Sie sind auf Seite 1von 33

Prokaryotes (2006) 1:475507

DOI: 10.1007/0-387-30741-9_18

CHAPTER 2.5
en i rM
a c i t ehtnysomehC seso i bmyS

Marine Chemosynthetic Symbioses


COLLEEN M. CAVANAUGH, ZOE P. MCKINESS, IRENE L.G. NEWTON AND
FRANK J. STEWART

Introduction ide or methane for carbon fixation and utiliza-


tion. On the basis of these unique biosynthetic
Bacteria and marine eukaryotes often coexist in capacities, notably the ability to synthesize C3
symbioses that significantly influence the ecol- compounds from C1 compounds, we refer collec-
ogy, physiology and evolution of both partners. tively to these bacterial symbionts as chemosyn-
De Bary (1879) defined symbiosis as the living thetic.
together of differently named organisms, imply- Given the sulfide-rich habitats in which
ing that the term encompasses both positive (e.g., chemoautotrophic symbioses occur, researchers
mutualism) and negative (e.g., parasitism) asso- infer that the bacterial symbionts oxidize
ciations. Many researchers now view symbiotic reduced inorganic sulfur compounds to obtain
interactions as those that persist over the major- energy and reducing power for autotrophic car-
ity of the lifespan of the organisms involved and bon fixation. While some endosymbionts (such as
that provide benefits to each partner beyond those in the protobranchs Solemya velum and S.
those obtained in the absence of association. This reidi; Cavanaugh, 1983; Anderson et al., 1987)
chapter describes such symbioses, specifically utilize thiosulfate (S2O3), an intermediate in sul-
those between marine invertebrate and protist fide oxidation, hydrogen sulfide is inferred to be
hosts and chemosynthetic bacterial symbionts. the preferred energy source in a variety of sym-
These bacteria, which cluster primarily within bioses (see review in Van Dover, 2000). But for
the Gammaproteobacteria (Fig. 1), are chemo- many symbioses the actual energy source has not
autotrophs or methanotrophs. In both chemo- been identified definitively; rather, only an
autotrophic and methanotrophic symbioses, the autotrophic metabolism has been confirmed.
hosts, through an astonishing array of physio- Indeed, chemosynthetic bacteria utilizing other
logical and behavioral adaptations, provide the energy sources (e.g., hydrogen or ammonia)
symbiont access to the substrates (i.e., electron could also serve similar nutritional roles in
donors and acceptors) necessary for the genera- symbiotic associations. In this review, bacterial
tion of energy and bacterial biomass. In symbionts that have been shown to use reduced
exchange, a portion of the carbon fixed by the sulfur compounds (H2S, HS, S2, S2O3, So) for
symbiont is used, either directly or indirectly, for energy metabolism are referred to as thio-
host energy and biosynthesis. These symbioses autotrophs, while the more general term
thereby increase the metabolic capabilities, and chemoautotroph is used to describe symbionts
therefore the number of ecological niches, of for which data supporting autotrophic CO2 fixa-
both the host and the bacterial symbiont. tion exist but for which the lithotrophic energy
In those symbioses for which the electron source is unknown.
donor has been explicitly identified, sulfide
and other inorganic reduced sulfur compounds
History
(e.g., thiosulfate) fuel energy generation by the
chemoautotrophic symbionts, serving as elec- The discovery of deep-sea hydrothermal vents
tron sources for oxidative phosphorylation. In and the flourishing ecosystems associated with
these symbioses, the ATP produced in electron them significantly advanced scientific under-
transport fuels autotrophic CO2 fixation via the standing of chemosynthetic symbioses. Oceanog-
Calvin cycle. In contrast, bacteria in marine raphers in the research submersible Alvin
invertebrate-methanotroph symbioses use meth- discovered hydrothermal vents along the Gal-
ane (CH4) as an energy, electron, and carbon apagos Rift in 1977. In stark contrast to percep-
source. Unlike their protist or metazoan hosts, tions of the deep benthos as a cold, food-limited
chemoautotrophs and methanotrophs share the habitat incapable of supporting substantial bio-
ability to use reduced inorganic compounds or mass, hydrothermal vents are oases, charac-
methane for energy generation and carbon diox- terized by high concentrations of free-living
476 C.M. Cavanaugh et al. CHAPTER 2.5

Chemoautotrophic symbionts:
Mussel symbionts
Vesicomyid clam symbionts
Solemyid clam symbionts
Lucinid clam symbionts
Thyasirid clam symbionts
Tubeworm symbionts
Oligochaete symbionts
Escherichia coli Nematode symbionts
Coxiella burnetii
Solemva reidi symbiont Methanotrophic symbionts
GAMMA JTB254 Epsilon symbionts
86 Achromatium oxaliferum
54 Inanidrilus leukodermatus symbiont
100 Olavius algarvensis gamma symbiont
65 Laxus sp symbiont
Olavius loisae gamma symbiont
82 Solemva relum symbiont
Solemva occidentalis symbiont
Lucinella nassula symbiont
97 Codakia orbicularis symbiont
98 Lucina floridana symbiont Chemoautotrophic
57 Codakia costata symbiont symbionts
68 Lucinoma annulata symbiont
Lucinoma aequizonata symbiont
94 100 Ridgeia piscesae symbiont
Rifna packyptria symbiont
100 Escorpia spicata symbiont
Lamembrachia colmnua symbiont
93 Anodontia phillipiana symbiont
64 70 Solemya terraeregina symbiont
69 Thyasira flexuosa symbiont
Solemya pusillu symbiont
Lucina pectinata symbiont
80 Pseudomonas fluorescens
100 Pseudomonas putida
88 Pseudomonas lundensis
81 Pseudomonas fragi
100 Marinobacter sp CAB
Marinobacter aquaeolei
Oceanospirillum kriegii
100 Thiothrix eikelboomii
52 Thiothrix unzii
98
Thiothrix nivea
Thiothrix ramosa
54 Bathymodiolus japonicus symbiont
Bathymodiolus childressi symbiont Methanotrophic
100
Bathymodiolus platifrons symbiont symbionts
100 Bathymodiolus puteoserpentis symbiont 2
100 76 Methylobacter sp BB5.1
53 Methylococcus luteus
60 Methylococcus capsulatus BARH
61 Methyllobacter whittenburyi
100 Methylomonas aurantiaca
Methylomonas rubra
Methylococcus thermophilus
GAMMA JTB35
100 Thiomicrospira sp.
82 Thiomicrospira crunogena
82 Thiomicrospira L12
100 Thiomicrospira milos T2
Thiomicrospira milos T1
69 100 Thiomicrospira thyasirae
Thiomicrospira pelophila
Maorithyas hadalis symbiont 1
Juan de Fuca mussel symbiont
55 Bathymodiolus septemdierum symbiont
52 70 Bathymodiolus thermophilus symbiont
Bathymodiolus puteoserpentis symbiont 1
Bathymodiolus aff. brevior symbiont
62 Calyptogena magnifica symbiont
Calyptogena fossajaponica symbiont Chemoautotrophic
100
82 Calyptogena phaseoliformis symbiont symbionts
Calyptogena FL sp. symbiont
94 Calyptogena pacifica symbiont
86 Vesicomya lepta symbiont
100
87 Vesicomya chordata symbiont
100 Calyptogena elongata symbiont
100 Calyptogena kilmeri symbiont
Vesicomya gigas symbiont
Ectenagena extenta symbiont
Maorithyas hadalis symbiont 2
100 Halothiobacillus kellyi
Halothiobacillus neapolitanus Non-gamma
Olavius loisae alpha symbiont oligochaete
Olavius loisae spirochete symbiont
100 Olavius algarvensis delta symbiont symbionts
Desulfosarcina variabilis
100 Alvinella pompejana epibiont APB13B
84
Alvinella pompejana epibiont APG44B Epsilon
87 Alvinella pompejana epibiont APG5B epibionts
Rimicaris exoculata symbiont
100 Alvinella pompejana epibiont APG56B
100 Sulfurospirillum sp 18.1
Sulfurospirillum Am-N

Fig. 1. Phylogeny showing the strict consensus of 46 trees obtained via parsimony analyses of 16S rRNA gene sequences
(1456 bp) from symbiotic and free-living bacteria. Results greater than 50% from a 500 replicate bootstrap analysis are
reported above respective branches. Chemosynthetic symbiont taxa are color-coded (see key on Figure).
CHAPTER 2.5 Marine Chemosynthetic Symbioses 477

source for R. pachyptila, implying a role for


chemoautotrophy in tubeworm metabolism.
Following confirmation of a chemosynthetic
endosymbiosis within the giant tubeworm,
researchers questioned the putative reliance on
filter feeding by other vent invertebrates. Ulti-
mately, anatomical, enzymological, and isotopic
analyses revealed the presence of sulfur-
oxidizing bacterial symbionts either within the
tissues (endosymbiotic) or attached to the sur-
faces (episymbiotic) of most vent taxa, including
vesicomyid clams, mytilid mussels, shrimp, and
polychaete worms (Cavanaugh, 1994; Nelson
and Fisher, 1995a; Van Dover, 2000).
Recognizing the ubiquity of chemoau-
totrophic symbioses at hydrothermal vents,
researchers searched for and discovered similar
symbiotic associations in other marine habitats,
including coastal and subtidal reducing sedi-
ments (e.g., Felbeck et al., 1981b; Southward et
al., 1981; Southward, 1982; Cavanaugh, 1983;
Giere, 1985; Bauer-Nebelsick et al., 1996), brine
and hydrocarbon seeps (Sibuet and Olu, 1998),
and whale skeletons (Bennet et al., 1994; Smith
and Baco, 2003), thereby extending the host
Fig. 2. Symbiont-containing host organisms from hydrother- taxa to include solemyid and lucinid bivalves,
mal vents and cold seeps. (A) Calyptogena magnifica shell; pogonophoran tubeworms, echinoids and cili-
courtesy of Emilio Jorge Power. (B) Filamentous bacteria on ates. In addition, methanotrophic bacteria were
sulfide deposits from the East Pacific Rise. (C) Bathymodio-
detected in a marine sponge (Vacelet et al.,
lus childressi from the Gulf of Mexico; courtesy of the
National Oceanic and Atmospheric Administration. (D)
1995), a pogonophoran tubeworm (Schmaljo-
Riftia pachyptila at the East Pacific Rise. hann and Flgel, 1987), and in vent and seep
mussels, sometimes co-occurring with sulfur-
oxidizing chemoautotrophs in a dual symbiosis
microorganisms and dense aggregations of inver- (Childress et al., 1986; Cavanaugh et al., 1987;
tebrates (Lonsdale, 1977; Grassle, 1985; Van Cavanaugh, 1993).
Dover, 2000). Researchers first argued that vent
invertebrates achieved high densities by filtering
organic matter, which was presumably trans-
Habitat Chemistry
ported to vent sites in hydrothermally-driven The seemingly disparate ecological niches where
convection cells (Lonsdale, 1977). A second these symbioses are found all possess a chemical
hypothesis suggested that the invertebrate com- gradient, or chemocline, which chemosynthetic
munity fed directly on locally dense populations bacteria exploit for energy production. Chemo-
of free-living chemoautotrophic bacteria clines form where reduced, high-energy com-
(Lonsdale, 1977; Corliss et al., 1979). pounds such as sulfide or methane (typically
However, studies of the giant tubeworm, Riftia produced in anoxic habitats, including vent fluids
pachyptila (Fig. 2), whose lack of mouth and gut and sediments) underlie an oxic water column.
precludes suspension feeding, suggested that Chemosynthetic microorganisms must bridge
sulfide-oxidizing endosymbiotic bacteria might the oxic-anoxic interface to access both the
contribute substantially to the vent food web. reduced compounds (e.g., H2S) used as an energy
Cavanaugh et al. (1981) proposed that symbiotic source and the oxygen to which electrons are
chemosynthetic bacteria occurred in R. pachyp- shuttled in aerobic energy metabolism.
tila. Microscopic and biochemical evidence The source of reduced compounds used in
indicated Gram-negative bacteria were present, chemoautotrophic energy metabolism differs
packed within the tubeworm trophosome among habitats. In marine sediments microbial
(Cavanaugh, 1981; Fig. 3), a highly vascularized sulfate reduction (in which SO42 is used as an
organ in the tubeworm trunk in which activities electron acceptor during the oxidation of organic
of enzymes involved in sulfide oxidation and car- matter) dominates, and the sulfide utilized by
bon fixation were also detected (Felbeck, 1981a). thioautotrophic symbioses (e.g., involving lucinid
In addition, Rau (1981) used stable isotope clams or solemyid protobranchs) in these habi-
signatures to show a nonphotosynthetic carbon tats is of biogenic origin. In contrast, sulfide at
478 C.M. Cavanaugh et al. CHAPTER 2.5

Fig. 3. Riftia pachyptila Jones (AC:


A D Galapagos Rift; D, E: 21N, East
Pacific Rise). (A) Photograph show-
ing elemental sulfur crystals (arrows)
scattered throughout trophosome;
courtesy of M. L. Jones. (B) Scanning
electron micrograph showing lobules
b of trophosome; arrow indicates area
of C (below) where surface epithelium
was removed to reveal symbionts
within trophosome. (C) Same, higher
magnification, showing symbionts
within trophosome; note spherical
cells as well as rod-shaped cells (small
arrows); large arrows indicate likely
B host cell membranes. (D) Cross-
section of portion of trophosome
lobule showing variable fine structure
b of symbionts, including membrane-
m
bound vesicles in many cells; all
symbionts contained within mem-
brane-bound vacuoles, either singly or
in groups of two or more; arrow: divid-
ing bacterium; b: bacteria; m: mito-
tc chondria; tc: trunk coelomic cavity.
(E) Same, higher magnification, show-
C E ing cell envelope of symbiont (resem-
bling that of Gram-negative bacteria),
pm intracytoplasmic vesicles, and peribac-
terial membrane; v: vesicle; cm: sym-
om biont cytoplasmic membrane; om:
symbiont outer membrane; pm: perib-
cm acterial membrane. Scale bars: A, 1
m; B, 250 m; C, 10 m; D, 3 m; and
E, 0.2 m. From Cavanaugh (1985),
with permission.

hydrothermal vents is produced by the geother- 7506500 mol/kg of Fe (Elderfield and Schultz,
mal reduction of seawater sulfate and by the 1996). As it exits the seafloor and mixes with the
interaction between geothermally heated water ambient bottom oxygenated seawater (pH, ca. 8;
and sulfur-containing rocks (e.g., basalt; Alt, temperature = 1.8C; [O2], ca. 110 M), metallic
1995; Elderfield and Schultz, 1996; Rouxel et al., sulfides precipitate out resulting in black
2004). Seawater that percolates into the develop- smokers (reviewed in Elderfield and Schultz,
ing crust becomes heated and reacts with oceanic 1996). Vent organisms are typically found clus-
basalt, becoming enriched with metals and sul- tered around more diffuse or low flow vents,
fide and charged with volcanic gases such as which are caused by ambient seawater mixing in
methane and carbon dioxide. Heated vent water the shallow subsurface with vent fluid. These
then exits with concentrations of reduced com- vents are characterized by a higher pH (ca. 6),
pounds orders of magnitude higher than in lower temperatures (1.8 to ca. 40C), and, conse-
ambient seawater. Hydrothermal effluent is hot quently, lower concentrations of reduced chemi-
(temperatures up to 350400C), acidic (pH ~ 3), cals (Van Dover, 2000).
anoxic, and can contain 312 mmol/kg of H2S, The relative acidity of vent fluid (black smok-
25100 mol/kg of CH4, and 0.051 mmol/kg of ers: pH, ca. 3; diffuse flow vents: pH, ca. 6) sig-
H2, as well as 3601140 mol/kg of Mn and nificantly impacts the concentration of inorganic
CHAPTER 2.5 Marine Chemosynthetic Symbioses 479

chemicals available to chemosynthetic symbio- of stable isotope signatures of symbiont-


ses. For example, carbon dioxide (CO2), bicar- containing and symbiont-free host tissue, pro-
bonate (HCO3), and carbonate (CO32), the vide valuable insight into the trophic dynamics
three distinct chemical species of the dissolved of symbiont-based communities. Molecular
inorganic carbon (DIC) pool, vary in relative approaches, such as polymerase chain reaction
abundance depending on pH; pKa values for (PCR)-based gene probing, 16S rRNA gene
these compounds are 6.4 for CO2 : HCO3 and analysis, and fluorescent in situ hybridization
10.3 for HCO3 : CO32 at 25C (Stumm and (FISH), are increasingly used to characterize the
Morgan, 1996). Thus CO2, which diffuses freely systematic relationships of symbiont and host
across biological membranes and is the DIC spe- species (e.g., Distel et al., 1995; Peek et al., 1998;
cies fixed by chemoautotrophic symbionts utiliz- Dubilier et al., 1999) and the metabolism and
ing the Calvin cycle, is readily available at vents gene flow of the bacterial symbionts (Robinson
(Cavanaugh and Robinson, 1996; Goffredi et al., et al., 1998; Lee et al., 1999; Millikan et al., 1999;
1997b). In addition, sulfide exists at three levels Podar et al., 2002). Molecular techniques have
of dissociation (H2S, HS, and S2) depending on also been used to detect symbiont transmission
pH, with pKa values of 7.0 for H2S : HS and 12.9 modes (Cary and Giovannoni, 1993a; Krueger,
for HS : S2 at 25C (Stumm and Morgan, 1996). 1996a) as well as symbiont abundance (Polz and
Therefore, in the relatively acidic vent fluids Cavanaugh, 1995).
sulfide occurs predominantly as H2S. In such
effluent, total sulfide (H2S, HS, and S2) concen-
Summary
tration correlates positively with temperature
(Johnson et al., 1988); conversely, higher temper- This chapter reviews symbiotic associations
atures (>30C) may facilitate reactions between between chemosynthetic bacteria and marine
sulfide and other dissolved elements (such as invertebrate and protist hosts. A bias toward
iron) that reduce free sulfide availability (Luther symbioses between chemoautotrophic bacteria
et al., 2001). The chemical environment (i.e., and hydrothermal vent invertebrates is evident,
concentrations of CO2, O2, H2S, CH4, H+, and primarily because our knowledge of marine bac-
dissolved metals) is therefore expected to signif- terial symbioses stems largely from studies of
icantly influence the ecology and evolution of vent fauna done in the 27 years following the
chemosynthetic symbioses. discovery of these unique organisms. But despite
this impressive body of research, much about
Methods for Studying these marine symbioses remains to be revealed.
In conjunction with several earlier reviews that
Chemosynthetic Symbioses
provide a thorough and thoughtful treatment of
To date, the bacteria involved in these symbiotic symbioses occurring at hydrothermal vents and
associations have not yet been isolated and cold seeps (Fisher, 1990; Felbeck and Distel,
grown in pure cultureperhaps because the 1991; Childress and Fisher, 1992; Cavanaugh,
unique environment encountered in situ by 1994; Nelson and Fisher, 1995a), the following
chemosynthetic symbionts has not been recre- chapter presents an overview of the ecology,
ated or because a reduced genome, characteristic physiology and evolution of chemosynthetic
of many endosymbionts, has precluded growth symbioses.
outside of the host. Symbiotic bacteria are there-
fore studied indirectly, using methods to assess
their physiology, ecology and phylogeny within Host Diversity
the context of the intact symbiosis. Traditionally,
researchers identify chemosynthetic symbioses Chemosynthetic bacteria are known to associ-
using a combination of microscopy (light, confo- ate with a diversity of invertebrate hosts (six
cal, scanning and transmission electron), which phyla), as well as with ciliate protists (Table 1
provides visual information on the location, mor- and references therein). To date, the majority
phology and ultrastructure of symbionts, and of the symbionts characterized via 16S rRNA
enzyme assays, which detect and quantify the phylogenetic analyses fall within the Gamma-
activity of key proteins involved in chemoau- proteobacteria division (Fig. 1; discussed fur-
totrophic (e.g., ribulose 1,5-bisphosphate car- ther below). The intimacy of these associations
boxylase-oxygenase) or methanotrophic (e.g., varies among taxa. The bacterial partners may
methanol dehydrogenase) metabolism. In addi- be episymbionts living on the surface of the
tion, tracing the incorporation of radiolabeled host (e.g., on shrimp, nematodes, sponges, lim-
substrates (e.g., carbon dioxide, methane and pets and ciliates; Figs. 46), or endosymbionts
nitrogen species) within the host helps define the living either extracellularly within host tissue
physiology of the host-bacteria partnership. Such (e.g., in oligochaetes; Fig. 7) or in specialized
physiological assays, in conjunction with analyses host cells and organs (e.g., in bivalves and
480

Table 1. List of invertebrate taxa hosting chemoautotrophic or methanotrophic bacterial symbionts.a


Common Symbiont-containing
Group name tissue Location Habitat Symbiont type References
Protozoa
Class Ciliata Ciliate NA Epibiotic Cold seeps, mangrove Chemoautotroph Bauer-Nebelsick et al., 1996
swamp Ott et al., 1998
C.M. Cavanaugh et al.

