Você está na página 1de 10

Sequential divergence and the multiplicative origin of

community diversity
Glen R. Hooda,1, Andrew A. Forbesb, Thomas H. Q. Powella,c, Scott P. Egana,d,e, Gabriela Hamerlinckb, James J. Smithf,
and Jeffrey L. Federa,d
a
Department of Biological Sciences, University of Notre Dame, Notre Dame, IN 46556; bDepartment of Biology, University of Iowa, Iowa City, IA 52242;
c
Department of Entomology and Nematology, University of Florida, Gainesville, FL 32611; dEnvironmental Change Initiative and Advanced Diagnostics and
Therapeutics, University of Notre Dame, Notre Dame, IN 46556; eDepartment of Biosciences, Anderson Biological Laboratories, Rice University, Houston,
TX 77005; and fDepartment of Entomology and Lyman Briggs College, Michigan State University, East Lansing, MI 48824

Edited by Douglas Futuyma, State University of New York, Stony Brook, NY, and approved September 18, 2015 (received for review December 24, 2014)

Phenotypic and genetic variation in one species can influence the Such “sequential” or “cascading” divergence events may be
composition of interacting organisms within communities and across particularly relevant to understanding why some groups of or-
ecosystems. As a result, the divergence of one species may not be an ganisms, like plants, the insects that feed on them, and the para-
isolated process, as the origin of one taxon could create new niche sitoids that attack the insects, are more diverse and species-rich than
opportunities for other species to exploit, leading to the genesis of other groups (8, 9, 12–15). Specifically, when phytophagous insects
many new taxa in a process termed “sequential divergence.” Here, diversify by adapting to new host plants, they create a new habitat
we test for such a multiplicative effect of sequential divergence in a for their parasitoids to exploit (Fig. 1). If a parasitoid shifts to the
community of host-specific parasitoid wasps, Diachasma alloeum, new habitat, it can encounter the same divergent ecological selec-
Utetes canaliculatus, and Diachasmimorpha mellea (Hymenoptera:
tion pressures as its insect host, which could result in the parallel
divergence of insect host and parasitoid (12–15) (Fig. 1A). More-
Braconidae), that attack Rhagoletis pomonella fruit flies (Diptera:
over, sequential divergence may have multiplicative effects in gen-
Tephritidae). Flies in the R. pomonella species complex radiated by
erating biodiversity, as the shift of an insect to a new plant may open
sympatrically shifting and ecologically adapting to new host plants, a new niche opportunity for not just one but the entire community
the most recent example being the apple-infesting host race of of parasitoids attacking the insect host (12, 13) (Fig. 1B). However,
R. pomonella formed via a host plant shift from hawthorn-infesting few convincing examples of sequential divergence exist (12–15), and
flies within the last 160 y. Using population genetics, field-based in no study is there both genetic and ecological evidence for
behavioral observations, host fruit odor discrimination assays, and sequential divergence multiplicatively amplifying biodiversity.
analyses of life history timing, we show that the same host-related Here, we test for the multiplicative effects of sequential di-
ecological selection pressures that differentially adapt and reproduc- vergence in the community of parasitoid wasps (Hymenoptera:
tively isolate Rhagoletis to their respective host plants (host-associated Braconidae) that attack fruit flies in the Rhagoletis pomonella
differences in the timing of adult eclosion, host fruit odor preference sibling species complex (Diptera: Tephritidae) (Fig. 2). Several
and avoidance behaviors, and mating site fidelity) cascade through the features of the biology and biogeography of the Rhagoletis−para-
ecosystem and induce host-associated genetic divergence for each of sitoid system make it ideal for investigating the multiplicative
the three members of the parasitoid community. Thus, divergent se- divergence hypothesis and allow us to directly test multiple criteria
lection at lower trophic levels can potentially multiplicatively and rap- supporting sympatric host race formation (16) and sequential di-
idly amplify biodiversity at higher levels on an ecological time scale, vergence (12, 13), summarized in Table 1. First, concerning the
which may sequentially contribute to the rich diversity of life. spatial context of divergence (criterion 1), R. pomonella complex

host plant adaptation | parasitoid | Rhagoletis | tritrophic interactions | Significance


ecological speciation
Understanding how new life forms originate is a central question
in biology. Population divergence is usually studied with respect
P opulation divergence is a fundamental evolutionary process
contributing to the diversity of life (1). Studies of how new life
forms originate typically focus on how barriers to gene flow evolve
to how single lineages diverge into daughter taxa. However,
populations may not always differentiate in isolation; divergence
in specific lineages, resulting in their divergence into descendent of one taxon could create new niche opportunities in higher tro-
phic levels, leading to the sequential origin of many new taxa.
daughter taxa. As a result, evolutionary biologists now have a good
Here, we show that this may be occurring for three species of
understanding of how variation within a population is trans-
parasitoid wasps attacking Rhagoletis fruit flies. As flies shift and
formed by selection into differences between taxa (1–3). What is
adapt to new host plants, wasps follow suit and diverge in kind,
less well understood is whether the divergence of one population resulting in a multiplicative increase of diversity as the effects of
has consequences that ripple through the trophic levels of an ecologically based divergent selection cascade through the eco-
ecosystem and affect entire communities of interacting organ- system. Biodiversity therefore may potentially beget increasing
isms. Studies in paleontology (4–6), community ecology (7, 8), levels of biodiversity.
systematics (8, 9), and ecosystem genetics (10, 11) suggest that
evolutionary change in one lineage can influence entire com- Author contributions: G.R.H., A.A.F., T.H.Q.P., S.P.E., and J.L.F. designed research; G.R.H.,
munities of organisms. For example, when the genotype/pheno- A.A.F., T.H.Q.P., S.P.E., G.H., J.J.S., and J.L.F. performed research; G.R.H., A.A.F., T.H.Q.P., and
type of a “foundation” species influences the relative fitness of J.L.F. contributed new reagents/analytic tools; G.R.H., A.A.F., S.P.E., G.H., and J.L.F. analyzed
data; and G.R.H., A.A.F., T.H.Q.P., S.P.E., G.H., J.J.S., and J.L.F. wrote the paper.
other species, evolutionary change(s) in this genotype/phenotype
The authors declare no conflict of interest.
may affect organisms in adjacent trophic levels (10, 11). If these
This article is a PNAS Direct Submission.
evolutionary changes are linked to ecological adaptation and re-
Data deposition: The mtDNA sequences reported in this paper have been deposited in the
productive isolation (RI), associated organisms may diverge in par- GenBank database (accession nos. KT761291–KT761497). Raw data are available from the
allel, potentially creating entire coevolved communities distinct from Dryad Digital Repository, dx.doi.org/10.5061/dryad.5n72m.
one another (12–15). Therefore, population divergence may not 1
To whom correspondence should be addressed. Email: ghood@nd.edu.
always be an isolated process, as the differentiation of one taxon This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
could beget the divergence of many others. 1073/pnas.1424717112/-/DCSupplemental.

E5980–E5989 | PNAS | Published online October 23, 2015 www.pnas.org/cgi/doi/10.1073/pnas.1424717112


PNAS PLUS
Multiplicative Co-speciation
Sequential Divergence Sequential Divergence in Allopatry
A B C
Fig. 1. Three scenarios of codivergence in a host−
parasitoid system. (A) A single sequential divergence
event, (B) sequential divergence with multiplicative
amplification of biodiversity, and (C) cospeciation in
allopatry. In A, codivergence is driven by the cas-
Allopatry cade of divergent ecological selection pressures
across trophic levels in sympatry. Here, a degree of
divergent ecological adaptation must accompany
Time

the host shift such that parasitoids are not merely


moving between geographically separated hosts. In
B, the multiplicative effects of sequential divergence
can be seen as several members of the parasitoid
community diverge in parallel with their host. In C,
Secondary contact codivergence (cospeciation) occurs after host plant,
fly, and parasitoid populations become jointly geo-
graphically isolated (black bar), resulting in parallel
allopatric speciation. Here, little differentiation need
accompany the initial host shift of fly or parasitoid.
Cospeciation is not necessarily driven by the creation
and adaptation to new niches but by the concordant
geographic and reproductive separation of hosts and

EVOLUTION
parasitoids.

