Uploaded by Sirley R

Bactérias em Esponjas

advertisement
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/263935949
Biotechnological Potential of Sponge-Associated Bacteria
Article in Current Pharmaceutical Biotechnology · July 2014
DOI: 10.2174/1389201015666140711115033 · Source: PubMed
CITATIONS
READS
30
1,188
4 authors:
Juliana Santos Gandelman
Marcia Giambiagi-deMarval
Federal University of Rio de Janeiro
Federal University of Rio de Janeiro
6 PUBLICATIONS 85 CITATIONS
77 PUBLICATIONS 1,276 CITATIONS
SEE PROFILE
SEE PROFILE
Walter M R Oelemann
Marinella S Laport
Federal University of Rio de Janeiro
Federal University of Rio de Janeiro
46 PUBLICATIONS 1,201 CITATIONS
85 PUBLICATIONS 717 CITATIONS
SEE PROFILE
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
Bioprospection of antibiotic and heavy metal resistance genes in two genera isolated from marine invertebrate organisms View project
Evaluation of anti-biofilm activity of marine sponges associated bacteria against Staphylococcus spp. View project
All content following this page was uploaded by Marinella S Laport on 11 August 2014.
The user has requested enhancement of the downloaded file.
Send Orders for Reprints to reprints@benthamscience.net
Current Pharmaceutical Biotechnology, 2014, 15, 143-155
143
Biotechnological Potential of Sponge-Associated Bacteria
Juliana F. Santos-Gandelman, Marcia Giambiagi-deMarval, Walter M.R. Oelemann and
Marinella S. Laport*
Instituto de Microbiologia Paulo de Góes, Universidade Federal do Rio de Janeiro, Av. Carlos Chagas Filho, 373,
Cidade Universitária, 21941-590, Rio de Janeiro, Brazil
Abstract: As sessile and filter-feeding metazoans, marine sponges represent an ecologically important and highly diverse
component of marine benthic communities throughout the world. It has been suggested that marine sponges are hosts to
many microorganisms which can constitute up to 40-60% of its biomass. Recently, sponges have attracted a high interest
from scientific community because two important factors. First there is the fact that sponges have a wide range of
associated bacteria; and, second, they are a rich source of bioactive substances. Since 1950, a number of bioactive substances with various pharmacological functions have been isolated from marine sponges. However, many of these substances were subsequently shown to be actually synthesized by sponge-associated bacteria. Bacteria associated with marine sponges constitute an interesting source of novel bioactive compounds with biotechnological potential such as antimicrobial substances, enzymes and surfactants. In addition, these bacteria may be biofilm forming and can act as bioindicators in bioremediation processes of environmental pollution caused by oil and heavy metals. This review focuses on the
biotechnological applications of these microorganisms.
Keywords: Bioactive substances, bioindicators, biofilm, bioremediation, biosurfactant, sponge-associated bacteria.
INTRODUCTION
As the simplest and most primitive metazoans, marine
sponges are important components of benthic communities
due to their biomass and potential influences on pelagic
processes [1-4]. Marine sponges belong to the phylum Porifera which consists of four classes, Hexactinellida, Calcarea,
Demospongiae, and Homoscleromorpha, the latter of which
was more recently established [5].
The architecture of sponges differs from that of any other
taxon. Sponges are the simplest form of multi-cellular animals. They do not possess typical tissues found in other multicellular animal species and their cells retain some degree of
totipotency and independence [6].
The basic body plan comprises of several different cell
layers (Fig. 1). The outer surface, or pinacoderm, is formed
by epithelial cells known as pinacocytes. Through pores
(ostia) on the sponge surface, these cells also extend along
the interior canals which permeate the sponge. Inside the
sponge, specialized flagellated cells (choanocytes) form a
series of chambers where feeding takes place. In these choanocyte chambers, collectively called the choanoderm, the
movement of the choanocytes’ flagella pumps in water
through the ostia and along the often elaborated aquiferous
systems within the sponge. Choanocytes also filter food particles (including bacteria and microalgae) from the water,
and these are transferred to the mesohyl, an extensive layer
*Address correspondence to this author at the Instituto de Microbiologia
Paulo de Góes - Universidade Federal do Rio de Janeiro, Av. Carlos Chagas
Filho, 373, Cidade Universitária, 21941-590, Rio de Janeiro, Brazil; Tel:
+5521-3938-8344; Fax: +5521-3938-8028;
E-mail: marinella@micro.ufrj.br
18-/14 $58.00+.00
of connective tissue (Fig. 1). In the mesohyl, food particles
are digested via phagocytosis by another group of sponge
cells, the archaeocytes (or amebocytes). These totipotent
cells are capable of differentiating into any of the other
sponge cell types. Mesohyls present in many sponges represent dense communities of microorganisms [6, 7]. The presence of these microorganisms near the amebocytes suggests
that the sponge cells either recognize different types of microorganisms or that the latter have developed mechanisms
to prevent being engulfed. Once filtered in the choanocyte
chambers, the water is expelled through an opening at the
top, the osculum [8].
In its natural habitat, the spatial distribution of Porifera is
strongly influenced by water quality, especially with respect
to its content of particles (organic and mineral), pollutants,
and dissolved organic matter. Therefore, marine sponges are
good indicators of water quality and have been used in environmental monitoring and bioremediation processes [9-11].
Sessile organisms such as sponges and other marine invertebrates, including corals and sea squirts, greatly rely on
the production of chemical compounds as defense against
natural predators, competitors, and invading organisms such
as bacteria, viruses, and eukaryotes. Therefore, marine
sponges have been attracting a particular interest in research,
and a wide variety of natural products with different pharmacological properties was identified and characterized [12].
Pharmaceutical interest in sponges began in 1950 with the
discovery of the nucleosides spongothymidine and spongouridine, which were isolated from the marine sponge
Cryptotethya crypta [13, 14]. These nucleosides formed the
basis for the synthesis of ara-C, the first anticancer agent
© 2014 Bentham Science Publishers
144 Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
Santos-Gandelman et al.
Pinacoderm
Choanocyte
Spicule
Mesohyl
Pinacocyte
Bacteria
Choanocyte chamber
Amebocyte
Secondary metabolites
Porocyte
Sclerocyte
Water flow
Fig. (1). Simplified schematic representation of a sponge. (Left) Cross-section of a sponge showing the mesohyl, pinacoderm and interconnected choanocyte chambers. Water enters the sponge through multiple small pores and exits through a large excurrent pore (arrows).
(Right) A close up of an area surrounding a choanocyte chamber. Flagellated choanocyte cells that capture food particles, which are then
phagocytosed by the mobile amebocytes, line the choanocyte chamber. Many members of the sponge symbiotic microbial community exist
extracellularly in the mesohyl, alongside food particles and potential pathogens. There, microorganisms and sponge cells likely interact for
the purposes of nutrient uptake, metabolite exchange and cell signaling and recognition.
derived of marine organisms, and ara-A, an antiviral compound [15].
Interestingly, in some cases compounds isolated from
sponges appear to be synthesized by their associated microorganisms [7]. Sponges interact with associated microorganisms in several ways. For the sponge, microorganisms may
represent either pathogens, or parasites, or symbionts, or
simply a different source of nutrients [16]. Microorganisms
may comprise 40 to 60% of the biomass of the sponge, with
densities exceeding 109 of cells per gram of sponge tissue
[17-19].
The ecologic and evolutionary importance of spongemicrobe associations is mirrored by their enormous biotechnological potential: marine sponges are among the animal
kingdom's most prolific producers of bioactive metabolites,
and in some cases, the compounds are synthesized by
sponge-associated microorganisms [7, 20-21]. In this review
we survey the discovery of products derived from spongeassociated bacteria with potent in vivo or in vitro activities.
Our objective is to highlight various biotechnological applications of these microorganisms (see also Fig. 2).
pharmaceutical importance of these substances is attributed
to their antimicrobial, anti-inflammatory, antitumor and anticancer properties. Among the large number of bioactive substances that have been isolated from marine sponges, some
bioactive substances have already undergone pre-clinical and
clinical screening as anticancer and anti-inflammatory
agents. Of the various chemical classes of this substances,
polyketides, alkaloids, fatty acids, peptides and terpenes are
the most abundant ones [26]. The identification of structurally similar bioactive substances from phylogenetically unrelated sponges suggests that these bioactive substances are
produced by sponge-associated bacteria, rather than the
sponge itself [7, 21]. This is further confirmed in case of
compounds particularly known to be produced exclusively
by bacteria, such as polyketides and non-ribosomal peptides.
Therefore, several studies have been conducted to investigate
bioactive substance produced by bacteria that are commonly
associated with marine sponges [26-32].
Production of Antimicrobial Substances
In this review, we do not intend to comprehensively review sponge-derived natural products; such reviews belong
in the field of chemistry rather than microbiology per se, and
many reviews dedicated to this topic already exist [12, 20,
33-43]. Therefore, we focus our attention on some relevant
examples of antimicrobial (antibacterial, antiviral, antifungal, antiprotozoal) substances produced by spongeassociated bacteria.