Buck et al., 2000


Fenchel and Finlay, 1989
Porifera
Class Demospongiae Sponge NA Extracellular Cold seeps Methanotroph Vacelet et al., 1995, 1996
Family Cladorhizidae
Nemata
Subfamily Stilbonematinae Nematode Cuticle Epibiotic Reducing Chemoautotroph Schiemer et al., 1990
sediments Polz et al., 1992
Mollusca
Class Bivalvia
Subclass Protobranchia
Family Solemyidae Clam Gills Intracellular Reducing sediments, Chemoautotroph Cavanaugh, 1983
hydrothermal vents,b Fisher and Childress, 1986
cold seepsc Conway et al., 1989
Subclass Heterodonta
Family Lucinidae Clam Gills Intracellular Reducing sediments, Chemoautotroph Giere, 1985
cold seepsc Schweimanns and Felbeck, 1985
Family Thyasiridae Clam Gills Intracellular Reducing sediments, Chemoautotroph Dando and Southward, 1986
cold seepsc Herry and Le Pennec, 1987
Family Vesicomyidae Clam Gills Intracellular Hydrothermal Chemoautotroph Cavanaugh, 1983
vents, cold seeps Rau, 1981
Subclass Pteriomorphia
Family Mytilidae Mussel Gills Extracellular Hydrothermal Chemoautotroph Childress et al., 1986
vents, cold seeps and/or Cavanaugh et al., 1987
Intracellular methanotroph Fisher et al., 1988
Cavanaugh et al., 1992
Class Gastropoda
Family Provannidae Snail Gills Intracellular Hydrothermal vents Chemoautotroph Stein et al., 1988
Endow and Ohta, 1989
CHAPTER 2.5
Common Symbiont-containing
Group name tissue Location Habitat Symbiont type References
Family Lepetodrilidae Limpet Gills Epibiotic Hydrothermal vents Chemoautotroph de Burgh and Singla, 1984
Fox et al., 2002
CHAPTER 2.5

Bates et al., 2004


Annelidad
Class Polychaeta Worm Dorsal Epibiotic Hydrothermal vents Chemoautotroph Desbruyeres et al., 1983, 1985
Family Alvinellidae surface Cary et al., 2003
Family Siboglinidae Tubeworm Trophosome Intracellular Deep-sea hydrothermal Chemoautotroph Cavanaugh et al., 1981
(Vestimentifera and vents, cold seeps, Felbeck, 1981
Pogonophora) reducing sediments, Southward et al., 1981
fjords Brooks et al., 1987
Schmaljohann and Flgel, 1987
Southward and Southward, 1988
de Burgh et al., 1989
Class Clitellata
Subfamily Phallodrilinae Oligochaete Subcuticular Extracellular Coralline sands Chemoautotrophe Felbeck et al., 1983
Giere, 1981, 1985
Giere and Langheld, 1987
Arthropoda
Class Crustacea
Family Alvinocarididae Shrimp Carapace Epibiotic Hydrothermal vents Chemoautotroph Van Dover et al., 1988
Polz and Cavanaugh, 1995
Polz et al., 1999
Echinodermata
Class Echinoidea Sea urchin Gut Extracellular Reducing sediments Chemoautotrophf Temara et al., 1993
Brigmon and de Ridder, 1998
Abbreviation: NA, not applicable.
a
Chemosynthetic status of symbionts inferred from ultrastructural, physiological, enzymatic, and molecular data.
b
A solemyid protobranch, Acharax alinae, has been described from the Lau Basin hydrothermal vents, but the symbiosis has not been characterized (Metivier and Voncosel, 1993).
c
Solemyid clams have been collected from cold seeps in the eastern Pacific, and lucinid and thyasirid clams have been collected from cold seeps in the Gulf of Mexico, Sagami Bay, and
Barbados Prism, though the presence of symbionts has not been formally described (Sibuet and Olu, 1998).
d
Though it is now accepted that the pogonophoran and vestimentiferan worms are not separate phyla but members of phylum Annelida, they are listed as separate groups for identification
purposes.
e
The oligochaete Olavius algarvensis has been shown to have an additional symbiont which is a sulfur reducing bacterium and member of the delta Proteobacteria (Dubilier et al., 2001).
Olavius loisae has been shown to host an alpha proteobacterium and a spirochete as well as a chemoautotroph (Dubilier et al., 1999).
f
This symbiont has been described as Thiothrix-like on the basis of morphology, physiology, and immunological assays (Temara et al., 1993; Brigmon and de Ridder, 1998).
Marine Chemosynthetic Symbioses
481
482 C.M. Cavanaugh et al. CHAPTER 2.5

A B

10 m
100 m

C D

1000 m 100 m

E F

5 m 1 m

G H

1 m 10 m

Fig. 4. Scanning electron micrographs showing the morphological diversity of ectosymbiotic bacteria on the colonial ciliate
Zoothamnium niveum (A, B), the shrimp Rimicaris exoculata (C, D), and the nematodes within the subfamily Stilbonema-
tinae (E-H). (A) Entire ciliate colony with zooids attached to a common stem, and (B) bacterial epigrowth on an individual
zooid. (C) Shrimp appendage covered by dense arrays of filamentous bacteria, and (D) detail of the hair-like bacterial
covering. Epigrowth on different species of nematodes showing (E) irregular epigrowth of two morphological types on
Robbea sp., (F) coccoid bacteria forming multilayers on Stibonema sp., (G) upright standing, longitudinally dividing rods on
Laxus oneistus, and (H) dense array of nonseptate filaments that can reach up to 100 mm in length on Eubostrichus dianae.
From Polz et al. (2000), with permission.

vestimentiferan tubeworms; Figs. 2, 3, and 8). fit from these intimate associations. Indeed,
In intracellular endosymbioses the symbionts as in the giant vent tubeworms, the digestive
are housed within specialized host cells called system of many endosymbiont-containing
bacteriocytes and are contained within a host- marine invertebrates is either reduced (e.g., in
derived membrane bound vacuole (Cavanaugh, coastal solemyid protobranchs) or absent alto-
1983; Cavanaugh, 1994; Fisher, 1990). Host gether (e.g., in oligochaetes and vestimentiferan
morphology clearly suggests a nutritional bene- and pogonophoran tubeworms), consistent with
CHAPTER 2.5 Marine Chemosynthetic Symbioses 483

host dependence on the symbiont for part or all In mollusk symbioses the bacteria occur only in
of its nutrition. the gills; bacteria have been found within gill
All of the members of the tubeworm family epithelial cells of solemyid protobranchs
Siboglinidae examined to date, including the ves- (Cavanaugh, 1983; Krueger et al., 1996b; Figs.
timentiferan and the smaller pogonophoran
tubeworms, contain intracellular symbionts.
Most of these symbionts are inferred to be A
chemoautotrophic, but methanotrophs have CU
been found in one host species (Siboglinum
poseidoni; Schmaljohann and Flgel, 1987). The
vent tubeworm Riftia pachyptila and other vesti-
mentiferan and pogonophoran tubeworm spe-
cies possess a unique morphological adaptation
to accommodate their symbionts. Tubeworm
bacteria reside within a lobular and highly vas-
cularized organ (the trophosome) that occupies S
most of the tubeworm trunk and functions spe- m
C
cifically to house bacteria (Cavanaugh, 1981;
Felbeck, 1981a; Figs. 3 and 9). The symbiosis is
obligate for these worms, as they are mouthless
and gutless as adults and depend on their inter-
nal bacteria for their nutrition.
Chemosynthetic symbioses are widespread
within the Mollusca and have been detected in
five bivalve and two gastropod families (Table 1).

CU

n bs
es

b
Fig. 7. Inanidrilus leukodermatus. (A) Light micrograph of a
10 m
cross section of an oligochaete worm. (B) Transmission elec-
tron micrograph of symbiont-containing region just below
Fig. 5. Lepetodrilus fucensis. Transverse section of gill tissue the cuticle. Note smaller and larger symbiont morphotypes
from the hydrothermal vent limpet showing episymbiotic fil- (smaller and larger arrows, respectively). c, coelomic cavity;
amentous bacteria partially embedded in the host epithelium. m, muscle tissue; s, symbiont-containing region between cuti-
b, bacteria; es, extracellular space; n, nucleus; bs, blood space. cle and epidermis; cu, cuticle. Scale bars: A, 20 m; B, 2 m.
Courtesy of Amanda Bates. From Dubilier et al. (1995), with permission.

A B

Fig. 6. Eubostrichus cf. parasitiferus. Scanning electron micrographs showing the symbiotic bacteria on the surface of the
nematode. (A) Anterior (bottom) and posterior (top) end with symbionts arranged in a characteristic helix. (B) Higher
magnification. Bacteria are attached with both ends to the worms cuticle. Note the increasing length of the cells from proximal
to distal along the worms surface. Scale bars: A, 20 m; B, 2 m. From Polz et al. (1992), with permission.
484 C.M. Cavanaugh et al. CHAPTER 2.5

Fig. 8. Calyptogena magnifica Boss


A mv and Turner (21N East Pacific Rise).
ni (A) Transmission electron micro-
graph of slightly oblique transverse
section of gill filament, showing
coccoid-shaped bacteria within gill
bacteriocyte and intercalary cells
lacking symbionts; b: bacteria; mv:
microvilli (of both cell-types); nb:
b nucleus of bacteriocyte; ni: nucleus of
intercalary cell. (B) Same, higher
magnification, transverse section of
coccoid-shaped symbionts, showing
cell ultrastructure typical of Gram-
negative bacteria and peribacterial
membrane (arrow). Scale bars: A,
5 m; B, 0.25 m. From Cavanaugh
(1985), with permission.

B
nb

A B

b Fig. 9. Escarpia spicata Jones (San


Clemente Fault). (A) Transmission
electron micrograph, portion of tro-
phosome lobule showing numerous
coccoid- to ovoid-shaped bacterial
nb symbionts, some of which appear
intracellular. (B) Same, higher
magnification, showing bacterial cell
envelope (resembling that of Gram-
negative bacteria) and intracytoplas-
mic membrane-bound vesicles; arrow:
peribacterial membrane; b: bacteria;
nb nb: nucleus of bacteriocyte. Scale
bars: A, 10 m; B, 1 m. From
Cavanaugh (1985), with permission.

1013), lucinid clams (Cavanaugh, 1983; Felbeck, totrophic symbionts have been detected only in
1983a), thyasirid clams (Felbeck et al., 1981b; members of the subfamily Bathymodiolinae,
Cavanaugh, 1983; Arp et al., 1984), vesicomyid which are found exclusively in the deep-sea.
clams (Boss and Turner, 1980; Rau, 1981; Arp et Further, dual symbioses involving both methan-
al., 1984; Fig. 8), mytilid mussels (Fiala-Mdioni, otrophs and chemoautotrophs are restricted to
1984; Le Pennec and Hily, 1984; Figs. 2 and 14), species of deep-sea bathymodioline mussels col-
and provannid gastropods (Stein et al., 1988; lected from methane seeps and hydrothermal
Windoffer and Giere, 1997). Within certain vents (Cavanaugh, 1994; Nelson and Fisher,
mollusk families (e.g., Solemyidae, Lucinidae 1995a; Van Dover, 2000; Fig. 14).
and Thyasiridae), all species examined form Chemosynthetic bacteria also occur as episym-
symbioses with chemoautotrophic bacteria. In bionts on marine invertebrates (Table 1; Fig. 4).
other families, such as the Mytilidae, chemoau- These symbionts include the Epsilonproteo-
CHAPTER 2.5 Marine Chemosynthetic Symbioses 485

Fig. 10. Solemya sp. (right hand) collected from deep-sea


vent sites (2380 m depth) along the subduction zone off Ore-
gon, and Solemya velum (left hand) collected from subtidal
reducing sediments (<1 m depth, mean low tide) of Massa-
chusetts eelgrass beds. Photo courtesy of Dr. Ruth D. Turner.

Fig. 11. Solemya velum. Characteris-


tic Y-shaped burrows dug by the
coastal protobranch clam to bridge
the oxic-anoxic interface and access
both reduced sulfur (from below) and
oxygen (from above). From Stanley
(1970), with permission. PLATE 3. SOLEMYA

A
mv

i b

Fig. 12. Solemya borealis. (A)


Transverse section of gill filaments
showing intracellular rod-shaped c
bacteria (arrows, rectangle). Bacterio-
cytes are confined to the region prox-
imal to the ciliated edge of the gill and bl
are flanked by symbiont-free interca-
lary cells that appear to comprise the
microvillar surface of the gill filament.
Light micrograph. b: bacteriocyte ci
nucleus; c: ciliated cell nucleus; i:
intercalary cell nucleus; bl: blood
B
space; ci: cilia; mv: microvilli. (B)
Higher magnification of symbionts
showing cell ultrastructure typical of
Gram-negative. Inset: Detail of sym-
biont cell envelope and peribacterial
membrane. p: peribacterial mem- cm
brane; cm: cell membrane; om: outer om
membrane. Scale bars: A, 20 m; B,
1 m; inset, 0.05 m. From Conway p
et al. (1992b), with permission.
486 C.M. Cavanaugh et al. CHAPTER 2.5

Fig. 13. Solemya velum. (A) Trans-


A B
m mission electron micrograph, trans-
ni
verse section of gill filament, showing
rod-shaped bacteria within gill bacte-
nb riocyte and intercalary cells lacking
symbionts; b: bacteria; mv: microvilli;
nb: nucleus of bacteriocyte; ni:
nucleus of intercalary cell. (B) Same,
higher magnification, transverse
section of rod-shaped bacterium,
showing cell ultrastructure typical of
b Gram-negative bacteria and peri-
bacterial membrane (arrows). Scale
mv bars: A, 3 m; B, 0.2 m. From
Cavanaugh (1985), with permission.

A C B

I
S

Fig. 14. Bathymodiolus puteoserpentis. Transverse section of


Mid-Atlantic Ridge (MAR) mytilid gill filament, showing
l symbiont-containing gill epithelial cells (bacteriocytes). (A)
n
Diagram of gill filament. Bacteriocytes are confined to the
region proximal to the ciliated border of the gill. Small box
shows positions of Figs. 14B. B: symbiont-containing bacte-
riocyte region; and C: symbiont-free ciliated region. (B)
Transmission electron micrograph. Large and small sym-
bionts (large and small arrows, respectively) are located in
the apical region of the cells, while nuclei and lysosomal
residual bodies occupy the region closest to the blood sinus.
Note centrally stacked intracytoplasmic membranes in large
symbionts. 1, Lysosomal residual body; n, bacteriocyte
nucleus; and s, blood sinus. Scale bar: 5 m. From Distel et
S al. (1995), with permission.

bacteria that cover the cuticle of Rimicaris and the surfaces of alvinellid polychaetes
shrimp, dominant members of the metazoan (Desbruyres et al., 1985; Cary et al., 1997).
fauna at vents on the Mid-Atlantic Ridge (MAR; Chemosynthetic episymbionts also associate
Polz et al., 1998) and the Central Indian Ridge with nematodes (Weiser, 1959; Ott et al., 1991;
(CIR; Van Dover et al., 2001; Van Dover, 2002a), Polz et al., 1992; Polz et al., 1994) and ciliates
CHAPTER 2.5 Marine Chemosynthetic Symbioses 487

(e.g., Fenchel and Finlay, 1989; Bauer-Nebelsick ferences in the lifecycle stage and metabolism
et al., 1996). In addition, methanotrophic epi- among symbiont cells (Bright et al., 2000).
symbionts have been found living on a deep-sea
sponge (Vacelet et al., 1995; Vacelet et al., 1996).
Vent limpet-bacteria associations seem to be an
Symbiont Phylogeny
intermediate between epi- and endosymbioses; While chemoautotrophic symbionts have consis-
bacteria exist partially embedded in the limpet tently evaded culture, the suite of cellular and
gill epidermis and may be endocytosed or fed on molecular methods used to characterize these
by the host (de Burgh and Singla, 1984; Bates bacteria has revealed startling evolutionary
et al., 2004; Fig. 5). Some epibiont communities, trends. Investigators have successfully sequenced
like those residing on the Rimicaris shrimp and 16S rRNA genes from symbiont-containing tis-
the nematode Laxus sp., are dominated by a sue and subsequently confirmed the symbiont
single phylotype (Polz et al., 1994; Polz and origin of these sequences via hybridization with
Cavanaugh, 1995), while others are quite diverse symbiont-specific probes. In contrast to the wide
(Polz et al., 1999; Campbell et al., 2003). But diversity of host taxa involved in these symbio-
given that morphological plasticity often belies ses, chemosynthetic symbionts cluster primarily
the phylogenetic identity of symbionts (Polz et within a single bacterial division, the Gamma-
al., 1999; Giere and Krieger, 2001), symbiont proteobacteria, on the basis of 16S rRNA
diversity estimates are only appropriate when gene sequences (Distel and Cavanaugh, 1994;
putative symbiont phylotypes are confirmed Dubilier et al., 1999; McKiness, 2004). Such anal-
using hybridization methods (e.g., FISH). yses have also shown that most host species typ-
ically form a relationship with a unique symbiont
phylotype. But this clearly is not always the case.
For example, while strain level variation may
Symbiont Diversity occur, vestimentiferan tubeworms belonging to
the genera Riftia, Tevnia, and Oasisia appear to
Morphology and Ultrastructure
share a single, or very similar, symbiont phylo-
Symbiont morphology varies among functional type based on 16S rRNA gene sequences
types (chemoautotroph vs. methanotroph), (Feldman et al., 1997; Laue and Nelson, 1997; Di
among phylotypes within the same functional Meo et al., 2000; Nelson and Fisher, 2000;
group, and among individuals in a population of McMullin et al., 2003), as do some species of
a single phylotype. The symbionts all have a tropical lucinid clams (Durand and Gros, 1996a;
Gram-negative cell envelope but range from Durand et al., 1996b).
small (ca. 0.25 m diameter) coccoid endosym- Recently, McKiness (2004) reported phyloge-
bionts within mussel gills (Cavanaugh, 1985; netic analyses of 16S rRNA gene sequences from
Dubilier et al., 1998) to large (>10 m length) chemosynthetic symbionts within the Gamma-
rod-shaped and filamentous episymbionts on proteobacteria. This study represented the most
vent shrimp (Hentschel et al., 1993b; Polz and comprehensive analysis of chemoautotrophic
Cavanaugh, 1995; Fig. 4). Some bathymodioline symbionts to date. It included 39 symbiont
mussels host two metabolically distinct sym- sequences and over 30 sequences from free-living
bionts: small (<0.5 m) chemoautotrophs and bacteria representatives of chemoautotrophs,
larger (1.52.0 m) methanotrophic bacteria methanotrophs, and marine bacterioplankton.
possessing stacked intracytoplasmic membranes, Here, an expanded phylogenetic analysis is pre-
which are typical of Type I methanotrophs (e.g., sented that includes the Epsilonproteobacteria
Childress et al., 1986; Cavanaugh et al., 1987; symbionts of shrimp and alvinellid worms (Fig.
Cavanaugh et al., 1992; Fiala-Mdioni et al., 1). This consensus tree illustrates the strong level
2002; Pimenov et al., 2002; Fig. 14). Morpholog- of resolution afforded by the 16S rRNA gene
ical diversity can also occur throughout mono- and shows that almost all of the chemosynthetic
specific populations within a single host animal. symbionts for which sequence data are available
For example, populations of sulfur-oxidizing cluster into two main clades. The first clade
chemoautotrophic symbionts in the tubeworm includes symbionts of lucinid and thyasirid
Riftia pachyptila contain distinct morphotypes clams, solemyid protobranchs, tubeworms, nem-
that vary in abundance depending on location atodes, and oligochaetes, and the second clade
within the trophosome lobule (Bright et al., includes the mytilid mussel and vesicomyid clam
2000); small, rod-shaped bacteria occur primarily symbionts.
in the innermost zone of the lobule nearest the Though the first clade as a whole has relatively
hosts axial blood vessel, while small and large low bootstrap support, smaller clusters within
cocci (1.610.7 m diameter) occupy zones the first clade are strongly supported. For exam-
nearer the periphery of the trophosome (Bright ple, the monophyletic cluster of nematode and
et al., 2000). Such variability may relate to dif- oligochaete symbionts has 100% bootstrap sup-
488 C.M. Cavanaugh et al. CHAPTER 2.5

