<<

Molecular Phylogenetics and Evolution 54 (2010) 746–762

Contents lists available at ScienceDirect

Molecular Phylogenetics and Evolution

journal homepage: www.elsevier.com/locate/ympev

Hennig’s orphans revisited: Testing morphological hypotheses in the ‘‘” (Diptera: )

Isaac S. Winkler a,d,*, Alessandra Rung b,c,d, Sonja J. Scheffer c,d a Department of Entomology, North Carolina State University, Raleigh, NC 27695, USA b Diagnositics Branch, California Department of Food and Agriculture, Sacramento, CA 95832, USA c Systematic Entomology Laboratory, USDA, Agricultural Research Service, Beltsville, MD 20705, USA d Department of Entomology, University of Maryland, College Park, MD 20742, USA article info abstract

Article history: The acalyptrate fly superfamily Opomyzoidea, as currently recognized, is a poorly-known group of 14 Received 26 February 2009 families. The composition of group and relationships among included families have been controver- Revised 15 December 2009 sial. Furthermore, the delimitation of two opomyzoid families, and , has Accepted 21 December 2009 been unstable with respect to placement of the genera , , and Planinasus. To test the Available online 28 December 2009 monophyly of Opomyzoidea, previously proposed relationships between families, and the position of the three problematic genera, we sequenced over 3300 bp of nucleotide sequence data from the 28S ribo- Keywords: somal DNA and CAD (rudimentary) genes from 29 taxa representing all opomyzoid families, as well as 13 Acalyptrate outgroup taxa. Relationships recovered differed between analyses, and only branches supporting well- Phylogeny Higher classification established monophyletic families were recovered with high support, with a few exceptions. Opomyzoi- Rapid radiation dea and its included subgroup, Asteioinea, were found to be non-monophyletic. Stenomicra, Cyamops, and Planinasus group consistently with Aulacigastridae, contrary to recent classifications. and , two small, monogeneric families placed in separate superfamilies, were strongly sup- ported as sister groups. Ó 2009 Elsevier Inc. All rights reserved.

1. Introduction 2008). The much-studied melanogaster represents a typical acalyptrate to many biologists, but the group 1.1. General overview also includes a few large and economically important families such as (fruit flies) and (leaf-mining flies), as Willi Hennig is rightly known as the founder of modern phylo- well as obscure and poorly-known families such as Strong- genetic systematics, but his equally groundbreaking contributions ylophthalmyiidae (Shatalkin, 1993), Gobryidae (McAlpine, 1997), to Diptera systematics are less widely appreciated (Meier, 2005). and Australimyzidae (Brake and Mathis, 2007). Acalyptrate flies Hennig’s extensive insight on relationships of the ‘‘acalyptrate” are very diverse and generally poorly collected; as a result, new flies is particularly relevant today, because these relationships re- taxa are continuously being added from all zoogeographic regions main largely unresolved. In several of his major works on Diptera, (e.g., Buck, 2006; Barraclough and McAlpine, 2006). Hennig gave much attention to several obscure acalyptrate fami- Determining phylogenetic relationships among families in this lies with especially problematic relationships, which he first listed assemblage is arguably the most difficult problem in the systematics as ‘‘families with unclear affinities” (Hennig, 1958), but later provi- of Diptera (McAlpine, 1989), and a consensus classification has not sionally grouped together as the ‘‘Anthomyzoidea” (Hennig, 1971). yet been achieved. According to the ‘‘Manual of Nearctic Diptera” In the classification which is most widely used today (McAlpine, (McAlpine, 1989), is monophyletic and is divided into 1989), this grouping remains largely intact in the superfamily 10 superfamilies. Of these, , Conopoidea, Opomyzoidea, though the evidence supporting it is still weak. (sometimes including ), and are also rec- ‘‘Acalyptratae” is a probably paraphyletic group within Schizo- ognized in competing classification systems (Hennig, 1958, 1971, phora of 27,000 described in about 70 families, represent- 1973; Colless and McAlpine, 1991; Griffiths, 1972). The limits, com- ing approximately 20% of total dipteran species (Evenhuis et al., position and phylogenetic relationships of the remaining superfam- ilies have been historically more problematic. This work is directed towards testing the monophyly of the * Corresponding author. Address: Deptartment of Entomology, North Carolina State University, Campus Box 7613, Raleigh, NC 27695, USA. Fax: +1 919 515 7746. superfamily Opomyzoidea, one of the most problematic acalyprate E-mail address: [email protected] (I.S. Winkler). superfamilies, and testing previous hypotheses of relationships be-

1055-7903/$ - see front matter Ó 2009 Elsevier Inc. All rights reserved. doi:10.1016/j.ympev.2009.12.016 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 747 tween its component families. The superfamily, as currently recog- Hennig (1971, 1973) grouped the majority of opomyzoid fami- nized, contains 14 families (excluding ; see be- lies under the name ‘‘Anthomyzoidea”. It included Acartophthalmi- low) and about 120 extant genera (see Table 2 and references dae, , , Aulacigastridae, , therein). The composition of Opomyzoidea, the limits of some of , Periscelididae, , and . its included families, and its position with respect to the remaining Hennig’s classification placed two remaining opomyzoid families, acalyptratae have been strongly debated. In to advance sensu J.F. McAlpine (1989), into a separate superfamily, Agromyzoi- knowledge of this little-known and controversial superfamily, we dea ( and Agromyzidae), and left Fergusoninidae un- present in this paper the first molecular phylogenetic analysis of placed. Xenasteiidae and were described after opomyzoid families, using partial sequences of two nuclear genes, Hennig’s publications. the 28S ribosomal DNA and the protein-coding gene CAD (rudi- D.K. McAlpine’s most comprehensive classification system (Col- mentary). We focus particularly on the composition of two in- less and McAlpine, 1991) defined Opomyzoidea as containing the cluded families, Periscelididae and Aulacigastridae. We then use families Acartophthalmidae, Agromyzidae, Clusiidae, Fergusonini- our results to evaluate the phylogenetic hypothesis of J.F. McAlpine dae, Odiniidae, Opomyzidae, Xenasteiidae and (Table 1). (1989) against alternative proposals. The remaining families of Opomyzoidea sensu J.F. McAlpine (1989) were classified in the superfamily Asteioidea. This classifi- 1.2. Taxonomic history and alternative hypotheses cation partially follows results of previous studies (e.g. McAlpine, 1978, 1983), but other proposed groupings were not explicitly jus- As defined by J.F. McAlpine (1989), Opomyzoidea includes 13 tified. D.K. McAlpine (1991) also tentatively added the Afrotropical families (Table 1, Fig. 1) that are grouped into four suprafamilies: Marginidae to Opomyzoidea. Later, D.K. McAlpine (1997) Clusioinea (Clusiidae and Acartophthalmidae), Agromyzoinea reaffirmed his earlier separation of Opomyzoidea and Asteioidea (Agromyzidae, Fergusoninidae, and Odiniidae), Opomyzoinea and questioned the putative synapomorphies leading to J.F. McAl- (Opomyzidae and Anthomyzidae), and Asteioinea (Asteiidae, pine’s broader concept of Opomyzoidea. Aulacigastridae, Xenasteiidae, Teratomyzidae, Neurochaetidae, Contrasting with Hennig (1971) and Colless and McAlpine and Periscelididae). Two families named more recently have been (1991), Griffiths’ general classification system has several unique associated with Opomyzoidea after McAlpine (1989): Neminidae characteristics. Griffiths (1972) divided Schizophora into five (Freidberg, 1994; described as a of Aulacigastridae by superfamilies, Lonchaeoidea, Lauxanioidea, Drosophiloidea, Nothy- McAlpine, 1983) and Marginidae (newly described and provision- boidea and . Muscoidea was further divided into nine ally placed in Opomyzoidea by McAlpine, 1991). In addition, McAl- ‘‘prefamilies,” one of which was . Anthomyzoinea con- pine (1989; following Hennig, 1971) suggested that Paraleucopis stituted another of the muscoid prefamilies in Griffiths’ system, Malloch (currently in ) and the unplaced Neotrop- including the following families currently placed in Opomyzoidea ical genera Mallochianamyia Santos–Neto (= Gayomyia Malloch; see (sensu McAlpine, 1989): Aulacigastridae, Anthomyzidae, Asteiidae, Wheeler, 2000) and Schizostomyia Malloch probably represent a and Opomyzidae. Griffiths’ Anthomyzoinea also included Chyr- separate lineage closely related to Asteiidae. Acartophthalmidae omyidae, , Trixoscelididae, Rhinotoridae, and Sph- is now generally considered to belong in the superfamily aeroceridae, all in the superfamily of McAlpine (Brake, 2000; Buck, 2006; see also Griffiths, 1972). (1989) (the last four families were recently included in a single

Table 1 Composition of the ‘‘Opomyzoidea” according to different authors. The original name of the superfamily is given below the author’s name. The families considered to belong in the ‘‘Opomyzoidea,” according to each author, are marked with an ‘‘X”. An ‘‘” indicates that the family was not described at the time. Putative outgroups (according to McAlpine) are listed below the line. Literature references are as follows: Hennig (1971, 1973), Griffiths (1972), McAlpine (1989), and Colless and McAlpine (1991).

Family Author (Group name) Hennig (Anthomyzoidea) Griffiths (Anthomyzoinea) J.F. McAlpine (Opomyzoidea) D.K. McAlpine (Opomyzoidea) Acartophthalmidaea X (Tephritoinea) X X Agromyzidae (Agromyzoidea) (Agromyzoinea) X X Asteiidae X X X (Asteioidea) Anthomyzidae X X X (Asteioidea) Aulacigastridae Xb Xc Xb,e (Asteioidea)e Clusiidae X (Agromyzoinea) X X Fergusoninidae (Unplaced) (Unplaced) X X Neurochaetidae d X (Asteioidea) Odiniidae (Agromyzoidea) (Tephritoinea) X X Opomyzidae X X X X Periscelididae X (Nothyboidea) X (Asteioidea)b Teratomyzidae X (Nothyboidea) X (Asteioidea) Xenasteiidae XX Carnidae (Chloropoidea) (Tephritoinea) (Carnoidea) Xf Chyromyidae X X (Sphaeroceroidea) (Heleomyzoidea) Heleomyzidaeg (unplaced) X (Sphaeroceroidea) (Heleomyzoidea) Sphaeroceridaeg (unplaced) X (Sphaeroceroidea) (Heleomyzoidea)

a Moved to Carnoidea by recent authors (Brake, 2000; Buck, 2006). b Includes Cyamops, Stenomicra, and Planinasus. c Includes Cyamops and Planinasus. d Hennig (1971) suggested the Baltic Anthoclusia may represent separate lineage in Anthomyzoidea. This species was placed in the new family Neurochaetidae by McAlpine (1978). e Including Nemininae. J.F. McAlpine (1989) mentions D.K. McAlpine’s inclusion of Ningulus and Nemo, but the putative synapomorphies he discusses for Aulacigastridae are not found in these genera. f Including Australimyza. g Subsumed within the expanded family Heteromyzidae by McAlpine (2007). 748 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762

