首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 250 毫秒
1.
Symbiotic associations with midgut bacteria have been commonly found in diverse phytophagous heteropteran groups, where microbiological characterization of the symbiotic bacteria has been restricted to the stinkbug families Acanthosomatidae, Plataspidae, Pentatomidae, Alydidae, and Pyrrhocoridae. Here we investigated the midgut bacterial symbiont of Cantao ocellatus, a stinkbug of the family Scutelleridae. A specific gammaproteobacterium was consistently identified from the insects of different geographic origins. The bacterium was detected in all 116 insects collected from 9 natural host populations. Phylogenetic analyses revealed that the bacterium constitutes a distinct lineage in the Gammaproteobacteria, not closely related to gut symbionts of other stinkbugs. Diagnostic PCR and in situ hybridization demonstrated that the bacterium is extracellularly located in the midgut 4th section with crypts. Electron microscopy of the crypts revealed a peculiar histological configuration at the host-symbiont interface. Egg sterilization experiments confirmed that the bacterium is vertically transmitted to stinkbug nymphs via egg surface contamination. In addition to the gut symbiont, some individuals of C. ocellatus harbored another bacterial symbiont in their gonads, which was closely related to Sodalis glossinidius, the secondary endosymbiont of tsetse flies. Biological aspects of the primary gut symbiont and the secondary Sodalis-allied symbiont are discussed.Insects are among the largest animal groups on the earth, embracing 750,000 to several millions of species (37, 52). Diverse insects are symbiotically associated with microorganisms, especially bacteria (5-7). In some insects, symbiotic bacteria are harbored in specialized host cells called bacteriocytes (or mycetocytes), constituting obligate mutualistic associations. For example, Buchnera aphidicola is harbored within bacteriocytes in the abdominal body cavity of almost all aphids and provides essential amino acids that are lacking in the phloem sap diet of the insects (9, 47). Wigglesworthia glossinidia is localized in a midgut-associated bacteriome of tsetse flies and plays pivotal roles in biosynthesis of B vitamins that are deficient in the vertebrate blood diet of the insects (2, 34). These obligate endocellular symbionts are often collectively referred to as “primary symbionts.”In contrast, there are facultative endosymbiotic microorganisms not essential for their host insects, often collectively called “secondary symbionts.” For example, many aphids are known to harbor various facultative symbionts, which belong to distinct lineages in the Gamma- and Alphaproteobacteria (33, 43) and the Mollicutes (10). While the majority of those facultative bacteria are either parasitic or commensalistic for their hosts, some of them affect the host fitness beneficially in particular ecological contexts (29, 32, 36, 44, 51). In addition to the obligate primary symbiont Wigglesworthia, tsetse flies harbor the facultative secondary symbiont Sodalis glossinidius, whose biological function for the hosts is currently elusive (3, 8).Members of the suborder Heteroptera, known as true bugs and consisting of over 38,000 described species, are characterized by their sucking mouthparts, half-membranous forewings, and incomplete metamorphosis (46). In the Heteroptera, symbiotic associations with bacteria are mainly found in phytophagous groups, especially in stinkbugs of the infraorder Pentatomomorpha. These stinkbugs generally possess many sacs or tubular outgrowths, called crypts or ceca, in a posterior region of the midgut, whose lumen is densely populated by a specific bacterial symbiont (7, 16). In some cases, experimental elimination of the symbiotic bacteria resulted in retarded growth and high mortality of the host insects (1, 13, 21, 26, 27, 39), indicating that these gut symbionts play important biological roles. Most of the gut symbionts are vertically transmitted through host generations by such mechanisms as egg surface contamination in the families Pentatomidae and Acanthosomatidae (1, 27, 39, 40, 42), coprophagy in the Cydnidae and Coreidae (22, 45), and capsule transmission in the Plataspidae (20), whereas a case of environmental acquisition has been reported from the Alydidae (26). Thus far, gut symbiotic bacteria of some members of the Acanthosomatidae, Plataspidae, Pentatomidae, Alydidae, and Pyrrhocoridae have been characterized using molecular techniques (21, 23, 25, 27, 38), while phylogenetic and biological aspects of gut symbiotic bacteria have been untouched in many other stinkbug groups.These gut symbiotic bacteria are, despite their extracellular localization, regarded as “primary symbionts” of the stinkbugs. On the other hand, some stinkbugs may, in addition to the gut symbiotic bacteria, also be associated with facultative “secondary symbionts.” For example, Wolbachia infections have been detected from diverse stinkbugs, most of which are probably of parasitic or commensalistic nature (24). Besides Wolbachia, there has been no report on facultative, secondary symbionts from stinkbugs.Members of the family Scutelleridae, often referred to as jewel bugs or shield-backed bugs, are stinkbugs characterized by their greatly enlarged convex scutellum that usually covers the entire abdomen. Some tropical species are also known for their vivid and beautiful body coloration (46). The family contains approximately 80 genera and 450 species, and in Japan, at least 7 genera and 9 species have been recorded (50). In the early 20th century, the presence of symbiotic bacteria was histologically described in midgut crypts of several scutellerid species (16, 31, 42). Since these pioneer works, however, no studies have been conducted on the symbiotic bacteria of scutellerid stinkbugs.Here we investigated the midgut symbiont of Cantao ocellatus, a scutellerid stinkbug widely distributed in Asian countries, including Japan, and known to guard their eggs and newborn nymphs (Fig. (Fig.1A)1A) (50). In addition to the gut symbiont, we also identified a Sodalis-allied facultative secondary symbiont from gonads of the insect.Open in a separate windowFIG. 1.(A) Adult female of Cantao ocellatus, guarding hatchlings under her body. (B) Dissected midgut from an adult female of C. ocellatus. 1st, midgut 1st section; 2nd, midgut 2nd section; 3rd, midgut 3rd section; 4th, midgut 4th section with crypts; hg, hindgut. (C) Enlarged image of the midgut 4th section with crypts. Arrowheads indicate three rows of crypts, while a fourth row is hidden behind. Glandular crypts (gc) are developed in adult females specifically, which may be involved in egg surface contamination with the symbiont. (D) An in situ hybridization image of the midgut 4th section, in which red and green signals indicate the gut symbiont and the host nucleus, respectively. Each arrow shows a crypt. (E) An enlarged image of the symbiotic bacteria in the crypts.  相似文献   

2.