flies have formed in the absence of geographic isolation. In par- vergence (14). Population genetic surveys, field observations,
ticular, the recent sympatric shift of R. pomonella from its ancestral behavioral assays of host choice, and studies of life history timing
host plant hawthorn (Crataegus spp.) to introduced domesticated support the existence of an ecologically derived population of
apple (Malus domestica) in the last 160 y is an example of host race D. alloeum attacking the recently formed apple-infesting host
formation in action (1–3, 17–19), the hypothesized initial stage of race of R. pomonella, meeting criteria 2, 3, 4, 6, and 7. We hy-
ecological speciation (16, 17). The short time frame and sympatric pothesize that if U. canaliculatus and D. mellea are undergoing
spatial context of R. pomonella’s shift to apple exclude passive sequential divergence, they will show similar patterns of host-
codivergence or speciation (Fig. 1C) as an explanation for differ- associated ecological and genetic divergence.
entiation. Specifically, fly and wasp populations could not have Two dimensions of divergent ecological selection generate RI
diverged in concert, because they became jointly geographically among host-associated populations of Rhagoletis and D. alloeum:
separated in the past (Fig. 1C). Rather, if flies and wasps display host-specific mating (habitat isolation) and differences in eclo-
concordant adaptations, it is likely due to the direct effects of di- sion phenology (temporal isolation). With respect to habitat
vergent ecological selection resulting from host shifts cascading isolation, Rhagoletis (26) and D. alloeum (14) court and mate on
from host plants to flies to parasitoids (Fig. 1 A and B). or near the fruit of their respective host plants. The most im-
Second, the apple and hawthorn host races of R. pomonella portant long- to intermediate-range cues that flies use to find and
belong to a closely related group of sibling species, including discriminate among plants are the volatile compounds emitted
Rhagoletis mendax (host: blueberry, Vaccinium spp.), Rhagoletis from the surface of ripening fruit (27–30). Flies display genetically
zephyria (host: snowberry, Symphoricarpos spp.), and the unde- based behavioral preference for natal fruit surface volatiles and
scribed flowering dogwood fly (host: Cornus florida). All of these avoid the volatiles of alternative fruit (29). Similarly, D. alloeum
taxa purportedly radiated via sympatric host shifts (17–23). In prefer natal and avoid nonnatal host fruit volatiles in behavioral
addition, other species in the genus, such as the eastern cherry assays (14), supporting criterion 3. Consequently, differences in
fly, Rhagoletis cingulata (host: black cherry, Prunus serotina) are host choice translate directly to mate choice, generating prezygotic
sympatric with R. pomonella group flies (23). Thus, the potential habitat-related RI for both flies and wasps, fulfilling criterion 4.
for sequential divergence in the Rhagoletis parasitoid community Additionally, host odor discrimination may also act as a postzygotic
extends beyond the host races, with multiple cooccurring fly re- barrier to gene flow in R. pomonella, as they suffer behavioral host
sources existing for wasps to attack, satisfying criterion 1. choice sterility mediated by a reduced chemosensory ability to find
Third, Rhagoletis in the eastern United States are attacked by a suitable host fruit for mating and oviposition (30). Whether or not
community of host-specific endoparasitoid wasps that include hybrid D. alloeum display a similar behavior is unknown. Lastly,
the species Diachasma alloeum, Utetes canaliculatus, and Diac- criteria 5 is partially met for Rhagoletis, as the host fruit envi-
hasmimorpha mellea (24, 25) (Fig. 2). All three species have a ronment has no effect on host odor discrimination behaviors for
free-living, sexually reproducing adult life stage. This life cycle hawthorn-origin R. pomonella reared in apple fruit, indicating
eliminates vertical transmission as a factor facilitating codi- that host selection and fidelity are under (partial) genetic control
vergence. U. canaliculatus oviposits into Rhagoletis eggs laid be-
neath the skin of ripe fruit, while D. alloeum and D. mellea
oviposit into late instar larvae feeding within fruit (24). A degree
of niche partitioning for oviposition sites therefore exists among A B C D E
species (25), potentially facilitating coexistence on the same fly
host. As a result, multiple host associations of wasps exist in close
geographic proximity, fulfilling the requirements of criterion 1
for sympatric race formation and sequential divergence for the Fig. 2. The community of host-specific parasitoids that attack members
wasp community as well. of the (A) Rhagoletis pomonella sibling species complex: (B) D. alloeum,
In addition, a previous study documented that one parasitoid (C) D. mellea, and (D) U. canaliculatus that (E) emerge from the fly pupal
attacking Rhagoletis, D. alloeum, is undergoing sequential di- case as adults following overwintering. (Scale bar, 1 mm.)

Hood et al. PNAS | Published online October 23, 2015 | E5981


Table 1. Summary of conditions (criteria) conducive to and supporting hypotheses of sympatric host race formation (modified from
ref. 16) and sequential divergence (modified from refs. 12 and 13)
Criteria supporting hypotheses of sympatric host race
formation and sequential divergence Rp Da Dm Uc

Criterion 1. Shift to new host resource and multiple host-associations occur in yes yes yes yes
sympatry or close geographic proximity
Criterion 2. Host-associated populations form distinct genetic clusters (spatially yes yes yes yes
replicable), but experience gene flow at appreciable rates
Criterion 3. Females, but also potentially males, display host preferences and yes yes yes yes
discriminate among alternate hosts
Criterion 4. Host choice is linked to mate choice facilitating assortative mating yes yes yes yes
and resulting in prezygotic habitat isolation
Criterion 5. Host selection and fidelity are under some degree of genetic yes yes, tested not not
control and not due solely to maternal, learning, or environmental effects in one direction tested tested
Criterion 6. Differences in insect phenologies tracks differences in the host yes yes yes yes
phenologies resulting in temporal (allochronic) isolation
Criterion 7. Insect phenology under some degree of genetic control and not yes yes yes yes
due solely to maternal or environmental effects
Criterion 8. Fitness tradeoffs exist between host-associated populations yes, sometimes not tested not tested not tested
resulting in migrants and hybrids having reduced fitness

Also shown is whether these criteria have been empirically tested and confirmed in R. pomonella (Rp) species complex flies and three members of the
parasitoid wasp community attacking the flies: D. alloeum (Da), D. mellea (Dm), and U. canaliculatus (Uc). Data for Da are from Forbes et al. (14) and the
current study (criterion 5).

(27, 30). Similar experiments have not yet been conducted in nonnatal fly and host plant environments to parental and nonnatal
D. alloeum (but see Reciprocal Rearing of Diachasma). host fruit volatiles. Third, to assess the degree of allochronic iso-
With respect to temporal isolation (criterion 6), the timing of lation due to variation in host phenology, we compared the timing
overwinter diapause is an important host-related ecological ad- of adult eclosion of U. canaliculatus and D. mellea attacking
aptation for Rhagoletis. The host plants of Rhagoletis fruit at different fly populations in sympatry (criterion 6). Lastly, to link
different times of the year (19, 31, 32). For example, apple va- host-associated genetic differentiation to divergence in life history
rieties favored by R. pomonella ripen 3–4 wk before native timing (criterion 7), we tested for associations between micro-
hawthorns in sympatry. Thus, flies must eclose to coincide with satellite genotypes and the timing of eclosion for U. canalicualtus
the availability of ripe fruit to find mates and oviposition sites. and D. mellea. Although host-associated fitness tradeoffs (criterion
Rhagoletis are univoltine, and their lifespan is short (1 mo). 8) have been inferred for several species of Rhagoletis feeding in
Differences in eclosion timing between races therefore results in natal versus nonnatal fruit (33, 34), difficulty in reciprocally
partial allochronic mating isolation (19, 29–32). The differences transplanting wasps precludes these experiments at this time but
in eclosion timing also confer a degree of postzygotic isolation remain an area for future study.
because hybrids will possess eclosion patterns asynchronous with
fruit ripening (29, 31, 32). Rhagoletis attacking blueberries and Results and Discussion
flowering dogwoods display similar differences in eclosion time mtDNA Divergence. To make a case for the sequential divergence
related to variation in host fruiting phenology (31). of U. canaliculatus and D. mellea, it must be shown that pop-
The life cycle of D. alloeum mirrors that of Rhagoletis, gen- ulations attacking the derived apple fly do not represent un-
erating the same divergent ecological selection pressures. As a
recognized sibling species associated with different Rhagoletis. In
result, populations of D. alloeum attacking different Rhagoletis
this case, differences could be erroneously attributed to ecolog-
eclose to match the phenology of fly larvae feeding within host
ical adaptation following a recent sympatric host shift when, in
fruit (14). In addition, longevity of D. alloeum (∼2 wk) is half that
of Rhagoletis, generating even more pronounced allochronic fact, they represent older speciation events that potentially occurred
mating isolation compared to the fly (14), supporting criterion 6. in allopatry. To test for the existence of cryptic sibling species, we
Significant allele frequency differences between sympatric pop- sequenced a 540-bp region of the mtDNA COI (cytochrome
ulations of D. alloeum attacking different fly hosts (criterion 2) oxidase subunit I) gene for specimens morphologically classified
were associated with differences in eclosion time (14), confirm- as U. canaliculatus and D. mellea attacking apple, hawthorn,
ing criterion 7. The same has also been found for Rhagoletis blueberry, flowering dogwood, snowberry, and black cherry flies.
(19, 31, 32), connecting host-related life history adaptation and The presence of sibling species would be indicated by diverged
RI to patterns of genetic differentiation among flies and wasps. mtDNA haplotypes that, upon analysis using nuclear-encoded
Here, we test for the multiplicative hypothesis of sequential microsatellites, would identify as genetically distinct populations,
divergence in the Rhagoletis−parasitoid system using the criteria implying an absence of gene flow.
in Table 1 to frame our experimental approach. For populations Utetes. A maximum parsimony gene tree of Utetes populations
of U. canaliculatus and D. mellea attacking different Rhagoletis, attacking apple, hawthorn, snowberry, and flowering dogwood
we first assessed mtDNA sequence variation and conducted flies indicated three distinct mtDNA haplotype groups distin-
population genetics surveys using nuclear-encoded microsatellites guished by ≥19 nucleotide substitutions (Fig. 3A). Haplotype A
to test for host-related genetic differentiation (criterion 2). Second, was restricted to wasps attacking the flowering dogwood fly, and
to test for host plant-related assortative mating caused by habitat haplotype B was found only in wasps attacking flowering dog-
isolation, we coupled field observations of mating behavior of wood and a single population of snowberry flies from State
U. canalicualtus and D. mellea with tests of host odor discrimination, College, PA (site 9). Haplotype C was most common, found in
including preference for and avoidance of fruit surface volatile U. canaliculatus attacking flowering dogwood, apple, hawthorn,
compounds (criteria 3 and 4). To determine if genetic effects and a single population of snowberry flies from East Lansing, MI
contributed to host odor discrimination (criterion 5), we com- (site 2). Thus, the possibility exists for three different cryptic
pared the behavioral responses of D. alloeum reared through Utetes sibling species, with haplotype C being the most informative

