Skip to main content

A novel bacterial symbiont in the nematode Spirocerca lupi

Abstract

Background

The parasitic nematode Spirocerca lupi (Spirurida: Thelaziidae), the canine esophageal worm, is the causative agent of spirocercosis, a disease causing morbidity and mortality in dogs. Spirocerca lupi has a complex life cycle, involving an obligatory coleopteran intermediate host (vector), an optional paratenic host, and a definitive canid host. The diagnosis of spirocercosis is challenging, especially in the early disease stages, when adult worms and clinical signs are absent. Thus, alternative approaches are needed to promote early diagnosis. The interaction between nematodes and their bacterial symbionts has recently become a focus of novel treatment regimens for other helminthic diseases.

Results

Using 16S rDNA-based molecular methods, here we found a novel bacterial symbiont in S. lupi that is closely related to Comamonas species (Brukholderiales: Comamonadaceae) of the beta-proteobacteria. Its DNA was detected in eggs, larvae and adult stages of S. lupi. Using fluorescent in situ hybridization technique, we localized Comamonas sp. to the gut epithelial cells of the nematode larvae. Specific PCR enabled the detection of this symbiont's DNA in blood obtained from dogs diagnosed with spirocercosis.

Conclusions

The discovery of a new Comamonas sp. in S. lupi increase the complexity of the interactions among the organisms involved in this system, and may open innovative approaches for diagnosis and control of spirocercosis in dogs.

Background

Vector-borne helminthic diseases, such as onchocerciasis and lymphatic filariasis, are major human diseases in endemic areas. Novel treatment approaches have been recently focusing on the interaction between the causative helminth agent and its bacterial symbiont. Consequently, antibiotics, such as doxycycline, are used instead of, or with, anti-helminthic drugs for treatment [1, 2]. However, because of difficulties in application, various bacterial targets are constantly studied [3]. This approach has also been adopted in veterinary helminthic diseases, such as bovine onchocerciasis and canine heartworm disease [4–6].

Spirocercosis is a vector-borne helminthic disease, mostly affecting carnivores, especially canids [7, 8]. It is caused by the esophageal nematode Spirocerca lupi (Spirurida: Thelaziidae) that has a wide distribution, but is mostly prevalent in warm, humid areas. The exact annual number of dogs affected annually worldwide has never been assessed. However, the disease has a wide distribution in the Mediterranean basin, Africa, Central and South America [9].

The definitive canid host of S. lupi is infected by ingesting an obligate intermediate coprophagous beetle vector, or a variety of paratenic hosts including birds, reptiles, amphibians and small mammals [10] that are infected by S. lupi (Figure 1). The main vector of S. lupi in Israel is the scarab beetle Onthophagus sellatus (Coleoptera: Scarabidae) [11]. The beetle ingests S. lupi eggs upon feeding on the definite host's feces, and within the beetle intermediate host, the infective larvae (L3) develop. Upon ingestion of the beetle, or the paratenic host, by the definitive host, L3 are released in the stomach, penetrate the gastric mucosa and migrate within blood vessel walls to the caudal thoracic aortic wall, where they develop to L4. From there, larvae migrate to the caudal esophagus, where they mature and sexually reproduce. In the esophageal wall the nematodes are surrounded by a nodule, comprised of fibroblasts. The female worms burrow a tunnel through the esophageal wall and pass their eggs, which contain larvae (L1) to the gastrointestinal tract, and into the feces. Dogs infected by S. lupi present variable clinical signs, depending on the stage of the disease. The esophageal nodule can undergo neoplastic transformation, resulting in development of sarcomas (Reviewed in 9). In Israel, spirocercosis is an emerging disease since the 1990's, with 50 dogs diagnosed with the disease annually at the Hebrew University Veterinary Teaching Hospital (HUVTH), most from the Greater Tel Aviv area [8]. Since then, the geographic distribution disease in Israel has widened, and during 2009, 91 dogs were diagnosed with spirocercosis at the HUVTH, of which 33 dogs had neoplastic esophageal disease, and died or were euthanized shortly post presentation. Additionally, the geographic distribution of the disease during this period had widened, and is no more restricted to the Greater Tel-Aviv area, but includes all the subtropical areas in the country (I. Aroch, unpublished data).

