Molecular Ecology (2014) 23, 1333–1347 doi: 10.1111/mec.12655

SPECIAL ISSUE: NATURE’S MICROBIOME Partner associations across sympatric broad-headed bug species and their environmentally acquired bacterial symbionts

J. R. GARCIA,* A. M. LAUGHTON,*† Z. MALIK,* B. J. PARKER,* C. TRINCOT,* S. S. L. CHIANG,* E. CHUNG* and N. M. GERARDO* *Biology Department, Emory University, O. Wayne Rollins Research Center, 1510 Clifton Rd, Atlanta, GA 30322, USA †School of Biological and Chemical Sciences, Queen Mary University of London, Mile End Road, London, E1 4NS, UK

Abstract Many organisms have intimate associations with beneficial microbes acquired from the environment. These host–symbiont associations can be specific and stable, but they are prone to lower partner specificity and more partner-switching than vertically transmit- ted mutualisms. To investigate partner specificity in an environmentally acquired symbiosis, we used 16S rRNA gene and multilocus sequencing to survey the bacterial population in the bacteria-harbouring organ (crypts) of 49 individuals across four sympatric broad-headed bug species (, A. conspersus, A. tomen- tosus and Megalotomus quinquespinosus). Similar to other insect–bacteria associations, Burkholderia spp. were the most common residents of the crypts in all four insect spe- cies (77.2% of recovered sequences). Burkholderia presence was associated with pro- longed survival to adulthood in A. tomentosus, suggesting a beneficial role of these specialized associations. Burkholderia were also found in environmental reservoirs in the ’ habitat, which may facilitate acquisition by insects by increasing Burk- holderia–insect encounters. Symbiont establishment could also be facilitated by resis- tance to insect defences; zone of inhibition assays demonstrated that Burkholderia and other bacteria isolated from crypts are resistant to insect defences that limit growth of Escherichia coli. Alternatively, the insects’ defences may not efficiently kill a broad range of bacteria. Although the symbiosis is targeted to Burkholderia, the insects’ crypts housed other bacteria, including non-Burkholderiaceae species. There is no sig- nificant effect of host insect species on Burkholderia distribution, suggesting a lack of strong partner specificity at finer scales. The presence of frequent partner-switching between sympatric insects and their symbionts likely prevents tight co-evolutionary dynamics.

Keywords: alydids, Burkholderia, insects, Lespedeza, mutualism Received 12 April 2013; revision received 12 December 2013; accepted 13 December 2013

genotype in offspring is largely determined by the sym- Introduction biont genotype present in parents, frequently leading to Most host beneficial symbiotic microbes that a long history of codiversification between host and are required for survival or provide key benefits under microbe (Baumann et al. 1995). Conversely, horizontal specific environmental or physiological conditions mutualisms are open systems where offspring must (McFall-Ngai et al. 2013). In vertical mutualisms, acquire symbiotic microbes from the environment each symbiotic microbes are passed from a parent, usually generation. In these mutualisms, the symbiont genotype the mother, to offspring. Consequently, the symbiont present in each host can be determined by a variety of host (Davidson et al. 2004; Sachs et al. 2010; Troll et al. Correspondence: Justine Garcia, Fax: 404 727 2880; 2010; Nyholm & Graf 2012), symbiont (Davidson et al. E-mail: [email protected] 2004; Ruby 2008; Troll et al. 2010) and environmental

© 2014 John Wiley & Sons Ltd 1334 J. R. GARCIA ET AL.

(Lee & Ruby 1992; Finney et al. 2010; Porter & Rice found in individuals across multiple locations. To find 2012) factors. In most systems, it is currently unclear potential environmental reservoirs for insect symbionts, how these variables interact to determine which symbi- we preferentially isolated and sequenced Burkholderia in otic microbes are present in an individual host. the soil and in association with the insects’ food plants Several pentatomomorphan stinkbugs and their sym- (Lespedeza spp.). Finally, through zone of inhibition bionts are ideal for investigating the determinants of assays, we tested whether Burkholderia and other crypt- partner specificity in hosts with horizontally acquired isolated bacteria were resistant to insect hemolymph- symbionts. Many stinkbugs in the Lygaeoidea and Co- based defences, one potential mechanism by which host reoidea superfamilies harbour bacteria of the genus factors could facilitate partner specificity. Burkholderia in midgut crypts, specialized structures of the gut (Fig. S1, Supporting information; (Olivier-Espe- Materials and methods jel et al. 2011; Kikuchi et al. 2011b; Boucias et al. 2012). Symbiont detection assays and rearing experiments in Isolation of bacteria and bacterial DNA from insects, Riptortus pedestris (Kikuchi et al. 2007) and Thasus neocal- soil and plants ifornicus (Olivier-Espejel et al. 2011) indicate insects hatch symbiont-free and acquire Burkholderia from the Male and female broad-headed bugs were collected soil or the rhizosphere later in juvenile development from 12 sites in Georgia and North Carolina, United (Kikuchi et al. 2011a). It is likely that this mode of trans- States (Table 1), in 2010, 2011 and 2012 by hand or with mission has been conserved in other insects that har- nets. Insects were frozen immediately at À80 °Cor bour Burkholderia as phylogenetic analyses suggest that maintained in plastic cages. Broad-headed bugs were there are no clear host–symbiont co-evolutionary pat- kept on a 16-h light: 8-h dark cycle at 25 °Cor28°C terns across the Lygaeoidea and superfami- and fed ad libitum autoclaved Lespedeza capitata seeds lies (Kikuchi et al. 2011b). It is unclear what each and one of two liquid diets: (i) water with 0.5 g/L partner receives from this association, although one ascorbic acid or (ii) 0.5 g/L ascorbic acid, 0.25 g/L cys- hypothesis is that Burkholderia provide nutrients lacking teine and 0.025% blue food colouring (added to easily in the hosts’ diet as one host, Riptortus pedestris, grows identify the digestive tract in dissected insects). Insects larger and heavier and has higher survival when it har- were killed and surface-sterilized in 70% ethanol for bours Burkholderia (Kikuchi et al. 2007, 2012), and other 4 min and then rinsed with sterilized water. Dissections hosts develop more quickly and live longer (Olivier-Es- were performed in one of two ways: (i) the abdomen pejel et al. 2011; Boucias et al. 2012). was cut open using sterile technique, and midgut crypts A broad survey of Burkholderia from 39 stinkbug spe- were removed from the rest of the gut, briefly rinsed in cies from the pentatomomorphan infraorder indicate 70% ethanol and placed in Carlson’s solution [0.7% that all species are phylogenetically constrained to three NaCl, 0.02% KCl, 0.02% CaCl2 hydrate, 0.01% MgCl2 clades largely composed of insect-associated Burkholde- hexahydrate, 0.02% NaH2PO4, 0.012% NaHCO3, 0.8% ria. Symbionts differ between insect species, but no glucose; (Mitsuhashi 2002)] or (ii) the entire abdomen coherent pattern of host–symbiont coassociation is was clipped with sterile microscissors from the thorax apparent (Kikuchi et al. 2011b). There has not, however, and placed in Carlson’s solution. Tissues from nonfroz- been a broad survey of bacteria isolated from sympatric en broad-headed bugs were homogenized with a sterile insect species collected from multiple locations, which pestle and plated on Luria-Bertani (LB) media. Plates could reveal host or geographical factors that may were incubated for 2 days at 27 °C, and nine random shape partner specificity. colonies were then subcultured and identified by We investigated this symbiosis in four sympatric sequencing the 16S rRNA gene. Tissues from frozen broad-headed bug species, Alydus conspersus, A. tomen- broad-headed bugs were thawed and used for DNA tosus, A. calcaratus and Megalotomus quinquespinosus extraction and cloning (see below). (order , superfamily Coreoidae, family Alydi- Broad-headed bugs frequently feed on the seeds of dae) from 12 meadows in the Southeastern United bush clover legumes (Lespedeza spp.), although they States. First, to determine whether Burkholderia provide have been reported to feed on additional food sources benefits to broad-headed bug hosts, as they do for other available in the natural mixed legume fields that they insects (Kikuchi et al. 2007; Olivier-Espejel et al. 2011; inhabit throughout North America (Schaefer 1980; Ven- Boucias et al. 2012), we assessed developmental time tura et al. 2000). Because Lespedeza spp. can host a vari- and survival of Alydus tomentosus with and without ety of Burkholderia spp. in nitrogen-fixing root nodules Burkholderia. Then, to address patterns of insect–bacteria (Palaniappan et al. 2010) and other tissues (Compant association across these sympatric species, we used 16S et al. 2008), we surveyed the bacteria within the root rRNA gene and MLST sequencing to identify bacteria nodules of this host plant to determine whether the

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1335

Table 1 Number of insect, Lespedeza and soil samples collected across sampling locations

