Open Access
How to translate text using browser tools
20 January 2022 Partial Ribosomal Nontranscribed Spacer Sequences Distinguish Rhagoletis zephyria (Diptera: Tephritidae) From the Apple Maggot, R. pomonella
J. J. Smith, P. Brzezinski, J. Dziedziula, E. Rosenthal, M. Klaus
Author Affiliations +
Abstract

The apple maggot, Rhagoletis pomonella (Walsh), was introduced into the apple-growing regions of the Pacific Northwest in the U.S.A. during the past 60–100 yr. Apple maggot (larvae, puparia, and adults) is difficult to distinguish from its morphologically similar sister species, Rhagoletis zephyria Snow, which is native and abundant in the Pacific Northwest. While morphological identifications are common practice, a simple, inexpensive assay based on genetic differences would be very useful when morphological traits are unclear. Here we report nucleotide substitution and insertion–deletion mutations in the nontranscribed spacer (NTS) of the ribosomal RNA gene cistron of R. pomonella and R. zephyria that appear to be diagnostic for these two fly species. Insertion–deletion variation is substantial and results in a 49 base-pair difference in PCR amplicon size between R. zephyria and R. pomonella that can be scored using agarose gel electrophoresis. PCR amplification and DNA sequencing of 766 bp of the NTS region from 38 R. pomonella individuals and 35 R. zephyria individuals from across their geographic ranges led to the expected PCR fragments of approx. 840 bp and 790 bp, respectively, as did amplification and sequencing of a smaller set of 26 R. pomonella and 16 R. zephyria flies from a sympatric site in Washington State. Conversely, 633 bp mitochondrial COI barcode sequences from this set of flies were polyphyletic with respect to R. pomonella and R. zephyria. Thus, differences in NTS PCR products on agarose gels potentially provide a simple way to distinguish between R. pomonella and R. zephyria.

Tephritid flies in the Rhagoletis pomonella species group (Diptera: Tephritidae) have been of great interest and utility for studying fundamental evolutionary patterns and processes (Bush 1969; Feder et al. 1988; Powell et al. 2014; Doellman et al. 2018, 2019). However, these flies are also important in apple-growing regions in eastern North America as insect pests of apple (Malus domestica Bork.). In the Pacific Northwest of the U.S.A., apple maggot (Rhagoletis pomonella (Walsh)) is an introduced, major quarantine pest of apple (Dowell et al. 1990) that was first reported in Oregon in 1979 (AliNiazee and Penrose 1981), with the initial infestations there arising from its introduction from eastern North America (McPheron 1990b, Arcella et al. 2015). In the apple-growing areas of Oregon, Washington, and British Columbia, quarantine zones are in effect, economic costs of potential infestation are high, and there is zero tolerance for the presence of apple maggot in apple shipments destined for export (WSDA 2001, Bush et al. 2002, CFIA 2017, Galinato et al. 2018).

Detection of apple maggot in and near apple orchards in the Pacific Northwest is complicated by the presence of Rhagoletis zephyria Snow, which is native to the Pacific Northwest (Gavrilovic et al. 2007), and common and abundant in apple-growing areas. Rhagoletis zephyria is a close relative of R. pomonella within the pomonella species group (Gavrilovic et al. 2007, Xie et al. 2007), and infests fruit of snowberries, Symphoricarpos spp. (Caprifoliaceae). Symphoricarpos albus var. laevigatus (Fern.), which is both common and abundant throughout the Pacific Northwest, is the primary host for R. zephyria in apple-growing regions.

Rhagoletis mendax Curran is a third species that is very closely related genetically and morphologically similar to R. pomonella and R. zephyria (Bush 1966). However, R. mendax is known only from Vaccinium spp. and Gaylussacia spp. hosts, and its geographic range is limited to the eastern half of North America (Payne and Berlocher 1995, Smith et al. 2001). Thus, R. mendax does not cause a threat to apples or create a problem for the apple industry in the Pacific Northwest.

Because R. zephyria is virtually identical morphologically to R. pomonella (Bush 1966), this creates problems for morphological identifications in areas where the ranges of these species overlap. Despite these similarities, Westcott (1982) was able to draw upon the work of Curran (1924), Wasbauer (1963), and Bush (1966, 1969) to establish morphological criteria for distinguishing between adults of the two species. Males of the two species can be distinguished from each other based on a difference in the surstylus shape, while females differ in mean ovipositor length. These characters have been used extensively as the basis for species diagnosis of field-caught adult flies in annual apple maggot surveys in Washington State. Ovipositor length can be used to distinguish between females of the two species, but large R. zephyria females can be difficult to distinguish from small R. pomonella females; males of the two species appear to be reliably distinguished by surstylus shape (Westcott 1982). Additionally, the introduction of R. pomonella into the Pacific Northwest over the past 50 yr has created problems for species identification, as introgression appears to have been ongoing (McPheron 1990a), presumably between the relatively abundant R. zephyria males and the introduced R. pomonella females (Arcella et al. 2015).

A simple, inexpensive diagnostic tool to distinguish between R. pomonella and R. zephyria would be very useful when morphological characters, especially in females, are unclear (i.e., ovipositor lengths are intermediate between typical specimens). DNA sequencing of the barcode or some other region of the mitochondrial DNA (Hebert et al. 2004) would appear to be the logical choice for this task, but published work to date (Smith and Bush 1997, Schwarz et al. 2005, Michel et al. 2007) has not revealed unique, diagnostic haplotypes for R. pomonella and R. zephyria based on mitochondrial DNA sequences. It is unclear whether the inability of mtDNA to distinguish between R. pomonella and R. zephyria is due to insufficient mitochondrial DNA sequence data (more nucleotides need to be sequenced per individual) or if the history and structure of R. pomonella and R. zephyria populations is such that their mitochondrial DNA haplotypes are shared because of ongoing introgression and/or incomplete lineage sorting.

To date, the use of multilocus genotypes to identify unknown field-caught flies as either R. zephyria or R. pomonella has had the highest potential for success (Schwarz et al. 2005, Michel et al. 2007). Green et al. (2013) described a quantitative real-time PCR approach for species diagnosis in R. zephyria and R. pomonella based on four single nucleotide polymorphisms (SNPs) derived from expressed sequence tags (ESTs). While the assay developed by Green et al. (2013) was useful in most cases for distinguishing whether a fly originated from apple, black hawthorn (Crataegus douglasii Lindl.), or snowberry (S. albus var. laevigatus), their assay was complicated by the presence of R. zephyria alleles that presumably had introgressed into the R. pomonella gene pool in areas where the two species were in close proximity. More recently, Doellman et al. (2020) developed a diagnostic test based on PCR amplification of alleles at five nuclear loci followed by restriction enzyme digestion to reveal differences between field-caught R. pomonella and R. zephyria, that appears to be diagnostic but is somewhat technically demanding.

Fig. 1.

Diagram of the eukaryotic ribosomal DNA repeat unit (cistron), showing the location of the NTS area sequenced in this study. The complete repeating unit of rDNA is shown, showing the relative locations and sizes of the different components. This basic structure, with minor variation, is conserved across the eukaryotes. Our overall rDNA alignment is 9,221 nucleotide positions, which comprise the area outlined with the bold box. We determined sequences (approx. 800 bp) from the 3' end of the nontranscribed spacer (NTS; shaded area within bold box) within the IGS region for 141 R. pomonella, R. mendax, and R. zephyria flies. Figure adapted from Hwang and Kim (1999).

img-z2-11_647.jpg

Here we describe genetic variation in the rRNA gene cistron of R. pomonella and R. zephyria (Fig. 1) that provides an additional useful tool for distinguishing between individuals in these two fly species, either as field-caught adults or larvae. As a part of a larger study in our lab on species relationships in the genus Rhagoletis (Hulbert 2018), we examined molecular variation in a 625-nu-cleotide fragment of the D2 expansion segment region of the 28S rRNA gene. This study revealed a single nucleotide insertion mutation in the 28S D2 expansion region of R. zephyria individuals that appeared to distinguish these flies from R. pomonella. Subsequent assemblies of larger portions of the rDNA cistron of R. pomonella and R. zephyria using Illumina NextGen sequences (Smith et al. in preparation) indicated that there may be several DNA mutations, both single nucleotide substitutions and larger scale insertion–deletion (indel) mutations, upstream of the 18S structural rRNA gene, in both the external transcribed spacer (ETS) and nontranscribed spacer (NTS) regions of the rDNA cistron.