port. Similarly, the vestimentiferan tubeworm The symbionts of the thyasirid clam Maorhith-
symbiont clade also has high bootstrap support, yas hadalis occupy unique positions in this phy-
corroborating prior evidence that these worms logenetic framework. On the basis of 16S rRNA
share a single or very similar symbiont phylo- sequence data and in situ hybridization, Fujiwara
type, which is consistent with environmental et al. (2001) described two different symbionts
transmission of symbionts (Feldman et al., 1997; within this clam. One of the symbionts shows
Laue and Nelson, 1997; Di Meo et al., 2000; evolutionary relatedness to the bathymodioline
Nelson and Fisher, 2000; McMullin et al., 2003; mussel and vesicomyid clam symbionts, occur-
see the section Ecology and Evolution in this ring basal to the clade containing these bacteria,
Chapter). In contrast, the clam symbionts exhibit and the other symbiont clusters with the free-
a more complicated relationship. The lucinid living Thiomicrospira spp. This free-living
clam symbionts form a paraphyletic group; some symbiont phylotype, however, may be a con-
are sister to the tubeworm symbionts while oth- taminant. Difficulties with in situ hybridization
ers group with thyasirid and solemyid symbionts. have precluded attempts to describe the micro-
The solemyid symbionts show similarly compli- distribution of the two symbiont types (Fujiwara
cated relationships, as they are polyphyletic and and Uematsu, 2002), bringing into question the
scattered throughout the first clade. Solemya phylogenetic identity of the clam symbionts.
velum and S. occidentalis symbionts cluster with The phylogenetic positions of the methan-
the nematode and oligochaete symbionts, while otrophic endosymbionts and the filamentous
S. terraeregina and S. pusilla symbionts cluster epibionts are shown also in Fig. 1. The methan-
with lucinid and thyasirid clam symbionts. Thus, otrophic symbionts characterized to date all
this disjointed distribution does not suggest belong to the Gammaproteobacteria, forming a
cospeciation between host taxa and symbionts in clade with 100% bootstrap support. The sister
this first clade and indicates that there were mul- group of this clade consists of free-living Type I
tiple initiations of symbiosis within the solemyid methanotrophs (Methylococcus, Methylobacter
and lucinid clams. and Methylomonas spp.). Given their mono-
The position of the S. reidi symbiont within phyly, the mussel symbionts apparently arose
this first main group is curious; this symbiont falls from a common ancestor. But the question of
at the base of this first main group, clustering whether these symbionts subsequently cospeci-
with an intracellular pathogen, Coxiella burnetti, ated with their hosts remains unanswered.
and an environmental clone from a Japanese The episymbionts in this analysis include the
cold seep, Gamma JTB254 (discussed further sulfur-oxidizing Epsilonproteobacteria found on
below). This symbiont sequence has held a basal the Mid-Atlantic Ridge shrimp Rimicaris exocu-
position in other analyses (see Bayesian analysis lata and the eastern Pacific polychaete Alvinella
in McKiness, 2004) and provokes questions con- pompejana. Interestingly, the shrimp epibiont
cerning the nature of symbiosis in protobranch clusters with the polychaete epibionts despite the
bivalves. As additional sequences become avail- fact that R. exoculata occurs on the Mid-Atlantic
able, it will be necessary to reassess the position Ridge while alvinellid polychaetes inhabit vents
of the S. reidi symbiont with respect to other in the eastern Pacific Ocean.
chemoautotrophic symbionts. Free-living microorganisms can potentially
The second clade of symbionts, which includes provide insight into the ancestral form of endo-
the mytilid mussel and vesicomyid clam sym- symbionts. For instance, the evolution of insect
bionts from vents and cold seeps, shows 100% endosymbionts (e.g., Wolbachia and Buchnera
bootstrap support (support for the mussel and spp.) is commonly studied by comparative anal-
vesicomyid clades being 70% and 99%, respec- yses with free-living, closely related microbes
tively). The symbiont from the Central Indian (Wernegreen, 2002; Moran, 2003). But until
Ridge mussel, Bathymodiolus aff. brevior, falls at recently, the chemoautotrophic symbiont clades
the base of the vesicomyid clam symbionts, with have not included any free-living bacteria. Fig-
82% bootstrap support. In contrast to the first ure 1 includes two species of bacteria that are
symbiont clade (Fig. 1, top), evidence suggests not chemoautotrophic symbionts (Coxiella bur-
that an ancestral symbiont initiated symbioses netti and Achromatium oxaliferum) and an envi-
with both the vesicomyid clams and the bathy- ronmental clone (JTB254). All three of these
modioline mussels and predated the split sequences fall within the first symbiont clade.
between these bivalve lineages. On the basis of Coxiella burnetii is an intracellular pathogenic
16S rRNA gene sequence data, the divergence bacterium (Woldehiwet, 2004) and the Gamma
of the clam and mussel symbionts has been dated JTB254 clone was recovered from a deep-sea
to 125300 million years ago (Mya). This is cor- cold seep in the Japan Trench (Li et al., 1999).
roborated by the fossil record, which dates the They both fall in a cluster with the S. reidi sym-
bathymodioline mussel hosts to 150 Mya and the biont. Both C. burnetii and the S. reidi symbionts
vesicomyid clams to 100 Mya (Distel, 1998). are able to maintain an intracellular existence in
CHAPTER 2.5 Marine Chemosynthetic Symbioses 489

eukaryotic hosts. Coxiella burnetii, however, is shrimp, sea urchins, and ciliates (Felbeck et al.,
capable of growth in animal cell lines (e.g., 1981b; Cavanaugh, 1983; Cavanaugh et al., 1988;
Woldehiwet, 2004) and pathogenically infects a Polz et al., 1992; Johnson et al., 1994; Nelson
wide range of hosts (Niemczuk and Kondracki, et al., 1995b; Bauer-Nebelsick et al., 1996;
2004; Watanabe, 2004; Woldehiwet, 2004). In Krieger et al., 2000; Elsaied et al., 2002; Fiala-
addition, A. oxaliferum, a freshwater sulfur- Mdioni et al., 2002). While RubisCO has not
oxidizing bacterium, falls out with the nematode been detected in the alvinellid epibionts, genes
and oligochaete symbionts clade. Achromatium encoding citrate lyase, a key enzyme of the
oxaliferum occurs in freshwater sediments along reductive tricarboxylic acid (TCA) cycle,
the redox zone where it has access to sulfide and recently have been detected via analyses of
oxygen (Head et al., 1996; Glockner et al., 1999; symbiont DNA sequences, suggesting that the
Gray et al., 1999). As cultivation methods Epsilonproteobacteria symbionts of alvinellid
improve and sequences are added to the 16 worms fix carbon via this pathway (Campbell
rRNA gene database, other free-living bacteria et al., 2003).
that are closely related to symbionts will likely Enzymes involved in chemosynthetic energy
be identified. Indeed, recent studies incorporat- generation have also been used to characterize
ing 16S rRNA gene sequences from unidentified these symbionts. Although the sulfur metabolism
environmental clones into phylogenetic analyses enzymes are not unique to sulfur oxidation,
of free-living and symbiotic bacteria suggest that certain enzymes such as ATP sulfurylase, when
chemosynthetic symbionts may in fact resolve detected in high activities, have been used to
into three distinct clades (N. Dubilier, personal infer sulfur-based chemolithotrophy (Felbeck,
communication; Duperron et al., 2004). Free- 1981a; Fisher et al., 1993b; Laue and Nelson,
living relatives of chemosynthetic symbionts 1994). As methane monoxygenase, the enzyme
should reveal much about the ecological and that catalyzes the first step in the oxidation of
evolutionary constraints on the symbiont as well methane in aerobic methanotrophs, is notori-
as about the potential for gene loss during the ously labile (Prior and Dalton, 1985; Cavanaugh,
transition from the free-living to the symbiotic 1993), methanol dehydrogenase (MeDH), the
state. enzyme that catalyzes the second oxidation step
(i.e., methanol to formaldehyde) and is known to
occur only in methylotrophs, has been used
Symbiont Characterization extensively to diagnose methanotrophy. MeDH
has been detected in gill extracts of mussels host-
Enzyme Activities Researchers routinely dem- ing methanotrophs or both methanotrophs and
onstrate chemoautotrophy or methanotrophy in thioautotrophs (Cavanaugh et al., 1992; Fisher et
symbionts by the activity or presence of diagnos- al., 1993b; Robinson et al., 1998; Fiala-Mdioni
tic enzymes. Indeed, given the inability to culture et al., 2002; Pimenov et al., 2002; Barry et al.,
chemoautotrophic symbionts, detection of such 2002) and in a deep-sea sponge (Vacelet et al.,
enzymes is often the only evidence used to 1996). Such enzymatic evidence strongly
infer symbiont metabolism. This characterization suggests methanotrophy, particularly when
often involves physiological assays using tissue coupled with ultrastructural observations show-
or purified protein extract, immunodetection, or ing symbionts with the complex intracytoplasmic
PCR-based gene probing. Such studies were membranes that are characteristic of Type I
initially conducted on the tubeworm Riftia methanotrophs.
pachyptila. For example, Felbeck (1981a)
assayed tubeworm trophosome tissue for the
activities of key enzymes of the Calvin cycle, the Stable Isotope Signatures
CO2-fixing enzyme, ribulose 1,5-bisphosphate
carboxylase-oxygenase (RubisCO), and phos- Carbon Isotopes In addition to enzymology, sta-
phoribulose kinase (PRK) as well as enzymes ble isotope data provided some of the first evi-
associated with the oxidation of reduced inor- dence in support of chemoautotrophy in marine
ganic sulfur compounds. The activity or presence invertebrate-bacteria symbioses (e.g., Rau and
of RubisCO has subsequently been used to diag- Hedges, 1979; Spiro et al., 1986) and continue to
nose symbiont autotrophy in a diversity of host be useful in assessing symbiont metabolism and
species, including all of the chemoautotroph- tracking energy and carbon transfer in chemo-
harboring invertebrates listed in Table 1 (exclud- synthetic symbioses (Colaco et al., 2002; Levin
ing alvinellid polychaetes; see below): shallow and Michener, 2002; Van Dover, 2002a;
water solemyid and lucinid bivalves, vent and Robinson et al., 2003; Scott et al., 2004). Because
seep tubeworms and bivalves (including mussels enzymes involved in distinct carbon fixation
hosting both methanotrophic and chemoau- pathways discriminate differently against the use
totrophic symbionts), nematodes, oligochetes, of the heavier carbon isotope (13C), the stable
490 C.M. Cavanaugh et al. CHAPTER 2.5

carbon isotope ratio comparing 13C to 12C (13C) the fossil lucinid hosted chemosynthetic sym-
can be used to help distinguish different bionts ca. 120,000 ya (CoBabe, 1991). Thus, sta-
autotrophic metabolisms. For example, whereas ble carbon isotope analysis may be an effective
13C values of marine phytoplankton typically tool for tracing the evolution of chemosynthetic
vary between 18 and 28 (Fry and Sherr, symbioses in the fossil record.
1984; Gearing et al., 1984; Goericke et al., 1994), One of the main factors affecting stable
carbon derived chemosynthetically at vents is carbon isotope signatures of chemoautotroph-
either considerably lighter (enriched in 12C), invertebrate symbioses appears to be the form of
with 13C values from 27 to 35, or heavier RubisCO used by the symbionts. While 13C
(depleted in 12C), with values from 9 to values for vent bivalves hosting sulfide-oxidizing
16 (Childress and Fisher, 1992; Robinson symbionts cluster between 27 and 35
and Cavanaugh, 1995; Robinson et al., 2003). and resemble values for free-living chemoau-
Depending on the source of methane, symbioses totrophic bacteria, values for vent tubeworms,
between mussels and methanotrophic bacteria shrimp episymbionts, and many free living bac-
may be even more depleted in 13C, with 13C terial mats at vents are significantly heavier,
ranging from 37 to 78 (Cavanaugh, 1993; ranging from 9 to 16 (Childress and
Nelson and Fisher, 1995a; Barry et al., 2002). Fisher, 1992; Van Dover and Fry, 1994; Robinson
Because consumers generally retain the car- and Cavanaugh, 1995; Cavanaugh and Robinson,
bon isotopic signature of their food (i.e., you are 1996; Robinson et al., 2003). The difference
what you eat; DeNiro and Epstein, 1979), com- between these groups relates to the form of
parisons between 13C signatures of symbiont- RubisCO used to fix CO2 by the symbionts, with
containing host tissue and symbiont-free host form I RubisCO occurring in most members of
tissue can be used to study the transfer of sym- the isotopically lighter group and form II in all
biont-derived carbon to the host. For example, members of the heavier tubeworm group
13C values (30.8 to 35.8) in symbiont- (Robinson and Cavanaugh, 1995). Corro-
containing gill tissue from the western Pacific borating this hypothesis, Robinson et al. (2003)
vent mussel Bathymodiolus brevior were signifi- showed that the kinetic isotope effect ( value),
cantly lower than values from symbiont-free foot the relative rate of 13CO2 to 13CO2 fixation
tissue, suggesting that B. brevior supplements (12k/13k) and a measure of discrimination against
13
its diet via filter feeding on photosynthetically C by the purified RubisCO enzyme in vitro, is
derived carbon (Dubilier et al., 1998). In con- significantly lower for form II RubisCO from
trast, other studies show a high dependence on Riftia pachyptila symbionts ( = 19.5) than for
symbiont carbon by host species, including the form I enzyme ( = 2230). Such variation
coastal solemyid protobranchs (Fisher and may have arisen from evolution under differing
Childress, 1986; Conway and Capuzzo, 1991), concentrations of CO2 and O2 (Robinson et al.,
thyasirid clams (Dando and Spiro, 1993; Fiala- 2003).
Mdioni et al., 1993), and vestimentiferan tube- But isotopic discrimination by RubisCO does
worms (Kennicutt et al., 1992), as well as suggest not fully account for 13C-enrichment in these
differences in the contribution of methanotrophs symbioses. For example, to explain the discrep-
and chemoautotrophs to host carbon in mussels ancy in R. pachyptila biomass 13C values, Scott
containing dual symbioses (Cavanaugh, 1993; (2003) used a mass balance model to show that
Trask and Van Dover, 1999; Fiala-Mdioni et al., steep gradients in [CO2] among symbiont, host,
2002; Yamanaka et al., 2003). and environmental pools may drive 13C enrich-
Stable carbon isotope signatures have also ment. RubisCO, which occurs in the symbiont
been used to detect chemosynthetic symbioses cytoplasm, preferentially fixes 12CO2, leaving
13
in fossil bivalves of the clam family Lucinidae, CO2 behind. If fixation is rapid, CO2 equilibra-
whose extant members all host chemosynthetic tion between the isotopically lighter host cyto-
symbionts. CoBabe (1991), by determining the plasm and the isotopically heavier symbiont
13C values of organic matrix material extracted cytoplasm cannot occur, causing RubisCO to
from lucinid fossils dating to ca. 120,000 ya, draw from a more enriched 13CO2 pool and
showed that fossilized lucinid (Epilucina sp.) accounting for the relatively heavy 13C of tube-
shells (13C = 25) were about 5 lighter than worm biomass. Further, stable carbon isotope
values from other bivalves collected in the same values are also affected by the 13C of the envi-
deposit (Pt. Loma, CA). Also, the organic matter ronmental carbon pool (Fisher, 1995; Colaco et
of lucinid fossils was similar to that from modern al., 2002). Scott et al. (2004) demonstrated that
samples, implying that the fossil organic matrix the light 13C values of the symbionts of the
did not decay or change significantly over time. coastal protobranch clam Solemya velum are
These results, along with strong evidence show- explained not only by the kinetic isotope effect
ing that shell 13C values are reasonable proxies of symbiont form I RubisCO ( = 24.5) but
for tissue values in extant species, suggest that also by the 13C value of the CO2 in the sediment.
CHAPTER 2.5 Marine Chemosynthetic Symbioses 491

Similarly, the source of methane production Similarly, 15N values, because they vary pre-
can significantly impact the isotopic signature dictably and largely between producer and con-
of methanotroph symbioses (Fisher, 1995; sumer trophic levels (increase of ca. 3.4 per
MacAvoy et al., 2002). The 13C of methane var- level), are particularly useful markers for study-
ies considerably depending on whether it is pro- ing aquatic food web interactions (Minagawa
duced thermogenically (13C of CH4 >45) or and Wada, 1984). In general, 15N values of
biologically by methanogens (13C of CH4 chemoautotrophic organisms are significantly
<60; Lilley et al., 1993; Fisher, 1995), and this lighter (<0; Van Dover and Fry, 1994) than
variability is reflected in the 13C values of values for photosynthetic organisms (>6; see
chemosynthetic symbioses (Fisher, 1995). There- Michener and Schell [1994] and Fisher [1995]).
fore, in instances where the 13C of the source Researchers have used this discrepancy and the
methane is unknown, conclusions about the con- predictable trophic level fractionation of 15N to
tribution of methanotrophic symbionts to host show host reliance on symbiont-derived organic
13C values should be interpreted with caution matter in a number of symbioses including the
(Fisher, 1995). In addition, interpreting 13C sig- coastal clams Solemya velum and S. borealis
natures may be especially problematic for dual (Conway et al., 1989; Conway et al., 1992b) and
symbioses in which both methanotrophic and in vent mussels from the Mid-Atlantic Ridge
thioautotrophic symbionts co-occur in the same (MAR) and the Galapagos Rift (Fisher et al.,
host cell. In these symbioses the 13C values of 1988; Trask and Van Dover, 1999). In addition,
the symbionts and the host reflect a mix of meth- 15N values have been used extensively in
anotrophic and thioautotrophic metabolism conjunction with 13C values to show the flow
(Fisher, 1995). These signatures are potentially of chemosynthetically derived organic matter
confounded in instances when the thioau- through vent food webs, including those on the
totrophic symbionts use the CO2 respired by the MAR (Vereshchaka et al., 2000; Colaco et al.,
methanotrophs, resulting in a second discrimina- 2002), the Central Indian Ridge (Van Dover,
tion against an already light pool of CO2 and an 2002a), and the Galapagos Rift (Fisher et al.,
anomalously light tissue 13C value (Fisher, 1994). As with 13C data, 15N values vary con-
1993a). siderably among sites; 15N may depend in part
Thus, while 13C values often provide the first on the 15N of the dissolved inorganic nitrogen
evidence that chemoautotrophic or methan- (DIN) pool, the proportions and 15N values of
otrophic symbioses occur in certain animal spe- different components (NH4+, NO32, NO2, and
cies, researchers must recognize that 13C is urea) in the DIN pool (Waser et al., 1998; Colaco
inherently responsive to physical, environmental et al., 2002), the uptake kinetics of different DIN
and enzymatic factors. Stable carbon isotope sig- assimilation pathways (Waser et al., 1998;
natures therefore should not be used apart from Krueger, 1996a), and, as shown for vent shrimp
other corroborating evidence (e.g., physiological (Vereshchaka et al., 2000) and mussels (Trask
and enzyme activity assays and genetic charac- and Van Dover, 1999), the ontogenetic stage of
terization) to identify carbon fixation pathways the host. Therefore, as noted above with stable
or methane oxidation in chemosynthetic symbi- carbon isotopes, in the absence of additional
oses (Fisher, 1995; Scott, 2003; Scott et al., 2004). enzymatic, genetic and environmental data, cau-
tion must be used when comparing 15N values
from different habitats and species.
Sulfur and Nitrogen Isotopes In addition to
carbon isotopes, stable isotopes of sulfur and
nitrogen are also used to study sources and
metabolism of these elements in symbioses. The Ecophysiology
extent to which different sources of reduced
sulfurgeothermal production in vent fluid Symbioses between chemosynthetic bacteria and
or microbial sulfate reduction in bottom sedi- marine invertebrates must acquire all of the sub-
mentsupport thioautotrophic metabolism has strates necessary for chemosynthetic meta-
been inferred from the 34S value of biological bolism: reduced sulfur or methane, oxygen,
samples. Such analyses revealed hydrothermally dissolved inorganic carbon (DIC, as CO2 or
derived sulfide as the dominant sulfide source for CH4), and other nutrients (e.g., nitrogen and
deep-sea vent symbioses (Fry et al., 1983; phosphorus) for use in biosynthesis. In particu-
Yamanaka et al., 2003). In contrast, symbiotic lar, to support energy generation, these symbio-
bacteria within a shallow water vestimentiferan ses must obtain substrates from both oxic and
tubeworm, Lamellibrachia satsuma (Miura et al., anoxic environments. To meet these demands,
2002), and the protobranch, Solemya velum the host-symbiont association relies on special-
(Conway et al., 1989), rely predominantly on sul- ized biochemistry, physiology and behavior.
fide derived from microbial sulfate reduction. These adaptations are best studied in thioau-
492 C.M. Cavanaugh et al. CHAPTER 2.5