Figure 1
figure 1

Schematic life-cycle of Spirocercal lupi . Eggs containing L1 larvae are found in the feces of the infected canid host (Feces: L1). The intermediate host, a dung beetle, consumes the feces and ingest the eggs (A). The eggs hatch and the larvae develop into L3 (Intermediate host: L1-L3). The intermediate host can either be consumed by paratenic hosts such as birds or small mammals (B), in which L3 arrest their development (paratenic host: L3), or by the definitive host (C) where the L3 larvae are released in the stomach, penetrate the gastric mucosa and migrate within blood vessel walls to the caudal thoracic aortic wall, where they develop to L4. From there, larvae migrate to the caudal esophagus, where they mature and sexually reproduce (E, Definitive host: L3-L5). Alternatively, the definitive host preys on L3 infected paratenic hosts (D). Adult worms are found in the esophageal wall, surrounded by a nodule. The female worms pass their eggs to the gastrointestinal tract, and into the feces (F, Definitive host: L5-eggs).

Diagnosis of spirocercosis is always challenging, because the clinical signs are variable and occur in advanced disease stages. Most animals are thus diagnosed only in the advanced stage of the disease, once nodules containing adult to egg shedding worms, are present in the esophagus [8]. The diagnosis of the disease in its early stages, prior to formation of esophageal nodules and egg shedding, is currently difficult and is almost impossible.

Recent studies have shown a relationship between bacterial symbionts of the genus Wolbachia and filarial pathogenic nematodes [12]. Wolbachia which is estimated to infect 66% of arthropods and nematodes [13] can manipulate various aspects of its arthropod hosts’ biology [14]. Wolbachia was found to be an obligatory symbiont of certain filarial nematodes, with a possible role in the pathogenesis and immune response to filarial infection in the mammalian host [4, 5, 15, 16].

In the current study, we tested for the presence of Wolbachia species and other specific symbionts in the nematode S. lupi, and detected a novel and stable infection in the worm. Our findings are expected to promote further understanding of the interactions among various organisms in complex systems such as spirocercosis, and may have clinical implications, because this stable bacterial infection can potentially be used for novel simple diagnostic methods of this disease and aid in its prevention and treatment.

Results and discussion

Identification of novel bacterial symbiont in S. lupi from the Thelazioidea super family

DNA of S. lupi adults and larvae was extracted as described below, and was used for the detection of possible bacterial symbiont species including Wolbachia, Cardinium and Rickettsia, by diagnostic PCR using specific primers. All S. lupi DNA samples were found to be negative for these bacteria, while all the control DNA samples were positive, as expected. This is in agreement with other studies, that have failed to detect Wolbachia in certain species of the super family Filarioidea [17], and in other previously tested non-filarial nematode groups ([18] and reference within). Thus, in order to detect other possible bacteria within the nematode, general 16S rDNA (rrs gene) primers able to detect most known Eubacteria were used in PCR. Adult nematode's DNA templates were positive for this bacterial gene, and the PCR products were cloned and sequenced. BLAST analysis (http://blast.ncbi.gov.il/) revealed initial similarity to sequences of the genus Comamonas, a beta-Proteobacterium of the Comamonadaceae family, as published in GenBank. Consensus sequence of the identified Comamonas sp. was determined, and deposited in GenBank under the accession number: JQ361660. In addition, for detection of other bacteria, rrs PCR-DGGE analysis was performed. DGGE separation resulted in a single product, suggesting that S. lupi probably carries only a single bacterial species (Figure 2a). Sequences of the excised DGGE band were also highly similar to the genus Comamonas. Based on the consensus sequence, Comamonas sp. specific primers were designed and used in nested PCR on DNA template extracted from S. lupi adults, larvae (L2 and L3) and eggs. All these stages were found to carry the same Comamonas bacterium (Figure 2b). These findings suggest that this novel Comamonas sp. is vertically transmitted, and suggests that a long-term association between Comamonas sp. and S. lupi exists.

