06.12.2012 Views

marine microbial thiotrophic ectosymbioses - HYDRA-Institute

marine microbial thiotrophic ectosymbioses - HYDRA-Institute

marine microbial thiotrophic ectosymbioses - HYDRA-Institute

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.

2727_C04.fm Page 95 Wednesday, June 30, 2004 12:00 PM<br />

MARINE MICROBIAL THIOTROPHIC ECTOSYMBIOSES<br />

J. OTT,* M. BRIGHT & S. BULGHERESI<br />

<strong>Institute</strong> of Ecology and Conservation Biology, University of Vienna,<br />

Althanstrasse 14, A-1090 Vienna, Austria<br />

*E-mail: joerg.ott@univie.ac.at<br />

Abstract A high diversity of <strong>thiotrophic</strong> symbioses is found in sulphide-rich <strong>marine</strong> habitats,<br />

involving several phyla of protists and invertebrates, as well as several subdivisions of the Proteobacteria.<br />

Whereas some of the better-known symbioses are highly evolved endosymbioses, the more<br />

primitive <strong>ectosymbioses</strong> are less well known. The sulphur-oxidising chemolithotrophic nature of<br />

the bacteria and their nutritive importance to the eukaryote host have been demonstrated for the<br />

ciliates Kentrophoros spp. and Zoothamnium niveum,<br />

the nematode subfamily Stilbonematinae, and<br />

the carid shrimp Rimicaris exoculata.<br />

For a number of other regular bacteria–eukaryote associations,<br />

such a symbiotic relationship has been hypothesised based on ecological, morphological, physiological<br />

or molecular data, but is still inconclusive.<br />

0-8493-2727-X/04/$0.00+$1.50<br />

Oceanography and Marine Biology: An Annual Review 2004, 42,<br />

95–118<br />

© R. N. Gibson, R. J. A. Atkinson, and J. D. M. Gordon, Editors<br />

The diversity of <strong>thiotrophic</strong> symbioses<br />

The interest in <strong>thiotrophic</strong> symbioses awakened by the discovery of the deep-sea hydrothermal<br />

vents has led to the discovery of an unexpected diversity of microbe/animal relationships in a<br />

variety of habitats from the intertidal zone to the deep sea (Cavanaugh 1985, Fisher 1990, Nelson<br />

& Fisher 1995). The fascination of food chains that operate without sunlight and the opportunity<br />

to find clues about the origin of life on this and probably other celestial bodies (Farmer 1998) have<br />

spurred research on hot vents and cold seeps in the deep sea and on continental slopes. In the wake<br />

of these expensive endeavours, research has been conducted in more easily accessible shallowwater<br />

sulphidic habitats and has revealed a comparable variety of symbiotic relationships (Ott 1996,<br />

Giere 1992). The deep-sea communities are unrivalled with regard to the importance that the<br />

<strong>thiotrophic</strong> symbioses play in an extremely food-limited setting. In shallow water the predominance<br />

of photoautotrophic production restricts <strong>thiotrophic</strong> symbioses to a more cryptic existence.<br />

To date symbioses with sulphur-oxidising chemolithoautotrophic bacteria have been recorded<br />

for protists (ciliates and probably also flagellates) and seven animal phyla (Platyhelminthes, Nematoda,<br />

Echiurida, Annelida, Mollusca, Arthropoda, and Echinodermata). With the exception of<br />

Platyhelminthes, Echiurida, and Echinodermata, the development of <strong>thiotrophic</strong> symbioses has<br />

occurred more than once in each phylum.<br />

The diversity of the <strong>microbial</strong> symbionts is as high as that of the hosts, and although they all<br />

belong to the Proteobacteria, there are representatives of the g-,<br />

e-,<br />

and a-subgroups.<br />

They occur<br />

as endosymbionts intracellularly in special organs such as the trophosome of the Vestimentifera<br />

(Siboglinidae, Annelida) and similar organs in Catenulida (Platyhelminthes) and the nematode<br />

Astomonema jenneri,<br />

or within organs of other functions, such as the gills of bivalves and gastropods,<br />

the vestigial gut in Astomonema southwardorum,<br />

or under the cuticle between epidermis cells<br />

of Oligochaeta (Fisher 1996, Giere 1996).<br />

95


2727_C04.fm Page 96 Wednesday, June 30, 2004 12:00 PM<br />

96 J. Ott, M. Bright & S. Bulgheresi<br />

In many cases, however, they appear ectosymbiotically, attached to the body surface of their<br />

eukaryote host, such as in flagellates, ciliates, Nematoda (Stilbonematinae), and in certain Annelida<br />

(Alvinellidae) and Arthropoda (Ott 1996, Polz & Cavanaugh 1996). This review summarises our<br />

knowledge about these <strong>ectosymbioses</strong>.<br />

Ectosymbioses differ from endosymbioses in many respects: the microbes are largely exposed<br />

to ambient conditions, although the eukaryote host is responsible for the position within gradients<br />

of environmental variables. Moreover, substances produced by the host may modify the environment<br />

and physiology of the bacterial partner. The animal hosts appear less modified than is the case in<br />

endosymbioses, and in most cases, the relationship to nonsymbiotic relatives can be traced with<br />

confidence. Rarely is a particular ectosymbiosis characteristic for taxa higher than genera. Although<br />

morphological modifications in conjunction with the symbiotic way of life are present in practically<br />

all <strong>ectosymbioses</strong>, they never reach the extent found in endosymbioses. In some cases the partners<br />

may be separated and kept alive at least for some time. These characteristics allow us to make<br />

inferences on how these symbioses originated, something that is extremely difficult to do in highly<br />

evolved endosymbioses.<br />

In many <strong>thiotrophic</strong> symbioses where the method of transmission has been clarified, there is<br />

no evidence of vertical transmission from parents to offspring. Apparently, the symbionts must be<br />

acquired in each generation, most probably from the free-living bacterial community in the respective<br />

habitat. The mechanisms of recognition, attachment, and internalisation of the <strong>microbial</strong> partner<br />

are still unclear.<br />

Thiobiotic habitats<br />

Hydrothermal Vents<br />

Hydrothermal springs are found in all oceans along the central rift valley of mid-oceanic ridges.<br />

Here, sea water percolates through the newly formed crust several kilometres deep. When it comes<br />

in contact and reacts with hot rocks near the underlying magma chamber it undergoes profound<br />

chemical changes (Alt 1995, Von Damm 1995). Most importantly sulphate is reduced to sulphide,<br />

the water becomes anoxic and the concentrations of heavy metals increase dramatically. The heated<br />

and chemically altered fluid then rises and flows warm (a few degrees above ambient deepwater<br />

temperatures) to extremely hot (350–400˚C) from cracks in the basaltic rocks covering the floor<br />

of the rift valley or seeps through sediments. Mineral precipitates may form chimneys dozens of<br />

metres high, from which the hydrothermal fluid emanates as black clouds coloured by precipitating<br />

metal sulphides (Goldfarb et al. 1983).<br />

Chemolithoautotrophic bacteria and Archaea already grow in the chemical gradients within the<br />

conduits in the basaltic rocks (Karl et al. 1980). Where the hydrothermal fluid is injected into the<br />

oxygen-containing, cold, deep-sea water a profusion of <strong>microbial</strong> production occurs on the surface<br />

of rocks and sediments. Most spectacular, however, is the animal life around the hydrothermal<br />

vents, which solely depends on the production of the chemolithoautotrophic microbes (Van Dover<br />

2000). Whereas many animals are suspension feeders or simply graze the <strong>microbial</strong> turf, others<br />

live in symbiosis with special kinds of bacteria. Hydrothermal vents are unpredictable environments,<br />

where fluid flows may vary on short temporal scales (Johnson et al. 1988). Successful survival<br />

strategies here include associations either with mobile animals that may follow gradients or with<br />

large sessile organisms that provide the necessary milieu for the bacteria. Both endo- and <strong>ectosymbioses</strong><br />

are found in these bizarre environments. The most prominent examples are the vestimentiferan<br />

tube worms, bivalves such as Calyptogena spp. and Bathymodiolus spp., and the shrimp<br />

Rimicaris spp.<br />

Hydrothermal vents are not restricted to the deep sea but also occur in shallow water in<br />

conjunction with volcanism. None of the few shallow hydrothermal vents studied so far, however,<br />

showed such a highly specialised fauna (Dando et al. 1995).


2727_C04.fm Page 97 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 97<br />

Cold seeps<br />

Wherever the sea sediments are subjected to pressure, pore fluid is squeezed from the interstices<br />

and seeps from the surface. This seepage may occur under a variety of tectonic forces: in accretion<br />

prisms formed during subduction of one lithospheric plate under another, at the margin of mud<br />

volcanoes, in places where salt diapirs rise through continental margin sediments, or in connection<br />

with hydrocarbon seeps. The expelled fluids differ chemically from the surrounding sea water. They<br />

are anoxic and may contain methane or sulphide, hydrocarbons, or high salt concentrations, but<br />

not the high heavy metal concentrations typical for hydrothermal vents (Suess et al. 1987).<br />

Flow speeds are generally low but steady and sulphide concentrations are often at the detection<br />

limit near the sediment surface. In contrast to the vents, the sulphide here is of biological origin,<br />

essentially having been produced by <strong>microbial</strong> sulphate reduction (Carney 1994).<br />

Symbiotic biota associated with cold seeps include vestimentiferan and frenulate tube worms,<br />

clams and mussels among the macrofauna, and stilbonematid nematodes among the meiofauna.<br />

Some of the mussels contain both <strong>thiotrophic</strong> and methanotrophic endosymbionts, sometimes within<br />

the same bacteriocyte (Fisher 1990). Methanotrophs are also found in the frenulate Siboglinum<br />

poseidoni (Schmaljohann & Flügel 1987).<br />

Shallow sheltered sediments<br />

This is by far the largest <strong>thiotrophic</strong> habitat. It extends from intertidal sand and mudflats, marsh<br />

and mangrove sediments, over essentially all of the shelf sediments, to dysoxic basins and upperslope<br />

sediments. It occurs under an oxic surface layer of variable thickness, ranging from a few<br />

millimetres to several centimetres, from which it is separated by a chemocline – the redox potential<br />

discontinuity layer (RPD) (Fenchel & Riedl 1970). Within the RPD, electron acceptors for the<br />

oxidation of organic material change in sequence from oxygen to nitrate, ferric iron, manganese<br />

and sulphate. Bacterial sulphate reduction produces sulphide in the deeper layers. Upward diffusion<br />

of sulphide leads to its oxidation and a variety of microbes use the free energy of this oxidation<br />

process for carbon fixation (Jørgensen 1989). Since sediment layers containing sulphide may be<br />

separated from those containing the best electron acceptor (oxygen) by several millimetres to<br />

centimetres, microorganisms are at a disadvantage when the sulphide/oxygen gradient is weak.<br />

Some of the larger sulphur bacteria such as Beggiatoa and especially the giant spaghetti bacterium<br />

Thioploca are mobile enough to bridge the gap (Gallardo 1977). Similar to what has been observed<br />

for hot vents and cold seeps, bacteria have succeeded in finding hosts that provide them with both<br />

oxygen and sulphide. Among the macrofauna, several families of bivalves, such as the Lucinidae,<br />

Thyasiridae, and Solemyidae, have species containing sulphur-oxidising chemoautotrophic bacteria<br />

in their gills (Allen 1958, Southward 1986). In those sediments with interstitial spaces, a variety<br />

of protists and meiofauna animals have symbiotic bacteria, either as endosymbionts (catenulid<br />

flatworms, phallodrilid oligochaetes, nematodes of the genus Astomonema)<br />

(Giere 1996) or as<br />

ectosymbionts (ciliates, stilbonematid nematodes) (Ott 1996). Recently, a number of flagellates<br />

living in the soft sediment of a dysoxic basin have been described to harbour potentially chemoautotrophic<br />

ectosymbionts (Buck et al. 2000, Bernhard et al. 2000).<br />

Macrophyte debris<br />

In shallow waters the debris originating from marsh and mangrove plants, algae, or sea grass may<br />

accumulate (Fenchel 1970, Mann 1976). The decomposition of this organic matter creates sulphidic<br />

habitats of various spatial and temporal extents. Mangrove peat is a relatively stable substratum<br />

having internal sulphide concentrations of up to 4 mM (McKee 1993). Diffusion of sulphide into<br />

the overlying water creates a few-millimetre-thick sulphidic boundary layer, whereby sulphide flux<br />

is highest in recently disturbed patches on the peat surface (Ott et al. 1998). Loose macrophyte