totrophic endosymbioses and are discussed pri- Table 2. Adaptations of thioautotrophs and methanotrophs
marily within this context below. for life at oxic-anoxic interfaces.a
Adaptation Example
Spanning the Oxic-Anoxic Interface Attachment Thiothrix
Motility, chemotaxis Beggiatoa Thioploca
Access to both oxygen and reduced chemicals Elemental sulfur deposition Beggiatoa Thiothrix
is necessary for aerobic respiration by chemo- Nitrate and sulfur storage Thiomargarita Thioploca
synthetic symbionts. Specifically, thioautotrophs Create own interface Thiovulum
shuttle electrons from reduced sulfur (e.g., Filamentous sulfur production Arcobacter sp.
sulfide) to a terminal electron acceptor during Resting cysts Methanotrophs
oxidative phosphorylation, generating a proton Associate with eukaryote Thioautotroph and
Methanotroph symbionts
gradient that drives ATP synthesis. Though some
thioautotrophic symbionts (such as those in the a
From Anthony (1982), Jrgensen and Postgate (1982),
tubeworm Riftia pachyptila [Hentschel and Cavanaugh (1985), Schulz et al. (1999), and Wirsen et al.
Felbeck, 1993a] and the clam Lucinoma aequizo- (2002).
nata [Hentschel et al., 1993b]) may use nitrate as
an electron acceptor during periods of anoxia,
most thioautotrophic symbionts typically use formation by methanotrophs; Table 2 and refer-
molecular oxygen for respiration. Similarly, ences therein).
methanotrophs must obtain oxygen for respira- Symbiosis thus may be viewed as an adapta-
tion as well as methane for both energy gen- tion to simultaneously obtain sulfide (or meth-
eration (via methane oxidation) and carbon ane) and oxygen from anoxic-oxic interfaces,
assimilation (Anthony, 1982). allowing thioautotroph or methanotroph sym-
This dual requirement for oxygen and bionts, via association with a eukaryotic host, to
reduced compounds poses unique problems for circumvent many of the problems of sulfide
thioautotrophs and methanotrophs. First, these acquisition (Cavanaugh, 1985). Similarly to free-
organisms must obtain energy substrates living sulfur bacteria, thioautotrophic symbioses
from mutually exclusive environmentsoxygen use specialized behavioral, anatomical or physi-
is absent or at very low levels in the anoxic zones ological mechanisms, either to spatially or tem-
from which sulfide or methane is typically porally bridge sulfidic and oxic zones or to
obtained. Second, sulfide, the predominant simultaneously sequester sulfide and oxygen
energy source for thioautotrophy, spontaneously (Cavanaugh, 1994; Fisher, 1996; Polz et al., 2000).
reacts with oxygen to form less-reduced sulfur For instance, the cold seep vestimentiferan tube-
compounds (S0, S2O32, or SO42; Zhang and worm Lamellibrachia cf. luymesi acquires oxy-
Millero, 1993), thereby decreasing the availabil- gen via its anterior plume while extending a
ity of substrates for thioautotrophy. Though such posterior section of its tube (the root) deep into
abiotic oxidation may be several orders of mag- the sediment to acquire sulfide (Julian et al.,
nitude slower than biological sulfide oxidation 1999; Freytag et al., 2001). Similar burrowing tac-
(Millero et al., 1987; Johnson et al., 1988), thio- tics occur in some species of symbiont-containing
autotrophic symbioses must still compete with thyasirid clams, which possess a superextensile
oxygen for free sulfide. Also, in habitats contain- foot (up to 30 times the length of the shell) that
ing both sulfide and methane, abiotic oxidation burrows into the sediment to access hydrogen
of sulfide may limit the oxygen available for sulfide (Dufour and Felbeck, 2003), and in pro-
methanotrophy. These limitations force free- tobranchs of the genus Solemya, which dig Y-
living thioautotrophs and methanotrophs into shaped burrows in reducing sediments to allow
microaerophilic zones at the interface, or simultaneous pumping of oxygenated water from
chemocline, between oxic (e.g., water column) above and sulfide-rich pore water from below
and anoxic (e.g., vent fluid and sediment pore (Stanley, 1970; Cavanaugh, 1983; Fig. 11). Also,
water) habitats. Such free-living bacteria demon- shrimp, nematodes and oligochaetes migrate
strate unique mechanisms to support life at the vertically along the oxygen-sulfide gradient or
oxic-anoxic interface; these adaptations may be between separate oxic and anoxic zones, thereby
behavioral (e.g., tracking the chemocline via enabling their symbionts to simultaneously
gliding by Beggiatoa), anatomical (e.g., keeping access both energy substrates or to store reduced
cells in the chemocline via veil formation by sulfur compounds for later oxidation (Polz et al.,
Thiovulum or creation of a filamentous sulfur 2000).
matrix by Arcobacter), biochemical (e.g., inter- The vent tubeworm Riftia pachyptila possesses
nal or external sulfur deposition that serves as a remarkable biochemical adaptation to simulta-
an electron source or sink when sulfide or neously acquire sulfide and oxygen. R. pachyptila
oxygen is limiting, as by Beggiatoa and Arco- produces coelomic and vascular hemoglobins
bacter), or developmental (e.g., resting stage that, in contrast to most invertebrate and verte-
CHAPTER 2.5 Marine Chemosynthetic Symbioses 493

brate hemoglobins, can bind oxygen in the pres- as the energy source in symbiotic carbon fixation.
ence of sulfide (Arp et al., 1985; Arp et al., 1987; Subsequently, researchers have demonstrated
Childress et al., 1991; Zal et al., 1996). R. pachyp- mitochondrial sulfide oxidation across a wide
tila appears to preferentially take up HS from range of organisms, including polychaete worms,
the surrounding fluid, despite a large H2S gradi- clams, fishes and chickens (Grieshaber and
ent from tubeworm blood to the environment Volkel, 1998; Yong and Searcy, 2001). These data
(Goffredi et al., 1997a). The HS diffuses across lend credence to the hypothesis that mitochon-
the plume of the worm (Goffredi et al., 1997a) dria evolved from sulfide-oxidizing endosymbi-
and then binds reversibly and independently of otic bacteria (Searcy, 1992).
O2 at two free cysteine residues, each located on Readers should consult several additional
a distinct globin type (Zal et al., 1997; Zal et al., reviews (e.g., Cavanaugh, 1994; Fisher, 1996; Polz
1998; Bailly et al., 2002). These residues are well et al., 2000) for a more extensive discussion of
conserved in both symbiont-containing and the remarkable adaptations used by chemoau-
symbiont-free annelids from sulfidic environ- totrophic symbioses to sequester both oxygen
ments but are absent in annelids from sulfide- and reduced chemicals across oxic-anoxic zones.
free habitats (Bailly et al., 2002; Bailly et al.,
2003). Bailly et al. (2003) suggest that the sulfide
Carbon Uptake and Transport
binding function may have been lost via positive
selection, if the sulfide-binding cysteine residues In addition to oxygen and reduced sulfur com-
react disadvantageously with other blood com- pounds, thioautotrophic symbionts utilizing the
ponents in the absence of sulfide. Calvin cycle require CO2 for autotrophic carbon
Extracellular hemoglobins that simulta- fixation. Acquisition of CO2 is not trivial given
neously bind sulfide and oxygen are absent in that relative concentrations of the three distinct
most other marine invertebrates that host sul- chemical species (CO2, HCO3 and CO32) in the
fide-oxidizing symbionts (Weber and Vinogra- dissolved inorganic carbon (DIC) pool can vary
dov, 2001); such organisms have evolved other considerably depending on pH (pKa of 6.4 for
mechanisms for regulating sulfide toxicity and CO2 : HCO3 at 25C; see the section Habitat
delivery. For instance, the vesicomyid clam Chemistry in this Chapter). In general, the
Calyptogena magnifica synthesizes a di-globular, majority of DIC in seawater (pH ~ 8.0) is HCO3.
non-heme molecule that readily binds free sul- But at vents the typically lower pH of the mixed
fide within the blood serum, perhaps via zinc vent fluid and ambient bottom water generates
residues (Arp et al., 1984; Zal et al., 2000). Also, higher concentrations of CO2, giving organisms
several thioautotroph-containing species, includ- that use the Calvin cycle a distinct advantage.
ing the vent mussel Bathymodiolus thermophilus The tubeworm Riftia pachyptila provides an
and the coastal clam Solemya velum, appear to interesting model in which to study the uptake
mediate detoxification in part by storage of sul- and transport of DIC. Goffredi et al. (1997b)
fur in amino acids (e.g., taurine and thiotaurine; demonstrated that for R. pachyptila, pH plays an
Conway and Capuzzo, 1992a; Pruski et al., 2000a; important role in DIC uptake. The acidity of dif-
Joyner et al., 2003; Pruski and Fiala-Mdioni, fuse vent fluid (pH ca. 6) around tubeworms
2003). Indeed, thiotaurine may be used effec- ensures that CO2 (pKa of 6.1 at in situ tempera-
tively as a biomarker of thioautotrophic symbio- ture and pressure of ca. 10C and 101.3 kPa;
ses (Pruski et al., 2000b). Dickson and Millero, 1987) is the dominant DIC
Other host organisms, including some bivalve form in the vent environment. This contrasts with
mollusks, apparently avoid sulfide toxicity via the vascular fluid of the worm, which has an alka-
mitochondrial oxidation of sulfide. Powell and line pH of 7.17.5, apparently because of the
Somero (1986) first demonstrated mitochondrial action of H+-ATPases (Goffredi et al., 1999;
sulfide oxidation in the coastal protobranch S. Goffredi and Childress, 2001; Girguis et al.,
reidi. The authors showed that mitochondria iso- 2002). The alkaline pH inside Riftia results in
lated from the gill and foot of S. reidi exhibit rapid conversion of CO2 to HCO3, which,
ADP-stimulated oxygen uptake and ATP syn- because of its negative charge, cannot diffuse out
thesis following the addition of sulfide. On the of the worm; this in effect creates a bicarbonate
basis of the effects of cytochrome and reduced trap (Childress et al., 1993). Thus, a gradient
nicotinamide adenine dinucleotide (NADH) oxi- of higher external [CO2] to lower internal [CO2]
dase inhibitors, electrons from sulfide oxidation develops across the tubeworm plume and drives
appear to enter the respiratory chain at cyto- diffusion of DIC into the blood (Childress et al.,
chrome c in S. reidi mitochondria (Powell and 1993; Goffredi et al., 1997b; Scott, 2003). Follow-
Somero, 1986). Further characterization of this ing diffusion into the plume, DIC (as CO2 and
system using 35S showed that sulfide is oxidized HCO3) is transported by the vascular system to
exclusively to thiosulfate (OBrien and Vetter, the symbiont-containing trophosome. Here, car-
1990), a nontoxic intermediate that can function bonic anhydrase, the enzyme that reversibly con-
494 C.M. Cavanaugh et al. CHAPTER 2.5

verts CO2 into HCO3 in both prokaryotes and eral other organic acids and sugars were excreted
eukaryotes, may play a role in converting HCO3 by purified suspensions of R. pachyptila sym-
into CO2, the DIC species used by RubisCO bionts, suggesting that these simple organic com-
(Kochevar and Childress, 1996; De Cian et al., pounds might be important intermediates in the
2003a; De Cian et al., 2003b). As discussed above transfer of fixed carbon from symbionts to host.
for Riftia, DIC incorporation into symbiont bio- Corroborating these data, Bright et al. (2000),
mass occurs via CO2 fixation by a form II using pulse labeling analysis, showed that the
RubisCO of the Calvin-Benson cycle. Rapid CO2 bulk of organic carbon assimilated into R. pac-
fixation rates create steep internal [CO2] gradi- hyptila tissue is first released by metabolically
ents between symbiont and host cytoplasm that active bacteria at the center of a trophosome
may, in combination with the relatively low dis- lobule. However, these authors also showed that
crimination of form II RubisCO against 13C, a smaller fraction of host carbon is obtained by
result in a 13C-enriched signature of symbiont digestion of bacterial cells at the lobule periph-
and host biomass (Robinson et al., 2003; Scott, ery (Bright et al., 2000). This evidence for diges-
2003). tion is supported by prior studies showing
In chemosynthetic endosymbioses the host degenerative stages of bacteria within the R.
benefits by obtaining part or all of its nutrition pachyptila trophosome (Bosch and Grass, 1984;
from the symbiont, via two potential transfer Hand, 1987). In addition, relatively high
mechanisms: the host may assimilate autotroph- lysozyme activity in Riftia tissue further suggests
ically fixed carbon that has been released by the that digestion of symbionts plays a role in tube-
symbiont and translocated to host cells in the worm nutrition (Boetius and Felbeck, 1995).
form of soluble organic molecules, or the host
may directly digest bacterial cells. Radiotracer
Nitrogen
analysis and microscopy have proven particu-
larly useful in studying host nutrition. For exam- The partners in a symbiosis must also acquire all
ple, Fisher and Childress (1986) showed a rapid of the other macro- and micronutrients, particu-
(within hours) appearance of radiolabeled car- larly nitrogen and phosphorus, for use in the
bon in the symbiont-free tissues of the host clam biosynthesis of organic compounds. Currently,
Solemya reidi following exposure to 14C-labeled very little is known about how various forms
bicarbonate, suggesting release of fixed carbon (inorganic and organic) of phosphorus are trans-
by the symbiont population. In contrast, a ferred to and among different pools within
slow (15 days) transfer of labeled organic chemoautotrophic endosymbioses. Most studies
carbon from methanotroph-containing tissue to have focused on nitrogen metabolism, using a
symbiont-free tissue of a seep mussel exposed combination of enzyme characterizations and
to 14C-labeled methane was inferred to be due to physiological experiments to elucidate nitrogen
initial 14CH4 incorporation by the symbionts with assimilation pathways. Nitrate (NO32), which is
host digestion of symbionts occurring later abundant at vents (in situ concentrations of
(Fisher and Childress, 1992). Electron micro- ~40 M; Johnson et al., 1988), appears to be the
scopy showing symbionts being degraded in the predominant nitrogen source for vent symbioses.
basal region of bacteriocytes in other methane- For example, Lee et al. (1999) demonstrated the
based and dual chemoautotroph-methanotroph activity of nitrate reductase, a bacterial enzyme
mussel symbioses supports this interpretation involved in converting nitrate to ammonia for
(Cavanaugh et al., 1992; Barry et al., 2002), as either assimilatory or respiratory purposes, in
does the detection of lysosomal enzymes in the the vent tubeworms Riftia pachyptila and Tevnia
gills of the vent bivalves Calyptogena magnifica jerichonana and the mussel Bathymodiolus
and Bathymodiolus thermophilus (Fiala-Mdioni thermophilus. In addition, the ammonia assimi-
et al., 1994; Boetius and Felbeck, 1995) and the lation enzymes glutamine synthetase (GS) and
shallow water clam Lucinoma aequizonata glutamate dehydrogenase (GDH) were detected
(Boetius and Felbeck, 1995). in these symbioses, and almost all GS activity in
In the R. pachyptila tubeworm symbiosis, the symbiont-containing tissue was shown to be due
transfer of carbon from symbiont to host appears to enzyme produced by the bacterial symbiont
to occur via both translocation and digestion and not the host (Lee et al., 1999). Supporting
(Bright et al., 2000). Felbeck (1985) and Felbeck these data, physiological experiments on R. pac-
and Turner (1995) documented a rapid (within hyptila kept in pressurized chambers showed
seconds) appearance of labeled succinate and that the symbiont population reduces nitrate to
malate in trophosome tissue and in vascular and ammonia not for respiratory purposes but for
coelomic blood following exposure of whole incorporation into both symbiont and host
worms (in pressure vessels) and plumes to 14C- biomass (Girguis et al., 2000). However, in R.
bicarbonate. Subsequently, Felbeck and Jarchow pachyptila, high GS activity also occured in
(1998) showed that succinate, malate, and sev- symbiont-free branchial plume tissue, suggesting
CHAPTER 2.5 Marine Chemosynthetic Symbioses 495

that the host may also be involved in assimilation later-Mesozoic and Cenozoic) and suggests an
of ammonia from the vent environment (Minic alternative hypothesis to vent taxa as living rel-
et al., 2001). But further enzymatic characteriza- ics: vents were recently populated from shallow
tion of Riftia tissues demonstrated that the tube- seeps or whale falls (Van Dover et al., 2002b;
worm depends on its symbionts for the de novo Hurtado, 2002). Indeed, the communities most
synthesis of pyrimidine nucleotides (Minic et al., similar to those of vents occur at seeps. Com-
2001) as well as for the biosynthesis of poly- pared to the spatially and temporally patchy
amines (Minic and Herve, 2003), suggesting that distribution of vent fossils (with most being
the trophosome is a primary site for nitrogen concentrated in the Silurian and Devonian rocks
assimilation and metabolism. of the Ural mountains), seep fossils are ubiqui-
In contrast, in the thioautotrophic symbiosis tous (Little and Vrijenhoek, 2003). At least 50,
involving the shallow-water clam Solemya reidi, and perhaps as many as 200, fossilized seep sites
inorganic nitrogen is readily assimilated in the dating from the Devonian to the Pleistocene
form of ammonia (Lee and Childress, 1994), have been uncovered. These specimens are much
which is abundant in the shallow water, nutrient- better preserved than most vent fossils and
rich habitats of the clam (e.g., sewage outfalls). include extant vent taxa not yet uncovered at
Ammonia incorporation rates are highest in the fossil vent sites (e.g., vesicomyids, thyasirids,
symbiont-containing gill tissue, and the sulfur- mytilids and solemyids; Little and Vrijenhoek,
containing amino acid taurine appears to be a 2003). This greater diversity supports the seeps-
major end product of ammonia assimilation (Lee to-vents hypothesis. However, opponents argue
et al., 1997). The mechanisms by which chemo- that the vent fossil record has been greatly
synthetic symbionts, particularly those contained affected by high calcium carbonate dissolution
within the cells of invertebrate hosts (such as rates (Little and Vrijenhoek, 2003).
Solemya and Riftia), acquire all of the other While the discrepancy between the evolution-
macro- and micronutrients for biosynthesis have ary histories suggested by the fossil and molecu-
yet to be characterized. lar evidence needs to be resolved, it must also be
stressed that these data are not evidence for
chemosynthetic symbioses. In a unique study,
CoBabe (1991) was able to deduce a chemosyn-
Ecology and Evolution thetic symbiosis by analyzing the organic matrix
from fossil lucinid shells using stable carbon iso-
History
topes. This result is encouraging and suggests
Prior to the use of molecular techniques, that both the age of these organisms and their
researchers considered vent taxa to be relic spe- symbiosis can be addressed using current
cies. These organisms, whose strange morpholo- methods.
gies suggest a primitive state, purportedly
survived past extinction events due to the rela-
tive isolation of vents from the photic zone
Organism Interactions
(McArthur and Tunnicliffe, 1998). This percep- In the relatively featureless and nutrient poor
tion of vents as ancient ecosystems is supported deep sea, vent and seep environments are eco-
by the fossil record, which shows that over 80% logical oases (Laubier, 1989). Initially, free-living
of vent species are found only at vent sites chemoautotrophic bacteria were hypothesized to
(Tunnicliffe, 1991; Tunnicliffe, 1992; Little et al., provide the bulk of primary production in these
1997; Little and Vrijenhoek, 2003) and that the communities (Lonsdale, 1977). Indeed, at some
oldest vent site dates to the Silurian (~430 Mya; vent sites, suspended bacteria or bacteria in
Little et al., 2004). But the fossil record for vents surface-attached mats are a large food source for
is relatively poor. There are only 19 known fos- higher trophic levels (Humes and Lutz, 1994;
silized vent sites on the planet, perhaps because Van Dover, 2000). But the dominant strategy for
calcium carbonate structures dissolve relatively the major vent and seep fauna is symbiosis with
quickly in vent fluids (Hunt, 1992; Kennish and chemoautotrophic bacteria (Cavanaugh, 1994),
Lutz, 1999). Also, studying vent fauna evolution and these symbioses significantly influence the
based on the morphological characters of fossils ecology of the nonsymbiotic community. Not
is problematic if much of the specimen has only are chemosynthetic symbioses a major and
degraded or if the preserved character is plastic stable source of organic carbon (Sarrazin and
or isomorphic. In particular, vestimentiferan Juniper, 1999), but as biogenic structures, they
tubeworms are known for the phenotypic plas- also provide living space for a diversity of species
ticity of their tubes (Southward et al., 1995; in an otherwise two-dimensional landscape of
Black et al., 1998). basalt or sediment (Bergquist et al., 2003). For
In contrast, molecular evidence suggests that example, the tubes of chemosynthetic vestimen-
vent taxa evolved more recently (22150 Mya; tiferans support mussels, sponges and limpets,
496 C.M. Cavanaugh et al. CHAPTER 2.5