Many aphids harbor a variety of endosymbiotic bacteria. The functions of these symbionts can range from an obligate nutritional role to a facultative role in protecting their hosts against environmental stresses. One such symbiont is “Candidatus Serratia symbiotica,” which is involved in defense against heat and potentially also in aphid nutrition. Lachnid aphids have been the focus of several recent studies investigating the transition of this symbiont from a facultative symbiont to an obligate symbiont. In a phylogenetic analysis of Serratia symbionts from 51 lachnid hosts, we found that diversity in symbiont morphology, distribution, and function is due to multiple independent origins of symbiosis from ancestors belonging to Serratia and possibly also to evolution within distinct symbiont clades. Our results do not support cocladogenesis of “Ca. Serratia symbiotica” with Cinara subgenus Cinara species and weigh against an obligate nutritional role. Finally, we show that species belonging to the subfamily Lachninae have a high incidence of facultative symbiont infection.Many insect species harbor heritable endosymbiotic bacteria. Among the best studied of these species are aphids. Almost all aphids are infected with the obligate nutritional symbiont Buchnera aphidicola, which is generally required for the survival of aphids and provides essential amino acids that are rare in their phloem sap diet (32). Many aphids also possess additional symbionts that may be facultative from the host''s perspective and that coexist with Buchnera (20).Three lineages of facultative symbionts that are prevalent in aphids belong to the Enterobacteriaceae. Two of these lineages (“Candidatus Hamiltonella defensa” and “Candidatus Regiella insecticola”) form well-defined clades distinct from free-living bacterial species (4, 20) and confer clear advantages to their hosts by protecting them against natural enemies. “Ca. Hamiltonella defensa” prevents wasp parasitism by arresting development of wasp larvae in pea aphids, and “Ca. Regiella insecticola” provides resistance against the fungal pathogen Pandora neoaphidis (24, 31). The third lineage, “Candidatus Serratia symbiotica,” is closely related to free-living members of the genus Serratia. This symbiont is distributed sporadically among aphid species and has been proposed to have a variety of effects on hosts. In pea aphids (Acyrthosiphon pisum; Macrosiphini), “Ca. Serratia symbiotica” ameliorates the deleterious fitness effects of heat shock by protecting symbiont-harboring bacteriocyte cells (2, 19, 29). Additionally, a strain of “Ca. Serratia symbiotica” provided some resistance to parasitoid wasp attack (24). “Ca. Serratia symbiotica” has been proposed to play a role in nutrition by producing amino acids for its aphid host and by decreasing its host''s reliance on Buchnera (10, 15, 16, 26). In contrast to most Buchnera strains, Buchnera strains from Cinara cedri (Lachnini) have lost the genes for biosynthesis of the essential amino acid tryptophan, while “Ca. Serratia symbiotica” in the same host possesses at least part of the pathway, suggesting that it has a mutualistic role in the nutrition of aphids (26).In A. pisum, “Ca. Serratia symbiotica” cells are rod-shaped bacteria that are present in the sheath cells, hemolymph, and bacteriocytes of some individuals. In contrast, in C. cedriCa. Serratia symbiotica” occurs in all individuals, and its cells are large, round, and pleomorphic, similar to the cells of many obligate bacterial aphid endosymbionts, including Buchnera (10, 26). Furthermore, “Ca. Serratia symbiotica” has consistently been present in other Cinara species sampled (28). Both the rod-shaped and pleomorphic forms are assigned to “Ca. Serratia symbiotica” based on phylogenetic analyses of several gene sequences, but they fall into two distinct sister clades of symbiont lineages that seem to coincide with bacterial morphology (17, 20).This diversity in “Ca. Serratia symbiotica” morphology, distribution, and functions may represent evolution of different features within lineages of a single symbiont clade. If “Ca. Serratia symbiotica” is an obligate nutritional symbiont in Cinara hosts, it is expected that Cinara-associated symbionts would form a clade in which the intraclade relationships mirror those of the hosts (cocladogenesis), as observed for Buchnera and other obligate nutritional symbionts of insects (13, 21, 38). Indeed, Lamelas et al. postulated that, based on their similar phylogenies, Serratia symbionts from aphids belonging to the subgenus Cinara have had a long-term relationship with their hosts (17).In addition to the three most common facultative symbiont types found in aphids described above, several other symbiont lineages with unknown functions have been identified by amplification of bacterial 16S rRNA gene sequences from various aphid species (10, 28, 39). Here we examine the diversity of Serratia and other facultative symbionts in aphids belonging to the subfamily Lachninae. We investigated the distribution of symbionts in aphid species and geographic locations and looked for coevolutionary patterns that may correspond to the functions of facultative symbionts within their hosts.  相似文献   

3.
Impacts of climate change on organisms are already apparent, with effects ranging from the individual to ecosystem scales. For organisms engaged in mutualisms, climate may affect population performance directly or indirectly through mediated effects on their mutualists. We tested this hypothesis for two stink bugs, Acrosternum hilare and Murgantia histrionica, and their gut-associated symbionts. We reared these species at two constant temperatures, 25 and 30°C, and monitored population demographic parameters and the presence of gut-associated symbionts with diagnostic PCR primer sets. Both stink bugs lost their respective gut symbionts within two generations at 30°C. In addition, the insect survivorship and reproductive rates of both A. hilare and M. histrionica at 30°C were lower than at 25°C. Other demographic parameters also indicated a decrease in overall insect fitness at the high temperature. Collectively our data showed that the decrease in host fitness was coupled with, and potentially mediated by, symbiont loss at 30°C. This work illustrates the need to better understand the biology of animal-symbiont associations and the consequences of local climate for the dynamics of these interactions.The effect of climate on organisms, communities, and the environment at large has become a pressing issue for biologists and environmental scientists. Recent studies indicate that previous forecasts were conservative in their predictions for the magnitude of global warming (29). Up-to-date models suggest that the global mean surface temperature will increase by 1.8 to 4°C by the year 2100 (11). The ecological impact of such warming is already apparent (35) in the effects seen on species fitness (24), range shifts (22), species interactions (10), and community structure (32).It is important to note that many macroorganisms live in symbiosis with microbes and that host fitness may be affected indirectly by higher temperatures due to the disruption of mutualistic relationships. Some corals, for example, have symbiotic relationships with photosynthesizing dinoflagellates (zoothanxellae) that provide them with nutrients (31). Higher water temperatures in reef ecosystems, among other factors, induce the expulsion of microbial symbionts by the host, resulting in coral bleaching (15). Therefore, it is plausible that observed effects of climate on species distribution or performance might stem from disruption of symbiotic interactions as much as from direct effects on host biology.Despite the current interest in insect-microbe symbioses, the vast majority of such systems have been poorly studied. A group of insects that has recently received some attention are the true bugs (Hemiptera, Pentatomomorpha). Studies in the early 1900s suggested that mutualistic bacteria colonized a portion of the gut of insects in different pentatomomorphan families (9). More specifically, monocultures of bacteria were present in high densities in the crypt- (or cecum-) bearing organ preceding the hindgut of hosts, with different bacterial taxa associated with different true bug families. Furthermore, studies have shown that these symbionts are mutualistic (1, 8, 13, 14, 26). Among these bug families, the stink bugs (Pentatomidae) have been shown to depend on their gut symbionts (1). Pentatomid symbionts are polyphyletic and closely related to plant-associated bacteria in the genera Erwinia and Pantoea (25). Although the mechanism of symbiont vertical transmission is poorly understood, females seem to smear the surface of eggs with bacteria while ovipositing (3). Aposymbiotic first instars hatch but remain on the surface of eggs and acquire the symbiont by probing on the egg surface, as evidenced by the fact that surface sterilization of egg masses generates aposymbiotic individuals (1, 26, 28).Climate change has already affected stink bug performance and geographic range (19, 20). In Japan, populations of two pentatomid species, Nezara viridula and N. antennata, have shifted northwards and to higher elevations, respectively, over the last 50 years (33). However, we found previously that for one of these species high temperature eliminated gut symbionts, without any clear decrease in host fitness (27). Thus, it remains unclear whether temperature change played a role, either directly or indirectly, in these geographic shifts. To better understand the extent to which temperature mediates stink bug ecology and prevalence of their gut bacteria, we conducted laboratory studies with two pentatomid species, Acrosternum hilare and Murgantia histrionica. We show that high temperature affects the symbiotic relationship, with concomitant reduction in insect fitness.  相似文献   

4.