E5982 | www.pnas.org/cgi/doi/10.1073/pnas.1424717112 Hood et al.


PNAS PLUS
for testing the sequential divergence hypothesis, attacking pop- Within haplotype C, significant microsatellite differentiation
ulations of all four Rhagoletis hosts parasitized by U. canaliculatus. was observed among populations of U. canaliculatus attacking
Diachasmimorpha. In contrast to U. canaliculatus, there was no evi- snowberry, flowering dogwood, hawthorn, and apple flies. No
dence for a distinct mtDNA matrilineage among D. mellea attacking fixed or high-frequency private allele was found for any of the 20
apple, hawthorn, blueberry, or black cherry flies (Fig. 3B). Similar to microsatellite loci that distinguished wasps attacking different fly
previous results from D. alloeum (14), several haplotypes were hosts. Instead, populations shared major alleles in common that
found in D. mellea displaying low levels of sequence divergence differed in frequency (see data in the Dryad database), implying
(≤1%). Haplotypes did not differentiate wasps in a host-specific that differentiation among populations is likely of relatively re-
manner and were generally shared among all populations. The cent origin with ongoing (or recently ceased) gene flow. Signif-
pattern implies that host-associated populations of D. mellea are icant allele frequency differences were observed in each of the
relatively recently derived, including those attacking apple flies, six sympatric pairwise comparisons between wasps attacking
and rules out the presence of long-established cryptic species. different fly hosts (Table S2). Seven loci (UC8, UC10, UC12,
UC16, UC22, UC47, UC54) displayed pronounced and consis-
Microsatellite Differentiation. tent differences (of similar magnitude and in the same direction)
Utetes. The pattern of microsatellite differentiation for 20 loci across at least three pairwise sympatric comparisons (Table S2).
suggested the existence of three cryptic species of Utetes likely Similarly, generalized linear model (GLM) analyses identified
having diverged before the introduction of apples to North seven loci (UC8, UC12, UC14, UC32, UC47, UC52, UC53)
America ∼400 y ago and the shift of R. pomonella to apples within showing a significant host effect across the five sympatric paired
the last 160 y. The three genetic clusters observed in the micro- sites of wasps attacking apple and hawthorn flies (Table S3).
satellite genetic distance network (Fig. S2) corresponded exactly There was little evidence for a major effect of geography in the
to the three major mtDNA haplotypes (Fig. 3A). Populations GLM analysis, as only two loci (UC14, UC61) had a significant
defined by mtDNA haplotypes A, B, and C each possessed alleles latitude and/or host × latitude effect (Table S3). This result
displaying high-frequency differences or private variants at many contrasts with previous findings for R. pomonella (18, 32) and
loci. For example, allele 173 at locus UC57 ranged in frequency D. alloeum (14) where latitudinal variation was pronounced. Thus,
from 0.45 to 0.76 among haplotype C wasp populations, was absent the direction and magnitude of host-related allele frequency

EVOLUTION
in haplotype B, and was present at low frequency (∼0.06) in differences were more consistent across sites for U. canal-
haplotype A (see data in the Dryad database). Thus, all sub- iculatus than for R. pomonella or D. alloeum. As a result, a
sequent analyses of U. canaliculatus focused on the widespread neighbor-joining (NJ) genetic distance network based on all 20
mtDNA haplotype C populations to test for recent sequential microsatellites distinguished populations attacking apple, haw-
divergence. thorn, flowering dogwood, and snowberry flies (Fig. 4A). Apple

A Utetes canaliculatus B Diachasmimorpha mellea


3(2), 6
1 substitution 1 1 substitution
1(3), 4(3), 6
1(2), 2(2), 5(2) 6
4(2)
2, 5 3
6
1(2), 2(2) 2
6
1 6
5(93) 1
6 2(2), 3(4), 6(3)
3
C
2, 3(2), 6
1, 2, 3(2), 6(2)
4(2), 6(3) 3(3)
3(4), 6, 8(3)
2(7) 6(2)
16(100)
2 6
5
2 3
1, 3, 5(2), 6, 7(3), 8
1(2)
1
7(2) 2(3)
6(2)
2(2)
9
4(3), 6(17) 6(2) Fig. 3. Maximum parsimony mtDNA COI gene trees
9 for (A) U. canaliculatus and (B) D. mellea attacking
4 6
hawthorn (red), apple (green), blueberry (blue), flow-
6 B 3 ering dogwood (yellow), snowberry (white), and black
4
9 3 cherry (black) flies. The three major haplotypes
9 1(5), 2(3), 3(3), 6(6) identified for U. canaliculatus are designated A, B,
14(100) 9(3) and C. Each tree is rooted with the outgroup taxon,
9 1(3), 2(3), 3(2), 6(6)
D. alloeum. Nonparenthetic numbers next to each
6(2)
6(3) circle indicate collecting site (see Table S1 and Fig. S1
4 for site designations). Numbers in parentheses in-
25(100) 6 4(3), 6(14)
3(2) dicate the number of identical haplotypes sequenced
6
6 A 3(3)
at each collecting site. The number of nucleotide sub-
stitutions (evolutionary steps) and bootstrap support
6(3) based on 10,000 replicates (in parentheses) are given
49(100) Diachasma alloeum 83(100)
Diachasma alloeum above longer branches.