Marine microbial bioactive substances have interesting
biomedical potential, pharmaceutical relevance and diverse
biotechnological applications [22-25]. The biomedical and
In the marine environment, competition for nutrients and
space is a powerful driving force that leads to the evolution
of effective strategies for marine microorganisms to colonize
BIOTECHNOLOGICAL POTENTIAL OF SPONGEASSOCIATED BACTERIA
Biotechnological Potential of Sponge-Associated Bacteria
Production of
145
Biological indicators
Production of enzymes
antimicrobial substances
a.
Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
b.
c.
Organic
matter
urease
Nitrogen
cycle
ammonia
contaminant
Biotechnological potential of sponge-associated bacteria
e.
d.
oil
biofilm
waste water
oil
clean
water
biofilm
oreanic matter
Production of biosurfactants
Biofilm generation
Fig. (2). Examples of the biotechnological potential of sponge-associated bacteria. (a) Production of antimicrobial substances: A number
of substances with various pharmacological functions have been described; (b) Production of enzymes: sponge-associated bacteria produce
hydrolytic enzymes as for example, urease. Hydrolysis of urea by endosymbionts represents a predominant ecological characteristic of
sponges-associated bacteria. Urease converts urea into ammonia that in turn participates in the nitrogen cycle; (c) Biological indicators: the
cultivable bacteria are ideal bioindicators, especially for monitoring marine pollution caused by heavy metals. In addition, they can be used
in bioremediation processes; (d) Production of biosurfactants: studies have indicated a function directed towards the bioremediation of
environments contaminated by oil and/or heavy metals; (e) Biofilm generation: Biofilms have been applied in wastewater remediation
processes.
and grow. Consequently, antimicrobials and other secondary
metabolites produced by microorganisms aim at antagonizing the growth of other microorganisms by affecting their
survival and reproduction [44]. The synthesis of these compounds depends on complex regulatory systems and is generally induced during the stationary phase of microbial
growth [45].
The major groups of microorganisms recognized as possible contributors of pharmacologically relevant secondary
metabolites of sponges include Actinobacteria, α, β, γ, δ Proteobacteria and Firmicutes. Many of the compounds produced by bacteria are not yet characterized. The phylum
Actinobacteria dominates in the production of therapeutic
compound, and is followed by Proteobacteria. The bioactive
potential of Firmicutes and Cyanobacteria is yet to be explored [26]. According to the previous studies, a number of
sponge-associated bacterial genera have been shown to produce antimicrobial substances [26, 27, 30, 41, 46-53, 73],
including Actinomyces, Aeromonas, Bacillus, Corynebacterium, Flavobacter, Micrococcus, Pseudoalteromonas, Pseudomonas, Streptomyces and Vibrio.
Table 1 [53-73] summarizes some examples of antimicrobial substances produced by sponge-associated bacteria.
Several of these substances possess a great potential for drug
development, but so far none resulted in a commercial medication. The most promising antimicrobial substances appear
to be 2-undecyl-4-quinolone, cyclic dipeptides (also known
as diketopiperazines, DKPs), lipopeptides (surfactins, iturins
and fengycins), manzamine A, organohalogens (2,4,4’trichloro-2’-hydroxydiphenylether), phenazine, pyrone I,
rifamycins and thiopeptides. Some of the substances produced by bacteria have been shown to be multifunctional.
Among these, manzamine A produced by an actinomycete
shows antibacterial, anti-malarial, anti-HIV, anti-tumor, insecticidal and anti-inflammatory activities [43] and appears
to hold great promise for the future.
Production of Enzymes
Apart from some exceptions, sponge-associated microbial communities appear to be relatively stable over time and
space [19]. Moreover, these associations are very specific for
the production of particular bioactive compounds. However,
the mechanism acting in compound production between host
and microbial associate is not well understood [26]. It became, clear that gene transfer frequencies are high among
sponge-associated bacteria communities. Gene transfer increases the genomic flexibility within these populations and
thereby facilitate their continuous adaptation to changing
environmental conditions [74]. Therefore, enzymes produced
by these bacteria present high potential for application in
various industrial processes [75]. Table 2 summarizes some
examples of enzymes produced by sponge-associated bacteria that possess potential biotechnological application.
146 Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
Table 1.
Santos-Gandelman et al.
Antimicrobial substances produced by sponge-associated bacteria.
Antimicrobial substances
Activity
Bacteria producing
Sponge
Reference
2,4,4’-trichloro-2’-hydroxy-diphenylether
(triclosan) and Acyl-1-(acyl-6’-mannobiosyl)3-glycerol (lutoside)
Antibacterial
Micrococcus luteus
R-1588-10
Xestospongia sp.
[53]
2-nonyl-4-hydroxyquinoline n-oxide
Antibacterial
Pseudomonas sp. 1537-E7
Homophymia sp.
[54]
2-undecyl-4-quinolone
Antimalarial anti-HIV
Pseudomonas sp. 1537-E7
Homophymia sp.
[54]
3,6-diisopropylpiperazin-2,5-dione
Antimicrobial
Pseudomonas sp. NJ6-3-1
Hymeniacidon perlevis
[55]
Acetic acid,-butylester ethanol, 2-(octyloxy)oxalic acid, Allylnonyl ester, 2-Isopropyl-5methyl-1-heptanol, Butylatedhydroxytoluene,
Cyclohexanecarboxylic acid, Hexyl ester,
Diethyl- phthalate, Pentadecanal- 1-tridecanol
and 9-Octadecenal
Antimicrobial
Nocardiopsis dassonvillei
Dendrilla nigra
[56]
Andrimid
Antibacterial
Vibrio sp. M22-1
Hyatella sp.
[57]
Chitinase
Antigungal
Streptomyces sp. DA11
Craniella australiensis
[58]
Cyclo-(glycyl-L-seryl-L-prolyl-L-glutamyl)
Antibacterial
Ruegeria sp.
Suberites domuncula
[59]
Cyclo-(L-proline-L-methionine)
Antibacterial
Pseudomonas aeruginosa
Isodictya setifera
[60]
Cyclo-(l-pro-l-phe)
Antimicrobial
Alcaligenes faecalis A72
Stelletta tenuis
[61]
Diketopiperazine
Antibacterial
Micrococcus spp.
Tedania ignis
[62]
Iturin
Antifungal
Bacillus subtilis A202
Aplysina aerophoba
[49]
Majusculamide c
Antifungal
Lyngbya majuscula
Ptilocaulis trachys
[63]
Manzamine A
Antimalarial
Micromonospora sp.
Acanthostrongylophora sp.
[64]
N-Hexadecanoic- acid
Nematicide
Nocardiopsis dassonvillei
Dendrilla nigra
[56]
MAD08
MAD08
Pumilacidin containing β-Hydroxy fatty-acid
Antifungal
Bacillus pumilus A586
Aplysina aerophoba
[49]
Pyrone-I
Antibacterial antifungal
Pseudomonas sp. F92S91
Unidentified
[65]
Rifamycin B and Rifamycin SV
Antibacterial
Salinospora sp.
Suberea clavata
[66]
Streptophenazines G and K
Antibacterial
Streptomyces sp. HB202
Halichondria panicea
[67]
Subtilomycin
Antibacterial
Bacillus subtilis MMA7
Haliclona simulans
[68]
Surfactin
Antifungal
Bacillus subtilis A190
Aplysina aerophoba
[49]
Surfactin, Iturin and Fengycin
Antibacterial antifungal
Bacillus subtilis A184
Aplysina aerophoba
[49]
Tetrabromo-diphenyl ethers
Antibacterial
Vibrio sp.
Dysidea sp.
[69]
Tetromycin B
Antiprotozoal
Streptomyces axinellae
Pol001
Axinella polypoides
[70]
Thiopeptide (YM-266183 and YM-266184)
Antibacterial
Bacillus cereus QN03323
Halichondria japonica
[71]
Trisindoline
Antibacterial
Vibrio sp.
Hyrtios altum
[72]
Valinomycin
Antiprotozoal
Streptomyces sp
Unidentified
[30]
Urauchimycin
Antibacterial
Streptomyces sp. Ni-80
Unidentified
[73]
Biotechnological Potential of Sponge-Associated Bacteria
Table 2.
Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
147
Enzymes produced by sponge-associated bacteria with potential biotechnological application.
Enzymes
Bacteria producing
Sponge
Reference
Acetylcholinesterase
Arthrobacter ilicis
Spirastrella sp.
[80]
Agarase
Cytophaga sp.
Halichondria panicea
[79]
Alkaline cellulase
Marinobacter MSI032
Dendrilla nigra
[84]
Alkalophilic amylase
Halobacter MMD047
Unidentified
[83]
Anticholinesterases
Bacillus subtilis M18SP4P
Fasciospongia cavernosa
[86]
Chitinase
Streptomyces sp. DA11
Craniella australiensis
[58]
Dehalogenase
Desulfovibrio
Aplysina aerophoba
[89]
Phospholipase A2 (PLA2
Streptomyces dendra sp. nov. MSI051
Dendrilla nigra
[56]
Psychrophilic alkaline lipase
Pseudomonas sp.