many of which host their own chemosynthetic larger and more genetically heterogeneous than
symbionts (Yamamoto et al., 2002; Bergquist et populations transmitted vertically. Comparisons
al., 2003; Bates et al., 2004). of 16S rRNA gene evolution between free-living
Vent symbioses may also significantly impact bacteria, in which significant recombination
the free-living bacterial community by providing occurs (Dykhuizen and Green, 1991; Levin and
increased surface area for attachment. Free- Bergstrom, 2000), and symbiotic chemosynthetic
living bacteria that cluster phylogenetically with bacteria revealed unexpected differences in rates
known chemoautotrophic and heterotrophic of evolution depending on mode of transmission
groups have been isolated from tubeworm sur- (Peek et al., 1998). While chemoautotrophic,
faces (Lopez-Garcia et al., 2002; Yamamoto et maternally transmitted endosymbionts did
al., 2002). On the Mid-Atlantic Ridge, a single exhibit rapid evolutionary rates, consistent with
phylotype of shrimp episymbionts, which appear their small population sizes, environmentally
to be transmitted among hosts via the environ- transmitted symbionts evolved more slowly than
ment, represented over 60% of the free-living their free-living counterparts (Peek et al., 1998).
bacteria (Polz and Cavanaugh, 1995). This sug- The authors suggest that this slower rate of evo-
gests that the host inoculates inanimate surfaces lution could be caused by purifying selection in
continuously, increasing the probability of a large population. These results, however, were
symbiont attachment relative to the free-living based on one gene across many lineages; a true
community (Polz and Cavanaugh, 1995). Such genomic analysis of evolution in chemosynthetic
environmental inoculation may also occur in endosymbionts is necessary to extend these
tubeworm and lucinid clam symbioses, in which findings.
the symbionts also appear to be transmitted Because chemosynthetic symbionts have yet
environmentally (Durand and Gros, 1996a; to be cultured and their hosts are difficult to
Durand et al., 1996b; Di Meo et al., 2000; Nelson maintain in the laboratory, the transmission
and Fisher, 2000; McMullin et al., 2003). strategy of a symbiosis has been inferred by phy-
logenetic analysis or PCR-based detection of
bacteria in host reproductive tissues or gametes.
Transmission Strategies and If the symbionts are maternally transmitted and
the symbioses stable, congruence of host and
Effects on Symbiosis
symbiont phylogenies should occur (e.g., Chen et
The transmission strategy of a symbiosis reveals al., 1999; Thao et al., 2000; Degnan et al., 2004)
much about the evolutionary dynamics between and bacterial symbionts should be found in ova-
host and symbiont. Symbiont transmission can ries or oviducts of the host. Using these tech-
occur environmentally (through a free-living niques, vertical transmission has been proposed
population of symbiotic bacteria), horizontally for the solemyid protobranchs (Cary, 1994;
(between contemporary organisms sharing the Krueger et al., 1996b) and vesicomyid clams
same habitat), or vertically (from parent to (Endow and Ohta, 1990; Cary and Giovannoni,
offspring). Vertically transmitted endosymbionts 1993a; Peek et al., 1998; Hurtado et al., 2003).
are effectively disconnected from their free- Interestingly, although bacteria have been
living counterparts. These symbionts experience detected via PCR in the gonads of female hosts,
elevated rates of mutation and fixation of slightly this does not necessarily imply direct bacterial
deleterious alleles because of genetic drift endocytic localization in host eggs. Indeed, in
(Wernegreen, 2002). For the most part, these Solemya reidi, the internal contents of oocytes do
evolutionary effects are due to a vastly different not contain bacteria, and instead the transmis-
selective regime inside the host and a severely sion mechanism is thought to occur via ingestion;
depreciated population size (Ohta, 1973); endo- the larvae ingest the bacteria, which are then
symbionts undergo a population bottleneck engulfed by hemocytes in the larval perivisceral
upon host colonization and another upon cavity and transported to the developing gill
transmission (Mira and Moran, 2002). But the (Gustafson and Reid, 1988). The oligochaetes
asexuality and lack of recombination in endo- also exhibit an interesting mechanism of vertical
symbionts exacerbate these genetic problems transmission. During oviposition, the eggs
through what is known as Mullers rachet appear to be infected with the symbiotic bacteria
(Muller, 1964; Moran, 1996). In Mullers ratchet, via the adults genital pad (Giere and Langheld,
wildtype recombinants cannot be introduced 1987). During the development of the larvae,
into the endosymbiont population (Moran and many of the bacteria exist intracellularly, but as
Baumann, 1994; Dale et al., 2003); genetic drift the animal matures, the symbionts take their pri-
therefore occurs quickly, and the population can- marily extracellular form.
not recover after fixation of deleterious alleles. However, of the putatively vertically transmit-
In contrast, symbiont populations that are ted symbioses, only associations involving vesi-
environmentally transmitted are effectively comyid clams show phylogenetic congruence
CHAPTER 2.5 Marine Chemosynthetic Symbioses 497

between host and symbiont (Peek et al., 1998; et al., 2003a; DeChaine et al., 2004). In addition,
Hurtado et al., 2003). Cospeciation does not McKiness (2004) provided the first assessment of
appear to have occurred in the solemyid proto- cospeciation between symbiont and host in
branchs (Durand et al., 1996b; Krueger and Bathymodiolus mussels, analyzing molecular
Cavanaugh, 1997) or in the mytilid mussels data for both methanotrophic and chemoau-
(McKiness, 2004). When evaluating phylogenetic totrophic symbionts and testing phylogenetic
congruence, however, other factors that influ- congruence with the hosts. The results showed
ence a phylogenetic reconstruction, such as geo- weak support for vertical transmission of the
graphic constraints, must be taken into account. chemoautotrophic symbionts but provided no
Also, robust phylogenies with adequate taxa evidence for vertical transmission of the
sampling for both host and symbiont are neces- methanotrophs.
sary; incomplete phylogenies may be hindering
analyses of the solemyid and mytilid symbioses.
Biogeography and Population Genetics
Lack of PCR-based evidence and phyloge-
netic incongruence has been used to infer an The view of hydrothermal vents as deep-sea
environmental mode of transmission for several islands frames questions of vent biogeography
of the chemosynthetic symbioses. For instance, and population genetics. Compared to the rela-
the lucinid clams exhibit environmental trans- tively uniform and stable environment of the
mission (Durand and Gros, 1996a; Gros et al., abyssal deep sea, hydrothermal vents are ephem-
1996; Gros et al., 1998; Gros et al., 2003a; Gros eral, dynamic and geographically fragmented. A
et al., 2003b). Researchers have even been able chain of vents along a mid-ocean ridge resembles
to exchange symbionts between lucinid species a chain of islands in an archipelago. However,
without affecting the development of the juve- genetic data for many vent species do not cleanly
nile animal (Gros et al., 2003a). In addition, vent fit an island or stepping-stone model of bio-
tubeworms appear to acquire their symbionts geography (Vrijenhoek et al., 1998). Some host
from the environment (Distel and Cavanaugh, taxa do exhibit a decline in gene flow with
1994; Feldman et al., 1997; Laue and Nelson, increasing distance between sites (Black et al.,
1997; Di Meo et al., 2000; Nelson and Fisher, 1994), as a stepping-stone model would predict
2000; McMullin et al., 2003), as evidenced in part (Kimura and Weiss, 1964), while others show a
by the presence of functional genes for sensing more widespread gene flux (Karl et al., 1996) or
and responding to the environment as well as a appear to encounter barriers to dispersal other
flagellin gene in the Riftia symbiont (Hughes et than distance (Black et al., 1998). These differ-
al., 1997; Hughes et al., 1998; Millikan et al., ences should be resolved with a greater under-
1999). Indeed the 16S rRNA phylotype has been standing of the major variables affecting vent
detected in vent environments via both PCR and biogeography, including larval development and
in situ hybridization, suggesting the vent tube- dispersal, symbiont distribution, oceanic flow,
worm symbionts are environmentally transmit- and past and current bathymetry. This section
ted (Harmer et al., 2004). While environmental focuses predominantly on host biogeography
transmission of tubeworm symbionts seems to be because research on the population genetics and
a potentially risky strategy, given the stochastic biogeography of bacterial symbionts is lacking.
nature of environmental transmission and the Understanding host population dynamics, how-
complete dependence of the adult tubeworms on ever, does provide valuable insight into the dis-
their symbionts for nutrition, detection of wild tribution of the chemosynthetic symbionts to
symbionts in conjunction with the phylogenetic which most vent fauna are tightly linked.
evidence supports environmental transmission in Vent habitats are highly ephemeral and sensi-
this species. tive to variations in tectonic activity, hydrother-
The mechanism of transmission for the mytilid mal inputs, and geologic events. Consequently,
mussels remains largely unresolved; on the basis the persistence of vent organisms, which are pre-
of varying evidence, researchers have suggested dominantly sessile as adults, depends on success-
both vertical and environmental transmission. ful larval dispersal to new sites. The dispersal
Vertical transmission in Bathymodiolus thermo- strategy of larvae can significantly impact the
philus, the thioautotroph-hosting mussel, was biogeography of the adult organism (Lalou and
suggested in 1993, but evidence supporting this Brichet, 1982; Fustec et al., 1987). On the basis
report is not yet published (Cary and of laboratory studies and comparisons with shal-
Giovannoni, 1993a). In contrast, a recent study low water species, researchers infer that some
based on genetic and ultrastructural data of the vent larvae are planktotrophic while others are
chemoautotrophic symbionts of B. azoricus, a lecithotrophic (Lutz et al., 1980; Turner et al.,
MAR mussel hosting both thioautotrophs and 1985; Young et al., 1996; Marsh et al., 2001).
methanotrophs, indicated environmental acqui- Although both forms are pelagic, planktotrophic
sition of the chemoautotrophic symbionts (Won larvae are positively buoyant and feed in the
498 C.M. Cavanaugh et al. CHAPTER 2.5

water column while lecithotrophic larvae are primarily NNW and SSE along the axis (Marsh
negatively buoyant and nonfeeding (Poulin et al., et al., 2001) may facilitate dispersal of larval B.
2001). Larvae of the large vesicomyid clam thermophilus, contributing to the homogeneity
Calyptogena magnifica typify a planktotrophic observed along the EPR. However, there is some
dispersal strategy successfully exploiting the vent genetic structure in the EPR mytilid populations;
plume to carry them many kilometers (Pradillon the westward currents across the ridge axis at
et al., 2001; Mullineaux et al., 2002). Although 15N and the Easter Microplate are obstacles for
planktotrophic larvae risk being carried off the planktotrophic larvae. At 15N, a westward cur-
ridge axis by cross currents, C. magnifica appar- rent flows across the ridge axis, partially isolating
ently encounters no significant barriers to dis- the 17S population from the other northern
persal across the equator on the East Pacific Rise populations. Further south, at the Easter Micro-
(EPR; Karl et al., 1996). Conversely, leci- plate, mussel populations are severely divergent
thotrophic larvae are less affected by cross cur- (Won et al., 2003). Although morphologically
rents but, because they are non-feeding, have indistinguishable, mussels north and south of the
relatively short larval stages and therefore lim- Microplate are genetically distinct (Won et al.,
ited time for dispersal. For example, in labora- 2003). The Easter Microplate therefore appears
tory studies, larval Riftia pachyptila exhibit a to be a significant topographic obstacle for larval
lecithotrophic strategy, surviving a maximum of dispersal. Such a feature can produce cross-axis
38 days (Marsh et al., 2001). Assuming flow rates currents, like those at 15N, that may sweep
characteristic of EPR currents, this interval sug- bathymodiolid larvae (which are positively
gests a maximum dispersal distance of 100 km buoyant) off the ridge axis (Fujio and Imasato,
(Marsh et al., 2001); however, the in situ dis- 1991; Mullineaux et al., 1995). The degree to
persal distance is unknown given that Riftia which such barriers also impact the genetic
larvae have not been detected in the wild. diversity and biogeography of chemosynthetic
The geology and tectonic activity associated symbiont populations remains an open question.
with mid-ocean ridges also impact the biogeog-
raphy of vent organisms. For instance, Iceland,
an active site of crust formation, rises out of the Summary
ocean along the northern MAR, forming a bar-
rier that prevents dispersal along the ridge axis Scientific understanding of chemosynthetic sym-
(Tyler and Young, 2003). Given that Iceland bioses continues to expand. The spectacular dis-
interrupted the MAR approximately 55 Mya, the covery of hydrothermal vents highlighted the
ridge axis north of Iceland constitutes one of the importance of chemosynthetic bacteria both in
most isolated vent systems on the planet, per- food webs and in symbioses with eukaryotes and
haps representing a new biogeographic province provided the impetus to examine less exotic
(Bilyard and Carey, 1980; Dunton, 1992; Svavar- environments for such associations. As other
sson et al., 1993). Similar dispersal barriers are oxic-anoxic environments (e.g., freshwater) and
seen among vent fields abutting the Azores Rise the invertebrates and protists that inhabit them
in the Atlantic (Tyler and Young, 2003) and also are explored, new symbioses will undoubtedly be
evident between the EPR and the Northeast discovered. Further, chemosynthetic bacteria
Pacific vent fields (Tunnicliffe, 1988; Tunnicliffe that use other sources of energy (e.g., hydrogen)
and Fowler, 1996). These barriers are insur- may be found in similar associations.
mountable and may provide the conditions for Current studies of these fascinating symbioses
allopatric speciation of both host and symbiont. involve a range of experimental and diagnostic
Research on EPR bathymodiolid mussels and tools, including physiological assays in special-
their symbionts provides a good example of how ized growth chambers, enzyme characterizations,
larval dispersal strategy, current regime, and immunodetections, and stable isotope analyses.
bathymetry interact to structure biogeography Increasingly, molecular techniques, such as PCR-
(Lonsdale, 1977; Corliss et al., 1979). Except at based gene probing, FISH, and 16 rRNA phylo-
Northern Pacific sites, which are separated from genetic analysis, are used to complement
the EPR by the North America landmass, vent traditional methods. These studies provide valu-
communities on the Pacific ridge axis appear able insight into the population dynamics,
relatively uniform. For example, the mussel evolutionary history, and carbon and nutrient
Bathymodiolus thermophilus, which undergoes a metabolism of symbionts. In addition, projects
planktotrophic larval stage, occurs over a dis- are currently underway to sequence the genomes
tance of 4900 km (from 13N to 32S) on the of some of the chemosynthetic symbionts
EPR. Mussel populations along 13N and 11S described in this review (e.g., symbionts of Riftia
are genetically indistinguishable, indicating no pachyptila and Solemya velum). Genomic ana-
population subdivision (Craddock et al., 1995; lysis, in conjunction with new technologies to
Won et al., 2003). Deep ocean currents that flow manipulate symbioses under in situ conditions
CHAPTER 2.5 Marine Chemosynthetic Symbioses 499

(e.g., via vascular catheters; Felbeck et al., 2004) Bailly, X., R. Leroy, S. Carney, O. Collin, F. Zal, A. Toulmond,
and to sample the physical environment (e.g., and D. Jollivet. 2003. The loss of the hemoglobin H2S-
electrochemical sampling; Luther et al., 2001), binding function in annelids from sulfide-free habitats
reveals molecular adaptation driven by Darwinian pos-
will contribute significantly to our understanding
itive selection. Proc. Natl. Acad. Sci. USA 100:58855890.
of symbiont biology. Scientists are now poised to Barry, J. P., K. R. Buck, R. K. Kochevar, D. C. Nelson,
reveal how interactions with the host and the Y. Fujiwara, S. K. Goffredi, and J. Hashimoto. 2002.
abiotic environment impact symbiotic chemo- Methane-based symbiosis in a mussel, Bathymodiolus
synthetic bacteria over both ecological and platifrons, from cold seeps in Sagami Bay, Japan. Inver-
evolutionary timescales. tebr. Biol. 121:4754.
Bauer-Nebelsick, M., C. F. Bardele, and J. A. Ott. 1996. Elec-
tron microscopic studies on Zoothamnium niveum
Acknowledgments. We thank our colleagues and (Hemprich & Ehrenberg, 1831) Ehrenberg 1838 (Oligo-
collaborators for active discussions on chemo- hymenophora, Peritrichida), a ciliate with ectosymbi-
synthetic symbioses and for their scientific con- otic, chemoautotrophic bacteria. Eur. J. Protistol.
tributions. Without the Chief Scientists, Captains 32:202215.
and crews of the research vessels (including R/V Bennett, B. A., C. R. Smith, B. Glaser, and H. L. Maybaum.
Atlantis II, R/V Atlantis, and R/V Knorr), and 1994. Faunal community structure of a chemoau-
the Expedition Leaders and crews of Deep Sub- totrophic assemblage on whale bones in the deep
mergence Vehicle (DSV) Alvin and the remotely northeast Pacific Ocean. Mar. Ecol. Progr. Ser. 108:205
223.
operated vehicles, we could not explore the vast
Bergquist, D. C., T. Ward, E. E. Cordes, T. McNelis, S.
unknown deep seato them we are grateful. Howlett, R. Kosoff, S. Hourdez, R. Carney, and C. R.
Research in my laboratory (CMC) on chemosyn- Fisher. 2003. Community structure of vestimentiferan-
thetic symbioses has been supported by grants generated habitat islands from Gulf of Mexico cold
from NSF (Biological Oceanography, RIDGE, seeps. J. Exp. Mar. Biol. Ecol. 289:197222.
Cell Biology), the Office of Naval Research, Bilyard, G. R., and A. G. Carey. 1980. Zoogeography of West-
NOAA National Undersea Research Center for ern Beaufort Sea Polychaeta (Annelida). Sarsia 65:19
the West Coast and Polar Regions, and NASA 25.
(Exobiology) and by graduate fellowships Black, M. B., R. A. Lutz, and R. C. Vrifenhoek. 1994. Gene
from the NIH, NSF (IGERT), and Howard flow among vestimentiferan tube worm (Riftia pachyp-
tila) populations from hydrothermal vents of the East-
Hughes Medical Institute, which we gratefully
ern Pacific. Mar. Biol. 120:3339.
acknowledge. Black, M. B., A. Trivedi, P. A. Y. Maas, R. A. Lutz, and R. C.
Vrijenhoek. 1998. Population genetics and biogeogra-
phy of vestimentiferan tube worms. Deep Sea Res. Pt.
Literature Cited II Top. Stud. Oceanogr. 45:365382.
Boetius, A., and H. Felbeck. 1995. Digestive enzymes in
Alt, J. C. 1995. Subseafloor processes in mid-ocean ridge marine-invertebrates from hydrothermal vents and
hydrothermal systems. In: S. E. Humphris, R. A. other reducing environments. Mar. Biol. 122:105113.
Zierenberg, L. S. Mullineaux, and R. E. Thomson Bosch, C., and P. P. Grass. 1984. Cycle partiel des bactries
(Eds.) Seafloor Hydrothermal Systems: Physical, Chem- chimioautotrophes symbiotiques et eurs rapports avec
ical, Biological, and Geological Interactions. American les bacteriocytes chez Riftia pachyptila Jones (Pogono-
Geophysical Union. Geophysical Monograph 91. phore Vestimentifre). II: Levolution des bactries sym-
Anderson, A., J. Childress, and J. Favuzzi. 1987. Net uptake botiques et des bacteriocytes. Comptes Rendus LAcad.
of CO2 driven by sulfide and thiosulfite oxidation in the Sci. Ser. III: Sciences de la VieLife Sciences 299:413
bacterial symbiont-containing clam Solemya reidi. J. 319.
Exp. Biol. 133:1131. Boss, K. J., and R. D. Turner. 1980. The giant white clam from
Anthony, C. 1982. The Biochemistry of Methylotrophs. Aca- the Galapagos Rift, Calyptogena magnifica species
demic Press. London, UK. novum. Malacologia 20:161194.
Arp, A. J., J. J. Childress, and C. R. J. Fisher. 1984. Metabolic Bright, M., H. Keckeis, and C. R. Fisher. 2000. An auto-
and blood gas transport characteristics of the hydrother- radiographic examination of carbon fixation, transfer
mal vent bivalve Calyptogena magnifica. Physiol. Zool. and utilization in the Riftia pachyptila symbiosis. Mar.
57:648662. Biol. 136:621632.
Arp, A. J., J. J. Childress, and C. R. Fisher. 1985. Blood gas Brigmon, R. L., and C. De Ridder. 1998. Symbiotic relation-
transport in Riftia pachyptila. Bull. Biol. Soc. Washing- ship of Thiothrix spp. with an echinoderm. Appl. Envi-
ton 6:289300. ron. Microbiol. 64:34913495.
Arp, A. J., J. J. Childress, and R. D. Vetter. 1987. The sul- Brooks, J. M., M. C. Kennicutt 2nd, C. R. Fisher, S. A. Macko,
phide-binding protein in the blood of the vestimentife- K. Cole, J. J. Childress, R. R. Bidigare, and R. D. Vetter.
ran tube-worm, Riftia Pachyptila, is the extracellular 1987. Deep-sea hydrocarbon seep communities: Evi-
haemoglobin. J. Exp. Biol. 128:139158. dence for energy and nutritional carbon sources. Science
Bailly, X., D. Jollivet, S. Vanin, J. Deutsch, F. Zal, F. Lallier, 238:11381142.
and A. Toulmond. 2002. Evolution of the sulfide-binding Buck, K. R., J. P. Barry, and A. G. B. Simpson. 2000.
function within the globin multigenic family of the deep- Monterey Bay cold seep biota: Euglenozoa with
sea hydrothermal vent tubeworm Riftia pachyptila. chemoautotrophic bacterial epibionts. Eur. J. Protistol.
Molec. Biol. Evol. 19:14211433. 36:117126.
500 C.M. Cavanaugh et al. CHAPTER 2.5

