26.12.2013 Views

Get PDF (1214K) - Wiley Online Library

Get PDF (1214K) - Wiley Online Library

Get PDF (1214K) - Wiley Online Library

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

RESEARCH ARTICLE<br />

Characterization and localization of Rickettsia sp. in water beetles of<br />

genus Deronectes (Coleoptera: Dytiscidae)<br />

Stefan Martin Küchler, Siegfried Kehl & Konrad Dettner<br />

Department of Animal Ecology II, University of Bayreuth, Bayreuth, Germany<br />

Correspondence: Stefan Martin Küchler,<br />

Department of Animal Ecology II, University of<br />

Bayreuth, Universitätsstraße 30, 95440<br />

Bayreuth, Germany. Tel.: 10049 0921 55<br />

2733; fax: 10049 0921 55 2743; e-mail:<br />

stefan.kuechler@uni-bayreuth.de<br />

Received 22 September 2008; revised 3<br />

February 2009; accepted 3 February 2009.<br />

First published online 16 March 2009.<br />

DOI:10.1111/j.1574-6941.2009.00665.x<br />

Editor: Michael Wagner<br />

Keywords<br />

Rickettsia; Deronectes; Dytiscidae; citrate<br />

synthase; electron microscopy.<br />

Abstract<br />

In the present study, Rickettsia sp. was detected in four water beetles of the genus<br />

Deronectes (Dytiscidae) for the first time. Rickettsiae were found in 100% of<br />

examined specimens of Deronectes platynotus (45/45), 39.4% of Deronectes aubei<br />

(28/71), 40% of Deronectes delarouzei (2/5) and 33.3% of Deronectes semirufus<br />

(1/3). Analysis of 16S rRNA gene sequences revealed a phylogenetic relationship<br />

with rickettsial isolates of Limonia chorea (Diptera), tentatively classified as<br />

members of the basal ancestral group. Phylogenetic analysis of the gltA (citrate<br />

synthase) gene sequences showed that Deronectes symbionts were closest to<br />

bacterial symbionts from spiders. Ultrastructural examinations revealed typical<br />

morphological features and intracellular arrangements of rickettsiae. The distribution,<br />

transmission and localization of Rickettsia sp. in D. platynotus were studied<br />

using a diagnostic PCR assay and FISH. Eggs from infected females of D. platynotus<br />

were all Rickettsia-positive, indicative of a vertical transmission.<br />

Introduction<br />

Gram-negative bacteria of the genus Rickettsia belong to the<br />

family Rickettsiaceae in the subdivision of the monophyletic<br />

order Rickettsiales (Alphaproteobacteria). All members of the<br />

genus Rickettsia are obligate intracellular bacteria that reside<br />

free in the cytoplasm of eukaryotic cells. Mostly known<br />

as pathogenic endosymbionts of arthropods (ticks, mites,<br />

fleas and lice), rickettsiae cause human and animal diseases<br />

with a worldwide distribution (Eremeeva & Dasch, 2000).<br />

Some species are also capable of infecting and causing<br />

disease in plants (Davis et al., 1998). Rickettsiae are usually<br />

transmitted by arthropods to vertebrates via feces or salivary<br />

secretions. Because of their considerable medical importance,<br />

the genus Rickettsia was traditionally divided into two<br />

groups (Weiss & Moulder, 1984): the spotted fever group,<br />

including Rickettsia rickettsii and allied species, causative<br />

agents of spotted fever; and the typhus group, including<br />

Rickettsia prowazekii and Rickettsia typhi, the agents of<br />

murine and endemic typhus, respectively. In recent years,<br />

new rickettsiae have been discovered that do not belong to<br />

the previously described subgroups. A third group named<br />

‘ancestral group’, partially subdivided into the bellii, leech<br />

and limoniae groups (Perlman et al., 2006; Perotti et al.,<br />

2006), was established (Stothard & Fuerst, 1995; Fournier &<br />

Raoult, 2007), and consists of Rickettsia bellii, Rickettsia<br />

canadensis and additional numerous rickettsial isolates that<br />

have been isolated from different terrestrial arthropods such<br />

as whiteflies, flies, aphids and beetles (Chen et al., 1996;<br />

Fukatsu et al., 2000; Campbell et al., 2004; Gottlieb et al.,<br />

2006; Zchori-Fein et al., 2006). The ancestral group is still an<br />

initial theoretical classification that is not based on<br />

morphological criteria or biochemical features concerning<br />

basal standing rickettsial isolates in the genus. Rickettsiae of<br />

the ancestral group were not shown to be pathogens for<br />

humans or their animal hosts. However, insects are not the<br />

only vector of rickettsiae. Recently, rickettsiae have been<br />

found in spider mite Tetranychus urticae (Hoy & Jeyaprakash,<br />

2005), various spider species (Goodacre et al., 2006),<br />

amoeba Nuclearia pattersoni (Dykova et al., 2003), marine<br />

ciliate Diophrys appendiculata (Vannini et al., 2005) and<br />

several leeches from freshwater environments (Kikuchi et al.,<br />

2002; Kikuchi & Fukatsu, 2005), which also cluster into the<br />

ancestral group. Analyses of 16S rRNA genes suggest that<br />

many of these rickettsial isolates could represent new genera<br />

within the order Rickettsiales. As already supposed (Yu &<br />

Walker, 2006), the family Rickettsiaceae could be more<br />

widespread than observed so far.<br />

FEMS Microbiol Ecol 68 (2009) 201–211<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved


202 S.M. Küchler et al.<br />

The purpose of this study was to investigate and identify<br />

symbiotic bacteria in selected water beetles of the palaearctic<br />

genus Deronectes. The association of endosymbiotic bacteria<br />

with water insects, especially water beetles, was already<br />

presumed in previous studies (Maddison et al., 1999; Shull<br />

et al., 2001). Representatives of the genus Deronectes are<br />

predacious and colored uniformly black or brown. Deronectes<br />

species usually live among gravel in small, swift or<br />

strongly flowing streams with sparse vegetation. Most<br />

species prefer mountainous regions, some occurring at a<br />

high altitude. To date 55 species have been described (Fery &<br />

Brancucci, 1997; Fery et al., 2001).<br />

In the present molecular study, we report the first finding<br />

of a coccobacillus bacterium in Deronectes, which could be<br />

identified as a member of the genus Rickettsia. A total of 136<br />

specimens from seven Deronectes species were examined for<br />

the existence of Rickettsia sp. using a diagnostic PCR assay.<br />

Phylogenetic analysis inferred by 16S rRNA gene and<br />

particularly the rickettsial gltA (citrate synthase) gene<br />

disclosed the phylogenetic position of beetle Rickettsia in<br />

the existing phylogenetic system of Rickettsia endosymbionts.<br />

FISH was used for the investigation of rickettsial<br />

distribution and vertical transmission rates. The morphological<br />

characteristics of the bacterium were analyzed by<br />

electron microscopic observations.<br />

Materials and methods<br />

Samples<br />

A total of 136 Deronectes specimens were collected between<br />

2003 and 2008 (Table 1) by kick-sampling (Schwoerbel,<br />

1994). Deronectes platynotus and Deronectes latus were<br />

obtained from Fichtelgebirge, a mountain range in northeastern<br />

Bavaria, Germany. Deronectes aubei was collected<br />

from Black Forest (Germany), Italy and French alps. Deronectes<br />

delarouzei, Deronectes semirufus, D. aubei sanfilippoi<br />

and Deronectes moestus inconspectus came from Spain, Italy<br />

and France. Beetles were brought to laboratory alive and<br />

embedded for histology/FISH or dissected for diagnostic<br />

PCR immediately.<br />

Histology<br />

Before all the beetles were fixed in 4% paraformaldehyde<br />

overnight, the elytra were carefully removed. The fixed<br />

beetles were washed in 1 phosphate-buffered saline and<br />

96% ethanol (1 : 1), dehydrated serially in ethanol (70%,<br />

90%, 2 100%) and embedded in UNICRYL TM (Plano<br />

GmbH, Germany). Serial sections (2 mm) were cut using a<br />

Leica Jung RM2035 rotary microtome (Leica Instruments<br />

GmbH, Germany), mounted on epoxy-coated glass slides<br />

and subjected to FISH.<br />

DNA extraction, cloning and sequencing<br />

DNA was extracted using a QUIAGEN DNeasy s Tissue Kit<br />

(QUIAGEN GmbH, Germany) following the protocol for<br />

animal tissue. The eubacterial 16S rRNA gene was PCR<br />

amplified using the primer set 07F (5 0 -AGAGTTT<br />

GATCMTGGCTCAG-3 0 ) and 1507R (5 0 -TACCTTGTTAC<br />

GACTTCAC-3 0 ) (Lane, 1991). All PCR reactions were<br />

performed in a Biometra thermal cycler with the following<br />

program: an initial denaturating step at 94 1C for 3 min,<br />

followed by 34 cycles of 94 1C for 30 s, 50 1C for 2 min and<br />

72 1C for 1 min. A final extension step of 72 1C for 10 min<br />

was included. PCR products of the expected sizes were<br />

cloned using the TOPO TA Cloning s Kit (Invitrogen, CA).<br />

Suitable clones for sequencing were selected by restriction<br />

fragment length polymorphism (RFLP). Inserts were digested<br />

by restriction endonucleases HaeIII and TaqI. Plasmids<br />

containing the DNA inserts of the expected sizes were<br />

sequenced at the DNA analytics core facility of the<br />

Table 1. Infection rates of Rickettsia sp. in natural populations of Deronectes spp.<br />

Species Locality Date of collection<br />

No. of<br />

individuals<br />

% of infection<br />

(infected/total)<br />

D. platynotus Fichtelgebirge, Bavaria, Germany 2007/2008 45 100 (45/45)<br />

D. aubei Black Forest, Germany, Italy, France 2003–2008 71 39.4 (28/71)<br />

D. latus Fichtelgebirge, Bavaria, Germany 17.10.2007 7 0 (0/7)<br />

D. delarouzei Bonansa, El pont de Suert, Spain 05.10.2007 2<br />

Llesp, El pont de Suert, Spain 10.08.2004 1 40 (2/5)<br />

05.10.2007 1<br />

Saldes, Spain 05.10.2007 1<br />

D. semirufus Grotta all’Onda, Casoli, Lucca, Tuscany, Italy 09.10.2003 1<br />

Sospel, Moulinet, Apuane-Maritimes, France 07.10.2003 1 33.3 (1/3)<br />

Fociomboli, Apuane Alps, Italy 07.09.2007 1<br />

D. aubei sanfilippoi Vernet les Bains, Pyr. Orientales, France 07.10.2007 3 0 (0/3)<br />

D. moestus inconspectus Saldes, Spain 17.10.2007 2 0 (0/2)<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved<br />