2727_C04.fm Page 98 Wednesday, June 30, 2004 12:00 PM<br />

98 J. Ott, M. Bright & S. Bulgheresi<br />

debris is much less stable and predictable than mangrove peat. It may, however, repeatedly collect<br />

in defined places such as depressions in the vicinity of algae or sea grass stands or in crevices and<br />

caves among rocks.<br />

Thiotrophic <strong>ectosymbioses</strong> have been reported from mangrove peat and decomposing sea grass<br />

and algae. In all cases the host is a sedentary peritrich ciliate belonging to the genus Zoothamnium.<br />

Kentrophoros<br />

Hosts<br />

Ciliates<br />

About 20 species of the ciliate genus Kentrophoros inhabit sheltered <strong>marine</strong> sands having an RPD<br />

several centimetres beneath the surface (Fenchel & Finlay 1989). Soon after the description of the<br />

first species (Sauerbrey 1928) it was recognised that the dorsal (left) surface of the cells is covered<br />

by rod-shaped bacteria containing sulphur granules (Kahl 1935). Raikov (1971, 1974) suggested<br />

a chemolithoautotrophic nature of the bacteria and showed that the bacteria are phagocytised by<br />

the ciliate.<br />

Specimens of Kentrophoros are ribbon shaped or tubularly involuted and, in the case of K.<br />

fistulosus, can be up to 3 mm long, but are only 2–3 mm thick (Figure 1). The extremely flattened<br />

shape is interpreted as an adaptation to provide ample space for the bacterial symbionts; it increases<br />

the surface-to-volume ratio by a factor of 6–7 compared with other similarly sized ciliates. The<br />

ventral (or right) side bears cilia arranged in longitudinal rows. The cells have only a vestigial<br />

cytostome (Foissner 1995). Rod-shaped bacteria occupy the unciliated dorsal (or left) side of the<br />

cell, which may be tubularly involuted in the central region (Figure 2).<br />

The ciliates glide sluggishly between the sand grains. In the sediment they concentrate in the<br />

oxic–anoxic chemocline. In an artificial oxygen gradient they aggregate around 5% saturation,<br />

avoiding high oxygen tensions. The ciliates, however, do not react to sulphide but appear to be<br />

randomly distributed in a sulphide gradient in the absence of oxygen (Fenchel & Finlay 1989).<br />

Under experimental conditions and under the assumption that the ectosymbiotic bacteria constitute<br />

its sole food, Kentrophoros was calculated to have a doubling time of 18 h at room<br />

temperature, which is low compared with similarly sized ciliates. This difference was attributed to<br />

suboptimal culture conditions (Fenchel & Finlay 1989).<br />

Zoothamnium<br />

The sedentary colonial peritrich ciliate Zoothamnium niveum (Hemprich & Ehrenberg 1831,<br />

Ehrenberg 1838) was originally described from the Red Sea. Although the authors were struck<br />

by its white appearance, they could not attribute it to the presence of bacteria. The ciliate has<br />

been redescribed by Bauer-Nebelsick et al. (1996a) based on material collected from mangrove<br />

peat in the Belize Barrier Reef system. There it grows on vertical to overhanging walls of tidal<br />

channels, and lagoons cut into mangrove peat of backreef islands. Since then, Zoothamnium<br />

niveum has been found in the western Mediterranean near Calvi, Corsica, where it grows on<br />

decomposing subtidal accumulations of leaf debris and adjacent rocks near stands of the large<br />

Mediterranean sea grass Posidonia oceanica.<br />

Other reports of large, white Zoothamnium colonies<br />

come from the Florida Keys (Bauer-Nebelsick et al. 1996a), Lanzarote (Canary Islands)<br />

(Wirtz & Debelius 2003), Elba (own unpublished observations), Giglio (T. Pillen, personal<br />

communication), and Greece (G. Scatolin, personal communication). The up to 15-mm-long<br />

colonies are feather-shaped fans with alternating branches growing from a central stalk (Figure<br />

3). Stalk and branches contain a contractile spasmoneme that allows the colony to contract<br />

rapidly. Colony contraction occurs spontaneously or upon disturbance.


2727_C04.fm Page 99 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 99<br />

Figure 1 Light microscopy (LM) micrograph of living specimen of Kentrophoros fistulosus;<br />

scale bar = 500<br />

mm.<br />

Figure 2 Scanning electron microscopy (SEM) micrograph of midbody region with rods (r) on the<br />

dorsal side and cilia on the ventral side; scale bar = 20 mm. (Courtesy of W. Foissner.)<br />

Figure 3 LM micrograph of Zoothamnium niveum colony with stalk (s) and terminal zooid (t) on its tip and<br />

alternate branches with microzooids (mi) and macrozooids (ma); scale bar = 100 mm. Figure 4 SEM<br />

micrograph of contracted colony showing several microzooids (mi) and one macrozooid (ma) covered by<br />

symbionts; scale bar = 50 mm.


2727_C04.fm Page 100 Wednesday, June 30, 2004 12:00 PM<br />

100 J. Ott, M. Bright & S. Bulgheresi<br />

Large colonies sprout secondary fans and may have about 200 branches bearing up to 20 feeding<br />

microzooids each, adding up to over 3000 microzooids. The tips of the stalk and of still-growing<br />

branches bear nonfeeding, terminal zooids that divide by unequal longitudinal fission. On some<br />

branches the proximal zooid develops into a globular nonfeeding macrozooid that eventually<br />

detaches as a motile swarmer. Except for the noncontractile basal part of the stalk, branches and<br />

zooids are densely covered by bacteria (Figure 4) (Bauer-Nebelsick et al. 1996a). Zoothamnium<br />

niveum occurs on the surfaces of macrophyte debris and peat where sharp gradients between<br />

sulphide and oxygen are developed within a few millimetres. The ciliary action of the microzooids<br />

effectively mixes sulphidic and oxic water (Vopel et al. 2001, 2002). In addition, the colonies<br />

contract into the sulphidic boundary layer and subsequently expand again into the surrounding<br />

oxygen-containing water (Ott et al. 1998). The life cycle of Z. niveum has been studied through<br />

several generations in the laboratory. Using sulphide as a cue, the swarmers of Z. niveum actively<br />

seek and colonise patches in the environment with high sulphide flux, such as areas where the peat<br />

surface has been recently disturbed. Upon settling, each swarmer changes into a terminal zooid<br />

that produces a stalk and starts to divide into microzooids and terminal zooids, which in turn grow<br />

into branches. After about 4 hours the first branch is formed, and within 4 days maximum size is<br />

reached. During the period of exponential growth on the second and third days, the terminal zooids<br />

divide approximately once every hour (own unpublished data). The colonies continue to live to a<br />

mean age of 7 days, showing loss of microzooids from proximal branches as signs of senescence.<br />

Starting with a colony size of about 10 branches, macrozooids are produced and released. Swarmers<br />

may settle at the same spot or colonise a new sulphide patch. The growth data obtained under<br />

laboratory conditions fit those observed on the peat wall in the field. Z. niveum is usually found in<br />

groups of a few to several hundred colonies. Small groups typically consist of young colonies and<br />

newly settled swarmers, large groups mainly of senescent colonies. A patch exists for approximately<br />

20 days, as has been determined for a population at Twin Cayes, Belize (Ott et al. 1998).<br />

Nematoda (Stilbonematinae)<br />

Invertebrates<br />

Ectosymbiotic chemoautotrophic bacteria have been reported for a group of eight closely related<br />

genera of free-living nematodes within the family Desmodoridae (Chromadoria, Adenophorea),<br />

classified as the subfamily Stilbonematinae. Originally thought to be parts of the worm (Greeff<br />

1869), fungal spores (Chitwood 1936) or epibiotic cyanobacteria (Gerlach 1950, Wieser 1959),<br />

they have been finally identified as sulphide-oxidising chemoautotrophic bacteria that coat the<br />

nematode surface in a species-specific pattern (Ott et al. 1991).<br />

Stilbonematinae occur in all intertidal and subtidal porous sediments where an oxidised surface<br />

layer overlies a deeper, reduced, sulphidic body of sediments. They are most abundant in and near<br />

the RPD (Ott & Novak 1989). Highest abundance and diversity are found in tropical calcareous<br />

sands. Special habitats include continental slope brine seeps (Jensen 1986), the shallow-water vents<br />

in the Bay of Plenty (New Zealand) (Kamenev et al. 1993), and the reduced sediments accumulating<br />

among the roots of the surf grass Phyllospadix spp. and in mussel banks on the wave-beaten U.S.<br />

West Coast (own unpublished observations). Stilbonematinae have been reported from all major<br />

oceans, the Mediterranean, the Red Sea, the Caribbean and the North Sea.<br />

Adult sizes of the elongated cylindrical worms range from less than 2 mm to about 15 mm<br />

(Figure 5 and Figure 6). The external appearance and the construction of the worm cuticle are<br />

highly diverse (Urbancik et al. 1996a,b). Unifying characters are the weak or absent buccal armature<br />

and the special construction of the foregut (pharynx), where muscles appear to be concentrated in<br />

the anterior-most part and the remainder is mostly glandular (Hoschitz et al. 2001). These similarities<br />

could be interpreted to reflect convergent evolution due to the symbiotic lifestyle and the<br />

specialisation on a single food item. The monophyly of the Stilbonematinae, however, is clearly


2727_C04.fm Page 101 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 101<br />

Figure 5 SEM micrograph of Laxus cosmopolitus (Stilbonematinae) with rod-shaped symbionts; arrow points<br />

to symbiont-free area showing rows of setae; scale bar = 100 mm. Figure 6 SEM micrograph of Eubostrichus<br />

dianae (Stilbonematinae) with nonseptate filaments (f); scale bar = 100 mm.<br />

supported by the presence of a unique glandular sense organ in the epidermis (Nebelsick et al.<br />

1992, Bauer-Nebelsick et al. 1995) and it forms a distinct clade within the Desmodoridae according<br />

to both morphological and molecular (18S rDNA sequence) characters (Kampfer et al. 1998).<br />

The worms move sluggishly through the sediments and often coil up and remain stationary for<br />

several hours. Riemann et al. (2003) even propose a hemisessile life strategy for Leptonemella spp.<br />

from intertidal sediments in the North Sea.<br />

No representative of the Stilbonematinae has been cultivated in the laboratory so far. The larger<br />

species may be kept in sand buckets and even in dishes with sea water for many days up to several<br />

weeks, but neither moulting nor egg laying has been observed here. They probably grow slowly<br />

and have long intermoult periods, high life expectancy and few offspring. Juveniles are rare (Ott<br />

et al. 1995) and moulting stages have only occasionally been found in field samples. The slow and<br />

sluggish lifestyle fits with a basal metabolism that is among the lowest ever measured in nematodes<br />

(Schiemer et al. 1990).<br />

Crustacea (Rimicaris)<br />

A number of decapod carid shrimps regularly occur at hydrothermal vents. They were originally<br />

placed into the family Bresiliidae but later a separate family, Alvinocarididae, was proposed<br />

(Christoffersen 1986). While species of the genus Alvinocaris are widespread scavengers, members<br />

of the genus Rimicaris have only been reported at Mid-Atlantic Ridge hydrothermal sites ( R.<br />

exoculata;<br />

Williams & Rona 1986) and recently at a vent field on the Central Indian Ridge ( R.<br />

kairei;<br />

Watabe & Hashimoto 2002). A second Atlantic species, R. aurantiaca (Martin et al. 1997)<br />

proved to be juveniles of R. exoculata (Shank et al. 1998).<br />

The well-studied species R. exoculata occurs in enormous densities of up to 50,000 specimens<br />

m–2<br />

on solid surfaces where hydrothermal fluids emanate (Segonzac et al. 1993). Adult R.<br />

exoculata are 40-to 60-mm-long whitish shrimps (Figure 7). Originally thought to be grazers on<br />

surface-living bacteria (Van Dover et al. 1989) or suspension feeders (Jannasch et al. 1991), the<br />

conspicuous and regular epigrowth of bacteria on the mouthparts and the inner surface of the