Campbell, B. J., J. L. Stein, and S. C. Cary. 2003. Evidence of otrophic marine molluscan (Bivalvia, Mytilidae) symbi-
chemolithoautotrophy in the bacterial community asso- osis: Mussels fueled by gas. Science 233:13061308.
ciated with Alvinella pompejana, a hydrothermal vent Childress, J. J., C. R. Fisher, J. A. Favuzzi, R. E. Kochevar,
polychaete. Appl. Environ. Microbiol. 69:50705078. N. K. Sanders, and A. M. Alayse. 1991. Sulfide-driven
Cary, S. C., and S. J. Giovannoni. 1993a. Transovarial inher- autotrophic balance in the bacterial symbiont-
itance of endosymbiotic bacteria in clams inhabiting containing hydrothermal vent tubeworm, Riftia pachyp-
deep-sea hydrothermal vents and cold seeps. Proc. Natl. tila Jones. Biol. Bull. 180:135153.
Acad. Sci. USA 90:56955699. Childress, J. J., and C. R. Fisher. 1992. The biology of hydro-
Cary, S. C., W. Warren, E. Anderson, and S. J. Giovannoni. thermal vent animals: Physiology, biochemistry, and
1993b. Identification and localization of bacterial endo- autotrophic symbioses. Oceanogr. Mar. Biol. Ann. Rev.
symbionts in hydrothermal vent taxa with symbiont- 30:337441.
specific polymerase chain reaction amplification and in Childress, J. J., R. W. Lee, N. K. Sanders, H. Felbeck, D. R.
situ hybridization techniques. Molec. Mar. Biol. Oros, A. Toulmond, D. Desbruyres, M. C. Kennicutt,
Biotechnol. 2:5162. and J. Brooks. 1993. Inorganic carbon uptake in hydro-
Cary, S. C. 1994. Vertical transmission of a chemoautotrophic thermal vent tubeworms facilitated by high environmen-
aymbiont in the protobranch bivalve, Solemya reidi. tal pCO2. Nature 362:147169.
Molec. Mar. Biol. Biotechnol. 3:121130. CoBabe, E. A. 1991. Lucinid Bivalve Evolution and the
Cary, S. C., M. T. Cottrell, J. L. Stein, F. Camacho, and D. Detection of Chemosymbiosis in the Fossil Record [PhD
Desbruyres. 1997. Molecular identification and local- thesis]. Harvard University. Cambridge, MA.
ization of filamentous symbiotic bacteria associated with Colaco, A., F. Dehairs, and D. Desbruyres. 2002. Nutritional
the hydrothermal vent annelid Alvinella pompejana. relations of deep-sea hydrothermal fields at the Mid-
Appl. Environ. Microbiol. 63:11241130. Atlantic Ridge: A stable isotope approach. Deep Sea
Cavanaugh, C. M., S. L. Gardiner, M. L. Jones, H. W. Res. Pt. I Oceanogr. Res. Pap. 49:395412.
Jannasch, and J. B. Waterbury. 1981. Prokaryotic cells in Conway, N., J. Capuzzo, and B. Fry. 1989. The role of endo-
the hydrothermal vent tube worm Riftia pachyptila symbiotic bacteria in the nutrition of Solemya velum:
Jones: Possible chemoautotrophic symbionts. Science Evidence from a stable isotope analysis of endosym-
213:340342. bionts and hosts. Limnol. Oceanogr. 34:249255.
Cavanaugh, C. M. 1983. Symbiotic chemoautotrophic bacte- Conway, N., and J. M. Capuzzo. 1991. Incorporation and
ria in marine invertebrates from sulfide-rich habitats. utilization of bacterial lipids in the Solemya velum sym-
Nature 302:5861. biosis. Mar. Biol. 108:277292.
Cavanaugh, C. M. 1985. Symbioses of chemoautotrophic bac- Conway, N. M., and J. E. M. Capuzzo. 1992a. High taurine
teria and marine invertebrates from hydrothermal vents levels in the Solemya velum symbiosis. Comp. Biochem.
and reducing sediments. Biol. Soc. Wash. Bull. 6:373 Physiol. B Biochem. Molec. Biol. 102:175185.
388. Conway, N., M. C. Kennicutt, and C. L. Van Dover. 1992b.
Cavanaugh, C. M., P. R. Levering, J. S. Maki, R. Mitchell, and Stable isotopes in the study of marine chemosynthesis-
M. E. Lidstrom. 1987. Symbiosis of methylotrophic bac- based ecosystems. In: K. Lajtha and R. Michener (Eds.)
teria and deep-sea mussels. Nature 325:346348. Stable Isotopes in Ecology. Blackwell. Oxford, UK.
Cavanaugh, C. M., M. S. Abbott, and M. Veenhuis. 1988. Corliss, J. B., J. Dymond, L. I. Gordon, J. M. Edmond,
Immunochemical localization of ribulose-1,5-bisphos- R. P. V. Herzen, R. D. Ballard, K. Green, D. Williams,
phate carboxylase in the symbiont-containing gills of A. Bainbridge, K. Crane, and T. H. Van Andel. 1979.
Solemya velum (Bivalvia:Mollusca). Proc. Natl. Acad. Submarine thermal springs on the Galapagos Rift. Sci-
Sci. USA 85:77867789. ence 203:10731083.
Cavanaugh, C. M., C. Wirsen, and H. J. Jannasch. 1992. Evi- Craddock, C., W. R. Hoeh, R. A. Lutz, and R. C. Vrijenhoek.
dence for methylotrophic symbionts in a hydrothermal 1995. Extensive gene flow among mytilid (Bathymodio-
vent mussel (Bivalvia:Mytilidae) from the Mid-Atlantic lis thermophilus) populations from hydrothermal vents
Ridge. Appl. Environ. Microbiol. 58:37993803. of the Eastern Pacific. Mar. Biol. 124:137146.
Cavanaugh, C. M. 1993. Methanotroph-invertebrate symbio- Dale, C., B. Wang, N. Moran, and H. Ochman. 2003. Loss of
ses in the marine environment: Ultrastructural, bio- DNA recombinational repair enzymes in the initial
chemical, and molecular studies. In: J. C. Murrell and stages of genome degeneration. Molec. Biol. Evol.
D. P. Kelly (Eds.) Microbial Growth on C1 Compounds. 20:11881194.
Intercept. Dando, P. R., and A. J. Southward. 1986. Chemoautotrophy
Cavanaugh, C. M. 1994. Microbial symbiosis: Patterns of in the bivalve molluscs of the genus Thyasira. J. Mar.
diversity in the marine environment. Am. Zool. 34:79 Biol. Assoc. UK 66:915929.
89. Dando, P. R., and B. Spiro. 1993. Varying nutritional depen-
Cavanaugh, C. M., and J. J. Robinson. 1996. CO2 fixation in dence of the thyasirid bivalves Thyasira sarsi and T.
chemoautotroph-invertebrate symbioses: Expression of equalis on chemoautotrophic symbiotic bacteria, dem-
Form I and Form II RuBisCO. In: M. E. Lidstrom and onstrated by isotope ratios of tissue carbon and shell
F. R. Tabita (Eds.) Microbial Growth on C1 Compounds. carbonate. Mar. Ecol. Progr. Ser. 92:151158.
Kluwer. Dordrecht, The Netherlands. De Bary, A. 1879. Die Erscheinung der Symbiose. Verlag von
Chen, X. A., S. Li, and S. Aksoy. 1999. Concordant evolution Karl J. Trubner. Strasburg, Germany.
of a symbiont with its host insect species: Molecular de Burgh, M. E., and C. L. Singla. 1984. Bacterial coloniza-
phylogeny of genus Glossina and its bacteriome- tion and endocytosis on the gill of a new limpet species
associated endosymbiont, Wigglesworthia glossinidia. J. from a hydrothermal vent. Mar. Biol. 84:16.
Molec. Evol. 48:4958. de Burgh, M. E., S. K. Juniper, and C. L. Singla. 1989. Bac-
Childress, J. J., C. R. Fisher, J. M. Brooks, M. C. Kennicutt terial symbiosis in northeast Pacific vestimentiferaa
2nd, R. Bidigare, and A. E. Anderson. 1986. A methan- TEM study. Mar. Biol. 101:97105.
CHAPTER 2.5 Marine Chemosynthetic Symbioses 501

De Cian, M. C., A. C. Andersen, X. Bailly, and F. H. Lallier. Dufour, S. C., and H. Felbeck. 2003. Sulphide mining by the
2003a. Expression and localization of carbonic anhy- superextensile foot of symbiotic thyasirid bivalves.
drase and ATPases in the symbiotic tubeworm Riftia Nature 426:6567.
pachyptila. J. Exp. Biol. 206:399409. Dunton, K. 1992. Arctic biogeography: The paradox of the
De Cian, M. C., X. Bailly, J. Morales, J. M. Strub, A. Van marine benthic fauna and flora. Trends Ecol. Evol.
Dorsselaer, and F. H. Lallier. 2003b. Characterization of 7:183189.
carbonic anhydrases from Riftia pachyptila, a symbiotic Duperron, S., T. Nadalig, J. Caprais, M. Sibuet, A. Fiala-
invertebrate from deep-sea hydrothermal vents. Prot. Medioni, R. Amann, and N. Dubilier. 2004. Dual sym-
Struct. Funct. Genet. 51:327339. biosis in a Bathymodiolus mussel from a methane seep
Degnan, P. H., A. B. Lazarus, C. D. Brock, and J. J. on the Gabon continental margin (South East Atlan-
Wernegreen. 2004. Host-symbiont stability and fast tic): 16S rRNA phylogeny and distribution of the sym-
evolutionary rates in an ant-bacterium association: bionts in the gills. Appl. Environ. Microbiol. 71(4):
Cospeciation of Camponotus species and their endo- 16941700.
symbionts, Candidatus blochmannia. System. Biol. 53: Durand, P., and O. Gros. 1996a. Bacterial host specificity of
95110. Lucinacea endosymbionts: Interspecific variation in
DeNiro, M. J., and S. Epstein. 1979. Relationship 16S rRNA sequences. FEMS Microbiol. Lett. 140:193
between the oxygen isotope ratios of terrestrial 198.
plant cellulose, carbon dioxide, and water. Science Durand, P., O. Gros, L. Frenkiel, and D. Prieur. 1996b. Phy-
204:5153. logenetic characterization of sulfur-oxidizing bacterial
Desbruyres, D., F. Gaill, L. Laubier, D. Prieur, and G. H. endosymbionts in three tropical Lucinidae by 16S rDNA
Rau. 1983. Unusual nutrition of the Pompeii worm sequence analysis. Molec. Mar. Biol. Biotechnol. 5:37
Alvinella pompejana (polychaetous annelid) from a 42.
hydrothermal vent environment: SEM, TEM, 13C and Dykhuizen, D. E., and L. Green. 1991. Recombination in
15
N evidence. Mar. Biol. 75:201205. Escherichia coli and the definition of biological species.
Desbruyres, D., F. Gaill, L. Laubier, and Y. Fouquet. 1985. J. Bacteriol. 173:72577268.
Polychaetous annelids from hydrothermal vent ecosys- Elderfield, H., and A. Schultz. 1996. Mid-ocean ridge hydro-
tems: An ecological overview. Biol. Soc. Wash. Bull. thermal fluxes and the chemical composition of the
6:103116. ocean. Ann. Rev. Earth Planet. Sci. 24:191224.
Dickson, A. G., and F. J. Millero. 1987. A comparison of the Elsaied, H., H. Kimura, and T. Naganuma. 2002. Molecular
equilibrium constants for the dissociation of carbonic characterization and endosymbiotic localization of the
acid in seawater media. Deep Sea Res. Pt. A Oceanogr. gene encoding D-ribulose 1,5-bisphosphate carboxylase-
Res. Pap. 34:17331743. oxygenase (RuBisCO) form II in the deep-sea vestimen-
Di Meo, C. A., A. E. Wilbur, W. E. Holben, R. A. Feldman, tiferan trophosome. Microbiology 148:19471957.
R. C. Vrijenhoek, and S. C. Cary. 2000. Genetic variation Endow, K., and S. Ohta. 1989. The symbiotic relationship
among endosymbionts of widely distributed vestimen- between bacteria and the mesogastropod snail, Alvini-
tiferan tubeworms. Appl. Environ. Microbiol. 66:651 concha hessleri, collected from hydrothermal vents of
658. the Mariana back-arc basin. Bull. Jpn. Soc. Microb. Ecol.
Distel, D. L., and C. M. Cavanaugh. 1994. Independent 3:7382.
phylogenetic origins of methanotrophic and chemoau- Endow, K., and S. Ohta. 1990. Occurrence of bacteria in the
totrophic bacterial endosymbioses in marine bivalves. J. primary oocytes of vesicomyid clam Calyptogena soyoae.
Bacteriol. 176:19321938. Mar. Ecol. Progr. Ser. 64:309311.
Distel, D. L., H. K. W. Lee, and C. M. Cavanaugh. 1995. Felbeck, H. 1981a. Chemoautotrophic potential of the
Intracellular coexistence of methano- and thioau- hydrothermal vent tube worm, Riftia pachyptila Jones
totrophic bacteria in a hydrothermal vent mussel. Proc. (Vestimentifera). Science 213:336338.
Natl. Acad. Sci. USA 92:95989602. Felbeck, H., J. J. Childress, and G. N. Somero. 1981b. Calvin-
Distel, D. 1998. Evolution of chemoautotrophic endosymbi- Benson cycle and sulfide oxidation enzymes in animals
oses in bivalves. BioScience 48:277286. from sulfide-rich habitats. Nature 293:291293.
Dubilier, N., O. Giere, and M. K. Grieshaber. 1995. Morpho- Felbeck, H. 1983a. Sulfide oxidation and carbon fixation by
logical and ecophysiological adaptations of the marine the gutless clam Solemya reidi: An animal-bacteria sym-
oligochaete Tubificoides benedii to sulfidic sediments. biosis. J. Comp. Physiol. 152:311.
Am. Zool. 35:163173. Felbeck, H., G. Liebezeit, R. Dawson, and O. Giere. 1983b.
Dubilier, N., R. Windoffer, and O. Giere. 1998. Ultrastructure CO2 fixation in tissues of marine oligochaetes (Phallo-
and stable carbon isotope composition of the hydrother- drilus leukodermatus and P. planus) containing symbi-
mal vent mussels Bathymodiolus brevior and B. sp. affi- otic, chemoautotrophic bacteria. Mar. Biol. 75:187191.
nis brevior from the North Fiji Basin, western Pacific. Felbeck, H. 1985. CO2 fixation in the hydrothermal vent tube
Mar. Ecol. Progr. Ser. 165:187193. worm Riftia pachyptila (Jones). Physiol. Zool. 58:272
Dubilier, N., R. Amann, C. Erseus, G. Muyzer, S. Y. Park, O. 281.
Giere, and C. M. Cavanaugh. 1999. Phylogenetic diver- Felbeck, H., and D. L. Distel. 1991. Prokaryotic symbionts of
sity of bacterial endosymbionts in the gutless marine marine invertebrates. In: A. Balows A., H. G. Truper, M.
oligochete Olavius loisae (Annelida). Mar. Ecol. Progr. Dworkin, W. Harder, and K.-H. Schleifer (Eds.) The
Ser. 178:271280. Prokaryotes. Springer-Verlag. New York, NY.
Dubilier, N., C. Mulders, T. Ferdelman, D. de Beer, A. Pern- Felbeck, H., and P. J. Turner. 1995. CO2 transport in cathe-
thaler, M. Klein, M. Wagner, C. Erseus, F. Thiermann, J. terized hydrothermal vent tubeworms, Riftia pachyptila
Krieger, O. Giere, and R. Amann. 2001. Endosymbiotic (Vestimentifera). J. Exp. Zool. 272:95102.
sulphate-reducing and sulphide-oxidizing bacteria in an Felbeck, H., and J. Jarchow. 1998. Carbon release from puri-
oligochaete worm. Nature 411:298302. fied chemoautotrophic bacterial symbionts of the hydro-
502 C.M. Cavanaugh et al. CHAPTER 2.5

thermal vent tubeworm Riftia pachyptila. Physiol. Zool. Fisher, C. R. 1995. Toward an appreciation of hydrothermal
71:294302. vent animals: Their environment, physiological ecol-
Felbeck, H., C. Arndt, U. Hentschel, and J. J. Childress. 2004. ogy, and tissue stable isotope values. In: Seafloor
Experimental application of vascular and coelomic cath- Hydrothermal Systems: Physical, Chemical, Biological,
eterization to identify vascular transport mechanisms for and Geological Interactions. American Geophysical
inorganic carbon in the vent tubeworm, Riftia pachyp- Union.
tila. Deep Sea Res. Pt. I Oceanogr. Res. Pap. 51:401411. Fisher, C. R. 1996. Ecophysiology of primary production at
Feldman, R., M. Black, C. Cary, R. Lutz, and R. Vrijenhoek. deep-sea vents and seeps. Biosystem. Ecol. Ser. 11:313
1997. Molecular phylogenetics of bacterial endosym- 336.
bionts and their vestimentiferan hosts. Molec. Mar. Biol. Fox, M., S. K. Juniper, and H. Vali. 2002. Chemoautotrophy
Biotechnol. 6:268277. as a possible nutritional source in the hydrothermal vent
Fenchel, T., and B. J. Finlay. 1989. Kentrophorosa mouth- limpet Lepetodrilus fucensis. Cahiers Biol. Mar. 43:371
less ciliate with a symbiotic kitchen garden. Ophelia 376.
30:7593. Freytag, J. K., P. R. Girguis, D. C. Bergquist, J. P. Andras,
Fiala-Mdioni, A. 1984. Ultrastructural evidence of abun- J. J. Childress, and C. R. Fisher. 2001. A paradox
dance of intracellular symbiotic bacteria in the gill of resolved: Sulfide acquisition by roots of seep tubeworms
bivalve mollusks of deep hydrothermal vents. Comptes sustains net chemoautotrophy. Proc. Natl. Acad. Sci.
Rendus LAcad. Sci. Ser. III Sciences de la VieLife USA 98:1340813413.
Sciences 298:487492. Fry, B., H. Gest, and J. M. Hayes. 1983. Sulfur isotopic com-
Fiala-Mdioni, A., J. Boulgue, S. Ohta, H. Felbeck, and A. positions of deep-sea hydrothermal vent animals.
Mariotti. 1993. Source of energy sustaining the Calypto- Nature 306:5152.
gena populations from deep trenches in subduction Fry, B., and E. B. Sherr. 1984. 13C measurements as indica-
zones off Japan. Deep Sea Res. 40:12411258. tors of carbon flow in marine and freshwater ecosystems.
Fiala-Mdioni, A., J.-C. Michalski, J. Jolls, C. Alonso, and J. Contrib. Mar. Sci. 27:1347.
Montreuil. 1994. Lysosomic and lysozyme activities in Fujio, S. Z., and N. Imasato. 1991. Diagnostic calculation for
the gill of bivalves from deep hydrothermal vents. circulation and water mass movement in the deep
Comptes Rendus LAcad. Sci. Ser. III Sciences de la Pacific. J. Geophys. Res. Oceans 96:759774.
VieLife Sciences 317:239244. Fujiwara, Y., C. Kato, N. Masui, K. Fujikura, and S. Kojima.
Fiala-Mdioni, A., Z. P. McKiness, P. Dando, J. Boulegue, A. 2001. Dual symbiosis in the cold-seep thyasirid clam
Mariotti, A. M. Alayse-Danet, J. J. Robinson, and C. M. Maorithyas hadalis from the hadal zone in the Japan
Cavanaugh. 2002. Ultrastructural, biochemical, and Trench, western Pacific. Mar. Ecol. Progr. Ser. 214:151
immunological characterization of two populations of 159.
the mytilid mussel Bathymodiolus azoricus from the Fujiwara, Y., and K. Uematsu. 2002. Microdistribution of two
Mid-Atlantic Ridge: Evidence for a dual symbiosis. Mar. endosymbionts in gill tissue from a hadal thyasirid clam
Biol. 141:10351043. Maorithyas hadalis. Cahiers Biol. Mar. 43:341343.
Fisher, C. R., and J. J. Childress. 1986. Translocation of fixed Fustec, A., D. Desbruyres, and S. K. Juniper. 1987. Deep-
carbon from symbiotic bacteria to host tissues in the sea hydrothermal vent communities at 13N on the East
gutless bivalve, Solemya reidi. Mar. Biol. 93:5968. Pacific Rise: Microdistribution and temporal variations.
Fisher, C., J. Childress, A. Arp, J. Brooks, D. Distel, J. Favuzzi, Biol. Oceanogr. 4:121164.
H. Felbeck, R. Hessler, K. Johnson, M. Kennicut 2nd, S. Gearing, J. N., P. J. Gearing, D. T. Rudnick, A. G. Requejo,
Macko, A. Newton, M. Powell, G. Somero, and T. Soto. and M. J. Hutchins. 1984. Isotopic variability of organic-
1988. Microhabitat variation in the hydrothermal vent carbon in a phytoplankton-based, temperate estuary.
mussel, Bathymodiolus thermophilus, at the Rose Gar- Geochim. Cosmochim. Acta 48:10891098.
den vent on the Galapagos Rift. Deep Sea Res. 35:1769 Giere, O. 1981. The gutless marine oligochaete Phallodrilus
1791. leukodermatus: Structural studies on an aberrant tubifi-
Fisher, C. R. 1990. Chemoautotrophic and methanotrophic cid associated with bacteria. Mar. Ecol. Progr. Ser.
symbioses in marine invertebrates. Rev. Aquat. Sci. 5:353357.
2:399436. Giere, O. 1985. The gutless marine tubificid Phallodrilus pla-
Fisher, C. R., and J. J. Childress. 1992. Organic carbon trans- nus, a flattened oligochaete with symbiotic bacteria.
fer from methanotrophic symbionts to the host hydro- Zool. Scripta 14:279286.
carbon seep mussel. Symbiosis 12:221235. Giere, O., and C. Langheld. 1987. Structural organiza-
Fisher, C. R. 1993a. Oxidation of methane by deep-sea mytil- tion, transfer and biological fate of endosymbiotic
ids in the Gulf of Mexico. In: R. S. Oremland (Ed.) bacteria in gutless oligochaetes. Mar. Biol. 93:641
Biogeochemistry of Global Change: Radiatively Active 650.
Trace Gases. Chapman and Hall. New York, NY. Giere, O., and J. Krieger. 2001. A triple bacterial endosym-
Fisher, C. R., J. M. Brooks, J. S. Vodenichar, J. M. Zande, biosis in a gutless oligochaete (Annelida): Ultrastruc-
J. J. Childress, and R. A. Burke Jr. 1993b. The co- tural and immunocytochemical evidence. Invertebr.
occurrence of methanotrophic and chemoautotrophic Biol. 120:4149.
sulfur-oxidizing bacterial symbionts in a deep-sea Girguis, P. R., R. W. Lee, N. Desaulniers, J. J. Childress, M.
mussel. Marine Ecology Pubblicazioni Della Stazione Pospesel, H. Felbeck, and F. Zal. 2000. Fate of nitrate
Zoologica Di Napoli I 14:277289. acquired by the tubeworm Riftia pachyptila. Appl. Envi-
Fisher, C. R., J. J. Childress, S. A. Macko, and J. M. Brooks. ron. Microbiol. 66:27832790.
1994. Nutritional interactions in Galapagos Rift hydro- Girguis, P. R., J. J. Childress, J. K. Freytag, K. Klose, and R.
thermal vent communities: Inferences from stable car- Stuber. 2002. Effects of metabolite uptake on proton-
bon and nitrogen isotope analyses. Mar. Ecol. Progr. equivalent elimination by two species of deep-sea
Ser. 103:4555. vestimentiferan tubeworm, Riftia pachyptila and Lamel-
CHAPTER 2.5 Marine Chemosynthetic Symbioses 503