Malloch, Parascutops Mathis and Papp, Loew, and Scutops Coquillett), and Aulacigastridae sensu stricto (containing Aulacigas- ter Macquart and Curiosimusca Rung, Mathis, and Papp) are gener- ally recognized as monophyletic units supported by a handful of putative synapomorphies (see Mathis and Rung, 2004; Rung et al., 2005). The debate has largely centered on the position of Ste- nomicra Coquillett, Cyamops Melander, and Planinasus Cresson, comprising the subfamily Stenomicrinae (Mathis and Papp, 1992; Mathis, in press). This group also now includes two newly de- scribed taxa closely related to Cyamops and Stenomicra, the fossil Procyamops succini Hoffeins and Rung from Dominican amber (Hoffeins and Rung, 2005) and the extant Stenocyamops Papp from Thailand (Papp et al., 2006). Cyamops had been previously placed in either Anthomyzidae or Periscelididae. Planinasus was originally described in and later transferred to Perisce- lididae. The genus Stenomicra has a particularly convoluted classi- fication history, having been placed into six different families since its description in the early twentieth century or classified as the type genus of a separate family (Papp, 1984; Khoo and Sabrosky, 1989). Prior to 1971 it was often included in Anthomyzidae. Hen- nig (1958, 1969, 1971), after a detailed discussion of these genera, was the first to include Cyamops, Planinasus, and later Stenomicra in Aulacigastridae, a placement followed by Griffiths (1972) and J.F. McAlpine (1989), the former with the exclusion of Stenomicra (pro- visionally retained in Anthomyzidae). Later, D.K. McAlpine (1978, 1983) transferred all three genera to the Periscelididae on the basis of similar antennal structure, a placement that has been followed by recent authors (Grimaldi and Mathis, 1993; Baptista and Mathis, 1994; Mathis and Papp, 1998; Mathis, in press).

1.3. Biology and diversity of Opomyzoidea

Fig. 1. Phylogeny and classification of the Opomyzoidea according to J.F. McAlpine Although a few opomyzoids, such as the Agromyzidae, are (1989). McAlpine’s Aulacigastridae includes Cyamops, Stenomicra, and Planinasus, widespread and abundant, most opomyzoid families are generally and presumably also Ningulus and Nemo (Nemininae of D.K. McAlpine, 1983), though the synapomorphies he discusses for Aulacigastridae are not present in the rarely collected and poorly-known, being among the most obscure latter two genera. groups of Diptera. In fact, five of the fifteen families putatively belonging to Opomyzoidea were undescribed and mostly unknown prior to 1970. For example, the diminutive Xenasteiidae were inde- family ‘‘Heteromyzidae” by McAlpine, 2007). Griffiths classified the pendently described from Mediterranean and Pacific seashores less families Agromyzidae and Clusiidae separately in the prefamily than three decades ago (Hardy, 1980; Papp, 1980) and are still lar- Agromyzoinea. A particularly innovative aspect of Griffiths’ work gely unknown. Neminidae, dubbed ‘‘nobody flies” by their discov- is the classification of Periscelididae and Teratomyzidae in the erer (Latin nemo = ‘‘nobody”; McAlpine, 1983, 1998), are now superfamily Nothyboidea, with the families and Nothybi- known from southern Africa, Australia, and New Guinea, but have dae (currently classified in the superfamily ; J.F. McAl- been very rarely collected, and nothing is known about their biol- pine, 1989; D.K. McAlphine, 1997), instead of in Anthomyzoinea. ogy. Although easily recognized by the laterally elongate head pro- Griffiths’ classification has been criticized on the grounds that it duced into eyestalks (Mathis and Rung, 2004), the bizarre emphasizes a single, highly variable character system (the male periscelidid genus Diopsosoma, with only one included species, D. postabdomen; see e.g. McAlpine, 1978), used inadequate primum Malloch, is still known only from the type series collected sampling, and some relationships he proposed are at odds with in Peru in 1931 (Mathis and Rung, 2004). Some opomyzoid taxa are those currently accepted. However, Griffiths’ work remains influ- restricted to very specific microhabitats (such as sap fluxes of ential, and evidence for many competing hypotheses is scant. wounded trees or phytotelmata on specific plant species) and are J.F. McAlpine (1989) summarized previous work on higher clas- thus infrequently encountered. Furthermore, many opomyzoid sification and phylogeny of cyclorrhaphan (‘‘higher”) flies, and pre- families exhibit limited diversity and distribution, especially in sented a fully resolved family-level phylogenetic hypothesis for the northern temperate regions (Table 2). Significant undescribed spe- Acalyptratae. His work was the first to propose Opomyzoidea as a cies diversity probably exists in many opomyzoid families. For monophyletic group including the taxa (except Chyromyidae) example, about twenty species of the widespread genus Stenomicra placed by Hennig (1971) in Anthomyzoidea and Agromyzoidea. (Periscelididae) have been described, including a few in north tem- We used J.F. McAlpine’s (1989) phylogeny in this study as the ma- perate regions, but over 100 undescribed tropical species are rep- jor source of hypotheses to be tested. resented in museum collections. Contrasting with the lack of robust resolution of interfamilial Because several opomyzoid families are either rare or simply relationships, the composition of families within Opomyzoidea is poorly collected, knowledge about their behavior and biology are now considered relatively stable, with the exception of two fami- fairly incomplete (Table 2). Phytophagy is known to occur in five lies: Aulacigastridae and Periscelididae. Genera have been trans- families: Agromyzidae (mostly leaf-miners on many different ferred between these two families and sometimes transferred to ) (Spencer, 1990; Scheffer et al., 2007), Anthomyzidae and seemingly distantly-related groups. Periscelididae sensu stricto Opomyzidae (larvae feeding in the stems of grasses) (Vockeroth, (containing Diopsosoma Malloch, Marbenia Malloch, Neoscutops 1987; Brunel, 1998; Rohácˇek, 1998a,b), Asteiidae (species of Asteia I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 749

Table 2 Biology, diversity, and distribution of families assigned to Opomyzoidea by McAlpine (1989) and subsequent authors. Species diversity estimates were obtained from a working version of the Biosystematic Database of World Diptera (Evenhuis et al., 2008), in addition to references listed below.

Family Diversity Distribution Known larval biologies References Acartophthalmidae 1 gen., 4 spp. 3 holarctic, 1 Australian Rotting fungi and carrion in damp woodlands Papp and Ozerov (1998) Agromyzidae 28 gen., ,800 spp. Widespread Phytophagous in many plants (leaf-miners, Spencer (1990), Scheffer et al. (2007) stem-borers, etc.) Anthomyzidae 19 gen., 94 spp. Widespread Mostly phytophagous or saprophagous in Rohácˇek, (1998a,b, 2006) (predominantly holarctic) graminoids, larvae of 2 spp. in fungi Asteiidae 10 gen., 135 spp. Widespread Fungivorous (Leiomyza), under bark of dead Murphy (1990), Papp (1998b), trees (Phlebosotera), saprophagous or Freidberg (2002); Freidberg, personal phytophagous in plants communication Aulacigastridae sensu 2 gen., ca 58 spp. Widespread (except Sap fluxes of decidous trees, phytotelmata of Papp (1998a), Rung et al. (2005); stricto (most undescribed) Australia and Oceania) bromeliads Rung and Mathis, unpublished Clusiidae 14 gen., 520 spp. Widespread Rotting fungi in damp wood Lonsdale and Marshall (2006, 2007, (some currently in 2008); Lonsdale, personal litt.) communication Fergusoninidae 1 gen., 31 + spp. Mostly Australia, also N. Gallers on in association with Scheffer et al. (2004), Taylor et al. Guinea, N. Zealand, (2004, 2005) tropical Asia Marginidae 1 gen., 2 spp. Southern Africa, Unknown McAlpine (1991) Madagascar Neminidae 3 gen., 14 spp. Australia, southern Africa Unknown (adults on tree trunks) McAlpine (1983), Freidberg (1994) Neurochaetidae 4 gen., 20 spp. Australia, southern Africa, Watery medium inside Alocasia inflorescence; McAlpine (1978, 1988a,b, 1993) tropical Asia probably phytotelmata of other plants Odiniidae 11 gen., 63 spp. Worldwide Saprophages and predators in galleries of Gaimari and Mathis (in press) wood-boring Opomyzidae 4 gen., 59 spp. Mostly holarctic, also in Phytophagous (in grass stems) Vockeroth (1987), Brunel (1998) Africa Periscelididae sensu 6 gen., 27 spp. Widespread (Periscelis), Sap fluxes of decidous trees, wood borer Mathis and Papp (1998), Mathis (in stricto others Neotropical tunnels press) Stenomicrinae 4 genera, 52 spp. (+ Widespread, mostly Plant phytotelmata (Stenomicra) Fish (1983), Baptista and Mathis (Aulacigastridae or many undescribed) tropics (1994), Papp et al. (2006) Periscelididae) Teratomyzidae 7 gen., 35 + spp. Australia, N. Zealand, E. Associated with ferns; larvae feed externally McAlpine and de Keyzer (1994) (only 8 described) Asia, S. America on microflora? Xenasteiidae 1 gen., 13 spp. Mediterranean, Oceania Unknown (vegetation near seashore) Hardy (1980), Papp (1980, 1998c)

