The Mitochondrial Genome of the Sipunculid Phascolopsis gouldii Supports Its Association with Annelida Rather than Mollusca

Jeffrey L. Boore and Joseph L. Staton

Department of Biology, University of Michigan;
DOE Joint Genome Institute and Lawrence Berkeley National Laboratory, Walnut Creek, California
Belle W. Baruch Institute for Marine Biology and Coastal Research, University of South Carolina


    Abstract
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 
We determined the sequence of about half (7,470 nt) of the mitochondrial genome of the sipunculid Phascolopsis gouldii, the first representative of this phylum to be so studied. All of the 19 identified genes are transcribed from the same DNA strand. The arrangement of these genes is remarkably similar to that of the oligochaete annelid Lumbricus terrestris. Comparison of both the inferred amino acid sequences and the gene arrangements of a variety of diverse metazoan taxa reveals that the phylum Sipuncula is more closely related to Annelida than to Mollusca. This requires reinterpretation of the homology of several embryological features and of patterns of animal body plan evolution.


    Introduction
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 
Complete mitochondrial genome sequences are available for 56 invertebrate animals (Boore 1999Citation ; from Genomic Diversity link at http://www.jgi.doe.gov). With a few notable exceptions, animal mitochondrial DNAs (mtDNAs) are circular molecules, about 16 kb in size, containing 37 genes: 13 for proteins of electron transport (cytochrome oxidase subunits I–III [cox1cox3], cytochrome b [cob], NADH dehydrogenase subunits 1–6 and 4L [nad1nad6 and nad4L], ATP synthase subunits 6 and 8 [atp6 and atp8]), 2 for rRNAs (large- and small-subunit rRNAs [rrnL and rrnS, respectively]), and 22 for tRNAs (abbreviated trnX, where "X" is the one-letter code for the corresponding amino acid, with anticodon where necessary for uniquely specifying identity). In a few lineages, the arrangements of these genes vary greatly, such as in bivalve (e.g., Hoffmann, Boore, and Brown 1992Citation ) and gastropod (e.g., Yamazaki et al. 1997Citation ) mollusks. However, in other lineages, these arrangements have remained unchanged for very long periods. For example, the arrangements of the 37 mitochondrial genes of the horseshoe crab (Staton, Daehler, and Brown 1997Citation ; Lavrov, Boore, and Brown 2000Citation ) and of the fruit fly (Clary and Wolstenholme 1985Citation ) differ by the location of only one tRNA gene.

Comparisons of mitochondrial gene arrangements have been very effective for reconstructing evolutionary relationships (Smith et al. 1993Citation ; Boore et al. 1995Citation ; Boore and Brown 1998, 2000Citation ; Boore, Lavrov, and Brown 1998Citation ; Dowton 1999Citation ; Stechmann and Schlegel 1999Citation ; Kurabayashi and Ueshima 2000Citation ). Despite the occasional finding of mtDNAs with "scrambled" gene arrangements, the data set remains nearly free of homoplasy (Boore 1999, 2000Citation ). Since the rate of rearrangement varies among lineages, these gene order characters may provide resolution at differing taxonomic levels. Within rapidly changing lineages, the signal may be best at lower levels, with anciently shared arrangements being eroded. Conversely, in more slowly evolving lineages the more ancient signal may be retained, but without rearrangements to resolve more recent branching. The merits of using gene rearrangements for recovering phylogenetic patterns lie not in their infrequency, as has been suggested (e.g., Saccone et al. 1999Citation ; Le et al. 2000Citation ), but in their complexity (i.e., there are a very large number of potential arrangements) and their near irreversibility (as judged by the infrequency of identified homoplasy). Thus, the ability to resolve any particular phylogenetic relationship may depend on the rate of gene rearrangement, but the confidence in a clade that is defined by a rearrangement, as analyzed by phylogenetic methods, does not.

In addition to providing gene order characters for inferring phylogeny, this data set also allows comparison of sequences as an independent estimator of evolutionary relationships. Furthermore, mitochondrial genome comparisons serve as a model system for genome evolution, where we can examine such factors as the following: What determines whether genes are all encoded by the same DNA strand versus being distributed between the two strands? How does mutation bias influence A+T-richness, dinucleotide frequency, codon usage, strand-skew between purines and pyrimidines, and amino acid composition of proteins? How do the structures of tRNAs evolve?

We describe here the partial mitochondrial genome of Phascolopsis gouldii, the first representative of the phylum Sipuncula to be so examined. The placement of this phylum within metazoan evolution has varied since its recognition as a taxon. Lamarck (1816)Citation confused sipunculans with sea cucumbers (Holothuroidea), and they were later considered to be a derived group of annelids (Delle Chiaje 1823Citation ). In his reorganization of metazoans, de Quatrefages (1847)Citation created the Gephyrea, or "bridge group," which contained sipunculans, echiurans, sternapsid annelids, and priapulans. With their lack of segmentation and simple internal anatomies, de Quatrefages envisioned these groups as the transitional forms from "lower" to "higher" metazoan archetypes. Sipunculans were later elevated to the phylum level (Sedgwick 1898Citation ) and associated with other spiralians either as a sister taxon to Annelida on the basis of biochemical properties (Florkin 1976Citation ; Henry 1987Citation ) or as sister to mollusks based on similarities in development (Scheltema 1993, 1996Citation ).

The developmental lynchpin that supports a sipunculan/molluscan affiliation is a shared "molluscan cross" cleavage pattern in the blastomeres (Heath 1899Citation ; Gerould 1907Citation ; Baba 1951Citation ; van Dongen and Geilenkirchen 1974Citation ; Rice 1975, 1985Citation ; Verdonk and van den Biggelar 1983Citation ), which has been interpreted as a synapomorphy uniting these two groups (Scheltema 1993, 1996Citation ). Annelids and echiurans possess a different arrangement of the blastomeres, termed an "annelid cross" (Gerould 1907Citation ). As an extension of this proposed relationship, Scheltema (1996)Citation hypothesized further homology of sipunculan larval characteristics to those of larval and adult mollusks, including the ventral buccal organ of sipunculan larvae with the odontophore of mollusks, the ciliated lip below the mouth of sipunculan pelagosphera stage larvae with the foot of a molluscan pediveliger larva, and the lip glands of sipunculan larvae with the pedal glands of larval chitons and adult neomenioid aplacophorans.