librachia cf luymesi: Proton elimination is a necessary brate-sulfur bacteria symbioses. Biol. Bull. 173:260
adaptation to sulfide-oxidizing chemoautotrophic sym- 276.
bionts. J. Exp. Biol. 205:30553066. Head, I. M., N. D. Gray, K. J. Clarke, R. W. Pickup, and J. G.
Glockner, F. O., H. D. Babenzien, J. Wulf, and R. Amann. Jones. 1996. The phylogenetic position and ultrastruc-
1999. Phylogeny and diversity of Achromatium oxaliferum. ture of the uncultured bacterium Achromatium
System. Appl. Microbiol. 22:2838. oxaliferum. Microbiol. SGM 142:23412354.
Goericke, R., J. P. Montoya, and B. Fry. 1994. Physiology of Hentschel, U., and H. Felbeck. 1993a. Nitrate respiration in
isotopic fractionation in algae and cyanobacteria. In: K. the hydrothermal vent tubeworm Riftia pachyptila.
Lajtha and R. H. Michener (Eds.) Stable Isotopes in Nature 366:338340.
Ecology and Environmental Science. Blackwell. Oxford, Hentschel, U., S. C. Cary, and H. Felbeck. 1993b. Nitrate
UK. respiration in chemoautotrophic symbionts of the
Goffredi, S. K., J. J. Childress, N. T. Desaulniers, and F. H. bivalve Lucinoma aequizonata. Mar. Ecol. Progr. Ser.
Lallier. 1997a. Sulfide acquisition by the hydrothermal 94:3541.
vent tubeworm Riftia pachyptila appears to be via Herry, A., and M. Le Pennec. 1987. Endosymbiotic bacteria
uptake of HS, rather than H2S. J. Exp. Biol. 200:2069 in the gills of the littoral bivalve molluscs Thyasira flex-
2616. uosa (Thyasiridae) and Lucinella divaricata (Lucinidae).
Goffredi, S. K., J. J. Childress, N. T. Desaulniers, R. W. Lee, Symbiosis 4:2536.
F. H. Lallier, and D. Hammond. 1997b. Inorganic carbon Hughes, D. S., H. Felbeck, and J. L. Stein. 1997. A histidine
acquisition by the hydrothermal vent tubeworm Riftia protein kinase homolog from the endosymbiont of the
pachyptila depends upon high external pCO2 and upon hydrothermal vent tubeworm Riftia pachyptila. Appl.
proton-equivalent ion transport by the worm. J. Exp. Environ. Microbiol. 63:34943498.
Biol. 200:883896. Hughes, D. S., H. Felbeck, and J. L. Stein. 1998. Signal trans-
Goffredi, S. K., J. J. Childress, F. H. Lallier, and N. T. duction and motility genes from the bacterial endosym-
Desaulniers. 1999. The ionic composition of the hydro- bionts of Riftia pachyptila. Cahiers Biol. Mar. 39:305
thermal vent tube worm Riftia pachyptila: Evidence for 308.
the elimination of SO42 and H+ and for a Cl/HCO3 Humes, A. G., and R. A. Lutz. 1994. Aphotopontius acanthi-
shift. Physiol. Biochem. Zool. 72:296306. nus, new species (Copepoda, Siphonostomatoida), from
Goffredi, S. K., and J. J. Childress. 2001. Activity and inhibi- deep-sea hydrothermal vents on the East Pacific Rise. J.
tor sensitivity of ATPases in the hydrothermal vent tube- Crustac. Biol. 14:337345.
worm Riftia pachyptila: A comparative approach. Mar. Hunt, S. 1992. Structure and composition of the shell of the
Biol. 138:259265. archaeogastropod limpet Lepetodrilus elevatus-elevatus
Grassle, J. F. 1985. Hydrothermal vent animals: Distribution (Mclean, 1988). Malacologia 34:129141.
and biology. Science 229:713717. Hurtado, L. A. 2002. Evolution and Biogeography of
Gray, N. D., R. Howarth, A. Rowan, R. W. Pickup, J. G. Jones, Hydrothermal Vent Organisms in the Eastern Pacific
and I. M. Head. 1999. Natural communities of Achro- Ocean [PhD thesis]. Rutgers University. New Brun-
matium oxaliferum comprise genetically, morphologi- swick, NC.
cally, and ecologically distinct subpopulations. Appl. Hurtado, L. A., M. Mateos, R. A. Lutz, and R. C. Vrijenhoek.
Environ. Microbiol. 65:50895099. 2003. Coupling of bacterial endosymbiont and host
Grieshaber, M. K., and S. Volkel. 1998. Animal adaptations mitochondrial genomes in the hydrothermal vent clam
for tolerance and exploitation of poisonous sulfide. Ann. Calyptogena magnifica. Appl. Environ. Microbiol. 69:
Rev. Physiol. 60:3353. 20582064.
Gros, O., A. Darrasse, P. Durand, L. Frenkiel, and M. Jacobs, D. K., and D. R. Lindberg. 1998. Oxygen and evolu-
Moueza. 1996. Environmental transmission of a sulfur- tionary patterns in the sea: Onshore/offshore trends and
oxidizing bacterial gill endosymbiont in the tropical recent recruitment of deep-sea faunas. Proc. Natl. Acad.
lucinid bivalve Codakia orbicularis. Appl. Environ. Sci. USA 95:93969401.
Microbiol. 62:23242330. Johnson, K. S., J. J. Childress, R. R. Hessler, C. M. Sakamoto-
Gros, O., P. De Wulf-Durand, L. Frenkiel, and M. Moueza. Arnold, and C. L. Beehler. 1988. Chemical and biologi-
1998. Putative environmental transmission of sulfur- cal interactions in the Rose Garden hydrothermal vent
oxidizing bacterial symbionts in tropical lucinid bivalves field, Galapagos spreading center. Deep Sea Res. 35:
inhabiting various environments. FEMS Microbiol. Lett. 17231744.
160:257262. Johnson, M., M. Diouris, and M. Le Pennec. 1994. Endosym-
Gros, O., M. Liberge, and H. Felbeck. 2003a. Interspecific biotic bacterial contribution in the carbon nutrition of
infection of aposymbiotic juveniles of Codakia orbicu- Loripes lucinalis (Mollusca : Bivalvia). Symbiosis 17:1
laris by various tropical lucinid gill endosymbionts. Mar. 13.
Biol. 142:5766. Jones, M. L. 1981. Riftia pachyptila Jones: Observations on
Gros, O., M. Liberge, A. Heddi, C. Khatchadourian, and H. the vestimentiferan worm from the Galapagos Rift. Sci-
Felbeck. 2003b. Detection of the free-living forms of ence 213:333336.
sulfide-oxidizing gill endosymbionts in the lucinid habi- Jones, J. L., and S. L. Gardiner. 1985. Light and scanning
tat (Thalassia testudinum environment). Appl. Environ. electron microscope studies of spermatogenesis in the
Microbiol. 69:62646267. vestimentiferan tube worm Riftia pachyptila
Gustafson, R. G., and R. G. B. Reid. 1988. Association of (Pogonophora : Obturata). Trans. Am. Microscop. Soc.
bacteria with larvae of the gutless protobranch bivalve 104:118.
Solemya reidi (Cryptodonta, Solemyidae). Mar. Biol. Jrgensen, B. B., and J. R. Postgate. 1982. Ecology of the
97:389401. bacteria of the sulfur cycle with special reference to
Hand, S. C. 1987. Trophosome ultrastructure and the char- anoxic oxic interface environments. Phil. Trans. Roy.
acterization of isolated bacteriocytes from inverte- Soc. Lond. Ser. B Biol. Sci. 298:543561.
504 C.M. Cavanaugh et al. CHAPTER 2.5

Joyner, J. L., S. M. Peyer, and R. W. Lee. 2003. Possible roles Lee, R. W., J. J. Childress, and N. T. Desaulniers. 1997. The
of sulfur-containing amino acids in a chemoautotrophic effects of exposure to ammonia on ammonia and taurine
bacterium-mollusc symbiosis. Biol. Bull. 205:331338. pools of the symbiotic clam Solemya reidi. J. Exp. Biol.
Julian, D., F. Gaill, E. Wood, A. J. Arp, and C. R. Fisher. 1999. 200:27972805.
Roots as a site of hydrogen sulfide uptake in the hydro- Lee, R. W., J. J. Robinson, and C. M. Cavanaugh. 1999. Path-
carbon seep vestimentiferan Lamellibrachia sp. J. Exp. ways of inorganic nitrogen assimilation in chemoau-
Biol. 202:22452257. totrophic bacteria-marine invertebrate symbioses:
Karl, S. A., S. Schutz, D. Desbruyres, R. Lutz, and R. C. Expression of host and symbiont glutamine synthetase.
Vrijenhoek. 1996. Molecular analysis of gene flow in the J. Exp. Biol. 202:289300.
hydrothermal vent clam (Calyptogena magnifica). Le Pennec, M., and A. Hily. 1984. Anatomy, structure and
Molec. Mar. Biol. Biotechnol. 5:193202. ultrastructure of the gill of a deep-sea hydrothermal vent
Kennicutt, M. C., R. A. Burke, I. R. MacDonald, J. M. mytilid. Oceanol. Acta 7:517523.
Brooks, G. J. Denoux, and S. A. Macko. 1992. Stable Levin, B. R., and C. T. Bergstrom. 2000. Bacteria are differ-
isotope partitioning in seep and vent organisms: Chem- ent: Observations, interpretations, speculations, and
ical and ecological significance. Chem. Geol. 101:293 opinions about the mechanisms of adaptive evolution in
310. prokaryotes. Proc. Natl. Acad. Sci. USA 97:69816985.
Kennish, M. J., and R. A. Lutz. 1999. Calcium carbonate Levin, L. A., and R. H. Michener. 2002. Isotopic evidence for
dissolution rates in deep-sea bivalve shells on the East chemosynthesis-based nutrition of macrobenthos: The
Pacific Rise at 21N: Results of an 8-year in situ exper- lightness of being at Pacific methane seeps. Limnol.
iment. Palaeogeogr. Palaeoclimatol. Palaeoecol. 154: Oceanogr. 47:13361345.
293299. Li, L., C. Kato, and K. Horikoshi. 1999. Microbial diversity
Kimura, M., and G. H. Weiss. 1964. The stepping stone model in sediments collected from the deepest cold-seep area,
of population structure and the decrease of genetic cor- the Japan Trench. Mar. Biotechnol. 1:391400.
relation with distance. Genetics 49:561576. Lilley, M. D., D. A. Butterfield, E. J. Olson, J. E. Lupton, S.
Kochevar, R. E., and J. J. Childress. 1996. Carbonic anhydrase A. Macko, and R. E. McDuff. 1993. Anomalous CH4 and
in deep-sea chemoautotrophic symbioses. Mar. Biol. NH4+ concentrations at an unsedimented mid-ocean-
125:375383. ridge hydrothermal system. Nature 364:4547.
Krieger, J., O. Giere, and N. Dubilier. 2000. Localization of Little, C. T. S., R. J. Herrington, V. V. Maslennikov, N. J.
RubisCO and sulfur in endosymbiotic bacteria of the Morris, and V. V. Zaykov. 1997. Silurian hydrothermal-
gutless marine oligochaete Inanidrilus leukodermatus vent community from the southern Urals, Russia.
(Annelida). Mar. Biol. 137:239244. Nature 385:146148.
Krueger, D. M. 1996a. Ecology and Evolution of Solemya Little, C. L. 2002. The fossil record of hydrothermal vent
spp., Marine Bivalves Living in Symbiosis with communities. Cahiers Biol. Mar. 43:313316.
Chemoautotrophic Bacteria [PhD thesis]. Harvard Uni- Little, C. T. S., and I. H. Thorseth. 2002. Hydrothermal vent
versity. Cambridge, MA. microbial communities: A fossil perspective. Cahiers
Krueger, D. M., N. Dubilier, and C. M. Cavanaugh. 1996b. Biol. Mar. 43:317319.
Chemoautotrophic symbiosis in the tropical solemyid Little, C. T. S., and R. C. Vrijenhoek. 2003. Are hydrothermal
Solemya occidentalis (Bivalvia : Protobranchia): Ultra- vent animals living fossils? Trends Ecol. Evol. 18:582
structural and phylogenetic analysis. Mar. Biol. 126:55 588.
64. Little, C. T. S., T. Danelian, R. J. Herrington, and R. M.
Krueger, D. M., and C. M. Cavanaugh. 1997. Phylogenetic Haymon. 2004. Early Jurassic hydrothermal vent com-
diversity of bacterial symbionts of Solemya hosts based munity from the Franciscan Complex, California. J.
on comparative sequence analysis of 16S rRNA genes. Paleontol. 78:542559.
Appl. Environ. Microbiol. 63:9198. Lonsdale, P. 1977. Clustering of suspension-feeding macro-
Lalou, C., and E. Brichet. 1982. Ages and implications of benthos near abyssal hydrothermal vents at oceanic
East Pacific Rise sulfide deposits at 21N. Nature spreading centers. Deep Sea Res. 24:857.
300:169171. Lopez-Garcia, P., F. Gaill, and D. Moreira. 2002. Wide bac-
Laubier, L. 1989. Ecosystmes benthiques profonds et chimi- terial diversity associated with tubes of the vent worm
osynthse bactrienne: Sources hydrothermales et Riftia pachyptila. Environ. Microbiol. 4:204215.
suintement froids [Deep benthic ecosystems and bacte- Luther, G. W. I., T. F. Rozan, M. Taillefert, D. B. Nuzzio, C.
rial chemosynthesys: Hydrothermal wells and cold seep- Di Meo, T. M. Shank, R. A. Lutz, and S. C. Cary. 2001.
ages]. In: M. Denis (Ed.) Oceanology: Today and in the Chemical speciation drives hydrothermal vent ecology.
Near Future. Nature 410:813816.
Laue, B. E., and D. C. Nelson. 1994. Characterization of the Lutz, R. A., D. Jablonski, D. C. Rhoads, and R. D. Turner.
gene encoding the autotrophic ATP sulfurylase from 1980. Larval dispersal of a deep-sea hydrothermal vent
the bacterial endosymbiont of the hydrothermal vent bivalve from the Galapagos Rift. Mar. Biol. 57:127133.
tubeworm Riftia pachyptila. J. Bacteriol. 176:3723 MacAvoy, S. E., S. A. Macko, and S. B. Joye. 2002. Fatty acid
3729. carbon isotope signatures in chemosynthetic mussels
Laue, B. E., and D. C. Nelson. 1997. Sulfur-oxidizing sym- and tubeworms from Gulf of Mexico hydrocarbon seep
bionts have not co-evolved with their hydrothermal vent communities. Chem. Geol. 185:18.
tubeworm hosts: An RFLP analysis. Molec. Mar. Biol. Marsh, A. G., L. S. Mullineaux, C. M. Young, and D. T.
Biotechnol. 6:180188. Manahan. 2001. Larval dispersal potential of the tube-
Lee, R., and J. Childress. 1994. Assimilation of inorganic worm Riftia pachyptila at deep-sea hydrothermal vents.
nitrogen by marine invertebrates and their chemoau- Nature 411:7780.
totrophic and methanotrophic symbionts. Appl. Envi- McArthur, A. G., and V. Tunnicliffe. 1998. Relics and antiq-
ron. Microbiol. 60:18521858. uity revisited in the modern vent fauna. In: R. Mills and
CHAPTER 2.5 Marine Chemosynthetic Symbioses 505

