Bacteria in molluscs: good and bad guys

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) __________________________...
Author: Shauna Little
1 downloads 2 Views 1MB Size
Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

Bacteria in molluscs: good and bad guys J.L. Romalde*, and J.L. Barja Department of Microbiology and Parasitology, CIBUS-Phaculty of Biology, University of Santiago de Compostela, Campus Sus s/n, 15782 Santiago de Compostela, Spain. *Corresponding author: [email protected] The aquaculture of molluscs is seriously affected worldwide by bacterial pathogens that cause high losses in hatcheries as well as in natural beds. The main responsible for the mortality outbreaks is a number of Vibrio species that are considered important pathogens in aquaculture. The pathologies caused by vibrios in bivalves have been described since the 1960s; however, over recent years, successive episodes of high mortality have been recorded due to these microorganisms. The present work provides an updated overview of main diseases and implicated Vibrio species affecting the different life stages of cultured bivalves. With the introduction of new species as the abalone (gastropod) and octopus (cephalopod), new potential pathogens have also appeared, such as members of the genus Pseudomonas or Serratia, among others. On the other hand, in the last years, special attention has been focussed on the use in hatcheries of bacteria with antimicrobial activity to control the composition of microbiota associated to the mollusc larvae in order to avoid pathogens and improve larvae survival. Different taxa such as Phaeobacter or Pseudoalteromonas have been considered as probiotics (extended definition for aquaculture) and tested in laboratory condition and in the field. Keywords: Bacteria, molluscs, pathogens, Vibrio, probiotics

1. Introduction One of the main problems in aquaculture of molluscs is the repetitive episodes of mortality, which seriously reduce the production. These outbreaks of disease affect larval and post-larval stages in hatcheries, as well as juvenile and adults cultured in natural environment. In the case of hatcheries, the massive mortalities involve the complete loss of the stocks of production, with serious economic consequences. In most of the cases, the studies have demonstrated that the problems are caused by bacterial pathologies, being the main etiological agents members of genus Vibrio [1, 2]. In relation to the stages cultured in natural beds, despite of the initial attention only to the pathologies caused by parasitic protozoa, in the last years special attention is being paid to diseases with bacterial origin, affecting survival of cultures. The lack of systematic and rigorous studies on the bacterial populations associated to the mollusc culture and, therefore, the scarce knowledge on the matter, has led to the search of solutions focussed to the complete elimination of the microbiota in the culture water during hatchery stages. The different methods employed, from water treatments to chemotherapy, have proved to be inadequate to avoid the episodes of mortality. The use of probiotic bacteria is the most promising alternative in aquaculture for the development of cultures, obtaining a well-balanced bacterial population with auto-regulation capacity. Moreover, their use avoid the hazards derived from antibiotics and other control measures. In this chapter, the actual knowledge about this subject is reviewed, detailing the main bacterial pathogens affecting larval and post-larval stages cultured in hatchery, as well as juveniles and adults cultured in natural environment. The characteristic signs of the diseases caused by the different bacterial species are described, as well as their host range and geographical distribution. Besides, the advantages of specific application of probiotics in hatcheries are analyzed, considering the special characteristics of this type of cultures. The need of their utilisation is justified by the disadvantages of the different systems for disease control usually applied, like treatment of water and chemotherapy. The different bacterial species suggested as potential probiotics are considered, as well as the way to select them, the modes of action, and their possible applications in hatcheries.

2. Bad guys 2.1. Pathogenic bacteria for bivalve larvae 2.1.1. Culture in hatchery. General view Hatcheries are nowadays the main source of seed for the aquaculture of many bivalve molluscs with high economic value, such as oysters, clams or scallops. Seed obtention by induced lay has been studied from a technical point of view in order to establish the optimum conditions for its performance. The influence of environmental factors including salinity, temperature, culture density, etc, has been the aim of numerous works. However, there is a lack of knowledge on the pathological problems, probably due to the shortage of rigurous systematic studies.

136

©FORMATEX 2010

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

The episodes of high mortality, which lead to the loss of whole production stocks, are of periodic frequency and usually the responsible agent is not determined [3-13], even though they imply an important economic loss and a lack of seed supply. The optimum conditions for bivalve culture also favour the growth of bacteria and the accumulation of their metabolites [7, 14, 15]. The disease process is favoured in many ocassions by a larval susceptibility increase due to external stress factors, including bad quality of food or water, organic contamination, etc. In addition, these factors also facilitate the growth of potential pathogenic bacteria [6, 16]. Therefore, in many occasions mortalities can be associated to the overgrowth of opportunist pathogens. 2.1.2. Vibriosis In a hatchery, although all larval stages are vulnerable, during the temporary fixation of the larvae on the bottom of the tank they are exposed to a high concentration of potential pathogenic bacteria associated with the tank surface, moribund larvae or organic detritus [17]. Guillard [18] was the first to report evidence of the involvement of a Vibrio sp. in the disruption of the velum and internal tissues of the clam larvae Mercenaria mercenaria, which produced a mortality of 70% of the population. In the study performed by Tubiash et al. [3], they observed “swarms” of bacteria appearing on the margins of the larvae (Fig. 1), which became progressively denser and after 8 hours, mortality occurred as a result of granular necrosis. This study was the first one to describe the term “bacillary necrosis” affecting numerous bivalves: Crassostrea virginica, Ostrea edulis, M. mercenaria, Argopecten irradians and Teredo navalis. Typical signs of “bacillary necrosis” included the extension of the velum, motility reduction or erratic movements in circles that appeared after 4-5 hours of exposure to Vibrio spp. The species V. alginolyticus, V. tubiashii and V. anguillarum were recognised as the main causal agents of the “bacillary necrosis” in later studies [16, 19]. The disease is characterized by bacterial colonization of the mantle, velum disruption, abnormal swimming, visceral atrophy, and lesions in the organs among other signs. Another characteristic sign of larval vibriosis in hatcheries is the appearance of the phenomenon called “spotting”, defined as an accumulation of moribund larvae agglutinated at the bottom of the tanks [6]. Table 1 provides a summary of the most recent studies on bivalve larvae vibriosis.

Fig. 1. Detail of the disease signs in experimentally infected larvae: bacterial “swarming” surrounding the larvae. Arrow indicates the bacterial cells. Bar = 200 µm.

Originally, V. anguillarum, V. alginolyticus, V. tubiashii and V. splendidus were the recognized agents associated to larval vibriosis. However, other species have been described in the last two decades [13, 20, 21]. This is the case of V. pectenicida [21], originally described as a Vibrio sp. strain able to produce a high mortality of the scallop larvae Pecten maximus after 48 hours of exposure due to the release of bacteria toxins which interrupt the digestive transit and degrade the tissues of the larvae [20]. A number of studies have been conducted with this species to reproduce experimentally the infection and to evaluate the pathogenesis and the interactions with haemocytes [2]. Prado et al. [13] investigated 3 Vibrio strains isolated from oyster larvae (O. edulis) in three different hatchery outbreaks in Galicia showing a disease compatible with a “bacillary necrosis”. One strain was identified as V. neptunius, being the first description of this bacterial species as pathogen of oyster larvae. The other two isolates were classified as Vibrio sp., because despite showing similarities to V. orientalis and V. vulnificus they could not be unequivocally assigned to these species. Further characterization studies showed that one of those Vibrio sp. strains constitutes a new species, for which the name V. ostreicida was proposed [unpublished results]. The species V. tubiashii reported as one of the causative agents of the “bacillary necrosis” [16, 19] has recently been described as a re-emergent pathogen in North America causing a decline of approximately 59% in larvae oyster production [22]. In addition, Hasegawa et al. [23] demonstrated that one of the critical factors in the pathogenicity of V. tubiashii in oysters (Crassostrea gigas) is the presence of a quorum sensing-regulated metalloprotease (VtpA), which is widespread among Vibrio species. It has been suggested that agents responsible for the development of disease in bivalve larvae seem to act synergistically and infection occurs as a result of stressed larvae [6, 13, 24]. This highlights the importance of maintaining optimum water quality and culture density in a hatchery. The use of antibiotics may be beneficial, although frequent use may lead to the appearance of bacteria resistant strains [25]. One possible measure against bacterial pathogens cultured bivalves could be the genetic selection of resistant larvae populations at the hatcheries [26].