We report the sequence of about half (7,470 nt) of the mtDNA of the first representative of the phylum Sipuncula, P. gouldii. We analyze genomic features in comparison with those of other animal mtDNAs and compare both gene arrangement and inferred amino acid sequences to resolve the phylogenetic placement of this phylum. These comparisons strongly support the closer relationship of Sipuncula with Annelida to the exclusion of Mollusca and other taxa.


    Materials and Methods
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 
Molecular Analysis
Live specimens of P. gouldii (Fisher 1950) were purchased from the Marine Biological Laboratory at Woods Hole, Mass. We isolated total DNA from excised retractor muscles using a CTAB method described in Collins et al. (1996)Citation . We also isolated total DNA from the upper "nuclear" fraction of the CsCl purification method for mitochondrial DNA (Dowling et al. 1996Citation ). The upper layer of that preparation contains both nuclear and nonsupercoiled (i.e., nicked) mitochondrial DNA.

Initially, primers designed to match generally conserved regions of the mtDNA were used in the PCR to amplify four short (450–710 nt) fragments from cox1 (primers LCO1490 and HCO2198; Folmer et al. 1994Citation ), cox3 (primers COIIIF and COIIIB; Boore and Brown 2000Citation ), cob (primers CytbF and CytbR; Boore and Brown 2000Citation ), and rrnL (primers 16SARL and 16SBRH; Palumbi 1996Citation ). DNA sequences obtained from these fragments were used to design oligonucleotides that were then employed in "long-PCR" (Barnes 1994Citation ) to amplify the portions of the mtDNA spanning cox1cox3, cox3cob, and cobrrnL. Reaction conditions and results, fragment purification, DNA sequence determination and assembly, and gene identifications were as in Boore and Brown (2000)Citation . All sequence was determined for both strands using synthetic oligonucleotides to "primer-walk" through the long-PCR-amplified fragments. This 7,470-nt sequence was deposited in GenBank under accession number AF374337.

Phylogenetic Analysis of Sequences and Gene Arrangements
The invertebrate mitochondrial genetic code was used to infer the amino acid sequences of the six protein-encoding genes identified in the studied portion of P. gouldii mtDNA. Those of Cob, Cox1, Cox2, and Cox3 were aligned to the homologs of 16 other, phylogenetically diverse animals (listed, along with citations, in table 1 ) using ClustalW as incorporated in MacVector 6.5 (Accelrys). The other two inferred protein sequences, Atp8 and Nad6, were judged to be too divergent to align with confidence and thus were not used in this phylogenetic analysis. The BLOSUM matrix was used to weight shared amino acids, with gap and extension penalties of 10 and 1, respectively. Amino acid alignments were then adjusted manually; these alignments can be viewed at EMBL accessions ALIGN_000119 and ALIGN_000121–ALIGN_000123.


View this table:
[in this window]
[in a new window]
 
Table 1 Complete Mitochondrial Gene Arrangements for the Taxa Considered in this Phylogenetic Analysis, with Those Matching Partial Phascolopsis gouldii Arrangements Shown in Boldface

 
Protein alignments were analyzed using three methods: (1) maximum likelihood (ML) quartet sampling with Tree-Puzzle 5.0 (Strimmer and von Haeseler 1996Citation ) using the mtREV24 model of substitution (Adachi and Hasegawa 1996Citation ), with gamma-distributed rates (eight categories) and parameters estimated from the data set and with no clock assumed; (2) equal-weighted parsimony (MP) (heuristic search of 1,000 random sequence additions); and (3) neighbor-joining (NJ) analysis of total pairwise distance of amino acids (Saitou and Nei 1987Citation ). The two latter methods were each implemented using PAUP*, version 4.0b4a (Swofford 2000Citation ). Gaps were treated as missing data. Data were also bootstrapped using the MP and NJ methods (1,000 replicates each). The single shortest tree for the heuristic parsimony analysis was used to calculate branch support values (Bremer 1994Citation ) using the program AutoDecay (Eriksson 1998Citation ) in combination with PAUP*.

Additional analyses specifically tested the effect of grouping P. gouldii with the two annelids versus with the mollusk/brachiopod group, first using a Kishino-Hasegawa test (Kishino and Hasegawa 1989Citation ) as implemented in Tree-Puzzle (http://www.tree-puzzle.de/) and second using a nonparametric Wilcoxon signed-ranks test (Templeton 1983Citation ) as implemented in PAUP*. For the partially determined mtDNA sequences of two other annelids, Galathealinum brachiosum and Helobdella robusta (Boore and Brown 2000Citation ), only the partial sequences of the cob genes are known, along with the complete sequences of cox1, cox2, and cox3. These were included in a separate analysis to verify that the phylogenetic placement of P. gouldii was not affected.

We also compared the gene arrangement of this portion of P. gouldii mtDNA with those completely determined for 13 other animals (table 1 ). These taxa were the same as those included in the sequence analysis except for three: Daphnia pulex and Ixodes hexagonus, were omitted since they had gene arrangements identical to those of Drosophila yakuba and Limulus polyphemus, respectively, and Locusta migratoria was omitted because it differed from the gene arrangement of D. yakuba by only a single tRNA position. A matrix was constructed that scored 36 characters, each as "upstream of X" or "downstream of X," where "X" refers to each of the 19 genes identified in the studied portion of P. gouldii mtDNA. Coded character states were the 5' or 3' end of the adjacent gene. This matrix of scored gene adjacencies was analyzed using parsimony criteria with the programs PAUP*, version 4.0b4a (Swofford 2000Citation ), and MacClade (Maddison and Maddison 1992Citation ).