† ‡ Collection site name (Abbreviation) Location Acal Acon Atom Mquin Les Soil

Chattahoochee Fish Hatchery (CFH) Blue Ridge, GA 1 Clairmont Campus (CM) Atlanta, GA 4 1 Frances Meadows Aquatic Center (FM) Gainesville, GA 4 1 Houston Mill Rd. (HM) Atlanta, GA 1 Joy Baptist Church (JOY) Wiley, GA 1 3 Laboratory-bred (LAB) NA 2 Lake Chatuge (LC) Hayesville, NC 1 Limestone Pkwy. (LIME) Gainesville, GA 2 2 Morningside Nature Preserve (MNP) Atlanta, GA 4 1 Skeenah Gap Rd. (SKEE) Blue Ridge, GA 1 Snapfinger Rd. (SNAP) Atlanta, GA 5 Songbird Meadow (SM) Stone Mountain, GA 4 5 7 5 10 U.S. 441 (WY) Wiley, GA 3 1 Total 15 13 16 5 5 15

Acal, Alydus calcaratus; Acon, Alydus conspersus; Atom, Alydus tomentosus; Mquin, Megalotomus quinquespinosus; Les, Lespedeza cuneata. † Number of root nodules. ‡ Number of 0.1 g subsamples taken from larger samples. plants could serve as a reservoir or alternative host for subsample was vortexed vigorously for 2 min and insect gut symbionts. Sericea lespedezas (Lespedeza cune- briefly centrifuged in a microcentrifuge. To increase the ata), common host plants of broad-headed bugs in likelihood of finding Burkholderia, 50 lL of supernatant Georgia, were collected from Songbird Meadow, a site from each subsample was plated on selective glucose- where many broad-headed bugs were collected. Plants based rhizobium-defined medium (Sachs et al. 2009) were excavated with a shovel and individually trans- with 200 mg/L cycloheximide and 2 mg/mL crystal ported in plastic bags. Dissections of the root nodules violet. Plates were incubated for 2 days at 27 °C, and were performed immediately upon return to the labora- unique colonies were then subcultured and identified tory, following previously established protocols (Sachs by sequencing the 16S rRNA gene. et al. 2009). Roots were separated from plants and DNA from cultivated bacteria isolated from insects, washed with deionized water to remove loose soil. soil and plants was extracted with the Qiagen DNeasy After drying, root nodules were cut from the roots with Blood and Tissue Kit from cultures grown in LB over- sterile tools and sterilized by washing in 5% bleach for night following the manufacturer’s protocol (lysis step 2 min. Nodules were then rinsed in distilled water, 30 min to 16 h). In addition, for direct cloning of bacte- placed in Carlson’s solution, homogenized with a sterile rial PCR products from insects, whole insect abdomens pestle and plated on LB media. An approximately 1-cm were surface-sterilized as above and placed in buffer section of root adjacent to each nodule was also ATL, frozen in liquid nitrogen and crushed with a ster- removed, sterilized, homogenized and plated on LB ile pestle before DNA extraction following the manufac- media to serve as a negative control. If there was bacte- turer’s protocol, with the lysing incubation step rial growth from a non-nodule root section, the corre- performed overnight. All DNA extracts were stored at sponding root nodule bacterial cultures were not used. À20 °C until further use. Plates were incubated for 2 days at 27 °C, and four ran- dom colonies were then subcultured and identified by Development and survival assays sequencing the 16S rRNA gene. Bulk soil near Lespedeza spp. patches (~8–15 cm away The effects of natural Burkholderia infections were tested from any plant tissue) was collected with a shovel and in A. tomentosus by comparing development time and transported in individual plastic bags from two sites in lifespan of Burkholderia-positive (symbiotic) and Burk- Georgia where broad-headed bugs are common (Song- holderia-negative (aposymbiotic) individuals. Eggs were bird Meadow and Snapfinger Drive). Each sample con- taken from a general laboratory stock population of sisted of the top six inches of soil, which was coarsely mixed genotypes, in which adults were allowed to mate homogenized in the plastic bag. Five 0.1 g subsamples freely. Two hundred and four eggs were divided into were taken from each soil sample and were diluted in two treatments: (i) a Burkholderia-positive treatment in 1 mL sterile phosphate-buffered saline (PBS). Each which insects were housed with soil that had

© 2014 John Wiley & Sons Ltd 1336 J. R. GARCIA ET AL. previously been found to harbour Burkholderia that was 16S rRNA gene sequencing and multilocus sequence collected from Songbird Meadow during insect collec- typing tions; and (ii) a Burkholderia-negative treatment in which The 16S rRNA gene was used to survey all bacterial insects were housed with autoclaved soil to rear them constituents in this study, facilitating comparison to aposymbiotically. Insects were inoculated through expo- previously sequenced insect symbionts (Kikuchi et al. sure to soil instead of pure bacterial culture because 2011b). Amplicons were produced using the 5 PRIME exposure to soil would allow the broad-headed bugs to MasterTaq PCR kit in 22 lL reactions containing 2.5 lL acquire ‘preferred’ species. Eggs for both treatments 2+ 5X MasterTaq Buffer with Mg ,10mM of each dNTP were sterilized in 70% ethanol for 2 min followed by (GenScript), 5 lM each primer [27F: 3′-AGA- 10% bleach for 2 min, rinsed in sterilized DI water and GTTTGATCCTGGCTCAG-5′; and a slightly modified dried on kimwipes. Each egg was placed individually 1492R: 3′-GGYTACCTTGTTACGACTT-5′; the under- in a sterile 3 cm 9 7.5 cm 9 7.5 cm plastic box with lined base was modified from a T to a Y; (Lane 1991)], autoclaved Lespedeza seeds, a sterile sponge saturated 5 lL5XTaqMaster PCR enhancer, 1 unit Taq DNA with nutrient solution (filter-sterilized water with 0.5 g/ polymerase and 2 lL genomic DNA. The 1492R primer L ascorbic acid, 0.25 g/L cysteine and 0.025% blue food was modified to better bind Burkholderia 16S rRNA gene colouring) and the appropriate type of soil. Broad- sequences by reducing mismatches. A negative control headed bugs are hemimetabolous insects that go was included for each PCR assay by replacing the DNA through five nymphal stages before reaching adulthood; with molecular-grade water. This reaction was dena- nymphs moult between each instar and between the tured at 94 °C for two min, then cycled for 35 cycles at fifth instar and adulthood. Insects were surveyed every 94 °C for 30 s, 55 °C for 30 s and 72 °C for 1 min, fol- day after hatching to identify changes in instar (indi- lowed by final extension at 72 °C for 3 min. Amplicons cated by the presence of exuvia) and for survival. The were purified using the QIAquick PCR Purification Kit Burkholderia infection status of each broad-headed bug (Qiagen) according to the manufacturer’s protocol. PCR was confirmed by Burkholderia-specific PCR with a pre- products from cultivated bacteria were then sequenced viously published thermocycling program and set of directly. PCR products amplified from insect crypts and primers specific to the Burkholderia genus (Kikuchi et al. abdomens were cloned using the Original TA Cloning 2005) on DNA extracted from the crypts. When a sam- Kit (Invitrogen) according to the manufacturer’s proto- ple had no amplification in this PCR assay, it was col. Ten white colonies were chosen randomly for assumed to be symbiont-negative. When the PCR sequencing in the forward direction with the 27F or results conflicted with the Burkholderia-exposure treat- M13F primer. ment, insects were reclassified according to the PCR To determine whether 16S rRNA sequencing cap- results (which occurred in 17% of insects). tured diversity and relationships found using other The number of days between moults was analysed gene sequences, a subset of the cultivated Burkholderia using linear mixed-effects (LME) models with the lme species isolated from broad-headed bug crypts, includ- function of the nlme package (Pinheiro et al. 2013) in R ing multiple representatives from OTU 2 (see Fig. 2), v. 2.11 (R Development Core Team 2009). Days were further characterized using the ATP synthase between moults were log-transformed and tested for (atpD), glutamate synthase (gltB), leader peptidase normality to ensure they fit model assumptions. The (lepA) and recombinase A (recA) genes based on a mul- model included the random effect of individual to tilocus sequence typing (MLST) assay previously devel- account for repeated measures of the same individuals. oped for the Burkholderia cepacia complex and other Burkholderia infection status and developmental stage similar species (Spilker et al. 2009). PCR amplification (instar) were treated as fixed effects. We derived a min- and purification were conducted as above except 10 lM imal model using stepwise term removal and deter- of each primer was used. All reactions were incubated mined the significance of each factor using likelihood as done previously (Spilker et al. 2009). Briefly, atpD ratio tests with the ANOVA function. Terms were was annealed at 56 °C, gltB and recA at 58 °C, and lepA retained if their removal significantly reduced the at 55 °C. PCRs for all genes were cycled for 30 cycles. explanatory power of a model. Survival data were anal- Amplicons were directly sequenced with the forward ysed using nonparametric survival models with a COXPH primer of each primer set. distribution using the survival package (Therneau 2013) in R. Data were tested for proportional hazards and for nonlinearity to ensure that data fit model assumptions. Sequence assembly and alignment A minimal model was derived using stepwise term Only 16S rRNA gene sequences that had at least 600 removal, and models were compared using chi-squared bases with Phred quality scores of 20 or greater were tests.