Figure 2
figure 2

Detection of a single bacterium, Comamonas sp., in Spirocerca lupi (a) Separation of DNA samples from 3 adult S. lupi after PCR analysis with general eubacterial primers on denaturing gradient gel electrophoresis (40% to 60% urea/formamide gradient) showing a single band result. (b) Detection of Comamonas sp. in DNA samples of S. lupi eggs, larvae (L2, L3), and adults (males and females), using PCR with Comamonas sp. specific primers.

Phylogenetic analysis of the S. lupi symbiont

Based on a nearly full length rrs gene from the above identified Comamonas symbiont of S. lupi, and other selected Comamonas spp. sequences, a phylogenetic tree was built. The phylogeny analysis showed that the current Comamonas sp. sequence is clustered in a separate branch, together with C. testosteroni, known to participate in steroid degradation [19], and other soil-derived Comamonas species, represented herein by C. composti[20] (Figure 3). Comamonas spp., however, are not strict soil bacteria, and have recently been described in several insect species. Interestingly, the S. lupi-dervied Comamonas sp. is clustered in the same clade of Comamonas spp. identified in blood feeding insects, such as mosquitoes [21, 22] and a flea [23]. This clade is separated from Comamonas spp. identified in non-blood feeders, namely the termite Odontotermes formosanus[24], a plant hopper, and a moth (Su and Li 2010: GenBank report GQ206315, Yin et al. 2008: GenBank report EU344924, respectively). The same clade also includes a Comamonas sp. identified in a soil nematode, Oscheius sp. (Deepa et al. 2010: GenBank report HQ200412). None of these studies, however, have suggested a role for these Comamonas spp. in their invertebrate hosts.

Figure 3
figure 3

Comamonas sp. from Spirocerca lupi is closely related to soil derived Comamonas spp. and to Comamonas spp. from blood feeding arthropods. Phylogenetic analysis based on maximum likelihood tree (1000 bootstraps) constructed with 16S rDNA sequences of various Comamonas species from different origin and host species. Host species are marked with asterisks. Published GenBank accession numbers are noted for each species. Bootstrap values are indicated on branches.

At present, the role that the identified Comamonas sp. plays in the biology of the nematode remains unknown, and so is its potential role in canine spirocercosis. A recent study, however, showed that benign infection with S. lupi induces an immune response that is atypical to chronic helminthic infection, but rather suggests a bacterial infection [25].

Localization of Comamonas sp. within S. lupi

Based on the rrs sequence of the novel Comamonas sp., a specific probe for fluorescent in situ hybridization (FISH) analysis was designed. Using the Comamonas-specific probe, we were able to demonstrate a specific signal in the gut epithelium of S. lupi larvae (Figure 4). The localization of the present Comamonas bacterium in the nematode's gut epithelium, and the phylogenetic proximity to other Comamonas spp. detected in blood-feeding insect hosts, may suggest that this novel Comamonas sp. plays a role in blood digestion or degradation within S. lupi, which feeds on its vertebrate hosts’ blood and tissues [9]. In addition, the FISH result, combined with the detection of Comamonas sp. in all the tested developmental stages of S. lupi using PCR, as described above, are in support of a stable, non- axenic infection of S. lupi by this bacterium.

Figure 4
figure 4

Comamonas sp. is restricted to the gut epithelium of Spirocerca lupi L3 larva. Images of fluorescence in-situ hybridization analysis of S. lupi L3 larva stained with Comamonas-specific probe (green), detected using confocal microscopy. (a) No-probe control; (b) Intact L3; (c, d) Ruptured L3; (e) Enlargement of (d), showing specific signal in the larval gut; (f) One optical section showing a specific signal only in the gut epithelium region. The arrow points to a specific focal point. All images but (f) are combined optical Z sections, overlaid on a bright-field image.