©FORMATEX 2010

137

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

2.1.3. Other bacterial diseases Members of the genus Pseudomonas have been described in some cases as pathogens for bivalve larvae, being in most cases isolated together with representatives of the genus Vibrio causing problems in larval cultures. Descriptions have been published from 1959 from different geographic areas and diverse mollusc species, including clams (M. mercenaria), oysters (O. edulis, C. virginica), and pectinids (T. gigas, A. purpuratus)[17, 27-29]. Garland et al. [10], studying the mortalities occurred in hatchery of C. gigas, obtained two isolates of the genus Alteromonas, which resulted pathogenic for the larvae. The genus Moraxella has been also occasionally described as pathogenic for T. gigas larvae [17]. Finally, the group Aeromonas – Plesiomonas comprises representatives isolated from moribund larvae of T. gigas [17]. The pathogenicity of an isolate of Aeromonas hydrophila for A. purpuratus larvae has been experimentally demonstrated [30]. 2.2. Diseases of juvenile and adult bivalves 2.2.1. Pathogenic Vibrio species The studies of Tubiash et al. [3] associated for the first time moribund adult bivalves (M. mercenaria, C. virginica, M. edulis and Mya arenaria) with the species V. tubiashii and V. alginolyticus. Further cases of vibriosis in oyster were described by Elston and Leibovitz [8], detecting anomalies in the shell and alterations in the function of the ligament and digestive processes. The classical Vibrio infections are Summer Mortality in juvenile oysters and Brown Ring Disease in adult clams. Table 1 lists recent studies of Vibrio spp. that have caused pathologies in juvenile and adult bivalves. In France, Summer Mortality (SM) affects juvenile populations of the Pacific oyster (C. gigas) during the warmer months when the water temperature is ≥18ºC and reproduction takes place [31]. This phenomenon has been associated to stress situations, low dissolved oxygen or presence of toxic substances in the sediment [32]. Lipp et al. [33] were the first to observe that the oyster haemolymph had high levels of vibrios that were causing death. Later studies identified V. splendidus as the causal agent of SM [34-37]. Garnier et al. [38] analysed the bacterial populations in the haemolymph of moribund oysters by phenotypic and molecular methods and found that the prevalent species were V. aestuarianus (56%), describing the new subspecies V. aestuarianus subsps. francensis, and V. splendidus (25%). The pathogenicity of V. aestuarianus for the oyster was demonstrated by Labreuche et al. [39], that described the immunosuppressive activity of the extracellular products of this species on the haemocyte functions. More recently, Allain et al. [40] suggested the possible role of V. harveyi as aetiological agent of SM, since it was detected in most samples of affected oysters during the 2008 warm season, and was able to produce mortality in experimental challenges. The most accepted theory today is that the oyster SM cannot be attributed to only one bacterial pathogen or to the oyster Herpesvirus, but to a complex interaction between the physiological and/or genetic state of the host, environmental factors and the presence of various opportunistic infectious Vibrio species [2, 39, 41]. Brown Ring Disease (BRD), caused by V. tapetis [42], has been widely studied since it is the main disease with bacterial aetiology in adult clams (R. phillipinarum and R. decussatus) [2, 24]. Susceptibility to V. tapetis infections is species-specific, causing greater physiological disturbances and mortality in R. philippinarum than in other species of clam (R. decussatus and M. mercenaria) or in the oyster C. virginica [43, 44]. The disease is characterized by the alteration of the calcification process on the inner surface of the valves and the appearance of a characteristic brown deposit consisting of conchiolin between the edge of the shell and the pallial line (Fig. 2a) [42, 45]. BRD is considered one of the main limiting factors in the culture of Manila clam (R. philippinarum) in Europe [45-49] and was also recently detected in Manila clams cultured in Korea [50]. Environmental factors (i.e. temperature and salinity) play a role in the development of BRD, which tends to be more frequent in the spring and winter as the optimum growth temperature for V. tapetis is 15ºC [2, 24]. Reid et al. [51] demonstrated in challenges with Manila clams, that the disease was more severe when performed at 20 ppt salinity than at 40 ppt. The diagnostic of BRD is currently based on the examination of the characteristic brown ring on the inner edge of the shell. For an early diagnosis in the absence of the brown ring, specific PCR detection protocols targeting the 16S rRNA gene have been designed [50, 52, 53]. Biochemical, serological and genetic intraspecific variability within V. tapetis had been described as well as three main subgroups that correlate with the host type [54]. This heterogeneity has been further demonstrated with MLSA (multilocus sequence analysis) and 2D-PAGE proteomic analyses [55, 56]. The pathogenic potential of the recently described species Vibrio celticus, a component of the microbiota of cultured clam populations in Galicia (NW Spain), has been confirmed in experimental challenges of juvenile clams [57].

138

©FORMATEX 2010

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

Table 1. Experimental infections of bivalve larvae, juvenile and adults with pathogenic Vibrio spp. *Abbreviations: C = Crassostrea, M = Mytilus, O = Ostrea, P= Pecten, R = Ruditapes, BRD= Brown Ring Disease, ECP= Extracellular products, NS= Not specified. Pathogenic species

Origin of the tested strain/s

Host*

V. alginolyticus

Culture collection

M. galloprovincialis

Larvae

58

Vibrio sp.

Mortality outbreak

C. gigas; O. edulis

Larvae

59

Mortality outbreaks

R. decussatus

Larvae

60

Mortality outbreaks

O. edulis

Larvae

13

Mortality outbreaks

P. maximus

Larvae

61

Vibrio sp.

Pacific oyster

C. virginica

Larvae

62

V. tubiashii

Pacific oyster

C. gigas

Larvae

23

V. tapetis

Symptomatic clams

V. splendidus

Moribund oysters

C. gigas

Juveniles

34

Mortality outbreak

C. gigas

Juveniles

35

C. gigas R. philippinarum

Adults Seed

37

C. gigas

Adults

Mortality outbreak

C. gigas

Adults

66

Moribund/ healthy oysters

C. gigas

Juveniles

36

Pacific oyster

C. virginica

Juveniles

62

V. splendidus biovar II V. alginolyticus V. neptunius; Vibrio sp. V. pectenicida V. splendidus-like

V. splendidus biovar II (V. chagasii) V. splendidus-like

V. aestuarianus V. aestuarianus subsp. francensis Vibrio sp.

Moribund/ healthy oysters

R. philippinarum R. decussatus

Mortality outbreak Diseased/ healthy oysters

Life stage

Adults

Reference

43, 44, 63 64

39, 65 38

2.2.2. Infections by Rickettsia-like organisms (RLO) The presence of intracellular prokaryotes (Fig. 2b) has been described in several marine invertabrates [67-76], being suggested in some occasions an association with mortalities [72]. However, if these infections produce significant alterations in the host has been controversial during decades [71]. Some authors indicated a key role in the appearance of disease and mortality [72, 77-82], whereas other auhtors suggested that these microorganisms do not cause damages to the host cells or cause just a limited pathology [73, 83]. In 1987, a massive mortality (aprox. 40%) of scallop (Pecten maximus) was detected in Britany (Francia). The anatomo-pathological study revealed the presence of a bacterial infection in the gills [77]. Intracellular basophylic bacterial colonies of different sizes were observed, in some cases blocking the blood vessels. Koch’s postulates could not be fulfilled, but the authors suggested that the intensity of infection supported the hypothesis that the functionality of gills was compromised [77]. In further studies, the microorganism was purified and characterized, demonstrating its relatedness with the Rickettsia group, as well as the production of enzimatic activities, such as catalase or acid phosphatase, which could be related with its pathogenicity [84]. Other cases were subsequentely described, also associated with mortalities in clam (Venerupis rhomboides) in Galicia (Spain) [82], and tropical pearl oyster (Pinctada maxima) and Australian oyster (Crassostrea ariakensis) in China [80, 85]. In these cases, the only pathogen observed was an intracellular prokaryote RLO, and therefore proposed as responsible of the mortalities. In clam the microorganism was located in the gills, whereas in oysters a more generalized infection was observed, the microorganism being detected in gills, mantle, and digestive tissues. The currently most accepted hypothesis is that these RLO would be of low virulence and that the degree of the pathological alterations would be related with the infection degree. Then, only massive infections, blocking important physiological functions by mecanic action, would cause a significative pathology

©FORMATEX 2010

139

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

a)

b)

Fig. 2.- BRD in Manila clam (R. philippinarum) caused by V. tapetis (a) and RLO inclusion in gills of carpet-shell clam (R. decussatus)(b).