    Results and Discussion
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 
Gene Content and Organization
This 7,470-nt portion of P. gouldii mtDNA contains six protein-encoding genes (cox1 is incomplete by only a few nucleotides at the 5' end), 11 tRNA genes, and 2 rRNA genes (rrnL is incomplete at the 3' end). Remarkably, the arrangement of these genes is very similar to that of the oligochaete annelid Lumbricus terrestris (Boore and Brown 1995Citation ; fig. 1 ). Among animal mtDNAs studied so far (see Boore [1999Citation ] and the Genomic Diversity link at http://www.jgi.doe.gov), the entire block of 11 genes spanning cox1 through cob is uniquely shared between P. gouldii and L. terrestris (table 1 ). Furthermore, the five genes in the block from trnC through rrnL are arranged identically between these two mtDNAs and also in the mtDNAs of the chiton Katharina tunicata (Boore and Brown 1994Citation ) and the brachiopod Terebratulina retusa (Stechmann and Schlegel 1999Citation ). Intervening between these two blocks of genes are trnP, trnE, and trnS2(uga). This is not similar to the arrangement of these genes in any other animal so far studied (Boore 1999Citation ). All of the genes identified in this portion of P. gouldii mtDNA are found in the same transcriptional orientation, as is true of the mtDNAs of L. terrestris (Boore and Brown 1995Citation ), the polychaete annelid Platynereis dumerilii (Boore and Brown 2000Citation ; GenBank AF178678), an ascidian (Yokobori et al. 1999Citation ), a mussel (Hoffmann, Boore, and Brown 1992Citation ), the two sampled brachiopods (Stechmann and Schlegel 1999Citation ; Noguchi et al. 2000Citation ), the four sampled nematodes (Okimoto et al. 1991, 1992Citation ; Keddie, Higazi, and Unnasch 1998Citation ), the seven sampled flatworms (Le et al. 2000Citation ; GenBank accession number AB018440), and a sea anemone (Beagley, Okimoto, and Wolstenholme 1998Citation ).



View larger version (22K):
[in this window]
[in a new window]
 
Fig. 1.—Gene map of the sequenced portion of the mtDNA of the sipunculid Phascolopsis gouldii compared with the complete map of the oligochaete annelid Lumbricus terrestris (Boore and Brown 1995Citation ), which has been graphically linearized at cox1. All genes are transcribed from left to right as depicted. Transfer RNA genes are designated by a single letter for the corresponding amino acid. There are two tRNAs for each of leucine and serine that are differentiated by numerals: L1 and L2 recognize the codons CUN and UUR, respectively, and S1 and S2 recognize the codons AGN and UCN, respectively. All other genes are designated by standard annotation and are not to scale. Lines connect homologous genes or blocks of genes for each mtDNA

 
In many mtDNAs, atp6 directly follows atp8. This is the case even for organisms that are unambiguously outgroup taxa to those considered here, including a sponge (Watkins and Beckenbach 1999Citation ) and yeast (e.g., Sekito et al. 1995Citation ). The explanation may be that the transcript of these two genes is not cleaved, such that these genes are cotranslated from a single bicistron, as has been shown for mammalian mitochondrial systems (Fearnley and Walker 1986Citation ). The gene for atp6 is not contained in the studied portion of P. gouldii mtDNA, but it clearly does not follow atp8. So far, these genes have been found separated (see Boore 1999Citation ) only for animals that are part of the proposed superphylum Eutrochozoa (Ghiselin 1988Citation ; Mollusca, Annelida, and others, but not including Arthropoda), leading to the speculation that the loss of cotranslation of these mRNAs could be a derived feature uniting this group.

Base Composition and Codon Usage
Overall, the 7,470 nt determined for P. gouldii mtDNA are 63.1% A+T, similar to the percentage found for the whole mtDNA sequence of L. terrestris (61.6%; Boore and Brown 1995Citation ). As is common among mtDNAs, CG is the least frequent dinucleotide, occurring at only 53% of expectation given the proportion of G and of C observed in this sequence.

Leucine is inferred to be the most frequent amino acid, present 243 times in these six inferred amino acid sequences, followed by four amino acids in nearly equal numbers, each about half as common as leucine: alanine (124), isoleucine (123), phenylalanine (135), and serine (125). Cysteine is the least frequently used amino acid, occurring only 14 times. These values are all similar to those for the amino acid usage of L. terrestris homologs (see table 2 ).


View this table:
[in this window]
[in a new window]
 
Table 2 Codon Usage for the Protein-Encoding Genes in the Sequenced Portion of Phascolopsis gouldii mtDNA (1,591 codons) Compared with that of the Homologous Portions of Lumbricus terrestris mtDNA

 
The DNA strand shown in figure 2 is rich in pyrimidines (C and T). One way to measure this is called skewness, an assessment of whether the G of GC pairs (GC-skew, calculated as [G - C]/[G + C]) and the A of AT pairs (AT-skew, calculated as [A - T]/[A + T]) are more commonly on the considered strand (Perna and Kocher 1995Citation ). GC-skew for the entire 7,470 nt is -0.239, that for the protein-encoding portion only is -0.265, and that for third codon positions (see below), which, presumably, are more free to change and therefore better reflect mutational bias, is a remarkable -0.754. AT-skew values are much less biased: that for the entire region is -0.080, that for the protein-encoding portion is -0.138, and that for the third codon positions is -0.028.



View larger version (57K):
[in this window]
[in a new window]
 
Fig. 2.—Abbreviated representation of the 7,470 nt determined of the mtDNA of the sipunculid Phascolopsis gouldii. Center portions of each protein- or rRNA-encoding gene have been replaced with a numeral indicating the number of omitted nucleotides. Nucleotides forming stop codons, partial or complete (see text), are marked with a caret

 
Gene Initiation and Termination
Most of the studied animal mitochondrial genomes contain one or more genes that initiate with some alternative to the standard ATG. Annelids appear as a notable exception: based on the few animals sampled, ATG is used as the initiation codon for all, or nearly all, genes (Boore and Brown 2000Citation ). Consistent with this and with the phylogenetic placement of Sipuncula advanced here (see below), each of the P. gouldii protein-encoding genes in this portion of the mtDNA appears to start with ATG (the initiator of cox1 is not present in this sequence).