2727_C04.fm Page 102 Wednesday, June 30, 2004 12:00 PM<br />

102 J. Ott, M. Bright & S. Bulgheresi<br />

Figure 7 LM micrograph of ventral and dorsal sides of Rimicaris exoculata specimens, approximately 40–60<br />

mm in length. (Courtesy of M. Segonzac and D. Desbruyères.) Figure 8 SEM of shrimp appendage; scale<br />

bar = 1 mm. (Courtesy of M.F. Polz.)<br />

modified carapace pointed to a symbiotic lifestyle (Gebruk et al. 1993). The carapace encloses<br />

the anterior body almost completely, forming voluminous chambers on either side. There is no<br />

rostrum and the first and second antennae are stout and strong. The exopodites of maxilla 2 and<br />

maxilliped 1 are greatly enlarged and densely covered with plumose setae (bacteriophores), which<br />

also occur on the proximal parts of the thoracic legs (Figure 8) (Gebruk et al. 1993, Segonzac<br />

et al. 1993, Casanova et al. 1993).<br />

The eyestalks are fused to form a large dorsal eye believed to be able to detect low levels of<br />

light emanating from vent chimneys (Van Dover et al. 1988, Van Dover & Fry 1994). This enables<br />

the shrimp to find the vents from a distance. A smooth cornea replaces the lenses of the compound<br />

eye, the photoreceptors in the fused retina are large with enlarged photosensitive regions, and the<br />

eye is underlain by a thick layer of white cells scattering light upwards (O’Neill et al. 1995, Nuckley<br />

et al. 1996, Chamberlain 2000). These modifications apparently sacrifice imaging ability in order<br />

to increase visual sensitivity. At close range the shrimp may additionally be guided by sensilla<br />

located on the second antennae, which show a concentration-dependent response to sulphide<br />

(Renninger et al. 1995).<br />

The shrimps form dense feeding swarms around hydrothermal chimneys and areas of “shimmering<br />

water”; some cling to the rock, forming layers several specimens thick and some move<br />

in and out of the thermal plumes. When dislodged by turbulence they rapidly move back to the<br />

chimneys. They ingest sulphide particles, attached bacteria and the bacteria growing on their<br />

cuticles. They are among the most important primary consumers and their ectosymbiotic <strong>thiotrophic</strong><br />

microbes are the dominant primary producers at certain sites (Van Dover 2002). In turn,<br />

Rimicaris<br />

is an important food for larger megafauna, such as macrourid and zoarcid fishes<br />

(Geistdoerfer 1994).<br />

The shrimp have small eggs and planktotrophic larvae (Ramirez Llodra et al. 2000). Juvenile<br />

shrimps have been found in midwater, where they spend an unknown period of time (Herring 1998).<br />

They are characterised by high amounts of wax esters as lipid reserves (Pond et al. 1977a, Allen<br />

et al. 1998, 2001). The fatty acid composition of these lipids points to a photosynthetic origin of


2727_C04.fm Page 103 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 103<br />

these substances. Such storage compounds are used during settlement and metamorphosis, while<br />

the shrimp develops the structures necessary to support the flora of ectosymbionts in adults (Gebruk<br />

et al. 2000).<br />

Microbial symbionts<br />

To date all <strong>microbial</strong> symbionts have resisted cultivation. The phylogenetic relationship of some<br />

of them to known bacteria was determined by 16S rRNA gene sequencing. The chemoautotrophic<br />

nature has been inferred on various grounds such as colour, ultrastructure, presence of ribulose-<br />

1,5-biphosphate carboxylase (RuBisCo, the enzyme necessary for carbon fixation), uptake of<br />

labelled inorganic carbon, presence of sulphur-oxidation enzymes and ecological data from the<br />

habitat.<br />

Kentrophoros<br />

According to Fenchel & Finlay (1989) the symbiotic bacteria are rod shaped, about 3.6 mm<br />

long<br />

and 0.8 mm<br />

wide. They appear brown to black in transmitted light due to sulphur inclusions that<br />

are contained in membrane-bound vesicles that occupy a large part of the cell volume. Among the<br />

cell organelles are probably carboxysomes, which contain RuBisCo. The rods are arranged perpendicular<br />

to the ciliate surface and show an unusual longitudinal division (Figure 9). They are<br />

embedded in a thick mucus layer produced by the ciliate that probably also covers the ciliated side<br />

(Foissner 1995). 14C<br />

incubations followed by autoradiography showed a carbon uptake rate that<br />

was equivalent to a doubling time of 5.3 h. The reduction of benzyl viologen in the presence of<br />

sulphide and uptake of 35S<br />

from labelled sulphide were indicative of sulphide oxidation. The bacteria<br />

are tightly packed with a density of 0.75 bacteria mm–2<br />

of ciliate surface. For a 170mm-long ciliate<br />

this amounts to 4500 bacteria with a total volume of 7650 mm3;<br />

this is roughly equivalent to the<br />

volume of the host or half of the volume of the symbiotic consortium. While K. fasciolata only<br />

contained one type of bacterium in transmission electron microscopy (TEM) sections, K. fistulosus<br />

also showed ectosymbiotic spirochaetes (Figure 9) (Foissner 1995) and K. latus intracellular<br />

prokaryotes of unknown function (Raikov 1974).<br />

Zoothamnium<br />

The bacteria found on stalk, branches, terminal zooids, and macrozooids are rod shaped, 1.4 mm<br />

long and 0.4 mm<br />

wide. They are attached along their longitudinal axis and are regularly arranged,<br />

resembling knitting patterns (Bauer-Nebelsick et al. 1996a). They completely cover the surface in<br />

a single layer except for the adhesive disc and the basal noncontractile part of the stalk. The rods<br />

also cover the basal (proximal) parts of the microzooids. Toward the peristomal disc the bacteria<br />

gradually change in shape, becoming more coccoid to slightly dumbbell shaped and growing larger<br />

(1.9 ¥ 1 mm)<br />

(Figure 10). Their arrangement becomes irregular and not all cells appear to be in<br />

contact with the microzooid surface. Especially when the microzooids are contracted the cocci<br />

seem to form more than one layer on the ciliate. The cocci have been observed to detach when the<br />

ciliates are active and become entrained in the feeding currents created by the paroral and adoral<br />

membranelles. Both extreme morphotypes are assumed to belong to the same species and represent<br />

a complex bacterial life cycle (rods/cocci coupled; Bright 2002). Both rods and cocci divide when<br />

they reach sizes of 2.2 and 2.6 mm,<br />

respectively.<br />

According to the 16S rRNA gene sequence, the bacteria belong to the g-proteobacteria<br />

(Molnar<br />

et al. 2000).<br />

The bacteria appear dark in transmitted light and pure white in incident light. When kept in<br />

seawater without supply of sulphide, the bacteria pale and eventually detach. Freshly collected Z.


2727_C04.fm Page 104 Wednesday, June 30, 2004 12:00 PM<br />

104 J. Ott, M. Bright & S. Bulgheresi<br />

Figure 9 Kentrophoros fistulosus with rod (r) and spirochaetes (s) on the dorsal body side; scale bar = 5 mm.<br />

(Courtesy of W. Foissner.) Figure 10 Zoothamnium niveum with cocci (c) on oral and rods (r) on aboral<br />

parts of the microzooids; scale bar = 10 mm. Figure 11 Stilbonema sp. with cocci (c); scale bar = 10<br />

mm. Figure 12 Laxus oneistus with rods (r); arrow points to dividing rod; scale bar = 5 mm. Figure 13<br />

Eubostrichus parasitiferus with nonseptate filaments (f) attached to the host’s cuticle in a spiral pattern; scale<br />

bar = 10 mm. Figure 14 Rimicaris exoculata with septate filaments (f) and rods (r); scale bar = 50 mm.<br />

(Courtesy of M.F. Polz) All figures are SEM micrographs.


2727_C04.fm Page 105 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 105<br />

niveum show a high rate of oxygen uptake of about 450 nl of O2<br />

mm–2<br />

colony surface. Within 4 h<br />

this drops to a sustained rate of 140–180 nl mm–2.<br />

Incubation of colonies, which have been kept<br />

in normoxic sea water for 24 h, in 100 mM<br />

sulphide resulted in a significant increase in respiration<br />

rate followed again by a subsequent decrease (Ott et al. 1998). The outer layer of the trilaminar<br />

cell envelope undulates in a manner similar to that in free-living thiobacilli. The cells contain large<br />

(diameter of 0.5 mm),<br />

electron-translucent, membrane-bound vesicles, which are indicative for<br />

elemental sulphur storage. Smaller (0.1 mm),<br />

electron-dense inclusions are interpreted as carboxysomes<br />

(Bauer-Nebelsick et al. 1996b), and RuBisCo has been found in the bacteria (H. Felbeck,<br />

personal communication). These physiological and morphological data, together with the ecological<br />

conditions, are strong evidence for a sulphide-oxidising chemoautotrophic nature of the symbionts.<br />

Density of bacteria is approximately 1.5 cells mm–2<br />

for rods and 0.5 cells mm–2<br />

for cocci. Since<br />

the bacteria-covered surface of a microzooid is 1900–2000 mm2,<br />

it supports 1900–2000 bacteria<br />

(assuming equal areas colonised by each morphotype). At an estimated volume of a microzooid of<br />

5700–6000 mm3<br />

and a bacterial volume (75% rods, 25% cocci) of 800–840 mm3,<br />

the <strong>microbial</strong><br />

symbionts amount to 12.1–12.3% of the volume of the symbiotic consortium. On stalks and<br />

branches the respective percentages are even lower, ranging between 5% on thinner and 2.5% on<br />

thicker parts.<br />

In old colonies, white filamentous bacteria grow on basal parts of the stalk and branches together<br />

with a diverse epigrowth of stalked bacteria and diatoms. This irregular fouling starts from the<br />

basal noncontractile part of the stalk and gradually extends to those parts of the stalk and branches<br />

where the symbiotic bacteria and microzooids have been lost (Bauer-Nebelsick et al. 1996a).<br />

Stilbonematinae<br />

A high diversity of ectosymbiotic bacteria is found within the Stilbonematinae. Form and size range<br />

from small (1–2 mm) cocci (Figure 11) through 2- to 5-mm-long rods (Figure 12) to nonseptate<br />

filaments of up to 100 mm in length (Figure 13 and Figure 14) containing approximately 50 nucleoids<br />

(DAPI staining; own unpublished observations). They appear dark brown to almost black in<br />

transmitted light and pure white in incident light due to sulphur inclusions contained in membranebound<br />

vesicles.<br />

Their arrangement on nematode cuticles may be genera or even species specific. In most cases<br />

they cover the whole body, leaving only the anterior-most part (head) and the tip of the tail free.<br />

In species of the nematode genus Eubostrichus the worm is entirely covered by bacterial filaments.<br />

In one species of the genus Laxus, L. oneistus, and in a yet undescribed species of the genus<br />

Catanema the bacterial coat starts a few 100 mm–1 mm posterior to the head at a defined level,<br />

where the diameter of the worm’s body decreases to accommodate the thickness of the bacterial<br />

coat (Figure 12). Several layers of cocci embedded in a gelatinous matrix surrounding the host’s<br />

body are typically found in species of the genera Stilbonema (Figure 11) and Leptonemella.<br />

Monolayers of rods are found in Laxus, Catanema, and some Leptonemella and Robbea species;<br />

in the latter genus two layers of rods in different orientations are present. In Laxus oneistus, L.<br />

cosmopolitus, and Catanema sp. the rods are arranged perpendicular to the worm cuticle and divide<br />

by longitudinal fission, a situation reminiscent of that in Kentrophoros. According to the 16S rRNA<br />

sequence the bacteria of Laxus oneistus belong to the g-proteobacteria (Polz et al. 1994).<br />

Uptake of 14 C-bicarbonate (Schiemer et al. 1990) and the presence of RuBisCo (Polz et al.<br />

1992) indicate an autotrophic nature of the bacteria. Uptake of 35 S-sulphide (Powell et al. 1979),<br />

the presence of sulphur metabolism key enzymes (ATP sulphurylase, sulphite-oxidase), and high<br />

amounts of elemental sulphur (Polz et al. 1992) have been demonstrated. The ultrastructure of the<br />

bacteria shows sulphur granules and possibly carboxysomes. Furthermore, d 13 C values of the<br />

symbiotic consortium were –24.9 to –27.5, which is similar to animals with sulphur-oxidising<br />

endosymbionts or thiobacilli (Ott et al. 1991).