Detection of S. lupi-derived Comamonas sp. in blood samples of infected dogs

DNA detection from the S. lupi-derived Comamonas sp. in infected dogs may potentially be important in understanding the pathogenesis and promoting the diagnosis of spirocercosis. Recently, the symbiotic bacterium Wolbachia was detected in blood samples of dogs infected by the heartworm Dirofilaria immitis[26]. In the present study, we used a diagnostic semi-nested PCR with Comamonas-specific primers on DNA extracted from blood samples of dogs definitely diagnosed with spirocercosis and of negative control dogs. Comamonas sp. DNA was detected in 9/18 (50%) samples obtained from dogs with spirocercosis, but in none of 11 negative control samples (Figure 5). The rather low detection rate of Comamonas sp. in the dogs infected with the nematode may be due to several reasons; an unavailable bacterial template; improper storage of blood samples, resulting in insufficient DNA preparation, or an undetectable symbiont template in standard PCR due to unknown PCR inhibitors on a low concentration of Comamonas DNA in the blood. Alternatively, detection of the symbiont in blood samples may depend on the specific interactions between the bacterium and the nematode within the definitive canine host. It may be speculated that bacteria are only released from the nematode upon its death and disintegration, or within a limited specific time-point during infection within the definitive canine host. Further studies are warranted, to assess the optimal blood storage protocols and DNA extraction methods of canine samples, along with spiking experiments with Comamonas sp. and quantitative PCR in order to devise an efficient molecular diagnostic protocol.

Figure 5
figure 5

DNA of Comamonas sp. can be detected in blood samples of dogs infected with Spirocerca lupi . PCR detection of Comamonas sp. in samples of DNA extracted from blood of dogs infected with S. lupi. 1-no template control, 2-Trichinella spiralis, 3-healthy dog, 4-21-sick dogs, 22- S. lupi L3.

Conclusions

In the present study, we detected an additional organism, a bacterial symbiont of the genus Comamonas, within the causative agent of spirocercosis, the nematode S. lupi. Recently, microbial symbiosis has been repetitively shown to be a driving force in the biology and evolution of many organisms. The present study adds yet additional evidence of this trend, in a highly complex system. Resolution of the complex interactions among the different organisms involved in the spirocercosis system may lead to novel, applicable methods for the early diagnosis, prevention and treatment of canine spirocercosis, in a similar manner as has been applied when the interaction between Wolbachia spp. symbionts with their filarial nematode hosts has been elucidated [3, http://a-wol.com].

Methods

Sample origin

Adult S. lupi worms were obtained from esophageal nodules of dogs diagnosed with spirocercosis at the Hebrew University Veterinary Teaching Hospital, at necropsy, and stored in −20°C pending analysis. Larvae (L2 and L3) were dissected under a stereoscope from O. sellatus beetles, isolated in the laboratory from dog fecal dungs, collected in a public park located in a S. lupi-endemic area in Central Israel [11]. These were either stored in absolute ethanol at −20°C, or freshly used. S. lupi eggs were concentrated through floatation [27], and stored as described above.

Blood samples were obtained from dogs diagnosed with spirocercosis through esophageal endoscopy and presence of eggs in the feces, and from puppies aged 2 to 4 months, housed in a breeding farm. Puppies were chosen as negative control because they were housed in a restricted kennel, and were thus unexposed to feces of other dogs.

DNA extraction, PCR, clone library and sequencing

DNA of adult S. lupi worms was extracted using hexadecyltrimethyl ammonium-bromide (CTAB) buffer [28], and were used in PCR with the 16S rDNA (rrs) gene primer set, targeting most known eubacteria (27F-1494R; [29]), under the following reaction conditions: 3 min at 95°C; 35 cycles of 1 min at 95°C, 1 min at 55°C, 1.5 min at 72°C; and 5 min at 72°C. The PCR products were run on 1% agarose gel, and were later extracted and cloned into pGEM-T easy vector (Promega, Madison, WI, USA), and transformed into competent Escherichia coli. Plasmids from 10 inserted clones were extracted from the gel and sequenced (HyLabs, Rehovot, Israel). As a control for DNA quality, PCR analysis was performed using primers for the S. lupi-specific cytochrome oxidase subunit 1 gene (cox1) as previously described [30].