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1337 used in downstream analyses. SeqMan Pro version except for the following: genthreshfortopoterm = 5000, 10.1.1 (DNASTAR) was used to remove vector sequence treerejectionthresholds = 50. The final MLST tree topol- and trim read ends. Trimmed sequences were aligned ogy, based on ML, was generated in Garli using a gen- to the standard Hugenholtz 7682 character alignment in eral time reversible model with empirical base Greengenes (DeSantis et al. 2006b) using the NAST frequencies, gamma-distributed rate variation among aligner (DeSantis et al. 2006a). All sequences were iden- sites, a proportion of invariant sites estimated from the tified with the Ribosomal Database Project, Greengenes data and four substitution rate categories. Starting trees and NCBI taxonomies in Greengenes Classifier. were obtained via stepwise addition. Likelihood boot- Sequences that were identified as Burkholderiaceae were strap values are based on 1000 bootstrap replicates and imported into mothur (Schloss et al. 2009). Thirty-five run with the same parameters as in the 16S tree. sequences that were not complete between Greengenes alignment positions 204 and 5809 were removed from Statistical comparison across insect hosts and further analysis. This included sequences that were geographical locations sequenced in the reverse direction with the M13F pri- mer. The remaining sequences, as well as Burkholderia- Community structure of the Burkholderiaceae 16S rRNA ceae 16S rRNA gene sequences from other studies of gene sequences from the four sympatric insect hosts true bug symbionts (refer to Table S1, Supporting infor- was analysed using the vegan package (Oksanen et al. mation), were clustered into operational taxonomic 2013) in R. Permutational MANOVA (the adonis function units (OTUs) with the average neighbour algorithm at in vegan) was used to test the effect of species and 97% similarity cut-off level. A heatmap was created location on OTU presence and absence. The matrix con- using mothur with a linear scale. sisted of each sample (individual bug) by OTU pres- All genes in the MLST scheme (atpD, gltB, lepA, recA) ence/absence for the four broad-headed bug species in were fully sequenced with the forward primer and this study. The full model included species, location trimmed as above in SeqMan Pro. Because most of the and a species–location interaction, and nonsignificant publicly available Burkholderia cepacia complex MLST variables were sequentially removed to arrive at the sequences were produced according to an earlier final model, which only retained location as a signifi- scheme that generated shorter gene fragments (Baldwin cant variable. Permutational MANOVA was run using the et al. 2005), the sequences produced in this study were Jaccard index, a robust method for binary data with trimmed by hand to the same length as those in the uneven sampling, to calculate pairwise distances from pubMLST database (http://pubmlst.org/bcc/) after the matrix. P-values were calculated from 10 000 per- they were aligned with MUSCLE. mutations. The analysis was restricted to only sequences classified as Burkholderiaceae. Permutational MANOVA was not performed with Burkholderiaceae from Phylogenetic analyses insect species outside our study because individual Separate phylogenies were reconstructed for representa- sample information was not available for these samples. tives of all Burkholderiaceae 16S rRNA gene OTUs and Using the same set of Burkholderiaceae sequences as for a subset of cultivated Burkholderia species based on for the PERMANOVA analysis, analysis of molecular vari- MLST sequence data. Nucleotide substitution models ance (AMOVA) in mothur was performed to test for sig- and parameters for phylogenetic analysis were chosen nificant symbiont genetic differentiation across the four using jModelTest (Darriba et al. 2012). Tree topology sympatric insect species and separately across locations. was explored using distance, parsimony, maximum- In addition, two AMOVAs were performed using samples likelihood and Bayesian models in mothur, PAUP* collected as part of this study and samples collected (Swofford 2002), Garli (Zwickl 2006) and Mr. Bayes previously from four other insect species (Table S1, (Huelsenbeck & Ronquist 2001). Tree topologies were Supporting information); one tested the significance of largely consistent between the models, and a final, boot- insect species, the other tested the significance of loca- strapped maximum-likelihood phylogeny is presented. tion (North America versus Asia). Nineteen host species The final 16S rRNA gene tree topology was a maxi- classified as ‘others’ in Fig. 3 (Kikuchi et al. 2011b) were mum-likelihood phylogeny generated using a general not included in these analyses because there was only time reversible model with empirical base frequencies, one sequence per species available. All AMOVAs were gamma-distributed rate variation among sites and four performed with sequences (from one or more libraries) substitution rate categories in Garli. Starting trees were pooled according to host species or location. Each AM- obtained via stepwise addition. Likelihood bootstrap OVA consisted of 10 000 iterations, and significance was values are based on 500 bootstrap replicates run in Garli determined after Bonferroni correction. Because the with a GTR+G model and default Garli parameters effort to culture bacteria from the various insect species

© 2014 John Wiley & Sons Ltd 1338 J. R. GARCIA ET AL. was not even, only sequences derived through PCR and Following this experiment, the susceptibility of five cloning from insect gut and crypt tissues were included additional bacteria to induced hemolymph defences of in these analyses. Alydus spp. was tested. The following additional bacte- ria were used in this assay: (i) Cupriavidus sp. BHJ32i (Fig. 2, Table S2, Supporting information), a bacterium Zone of inhibition assays that is less common but still found in broad-headed The susceptibility of multiple bacteria to defences in the bug crypts; (ii) Lactococcus sp. BHJ32a, a non-Burkholder- hemolymph of Alydus spp. was tested using zone of iaceae bacterium isolated from a crypt sample; (iii) Burk- inhibition (ZOI) plate assays. As many components of holderia fungorum str. Snap2c (OTU 21 in Fig. 3, Table the insect immune system [e.g. antimicrobial peptides S2, Supporting information), a bacterium isolated from (Lemaitre & Hoffmann 2007)] are induced in response to soil, but not crypts; (iv) Burkholderia sp. Les1n2i (OTU 7 signals of infection such as cell wall components like in Fig. 3, Table S2, Supporting information), a bacte- peptidoglycan, the susceptibilities of bacteria to induced rium isolated from a Lespedeza root nodule that clusters and constitutive host defences were tested first. Twenty with insect symbionts; and (v) Serratia marcescens str. Alydus conspersus were split into two treatment groups: RHoD, a pathogenic bacterium of many insects. Sixty- (i) immune-challenged (injected with 1 lL of 0.5 mg/mL five insects were injected with 1 lL LPS as described E. coli-derived lipopolysaccharide (LPS), a bacterial cell above and incubated for 24 h. Hemolymph was col- wall component, dissolved in insect Ringer’s solution lected via a neck wound as above and then randomly

[128 mM NaCl, 18 mM CaCl2, 1.3 mM KCl, 2.3 mM NaH- assigned to one of the five bacteria listed above. In

CO3; (Mitsuhashi 2002)]; and (ii) no-stab controls. addition, three samples were assigned to an E. coli plate Immune-challenged insects were injected with LPS to ensure inhibition to this bacterium was present in between the second and third abdominal tergites using a these samples as before. Four samples were assigned to pulled glass needle and incubated under normal rearing a Burkholderia 11BH497 plate to ensure there was no conditions for 24 h post-treatment. Control insects were inhibition to this bacterium. When more than 1 lL was handled in the same manner, but not stabbed, and incu- present in a sample, the remaining hemolymph was bated for 24 h. Hemolymph from insects in both treat- randomly assigned to another of the five bacteria plates. ments was extracted via a neck wound swabbed with Phosphate-buffered solution (PBS) was added to one 70% ethanol and immediately frozen until required well on each plate to serve as a negative control. ZOI (À80 °C). ZOI assays were carried out following Moret plates were incubated and scored as above. and Schmid-Hempel (2000). Briefly, samples were defrosted on ice, and 1 lL of hemolymph was plated Results into small wells punched into 1% LB agar saturated with either E. coli MG1655 or Burkholderia 11BH497 (species Effect of Burkholderia on host survival and isolated from an Alydus calcaratus crypt which groups in development an insect symbiont clade) and incubated overnight at 37 °C and 27 °C, respectively. Plates were scored for For Alydus tomentosus, the number of days between mo- inhibition 24 h later. Wells with no inhibition were ults was significantly influenced by instar (P = 0.052), scored as zero. When inhibition was present, the radius but not by Burkholderia infection status (P = 0.304; and area of the zone of clearance were measured under Fig. 1A). However, Burkholderia infection significantly a dissecting scope and calculated with ImageJ (Abramoff increased survival (v2 = 4.42, d.f. = 1, P = 0.036; & Magalhaes~ 2004). Fig. 1B). Infection with Burkholderia increased survival The susceptibility of E. coli and Burkholderia 11BH497 to adulthood in A. tomentosus by 2 days (Burkholderia- to immune defences induced by E. coli-derived LPS and positive median lifespan = 22 days; Burkholderia-nega- heat-killed Burkholderia was compared to ensure that the tive median lifespan = 20 days). Although we did previous results were not specific to E. coli-derived LPS. screen every insect for the presence or absence of Burk- Eighteen Alydus spp. were divided into two treatments: holderia, we did not screen for other bacteria in the (i) injected with LPS as done above; and (ii) injected crypts, and Burkholderia may not be the only bacteria with heat-killed Burkholderia 11BH497. The heat-killed responsible for this effect. Burkholderia solution was made by growing a culture of Burkholderia 11BH497 in LB to OD = 1(~108 bacterial 600 Bacteria in insect crypts and abdomens cells per ml), resuspending the culture in 50 lL of PBS and autoclaving for 15 min. Injected insects were incu- The majority of insect-associated bacterial sequences, bated for 24 h, and extracted hemolymph was plated on both from crypts and whole abdomens, were bacteria in E. coli and Burkholderia 11BH497 ZOI plates as above. the Burkholderiaceae. A total of 183 Burkholderiaceae