Mitochondrial genes often terminate with abbreviated stop codons, where a single T or TA is presumably completed by polyadenylation to a TAA stop codon (Ojala, Montoya, and Attardi 1981Citation ). This appears to be the case for three genes in P. gouldii mtDNA. The least certain of these is for atp8, since the next nucleotide (A), otherwise part of trnY, would complete the stop codon (fig. 2 ). If cob were to extend to the first complete stop codon, it would overlap trnP by 41 nt, and if nad6 were to do so, it would overlap cob by 20 nt. In each case, homologous genes of related animals are similar in predicted amino acid sequences to the abbreviated forms in P. gouldii and have no similarity to the overlapping extensions.

All genes are in a compact arrangement, with a total of only 13 noncoding nucleotides. These are distributed in six regions of 1–4 nt each. There is a CC between trnY and trnG, but otherwise, all are A or T. No genes are inferred to overlap except for the possibility of a single nucleotide shared between atp8 and trnY (see above).

Transfer RNAs
There are 11 sequences identified with the potential for folding into tRNA-like structures (fig. 3 ). Each has an anticodon matching exactly one tRNA in L. terrestris mtDNA. Each has a seven-member amino-acyl acceptor stem (four with a single mismatch each) and a five-member anticodon stem (again, four with a single mismatch each). Two have 5 nt in the extra arm, and all others have 4 nt. All but four have A immediately preceding the anticodon arm. For all except tRNA(N) and tRNA(S2), there are three to five nucleotide pairs in both the DHU and the T{Psi}C arms. All but four tRNAs have TA immediately preceding the DHU arm. The nucleotides preceding the anticodon are YT for all tRNAs, and the nucleotide following the anticodon is A for all but tRNA(P).



View larger version (24K):
[in this window]
[in a new window]
 
Fig. 3.—DNA sequences of the 11 tRNA genes proposed for Phascolopsis gouldii mtDNA folded into the standard cloverleaf structure. Two serine-specifying tRNAs are normally encoded by animal mtDNAs, and it is the one predicted to recognize the codon UCN that is found here; this tRNA is designated S2 for consistency with our earlier work, although there is no universal convention for this. Potential base-pairing for an unusual DHU arm of tRNA(S2) that would have only 1 nt after the amino-acyl acceptor arm and 2 nt before the anticodon arm (see text) is indicated by connecting lines. Structural elements are identified for tRNA(V)

 
It is common for tRNA(S) in many mtDNAs to lack a paired DHU arm and, in some cases, to have a six-member anticodon stem (Yokogawa et al. 1991Citation ). There is potential for P. gouldii tRNA(S2) to have such an extended anticodon stem and also to have a paired DHU arm with an unusual structure, with only one nucleotide separating the amino-acyl acceptor arm from the DHU arm. Several other serine-specifying tRNAs have been identified in animal mtDNAs that have this identical pairing potential (Boore and Brown 2000Citation ).

Phylogenetic Reconstruction
Analyses based on maximum likelihood (ML), maximum parsimony (MP), and neighbor-joining of pairwise distances (NJ) of inferred amino acids were in almost complete agreement (fig. 4 ). In all cases, P. gouldii was sister to the two annelids, and the sipunculan, annelids, mollusks, and brachiopods were placed within a well-supported Eutrochozoa (Ghiselin 1988Citation ). Furthermore, there was strong support for the monophyly of several traditionally recognized groups (at least as composed by this small sampling), such as Insecta, Chelicerata, Arthropoda, and Echinodermata. Of the 1,396 aligned inferred amino acids, 673 were parsimony-informative. A heuristic search with 1,000 random stepwise additions of taxa yielded a single tree (consistency index = 0.618, retention index = 0.366, and rescaled consistency index = 0.226), 4,552 steps in length, identical to that in figure 4 except for the resolution of two trichotomies.



View larger version (27K):
[in this window]
[in a new window]
 
Fig. 4.—Phylogenetic tree based on comparisons of amino acid sequences for Cob, Cox1, Cox2, and Cox3 for 17 taxa. The tree was rooted with the deuterostome sequences, and branch lengths were assigned using Tree-Puzzle (Strimmer and von Haeseler 1996Citation ). Numbers above branches are the percentage of puzzle quartets supporting this branch, followed by bootstrap support using distance criteria, then parsimony criteria (a dash denotes <50% bootstrap support), then the parsimony branch support value (Bremer 1994Citation ). The two sets of relationships for which the methods give differing results (monophyly vs. paraphyly of the crustaceans Daphnia pulex and Artemia franciscana and the placement of Terebratulina retusa as sister to Loligo bleekeri vs. as sister to both mollusks) were subsequently collapsed to polytomies, so this is a strict consensus tree of the three phylogenetic methods employed (maximum likelihood, parsimony, and neighbor joining; see text)

 
Discrepancies among the ML, MP, and NJ analyses were in the placement of the brachiopod T. retusa and the relationship among the crustaceans D. pulex and Artemia franciscana. In MP and NJ trees, T. retusa was placed as sister to Loligo bleekeri, albeit with less than 52% bootstrap support in either case, whereas in the ML tree, T. retusa was placed as sister to a clade of the two mollusks (69% of quartets sampled). The two crustaceans were paraphyletic in the MP analysis, albeit with poor (<50%) bootstrap support, but grouped together in both the ML and NJ analyses. This association was found in 95% of the quartets in ML and with 82% of the bootstrap resamplings in NJ.