2.2.3. Juvenile oyster disease (JOD) Juvenile oyster disease (JOD) was detected for the first time in cultured populations of Crassostrea virginica in USA during the 1980s [86]. The disease affects to oyster shorter than 25 mm high, and mortalities may reach 90% of annual production [86, 87]. In affected areas, the disease has annual frequency, when water temperature is ≥ 20ºC [86-88]. The main signs are the growth cessation, irregular edges of the valves and a great bukging of the inferior valve followed a conchiolin deposit in the inner shells [86, 88]. Cohabitation experiments with infected oysters demonstrated the transmissibility of the disease [89, 90], although its etiology was more difficult to evidenced. Its was first suggested that bacteria or parasites occasionally observed in the lesions and/or conchiolin could be the causive agents [91], confirming the bacterial hypothesis some experiments with antibacterial agents which decreased the mortality [92]. Further studies revealed that a new species of α-proteobacteria within the Roseobacter branch [92], Roseovarius crassostreae [93], showing a strong colonization in all the infected oysters. Koch’s postulates could be than fulfilled [94, 95]. In fact, a change of the disease name to “Roseovarius oyster disease (ROD)” was proposed in order to avoid confussion with other diseases of juvenile oysters [95]. 2.2.4. Nocardiosis Around 60 years ago, mortalities in Pacific oyster (C. gigas) were registered in Japan and USA during the summer months [96, 97]. The presence of pathological alterations in the conective tissue around the digestive tract suggested their infectious nature, probably due to a Gram positive bacterium [98]. It was in 1991, when Friedman and Hedrick isolated a bacterium belonging to the genus Nocardia from diseased cultured oysters in Washington (USA) and British Columbia (Canada), naming the disease as Pacific Oyster Nocardiosis (PON) [99]. The phenotypic and genetic characterization of the isolates demonstrated that they constituted a new species for which the name Nocardia crassostreae was proposed [100]. The disease, both in natural and experimental infections, shows practically no external signs. In heavy infections, the presence of greenish/yellowish nodules was observed on most tissues, including adductor muscle, gills, heart and mantle [99]. These nodules are formed by haemocytes and Gram positive acid-resistant bacteria. 2.2.5. Other bacterial diseases Infections by members of the genera Chlamydia and Mycoplasma have been described in a variety of adult bivalve molluscs, such as oyster, scallop, clam, mussel, and cockle (Cerastoderma edule) [67, 76, 79, 101-109]. Most descriptions are based on histological studies observing variable prevalences and, in some cases, pathological alterations in the host tissues related, as in the case of RLO, to the infection intensity [107, 108]. The genus Cytophaga has also been related to lesions in Crassostrea gigas juveniles. Dungan and Elston [110] have described homogeneous bacterial populations associated to the 'resilium' showing a degradation process. The 'resilium' is one of the estructural components of the ligament, responsible of the valves opening, and therefore, the degradation of the 'resilium' occasionate problems in the feeding, breathing, and other essential functions, resulting in a general physiologic weakening. In addition, the mecanic barrier of the valves is lost, being easier the colonization of the animal by pathogenic microorganisms. 2.3. Pathogenic bacteria for other molluscs Abalone are gastropod molluscs of the family Haliotidae that inhabit coastal reefs in tropical, subtropical and temperate areas. In the last years, their culture has gained great importance in different countries throughout the world due to their high economic market value. Several infectious diseases have been reported in Haliotis spp., mostly bacterial and parasitic [111]. Pathogenic Vibrio alginolyticus, V. harveyi and V. parahaemolyticus have been isolated from diseased

140

©FORMATEX 2010

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

abalones in Japan, Taiwan, China, and France [112-118], causing in some cases mass mortlities. The main clinical signs observed in the affected juvenile and adult individuals were abscessing or ulceration in the mantle, white spots on the foot, general whitening, as well as loss of ability to adhere to surfaces [119]. More recently, Shewanella alga and Klebsiella oxytoca have been described as causative agents of mass mortalities in post-larvae abalone in China, being their pathogenicity confirmed in experimental infections [120, 121]. In 2008 during an experiment of acclimatization, mortalities were recorded in wild individuals of abalone (Haliotis tuberculata) captured to constitute a broodstock for hatching. The phylogenetic analysis of the isolates indicated that they constitute a new genus and two new species within the Family Oceanospirillaceae, close to the genera Neptunomonas and Oceanospirillum, being proposed the names Nuadamonas halioticida and N. abalonii. Experimental infections in healthy abalone individuals demonstrated their pathogenic potential for this gastropod mollusc [122]. Octopus is a cephalopod mollusc which has also received great attention in the last years due to its interest for Aquaculture. There are some reports describing infectious agents in different species of octopus (Octopus vulgaris, O. joubini, or O. briareus), mainly from cultured animals but also from wild individuals [123-125]. The main signs observed in diseased octopus are skin ulcers, which can derive in deep wounds in head mantle or arms. Different Vibrio species, including V. alginolyticus, V. anguillarum, V. harveyi or V. parahaemolyticus, Aeromonas cavieae and A. hydrophila, or Pseudomonas stutzeri have been associated with the appearance of these lesions in cultured octopus [123, 124]. In addition, Farto and coworkers [125] isolated the species Vibrio lentus from gill heart and skin lesions of wild octopus, confirming its pathogenic potential.

3. Good guys 3.1.

Classical methods for control of pathogens

The influence of the environment on the hatchery cultures is enormous. It is important to point out that larvae are released in early ontogenic stages which are highly sensitive to infections. In the case of bivalve larvae the influence is even higher since feeding of these animals is through filtration and therefore, a continuous water flow exist through the organisms [126, 127]. Different water treatments, such as filtration, pasteurization, ozone and UV radiation have been employed in mollusc hatcheries with the aim of eliminating potential pathogens [126]. The use of chemotherapeutic agents has been widespread in mollusc hatcheries, although showing inconsistant results [127]. In addition, drug usage present also other disadvantages including the high economic costs and, more important, the potential appearance of resistences and their transference to human pathogens. In fact, laws are becoming more restrictive for the use of antibiotics, being some of them forbbiden for their application to animals for human consumption. The final objective of all the methods cited above is to eliminate the entire water microbiota within the hatcheries, but this objective does not seem to reasonable, since the bacterial population, or at least part of it, may have a benefitial effect on the larval development. In addition, a lack of microbiota can favour the colonization of the system by nondesirable microrganisms, due to a lack of natural competitors, in an environment with a regular input of organic matter and optimal physical conditions. Vaccination, an alternative with excellent results in the fish culture [128, 129] is not applicable to molluscs, since these organisms do not possess a real immunologic system. Taking into acount all these facts, it seems clear that different approaches for the control of molluscan bacterial diseases are needed, especially at larvar stages. 3.2.