Unidentified
[81]
SGNH hydrolase
Unknown
Hyrtios erecta
[85]
Thermo-stable amylase
Nocardiopsis dassonveille MAD04
Unidentified
[82]
Thermo-tolerant alkaline protease
Roseobacter MMD040
Unidentified
[82]
Urethanase
Micrococcus sp.
Spirastrella sp.
[87]
Phospholipase A2 (PLA2) is an ubiquitous defense enzyme that was first described in snake and bee venoms and
later shown to be widely distributed among both, plant and
animal kingdoms [76]. Previous studies have demonstrated
different levels of PLA2 in a number of marine invertebrates
including Cnidaria [77], Porifera [78] and Echinodermata
[56]. In addition to secondary metabolites, sponge-associated
microorganisms also can produce PLA2. However, there is
only one report on the production of PLA2 by the bacterium
Streptomyces dendra sp. nov. MSI051 which in turn is associated with the sponge Dendrilla nigra. Both, the sponge and
its bacterial symbiont, contain high levels (about 1,032 U/L)
of PLA2. It has been speculated that PLA2 may play a functional role in the protection of the sponge against fouling
[56].
Because marine sponges are in constant contact with
large amounts of water and organic matter, one hypothesis
suggests that some sponge-associated microorganisms produce hydrolytic enzymes to convert this organic matter into
nutrients for the sponge. Some evidence from studies on enzyme production by different microorganisms isolated from
marine sponges support this hypothesis. Bacteria of the genus Cytophaga associated with the sponge Halichondria
panicea were shown to hydrolyze agar [79]. Other bacteria
isolated from six marine sponges (Spirastrella sp., Phyllospongia sp., Ircinia sp., Aaptos sp., Azorica sp. and Axinella sp.) were found to produce amylase, protease, and carboxymethylcellulase [80]. In addition, other studies have
demonstrated that the cultivable endosymbionts associated
with the sponges Fasciospongia cavernosa and Dendrilla
nigra produce valuable enzymes for the production of nutrients, including amylase, cellulase, lipase and protease [8183].
Cellulases are important biocatalysts with many industrial applications, for instance in textile and paper industry,
in laundry detergents, as well as in the grains processing, like
coffee and beans. Apart from cellulases, esterases and lipases
are also important biocatalysts used in organic synthesis.
They have been used in a broad range of industrial applications because their stability, their high activity in organic
solvents, and their enantio-⁄stereoselectivity. Proteases represent one of the largest groups of industrial enzymes and possess a variety of applications ranging from use in detergents,
in leather preparation and in food processing [75].
Enzyme production under in vitro conditions has been
observed in numerous studies, such as the synthesis of alkalophilic amylase by Halobacter MMD047 [83], psychrophilic alkaline lipase by Pseudomonas sp. [81], thermotolerant alkaline protease by Roseobacter MMD040, alkaline
cellulase by Marinobacter MSI032 [84], and thermo-stable
amylase by Nocardiopsis dassonveille MAD04. These enzymes can be obtained from cultured bacteria and possess
potential application in the production of drugs, foods, beverages, detergents, textiles, and in the processing and treatment of contaminated water [82].
Using a metagenomic library constructed from bacteria
associated with a marine sponge Hyrtios erecta, a novel esterase was identified that belongs to the SGNH hydrolase
superfamily of esterases. This esterase shows thermal stability and salt tolerance necessary for its use as an industrial
enzyme [85].
Enzymes produced by sponge-associated bacteria may
also have clinical relevance, for example in the degradation
of the neurotransmitter acetylcholine in the brain, where acetylcholinesterase (AChE) inhibitors or anticholinesterases
decrease the activity of enzyme acetylcholinesterase [86].
These inhibitors play an important pharmacological role in
neurodegenerative diseases like Alzheimer and Parkinson.
Recently, a potent AChE inhibitor was identified [86] that is
148 Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
produced by Bacillus subtilis strain M18SP4P, isolated from
the marine sponge Fasciospongia cavernosa. The bacterium
Arthrobacter ilicis isolated from the sponge Spirastrella sp.
produces the enzyme acetylcholinesterase. This enzyme was
shown to be heat tolerant and its activity is not affected by
the relatively high concentration of the major cations found
in seawater, such as Na+, Ca2+, and Mg2+. The bacterium
Micrococcus sp. is associated with the sponge Spirastrella
sp. and is capable of producing urethanase that could potentially be used to remove the chemical carcinogen urethane
from alcoholic beverages [87].
Selvin and colleagues [88] reported that about 65% of the
microorganisms isolated from marine sponges released the
enzyme urease at an early stage of growth. They assumed
that the sponges were exposed in an area contaminated with
urea which is toxic to microorganisms. Thus, effective hydrolysis of urea by endosymbionts represents a predominant
ecological characteristic of sponges-associated bacteria.
Urease converts urea into ammonia that in turn participates
in the nitrogen cycle. Moreover, ammonia generated by hydrolysis can be used by the sponge in protein synthesis.
These findings suggest that these bacteria play multiple roles
in nutrition, physiology and ecology of the marine sponge
and its environment [88].
Bioremediation
Bioremediation is the process of using microorganisms in
situ or ex situ to clean up contaminated environments. The
metabolic ability of the microorganisms to mineralize or
transform organic contaminants into less harmful substances
can be integrated into natural biogeochemical cycles. Bioremediation is an attempt to accelerate naturally occurring degradation by optimizing the limiting conditions. Bioremediation is nondestructive, treatment- and cost-effective and
therefore represents a logistically favorable clean-up technology [90]. Some common microorganisms which are able
to perform bioremediation in the marine environment are
Pseudomonas, Flavobacterium, Arthrobacter, Azotobacter,
Rhodococcus, and Bacillus [91].
To better illustrate the biotechnological application of
sponge-associated bacteria in the bioremediation process,
this section is subdivided into three topics: biological indicators, production of biosurfactants and biofilm generation.
Santos-Gandelman et al.
arise within a standard time pattern [97]. This sub-lethal exposure can affect physiological functions and behavior of
organisms [98]. Thus, the effects of heavy metal pollution in
the environment could be detected directly in spongeassociated bacteria instead of in situ monitoring of changes
in the sponge host organisms [96].
Free-living marine bacteria are constantly exposed to
environmental contaminants and could preferentially be used
as a model for monitoring of contamination in marine ecosystems. However, the population structure of these microorganisms changes drastically as a consequence of excessive
exposure to contaminants and to external disturbances, such
as the movement of migratory animals [94, 99]. On the other
hand, the cultivation of sponge-associated bacteria would
allow to estimate these effects since sponges have a high
filtration capability with retention of about 80% of the particles, including free-living bacteria [100].
According to a study by Selvin and colleagues [96] and
De and colleagues [101], marine pollution caused by heavy
metals can be monitored by culture of bacteria associated
with sponges. In addition to their use as bioindicators,
sponge-associated bacteria can also be used as actors in
bioremediation processes. Significant reduction of cadmium
in culture supernatants of marine bacteria suggests microbial
absorption of this metal [102], and an absorption rate of
about 110 mg of lead per gram of dry weight has been reported for Pseudomonas aeruginosa [103].
Heavy metals have remarkable influences on mammalian
cells and have the ability to modulate the function of multicellular networks, such as the immune system. Certain metals, such as zinc, manganese, and copper are necessary for
normal physiological processes, whereas others, including
lead, cadmium, chromium and mercury are environmental
pollutants and may adversely affect human health. Among
the several heavy metals related with health risks, mercury is
notable for its wide distribution in the environment and the
wide spectrum of effects that it can induce [104].
The organic form of mercury, methylmercury (MeHg),
acts as a potent neurotoxin [105]. MeHg accumulates in
various levels throughout the aquatic food chain and humans
are exposed primarily through consumption of contaminated
fish.
Numerous toxicological studies examined the bacterial
sensitivity and resistance to metals [93-95]. Generally, the
introduction of heavy metals in the environment can produce
substantial changes in microbial communities and their activities, and bioavailability and bioaccumulation of these
metals in aquatic ecosystems have tremendous importance
worldwide [96].
Several studies have been conducted with the aim of reducing the impact of contamination by heavy metals in various environments. Regarding the removal of heavy metals,
bioremediation processes are more attractive when compared
to physicochemical methods because the overall cost of
bioremediation is generally significantly lower than physicochemical methods and it is highly efficient [106]. The
mechanisms of resistance to heavy metals that evolved in
microorganisms include: precipitation of metals as phosphates, carbonates and/or sulfates; volatilization of metals by
methylation or ethylation; physical exclusion of electronegative components in membranes and in extracellular polymeric compounds; and energy-dependent efflux systems and
intracellular sequestration by low molecular weight cysteinerich proteins [107].
The harmful effects of the accumulation of sub-lethal
amounts of heavy metals are not obvious because they do not
Certain environmental strains of bacteria have developed
resistance mechanisms that are highly specific for mercury.