FEMS Microbiol Ecol 68 (2009) 201–211


Rickettsia endosymbiont in water beetles<br />

203<br />

University of Bayreuth with M13 forward and M13 reverse<br />

sequencing primers (Invitrogen).<br />

For diagnostic PCR, the 16S rRNA gene was amplified using<br />

the Rickettsia-specific primers Ri_170F (5 0 -GGGCTTGCTCTA<br />

AATTAGTTAGT-3 0 ) and Ri_1500R (5 0 -ACGTTAGCTCACC<br />

ACCTTCAGG-3 0 ). The citrate synthase gene (gltA) wasdetected<br />

using the primers CS5_F (5 0 -TCTTTATGGGGACC<br />

AGCC-3 0 )andCS4_R(5 0 -TTTCCATTGTGCCATCCAG-3 0 ).<br />

PCR primers were designed using the probe design tool of the<br />

ARB software package (Ludwig et al., 2004). Diagnostic PCR<br />

was performed under the temperature profile described above.<br />

FISH<br />

The following probes were used for FISH targeted to the 16S<br />

rRNA gene: eubacterial probe EUB338 [5 0 -(Cy5)-GCTGCCT<br />

CCCGTAGGAGT-3 0 ](Amannet al., 1995), EUB388 II [5 0 -<br />

(Cy3)-GCAGCCACCCGTAGGTGT-3 0 ], EUB338 III [5 0 -(Cy3)-<br />

GCTGCCACCCGTAGGTGT-3 0 ] (Daims et al., 1999) and<br />

Rickettsia-specific probe Rick_B1 [5 0 -(Cy3)-CCATCATCCC<br />

CTACTACA-3 0 ] (Perotti et al., 2006). Additionally, a nonsense<br />

probe complementary to EUB338, NON338 [5 0 -(Cy3)-<br />

ACTCCTACGGGAGGCAGC-3 0 ] (Manz et al., 1992) was<br />

used as a negative control of the hybridization protocol.<br />

Slides were hybridized in hybridization buffer [20 mM Tris-<br />

HCl (pH 8.0), 0.9 M NaCl, 0.01% sodium dodecyl sulfate<br />

(SDS), and 20% formamide] containing 10 pmol of fluorescent<br />

probes per milliliter, incubated at 46 1C for 90 min,<br />

rinsed in washing buffer [20 mM Tris-HCl (pH 8.0),<br />

225 mM NaCl, 5 mM EDTA, and 0.01% SDS], mounted<br />

with antibleaching solution (VECTASHIELD s Mounting<br />

Medium; Vector Laboratories, UK) and viewed under a<br />

fluorescent microscope.<br />

Electron microscopy<br />

Male accessory glands of D. platynotus were fixed in 2.5%<br />

glutaraldehyde in 0.1 M cacodylate buffer (pH 7.3) for 1 h,<br />

embedded in 2% agarose and fixed again in 2.5% glutaraldehyde<br />

in 0.1 M cacodylate buffer (pH 7.3) overnight.<br />

Specimens were washed in 0.1 M cacodylate three times for<br />

20 min. Following postfixation in 2% osmium tetroxide for<br />

2 h, the specimens were washed and stained en bloc in 2%<br />

uranyl acetate for 90 min. After fixation, specimens were<br />

dehydrated serially in ethanol (30%, 50%, 70%, 95% and<br />

3 100%), transferred to propylene oxide and embedded in<br />

Epon. Ultrathin sections (70 nm) were cut with a diamond<br />

knife (MicroStar, Huntsville, TX) on a Leica Ultracut UCT<br />

microtome (Leica Microsystems, Vienna, Austria). Ultrathin<br />

sections were mounted on pioloform-coated copper<br />

grids, and stained with saturated uranyl acetate, followed by<br />

lead citrate. The sections were viewed with a Zeiss CEM 902<br />

A transmission electron microscope (Carl Zeiss, Oberkochen,<br />

Germany) at 80 kV. Micrographs were taken using an<br />

SO-163 EM film (Eastman Kodak, Rochester, NY).<br />

Phylogenetic analysis<br />

High-quality sequences of the 16S rRNA gene and the gltA<br />

gene for citrate synthase were aligned with the CLUSTALW<br />

software in BIOEDIT (Hall, 1999) and transferred to the ARB<br />

software package and edited manually. Phylogenetic analyses<br />

using maximum parsimony were performed using<br />

PAUP version 4.0b10 (Swofford, 2000). The heuristic<br />

search included 10 000 random addition replicates and tree<br />

bissection–reconnection (TBR) branch swapping with the<br />

Multrees option. A 50% majority-rule consensus tree of the<br />

most parsimonious tree was constructed and exported to<br />

ARB program package, where branch lengths were calculated.<br />

Bootstrap analyses were performed with 500 replicates<br />

with TBR branch swapping, 25 random addition replicates<br />

and the Multrees option in PAUP . The gltA citrate<br />

synthase sequences were checked for chimeras using<br />

CCODE (Gonzalez et al., 2005).<br />

Sequence data<br />

Clone sequences of 16S rRNA gene and gltA sequences for the<br />

D. platynotus-, D. aubei-, D. semirufus- andD. delarouzeiassociated<br />

Rickettsia sp. were deposited in the DDBJ/EMBL/<br />

GenBank nucleotide sequence databases with the accession<br />

numbers FM177868–FM177878 and FM955310–FM955315,<br />

respectively.<br />

Results<br />

Identification of bacterial symbiont<br />

A 1.5-kb segment of the eubacterial 16S rRNA gene was<br />

amplified by PCR and was subjected to cloning and RFLP<br />

typing. Overall, 19 clones were sequenced and compared with<br />

other sequences found in GenBank. The nucleotide sequences<br />

of clones exhibited 99% similarity to the sequence of Rickettsia<br />

limoniae (AF322442, AF322443) from cranefly L. chorea<br />

(Diptera, Limoniidae). The validity of the sequences was<br />

confirmed by FISH performed with probe Rick_B1. Furthermore,<br />

we sequenced a 416-bp fragment of the gltA gene<br />

(citrate synthase) from four Rickettsia-positive D. platynotus<br />

beetles as well as from D. aubei (404 bp), D. semirufus<br />

(367 bp) and D. delarouzei (401 bp). The gltA sequences from<br />

D. platynotus rickettsiae were 100% identical to rickettsial<br />

isolate (DQ 231491) from spider Pityophantes phrygianus<br />

(Goodacre et al., 2006). The gltA gene sequences from<br />

rickettsial symbionts of D. aubei and D. delarouzei exhibited<br />

99% similarity to the sequence of P. phrygianus, whereas<br />

rickettsial gltA sequences from D. semirufus showed 98%<br />

similarity to Rickettsia endosymbionts of spider Meta menegi.<br />

FEMS Microbiol Ecol 68 (2009) 201–211<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved


204 S.M. Küchler et al.<br />

Screening for the presence of Rickettsia sp.<br />

Six Deronectes species and one subspecies were examined<br />

using Rickettsia-specific primers in a diagnostic PCR assay<br />

(Table 1). A total of 45 D. platynotus were screened. Rickettsia<br />

sp. could be detected in 100% of all tested D. platynotus. In<br />

contrast, only 39.4% of 71 examined individuals of D. aubei<br />

were infected. Both species were collected in different seasons<br />

and populations. Just two individuals of D. delarouzei and<br />

one individual of D. semirufus werefoundtobepositivefor<br />

the Rickettsia symbiont. For elaboration of infection rates in<br />

these two species, more specimens from different populations<br />

have to be examined for rickettsiae. No rickettsiae were<br />

detected in seven individuals of D. latus. In the same way,<br />

three surveys of D. aubei sanfilippoi indicated a negative signal<br />

for Rickettsia. Two examined individuals of D. moestus<br />

inconspectus also did not show any infection of Rickettsia.<br />

However, only a few individuals of D. latus, D. aubei<br />

sanfilippoi and D. moestus inconspectus could be examined<br />

for the presence of rickettsiae. All examined species of<br />

D. platynotus and D. aubei exhibited a well-balanced sex ratio.<br />

Furthermore, there was no evidence for male killing, parthenogenesis<br />

or other sex ratio distortion in numerically examined<br />

species of D. platynotus and D. aubei. Currentstudies<br />

indicated that further water beetles are associated with<br />

Rickettsia sp. too (data not shown). Examined species of<br />

Agabus melanarius, Agabus wasastjernae, Agabus guttatus,<br />

Hydroporus gyllenhalii, Hydroporus tristis, Hydroporus umbrosus<br />

and Hydroporus obscurus also exhibited Rickettsia infections.<br />

In situ hybridization of Rickettsia symbiont<br />

FISH revealed a remarkable affection of the Rickettsia<br />

symbiont (Fig. 1a). All tissues that are in contact with the<br />

hemolymph are affected by rickettsiae. The symbionts could<br />

be identified in all compartments of Deronectes, including<br />

the head (ommatidia) (Fig. 1b) and legs. Actually, internal<br />

parts of elytra with soft tissue exhibited bacteria. Especially,<br />

numerous rickettsiae could be detected in tissue of active<br />

metabolism, such as the fatbody and internal reproduction<br />

organs. A lower number of rickettsiae was found in muscles.<br />

In addition, Rickettsia sp. is more abundant in females than<br />

males. Accessory glands of males will be affected predominantly<br />

(Fig. 1c). The strongest signals of a Rickettsia-specific<br />

probe were found in females of D. platynotus. Minor signals<br />

of symbionts were detected in D. aubei, D. semirufus and<br />

D. delarouzei (data not shown).<br />

The distribution of the Rickettsia was also investigated in<br />

eggs and larvae of D. platynotus. The presence of large<br />

numbers of symbionts in oocytes and follicle cells (Fig. 1d),<br />

and the detection of Rickettsia in eggs (Fig. 1e) are indicative<br />

of the vertical transmission of symbionts. In the second and<br />

third larval stages of Deronectes, rickettsiae could be detected<br />

in the entire body. The number of symbionts increased from<br />

stage to stage.<br />

Fig. 1. FISH with specific probe Rick_B1 (Cy3) of<br />

square sections of water beetle Deronectes<br />

platynotus adults. (a) Rickettsia sp. infection of<br />

fatbody in a female abdomen. Scale bar = 40 mm.<br />

(b) Square section of head; numerous Rickettsia<br />

signals from ommatidia (arrows). Rickettsia sp.<br />

clustered around the cerebral ganglia. Scale<br />

bar = 40 mm. (c) Male accessory glands infected<br />

by intracytosolic Rickettsia sp. Bacteria are<br />

located near the connective tissue (arrow). Scale<br />

bar = 20 mm. (d) Ovariole infected by Rickettsia<br />

sp. symbiont. Scale bar = 20 mm. (e) Bacterial<br />

symbionts are also present within the egg<br />

(arrow). Surrounding tissue of ovarioles habouring<br />

Rickettsia. Scale bar = 50 mm. FB, fat body;<br />

B, brain; E, egg; AGT, accessory glandular tissue;<br />

M, muscle; O, ovariole; OM, ommatidia. (a), (b)<br />

and (c) were taken by confocal microscope.<br />

(d) and (e) were taken by a conventional<br />

fluorescence microscope.<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved<br />

FEMS Microbiol Ecol 68 (2009) 201–211


Rickettsia endosymbiont in water beetles<br />

205<br />

Fig. 2. FISH double hybridization of the<br />

Rickettsia symbiont on tissue (fatbody) crosssection<br />

from Deronectes platynotus (a–c, scale<br />

bar = 40 mm) and Deronectes aubei (d–f, scale<br />

bar = 20 mm). The left panels (a, d) show FISH<br />

with the probe Cy3-Rick_B1, which visualizes the<br />

Rickettsia symbiont, while middle panels (b, e)<br />

show FISH with the probe Cy5-EUB338, which<br />

recognizes diverse eubacteria universally. The<br />

overlay (c, f) does not show other bacteria.<br />

Double hybridization performed with Rick_B1, together<br />

with the universal probe EUB338, showed the absence of any<br />

other kind of endosymbiotic bacteria (Fig. 2), with the<br />

exception of gut bacteria inside adult and larval specimens<br />

of Deronectes. FISH accomplished with EUB388 II and<br />

EUB338 III did not show any positive signals (data not<br />

shown).<br />

Electron microscopy of the Rickettsia symbiont<br />

Ultrastructural examinations of male accessory glands of<br />

D. platynotus revealed coccobacillary rickettsiae that were<br />

observed free in the cytoplasm (Fig. 3a). Most rickettsiae<br />

ranged from 0.35 mm in width to 0.65 mm in length. Rickettsiae<br />

were delineated by an inner periplasmatic membrane,<br />

a periplasmatic space and a trilaminar cell wall with a thicker<br />

inner leaflet, characteristic of the genus Rickettsia. ARickettsia-typical<br />

slime layer or an electron-translucent zone, up<br />

to 60 nm thick, surrounded the cell wall and separated the<br />

bacterium from the cytoplasm of the host cell (Fig. 3b).<br />

Isolated bacteria were also found in the musculature,<br />

enclosing the male accessory gland (Fig. 3c). Some of the<br />

symbionts were much longer and had a ‘filamentous’<br />

appearance. They could be as long as 2.74 mm (Fig. 3d).<br />

Phylogenetic analysis of the Rickettsia symbiont<br />

Phylogenetic analyses on the basis of the 16S rRNA gene<br />

sequence revealed that Rickettsia sp. from Deronectes belong<br />

neither to the ‘spotted fever group’ nor the ‘typhus group,’<br />

but was placed within the ancestral group (100% bootstrap<br />

support) (Fig. 4). Clones of the endosymbiotic bacteria of<br />

D. platynotus, D. aubei and D. delarouzei were placed in the<br />

clade with R. limoniae, Rickettsia endosymbionts of Cerobasis<br />

guestifalica (Psocoptera, Trogiidae) and Lutzomyia apache<br />

(Diptera, Psychodinae), whereas Rickettsia sp. of D. semirufus<br />

cluster basal with rickettsiae from leeches. By phylogenetic<br />

analysis based on the gltA gene (Fig. 5), the Rickettsia strain<br />

from Deronectes clustered with Rickettsia endosymbionts of<br />

L. apache and different spider species in a terminal group<br />

(99% bootstrap support) that is distinctly separated from<br />

other Rickettsia species of the spotted fever, typhus and bellii<br />

groups. The phylogenetic position of Deronectes rickettsiae<br />

was the same in the 16S rRNA gene and gltA trees when the<br />

maximum-likelihood, parsimony and distance methods<br />

were used (data not shown). Only the position of the<br />

Rickettsia endosymbiont of D. semirufus varied with different<br />

methods (MrBayes, data not shown), and was sometimes<br />

added to the leech group or to the other rickettsial<br />

endosymbionts of the Deronectes species of the limoniae<br />

group. This explains the low bootstrap support values<br />

in the ancestral group in the phylogenetic tree of 16S rRNA<br />

gene.<br />

Discussion<br />

The present study reports the first molecular identification<br />

of Rickettsia in four water beetle species of the genus<br />

Deronectes and generally for adephagan beetles. Evaluable<br />

results have been obtained of D. platynotus and D. aubei.<br />

Rickettsia infection was detected in 100% (45/45) of<br />

D. platynotus and 39.4% (28/71) of D. aubei. The frequencies<br />

of Rickettsia infection were maintained constant in<br />

different seasons. Individuals of D. delarouzei and D. semirufus<br />

were also tested positive for rickettsiae, whereas<br />

individuals of D. latus, D. aubei sanfilippoi and D. moestus<br />

inconspectus did not show any signal for symbiotic bacteria.<br />

Remarkably, Rickettsia-positive species of Deronectes do not<br />

cluster in a single group within the phylogenetic tree of<br />

Deronectes. In addition, the current analysis indicated six<br />

more species of Agabus and Hydroporus (Dytiscidae) that are<br />

FEMS Microbiol Ecol 68 (2009) 201–211<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved


206 S.M. Küchler et al.<br />

Fig. 3. Electron photomicrographs of male<br />

accessory gland tissue from water beetle<br />

Deronectes platynotus infected by intracytosolic<br />

Rickettsia sp. (a) Coccobacillary rickettsiae<br />

resided free in the cytoplasm of accessory gland<br />

tissue (asterisk), located near connective tissue<br />

(arrow), which close about the glandular tissue.<br />

Scale bar = 1.1 mm. (b) High magnification of<br />

Rickettsia sp. showing the typical rickettsial<br />

morphology, with a periplasmic membrane<br />

(white arrowhead), an electron-lucent<br />

periplasmic space (black arrowhead), the<br />

trilaminar cell wall (black arrow) and with an<br />

electron-lucent ‘halozone’ (asterisk). Scale<br />

bar = 0.09 mm. (c) Rickettsia within accessory<br />

gland musculature. Scale bar = 0.6 mm.<br />

(d) A filamentous Rickettsia (asterisk), possessing<br />

a length of 2.74 mm, before extending out of the<br />

plane of section (arrow). Scale bar = 0.4 mm.<br />

associated with Rickettsia sp. too. To our knowledge, only six<br />

verifications of Rickettsia have been documented in Coleoptera<br />

so far (Coccinellidae: Werren et al., 1994; von der<br />

Schulenburg et al., 2001; Bruchidae: Fukatsu et al., 2000;<br />

Buprestidae: Lawson et al., 2001; Curculionidae: Zchori-Fein<br />

et al., 2006; and Mordellidae: Duron et al., 2008). All<br />

rickettsial isolates from beetles, with the exception of<br />

Mordellistena sp., Adalia bipunctata and Adalia decempunctata,<br />

have been allocated to the rickettsial ancestral group,<br />

which includes all members of a clade that are most basal in<br />

the genus Rickettsia. Interestingly, the ancestral group comprised<br />

various symbionts from aquatic hosts, such as leeches<br />

(Kikuchi et al., 2002; Kikuchi & Fukatsu, 2005), amphizoic<br />

amoeba N. pattersoni (Dykova et al., 2003) and marine<br />

ciliate D. appendiculata (Vannini et al., 2005). Even insects<br />

that spend most of their development in water, i.e. the<br />

cranefly L. chorea (Diptera, Limoniidae; AF322443), mothfly<br />

L. apache (Diptera, Psychodidae; EU223247) and biting<br />

midge Culicoides sonorensis (Diptera, Ceratopogonidae)<br />

(Campbell et al., 2004), have been shown to be infected with<br />

rickettsiae. The exact classification of these rickettsial isolates<br />

that are allocated to the ancestral group is still unclear.<br />

Only 16S rRNA gene sequences are available for most<br />

members of the ancestral group. However, for detailed<br />

phylogenetic analysis within the genus Rickettsia, 16S rRNA<br />

gene sequences are not convenient because of their conservatism<br />

(Roux et al., 1997). Nevertheless, rickettsial 16S<br />

rRNA clone sequences from D. platynotus are different<br />

in two to six positions from each other, mainly in two<br />

variable regions. In all sequenced clones, the sequence<br />

similarity ranges from 99.1% to 100%. Possible reasons<br />

for the microdiversity of rickettsial 16S rRNA gene clones<br />

in D. platynotus may be the presence of multiple different<br />

copies of the 16S rRNA gene in one genome, which has<br />

been demonstrated for numerous species of bacteria (Cilia<br />

et al., 1996).<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved<br />

FEMS Microbiol Ecol 68 (2009) 201–211


Rickettsia endosymbiont in water beetles<br />

207<br />

Fig. 4. Phylogenetic position of different clones of the endosymbiotic bacteria of water beetles Deronectes platynotus, Deronectes aubei,<br />

Deronectes semirufus and Deronectes delarouzei (bold) based on 55 16S rRNA gene sequences (1350 bp, maximum parsimony heuristic search, 50%<br />

majority-rule consensus tree of 840 most parsimonious trees, tree length = 497, CI = 0.71, RI = 0.86). The tree has been rooted with Orientia<br />

tsutsugamushi as an outgroup. Only bootstrap values of at least 50% are shown at nodes (n = 500).<br />

FEMS Microbiol Ecol 68 (2009) 201–211<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved


208 S.M. Küchler et al.<br />

Fig. 5. Phylogenetic position of the endosymbiotic bacteria of water beetles Deronectes platynotus, Deronectes aubei, Deronectes semirufus and<br />

Deronectes delarouzei (bold) based on 46 gltA gene sequences (275 bp, maximum parsimony heuristic search, 50% majority-rule consensus tree of 30<br />

most parsimonious trees, tree length = 459, CI = 0.55, RI = 0.88). The tree has been rooted with Rhodopirellula baltica as an outgroup. Only bootstrap<br />

values of at least 50% are shown at nodes (n = 500).<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved<br />