Probiotics

Verschuere et al. [126] gave a definition of probiotic appropriated to be used within the field of Aquaculture, taking into account the close interaction of the animal with the environment. According to these authors a probiotic would be a live microbial additive with a beneficial effect on the host, modifying the microbiota associated with the host or the environment, ensuring an optimal use of the feed or improving its nutritional value, improving the host response against the disease, or getting a better quality of its environment. The desirable characteristics of a probiotic bacterium for aquaculture would include: i) to be isolated from the environment where it will be employed, which means a marine bacterium able to grow in the hatchery environment avoiding, in addition, the risks of introduction of alloctonous organisms in the system; ii) to have benecitial effects on the target host; and iii) to be non pathogenic or toxic, not only for the target host, but also for other live organisms present in the system, such as phytoplankton or other live food. The majority of studies published on the use of probiotics in aquaculture correspond to works on fish and crustaceans, being scarce those focussed on molluscs. Lodeiros et al. [130] first published a study on the effect of antibiotic-producing marine bacteria on the larval survival of the scallop Pecten ziczac. Riquelme and coworkers developed deeper studies on the application of a probiotic bacterium Pseudoalteromonas haloplanktis (formerly Alteromonas haloplanktis) to Argopecten purpuratus larvae [131], determining its spectrum of activity against different bacterial species, including members of the genus Vibrio. They performed also a preliminary characterization of the

©FORMATEX 2010

141

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

active compound concluding that it probably is an intracellular component, produced as secondary metabolite, that is secreted during the stationary phase of growth. In vivo experiments with larvae confirmed the protection conferred by the probiotic bacterium to larvae infected by a pathogenic V. anguillarum strain. The authors suggested its possible use as profilactic meassure against the oportunistic pathogens in aquaculture. Further works of the same group [132] identified Pseudomonas sp. and Vibrio sp. strains with a strong antibacterial activity, and studied the use of axenic microalgal cultures (Isochrysis galbana) as a way of incorporating the active bacteria to the culture [133]. Finally, they also studied the incorporation of the probiotic bacteria to massive larval cultures [134], inoculating periodically a mixture of antibiotic-producing strains to the system and demonstrating that the larval phase could be completed without adding commercial chemotherapeutic agents. Such treatment managed a modification of the microbiota associated to the larvae, eliminating potential pathogens. Other bacteria with probiotic potential, including isolates of Aeromonas media and Pseudoalteromonas sp., were described by Gibson et al. [135] and Longeon et al. [136] to be applied in larval cultures of C. gigas and P. maximus. In the last years, different isolates of Phaeobacter gallaeciensis (formerly Roseobacter gallaeciensis) have received special attention by different research groups, due to their great spectrum of inhibition against pathogenic bacteria from aquaculture systems [127, 137, 138]. Prado et al. [138] in experiments performed in marine water, with phytoplankton cultures and with larval oyster (O. edulis) and clam (R. philippinarum) cultures confirmed the results obtained in vitro on the inhibition of pathogenic vibrios by the P. gallaeciensis isolate. In both cases the same pattern of activity was observed, being necesary the introduction of the probiotic isolates in the system previously or simultaneously to the pathogens. These results indicate the potential use of P. gallaeciensis as control method in mollusc hatcheries, if its action is allowed before the pathogens reach high concentrations in the system.

Fig. 3.- Detection of in vitro antibacterial activity of a probiotic isolate of Phaeobacter gallaeciensis.

Fixation, irreversible process of adherence to the substrate at the end of the larval phase, is a critical moment during the larval development when important mortalities are usually observed. It is followed by the metamorphosis, when adult structures and bentonic life are acquired. The implication of bacteria in the processes of fixationa and metamorphosis od marine invertebrates is well known. In most cases, an induction mediated by bacterial biofilms was observed. Weiner and Colwell [139] demsontrated that biofilms of a pigmented marine bacterium stimulate the larval fixation in C. virginica. The bacterium also produces a hydrosoluble exopolysaccharide [140], PAVE (polysaccharide adhesive viscous exopolymer)-like, promoting the adhesion of microorganisms to surfaces [141]. Further characterization of the bacterium identified it as Alteromonas colwelliana [142], later reclassified as Shewanella colwelliana [143]. Other works demonstrated that the biofilms of Sh. colwelliana could also induce the fixation of larvae of Ostrea edulis, but not of Pecten maximus [144]. Sh. colwelliana was successfully employed to induce the fixation of C. gigas and C. virginica larvae in hatcheries [145].

4. Concluding remarks In this work, a revision of the current knowledge on molluscan bacterial pathology has been performed. It is expected that in a near future, with the diversification of the cultured species of molluscs and the incorporation of the intensive culture of these organisms to new geographic areas, new bacterial species with pathogenic potential emerge. The rapid identification of the new pathogens will be very helpful for their control which has to be based, due to the special characteristics of the culture of juvenile and adult molluscs, in the establishment of adequate preventive measures and the limitation of the movement of affected individuals. The development of molecular methods for diagnosis will be crucial to achieve such objectives. In addition, it will also allow the correct etiology of the diseases, as well as the study of the infection routes and the mode of action of the pathogens. In the hatcheries, there is a constant entry of bacteria through the different compartments of the system (broodstock, water, phytoplankton, larvae, tanks, etc), which imply multiple routes of incorporation of potential pathogens. The seawater, where the larval cultures are maintained, determine their microbiota. Pathogen colonization can occur through any part of the organism, including the shell, and spread fast and easily. The best way to avoid this fact is the control of the environmental (water) microbiota to maintain a composition beneficial to the larval development and survival, or at

142

©FORMATEX 2010

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

least, to hamper the proliferation of opportunistic pathogens. Therefore, more studies on the practical utilization of probiosis in mollusc larval cultures are needed in order to know the interaction of probiotics with other live organisms present in the system, to establish their capacity to grow and the effective doses, and to get a method for their conservation and handling in the hatcheries. Their generalized application will allow not only a decrease in the use of antibiotics, with the subsequent beneficial effects on environment and human health, but also a more sustainable aquaculture. Acknowledgements. The support by Spanish Ministry of Science and Innovation and Xunta de Galicia (Spain) is gratefully appreciated.