Biological Indicators
Marine sponges filter a large volume of water and accumulate heavy metals and other contaminants from the surrounding environment. Therefore, bacteria associated with
sponges have been used over a long period as indicators of
contamination in marine ecosystems [92].
Biotechnological Potential of Sponge-Associated Bacteria
The mechanism of bacterial resistance to mercury involves
two enzymes that operate sequentially: an organomercurial
lyase (MerB) that cleaves the carbon-mercury bonds of organomercury compounds, and a mercuric reductase (MerA)
that reduces Hg2+ to Hg0 (volatile) [108]. Under anaerobic
conditions, some bacteria are not affected by mercury, since
it is directly used for production of mercury sulfide (HgS)
which is apparently non-toxic [109].
Mobile genetic elements such as plasmids and transposons may carry multiple genes encoding proteins involved in
metal resistance. Thus, exposure to these agents can select
microorganisms resistant to a variety of heavy metals. These
microorganisms are useful in bioremediation in contaminated environments, like a genetically engineered strain of
Escherichia coli which expresses metallothionein and an
Hg2+ transport system. Metallothioneins are cysteine-rich
proteins that have high affinity to metals [110], and bioaccumulation of toxic metals using genetically engineered microorganisms have been reported in a number of studies
[111-114]. Another example is the extremophile bacterium
Deinococcus radiodurans that has been employed to treat
mixed radioactive wastes and to volatilize mercury [115].
All mercury -resistant strains isolated from either terrestrial or fresh water environments harbor genes with homology to merTn21 [116] or to merTn501 [117]. However, in
mercury -resistant marine strains, the observed frequency
was lower [118-119]. In the study conducted by De and colleagues [101], only 9 of the 11 mercury-resistant marine
strains were positive of merA, suggesting the presence of
other genes involved in mercury resistance.
In Alcaligenes eutrophus CH34, Liesegang and colleagues [120] identified various genes involved in metal resistance, including three genes for mercury resistance, one
for chromium resistance and two for resistance to bivalent
cations. Furthermore, the gene cluster czr in P. aeruginosa
was reported to contribute to the resistance to cadmium and
zinc [121].
Microorganisms resistant to multiple heavy metals are of
great importance because they can be used in bioremediation
of environments contaminated with various metals, as previously reported [101].
Production of Biosurfactants
Biosurfactants are a structurally diverse group of tensioactive substances produced by microorganisms. All biosurfactants are amphiphilic, with a polar (hydrophilic) portion
and a nonpolar (hydrophobic) portion. The hydrophilic group
is composed of mono-, oligo- or polysaccharides, peptides or
proteins, while the hydrophobic portion usually contains
saturated or unsaturated fatty acids or aliphatic alcohols
[122]. Due to their amphiphilic structure, biosurfactants increase the surface area of hydrophobic water-insoluble substances, leading to an increase in the bioavailability of such
substances in water. Furthermore, they change the properties
of the bacterial cell surface. The surface activity of biosurfactants makes them excellent emulsifying agents, dispersants, and defoamers [123]. Compared with their chemically
synthesized equivalents, biosurfactants have many advantages because they are biodegradable, less toxic, and show
higher selectivity and increased foaming capacity. Further-
Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
149
more, biosurfactants can be active under extreme temperature, pH and salinity conditions and may also be produced
from industrial waste and sub-products [124-125].
Several studies have demonstrated the biosurfactant potential of sponge-associated bacteria. Most of these studies
indicated a possible application in the bioremediation of environments contaminated by oil and/or heavy metals [126131].
The potential of biosurfactants for decontamination of
soil containing heavy metals was confirmed by Juwarkar and
colleagues [132]. The authors showed that a dirhamnolipid
biosurfactant produced by Pseudomonas aeruginosa strain
BS2 was able to selectively remove heavy metals from contaminated soil in the following order of efficiency: chrome/
cadmium ≥ copper/lead ≥ nickel [132].
The role of biosurfactants produced by marine bacteria in
the remediation of polyaromatic hydrocarbons has also been
previously reported [125]. Das and colleagues [133] further
investigated the possibility of using biosurfactants produced
by marine bacteria in the removal of heavy metals from solutions. Their study revealed that the biosurfactants tested were
able to bind metal ions, and that the removal percentage of
cadmium and lead was variable depending on the different
concentrations of metals and biosurfactants. The ability of
marine biosurfactants to sequester toxic heavy metals and to
form an insoluble precipitate may be useful in the treatment
of wastewater containing such metals [133].
In 2010, Gnanamani and colleagues [128] found that the
marine strain Bacillus sp. MTCC 5514 reduced the concentration of hexavalent chromium and developed tolerance to
trivalent chromium through the synthesis of biosurfactants
by the extracellular enzyme chromium-reductase. The hypothesis is that trivalent chromium generated by the reduction of hexavalent chromium would be entrapped in micelles
formed by the biosurfactants synthesized by the cell. Thus,
the micelles would prevent microbial cells from being exposed to trivalent chromium, and microbial growth and tolerance to chromium were observed even at increased concentrations of chromium.
In this context, our group observed that six out of 21
mercury-resistant isolates from marine sponges produced
biosurfactants. However, no merA gene was detected in
these strains, suggesting that tolerance to mercury may involve sequestration [134]. Recently, Santos-Gandelman and
colleagues [135] described the potential mercury bioremediation by Bacillus cereus strain Pj1 isolated from the marine
sponge Polymastia janeirensis.B. cereus Pj1 was found to be
resistant to 100 µM HgCl2 and to 10 µM methylmercury. Pj1
was also highly resistant to other heavy metal salts, including
CdCl2 and Pb(NO3)2, either alone or in combination. The
mer operon is located on the bacterial chromosome, and the
volatilization test indicated that the B. cereus Pj1 was able to
reduce Hg2+ to Hg0. This strain demonstrated a potential for
biosurfactant production and presented a higher emulsification activity than synthetic surfactants.
Biosurfactants of the lipopeptide type are also produced
by sponge-associated bacteria. The production of lipopeptide
biosurfactant from sponge-associated actinomycetes Nocardiopsis alba MSA10 has been characterized [136]. Recently,
150 Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
Lawrance and colleagues [137] investigated the production
of a lipopeptide surfactant from the sponge-associated Bacillus licheniformis NIOT-AMKV06, isolated from the Andaman and Nicobar Islands. The purified surfactant showed
excellent emulsification activity with crude oil, kerosene and
diesel. The lipopeptide surfactant was highly stable over a
pH range of 5.0-10.0 and a temperature range of 20-70°C in
the presence of high NaCl concentrations. Besides, the surfactant biosynthesis gene cluster (sfp, sfpO and srfA) from B.
licheniformis NIOT-AMKV06 was heterologously expressed
in Escherichia coli and the production was increased threefold when compared to the original strain. These results confirmed the potential of the surfactant for use in bioremediation of hydrocarbons in a marine environment and for enhanced oil recovery.
Biofilm Generation
The high density of microbial communities associated
with marine sponges prompted several groups to investigate
their ability to form biofilm in the sponge host [27, 134, 138139]. Biofilm, by definition, is composed of bacterial communities surrounded by an extracellular polymeric matrix
and attached to surfaces [140]. The polymeric matrix can
present different structures and functions depending on the
bacterial community and/or environmental conditions involved. The biofilm matrix can physically prevent the penetration of antimicrobial agents, especially those that are hydrophilic and positively charged. In some cases, biofilms are
capable of sequestering cations, metals and toxins. Additionally, biofilms can act to protect against ultraviolet radiation,
changes in pH, osmotic shock and desiccation [141].
Polysaccharides are a major component of the polymeric
matrix, but proteins, lipids and nucleic acids may also be
present in the biofilm. In a study by Whitchurch and colleagues [142], DNA was observed in biofilms formed by P.
aeruginosa. Allesen-Holm and colleagues [143] described
evidence that the DNA was liberated by lysis of a portion of
the bacterial population and its presence depended on quorum sensing.
The construction of a marine biofilm requires a series of
steps that are regulated during the maturation of the structure
[144]. Initially, free bacteria in seawater interact with organic and inorganic particles on a surface to promote an initial adhesion. After primary adhesion, bacteria accumulate in
the biofilm through growth, resulting in bacterial colonies
that can synthesize an extracellular matrix, which in turn can
act as a substrate for the adhesion of more microorganisms,
the so called secondary colonizers. These secondary colonizers can adhere either directly to the primary film, or they
may promote the co-aggregation with other microorganisms
and then adhere to the primary film [145].
Biological processes for treating toxic effluents are better
than chemical and physical methods in terms of their efficiency and economy and researchers became aware of the
potential of biofilm communities for bioremediation processes [144, 146]. Many reports have shown the enhanced
bioremediation capability of biofilm forming isolates [147,
148]. Biofilm-mediated bioremediation has showed high
efficiency and safety, since the cells in a biofilm have a
higher possibility of adaptation and survival (especially
Santos-Gandelman et al.
under stress) because they are protected within the matrix
[149]. In early 1980, Akinson [150] described the use of
biofilms for the treatment of water and sewage. However,
only in recent decades researchers in the field of
bioremediation became interested in biofilm reactors [146].