Bacteria in the genus Rickettsia are intracellular symbionts of disparate groups of organisms. Some Rickettsia strains infect vertebrate animals and plants, where they cause diseases, but most strains are vertically inherited symbionts of invertebrates. In insects Rickettsia symbionts are known to have diverse effects on hosts ranging from influencing host fitness to manipulating reproduction. Here we provide evidence that a Rickettsia symbiont causes thelytokous parthenogenesis (in which mothers produce only daughters from unfertilized eggs) in a parasitoid wasp, Pnigalio soemius (Hymenoptera: Eulophidae). Feeding antibiotics to thelytokous female wasps resulted in production of progeny that were almost all males. Cloning and sequencing of a fragment of the 16S rRNA gene amplified with universal primers, diagnostic PCR screening of symbiont lineages associated with manipulation of reproduction, and fluorescence in situ hybridization (FISH) revealed that Rickettsia is always associated with thelytokous P. soemius and that no other bacteria that manipulate reproduction are present. Molecular analyses and FISH showed that Rickettsia is distributed in the reproductive tissues and is transovarially transmitted from mothers to offspring. Comparison of antibiotic-treated females and untreated females showed that infection had no cost. Phylogenetic analyses of 16S rRNA and gltA gene sequences placed the symbiont of P. soemius in the bellii group and indicated that there have been two separate origins of the parthenogenesis-inducing phenotype in the genus Rickettsia. A possible route for evolution of induction of parthenogenesis in the two distantly related Rickettsia lineages is discussed.The genus Rickettsia contains a group of obligate intracellular symbionts of eukaryotic cells and belongs to the family Rickettsiaceae in the order Rickettsiales of the Alphaproteobacteria (58, 90). Many species have medical importance as they are pathogens of humans and other vertebrates; pathogenic Rickettsia species infect their hosts through blood-feeding arthropods, including lice, fleas, ticks, and mites (51, 80). In addition to Rickettsia species that cause infectious diseases in vertebrates, symbiotic species have been found in disparate groups of organisms, including arthropods, annelids, amoebae, hydrozoa, and plants (53). Rickettsia appears to be especially common in arthropods, having been found in a wide range of taxa in the classes Entognatha (springtails), Insecta (booklice, lice, bugs, leafhoppers, aphids, whiteflies, fleas, flies, lacewings, moths, beetles, and wasps), and Acarina (ticks and mites) (86). However, in most cases, the effect of Rickettsia on the invertebrate host has not been established yet. In general, Rickettsia bacteria are facultative symbionts, but in the booklouse Liposcelis bostrychophila the association is strictly obligate and Rickettsia has an essential role in oocyte development (54, 92). Facultative symbiotic Rickettsia strains have been reported to negatively affect some aspects of host fitness, causing reductions in body weight, fecundity, and longevity in the pea aphid (16, 60, 64), reductions in viability in some blood-feeding arthropod vectors (5, 46), and increased susceptibility to insecticides in the sweet potato whitefly (41). There is also evidence that Rickettsia has positive effects on host fitness, such as a larger body size in infected leeches (40) and a possible role in the oogenesis of a bark beetle (93). Finally, facultative symbiotic rickettsiae can be reproductive parasites of insects. Rickettsia strains are the causal agents of male killing (infected male embryos die) in some ladybird (79, 88) and buprestid leaf-mining (42) beetles. They are also the cause of thelytokous parthenogenesis (in which mothers produce only daughters from unfertilized eggs) in a parasitoid wasp (32). Both kinds of reproductive manipulation bias the host sex ratio toward females and favor the spread of the transovarially inherited Rickettsia strains in the infected populations. In general, Rickettsia is transmitted primarily vertically to host progeny, but in pathogenic species there is concomitant horizontal transmission via intermediate vertebrate hosts, which plays an important role in maintaining the infection in populations of blood-feeding arthropods (53, 57). An exception is Rickettsia prowazekii, the epidemic typhus agent, which spreads only via horizontal transmission in louse host populations (5). Only one Rickettsia is known to be a plant pathogen, and leafhoppers transfer this pathogen horizontally between plants (20). The fact that Rickettsia can be transmitted horizontally and then perpetuated vertically in host descendants has probably been one of the most important factors determining the enormous diversity of Rickettsia symbiotic associations. This point has been emphasized by analyses that have revealed considerable incongruence between Rickettsia and host phylogenies, indicating that horizontal transfer has occurred multiple times over evolutionary timescales (53, 54, 86).In addition to Rickettsia, diverse heritable bacteria are known to manipulate host reproduction to enhance their transmission in arthropods (12, 23). Wolbachia (order Rickettsiales, family Anaplasmataceae), a close relative of Rickettsia (90), is able to induce all known forms of manipulation of reproduction, including cytoplasmic incompatibility, feminization of genetic males, male killing, and parthenogenesis (68). Previously, only Cardinium (Sphingobacteria) has been shown to cause a similar range of reproductive phenotypes, except for male killing (35). The emerging diversity of Rickettsia associated with arthropods (53, 86), combined with evidence that it can manipulate host reproduction in more than one way, suggests that this symbiont may also be a master manipulator.In the Hymenoptera, the dominant mode of reproduction is arrhenotoky; that is, diploid females develop from fertilized eggs, and haploid males develop from unfertilized eggs (76). However, thelytokous parthenogenesis is common, and in some lineages, like the superfamilies Chalcidoidea and Cynipoidea, it is strongly associated with Wolbachia or Cardinium infection (33, 35). Parthenogenesis-inducing (PI) bacteria cause restoration of diploidy in unfertilized haploid eggs, which results in female offspring (28, 50, 69). PI Wolbachia and PI Cardinium also occur in other groups of haplodiploid arthropods, such as mites (82), scale insects (56), and thrips (4). Previously, the only example of PI caused by Rickettsia was found in the eulophid parasitoid wasp Neochrysocharis formosa (1, 32). Besides PI bacteria, uniparental (thelytokous) reproduction in haplodiploid arthropods can also be caused by feminizing bacteria that are able to interact with the host sex determination system and force the development of genotypic males toward functional phenotypic females. To date, only Cardinium has been reported to be a causal agent of feminization in haplodiploid arthropods, and only two examples are known: a mite in which Cardinium causes haploid male embryos to develop as haploid females (18, 83) and a parasitoid wasp in which diploid males are converted to females (27).In this paper, thelytokous reproduction in a parasitoid wasp, Pnigalio soemius (Hymenoptera: Eulophidae), was studied. This wasp, which is probably a complex of cryptic species (8), is a solitary ectoparasitoid that attacks larvae of many leafminer insect species in the orders Coleoptera, Diptera, Hymenoptera, and Lepidoptera (48), some of which are pests of agricultural crops (37, 61). Female P. soemius wasps paralyze host larvae by injection of venom and subsequently lay a single egg next to the host inside a leaf mine; then the parasitoid larva eats the killed host (7). Commonly, P. soemius reproduces biparentally, and the occurrence of thelytoky has not been reported previously. The aims of this study were to determine whether symbiotic bacteria are involved in manipulating the reproduction of P. soemius and then to determine the taxonomic affiliation and phenotype of the manipulators of reproduction discovered. By using antibiotic treatments and karyological analysis of the insect studied, molecular and phylogenetic characterization of the symbiotic bacteria, and detection of intracellular symbionts by means of fluorescence in situ hybridization, it was demonstrated that a PI Rickettsia causes thelytokous reproduction in P. soemius.  相似文献   

5.
Insect endosymbiont genomes reflect massive gene loss. Two Blattabacterium genomes display colinearity and similar gene contents, despite high orthologous gene divergence, reflecting over 140 million years of independent evolution in separate cockroach lineages. We speculate that distant homologs may replace the functions of some eliminated genes through broadened substrate specificity.Obligate symbionts of insects exhibit extreme patterns of genome evolution and include the smallest known bacterial genomes (10, 11, 14). Two recently published sequences of Blattabacterium, the obligate symbiont of cockroaches (7, 16), present the opportunity to analyze genome evolution in an additional symbiont lineage with extreme genome reduction.  相似文献   

6.