FEMS Microbiol Ecol 68 (2009) 201–211


Rickettsia endosymbiont in water beetles<br />

209<br />

Referred to rickettsiae of the ancestral group, the citrate<br />

synthase-coding gltA gene sequences, also used for phylogenetic<br />

analysis, are only available for Rickettsia sp. of L. apache<br />

(EU368001) and rickettsial isolates from spiders (Goodacre<br />

et al., 2006). Sequences of further protein-coding genes,<br />

mostly those belonging to the surface cell antigen (sca)<br />

family, which have been used for the classification of<br />

rickettsial isolates over the past few years (Fournier et al.,<br />

1998; Roux & Raoult, 2000; Sekeyova et al., 2001; Ngwamidiba<br />

et al., 2006), are not available for rickettsiae of the<br />

ancestral group (Fournier et al., 2003).<br />

Phylogenetic analysis based on the 16S rRNA gene<br />

showed that Rickettsia sp. from Deronectes cluster into the<br />

‘limoniae group’, the basal subgroup of the ancestral group.<br />

Furthermore, we succeeded in partial sequencing of the gltA<br />

gene of these rickettsiae. Sequences from Deronectes symbionts<br />

resulted in a high concordance with rickettsial<br />

isolates from spiders P. phrygianus and M. mengei (Goodacre<br />

et al., 2006). However, 16S rRNA gene sequences are not<br />

available for spider rickettsiae in GenBank. It is interesting<br />

that sequences of the gltA gene of Rickettsia sp. isolated from<br />

water beetles are quite divergent from all previous valid<br />

Rickettsia strains from the spotted fever, typhus and bellii<br />

groups. Furthermore, gltA sequences of Rickettsia sp. show a<br />

high similarity to the related genus Bartonella, whose<br />

members are known as emerging human pathogens (Minnick<br />

& Anderson, 2006), but do not cluster with Bartonella<br />

sequences. Goodacre et al. (2006) presumed that spiderspecific<br />

Rickettsia form an isolated, terminal group that<br />

contains no symbionts of other arthropod hosts. They<br />

further assumed that frequent horizontal transmission of<br />

rickettsiae between spiders and their insect prey is unlikely<br />

because this would produce a heterogeneous assortment of<br />

non-spider-specific Rickettsia in spiders. In contrast, we were<br />

able to connect molecular data of the 16S rRNA gene<br />

sequences from basal rickettsiae strains and spider-specific<br />

Rickettsia by sequencing of 16S rRNA and gltA genes of the<br />

Deronectes symbiont. We assume that other Rickettsia strains<br />

of the ancestral group could exhibit comparable gltA<br />

sequences. Based on this assumption, Rickettsia from water<br />

beetles and spiders could belong to a new taxon, which might<br />

represent a near relative of the genus Rickettsia. Itcanbe<br />

speculated that sequences of the ompA, ompB genes and gene<br />

D will show appreciable deviations like gltA, if these genes are<br />

present in rickettsiae of the ancestral group in general.<br />

However, ultrastructural observations revealed typical<br />

morphological features and intracellular arrangements of<br />

rickettsiae, such as a thickened inner leaflet or electronlucent<br />

‘halos’ (Silverman, 1991). Additionally, some rickettsiae<br />

had a ‘filamentous’ appearance, similar to R. bellii<br />

from Amblyomma cooperi ticks (Labruna et al., 2004). The<br />

actual length of these bacteria cells could not be revealed by<br />

transmission electron microscopical sections.<br />

Furthermore, Rickettsia could be detected in all tissues of<br />

all investigated Rickettsia-positive Deronectes species. Presumably,<br />

symbionts have been transported by the hemolymph,<br />

comparable to observations in aphids (Chen et al.,<br />

1996). This would support their appearance in the eyes and<br />

legs. However, there is no evidence of harboring of Rickettsia<br />

in specific sheath cells, secondary mycetocytes or mycetomelike<br />

structures in aphids or booklice (Sakurai et al., 2005;<br />

Perotti et al., 2006). Likewise, Rickettsia could not be<br />

detected in high densities close to midgut epithelia or<br />

malpighian tubules (Perotti et al., 2006). In contrast, the<br />

internal genitals and fatbody exhibited numerous symbionts.<br />

Rickettsiae depend on precursors of amino acids<br />

from their host. Equally, the de novo synthesis of purines as<br />

well as pyrimidine is not achieved by rickettsiae (Renesto<br />

et al., 2005). Hence, tissue of high metabolic activity, such as<br />

the fatbody, could be preferably infested with Rickettsia. The<br />

presence of bacteria within ovariols and eggs also indicates<br />

strong evidence for transovarial transmission, which must<br />

contribute to the maintenance of Rickettsia sp. infection in<br />

Deronectes populations (Whitworth et al., 2003). This is<br />

supported by the existence of Rickettsia sp. in the second and<br />

third larval stages of Deronectes, whereas the quantity of<br />

symbionts increased gradually in each stage.<br />

Thus, the amount of bacteria seems to be determined by<br />

several factors: (1) fidelity of vertical transmission, (2)<br />

frequency of horizontal transmission (Fukatsu et al., 2000),<br />

(3) fitness effect on the host and (4) selfish manipulation of<br />

the host reproduction (Tsuchida et al., 2002). Moreover,<br />

interactions between the endosymbionts and their hosts are<br />

complicated due to subpopulation structures and fluctuating<br />

environmental conditions. The resulting bottleneck<br />

during vertical transmission could be a reason for the<br />

fluctuated rate of rickettsial affection in D. aubei. More<br />

individuals of D. delarouzei and D. semirufus have to be<br />

investigated to verify the infection rates. Because of the<br />

predatory nature of Deronectes (Dettner et al., 1986), a<br />

horizontal transmission of Rickettsia sp. seems to be possible.<br />

Current research should demonstrate whether aquatic<br />

prey of Deronectes is also associated with Rickettsia sp.<br />

On the basis of the present results, there are no indications<br />

that Rickettsia sp. affection has an effect on the fitness<br />

of Deronectes. Neither reduced body weight and fecundity<br />

like in Rickettsia-infected host aphids (Sakurai et al., 2005)<br />

nor remarkable increases in host size as observed in leeches<br />

(Kikuchi & Fukatsu, 2005) could be observed. However, the<br />

research of fitness effects is quite difficult, because Deronectes<br />

cannot be bred under laboratory conditions.<br />

Parasitic living bacteria, such as Rickettsia, Spiroplasma,<br />

Cardinium and especially Wolbachia, have a tendency to<br />

manipulate host reproduction for their own benefit. Reproductive<br />

phenomena such as parthenogenesis, cytoplasmatic<br />

incompatibility, feminization and male killing have been<br />

FEMS Microbiol Ecol 68 (2009) 201–211<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved


210 S.M. Küchler et al.<br />

documented (O’Neill et al., 1997). Male killing caused by<br />

Rickettsia was described until now in ladybird beetles A.<br />

bipunctata and A. decempunctata (Werren et al., 1994; von<br />

der Schulenburg et al., 2001) and buprestid beetle Brachys<br />

tessellates (Lawson et al., 2001). Probably, rickettsiae also<br />

influence oogenesis in bark beetle Coccotrypes dactyliperda<br />

(Zchori-Fein et al., 2006). Rickettsia-induced parthenogenesis<br />

was documented from eulophid parasitoid Neochrysocharis<br />

formosa (Hagimori et al., 2006). In contrast, all investigated<br />

species of Deronectes indicated a balanced sex ratio.<br />

In this context, further research is required to investigate<br />

in detail more about the influence of Rickettsia on Deronectes<br />

species, especially at the biochemical level. Furthermore,<br />

transmission and maintenance of Rickettsia in different<br />

populations of Deronectes species must be examined. Future<br />

studies will show, whether other Dytiscidae as well as their<br />

aquatic prey are also infected by Rickettsia sp. symbiont.<br />

Acknowledgements<br />

We thank M. Meyering-Vos, G. Rambold and D. Peršoh for<br />

providing equipment and assistance in cloning, sequencing<br />

and phylogenetic analysis. S. Geimer and R. Grotjahn are<br />

gratefully acknowledged for assistance in electron microscopy<br />

analysis, as well as S. Heidmann for the opportunity<br />

to use the confocal microscope, and for providing help. We<br />

also thank A. Kirpal for technical assistance.<br />

References<br />

Amann RI, Ludwig W & Schleifer KH (1995) Phylogenetic<br />

identification and in situ detection of individual microbial cells<br />

without cultivation. Microbiol Rev 59: 143–169.<br />

Campbell CL, Mummey DL, Schmidtmann ET & Wilson WC<br />

(2004) Culture-independent analysis of midgut microbiota in<br />

the arbovirus vector Culicoides sonorensis (Diptera:<br />

Ceratopogonidae). J Med Entomol 41: 340–348.<br />

Chen DQ, Campbell BC & Purcell AH (1996) A new rickettsia<br />

from a herbivorous insect, the pea aphid Acyrthosiphon pisum<br />

(Harris). Curr Microbiol 33: 123–128.<br />

Cilia V, Lafay B & Christen R (1996) Sequence heterogeneities<br />

among 16S ribosomal RNA sequences, and their effect on<br />

phylogenetic analyses at the species level. Mol Biol Evol 13:<br />

451–461.<br />

Daims H, Brühl A, Amann R, Schleifer K & Wagner M (1999) The<br />

domain-specific probe EUB338 is insufficient for the detection<br />

of all Bacteria: development and evaluation of a more<br />

comprehensive probe set. Syst Appl Microbiol 22: 434–444.<br />

Davis MJ, Ying Z, Brunner BR, Pantoja A & Ferwerda FH (1998)<br />

Rickettsial relative associated with papaya bunchy top disease.<br />

Curr Microbiol 36: 80–84.<br />

Dettner K, Hübner M & Classen R (1986) Age structure,<br />

phenology and prey of some rheophilic Dytiscidae<br />

(Coleoptera). Ent Basil 11: 343–370.<br />

Duron O, Bouchon D, Boutin S, Bellamy L, Zhou L, Engelstädter<br />

J & Hurst G (2008) The diversity of reproductive parasites<br />

among arthropodes: Wolbachia do not walk alone. BMC Biol<br />

6: 27.<br />

Dykova I, Veverkova M, Fiala I, Machackova B & Peckova H<br />

(2003) Nuclearia pattersoni sp. n. (Filosea), a new species of<br />

amphizoic amoeba isolated from gills of roach (Rutilus<br />

rutilus), and its rickettsial endosymbiont. Folia Parasit 50:<br />

161–170.<br />

Eremeeva ME & Dasch GA (2000) Rickettsiae. Encyclopedia of<br />

Microbiology, Vol. 4 (Lederberg J, ed.), pp. 140–179. Academic<br />

Press, New York.<br />

Fery H & Brancucci M (1997) A taxonomic revision of Deronectes<br />

(Sharp), 1882 (Insecta: Coleoptera: Dytiscidae) (partI). Ann<br />

Naturhist Mus Wien 99B: 17–302.<br />

Fery H, Köksal Erman Ö & Hosseinie S (2001) Two new<br />

Deronectes (Sharp), 1882 (Insecta: Coleoptera: Dytiscidae) and<br />

notes on other species of the genus. Ann Naturhist Mus Wien<br />

103B: 341–351.<br />

Fournier P & Raoult D (2007) Identification of rickettsial isolates<br />

at the species level using multi-spacer typing. BMC Microbiol<br />

7: 72.<br />

Fournier PE, Roux V & Raoult D (1998) Phylogenetic analysis of<br />

spotted fever group rickettsiae by study of the outer surface<br />

protein rOmpA. Int J Syst Bacteriol 48: 839–849.<br />

Fournier PE, Dumler JS, Greub G, Zhang J, Wu Y & Raoult D<br />

(2003) Gene sequence-based criteria for identification of new<br />

rickettsia isolates and description of Rickettsia heilongjiangensis<br />

sp. nov. J Clin Microbiol 41: 5456–5465.<br />

Fukatsu T, Nikoh N, Kawai R & Koga R (2000) The secondary<br />

endosymbiotic bacterium of the pea aphid Acyrthosiphon<br />

pisum (Insecta: Homoptera). Appl Environ Microbiol 66:<br />

2748–2758.<br />

Gonzalez J, Zimmermann J & Saiz-Jimenez C (2005) Evaluating<br />

putative chimeric sequences from PCR-amplified products.<br />

Bioinformatics 21: 333–337.<br />

Goodacre SL, Martin OY, Thomas CF & Hewitt GM (2006)<br />

Wolbachia and other endosymbiont infections in spiders. Mol<br />

Ecol 15: 517–527.<br />

Gottlieb Y, Ghanim M, Chiel E et al. (2006) Identification and<br />

localization of a Rickettsia sp. in Bemisia tabaci (Homoptera:<br />

Aleyrodidae). Appl Environ Microbiol 72: 3646–3652.<br />

Hagimori T, Abe Y, Date S & Miura K (2006) The first finding of a<br />

Rickettsia bacterium associated with parthenogenesis<br />

induction among insects. Curr Microbiol 52: 97–101.<br />

Hall TA (1999) BioEdit. A user-friendly biological sequence<br />

alignment editor analysis program for Windows 95/98/NT.<br />

Nucl Acids Symp Ser 41: 95–98.<br />

Hoy MA & Jeyaprakash A (2005) Microbial diversity in the<br />

predatory mite Metaseiulus occidentalis (Acari: Phytoseiidae)<br />

and its prey, Tetranychus urticae (Acari: Tetranychidae). Biol<br />

Control 32: 427–441.<br />

Kikuchi Y & Fukatsu T (2005) Rickettsia infection in natural leech<br />

populations. Microb Ecol 49: 265–271.<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved<br />