2727_C04.fm Page 106 Wednesday, June 30, 2004 12:00 PM<br />

106 J. Ott, M. Bright & S. Bulgheresi<br />

An 8-mm-long Stilbonema majum with a diameter of 60 mm covered by a mucus sheath of 7.5<br />

mm thickness and containing 10 layers of 1.3 ¥ 0.6 mm cocci carries about 21 ¥ 10 6 bacteria. This<br />

makes up 22% of the volume of the symbiotic consortium. In Laxus oneistus the density of the<br />

upright rods is approximately 3.5 cells mm –2 . A 9-mm-long male with a diameter of 50 mm and an<br />

8-mm-long bacterial coat is covered by 4.5 ¥ 10 6 rods (size 2.1 ¥ 0.6 mm each). This represents<br />

12.3% of the consortium volume.<br />

In the genus Eubostrichus, two types of arrangement of the bacteria are found: in E. parasitiferus<br />

and several similar undescribed species the bacteria are crescent-shaped nonseptate filaments, 0.6<br />

¥ 30 mm in size, that are attached to the cuticle with both ends oriented parallel to the worm’s<br />

longitudinal axis. About 80 bacteria are arranged in a spiral fashion around the circumference of<br />

each worm, giving it the appearance of a rope. In cross section the bacteria appear to form several<br />

layers, when, in fact, all bacteria are in contact with the worm surface. A 3-mm-long E. parasitiferus<br />

with a diameter of 20 mm carries about 8000 bacteria, which make up only 7% of the volume of<br />

the symbiotic consortium, despite their spectacular appearance. In E. dianae the bacteria form up<br />

to 120-mm-long and 0.4-mm-thick filaments, which are attached to the cuticle by one end (Figure<br />

6) and form a dense fur-like coat that in live worms appears nicely groomed. With a size similar<br />

to that of E. parasitiferus, E. dianae carries 40–60 ¥ 10 3 filaments, contributing substantially<br />

(36–44%) to the consortium volume. The dense <strong>microbial</strong> coat is colonised by additional bacterial<br />

epibionts (Polz et al. 1999a).<br />

Rimicaris<br />

Three morphological types of bacteria have been described from Rimicaris by various authors (Van<br />

Dover et al. 1988, Casanova et al. 1993, Gebruk et al. 1993, Polz & Cavanaugh 1995): rods with<br />

a diameter from 0.2–0.4 mm and 0.5–3 mm in length and two kinds of filaments, a rare form with<br />

a diameter of 0.2–0.5 mm and a common larger form with a diameter of 0.8–3 mm (Figure 14).<br />

Several septate filaments grow from a common basal attachment disc and, in the large form, may<br />

attain a length of 1.5 mm. The filaments consist of cylindrical cells of approximately the same<br />

length as their diameter. These bacteria densely cover the inner surface of the extended carapace<br />

and also cover the bacteriophores on the enlarged exopodites of the second maxilla and first<br />

maxilliped and on the bases of the thoracic appendages. The rods co-occur with the filaments and<br />

are especially abundant in juveniles. They are attached along their whole length to the cuticle of<br />

the shrimp.<br />

Using a 16S rRNA-specific fluorescent hybridisation probe, Polz & Cavanaugh (1995)<br />

demonstrated that all three morphotypes belong to the same phylotype of e-proteobacteria. A<br />

number of parasites, but also sulphur bacteria such as Thiovolum sp., are found among the eproteobacteria.<br />

Recently, bacteria related to Rimicaris symbionts have been detected with<br />

molecular methods in <strong>marine</strong> anoxic water and sediments (Madrid et al. 2001, Lee et al. 2001).<br />

Elemental sulphur within the cells and RuBisCo activity (Gebruk et al. 1993) strongly suggest<br />

a <strong>thiotrophic</strong> nature of the bacteria. Polz & Cavanaugh (1995) estimate that an average shrimp<br />

may carry 8.5 ¥ 10 6 bacteria.<br />

Mutual benefits<br />

The consensus is that the above symbioses are largely nutritional. On one hand, the bacteria provide<br />

organic matter from their own primary chemoautotrophic production. On the other hand, the<br />

eukaryote partner facilitates access to reduced sulphur compounds and electron acceptors, which<br />

may be separated in space and time. Since sulphide is toxic for aerobic metazoans, the idea has<br />

been proposed that the bacteria may act as a detoxification mechanism, oxidising sulphide into<br />

elemental sulphur and finally to sulphate (Somero et al. 1989). Powell (1989) argued that in


2727_C04.fm Page 107 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 107<br />

meiofauna, due to the small size of the animals, sulphur detoxification could not work and they<br />

should be sulphide insensitive. In fact, high concentrations of thioles can be found in the tissues<br />

of Stilbonematinae (Hentschel et al. 1999). A certain detoxification may nevertheless be provided,<br />

because the bacterial symbionts increase heat tolerance in Stilbonematinae in the presence of<br />

simultaneous sulphide stress (Ott 1995). Sulphide detoxification probably occurs in sulphideoxidising<br />

bodies in the gills of Rimicaris (Compere et al. 2002) without any involvement of the<br />

ectosymbionts.<br />

Nutrition<br />

The limited success in culturing <strong>thiotrophic</strong> symbioses has precluded direct evidence that the<br />

animals feed on the symbiotic bacteria and may grow with them as their sole food. The presence<br />

of bacteria in feeding and digestive vacuoles has been reported for Kentrophoros (Raikov 1971,<br />

1974, Fenchel & Finlay 1989) and Zoothamnium niveum (Bauer-Nebelsick et al. 1996b) and in the<br />

gut lumen in several species of Stilbonematinae (Wieser 1959, Ott & Novak 1989, Riemann et al.<br />

2003). In all cases the bacteria have the distinct morphology and fine structure of the respective<br />

ectosymbionts and dominate the content of the digestive tract.<br />

Transfer of labelled carbon from symbionts to host has been shown for Z. niveum by Rinke<br />

(2002) in pulse and chase experiments. Most evidence for a nutritional interaction comes from<br />

studies of natural tracers in food chains, such as fatty acids and stable isotopes. High proportions<br />

of n-4 fatty acids, which are indicative for a bacterial origin, have been found in tissues of Rimicaris,<br />

whereas in the closely related genera Alvinocaris and Mirocaris, n-3 fatty acids characteristic for<br />

photosynthetically derived carbon are more abundant (Pond et al. 1997b,c). Muscles of Rimicaris<br />

contain n-7 fatty acids, which are closer in d 13 C (–13‰) to the ectosymbionts (–12‰) than to<br />

bacteria scraped from hydrothermal chimneys (–21‰) (Rieley et al. 1999). As a relic of its early<br />

planktotrophic life, however, Rimicaris contains high levels of polyunsaturated fatty acids in storage<br />

compounds (wax esters) in reproductive tissues. These fatty acids are thought to be important for<br />

reproduction in Crustacea (Pond et al. 2000, Allen et al. 2001). Gebruk et al. (1993) report d 13 C<br />

values for various tissues ranging from –10.5 to –12.5‰, which is similar to values reported for<br />

other hydrothermal vent animals. Stilbonematinae have d 13 C of –25.9‰ without and –24.9 to<br />

–27.5‰ with their symbionts, whereas nonsymbiotic nematodes and detritus from the same habitat<br />

show values of –10.3 and –10.5‰, respectively (Ott et al. 1991).<br />

In some cases the biomass of the bacterial symbionts dominates the food availability in the<br />

environment (Gebruk et al. 1993, Van Dover 2002). Polz & Cavanaugh (1995) estimated that at a<br />

density of 25,000 shrimp m –2 the number of symbiotic bacteria attached to Rimicaris would be 2.1<br />

¥ 10 11 , which is almost three orders of magnitude higher than the 4.9 ¥ 10 8 bacteria attached to a<br />

square meter of sulphide chimney surface.<br />

Dependence on a special resource is implied by morphological changes in feeding structures,<br />

such as the near disappearance of a functional mouth in Kentrophoros (Foissner 1995) or the<br />

reduction in dentition and the weakening and rearrangement of pharynx musculature in the Stilbonematinae<br />

(Hoschitz et al. 2001).<br />

Fenchel & Finlay (1989) calculated that the bacterial production in Kentrophoros allows a<br />

doubling time of the ciliate of 18 h, which is low for a ciliate of this size, but would allow the<br />

protozoan to grow with the symbionts as a unique food source. No such data are available for the<br />

other symbioses. Zoothamnium niveum, however, can maintain the high growth speed to reach the<br />

large size only in the presence of the bacteria. The growth rate and maximum size reached by<br />

aposymbiotic colonies reared from macrozooids that had lost their bacteria is only about 10% of<br />

those of symbiotic colonies. Stilbonematinae have an extremely low metabolism (Schiemer et al.<br />

1990) and could probably be easily sustained on the production of their bacteria. Note also that<br />

the bacteria appear to divide more rapidly on the margin of presumed feeding patches on the<br />

nematode cuticle (Polz et al. 1992).


2727_C04.fm Page 108 Wednesday, June 30, 2004 12:00 PM<br />

108 J. Ott, M. Bright & S. Bulgheresi<br />

Access to sulphide and electron acceptors<br />

The main benefit for the bacteria is apparently that the association with a mobile host allows them<br />

to exploit redox gradients that would otherwise not be available for their growth. The sharpness of<br />

the gradient and the distances covered by the carrier host vary greatly (Figure 15 to Figure 17).<br />

Zoothamnium niveum exploits the sharpest gradient, which develops within a boundary layer<br />

between a sulphide source and ambient water on the surface of macrophyte debris accumulations<br />

(Figure 15). This boundary layer is only a few millimetres thick but oxygen concentrations drop<br />

from near saturation to virtually zero while sulphide concentrations increase from undetectable to<br />

several hundred micromolar (Ott et al. 1998, Vopel et al. 2001). The ciliates effectively mix sulphidic<br />

and oxygen-containing water and create high current speeds (up to 11 mm s –1 ) over the zooid<br />

surfaces. Rapid contractions with speeds up to 520 mm s –1 exchange water along the colony, whereas<br />

slow expansions (300 mM make the<br />

worms turn around and crawl in the direction of the surface. In this ping-pong fashion, worms<br />

alternately visit deeper sulphidic layers and surface layers with abundant electron acceptors (oxygen,<br />

nitrate) (Ott et al. 1991). Schiemer et al. (1990) showed that the symbiotic bacteria can take<br />

advantage of this behaviour by storing reduced sulphur compounds and oxidising them upon<br />

availability of electron acceptors.<br />

Rimicaris is the most mobile carrier host, moving rapidly in and out of hydrothermal fluid<br />

plumes where sulphidic and oxic water mixes (Figure 17). In addition, the ventilation of the chamber<br />

formed by the enlarged carapace with the large scaphognathites is thought to create a favourable<br />

environment for chemoautotrophic growth. Polz et al. (1999b, 2000) have put forward an interesting<br />

Figure 15 Model of access to sulphide and oxygen in <strong>ectosymbioses</strong>. Zoothamnium niveum growing on peat<br />

wall has access to sulphide leaking from the peat when contracted and access to oxygen from the overlying<br />

sea water when expanded; scale bar = 1 mm. Figure 16 Model of access to sulphide and oxygen in<br />

<strong>ectosymbioses</strong>. Interstitial Kentrophoros spp. and Stilbonematinae migrate between sulphidic deeper sediment<br />

layers and oxic surface layers; scale bar = 1 cm. Figure 17 Model of access to sulphide and oxygen in<br />

<strong>ectosymbioses</strong>. Rimicaris exoculata swims in and out of sulphidic hydrothermal fluid and oxic ambient deepsea<br />

water; scale bar = 1 m.