Heritable bacterial symbionts are widespread in insects and can have many important effects on host ecology and fitness. Fungal symbionts are also important in shaping their hosts'' behavior, interactions, and evolution, but they have been largely overlooked. Experimental tests to determine the relevance of fungal symbionts to their insect hosts are currently extremely rare, and to our knowledge, there have been no such tests for strictly predacious insects. We investigated the fitness consequences for a parasitic wasp (Comperia merceti) of an inherited fungal symbiont in the Saccharomycotina (Ascomycota) that was long presumed to be a mutualist. In comparisons of wasp lines with and without this symbiont, we found no evidence of mutualism. Instead, there were significant fitness costs to the wasps in the presence of the yeast; infected wasps attacked fewer hosts and had longer development times. We also examined the relative competitive abilities of the larval progeny of infected and uninfected mothers, as well as horizontal transmission of the fungal symbiont among larval wasps that shared a single host cockroach egg case. We found no difference in larval competitive ability when larvae whose infection status differed shared a single host. We did find high rates of horizontal transmission of the fungus, and we suggest that this transmission is likely responsible for the maintenance of this infection in wasp populations.The majority of heritable bacterial symbionts associated with insects either provide nutritional benefits for hosts that feed on nutrient-poor diets, such as blood (e.g., Wigglesworthia sp. [1]) or sap (e.g., Buchnera spp. [33]), or manipulate the hosts'' reproduction to benefit their own transmission (e.g., Wolbachia spp. [38] or Cardinium sp. [40]). Thanks in part to these examples, research efforts have become more diverse, leading to the discovery of additional benefits, such as heat tolerance (29) and protection from parasitism (26).Despite growing interest in the cryptic roles of insect associates, fungal symbionts have largely been overlooked, and their prevalence, ecological importance, and evolutionary implications for hosts are still poorly understood. Yet we have reason to suspect that fungal symbionts may be as diverse and functionally important as bacteria in insects. Buchner''s (5) foundational work on arthropod-microbe symbioses included many fungi, and anecdotal reports of such symbioses are scattered throughout the literature (e.g., fire ants [3]; stingless bees [28]; earwigs, scale insects, flies, andrenid bees, and ants [39]; and leafhoppers [30]). Recent surveys of insects for fungi have resulted in an astonishing diversity, including fungi in beetles (35), a cockroach and five other neuropteran families (24), sap-feeding beetles, and flies and bees (15), and it has been suggested that the majority of unicellular fungal diversity may be in insects (35). It is often suggested that such associations are mutualistic, with the fungus presumably providing enzymes, essential amino acids, vitamins, or sterols (37) and the insect vectoring and providing a habitat for the fungus. Fitness consequences of these associations have been assessed in only a few cases, including associations in planthoppers (31), anobiid beetles (23, 32), and scolytid beetles (2). In most instances the significance of the relationship is not clear, especially in the many cases where the fungi are not obligate associates.In 1985, LeBeck (18) reported a unicellular fungal symbiont in Comperia merceti (Compere) (Hymenoptera: Encyrtidae), a gregarious endoparasitoid wasp that specializes on the egg cases of brown-banded cockroaches [Supella longipalpa (Serville) (Blattaria: Blattellidae)]. The fungus is found throughout the hemocoel in juvenile wasps, in adult males, and in the venom gland of adult females (18). In addition, the fungus is vertically transmitted from mother to offspring via the external surface of wasp eggs during oviposition into cockroach egg cases. Vertical transmission via the egg surface is a common method in other fungal symbiont systems (e.g., planthoppers [19]; lacewings [10]; and wood wasps, anobiid beetles, and cerambycid beetles [5]). LeBeck (18) characterized the fungus as a Candida sp. and suggested that it might alter the nutritional value of the host cockroach egg case for the benefit of the developing wasp larvae. However, this claim has never been tested. Further, the predacious diet of immature parasitic wasps would make them unusual candidates for nutritional symbionts; parasitic wasps consume other insects and do not ordinarily require the complementary nutrients that many fungal and bacterial symbionts provide to insects with unbalanced diets. To our knowledge, our study is the first to specifically test the role of an inherited fungus in an insect with a strictly predacious diet.C. merceti wasps house a single known fungal symbiont belonging to the Ascomycota (Saccharomycotina) and no detectable bacterial symbionts (9). Further, these wasps do not become infected with any of their host cockroaches'' symbionts (9). In in vitro trials of the C. merceti wasp fungus with other microbes there was no evidence of inhibition or any type of interaction (C. M. Gibson, unpublished). The current research tests the hypothesis that the wasps'' fungal symbiont is a mutualist and explores alternative means by which this fungus could be maintained in wasp populations.  相似文献   

7.
Bacteria often infect their hosts from environmental sources, but little is known about how environmental and host-infecting populations are related. Here, phylogenetic clustering and diversity were investigated in a natural community of rhizobial bacteria from the genus Bradyrhizobium. These bacteria live in the soil and also form beneficial root nodule symbioses with legumes, including those in the genus Lotus. Two hundred eighty pure cultures of Bradyrhizobium bacteria were isolated and genotyped from wild hosts, including Lotus angustissimus, Lotus heermannii, Lotus micranthus, and Lotus strigosus. Bacteria were cultured directly from symbiotic nodules and from two microenvironments on the soil-root interface: root tips and mature (old) root surfaces. Bayesian phylogenies of Bradyrhizobium isolates were reconstructed using the internal transcribed spacer (ITS), and the structure of phylogenetic relatedness among bacteria was examined by host species and microenvironment. Inoculation assays were performed to confirm the nodulation status of a subset of isolates. Most recovered rhizobial genotypes were unique and found only in root surface communities, where little bacterial population genetic structure was detected among hosts. Conversely, most nodule isolates could be classified into several related, hyper-abundant genotypes that were phylogenetically clustered within host species. This pattern suggests that host infection provides ample rewards to symbiotic bacteria but that host specificity can strongly structure only a small subset of the rhizobial community.Symbiotic bacteria often encounter hosts from environmental sources (32, 48, 60), which leads to multipartite life histories including host-inhabiting and environmental stages. Research on host-associated bacteria, including pathogens and beneficial symbionts, has focused primarily on infection and proliferation in hosts, and key questions about the ecology and evolution of the free-living stages have remained unanswered. For instance, is host association ubiquitous within a bacterial lineage, or if not, do host-infecting genotypes represent a phylogenetically nonrandom subset? Assuming that host infection and free-living existence exert different selective pressures, do bacterial lineages diverge into specialists for these different lifestyles? Another set of questions addresses the degree to which bacteria associate with specific host partners. Do bacterial genotypes invariably associate with specific host lineages, and is such specificity controlled by one or both partners? Alternatively, is specificity simply a by-product of ecological cooccurrence among bacteria and hosts?Rhizobial bacteria comprise several distantly related proteobacterial lineages, most notably the genera Azorhizobium, Bradyrhizobium, Mesorhizobium, Rhizobium, and Sinorhizobium (52), that have acquired the ability to form nodules on legumes and symbiotically fix nitrogen. Acquisition of nodulation and nitrogen fixation loci has likely occurred through repeated lateral transfer of symbiotic loci (13, 74). Thus, the term “rhizobia” identifies a suite of symbiotic traits in multiple genomic backgrounds rather than a taxonomic classification. When rhizobia infect legume hosts, they differentiate into specialized endosymbiotic cells called bacteroids, which reduce atmospheric nitrogen in exchange for photosynthates from the plant (35, 60). Rhizobial transmission among legume hosts is infectious. Rhizobia can spread among hosts through the soil (60), and maternal inheritance (through seeds) is unknown (11, 43, 55). Nodule formation on hosts is guided by reciprocal molecular signaling between bacteria and plant (5, 46, 58), and successful infection requires a compatible pairing of legume and rhizobial genotypes. While both host and symbiont genotypes can alter the outcome of rhizobial competition for adsorption (34) and nodulation (33, 39, 65) of legume roots, little is known about how this competition plays out in nature.Rhizobia can achieve reproductive success via multiple lifestyles (12), including living free in the soil (14, 44, 53, 62), on or near root surfaces (12, 18, 19, 51), or in legume nodules (60). Least is known about rhizobia in bulk soil (not penetrated by plant roots). While rhizobia can persist for years in soil without host legumes (12, 30, 61), it appears that growth is often negligible in bulk soil (4, 10, 14, 22, 25). Rhizobia can also proliferate in the rhizosphere (soil near the root zone) of legumes (4, 10, 18, 19, 22, 25, 51). Some rhizobia might specialize in rhizosphere growth and infect hosts only rarely (12, 14, 51), whereas other genotypes are clearly nonsymbiotic because they lack key genes (62) and must therefore persist in the soil. The best-understood rhizobial lifestyle is the root nodule symbiosis with legumes, which is thought to offer fitness rewards that are superior to life in the soil (12). After the initial infection, nodules grow and harbor increasing populations of bacteria until the nodules senesce and the rhizobia are released into the soil (11, 12, 38, 40, 55). However, rhizobial fitness in nodules is not guaranteed. Host species differ in the type of nodules they form, and this can determine the degree to which differentiated bacteroids can repopulate the soil (11, 12, 38, 59). Furthermore, some legumes can hinder the growth of nodules with ineffective rhizobia, thus punishing uncooperative symbionts (11, 27, 28, 56, 71).Here, we investigated the relationships between environmental and host-infecting populations of rhizobia. A main objective was to test the hypothesis that rhizobia exhibit specificity among host species as well as among host microenvironments, specifically symbiotic nodules, root surfaces, and root tips. We predicted that host infection and environmental existence exert different selective pressures on rhizobia, leading to divergent patterns of clustering, diversity, and abundance of rhizobial genotypes.  相似文献   

8.