We report here genetic variation in a portion of the ribosomal NTS region that appears to be useful for distinguishing between R. pomonella and R. zephyria, with indel mutations that potentially provide the basis for a simple diagnostic test for these two species based on a 49 bp size difference in their PCR amplicons. The ribosomal region that we examined (the 3' end of the NTS region; Fig. 1) contains both DNA substitutions and indel variation that appear to be fixed between R. pomonella and R. zephyria. On the other hand, R. pomonella and R. mendax, the third named species in the pomonella group sibling species complex, appear to be polyphyletic with respect to each other based on the NTS gene region, while R. zephyria remains monophyletic with respect to both R. pomonella and R. mendax.

Materials and Methods

Taxon Sample

The taxon sample for this study consists of both a geographic sample of 101 flies collected from across the northern US, and a sample collected mainly from Beacon Rock State Park near the Columbia River in Washington State. The geographic sample consists of 101 individual flies (38 R. pomonella, 35 R. zephyria, and 26 R. mendax) from populations across their respective geographic ranges (Table 1; Fig. 2), plus two R. pomonella Mex. individuals from the Eje Volcanico Trans Mexicano (EVTM; Michel et al. 2007). Following Green et al. (2013), species names were assigned to individuals collected from apple or hawthorn as R. pomonella, individuals collected from snowberry as R. zephyria, and individuals collected from blueberry as R. mendax. This set of 101 flies includes 14 flies from which we have obtained approx. 9,000 bp sequences of the ribosomal DNA cistron (partial NTS, ETS, 18S, ITS1, 5.8S, ITS2, and 28S; GenBank Accession #'s MN507539–MN507552) and 14,000 bp of mitochondrial DNA (all 37 genes; GenBank Accession #'s MN443932–MN443945) via Illumina NextGen sequencing (Table 1; Smith et al. in preparation). The Beacon Rock sample includes 26 individual R. pomonella flies collected from C. douglasii and 10 individual R. zephyria flies collected from S.albus at Beacon Rock State Park (WA) in 2000, and also includes six individual R. zephyria flies collected nearby from S. albus at St. Cloud Ranch (WA) in 2000. Beacon Rock St. Park and St. Cloud Ranch correspond to sites 17 and 18, respectively, in Mattson et al. (2022).

Table 1.

Rhagoletis pomonella, R. zephyria, and R. mendax analyzed in this study, including collection information (locality, host plant, date)

img-AqBF_647.gif

Fig. 2.

Map showing collection locations of fly samples used in this study. Locations where fly samples were collected from hawthorns or apples (R. pomonella), from snowberries (R. zephyria), and from blueberries or huckleberries (R. mendax) are shown. The location where flies were collected from hawthorn in Mexico (R. pomonella Mex. EVTM) is not shown. Further information on each collection locality can be found in Table 1.

img-z4-1_647.jpg

PCR Primers

We designed PCR primers and developed PCR conditions that reliably amplify an approximately 800 bp portion of the NTS gene region of the rDNA cistron in R. pomonella, R. zephyria, and R. mendax individuals (Fig. 1). In preliminary work, examination of approx. 9,000 bp of the rDNA cistron from 14 individual Rhagoletis individuals (six R. zephyria, five R. pomonella, one R. mendax, and two R. pomonella Mex. flies; GenBank Accession #'s MN507539–MN 507552) revealed marked indel variation between the R. pomonella and R. zephyria individuals in the ETS and NTS regions. We examined this further by designing PCR primers that could amplify the DNA segments in which these indels occurred. Two primer pairs were designed using Primer-BLAST as implemented by the NCBI website (Ye et al. 2012). The first pair (PMZrDNA1F/PMZrDNA1R; Table 2) was designed to amplify approx. 800 bp of DNA located at the 3' end of the nontranscribed spacer (NTS) of the rDNA cistron. The second pair (PMZrDNA2F/PMZrDNA2R) was designed to amplify approx. 800 bp of DNA located at the 3' end of the external transcribed spacer (ETS) and the 5' end of the 18S rDNA structural gene. Examination of variation in these two rDNA regions across the 14 individual Rhagoletis individuals indicated that the NTS region amplified using PMZrDNA1F/PMZrDNA1R was more likely to provide a marked size difference in fragments between R. pomonella and R. zephyria individuals. All the work presented here is thus based on amplifications of the NTS region using the PMZrDNA1F/PMZrDNA1R primer pair.

The mitochondrial COI barcode primers of Hebert et al. (2004), LEP F1 and LEP R1 (Table 2), were used to amplify the mitochondrial COI region from the same set of flies from which we amplified the NTS and ETS regions. As observed by Barr et al. (2021) in European cherry fruit fly Rhagoletis cerasi (L.), the quality of some COI sequence reads from Rhagoletis flies in our sample was low, leading to ‘dirty’ sequences. In two cases, PCR of whole fly Rhagoletis extracts using LEP F1 and LEP R1 resulted in the amplification of COI fragments that were very clean upon sequencing, but yielded a DNA sequence that matched exactly the COI sequence from Wolbachia pipientis (GenBank Accession CP 042445.1). This phenomenon has been observed before (Smith et al. 2012), and we thus followed the recommendation of Barr et al. (2021) and used the primer pair TY-J-1460 and C1-N-2191 (Simon et al. 1994) for a number of subsequent COI amplifications, modifying the TY-J-1460 primer to match the known Rhagoletis sequence of the tRNATyr gene (TY-J-1460_Rhag; Table 2).

Table 2.

DNA sequences of PCR primers used and/or discussed in this study

img-z5-2_647.gif

PCR

All PCR amplification was carried out using a BioRad T100 Thermocycler in 25 µl reactions using GotaqFlexi (Promega, Madison, WI) with the following reagents (and concentrations): reaction buffer (1X), MgCl2 (8 mM), dNTP (0.5 mM each), forward and reverse primers (0.5 mM each), DNA polymerase (2.5 u), and DNA template (∼18–92 ng). The forward and reverse primers used in the PCR are listed in Table 2. Thermocycler conditions used to amplify the approx. 800 bp ribosomal NTS fragment using the PMZrDNA1F/PMZrDNA1R primer pair were as follows: initial denaturation was at 95°C for 3 min, followed by 35 cycles of denaturation at 95°C for 30 s, primer annealing at 58°C for 30 s, and product elongation at 72°C for 1 min (×35), followed by a final product elongation at 72°C for 5 min. To amplify the approx. 650 bp mitochondrial COI fragment using the LEP F1/LEP R1 primer pair, the conditions used were as described above except the annealing temperature was set at 46°C and the product elongation step was 2 min. These latter conditions were also used to amplify COI using the TY-J-1460_Rhag/C1-N-2191 primer pair.

Agarose Gel Electrophoresis

Agarose gel electrophoresis for determining if PCR was successful in preparation for DNA sequencing was carried out with a MiniOne gel apparatus using 1.5–2.0% (w/v) agarose in a gel volume of 15–20 ml. MiniOne gels were run in 1X TBE buffer for 15–20 min. For the higher resolution of PCR products in the 700–900 bp range, larger 2% agarose gels were used and run at 80 V for 4 hr. DNA bands were visualized using SYBR Green (Invitrogen S33102; Waltham, MA) with a UV transilluminator and photographed using a mobile phone. A 100 bp ladder (Gold Biotechnology D001-500; St. Louis, MO) was used to estimate band sizes.