The tree shown in figure 4 was specifically compared with the alternative which repositions P. gouldii to be more closely related to mollusks than to annelids. Using ML, a Kishino-Hasegawa test (Kishino and Hasegawa 1989Citation ) demonstrated that the grouping of P. gouldii with the two annelids was significantly better (at the 5% significance level) than the grouping of P. gouldii with the brachiopod/mollusk clade. Using MP, nonparametric tests (Templeton 1983Citation ) of the single tree with P. gouldii as sister to annelids (4,552 steps) and the single most-parsimonious tree with P. gouldii sister to mollusks (4,582 steps) demonstrated that the (P. gouldii, P. dumerilii, L. terrestris) tree was significantly shorter (P < 0.0001). In case of bias from the inclusion of the ambiguously placed T. retusa among the mollusks, a second test omitting T. retusa was performed. The resultant sipunculan/annelid tree in this analysis was also significantly shorter (P < 0.0001) than (P. gouldii, L. bleekeri, K. tunicata).

Two other annelids, Galathealinum brachiosum and Helobdella robusta, were included in a separate analysis, since each lacks a complete sequence for cob (Boore and Brown 2000Citation ). Their inclusion had no effect on the phylogenetic placement of P. gouldii, and these two taxa joined P. dumerilii and L. terrestris in a topology consistent with the results of the earlier study (not shown).

Gene arrangements were also compared for phylogenetic reconstruction by scoring gene adjacencies as characters for cladistic analysis. The strict consensus of 30 equally parsimonious trees is shown in figure 5 . The unusual placement of Alligator mississippiensis as basal to the other included deuterostomes is not well supported, since it is based on the sharing of a single gene boundary (trnI, trnM) between Branchiostoma floridae and Balanoglossus carnosus in contrast to an alternative arrangement (trnI, -trnQ, trnM) in A. mississippiensis and two outgroup taxa (D. yakuba and L. polyphemus). This shared gene boundary has evidently been created by two independent translocations of trnQ from its primitive position (between trnI and trnM), leading in one case to the arrangement nad1, -trnQ, trnI, trnM, nad2 in B. carnosus and in the other to nad1, trnI, trnM, -trnQ, nad2 in B. floridae.



View larger version (20K):
[in this window]
[in a new window]
 
Fig. 5.—Phylogenetic tree based on gene arrangement comparisons using an adjacency matrix. This is the strict consensus of 30 equally parsimonious trees. The unusual placement of Alligator mississippiensis is poorly supported (see text)

 
Moving P. gouldii to make it a sister taxon to an (L. bleekeri, K. tunicata) clade or to an ((L. bleekeri, K. tunicata), T. retusa) clade increases tree lengths by seven steps in each case for this gene arrangement analysis.

Our findings are in general agreement with several published studies (Winnepenninckx, Backeljau, and De Wachter 1995Citation [parsimony-based analysis]; Giribet et al. 2000Citation [18S-only tree]; Regier and Shultz 1998) in that sipunculans are closely associated with annelids to the exclusion of mollusks. Other studies, or portions thereof, give results to the contrary. Comparisons of 18S rDNA sequences have placed Sipuncula as basal and sister to an assemblage of worms and mollusks (Field et al. 1988Citation ; Winnepenninckx, Backeljau, and De Wachter 1995Citation [neighbor-joining analysis]; Giribet et al. 2000Citation [when combined with morphological characters]) or as sister to only one of two ectoprocts analyzed (Mackey et al. 1996Citation ). The conflicts among these studies indicate that comparisons of short DNA sequences lack the necessary resolving power at this level of relationship. Furthermore, some of the associations between sipunculans and annelids found in these past studies were probably discounted based on the problematica associated with these data (Maley and Marshall 1998Citation ).


    Conclusions
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 
The results of this study constitute the first genomic-level sequence from the phylum Sipuncula and the first molecular evolutionary treatment of the group through comparison of mitochondrial genomes. The shared mitochondrial gene orders suggest a close phylogenetic affinity between sipunculan and annelid taxa, and phylogenetic analysis of their combined amino acid sequences also reflects this relationship. In short, sipunculans and annelids form a natural group within the Eutrochozoa separate from that of mollusks. While it might seem controversial from a modern evolutionary perspective, a vermiform baüplan may not represent the simplest ancestral holdover or a failure to develop elaboration beyond the simplest hydrostatic skeleton, but, rather, it may ultimately prove to be a unifying characteristic among eutrochozoan worms.

Likewise, many characteristics that have been hypothesized to link sipunculans with mollusks, e.g., developmental pattern, lack of segmentation, etc., must be reevaluated. The presence of a "molluscan cross" may be ancestral in all eutrochozoans and subsequently lost in the annelids, or possibly is a convergent pattern that has independent origins in the two phyla. Recent reports of segmental development in chitons (Jacobs et al. 2000Citation ) suggest that segmentation in some form may be present in all spiralians, so its loss in other mollusks and worm groups may not be surprising. Certainly, the detailed examination of hypothesized synapomorphic larval and adult morphology in the sipunculans and molluscan groups (Scheltema 1993, 1996Citation ) will need careful reconsideration to exclude convergence or oversight of annelid larval and adult homologs.


    Acknowledgements
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 
Thanks to Inaki Ruiz-Trillo for assistance with the program Tree-Puzzle and to Kevin Helfenbein for helpful comments on the manuscript. This work was supported by NSF grants DEB-9807100 and DEB-9303301. Part of this work was performed under the auspices of the United States Department of Energy Office of Biological and Environmental Research by Lawrence Livermore National Laboratory under contract DE-AC03-76SF00098.