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1339

(A) (B) Burkholderia-positive 100 Burkholderia-positive Burkholderia-negative Burkholderia-negative 40 80

30 60

20 40 Days to molt Percent survival 10 20

0 0 First Second Third Fourth 01020304050 Instar Days post-hatching

Fig. 1 The effect of Burkholderia acquisition on Alydus tomentosus. (A) Broad-headed bugs with and without Burkholderia have similar nymphal development times but (B) Error bars represent standard error. cloned sequences from 34 broad-headed bugs clustered c-subdivisions of the proteobacteria (Table 3). Wolbachia into 18 OTUs (OTUs 1–13, 17–19, 29, 30; Figs 2 and 3, spp., common symbionts, and a Rhizobium Table 2; Table S1, Supporting information). Thirteen of sp. accounted for nearly all of the a-proteobacteria. The the OTUs (1–13), which accounted for 90.8% of the Wolbachia sequences were 99% similar to a number of sequences we recovered, grouped together in clades bacteria isolated from ant-lions. The Rhizobium largely composed of Burkholderia spp. from Coreoidea sequences were most closely related to a bacterium and Lygaeoidea stinkbugs, which we refer to as the from the root nodule of a woody shrub (99% similarity), ‘insect symbiont clades’ (Figs 2 and 3). Two OTUs (29, and most of the top BLAST hits (99–98% similarity) 30) closely related to Cupriavidus bacteria, another genus were bacteria isolated from plant or soil habitats. All of in the Burkholderiaceae family, accounted for 5.4% of the c-proteobacteria belonged to the Chromatiaceae,a recovered sequences. The remaining three OTUs family of anoxygenic photolithoautotrophic and grouped with other free-living or opportunistic, patho- sulphur-metabolizing bacteria. genic Burkholderia species and together accounted for 3.8% of the recovered sequences. More than half of the Distribution of Burkholderiaceae among insect hosts broad-headed bugs (19/34) were coinfected with two or and across locations more Burkholderiaceae OTUs (Table 2). Most of the coin- fections (15/22) were composed of the most prevalent Three Burkholderiaceae OTUs (1, 2, 3) were shared broad-headed bug OTU (OTU 2) and other less com- among all the host insect species in this study (Alydus mon OTUs. We used a MLST assay to characterize a spp. and M. quinquespinosus; Fig. 3). All three of these subset of the bacteria for which we sequenced the 16S OTUs fell within the ‘insect symbiont clades’, groups of rRNA gene to ensure that the phylogenetic placement bacteria that contain most true bug Burkholderia symbio- of OTU representatives was not specific to the 16S nts (Fig. 3). In addition, there were five OTUs that were rRNA gene. Overall, the MLST OTU representatives shared between at least two of the host species from grouped into the same clades as in the 16S rRNA gene this study. Ten OTUs were composed of sequences phylogeny (Fig. S3, Supporting information). from only one insect species. Cultivation of bacteria from crypts resulted in bacte- We used two approaches to detect differences ria similar to the cloned sequences. Twenty-five Burkhol- between the symbionts distributed across the four sym- deriaceae bacterial sequences were recovered from seven patric host species and across meadows where they individual insects (Table 2). These sequences fell into were collected. First, we tested for differences in com- four major groups: the insect symbiont clade (with OTU munity structure of Burkholderiaceae OTUs by running 2), the Burkholderia cepacia complex (isolate BHJ35g), the PERMANOVA on a sample X OTU presence/absence plant-associated beneficial and environmental Burkholde- matrix. In this analysis, the final model retained only ria group (isolate BHJ34h) and Cupriavidus spp. (isolate location as a significant variable (PERMANOVA, F = 1.78, BHJ32i) (Fig. 2). d.f. = 8, P = 0.009). Host species was not significant (PER- Eighty-five non-Burkholderiaceae bacterial sequences MANOVA, F = 1.59, d.f. = 3, P = 0.078). Second, we used were recovered from the crypts and abdomens of AMOVAs, which test for differences in Burkholderiaceae the broad-headed bugs. The most abundant genetic diversity as opposed to Burkholderiaceae OTU non-Burkholderiaceae bacteria were species in the a- and presence/absence, to test for differences based on host

© 2014 John Wiley & Sons Ltd 1340 J. R. GARCIA ET AL.

diversity in comparison with other American and Asian 70 symbiont of Blissus insularis (OTU16) 50 symbiont of Thasus neocalifornicus (OTU15) Lygaeoidea and Coreoidae stinkbugs (Kikuchi et al. OTU12 symbiont of Riptortus pedestris Candidatus Burkholderia calva 2005; Olivier-Espejel et al. 2011; Boucias et al. 2012). AM- symbiont of Leptocorisa chinensis OTU1 OVA testing for differences in Burkholderiaceae genetic OTU4 symbiont of Homoeocerus unipunctatus symbiont of Togo hemipterus diversity across the species from this study and the pre- viously investigated insects indicated a significant dif- 59 ference in Burkholderiaceae across host species (AMOVA, symbiont of Daclera levana Insect = = < 65 F 17.85, d.f. 7, P 0.0001) and across continents Symbiont AMOVA F = = P = Clades (America vs. Asia; , 14.11, d.f. 1, 0.049).

55

symbiont of Thasus neocalifornicus Presence of Burkholderiaceae in root nodules and soil OTU3 OTU8 OTU6 OTU9 The eight Burkholderiaceae sequences isolated from the OTU2 OTU2 root nodules of three Lespedeza plants clustered into five 87 OTU5 OTU7 OTUs (7, 17, 20, 21, 23; Figs 2 and 3). One sequence 57 symbiont of symbiont of Molipteryx fuliginosa symbiont of Acanthocoris sordidus (OTU 7) fell within the insect symbiont clades, and the symbiont of Plinachtus bicoloripes Burkholderia glathei remaining four grouped with free-living and epiphytic 92 isolate BHJ35g Burkholderia glumae 56 Burkholderia ubonensis Burkholderia species. The seven Burkholderiaceae Burkholderia 58 Burkholderia vietnamensis sequences isolated from soil clustered into six OTUs (1, Burkholderia pyrrocinia cepacia OTU17 – Burkholderia ambifaria complex 17, 20 22, 24; Fig. 3). One soil sequence (OTU 1) fell 66 Burkholderia cenocepacia Burkholderia cepacia within the insect symbiont clades, and the remaining 93 Burkholderia pseudomallei Burkholderia mallei 71 Burkholderia thailandensis soil sequences grouped with Lespedeza root nodule 93 symbiont of Blissus insularis (OTU26) symbiont of Blissus insularis (OTU25) sequences or with other Burkholderia (Figs 2 and 3). 50 Burkholderia mimosarum Plant-associated Burkholderia nodosa isolate BHJ34h Beneficial & Two OTUs were shared between plants and broad- 100 Burkholderia caledonica OTU21 Environmental headed bugs (OTUs 7 and 17) and between soil and 79 Burkholderia xenovorans Burkholderia fungorum Group (PBE) Burkholderia phytofirmans broad-headed bugs (OTUs 1 and 17). Investigation of Burkholderia terrae OTU20 these sequences determined that while highly similar, 62 64 Burkholderia caribensis 91 Cupriavidus pauculus 100 Cupriavidus taiwanensis A. calcaratus the plant and bug sequences, as well as the soil and 94 OTU29 A. conspersus 92 isolate BHJ32i A. tomentosus bug sequences, were not identical. Other non-Burkholde- Ralstonia solanacearum Pandoraea norimbergensis M. quinquespinosus ria bacteria were recovered from soil and root nodules, L. cuneata 0.2 soil including Rhizobium spp., the typical root nodule symbi- ont for Lespedeza spp., but those data are not shown Fig. 2 Maximum-likelihood phylogenetic analysis of the Burk- holderiaceae 16S rRNA gene sequences, including all nonsingle- here. ton OTUs in this study (in bold). Bold sequences labelled as ‘isolate’ are bacteria cultivated from crypts. Circle, plant and Susceptibility of bacteria to hemolymph defences soil icons placed after each OTU number indicate where each OTU was recovered. Clades without OTU representatives from In a test of bacterial susceptibility to induced and con- this study were collapsed. The tree was rooted with Escherichia stitutive immune defences in A. conspersus hemolymph, coli. Numbers at nodes are ML bootstrap values (500 repli- E. coli, but not Burkholderia sp., was susceptible and cates). Scale bar at bottom represents 0.2 nucleotide changes per site. A phylogeny with accession nos for each sequence only under the induced treatment (Fig. 4A). E. coli was and without collapsed clades is available in the supplementary not inhibited by noninduced host hemolymph, and information (Fig. S2, Supporting information). Burkholderia was not inhibited by either hemolymph treatment (Fig. 4A). To ensure this result was not due species and location. There was a significant difference to the activation of a limited repertoire of immune in Burkholderiaceae genetic diversity across host species defences by E. coli-derived LPS, the susceptibility of (AMOVA, F = 25.66, d.f. = 3, P = 0.008), but no significant both bacteria was also tested with hemolymph from difference across locations (AMOVA, F = À1.19, d.f. = 8, Alydus spp. that was stimulated with heat-killed Burk- P = 0.747). The significant difference across host species holderia. E. coli was similarly susceptible to hemolymph was largely driven by the pairwise difference between stimulated by heat-killed Burkholderia and E. coli- Alydus calcaratus and A. conspersus (AMOVA, F = 53.79, derived LPS. Burkholderia was not susceptible to hemol- d.f. = 1, P = 0.007). ymph stimulated by either immune elicitor (Fig. 4B). The four broad-headed bug species and Blissus The susceptibility of five additional bacteria (Serratia insularis, a species that has an overlapping range with marcescens, Lactococcus lactis,aBurkholderia isolate from the broad-headed bug species, have more symbiont a root nodule, a Burkholderia isolate from soil (OTU 21)