DNA Sequencing and Sequence Alignment

PCR products for DNA sequencing were purified using QIAquick PCR Purification Kits (Qiagen; Hilden, Germany) according to the manufacturer's recommendations. Sanger sequencing was performed at the Michigan State University Research Technology Support Facility via BigDye Terminator Sequencing on an Applied Biosystems 3730xl DNA Analyzer (Foster City, CA, USA) using the PCR primers as sequencing primers. Base calls were made within MEGA (vers. 7.2; Kumar et al. 2016) based on the sequences of both the forward and reverse strands. DNA nucleotide substitutions observed within the data sets were verified by visual examination of .ab1 file tracings within MEGA. DNA sequence alignment of the NTS amplicons was carried out using a TCS alignment (Chang et al. 2015) of 14 complete ribosomal DNA cistron sequences from R. pomonella, R. zephyria, and R. mendax (described above) as a scaffold. All of the NTS sequences generated for this study via Sanger sequencing readily and unambiguously aligned to this scaffold. Mitochondrial COI sequences obtained for this study readily aligned to a set of published COI sequences from R. pomonella and R. mendax (Frey et al. 2013).

Phylogenetic Analysis and Network Construction

Phylogenetic analysis of the aligned NTS sequences was carried out via the neighbor-joining algorithm of Saitou and Nei (1987) as implemented in PAUP*4.0 Version 4.0a (build 167) for Macintosh (X86) (Swofford 2000). The minimum evolution criterion was used as this has been shown to be among the most efficient methods when sequences are closely related (Nei 1991). The NTS sequences from two R. pomonella Mex. EVTM were used as the outgroup. Neighbor-joining topologies were examined in Parsimony mode in PAUP to identify unique alleles for generating input files for minimum-spanning network construction. Character changes on branches were identified in PAUP using ‘Show Reconstructions’. Minimum spanning networks (Bandelt et al. 1999) were generated using the computer program PopArt ( http://popart.otago.ac.nz) with default settings.

Results

Preliminary PCR of the NTS/ETS region of the ribosomal DNA with both primer pairs PMZrDNA1F/PMZrDNA1R and PMZrDNA2F/PMZrDNA2R using a selected number of R. pomonella and R. zephyria individuals to provide template DNA led to amplification of the predicted NTS and ETS regions of the rDNA cistron sequences that had been obtained by Illumina NextGen sequencing (GenBank Accession #'s MN507539–MN507552). Because indel variation in the NTS sequences appeared to provide an opportunity to distinguish between R. pomonella and R. zephyria based on PCR product length differences, we chose to focus our work on the NTS sequences obtainable using the primer pair PMZrDNA1F/PMZrDNA1R.

NTS Sequences From the Geographic Sample

PCR-amplification using the primer pair PMZrDNA1F/PMZrDNA1R led to a PCR product of approx. 840 bp in the R. pomonella and R. mendax individuals tested from the geographic sample (n = 64), and approx. 790 bp in the R. zephyria individuals from the geographic sample (n = 35), based on visualization of fragments via agarose gel electrophoresis (Fig. 3) and the size of the sequenced PCR amplicons (see below).

We obtained DNA sequences from the amplified NTS region from 38 R. pomonella, 35 R. zephyria, and 26 R. mendax individuals in the geographic sample. DNA sequencing of each PCR product indicated that these PCR products were 839 bp in R. pomonella and R. mendax, and 792 bp in R. zephyria, respectively, with minor size variation within species due to the presence of short 1–2 bp indels in some sequences. Base calling and alignment of the sequences were straightforward and unambiguous, resulting in 766 aligned nucleotide positions, which is included in .fas format in  Supp File 1a (toab264_suppl_supplementary_file_1a.txt) (online only).

The PCR products obtained with primers PMZrDNA1F and PMZrDNA1R from all flies studied here yielded DNA sequences that were very clean. There was no evidence of PCR product size heterogeneity, as would be evidenced by sequences that become unreadable at a particular point. This would appear to indicate that the NTS indel structure did not vary within individuals for this PCR amplicon, and that the ribosomal DNA cistrons are well homogenized within the three species examined here. On the other hand, we did observe what could be interpreted as evidence of DNA substitution heterozygosity within R. mendax: some changes observed within the R. mendax individuals appeared to involve heterozygous states (double peaks; state assigned as the higher of the two peaks). However, no double peaks were observed in the NTS sequences from R. pomonella and R. zephyria individuals; all of the base calls were clean and we had no difficulty assigning the character states within these sequences.

The overall size difference of the NTS fragment between R. pomonella and R. mendax individuals compared to R. zephyria individuals is accounted for by the presence of a characteristic set of indels in the NTS alignment, each of which appears to have been uniquely derived (relative to the R. pomonella Mex. EVTM outgroup) in either all R. zephyria individuals, or in all R. pomonella and R. mendax individuals (Table 3). For example, all R. zephyria individuals examined contained a single deletion of 15 nucleotides from positions 318–332 in the NTS alignment relative to R. pomonella and R. mendax (Fig. 4). Conversely, all R. pomonella and R. mendax individuals contain a 20-nu-cleotide insertion from positions 352–371. Consideration of all of the indels in the PCR amplicon corresponding to the NTS region examined accounts for the size difference of 49 nucleotides between the ribosomal NTS PCR products obtained from R. pomonella and R. mendax individuals, compared to those obtained from R. zephyria individuals.

Fig. 3.

Agarose gel showing the NTS size difference between the PCR products obtained from R. pomonella, R. mendax, and R. zephyria individuals. PCR products were obtained using primer pair PMZrDNA1F and PMZrDNA1R. Lanes: MW, 100 bp ladder; Lanes 1–4, PNY 12–15; Lanes 5–7, ZWI 12, 15, and 16; Lanes 8–11 MEL 11 16–18. Agarose gel (2%) run for 4 hr at 80 V. Stained with SYBR Green.

img-Aa9_647.jpg

We also observed single-nucleotide substitution variation within the R. pomonella, R. mendax, and R. zephyria individuals of the geographic sample at 12 nucleotide positions in the aligned data, with 11 of these being phylogenetically informative. A minimum evolution neighbor-joining tree based only on DNA substitution mutations shows the relationships of the NTS sequences obtained from the 38 R. pomonella, 35 R. zephyria, and 26 R. mendax individuals in the geographic sample (Fig. 5). Even though DNA substitution variation was slight, the NTS sequences from the 35 R. zephyria individuals formed a monophyletic group supported by a bootstrap value of 96. Three of the observed DNA substitutions (position 4, A → T; position 41, G → C; position 113, A → G) are uniquely derived in all of the R. zephyria individuals, and only in R. zephyria individuals, providing what appear to be diagnostic nucleotide substitutions (character states) for R. zephyria, above and beyond the set of indel mutations described above. Within the R. zephyria cluster, there is also a group of 9 sequences, also with strong bootstrap support, that is mainly composed of R. zephyria individuals from the eastern U.S. and differs from the other R. zephyria sequences by a single nucleotide change at position 696 in the alignment. While R. zephyria was monophyletic in the NTS tree, R. pomonella and R. mendax were not, appearing polyphyletic with respect to each other in the neighbor-joining tree (Fig. 5). Three R. mendax individuals (MJP5, MEL8, and MEL20) share the sequence that is most common for R. pomonella, while one R. pomonella individual (PNY7) shares its sequence with four R. mendax individuals.

Using the neighbor-joining tree as a basis, we defined 15 alleles (genotypes) within the NTS data set. Construction of a minimum-spanning network (Fig. 6) provides a visual representation of the relationships of these 15 alleles and their relative abundance within each of the three ingroup species. Allelic diversity at the NTS locus was higher in R. mendax than it was in either R. zephyria or R. mendax. While R. zephyria and R. pomonella had three alleles each, R. mendax had nine, despite the fact that there were fewer R. mendax individuals in the sample (n = 26) than R. zephyria (n = 35) or R. pomonella (n = 38). All of the R. zephyria individuals had one of three alleles (Hap11, Hap12, Hap13), and each of these is found in R. zephyria individuals only. R. pomonella individuals also had one of three alleles, with R. pomonella and R. mendax sharing two of these alleles (Hap1 and Hap5). Hap1 was the most frequent allele in the data set, being found in 36 R. pomonella individuals and three R. mendax individuals.

