Traditionally, it has been assumed that Hex-
apoda evolved from a myriapodlike ancestor and that
LITERATURE CITED
Crustacea formed a sister group to this ‘‘Atelocerate
clade’’ (Kristensen 1991). However, almost all molecu- ADACHI, J., and M. HASEGAWA. 1996a. Model of amino acid
lar analyses indicate that Crustacea, rather than Myri- substitution in proteins encoded by mitochondrial DNA. J.
apoda, are sister to Insecta (Wheeler, Cartwright, and Mol. Evol. 42:459–468.
Hayashi 1993; Friedrich and Tautz 1995; Boore, Lavrov, ———. 1996b. MOLPHY version 2.3: programs for molecular
and Brown 1998), and arguments have also been put phylogenetics based on maximum likelihood. Compu. Sci.
forward on morphological grounds, taking into account Monogr. 28:1–150.
key features of Myriapoda, Hexapoda, and Crustacea, ANDERSON, D. T. 1982. Embryology. Pp. 1–41 in L. G. ABELE,
ed. Embryology, morphology and genetics. Academic
that make this scenario plausible (Averof and Akam Press, New York.
1995; Osorio, Averof, and Bacon 1995; Telford and ANDERSON, S., A. T. BANKIER, B. G. BARRELL et al. (11 co-
Thomas 1995). authors). 1981. Sequence and organization of the human
One conclusion that is clear from the present paper mitochondrial genome. Nature 290:457–470.
and the recent completion of the D. pulex mitochondrial AVEROF, M., and M. AKAM. 1995. Insect-crustacean relation-
genome sequence (Crease 1999) is that while identity of ships: insights from comparative developmental and molec-
mitochondrial gene order may be a strong indicator of ular studies. Philos. Trans. R. Soc. Lond. B Biol. Sci. 347:
phylogenetic relatedness due to the low probability of 293–303.
convergence (Boore and Brown 1998), differences in BEARD, C. B., H. MILLS, and F. H. COLLINS. 1993. The mito-
mitochondrial gene order do not necessarily reflect ma- chondrial genome of the mosquito Anopheles gambiae:
DNA sequence, genome organization, and comparisons
jor phylogenetic differences. This is illustrated in the with mitochondrial sequences of other insects. Insect Mol.
Arthropoda by the fact that two widely divergent crus- Biol. 2:103–124.
taceans and three dipteran flies all share the same mi- BLACK, W., and R. ROEHRDANZ. 1998. Mitochondrial gene or-
tochondrial gene order, while those of other dipterans der is not conserved in arthropods: prostriate and metas-
and hexapods differ in the placement of one or more triate tick mitochondrial genomes. Mol. Biol. Evol. 15:
tRNAs. Most strikingly, analysis of the mitochondrial 1772–1785.
Complete mtDNA Sequence of the Shrimp P. monodon 873

BOORE, J. L., and W. M. BROWN. 1995. Complete DNA se- FELSENSTEIN, J. 1993. PHYLIP (phylogeny inference package).
quence of the mitochondrial DNA of the annelid worm Version 3.57c. Distributed by the author, Department of Ge-
Lumbricus terrestris. Genetics 141:305–319. netics, University of Washington, Seattle.
———. 1998. Big trees from little genomes: mitochondrial FLOOK, P., C. ROWELL, and G. GELLISSEN. 1995. The sequence,
gene order as a phylogenetic tool. Curr. Opin. Genet. Dev. organization and evolution of the Locusta migratoria mi-
8:668–674. tochondrial genome. J. Mol. Evol. 41:928–941.
BOORE, J. L., T. M. COLLINS, D. STANTON, L. L. DAEHLER, FOSTER, P. G., L. S. JERMIIN, and D. A. HICKEY. 1997. Nucle-
and W. M. BROWN. 1995. Deducing the pattern of arthropod otide composition bias affects amino acid content in pro-
phylogeny from mitochondrial DNA rearrangements. Na- teins coded by animal mitochondria. J. Mol. Evol. 44:282–
ture 376:163–165. 288.
BOORE, J. L., D. V. LAVROV, and W. M. BROWN. 1998. Gene FRIEDRICH, M., and D. TAUTZ. 1995. Ribosomal DNA phylog-
translocation links insects and crustaceans. Nature 392:667– eny of the major extant arthropod classes and the evolution
668. of myriapods. Nature 376:165–167.
BOUCHON, D., C. SOUTY-GROSSET, and R. RAIMOND. 1994. GALTIER, N., and M. GOUY. 1995. Inferring phylogenies from
Mitochondrial DNA variation and markers of species iden- DNA sequences of unequal base compositions. Proc. Natl.
tity in two penaeid shrimp species: Penaeus monodon Fa- Acad. Sci. USA 92:11317–11321.
bricius and P. japonicus Bate. Aquaculture 127:131–144. GARCÍA-MACHADO, E., M. PEMPERA, N. DENNEBOUY, M. OLI-
BRIGGS, D. E. G. 1983. Affinities and early evolution of the VA-SUAREZ, J.-C. MOUNOLOU, and M. MONNEROT. 1999.