Direct probing of known invertebrate symbiont

DNA of S. lupi was used in PCR with specific primers and conditions to identify Wolbachia, Cardinium and Rickettsia spp., as previously described [31]. DNA extracted from Bemisia tabaci, harboring Wolbachia and Rickettsia spp., and from Plagiumerus diaspidis containing Cardinium sp. were used as positive controls.

Denaturating gradient gel electrophoresis (DGGE)

PCR was performed on adult S. lupi DNA using primers GC-clamp 341 F-907R, targeting the bacterial rrs gene, with PCR conditions permitting its amplification from most known bacteria [32]. DGGE was performed using a 40% to 60% urea/formamide gradient for standard reactions. After the electrophoresis, gel was incubated in ethidium-bromide solution (250 ng/ml) for 10 min, rinsed in distilled water, and photographed under UV illumination. Bands were extracted, and sent for direct sequencing (HyLab, Rehovot, Israel).

Phylogenetic analysis

Nine nearly-full length sequences of the rrs gene of Comamonas sp. were obtained from five adult S. lupi worms. All clones were sequenced from both directions. Sequences were edited using DNAMAN software (Lynnon Corporation, Canada) and a consensus sequence was determined. The Comamonas sp. rrs sequence was aligned, using MUSCLE 3.7, with other published Comamonas spp. sequences, selected based on BLAST results, and based on their invertebrate host origin. The rrs gene sequence of Verminephrobacter eiseniae was used as an out group. A maximum-likelihood tree was constructed using PhyML 3.0 software. Bootstrap analyses with 1000 re-samplings were performed to test branching robustness. The tree was illustrated using TreeDyn 198.3. All software packages are available at http://www.phylogeny.fr/.

Direct probing of Comamonas sp.

To confirm the presence of Comamonas sp. in the various S. lupi developmental stages (eggs, larvae and adults), and in blood samples obtained from S. lupi- infected dogs, a diagnostic PCR was planned. Based on the rrs sequence established, specific primers were designed; /ComF323/ 5’-CCTCGGGTTGTAAACTGCTT-3’ and /ComR1393/ 5’-TCTCTTTCGAGCACGAATCC-3’. The primers were used in a standard PCR, under the following conditions: 3 min at 95°C; 35 cycles of 1 min at 95°C, 1 min at 58°C, 1 min at 72°; and a final 5 min at 72°C. The PCR product size was expected to be ca. 1000 bp. Positive and negative PCR products were retested using semi-nested PCR, with the forward primer /ComNest F/ 5’- ACTGCCATTGTGACTGCAAG-3’ and the ComR1393 reverse primer, with PCR conditions as described above, resulting in ca. 600 bp product. Three PCR products from each sample category were directly sequenced in order to confirm the Comamonas specific sequence.

Fluorescent in-situ hybridization (FISH)

FISH was performed as previously described [33]. Briefly, larvae were fixed in Carnoy’s fixative (6:3:1 parts of chloroform: ethanol: acetic acid), and later hybridized with the rrs-based designed probe: Com-probe /Cy3/ 5’- TGTGCTACTAGAGCGGCTGA-3’, in hybridization buffer. Since intact larvae could not uptake the probe, larvae were first ruptured using sterile insect pins, and their content was removed from the cuticle. Specimens were viewed under an IX81Olympus FluoView500 confocal microscope. Signal specificity was confirmed based on sequence comparison in the ‘Probe Match’ function in the Ribosomal Database Project website (http://rdp.cme.msu.edu/), and using a no-probe control, and hybridization to a non-target nematode, Trichinella spiralis.