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1341

(A)North America Asia (B) Acal Acon Atom Mquin Binsul Tneo Lchin Rped others Lesp soil n = 63 n = 62 n = 46 n = 12 n = 243 n = 9 n = 11 n = 11 n = 19 n = 8 n = 7 1 1 2 2 3 3 4 4 5 5 6 6 spp. 7 7 Burkholderia 8 8 insect symbiont 9 9 10 10 clades 11 11 12 12 13 13 14 14 15 15 16 16 17 17 18 18 19 19 20 20 Bcc & 21 21 PBE Group 22 22 23 23 24 24 25 25 26 26 27 27 28 28 spp. 29 29 Cupriavidus 30 30

0.2 0.4 0.6 0.8 1

Fig. 3 Heatmap of Burkholderiaceae sequence abundance. (A) Sequences from cloned amplicons for the four sympatric broad-headed bug species from this study and 23 other insect species characterized in four previous studies. (B) Sequences from direct sequencing of amplicons from bacterial cultures for the host plant L. cuneata and soil. Each numbered row represents an OTU defined at 97% similarity. The most abundant OTU is relativized to equal an abundance of 1. The abundances of the remaining OTUs are expressed as a percentage of the most abundant OTU. All species and sample types to the left of the dashed line are from this study; species to the right are from other studies as cited below. The black lines above the sample names indicate the location. The number of sequences is indicated under each sample name. Bcc is an abbreviation for Burkholderia cepacia complex, and PBE is an abbreviation for plant-associated beneficial and environmental Burkholderia group as previously outlined (Suarez-Moreno et al. 2011). Acal, Alydus calcaratus; Acon, Alydus conspersus; Atom, Alydus tomentosus; Mquin, Megalotomus quinquespinosus; Binsul, Blissus insularis (Boucias et al. 2012); Tneo, Thasus neocalifornicus (Olivier-Espejel et al. 2011); Lchin, Leptocorisa chinensis (Kikuchi et al. 2005); Rped, Riptortus pedestris (Kikuchi et al. 2005); others, pooled sequences of 19 Asian pentatomomorphan symbionts from a broad survey of symbiont diversity (Kikuchi et al. 2011b); Lesp, Lespedeza cuneata (bush clover); and soil, soil collected near L. cuneata plants. Sample informa- tion for sequences obtained from other studies is available in Table S1 (Supporting information). and a Cupriavidus isolate from crypts) was tested using were no differences in development time between Burk- LPS-stimulated hemolymph. Of these five bacteria, only holderia-positive and Burkholderia-negative A. tomentosus, the Burkholderia sp. isolated from a Lespedeza root nod- but acquiring Burkholderia increased survival. Although ule (OTU 7) was susceptible. This species was less sus- it was not possible to attribute the survival benefits ceptible than E. coli as evidenced by smaller zones of solely to Burkholderia, it is consistent with the evidence inhibition (Fig. 5). The remaining four bacteria, were presented here and in the previous studies. Differences not inhibited by LPS-stimulated hemolymph. in the measured benefits across these studies may be methodological, or they may be due to differences in the benefits that are conferred by different Burkholderia Discussion species or to intrinsic differences between host species. Burkholderia spp. are the most frequent bacteria isolated While fitness differences between Burkholderia-positive from the midgut crypts of the four sympatric broad- and Burkholderia-negative broad-headed bugs in these headed bug species, similar to previously characterized studies presumably are due to Burkholderia, the acquisi- bacterial communities in other hemipterans (Kikuchi tion or maintenance of other microbes may have been et al. 2005, 2011b; Olivier-Espejel et al. 2011; Boucias impacted by the experimental conditions, and future et al. 2012). Associations with gut symbionts can benefi- studies should assess the effects of other gut microbes cially impact hemipteran development, growth and as well. survival (Fukatsu & Hosokawa 2002; Douglas 2009; Ecologically similar, sympatric broad-headed bug Prado & Almeida 2009). Specifically, establishment of species presumably are exposed to the same environ- Burkholderia in midgut crypts decreases development mental microbes, but many factors could lead to host– time and increases adult size and survival in R. pedestris symbiont partner specificity nonetheless. Specificity can (Kikuchi et al. 2007, 2012). In Thasus neocalifornicus, be considered at a number of different scales. For exam- Burkholderia acquisition increases nymphal survival ple, we can ask whether broad-headed bugs as a whole (Olivier-Espejel et al. 2011), and in Blissus insularis, anti- preferentially associate with some families of bacteria biotic clearance of Burkholderia retards development and over others. Then, at a finer scale, we can ask whether increases mortality (Boucias et al. 2012). Here, there broad-headed bugs as a whole preferentially associate

© 2014 John Wiley & Sons Ltd 1342 J. R. GARCIA ET AL.

Table 2 Summary of Burkholderiaceae sequences found in insect, soil and plant samples

Species Sample Location Method Sample type No. seqs OTUs

Alydus calcaratus 11BH CM Culturing Abdomen 1 n/a BHB6 SM Cloning Crypts 1 2 BHJ35 LIME Culturing Crypts 1 n/a BHJ36 SM Cloning Crypts 4 2, 5 C7 CM Cloning Abdomen 8 2, 3, 6 C8 CM Cloning Abdomen 3 2, 10 FMAC3 FM Cloning Crypts 8 1, 7, 12 FMAC4 FM Cloning Crypts 9 2, 9 FMAC6 FM Cloning Crypts 10 1, 4 FMAC7 FM Cloning Crypts 10 1, 7, 13 Joy1 JOY Cloning Crypts 6 2, 5 Lime2 LIME Cloning Crypts 4 2 Alydus conspersus 3BH SM Culturing Abdomen 2 n/a BHJ27 MNP Cloning Crypts 7 2, 17 BHJ28 MNP Cloning Crypts 6 2, 17, 29 BHJ29 MNP Cloning Crypts 7 2, 3, 29 BHJ30 MNP Cloning Crypts 6 2, 29, 30 FMAC8 FM Cloning Crypts 1 18 SM13 SM Cloning Abdomen 3 1, 4 SM2 SM Cloning Crypts 7 1, 2 SM8 SM Cloning Crypts 7 1 W3 WY Cloning Crypts 1 2 W4 WY Cloning Crypts 6 2, 11 W5 WY Cloning Crypts 11 2 Alydus tomentosus 2BH SM Culturing Abdomen 3 n/a BHB5 SM Cloning Abdomen 3 2 BHB7 SM Cloning Abdomen 1 2 BHB11 SM Cloning Abdomen 2 2 BHJ32 MNP Culturing Crypts 5 n/a BHJ38 SM Cloning Crypts 8 2 BHJ39 SM Cloning Abdomen 10 2, 5, 8, 9 ELF1 LC Cloning Crypts 1 2 Joy2 JOY Cloning Crypts 1 1 Joy3 JOY Cloning Crypts 9 2, 3 Joy4 JOY Cloning Crypts 4 2 Lime1 LIME Cloning Crypts 5 2, 6, 19 Lime4 LIME Cloning Crypts 2 2, 8 Megalotomus quinquespinosus BHJ23 LAB Cloning Crypts 2 1, 8 BHJ31 HM Culturing Crypts 11 n/a BHJ34 CFH Culturing Crypts 2 n/a W2 WY Cloning Crypts 10 2, 3 Lespedeza cuneata Les1 SM Culturing Nodule 1 7 Les2 SM Culturing Nodule1 1 20 Culturing Nodule2 2 20 Culturing Nodule3 3 17, 21, 23 Les4 SM Culturing Nodule 1 21 Soil SML SM Culturing n/a 3 29, 20 SMT SM Culturing n/a 1 22 Snap SNAP Culturing n/a 3 1, 21, 24 with some bacterial species over others. Finally, we can Burkholderia associations are more frequently established ask whether broad-headed bug species differ in their over associations with environmental bacteria of other associations with bacterial species. Consistent with genera, of which there can be many [e.g. more than a previous studies, a majority of bacteria sequenced from 1 000 000 bacterial species in a gram of soil (Gans & the broad-headed bug crypts were Burkholderia (77.2% Wolinsky 2005)]. We did, however, uncover some Burk- of sequences), suggesting some mechanism by which holderiaceae bacteria outside the genus Burkholderia (see