    Footnotes
 
Richard Thomas, Reviewing Editor

Keywords: Phascolopsis mitochondria evolution phylogeny genome Annelida Back

Address for correspondence and reprints: Jeffrey L. Boore, DOE Joint Genome Institute and Lawrence Berkeley National Laboratory, 2800 Mitchell Drive, Walnut Creek, California 94598. jlboore{at}lbl.gov . Back


    References
 TOP
 Abstract
 Introduction
 Materials and Methods
 Results and Discussion
 Conclusions
 Acknowledgements
 References
 

    Adachi J., M. Hasegawa, 1996 Model of amino acid substitution in proteins encoded by mitochondrial DNA J. Mol. Evol 42:459-468[ISI][Medline]

    Asakawa S., H. Himeno, K. Miura, K. Watanabe, 1995 Nucleotide sequence and gene organization of the starfish Asterina pectinifera mitochondrial genome Genetics 140:1047-1060[Abstract/Free Full Text]

    Baba K., 1951 General sketch of the development of a solenogastre, Epimenia verrucosa (Nierstrasz) J. Dept. Agric. Kyusyu Imp. Univ 6:21-40

    Barnes W. M., 1994 PCR amplification of up to 35-kb DNA with high fidelity and high yield from bacteriophage templates Proc. Natl. Acad. Sci. USA 91:2216-2220[Abstract]

    Beagley C. T., R. Okimoto, D. R. Wolstenholme, 1998 The mitochondrial genome of the sea anemone Metridium senile (Cnidaria): introns, a paucity of tRNA genes, and a near-standard genetic code Genetics 148:1091-1108[Abstract/Free Full Text]

    Black W. C., R. L. Roehrdanz, 1998 Mitochondrial gene order is not conserved in arthropods: prostriate and metastriate tick mitochondrial genomes Mol. Biol. Evol 15:1772-1785[Abstract/Free Full Text]

    Boore J. L., 1999 Animal mitochondrial genomes Nucleic Acids Res 27:1767-1780[Abstract/Free Full Text]

    ———. 2000 The duplication/random loss model for gene rearrangement exemplified by mitochondrial genomes of deuterostome animals, Pp. 133–147 in D. Sankoff and J. Nadeau, eds. Computational biology series, Vol. 1. Comparative genomics. Kluwer Academic Publishers, Dordrecht, the Netherlands

    Boore J. L., W. M. Brown, 1994 Complete DNA sequence of the mitochondrial genome of the black chiton, Katharina tunicata Genetics 138:423-443[Abstract/Free Full Text]

    ———. 1995 Complete DNA sequence of the mitochondrial genome of the annelid worm, Lumbricus terrestris Genetics 141:305-319[Abstract/Free Full Text]

    ———. 1998 Big trees from little genomes: mitochondrial gene order as a phylogenetic tool Curr. Opin. Genet. Dev 8:668-674[ISI][Medline]

    ———. 2000 Mitochondrial genomes of Galathealinum, Helobdella, and Platynereis: sequence and gene arrangement comparisons indicate that Pogonophora is not a phylum and Annelida and Arthropoda are not sister taxa Mol. Biol. Evol 17:87-106[Abstract/Free Full Text]

    Boore J. L., T. M. Collins, D. Stanton, L. L. Daehler, W. M. Brown, 1995 Deducing arthropod phylogeny from mitochondrial DNA rearrangements Nature 376:163-165[ISI][Medline]

    Boore J. L., L. L. Daehler, W. M. Brown, 1999 Complete sequence, gene arrangement and genetic code of mitochondrial DNA from the cephalochordate Branchiostoma floridae ("amphioxus") Mol. Biol. Evol 16:410-418[Abstract]

    Boore J. L., D. Lavrov, W. M. Brown, 1998 Gene translocation links insects and crustaceans Nature 393:667-668

    Bremer K., 1994 Branch support and tree stability Cladistics 10:295-304[ISI]

    Cantatore P., M. Roberti, G. Rainaldi, M. N. Gadaleta, C. Saccone, 1989 The complete nucleotide sequence, gene order and genetic code of the mitochondrial genome of Paracentrotus lividus J. Biol. Chem 264:10965-10975[Abstract/Free Full Text]

    Castresana J., G. Feldmaier-Fuchs, S.-I. Yokobori, N. Satoh, S. Pääbo, 1998 The mitochondrial genome of the hemichordate Balanoglossus carnosus and the evolution of deuterostome mitochondria Genetics 150:1115-1123[Abstract/Free Full Text]

    Clary D. O., D. R. Wolstenholme, 1985 The mitochondrial DNA molecule of Drosophila yakuba: nucleotide sequence, gene organization, and genetic code J. Mol. Evol 22:252-271[ISI][Medline]

    Collins T. M., K. Frazer, A. R. Palmer, G. J. Vermeij, W. M. Brown, 1996 Evolutionary history of northern hemisphere Nucella (Gastropoda Muricidae): molecular, morphological, ecological, and paleontological evidence Evolution 50:2287-2304[ISI]

    Crease T. J., 1999 The complete sequence of the mitochondrial genome of Daphnia pulex Gene 233:89-99[ISI][Medline]

    de Quatrefages M. A., 1847 Études sur les types inférieurs de l'embranchement des Annelés. Mémoire sur l’échiure de Gaertner (Echiurus Gaertnerii NOB.) Ann. Sci. Nat. Zool. (Paris) 7:307-343

    Delle Chiaje S., 1823 Memorie sulla storia e notomina degli animali senza vertebre del Regno di Napoli. Vol. I Fratelli Fernandes, Naples, Italy

    Dowling T. E., C. Moritz, J. D. Palmer, L. H. Riesenberg, 1996 Nucleic acids III: analysis of fragments and restriction sites Pp. 249–320 in D. M. Hillis, C. Moritz, and B. K. Mable, eds. Molecular systematics. Sinauer, Sunderland, Mass

    Dowton M., 1999 Relationships among the cyclostome braconid (Hymenoptera: Braconidae) subfamilies inferred from a mitochondrial tRNA gene rearrangement Mol. Phylogenet. Evol 11:283-287[ISI][Medline]

    Eriksson T., 1998 AutoDecay. Version 4.0 Distributed by the author, Bergius Foundation, Royal Swedish Academy of Sciences, Stockholm, Sweden