Mitochondrial COI Sequences From the Geographic Sample

A different picture emerged when the geographic sample was analyzed using the mitochondrial COI barcode region. Mitochondrial COI sequences were obtained from the same set of individuals in the Geographic sample from which NTS sequences were obtained. Base calling and alignment of the sequences again were straightforward, resulting in 633 aligned nucleotide positions corresponding to the BOLD barcode region (Ratnasingham and Hebert 2007). The complete set of COI sequences is included in .fas format in  Supp File 1b (toab264_suppl_supplementary_file_1b.txt) (online only). As was observed in the NTS data set, DNA sequence variation was slight within the COI data set, with observed variation at 14 nucleotide positions within the ingroup, and 8 of these being phylogenetically informative. We included in the COI data set six R. mendax and R. pomonella COI sequences from the literature (Frey et al. 2013) and we observed COI sequences in the geographic sample that matched each of these. Rhagoletis zephyria is polyphyletic in the minimum evolution neighbor-joining COI tree with respect to both R. pomonella and R. mendax (Fig. 7). Three R. zephyria individuals (ZWI10, ZPA2, and ZPA10) appear in a cluster that contains predominantly R. mendax and R. pomonella, while two R. pomonella individuals (PNY5 and PWA19) are grouped with the majority of the R. zephyria individuals in the Geographic sample. While the neighbor-joining tree shows some apparent structure in the R. pomonella, R. mendax, and R. zephyria individuals included in the geographic sample, bootstrap support overall was very weak, with values >50 obtained for only two clusters, one containing three R. mendax individuals and the other containing two R. pomonella individuals.

Table 3.

Insertion/deletion mutations in the NTS region that distinguish between R. pomonella and R. zephyria

img-z7-6_647.gif

Fig. 4.

Selected aligned NTS DNA sequences from the R. pomonella/R. mendax/R. zephyria taxon sets showing consistent differences between R. pomonella and R. zephyria. Shown are three NTS sequences from R. pomonella (PNY2, PIL4, PCO1), three from R. zephyria, (ZPA10, ZWI10, ZCO16), and two from R. mendax (MWB9, MJP5) from the Geographic sample, plus one sequence from hawthorn (MK92_haw) and one from snowberry (MK53_snow) from the Beacon Rock sample, and two sequences (PMEX4n, PMEX7n) from the R. pom. Mex. outgroup. A) Positions 309–423 in the 766 bp NTS alignment. This segment of the alignment shows a set of indels that appear to be diagnostic between R. pomonella and R. zephyria (R. mendax shares the R. pomonella indels in all instances). B) Positions 40–154 in the alignment. This part of the alignment has four single nucleotide substitution mutations (41, 52, 119, and 152) that appear to be derived (based on outgroup comparison) in R. zephyria (positions 41 and 113) or R. pomonella and R. mendax (positions 119 and 152). MEL 25 and MJP2 share the ancestral state at position 119, and MEL25, MJP2, MEL19, and MEL29 share the ancestral state at position 152, with all R. zephyria. Variation at position 67 is homoplasious.

img-z7-10_647.jpg

Using the neighbor-joining tree as a basis, we defined 17 COI haplotypes. Construction of a minimum-spanning network (Fig. 8) shows extensive COI haplotype sharing between R. pomonella, R. mendax, and R. zephyria, with none of the three ingroup species appearing as a monophyletic group. While most of the R. zephyria individuals have ‘private’ haplotypes (Haplotypes 8, 11, or 12), some R. zephyria individuals share haplotypes with R. pomonella (Haplotype 7), R. mendax (Haplotype 2), or both (Haplotype 1). Unlike the NTS locus, where allele diversity was higher in R. mendax than it was in R. zephyria or R. mendax., mitochondrial COI haplotype diversity is highest in R. pomonella, which has 10 haplotypes while R. mendax has six and R. zephyria has five. At the NTS locus, R. pomonella individuals only had one of three closely related alleles, while at the COI locus the ten haplotypes found in R. pomonella individuals are spread across the haplotype network.

Fig. 5.

Neighbor joining tree showing relationships of 766 bp NTS sequences obtained from R. pomonella, R. zephyria, and R. mendax individuals in the Geographic Sample. Sequences drawn from R. pomonella, “P”; R. mendax, “M”; R. zephyria, “Z”; and R. pomonella Mex. EVTM (PMex; outgroup), are shown. Numbers on branches are bootstrap values (1,000 replicates). Scale bar = p-distance.

img-z8-1_647.jpg

NTS and COI Sequences From the Beacon Rock Sample

Incorporation into the analysis of a set of 27 individual Rhagoletis flies reared from black hawthorn and 17 individuals reared from snowberry from Beacon Rock State Park and St. Cloud Ranch (the Beacon Rock sample;  Supp Table S1 (toab264_suppl_supplementary_table_s1.docx) [online only]) served as a preliminary ‘field-test’ of the ability of the NTS PCR product differences to serve as a diagnostic for identification purposes in areas where R. pomonella and R. zephyria co-exist. All of the flies in the Beacon Rock sample were collected as larvae within infested fruit in 2000 by one of the authors (MK), sent to another of the authors (JJS), overwintered as pupae and reared to adults in the lab (with emerged adults randomly labeled to mask host of origin) and then sent back to the Washington State Department of Agriculture (WSDA) for species identification using morphological characteristics. Locality information and morphological scoring data for these flies is shown in  Supp Table S1 (toab264_suppl_supplementary_table_s1.docx) (online only). All 12 of the males reared from black hawthorn were identified as R. pomonella based on morphological criteria, while 6 of the 15 females reared from black hawthorn were identified as R. pomonella and the other 9 were identified as R. zephyria (based on ovipositor length). On the other hand, all 17 of the flies (9 females and 8 males) reared from snowberry in the Beacon Rock sample were identified as R. zephyria based on morphological criteria.

Fig. 6.

Minimum spanning network showing relationships of rDNA NTS sequences drawn from R. pomonella, R. mendax, and R. zephyria. The snowberry fly, R. zephyria, is distinguished from R. mendax and R. pomonella by four nucleotide substitutions, while R. pomonella and R. mendax share alleles (1 and 5).

img-z9-1_647.jpg

When we sequenced the PMZrDNA1F/PMZrDNA1R PCR product from these flies, all 26 NTS sequences obtained from individuals from Beacon Rock that were reared from black hawthorn had the unique set of NTS indels and DNA substitutions (described above) that were associated with the R. pomonella flies in the geographic sample, while all 16 NTS sequences of individuals from Beacon Rock and St. Cloud Ranch that were reared from snowberry had the unique set of NTS indels and DNA substitutions that were associated with all of the R. zephyria flies in the geographic sample ( Supp Figs. S1 (toab264_suppl_supplementary_figure_s1.pdf) and  S2 (toab264_suppl_supplementary_figure_s2.pdf) [online only]). Thus, the 49 bp PCR-amplicon length difference that was observed between R. pomonella and R. zephyria individuals in the geographic sample was also observed in the flies reared from black hawthorns (R. pomonella) and flies reared from snowberries (R. zephyria) in the Beacon Rock sample.

Fig. 7.

Neighbor-joining tree showing relationships of 633 bp mitochondrial COI sequences obtained from R. pomonella, R. zephyria, and R. mendax individuals in the Geographic Sample. Sequences drawn from R. pomonella individuals, “P”; R. mendax, “M"; R. zephyria, “Z". R. pomonella Mex. EVTM is shown as the outgroup (R. pom Mex.). Numbers on branches are bootstrap values (500 replicates). Scale bar = p-distance.

img-z10-1_647.jpg

A neighbor-joining tree with the combined Geographic sample and Beacon Rock sample is shown as  Supp Fig. S1 (toab264_suppl_supplementary_figure_s1.pdf) (online only), and the NTS allele network including flies from both the Geographic sample and the Beacon Rock sample is shown as  Supp Fig. S2 (toab264_suppl_supplementary_figure_s2.pdf) (online only). All of the individuals reared from snowberry in the Beacon Rock sample clustered with R. zephyria in the NTS NJ tree, while all of the individuals reared from black hawthorn in the Beacon Rock sample clustered with R. pomonella and R. mendaxSupp Fig. 1 (toab264_suppl_supplementary_figure_s1.pdf) [online only]). There were no unique NTS sequences (alleles) observed in the Beacon Rock sample ( Supp Fig. 2 (toab264_suppl_supplementary_figure_s2.pdf) [online only]). Twenty-five of the 26 individuals reared from black hawthorn in the Beacon Rock sample carried NTS allele 1, while one individual (MK34) carried allele 9, which is shared with two R. mendax individuals (MEL19 and MEL29). Conversely, all of the 16 individuals reared from snowberry from Beacon Rock State Park and St. Cloud Ranch carried NTS allele 13, which was the main allele exclusively associated with R. zephyria in the geographic sample.