Meigen reared as primary phytophages from Chenopodiaceae, 2. Materials and methods Asteraceae, and Acanthaceae; Murphy, 1990; Freidberg, personal communication) and Fergusoninidae (gallers on Myrtaceae) (Cur- 2.1. Taxon sampling rie, 1937; Taylor et al., 2004, 2005). Fungal associations are docu- mented for Acartophthalmidae (Papp and Ozerov, 1998), some In accordance with our stated goals, the set of exemplar spe- Asteiidae (Leiomyza Macquart; Papp, 1998b) and two species of cies selected (Appendix A) comprises several elements. First, a Anthomyzidae (Fungomyza Rohácˇek; Rohácˇek, 2006). Larvae and putative ingroup sample was constructed that included 29 adults of temperate species of Periscelididae sensu stricto (Mathis opomyzoid taxa distributed in all 15 putative component families. and Papp, 1998) and Aulacigastridae (Papp, 1998a), as well as some Outgroups were chosen initially on the basis of McAlpine’s (1989) adult Odiniidae (Gaimari and Mathis, in press) have been collected hypothesis of a sister-group relationship between Opomyzoidea from sap fluxes of deciduous trees. Odiniid flies display mixed and Carnoidea. Five genera of the latter, representing the families feeding strategies. Some species apparently can feed and fully de- Carnidae, Australimyzidae, , and were velop solely on frass and fungi within galleries of wood-boring - sampled. tles, lepidopterans and flies, but they sometimes attack and To this list of putative outgroups were added exemplars in- predate upon its residents. Some species have also been reared cluded in other acalyptrate superfamilies by McAlpine (1989), from the egg masses of mealybugs associated with tree bark (Gai- which have been postulated by other authors to be related to var- mari and Mathis, in press). Aquatic or semi-aquatic larvae are ious ‘‘opomyzoid” families. These are as follows: (1) Representa- found within three families: Aulacigastridae, Periscelididae and tives of Chyromyidae (two genera), placed in Sphaeroceroidea by Neurochaetidae. Undescribed New World species of Aulacigaster McAlpine (1989), were included to test Hennig’s (1971) and Grif- (Rung and Mathis, unpublished) have fully aquatic larvae, develop- fiths’ (1972) suggestion that the family instead lies close to Aulac- ing in the phytotelmata of bromeliads. The same may be true for igastridae. (2) A single representative of was added several undescribed species of Stenomicra (Periscelididae) collected to test Griffiths’ (1972) placement of this and related taxa (‘‘Het- from these plants (see Fish, 1983). Two other genera of Periscelid- eromyzidae”; see McAlpine, 2007) in Anthomyzoinea. (3) Repre- idae, Cyamops and Planinasus, have been collected near streams, sentatives of Psilidae (two genera) were included to test Griffiths’ bogs, and marshes, and larvae may be aquatic or semi-aquatic. Lar- (1972) hypothesis that the families Teratomyzidae and Periscelid- vae of Neurochaeta inversa McAlpine (Neurochaetidae) develop idae sensu stricto are more closely related to this family than to within a watery liquid between developing fruitlets of Alocasia other opomyzoid families. (4) We also included one species of (Schott) (McAlpine, 1978). Other species are presumed to similarly Tephritidae (superfamily ) to test Griffiths’ (1972) ex- develop in phytotelmata in other plants on which the adults have panded concept of this superfamily, which included Odiniidae and been collected, including Pandanus Parkinson, Ravenala Adans., Zin- Acartophthalmidae, as well as members of McAlpine’s ‘‘Carnoidea”. giber Boehm., Musa L., Strelitzia Aiton, and Xanthorrhoea Smith Finally, sequences of Musca domestica L. (Calyptratae: Muscoidea: (McAlpine, 1988b; Britton, 2009). ) and nasica Say, belonging to the non-schizopho- 750 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 ran family Syrphidae, were obtained from Genbank and included to 28H and rc28D/28Z for the 28S gene (Wiegmann et al., 2000), allow unambiguous rooting of the tree. and 54F/405R for CAD (Moulton and Wiegmann, 2004). For 28S, a standard touchdown PCR protocol was used: 94 °C for 4 min fol- 2.2. Sequence data collection lowed by 30 cycles of 94 °C for 1 min, 48–56 °C for 1 min, and 72 °C for 1 min 30 s with a final extension at 72 °C for 4 min. The same Adults representing the various species included (see Appendix program was used for CAD, with the addition of five extra cycles A) were collected and preserved in 70–100% ethanol for various before the final extension to obtain a higher yield of product. PCR amounts of time. In the lab, specimens were transferred to vials product was purified using the QIAquick PCR purification kits (QIA- containing 100% ethanol and stored in a 80 °C freezer until DNA GEN). Sequencing reactions were carried out with PE Applied Bio- extraction. Two rare taxa ( clausa McAlpine and Nemula lon- systems Big Dye Terminator sequencing kits (Perkin-Elmer Applied garista Freidberg) were sequenced from recently collected pinned Biosystems, Foster City, CA). Besides the standard 28S and CAD specimens and DNA extracted by soaking the entire specimen in primers (Wiegmann et al., 2000; Moulton and Wiegmann, 2004), a buffer/proteinase solution for 36–48 h, followed by removal of a few additional primers were used for sequencing portions of the specimen and adherence to standard protocol using the 28S in some taxa (Table 3). Sequence data were obtained by ana- DNeasy Blood and Tissue Kit (QIAGEN, Valencia, CA). lyzing samples on ABI 377 and ABI 3130 automated sequencers. DNA from remaining specimens was extracted from individual Sequences from both strands were assembled and edited as fly specimens following standard protocols in the DNEasy Blood necessary using Sequencher (Gene Codes Corp., Ann Arbor, Michi- and Tissue Kit. Before DNA extraction, the postabdomen and wings gan, U.S.A.). Multiple alignment of 28S was carried out with Muscle were preserved in microvials containing glycerin. These vials have v.3.6 (Edgar, 2004) using default parameters. The program Gblocks been kept as vouchers and are deposited at the National Museum (Castresana, 2000, 2002) was then used to identify regions of of Natural History (Smithsonian Institution, USA) collection. ambiguous alignment to exclude. Several different parameter com- Polymerase chain reaction (PCR) was carried out using TaKaRa binations were tried in Gblocks, but the final data set used the ExTaq (TaKaRa, U.S.A.) following the manufacturer’s protocol with parameters 27 (minimum number of sequences for a conserved 1–2 ll of DNA extract in a total of 50 ll solution. A large position, out of 42 total), 39 (minimum number of sequences for (2700 bp) region of 28S ribosomal DNA was amplified in 2–3 a flanking position), 5 (maximum number of contiguous noncon- overlapping pieces and sequenced for all taxa, and a small served positions), 3 (minimum length of a block), and half (allowed (680 bp) region of CAD (rudimentary) representing the initial 5’ gap positions). The set of included characters was adjusted slightly portion was obtained for 29 of the 42 taxa. PCR primers used were by hand to account for incongruous handling of positions with standard primer pairs used for other Diptera. These were rc28C/ gaps. Nucleotide sequences for CAD were aligned using Clustal X (Thompson et al., 1994). The portion of CAD used excluded a near- by intron, and very minimal length variation was observed. All se- Table 3 quences are deposited in Genbank under accession numbers Sequencing primers designed for this study. Position numbers indicate the annealing GU299217–GU299280. site of the 50 end to the ribosomal RNA sequence published by Tautz et al. (1988). Primer direction is indicated by an F (forward) or R (reverse) in the primer name. 2.3. Data analysis Primer Sequence Position 28P1F AAC AAC TCA CCT GCC GAA GCA ACT AG 4747 Data sets were first examined for base composition bias in 28P2F GTG AAC AGT GGT TGA TCA CGA G 5005 PAUP* 4.0b10 (Swofford, 2001). Uncorrected (P) distances and cor- 28P2R CTC GTG ATC AAC CAC TGT TCA C 5027 rected (ML, GTR + I + G) distances were also calculated in PAUP* for 28JR GGT YGT TCC TGT TRC CAG GAT GAG 5200 all pairwise comparisons and plotted to examine levels of diver-

Fig. 2. Uncorrected pairwise (P) distances plotted against corrected (ML, GTR + I + G) distances for each of the three gene partions (28S, CAD positions 1 and 2, CAD position 3). Third positions of CAD show very high divergence levels and evidence of saturation at deep (interfamilial) comparisons. I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 751

Table 4 Table 5 Base frequecies observed for the gene partitions examined in this study. 28S was Results of topology tests in CONSEL (Shimodaira and Hasegawa, 2002). Results analyzed both with the entire included sequence (i.e. excluding hypervariable significant at the 0.05 level are marked by , and at the 0.10 level by . regions), and with only those sites which showed variation (also excluding AU = approximately unbiased test; SH = Shimodaira–Hasegawa test. hypervariable regions). Base frequencies are exceptionally variable in third codon positions of CAD, which were excluded from further analyses. Hypothesized clade ln L AU SH Unconstrained phylogeny 18,239.3 N/A N/A Taxon Proportion G + C Opomyzoidea (excl. Margo) 18,322.0 <0.001** 0.012** 28S (all excluding 28S (included CAD pos. CAD Above—Xenasteiidae 18,291.9 0.047** 0.065* ypervariable) variable sites) 1&2 pos. 3 Above—Clusioinea 18,287.8 0.051* 0.093* Above—Opomyzoinea 18,257.1 0.281 0.544 Rhingia 0.428 0.407 0.463 0.460 Above—Periscelis 18,247.9 0.402 0.688 Musca 0.434 0.434 0.442 0.476 Asteioinea 18,305.8 <0.001** 0.005** Mumetopia 0.433 0.434 — — Above—Xenasteiidae 18,272.3 0.075* 0.118 0.433 0.434 0.449 0.422 Above—Periscelis 18,263.3 0.154 0.243 0.435 0.438 — — Asteioinea grp. 1 18,271.0 0.117 0.194 Xenasteia 0.430 0.416 — — Above—Neminidae 18,255.7 0.243 0.447 Australimyza 0.430 0.417 0.486 0.467 Above—Neurochaetidae 18,249.2 0.523 0.620 Teratomyza 0.428 0.410 0.461 0.342 Above—Aulacigaster (= Periscelididae s. lat.) 18,257.3 0.128 0.405 Neurochaeta 0.429 0.413 0.438 0.235 Agromyzoinea 18,247.3 0.363 0.375 Leiomyza 0.422 0.392 0.464 0.586 Opomyzoinea 18,257.9 0.325 0.302 Asteia 0.429 0.411 0.462 0.472 Clusioinea 18,249.2 0.346 0.334 0.423 0.385 0.443 0.293 Periscelis 0.417 0.366 0.472 0.398 0.428 0.409 0.448 0.381 0.429 0.411 0.426 0.333 ations at which point the convergence statistic (standard deviation Margo 0.424 0.389 — — 0.430 0.420 0.464 0.482 of split frequencies between independent runs) was approximately 0.429 0.416 0.465 0.592 0.04. 0.431 0.420 0.460 0.565 Alternative topologies representing a priori hypotheses of rela- Aphaniosoma 0.422 0.399 — — tionships were tested to see if they differed significantly from the Gymnochiromyia 0.431 0.420 0.490 0.553 Nemula 0.425 0.395 — — ML tree based on the combined data. For some hypotheses of Meoneura 0.432 0.437 0.455 0.414 monophyly (e.g. Opomyzoidea), additional hypotheses represent- sp. 0.424 0.388 0.451 0.450 ing sequential modification of the original hypothesis were also F. purcelli 0.422 0.389 0.457 0.317 tested (i.e. subtracting families; see table 5). Hypothetical clades 0.428 0.414 0.434 0.264 tested were Opomyzoidea exclusive of Marginidae, Asteioinea, Neoalticomerus 0.430 0.418 — — 0.425 0.394 0.461 0.450 subgroup 1 of Asteioinea (see Fig. 1), Agromyzoinea, Anthomyzoi- 0.416 0.354 — — nea, Clusioinea, and Periscelididae sensu lato (including Stenomicra, 0.420 0.372 0.438 0.296 Cyamops, and Planinasus). To accomplish this, ML analyses were 0.422 0.389 0.470 0.566 first repeated in Garli (20 runs) with the hypothesized group con- 0.418 0.363 0.477 0.435 0.436 0.441 0.452 0.265 strained as monophyletic. Site likelihoods were obtained sepa- Chaetochlorops 0.431 0.419 0.424 0.140 rately for these topologies and for the unconstrained ML 0.425 0.393 0.428 0.246 topology in PAUP* (Swofford, 2001) using the model parameters Planinasus 0.424 0.388 — — and branch lengths estimated by Garli for each. These site likeli- Stenomicra 0.425 0.395 0.360 0.094 hoods served as input for the program CONSEL (Shimodaira and Cyamops sp. 0.425 0.399 — — C. nebulosus 0.423 0.385 — — Hasegawa, 2001), which performed the Shimodaira–Hasegawa Aulacigaster ecuadoriensis 0.420 0.378 0.421 0.274 (SH) test (Shimodaira and Hasegawa, 1999) and the Approximately A. leucopeza 0.422 0.378 0.430 0.238 Unbiased (AU) test (Shimodaira, 2002). P-Values resulting from A. neoleucopeza 0.423 0.386 0.424 0.202 these tests (Table 5) represent the probability that a single con- Mean 0.426 0.403 0.447 0.367 strained phylogeny represents the true optimal tree given the pro- vided set of topologies and bootstrapped data sets generated by the program. Topologies representing related, nested hypotheses were tested simultaneously; testing these independently changed the P- gence and possible signal saturation. Based on these results, third values slightly, but had no effect on the final inferences. codon positions of CAD, which were highly divergent and extre- mely biased (Fig. 2, Table 4), were excluded from further analyses. The final data set included 2889 bp (2433 of 28S and 456 of CAD) of 3. Results which 688 were parismony informative (367 of 28S and 321 of CAD). Analyses were then performed on the separate and com- Although containing some common elements, trees from each bined data sets using maximum parsimony (MP) in PAUP* with analysis and gene partition differed substantially from each other. TBR swapping on 100 random addition trees. The maximum likeli- Also, exploratory analyses with different alignments, inclusion hood (ML) criterion was implemented with a GTR + I + G model in sets, and taxon sampling resulted in different topologies, though Garli 0.951 (Zwickl, 2006). One hundred separate runs were per- sometimes only slightly so. Each of the 11 families represented formed with each data set, as this was found to be necessary due by two or more species (except for Periscelididae sensu lato) were to differences in topology (with small differences in likelihood generally recovered as monophyletic, except for Psilidae, which score) obtained in successive runs. Bootstrap analyses (500 reps.) was recovered only by the combined ML and BI analyses. Opomy- were conducted using both MP and ML criteria. Next, the combined zoidea and the four included suprafamilies were not monophyletic data set was subjected to Bayesian inference (BI) using Mr. Bayes in any of the analyses. version 3.1.2p (Ronquist and Huelsenbeck, 2003) with two inde- Divergence levels for 28S and for first and second positions of pendent runs of four chains each run on eight processors. For this CAD were low to moderate (P distance < 0.25; Fig. 2) and the plot analysis, the genes were modeled separately, each using the of P distance vs. ML distance showed little evidence of saturation. GTR + I + G model. The analysis was extended to 10 million gener- Third positions of CAD were, however, highly divergent (P distance 752 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 mostly > 0.4) and the plot showed strong evidence of saturation different tree (Fig. 4). Bootstrapping showed that, with few excep- past the most recent divergences (Fig. 2). tions, only relationships uniting families were strongly supported Base frequencies at third positions of CAD were exceptionally by either of these analyses. Fig. 5 shows the Bayesian majority rule variable (9–59% G + C; see Table 4), but frequencies were not sig- consensus tree, which showed levels of support generally compa- nificantly variable in the included portions of both genes, as calcu- rable to the ML analysis, but with moderate to strong support for lated in PAUP* (ignoring phylogenetic correlation). However, G + C three additional relationships recovered with weak support in content does vary in slight but potentially consequential ways the ML analysis (see below). Although Bayesian values may be sub- when only the included portions of 28S and CAD are considered. ject to artificial inflation (Suzuki et al., 2002), this difference may For example, the families Anthomyzidae, Sphaeroceridae, Musci- also reflect the desirable effect of modeling genes separately. dae, and Carnidae share a slightly elevated (3% higher) G + C con- Although 28S and CAD had approximately the same number of tent in 28S compared to the average value, and these families parsimony informative sites, CAD performed much more poorly as grouped together in the 28S ML analysis (Fig. 5). measured by the number of families recovered in both the MP and Parsimony analysis of the combined data set resulted in two ML individual analyses (four of eight for CAD and 10 of 11 for 28S). most parsimonious trees, the consensus of which (Fig. 3) was fully Because of this, the CAD only analyses were not further considered. resolved except for a single node. ML analysis resulted in a much The 28S MP analysis resulted in 40 MP trees, the consensus of