K. Harrison (Eds.) Modern Ocean Floor Processes and Nelson, D. C., K. D. Hagen, and D. B. Edwards. 1995b. The
the Geological Record. Geological Society of London. gill symbiont of the hydrothermal vent mussel
London, UK. Bathymodiolus thermophilus is a psychrophilic, chemo-
McKiness, Z. P. 2004. Evolution of Endosymbioses in Deep- autotrophic, sulfur bacterium. Mar. Biol. 121: 487
sea Bathymodioline Mussels (Mollusca:Bivalvia) [PhD 495.
thesis]. Harvard University. Cambridge, MA. Nelson, K., and C. R. Fisher. 2000. Absence of cospeciation
McMullin, E. R., S. Hourdez, S. W. Schaeffer, and C. R. in deep-sea vestimentiferan tubeworms and their bacte-
Fisher. 2003. Phylogeny and biogeography of deep-sea rial endosymbionts. Symbiosis 28:115.
vestimentiferan tubeworms and their bacterial sym- Niemczuk, K., and M. Kondracki. 2004. Q fever as zoonosis.
bionts. Symbiosis 34:141. Medycyna Weterynaryjna 60:129131.
Metivier, B., and R. Voncosel. 1993. Acharax alinae n sp., a OBrien, J., and R. D. Vetter. 1990. Production of thiosul-
giant solemyid (Mollusca, Bivalvia) from the Lau Basin. phate during sulphide oxidation by mitochondria of the
Comptes Rendus LAcad. Sci. Ser. III Sciences de la symbiont-containing bivalve Solemya reidi. J. Exp. Biol.
VieLife Sciences 316:229237. 149:133148.
Michener, R. H., and D. M. Schell. 1994. Stable isotope ratios Ohta, T. 1973. Slightly deleterious mutant substitutions in
as tracers in marine aquatic food webs. In: K. Lajtha and evolution. Nature 246:9698.
R. H. Michener (Eds.) Stable Isotopes in Ecology and Ott, J. A., R. Novak, F. Schiemer, U. Hentschel, M. Nebel-
Environmental Science. Blackwell. Oxford, UK. sick, and M. Polz. 1991. Tackling the sulfide gradient:
Millero, F. J., T. Plese, and M. Fernandez. 1987. The A novel strategy involving marine nematodes and
dissociation of hydrogen-sulfide in seawater. Limnol. chemoautotrophic ectosymbionts. Marine Ecology
Oceanogr. 33:269274. Pubblicazioni Della Stazione Zoologica Di Napoli I
Millikan, D. S., H. Felbeck, and J. L. Stein. 1999. Identifica- 12:261279.
tion and characterization of a flagellin gene from the Ott, J. A., M. Bright, and F. Schiemer. 1998. The ecology of
endosymbiont of the hydrothermal vent tubeworm a novel symbiosis between a marine peritrich ciliate and
Riftia pachyptila. Appl. Environ. Microbiol. 65:3129 chemoautotrophic bacteria. Marine EcologyPubbli-
3133. cazioni Della Stazione Zoologica Di Napoli I 19:229
Minagawa, M., and E. Wada. 1984. Stepwise enrichment of 243.
15
N along food chainsfurther evidence and the rela- Peek, A. S., R. C. Vrijenhoek, and B. S. Gaut. 1998. Acceler-
tionship between delta 15N and animal age. Geochim. ated evolutionary rate in sulfur-oxidizing endosymbiotic
Cosmochim. Acta 48:11351140. bacteria associated with the mode of symbiont transmis-
Minic, Z., V. Simon, B. Penverne, F. Gaill, and G. Herve. 2001. sion. Molec. Biol. Evol. 15:15141523.
Contribution of the bacterial endosymbiont to the bio- Pimenov, N. V., M. G. Kalyuzhnaya, V. N. Khmelenina, L. L.
synthesis of pyrimidine nucleotides in the deep-sea tube- Mityushina, and Y. A. Trotsenko. 2002. Utilization of
worm Riftia pachyptila. J. Biol. Chem. 276:2377723784. methane and carbon dioxide by symbiotrophic bacteria
Minic, Z., and G. Herve. 2003. Arginine metabolism in the in gills of mytilidae (Bathymodiolus) from the rainbow
deep-sea tube worm Riftia pachyptila and its bacterial and Logachev hydrothermal fields on the Mid-Atlantic
endosymbiont. J. Biol. Chem. 278:4052740533. Ridge. Microbiology 71:587594.
Mira, A., and N. A. Moran. 2002. Estimating population size Podar, M., L. Mullineaux, H. R. Huang, P. S. Perlman, and
and transmission bottlenecks in maternally transmitted M. L. Sogin. 2002. Bacterial group II introns in a deep-
endosymbiotic bacteria. Microb. Ecol. 44:137143. sea hydrothermal vent environment. Appl. Environ.
Miura, T., M. Nedachi, and A. Hashimoto. 2002. Sulphur Microbiol. 68:63926398.
sources for chemoautotrophic nutrition of shallow water Polz, M. F., H. Felbeck, R. Novak, M. Nebelsick, and J. A.
vestimentiferan tubeworms in Kagoshima Bay. J. Mar. Ott. 1992. Chemoautotrophic, sulfur-oxidizing symbi-
Biol. Assoc. UK 82:537540. otic bacteria on marine nematodes: Morphological and
Moran, N., and P. Baumann. 1994. Phylogenetics of cytoplas- biochemical characterization. Microb. Ecol. 24:313
mically inherited microorganisms of arthropods. Trends 329.
Ecol. Evol. 9:1520. Polz, M. F., D. L. Distel, B. Zarda, R. Amann, H. Felbeck,
Moran, N. A. 1996. Accelerated evolution and Mullers J. A. Ott, and C. M. Cavanaugh. 1994. Phylogenetic anal-
rachet in endosymbiotic bacteria. Proc. Natl. Acad. Sci. ysis of a highly specific association between ectosymbi-
USA 93:28732878. otic, sulfur-oxidizing bacteria and a marine nematode.
Moran, N. A. 2003. Tracing the evolution of gene loss in Appl. Environ. Microbiol. 60:44614467.
obligate bacterial symbionts. Curr. Opin. Microbiol. Polz, M. F., and C. M. Cavanaugh. 1995. Dominance of one
6:512518. bacterial phylotype at a Mid-Atlantic Ridge hydrother-
Muller, H. J. 1964. The relation of recombination to muta- mal vent site. Proc. Natl. Acad. Sci. USA 92:72327236.
tional advance. Mut. Res. 1:29. Polz, M. F., J. J. Robinson, C. M. Cavanaugh, and C. L. Van
Mullineaux, L. S., P. H. Wiebe, and E. T. Baker. 1995. Larvae Dover. 1998. Trophic ecology of massive shrimp aggre-
of benthic invertebrates in hydrothermal vent plumes gations at a Mid-Atlantic Ridge hydrothermal vent site.
over Juan de Fuca Ridge. Mar. Biol. 122:585596. Limnol. Oceanogr. 43:16311638.
Mullineaux, L. S., K. G. Speer, A. M. Thurnherr, M. E. Polz, M. F., C. Harbison, and C. M. Cavanaugh. 1999. Diver-
Maltrud, and A. Vangriesheim. 2002. Implications of sity and heterogeneity of epibiotic bacterial communi-
cross-axis flow for larval dispersal along mid-ocean ties on the marine nematode Eubostrichus dianae. Appl.
ridges. Cahiers Biol. Mar. 43:281284. Environ. Microbiol. 65:42714275.
Nelson, D., and C. Fisher. 1995a. Chemoautotrophic and Polz, M. F., J. A. Ott, M. Bright, and C. M. Cavanaugh. 2000.
methanotrophic endosymbiotic bacteria at deep-sea When bacteria hitch a ride. ASM News 66:531539.
vents and seeps. In: D. Karl (Ed.) Microbiology of Deep- Poulin, E., S. von Boletzky, and J. P. Feral. 2001. Combined
sea Hydrothermal Vents. CRC Press. Boca Raton, FL. ecological factors permit classification of developmental
506 C.M. Cavanaugh et al. CHAPTER 2.5

patterns in benthic marine invertebrates: A discussion Scott, K. M. 2003. A d13C-based carbon flux model for the
note. J. Exp. Mar. Biol. Ecol. 257:109115. hydrothermal vent chemoautotrophic symbiosis Riftia
Powell, M. A., and G. N. Somero. 1986. Adaptations to sul- pachyptila predicts sizeable CO2 gradients at the host-
fide by hydrothermal vent animals: Sites and mecha- symbiont interface. Environ. Microbiol. 5:424432.
nisms of detoxification and metabolism. Biol. Bull. Scott, K. M., J. Schwedock, D. P. Schrag, and C. M.
171:274290. Cavanaugh. 2004. Influence of form IA RubisCO and
Pradillon, F., B. Shillito, C. M. Young, and F. Gaill. 2001. environmental dissolved inorganic carbon on the
Deep-sea ecology: Developmental arrest in vent worm delta13C of the clam-chemoautotroph symbiosis Sole-
embryos. Nature 413:698699. mya velum. Environ. Microbiol. 6(12):12101219.
Prior, S. D., and H. Dalton. 1985. The effect of copper ions Searcy, D. G. 1992. Origins of mitochondria and chloroplasts
on membrane content and methane monooxygenase from sulfur-based symbioses. H. Hartman and K.
activity in methanol-grown cells of Methylococcus cap- Matsuno (Eds.) The Origin and Evolution of the Cell:
sulatus (Bath). J. Gen. Microbiol. 131:155163. World Scientific Press.
Pruski, A. M., A. Fiala-Mdioni, C. R. Fisher, and J. C. Sibuet, M., and K. Olu. 1998. Biogeography, biodiversity
Colomines. 2000a. Composition of free amino acids and and fluid dependence of deep-sea cold-seep communi-
related compounds in invertebrates with symbiotic bac- ties at active and passive margins. Deep Sea Res. II
teria at hydrocarbon seeps in the Gulf of Mexico. Mar. 45:517567.
Biol. 136:411420. Smith, C. R., and A. R. Baco. 2003. Ecology of whale falls at
Pruski, A. M., A. Fiala-Mdioni, R. Prodon, and J. C. the deep-sea floor. Oceanogr. Mar. Biol. 41:311354.
Colomines. 2000b. Thiotaurine is a biomarker of sulfide- Southward, A. J., E. C. Southward, P. R. Dando, G. H. Rau,
based symbiosis in deep-sea bivalves. Limnol. Oceanogr. H. Felbeck, and H. Flugel. 1981. Bacterial symbionts and
45:18601867. low 13C/12C ratios in tissues of Pogonophora indicate
Pruski, A. M., and A. Fiala-Mdioni. 2003. Stimulatory effect unusual nutrition and metabolism. Nature 293:616620.
of sulphide on thiotaurine synthesis in three hydrother- Southward, E. C. 1982. Bacterial symbionts in Pogonophora.
mal vent species from the East Pacific Rise. J. Exp. Biol. J. Mar. Biol. Assoc. UK 62:889906.
206:29232930. Southward, A. J., and E. C. Southward. 1988. Pogonophora:
Rau, G. A., and J. I. Hedges. 1979. Carbon-13 depletion in a Tube worms dependent on endosymbiotic bacteria. ISI
hydrothermal vent mussel: Suggestion of a chemosyn- Atlas of Science, Animal & Plant Sciences 1:203207.
thetic food source. Science 203:648649. Southward, E. C., V. Tunnicliffe, and M. Black. 1995. Revi-
Rau, G. H. 1981. Hydrothermal vent clam and tube worm sion of the species of Ridgeia from North-East Pacific
13
C/12C: Further evidence of non photosynthetic food hydrothermal vents, with a redescription of Ridgeia pisc-
sources. Science 213:338340. esae Jones (Pogonophora : Obturata = Vestimentifera).
Robinson, J. J., and C. M. Cavanaugh. 1995. Expressions of Can. J. Zool. 73:282295.
form I and form II RubisCO in chemoautotrophic Spiro, B., P. B. Greenwood, A. J. Southward, and P. R. Dando.
symbioses: Implications for the interpretation of stable 1986. 13C/12C ratios in marine invertebrates from reduc-
carbon isotope values. Limnol. Oceanogr. 40:1496 ing sediments: Confirmation of nutritional importance
1502. of chemoautotrophic endosymbiotic bacteria. Mar.
Robinson, J. J., M. F. Polz, A. Fiala-Medioni, and C. M. Ecol. Progr. Ser. 28:233240.
Cavanaugh. 1998. Physiological and immunological Stanley, S. M. 1970. Shell form and life habits of the Bivalvia.
evidence for two distinct C-1-utilizing pathways in Geol. Soc. Am. Mem. 125:119121.
Bathymodiolus puteoserpentis (Bivalvia : Mytilidae), a Stein, J., S. Cary, R. Hessler, S. Ohta, R. Vetter, J. Childress,
dual endosymbiotic mussel from the Mid-Atlantic and H. Felbeck. 1988. Chemoautotrophic symbiosis in
Ridge. Mar. Biol. 132:625633. a hydrothermal vent gastropod. Biol. Bull. 174:373
Robinson, J. J., K. M. Scott, S. T. Swanson, M. H. OLeary, 378.
K. Horken, and F. R. Tabita. 2003. Kinetic isotope effect Stumm, W., and J. J. Morgan. 1996. Aquatic Chemistry, 3rd
and characterization of form II RubisCO from the ed. Wiley-Interscience.
chemoautotrophic endosymbionts of the hydrothermal Svavarsson, J., J. O. Stromberg, and T. Brattegard. 1993. The
vent tubeworm Riftia pachyptila. Limnol. Oceanogr. deep-sea asellote (Isopoda, Crustacea) fauna of the
48:4854. Northern Seas: Species composition, distributional pat-
Rouxel, O., Y. Fouquet, and J. N. Ludden. 2004. Subsurface terns and origin. J. Biogeogr. 20:537555.
processes at the Lucky Strike hydrothermal field, Mid- Temara, A., C. de Ridder, J. G. Kuenen, and L. A. Robertson.
Atlantic Ridge: Evidence from sulfur, selenium, and 1993. Sulfide-oxidizing bacteria in the burrowing echi-
iron isotopes. Geochim. Cosmochim. Acta 68:2295 noid, Echinocardium cordatum (Echinodermata). Mar.
2311. Biol. 115:179185.
Sarrazin, J., and S. K. Juniper. 1999. Biological characteristics Thao, M., N. Moran, P. Abbot, E. Brennan, D. Burckhardt,
of a hydrothermal edifice mosaic community. Mar. Ecol. and P. Baumann. 2000. Cospeciation of psyllids and their
Progr. Ser. 185:119. primary prokaryotic endosymbionts. Appl. Environ.
Schiemer, F., R. Novak, and J. Ott. 1990. Metabolic studies Microbiol. 66:28982905.
on thiobiotic free-living nematodes and their symbiotic Trask, J. L., and C. L. Van Dover. 1999. Site-specific and
microorganisms. Mar. Biol. 106:129137. ontogenetic variations in nutrition of mussels (Bathymo-
Schmaljohann, R., and H. J. Flgel. 1987. Methane-oxidizing diolus sp.) from the Lucky Strike hydrothermal vent
bacteria in pogonophora. Sarsia 72:9198. field, Mid-Atlantic Ridge. Limnol. Oceanogr. 44:334
Schweimanns, M., and H. Felbeck. 1985. Significance of the 343.
occurrence of chemoautotrophic bacterial endosym- Tunnicliffe, V. 1988. Biogeography and evolution of hydro-
bionts in lucinid clams from Bermuda. Mar. Ecol. Progr. thermal-vent fauna in the eastern Pacific Ocean. Proc.
Ser. 24:113120. Roy. Soc. Lond B 233:347366.
CHAPTER 2.5 Marine Chemosynthetic Symbioses 507

Tunnicliffe, V. 1991. The biology of hydrothermal vents: Wernegreen, J. J. 2002. Genome evolution in bacterial endo-
Ecology and evolution. Oceanogr. Mar. Biol. 29:319 symbionts of insects. Nature Rev. Genet. 3:850861.
407. Windoffer, R., and O. Giere. 1997. Symbiosis of the hydro-
Tunnicliffe, V., and C. Fowler. 1996. Influence of sea-floor thermal vent gastropod Ifremeria nautilei (Provan-
spreading on the global hydrothermal vent fauna. nidae) with endobacteria: Structural analyses and
Nature 379:531533. ecological considerations. Biol. Bull. 193:381392.
Turner, R. D., R. A. Lutz, and D. Jablonski. 1985. Modes of Wirsen, C. O., S. M. Sievert, C. M. Cavanaugh, S. J.
molluscan larval development at deep-sea hydrothermal Molyneaux, A. Ahmad, L. T. Taylor, E. F. DeLong, and
vents. Biol. Soc. Wash. Bull. 6:167184. C. D. Taylor. 2002. Characterization of an autotrophic
Tyler, P. A., and C. M. Young. 2003. Dispersal at hydrother- sulfide-oxidizing marine Arcobacter sp. that produces
mal vents: A summary of recent progress. Hydrobiologia filamentous sulfur. Appl. Environ. Microbiol. 68:316
503:919. 325.
Vacelet, J., N. Boury-Esnault, A. Fiala-Medioni, and C. R. Woldehiwet, Z. 2004. Q fever (coxiellosis): Epidemiology
Fisher. 1995. A methanotrophic carnivorous sponge. and pathogenesis. Res. Vet. Sci. 77:93100.
Nature 377:296. Won, Y. J., S. J. Hallam, G. D. OMullan, I. L. Pan, K. R. Buck,
Vacelet, J., A. Fiala-Mdioni, C. R. Fisher, and N. Boury- and R. C. Vrijenhoek. 2003a. Environmental acquisition
Esnault. 1996. Symbiosis between methane-oxidizing of thiotrophic endosymbionts by deep-sea mussels of
bacteria and a deep-sea carnivorous cladorhizid sponge. the genus Bathymodiolus. Appl. Environ. Microbiol.
Mar. Ecol. Progr. Ser. 145:7785. 69:67856792.
Van Dover, C. L., B. Fry, J. F. Grassle, S. Humphris, and P. A. Won, Y., C. R. Young, R. A. Lutz, and R. C. Vrijenhoek.
Rona. 1988. Feeding biology of the shrimp Rimicaris 2003b. Dispersal barriers and isolation among deep-sea
exoculata at hydrothermal vents on the Mid-Atlantic mussel populations (Mytilidae:Bathymodiolus) from
Ridge. Mar. Biol. 98:209216. eastern Pacific hydrothermal vents. Molec. Ecol. 12:169
Van Dover, C. L., and B. Fry. 1994. Microorganisms as food 184.
resources at deep-sea hydrothermal vents. Limnol. Yamamoto, H., K. Fujikura, A. Hiraishi, K. Kato, and Y.
Oceanogr. 39:5157. Maki. 2002. Phylogenetic characterization and biomass
Van Dover, C. 2000. The Ecology of Deep-sea Hydrothermal estimation of bacterial endosymbionts associated with
Vents. Princeton University Press. Princeton, NJ. invertebrates dwelling in chemosynthetic communities
Van Dover, C. L., S. E. Humphris, D. Fornari, C. M. of hydrothermal vent and cold seep fields. Mar. Ecol.
Cavanaugh, R. Collier, S. K. Goffredi, J. Hashimoto, M. Progr. Ser. 245:6167.
Lilley, A. L. Reysenbach, T. M. Shank, K. L. Von Damm, Yamanaka, T., C. Mizota, H. Satake, F. Kouzuma, T. Gamo,
A. Banta, R. M. Gallant, D. Gtz, D. Green, J. Hall, U. Tsunogai, T. Miwa, and K. Fujioka. 2003. Stable
T. L. Harmer, L. A. Hurtado, P. Johnson, Z. P. McKiness, isotope evidence for a putative endosymbiont-based
C. Meredith, E. Olson, I. L. Pan, M. Turnipseed, Y. Won, lithotrophic Bathymodiolus sp. mussel community atop
C. R. Young 3rd, and R. C. Vrijenhoek. 2001. Biogeog- a serpentine seamount. Geomicrobiol. J. 20:185197.
raphy and ecological setting of Indian Ocean hydrother- Yong, R., and D. G. Searcy. 2001. Sulfide oxidation coupled
mal vents. Science 294:818822. to ATP synthesis in chicken liver mitochondria. Comp.
Van Dover, C. 2002a. Trophic relationships among inverte- Biochem. Physiol. Pt. B 129:129137.
brates at the Kairei hydrothermal vent field (Central Young, C. M., E. Vazquez, A. Metaxas, and P. A. Tyler. 1996.
Indian Ridge). Mar. Biol. 141:761772. Embryology of vestimentiferan tube worms from deep-
Van Dover, C. L., C. R. German, K. G. Speer, L. M. Parson, sea methane/sulphide seeps. Nature 381:514516.
and R. C. Vrijenhoek. 2002b. Marine biology: Evolution Zal, F., F. H. Lallier, J. S. Wall, S. N. Vinogradov, and A.
and biogeography of deep-sea vent and seep inverte- Toulmond. 1996. The multi-hemoglobin system of the
brates. Science 295:12531257. hydrothermal vent tube worm Riftia pachyptila.1: Reex-
Vereshchaka, A. L., G. M. Vinogradov, A. Y. Lein, S. Dalton, amination of the number and masses of its constituents.
and F. Dehairs. 2000. Carbon and nitrogen isotopic com- J. Biol. Chem. 271:88698874.
position of the fauna from the Broken Spur hydrother- Zal, F., B. N. Green, F. H. Lallier, and A. Toulmond. 1997.
mal vent field. Mar. Biol. 136:1117. Investigation by electrospray ionization mass spectrom-
Vrijenhoek, R., T. Shank, and R. Lutz. 1998. Gene flow etry of the extracellular hemoglobin from the polycha-
and dispersal in deep-sea hydrothermal vent animals. ete annelid Alvinella pompejana: An unusual hexagonal
Cahiers Biol. Mar. 39:363366. bilayer hemoglobin. Biochemistry 36:1177711786.
Waser, N. A. D., P. J. Harrison, B. Nielsen, S. E. Calvert, and Zal, F., E. Leize, F. H. Lallier, A. Toulmond, A. Van Dorsse-
D. H. Turpin. 1998. Nitrogen isotope fractionation dur- laer, and J. J. Childress. 1998. S-sulfohemoglobin and
ing the uptake and assimilation of nitrate, nitrite, ammo- disulfide exchange: The mechanisms of sulfide binding
nium, and urea by a marine diatom. Limnol. Oceanogr. by Riftia pachyptila hemoglobins. Proc. Natl. Acad. Sci.
43:215224. USA 95:89979002.
Watanabe, A. 2004. Various clinical types of Q-fever disease. Zal, F., E. Leize, D. R. Oros, S. Hourdez, A. Van Dorsselaer,
Int. Med. 43:12. and J. J. Childress. 2000. Haemoglobin structure and
Weber, R. E., and S. N. Vinogradov. 2001. Nonvertebrate biochemical characteristics of the sulphide-binding com-
hemoglobins: Functions and molecular adaptations. ponent from the deep-sea clam Calyptogena magnifica.
Physiol. Rev. 81:569628. Cahiers Biol. Mar. 41:413423.
Weiser, W. 1959. Eine ungewhnliche Assoziation zwischen Zhang, J. Z., and F. J. Millero. 1993. The products from the
Blaualagen und freilebenden marinen Nematoden. oxidation of H2S in seawater. Geochim. Cosmochim.
sterr. bot. Zeitschr. 106:8187. Acta 57:17051718.

Das könnte Ihnen auch gefallen