Ethics statement

Samples (nematodes and blood) were obtained from S. lupi-infected dogs presented to the Hebrew University Veterinary Teaching Hospital, Koret School of Veterinary Medicine, Hebrew University of Jerusalem with their owners' consent, during diagnosis, treatment and necropsy. Samples obtained from control dogs were obtained with their owner's consent. This study was approved by the Institutional Committee of Animal Handling and Experimentation.

References

  1. Brouqui P, Fournier PE, Raoult D: Doxycycline and eradication of microfilaremia in patients with loiasis. Emerg Infect Dis. 2001, 7: 604-605.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  2. Hoerauf A, Specht S, Buttner M, Pfarr K, Mand S, Fimmers R, Marfo-Debrekyei Y, Konadu P, Debrah AY, Bandi C, Brattig N, Albers A, Larbi J, Batsa L, Taylor MJ, AdJei O, Buttner DW: Wolbachia endobacteria depletion by doxycycline as antifilarial therapy has macrofilaricidal activity in onchocerciasis: a randomized placebo-controlled study. Med Microbiol Immunol. 2008, 197: 295-311. 10.1007/s00430-007-0062-1.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  3. Slatko B, Taylor M, Foster J: The Wolbachia endosymbiont as an anti-filarial nematode target. Symbiosis. 2010, 51: 55-65. 10.1007/s13199-010-0067-1.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Hansen RDE, Trees AJ, Bah GS, Hetzel U, Martin C, Bain O, Tanya VN, Makepeace BL: A worm's best friend: Recruitment of neutrophils by Wolbachia confounds eosinophil degranulation against the filarial nematode Onchocerca ochengi. Proc R Soc B. 2011, 278: 2293-2302. 10.1098/rspb.2010.2367.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Kramer L, Grandi G, Leoni M, Passeri B, McCall J, Genchi C, Mortarino M, Bazzocchi C: Wolbachia and its influence on the pathology and immunology of Dirofilaria immitis infection. Vet Parasitol. 2008, 158: 191-195. 10.1016/j.vetpar.2008.09.014.

    Article  PubMed  CAS  Google Scholar 

  6. McCall JW, Kramer L, Genchi C, Guerrero J, Dzimianski MT, Supakorndej P, Mansour A, McCall SD, Supakorndej N, Grandi G, Carson B: Effects of doxycycline on early infections of dirofilaria immitis in dogs. Vet Parasitol. 2011, 176: 361-367. 10.1016/j.vetpar.2011.01.022.

    Article  PubMed  CAS  Google Scholar 

  7. Mylonakis ME, Rallis T, Koutinas AF, Leontides LS, Patsikas M, Florou M, Papadopoulos E, Fytianou A: Clinical signs and clinicopathologic abnormalities in dogs with clinical spirocercosis: 39 cases (1996–2004). J Am Vet Med Assoc. 2006, 228: 1063-1067. 10.2460/javma.228.7.1063.

    Article  PubMed  Google Scholar 

  8. Mazaki-Tovi M, Baneth G, Aroch I, Harrus S, Kass PH, Ben-Ari T, Zur G, Aizenberg I, Bark H, Lavy H: Canine spirocercosis: clinical, diagnostic, pathologic, and epidemiologic characteristics. Vet Parasitol. 2002, 107: 235-250. 10.1016/S0304-4017(02)00118-8.

    Article  PubMed  Google Scholar 

  9. van der Merwe LL, Kirberger RM, Clift S, Williams M, Heller N, Naidoo V: Spirocerca lupi infection in the dog: a review. Vet J. 2007, 176: 294-309.

    Article  PubMed  Google Scholar 

  10. Fox SM, Burns J, Hawkins J: Spirocercosis in dogs. Comp Cont Educ Pract Vet. 1988, 10: 807-824.

    Google Scholar 

  11. Gottlieb Y, Markovics A, Klement E, Naor S, Samish M, Aroch I, Lavy E: Characterization of Onthophagus sellatus as the major intermediate host of the dog esophageal worm Spirocerca lupi in Israel. Vet Parasitol. 2011, 180: 378-382. 10.1016/j.vetpar.2011.03.008.