2727_C04.fm Page 109 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 109<br />

hypothesis for the predominance of the symbiotic bacterial phylotype in the free-living community<br />

attached to the sulphidic rocks, where it contributes to >60% of the bacterial nucleic acids recovered<br />

from scrapings (Polz & Cavanaugh 1996). The constant inoculum through bacteria falling off the<br />

shrimps provides additional positive feedback to the free-living population of the symbiotic bacteria,<br />

helping it to rapidly reach high cell numbers and to outcompete other bacterial species. In turn,<br />

the abundance and stability of the free-living stock increase the chance that newly arrived juveniles<br />

or freshly moulted individuals will pick up the right symbiont.<br />

Maintenance and evolution of <strong>thiotrophic</strong> <strong>ectosymbioses</strong><br />

A vertical transmission to subsequent generations has only been described in the ciliates. When a<br />

Kentrophoros cell divides, both daughter cells get their share of the symbiotic bacteria (W. Foissner,<br />

personal communication). In Zoothamnium the motile macrozooids (swarmers) are covered with<br />

rods when they detach (Bauer-Nebelsick et al. 1996a), with a typical swarmer of 150 mm in diameter<br />

carrying about 90,000 bacteria. During colony growth after settlement, all surfaces, except for the<br />

basal 300–500 mm of the stalk, remain covered by the bacterial coat, which keeps pace with the<br />

production of new surfaces by the terminal zooids.<br />

In the nematodes there is no evidence of vertical transmission from parents to offspring.<br />

Nevertheless, even very small (stage 1) juveniles have been observed to carry a complete <strong>microbial</strong><br />

coat (own unpublished observations). Since all ectosymbionts are attached to the worm cuticle,<br />

they are shed when the host moults. In the nematodes this usually occurs four times in a life cycle.<br />

Moulting stages of several species have been observed where all bacteria are left behind on the<br />

exuvia (Wieser 1959, own unpublished observations). Recolonisation, however, must be a rapid<br />

process, since aposymbiotic Stilbonematinae are rarely found in field collections. These facts argue<br />

in favour of an environmental transmission with immediate colonisation of newly hatched and rapid<br />

recolonisation of moulted worms. In the case of Laxus oneistus mannose-binding lectins expressed<br />

on its cuticle could enable it to specifically select bacteria from the surrounding sediment<br />

(Nussbaumer et al. 2004).<br />

Larvae and juveniles of Rimicaris do not carry the symbiotic bacteria and feed on photosynthetic<br />

microplankton (Pond et al. 1997b, Dixon et al. 1998). When returning to the vents they must acquire<br />

the bacteria to populate the growing morphological structures that support the symbiotic population<br />

of adult shrimp. The mechanism of this process is unknown. As in nematodes, the shrimps must<br />

moult in order to grow and therefore repeatedly lose the symbionts during their life. The high<br />

specificity of the symbiotic association suggests a precise recognition mechanism as postulated for<br />

the Stilbonematinae.<br />

Comparison with close relatives allows inferences about the evolution of the symbioses. There<br />

are a number of prerequisites for the formation of such close associations. The ancestors of today’s<br />

partners must have lived in the same environment, and loose, nonobligate relationships must have<br />

preceded the tight and specific bonds of today. Morphological and physiological characters must<br />

have preadapted the future partners for the ability to make and maintain contact, and behavioural<br />

traits were necessary to select one of the many possible partners.<br />

The most complete line of evidence exists for the Stilbonematinae symbiosis. Species of the<br />

nematode family Desmodoridae are characteristic for oxygen-poor layers of <strong>marine</strong> sediment. In<br />

fact, the species Spirinia gnaigeri has the lowest weight-specific respiration rate of all <strong>marine</strong> freeliving<br />

nematodes (Schiemer et al. 1990). Occurring in or close to the RPD, they share the habitat<br />

with sulphur-oxidising chemoautotrophic bacteria, on which they probably feed together with other<br />

bacteria occurring in the sand. Furthermore, desmodorids are frequently fouled by a diverse<br />

assemblage of microorganisms, including a variety of bacterial morphotypes, stalked diatoms, and<br />

suctorians (Ott 1996). This is in contrast to most other nematode taxa, which rarely show <strong>microbial</strong><br />

epigrowth. We may expect that sulphur bacteria were among this fortuitous <strong>microbial</strong> cover on the


2727_C04.fm Page 110 Wednesday, June 30, 2004 12:00 PM<br />

110 J. Ott, M. Bright & S. Bulgheresi<br />

ancestors of the Stilbonematinae. Moreover, behavioural traits such as migration through the<br />

chemocline — possibly to feed on dissolved organic matter (Riemann et al. 1990) — no doubt<br />

selected for the sulphur bacteria because they not only tolerated but benefited from alternate<br />

exposure to sulphide and oxygen. In the case of weak or unstable chemoclines, the association<br />

with a mobile host may more than compensate for the grazing loss to the worms.<br />

The Zoothamnium niveum symbiosis may have a similar history. Members of the large genus<br />

Zoothamnium occur in a great variety of <strong>marine</strong> and limnic habitats, including those low in oxygen.<br />

Zoothamnium alternans, a close relative of Z. niveum, is regularly found in the same habitat.<br />

Zoothamnium species are notorious for <strong>microbial</strong> fouling. For the association between Z. pelagicum<br />

and cyanobacteria, a symbiotic relationship has even been suggested (Laval-Peuto & Rassoulzadegan<br />

1988). The spontaneous contractions of the colonies, which are typical for many sessile peritrich<br />

ciliates, and the feeding currents that mix sulphidic and oxic water provided the selective force for<br />

the association with sulphur bacteria.<br />

The Alvinocarididae exhibit several modes of nutrition. While Alvinocaris appears to be an<br />

unspecialised scavenger and predator, Chorocaris feeds on a mixed diet and has already modified<br />

mouthparts overgrown with bacteria, although not to the extent of Rimicaris, which presumably is<br />

an obligate bacteriovore. The high attractiveness of chemoautotrophic primary production in the<br />

otherwise food-limited deep sea may have brought the ancestors of Rimicaris and its symbiont<br />

together.<br />

Suspected symbioses<br />

In addition to the symbioses described above, a great variety of associations between protists or<br />

invertebrates and ectosymbiotic bacteria have been described from <strong>marine</strong> sulphidic habitats. In<br />

most of these cases the nature of the bacteria and their function in the symbiosis are unknown.<br />

Among the protists, epibiotic bacteria in a regular arrangement suggesting a symbiotic association<br />

have been described from flagellates and several ciliates. Euglenozoans from a Monterey<br />

Bay cold seep and the dysoxic Santa Barbara Basin (California) are densely covered by rod-shaped<br />

bacteria. Putative sulphur vesicles in the bacteria and the presence of sulphide in the sediment<br />

suggest a sulphur-oxidising metabolism for the symbionts. No evidence for ingestion of the<br />

bacteria by the protists has been found (Buck et al. 2000, Bernhard et al. 2000). Several cases of<br />

ciliates belonging to the genera Parablepharisma, Metopus, Caenomorpha, and Sonderia having<br />

ectobiotic bacteria have been reported from anoxic and sulphidic sediments (Fenchel et al. 1977).<br />

The bacteria are in all cases curved rods and probably utilise products of the ciliates’ fermentative<br />

metabolism. Density varies from 1,000 to 100,000 bacteria per ciliate. In the ciliate Geleia fossata<br />

from a tidal flat on the U.S. East Coast, short Gram-negative rods are positioned in and along the<br />

ciliated grooves. They are embedded in deep cell membrane invaginations and some seem to be<br />

enclosed in membrane-bound vesicles. There is no indication that the bacteria are chemolithoautotrophs<br />

as in the closely related genus Kentrophoros, and their low density and biomass (2–10<br />

¥ 10 3 per ciliate, amounting to only a few percent of the ciliate biovolume) make a nutritive<br />

dependence unlikely. Several other related ciliates belonging to the genera Tracheloraphis,<br />

Paraspathidium, Loxophyllum, and Cyclidium from the same habitat showed scattered ectobiotic<br />

bacteria, indicating that associations with bacteria are probably widespread in <strong>marine</strong> sediments<br />

(Epstein et al. 1998).<br />

Among the invertebrates, a yet undescribed flatworm belonging to the Monocelidae (Proseriata,<br />

Platyhelminthes) dominates the meiofauna of sediments at a temperature of 30–40˚C under the<br />

influence of shallow-water hydrothermal vents on Deception Island, Antarctica. The outer surface<br />

of the worm is covered by a single morphotype of straight to slightly curved rods (0.7 ¥ 2 mm)<br />

embedded in mucus at the level of the cilia tips. There is no indication for a chemoautotrophic<br />

nature of the bacteria because tests for RuBisCo were negative. Ectobiotic bacteria have been


2727_C04.fm Page 111 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 111<br />

occasionally reported for a number of free-living <strong>marine</strong> flatworms, but no indication of specificity<br />

and persistence has been given. The uniformity of the bacteria and the fact that all specimens<br />

carried the same dense coat suggest a symbiotic relationship of an as yet unknown function in the<br />

Antarctic monocelid (Bright et al. 2003).<br />

Similarly, the small bacteria that apparently represent a single morphotype covering the entire<br />

surface of the annelid Xenonerilla bactericola (Nerillidae) from the dysoxic Santa Barbara Basin<br />

off California are suspected to be symbionts. The nature of this symbiosis also remains unresolved<br />

(Müller et al. 2001). In the same habitat the nematode Desmodora masira regularly has epibiotic<br />

bacteria (Bernhard et al. 2000) a feature that is characteristic for many species of the family<br />

Desmodoridae, to which the Stilbonematinae also belong (Ott 1996).<br />

In the best-studied case, the hydrothermal vent annelid Alvinella pompejana, three morphotypes<br />

of bacteria occur on epidermal expansions and cuticular protrusions on the intersegmental<br />

spaces. The cylindro-conical expansions, which may be 1 cm long, are peculiar to Alvinella. They<br />

are glandular, not cuticularised, and are covered by two types of filamentous bacteria. Two<br />

phylotypes of e-proteobacteria have been identified (Haddad et al. 1995, Cary et al. 1997). Whereas<br />

one of these is also a major component of the free-living bacterial community, the other is<br />

exclusively found on Alvinella and on the inside of its tube. Although uptake of inorganic carbon<br />

and RuBisCo activity have been reported (Alayse-Danet et al. 1986) their low levels make an<br />

important contribution of autotrophs unlikely. The presence of bisulphate reductase genes in<br />

bacteria suggests that they are anaerobic sulphate reducers (Cottrell & Cary 1999) that probably<br />

play a role in the formation of the typical “white smokers” associated with the presence of<br />

Alvinella spp. Recently, abundant spirochetes (Campbell & Cary 2001) and a vibrio (Raguenes<br />

et al. 1997) that probably are heterotrophs have been detected among the ectosymbionts by<br />

molecular methods.<br />

It is unclear to what extent A. pompejana feeds on its episymbiotic bacteria. Stable isotope<br />

data point to a bacterial food source (Desbruyàres et al. 1983) but behavioural observation and gut<br />

content analysis suggest that grazing on the tube may be the usual mode of nutrition in this worm.<br />

The modifications found on those segments of Alvinella that carry bacteria are evidence for a close<br />

relationship between the microbes and their host. The specific nature of the association — be it<br />

nutritional or related to sulphide detoxification — has yet to be elucidated (Alayse-Danet et al.<br />

1987). The biology of A. pompejana, which is an early coloniser of vent chimneys and has a<br />

remarkably high temperature tolerance, has been exhaustively summarised by Desbruyàres et al.<br />

(1998).<br />

The priapulid Halicryptus spinulosus has a modified outer cuticle layer forming minute ridges<br />

that greatly enlarge the cuticular surface. The crevices formed by these ridges are densely populated<br />

by bacteria of three distinct morphological types embedded in mucus in a characteristic arrangement.<br />

Cluster-forming bacteria deep in the crevices precipitate iron probably as Fe-sulphide.<br />

Oxidation with iron may bind small amounts of sulphide, and this may help the worm to survive<br />

intertidal exposure until metabolic adaptations take over. The bacteria may be of potential nutritional<br />

significance. No indication, however, of transepidermal transport was observed (Oeschger & Janssen<br />

1991, Oeschger & Schmaljohann 1988).<br />

Among the Crustacea, stalked barnacles (“Lau A,” Scalpellomorpha, Neolepadinae) from hydrothermal<br />

vents have dense, elongated cirral setae that are covered by white bacterial filaments. The<br />

filaments have a diameter of 1.15 mm, are composed of cells approximately 1 mm long, and may<br />

be up to 220 mm long. The mandibles appear modified compared with other stalked vent barnacles,<br />

probably to facilitate combing of bacteria from the cirri. Except for the white colour, no other<br />

indication of a sulphur metabolism of the bacteria is given. Ecological observations showing the<br />

barnacles extending their cirri into diffuse hydrothermal flow could be taken to indicate a chemoautotrophic<br />

nature of the bacteria (Southward & Newman 1998).