Chemotaxis and biosurfactant production are physiological properties of the microorganisms in a biofilm that elevate
the degradation levels of hydrophobic compounds [151-152].
Microorganisms that form biofilm and secrete polymers on
the surface of hydrocarbons are very suitable for the
degradation of recalcitrant or slowly degrading compounds,
due to their high biomass and the ability to immobilize compounds by bioaccumulation (enhanced accumulation of
microorganisms under influence), biosorption (sequestration
by interactions with biological matter) and biomineralization
(formation of inorganic materials with complex form by
interactions with microbial metabolic products) [153].
Most of the recent work published has focused on the
adsorption of heavy metals to bacteria or tobiofilm formed
by a single bacterial strain intentionally grown in the laboratory [154-162]. In order to exemplify this statement, two
mercury-resistant strains, Bacillus thuringiensis PW-05
[163] and Pseudomonas putida SP1 [164] isolated from marine environments were shown to posses an interesting potential for bioremediation of mercury. These bacteria
formed biofilm in the presence of high concentrations of
HgCl2 [163, 164].
In relation to sponge-associated bacteria, SantosGandelman and colleagues [134] observed that 71 of 100
strains were able to produce biofilm. Among these bacteria,
which belonged mainly to Proteobacteria and Firmicutes, 21
strains were resistant to HgCl2, (unpublished data). In addition, the biofilm-forming bacterium Bacillus cereus Pj1 isolated from the sponge is a strain that
can potentially be applied in the bioremediation of HgCl2
and MeHg contamination in aquatic environments [135].
These studies suggest that marine sponge-associated bacteria able to produce biofilm may represent an important tool
for remediation of contaminated sites either through reduction or sequestration
Biofilm has high adsorption capacities and low production cost. The utilization of biofilm as an adsorbent of polluting ions is one of the promising technologies for treatment of
contaminated water. Marine bacteria provide a useful source
for bioremediation, and there are many advantages of using
them under extreme conditions. When a bacterium is capable
of forming biofilms, it increases its bioremediation capability
due to the presence of many extracellular polymeric substances like neutral polysaccharides, amyloids, extracellular
enzymes and biosurfactants [163].
CONCLUDING REMARKS
The mesohyl of marine sponges provides a favorable
environment for interactions of a large number of taxonomically diverse bacteria [144]. Several review articles have
emphasized the need for understanding symbiotic functions
in marine sponges [7, 88, 165-166]. The biotechnological
potential of these microorganisms (Fig. 2) remains little investigated and discussed. As shown in this mini-review, the
Biotechnological Potential of Sponge-Associated Bacteria
high diversity of marine bacteria is underexplored. These
microorganisms constantly metabolize sponge products and
synthetize numerous specific enzymes and secondary metabolites. From their immense diversity and their constant
activity stems their great potential as source of novel and
original metabolites and enzymes. In addition, it is likely that
disruption of the microbial symbiosis as a result from climate
change and/or environmental stress will have a significant
impact on the growth of marine sponges and their protection
against contamination, predation, and diseases. In an era of
rapid environmental change and degradation of marine ecosystems, sponge microbiology investigating the interaction
of marine sponges and associated microorganisms is fundamental for unraveling the maximum potential of microorganism for human use.
Advances have been made in the field of marine microbial biotechnology but a more extensive and focused approach is needed to investigate what else the marine microbes have to offer. It is important to seriously consider the
exploitation of marine microbial life and its associated secondary metabolites. Such studies can be aided by genomic
analyses, applying metabolomic approaches and employing
combined biomedical and biotechnological efforts, which
would lead to discovery of new compounds with a variable
degree of bioactivity [41]. The novel antimicrobial substances isolated and characterized from sponge-associated
bacteria seem to be very useful and promising for biomedical
research in the design of very specific and potent new pharmaceuticals for a wide variety of diseases [21]. The enzymes
produced by these microorganisms have the potential capacity to be uniquely suited for many industrial processes [75].
Furthermore, the treatment of environmental pollution by
employing microorganisms is a promising technology, either
by using bacteria associated with sponges as indicators of
contamination or for forming biofilms, or for biosurfactants
produced by them in marine ecosystems. Certain bioactive
metabolites may also be beneficial in ensuring environmental
hygiene, as such antifouling compounds [167]. Unfortunately, there are few studies on these compounds isolated
from sponge-associated bacteria [168-169].
Although new compounds are being added day by day,
our knowledge of marine microbial bioactive metabolites is
very small considering what exists in the deep ocean [41].
Clearly, coordination among microbiologists, natural product
chemists, and bioengineers will contribute significantly to
investigate the biotechnological potential of spongeassociated bacteria to the pharmaceutical and enzyme industries, and for bioremediation of contaminated aquatic environments.
Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
REFERENCES
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
CONFLICT OF INTEREST
The authors confirm that this article content has no conflict of interest.
[21]
[22]
ACKNOWLEDGEMENTS
This work was supported by a grant from the CAPES,
CNPq and FAPERJ to M.S. Laport. J. F. Santos-Gandelman
is the recipient of a CAPES and FAPERJ fellowship.
151
[23]
[24]
Dayton, P.K. Kelp communities of southern South America. Antarctic J. of the U.S. 1974, 9, 22-23.
Dayton, P.K. Interdecadal variation in an Antarctic sponge and its
predators from oceanographic climate shifts. Science, 1989, 245,
1484-1496.
Gili, J.M.; Coma, R. Benthic suspension feeders: Their paramount
role in littoral marine food webs. Trends Ecol. Evol., 1998, 13,
316-321.
Maldonado, M.; Cortadellas, N.; Trillas, M.I.; Rutzler, K. Endosymbiotic yeast maternally transmitted in a marine sponge. Biol.
Bull., 2005, 209, 94-106.
Gazave, E.; Lapébie, P.; Ereskovsky, A.V.; Vacelet, J.; Renard, E.;
Cárdenas, P.; Borchiellini, C. No longer Demospongiae: Homoscleromorpha formal nomination as a fourth class of Porifera.
Hydrobiologia, 2011, 687, 3-10.
Hooper, J.N.A.; Van Soest, R.W.M. Systema Porifera: a guide to
the classification of sponges, 1th ed. Plenum Publishers: New
York, USA, 2002.
Taylor, M.W.; Radax, R.; Steger, D.; Wagner, M. Spongeassociated microorganisms: evolution, ecology, and biotechnological potential. Microbiol. Mol. Biol. Rev., 2007, 71(2), 295-347.
Wilkinson, C.R. Immunological evidence for the Precambrian
origin of bacterial symbioses in marine sponges. Proc. R. Soc.
Lond. B, 1984, 220, 509-517.
Berthet, B.; Mouneyrac, C.; Pérez, T.; Amiard-Triquet, C. Metallothionein concentration in sponges (Spongia officinalis) as a biomarker of metal contamination. Comp. Bioch. Physiol., 2005,
C141, 306-313.
Rao, J.V.; Kavitha, P.; Chakra Reddy, N.; Rao, T.G. Petrosia
testudinaria as a biomarker for metal contamination at Gulf of
Mannar, southeast coast of India. Chemosphere, 2006, 65, 634-638.
Rao, J.V.; Srikanth, K.; Pallela, R.; Rao, T.G. The use of marine
sponge, Haliclona tenuiramosa as bioindicator to monitor heavy
metal pollution in the coasts of Gulf of Mannar, India. Environ.
Monit. Assess, 2009, 156(1-4), 451-459.
Blunt, J.W.; Copp, B.R.; Munro, M.H.; Northcote, P.T.; Prinsep,
M.R. Marine natural products. Nat. Prod. Rep., 2006, 23, 26-78.
Bergmann, W.; Feeney, R.J. The isolation of a new thymine pentoside from sponges. J. Am. Chem. Soc., 1950, 72, 2809-2810.
Bergmann, W.; Feeney, R.J. Contribution to the study of marine
sponges. 32. The nucleosides of sponges. J. Org. Chem., 1951, 16,
981-987.
Proksch, P.; Edrada, R.A.; Ebel, R. Drugs from the sea - Current
status and microbiological implications. Appl. Microbiol. Biotechnol., 2002, 59 (2-3), 125-134.
Wilkinson, C.R. Symbiotic interactions between marine sponges
and algae. In Algae and Symbioses: Plants, Animals, Fungi, Viruses, Interactions Explored. Reisser, W., Ed., Biopress Ltd.: Bristol, England, 1992, pp. 112-128.
Friedrich, A.B.; Hacker, J.; Fischer, I.; Proksch, P.; Hentschel, U.
Temporal variations of the microbial community associated with
the Mediterranean sponge Aplysina aerophoba. FEMS Microbiol.
Ecol., 2001, 38, 105-113.
Hentschel, U.; Hopke, J.; Horn, M.; Friedrich, A.B.; Wagner, M.;
Hacker, J.; Moore, B.S. Molecular evidence for a uniform microbial community in sponges from different oceans. Appl. Environ.