Fig. 3. Strict consensus cladogram of the two most parsimonious trees obtained from analysis of the combined dataset (length 2928, CI = 0.3914, RI = 0.3953). Bootstrap values (500 reps.) are noted if greater than 50%. Taxa included by McAlpine (1989) in Opomyzoidea are in bold type. I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 753

Fig. 4. Maximum likelihood phylogeny from the combined 28S and CAD data set (ln L = 18,239.3). Bootstrap values (500 reps.) are noted if greater than 50%. Taxa included by McAlpine (1989) in Opomyzoidea are in bold type. which (not shown) was largely compatible with the 28S ML tree was found to be sister group to Aulacigaster in most trees for (Fig. 6) and Bayesian consensus tree (Fig. 5). the 28S ML analysis (90%) and combined ML analysis (80%). A consistent result, found in all analyses of the 28S and com- Within this clade (Aulacigastridae s.lat.), Cyamops and Aulacigas- bined data sets, was the placement of Cyamops, Stenomicra, and ter grouped together with low support, and Planinasus and Ste- Planinasus in a clade with Aulacigaster, as suggested by Hennig nomicra were recovered as sister taxa with moderate to strong (1958, 1969, 1971). This clade received strong support (100% support. Xenasteia Hardy and Australimyza Harrison grouped to- PP) in the combined BI analysis, though support was weak in gether consistently, with strong support values comparable to the MP and ML bootstrap analyses. Upon further inspection of those uniting members of the same family. A sister-group rela- the individual trees for each bootstrap replicate (not shown), tionship of Fergusoninidae and Asteiidae was observed in all however, one or more of the three genera mentioned above 28S and combined analyses. This relationship was moderately 754 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762

Fig. 5. Majority rule consensus tree resulting from Bayesian analysis (107 generations) of the combined data set, with posterior probabilities labeled above branches. Taxa included by McAlpine (1989) in Opomyzoidea are in bold type. supported (0.87 PP) in the BI analysis, but poorly supported not show close relationship with the families in which they were (<50%) in remaining analyses. originally included (Clusiidae, Aulacigastridae, and Agromyzidae, A number of previously hypothesized sister-group relationships respectively), despite the retention of these hypothesized relation- in ‘‘Opomyzoidea” (McAlpine, 1989) were never recovered by our ships by later authors after they were given family status. Finally, data, though we cannot statistically disprove monophyly of these the position of Periscelis was variable, but it was never placed near putative clades. For example, Opomyzidae and Anthomyzidae were Aulacigaster nor did it ever group with Cyamops, Stenomicra,orPla- distant from each other on every observed topology. Three small ninasus. It is notable also that the Carnoidea of McAlpine and later families, Acartophthalmidae, Neminidae, and Fergusoninidae, did authors (Brake, 2000; Buck, 2006) was not supported, as Carnidae, I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 755

Fig. 6. Maximum likelihood phylogeny from analysis of the 28S data set (ln L = 12,508.7). Taxa included by McAlpine (1989) in Opomyzoidea are in bold type. Bootstrap values (500 reps.) are noted if greater than 50%, for both ML (above branches) and MP (below branches) analyses. Nodes also recovered in the 28S parsimony analysis, but supported by less than 50% bootstrap values are marked by an asterisk .

Australimyzidae and Acartophthalmidae never appeared near (a = 0.10), two additional groups were rejected: Opomyzoidea Milichiidae + Chloropidae. (excluding Xenasteia and Clusioinea), and Asteioinea (excluding The topology tests in CONSEL (Table 5) resulted in rejection of Xenasteia). The SH test is known to be overly conservative, espe- monophyly for several hypothesized clades. At the conservative le- cially when multiple topologies are compared (Shimodaira and vel of a = 0.05, monophyly of only the Opomyzoidea exclusive of Hasegawa, 1999), and rejected only Opomyzoidea (including - Marginidae (including or excluding Xenasteia) and Asteioinea steia), Asteioinea (including Xenasteia), and (with a = 0.10) Opomy- (including Xenasteia) could be rejected using the Approximately zoidea (excluding Xenasteia) and Opomyzoidea (excluding Unbiased test. However, with a less conservative criterion Xenasteia and Clusioinea). 756 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762