2727_C04.fm Page 112 Wednesday, June 30, 2004 12:00 PM<br />

112 J. Ott, M. Bright & S. Bulgheresi<br />

Summary and outlook<br />

The importance of the <strong>thiotrophic</strong> <strong>ectosymbioses</strong> varies greatly between the different cases. Fenchel<br />

& Finlay (1989) calculated that the biomass of bacteria associated with specimens of Kentrophoros<br />

in their sediments is only 3.7 ¥ 10 –4 g m –2 compared with a biomass of free-living sulphur-oxidising<br />

bacteria of 5–20 g m –2 . It may therefore be merely “an exotic phenomenon which makes only a<br />

symbolic contribution” (Fenchel & Finlay 1989). In Zoothamnium niveum it may be important at<br />

a microspatial scale: with an estimated weight of the bacteria on an average colony of 1000<br />

microzooids of 1 mg, the symbionts would contribute only 1.2 mg m –2 (assuming 1200 colonies;<br />

Ott et al. 1998). Because the colonies are highly aggregated, this value increases 100-fold when<br />

only the patches are taken into account (assuming 100 colonies on an area of 10 cm 2 ). In tropical<br />

calcareous sands, where Stilbonematinae are a dominant element of the meiofauna (Ott & Novak<br />

1989), the weight of the symbiotic bacteria may be in the same order of magnitude as that of freeliving<br />

sulphur bacteria. At a density of about 300 ¥ 10 3 worms m –2 (consisting of equal numbers<br />

of Laxus oneistus and Stilbonema majum, carrying on average half of the bacteria calculated for<br />

an adult worm), the combined weight of the symbiotic bacteria would amount to 0.7–0.8 g. In the<br />

case of Rimicaris the ectosymbionts are (with estimated 2.1–4.2 ¥ 10 11 bacterial cells m –2 ) significantly<br />

more abundant than the sulphur-oxidising bacteria attached to the chimney surface (approximately<br />

4.9 ¥ 10 8 m –2 ), but also dominate the water close to the chimney, which contains about 5<br />

¥ 10 8 bacteria l –1 .<br />

Sulphide symbioses are apparently a frequent outcome of the various associations of protists<br />

and invertebrates with bacteria. The most common type of association leading to <strong>ectosymbioses</strong> is<br />

fouling of surfaces, which originally involved a diversity of microbes, microalgae, and protists.<br />

Evidence for this evolutionary intermediate stage is the many cases of irregular epigrowth reported<br />

from ciliates, including relatives of Kentrophoros and Zoothamnium niveum, and nematodes that<br />

are closely related to the Stilbonematinae. These fouled organisms may be regarded as models for<br />

the ancestors of extant symbioses and apparently lack the ability to keep their surface free of<br />

epibionts, which at high densities may become a nuisance (Ott 1996). This imperfection, however,<br />

was a necessary precondition for the evolution of a successful mutualistic interaction. There is little<br />

evidence pertaining to the age of <strong>thiotrophic</strong> symbioses. Most of the above-described hosts do not<br />

fossilise well enough to leave an indication of bacterial symbioses. Only in the arthropods is there<br />

some evidence that a <strong>microbial</strong> and possibly <strong>thiotrophic</strong> symbiosis existed in the Ordovician olenid<br />

trilobites. These now extinct animals lived on reduced sediments, probably in dysoxic and sulphidic<br />

bottom water. Their extended carapace has been interpreted as an incubation chamber for sulphur<br />

bacteria, similar to that of Rimicaris (Fortey 2000).<br />

The spectacular endosymbioses of the Vestimentifera and the large clams and mussels at hot<br />

vents have directed much attention to these conspicuous animals. Until the discovery of the<br />

Rimicaris symbiosis the study of <strong>ectosymbioses</strong> had not been pursued with the same effort. There<br />

is, however, still much to discover both in shallow water and in the deep sea that will shed new<br />

light on the fascinating functioning and evolution of <strong>thiotrophic</strong> symbioses.<br />

References<br />

Alayse-Danet, A.M., Desbruyères, D. & Gaill, F. 1987. The possible nutritional or detoxification role of the<br />

epibiotic bacteria of alvinellid polychaetes: review of current data. Symbiosis 4, 51–62.<br />

Alayse-Danet, A.M., Gaill, F. & Desbruyères, D. 1986. In situ bicarbonate uptake by bacteria-Alvinella<br />

associations. Pubblicazioni della Stazione Zoologica di Napoli. I. Marine Ecology 7, 233–240.<br />

Allen, C.E. 1958. On the basic form and adaptations to the habitat in Lucinaceae (Eulamellibranchia).<br />

Philosophical Transactions of the Royal Society of London Biological Sciences 241, 421–484.<br />

Allen, C.E., Copley, J.T. & Tyler, P.A. 2001. Lipid partitioning in the hydrothermal vent shrimp Rimicaris<br />

exoculata. Pubblicazioni della Stazione Zoologica di Napoli. I. Marine Ecology 22, 241–253.


2727_C04.fm Page 113 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 113<br />

Allen, C.E., Copley, J.T., Tyler, P.A. & Varney, M. 1998. Lipid profiles of hydrothermal vent shrimps. Cahiers<br />

de Biologie Marine 39, 229–231.<br />

Alt, J.C. 1995. Subsea floor processes in mid-ocean ridge hydrothermal systems. In Sea-Floor Hydrothermal<br />

Systems: Physical, Chemical, Biological and Geological Interactions, S.E. Humphris et al. (eds).<br />

Washington, DC: American Geophysical Union, Geophysical Monograph 91, 85–114.<br />

Bauer-Nebelsick, M., Bardele, C.F. & Ott, J.A. 1996a. Redescription of Zoothamnium niveum (Hemprich &<br />

Ehrenberg, 1831) Ehrenberg 1838 (Oligohymenophora, Peritrichida), a ciliate with ectosymbiotic,<br />

chemoautotrophic bacteria. European Journal of Protistology 32, 18–30.<br />

Bauer-Nebelsick, M., Bardele, C.F. & Ott, J.A. 1996b. Electron microscopic studies on Zoothamnium niveum<br />

(Hemprich & Ehrenberg, 1831) Ehrenberg 1838 (Oligohymenophora, Peritrichida), a ciliate with<br />

ectosymbiotic, chemoautotrophic bacteria. European Journal of Protistology 32, 202–215.<br />

Bauer-Nebelsick, M., Blumer, M., Urbancik, W. & Ott, J.A. 1995. The glandular sensory organ of Desmodoridae<br />

(Nematoda): ultrastructure and phylogenetic implications. Invertebrate Biology 114, 211–219.<br />

Bernhard, J.M., Buck, K.R., Farmer, M.A. & Bowser, S.S. 2000. The Santa Barbara Basin is a symbiosis<br />

oasis. Nature 403, 77–80.<br />

Bright, M. 2002. Life strategies of <strong>thiotrophic</strong> <strong>ectosymbioses</strong>. In The Vienna School of Marine Biology: A<br />

Tribute to Jörg Ott, M. Bright et al. (eds). Wien: Facultas Universitätsverlag, pp. 19–32.<br />

Bright, M., Arndt, C., Keckeis, H. & Felbeck, H. 2003. A temperature-tolerant interstitial worm with associated<br />

epibiotic bacteria from the shallow water fumaroles of Deception Island, Antarctica. Deep-Sea<br />

Research II 50, 1859–1871.<br />

Buck, K.R., Barry, J.P. & Simpson, A.G.B. 2000. Monterey Bay cold seep biota: Euglenozoa with chemoautotrophic<br />

bacterial epibionts. European Journal of Protistology 36, 117–126.<br />

Campbell, J.B. & Cary, C.S. 2001. Characterization of a novel spirochete associated with the hydrothermal<br />

vent polychaete annelid, Alvinella pompejana. Applied Environmental Microbiology 67, 110–117.<br />

Carney, R.S. 1994. Consideration of the oasis analogy for chemosynthetic communities at Gulf of Mexico<br />

hydrocarbon vents. Geo-Marine Letters 14, 149–159.<br />

Cary, C.S., Cottrell, M.T., Stein, J.L., Camacho, F. & Desbruyàres, D. 1997. Molecular identification and<br />

localization of filamentous symbiotic bacteria associated with the hydrothermal vent annelid Alvinella<br />

pompejana. Applied Environmental Microbiology 63, 1124–1130.<br />

Casanova, B., Brunet, M. & Segonzac, M. 1993. L’impact d’une épibiose bactérienne sur la morphologie<br />

fonctionnelle de crevettes associées à l’hydrothermalisme médio-Atlantique. Cahiers de Biologie<br />

Marine 34, 573–588.<br />

Cavanaugh, C.M. 1985. Symbioses of chemoautotrophic bacteria and <strong>marine</strong> invertebrates from hydrothermal<br />

vents and reducing sediments. Bulletin of the Biological Society of Washington 6, 373–388.<br />

Chamberlain, S. 2000. Vision in hydrothermal vent shrimp. Philosophical Transactions of the Royal Society<br />

of London Biological Sciences 355, 1151–1154.<br />

Chitwood, B.G. 1936. Some <strong>marine</strong> nematodes from North Carolina. Proceedings of the Helminthological<br />

Society of Washington 3, 1–16.<br />

Christoffersen, M.L. 1986. Phylogenetic relationships between Oplophoridae, Atyidae, Pasipheidae, Alvinocarididae<br />

fam.n., Bresiliidae, Psalidopodidae and Disciadidae (Crustacea Caridea Atyoidea). Boletim<br />

de Zoologia, Universidada de Sao Paulo 10, 273–281.<br />

Compere, P., Martinez, A.S., Charmantier, D.M., Toullec, J.Y., Goffinet, G. & Gaill, F. 2002. Does sulphide<br />

detoxication occur in the gills of the hydrothermal vent shrimp, Rimicaris exoculata? Comptes Rendus<br />

Biologies 325, 591–596.<br />

Cottrell, T.M. & Cary, C. 1999. Diversity of dissimilatory bisulfite reductase genes of bacteria associated with<br />

the deep-sea hydrothermal vent polychaete annelid Alvinella pompejana. Applied Environmental<br />

Microbiology 65, 1127–1132.<br />

Dando, P.R., Hughes, J.A. & Thiermann, F. 1995. Preliminary observations on biological communities at<br />

shallow hydrothermal vents in the Aegean Sea. In Hydrothermal Vents and Processes, L.M. Parson<br />

et al. (eds). Geological Society Special Publications 87, 303–317.<br />

Desbruyères, D., Chevaldonne, P., Alayse, A.-M., Jollivet, D., Lallier, F.H., Jouin-Toulmond, C., Zal, F.,<br />

Sarradin, P.-M., Cosson, R., Caprais, J.-C., Arndt, C., O’Brien, J., Guezennec, J., Hourdez, S., Riso,<br />

R., Gaill, F., Laubier, L., Toulmond, A. 1998. Biology and ecology of the ‘Pompeii worm’ (Alvinella<br />

pompejana: Desbruyères & Laubier), a normal dweller of an extreme deep-sea environment: a<br />

synthesis of current knowledge and recent developments. Deep-Sea Research II 45, 383–422.


2727_C04.fm Page 114 Wednesday, June 30, 2004 12:00 PM<br />

114 J. Ott, M. Bright & S. Bulgheresi<br />

Desbruyàres, D., Gaill, F., Laubier, L., Prieur, D. & Rau, G.H. 1983. Unusual nutrition of the ‘Pompeii worm’<br />

Alvinella pompejana (polychaetous annelid) from a hydrothermal vent environment: SEM, TEM, 13 C<br />

and 15 N evidence. Marine Biology 75, 201–205.<br />

Dixon, D.R., Dixon, L.R. & Pond, D.W. 1998. Recent advances in our understanding of the life history of<br />

bresiliid vent shrimps on the MAR. Cahiers de Biologie Marine 39, 383–386.<br />

Epstein, S.S., Bazylinski, D.A. & Fowle, W.H. 1998. Epibiotic bacteria on several ciliates from <strong>marine</strong><br />

sediments. The Journal of Eukariotic Microbiology 45, 64–70.<br />

Farmer, J.D. 1998. Thermophiles, early biosphere evolution and the origin of life on Earth: implications for<br />

the exobiological exploration of Mars. Journal of Geophysical Research 103, 28, 457, 461.<br />

Fenchel, T. 1970. Studies on the decomposition of organic detritus derived from the turtle grass Thalassia<br />

testudinum. Limnology and Oceanography 15, 14–20.<br />

Fenchel, T. & Finlay, B.J. 1989. Kentrophoros: a mouthless ciliate with a symbiotic kitchen garden. Ophelia<br />

30, 75–93.<br />

Fenchel, T., Perry, T. & Thane, A. 1977. Anaerobiosis and symbiosis with bacteria in free-living ciliates.<br />

Journal of Protozoology 24, 154–163.<br />

Fenchel, T. & Riedl, R. 1970. The sulfide system: a new biotic community underneath the oxidized layer of<br />

<strong>marine</strong> sand bottoms. Marine Biology 7, 255–268.<br />

Fisher, C.R. 1990. Chemoautotrophic and methanotrophic symbioses in <strong>marine</strong> invertebrates. Reviews in<br />

Aquatic Sciences 2, 399–436.<br />

Fisher, C.R. 1996. Ecophysiology of primary production at deep-sea vents and seeps. In Deep-Sea and Extreme<br />

Shallow-Water Habitats: Affinities and Adaptations, F. Uiblein et al. (eds). Vienna: Austrian Academy<br />

of Sciences Press, Biosystematics and Ecology Series 11, 313–336.<br />

Foissner, W. 1995. Kentrophoros (Ciliophora, Karyorelictea) has oral vestiges: a reinvestigation of K. fistulosus<br />

(Faure-Fremiet, 1950) using protargol impregnation. Archiv für Protistenkunde 146, 165–179.<br />