Hood et al. PNAS | Published online October 23, 2015 | E5983


and hawthorn fly-attacking wasps were positioned adjacently in (ii) at the sympatric site in Fennville, MI, more loci displayed
the network, consistent with the apple fly-attacking populations significant differences between wasps attacking blueberry and
having originated via a recent shift from hawthorn flies. Al- apple flies (7 loci) than between blueberry and hawthorn flies (2
though a shift of U. canaliculatus from the snowberry to apple fly loci), and (iii) D. mellea is a relatively rare parasitoid of blue-
cannot be completely discounted, this scenario appears unlikely berry flies (24, 25). Thus, a hawthorn or black cherry fly ancestry
because (i) the genetic distance from apple to snowberry fly- cannot be discounted.
attacking wasps is greater than that from apple to hawthorn fly- The origins of the recently formed apple fly-attacking parasitoid
attacking wasps and (ii) populations of the snowberry fly are community still need to be established. Genetic results imply that
relatively rare and present at low densities in the eastern apple-infesting populations of D. alloeum are genetically more
United States compared with R. pomonella (35). closely related to wasps parasitizing the blueberry fly, R. mendax,
Diachasmimorpha. The pattern of microsatellite differentiation for than the hawthorn-infesting race of R. pomonella (14). Although
21 loci for D. mellea also supported the existence of a recently inferences can be made regarding the origins of the apple-
derived population attacking apple flies. Similar to U. canal- infesting populations of U. canaliculatus and D. mellea, the genetic
iculatus, D. alloeum (14), and R. pomonella (18, 22, 32), no fixed results are more ambiguous than for D. alloeum. Regardless,
or high-frequency private alleles were found that distinguish Rhagoletis and its parasitoids may not always cospeciate in a strict
D. mellea populations attacking different fly hosts (see data in the 1:1 follow-the-leader fashion. Wasps attacking different flies in
Dryad database). However, significant and consistent allele fre- the community may take advantage of the new niche opportunity
quency differences were observed for eight loci (DM19, DM32, provided by Rhagoletis host shifts, not necessarily just the para-
DM42, DM48, DM51, DM61, DM62, DM65) in at least three sitoid infesting ancestral fly hosts. Thus, adaptive starbursts of
sympatric pairwise comparisons between D. mellea attacking sequential divergence may be the result of biodiversity radiating
different fly hosts (Table S2). GLM analyses revealed four loci from several different origins within the community. Regardless,
(DM14, DM15, DM25, DM61) displaying a significant host ef- the genetic results support criterion 2 and multiplicative sequential
fect across the four sympatric pairs of apple and hawthorn fly- divergence by establishing the existence of host-associated genetic
attacking wasp populations (Table S3). Unlike U. canaliculatus differentiation among both U. canaliculatus and D. mellea.
but similar to Rhagoletis (18, 22, 32) and D. alloeum (14), a high
proportion (28.6%) of loci (DM1, DM6, DM14, DM15, DM19, Site of Mating Assembly.
DM28) showed a significant host × latitude effect for D. mellea Utetes. Field observations of mtDNA haplotype C U. canaliculatus
(Table S3). An NJ network based on all 21 loci clustered all in stands of apple and hawthorn trees confirmed the use of host
apple fly-attacking D. mellea separately from populations para- fruit as the site of mating assembly. A total of 23 mating pairs were
sitizing hawthorn, blueberry, and black cherry flies (Fig. 4B). detected over 45 h. Eleven of the 23 mating pairs (48%) were de-
Similar to previous findings for D. alloeum (14), the network tected directly on apple or hawthorn fruit, whereas the remaining
suggested that D. mellea attacking blueberry flies could be the 12 pairs (52%) were observed on branches or leaves an average of
source of the population parasitizing apple flies. However, this 8.52 ± 2.11 cm from the nearest host fruit (range = 5–30 cm). The
mean time from the first observation of a male and female to the
conclusion is preliminary because (i) only a single population of
initiation of mating was 143 ± 19 s (range = 4–411 s). No mating
D. mellea attacking blueberry flies was genotyped in the study,
pair or individual U. canaliculatus was observed on any non-Rha-
goletis plant, underscoring the host-specific nature of mating.
Diachasmimorpha. Field observations of D. mellea in stands of
A Utetes canaliculatus B Diachasmimorpha mellea apple and hawthorn trees also confirmed the use of host fruit as
the site of mating assembly similar to U. canaliculatus, D. alloeum
N 6
(14), and R. pomonella (26). A total of 19 mating pairs were
8 observed over 45 h. Five of the 19 mating pairs (26%) were
detected directly on apple or hawthorn fruit, whereas the
100 1 remaining 14 pairs (74%) were observed on branches or leaves
an average of 12.37 ± 3.25 cm from the nearest host fruit (range
3
7 = 3–45 cm). The mean time from first observation to initiation of
4 mating was 209 ± 35.1 s (range = 12–516 s). No mating pair or
2
individual D. mellea was observed on any non-Rhagoletis plant.
6
6 Host Plant Odor Discrimination.
2 5 3
Utetes. Results from Y-tube behavioral assays supported the ex-
4 1 istence of divergent host choice based on fruit odor discrim-
6 ination in U. canaliculatus similar to Rhagoletis (27, 30) and
3
D. alloeum (14). Wasps displayed no tendency to orient to the
1 2 right or left arm of the Y tube when no fruit odor was present
(X2 = 1.2; P = 0.27). However, odor-naïve, mtDNA haplotype
1 C U. canaliculatus reared from snowberry, flowering dogwood,
2
apple, and hawthorn flies all preferentially oriented to the arm
3
6 of the Y tube containing their respective natal fruit volatiles
100 compared with the odorless control arm (Fig. 5 and Table S4).
Wasps also avoided the odors of alternative hosts, orienting less
2 3 often to nonnatal fruit volatiles compared with the odorless
0.01 0.01 control (Fig. 5 and Table S4).
Fig. 4. NJ genetic distance networks of Nei’s D based on all microsatellite loci
Diachasmimorpha. Results from behavioral assays also supported
for (A) U. canaliculatus and (B) D. mellea attacking hawthorn (red), apple the existence of divergent host choice in D. mellea. Wasps showed
(green), blueberry (blue), flowering dogwood (yellow), snowberry (white), and no tendency to orient to either arm of the Y tube when no fruit
black cherry (black) flies. Bootstrap support values ≥ 50% based on 10,000 odor was present (X2 = 0.10; P = 0.76). However, odor-naïve
replicates across loci are given along branches. Numbers next to or inside cir- D. mellea attacking apple, hawthorn, blueberry, and black cherry
cles indicate collecting sites (see Table S1 and Fig. S1 for site designations). The flies all preferred their respective natal host fruit volatiles and
node of U. canaliculatus labeled “N” refers to a northern population of wasps avoided nonnatal fruit odors compared with the odorless control
attacking flowering dogwood that was pooled across multiple collecting sites. arm (Fig. 5 and Table S4).

E5984 | www.pnas.org/cgi/doi/10.1073/pnas.1424717112 Hood et al.


PNAS PLUS
Fig. 5. Host fruit odor discrimination of (A) D. alloeum,
(B) U. canaliculatus, and (C) D. mellea. Values represent
the percent increase (preference) or decrease (avoid-
ance) in the orientation of wasps to the arm of the
Y tube containing different host fruit volatiles relative
to the blank (odorless) control arm. Each host associ-
ated population of wasps displayed avoidance
behavior to all nonnatal fruit odors. We therefore
averaged avoidance values ± SEMs (gray bars) across
all nonnatal fruit assays (see Table S4 for individual
values and statistical significance). Data for D. alloeum
are taken from Forbes et al. (14). *P = 0.05; **P < 0.01;
***P < 0.001; ****P < 0.0001.

In many models of ecological speciation (1, 12, 13, 16, 17), in- contend they are unlikely to be major contributors to host choice