References [1] Beaz-Hidalgo R, Balboa S, Romalde JL, Figueras MJ.. Diversity and pathogenicity of Vibrio species in cultured bivalve molluscs. Environ Microbiol Reports 2010;2: 34-43. [2] Paillard C, Le Roux F, Borrego JJ. Bacterial disease in marine bivalves, a review of recent studies: Trends and evolution. Aquat Living Resour 2004;17: 477-498. [3] Tubiash HS, Chanley PE, Leifson E. Bacillary necrosis, a disease of larval and juvenile bivalve mollusks. I. Etiology and epizootiology. J Bacteriol 1965; 90:1036-1044. [4] Tubiash HS. Bacterial pathogens associated with cultured bivalve mollusk larvae. In: Smith WL, Chanley CMH, eds. Culture of Marine Invertebrate Animals. New York, NY: Plenum Press; 1975:61-71. [5] Brown C, Losee E. Observations on natural and induced epizootics of vibriosis in Crassostrea virginica larvae. J Invertebr Pathol;1978;31: 41-47. [6] Di Salvo LH, Blecka J, Zebal R. Vibrio anguillarum and larval mortality in a California coastal shellfish hatchery. Appl Environ Microbiol 1978;35: 219-221. [7] Prieur D, Carval YJP. Bacteriological and physico-chemical analysis in a bivalve hatchery: Techniques and preliminary results. Aquaculture 1979;17: 359-374. [8] Elston R, Leibovitz L. Pathogenesis of experimental vibriosis in larval American oysters, Crassostrea virginica. Can J Fish Aquat Sci 1980;37: 964-978. [9] Brown C. A study of two shellfish-pathogenic Vibrio strains isolated from a Long Island hatchery during a recent outbreak of disease. J Shellfish Res 1981;1: 83-87. [10] Garland CD, Nash GV, Sumner CE, Mcmeekin TA. Bacterial pathogens of oyster larvae (Crassostrea gigas) in a Tasmanian hatchery. Aust. J. Mar. Fresw. Res. 1983;34: 483-487. [11] Elston R. Prevention and management of infectious diseases in intensive mollusc husbandry. J World Maricul Soc 1984;15: 284-300. [12] Sugumar G, Nakai T, Hirata Y, Matsubara D, Muroga K. Vibrio splendidus biovar II as the causative agent of bacillary necrosis of Japanese oyster Crassostrea gigas larvae. Dis Aquat Org 1998;33: 111-118. [13] Prado S, Romalde JL, Montes J, Barja JL. Pathogenic bacteria isolated from disease outbreaks in shellfish hatcheries. First description of Vibrio neptunius as an oyster pathogen. Dis Aquat Org 2005;67: 209-215. [14] Murchelano RA, Brown C, Bishop C. Quantitative and qualitative studies of bacteria isolated from sea water used in the laboratory culture of the American oyster, Crassostrea virginica. J Fish Res Board Can 1975;32: 739-745. [15] Brown C, Tettelbach LP. Characterization of a nonmotile Vibrio sp. pathogenic to larvae of Mercenaria mercenaria and Crassostrea virginica. Aquaculture 1988;74: 195-204. [16] Tubiash HS, Otto SV. Bacterial problems in oysters. A review. In: Vivarès CP, Bonami JR, Jasper E, eds. Pathology in Marine Aquaculture. Bredene, Belgium, European Aquaculture Society, Spec. Publ. 1986;9: 233-242. [17] Sutton DC, Garrik R. Bacterial disease of cultured giant clam Tridacna gigas larvae. Dis Aquat Org 1993;16: 47-53. [18] Guillard RRL. Further evidence of the destruction of bivalve larvae by bacteria. Biol Bull 1959;117: 258-266. [19] Tubiash HS, Colwell RR, Sakazaki R. Marine vibrios associated with bacillary necrosis, a disease of larval and juvenile bivalve molluscs. J Bacteriol 1970;103: 272-273. [20] Nicolas JL, Corre S, Gauthier G, Robert R, Ansquer D. Bacterial problems associated with scallop Pecten maximus larval culture. Dis Aquat Org 1996;27: 67-76. [21] Lambert C, Nicolas JL, Cilia V, Corre S. Vibrio pectenicida sp. nov., a pathogen of scallop (Pecten maximus) larvae. Int J Syst Bacteriol 1998;48: 481-487. [22] Elston RA, Hasegawa H, Humphrey KL, Polyak I.K, Häse CC. Re-emergence of Vibrio tubiashii in bivalve shellfish aquaculture: severity, environmental drivers, geographic extent and management. Dis Aquat. Org 2008;82: 119-134. [23] Hasegawa H, Lind EJ, Boin MA, Häse C. The extracelular metalloprotease of Vibrio tubiashii is a major virulence factor for pacific oyster (Crassostrea gigas) larvae. Appl Environ Microbiol 2008;74: 4101-4110. [24] Paillard C, Maes P, Oubella R. Brown ring disease in clams. Ann Rev Fish Dis 1994;4: 1–22. [25] Karunasagar I, Pai R, Malathi GR, Karunasagar I. Mass mortality of Penaeus monodon larvae due to antibiotic resustat Vibrio harveyi infection. Aquaculture 1994;128: 203-209. [26] Tanguy A, Bierne N, Saavedra C, Pina B, Bachérre E, Kube M, et al. Increasing genomic information in bivalves through new EST collections in four species: Development of new genetic markers for environmental studies and genome evolution. Gene 2008;408: 27-36. [27] Helm MM, Smith FM. Observations on a bacterial disease in laboratory-cultured larvae of the European flat oyster, Ostrea edulis L. Int. Counc. Explor. Sea, CM 1971 / k:10 1971;9 pp.

©FORMATEX 2010

143

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

[28] Lodeiros C, Bolinches J, Dopazo CP, Toranzo AE. Bacillary necrosis in hatcheries of Ostrea edulis in Spain. Aquaculture 1987;65: 15-29 [29] Riquelme C, Hayashida G, Toranzo AE, Vilches J, Chávez P. Pathogenicity studies on a Vibrio anguillarum-related (VAR) strain causing an epizootic in Argopecten purpuratus larvae cultured in Chile. Dis Aquat Org 1995;22: 135-141. [30] Riquelme C, Hayashida G, Araya R, Uchida A, Satomi M, et al. Isolation of a native bacterial strain from the scallop Argopecten purpuratus with inhibitory effects against pathogenic vibrios. J Shellfish Res 1996;15: 369-374. [31] Berthelin C, Kellner K, Mathieu M. Storage metabolism in the Pacific oyster (Crassostrea gigas) in relation to summer mortalities and reproductive cycle (west coast of France). Comp Biochem Physiol B Biochem Mol Bio 2000;125: 359-169. [32] Cheney DP, McDonald BF, Elston RA. Summer mortality of Pacific oysters, Crassostrea gigas (Thunberg): initial findings on multiple environmental stressors in Puget Sound, Washington, 1998. J Shellfish Res 2000;19: 353-380. [33] Lipp PR, Brown B, Liston J, Chew K. Recent findings on the summer diseases of Pacific oysters. Proc Nat Shellfish Assoc 1976;65: 9-10. [34] Lacoste A, Jalabert F, Malham S, Cueff A, Gelebart F, Cordevant C, et al. A Vibrio splendidus strain is associated with summer mortality of juvenile oysters Crassostrea gigas in the Bay of Morlaix (North Brittany, France). Dis Aquat Org 2001;46: 139145. [35] Waechter M, Le Roux F, Nicolas JL, Marissal E, Berthe F. Characterization of pathogenic bacteria of the cupped oyster Crassostrea gigas. C R Biol 2002;325: 231 [36] Gay M, Berthe FCJ, Le Roux F. Screening of Vibrio isolates to develop an experimental infection model in the Pacific oyster Crassostrea gigas. Dis Aquat Org 2004;59: 49-56. [37] Gay M, Renault T, Pons AM, Le Roux F. Two Vibrio splendidus related strains collaborate to kill Crassostrea gigas: taxonomy and host alterations. Dis Aquat Org 2004;62: 65-74. [38] Garnier M, Labreuche Y, García C, Robert M, Nicolas JL. Evidence for the involvement of pathogenic bacteria in summer mortalities of the Pacific oyster Crassostrea gigas. Microb Ecol 2007;53: 187-196. [39] Labreuche Y, Soudant P, Goncalves M, Lambert C, Nicolas JL. Effects of extracellular products from the pathogenic Vibrio aestuarianus strain 01/32 on lethality and cellular immune responses of the oyster Crassostrea gigas. Dev Comp Immunol 2006;30: 367–379. [40] Allain G, Arzul I, Chollet B, Cobret L, de Decker S, Faury N, Ferrand S, et al. Summer mortality outbreaks of French Pacific oysters, Crassostrea gigas in 2008: research and detection of pathogens. 14th EAFP International Conference. Diseases of Fish and Shellfish. Prague (Czech Republic). 2009;pp. 127. [41] Pruzzo C, Gallo G, Canesi L. Persistence of vibrios in marine bivalves: the role of interactions with haemolymph components. Environ Microbiol 2005;6: 761-772. [42] Borrego JJ, Castro D, Luque A, Paillard C, Maes P, García M, Ventosa A. Vibrio tapetis sp. nov., the causative agent of the brown ring disease affecting cultured clams. Int J Syst Bacteriol 1996;46: 480-484. [43] Allam B, Ashton-Alcox A, Ford SE. Hemocyte activities associated with resistance to brown ring disease in Ruditapes spp. clams. Dev Comp Immunol 2001;25: 365-375. [44] Allam B, Paillard C, Auffre M, Ford SE. Effects of the pathogenic Vibrio tapetis on defence factors of susceptible and nonsusceptible bivalves species: II. Cellular and biochemical changes following in vivo challenges. Fish Shellfish Immunol 2006;20: 384-397. [45] Paillard C, Maes P. Étiologie de la maladie de l’anneau brun chez Tapes philippinarum: Pathogenicité d’un Vibrio sp. CR Acad Sci Paris 1990 ;310: 15–20. [46] Figueras A, Robledo JAF, Novoa B. Brown ring disease and parasites in clams (Ruditapes decussatus and Ruditapes philippinarum) from Spain and Portugal. J Shellfish Res 1996;15: 363-368. [47] Allam B, Paillard C, Howard A, Le Pennec M. Isolation of the pathogen Vibrio tapetis and defence parameters in brown ring diseased Manila clams Ruditapes philippinarum cultivated in England. Dis Aquat Org 2000;41:105–113. [48] Drummond LC, Balboa S, Beaz R, Mulcahy MF, Barja JL, Culloty SC, Romalde JL. The susceptibility of Irish-grown and Galician-grown Manila clams, Ruditapes philippinarum, to Vibrio tapetis and brown ring disease. J Invertebr Pathol 2007;95:1–8. [49] Paillard C, Korsnes K, Le Chevalier P, Le Boulay C, Harkestad L, Eriksen AG, et al., Vibrio tapetis-like strain isolated from introduced Manila clams Ruditapes philippinarum showing symptoms of brown ring disease in Norway. Dis Aquat Org 2009;81: 153-161. [50] Park KI, Paillard C, Le Chevalier P, Choi KS. Report on the occurrence of brown ring disease (BRD) in Manila clam, Ruditapes philippinarum, on the west coast of Korea. Aquaculture 2006;200: 610-613. [51] Reid HI, Soudant P, Lambert C, Paillard C, Birkbeck TH. Salinity effects on immune parameters of Ruditapes philippinarum challenged with Vibrio tapetis. Dis Aquat Org 2003;56: 249–258. [52] Paillard C, Gausson S, Nicolas JL, Le Pennec JP, Haras, D. Molecular identification of Vibrio tapetis, the causative agent of the BRD of Ruditapes philippinarum. Aquaculture 2006;253: 25-38. [53] Romalde JL, Rodríguez JM, Borrego JJ. Protocolo de PCR (reacción en cadena de la polimerasa) para detección e identificación de Vibrio tapetis a partir de cultivos puros del microorganismo y de tejidos de moluscos. Patent ES 2 265 707. Oficina Española de Patentes y Marcas. 2007. [54] Rodríguez JM, López-Romalde S, Beaz R, Alonso C, Castro D, Romalde JL. Molecular fingerprinting of Vibrio tapetis strains using three PCR-based methods: ERIC-PCR, REP-PCR and RAPD. Dis Aquat Org 2006;69: 175-183. [55] Balboa S, Bermúdez-Crespo J, Gianzo C, Doce A, López JL, Romalde JL. Proteomic analysis of Vibrio tapetis, the etiological agent of Brown Ring Disease. Vibrio Congress 2007. [56] Balboa S, Doce A, Diéguez AL, Romalde JL. Estudio de la variabilidad intraespecífica del patógeno de almeja Vibrio tapetis mediante MLSA. VII Reunión del grupo de Microbiología del Medio Acuático (S.E.M.). Bilbao, Spain. 2008 [57] Beaz-Hidalgo R, Diéguez AL, Cleenwerck I, Balboa S, Doce A, de Vos P, Romalde JL. Vibrio celticus sp. nov., a new Vibrio species belonging to the Splendidus clade with pathogenic potential for clams. Syst Appl Microbiol 2010; in press.