4. Discussion characters (McAlpine, 1978, 1983; see below) which, in view of our results, must be interpreted as convergence. Additionally, our 4.1. Phylogenetic conclusions data do not support a close relationship of Neminidae with Aulacig- aster, as suggested by McAlpine (1983). In conclusion, based on our Relationships among ‘‘acalyptrate” flies are notably difficult be- results we favor an expanded concept of Aulacigastridae that in- cause of their large diversity and the high level of homoplasy ob- cludes Stenomicra, Cyamops, and Planinasus, in addition to Stenoc- served in most studied character systems. DNA sequence data yamops and Curiosimusca (not sampled in this analysis), and have proven to be no exception to this latter difficulty. However, excludes Neminidae. Morphological support for Aulacigastridae, our results permit a few definitive conclusions. Most importantly, as defined here, might depend on the exploration of alternative Opomyzoidea sensu McAlpine and variants thereof proposed by character systems. The sister-group relationship between Aulacig- other authors are decidedly non-monophyletic, even if Xenasteia aster and Curiosimusca is supported by a handful of putative syna- is removed. In our recovered topologies, Opomyzoidea is not sim- pomorphies (Rung et al., 2005). In contrast, all synapomorphies ply paraphyletic or diphyletic, but lacks cohesion in any sense. This advanced by Hennig (1971) to support an expanded Aulacigastri- failure of monophyly cannot be attributed simply to lack of phylo- dae are highly homoplasious and none of them is present in all genetic resolution because it is statistically supported by the re- aulacigastrids as the family is here defined. The most important sults of the topology tests, and recovery of monophyletic families of these characters (McAlpine, 1989), ‘‘ocellar setae weak or ab- through analysis of our data suggests the presence of some useful sent,” has intermediate states amongst undescribed species of phylogenetic signal. This result should not be wholly surprising, gi- Aulacigaster, and the ocellar seta is fully developed in Curiosimusca. ven the previously ambiguous morphological support for this Within an expanded Aulacigastridae, we did not find the genera group and the earlier status of many component families as Cyamops, Stenomicra, and Planinasus to form a monophyletic group. ‘‘unplaced”. Instead, Cyamops appeared more closely related to Aulacigaster In the case of McAlpine’s (1989) opomyzoid suprafamilies, our than to Stenomicra and Planinasus, though with low support. results are not as decisive. Non-monophyly for these groups was Although Cyamops, Stenomicra, and Planinasus are now generally statistically significant according to the topology tests only in the acknowledged to be closely related, the presumption that they case of Asteioinea; monophyly for the other three suprafamilies form a monophyletic group is not universal. For instance, Stenom- (Opomyzoinea, Agromyzoinea, Clusioinea) cannot be ruled out. icra is morphologically quite reduced, and its relationship with Furthermore, alternative arrangements for families in these supra- Cyamops and Planinasus was not initially apparent to Hennig families were weakly supported and inconsistent among analyses. (1958, 1959) or Griffiths (1972). Hennig (1971) suggested, with Morphological evidence has been ambivalent at best in support for some reservations, placement in Aulacigastridae as the sister group these groups. For example, Acartophthalmidae (Clusioinea) is now to Cyamops, Planinasus, and Aulacigaster. On the other hand, Baptis- generally included in the superfamily Carnoidea (Brake, 2000; ta and Mathis (1994), while retaining the concept of Stenomicrinae Buck, 2006), and both Hennig (1971, 1973) and Griffiths (1972) within Periscelididae, thought it likely that Planinasus was in fact considered Fergusoninidae (Agromyzoinea) as too highly derived more closely related to Periscelidinae, some genera of which show to place anywhere. Opomyzoinea represents a special enigma. a strong resemblance to Planinasus. From the above review, we The close relationship of Opomyzidae and Anthomyzidae has been conclude that relationships recovered by the current study within accepted by most recent authors (Hennig, 1971; McAlpine, 1989; Aulacigastridae, while not strongly supported, are plausible con- Rohácˇek, 1998a, 2006; but see Colless and McAlpine, 1991) and sidering the morphological evidence. is supported by similar larval biologies and a similarly modified The enigmatic genus Periscelis does not group with other Aste- structure of the male aedeagal complex, among other characters. ioinea in any analysis, and its position varies widely among the Our failure to recover Opomyzoinea should thus be considered sus- separate analyses. The hypothesis of a close relationship between pect, especially in light of the observed 28S base composition bias Periscelis, Stenomicrinae, and Neurochaetidae is based primarily in Anthomyzidae, which may have resulted in spurious placement. on a distinctive antennal form in which the pedicel is cap-like with However, we note that a close relationship between Anthomyzidae a dorsal seam; the first flagellomere (third antennal segment or ba- and other families (notably members of ‘‘Heteromyzidae” sensu sal flagellomere) is geniculate, or reflexed downward; and the aris- McAlpine, 2007) has been suggested before based on morphology ta has alternate, long rays (McAlpine, 1978). These taxa also have a (Speight, 1969; Griffiths, 1972). Additionally, in a study of the fe- unique type of articulation between the pedicel and first flagello- male internal genital tract of the Opomyzidae, Kotrba and Baptista mere in which a basal process of the first flagellomere is inserted (2002) failed to find any putative synapomorphies between Opo- into a cavity in the pedicel (McAlpine, 1978; Rung, Mathis, and myzidae and Anthomyzidae. Instead, the authors advanced a few Mitter, unpublished); in most other Schizophora, a membranous putative synapomorphies that may unite each of these two fami- extension of the pedicel is instead inserted into a cavity in the fla- lies with different groups within the Opomyzoidea. In light of these gellum. The placement of Aulacigastridae in this group by McAl- considerations, our results cast further doubt upon the validity of pine (Asteioinea group 1, see Fig. 1) implies a reversal of these McAlpine’s (1989) opomyzoid suprafamilies, despite lack of statis- antennal characters in this family. Because of this, Grimaldi and tical support. Mathis (1993) suggest instead a sister-group relationship between The genera Stenomicra, Cyamops, and Planinasus evidently are Periscelididae (including Stenomicrinae) and Neurochaetidae. De- not closely related to Periscelis, but instead show close affinity with spite these striking similarities in antennal structure, it is not clear Aulacigaster, as proposed by Hennig (1958, 1969, 1971). We there- if the antennae of Periscelis are substantially different from mem- fore follow Hennig (1971) and McAlpine (1989) in considering bers of the Diopsoidea (including part of ‘‘Nothyboidea” of Grif- these genera (as well as the related Procyamops and Stenocyamops) fiths, 1972) and some other acalyptrates in these structural as belonging to Aulacigastridae. Although support for this ex- details. For example, McAlpine (1997) considers the diopsoid panded concept of Aulacigastridae was poor in the ML and MP groundplan to include a deeply concave pedicel with a dorsal seam, analyses, its strong (100% PP) support in the BI analysis and the deflexion of the first flagellomere, and articulation of the first flag- consistent grouping of these genera in the ML analyses are suffi- ellomere by means of a dorsobasal process inserted into a cavity on cient in our view to uphold Hennig’s classification. Inclusion of the pedicel. Similarly, the antennal similarity between Periscelidi- these genera in Periscelididae was based primarily on antennal dae sensu stricto and Cyamops, Stenomicra, and Planinasus may be due to convergence. I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 757

One strongly supported result from our data is that Xenasteii- in Mr. Bayes (Fig. 5). Models that account for taxon-specific base dae is closely related to Australimyzidae and does not belong with frequency variation also exist (Blanquart and Lartillot, 2006; Gow- Asteiidae or any other ‘‘opomyzoid” family. These small, enigmatic ri-Shankar and Rattray, 2007), but their application has not yet families share a common habitat (seashore), and proposed charac- been well-developed. In the future, such improved models will ters uniting the Xenasteiidae and Asteiidae (Hardy, 1980), most probably be important in more confidently resolving relationships notably in wing venation, can probably be attributed to morpho- in the Schizophora and other phylogenetically difficult groups. logical reduction associated with reduction in size. This result also Cladistic (and non-cladistic) analysis of morphological charac- confirms the distinctness of Australimyza from Carnidae, as first ters faces similar problems, and the highly homoplastic nature of proposed by Griffiths (1972). Although later authors (McAlpine, many characters used in acalyptrate higher classification has been 1989; Buck, 2006) supported its inclusion as a separate family in acknowledged by previous workers (Hennig, 1971; McAlpine, Carnoidea, Griffiths found it to be unrelated to any other muscoid 1989) and demonstrated in work by the second author (Rung, lineage and accordingly designated a separate prefamily for Austra- Mathis, and Mitter, unpublished). Debates in the literature thus limyza. Furthermore, the concept of Carnoidea fares little better largely center on the a priori polarity of characters and on the sub- than Opomyzoidea in our analyses and morphological support for jective weight that should be given to various characters or charac- this group is similarly weak (Buck, 2006). More systematic work ter systems (Yeates and Wiegmann, 1999). Taxa of the including more characters and additional taxa is desirable. ‘‘Opomyzoidea” may be especially prone to morphological bias due to frequent reduction in size, with accompanied morphological 4.2. Causes and implications of phylogenetic uncertainty changes. This is especially true with characters of the wing vena- tion, but may also apply to others such as thoracic structure, chae- The differences between the topologies presented here and pre- totaxy, and even genitalic structures. Other characters of the male vious morphological hypotheses may be explained partly by a high and female postabdomen may vary rapidly, and a correlation with degree of phylogenetic ‘‘noise” (i.e. large fraction of highly homo- mating system can be expected. In general, it is agreed that novel plastic characters) in both molecular and morphological data sets. character systems, including molecular sequence data, will be Because both molecular and previous morphological hypotheses important for substantial future progress in schizophoran phylog- are subject to this random error and generally poorly supported, enetics (Yeates et al., 2007). it is difficult to evaluate the accuracy of many relationships pre- Regardless of the quality of data available, deeper relationships sented here. In our exploratory analyses (not shown), we tried var- of Schizophora may prove difficult because this represents an an- ious means of maximizing the signal contained in our data, cient, rapid radiation comparable to those of similar age in other including attempting to extract signal from the excluded hyper- orders (e.g. Ditrysian moths, parasitic ; Whit- variable regions (Lutzoni et al., 2000; Wheeler et al., 2006) and field and Kjer, 2008). Although the hypothesis of rapid radiation implicitly downweighting homoplastic characters (Goloboff, may be invoked more often than warranted for intransigent data 1997). None of these methods produced demonstrably better re- sets, there are reasons to believe that the lack of resolution we ob- sults than the approach presented here of excluding rapidly evolv- served reflects, in part, a true rapid radiation of schizophoran flies. ing regions and positions that are more likely to be saturated and To begin with, the markers used by us have been successful in biased. In our experience during this study, using Gblocks (Castre- resolving dipteran phylogeny at several different levels (Moulton sana, 2002) to delimit the hypervariable regions of 28S seems to and Wiegmann, 2007; Bertone et al., 2008; Scheffer et al., 2007; facilitate a good balance between including too many ‘‘noisy” sites Winterton et al., 2006; Petersen et al. 2007), and observed diver- and excluding too many informative sites, depending on the gence levels do not suggest that saturation is an important prob- parameters used. However, even with hypervariable regions in- lem in our data (Fig. 2; except for third positions of CAD). The cluded, a 28S ML analysis (not shown) resulted in a substantially conjecture of a rapid Tertiary radiation of Schizophora is also sup- similar tree, with all but one (Psilidae) of the included families ported by the fossil record (Blagoderov et al., 2002; Grimaldi and recovered and including the clade Stenomicrinae + Aulacigastridae. Engel, 2005). Despite a few possible earlier records (McAlpine, Despite this encouraging observation, inclusion of these regions 1970; Grimaldi et al., 1989), most extant family-level diversity is may exacerbate base composition bias, and they are more likely first encountered in the extensively studied Baltic amber deposits to be saturated, and thus misleading, at deeper levels in the (Eocene, 50 mya; Hennig, 1965; Evenhuis, 1994; von Tschirnhaus phylogeny. and Hoffeins, 2009). This is true of many families treated here, Systematic biases may also contribute to lack of phylogenetic including Clusiidae, Acartophthalmidae, Odiniidae, Anthomyzidae, resolution among families included here. For example, these may Aulacigastridae, Neurochaetidae, and Carnidae. If this picture of take the form of base composition bias or long branch artifacts in schizophoran evolution is accurate, it may be difficult to resolve the molecular data or repeated parallel reduction in morphological the phylogeny of Schizophora even with a large number of charac- characters. Systematic biases are more problematic than random ters. However, this degree of pessimism is not yet warranted, as error because they cannot be resolved by the inclusion of more lack of resolution due to rapid radiation, systematic biases, or use characters. Furthermore, detecting systematic error in phylogeny of markers with inappropriate evolutionary rates may be overcome construction is generally difficult (Bergsten, 2005), and we do not with addition of more taxa or genes (e.g. Ortiz-Rivas et al., 2004; know which, if any, results of our study may be attributable to sys- Baurain et al., 2007; Jian et al., 2008). tematic error. Even in cases where non-stationarity of base fre- Molecular phylogenetics of the ‘‘Acalyptratae” is still in its in- quencies is observed, we are unsure whether similar base fancy and is now primarily limited to work within individual frequencies represent convergence or common inheritance. More genera or families, for example, (Remsen and realistic models of sequence evolution are one possible solution O’Grady, 2002), (Meier and Wiegmann, 2002), Agro- to this problem. For example, Baurain et al. (2007) and Lartillot myzidae (Scheffer et al., 2007), Diopsidae (Meier and Baker, et al. (2007) found that using a model (CAT model; Lartillot and 2002), (Su et al., 2008), and Tephritidae (Han and Philippe, 2004) that accounts for site-specific variation in base fre- McPheron, 1997). Although calyptrate flies have been slightly quencies corrected systematic biases in inferring phylog- better studied (Bernasconi et al., 2000; Nirmala et al., 2001; Pet- eny. Our exploratory analyses implementing the CAT model in ersen et al., 2007; Kutty et al., 2008), to our knowledge only one PhyloBayes v. 2.3 (Lartillot et al., 2004) for our 28S data (results other published molecular phylogenetic study has investigated not shown) found largely the same result as the combined analysis relationships between acalyptrate families: that of Han and Ro 758 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762