EVOLUTION
cluding R. pomonella (1, 12, 20, 23, 29), host choice and mate choice for these wasps. In this regard, U. canaliculatus and D. mellea are
(the site where courtship and mating occurs) are linked, generating endoparasitoids that oviposit into fly eggs and larvae, respec-
habitat-related prezygotic RI among diverging populations. When tively, in fruit. Wasp larvae do not feed on their Rhagoletis hosts
we couple the results from host fruit odor discrimination assays and until larvae have left fruit and pupariated in the soil. Thus, wasps
field observations, the same is true for the community of wasps do not come in direct contact with the surface volatiles of rip-
attacking different Rhagoletis hosts, supporting criteria 3 and 4 and ening host fruit before returning to their respective host fruit to
multiplicative sequential divergence. We estimate that fruit odor mate and oviposit as adults the following year. In addition, wasps
discrimination potentially translates into ≥84% prezygotic RI eclose as adults from fly pupal cases buried in the soil for >300 d.
among U. canaliculatus attacking different Rhagoletis hosts at sym- Following eclosion, flies must move quickly to the soil surface or
patric sites (Table 2 and SI Text). Similarly, estimates for D. mellea be trapped underground, therefore making it difficult for wasps
ranged from 68% to 88% across populations (Table 2). to antennate fly pupal cases for host cues. Moreover, any volatile
We caution, however, that observations of site of mating may compound from the surface of ripe fruit, even if originally pre-
be skewed unintentionally by observers’ bias by focusing on lo- sent (which is unlikely, as fly larvae do not consume the skin of
cations in the field where insects are more likely to be present fruit but the flesh, which is chemically different), would have
(on host fruit) than not (on other nonhost plant tissue). Thus, it long since dissipated from the pupal case and surrounding soil.
is possible that we missed mating pairs behaving in a non-host- Furthermore, it is unlikely wasps would also learn to avoid odors
specific manner, thereby overestimating the importance of host emitted from nonnatal fruits from exposure to the natal host fly
choice and RI. However, we spent more time surveying nonhost pupal case or a similar host-related stimuli only. Thus, the bi-
plants (30 h) than host trees (15 h) for parasitoids (SI Text). ologies of U. canaliculatus and D. mellea suggest that fruit odor
Moreover, all of the wasps are conspicuous in color (bright red or
reddish black) and large compared with other parasitoid species
(Fig. 2 B−E). Consequently, wasps or mating pairs residing on a Table 2. Pairwise estimates of habitat isolation due to
nonhost plant would have likely been observed if present. differences in fruit odor discrimination behavior calculated
Reciprocal rearing of Diachasma. In the current study, we reared 22 between wasps attacking different host populations of
hawthorn fly-origin and five blueberry fly-origin D. alloeum through Rhagoletis
a single generation in R. pomonella infesting apples. Both hawthorn-
and blueberry-origin D. alloeum retained preferences for their re- Habitat isolation, %
spective natal host fruit odors, while avoiding apple volatiles (Fig. 6 Host comparison Da Uc Dm
and Table S4). Thus, rearing studies support a genetic basis for
behavioral differences in host fruit odor discrimination for both A vs. H 69 85 77
Rhagoletis (27) and D. alloeum (14), supporting, partially, criteria 5. A vs. B 68 68
At present, we cannot rear U. canaliculatus or D. mellea in A vs. S 79 84
alternate fly hosts to test for environmental or maternal effects A vs. C 85
on fruit odor discrimination, a shortcoming of our results and an A vs. D 89
important avenue for future research. However, examples of H vs. B 74 77
larval conditioning affecting adult preference are rare for holo- H vs. S 56 85
metabolous insects (36) and, when found, usually involve cases H vs. C 88
where juvenile life stages come in direct contact with plant tissues H vs. D 85
during development (37). Nevertheless, instances do exist in which
B vs. S 66
parasitoids learn cues from their insect host (38). For example, the
B vs. C 79
braconid wasp Aphidius colemani learns natal host preference by
S vs. D 85
antennating the mummy of its aphid host following adult emer-
gence. However, if A. colemani are dissected from their hosts Da, D. alloeum; Uc, U. canaliculatus; Dm, D. mellea; A, apple flies; H,
before adult emergence and allowed to complete development, hawthorn flies; S, snowberry flies; B, blueberry flies; C, black cherry flies.
natal host preference is reduced. Although we cannot completely Estimates for Da are calculated from host odor discrimination values from
discount such effects for D. mellea and U. canaliculatus, we Forbes et al. (14).

Hood et al. PNAS | Published online October 23, 2015 | E5985


discrimination may be (partially) genetically based, as is the case
Odor Sources

Preference (+)
for Rhagoletis (27, 30) and D. alloeum (14).
Hawthorn

Response to odor (%)


Eclosion Timing. 200
Utetes. Eclosion curves differed significantly among mtDNA
haplotype C U. canaliculatus attacking different fly hosts at Apple
sympatric sites, tracking the eclosion times of their Rhagoletis
hosts and the fruiting times of their respective host plants (Fig. Blueberry
7, Fig. S3, and Table S5). The one exception was at Dowagiac, 100
MI, where eclosion times of U. canaliculatus infesting apple
(102.83 ± 1.05 d; n = 35) and hawthorn (102.71 ± 1.18 d; n = 38)
flies did not differ (Fig. S3 and Table S5). Interestingly, similar
results were previously found for both R. pomonella and
D. alloeum at this site (Table S5) (14). Mean adult longevity for
U. canaliculatus was 9.54 ± 0.18 d (n = 167) in the laboratory. 0

Avoidance (-)
Assuming Utetes have the same life span in nature (likely an
overestimate), we estimate that the observed differences in eclosion
time translate into allochronic prezygotic RI between host-associated
populations of U. canaliculatus at all sites (excluding Dowagiac, MI)
(11–96%) (Table 3).
Diachasmimorpha. Eclosion curves for D. mellea also differed sig-
-100
nificantly among populations attacking different fly hosts at Haw Blue
sympatric sites (Fig. 7, Fig. S3, and Table S5). Mean adult lon- Parental host
gevity for D. mellea was 12.21 ± 0.30 d (n = 67) in the laboratory. Fig. 6. Host fruit odor discrimination of hawthorn fly-origin and blueberry
We estimate that the differences in eclosion times translate into fly-origin D. alloeum reared for a single generation on apple-infesting
prezygotic RI between host-associated populations of D. mellea R. pomonella in apple fruits. Values represent the percent increase (prefer-
at all sites (12–55%) (Table 3). We note that wasps may not ence) or decrease (avoidance) in the orientation of hawthorn-origin or
reach sexual maturity until several days after adult eclosion, po- blueberry-origin D. alloeum to the arm of the Y-tube olfactometer con-
tentially increasing temporal isolation among host-associated pop- taining the indicated host fruit odor relative to blank odorless control
ulations of all three species of parasitoids. treatment. **P < 0.01; ***P < 0.001.

Genetic Correlations with Eclosion Time.


Utetes. Eight loci in females (UC10, UC12, UC14, UC18, UC47, We caution, however, that our results represent a single da-
UC52, UC59, UC60) and eight in males (UC14, UC32, UC48, tum. Additional studies are needed to assess how common and
UC52, UC53, UC54, UC59, UC60) displayed significant geno- taxonomically widespread sequential divergence, evaluate its
type or genotype × host effects with eclosion time (Table S6). Of importance for generating patterns of biodiversity, and assess its
these loci, six showed a significant host effect between wasps role in assembling novel guilds or communities de novo. Condi-
attacking apple and hawthorn flies in GLM analyses (Table S3), tions promoting sequential divergence (Table 1 and Fig. 1) may
linking the pattern of genetic differentiation with an important not be common for many species and may be difficult to empir-
host-related adaptation causing prezygotic RI. Overall, a signif- ically test. Regardless, even if sequential divergence is rare, it could
icant stepwise multiple regression was found between seven loci still, through a multiplicative effect, make a nontrivial contribu-
for females (r = 0.49; P < 0.0001) and six for males (r = 0.40; P = tion to the origin of new taxa. Thus, for organisms such as insects
0.03) with eclosion time. Similar relationships between multiple and their parasites that experience and partition resources on a
microsatellites and eclosion time were found for Rhagoletis (18, fine scale (39), the effects of new niche construction may cascade
22) and D. alloeum (14). through ecosystems and have an important effect on biodiversity.
Diachasmimorpha. Three microsatellite loci for females (DM18, Given that an estimated 10–30 million plant-feeding insect spe-
DM25, DM61) and five loci for males (DM18, DJ19, DM28, cies exist (40), that each phytophagous insect is likely attacked by
DM33, DM36) displayed significant genotype or genotype × host ≥1 parasitoid (41), and that host-specific parasitoids account for
effects with eclosion time (Table S6). Of these loci, two displayed an estimated 25% of insects (40), sequential divergence may be
a significant host effect between apple and hawthorn fly-attack-
an underappreciated process (12–15).
ing wasps in GLM analyses (Table S3). Consequently, evidence
Additional concerns regarding sequential divergence require
supporting criteria 6 and 7 was found for both D. mellea and
clarification. In our study, we associated the process with sym-
U. canaliculatus. However, unlike U. canaliculatus, D. alloeum
(14), and R. pomonella (18, 22), stepwise multiple regressions patric divergence with gene flow (criterion 1). This need not
revealed no significant relationship between microsatellites and always be the case. The critical element is that divergent eco-
eclosion time for female or male D. mellea. logical selection pressures transcend trophic levels to generate
new biodiversity. Thus, sequential divergence may occur in any
Conclusions geographic context. For example, sequential divergence is also
The concept that “biodiversity begets biodiversity” is deeply possible in allopatry if the evolution of RI is driven by adaptation
engrained in biology (4–9, 12–15). Sequential divergence has to a new host following migration of a parasite into a different
been proposed to help explain a number of diverse patterns, geographic area (Fig. 1C). However, verifying that population
including radiations following mass extinctions (4–6), species divergence was due to the cascading effects of selection from
diversity in the tropics (7), and macroevolutionary patterns of plant, to insect, to parasitoid, as opposed to genomic conflict and
diversification in herbivorous insects (8, 9). However, the process RI arising as an inadvertent by-product of mutation order pro-
has rarely been explicitly studied. Here, we document for the first cesses (42, 43), is difficult in the absence of gene flow. Conse-
time, to our knowledge, the genetic signatures of and ecological quently, we focused on an example of sequential divergence involving
mechanisms promoting rapid, sequential divergence for three dis- sympatric host shifts, allowing us to directly document the mul-
tinct species, as the same host-related ecological adaptations asso- tiplicative potential of the process in action. Thus, we excluded
ciated with host choice and life history timing in Rhagoletis appear mtDNA haplotypes A and B in Utetes from further analysis here,
to have evolved concordantly and multiplicatively in the parasitoid but these taxa may represent cases of sequential divergence when
community. considered from a broader geographic perspective.