144

©FORMATEX 2010

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

[58] Anguiano-Beltrán C, Lizárraga-Partida ML, Searcy-Bernal R. Effect of Vibrio alginolyticus on larval survival of the blue mussel Mytilus galloprovincialis. Dis Aquat Org 2004;59: 119-123. [59] Estes RM, Friedman CS, Elston RA, Herwig RP. Pathogenicity testing of shellfish hatchery bacterial isolates on Pacific oyster Crassostrea gigas larvae. Dis Aquat Org 2004;58: 223-230. [60] Gómez-León J, Villamil L, Lemos ML, Novoa B. Isolation of Vibrio alginolyticus and Vibrio splendidus from aquacultured carpet shell clam (Ruditapes decussatus) larvae associated with mass mortalities. Appl Environ Microbiol 2005;71: 98-104. [61] Sandlund N, Torkildsen L, Maqnesen T, Mortesen S, Bergh O. Immunohistochemistry of great scallop Pecten maximus larvae experimentally challenged with pathogenic bacteria. Dis Aquat Org 2006;69: 163-173. [62] Gómez- León J, Villamil L, Salger SA, Sallum RH, Remacha-Triviño A, Leavitt DF, Gómez-Chiarri M. Survival of eastern oysters Crassostrea virginica from three lines following experimental challenge with bacterial pathogens. Dis Aquat Org 2008;79: 95-105. [63] Allam B, Paillard C, Ford SE. Pathogenicity of Vibrio tapetis, the etiological agent of brown ring disease in clams. Dis Aquat Org 2002;48: 221-231. [64] Allam B, Ford SE. Effects of the pathogenic Vibrio tapetis on defence factors of susceptible and non-susceptible bivalve species: I. Haemocyte changes following in vitro challenge. Fish Shellfish Immunol 2006;20: 374-383. [65] Labreuche Y, Lambert C, Soudant P, Boulo V, Huvet A, Nicolas JL. Cellular and molecular hemocyte responses of the Pacific oyster, Crassostrea gigas, following bacterial infection with Vibrio aestuarianus strain 01/32. Microbes Infect 2006 ;8: 27152724. [66] Garnier M, Laubreche Y, Nicolas JL. Molecular and phenotypic characterization of Vibrio aestuarianus subsp. francensis subsp. nov., a pathogen of the oyster Crassostrea gigas. Syst Appl Microbiol 2008;31: 358-365. [67] Harshbarger JC, Chang SC. Chlamydiae (with phages), mycoplasmas, and rickettsiae in Chesapeake Bay bivalves. Science 1977;196: 666-668. [68] Buchanan JS. Cytological studies on a new species of rickettsia found in association with a phage in the digestive gland of the marine bivalve mollusc, Tellina tenuis (da Costa). J Fish Dis 1978;1: 27-43. [69] Otto SV, Harshbarger JC, Chang SC. Status of selected unicellular eucaryote pathogens, and prevalence and histopathology of inclusions containing obligate prokaryote parasite, in commercial bivalve mollusks from Maryland estuaries. Haliotis 1979;8: 285-295. [70] Meyers TR. Endemic diseases of cultured shellfish of Long Island, New York: Adult and juvenile American oysters (Crassostrea virginica) and hard clams (Mercenaria mercenaria). Aquaculture 1981;22: 305-330. [71] Comps M. Recherches histologiques et cytologiques sur les infections intracellulaires des mollusques bivalves marins. Thèse de Doctorat d’Etat, Université des Sciences et Téchniques du Languedoc, Montpellier. 1983. [72] Gulka G, Chang PW, Marti KA. Prokaryotic infection associated with a mass mortality of the sea scallop, Placopecten magellanicus. J Fish Dis 1983;6: 355-364. [73] Morrison C, Shum G. Chlamydia-like organisms in the digestive diverticula of the bay scallop, Argopecten irradians (Lmk). J Fish Dis 1983;6: 537-541. [74] Elston R A, Peacock MG. A rickettsiales-like infection in the Pacific razor clam, Siliqua patula. J Invertebr Pathol 1984;47: 93104. [75] Mialhe E, Chagot D, Boulo V, Comps M, Ruano F, et al. An infection of Ruditapes decussatus (Bivalvia) by Rickettsia. Aquaculture 1987;67: 258-259. [76] Hine PM, Diggles BK. Prokaryote infections in the New Zealand scallops Pecten novaezelandiae and Chlamys delicatula. Dis Aquat Org 2002;50: 137-144. [77] Le Gall G, Chagot D, Mialhe E, Grizel H. Branchial rickettsiales-like infection associated with a mass mortality of sea scallop Pecten maximus. Dis Aquat Org 1988;10: 139-145. [78] Norton JH, Shepherd MA, Abdon-Naguit MR, Lindsay S. Mortalities in the Giant clam Hippopus hippopus associated with Rickettsiales-like organisms. J Invertebr Pathol 1993;62: 207-209. [79] Norton JH, Shepherd MA, Prior HC. Intracellular bacteria associated with winter mortality in juvenile giant clam, Tridacna gigas. J Invertebr Pathol 1993;62: 204-206. [80] Wu XZ, Pan JP. Studies on Rickettsia-like organism disease of the Tropical marine pearl oyster. I: The fine structure and morphogenesis of Pinctada maxima pathogen Rickettsia-like organism. J. Invertebr. Pathol. 1999;73: 162-172. [81] Wu XZ, Pan JP. An intracellular prokaryotic microorganism associated with lesions in the oyster Crassostrea ariakensis Gould. J Fish Dis 2000;23: 409-414. [82] Villalba A, Carballal MJ, López C, Cabada A, Corral L, et al. Branchial rickettsia-like infection associated with clam Venerupis rhomboides mortality. Dis Aquat Org 1999;36: 53-60. [83] Fries CR, Grant DM. Rickettsiae in gill epithelial cells of the hard clam Mercenaria mercenaria. J Invertebr Pathol 1991;57: 166-171. [84] Le Gall G, Mialhe E. Purification of Rickettsiales-like organisms associated with Pecten maximus (Mollusca: Bivalvia): serological and biochemical characterization. Dis Aquat Org 1992;12: 215-220. [85] Sun J, Wu X. Histology, ultrastructure, and morphogenesis of a Rickettsia-like organism causing disease in the oyster, Crassostrea ariakensis Gould. J Invertebr Pathol 2004;86: 77-86. [86] Bricelj VM, Ford SE, Borrero FJ, Perkins FO, Rivara G, et al. Unexplained mortalities of hatchery-reared juvenile oysters, Crassostrea virginica (Gmelin). J Shellfish Res 1992;11: 331-347. [87] Davis CV, Barber BJ. Size-dependent mortality in hatchery-reared populations of oysters, Crassostrea virginica, Gmelin 1791, affected by juvenile oyster disease. J Shellfish Res 1994;13: 137–142. [88] Ford SE, Borrero FJ. Epizootiology and pathology of juvenile oyster disease in the easter oyster, Crassostrea virginica. J Invertebr Pathol 2001;78: 141-154. [89] Lewis EJ, Farley CA, Small EB, Baya AM. A synopsis of juvenile oyster disease (JOD) experimental studies in Crassostrea virginica. Aquat Living Resour 1996;9: 169-178.