Fortey, R. 2000. Olenid trilobites: the oldest known chemoautotrophic symbionts? Proceedings of the National<br />

Academy of Sciences of the United States of America 97, 6574–6578.<br />

Gallardo, V.A. 1977. Large benthic <strong>microbial</strong> communities in sulphide biota under Peru–Chile subsurface<br />

countercurrent. Nature 268, 331–332.<br />

Gebruk, A.V., Pimenov, N.V. & Savvichez, A.S. 1993. Feeding specialization of bresiliid shrimps in the TAG<br />

site hydrothermal community. Marine Ecology Progress Series 98, 247–253.<br />

Gebruk, A.V., Southward, E.C., Kennedy, H. & Southward, A.J. 2000. Food sources, behaviour, and distribution<br />

of hydrothermal vent shrimps at the Mid-Atlantic Ridge. Journal of the Marine Biological Association<br />

of the United Kingdom 80, 485–499.<br />

Geistdoerfer, P. 1994. Pachycara thermophilum, a new zoarcid fish species from hydrothermal ecosystems of<br />

the Mid-Atlantic Ridge. Cybium 18, 109–115.<br />

Gerlach, S.A. 1950. Über einige Nematoden aus der Familie der Desmodoriden. Neue Ergebnisse und Probleme<br />

der Zoologie (Klatt-Festschrift), Leipzig 1950, pp. 178–198.<br />

Giere, O. 1992. Benthic life in sulfidic zones of the sea: ecological and structural adaptations to a toxic<br />

environment. Verhandlungen der deutschen Zoologischen Gesellschaft 85, 77–93.<br />

Giere, O. 1996. Bacterial endosymbionts in <strong>marine</strong> littoral worms. In Deep-Sea and Extreme Shallow-Water<br />

Habitats: Affinities and Adaptations, F. Uiblein et al. (eds). Vienna: Austrian Academy of Sciences<br />

Press, Biosystematics and Ecology Series 11, 369–382.<br />

Goldfarb, M.S., Converse, D.R., Holland, H.D. & Edmond, J.M. 1983. The genesis of the hot spring deposits<br />

on the East Pacific Rise, 21˚N. In The Kuroko and Related Volcanic Massive Sulfide Deposits, H.<br />

Ohmoto & B.J. Skinner (eds). Economical Geology Monograph 5, 184–197.<br />

Greeff, R. 1869. Untersuchungen über einige merkwürdige Formen des Arthropoden- und Wurmtypus. Archiv<br />

für Naturgeschichte 35, 71–121.<br />

Haddad, A., Camacho, F., Durand, P. & Cary, S.C. 1995. Phylogenetic characterization of the epibiotic bacteria<br />

associated with the hydrothermal vent polychaete Alvinella pompejana. Applied and Environmental<br />

Microbiology 61, 1679–1687.<br />

Hemprich, F.W. & Ehrenberg, C.G. 1831. Symbolae Physicae. Evertebrata. I. Phytozoa. Abhandlungen der<br />

Akademie der Wissenschaften zu Berlin.<br />

Hentschel, U., Berger, E.C., Bright, M., Felbeck, H. & Ott, J.A. 1999. Metabolism of nitrogen and sulfur in<br />

ectosymbiotic bacteria of <strong>marine</strong> nematodes (Nematoda, Stilbonematinae). Marine Ecology Progress<br />

Series 183, 149–158.


2727_C04.fm Page 115 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 115<br />

Herring, P.J. 1998. North Atlantic midwater distribution of the juvenile stages of hydrothermal vent shrimps<br />

(Decapoda: Bresiliidae). Cahiers de Biologie Marine 39, 387–390.<br />

Hoschitz, M., Bright, M. & Ott, J.A. 2001. Ultrastructure and reconstruction of the pharynx of Leptonemella<br />

juliae (Nematoda, Adenophorea). Zoomorphology 121, 95–107.<br />

Jannasch, H.W., Wirsen, C.O. & Molyneaux, S.J. 1991. Chemosynthetic oxidation of polymetal sulfides and<br />

the formation of biomass at the 23˚ and 26˚N Mid-Atlantic Ridge hydrothermal vent sites. EOS 73, 585.<br />

Jensen, P. 1986. Nematode fauna in the sulfide-rich brine seep and adjacent bottoms of the East Flower Garden,<br />

NW Gulf of Mexico. IV. Ecological aspects. Marine Biology 92, 489–503.<br />

Johnson, K.S., Childress, J.J. & Beehler, C.L. 1988. Short-term temperature variability in the Rose Garden<br />

hydrothermal vent field: an unstable deep-sea environment. Deep-Sea Research 35, 1711–1722.<br />

Jørgensen, B.B. 1989. Biogeochemistry of chemoautotrophic bacteria. In Autotrophic Bacteria, H.G. Schlegel<br />

& B. Bowien (eds). Madison, WI: Science Technology Publishers & Springer Verlag, pp. 117–146.<br />

Kahl, A. 1935. Urtiere oder Protozoa I: Wimpertiere oder Ciliata (Infusoria) 4. Peritricha und Chonotricha.<br />

Die Tierwelt Deutschlands 30, 651–686.<br />

Kamenev, G.M., Fadeev, V.I., Selin, N.I., Tarasov, V.G. & Malakhov, V.V. 1993. Composition and distribution<br />

of macro- and meiobenthos around sub-littoral hydrothermal vents in the Bay of Plenty, New Zealand.<br />

New Zealand Journal of Marine and Freshwater Research 27, 407–418.<br />

Kampfer, S., Sturmbauer, C. & Ott, J. 1998. Phylogenetic analysis of rDNA sequences from adenophorean<br />

nematodes and implications for the Adenophorea-Secernentea controversy. Invertebrate Biology 117,<br />

29–36.<br />

Karl, D.M., Wirsen, C.O. & Jannasch, H.W. 1980. Deep-sea primary production at the Galapagos hydrothermal<br />

vents. Science 207, 1345–1347.<br />

Laval-Peuto, M. & Rassoulzadegan, F. 1988. Autofluorescence of <strong>marine</strong> planktonic Oligotrichina and other<br />

ciliates. Hydrobiologia 159, 99–110.<br />

Lee, J., Kwon, K., Hyun, J., Kim, S. & Lee, H. 2001. Analysis of bacterial community structure in <strong>marine</strong><br />

sediments using terminal-restriction fragment length polymorphism (T-RFLP) and sequencing of 16S<br />

rDNA clones. Abstracts of the General Meeting of the American Society for Microbiology 101, 496.<br />

Madrid, V.M., Taylor, G.T., Scranton, M.I. & Chistoserdov, A.Y. 2001. Phylogenetic diversity of bacterial and<br />

archaeal communities in the anoxic zone of the Cariaco Basin. Applied Environmental Microbiology<br />

67, 1663–1674.<br />

Mann, K.H. 1976. Decomposition of <strong>marine</strong> macrophytes. In The Role of Terrestrial and Aquatic Organisms<br />

in Decomposition Processes, J.M. Anderson & A. Macfadyen (eds). Oxford: Blackwell Scientific<br />

Publications, pp. 247–267.<br />

Martin, J.W., Signorovitch, J. & Patel, H. 1997. A new species of Rimicaris (Crustacea: Decapoda: Bresiliidae)<br />

from the Snake Pit hydrothermal vent field on the Mid-Atlantic Ridge. Proceedings of the Biological<br />

Society of Washington 110, 399–411.<br />

McKee, K.L. 1993. Soil physicochemical patterns and mangrove species distribution: reciprocal effects?<br />

Journal of Ecology 81, 477–487.<br />

Molnar, D.A., Nikolausz, M., Bright, M., Márialigeti, K., Vanura, K., Buchholz, T.G. & Ott, J.A. 2000.<br />

Phylogenetic characterization of Zoothamnium niveum’s (Ciliophora, Peritrichida) <strong>thiotrophic</strong> ectosymbiotic<br />

bacterial community based on comparative 16S rRNA sequence analysis. In Programs,<br />

Abstracts and Papers of the Third International Congress on Symbiosis, H.C. Weber et al. (eds).<br />

Philipps University of Marburg, Germany, p. 153.<br />

Müller, M.C., Bernhard, J.M. & Jouin-Toulmond, C. 2001. A new member of Nerillidae (Annelida: Polychaeta),<br />

Xenonerilla bactericola gen. et sp. nov., collected off California, USA. Cahiers de Biologie<br />

Marine 42, 203–217.<br />

Nebelsick, M., Blumer, M., Novak, R. & Ott, J. 1992. A new glandular sensory organ in Catanema sp.<br />

(Nematoda, Stilbonematinae). Zoomorphology 112, 17–26.<br />

Nelson, D.C. & Fisher, C. 1995. Chemoautotrophic and methanotrophic endosymbiotic bacteria at deep-sea<br />

vents and seeps. In The Microbiology of Deep-Sea Hydrothermal Vents, D.M. Karl (ed.). Boca Raton,<br />

FL: CRC Press, pp. 125–167.<br />

Nuckley, D.J., Jinks, R.N., Battelle, B.A., Herzog, E.D., Kass, L., Renninger, G.H. & Chamberlain, S. 1996.<br />

Retinal anatomy of a new species of bresiliid shrimp from a hydrothermal vent field on the Mid-<br />

Atlantic Ridge. Biological Bulletin 190, 98–110.


2727_C04.fm Page 116 Wednesday, June 30, 2004 12:00 PM<br />

116 J. Ott, M. Bright & S. Bulgheresi<br />

Nussbaumer, A.D., Bright, M., Baranyi, Ch., Beisser, Ch.J. & Ott, J.A. 2004. Attachment mechanisms in a<br />

highly specific association between ectosymbiotic bacteria and <strong>marine</strong> nematodes. Aquatic Microbial<br />

Ecology, 34, 239–246.<br />

Oeschger, R. & Janssen, H.H. 1991. Histological studies on Halicryptus spinulosus (Priapulida) with regard<br />

to environmental hydrogen sulfide resistance. Hydrobiologica 122, 1–12.<br />

Oeschger, R. & Schmaljohann, R. 1988. Association of various types of epibacteria with Halicryptus spinulosus<br />

(Priapulida). Marine Ecology Progress Series 48, 285–293.<br />

O’Neill, P.A., Jinks, R.N., Herzog, E.D., Battelle, B.A., Kass, L., Renninger, G.H. & Chamberlain, S.C. 1995.<br />

The morphology of the dorsal eye of the hydrothermal vent shrimp, Rimicaris exoculata. Visual<br />

Neuroscience 12, 861–875.<br />

Ott, J. 1995. Sulphide symbioses in shallow sands. In Biology and Ecology of Shallow Coastal Waters, A.<br />

Eleftheriou et al. (eds). Fredensbourg, Denmark: Olsen & Olsen, pp. 143–147.<br />

Ott, J. 1996. Sulphide <strong>ectosymbioses</strong> in shallow <strong>marine</strong> habitats. In Deep-Sea and Extreme Shallow-Water<br />

Habitats: Affinities and Adaptations, F. Uiblein et al. (eds). Vienna: Austrian Academy of Sciences<br />

Press, Biosystematics and Ecology Series 11, 369–382.<br />

Ott, J. & Novak, R. 1989. Living at an interface: meiofauna at the oxygen-sulfide boundary of <strong>marine</strong> sediments.<br />

In Reproduction, Genetics and Distribution of Marine Organisms, J.S. Ryland & P.A. Tyler (eds).<br />

Fredensbourg, Denmark: Olsen & Olsen, pp. 415–422.<br />

Ott, J.A., Bauer-Nebelsick, M. & Novotny, V. 1995. The genus Laxus Cobb, 1894 (Stilbonematinae: Nematoda):<br />

description of two new species with ectosymbiotic chemoautotrophic bacteria. Proceedings of<br />

the Biological Society of Washington 108, 508–527.<br />

Ott, J.A., Bright, M. & Schiemer, F. 1998. The ecology of a novel symbiosis between a <strong>marine</strong> peritrich ciliate<br />

and chemoautotrophic bacteria. Pubblicazioni della Stazione Zoologica di Napoli. I. Marine Ecology<br />

19, 229–243.<br />

Ott, J.A., Novak, R., Schiemer, F., Hentschel, U., Nebelsick, M. & Polz, M. 1991. Tackling the sulfide gradient:<br />

a novel strategy involving <strong>marine</strong> nematodes and chemoautotrophic ectosymbionts. Pubblicazioni<br />

della Stazione Zoologica di Napoli. I. Marine Ecology 12, 261–279.<br />

Polz, M.F. & Cavanaugh, C.M. 1995. Dominance of one bacterial phylotype at a Mid-Atlantic Ridge hydrothermal<br />

vent site. Proceedings of the National Academy of Sciences of the United States of America<br />