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1343

Table 3 Percentages of sequences recovered from insect abdomens and crypts that were various non-Burkholderiaceae

Insect species Tissue type

Bacteria species Acal Acon Atom Mquin Abdomen Crypts

† ‡ a-Proteobacteria, Bartonella sp. 2.2 1.0 a-Proteobacteria, Brucella sp. 2.2 1.0 a-Proteobacteria, Ochrobactrum sp. 2.2 1.0 a-Proteobacteria, Rhizobium sp. 2.8 13.3 3.1 a-Proteobacteria, Wolbachia spp. 14.8 1.4 3.3 16.0 0.5 a-Proteobacteria, Sphingomonas sp. 2.5 2.5 b-Proteobacteria, Bergeriella sp. 1.1 1.4 2.5 c-Proteobacteria, Allochromatium sp. 1.1 2.8 3.7 c-proteobacteria, Thiocapsa sp. 4.3 1.4 6.2 c-Proteobacteria, Thiococcus sp. 2.2 2.5 c-Proteobacteria, Thiodictyon sp. 5.4 1.4 7.4 c-Proteobacteria, Thiorhodococcus sp. 5.4 6.2 Firmicutes, Enterococcus sp. 1.1 2.8 3.7 Firmicutes, Lactococcus sp. 11.1 7.4 1.0

Acal, Alydus calcaratus; Acon, Alydus conspersus; Atom, Alydus tomentosus; Mquin, Megalotomus quinquespinosus. † Identifications were made according to the NCBI using Greengenes Classifier. ‡ Indicates the percentage of sequences including Burkholderiaceae and non-Burkholderiaceae sequences. We do not include any non- Burkholderiaceae sequences that were recovered only once.

below) and non-Burkholderiaceae as well. A majority of which is nearly absent from all other previously charac- the non-Burkholderiaceae sequences were isolated from terized insects that host Burkholderia in crypts. We also whole abdomens rather than from the symbiont-hous- isolated Cupriavidus, another genus within the Burkhol- ing crypts in isolation, making it likely that these bacte- deriaceae, (Figs 2 and 3) from A. conspersus and A. tomen- ria were in other abdominal tissues or, despite careful tosus. These bacteria have not previously been reported sterilization, present on the exterior of the insects. A as insect symbionts and are typically associated with few of the non-Burkholderiaceae sequences, however, plants, especially C. taiwanensis, a widespread and dom- were isolated from crypts. Some of these bacteria (e.g. inant symbiont of Mimosa spp. root nodules (Chen et al. Wolbachia sp. and Thiodictyon sp.) were recovered from 2003). While we found a significant difference in multiple individuals and species indicating that these Burkholderiaceae genetic diversity due to host species bacteria may be stable and recurring constituents of the when we considered both American and Asian insect insects’ microbiomes, although perhaps they occur on species, when we focused on only the four sympatric the outside of the crypts rather than within. Coinfec- hosts, we found no significant impact of host species on tions with these other bacteria, although rare, could Burkholderiaceae presence and absence, although there result in bacterial interactions that alter host or symbi- was evidence of significant differences in the genetic ont fitness (Goto et al. 2006; Jaenike et al. 2009). For diversity of Burkholderia across these species, which was example, Wolbachia growth is suppressed in Drosophila possibly driven by the isolation of Cupriavidus from melanogaster in coinfections with Spiroplasma compared only A. conspersus. These results suggest that these to Wolbachia-only infections (Goto et al. 2006). sympatric species encounter and then sequester similar Focusing more closely on specificity in relation symbionts. Significant differences between the symbio- to associations with Burkholderiaceae, a majority of nts present in insects from different meadows does microbes isolated from Alydus spp. and Megalotomus suggest the potential for different host–symbiont quinquespinosus were Burkholderia spp. similar to those dynamics across natural habitats, although this may be isolated from other insect species (i.e. Burkholderia from weak as there was no difference in genetic diversity the ‘insect symbiont clades’; Figs 2 and 3). One across habitats. Burkholderia OTU (OTU 1), abundant in a number of Many insects were coinfected with two or more Burk- previously investigated insect species (Kikuchi et al. holderia OTUs (Table 2). Coinfection is present in Riptor- 2007, 2011b; Olivier-Espejel et al. 2011; Boucias et al. tus pedestris and Leptocorisa chinenesis (Kikuchi et al. 2012), was recovered from our focal species, although at 2005), but at a much lower frequency than found in a low prevalence compared to Burkholderia OTU 2, Alydus spp. and M. quinquespinosus. It is unclear

© 2014 John Wiley & Sons Ltd 1344 J. R. GARCIA ET AL.

(A) LPS no stab control 0.1 0.2 0.3 0.4 Radius of inhibition zone (cm) 0 0 0 0 0

0.0 crypts cryptsnodule soil crypts 0 0 0 sp. E. coli sp. sp. sp.

0.0 0.1 0.2 0.3 0.4 0.5 0.6 L. lactis E.coli Burkholderia S. marcescens 11BH497 BurkholderiaBurkholderia Burkholderia Cupriavidus

(B) Fig. 5 Bacteria differ in their susceptibility to LPS-stimulated LPS Alydus spp. hemolymph defences, as measured by zone of heat-killed inhibition assays. Bacteria species are as follows (see Table 2; Burkholderia Table S1, Supporting information for details of isolation): E. coli MG1655, Serratia marcescens RHoD, Lactococcus lactis BHJ32a, Burkholderia sp. 11BH497, Burkholderia sp. Les1n2i, Burkholderia sp. Snap2c and Cupriavidus sp. BHJ32i. Error bars represent standard error.

studied species. We isolated Burkholderia species from Alydus spp. and M. quinquespinosus individuals that fell Radius of inhibition zone (cm)Radius of inhibition zone (cm) Radius of inhibition zone 0.1 0.2 0.3 0.4 0.5 0.6 Burkholderia 0 0 outside the clade dominated by insect sym-

0.0 bionts (Figs 2 and 3). The southern chinch bug also E.coli Burkholderia 11BH497 occasionally harbours these Burkholderia (Boucias et al. 2012), but neither the giant mesquite bug (Olivier-Espe- Fig. 4 Susceptibility of E. coli MG1655 and Burkholderia sp. jel et al. 2011) nor 19 Asian insect species do (Kikuchi 11BH497 to Alydus spp. hemolymph defences measured by et al. 2011b). While this could be an artefact of sampling zone of inhibition assays. (A) Bacterial susceptibility to hemol- ymph defences of insects injected with LPS or insects handled depth in some cases, extensive sampling in two Asian but not injected. (B) Bacterial susceptibility to hemolymph species, Riptortus pedestris and Leptocorisa chinensis, did defences of insects injected with LPS or heat-killed Burkholderia not uncover any symbiont associations outside the 11BH4974. Error bars represent standard error. ‘insect symbiont clades’ (Kikuchi et al. 2005). Therefore, greater symbiont diversity associated with some insect whether mixed Burkholderia symbiont infections are ben- species compared to others could reflect ecological dif- eficial to the insect host. In most of the coinfections, the ferences. For example, R. pedestris were collected from common Burkholderia sp. (OTU 2) was present, although soya bean fields (Kikuchi et al. 2007, 2011b), where agri- it is not clear whether this species was at low or high cultural practices in a monoculture setting may limit abundance relative to its coinfecting species. Coinfecting the diversity of bacteria present in the environment or species could represent cheaters (Sachs et al. 2010) or where there may be selection for a narrower range of could represent symbionts that provide different bene- symbiotic partners [e.g. those that provide insecticide fits. However, coinfection could also lead to within-host resistance (Kikuchi et al. 2012)]. Within more diverse competition that could have negative fitness effects for natural settings of mixed plant species, variation in coinfected hosts (Sakurai et al. 2005; Oliver et al. 2006). symbionts, which could have alternative metabolic Most guts are coinfected with many different capabilities including detoxification of specific plant bacterial species, and the interactions between these compounds or protection against more diverse patho- species likely have important implications for both the gens, could select for the maintenance of symbiont spe- host and the bacteria. Broad-headed bug–Burkholderia cies outside the insect symbiont clade. Further research associations provide insect systems in which to test on the maintenance of Burkholderia genetic and pheno- these hypotheses. typic diversity in alternative ecosystems is needed. Overall, symbiont diversity in broad-headed bug Despite the fact that we isolated diverse Burkholderia- populations is greater than that in some previously ceae and other bacteria from our focal species, several