Microbiol., 2002, 68, 4431-4440.
Hentschel, U.; Usher, K.M.; Taylor, M.W. Marine sponges as
microbial fermenters. FEMS Microbiol. Ecol., 2006, 55, 167-177.
Blunt, J.W.; Copp, B.R.; Hu, W.P.; Munro, M.H.; Northcote, P.T.;
Prinsep, M.R. Marine natural products. Nat. Prod. Rep., 2009,
26(2), 170-244.
Laport, M.S.; Santos, O.C.S.; Muricy, G. Marine sponges: potential
sources of new antimicrobial drugs. Curr. Pharm. Biotechnol.,
2009, 10(1), 86-105.
Lee, Y.K.; Lee, J.H.; Lee, H.K. Microbial symbiosis in marine
sponges. J. Microbiol., 2001, 39, 254-264.
Jensen, P.R.; Fenical, W. Strategies for the discovery of secondary
metabolites from marine bacteria: ecological perspectives. Ann.
Rev. Microbiol., 1994, 48, 559-584.
Bernan, V.S.; Greenstein, M.; Maise, W.M. Marine microorganisms as a source of new natural products. Adv. Appl. Microbiol.,
1997, 43, 57-89.
152 Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
Haygood, M.G.; Schmidt, E.W.; Davidson, S.K.; Faulkner, D.J.
Microbial symbionts of marine invertebrates: opportunities for microbial biotechnology. J. Molec. Microbiol. Biotechnol., 1999, 1,
33-43.
Thomas, T.R.; Kavlekar, D.P.; LokaBharathi, P.A. Marine drugs
from sponge-microbe association-a review. Mar. Drugs, 2010,
8(4), 1417-1468.
Santos, O.C.S.; Pontes, P.V.M.L.; Santos, J.F.M.; Muricy, G.;
Giambiagi-Demarval, M.; Laport, M.S. Isolation, characterization
and phylogeny of sponge-associated bacteria with antimicrobial activities from Brazil. Res. Microbiol., 2010, 161, 604-612.
Kennedy, J.; Baker, P.; Piper, C.; Cotter, P.D.; Walsh, M.; Mooij,
M.J.; Bourke, M.B.; Rea, M.C.; O’Connor, P.M.; Ross, R.P.; Hill,
C.; O’Gara, F.; Marchesi, J.R.; Dobson, A.D. Isolation and analysis
of bacteria with antimicrobial activities from the marine sponge
Haliclona simulans collected from Irish waters. Mar. Biotechnol.,
2009, 11(3), 384-396.
Zhang, W.; Zhang, F.; Li, Z.; Miao, X.; Meng, Q.; Zhang, X. Investigation of bacteria with polyketide synthase genes and antimicrobial activity isolated from South China Sea sponges. J. Appl. Microbiol., 2009, 107(2), 567-575.
Pimentel-Elardo, S.M.; Kozytska, S.; Bugni, T.S.; Ireland, C.M.;
Moll, H.; Hentschel, U. Anti-parasitic compounds from Streptomyces sp. Strains isolated from Mediterranean sponges. Mar. Drugs,
2010, 8(2), 373-380.
Abdelmohsen, U.R.; Pimentel-Elardo, S.M.; Hanora, A.; Radwan,
M.; Abou-El-Ela, S.H.; Ahmed, S.; Hentschel, U. Isolation, phylogenetic analysis and anti-infective activity screening of marine
sponge-associated actinomycetes. Mar. Drugs, 2010, 8, 399-412.
Flemer, B.; Kennedy, J.; Margassery, L.M.; Morrissey, J. P.;
O'gara, F.; Dobson, A.D. Diversity and antimicrobial activities of
microbes from two Irish marine sponges, Suberites carnosus and
Leucosolenia sp. J. Appl. Microbiol., 2012, 11(2), 289-301.
Enticknap, J.J.; Kelly, M.; Peraud, O.; Hill, R.T. Characterization
of a culturable alphaproteobacterial symbiont common to many
marine sponges and evidence for vertical transmission via sponge
larvae. Appl. Environ. Microbiol., 2006, 72, 3724-3732.
Thiel, V.; Imhoff, J.F. Phylogenetic identification of bacteria with
antimicrobial activities isolated from Mediterranean sponges. Biomol. Eng., 2003, 20, 421-423.
Müller, W.E.G.; Brummer, F.; Batel, R.; Müller, I.M.; Schröder,
H.C. Molecular biodiversity. 1Case study: Porifera (Sponges).
Naturwissenschaften, 2003, 90, 103-120.
Blunt, J.W.; Copp, BR.; Keyzers, R.A.; Munro, M.H.; Prinsep,
M.R. Marine natural products. Nat. Prod. Rep., 2014, 31(2), 160258.
Blunt, J.W.; Copp, BR.; Keyzers, R.A.; Munro, M.H.; Prinsep,
M.R. Marine natural products. Nat. Prod. Rep., 2013, 30(2), 237323.
Blunt, J.W.; Copp, BR.; Keyzers, R.A.; Munro, M.H.; Prinsep,
M.R. Marine natural products. Nat. Prod. Rep., 2012, 29(2), 144222.
Blunt, J.W.; Copp, BR.; Keyzers, R.A.; Munro, M.H.; Prinsep,
M.R. Marine natural products. Nat. Prod. Rep., 2011, 28(2), 196268.
Blunt, J.W.; Copp, BR.; Keyzers, R.A.; Munro, M.H.; Prinsep,
M.R. Marine natural products. Nat. Prod. Rep., 2010, 27(2), 165237.
Bhatnagar, I.; Kim, S.K. Immense essence of excellence: marine
microbial bioactive compounds. Mar. Drugs, 2010, 8, 2673-2701.
Yasuhara-Bell, J.; Lu, Y. Marine compounds and their antiviral
activities. Antiviral. Res., 2010, 86(3), 231-240.
Radwan, M.; Hanora, A.; Khalifa, S.; Abou-El-Ela, S.H. Manzamines: A potential for novel cures. Cell Cycle, 2012 11(9), 17651772.
Burgess, J.G.; Jordan, E.M.; Bregu, M.; Mearns-Spragg, A.; Boyd,
K.G. Microbial antagonism: A neglected avenue of natural products research. J. Biotechnol., 1999, 70, 27-32.
Martin, J.F.; Liras, P. Organization and expression of genes involved in the biosynthesis of antibiotics and other secondary metabolites. Ann. Rev. Microbiol., 1989, 43, 173-206.
Hentschel, U.; Schmid, M.; Wagner, M.; Fieseler, L.; Gernert, C.;
Hacker, J. Isolation and phylogenetic analysis of bacteria with antimicrobial activities from the Mediterranean sponges Aplysina
aerophoba and Aplysina cavernicola. FEMS Microbiol. Ecol.,
2001, 35, 305-312.
Santos-Gandelman et al.
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
Chelossi, E.; Milanese, M.; Milano, A.; Pronzato, R.; Riccardi, G.
Characterisation and antimicrobial activity of epibiotic bacteria
from Petrosia ficiformis (Porifera, Demospongiae). J. Exp. Mar.
Biol. Ecol., 2004, 309, 21-33.
Marinho, P.R.; Moreira, A.P.; Pellegrino, F.L.; Muricy, G.; Bastos,
M.C.; Santos, K.R.; Giambiagi-deMarval, M.; Laport, M.S. Marine
Pseudomonas putida: a potential source of antimicrobial substances
against antibiotic-resistant bacteria. Mem. Inst. Oswaldo Cruz,
2009, 104(5), 678-682.
Pabel, C.T.; Vater, J.; Wilde, C.; Franke, P.; Hofemeister, J.; Adler,
B.; Bringmann, G.; Hacker, J.; Hentschel, U. Antimicrobial activities and matrix-assisted laser desorption/ionization mass spectrometry of Bacillus isolates from the marine sponge Aplysina
aerophoba. Mar. Biotechnol., 2003, 5(5), 424-434.
Selvin, J.; Joseph, S.; Asha, K.R.T.; Manjusha, W.A.; Sangeetha,
V.S.; Jayaseema, D.M.; Antony, M.C.; Vinitha, A.J.D. Antibacterial potential of antagonistic Streptomyces sp. isolated from marine
sponge Dendrilla nigra. FEMS Microbiol. Ecol., 2004, 50(2), 117122.
Thakur, A.N.; Thakur, N.L.; Indap, M.M.; Pandit, R.A.; Datar,
V.V.; Müller, W.E.G. Antiangiogenic, antimicrobial, and cytotoxic
potential of sponge-associated bacteria. Mar. Biotechnol., 2005,
7(3), 245-252.
Anand, T.P.; Bhat, A.W.; Shouche, Y.S.; Roy, U.; Siddharth, J.;
Sarma, S.P. Antimicrobial activity of marine bacteria associated
with sponges from the waters off the coast of South East India. Microbiol. Res., 2006, 161(3), 252-262.
Bultel-Poncé, V.; Debitus, C.; Berge, J.; Cerceau, C.; Guyot, M.