    Article  PubMed  Google Scholar 

  12. Fenn K, Blaxter M: Wolbachia. Edited by: Hoerauf A, Rao R. 2007, Basel: Karger, 66-76.Coexist, cooperate and thrive: Wolbachia as long-term symbionts of filarial nematodes,Issues Infect Dis,

    Google Scholar 

  13. Hilgenboecker K, Hammerstein P, Schlattmann P, Telschow A, Werren JH: How many species are infected with Wolbachia? – a statistical analysis of current data. FEMS Microbiol Lett. 2008, 281: 215-220. 10.1111/j.1574-6968.2008.01110.x.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  14. Werren JH, Baldo L, Clark ME: Wolbachia: Master manipulators of invertebrate biology. Nat Rev Microbiol. 2008, 6: 741-751. 10.1038/nrmicro1969.

    Article  PubMed  CAS  Google Scholar 

  15. Saint André AV, Blackwell NM, Hall LR, Hoerauf A, Brattig NW, Volkmann L, Taylor MJ, Ford L, Hise AG, Lass JH, Diaconu E, Pearlman E: The role of endosymbiotic Wolbachia bacteria in the pathogenesis of river blindness. Science. 2002, 295: 1892-1895. 10.1126/science.1068732.

    Article  PubMed  Google Scholar 

  16. Tamarozzi F, Halliday A, Gentil K, Hoerauf A, Pearlman E, Taylor MJ: Onchocerciasis: The role of Wolbachia bacterial endosymbionts in parasite biology, disease pathogenesis, and treatment. Clin Microbiol Rev. 2011, 24: 459-468. 10.1128/CMR.00057-10.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  17. Ferri E, Bain O, Barbuto M, Martin C, Lo N, Uni S, Landmann F, Baccei SG, Guerrero R, de Souza Lima S, Bandi C, Wanji S, Diagne M, Casiraghi M: New insights into the evolution of Wolbachia infections in filarial nematodes inferred from a large range of screened species. PLoS One. 2011, 6: e20843-10.1371/journal.pone.0020843.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  18. Foster JM, Kumar S, Ford L, Johnston KL, Ben R, Graeff-Teixeira C, Taylor MJ: Absence of Wolbachia endobacteria in the non-filariid nematodes Angiostrongylus cantonensis and A. costaricensis. Parasites & Vectors. 2008, 1: 31-35. 10.1186/1756-3305-1-31.

    Article  Google Scholar 

  19. Horinouchi M, Hayashi T, Kudo T: Steroid degradation in Comamonas testosteroni. J Steroid Biochem Mol Biol. 2012, 129: 4-14. 10.1016/j.jsbmb.2010.10.008.

    Article  PubMed  CAS  Google Scholar 

  20. Young C-C, Chou J-H, Arun AB, Yen W-S, Sheu S-Y, Shen F-T, Lai W-A, Rekha PD, Chen W-M: Comamonas composti sp. Nov., isolated from food waste compost. ISME J. 2008, 58: 251-256.

    CAS  Google Scholar 

  21. Lindh JM, Borg-Karlson AK, Faye I: Transstadial and horizontal transfer of bacteria within a colony of Anopheles gambiae (Diptera: Culicidae) and oviposition response to bacteria-containing water. Acta Trop. 2008, 107: 242-250. 10.1016/j.actatropica.2008.06.008.

    Article  PubMed  CAS  Google Scholar 

  22. Zouache K, Voronin D, Tran-Van V, Mousson L, Failloux A-B, Mavingui P: Persistent Wolbachia and cultivable bacteria infection in the reproductive and somatic tissues of the mosquito vector Aedes albopictus. PLoS One. 2009, 4: e6388-10.1371/journal.pone.0006388.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Jones RT, McCormick KF, Martin AP: Bacterial communities of Bartonella-positive fleas: Diversity and community assembly patterns. Appl Environ Microbiol. 2008, 74: 1667-1670. 10.1128/AEM.02090-07.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  24. Chou J-H, Sheu S-Y, Lin K-Y, Chen W-M, Arun AB, Young C-C: Comamonas odontotermitis sp. Nov., isolated from the gut of the termite Odontotermes formosanus. IJSEM. 2007, 57: 887-891.