©FORMATEX 2010

145

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

[90] Paillard C, Ashton-Alcox KA, Ford SE. Changes in bacterial densities and hemocyte parameters in Eastern oysters, Crassostrea virginica, affected by juvenile oyster disease. Aquat Living Resour 1996;9: 145-158. [91] Boardman CL. Host-pathogen interactions between Eastern oysters (Crassostrea virginica) and the bacterial agent of juvenile oyster disease (Roseovarius crassostreae). Master’s thesis, Dept. Biochem. Microbiol. Mol. Biol. University of Maine, Orono, Maine. 2005. [92] Boettcher KJ, Barber BJ, Singer JT. Use of antibacterial agents to elucidate the etiology of juvenile oyster disease (JOD) in Crassostrea virginica and numerical dominance of an -Proteobacterium in JOD-affected animals. Appl Environ Microbiol 1999;65: 2534–2539. [93] Boettcher KJ, Geaghan KK, Maloy AP, Barber JB. Roseovarius crassostreae sp. nov., a member of the Roseobacter clade and the apparent cause of juvenile oyster disease (JOD) in cultured Eastern oysters. Int J Syst Evol Microbiol 2005;55: 1531-1537. [94] Boettcher KJ, Barber BJ, Singer JT. Additional evidence that juvenile oyster disease is caused by a member of the Roseobacter group and colonization of nonaffected animals by Stappia stellulata-like strains. Appl Environ Microbiol 2000;66: 3924–3930. [95] Maloy AP, Ford SE, Karney RC, Boettcher KJ. Roseovarius crassostreae, the etiological agent of Juvenile Oyster Disease (now to be known as Roseovarius Oyster Disease) in Crassostrea virginica. Aquaculture 2007;269: 71-83. [96] Imai T, Numachi K-I, Oizimu J, Sato S. Studies on the mass mortality of the oyster in Matsushima Bay. II. Search for the cause of mass mortality and the possibility to prevent it by transplantation experiment. Bull Tohoku Reg Fish Res Lab 1965;25: 25-38. [97] Perdue JA, Beattie JH, Chew KK. Some relationships between gametogenic cycle and summer mortality phenomenon in the Pacific oyster (Crassostrea gigas) in Washington State. J Shellfish Res 1981;1: 9-16. [98] Numachi K-I, Oizimu J, Sato S, Imai T. Studies on the mass mortality of the oyster in Matsushima Bay. III. The pathological changes of the oyster caused by a gram-positive bacteria and the frequency of their infection. Bull Tohoku Reg Fish Res Lab 1965;25: 39-47. [99] Friedman CS, Hedrick RP. Pacific oyster nocardiosis: Isolation of the bacterium and induction of laboratory infections. J Invertebr Pathol 1991;57: 109-120. [100] Friedman CS, Beaman BL, Chun J, Goodfellow M, Gee A, et al. Nocardia crassostreae sp. nov., the casual agent of nocardiosis in Pacific oysters. Int J Syst Bacteriol 1998;48: 237-246. [101] Joly JP, Comps M. Étude d’un micro-organisme de tupe chlamydien chez la palourde Ruditapes decussatus L. Rev Trav Inst Pêches Marit 1980;44: 285-287. [102] Bower SM, Meyer GR. Disease of Japanese scallops (Patinopecten yessoensis) caused by an intracellular bacterium. J Shellfish Res 1991;10: 513. [103] Cajaraville MP, Angulo E. Chlamydia-like organisms in digestive and duct cells of mussels from the Basque coast. J Invertebr Pathol 1991;58: 381-386. [104] Bower SM. Diseases of cultured Japanese scallops (Patinopecten yessoensis) in British Columbia, Canada. Aquaculture 1992;107: 201-210. [105] Azevedo C. Occurrence of an unusual branchial mycoplasma-like infection in cockle Cerastoderma edule (Mollusca, Bivalvia). Dis Aquat Org 1993;16: 55-59. [106] Renault T, Cochennec N. Chlamydia-like organisms in ctenidia and mantle cells of the Japanese oyster Crassostrea gigas from the French Atlantic coast. Dis Aquat Org 1995;23: 153-159. [107] Svärdh L. Bacteria, granulocytomas, and trematode metacercariae in the digestive gland of Mytilus edulis: seasonal and interpopulation variation. J Invertebr Pathol 1999;74: 275-280. [108] Ward ME, Shields JD, van Dover CL. Parasitism in species of Bathymodiolus (Bivalvia: Mytilidae) mussels from deep-sea and hydrothermal vents. Dis Aquat Org 2004;62: 1-16. [109] Moss JA, Burreson EM, Cordes JF, Dungan CF, Wang A, et al. Pathogens in Crassostrea ariakensis and other Asian oyster species: implications for non-native oyster introduction to Chesapeake Bay. Dis Aquat Org 2007;77: 207-223. [110] Dungan CF, Elston RA. Histopathological and ultrastructural characteristics of bacterial destruction of the hinge ligaments of cultured juvenile Pacific oysters, Crassostrea gigas. Aquaculture 1988;72: 1-14. [111] Bower S. Infectious diseases of abalone (Haliotis spp.) and risks associated with transplantation. In: Campbell A, ed. Workshop on rebuilding abalone stocks in British Columbia. Can Spec Publ Fish Aquatic Sci 2000;130, Canada. pp: 111-122. [112] Nishimori E, Hasgawa O, Numata T and Wakabayashi H. Vibrio carchariae (harveyi) causes mass mortalities in Japanese abalone, Suculus diversicolor suprasexta. Fish Pathol 1998;33: 495-502. [113] Liu PC, Chen YC, Huang CY and Lee KK. Virulence of Vibrio parahaemolyticus isolated from cultured small abalone Haliotis diversicolor suprasexta, with whitering syndrome. Lett Appl Microbiol 2000;31: 433-437. [114] Liu PC, Chen YC, and Lee KK. Pathogenicity of Vibrio alginolyticus isolated from diseased small abalone, Haliotis diversicolor suprasexta. Microbios 2001;104: 71-77. [115] Huang CY, Liu PC, and Lee KK. Whitering syndrome of the small abalone Haliotis diversicolor suprasexta, is caused by Vibrio parahaemolyticus and associated with thermal induction. Z. Naturforsch 2001;56C: 898-901. [116] Lee KK, Liu PC, Chen YC and Huang CY. The implication of ambient temperature with outbreak of vibriosis in cultured small abalone Haliotis diversicolor suprasexta Lischke. J Therm Biol 2001;26: 585-587. [117] Nicolas JL, Basuyaux O, Mazurie J, Thebault A. Vibrio carchariae (syn. Vibrio harveyi), a pathogen of the abalone Haliotis tuberculata. Dis Aquat Org 2002;50: 35-43. [118] Cai J, Han Y, Wang Z. Isolation of Vibrio parahaemolyticus from abalone (Haliotis diversicolor suprasexta L.) postlarvae associated with mass mortalities. Aquaculture 2006;257: 161-166. [119] Cai J, Li J, Thompson KD, Li C, Han H. Isolation and characterization of pathogenic of Vibrio parahaemolyticus from diseased post-larvae of abalone Haliotis diversicolor suprasexta. J Basic Microbiol 2007;47: 84-86. [120] Cai J, Chan H, Thompson KD, Li C. Isolation and identification of Shewanella alga and its pathogenic effects on postlarvae of abalone Haliotis diversicolor suprasexta. J Fish Dis 2006;29: 505-508. [121] Cai J, Wang Z, Cai C, Zhou Y.. Characterization and identification of virulent Klebsiella oxytoca isolated from abalone (Haliotis diversicolor suprasexta) postlarvae with mass mortality in Fujian, China. J Invertebr Pathol 2008;97: 70-75.