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1345 lines of evidence suggest that these broad-headed bugs, bacteria play a role in symbiont establishment and like other true bugs, preferentially associate with a sub- maintenance. To understand how bacterial growth set of Burkholderiacae. First, Burkholderia species isolated could be suppressed elsewhere in symbiotic hosts, from Alydus spp. and M. quinquespinosus were concen- using traditional zone of inhibition assays, we found trated in the ‘insect symbiont clades’ instead of being that bacteria exhibit different susceptibilities to induced evenly dispersed across the Burkholderia phylogeny. Sec- defences of the host hemolymph (Figs 4 and 5). While ond, while a few Burkholderia OTUs were isolated from E. coli was susceptible to inhibition, bacteria from four both broad-headed bugs and soil or plant tissues, many other genera were not. This is in contrast to most Burkholderia isolated from the broad-headed bugs’ envi- insects, where antibacterial activity after immune stimu- ronment were not found within broad-headed bugs. lation is strong, long lasting and general to broad cate- These results, consistent with previous studies of some gories of microbes [e.g. Gram-positive or Gram-negative other insect–Burkholderia associations (Kikuchi et al. bacteria; (Lemaitre & Hoffmann 2007; Haine et al. 2007), suggest a process in which the insects associate 2008)]. The limited antimicrobial activity seen in assays with only a subset of available Burkholderia. with broad-headed bugs suggests that they may have We examined several factors that could facilitate such narrower antimicrobial responses than other insects. specificity. Maintenance of Burkholderia within other While zone of inhibition assays indicated that a broad habitats or hosts, for example, could function as an eco- range of bacteria are resistant to broad-headed bug host logical mechanism leading to more frequent insect immune factors, some Burkholderia species were resis- encounters with Burkholderia. We isolated the same tant, while others were not. Specifically, a Burkholderia Burkholderia OTU from both the broad-headed bug Aly- species isolated from Lespedeza and genetically similar dus calcaratus and from rhizobium-harbouring root nod- to Burkholderia from insect guts was inhibited by host ules of its leguminous food plant, Lespedeza cuneata.Itis hemolymph, while other Burkholderia were not. These unclear whether this Burkholderia species is a beneficial differences are likely due to differential susceptibility to root nodule symbiont (i.e. a bacterium that provides host antimicrobial peptides, but could also be due to fixed nitrogen to the plant), but Lespedeza does have an differential susceptibility to reactive oxygen species, intimate relationship with this bacterium; multiple Burk- phenoloxidase or cellular defences. Further work will holderia species have previously been recovered from be needed to understand whether susceptibility to Lespedeza spp. nodules (Compant et al. 2008; Palaniap- hemolymph-based defences in zone of inhibition assays pan et al. 2010). Although Palaniappan et al. (2010) is correlated with the ability to colonize crypts or main- found that Burkholderia could not induce nodule forma- tain long-term associations with insects. Given that the tion in the absence of nodule-competent bacteria such four sympatric host species studied here harbour simi- as Rhizobium, Burkholderia did have plant growth- lar symbionts, the mechanisms underlying specific asso- promoting properties. While it is unlikely that broad- ciations with particular Burkholderia are likely universal. headed bugs directly acquire bacteria from nodules, Our goal was to sample symbionts from multiple plant associations with Burkholderia could maintain a individuals of sympatric insect species to determine higher concentration of Burkholderia in the soil sur- whether fine-scale sampling could reveal partner speci- rounding host plants. To determine whether these hosts ficity. While we hypothesized that ecological or physio- are sharing a Burkholderia symbiont pool, however, fur- logical differences between sympatric hosts could lead ther investigation is needed to determine whether Burk- to differences in their symbiont populations, we found holderia species isolated from plants can establish and no evidence for differences in their symbiont popula- be maintained in insects and, similarly, whether plants tions based on the presence and absence of particular can uptake Burkholderia isolated from insects. OTUs and only limited evidence of differences in their Many host and symbiont traits could also facilitate symbiont populations based on genetic diversity. Lack partner specificity. Recent research, for example, of partner specificity across these hosts could be driven suggests that in another true bug, R. pedestris, by similarities in behaviour, in host plant use or in life antimicrobial activity within a region of the gut adjacent cycle or by shared ancestry. Regardless of the reason, to the crypts, but not within the crypts themselves, sup- partner-switching likely prevents tight co-evolutionary presses Burkholderia (Kim et al. 2013a). Furthermore, dynamics between hosts and symbionts in this system. Burkholderia with loss of function mutations in the uppP gene, which is involved in biosynthesis of cell wall Acknowledgements components that trigger host immune responses (Loutet & Valvano 2011), are able to colonize but not establish We are grateful to N. Lowe, E. Akhirome, N. Couret, L. Griffin, in the crypts of R. pedestris (Kim et al. 2013b). These M. Garcia, B. Hall, V. Harris, E. Quaki, R. Lowe, T. Acevedo, findings suggest that host immune responses towards G. Fricker and A. Chang for help with collections and

© 2014 John Wiley & Sons Ltd 1346 J. R. GARCIA ET AL. experiments. We thank D. Boucias for the MLST primers and coexisting in the same insect host. Applied and Environmental advice on their use. C. T. was supported by Emory’s SURE Microbiology, 72, 4805–4810. program. J. G. and B. P. were funded by National Science Haine ER, Pollitt LC, Moret Y, Siva-Jothy MT, Rolff J (2008) Foundation Graduate Research Fellowships. Research was sup- Temporal patterns in immune responses to a range of micro- ported by National Science Foundation grant IOS-1149829 to bial insults (Tenebrio molitor). Journal of Insect Physiology, 54, N.M.G. 1090–1097. Huelsenbeck JP, Ronquist F (2001) MRBAYES: Bayesian infer- ence of phylogenetic trees. Bioinformatics, 17, 754–755. References Jaenike J, Stahlhut JK, Boelio LM, Unckless RL (2009) Associa- tion between Wolbachia and Spiroplasma within Drosophila Abramoff MD, Magalhaes~ PJ (2004) Image processing with neotestacea: an emerging symbiotic mutualism? Molecular ImageJ. Biophotonics International, 11,36–42. Ecology, 19, 414–425. Baldwin A, Mahenthiralingam E, Thickett KM et al. (2005) Mul- Kikuchi Y, Meng X-Y, Fukatsu T (2005) Gut symbiotic bacteria tilocus sequence typing scheme that provides both species of the genus Burkholderia in the broad-headed bugs Riptortus and strain differentiation for the Burkholderia cepacia complex. clavatus and Leptocorisa chinensis (Heteroptera: ). Journal of Clinical Microbiology, 43, 4665–4673. Applied and Environmental Microbiology, 71, 4035–4043. Baumann P, Baumann L, Lai C-Y, Rouhbakhsh D (1995) Genet- Kikuchi Y, Hosokawa T, Fukatsu T (2007) Insect-microbe ics, physiology, and evolutionary relationships of the genus mutualism without vertical transmission: a stinkbug acquires Buchnera: intracellular symbionts of aphids. Annual Review of a beneficial gut symbiont from the environment every gener- Microbiology, 49,55–94. ation. Applied and Environmental Microbiology, 73, 4308–4316. Boucias DG, Garcia-Maruniak A, Cherry R et al. (2012) Kikuchi Y, Hosokawa T, Fukatsu T (2011a) Specific develop- Detection and characterization of bacterial symbionts in the mental window for establishment of an insect-microbe gut Heteropteran, Blissus insularis. FEMS Microbiology Ecology, 82, symbiosis. Applied and Environmental Microbiology, 77, 629–641. 4075–4081. Chen WM, Moulin L, Bontemps C et al. (2003) Legume symbi- Kikuchi Y, Hosokawa T, Fukatsu T (2011b) An ancient but pro- otic nitrogen fixation by Beta-proteobacteria is widespread in miscuous host-symbiont association between Burkholderia gut nature. Journal of Bacteriology, 185, 7266–7272. symbionts and their heteropteran hosts. The ISME Journal, 5, Compant S, Nowak J, Coenye T, Clement C, Barka EA (2008) 446–460. Diversity and occurrence of Burkholderia spp. in the natural Kikuchi Y, Hayatsu M, Hosokawa T et al. (2012) Symbiont- environment. FEMS Microbiology Reviews, 32, 607–626. mediated insecticide resistance. Proceedings of the National Darriba D, Taboada GL, Doallo R, Posada D (2012) jModelTest Academy of Sciences, 109, 8618–8622. 2: more models, new heuristics and parallel computing. Kim JK, Kim NH, Jang HA et al. (2013a) A specified midgut Nature Methods, 9, 772–772. region controlling the symbiont titer in an insect-microbe gut Davidson SK, Koropatnick TA, Kossmehl R, Sycuro L, McFall- symbiotic association. Applied and Environmental Microbiology Ngai MJ (2004) NO means “yes” in the squid-vibrio symbio- 79, 7229–7233. sis: nitric oxide (NO) during the initial stages of a beneficial Kim JK, Lee HJ, Kikuchi Y et al. (2013b) The bacterial cell wall association. Cellular Microbiology, 6, 1139–1151. synthesis gene uppP is required for Burkholderia colonization DeSantis TZ, Hugenholtz P, Keller K et al. (2006a) NAST: a of the stinkbug gut. Applied and Environmental Microbiology, multiple sequence alignment server for comparative 79, 4879–4886. analysis of 16S rRNA genes. Nucleic Acids Research, 34, Lane DJ (1991) 16S/23S rRNA sequencing. In: Nucleic Acid W394–W399. Techniques in Bacterial Systematics (eds Stackebrandt E, Good- DeSantis TZ, Hugenholtz P, Larsen N et al. (2006b) Greeng- fellow M), pp.115–175. Wiley, New York. enes, a chimera-checked 16S rRNA gene database and work- Lee K-H, Ruby EG (1992) Detection of the light organ symbi- bench compatible with ARB. Applied and Environmental ont, Vibrio fischeri, in Hawaiian seawater by using lux gene Microbiology, 72, 5069–5072. probes. Applied and Environmental Microbiology, 58, 942–947. Douglas AE (2009) The microbial dimension in insect nutri- Lemaitre B, Hoffmann J (2007) The host defense of Drosophila tional ecology. Functional Ecology, 23,38–47. melanogaster. Annual Review of Immunology, 25, 697–743. Finney JC, Pettay DT, Sampayo EM et al. (2010) The relative Loutet SA, Valvano MA (2011) Extreme antimicrobial peptide significance of host–habitat, depth, and geography on the and polymyxin B resistance in the genus Burkholderia. Fron- ecology, endemism, and speciation of coral endos- tiers in Microbiology, 2,1–8. ymbionts in the genus Symbiodinium. Microbial Ecology, 60, McFall-Ngai M, Hadfield MG, Bosch TCG et al. (2013) Animals 250–263. in a bacterial world, a new imperative for the life sciences. Fukatsu T, Hosokawa T (2002) Capsule-transmitted gut symbi- Proceedings of the National Academy of Sciences of the United otic bacterium of the Japanese common plataspid stinkbug, States of America, 110, 3229–3236. Megacopta punctatissima. Applied and Environmental Microbiol- Mitsuhashi J (2002) Composition of salt solutions and culture ogy, 68, 389–396. media. In: Invertebrate Tissue Culture Methods (eds Mitsuhashi J), Gans J, Wolinsky M (2005) Computational improvements pp. 401–420. Springer Lab Manual, Tokyo. reveal great bacterial diversity and high metal toxicity in Moret Y, Schmid-Hempel P (2000) Survival for immunity: the soil. Science, 309, 1387–1390. price of immune system activation for bumblebee workers. Goto S, Anbutsu H, Fukatsu T (2006) Asymmetrical interac- Science, 290, 1166–1168. tions between Wolbachia and Spiroplasma endosymbionts