(2005) on Tephritoidea. Molecular phylogenetic studies of other Although we were not able to definitively resolve any clades groups face greater obstacles, including less robust initial mor- above the family-level (except for Xenasteiidae + Australimyzi- phological hypotheses of internal relationships, and more impor- dae), our results are valuable in corroborating some previous tantly, weaker evidence for monophyly of clades to begin with. hypotheses and contradicting others. Most notably, Hennig’s As a result, it is difficult to obtain a reliable initial frame of ref- (1971) placement of Stenomicra, Cyamops, and Planinasus within erence from which to design taxon sampling and identify egre- Aulacigastridae is confirmed by molecular data. Additionally, gious examples of systematic biases. Our study highlights these some novel hypotheses emerged from analysis of our data which problems, which we tried to alleviate by using extensive sam- deserve further investigation. Accordingly, we view this study as pling of possible outgroups, including one non-schizophoran, an important initial step in using molecular data to evaluate the and by evaluating alternative analyses based on recovery of rec- plethora of morphological hypotheses available for higher rela- ognized families. tionships of the Schizophora. As schizophoran flies represent a Although some groups appear to be based on firm morphlog- major clade including important model organisms and agricul- ical grounds (e.g. Ephydroidea, Sciomyzoidea), a number of tural pests, further work to resolve the phylogenetic history of schizophoran superfamilies proposed by Hennig (1971) and oth- Schizophora is warranted and will yield rich fruit for comparative ers were originally based on weak evidence and intended as ten- evolutionary biology. tative working hypotheses. Some of these hypotheses have been given more weight than originally intended as they have been Acknowledgments repeated in the literature. Given the rampant polyphyly of the Opomyzoidea observed in our study, the existence of a paraphy- We thank Charlie Mitter and the Department of Entomology letic residue of ‘‘orphan” lineages within Schizophora is possible, at the University of Maryland for encouraging and assisting us and dipterists should be wary of accepting the monophyly of to begin this project. Matthew Lewis (USDA) was very helpful other proposed acalyptrate superfamilies without a careful re- in providing assistance and instruction in the laboratory. Brian view of the evidence. Consequently, future studies of schizopho- Wiegmann supported the final stages of the study under the ran phylogeny should include as broad a taxon sample as NSF-funded FLYTREE project (EF-03-34948). Special thanks are possible, at least until monophyletic groups are identified with extended to those who donated specimens: Amnon Freidberg, a greater degree of confidence. Increased character sampling, Wayne Mathis, Peter Kerr, Brian Wiegmann, Marion Kotrba, Gary including more protein-coding genes, is also desirable, though Taylor, and Riley Nelson. Owen Lonsdale, Bob Kula, Wayne the poor performance of CAD here suggests this may not be a Mathis, and two anonymous reviewers gave helpful comments straightforward solution. on the manuscript.

Appendix A

Exemplar specimens used in the phylogenetic analysis.

Family Species Collection data Genbank No. (28S/CAD) Agromyzidae Agromyza ambrosivora USA, Maryland, Prince George’s Co., EF104822 EF104735 Spencer SE of Clinton, IX.2000, S.J. Scheffer South Africa, Western Cape Province, EF104876 EF104788 (Blanchard) Lambert’s Bay, XI.1999, D. Visser, ex. potato (Loew) USA, Colorado: Gunnison Co., Irwin, EF104859 EF104771 VI.1997, S.J. Scheffer

Acartophthalmidae Acartophthalmus bicolor UK, Gwynedd Co., Afon Eden, Coed y GU299253 GU299254 Oldenberg Brenin Park, 19.VI.2004, I. Winkler

Anthomyzidae Fallén Germany, Berlin, Charlottenburg, GU299220 — 1.VII.2002, M. Kotrba & A. Rung sp. USA, Maryland, Montgomery Co., GU299251 GU299270 Beltsville, 7.VIII.2002, I. Winkler & A Rung

Asteiidae Meigen Switzerland, Genève, Jussy Prés-de- GU299218 GU299255 Villette, 2.VIII.2002, B. Merz Leiomyza curvinervis USA, Washington, Pend Oreille Co., NE of GU299232 GU299268 Zetterstedt Sullivan Lake, 7.VI.2003, I. Winkler

Aulacigastridae s.str. Aulacigaster ecuadoriensis Ecuador, Puerto Orellana, Rio Tiputini, GU299237 GU299256 (Hennig) 12–26.VIII.1999, W.N. Mathis, A. Rung, M. Kotrba Aulacigaster leucopeza Germany, Berlin, Tiergarten, 28.VI.1999, GU299238 GU299257 (Meigen) M. Kotrba Aulacigaster neoleucopeza USA, Maryland, Prince George’s Co., GU299224 GU299258 Mathis and Freidberg Greenbelt, 27.VII.2002, A. Rung I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 759

Appendix A (continued) Family Species Collection data Genbank No. (28S/CAD) Clusiidae Clusia czernyi Johnson USA, Maryland, Montgomery Co., GU299246 GU299263 Plummers Island, 7–15.V.2001, A. Rung Clusiodes johnsoni Malloch USA, Maryland, Prince George’s Co., GU299240 GU299264 College Park, 25.VIII.2002, I. Winkler & A. Rung

Fergusoninidae Fergusonina purcelli Taylor Australia, Queensland, Daintree Ferry, GU299241 GU299265 (FC24) 76-MC-F3, ex. cajuputi Fergusonina sp. (FC73) Australia, WINC2004, ex. GU299242 GU299266 intertexta

Marginidae Margo clausa McAlpine Madagascar, Andasibe, Analamazaotra GU299230 — Forest, 31.X-4.XI.2007, A. Freidberg

Neminidae Nemula longarista Freidberg Madagascar, Andasibe, Analamazaotra GU299223 — Forest, 31.X-4.XI.2007, A. Freidberg

Neurochaetidae Neurochaeta inversa Australia, Upper Cedar Cr., nr. Sanford, GU299217 GU299271 McAlpine 5.X.2002, M. Kotrba

Odiniidae Odinia boletina (Zetterstedt) UK, Hertfordshire, Aldbury Common, GU299247 GU299272 14.VI.1997 Neoalticomerus seamansi UTAH, Utah Co., American Fork Cyn., 1– GU299229 — Shewell 15.VII.2005, C.R. Nelson et al.

Opomyzidae Fallén Switzerland, Genève, Bernex Chante- GU299222 GU299267 Merle 16.VIII.2002, B. Merz Opomyza sp. Germany, Bradenburg, Kloster Chorin, GU299221 GU299275 VII.2000, M. Kotrba & A. Rung

Periscelididae s.str. Periscelis flinti (Malloch) USA, Maryland, Prince George’s Co., GU299228 GU299274 Greenbelt, 29.VII.2002, A. Rung

Stenomicrinae (included in Cyamops sp. Australia, Queensland, Josephine Falls, GU299236 — Periscelididae or 23.IX.2002, W.N. & D. Mathis Aulacigastridae) Cyamops nebulosus USA, Maryland, Montgomery Co., GU299226 — Melander Pautuxent Refuge, S.J. Scheffer et al. Stenomicra angustata USA, Maryland, Montgomery Co., GU299244 GU299277 Coquillett Beltsville, 7.VIII.2002, I. Winkler & A. Rung Planinasus sp. Brazil, Rio de Janeiro, Floresta da Tijuca, GU299243 — VIII.1998, A. Rung, R. Baptista

Teratomyzidae Teratomyza (Vitila) sp. Australia, Queensland, Millaa Millaa, GU299227 GU299279 24.IX.2002, W.N. & D. Mathis

Xenasteiidae Xenasteia shalam Israel, Enot, 3.IX.2007, A. Freidberg GU299225 — (Freidberg)

Outgroups Australimyzidae (Carnoidea) Australimyza salicorniae New Zealand, N. Isl., WN, Waikanae GU299219 GU299259 Harrison Beach, 6.I.2004, W.N. Mathis

Carnidae (Carnoidea) Meoneura sp. USA, California, Inyo Co., South Fork GU299239 GU299260 Bishop Cr., 4-Jeffrey campground, 3.III.2003, I. Winkler

Chloropidae (Carnoidea) Chaetochlorops inquilinus USA, Maryland, Montgomery Co., GU299234 GU299261 (Coquillett) Plummers Island, 7-15.V.2001, P. Kerr Thaumatomyia sp. USA, Maryland, Prince George’s Co., GU299235 GU299280 Greenbelt, 2.VII.2002, I. Winkler & A. Rung

Chyromyidae (Sphaeroceroidea) Aphaniosoma sp. USA, Colorado, Adams Co., Barr Lake St. GU299231 — Park, 3.III.2004, I. Winkler

(continued on next page) 760 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762

Appendix A (continued) Family Species Collection data Genbank No. (28S/CAD) Gymnochiromyia sp. USA, Maryland, Prince George’s Co., GU299250 — Adelphi, 19.IV.2004, I. Winkler

Milichiidae (Carnoidea) Paramyia nitens (Loew) USA, Maryland, Prince George’s Co., GU299233 GU299273 Adelphi, 3.VIII.2002, I. Winkler

Muscidae (Calyptratae) Musca domestica Linnaeus (Genbank) AF503004 AY280689 AF503025 AY123358

Psilidae (Diopsoidea) Chyliza sp. USA, Maryland, Prince George’s Co., GU299249 GU299262 Greenbelt, 29.VIII.2002, A. Rung Loxocera sp. USA, Maryland, Montgomery Co., GU299245 GU299269 Plummers Island, 7–15.V.2001, P. Kerr

Sphaeroceridae (Sphaeroceroidea) Pterogramma sp. USA, Maryland, Montgomery Co., GU299248 GU299276 Plummers Island, 4–18.VI.2001, P. Kerr

Syrphidae () Rhingia nasica Say (Genbank) AF502977 AY280697 AF502998 AF503019

Tephritidae (Tephritoidea) Stone Australia, Queensland, 2001, A.L. GU299252 GU299278 Norrbom, ex. Ageratina adenophora