FEMS Microbiol Ecol 68 (2009) 201–211


Rickettsia endosymbiont in water beetles<br />

211<br />

Kikuchi Y, Sameshima S, Kitade O, Kojima J & Fukatsu T (2002)<br />

Novel clade of Rickettsia spp. from leeches. Appl Environ<br />

Microbiol 68: 999–1004.<br />

Labruna MB, Whitworth T, Horta MC, Bouyer DH, McBride JW,<br />

Pinter A, Popov V, Gennari SM & Walker DH (2004) Rickettsia<br />

species infecting Amblyomma cooperi ticks from an area in the<br />

state of Sao Paulo, Brazil, where Brazilian spotted fever is<br />

endemic. J Clin Microbiol 42: 90–98.<br />

Lane DJ (1991) 16S and 23S rRNA sequencing. Nucleic Acid<br />

Techniques in Bacterial Systematics (Stackenbrandt E &<br />

Goodfellow M, eds), pp. 115–148. John <strong>Wiley</strong> and Sons,<br />

New York.<br />

Lawson ET, Mousseau TA, Klaper R, Hunter MD & Werren JH<br />

(2001) Rickettsia associated with male-killing in a buprestid<br />

beetle. Heredity 86: 497–505.<br />

Ludwig W, Strunk O, Westram R et al. (2004) ARB: a software<br />

environment for sequence data. Nucleic Acids Res 32:<br />

1363–1371.<br />

Maddison D, Baker M & Ober K (1999) Phylogeny of Carabid<br />

beetles as inferred from 18S ribosomal DNA (Coleoptera:<br />

Carabidae). Syst Entomol 24: 1–36.<br />

Manz W, Amann R, Ludwig W, Wagner M & Schleifer K-H<br />

(1992) Phylogenetic oligodeoxynucleotide probes for the<br />

major subclasses of proteobacteria: problems and solutions.<br />

Syst Appl Microbiol 15: 593–600.<br />

Minnick MF & Anderson BE (2006) The genus Bartonella.<br />

Prokaryotes, Vol. 5 (Dworkin M, Falkow S, Rosenberg E,<br />

Schleifer K-H & Stackebrandt E, eds), pp. 467–492. Springer,<br />

New York.<br />

Ngwamidiba M, Blanc G, Raoult D & Fournier PE (2006) Sca1, a<br />

previously undescribed paralog from autotransporter proteinencoding<br />

genes in Rickettsia species. BMC Microbiol 6: 12.<br />

O’Neill SL, Hoffmann AA & Werren JH (1997) Influential<br />

Passengers: Inherited Microorganisms and Arthropod<br />

Reproduction. Oxford University Press, Oxford.<br />

Perlman SJ, Hunter MS & Zchori-Fein E (2006) The emerging<br />

diversity of Rickettsia. Proc Biol Sci 273: 2097–2106.<br />

Perotti MA, Clarke HK, Turner BD & Braig HR (2006) Rickettsia<br />

as obligate and mycetomic bacteria. FASEB J 20: 2372–2374.<br />

Renesto P, Ogata H, Audic S, Claverie JM & Raoult D (2005)<br />

Some lessons from Rickettsia genomics. FEMS Microbiol Rev<br />

29: 99–117.<br />

Roux V & Raoult D (2000) Phylogenetic analysis of members of<br />

the genus Rickettsia using the gene encoding the outermembrane<br />

protein rOmpB (ompB). Int J Syst Evol Microbiol<br />

50: 1449–1455.<br />

Roux V, Rydkina E, Eremeeva M & Raoult D (1997) Citrate<br />

synthase gene comparison, a new tool for phylogenetic<br />

analysis, and its application for the rickettsiae. Int J Syst<br />

Bacteriol 47: 252–261.<br />

Sakurai M, Koga R, Tsuchida T, Meng XY & Fukatsu T (2005)<br />

Rickettsia symbiont in the pea aphid Acyrthosiphon pisum:<br />

novel cellular tropism, effect on host fitness, and interaction<br />

with the essential symbiont Buchnera. Appl Environ Microbiol<br />

71: 4069–4075.<br />

Schwoerbel J (1994) Methoden der Hydrobiologie,<br />

Süßwasserbiologie. Gustav Fischer, Stuttgart.<br />

Sekeyova Z, Roux V & Raoult D (2001) Phylogeny of Rickettsia<br />

spp. inferred by comparing sequences of ‘gene D’, which<br />

encodes an intracytoplasmic protein. Int J Syst Evol Microbiol<br />

51: 1353–1360.<br />

Shull VL, Vogler AP, Baker MD, Maddison DR & Hammond PM<br />

(2001) Sequence alignment of 18S ribosomal RNA and the<br />

basal relationships of Adephagan beetles: evidence for<br />

monophyly of aquatic families and the placement of<br />

Trachypachidae. Syst Biol 50: 945–969.<br />

Silverman DJ (1991) Some contributions of electron microscopy<br />

to the study of rickettsiae. Eur J Epidemiol 7: 200–206.<br />

Stothard DR & Fuerst PA (1995) Evolutionary analysis of the<br />

spotted fever and typhus groups of Rickettsia using 16 rRNA<br />

gene sequences. Syst Appl Microbiol 18: 52–61.<br />

Swofford DL (2000) PAUP . Phylogenetic Analysis Using<br />

Parsimony ( and Other Methods). Sinauer Associates,<br />

Sunderland, MA.<br />

Tsuchida T, Koga R, Shibao H, Matsumoto T & Fukatsu T (2002)<br />

Diversity and geographic distribution of secondary<br />

endosymbiotic bacteria in natural populations of the pea<br />

aphid, Acyrthosiphon pisum. Mol Ecol 11: 2123–2135.<br />

Vannini C, Petroni G, Verni F & Rosati G (2005) A bacterium<br />

belonging to the Rickettsiaceae family inhabits the cytoplasm<br />

of the marine ciliate Diophrys appendiculata (Ciliophora,<br />

Hypotrichia). Microb Ecol 49: 434–442.<br />

von der Schulenburg JH, Habig M, Sloggett JJ, Webberley KM,<br />

Bertrand D, Hurst GD & Majerus ME (2001) Incidence of<br />

male-killing Rickettsia spp. (alpha-proteobacteria) in the tenspot<br />

ladybird beetle Adalia decempunctata L. (Coleoptera:<br />

Coccinellidae). Appl Environ Microbiol 67: 270–277.<br />

Weiss E & Moulder JW (1984) Order I. Rickettsiales<br />

Gieszczkiewicz 1939, 25AL. Bergey’s Manual of Systematic<br />

Bacteriology, Vol. 1 (Krieg NR & Holt JG, eds), pp. 687–701.<br />

Williams and Wilkins, Baltimore, MD.<br />

Werren JH, Hurst GD, Zhang W, Breeuwer JA, Stouthamer R &<br />

Majerus ME (1994) Rickettsial relative associated with male<br />

killing in the ladybird beetle (Adalia bipunctata). J Bacteriol<br />

176: 388–394.<br />

Whitworth T, Popov V, Han V, Bouyer D, Stenos J, Graves S, Ndip<br />

L & Walker D (2003) Ultrastructural and genetic evidence of a<br />

reptilian tick, Aponomma hydrosauri, as a host of Rickettsia<br />

honei in Australia: possible transovarial transmission. Ann NY<br />

Acad Sci 990: 67–74.<br />

Yu X-J & Walker DH (2006) The order Rickettsiales.<br />

Prokaryotes, Vol. 5 (Dworkin M, Falkow S, Rosenberg E,<br />

Schleifer K-H & Stackebrandt E, eds), pp. 493–528. Springer,<br />

New York.<br />

Zchori-Fein E, Borad C & Harari A (2006) Oogenesis in the date<br />

stone beetle, Coccotrypes dactyliperda, depends on symbiotic<br />

bacteria. Phy Entomol 31: 164–169.<br />

FEMS Microbiol Ecol 68 (2009) 201–211<br />

c 2009 Federation of European Microbiological Societies<br />

Published by Blackwell Publishing Ltd. All rights reserved

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!