Metabolites from the sponge associated bacterium Micrococcus luteus. J. Mar. Biotechnol., 1998, 6, 233-236.
Bultel-Poncé, V.; Berge, J.; Debitus, C.; Nicolas, J.; Guyot, M.
Metabolites from the sponge associated bacterium Pseudomonas
species. Mar. Biotechnol., 1999, 1, 384-390.
Zheng, L.; Yan, X.; Xu, J.; Chen, H. and Lin, W. Hymeniacidon
perleve associated bioactive bacterium Pseudomonas sp. NJ6-3-1.
Prikl. Biokhim. Mikrobiol., 2005, 41, 35-39.
Selvin, J. Exploring the antagonistic producer Streptomyces
MSI051: implications of polyketide synthase gene type II and a
ubiquitous defense enzyme phospholipase A2 in host sponge Dendrilla nigra. Curr. Microbiol., 2009, 58(5), 459-463.
Oclarit, J.M.; Okada, H.; Ohta, S.; Kaminura, K.; Yamaoka, Y.;
Iizuka, T.; Miyashiro, S.; Ikegami, S. Anti-bacillus substance in the
marine sponge, Hyatella species, produced by an associated Vibrio
species bacterium. Microbios, 1994, 78, 7-16.
Han, Y.; Yang, B.; Zhang, F.; Miao, X.; Li, Z. Characterization of
antifungal chitinase from marine Streptomyces sp. DA11 associated
with south China sea sponge Craniella australiensis. Mar. Biotechnol., 2009, 11, 132-140.
Mitova, M.; Popov, S.; De Rosa, S. Cyclic peptides from a
Ruegeria strain of bacteria associated with the sponge Suberites
domuncula. J. Nat. Prod., 2004, 67(7), 1178-1181.
Jayatilake, G.S.; Thornton, M.P.; Leonard, A.C.; Grimwade, J.E.;
Baker, B.J. Metabolites from an Antarctic sponge associated bacterium, Pseudomonas aeruginosa. J. Nat. Prod. 1996, 59, 293-296.
Li, Z. Advances in marine microbial symbionts in the China Sea
and related pharmaceutical metabolites. Mar. Drugs, 2009, 7, 113129.
Stierle, A.C.; Cardellina, J.H. A marine Micrococcus produces
metabolites ascribed to the sponge Tedania ignis. 2nd, Singleton,
F.L. Experientia, 1988, 44(11-12), 1021.
Dunlap, W.C.; Battershill, C.N.; Liptrot, C.H.; Cobb, R.E.; Bourne,
D.G.; Jaspars, M.; Long, P.F.; Newman, D.J. Biomedicinals from
the phytosymbionts of marine invertebrates: A molecular approach.
Methods, 2007, 42, 358-376.
Ang, K.K.H.; Holmes, M.J.; Higa, T.; Hamann, M.T.; Kara,
U.A.K. In vivo antimalarial activity of the beta-carboline alkaloid
manzamine A. Antimicrob. Agents Chemother., 2000, 44, 16451649.
Singh, M.P.; Kong, F.; Janso, J.E.; Arias, D.A.; Suarez, P.A.; Bernan, V.S.; Petersen, P.J.; Weiss, W.J.; Carter, G.; Greenstein, M.
Novel alpha-pyrones produced by a marine Pseudomonas sp.
F92S91: taxonomy and biological activities. J. Antibiot., 2003, 56,
1033-1044.
Kim, T.K.; Hewavitharana, A.K.; Shaw, P.N.; Fuerst, J.A. Discovery of a new source of rifamycin antibiotics in marine sponge Acti-
Biotechnological Potential of Sponge-Associated Bacteria
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
nobacteria by phylogenetic prediction. Appl. Environ. Microbiol.,
2006, 72, 2118-2125.
Kunz, A.L.; Labes, A.; Wiese, J.; Bruhn, T. ; Bringmann,
G.; Imhoff, J.F. Nature's lab for derivatization: new and revised
structures of a variety of streptophenazines produced by a spongederived Streptomyces strain. Mar. Drugs, 2014 12(4), 1699-1714.
Phelan, R.W.; Barret, M.; Cotter, P.D.; O'Connor, P.M.; Chen, R.;
Morrissey, J.P.; Dobson, A.D.; O'Gara, F.; Barbosa, T.M. Subtilomycin: a new lantibiotic from Bacillus subtilis strain MMA7 isolated from the marine sponge Haliclona simulans. Mar. Drugs,
2013, 11(6), 1878-1898.
Elyakov, G.B.; Kuznetsova, T.; Mikhailov, V.V.; Maltsev, I.I.;
Voinov, V.G.; Fedoreyev, S.A. Brominated diphenyl ethers from a
marine bacterium associated with the sponge Dysidea sp. Cell. Mol.
Life Sci., 1991, 47, 632-633.
Pimentel-Elardo, S.M.; Buback, V.; Gulder, T.A.; Bugni, T.S.;
Reppart, J.; Bringmann, G.; Ireland, C.M.; Schirmeister, T.;
Hentschel, U. New tetromycin derivatives with anti-trypanosomal
and protease inhibitory activities. Mar. Drugs, 2011, 9(10), 16821697.
Nagai, K.; Kamigiri, K.; Arao, N.; Suzumura, K.; Kawano, Y.;
Yamaoka, M.; Zhang, H.; Watanabe, M.; and Suzuki, K. YM266183 and YM-266184, novel thiopeptide antibiotics produced by
Bacillus cereus isolated from a marine sponge. I. Taxonomy, fermentation, isolation, physico-chemical properties and biological
properties. J. Antibiot., 2003, 56, 123-128.
Kobayashi, M.; Aoki, S.; Gato, K.; Matsunami, K.; Kurosu, M.;
Kitagawa, I. Marine natural products. XXXIV. Trisindoline, a new
antibiotic indole trimer, produced by a bacterium of Vibrio sp.
separated from the marine sponge Hyrtios altum. Chem. Pharm.
Bull., 1994, 42, 2449-2451.
Imamura, N.; Nishijima, M.; Adachi, K.; Sano, H. Novel antimycin
antibiotics, urauchimycins A and B, produced by marine actinomycete. J. Antibiot., 1993, 46, 241-246.
McDaniel, L.D.; Young, E.; Delaney, J.; Ruhnau, F.; Ritchie, K.B.;
Paul, J.H. High frequency of horizontal gene transfer in oceans.
Science, 2010, 330, 50.
Kennedy, J.; O'Leary, N.D.; Kiran, G.S.; Morrissey, J.P.; O'Gara,
F.; Selvin, J.; Dobson, A.D. Functional metagenomic strategies for
the discovery of novel enzymes and biosurfactants with biotechnological applications from marine ecosystems. J. Appl. Microbiol.,
2011, 111(4), 787-799.
Stahl, U.; Lee, M.; Sjodahl, S.; Archer, D.; Cellini, F; Ek, B. Plant
low molecular-weight phospholipase A2s (PLA2s) are structurally
related to the animal secretory PLA2s and are present as a family of
isoformas in rice (Oryza sativa). Plant Mol Biol., 1999, 41(4), 481490.
Nevalainen, T.J.; Peuravuori, H.J.; Quinn, R.J.; Llewellyn, L.E.;
Benzie, J.A.H.; Fenner, P.J.; Winkel, K.D. Phospholipase A2 in
Cnidaria. Comp. Biochem. Physiol. Part B, 2004a, 139, 731-735.
Nevalainen, T.J.; Quinn, R.J.; Hooper, J.N.A. Phospholipase A2 in
Porifera. Comp. Biochem. Physiol. Part B, 2004b, 137, 413-420.
Imhoff, J.F.; Stöhr, R. Sponge-associated bacteria: General overview and special aspects of the diversity of bacteria associated with
Halichondria panicea. In Marine Molecular Biotechnology,
W.E.G. Müller Ed.; Springer: New York, USA, 2003, Vol. 1
Sponges (Porifera), pp. 35-57.
Mohapatra, B.R., Bapuji, M.; Sree, A. Production of industrial
enzymes (amylase, carboxymethylcellulase and protease) by bacteria isolated from marine sedentary organisms. Acta Biotechnol.,
2003, 23, 75-84.
Kiran, G.S.; Shanmughapriya, S.; Jayalakshmi, J.; Selvin, J.;
Gandhimathi, R.; Sivaramakrishnan, S.; Arunkumar, M.; Thangavelu, T.; Natarajaseenivasan, K. Optimization of extracellular
psychrophilic alkaline lipase produced by marine Pseudomonas sp.
(MSI057). Bioproc. and Biosyst. Engi., 2008, 31, 483-492.
Shanmughapriya, S.; Krishnaveni, J.; Selvin, J.; Gandhimathi, R.;
Arunkumar, M.; Thangavelu, T.; Seghal Kiran, G.; Natarajaseenivasan, K. Optimization of extracellular thermotolerant alkaline protease produced by marine Roseobacter sp (MMD040). Bioprocess.
Biosyst. Eng., 2008, 31, 427-433.