Protozoa play host for many intracellular bacteria and are important for the adaptation of pathogenic bacteria to eukaryotic cells. We analyzed the genome sequence of “Candidatus Amoebophilus asiaticus,” an obligate intracellular amoeba symbiont belonging to the Bacteroidetes. The genome has a size of 1.89 Mbp, encodes 1,557 proteins, and shows massive proliferation of IS elements (24% of all genes), although the genome seems to be evolutionarily relatively stable. The genome does not encode pathways for de novo biosynthesis of cofactors, nucleotides, and almost all amino acids. “Ca. Amoebophilus asiaticus” encodes a variety of proteins with predicted importance for host cell interaction; in particular, an arsenal of proteins with eukaryotic domains, including ankyrin-, TPR/SEL1-, and leucine-rich repeats, which is hitherto unmatched among prokaryotes, is remarkable. Unexpectedly, 26 proteins that can interfere with the host ubiquitin system were identified in the genome. These proteins include F- and U-box domain proteins and two ubiquitin-specific proteases of the CA clan C19 family, representing the first prokaryotic members of this protein family. Consequently, interference with the host ubiquitin system is an important host cell interaction mechanism of “Ca. Amoebophilus asiaticus”. More generally, we show that the eukaryotic domains identified in “Ca. Amoebophilus asiaticus” are also significantly enriched in the genomes of other amoeba-associated bacteria (including chlamydiae, Legionella pneumophila, Rickettsia bellii, Francisella tularensis, and Mycobacterium avium). This indicates that phylogenetically and ecologically diverse bacteria which thrive inside amoebae exploit common mechanisms for interaction with their hosts, and it provides further evidence for the role of amoebae as training grounds for bacterial pathogens of humans.Free-living amoebae, such as Acanthamoeba spp., are ubiquitous protozoa which can be found in such diverse habitats as soil, marine water, and freshwater and in many engineered environments (62, 100). They are important predators of prokaryotic and eukaryotic microorganisms, thereby having great influence on microbial community composition, soil mineralization, plant growth, and nutrient cycles (14, 100). Interestingly, many well-known pathogens of humans are able to infect, survive, and multiply within amoebae (39, 51). These protozoa can thus serve as reservoirs and vectors for the transmission of pathogenic bacteria to humans, as demonstrated for L. pneumophila and Mycobacterium avium (2, 115). It is also increasingly being recognized that protozoa are important for the adaptation of (pathogenic) bacteria to higher eukaryotic cells as a niche for growth (2, 24, 42, 78, 89).In addition to the many recognized transient associations between amoeba and pathogens, stable and obligate relationships between bacteria and amoebae also were described for members of the Alphaproteobacteria (11, 34, 48), the Betaproteobacteria (49), the Bacteroidetes (50), and the Chlamydiae (4, 12, 35, 52). These obligate amoeba symbionts show a worldwide distribution, since phylogenetically highly similar strains were found in amoeba isolates from geographically distant sources (51, 107). The phylogenetic diversity and the different lifestyles of these obligate intracellular bacteria—some are located directly in the host cell cytoplasm (11, 34, 48-50, 52), while others are enclosed in host-derived vacuoles (4, 35, 44)—suggest fundamentally different mechanisms of host cell interaction. However, with the exception of chlamydia-related amoeba symbionts (37, 46, 47), our knowledge of the biology of obligate intracellular symbionts of amoebae is still scarce.Comparative genomics has been extremely helpful for the analysis of intracellular bacteria. Numerous genome sequences from the Alpha- and Gammaproteobacteria and the Chlamydiae are available and have contributed significantly to our understanding of genome evolution, the biology of intracellular bacteria, and the interactions with their host cells (24, 26, 46, 79, 82). In this study, we determined and analyzed the complete genome sequence of “Candidatus Amoebophilus asiaticus” strain 5a2 in order to gain novel insights into its biology. “Ca. Amoebophilus asiaticus” is a Gram-negative, obligate intracellular amoeba symbiont belonging to the Bacteroidetes which has been discovered within an amoeba isolated from lake sediment (107). “Ca. Amoebophilus asiaticus” shows highest 16S rRNA similarity to “Candidatus Cardinium hertigii,” an obligate intracellular parasite of arthropods able to manipulate the reproduction of its hosts (131). According to 16S rRNA trees, both organisms are members of a monophyletic group within the phylogenetically diverse phylum Bacteroidetes, consisting only of symbionts and sequences which were directly retrieved from corals (113). Among members of the Bacteroidetes, the genome sequences of only three symbionts, which are only distantly related (75 to 80% 16S rRNA sequence similarity) to “Ca. Amoebophilus asiaticus,” have been determined to date: two strains of “Candidatus Sulcia muelleri, a symbiont of sharpshooters, and “Azobacteroides pseudotrichonymphae,” a symbiont of an anaerobic termite gut ciliate (45, 72, 74, 127).The genome of “Ca. Amoebophilus asiaticus” is only moderately reduced in size compared to those of many other obligate intracellular bacteria (75, 123), but nevertheless, its biosynthetic capabilities are extremely limited. A large fraction of the genome consists of IS elements and an unparalleled high number of proteins with eukaryotic domains, such as ankyrin repeats, TPR/SEL1 repeats, leucine-rich repeats, and domains from the eukaryotic ubiquitin system, all of them most likely important for host cell interaction. Feature enrichment analysis across a nonredundant data set of all bacterial genomes showed that these domains are enriched in the genomes of bacteria (including several pathogens of humans) known to be able to infect amoebae, providing further evidence for an important role of amoebae in the evolution of mechanisms for host cell interaction in intracellular bacteria.  相似文献   

9.
10.
11.
Almost all lumbricid earthworms (Oligochaeta: Lumbricidae) harbor species-specific Verminephrobacter (Betaproteobacteria) symbionts in their nephridia (excretory organs). The function of the symbiosis, and whether the symbionts have a beneficial effect on their earthworm host, is unknown; however, the symbionts have been hypothesized to enhance nitrogen retention in earthworms. The effect of Verminephrobacter on the life history traits of the earthworm Aporrectodea tuberculata (Eisen) was investigated by comparing the growth, development, and fecundity of worms with and without symbionts given high (cow dung)- and low (straw)-nutrient diets. There were no differences in worm growth or the number of cocoons produced by symbiotic and aposymbiotic worms. Worms with Verminephrobacter symbionts reached sexual maturity earlier and had higher cocoon hatching success than worms cured of their symbionts when grown on the low-nutrient diet. Thus, Verminephrobacter nephridial symbionts do have a beneficial effect on their earthworm host. Cocoons with and without symbionts did not significantly differ in total organic carbon, total nitrogen, or total hydrolyzable amino acid content, which strongly questions the hypothesized role of the symbionts in nitrogen recycling for the host.Symbiosis has long been recognized as a source of evolutionary innovation (24), and the acquisition of symbionts can enable animal hosts to exploit previously inaccessible niches (3). The phylum Annelida is no exception to this; chemosynthetic symbionts in marine annelids (e.g., the giant tubeworm Riftia sp. and other gutless marine oligochaetes [9]) gain energy from the oxidation of reduced sulfur compounds and fix CO2 and supply their animal host with fixed carbon. A more obscure partnership is known from the bone-eating annelid Osedax sp., where endosymbionts help degrade the bones of whale carcasses, the only known habitat of the worms (33). The medicinal leech Hirudo sp., like other blood-feeding animals (3, 7), has symbionts that are thought to produce essential vitamins missing from a blood meal (13). In addition, leeches have a number of symbionts of unknown function in their nephridia (excretory organs) (18). Earthworms (Oligochaeta: Lumbricidae) have also long been known to harbor symbiotic bacteria in their nephridia (19, 36). The function of this symbiosis, however, is still not known, but the stability of the symbiosis over evolutionary time (23) suggests that the symbionts benefit the host.The earthworm symbionts reside in the nephridia and have therefore been proposed to be involved in internal recycling of nitrogen in the host (29). The earthworm nephridia play an important role in both nitrogenous waste excretion and osmoregulation (20). The nephridia are found in pairs in each segment of the worm and consist of a coiled tube leading from the coelom to the exterior (Fig. (Fig.1).1). The tube forms three loops, and the symbiotic bacteria are situated in the ampulla in the second loop, where they form a dense population lining the lumen wall (19, 36).Open in a separate windowFIG. 1.The nephridia are found as paired organs in each segment of the worm. The nephridostome (the inlet to the nephridia) protrudes into the previous segment. The nephridial tube forms three loops and finally empties out through the body wall via the nephridopore. The symbionts (black) reside in the ampulla in the second loop. (Modified from reference 23 with permission.)The symbionts form the monophyletic genus Verminephrobacter (Betaproteobacteria) (30, 36); they are species specific and present in almost all lumbricid earthworms (23). The Verminephrobacter symbionts are transmitted vertically via the cocoon, where they are deposited along with eggs and sperm (5). During embryogenesis, the symbionts migrate into the developing nephridia, and after the worm hatches, the symbionts can no longer infect (5, 6). By taking advantage of the vertical transmission mode, it has been possible to establish symbiont-free earthworm cultures in the laboratory through controlled antibiotic treatment of newly deposited cocoons (5; this study). Separation of the symbiotic partners allows studies of the effect of the symbionts on their earthworm hosts.Pandazis (29) hypothesized that the symbionts enhance earthworm nitrogen retention by excreting proteolytic enzymes that will degrade peptides and proteins lost in urine; this would allow the earthworm to reabsorb the resulting amino acids. As a consequence, earthworms cured of their symbionts should have a lower fitness level than control worms when grown under nitrogen-limiting conditions. To test this hypothesis, growth, development, fecundity, and cocoon hatching success were compared for symbiotic and aposymbiotic earthworms of the species Aporrectodea tuberculata (Eisen) under high and low nutrient availability conditions.  相似文献   

12.