Fig. 8.

Haplotype network showing relationships of mitochondrial COI sequences drawn from R. pomonella, R. mendax, and R. zephyria in the Geographic sample.Tick marks represent individual DNA substitution mutations between haplotypes. While most R. zephyria flies have Haplotype 8, which is private, other R. zephyria flies share haplotypes with both R. pomonella and R. mendax individuals elsewhere in the network.

img-z11-1_647.jpg

The COI neighbor-joining tree based on the combined Geographic and Beacon Rock samples is shown as  Supp Fig. S3 (toab264_suppl_supplementary_figure_s3.pdf) (online only), and the COI haplotype network including flies from both the Geographic sample and the Beacon Rock sample is shown as  Supp Fig. S4 (toab264_suppl_supplementary_figure_s4.pdf) (online only). No new COI sequences were observed when the Beacon Rock sample was added to the analysis, and the neighbor-joining tree shows that the COI sequences from the flies reared from black hawthorn and snowberry were not strictly associated with COI sequences drawn from R. pomonella and R. zephyria, respectively ( Supp Fig. S3 (toab264_suppl_supplementary_figure_s3.pdf) [online only]). While most Beacon Rock sample flies from snowberry (12/16) carried COI haplotype 8, MK97, which was reared from snowberry and identified as R. zephyria based on morphology, carried COI haplotype 1, which was predominantly associated with R. pomonella and R. mendax (ZWI10 also had COI haplotype 1). Similarly, Beacon Rock sample flies reared from black hawthorn carried four haplotypes. While most of these flies carried COI haplotype 1 (16/26), others showed COI haplotypes that were more closely related to those carried mainly by R. zephyria flies in the geographic sample (e.g., haplotype 10;  Supp Figs. S3 (toab264_suppl_supplementary_figure_s3.pdf) and  S4 (toab264_suppl_supplementary_figure_s4.pdf) [online only]).

Discussion

NTS Sequences Distinguish R. pomonella From R. zephyria

All of the R. zephyria individuals analyzed in this study formed a monophyletic group based on DNA sequences of an 800 bp PCR product amplified from the nuclear rDNA NTS region ( Supp Fig. 1 (toab264_suppl_supplementary_figure_s1.pdf) [online only]). This monophyletic group was supported unambiguously by three DNA substitutions in the 766 bp NTS alignment (positions 4, 41, and 113). In addition, R. zephyria was supported as a monophyletic group by a set of four unique insertion–deletion mutations within the alignment (Table 3), which serve as synapomorphies defining this clade, using R. pomonella Mex. EVTM as the outgroup, and appear to unambiguously distinguish R. zephyria from R. pomonella and R. mendax.

The indel structure occurring between positions 318 and 371 in the alignment contributes the bulk of the size difference of 49 nucleotides in the PCR products obtained from R. zephyria individuals compared to the PCR products obtained from R. pomonella and R. mendax individuals (Fig. 4). This DNA size fragment difference between R. zephyria, R. pomonella, and R. mendax individuals is readily apparent on agarose gels of the PCR products from these three different fly species (Fig. 2), and provides the potential for the development of a rapid, simple, inexpensive agarose gel-based test to use in efforts to distinguish between R. zephyria and R. pomonella in the apple-growing regions of the Pacific Northwest. A method based on these NTS differences would allow discrimination between R. pomonella and R. zephyria simply by scoring size differences between amplified PCR products via agarose gel electrophoresis.

The availability of the rDNA NTS region adjacent to the ETS as a molecular marker will complement and extend other tools available to distinguish R. pomonella from R. zephyria for quarantine purposes. Traditionally, adults obtained via trap catches in apple maggot surveys have been analyzed for morphological differences in ovipositor length and surstylus shape (Westcott 1982, Yee et al. 2011). The PCR assay described here based on size differences in the NTS PCR amplicon would be particularly useful for individuals in the ‘grey area’ in these analyses. The R. pomonella individuals and R. zephyria individuals from Beacon Rock State Park and St. Cloud Ranch analyzed in this paper were scored by morphology in the WSDA lab ( Supp Table 1 (toab264_suppl_supplementary_table_s1.docx) [online only]), which resulted in scoring nine R. pomonella females as R. zephyria (ovipositor length below threshold) as judged by the NTS sequence data. While the advanced morphometric techniques now available based on wing shape and aculeus shape (Yee et al. 2009, 2011) would likely correctly identify these flies as R. pomonella, all of the flies reared from black hawthorn in the Beacon Rock sample carry the set of diagnostic DNA nucleotide substitutions and indel mutations characteristic of R. pomonella and not R. zephyria.

Green et al. (2013), and more recently Doellman et al. (2020), described SNP-based diagnostic tools that would also identify all individuals in this study to their correct species. The method described by Doellman et al. (2020) is based on PCR-amplification and restriction digestion of PCR-amplified products at five nuclear loci. While accurate, the method is potentially more time and labor intensive than the assay described here. Agarose gel separation of DNA fragments following PCR amplifications is standard, and the possibility of errors associated with a restriction digestion step would be avoided in an NTS-based assay. No DNA sequencing would be required for distinguishing between R. pomonella and R. zephyria, and the known sequence differences between the two species provide opportunity for the development of a qPCR assay.

While successful PCR-amplification is required for the NTS assay to distinguish between R. zephyria, R. pomonella, and R. mendax, no further processing is necessary beyond running and visualizing the agarose gel. The primers designed for use in the current study (PMZrDNA1F and PMZrDNA1R), and the PCR conditions used, appear to be specific with respect to amplification of NTS only from R. zephyria, R. pomonella, and R. mendax. However, one shortcoming of the current study is that the specificity of these primers beyond R. zephyria, R. pomonella, and R. mendax has not been determined. There is one mismatch between the sequence of the forward primer PMZrDNA1F and the corresponding sequence in R. pomonella Mex. EVTM, and either four mismatches (R. tabellaria (Fitch) and R. electromorpha Berlocher) or five (R. persimilis Bush) in PMZrDNA1F in Rhagoletis tabellaria species group flies ( Supp Table 2 (toab264_suppl_supplementary_table_s2.docx) [online only]), which is a sister group to the R. pomonella species group within the genus Rhagoletis (Hulbert 2018). Sequences corresponding to the PMZrDNA1F primer have not yet been found in either R. cornivora Bush or R. bushi Hulbert and Smith. Whether the PMZrDNA1F and PMZrDNA1R primer pair can be used to amplify the NTS region in R. pomonella Mex. EVTM, R. cornivora, and the four R. tabellaria group species, and under what conditions, remain to be empirically determined.

NTS as a Molecular Marker

One significant aspect of the current study is the possible use of the NTS region of the ribosomal DNA just upstream from the ETS to explore systematics and evolutionary questions in other insect groups. The NTS region may prove to be particularly useful in circumstances like R. pomonella and R. zephyria, where closely related and/or cryptic species present taxonomic problems, and positive species ID is necessary for pest regulatory purposes. However, the usefulness of NTS as a marker may be limited by the history of gene flow between the species examined. While NTS appears to have fixed differences between R. zephyria and both R. pomonella and R. mendax, the NTS region is not diagnostic for these latter two sister species, based either on indel variation or DNA substitutions. Until the introduction of R. pomonella into western North America, R. zephyria had been geographically separated from R. pomonella except for along a narrow band of overlap from Minnesota into the eastern U.S. (Gavrilovic et al. 2007, Arcella et al. 2015). This historical lack of gene flow may have been necessary for the NTS divergence observed in this study to have occurred. However, there has been some gene flow between R. mendax and R. zephyria where they overlap in eastern North America (Schwarz et al. 2005), and NTS remained diagnostic in eastern North American samples of those species examined here. On the other hand, R. mendax and R. pomonella are broadly sympatric in eastern North America, with ample opportunities for gene flow between them, and the NTS region is not diagnostic for these two sister species, either based on indel variation or DNA substitutions.