146

©FORMATEX 2010

Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology A. Méndez-Vilas (Ed.) _______________________________________________________________________________________

[122] Doce A, Balboa S, Diéguez AL, Barja JL, Romalde JL.. Estudio taxonómico de aislados asociados a mortalidad en Haliotis tuberculata. XII Reunión del Grupo de Taxonomía Filogenia y Biodiversidad Microbiana (S.E.M.). Tarragona (Spain). 2008, pp: 53 . [123] Hanlon RT, Forsythe JW. Diseases of Mollusca: Cephalopoda. Diseases caused by microorganisms. In: Kinne O, ed. Diseases of marine animals, vol III. Biologische Anstalt Helgoland, Hamburg, Germany, 1990, pp. 21-39. [124] Hanlon, RT, Forsythe JW, Cooper KM, Dinuzzo AR, Folse DS, Kelly MT. Fatal penetrating ulcers in laboratory-reared octopuses. J Invert Pathol 1984;44: 67-83. [125] Farto, R, Armada SP, Montes M, Guisande JA, Pérez MJ, Nieto TP. Vibrio lentus associated with diseased wild octopus (Octopus vulgaris). J Invertebr Pathol 2003;83: 149-156. [126] Verschuere L, Rombaut G, Sorgeloos P, Verstraete W. Probiotic bacteria as biological control agents in aquaculture. Microbiol Mol Biol Rev 2000;64: 655-671. [127] Prado S. Microbiota asociada a criaderos de moluscos: patogénesis y probiosis. Ph.D. Thesis. Universidad de Santiago de Compostela, Spain. 2006 [128] Toranzo AE, Santos Y, Barja JL. Immunization with bacterial antigens: Vibrio infections In: Gudding R, Lillehaug A, Midtlyng PJ, Brown F, eds. Fish Vaccinology. Karger, Basilea, Suiza. 1996, pp. 93-105. [129] Romalde JL, Ravelo C, López-Romalde S, Avendaño-Herrera R, Magariños B, et al. Vaccination strategies to prevent emerging diseases for Spanish aquaculture. In: Midtlyng PJ, ed. Progress in Fish Vaccinology. Karger, Basilea, Suiza. 2003, pp. 85-95 [130] Lodeiros C, Freites L, Fernández E, Vélez A, Bastardo J. Efecto antibiótico de tres bacterias marinas en la supervivencia de larvas de la vieira Pecten ziczac infectadas con el germen Vibrio anguillarum. Bol Inst Oceanogr Venezuela Univ Oriente 1989;28: 165-169. [131] Riquelme C, Hayashida G, Araya R, Uchida A, Satomi M, et al. Isolation of a native bacterial strain from the scallop Argopecten purpuratus with inhibitory effects against pathogenic vibrios. J Shellfish Res 1996;15: 369-374 [132] Riquelme C, Araya R, Vergara N, Rojas A, Guaita M, et al. Potential probiotic strain in the culture of the Chilean scallop Argopecten purpuratus (Lamarck, 1819). Aquaculture 1997;154: 17-26. [133] Avendaño RE, Riquelme CE. Establishment of mixed-culture probiotics and microalgae as food for bivalve larvae. Aquaculture Res 1999;30: 893-900. [134] Riquelme CE, Jorquera MA, Rosas AI, Avendaño RE, Reyes N. Addition of inhibitor-producing bacteria to mass cultures of Argopecten purpuratus larvae (Lamarck, 1819). Aquaculture 2001;192: 111-119. [135] Gibson LF, Woodworth J, George AM. Probiotic activity of Aeromonas media on the Pacific oyster, Crassostrea gigas, when challenged with Vibrio tubiashii. Aquaculture 1998;169: 111-120. [136] Longeon A, Peduzzi J, Barthélemy M, Corre S, Nicolas JL, et al. Purification and partial identification of novel antimicrobial protein from marine bacterium Pseudoalteromonas species strain X153. Mar Biotech 2004;6: 633-641. [137] Ruíz-Ponte C, Samain JF, Sánchez JL, Nicolas JL. The benefit of a Roseobacter species on the survival of the scallop larvae. Mar Biotech 1999;1: 52-59. [138] Prado S, Montes J, Romalde JL, Barja JL. Inhibitory activity of Phaeobacter strains against aquaculture pathogenic bacteria. Int Microbiol 2009 ;12: 107-114. [139] Weiner RM, Colwell RR. Induction of settlement and metamorphosis in Crassostrea virginica by a melanin-synthesizing bacterium. Technical Report Maryland Sea Grant Program, UM-SG-TS-82-05, 1982;pp.1-44. [140] Abu GO, Weiner RM, Bonar DB, Colwell RR. Extracellular polysaccharide production by a marine bacterium. In: Barry S, Houghton DR, eds. Proc. Sixth Int. Biodet. Sympos., Cambrian News LTD. Eng. 1986, pp. 543-549. [141] Zobell CE. The effect of solid surfaces upon bacterial activity. J Bacteriol 1943;46: 39-56. [142] Weiner RM, Walch M, Labare MP, Bonar DB, Colwell RR. Effect of biofilms of the marine bacterium Alteromonas colwelliana on set of the oysters Crassostrea gigas and Crassostrea virginica. J Shellfish Res 1989;8: 117-123. [143] Coyne VE, Pillidge CJ, Sledjeski DD, Hori H, Ortiz-Conde BA, et al. Reclassification of Alteromonas colwelliana to the genus Shewanella by DNA-DNA hybridization, serology and 5S ribosomal RNA sequence data. Syst Appl Microbiol 1989;12: 275279. [144] Tritar S, Prieur D, Weiner R. Effects of bacterial films on the settlement of the oysters, Crassostrea gigas, and Ostrea edulis and the scallop Pecten maximus. J Shellfish Res 1992;11: 325-330. [145] Walch M, Weiner RM, Colwell RR, Coon SL. Use of L-DOPA and soluble bacterial products to improve set of Crassostrea virginica (Gmelin, 1791) and C.gigas (Thunberg, 1793). J Shellfish Res 1999;18: 133-138.

©FORMATEX 2010

147