    Fearnley I. M., J. E. Walker, 1986 Two overlapping genes in bovine mitochondrial DNA encode membrane components of ATP synthase EMBO J 5:2003-2008[Abstract]

    Field K. G., G. J. Olsen, D. J. Lane, S. J. Giovannoni, M. T. Ghiselin, E. C. Raff, N. R. Pace, R. A. Raff, 1988 Molecular phylogeny of the animal kingdom Science 239:748-753[ISI][Medline]

    Flook P., C. H. F. Rowell, G. Gellissen, 1995 The sequence, organization, and evolution of the Locusta migratoria mitochondrial genome J. Mol. Evol 41:928-941[ISI][Medline]

    Florkin M., 1976 Biochemical evidence for the phylogenetic relationships of the Sipuncula Pp. 95–108 in M. E. Rice and M. Todorovic, eds. Proceedings of the International Symposium on the Biology of the Sipuncula and Echiura. Vol. 2. Naucno Delo Press, Belgrade

    Folmer O., M. Black, W. Hoeh, R. Lutz, R. Vrijenhoek, 1994 DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates Mol. Mar. Biol. Biotech 3:294-299[Medline]

    Gerould J. H., 1907 The development of Phascolosoma. Studies on the embryology of the Sipunculidae. I Zool. Jahrb. Anat 23:77-162

    Ghiselin M. T., 1988 The origin of molluscs in the light of molecular evidence Pp. 66–95 in P. Harvey and L. Partridge, eds. Oxford Surveys in evolutionary biology. Oxford University Press, Oxford, England

    Giribet G., D. L. Distel, M. Polz, W. Sterrer, W. C. Wheeler, 2000 Triploblastic relationships with emphasis on the acoelomates and the position of Gnathostomulida, Cycliophora, Plathelminthes, and Chaetognatha: a combined approach of 18S rDNA sequences and morphology Syst. Biol 49:539-562[ISI][Medline]

    Heath H., 1899 The development of Ischnochiton Zool. Jahrb. Anat 12:567-656

    Henry R. P., 1987 Invertebrate red blood cell carbonic anhydrase J. Exp. Zool 242:113-116[ISI][Medline]

    Hoffmann R. J., J. L. Boore, W. M. Brown, 1992 A novel mitochondrial genome organization for the blue mussel, Mytilus edulis Genetics 131:397-412[Abstract/Free Full Text]

    Jacobs D. K., C. G. Wray, C. J. Wedeen, R. Kostriken, R. DeSalle, J. L. Staton, R. D. Gates, D. R. Lindberg, 2000 Molluscan engrailed expression, serial organization, and shell evolution Evol. Dev 2:340-347[ISI][Medline]

    Janke A., U. Arnason, 1997 The complete mitochondrial genome of Alligator mississippiensis and the separation between recent Archosauria (birds and crocodiles) Mol. Biol. Evol 14:1266-1272[Abstract]

    Keddie E. M., T. Higazi, T. R. Unnasch, 1998 The mitochondrial genome of Onchocerca volvulus: sequence, structure and phylogenetic analysis Mol. Biochem. Parasitol 95:111-127[ISI][Medline]

    Kishino H., M. Hasegawa, 1989 Evaluation of the maximum likelihood estimate of the evolutionary tree topologies from DNA sequence data and the branching order in Hominoidea J. Mol. Evol 29:170-179[ISI][Medline]

    Kurabayashi A., R. Ueshima, 2000 Complete sequence of the mitochondrial DNA of the primitive opisthobranch Pupa strigosa: systematic implications of the genome organization Mol. Biol. Evol 17:266-277[Abstract/Free Full Text]

    Lamarck J. A., 1816 Histoire naturelle des animaux sans vertèbres, Vol. 3 Radiaires, vers, crustacés, insectes. Verdière, Paris

    Lavrov D., J. L. Boore, W. M. Brown, 2000 The complete mitochondrial DNA sequence of the horseshoe crab Limulus polyphemus Mol. Biol. Evol 17:813-824[Abstract/Free Full Text]

    Le T. H., D. Blair, T. Agatsuma, et al. (14 co-authors) 2000 Phylogenies inferred from mitochondrial gene orders—a cautionary tale from the parasitic flatworms Mol. Biol. Evol 17:1123-1125[Free Full Text]

    Mackey L. Y., B. Winnepenninckx, R. De Wachter, T. Backeljau, P. Emschermann, J. R. Garey, 1996 18S rRNA suggests that Entoprocta are protostomes, unrelated to Ectoprocta J. Mol. Evol 42:552-559[ISI][Medline]

    Maddison W. P., D. R. Maddison, 1992 Analysis of phylogeny and character evolution Sinauer, Sunderland, Mass

    Maley L. E., C. R. Marshall, 1998 Evolution—the coming of age of molecular systematics Science 279:505-506[Free Full Text]

    Noguchi Y., K. Endo, F. Tajima, R. Ueshima, 2000 The mitochondrial genome of the brachiopod Laqueus rubellus Genetics 155:245-259[Abstract/Free Full Text]

    Ojala D., J. Montoya, G. Attardi, 1981 tRNA punctuation model of RNA processing in human mitochondria Nature 290:470-474[ISI][Medline]

    Okimoto R., H. M. Chamberlin, J. L. Macfarlane, D. R. Wolstenholme, 1991 Repeated sequence sets in mitochondrial DNA molecules of root knot nematodes (Meloidogyne): nucleotide sequences, genome location and potential for host race identification Nucleic Acids Res 19:1619-1626[Abstract]

    Okimoto R., J. L. Macfarlane, D. O. Clary, D. R. Wolstenholme, 1992 The mitochondrial genomes of two nematodes, Caenorhabditis elegans and Ascaris suum Genetics 130:471-498[Abstract/Free Full Text]

    Palumbi S. R., 1996 Nucleic acids II: the polymerase chain reaction Pp. 205–247 in D. M. Hillis, C. Moritz, and B. K. Mable, eds. Molecular systematics. Sinauer, Sunderland, Mass