13.
The Caribbean reef sponge Svenzea zeai was previously found to contain substantial quantities of unicellular photosynthetic and autotrophic microbes in its tissues, but the identities of these symbionts and their method of transfer from adult to progeny are largely unknown. In this study, both a 16S rRNA gene-based fingerprinting technique (denaturing gradient gel electrophoresis [DGGE]) and clone library analysis were applied to compare the bacterial communities associated with adults and embryos of S. zeai to test the hypothesis of vertical transfer across generations. In addition, the same techniques were applied to the bacterial community from the seawater adjacent to adult sponges to test the hypothesis that water column bacteria could be transferred horizontally as sponge symbionts. Results of both DGGE and clone library analysis support the vertical transfer hypothesis in that the bacterial communities associated with sponge adults and embryos were highly similar to each other but completely different from those in the surrounding seawater. Sequencing of prominent DGGE bands and of clones from the libraries revealed that the bacterial communities associated with the sponge, whether adult or embryo, consisted of a large proportion of bacteria in the phyla Chloroflexi and Acidobacteria, while most of the sequences recovered from the community in the adjacent water column belonged to the class Alphaproteobacteria. Altogether, 21 monophyletic sequence clusters, comprising sequences from both sponge adults and embryos but not from the seawater, were identified. More than half of the sponge-derived sequences fell into these clusters. Comparison of sequences recovered in this study with those deposited in GenBank revealed that more than 75% of S. zeai-derived sequences were closely related to sequences derived from other sponge species, but none of the sequences recovered from the seawater column overlapped with those from adults or embryos of S. zeai. In conclusion, there is strong evidence that a dominant proportion of sponge-specific bacteria present in the tissues of S. zeai are maintained through vertical transfer during embryogenesis rather than through acquisition from the environment (horizontal transfer).Besides being the oldest metazoans, sponges are the simplest multicellular animals and possess a low degree of tissue differentiation and coordination (54). Sponges are sessile, filter-feeding organisms that may harbor within their tissues a remarkable array of microorganisms, including bacteria (19, 59, 64), archaea (41), zooxanthellae (22), diatoms (63), and fungi (35). In some cases, microbial consortia can make up to 40 to 60% of the sponge tissue volume (21, 61) and exceed a density of 109 microbial cells per ml of sponge tissue (62), which is several orders of magnitude higher than that found in seawater. Apart from being a source of food (43), bacterial symbionts may participate in the acquisition and transfer of nutrients inside sponges (67, 68), the recycling of insoluble protein (69), the stabilization of the sponge skeleton (44), and the processing of metabolic waste (4, 65). Many antimicrobial compounds have been isolated from sponge bacterial symbionts (24, 47, 53), suggesting the involvement of symbiotic bacteria in sponge chemical defenses. In some cases, bacterial symbionts have been found to be the source of bioactive compounds that were isolated from sponges, which has opened up new research directions in marine natural product chemistry, biotechnology, and pharmaceutical development (18, 23, 40).Based on immunological evidence from the 1980s (66), sponge-bacterium symbioses are thought to have originated in the Precambrian, when bacteria evolved to form a single clade of sponge-specific bacteria that were distinct from isolates found in the surrounding seawater. Since then, many studies have similarly documented a high level of consistency and specificity in sponge-bacterium associations (20, 27, 59). Nevertheless, questions remain about the acquisition and maintenance of symbionts in host sponges. In general, the following two hypotheses have been proposed: (i) a recently metamorphosed sponge selectively retains specific groups of bacteria from the diverse pool of bacteria present in the water column as it begins filter feeding (horizontal transfer) or (ii) specific bacterial strains are transmitted by the maternal sponge to developing embryos and are already present in the metamorphosing sponge (vertical transfer) (58). The first hypothesis requires some recognition of specific microbes by the sponge, perhaps through an innate immune system (36) or other means to distinguish symbiont strains from food bacteria (70).Vertical transfer of bacterial symbionts in sponges was first proposed by Lévi and Porte (29), who demonstrated the presence of bacteria inside the larvae of the sponge Oscarella lobularis. Later, in 1976, Lévi and Lévi (30) studied the transmission of bacteria in the sponge Chondrosia reniformis via sponge oocytes. Since then, vertical transmission of bacterial symbionts via eggs or larvae has been documented for several sponge species, including Tethya citrina (15), Geodia cydonium (50), Stelletta grubii (49), Hippospongia sp. (25), Spongia sp. (25), Halisarca dujardini (10), and Corticium candelabrum (8). However, all of these studies employed transmission and scanning electron microscopy and could only examine the presence of bacteria in maternal sponges, oocytes, or larvae at the morphological level, with no determination of microbial identity. With advances in molecular techniques, Enticknap et al. (9) were the first to report the successful isolation of an alphaproteobacterial symbiont, strain NW001, from both the adult sponge Mycale laxissima and its larvae. They also did a preliminary denaturing gradient gel electrophoresis (DGGE) analysis of the bacterial community in seawater and compared that with the community in the sponge larval sample. However, such a comparison was not extended to the sponge adult, and no solid conclusion can be drawn for the horizontal transfer mechanism of sponge symbionts. More recently, Sharp et al. (52) used fluorescence in situ hybridization (FISH) and clone library techniques to demonstrate the presence of proteobacteria, actinobacteria, and a clade of sponge-associated bacteria in the embryos and mesohyl of the tropical sponge Corticium sp. By clone library and DGGE analyses, Schmitt et al. (48a) identified 28 vertical-transmission clusters in five different Caribbean sponge species and demonstrated that the complex sponge adult microbial community was collectively transmitted through reproductive stages. While these recent studies support the vertical transfer hypothesis, they did not fully address the identities of microbes in the water column surrounding the sponges, which is key to determining whether horizontal transfer may also take place.The Caribbean reef sponge Pseudaxinella zeai was reclassified into a new genus, Svenzea (Demospongiae, Halichondria, Dictyonellidae), in 2002 because it has an unusual skeleton arrangement consisting mainly of short stout styles that are arranged in an isodictyal reticulation (2). It is a viviparous sponge that produces the largest embryos (>1 mm in diameter) and larvae (6 mm long) recorded for the phylum Porifera (45). Svenzea zeai has also been classified as a bacteriosponge because it contains substantial amounts of unicellular photosynthetic and autotrophic microbial symbionts in its tissues (2, 45). Although bacteria were observed in the embryos and larvae of this sponge based on transmission electron microscopy studies (45), neither the direct linkage between the maternal sponge and the propagules nor the identity of the microbial symbionts had been established.In this study, our objective was to examine vertical versus horizontal transfer of bacterial symbionts in Svenzea zeai. This was achieved by comparing the bacterial community profiles of the adults and embryos of the sponge by use of a combination of molecular techniques, including DGGE and clone library analysis. More than one technique was employed to compensate for deficiencies of each technique in revealing bacterial community structure. Additionally, we used the same techniques to examine the bacterial community in the seawater that surrounded the sponge to determine whether horizontal transfer was evident.  相似文献   

14.