References Evenhuis, N.L., 1994. Catalogue of the Fossil of the World (Insecta: Diptera). Available from: (accessed Oct. 2008). Baptista, A.R., Mathis, W.N., 1994. A revision of New World Cyamops Melander Evenhuis, N.L., Pape, T., Pont, A.C., Thompson, F.C. (Eds.), 2008. Biosystematic (Diptera, Periscelididae). Smithson. Contrib. Zool. 563, 1–28. Database of World Diptera, Working Version. Available from: (accessed Dec. 2008). acalyptrate flies (Diptera: Schizophora: Sciomyzoidea). Afr. Invertebr. 47, 117– Fish, D., 1983. Phytotelmata: Flora and Fauna. In: Frank, J.H., Lounibos, L.P. (Eds.), 134. Phytotelmata: Terrestrial Plants as Hosts for Aquatic Insect Communities. Baurain, D., Brinkmann, H., Philippe, H., 2007. Lack of resolution in the animal Plexus, Medford, New Jersey, pp. 1–27. phylogeny: closely spaced cladogeneses or undetected systematic errors? Mol. Freidberg, A., 1994. Nemula, a new genus of Neminidae (Diptera) from Madagascar. Biol. Evol. 24, 6–9. Proc. Entomol. Soc. Wash. 96, 471–482. Bergsten, J., 2005. A review of long branch attraction. Cladistics 21, 163–193. Freidberg, A., 2002. Recent advances in the study of Asteiidae. In: Abstracts Volume, Bernasconi, M.V., Valsangiacomo, C., Piffaretti, J.C., Ward, P.I., 2000. Phylogenetic Fifth International Congress of Dipterology, Brisbane, Australia. University of relationships among Muscoidea (Diptera: Calyptratae) based on mitochondrial Queensland, Brisbane, p. 73. DNA sequences. Insect Mol. Biol. 9, 67–74. Gaimari, S.D., Mathis, W.N., in press. World catalog and conspectus of the family Bertone, M.A., Courtney, G.W., Wiegmann, B.M., 2008. Phylogenetics and temporal Odiniidae (Diptera: Schizophora). Myia. diversification of the earliest true flies (Insecta: Diptera) based on multiple Goloboff, P.A., 1997. Self-weighted optimization: tree searches and character nuclear genes. Syst. Entomol. 33, 668–687. state reconstructions under implied transformation costs. Cladistics 13, 225– Blagoderov, V.A., Lukashevich, E.D., Mostovski, M.B., 2002. Order Diptera Linné, 245. 1758. The true flies (= Muscida Laicharting, 1781). In: Rasnitsyn, A.P., Quicke, Gowri-Shankar, V., Rattray, M., 2007. A reversible jump method for Bayesian D.L.J. (Eds.), History of Insects. Kluwer Academic Publishers, Dordrecht, pp. 227– phylogenetic inference with a nonhomogeneous substitution model. Mol. Biol. 240. Evol. 24, 1286–1299. Blanquart, S., Lartillot, N., 2006. A Bayesian compound stochastic process for Griffiths, G.C.D., 1972. The phylogenetic classification of Diptera with modeling non-stationary and nonhomogeneous sequence evolution. Mol. Biol. special reference to the structure of the male postabdomen. Ser. Entomol. 8, 1– Evol. 23, 2058–2071. 340. The Hague. Brake, I., 2000. Phylogenetic systematics of the Milichiidae (Diptera, Schizophora). Grimaldi, D.A., Engel, M.S., 2005. Evolution of the Insects. Cambridge University Entomol. Scand. 57, 1–58. Press, Cambridge, UK. Brake, I., Mathis, W.N., 2007. Revision of the genus Australimyza Harrison (Diptera: Grimaldi, D.A., Mathis, W.N., 1993. Fossil Periscelididae (Diptera). Proc. Entomol. Australimyzidae). Syst. Entomol. 32, 252–275. Soc. Wash. 95, 383–403. Britton, D., 2009. Rare – or just hidden? Australian Museum Online. Available from: Grimaldi, D.A., Beck, C.W., Boon, J.J., 1989. Occurrence, chemical characteristics, and paleontology of the fossil resins from New Jersey. Am. Mus. Novit. 2948, 1–28. (accessed Jan. 2010). Han, H.Y., McPheron, B.A., 1997. Molecular phylogenetic study of Tephritidae Brunel, É., 1998. Family Opomyzidae. In: Soós, A., Papp, L. (Eds.), Catalogue of (Insecta: Diptera: Tephritidae) using partial sequences of the mitochondrial 16S Palearctic Diptera. Elsevier Science Publishers and Akadémiai Kiadó, Budapest, ribosomal DNA. Mol. Phylogenet. Evol. 7, 17–32. pp. 259–266. Han, H.Y., Ro, K.E., 2005. Molecular phylogeny of the superfamily Tephritoidea Buck, M., 2006. A new family and genus of acalypterate flies from the Neotropical (Insecta: Diptera): new evidence from the mitochondrial 12S, 16S, and COII region, with a phylogenetic analysis of Carnoidea family relationships (Diptera, genes. Mol. Phylogenet. Evol. 34, 416–430. Schizophora). Syst. Entomol. 31, 377–404. Hardy, D.E., 1980. Xenasteiidae, a new family of Schizophora from the Pacific and Castresana, J., 2000. Selection of conserved blocks from multiple alignments for Indian oceans. Proc. Hawaii Entomol. Soc. 23, 205–225. their use in phylogenetic analysis. Mol. Biol. Evol. 17, 540–552. Hennig, W., 1958. Die Familien der Diptera Schizophora und ihre phylogenetischen Castresana, J., 2002. Gblocks v.0.91b. Available from: . Hennig, W., 1965. Die Acalyptraten des Baltischer Bernsteins. Stuttgarter Beitr. Colless, D.H., McAlpine, D.K., 1991. Diptera (flies), second ed.. In: Naumann, I.D. Naturkd. 145, 1–215. (Ed.), The Insects of Australia. A Textbook for Students and Research Workers, Hennig, W., 1969. Neue Gattungen und Arten der Acalypteratae. Can. Entomol. 101, vol. II Melbourne University Press, Carlton, Victoria, Australia, pp. 717–786. 589–633. Currie, G.A., 1937. Galls on Eucalyptus trees: a new type of association between flies Hennig, W., 1971. Neue Untersuchungen über die Familien der Diptera Schizophora and nematodes. Proc. Linn. Soc. N. S. W. 62, 147–174. (Diptera: Cyclorrhapha). Stuttgarter Beitr. Naturkd. 226, 1–76. Edgar, R.C., 2004. MUSCLE: multiple sequence alignment with high accuracy and Hennig, W., 1973. Diptera. In: Kukenthal, W. (Ed.), Handbuch der Zoologie, IV: high throughput. Nucleic Acids Res. 32, 1792–1797. Available from: Arthropoda, 2, Part 2, vol. 31. De Gruyter, New York, pp. 1–337. . I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762 761