© 2014 John Wiley & Sons Ltd BROAD-HEADED BUG–BURKHOLDERIA SYMBIOSIS 1347

Nyholm SV, Graf J (2012) Knowing your friends: invertebrate Swofford D (2002) PAUP*. Phylogenetic Analysis Using Parsi- innate immunity fosters beneficial bacterial symbioses. mony (*and Other Methods). Sinauer Associates, Sunderland, Nature Reviews Microbiology, 10, 815–827. MA. Oksanen J, Guillaume Blanchet F, Kindt R et al. (2013) Vegan: Therneau T (2013) A Package for Survival Analysis in S. Mayo Community Ecology Package. Foundation, Rochester, MN. Oliver KM, Moran NA, Hunter MS (2006) Costs and benefits Troll JV, Bent EH, Pacquette N et al. (2010) Taming the of a superinfection of facultative symbionts in aphids. symbiont for coexistence: a host PGRP neutralizes a bac- Proceedings of the Royal Society B: Biological Sciences, 273, terial symbiont toxin. Environmental Microbiology, 12, 1273–1280. 2190–2203. Olivier-Espejel S, Sabree ZL, Noge K, Becerra JX (2011) Gut Ventura MU, Silva JJ, Panizzi AR (2000) Phyophagous Neom- microbiota in nymph and adults of the giant mesquite bug egalotomus parvus (Westwood) (Hemiptera: Alydidae) feeding (Thasus neocalifornicus) (Heteroptera: Coreidae) is dominated on carrion and feces. Annals of the Society of Entomology by Burkholderia acquired de novo every generation. Environ- Brasil, 29, 839–841. mental Entomology, 40, 1102–1110. Zwickl DJ (2006) Genetic algorithm approaches for the phylogenetic Palaniappan P, Chauhan PS, Saravanan VS, Anandham R, Sa T analysis of large biological sequence datasets under the maximum (2010) Isolation and characterization of plant growth promot- likelihood criterion. PhD Dissertation. University of Texas at ing endophytic bacterial isolates from root nodule of Lespe- Austin, Austin, Texas. deza sp. Biology and Fertility of Soils, 46, 807–816. Pinheiro J, Bates D, DebRoy S, Sarvat D, Team RDC (2013) NLME: Linear and Nonlinear Mixed Effects Model. Porter SS, Rice KJ (2012) Trade-offs, spatial heterogeneity, and J.G., A.L., C.T., E.C., Z.M. and S.C. conducted experi- the maintenance of microbial diversity. Evolution, 67, 599– ments. J.G. and B.P. performed analyses. J.G., A.L., E.C., 608. N.M.G. and Z.M. designed experiments. J.G. and Prado SS, Almeida RPP (2009) Role of symbiotic gut bacteria in N.M.G. wrote the manuscript. All authors collected the development of Acrosternum hilare and Murgantia histrio- insects and approved the final manuscript. nica. Entomologia Experimentalis et Applicata, 132,21–29. R Core Team (2009) R: A Language and Environment for Statisti- cal Computing. R Foundation for Statistical Computing, Vienna, Austria. Ruby EG (2008) Symbiotic conversations are revealed under Data accessibility genetic interrogation. Nature Reviews Microbiology, 6, 752–762. All 16S rRNA gene sequences are deposited in GenBank Sachs JL, Kembel SW, Lau AH, Simms EL (2009) In situ – phylogenetic structure and diversity of wild Bradyrhizobium under accession nos KF931405 KF931627. All other data communities. Applied and Environmental Microbiology, 75, from this manuscript including MLST assay gene 4727–4735. sequences, tree files with alignments and R and mothur Sachs JL, Russell JE, Lii YE et al. (2010) Host control over infec- scripts are accessible through Dryad doi 10.5061/dryad. tion and proliferation of a cheater symbiont. Journal of Evolu- dp7b8. tionary Biology, 23, 1919–1927. Sakurai M, Koga R, Tsuchida T, Meng XY, Fukatsu T (2005) Rickettsia symbiont in the pea aphid Acyrthosiphon pisum: Supporting information novel cellular tropism, effect on host fitness, and interaction with the essential symbiont Buchnera. Applied and Environ- Additional supporting information may be found in the online mental Microbiology, 71, 4069–4075. version of this article. Schaefer CW (1980) The host plants of the , with a Fig. S1. Picture of broad-headed bug midgut with detailed note on heterotypic feeding aggregations (Hemiptera: Co- picture of the symbiont-harbouring crypts. reoidea: Alydidae). Journal of the Kansas Entomological Society, 53, 115–122. Fig. S2. Phylogenetic analysis of the Burkholderiaceae 16S rRNA Schloss PD, Westcott SL, Ryabin T et al. (2009) Introducing mo- gene sequences from all nonsingleton OTUs in this study (in thur: open-source, platform-independent, community-sup- bold). ported software for describing and comparing microbial communities. Applied and Environmental Microbiology, 75, Fig. S3. Phylogenetic analysis of eight cultivated Burkholderia 7537–7541. species using a concatenation of four housekeeping genes: Spilker T, Baldwin A, Bumford A et al. (2009) Expanded mul- atpD, gltB, lepA and recA. tilocus sequence typing for Burkholderia species. Journal of Table S1. Sample information for Burkholderiaceae sequences Clinical Microbiology, 47, 2607–2610. obtained from previous studies. Suarez-Moreno ZR, Caballero-Mellado J, Coutinho BG et al. (2011) Common features of environmental and potentially Table S2. Sequence information for all Burkholderiaceae beneficial plant-associated Burkholderia. Microbial Ecology, 63, sequences recovered from broad-headed bugs, root nodules – 249 266. and soil.

© 2014 John Wiley & Sons Ltd