The brown planthopper (Nilaparvata lugens Stål), the most destructive pest of rice, has been identified, including biotypes with high virulence towards previously resistant rice varieties. There have also been many reports of a yeast-like symbiont of N. lugens, but little is known about the bacterial microbes. In this study, we examined the bacterial microbes in N. lugens and identified a total of 18 operational taxonomic units (OTUs) representing four phyla (Proteobacteria, Firmicutes, Actinobacteria, and Bacteroidetes) by sequencing and analyzing 16S rRNA gene libraries obtained from three populations of N. lugens, which were maintained on the rice varieties TN1, Mudgo, and ASD7. Several of the OTUs were similar to previously reported secondary symbionts of other insects, including an endosymbiont of the psyllid Glycapsis brimblecombei, an Asaia sp. found in the mosquito Anopheles stephensi, and Wolbachia, found in the mite Metaseiulus occidentalis. However, the species and numbers of the detected OTUs differed substantially among the N. lugens populations. Further, in situ hybridization analysis using digoxigenin-labeled probes indicated that OTU 1 was located in hypogastrium tissues near the ovipositor and ovary in biotype 1 insects, while OTU 2 was located in the front of the ovipositor sheath in biotype 2 insects. In addition, masses of bacterium-like organisms were observed in the tubes of salivary sheaths in rice plant tissues that the insects had fed upon. The results provide indications of the diversity of the bacterial microbes harbored by the brown planthopper and of possible associations between specific bacterial microbes and biotypes of N. lugens.Close associations between insects and the microbes they harbor appear to be common. Symbionts have been found to contribute to the nutrition, development, reproduction, speciation, and defense against natural enemies of their host insects (1, 11, 18, 30, 39). The small brown planthopper (Laodelphax striatellus) and the white-backed planthopper (Sogatella furcifrea) also reportedly harbor an alphaproteobacterial Wolbachia symbiont (29) that can be transferred horizontally between different insect species and that affects its hosts'' sexual reproduction, cytoplasmic incompatibility, and immune responses (21, 38, 39).The brown planthopper, Nilaparvata lugens Stål (Homoptera: Delphacidae), is a monophagous insect herbivore of rice (13) that feeds on rice phloem and causes serious damage to rice crops. N. lugens reportedly harbors an intracellular, eukaryotic “yeast-like symbiont” (YLS) in the fat body, which plays a key role in recycling uric acid (3, 33). However, little is known about bacterial symbionts in N. lugens.It has been well recognized that diversity exists within insect species and that “biotypes” or populations that are adapted to or that prefer a particular host can frequently develop (10, 12). The behavioral and physiological responses during insect establishment on plants are feeding, metabolism of ingested food, growth, adult survival, egg production, and oviposition (34). In N. lugens, the biotype is assigned to a population with the ability to damage varieties of rice that carry resistance genes and that were previously resistant to it (5). It has been claimed that some biotypes of N. lugens differ in small morphological features, isozymes, and DNA polymorphisms (6, 25, 36). However, the precise nature of the virulence-conferring mechanisms in N. lugens biotypes (and their modes and stability of inheritance) is not clear. It is interesting to survey symbionts in different biotype populations of N. lugens.Generally, the 16S rRNA gene has been used as a molecular marker enabling the detection of as-yet-uncultured microbes, and it facilitates a profound investigation of microbial diversity (2, 22, 44). We initiated a study using molecular methods to investigate the bacterial symbionts of N. lugens. The major objective of this study was to identify bacterial microbes in N. lugens. The identified bacterial microbes appeared to be associated with different populations of N. lugens and in some cases were located in specific tissues, according to in situ hybridization (ISH) analyses.  相似文献   

15.
The European medicinal leech, Hirudo verbana, harbors simple microbial communities in the digestive tract and bladder. The colonization history, infection frequency, and growth dynamics of symbionts through host embryogenesis are described using diagnostic PCR and quantitative PCR. Symbiont species displayed diversity in temporal establishment and proliferation through leech development.The hermaphroditic European medicinal leech (Hirudo spp.) is one of the most extensively examined animal models in neurophysiological, developmental, and behavioral studies (14) and has recently been used as a naturally occurring simple model for beneficial symbioses (5, 13). A fundamental question of microbial symbioses is how to determine the transmission mode of the symbionts between generations. Hirudo verbana reproduces by depositing eggs, which are surrounded by a cocoon. The cocoon is secreted from glandular cells of the parental mouth and usually contains 5 to 25 eggs. Each individual egg is encased by a self-enclosed vitelline membrane, referred to as the larval sac, and is bathed in a nutritious albumenous fluid (14). Complete embryonic development occurs within the cocoon and is composed of two distinct life stages, cryptolarva and juvenile. The cryptolarva transitions into a juvenile approximately midway into embryogenesis. The temporal acquisition of morphological attributes during embryonic development have been well described (3, 12, 16) (Fig. (Fig.11).Open in a separate windowFIG. 1.Paradigm of percent embryonic development (% ED) of the European medicinal leech, H. verbana, relative to the acquisition of digestive tract features. At 20 ± 1°C, 24 h is equivalent to 3.33% ED, with complete embryogenesis (spanning from cocoon deposition to the emergence of adult-like juveniles) requiring approximately 30 h. Staging scheme based on references 3 and 12. *, sampling time point; PD, days postcocoon deposition; prs, pairs; d, days. (Adapted from reference 12 with permission of John Wiley & Sons. Copyright 1998 Wiley-Liss, Inc.)The medicinal leech houses distinct microbial communities within its digestive tract and secretory bladders. Culturing and culture-independent profiling of the European medicinal leech, H. verbana, through fluorescence in situ hybridization, study of 16S rRNA gene clone libraries, and terminal restric-tion length polymorphism, revealed a simple and stable microbial community within the adult midgut (2, 4, 7, 8, 18). The gammaproteobacterium Aeromonas veronii and a member of the Bacteroidetes, Rikenella, were identified as consistent and dominant extracellular residents of the medicinal leech crop and intestinum. An early culture-based study detected a bacterium that is now considered to be A. veronii in the cocoon albumen and in young leeches after hatching (1). In previous electron microscopy work investigating the embryonic development of the bladders, intracellular bacteria were detected within the bladder wall and extracellular bacteria within the lumen (2, 16, 17). A recent study revealed that this microbiota is organized in distinct layers and is composed of the deltaproteobacterium Bdellovibrio, betaproteobacteria Comamonas and Sterolibacterium, members of the Bacteroidetes, Sphingobacterium and Niabella, and alphaproteobacterium Ochrobactrum spp. (10). Although the microbial constituents of the adult H. verbana digestive tract have been previously characterized, the succession, inoculum sizes, frequency of infection, and growth dynamics of these symbiont species during embryogenesis remain to be described.Putative functional roles for the crop/intestinum symbionts of the leech host include aiding in digestion, provisioning essential nutrients to the host, which are lacking in the blood meal (14), and preventing the establishment of foreign microbiota (1, 15). The symbionts localized in the bladders are suspected to play a role in the recycling of host metabolic waste into ammonia (10). The digestive tract symbionts may also display nutritional syntrophy, and possibly, A. veronii primes the host''s digestive tract to enable the establishment and persistence of the obligate anaerobic Rikenella-like bacterium, thereby contributing to the selection of the future microbiota (reviewed in reference 13). This paper details the microbial colonization patterns relative to H. verbana embryogenesis using a combination of species-specific diagnostic PCR and quantitative PCR (qPCR) analyses.  相似文献   

16.
17.
18.