    Perna N. T., T. D. Kocher, 1995 Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes J. Mol. Evol 41:353-358[ISI][Medline]

    Regier J. C., J. W. Shultz, 1998 Molecular phylogeny of arthropods and the significance of the Cambrian ‘explosion’ for molecular systematics Am. Zool 38:918-928[ISI]

    Rice M. E., 1975 Sipuncula Pp. 67–127 in A. Giese and J. Pearse, eds. Reproduction of marine invertebrates, Vol. 2. Entoprocts and lesser coelomates. Academic Press, New York

    ———. 1985 Sipuncula: developmental evidence for phylogenetic inference Pp. 274–296 in S. Conway Morris, J. D. George, R. Gibson, and H. M. Platt, eds. The origins and relationships of lower invertebrates. Oxford University Press, Oxford, England

    Saccone C., C. DeGiorgi, C. Gissi, G. Pesole, A. Reyes, 1999 Evolutionary genomics in Metazoa: the mitochondrial DNA as a model system Gene 238:195-209[ISI][Medline]

    Saitou N., M. Nei, 1987 The neighbor-joining method: a new method for reconstructing phylogenetic trees Mol. Biol. Evol 4:1406-425

    Sasuga J., S.-I. Yokobori, M. Kaifu, T. Ueda, K. Nishikawa, K. Watanabe, 1999 Gene contents and organization of a mitochondrial DNA segment of the squid Loligo bleekeri J. Mol. Evol 48:692-702[ISI][Medline]

    Scheltema A. H., 1993 Aplacophora as progenetic aculiferans and the coelomate origin of mollusks as the sister taxon of Sipuncula Biol. Bull 184:57-78[Abstract/Free Full Text]

    ———. 1996 Phylogenetic position of the Sipuncula, Mollusca and the progenetic Aplacophora Pp. 53–58 in J. D. Taylor, ed. Origin and evolutionary radiation of the Mollusca. Oxford University Press, Oxford, England

    Sedgwick A., 1898 Sipunculoidea (Gephrea, Achaeta) Pp. 534–539 in A student's textbook of zoology. Swan Sonnenschein, London

    Sekito T., K. Okamoto, H. Kitano, K. Yoshida, 1995 The complete mitochondrial DNA sequence of Hansenula wingei reveals new characteristics of yeast mitochondria Curr. Genet 28:39-53[ISI][Medline]

    Smith M. J., A. Arndt, S. Gorski, E. Fajber, 1993 The phylogeny of echinoderm classes based on mitochondrial gene arrangements J. Mol. Evol 36:545-554[ISI][Medline]

    Staton J. L., L. L. Daehler, W. M. Brown, 1997 Mitochondrial gene arrangement of the horseshoe crab Limulus polyphemus L.: conservation of major features among arthropod classes Mol. Biol. Evol 14:867-874[Abstract]

    Stechmann A., M. Schlegel, 1999 Analysis of the complete mitochondrial DNA sequence of the brachiopod Terebratulina retusa places Brachiopoda within the protostomes Proc. R. Soc. Lond. B Biol. Sci 260:2043-2053

    Strimmer K., A. von Haeseler, 1996 Quartet puzzling: a quartet maximum likelihood method for reconstructing tree topologies Mol. Biol. Evol 13:964-969[Free Full Text]

    Swofford D. L., 2000 PAUP*. Phylogenetic analysis using parsimony (*and other methods). Version 4 Sinauer, Sunderland, Mass

    Templeton A. R., 1983 Phylogenetic inference from restriction endonuclease cleavage site maps with particular reference to humans and apes Evolution 37:221-244[ISI]

    Valverde J., B. Batuecas, C. Moratilla, R. Marco, R. Garesse, 1994 The complete mitochondrial DNA sequence of the crustacean Artemia franciscana J. Mol. Evol 39:400-408[ISI][Medline]

    van Dongen C. A. M., W. L. M. Geilenkirchen, 1974 The development of Dentalium with special reference to the significance of the polar lobe. I, II and III. Division chronology and development of the cell pattern in Dentalium dentale (Scaphopoda) Proc. K. Nederl. Akad. Wetens. C 77:57-100

    Verdonk N. H., J. A. M. van den Biggelar, 1983 Early development and the formation of the germ layers Pp. 91–122 in N. H. Verdonk, J. A. M. van den Biggelar, and A. S. Tompa, eds. The Mollusca, Vol. 3. Development. Academic Press, New York

    Watkins R. F., A. T. Beckenbach, 1999 Partial sequence of a sponge mitochondrial genome reveals sequence similarity to Cnidaria in cytochrome oxidase subunit II and the large ribosomal RNA subunit J. Mol. Evol 48:542-554[ISI][Medline]

    Winnepenninckx B., T. Backeljau, R. De Wachter, 1995 Phylogeny of protostome worms derived from 18S rRNA sequences Mol. Biol. Evol 12:641-649[Abstract]

    Yamazaki N., R. Ueshima, J. Terrett, et al. (12 co-authors) 1997 Evolution of pulmonate gastropod mitochondrial genomes: comparisons of gene organizations of Euhadra, Cepaea and Albinaria and implications of unusual tRNA secondary structures Genetics 145:749-758[Abstract/Free Full Text]

    Yokobori S.-I., U. Takuya, G. Feldmaier-Fuchs, S. Pääbo, R. Ueshima, A. Kondow, K. Nishikawa, K. Watanabe, 1999 Complete DNA sequence of the mitochondrial genome of the ascidian Halocynthia roretzi (Chordata, Urochordata) Genetics 153:1851-1862[Abstract/Free Full Text]

    Yokogawa T., Y.-I. Watanabe, Y. Kumazawa, T. Ueda, I. Hirao, K.-I. Miura, K. Watanabe, 1991 A novel cloverleaf structure found in mammalian mitochondrial tRNAser(UCN) Nucleic Acids Res 19:6101-6105[Abstract]

Accepted for publication June 26, 2001.