Nonetheless, analysis of NTS regions may be useful for addressing questions in instances where there are shared mitochondrial DNA haplotypes across species boundaries (e.g., the mealybugs Planococcus citri Russo (Hemiptera: Pseudococcidae) and P. minor (Maskell), Rung et al. 2009; several species of Neodiprion sawflies (Hymenoptera: Diprionidae), Linnen and Farrell 2007; the blow flies Lucilia coeruleiviridis Macquart (Diptera: Calliphoridae) and Lucilia Mexicana Macquart, Debry et al. 2013), and in insect taxa where two or more divergent mitochondrial lineages exist within a single species (e.g., the tree weta Hemideina crassidens (Blanchard) (Orthoptera: Anostostomatidae), Morgan-Richards et al. 2017).

While the 18S and 28S ribosomal RNA structural genes, and the internal transcribed spacers (ITS1 and ITS2), of the ribosomal DNA (rDNA) cistron have been used extensively for phylogenetics and population-level studies (the review of Hillis and Dixon (1991) has been cited over 2000 times), the external transcribed spacer (ETS) and nontranscribed spacer (NTS) regions, which together comprise the ribosomal intergenic spacer (IGS), have been relatively underutilized in systematics and species identification in insects, and exploration of these uses may prove fruitful. Shaikevich et al. (2013) described IGS sequences from mosquitoes (Culex spp.; Diptera: Culicidae) and indicated that these might prove useful for population genetics and phylogenetic analyses, but this study has been cited only once in the literature since publication (Web of Science; accessed 22 June 2021). On the other hand, the NTS and ETS gene regions seem to have found some use in phylogenetics studies, mainly in plants, with most of these studies employing the ETS region (e.g., Linder et al. 2000, Valtuena et al. 2017).

For reasons that are not entirely clear, the work of Shaikevich et al. (2013), who showed that the IGS regions of Culex species provided promising molecular markers for population genetics studies and phylogenetic analysis, has not seen broad applicability. This may be, in part, because the ribosomal IGS regions are not as easy to use as the 18S, 28S, ITS1, and ITS2 sequences, which are readily accessible and easy to amplify via PCR, and have proven to be successful in a range of phylogenetics studies (e.g., Si et al. 2020). In addition, because they are not transcribed, the NTS region will not be included in results obtained using RNASeq protocols. Our work allowed us to find the NTS regions by virtue of the use of whole genome NextGen sequencing, originally targeting high coverage sequences to assemble nearly complete mtDNA sequences (Smith et al. in preparation). The IGS regions also appear to be fast evolving (Stage and Eickbush 2007), so they are not readily obtained via homology searches. The DNA sequences of the NTS regions analyzed in this study from R. pomonella, R. mendax, and R. zephyria are apparently unique: BLAST searches of GenBank using the NTS PCR product sequences did not return any matches. Nonetheless, we were able to align our NTS sequences, at least partially, to the NTS sequence from Bactrocera tryoni (Froggatt) (Diptera: Tephritidae) (Gilchrist et al. 2014) and used this as the basis for hypothesizing the position of the ribosomal binding site.

The data presented here also illustrate why mtDNA has not proven useful for species diagnostics in R. pomonella and R. zephyria, as there are shared COI mtDNA haplotypes between these two species (Fig. 8). One possible explanation for this observation is that we simply have not sequenced enough mtDNA characters; the dataset here includes only 633 of the mitochondrion's approximately 15,000 bp. However, NextGen sequences of complete coding regions of the mtDNA (37 genes, 14,942 bp) from five R. pomonella, six R. zephyria, and one R. mendax individuals included in this study (GenBank Accession #'s MN443934–MN443945) did not provide species-level resolution (Smith et al. in preparation); instead, we observed what appears to be cytonuclear discordance between R. zephyria and R. mendax in the eastern U.S. This discordance presumably arises via introgression of R. mendax mtDNA into R. zephyria, which was described by Schwarz et al. (2005) in their report of the ‘speciation by hybridization’ of an (undescribed) Rhagoletis species, the Lonicera fly. The extent to which these introgressions have occurred (and are occurring) in R. pomonella, R. mendax, and R. zephyria is unclear and further work remains to be done to provide a clearer picture of these dynamics.

Future Work

Another shortcoming of the current study is that we have not determined the extent of fixation of the IGS differences in a large sample such as was employed by Doellman et al. (2020). In our study, we analyzed 38 R. pomonella, 26 R. mendax, and 35 R. zephyria in the geographic sample, and an additional 27 R. pomonella and 17 R. zephyria in the Beacon Rock sample. We also have not yet analyzed F1 hybrids between R. pomonella and R. zephyria (Yee and Goughnour 2011, Yee et al. 2013), or backcross individuals. Work to expand the sample size and experiments to determine the outcomes given hybrid and introgressed individuals remains to be carried out. Results of such studies may inform questions about the inheritance of rDNA cistrons and the homogenization of the different regions within them.

We also did not take into consideration or try to determine the nature of the variation leading to the dual NTS sequence signals observed at some nucleotide positions in some R. mendax sequences. It is unclear whether these dual signals occurred because of variation within the multiple copies of the ribosomal DNA cistron within the same haploid genome or if there are differences at some positions due to heterozygosity. In either case, we did not observe any differences in the indel patterns within any individual sequences, which would have been observable as sequences that fell out of register from each other at a certain point in a particular read. Because R. mendax was not a major focus of this study, and because the sequences in question did not impact the finding that R. zephyria is diagnosable based on its indel pattern, we did not pursue these lines of inquiry here and leave them to a future study.

Supplementary Data

Supplementary data are available at Journal of Economic Entomology online.

Acknowledgments

We thank Dr. Daniel L. Hulbert for designing the primers used in this study. We thank Dr. Wee L. Yee and Dr. Ron Debry for providing valuable feedback and comments on an earlier version of the manuscript. We also thank Dr. Pat Edger for co-teaching, and the Lyman Briggs College for funding, the LB348 undergraduate research course (CURE) in which the NGS data that provided the basis of the work done in this publication were produced.

References Cited

1.

AliNiazee, M. T., and R. L. Penrose. 1981. Apple maggot in Oregon: a possible new threat to the northwest apple industry. Bull. Entomol. Soc. Am. 27: 245–246. Google Scholar

2.

Arcella, T., G. R. Hood, T. H. Powell, S. B. Sim, W. L. Yee, D. Schwarz, S. P. Egan, R. B. Goughnour, J. J. Smith, and J. L. Feder. 2015. Hybridization and the spread of the apple maggot fly, Rhagoletis pomonella (Diptera: Tephritidae), in the northwestern United States. Evol. Appl. 8: 834–846. Google Scholar

3.

Bandelt, H. J., P. Forster, and A. Röhl. 1999. Median-joining networks for inferring intraspecific phylogenies. Mol. Biol. Evol. 16: 37–48. Google Scholar

4.

Barr, N. B., D. Garza, L. A. Ledezma, and D. A. Salinas. 2021. Using the rDNA internal transcribed spacer 1 to identify the invasive pest Rhagoletis cerasi (Diptera: Tephritidae) in North America. J. Econ. Entomol. 114: 360–370. Google Scholar

5.

Bush, G. L. 1966. The taxonomy, cytology and evolution of the genus Rhagoletis in North America (Diptera, Tephritidae). Bull. Mus. Comp. Zool. 134: 431–562. Google Scholar

6.

Bush, G. L. 1969. Sympatric host race formation and speciation in frugivorous flies of the genus Rhagoletis (Diptera, Tephritidae). Evolution. 23: 237–251. Google Scholar

7.

Bush, M. R., M. Klaus, A. Antonelli, and C. Daniels. 2002. Protecting backyard apple trees from apple Maggot. Washington State University Extension Bulletin #1928, Pullman, Washington. Google Scholar

8.