Crustacea: the evidence of the Cambrian. Pp. 1–22 in F. R. Mitochondrial genes collectively suggest the paraphyly of
SCHRAM, ed. Crustacean phylogeny. A. A. Balkema, Rot- Crustacea with respect to Insecta. J. Mol. Evol. 49:142–149.
terdam, the Netherlands. GENETICS COMPUTER GROUP. 1994. Program manual for the
BRIGGS, D. E. G., and R. A. FORTEY. 1989. The early radiation Wisconsin package. Version 8. Genetics Computer Group,
and relationships of the major arthropod groups. Science Madison, Wis.
246:241–243. HERTZLER, P. L., and W. H. CLARK. 1992. Cleavage and gas-
BRIGGS, D. E. G., R. A. FORTEY, and M. A. WILLIS. 1992. trulation in the shrimp Sicyonia ingentis: invagination is
Morphological disparity in the Cambrian. Science 256:
accompanied by oriented cell division. Development 116:
1670–1673.
127–140.
CAMPBELL, N. J. H., and S. C. BARKER. 1998. An unprece-
HILLIS, D. M., and J. J. BULL. 1993. An empirical test of boot-
dented major rearrangement in an arthropod mitochondrial
strapping as a method for assessing confidence in phylo-
genome. Mol. Biol. Evol. 15:1786–1787.
genetic analysis. Syst. Biol. 42:182–192.
CHAPMAN, R. F. 1991. General anatomy and function. Pp. 33–
HILLIS, D. M., C. MORITZ, and B. K. MABLE. 1996. Molecular
67 in C.S.I.R.O. Division of Entomology, ed. The insects
systematics. Sinauer, Sunderland, Mass.
of Australia. A textbook for students and research workers.
KADOWAKI, K.-I., K. OZAWA, S. KAZAMA, N. KUBO, and T.
Melbourne University Press, Melbourne, Australia.
AKIHAMA. 1995. Creation of an initiation codon by RNA
CISNE, J. L. 1982. Origin of the Crustacea. Pp. 65–92 in L. G.
editing in the cox1 transcript from tomato mitochondria.
ABELE, ed. The biology of Crustacea: systematics, the fossil
Curr. Genet. 28:415–422.
record and biogeography. Academic Press, New York.
KLINBUNGA, S., D. J. PENMAN, B. J. MCANDREW, and A. TAS-
CLARY, D. O., and D. R. WOLSTENHOLME. 1983. Genes for
SANAKAJON. 1999. Mitochondrial DNA diversity in three
cytochrome c oxidase subunit I, URF2, and three tRNAs in
Drosophila mitochondrial DNA. Nucleic Acids Res. 11: populations of the giant tiger shrimp, Penaeus monodon.
6859–6872. Mar. Biotechnol. 1:113–121.
———. 1985. The mitochondrial molecule of Drosophila yak- KRISTENSEN, N. P. 1991. Phylogeny of extant hexapods. Pp.
uba: nucleotide sequence, gene organisation, and genetic 125–140 in C. Division of Entomology, ed. The insects of
code. J. Mol. Evol. 22:252–271. Australia. A textbook for students and research workers.
COCKBURN, A. F., S. E. MITCHELL, and J. A. SEAWRIGHT. 1990. Melbourne University Press, Melbourne, Australia.
Cloning of the mitochondrial genome of Anopheles quad- KUKALOVÁ-PECK, J. 1991. Fossil history and the evolution of
rimaculatus. Arch. Insect Biochem. Physiol. 14:31–36. hexapod structures. Pp. 141–179 in C.S.I.R.O. Division of
CREASE, T. J. 1999. The complete sequence of the mitochon- Entomology, ed. The insects of Australia. A textbook for
drial genome of Daphnia pulex (Cladocera: Crustacea). students and research workers. Melbourne University Press,
Gene 233:89–99. Melbourne, Australia.
CROZIER, R. H., and Y. C. CROZIER. 1993. The mitochondrial ———. 1992. The ‘Uniramia’ do not exist: the ground plan
genome of the honeybee Apis mellifera: complete sequence of the Pterygota as revealed by Permian Diaphanopterodea
and genome organization. Genetics 133:97–117. from Russia (Insecta: Paleodictyopteroidea). Can. J. Zool.
DAHL, E. 1983. Malacostracan phylogeny and evolution. Pp. 70:236–255.
189–212 in F. R. SCHRAM, ed. Crustacean phylogeny. A. A. KUMAR, S., K. TAMURA, and M. NEI. 1993. MEGA: molecular
Balkema, Rotterdam, the Netherlands. evolutionary genetics analysis. Version 1.0. The Pennsyl-
DE BRUIJN, M. H. L. 1983. Drosophila melanogaster mito- vania State University, Philadelphia.
chondrial DNA, a novel organization and genetic code. Na- LAWRENCE, J. F., E. S. NIELSEN, and I. M. MACKERRAS. 1991.
ture 304:234–241. Skeletal Anatomy. Pp. 3–32 in C. Division of Entomology,
DOWLING, T. E., C. E. MORITZ, and J. D. PALMER. 1990. Nu- ed. The insects of Australia. A textbook for students and
cleic acids II: restriction site analysis. Pp. 250–317 in D. research workers. Melbourne University Press, Melbourne,
M. HILLIS, C. MORITZ, and J. D. PALMER, eds. Molecular Australia.
systematics. Sinauer, Sunderland, Mass. LEWIS, D., C. FARR, and L. KAGUNI. 1995. Drosophila melan-
FEARNLEY, I. M., and J. E. WALKER. 1987. Initiation codons ogaster mitochondrial DNA: completion of the nucleotide
in mammalian mitochondria; differences in genetic code in sequence and evolutionary comparisons. Insect Mol. Biol.
the organelle. Biochemistry 26:8247–8251. 4:263–278.