    PubMed  CAS  Google Scholar 

  25. Dvir E, Mellanby RJ, van der Merwe LL, Kjelgaard-Hansen M, Schoeman JP:Differences in the plasma cytokine milieu between dogs with benign and malignant spirocercosis. The 21th Congress of the European College of Veterinary Internal Medicine Companion Animals (ECVIM-CA), September 2011, Seville, Spain. 2011,

    Google Scholar 

  26. Rossi MID, Aguiar-Alves F, Santos S, Paiva J, Bendas A, Fernandes O, Labarthe N: Detection of Wolbachia DNA in blood from dogs infected with Dirofilaria immitis. Exp Parasitol. 2010, 126: 270-272. 10.1016/j.exppara.2010.05.002.

    Article  PubMed  CAS  Google Scholar 

  27. Markovics A, Medinski B: Improved diagnosis of low intensity Spirocerca lupi infection by sugar flotation method. J Vet Diagn Invest. 1996, 8: 400-401. 10.1177/104063879600800326.

    Article  PubMed  CAS  Google Scholar 

  28. Chen DH, Ronald PC: A rapid DNA minipreparation method suitable for AFLP and other PCR applications. Plant Mol Biol Rep. 1999, 17: 53-57. 10.1023/A:1007585532036.

    Article  CAS  Google Scholar 

  29. Weisburg WG, Barns SM, Pelletier DA, Lane DJ: 16S ribosomal DNA amplification for phylogenetic study. J Bacteriol. 1991, 173: 697-703.

    PubMed  CAS  PubMed Central  Google Scholar 

  30. Traversa D, Costanzo F, Iorio R, Aroch I, Lavy E: Mitochondrial cytochrome C oxidase subunit 1 (cox1) gene sequence of Spirocerca lupi (Nematoda, Spirurida): Avenues for potential implications. Vet Parasitol. 2007, 146: 263-270. 10.1016/j.vetpar.2007.03.015.

    Article  PubMed  CAS  Google Scholar 

  31. Chiel E, Gottlieb Y, Zchori-Fein E, Mozes-Daube N, Katzir N, Inbar M, Ghanim M: Biotype-dependent secondary symbiont communities in sympatric populations of Bemisia tabaci. Bull Entomol Res. 2007, 97: 407-413. 10.1017/S0007485307005159.

    Article  PubMed  CAS  Google Scholar 

  32. Muyzer G, de Waal EC, Uitterlinden AG: Profiling of complex microbial populations by denaturing gradient gel electrophoresis analysis of polymerase chain reaction-amplified genes coding for 16S rRNA. Appl Environ Microbiol. 1993, 59: 695-700.

    PubMed  CAS  PubMed Central  Google Scholar 

  33. Gottlieb Y, Ghanim M, Gueguen G, Kontsedalov S, Vavre F, Fleury F, Zchori-Fein E: Inherited intracellular ecosystem: Symbiotic bacteria share bacteriocytes in whiteflies. FASEB J. 2008, 22: 2591-2599. 10.1096/fj.07-101162.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgments

We would like to Dr. Shachar Naor and Dr. Zippi Prize for their technical assistance in the laboratory work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yuval Gottlieb.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

YG conceived the study, participated in its design, performed the molecular identification as well as phylogenetic and FISH analyses, and wrote the paper. MK performed the specific molecular detection, and analyzed the data. AM participated in the study design and perform the eggs and larvae collections. MG participated in the study design, provided controls and drafted the paper. EL and IA conceived the study, participated in its design, provided adult worms, and drafted the paper. All authors read and approved the final manuscript.

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Gottlieb, Y., Lavy, E., Kaufman, M. et al. A novel bacterial symbiont in the nematode Spirocerca lupi. BMC Microbiol 12, 133 (2012). https://doi.org/10.1186/1471-2180-12-133

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2180-12-133

Keywords