CFIA. 2017. RMD-16-01: pest risk management document—determining the regulatory status of Rhagoletis pomonella (apple maggot). Government of Canada, Canadian Food Inspection Agency, Ottawa, ON, Canada. Google Scholar

9.

Chang, J. M., P. Di Tommaso, V. Lefort, O. Gascuel, and C. Notredame. 2015. TCS: a web server for multiple sequence alignment evaluation and phylogenetic reconstruction. Nucleic Acids Res. 43: W3–W6. Google Scholar

10.

Curran, C. H. 1924. Rhagoletis symphoricarpi, a new trypaneid from British Columbia (Dipt.). Can. Ent. 61: 62–63. Google Scholar

11.

DeBry, R. W., A. Timm, E. S. Wong, T. Stamper, C. Cookman, and G. A. Dahlem. 2013. DNA-based identification of forensically important Lucilia (Diptera: Calliphoridae) in the continental United States. J. Forensic Sci. 58: 73–78. Google Scholar

12.

Doellman, M. M., G. J. Ragland, G. R. Hood, P. J. Meyers, S. P. Egan, T. H. Q. Powell, P. Lazorchak, M. M. Glover, C. Tait, H. Schuler, et al. 2018. Genomic differentiation during speciation-with-gene-flow: comparing geographic and host-related variation in divergent life history adaptation in Rhagoletis pomonella. Genes. 9: 262. Google Scholar

13.

Doellman, M. M., S. P. Egan, G. J. Ragland, P. J. Meyers, G. R. Hood, T. H. Q. Powell, P. Lazorchak, D. A. Hahn, S. H. Berlocher, P. Nosil, et al. 2019. Standing geographic variation in eclosion time and the genomics of host race formation in Rhagoletis pomonella fruit flies. Ecol. Evol. 9: 393–409. Google Scholar

14.

Doellman, M. M., G. R. Hood, J. Gersfeld, A. Driscoe, C. C. Y. Xu, R. N. Sheehy, N. Holmes, W. L. Yee, J. L. Feder. 2020. Identifying diagnostic genetic markers for a cryptic invasive agricultural pest: a test case using the apple maggot fly (Diptera: Tephritidae). Ann. Entomol. Soc. Am. 113: 246–256.  https://doi-org.proxy2.cl.msu.edu/10.1093/aesa/saz069Google Scholar

15.

Dowell, R. N., L. T. Wilson, and V. P. Jones, (eds). 1990. Apple Maggot in the west: history, biology, and control. University of California (System). Division of Agriculture and Natural Resources, Publ. 3341. Google Scholar

16.

Feder, J. L., C. A. Chilcote, and G. L. Bush. 1988. Genetic differentiation between sympatric host races of the apple maggot fly Rhagoletis pomonella. Nature. 336: 61–64. Google Scholar

17.

Folmer, O., M. Black, W. Hoeh, R. Lutz, and R. Vrijenhoek. 1994. DNA primers for amplification of mitochondrial cytochrome c oxidase I from diverse metazoan invertebrates. Mol. Mar. Biol. Biotechnol. 3: 294–299. Google Scholar

18.

Frey, J. E., L. Guillén, B. Frey, J. Samietz, J. Rull, and M. Aluja. 2013. Developing diagnostic SNP panels for the identification of true fruit flies (Diptera: Tephritidae) within the limits of COI-based species delimitation. BMC Evol. Biol. 13: 106. Google Scholar

19.

Galinato, S. P., R. K. Gallardo, D. M. Granatstein, and M. Willett. 2018. Economic impact of a potential expansion of pest infestation: Apple Maggot in Washington State. HortTechnology. 28: 651–659. Google Scholar

20.

Gavrilovic, V., G. L. Bush, D. Schwarz, and J. J. Smith. 2007. Rhagoletis zephyria Snow (Diptera: Tephritidae) in the Great Lakes basin: a native insect on native hosts? Ann. Entomol. Soc. Am. 100: 474–482. Google Scholar

21.

Gilchrist, A. S., D. C. Shearman, M. Frommer, K. A. Raphael, N. P. Deshpande, M. R. Wilkins, W. B. Sherwin, and J. A. Sved. 2014. The draft genome of the pest tephritid fruit fly Bactrocera tryoni: resources for the genomic analysis of hybridising species. BMC Genomics. 15: 1153. Google Scholar

22.

Green, E., K. Almskaar, S. B. Sim, T. Arcella, W. L. Yee, J. L. Feder, and D. Schwarz. 2013. Molecular species identification of cryptic apple and snowberry maggots (Diptera: Tephritidae) in western and central Washington. Environ. Entomol. 42: 1100–1109. Google Scholar

23.

Hebert, P. D., E. H. Penton, J. M. Burns, D. H. Janzen, and W. Hallwachs. 2004. Ten species in one: DNA barcoding reveals cryptic species in the neotropical skipper butterfly Astraptes fulgerator. Proc. Natl. Acad. Sci. U. S. A. 101: 14812–14817. Google Scholar

24.

Hillis, D. M., and M. T. Dixon. 1991. Ribosomal DNA: molecular evolution and phylogenetic inference. Q. Rev. Biol. 66: 411–453. Google Scholar

25.

Hulbert, D. L. 2018. Systematics of the genus Rhagoletis (Diptera: Tephritidae): New species, phylogeny, and justifications. Unpublished Ph. D. Dissertation, Michigan State University, East Lansing, MI. Google Scholar

26.

Hulbert, D., M. D. Jackson, G. R. Hood, and J. J. Smith. 2018. Description of a new Rhagoletis (Diptera: Tephritidae) species in the tabellaria species group. Insect Syst. Divers. 2:1–14. Google Scholar

27.

Hwang, U. W., and W. Kim. 1999. General properties and phylogenetic utilities of nuclear ribosomal DNA and mitochondrial DNA commonly used in molecular systematics. Korean J. Parasitol. 37: 215–228. Google Scholar

28.

Kumar, S., G. Stecher, and K. Tamura. 2016. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger data sets. Mol. Biol. Evol. 33: 1870–1874. Google Scholar

29.

Linder, C. R., L. R. Goertzen, B. V. Heuvel, J. Francisco-Ortega, and R. K. Jansen. 2000. The complete external transcribed spacer of 18S-26S rDNA: amplification and phylogenetic utility at low taxonomic levels in asteraceae and closely allied families. Mol. Phylogenet. Evol. 14: 285–303. Google Scholar

30.

Linnen, C. R., and B. D. Farrell. 2007. Mitonuclear discordance is caused by rampant mitochondrial introgression in Neodiprion (Hymenoptera: Diprionidae) sawflies. Evolution. 61: 1417–1438. Google Scholar

31.

Mattson, M., G. R. Hood, W. L. Yee, M. M. Doellman, D. J. Bruzzese, R. B. Goughnour, A. L. Driscoe, S. Van Dexter, C. Tait, M. M. Glover, et al. 2022. Recursive adaptation in action: allochronic isolation and divergence of host-associated populations of Rhagoletis pomonella following its recent introduction to western USA. Entomologia Experimentalis et Applicata 170: 48–63. Google Scholar

32.

McPheron, B. A. 1990a. Genetic structure of apple maggot fly (Diptera: Tephritidae) populations. Ann. Entomol. Soc. Am. 83: 568–577. Google Scholar

33.

McPheron, B. A. 1990b. Implications of genetic variation in western apple maggots for understanding population biology. Ch. 3 In R. N. Dowell, L. T. Wilson, V. P. Jones (eds.), Apple Maggot in the west: history, biology, and control, University of California (System). Division of Agriculture and Natural Resources Publ. 3341. University of California. 95 pp. Google Scholar

34.

Michel, A. P., J. Rull, M. Aluja, and J. L. Feder. 2007. The genetic structure of hawthorn-infesting Rhagoletis pomonella populations in Mexico: implications for sympatric host race formation. Mol. Ecol. 16: 2867–2878. Google Scholar

35.

Morgan-Richards, M., M. Bulgarella, L. Sivyer, E. J. Dowle, M. Hale, N. E. McKean, and S. A. Trewick. 2017. Explaining large mitochondrial sequence differences within a population sample. R. Soc. Open Sci. 4: 170730. Google Scholar

36.