E5986 | www.pnas.org/cgi/doi/10.1073/pnas.1424717112 Hood et al.


PNAS PLUS
Recent theory suggests that, during ecological divergence, pop-
ulations may diverge incrementally through time before a threshold
level of differentiation is reached and the genomes of taxa rapidly
“congeal” into new species (45). Thus, populations that appear to
be in evolutionary limbo may be slowly accumulating new mutations
of small effect and progressively inching their way toward species
status. In this regard, the two cryptic Utetes taxa we discovered (Fig.
2A) imply that the wasps have the potential to diverge further.
However, it remains to be determined whether sequential di-
vergence alone (cascading ecological selection pressures), or in
conjunction with other mechanisms, contributes to the transition
from races or biotypes to species.
With respect to the current taxonomic status of wasp pop-
ulations, we cast our experimental design in terms of eight cri-
teria supporting sympatric host race formation and sequential
divergence (Table 1). Our results support, partially or in whole,
criteria 1–7 for Rhagoletis and its parasitoid community. Due to
difficulties in wasp husbandry, we could not test for host fly-
related fitness tradeoffs among populations (criterion 8). Re-
ciprocal fruit transplant studies of apple- and hawthorn-infesting
R. pomonella have failed to detect host-related differences in
larval feeding performance with respect to survivorship to pupar-
iation (46). However, tradeoffs in larval performance have been
documented for R. pomonella sibling species (33, 34) that may
distinguish species from the apple and hawthorn host races. If the

EVOLUTION
same is true for wasps, experiments investigating performance
tradeoffs may help establish the taxonomic status of D. alloeum,
D. mellea, and U. canaliculatus. Nevertheless, estimates of habitat
isolation via host odor discrimination and temporal isolation
U. canaliculatus via differences in eclosion timing substantially reproductively
isolate host-associated populations of D. alloeum, D. mellea, and
U. canaliculatus (Tables 2 and 3). Taking into account the relative
contribution of each measure of RI (SI Text), temporal isolation
and habitat isolation can act in concert to produce prezygotic RI
among pairs of D. alloeum (70–94%), U. canaliculatus (85–99%),
and D. mellea (79–90%). Moreover, other cues in addition to fruit
surface volatiles also likely increase host plant fidelity and habitat
isolation for the wasps (e.g., differences in fruit color, size, and
taste), as in Rhagoletis (47). Host fruit odor discrimination could
Rhagoletis also contribute to postzygotic RI for the wasps due to hybrid
behavioral sterility, as it may for Rhagoletis (30).
Finally, competition may limit the potential for multiplicative
sequential divergence (25), placing limits on the overall number
Host association of parasitoids that can simultaneously share the same host.
However, we discovered two cryptic, highly diverged haplotypes
Fig. 7. Mean eclosion times averaged across collecting sites for Rhagoletis of egg-attacking Utetes sharing snowberry and flowering dog-
(circles) infesting blueberry (blue), apple (green), snowberry (white), black wood hosts, suggesting that wasps can subdivide the fly resource
cherry (black), hawthorn (red), and flowering dogwood (yellow) host plants, and avoid extinction. Regardless, competition does occur among
and D. alloeum (squares), U. canaliculatus (triangles), and D. mellea (dia- Rhagoletis parasitoids, as only one adult wasp ecloses when a
monds) attacking each fly host. Data for D. alloeum are from Forbes et al.
single fly is attacked by multiple parasitoids (25). Furthermore,
(14). See Table S5 for individual site values ± SEMs.
larval-attacking D. mellea and D. alloeum overlap in host use.
Thus, several species cooccur, and it is possible that interspecific
interactions may foster diversity by selecting for newly shifting
We also defined the process with respect to “population di-
wasps to differentiate to avoid competition on novel hosts, re-
vergence” rather than using the previously defined term “se- ducing temporal overlap and increasing their RI from ancestral
quential speciation” (14) to convey a focus on recent sympatric population (27). Indeed, eclosion time differences exist not only
host shifts for empirical testing and to remove any presumption within each species attacking different fly hosts but also between
of the evolutionary outcome of our study. Although our results parasitoid species sharing the same host (Table 2 and Fig. 7).
support the sequential divergence hypothesis, it is not possible to Therefore, a degree of temporal niche partitioning for oviposi-
discern the long-term evolutionary fate of the species we studied. tion sites exists among species, a common phenomenon among
We are unable to determine, with certainty, if these taxa will parasitoids sharing the same host (48, 49). Although speculative,
diverge further along the “speciation continuum” (44), represent this may permit coexistence of multiple species on shared fly
“frozen” multistate polymorphisms in evolutionary equilibrium hosts. Moreover, escape from parasitism itself may also favor
between selection and gene flow, or have already attained species host shifting in Rhagoletis (50). Thus, host plant, fly, and wasp
status. Thus, case studies of speciation must rely on inferences may play a tritrophic coevolutionary game of hide and seek that
drawn from comparisons involving a series of related population spins off several new life forms.
pairs at varying stages along the speciation continuum (44). We
therefore contend that host-specific populations of D. alloeum, Materials and Methods
D. mellea, and U. canaliculatus provide an opportunity to make Specimen Collection. Host fruits infested with parasitized flies were collected
inferences concerning the critical ecological attributes and adapta- from nine sites in the eastern United States from 2006 to 2012 (Fig. S1, Table
tions initiating sequential divergence that may result in speciation. S1, and SI Text). Six of these sites for U. canaliculatus and four for D. mellea

Hood et al. PNAS | Published online October 23, 2015 | E5987


Table 3. Pairwise estimates of both intraspecific (within) and interspecific (between) temporal isolation for wasps
attacking different and the same Rhagoletis host(s), respectively, at sympatric sites
Within-species
comparisons, % Between-species comparisons, %

Site no. Location Host comparison Da Uc Dm Host Da vs. Uc Da vs. Dm Uc vs. Dm

1 Grant, MI A vs. H 17 11 55 A 11 9 13
H 30 8 22
2 E. Lansing, MI A vs. H 34 30 12 A 31 51 17
A vs. S 5 52 H 74 74 4
H vs. S 53 47 S 85
3 Fennville, MI A vs. H 30 15 54 A 21 18 12
A vs. B 29 33 H 40 37 21
H vs. B 75 38 B 19
4 Dowagiac, MI A vs. H 5 3 A 53
H 31
5 Three Oaks, MI A vs. H 28

6 S. Bend, IN A vs. H 96 17 A 67
A vs. C 21 H 21
H vs. C 16
7 Urbana, IL A vs. H H 85

See Table S1 and Fig. S1 for site designations. Da, D. alloeum; Uc, U. canaliculatus; Dm, D. mellea; A, apple flies; H, hawthorn flies;
S, snowberry flies; B, blueberry flies; C, black cherry flies. Data for Da from Fennville, MI, are modified from Forbes et al. (14).