Hoffeins, C., Rung, A., 2005. Procyamops succini, a new genus and new species of Moulton, J.K., Wiegmann, B.M., 2004. Evolution and phylogenetic utility of CAD Periscelididae (Diptera, ) from Baltic amber (Tertiary, Eocene). Stud. (rudimentary) among Mesozoic-aged eremoneuran Diptera (Insecta). Mol. Dipterol. 12, 23–27. Phylogenet. Evol. 31, 363–378. Jian, S., Soltis, P.S., Gitzendanner, M.A., Moore, M.J., Li, R., Hendry, T.A., Qiu, Y.-L., Moulton, J.K., Wiegmann, B.M., 2007. The phylogenetic relationships of flies in the Dhingra, A., Bell, C.D., Soltis, D.E., 2008. Resolving an ancient rapid radiation in superfamily (Insecta: Diptera). Mol. Phylogenet. Evol. 43, 701–713. Saxifragales. Syst. Biol. 57, 38–57. Murphy, D.H., 1990. The natural history of insect herbivory on mangrove trees in Khoo, K.C., Sabrosky, C.W., 1989. Family . In: Evenhuis, N.L. (Ed.), and near Singapore. Raffles Bull. Zool. 38, 119–203. Catalog of the Diptera of the Australasian and Oceanian Regions. B.P. Bishop Nirmala, X., Hypsa, V., Zurovec, M., 2001. Molecular phylogeny of Calyptratae Museum Special Publication No. 86.. E.J. Brill and Bishop Museum Press, (Diptera: Brachycera): the evolution of 18S and 16S ribosomal rDNAs in higher Honolulu. p. 551. dipterans and their use in phylogenetic inference. Insect Mol. Biol. 10, 475–485. Kotrba, M., Baptista, A.R., 2002. The internal reproductive tract of the Opomyzidae Ortiz-Rivas, B., Moya, A., Martínez-Torres, D., 2004. Molecular systematics of (Diptera, Schizophora). Stud. Dipterol. 9, 57–71. (Homoptera: Aphididae): new insights from the long-wavelength opsin gene. Kutty, S.N., Pape, T., Pont, A., Wiegmann, B.M., Meier, R., 2008. The Muscoidea Mol. Phylogenet. Evol. 30, 24–37. (Diptera: Calyptratae) is paraphyletic: evidence from four mitochondrial and Papp, L., 1980. New taxa of acalyptrate flies (Diptera: Tunisimyiidae fam. n., Risidae, four nuclear genes. Mol. Phylogenet. Evol. 49, 639–652. Ephydridae: Nannodastiinae subfam. n.). Acta Zool. Acad. Sci. Hung. 26, 415–431. Lartillot, N., Philippe, H., 2004. A Bayesian mixture model for across-site Papp, L., 1984. Family Aulacigastridae. In: Soós, A., Papp, L. (Eds.), Catalogue of heterogeneities in the amino-acid replacement process. Mol. Biol. Evol. 21, Palearctic Diptera. Elsevier Science Publishers and Akadémiai Kiadó, Budapest, 1095–1109. pp. 60–61. Lartillot, N., Blanquart, S., Lepage, T., 2004. PhyloBayes 2.3: a Bayesian software for Papp, L., 1998a. Family Aulacigastridae. In: Papp, L., Darvas, B. (Eds.), Contributions phylogenetic reconstruction using mixture models. Available from: . Herald, Budapest, pp. 279–284. Lartillot, N., Brinkmann, H., Philippe, H., 2007. Suppression of long-branch attraction Papp, L., 1998b. Family Asteiidae. In: Papp, L., Darvas, B. (Eds.), Contributions to a artefacts in the animal phylogeny using a site-heterogeneous model. BMC Evol. Manual of Palaearctic Diptera, Volume 3: Higher Brachycera. Science Herald, Biol. 7 (Suppl. 1), S4. Budapest, pp. 295–303. Lonsdale, O., Marshall, S.A., 2006. Redefinition of the Clusiinae and Clusiodinae, Papp, L., 1998c. Family Xenasteiidae. In: Papp, L., Darvas, B. (Eds.), Contributions to a description of the new subfamily Sobarocephalinae, revision of the genus Manual of Palaearctic Diptera, Volume 3: Higher Brachycera. Science Herald, Chaetoclusia and a description of Procerosoma gen. n. (Diptera: Clusiidae). Eur. J. Budapest, pp. 305–308. Entomol. 103, 163–182. Papp, L., Ozerov, A.L., 1998. Family Acartophthalmidae. In: Papp, L., Darvas, B. (Eds.), Lonsdale, O., Marshall, S.A., 2007. Redefinition of the genera Clusiodes and Hendelia Contributions to a Manual of Palaearctic Diptera, Volume 3: Higher Brachycera. (Diptera: Clusiidae: Clusiodinae) with a review of Clusiodes. Stud. Dipterol. 14, Science Herald, Budapest, pp. 227–232. 117–159. Papp, L., Merz, B., Földvári, M., 2006. Diptera of Thailand: a summary of the families Lonsdale, O., Marshall, S.A., 2008. Synonymy within Clusia and description of the and genera with references to the species representations. Acta Zool. Acad. Sci. new genus Melanoclusia (Diptera: Clusiidae: Clusiinae). Ann. Entomol. Soc. Am. Hung. 52, 97–268. 101, 327–330. Petersen, F.T., Meier, R., Kutty, S.N., Wiegmann, B.M., 2007. The phylogeny and Lutzoni, F., Wagner, P., Reeb, V., Zoller, S., 2000. Integrating ambiguously aligned evolution of host choice in the (Diptera) as reconstructed using regions of DNA sequences in phylogenetic analyses without violating positional four molecular markers. Mol. Phylogenet. Evol. 45, 111–122. homology. Syst. Biol. 49, 628–651. Remsen, J., O’Grady, P.M., 2002. Phylogeny of Drosophilidae (Diptera), with Mathis, W.N., in press. World Catalog and Conspectus on the Family Periscelididae comments on combined analysis and character support. Mol. Phylogenet. (Diptera: Schizophora). Myia. Evol. 24, 248–263. Mathis, W.N., Papp, L., 1992. A new genus of Periscelididae (Diptera) from the Rohácˇek, J., 1998a. Taxonomic limits, phylogeny and higher classification of Neotropics. Proc. Biol. Soc. Wash. 105, 366–372. Anthomyzidae (Diptera), with special regard to fossil record. Eur. J. Entomol. Mathis, W.N., Papp, L., 1998. Family Periscelididae. In: Papp, L., Darvas, B. (Eds.), 95, 141–177. Contributions to a Manual of Palaearctic Diptera, Volume 3: Higher Brachycera. Rohácˇek, J., 1998b. Family Anthomyzidae. In: Papp, L., Darvas, B. (Eds.), Science Herald, Budapest, pp. 285–294. Contributions to a Manual of Palaearctic Diptera, Volume 3: Higher Mathis, W.N., Rung, A., 2004. Redescription of the genus Diopsosoma Malloch Brachycera. Science Herald, Budapest, pp. 267–278. (Diptera, Periscelididae). Rev. Bras. Entomol. 48, 303–309. Rohácˇek, J., 2006. A monograph of Palaearctic Anthomyzidae (Diptera) part 1. Cˇas. McAlpine, J.F., 1970. First record of calypterate flies in the Mesozoic era (Diptera: Slez. Muz. Opava (A) 55 (suppl. 1), 1–328. Caliphoridae). Can. Entomol. 102, 342–346. Ronquist, F., Huelsenbeck, J.P., 2003. MrBayes 3: Bayesian phylogenetic inference McAlpine, D.K., 1978. Description and biology of a new genus of flies related to under mixed models. Bioinformatics 19, 1572–1574. Anthoclusia and representing a new family (Diptera, Schizophora, Rung, A., Mathis, W.N., Papp, L., 2005. Curiosimusca, gen. nov., and three new species Neurochaetidae). Ann. Natal Mus. 23, 273–295. in the family Aulacigastridae from the Oriental Region (Diptera: Opomyzoidea). McAlpine, D.K., 1983. A new subfamily of Aulacigastridae (Diptera: Schizophora), Zootaxa 1009, 21–36. with a discussion of aulacigastrid classification. Aust. J. Zool. 31, 55–78. Scheffer, S.J., Giblin-Davis, R.M., Davies, K.A., Taylor, G.S., Purcell, M., Lewis, M.L., McAlpine, D.K., 1988a. Studies in upside-down flies (Diptera: Neurochaetidae). Part Goolsby, J., Center, T.D., 2004. Phylogenetic relationships, species limits, and I. Systematics and phylogeny. Proc. Linn. Soc. N. S. W 110, 31–58. host specificity of gall-forming Fergusonina flies (Diptera: Fergusoninidae) McAlpine, D.K., 1988b. Studies in upside-down flies (Diptera: Neurochaetidae). Part feeding on Melaleuca (Myrtaceae). Ann. Entomol. Soc. Am. 97, 1216–1221. II. Biology, adaptations and specific mating mechanisms. Proc. Linn. Soc. N. S. W Scheffer, S.J., Winkler, I.S., Wiegmann, B.M., 2007. Phylogenetic relationships within 110, 59–82. the leaf-mining flies (Diptera: Agromyzidae) inferred from sequence data from McAlpine, J.F., 1989. Phylogeny and classification of the . In: multiple genes. Mol. Phylogenet. Evol. 42, 756–775. McAlpine, J.F., Wood, D.M. (Eds.), Manual of Nearctic Diptera, Volume 3: Shatalkin, A.I., 1993. On the of the flies of the family Research Branch, Agriculture Canada, Monograph 32, pp. 1397–1518. (Diptera). Zool. Zh. 72, 124–131. McAlpine, D.K., 1991. Marginidae, a new Afrotropical family of Diptera Shimodaira, H., 2002. An approximately unbiased test of phylogenetic tree (Schizophora: Opomyzoidea). Ann. Natal Mus. 32, 167–177. selection. Syst. Biol. 51, 492–508. McAlpine, D.K., 1993. Review of the upside-down flies (Diptera: Neurochaetidae) of Shimodaira, H., Hasegawa, M., 1999. Multiple comparisons of log-likelihoods with Madagascar and Africa, and evolution of neurochaetid host plant associations. applications to phylogenetic inference. Mol. Biol. Evol. 16, 1114–1116. Rec. Aust. Mus. 45, 221–239. Shimodaira, H., Hasegawa, M., 2001. CONSEL: for assessing the confidence of McAlpine, D.K., 1997. Gobryidae, a new family of acalyptrate flies (Diptera: phylogenetic tree selection. Bioinformatics 17, 1246–1247. Available from: Diopsoidea), and a discussion of relationships of the diopsoid families. Rec. . Aust. Mus. 49, 167–194. Speight, M.C.D., 1969. The prothoracic morphology of Acalyptrates (Diptera) and its McAlpine, D.K., 1998. Neminidae (Nobody flies). In: Oosterbroek, P. (Ed.), The use in systematics. Trans. R. Entomol. Soc. Lond. 121, 325–421. Families of Diptera of the Malay Archipelago. Fauna Malesiana Handbook 1. Spencer, K.A., 1990. Host specialization in the world Agromyzidae (Diptera). Ser. Brill, Leiden, pp. 123–124. Entomol. 45, 1–444. Kluwer Academic Publishers, Dordrecht. McAlpine, D.K., 2007. Review of the Borboroidini or wombat flies (Diptera: Su, F.-Y.K., Kutty, S., Meier, R., 2008. Morphology versus molecules: the Heteromyzidae), with reconsideration of the status of families Heleomyzidae phylogenetic relationships of Sepsidae (Diptera: Cyclorrhapha) based on and Sphaeroceridae, and descriptions of femoral gland-baskets. Rec. Aust. Mus. morphology and DNA sequence data from ten genes. Cladistics 24, 902–916. 59, 143–219. Suzuki, Y., Glazko, G.V., Nei, M., 2002. Overcredibility of molecular phylogenies McAlpine, D.K., de Keyzer, R.G., 1994. Generic classification of the fern flies (Diptera: obtained by Bayesian phylogenetics. Proc. Natl. Acad. Sci. USA 99, 16138–16143. Teratomyzidae) with a larval description. Syst. Entomol. 19, 305–326. Swofford, D.L., 2001. PAUP* 4.0 beta 10. Phylogenetic Analysis Using Parsimony Meier, R., 2005. Role of dipterology in phylogenetic systematics: the insight of Willi (*and Other Methods). Sinauer Associates, Massachussets. Hennig. In: Yeates, D.K., Wiegmann, B.M. (Eds.), The Evoutionary Biology of Tautz, D., Hancock, J.M., Webb, D.A., Tautz, C., Dover, G.A., 1988. Complete Flies. Columbia University Press, New York, pp. 45–62. sequences of the rRNA genes of Drosophila melanogaster. Mol. Biol. Evol. 5, Meier, R., Baker, R., 2002. A cladistic analysis of Diopsidae (Diptera) based on 366–376. morphological and DNA sequence data. Insect Syst. Evol. 33, 325–336. Taylor, G.S., Davies, K.A., Giblin-Davis, R.M., 2004. Species-richness in gall-flies Meier, R., Wiegmann, B.M., 2002. A phylogenetic analysis of Coelopidae (Diptera) (Diptera: Fergusoninidae), nematodes (Nematoda: Sphaerulariidae) and based on morphological and DNA sequence data. Mol. Phylogenet. Evol. 25, associated and inquilines on Myrtaceae. Rec. S. Aust. Mus. 393–407. Monogr. Ser. 7, 249–256. 762 I.S. Winkler et al. / Molecular Phylogenetics and Evolution 54 (2010) 746–762

Taylor, G.S., Head, E.R., Davies, K.A., 2005. Gall flies (Diptera: Fergusoninidae) on Systematics: A Unified Approach Using POY. American Museum of Natural Myrtaceae: a mutualistic association between flies and nematodes. In: Raman, History, New York. pp 365. A., Schaefer, C.W., Withers, T.M. (Eds.), Biology, , and Evolution of Gall- Whitfield, J.B., Kjer, K.M., 2008. Ancient rapid radiations of insects: challenges for inducing . Science Publishers, Enfield, New Hampshire, pp. 643–671. phylogenetic analysis. Annu. Rev. Entomol. 53, 449–472. Thompson, J.D., Higgins, D.G., Gibson, T.J., 1994. Clustal W: improving the sensitivity Wiegmann, B.M., Tsaur, S.C., Webb, D.W., Yeates, D.K., Cassel, B.K., 2000. Monophyly of progressive multiple sequence alignment through sequence weighting, and relationships of the (Diptera: Brachycera) based on 28S position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 22, ribosomal gene sequences. Ann. Entomol. Soc. Am. 93, 1031–1038. 4673–4680. Winterton, S.L., Wiegmann, B.M., Schlinger, E.I., 2006. Phylogeny and Vockeroth, J.R., 1987. Family Opomyzidae. In: McAlpine, J.F., Wood, D.M. (Eds), Bayesian divergence time estimations of small-headed flies (Diptera: Manual of Nearctic Diptera, Volume 2: Research Branch. Agriculture Canada, ) using multiple molecular markers. Mol. Phylogenet. Evol. Monograph 32, pp. 881–885. 43, 808–832. von Tschirnhaus, M., Hoffeins, C., 2009. Fossil flies in Baltic amber—insights in the Yeates, D.K., Wiegmann, B.M., 1999. Congruence and controversy: Toward a higher- diversity of Tertiary Acalyptratae (Diptera, Schizophora), with new level phylogeny of Diptera. Annu. Rev. Entomol. 44, 397–428. morphological characters and a key based on 1000 collected inclusions. Denisia Yeates, D.K., Wiegmann, B.M., Courtney, G.W., Meier, R., Lambkin, C., Pape, T., 2007. 26, 171–212. Phylogeny and systematics of Diptera: two decades of progress and prospects. Wheeler, T.A., 2000. Systematics of Mallochianamyia Santos-Neto (= Gayomyia Zootaxa 1668, 565–590. Malloch), an unplaced genus of Neotropical acalyptrate flies (Diptera: Zwickl, D.J., 2006. Genetic algorithm approaches for the phylogneetic analysis of Schizophora). Insect Syst. Evol. 31, 209–224. large biological sequence datasets under the maximum likelihood criterion. Wheeler, W.C., Aagesen, L., Arango, C.P., Faivovich, J., Grant, T., D’Haese, C., Janies, D., Ph.D. Dissertation, the University of Texas at Austin. Available from: .