874 Wilson et al.

LOWE, T., and S. EDDY. 1997. tRNAscan-SE: a program for SCHRAM, F., and M. J. EMERSON. 1991. Arthropod pattern the-
improved detection of transfer RNA genes in genomic se- ory: a new approach to arthropod phylogeny. Mem. Qld.
quence. Nucleic Acids Res. 25:955–964. Mus. 31:1–18.
OSORIO, D., M. AVEROF, and J. P. BACON. 1995. Arthropod SNODGRASS, R. 1938. Evolution of the Annelida, Onychophora
evolution: great brains, beautiful bodies. Trends Ecol. Evol. and Arthropoda. Smithson. Misc. Coll. 97:1–159.
10:449–454. SPANOS, L., G. KOUTROUMBAS, M. KOTSYFAKIS, and C. LOUIS.
2000. The mitochondrial genome of the Mediterranean
PALUMBI, S. R., and J. A. H. BENZIE. 1991. Large mitochon-
fruitfly, Ceratitis capitata. Insect Mol. Biol. 9:139–144.
drial DNA differences between morphologically similar STATON, J. L., L. L. DAEHLER, and W. M. BROWN. 1997. Mi-
Penaeid shrimp. Mol. Mar. Biol. Biotechnol. 1:27–34. tochondrial gene arrangement of the horseshoe crab Limulus
POPADIC, A., A. ABZHANOV, D. ROUSCH, and T. KAUFMAN. polyphemus L.: conservation of major features among ar-
1998. Understanding the genetic basis of morphological thropod classes. Mol. Biol. Evol. 14:867–874.
evolution: the role of homeotic genes in the diversification SWOFFORD, D. L. 1993. PAUP: phylogenetic analysis using
of the arthropod bauplan. Int. J. Dev. Biol. 42:453–461. parsimony. Version 3.1.1. Illinois Natural History Survey,
REGIER, J. C., and J. W. SHULTZ. 1997. Molecular phylogeny Champaign.
of the major arthropod groups indicates polyphyly of crus- TAMARELLE, M. 1984. Transient rudiments of second antennae
taceans and a new hypothesis for the origin of hexapods. on the ‘‘intercalary’’ segment of embryos of Anurida mar-
Mol. Biol. Evol. 14:902–903. itima Guer. (Collembola: Arthropleona) and Hyphantria cu-
———. 1998. Molecular phylogeny of arthropods and the sig- nea Drury (Lepidoptera: Arctiidae). Int. J. Insect Morphol.
Embryol. 13:331–336.
nificance of the Cambrian ‘‘explosion’’ for molecular sys-
TELFORD, M. J., and R. H. THOMAS. 1995. Demise of the Ateo-
tematics. Am. Zool. 38:918–928. cerata? Nature 376:123–124.
RIVERA, M. C., R. JAIN, J. E. MOORE, and J. A. LAKE. 1998. TIEGS, O. W., and S. M. MANTON. 1958. The evolution of the
Genomic evidence for two functionally distinct gene clas- Arthropoda. Biol. Rev. 33:255–337.
ses. Proc. Natl. Acad. Sci. USA 95:6239–6244. VALVERDE, J. R., B. BATUECAS, C. MORATILLA, R. MARCO,
SCHOLTZ, G. 1998. Cleavage, germ band formation and head and R. GARESSE. 1994. The complete mitochondrial DNA
segmentation: the ground pattern of the Euarthropoda. Pp. sequence of the crustacean Artemia franciscana. J. Mol.
317–332 in R. A. FORTEY and R. H. THOMAS, eds. Arthro- Evol. 39:400–408.
pod phylogeny. Chapman and Hall, London. WHEELER, W., P. CARTWRIGHT, and C. HAYASHI. 1993. Ar-
SCHOLTZ, G., B. MITTMANN, and M. GERBERDING. 1998. The thropod phylogeny: a combined approach. Cladistics 9:1–
39.
pattern of Distal-less expression in the mouthparts of crus- WOLSTENHOLME, D. R. 1992. Animal mitochondrial DNA:
taceans, myriapods and insects: new evidence for a gnatho- structure and evolution. Int. Rev. Cytol. 141:173–216.
basic mandible and the common origin of Mandibulata. Int.
J. Dev. Biol. 42:801–810. B. FRANZ LANG, reviewing editor
SCHRAM, F. R. 1986. Crustacea. Oxford University Press, New
York and Oxford. Accepted February 1, 2000

Anda mungkin juga menyukai