represent locations where two or more Rhagoletis taxa were sympatric allowed us to combine data across hosts and sites to test for genetic associa-
(<1.5 km apart) or in close proximity (<10 km apart). tions. We report the results for the main genotype effect and the genotype ×
host interaction because the host effect was significant in the ANOVA for all
Genetic Methods. Twenty and 21 pairs of unique microsatellite primers were loci. We also performed stepwise multiple regressions for all microsatellites
developed for U. canaliculatus and D. mellea, respectively (see data in the against normalized eclosion times separately for females and males.
Dryad database). Compared with other types of markers, microsatellites We tested for significant allele frequency differences between host-
have high mutation rates and levels of polymorphism, characteristics useful associated populations at sympatric sites using a nonparametric, Monte Carlo
for revealing subtleties in population genetic structure associated with rapid approach. Microsatellite genotypes were randomly resampled with re-
and recent divergence. We also PCR-amplified and sequenced a 540-bp region placement from the combined data set for a locus at a site to determine the
of the mtDNA Cytochrome Oxidase subunit I (COI). For further details on ge- probability from 100,000 replicates of generating a Nei’s genetic distance D
netic methodology, see SI Text. (54) greater than the observed value between populations by chance. We
also performed GLM analyses at sympatric sites to test for significant host
Analysis of Genetic Data. Microsatellite loci were tested for conformity to and latitude effects on microsatellite allele frequencies for each species. For
Hardy−Weinberg equilibrium (HWE) in females (diploid) using Markov chain the GLM analyses, allele frequency was modeled as a quasi-binomial vari-
randomization methods with default parameters in Micro-Checker (51). Of able, with host as a discrete factor and latitude as a continuous factor. The
the 41 microsatellites scored, no locus displayed significant deficiencies of GLM analysis required that we collapse the microsatellites to a diallelic
heterozygosity in more than two populations, implying markers represented dataset. To accomplish this, we pooled different combinations of alleles at
single-copy sequences that did not possess high frequencies of null alleles. each locus into two variant classes, using the diallelic combination that (i)
Furthermore, every haploid male possessed only a single allele for each generated the highest level of LD with other loci (Fig. S4) and (ii) explained
microsatellite, and we observed no obvious null allele at each locus. We the largest amount of variation in adult eclosion time (Table S6), similar to
estimated linkage disequilibrium (LD) between pairs of microsatellites sep- Michel et al. (18) and Powell et al. (22) (see data in the Dryad database).
arately for each host-associated population using Burrow’s composite Δ (52), We assessed the genetic relationships among populations for COI mtDNA
which does not assume HWE or require phased data but instead provides a by constructing maximum parsimony gene trees for U. canaliculatus and
joint metric of intralocus and interlocus disequilibria based solely on geno- D. mellea using Phylogeny Inference Package (PHYLIP) 3.69 (55). We exam-
type frequencies. Thus, Δ is equivalent to the LD parameter D under HWE (52). ined relationships for the nuclear-encoded microsatellites by constructing
We combined multilocus haplotypes of individual males together at random, unrooted NJ networks for U. canaliculatus and D. mellea based on Nei’s
based on their frequencies in a population, to construct diploid genotypes in pairwise genetic distances calculated between populations using PHYLIP
HWE to combine with females genotypes to estimate a single Δ value for each 3.69 (55), with bootstrap support determined by 10,000 replicates across loci.
population. These values were then transformed to a standardized correlation
coefficient and combined across populations using the method of Fisher (53) to Eclosion Study. The timing of adult eclosion was determined by monitoring fly
derive overall measures that were tested for significance by χ2 tests, as de- pupae once daily, following overwinter, for emerging flies and wasps. Adult
scribed by Weir (52). In general, LD was low between loci of both species, an wasp longevity was calculated by measuring the interval, in days, between
indication that most markers were evolving independently. However, certain adult emergence and death. We estimated the degree of allochronic isolation
microsatellites did display significant LD, and these loci were (i) arranged between wasp populations due to differences in eclosion time and adult
stepwise with one another to form arrays interconnecting the majority of according to the formula in SI Text.
markers genotyped for each species (Fig. S4) and (ii) used in association with
eclosion time relationships to pool alleles at loci for GLM analysis (Table S3). Field Observations and Behavioral Assays. To assess the site of assembly for
We tested for a relationship between eclosion timing and microsatellite mating, we conducted 45 h of observations of behavior for both U. canal-
genotypes separately for males and females by a two-way ANOVA with iculatus and D. mellea in stands of apple and hawthorn trees, and nearby
genotype and host as main effects. Our analyses were performed on the non-Rhagoletis host plants, at sympatric study sites in Fennville and Grant,
subset of individuals genotyped (Table S1) that were also phenotyped for MI (14) (SI Text). The responses of U. canaliculatus, D. mellea, and D. alloeum
eclosion timing (see data in the Dryad database). We concentrated on the to natal and nonnatal fruit volatiles were determined using a Y-tube olfac-
sympatric pairs of wasps attacking hawthorn and apple flies, transforming tometer based on the protocol previously developed for D. alloeum (14) (SI
the eclosion time for each individual by its SD from the mean eclosion time for Text). We estimated the degree of habitat isolation between wasp populations
each sex at the site from which it was collected. Normalizing eclosion times due to host fruit odor discrimination according to the formula in SI Text.

E5988 | www.pnas.org/cgi/doi/10.1073/pnas.1424717112 Hood et al.


PNAS PLUS
ACKNOWLEDGMENTS. We thank E. Archie, S. Berlocher, M. Brueseke, Funding was provided to J.L.F., A.A.F., and S.P.E. by the National Sci-
T. Brown, M. Doellman, C. Finn, M. Glover, K. Gomez, T. Kubler, ence Foundation (NSF) (DEB-1145573) and to G.R.H. by the Indiana Academy
J. McLachlan, P. Meyers, M. Pfrender, G. Ragland, D. Sanford, H. Schuler, of Science, the Entomological Society of America, Sigma Xi, and NSF (DEB-
C. Tait, and especially P. Morton for helpful comments and support. 1310850).

1. Coyne JA, Orr HA (2004) Speciation (Sinauer, Sunderland, MA). 34. Ragland GJ, et al. (2015) Differences in performance and transcriptome-wide gene
2. Feder JL, Flaxman SM, Egan SP, Comeault AA, Nosil P (2013) Geographic mode of expression associated with Rhagoletis (Diptera: Tephritidae) larvae feeding in alter-
speciation and genomic divergence. Annu Rev Ecol Syst 44:73–97. nate host fruit environments. Mol Ecol 24(11):2759–2776.
3. Pinho C, Hey J (2010) Divergence with gene flow: Models and data. Annu Rev Ecol Syst 35. Gavrilovic V, et al. (2007) Rhagoletis zephyria (Diptera: Tephritidae) in the Great Lakes
41:215–230. Basin: A native insect on native hosts? Ann Entomol Soc Am 100(4):474–482.
4. Erwin DH (2008) Macroevolution of ecosystem engineering, niche construction and 36. Barron AB (2001) The life and death of Hopkins’ host-selection principle. J Insect
diversity. Trends Ecol Evol 23(6):304–310. Behav 14(6):725–737.
5. Butterfield NJ (2001) The Ecology of the Cambrian Region (Columbia Univ Press, New 37. Turlings TCJ, et al. (1993) Learning of host-finding cues by hymenopterous parasit-
York). oids. Insect Learning: Ecological and Evolutionary Perspectives, eds Papaj DR, Lewis A
6. Bush AM, Bambach RK, Erwin DH (2011) Ecosphere utilization during the Ediacaran (Chapman & Hall, New York), pp 51–78.
Radiation and the Cambrian Eco-explosion. Quantifying the Evolution of Early Life: 38. Storeck K, Poppy GM, van Emden HF, Powell W (2000) The role of plant chemical cues
Numerical Approaches to the Evalucation of Fossils and Ancient Ecosystems, eds in determining host preference in the generalist aphid parasitoid Aphidius colemani.
Laflamme M, Schiffbauer J, Dornbos S (Springer, Doredrecht, The Netherlands), pp Entomol Exp Appl 97(1):41–46.
111–133. 39. Bush GL (1993) A reaffirmation of Santa Rosalia, or why there are so many kinds of
7. Emerson BC, Kolm N (2005) Species diversity can drive speciation. Nature 434(7036): small animals? Evolutionary Patterns and Processes, eds Edwards D, Lees DR (Aca-
demic, New York), pp 229–249.
1015–1017.
40. Price PW (1980) Evolutionary Biology of Parasites (Princeton Univ Press, Princeton).
8. Strong DR, Lawton JA, Southwood TRE (1984) Insects on Plants: Community Patterns
41. Hawkins BA, Lawton JH (1987) Species richness for parasitoids of British phytopha-
and Mechanisms (Blackwell Sci, Oxford).
gous insects. Nature 326(6115):788–790.
9. Mitter C, Farrell B, Wiegmann B (1988) The phylogenetic study of adaptive zones: Has
42. Schluter D (2009) Evidence for ecological speciation and its alternative. Science
phytophagy promoted insect diversification? Am Nat 132(1):107–128.
323(5915):737–741.
10. Whitham TG, et al. (2006) A framework for community and ecosystem genetics: From
43. Nosil P, Flaxman SM (2011) Conditions for mutation-order speciation. Proc Biol Sci
genes to ecosystems. Nat Rev Genet 7(7):510–523.
278(1704):399–407.
11. Whitham TG, et al. (2008) Extending genomics to natural communities and ecosys-
44. Feder JL, Egan SP, Nosil P (2012) The genomics of speciation-with-gene-flow. Trends