Following cultivation-dependent and -independent techniques, we investigated the microbiota associated with Bactrocera oleae, one of the major agricultural pests in olive-producing countries. Bacterial 16S rRNA gene libraries and ultrastructural analyses revealed the presence of several bacterial taxa associated with this insect, among which Acetobacter tropicalis was predominant. The recent increased detection of acetic acid bacteria as symbionts of other insect model organisms, such as Anopheles stephensi (G. Favia et al., Proc. Natl. Acad. Sci. USA 104:9047-9051, 2007) or Drosophila melanogaster (C. R. Cox and M. S. Gilmore, Infect. Immun. 75:1565-1576, 2007), prompted us to investigate the association established between A. tropicalis and B. oleae. Using an A. tropicalis-specific PCR assay, the symbiont was detected in all insects tested originating from laboratory stocks or field-collected from different locations in Greece. This acetic acid bacterium was successfully established in cell-free medium, and typing analyses, carried out on a collection of isolates, revealed that different A. tropicalis strains are present in fly populations. The capability to colonize and lodge in the digestive system of both larvae and adults and in Malpighian tubules of adults was demonstrated by using a strain labeled with a green fluorescent protein.Associations of insects with bacteria, protozoa, and fungi are complex and intimate, ranging from parasitism to mutualism, and may be extracellular or intracellular and may play a role in the nutrition, the physiology, or the reproduction of the insect host (10). Petri (1909 to 1910) described one of the first bacterial symbiotic associations in an insect species, the olive fly, Bactrocera (Dacus) oleae (31, 32).The olive fruit fly B. oleae is one of the major pests of the olive tree, strongly affecting olive production worldwide, especially in the Mediterranean area, where more than 90% of the world''s olive tree cultivation takes place (24, 27). Although there have been reports on the isolation of potentially effective Bacillus thuringiensis strains against B. oleae, olive fly control strategies remain almost exclusively based on insecticides, despite the awareness of a need for the use of more environmentally friendly control methods (29). Recently, new concepts are emerging, among which the symbiotic control approach is particularly noteworthy (4). This strategy includes the use of symbionts as vectors of antagonistic factors able to block the life cycle of the plant pathogen in the insect host or, alternatively, their use for the suppression of host natural populations (45). In any case, a prerequisite for developing a symbiotic control approach is the knowledge of the microbiota associated with the insect pest.The nature of the olive fruit fly-associated microbiota is controversial. The culturable bacterium Pseudomonas savastanoi has been suspected to be a mutualist of B. oleae for more than 50 years (6, 17, 22, 25, 32, 33). In addition, traditional microbiological approaches have identified other bacteria of the genera Bacillus, Erwinia, Lactobacillus, Micrococcus, Pseudomonas, Streptococcus, Citrobacter, Proteus, Providencia, Enterobacter, Hafnia, Klebsiella, Serratia, and Xanthomonas as associated with the olive fruit fly (3, 14, 19, 37). Recently, it was suggested that the bacterium housed within the esophageal bulb and the midgut of B. oleae is unculturable, and the novel name “Candidatus Erwinia dacicola” was proposed (7). The presence of “Ca. Erwinia dacicola” was confirmed in Italian natural populations (36).The contradictory results obtained in previous studies prompted us to investigate the microbiota associated with both laboratory and natural populations of the olive fruit fly by employing both cultivation-independent and -dependent methods.  相似文献   

19.
The present study describes an accurate quantitative method for quantifying the adherence of conidia to the arthropod cuticle and the dynamics of conidial germination on the host. The method was developed using conidia of Metarhizium anisopliae var. anisopliae (Metschn.) Sorokin (Hypocreales: Clavicipitaceae) and engorged Rhipicephalus annulatus (Say) (Arachnida: Ixodidae) females and was also verified for M. anisopliae var. acridum Driver et Milner (Hypocreales: Clavicipitaceae) and Alphitobius diaperinus (Panzer) (Coleoptera: Tenebrionidae) larvae. This novel method is based on using an organic solvent (dichloromethane [DCM]) to remove the adhered conidia from the tick cuticle, suspending the conidia in a detergent solution, and then counting them using a hemocytometer. To confirm the efficacy of the method, scanning electron microscopy (SEM) was used to observe the conidial adherence to and removal from the tick cuticle. As the concentration of conidia in the suspension increased, there were correlating increases in both the number of conidia adhering to engorged female R. annulatus and tick mortality. However, no correlation was observed between a tick''s susceptibility to fungal infection and the amount of adhered conidia. These findings support the commonly accepted understanding of the nature of the adhesion process. The mechanism enabling the removal of the adhered conidia from the host cuticle is discussed.The entomopathogenic fungus Metarhizium anisopliae var. anisopliae (Metschn.) Sorokîn (1883) infects a broad range of arthropod hosts and can be used as a biopesticide against different insect and tick species (8, 22, 35, 36). The adhesion of the conidia of entomopathogenic fungi to the host cuticle is the initial stage of the pathogenic process and includes both passive and active events (5, 10). The hydrophobic epicuticular lipid layer plays an important role during both the attachment process and the germination of the conidia on the surface of the host (15, 19). According to Boucias et al. (7), the attachment of conidia to the host cuticle is based on nonspecific hydrophobic and electrostatic forces. The conidia of most entomopathogenic fungi, including M. anisopliae, have an outer cell layer made up of rodlets (6). The hydrophobins, specific proteins present in the rodlet layer, mediate the passive adhesion of conidia to hydrophobic surfaces, such as the cuticles of arthropods (16, 45, 46). However, as germination commences, the hydrophobins are replaced by an adhesion-like protein, Mad1, which promotes tighter and more-specific adhesion between the conidia and the host (44). Many factors may affect the adhesion and persistence of conidia on the host cuticle (i.e., characteristics of the pathogen, including its virulence [2, 18, 48], conditions under which the pathogen is cultured [17], type of spores [7, 16], topographical and chemical properties of the host cuticle [9, 38, 42], host surface hydrophobicity [15, 23], host behavior [31, 33], and environmental conditions [33]). Conidia of M. anisopliae have shown an affinity to cuticular regions containing setae or spines (7, 38) and to highly hydrophobic cuticle regions, such as mosquitoes'' siphon tubes (23) and intersegmental folds (43). Sites with higher numbers of adhered conidia varied among host species. However, in general, the membranous intersegmental regions were often particularly attractive sites for conidial attachment (26). Variation in the distribution of conidia across different anatomical regions has also been noted in studies of several tick species inoculated with entomopathogenic fungi (3, 21, 22). An evaluation of the attachment of Beauveria bassiana conidia to three tick species, Dermacentor variabilis, Rhipicephalus sanguineus, and Ixodes scapularis, demonstrated that the distribution patterns of the different conidia on the ticks'' bodies were not uniform (22). The density of the conidia and their germination varied dramatically across different anatomical regions of Amblyomma maculatum and A. americanum that had been inoculated with B. bassiana (21). Arruda et al. (3) demonstrated that mass adhesion of M. anisopliae conidia to engorged Boophilus microplus females occurs predominantly on ticks'' legs, suggesting its association with the presence of setae.There are a few approaches for assessing the adhesion of conidia to the host cuticle that are based on direct observation of the conidia on the arthropod cuticle. They involve examining a few areas on the surface of an arthropod by means of scanning electron microscopy (SEM) (11, 15, 30), transmission electron microscopy (TEM) (4), or fluorescence microscopy following vital staining of the conidia (2, 28, 29, 37). These methods are expensive, time-consuming, and relatively inaccurate due to the uneven distribution of conidia on the host surface.In this work, we describe a quantitative method for determining the total amount of conidia that have adhered to a whole host cuticle. This method is based on removing adhered conidia from the tick cuticle using an organic solvent, separating the conidia from the extract by centrifugation, resuspending the conidia in a detergent solution, and then counting the conidia in a hemocytometer. The efficacy of the method was evaluated by comparing the results of this procedure with those of a supplementary examination of conidial removal using SEM.The term “adhered” is often used to define conidia in different states: washed or unwashed after inoculation, present on the host cuticle immediately after inoculation, or kept for several hours (1, 2, 38). In this paper, the term “adhered conidia” refers to conidia that remained on the cuticle after washing by vortexing the inoculated and dried host in an aqueous solution of Triton X-100 and rinsing of the material under tap water.  相似文献   

20.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号