Nei, M. 1991. Relative efficiencies of different tree-making methods for molecular data, pp. 90–128 In M. M. Miyamoto, and J. Cracraft (eds). Phylogenetic analysis of DNA sequences, Oxford University Press, New York, NY. Google Scholar

37.

Payne, J. A., and S. H. Berlocher. 1995. Distribution and host plants of the blueberry maggot, Rhagoletis mendax (Diptera: Tephritidae) in southeastern North America. J. Kans. Entomol. Soc. 68: 133–142. Google Scholar

38.

Powell, T. H., A. A. Forbes, G. R. Hood, and J. L. Feder. 2014. Ecological adaptation and reproductive isolation in sympatry: genetic and phenotypic evidence for native host races of Rhagoletis pomonella. Mol. Ecol. 23: 688–704. Google Scholar

39.

Ratnasingham, S., and P. D. Hebert. 2007. bold: the barcode of life data system ( http://www.barcodinglife.org). Mol. Ecol. Notes. 7: 355–364. Google Scholar

40.

Rung, A., D. R. Miller, and S. J. Scheffer. 2009. Polymerase chain reaction-restriction fragment length polymorphism method to distinguish three mealybug groups within the Planococcus citriP. minor species complex (Hemiptera: Coccoidea: Pseudococcidae). J. Econ. Entomol. 102: 8–12. Google Scholar

41.

Saitou, N., and M. Nei. 1987. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol. Biol. Evol. 4: 406–425. Google Scholar

42.

Schwarz, D., B. M. Matta, N. L. Shakir-Botteri, and B. A. McPheron. 2005. Host shift to an invasive plant triggers rapid animal hybrid speciation. Nature. 436: 546–549. Google Scholar

43.

Shaikevich, E. V., M. V. Zagoskin, and D. V. Mukha. 2013. Comparative characteristics of the intergenic spacer of the ribosomal RNA gene cluster in mosquitoes of the genus Culex (Diptera: Culicidae). Mol. Biol. 47: 364–372. Google Scholar

44.

Si, C., K. Chen, R. Tao, C. Su, J. Ma, L. Li, J. Hao, and Q. Yang. 2020. Genetic differentiation and divergence time of Chinese Parnassius (Lepidoptera: Papilionidae) species based on nuclear internal transcribed spacer (ITS) sequence data. J. Entomol. Sci. 55: 520–546. Google Scholar

45.

Simon, C., F. Frati, A. Beckenbach, B. Crespi, H. Liu, and P. Flook. 1994. Evolution, weighting, and phylogenetic utility of mitochondrial gene sequences and a compilation of conserved polymerase chain reaction primers. Ann. Entomol. Soc. Am. 87: 651–701. Google Scholar

46.

Smith, J. J., and G. L. Bush. 1997. Phylogeny of the genus Rhagoletis (Diptera: Tephritidae) inferred from DNA sequences of mitochondrial cytochrome oxidase II. Mol. Phylogenet. Evol. 7: 33–43. Google Scholar

47.

Smith, J. J., V. Gavrilovic, and D. R. Smitley. 2001. Native Vaccinium spp. infested by Rhagoletis mendax (Blueberry maggot; Diptera: Tephritidae) in the Great Lakes region: a potential source of inoculum for infestation of cultivated blueberries. J. Econ. Entomol. 94: 1378–1385. Google Scholar

48.

Smith, M. A., D. M. Wood, D. H. Janzen, W. Hallwachs, and P. D. Hebert. 2007. DNA barcodes affirm that 16 species of apparently generalist tropical parasitoid flies (Diptera, Tachinidae) are not all generalists. Proc. Natl. Acad. Sci. U. S. A. 104: 4967–4972. Google Scholar

49.

Smith, M. A., C. Bertrand, K. Crosby, E. S. Eveleigh, J. Fernandez-Triana, B. L. Fisher, J. Gibbs, M. Hajibabaei, W. Hallwachs, et al. 2012. Wolbachia and DNA barcoding insects: patterns, potential, and problems. PLoS One. 7: e36514. Google Scholar

50.

Stage, D. E., and T. H. Eickbush. 2007. Sequence variation within the rRNA gene loci of 12 Drosophila species. Gen. Res. 17: 1888–1897. Google Scholar

51.

Swofford, D. 2000. PAUP*. Phylogenetic analysis using parsimony (*and other methods). Version 4. Sinauer Associates, Sunderland, MA. Google Scholar

52.

Valtueña, F. J., T. Rodriguez-Riaño, J. Lopez, C. Mayo, and A. Ortega-Olivencia. 2017. Peripatric speciation in an endemic Macaronesian plant after recent divergence from a widespread relative. PLoS One. 12: e0178459. Google Scholar

53.

Wasbauer, M. S. 1963. Taxonomic discrimination of Rhagoletis pomonella from the eastern United States and Rhagoletis zephyria from California. Calif. Dept. Agric. T-9: 1–6. Google Scholar

54.

Westcott, R. L. 1982. Differentiating adults of apple maggot, Rhagoletis pomonella (Walsh) from snowberry maggot, Rhagoletis zephyria Snow (Diptera: Tephritidae) in Oregon. Pan-Pac. Entomol. 58: 25–30. Google Scholar

55.

WSDA. 2001. Washington Administrative Code 16-470-108. Distribution of infested or damaged fruit is prohibited. Washington State Department of Agriculture. Available from:  https://apps.leg.wa.gov/WAC/default.aspx?cite=16-470-108. Accessed 29 December 2021. Google Scholar

56.

Xie, X., J. Rull, A. P. Michel, S. Velez, A. A. Forbes, N. F. Lobo, M. Aluja, and J. L. Feder. 2007. Hawthorn-infesting populations of Rhagoletis pomonella in Mexico and speciation mode plurality. Evolution (N. Y) 61: 1091–1105. Google Scholar

57.

Yee, W. L., and R. B. Goughnour. 2011. Mating frequencies and production of hybrids by Rhagoletis pomonella and R. zephyria (Diptera: Tephritidae) in the laboratory. Can. Ent. 143: 82–90. Google Scholar

58.

Yee, W. L., P. S. Chapman, H. D. Sheets, and T. R. Unruh. 2009. Analysis of body measurements and wing shape to discriminate Rhagoletis pomonella and Rhagoletis zephyria (Diptera: Tephritidae) in Washington State. Ann. Entomol. Soc. Am. 102: 1013–1028. Google Scholar

59.

Yee, W. L., H. D. Sheets, and P. S. Chapman. 2011. Analysis of surstylus and aculeus shape and size using geometric morphometrics to discriminate Rhagoletis pomonella and Rhagoletis zephyria (Diptera: Tephritidae). Ann. Entomol. Soc. Am. 104: 105–114. Google Scholar

60.

Ye, J., G. Coulouris, I. Zaretskaya, I. Cutcutache, S. Rozen, and T. L. Madden. 2012. Primer-BLAST: a tool to design target-specific primers for polymerase chain reaction. BMC Bioinform. 13: 134. Google Scholar

61.

Yee, W. L., P. S. Chapman, and H. D. Sheets. 2013. Comparative body size and shape analyses of F1 hybrid Rhagoletis pomonella and Rhagoletis zephyria (Diptera: Tephritidae). Ann. Entomol. Soc. Am. 106: 410–423. Google Scholar
© The Author(s) 2022. Published by Oxford University Press on behalf of Entomological Society of America.
J. J. Smith, P. Brzezinski, J. Dziedziula, E. Rosenthal, and M. Klaus "Partial Ribosomal Nontranscribed Spacer Sequences Distinguish Rhagoletis zephyria (Diptera: Tephritidae) From the Apple Maggot, R. pomonella," Journal of Economic Entomology 115(2), 647-661, (20 January 2022). https://doi.org/10.1093/jee/toab264
Received: 29 October 2021; Accepted: 13 December 2021; Published: 20 January 2022
KEYWORDS
apples
fruit flies
molecular diagnostics
rDNA
snowberries
Back to Top