EVOLUTION
tems. Science 320(5875):492–495.
Genet 28(7):342–350.
12. Abrahamson WG, Blair CP (2008) Sequential radiation through host-race formation:
45. Flaxman SM, Wacholder AC, Feder JL, Nosil P (2014) Theoretical models of the in-
Herbivore diversity leads to diversity in natural enemies. Specialization, Speciation, fluence of genomic architecture on the dynamics of speciation. Mol Ecol 23(16):
and Radiation: The Evolutionary Biology of Herbivorous Insects, ed Tilmon KJ (Univ 4074–4088.
California Press, Oakland, CA), pp 188–202. 46. Reissig WH, Smith DC (1978) Bionomics of Rhagoletis pomonella in Crataegus. Ann
13. Feder JL, Forbes AA (2010) Sequential divergence and the diversity of parasitic insects. Entomol Soc Am 71(2):155–159.
Ecol Entomol 35(Suppl 1):67–76. 47. Prokopy RJ, Roitberg BD (1984) Foraging behavior of true fruit flies. Am Sci 72(1):
14. Forbes AA, Powell THQ, Stelinski LL, Smith JJ, Feder JL (2009) Sequential sympatric 41–49.
speciation across trophic levels. Science 323(5915):776–779. 48. Weis AR, Abrahamson WG (1985) Potential selective pressures by parasitoids on a
15. Stireman JO, Nason JD, Heard SB, Seehawer JM (2006) Cascading host-associated plant-herbivore interaction. Ecology 66(4):1261–1269.
genetic differentiation in parasitoids of phytophagous insects. Proc Biol Sci 273(1586): 49. Hackett-Jones E, Cobbold C, White A (2009) Coexistence of multiple parasitoids on a
523–530. single host due to differences in parasitoid phenology. Theor Ecol 2(1):19–31.
16. Drès M, Mallet J (2002) Host races in plant-feeding insects and their importance in 50. Feder JL (1995) The effects of parasitoids on sympatric host races of the apple maggot
sympatric speciation. Philos Trans R Soc Lond B Biol Sci 357(1420):471–492. fly, Rhagoletis pomonella (Diptera: Tephritidae). Ecology 76(3):801–813.
17. Berlocher SH, Feder JL (2002) Sympatric speciation in phytophagous insects: Moving 51. Van Oosterhout C, Hutchinson WF, Wills DPM, Shirley P (2004) Micro-Checker: Soft-
beyond controversy? Annu Rev Entomol 47:773–815. ware for identifying and correcting genotyping errors in microsatellite data. Mol Ecol
18. Michel AP, et al. (2010) Widespread genomic divergence during sympatric speciation. Notes 4(3):535–538.
Proc Natl Acad Sci USA 107(21):9724–9729. 52. Weir BS (1979) Inferences about linkage disequilibrium. Biometrics 35(1):235–254.
19. Egan SP, et al. (2015) Experimental evidence of genome-wide impact of ecological 53. Sokal RR, Rohlf FJ (2012) Biometry: the Principles and Practice of Statistics in Biological
selection during early stages of speciation-with-gene-flow. Ecol Lett 18(8):817–825. Research (Freeman, New York).
20. Berlocher SH (2000) Radiation and divergence in the Rhagoletis pomonella species 54. Nei M (1972) Genetic distance between populations. Am Nat 106(949):283–292.
group: Inferences from allozymes. Evolution 54(2):543–557. 55. Felsenstein J (2005) PHYLIP (Phylogeny Inference Package) Version 3.6 (Dep Genome
21. Xie X, et al. (2008) Radiation and divergence in the Rhagoletis pomonella species Sci, Univ Washington, Seattle, WA).
complex: Inferences from DNA sequence data. J Evol Biol 21(3):900–913. 56. Lathrop FH, Newton RC (1933) The biology of Opius melleus Gahan, a parasite of the
22. Powell THQ, et al. (2013) Genetic divergence along the speciation continuum: The blueberry maggot. J Agric Res 46(2):143–160.
transition from host race to species in rhagoletis (Diptera: tephritidae). Evolution 57. Wharton WA, Marsh PM (1978) New world Opiinae (Hymenoptera: Braconidae)
67(9):2561–2576. parasitic on Tephritidae (Diptera). J Wash Acad Sci 68(4):147–167.
23. Bush GL (1969) Sympatric host race formation and speciation in frugivorous flies of 58. Wharton RA (1997) Generic relationships of opine Braconidae (Hymenoptera) para-
the genus Rhagoletis (Diptera: Tephritidae). Evolution 23(2):237–251. sitic on fruit infesting Tephritidae (Diptera). Contrib Am Entomol Inst 30:1–53.
24. Forbes AA, Hood GR, Feder JL (2010) Geographic ranges and host breadths of para- 59. Wharton RA, Yoder MJ (2015) Parasitoids of fruit-infesting tephritidae. Available at
paroffit.org/public/site/paroffit/home. Accessed July 2, 2015.
sitoid wasps associated with the Rhagoletis pomonella (Diptera: Tephritidae) species
60. Simon C, et al. (1994) Evolution, weighting and phylogenetic utility of mitochondrial
complex. Ann Entomol Soc Am 103(6):908–915.
gene sequences and a compilation of conserved polymerase chain reaction primers.
25. Hood GR, Egan SP, Feder JL (2012) Interspecific competition and speciation in en-
Ann Entomol Soc Am 87(6):651–701.
doparasitoids. Evol Biol 39(2):219–230.
61. Nojima S, Linn C, Jr, Morris B, Zhang A, Roelofs W (2003) Identification of host fruit
26. Prokopy RJ, Bennett EW, Bush GL (1971) Mating behavior in Rhagoletis pomonella
volatiles from hawthorn (Crataegus spp.) attractive to hawthorn-origin Rhagoletis
(Diptera: Tephritidae). I. Site of assembly. Can Entomol 103(10):1405–1409.
pomonella flies. J Chem Ecol 29(2):321–336.
27. Linn C, Jr, et al. (2003) Fruit odor discrimination and sympatric host race formation in
62. Nojima S, Linn C, Jr, Roelofs W (2003) Identification of host fruit volatiles from
Rhagoletis. Proc Natl Acad Sci USA 100(20):11490–11493.
flowering dogwood (Cornus florida) attractive to dogwood-origin Rhagoletis po-
28. Forbes AA, Fisher J, Feder JL (2005) Habitat avoidance: Overlooking an important
monella flies. J Chem Ecol 29(10):2347–2357.
aspect of host-specific mating and sympatric speciation? Evolution 59(7):1552–1559. 63. Powell THQ, Cha DH, Linn CE, Jr, Feder JL (2012) On the scent of standing variation for
29. Feder JL, et al. (1994) Host fidelity is an effective premating barrier between sym-
speciation: Behavioral evidence for native sympatric host races of Rhagoletis pomo-
patric races of the apple maggot fly. Proc Natl Acad Sci USA 91(17):7990–7994. nella (Diptera: Tephritidae) in the southern United States. Evolution 66(9):2739–2756.
30. Linn CE, Jr, et al. (2004) Postzygotic isolating factor in sympatric speciation in Rha- 64. Powell THQ, Forbes AA, Hood GR, Feder JL (2014) Ecological adaptation and re-
goletis flies: Reduced response of hybrids to parental host-fruit odors. Proc Natl Acad productive isolation in sympatry: Genetic and phenotypic evidence for native host
Sci USA 101(51):17753–17758. races of Rhagoletis pomonella. Mol Ecol 23(3):688–704.
31. Dambroski HR, Feder JL (2007) Host plant and latitude-related diapause variation in 65. Ramsey J, Bradshaw HD, Jr, Schemske DW (2003) Components of reproductive iso-
Rhagoletis pomonella: A test for multifaceted life history adaptation on different lation between the monkeyflowers Mimulus lewisii and M. cardinalis (Phrymaceae).
stages of diapause development. J Evol Biol 20(6):2101–2112. Evolution 57(7):1520–1534.
32. Filchak KE, Roethele JB, Feder JL (2000) Natural selection and sympatric divergence in 66. Stelinski LL, Oakleaf R, Rodriguez-Saona C (2007) Oviposition-deterring pheromone
the apple maggot Rhagoletis pomonella. Nature 407(6805):739–742. deposited on blueberry fruit by the parasitic wasp, Diachasma alloeum. Behaviour
33. Bierbaum TJ, Bush GL (1990) Genetic differentiation in the viability of sibling species 144(4):429–445.
of Rhagoletis fruit-flies on host plants, and the influence of reduced hybrid viability 67. Jones TH, Godfray HCJ, Hassell MP (1996) Relative movement of a tephritid fly and its
on reproductive isolation. Entomol Exp Appl 55(2):105–118. parasitoid wasps. Oecologia 106(3):317–324.

Hood et al. PNAS | Published online October 23, 2015 | E5989

Você também pode gostar