92, 7232–7236.<br />

Polz, M.F. & Cavanaugh, C.M. 1996. The ecology of ectosymbiosis at a Mid-Atlantic Ridge hydrothermal<br />

vent site. In Deep-Sea and Extreme Shallow-Water Habitats: Affinities and Adaptations, F. Uiblein et<br />

al. (eds). Vienna: Austrian Academy of Sciences Press, Biosystematics and Ecology Series 11,<br />

337–352.<br />

Polz, M.F., Distel, D.L., Zarda, B., Amann, R., Felbeck, H., Ott, J.A. & Cavanaugh, C.M. 1994. Phylogenetic<br />

analysis of a highly specific association between ectosymbiotic, sulfur-oxidizing bacteria and a <strong>marine</strong><br />

nematode. Applied Environmental Microbiology 60, 4461–4467.<br />

Polz, M.F., Felbeck, H., Novak, R., Nebelsick, M. & Ott, J.A. 1992. Chemoautotrophic, sulfur-oxidizing<br />

symbiotic bacteria on <strong>marine</strong> nematodes: morphological and biochemical characterizations. Microbial<br />

Ecology 24, 313–329.<br />

Polz, M.F., Harbison, C. & Cavanaugh, C.M. 1999a. Diversity and heterogeneity of epibiotic bacterial<br />

communities on the <strong>marine</strong> nematode Eubostrichus dianae. Applied Environmental Microbiology 65,<br />

4271–4275.<br />

Polz, M.F., Ott, J.A., Bright, M. & Cavanaugh, C.M. 2000. When bacteria hitch a ride: associations between<br />

sulfur-oxidizing bacteria and eucaryotes represent spectacular adaptations to environmental gradients.<br />

ASM News 66, 531–532.<br />

Polz, M.F., Robinson, J.J., Cavanaugh, C.M. & Van Dover, C.L. 1999b. Trophic ecology of massive shrimp<br />

aggregations at a Mid-Atlantic Ridge hydrothermal vent site. Limnology and Oceanography 43,<br />

1631–1638.<br />

Pond, D., Dixon, D. & Sargent, J. 1997a. Wax-ester reserves facilitate dispersal of hydrothermal vent shrimps.<br />

Marine Ecology Progress Series 146, 289–290.<br />

Pond, D.W., Dixon, D.R., Bell, M.V., Fallick, A.E. & Sargent, J.R. 1997b. Occurrence of 16:2 (n-4) and 18:2<br />

(n-4) fatty acids in the lipids of the hydrothermal vent shrimps Rimicaris exoculata and Alvinocaris<br />

markensis: nutritional and trophic implications. Marine Ecology Progress Series 156, 167–174.


2727_C04.fm Page 117 Wednesday, June 30, 2004 12:00 PM<br />

Marine Microbial Thiotrophic Ectosymbioses 117<br />

Pond, D.W., Gebruk, A., Southward, E.C., Southward, A.J., Fallick, A.E., Bell, M.V. & Sargent, J.R. 2000.<br />

Unusual fatty acid compositions of storage lipids in the bresiliid shrimp Rimicaris exoculata couples<br />

the photic zone with MAR hydrothermal vent sites. Marine Ecology Progress Series 198, 171–179.<br />

Pond, D.W., Segonzac, M., Bell, M.V., Dixon, D.R., Fallick, A.E. & Sargent, J.R. 1997c. Lipid and lipid<br />

carbon stable isotope composition of the hydrothermal vent shrimp Mirocaris fortunata: evidence for<br />

nutritional dependence on photosynthetically fixed carbon. Marine Ecology Progress Series 157,<br />

221–231.<br />

Powell, E.N. 1989. Oxygen, sulfide and diffusion: why thiobiotic meiofauna must be sulfide-insensitive firstorder<br />

respirers. Journal of Marine Research 47, 887–932.<br />

Powell, E.N., Crenshaw, M. & Rieger, R.M. 1979. Adaptions to sulfide in the meiofauna of the sulfide system.<br />

I. 35 S-sulfide accumulation and the presence of a sulfide detoxification system. Journal of Experimental<br />

Marine Biology and Ecology 37, 57–76.<br />

Raguenes, G., Christen, R., Guezennec, J., Pignet, P. & Barbier, G. 1997. Vibrio diabolicus sp. nov., a new<br />

polysaccharide-secreting organism isolated from a deep-sea hydrothermal vent polychaete annelid,<br />

Alvinella pompejana. International Journal of Systematic Bacteriology 47, 989–995.<br />

Raikov, I.B. 1971. Bactéries épizoiques et mode de nutrition du ciliés psammophile Kentrophoros fistulosum<br />

Faure-Fremiet (étude du microscope électronique). Protistologica 7, 365–378.<br />

Raikov, I.B. 1974. Étude ultrastructurele de bactéries épizoiques et endozoiques de Kentrophoros latum Raikov,<br />

ciliés holotriche mésopsammique. Cahiers de Biologie Marine 15, 379–393.<br />

Ramirez Llodra, E., Tyler, P.A. & Copley, J.T.P. 2000. Reproductive biology of three caridean shrimp, Rimicaris<br />

exoculata, Chorocaris chacei and Mirocaris fortunata (Caridea: Decapoda), from hydrothermal vents.<br />

Journal of the Marine Biology Association of the United Kingdom 80, 473–484.<br />

Renninger, G.H., Kass, L., Gleeson, R.A., Van Dover, C.L., Battelle, B.-A., Jinks, R.N., Herzog, E.D. &<br />

Chamberlain, S.C. 1995. Sulfide as a chemical stimulus for deep-sea hydrothermal vent shrimp.<br />

Biological Bulletin 189, 69–76.<br />

Rieley, G., Van Dover, C.L., Hedrick, D.B. & Eglinton, G. 1999. Trophic ecology of Rimicaris exoculata: a<br />

combined lipid abundance/stable isotope approach. Marine Biology 133, 495–499.<br />

Riemann, F., Ernst, W. & Ernst, R. 1990. Acetate uptake from ambient water by the free-living <strong>marine</strong> nematode<br />

Adoncholaimus thalassophygas. Marine Biology 104, 453–457.<br />

Riemann, F., Thiermann, F. & Bock, L. 2003. Leptonemella species (Desmodoridae, Stilbonematinae), benthic<br />

<strong>marine</strong> nematodes with ectosymbiotic bacteria, from littoral sand of the North Sea island of Sylt.<br />

Taxonomy and ecological aspects. Helgoland Marine Research 57, 118–131.<br />

Rinke, Ch. 2002. Carbon Fixation, Incorporation, and Transfer in the Chemoautotrophic Zoothamnium niveum<br />

Symbiosis with 14C Bicarbonate Autoradiography. Diploma thesis, University of Vienna.<br />

Sauerbrey, E. 1928. Beobachtungen über einige neue oder wenig bekannte <strong>marine</strong> Ciliaten. Archiv für Protistenkunde<br />

62, 355–407.<br />

Schiemer, F., Novak, R. & Ott, J. 1990. Metabolic studies on thiobiotic free-living nematodes and their<br />

symbiotic microorganisms. Marine Biology 106, 129–137.<br />

Schmaljohann, R. & Flügel, H.J. 1987. Methane-oxidizing bacteria in Pogonophora. Sarsia 72, 91–98.<br />

Segonzac, M., de Saint Laurent, M. & Casanova, B. 1993. L’énigme du comportement trophique des crevettes<br />

Alvinocarididae des sites hydrothermaux de la dorsale médio-atlantique. Cahiers de Biologie Marine<br />

34, 535–571.<br />

Shank, T.M., Lutz, R.A. & Vrijenhoek, R.C. 1998. Molecular systematics of shrimp (Decapoda: Bresiliidae)<br />

from deep-sea hydrothermal vents. I. Enigmatic “small orange” shrimp from the Mid-Atlantic Ridge<br />

are juvenile Rimicaris exoculata. Molecular Marine Biology and Biotechnology 7, 88–96.<br />

Somero, G.N., Childress, J.J. & Anderson, A.E. 1989. Transport, metabolism and detoxification of hydrogen<br />

sulphide in animals from sulphide-rich <strong>marine</strong> environments. Critical Reviews in Aquatic Sciences 1,<br />

591–614.<br />

Southward, A.J. & Newman W.A. 1998. Ectosymbiosis between filamentous sulfur bacteria and a stalked<br />

barnacle (Scalpellomorpha, Neolepadinae) from the Lau Back Arc Basin, Tonga. Cahiers de Biologie<br />

Marine 39, 259–262.<br />

Southward, E.C. 1986. Gill symbionts in thyasirids and other bivalve molluscs. Journal of the Marine<br />

Biological Association of the United Kingdom 66, 889–914.<br />

Suess, E., Kulm, L.D., Carson, B. & Whiticar, M.J. 1987. Fluid flow and methane fluxes from vent sites at<br />

the Oregon subduction zone. EOS: Transactions of the American Geophysical Union 68, 1486.


2727_C04.fm Page 118 Wednesday, June 30, 2004 12:00 PM<br />

118 J. Ott, M. Bright & S. Bulgheresi<br />

Urbancik, W., Bauer-Nebelsick, M. & Ott, J.A. 1996a. The ultrastructure of the cuticle of Nematoda. I. The<br />

body cuticle within the Stilbonematinae (Adenophorea, Desmodoridae). Zoomorphology 116, 51–64.<br />

Urbancik, W., Novotny, V. & Ott, J.A. 1996b. The ultrastructure of the cuticle of Nematoda. II. The cephalic<br />

cuticle of Stilbonematinae (Adenophorea, Desmodoridae). Zoomorphology 116, 65–75.<br />

Van Dover, C.L. 2000. The Ecology of Deep-Sea Hydrothermal Vents. Princeton, NJ: Princeton University<br />

Press.<br />

Van Dover, C.L. 2002. Trophic relationships among invertebrates at the Kairei hydrothermal vent field (Central<br />

Indian Ridge). Marine Biology 141, 761–772.<br />

Van Dover, C.L. & Fry, B. 1994. Microorganisms as food resources at deep-sea hydrothermal vents. Limnology<br />

and Oceanography 39, 51–57.<br />

Van Dover, C.L., Fry, B., Grassle, J.F., Humphris, S. & Rona, P.A. 1988. Feeding biology of the shrimp<br />

Rimicaris exoculata at hydrothermal vents on the Mid-Atlantic Ridge. Marine Biology 98, 209–216.<br />

Van Dover, C.L., Szuts E.Z., Chamberlain, S.C. & Cann, J.R. 1989. A novel eye in ‘eyeless’ shrimp from<br />

hydrothermal vents of the Mid-Atlantic Ridge. Nature 337, 458–460.<br />

Von Damm, K.L. 1995. Controls on the chemistry and temporal variability of seafloor hydrothermal fluids.<br />

In Sea-Floor Hydrothermal Systems: Physical, Chemical, Biological and Geological Interactions,<br />

S.E. Humphris et al. (eds). Washington, DC: American Geophysical Union, Geophysical Monograph<br />

91, 222–247.<br />

Vopel, K., Pöhn, M., Sorgo, A. & Ott, J. 2001. Ciliate-generated advective seawater transport supplies<br />

chemoautotrophic ectosymbionts. Marine Ecology Progress Series 210, 93–99.<br />

Vopel, K., Reick, C.H., Arlt, G., Pöhn, M. & Ott, J.A. 2002. Flow microenvironment of two <strong>marine</strong> peritrich<br />

ciliates with ectobiotic chemoautotrophic bacteria. Aquatic Microbial Ecology 29, 19–28.<br />

Watabe, H. & Hashimoto, J. 2002. A new species of the genus Rimicaris (Alvinocarididae, Caridea, Decapoda)<br />

from the active hydrothermal vent field, “Kairei Field,” on the Central Indian Ridge, the Indian Ocean.<br />

Zoological Science Tokyo 19, 1167–1174.<br />

Wieser, W. 1959. Eine ungewöhnliche Assoziation zwischen Blaualgen und freilebenden <strong>marine</strong>n Nematoden.<br />

Österreichische Botanische Zeitschrift 106, 81–87.<br />

Williams, A.B. & Rona, P.A. 1986. Two new caridean shrimps (Bresiliidae) from a hydrothermal field on the<br />

Mid-Atlantic Ridge. Journal of Crustacean Biology 6, 446–462.<br />

Wirtz, P. & Debelius, H. 2003. Niedere Tiere. Mittelmeer und Atlantik. Hamburg, Germany: Jahr Top Special<br />

Verlag.

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

Saved successfully!

Ooh no, something went wrong!