Shanmughapriya, S.; Seghal Kiran, G.; Selvin, J.; Gandhimathi, R.;
Bastin Baskar, T.; Manilal, A.; Sujith, S. Optimization, production
and partial characterization of an alkalophilic amylase produced by
sponge associated marine bacterium Halobacterium salinarum
MMD047. Biotechnol. Bioprocess. Eng., 2009, 14, 67-75.
Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
153
Shanmughapriya, S.; Kiran, G.S.; Selvin, J.; Thomas, T.A.; Rani,
C. Optimization, purification, and characterization of extracellular
mesophilic alkaline cellulase from sponge-associated Marinobacter
sp. MSI032. Appl. Biochem. Biotechnol., 2010, 162(3), 625-640.
Okamura, Y.; Kimura, T.; Yokouchi, H.; Meneses-Osorio, M.;
Katoh, M.; Matsunaga, T.; Takeyama, H. Isolation and characterization of a GDSL esterase from the metagenome of a marine
sponge-associated bacteria. Mar. Biotechnol., (NY) 2010, 12(4),
395-402.
Pandey, S.; Sree, A.; Sethi, D.P.; Kumar, C.G.; Kakollu, S.;
Chowdhury, L.; Dash, S.S. A marine sponge associated strain of
Bacillus subtilis and other marine bacteria can produce anticholinesterase compounds. Microb. Cell Fact, 2014, 13(1), 24.
Mohapatra, B.R.; Bapuji, M. Characterization of urethanase from
Micrococcus species associated with the marine sponge (Spirastrella species). Lett. Appl. Microbiol., 1997, 25, 393-396.
Selvin, J.; Ninawe, A.S.; Seghal Kiran, G.; Lipton, A.P. Spongemicrobial interactions: Ecological implications and bioprospecting
avenues. Crit. Rev. Microbiol., 2010, 36(1), 82-90.
Ahn, Y.B.; Rhee, S.K.; Fennell, D.E.; Kerkhof, L.J.; Hentschel, U.;
Haggblom, M.M. Reductive dehalogenation of brominated phenolic compounds by microorganisms associated with the marine
sponge Aplysina aerophoba. Appl. Environ. Microbiol., 2003, 69,
4159-4166.
Margesin, R.; Schinner, F. Biodegradation and bioremediation of
hydrocarbons in extreme environments. Appl. Microbiol. Biotechnol., 2001, 56, 650-663.
Morikawa, M. Beneficial Biofilm Formation by Industrial Bacteria
Bacillus subtilis and Related Species. J. Biosci. Bioeng., 2006,
101(1), 1-8.
Kefalas, E.; Castritsi-Catharios, J.; Miliou, H. Bacteria associated
with the sponge Spongia officinalis as indicators of contamination.
Ecol. Indicators, 2003, 2, 339-343.
Hiroki, M. Effects of heavy metal contamination on soil microbial
population. Soil Sci. Plant Nutr., 1992, 38, 141-147.
Doelman, P.; Jansen, E.; Michels, M.; Van Til, M. Effects of heavy
metals in soil on microbial diversity and activity as shown by the
sensitivity-resistance index, an ecologically relevant parameter.
Biol. Fertil. Soil, 1994, 17, 177-184.
Lu, W.B.; Shi, J.J.; Wang, C.H.; Chang, J.S. Biosorption of lead,
copper and cadmium by an indigenous isolate Enterobacter sp. J1
possessing high heavy-metal resistance. J. Hazard. Mater., 2006,
B134, 80-86.
Selvin, J.; Shanmugha Priya, S.; Seghal Kiran, G.; Thangavelu, T.;
Sapna Bai, N. Sponge-associated marine bacteria as indicators of
heavy metal pollution. Microbiol. Res., 2007, 164(3), 352-633.
Cebrain, E.; Marti, R.; Uriz, J.M.; Turon, X. Sublethal effects of
contamination on the Mediterranean sponge Crambe crambe: metal
accumulation and biological responses. Mar. Pollut. Bull., 2003,
46, 1273-1284.
Agell, G.; Uriz, M.L.; Cebrain, E.; Martz, R. Does stress proteins
induction by copper modify natural toxicity in sponges? Environ.
Toxicol. Chem., 2001, 20, 2588-2593.
Guzzo, A.; Du Bow, M.; Bauda, P. Identification and characterization of genetically programmed responses to toxic metal exposure
in Escherichia coli. Metals and microorganisms: relationships and
application. FEMS Microbiol. Rev., 1994, 14, 369-374.
Milanese, M.; Chelossi, E.; Manconi, R.; Sara, A.; Sidri, M.; Pronzato, R. The marine sponge Chondrilla nucula Schmidt, 1862 as an
elective candidate for bioremediation in integrated aquaculture.
Biomol. Eng., 2003, 2, 363-368.
De, J.; Ramaiah, N.; Vardanyan, L. Detoxification of toxic heavy
metals by marine bacteria highly resistant to mercury. Mar. Biotechnol., 2008, 10, 471-477.
Nies, D.H. Microbial heavy-metal resistance. Appl. Microbiol.
Biotechnol., 1999, 51, 730-750.
Chang, J.S.; Law, R.; Chang, C.C. Biosorption of lead, copper and
mercury by biomass of Pseudomonas aeruginosa PU21. Wat. Res.,
1997, 31, 1651-1658.
Selin, N.E. Global biogeochemical cycling of mercury: A review.
Ann. Rev. Environ. Resour., 2009, 34, 43-63.
Vas, J.; Monestier, M. Immunology of mercury. Ann. NY Acad.
Sci., 2008, 1143, 240-267.
Gadd, G.M.; White, C. Microbial treatment of metal pollution: A
working biotechnology? Trends Biotechnol., 1993, 11, 353-359.
154 Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
[107]
[108]
[109]
[110]
[111]
[112]
[113]
[114]
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]
[125]
[126]
[127]
[128]
[129]
Silver, S. Bacterial resistances to toxic metals: A review. Gene,
1996, 179, 9-19.
Nakamura, K.; Nakahara, H. Simplified X-ray film method for
detection of bacterial volatilization of mercury chloride by Escherichia coli. Appl. Environ. Microbiol., 1988, 54, 2871-2873.
Gerlach, A.S. Marine Pollution: Diagnosis and Therapy. Springer:
New York, USA, 1981; pp. 218.
Hall, J.L. Cellular mechanisms for heavy metal detoxification and
tolerance. J. Exp. Bot., 2002, 53(366), 1-11.
Deng, X.; Wilson, D.B. Bioaccumulation of mercury from wastewater by genetically engineered Escherichia coli. Appl. Microbiol.
Biotechnol., 2001, 56, 276-279.
Guo, W. J.; Bundithya, W.; Goldsbrough, P.B. Characterization of
the Arabidopsis metallothionein gene family: tissue-specific expression and induction during senescence and in response to copper. New Phytologist., 2003, 159, 369-381.
Mir, G.; Domènech, J.; Huguet, G.; Guo, W.J.; Goldsbrough, P.;
Atrian, S.; Molinas, M. A plant type 2 metallothionein (MT) from
cork tissue responds to oxidative stress. J. Exp. Bot., 2004, 55,
2483-2493.
Eapen, S.; D’Souza, S. F. Prospects of genetic engineering of
plants for phytoremediation of toxic metals. Biotechnol. Adv.,
2005, 23, 97-114.
Brim, H.; Mcfarlan, S.C.; Fredrickson, J.K.; Minton, K.; Zhai, M.;
Wackett, L.P.; Daly, M.J. Engineering Deinococcus radiodurans
for metal remediation in radioactive mixed waste environments.
Nat. Biotechnol., 2000, 18, 85-90.
Barkay, T.; Liebert, C.; Gillman, M. Environmental significance of
the potential for mer(Tn21)-mediated reduction of Hg2+ to Hg0 in
natural waters. Appl. Environ. Microbiol., 1989, 55(5), 1196-1202.
Bruce, K.D.; Osborn, A.M.; Pearson, A.J.; Strike, P.; Ritchie, D.A.
Genetic diversity within mer genes directly amplified from communities in noncultivated soil and sediment bacteria. Mol. Ecol.,
1995, 4(5), 605-612.
Rasmussen, L.D.; Sørensen, S.J. The effect of long-term exposure
of mercury in the bacterial community in marine sediment. Curr.
Microbiol., 1998, 36, 291-297.
Reyes, N.S.; Frischer, M.E.; Sobecky, P.A. Characterization of
mercury resistance mechanisms in marine sediment microbial
communities. FEMS Microbiol. Ecol., 1999, 30, 273-284.
Liesegang, H.; Lemke, K.; Siddiqui, R.A.; Schlegel, H.G. Characterization of the inducible nickel and cobalt resistance determinant
cnr from pMOL28 of Alcaligenes eutrophus CH34. J. Bacteriol.,
1993, 175, 767-778.
Hassan, M.; Der Lelie, D.V.; Springael, D.; Römling, N.; Ahmed,
N.; Mergeay, M. Identification of a gene cluster, czr, involved in
cadmium and zinc resistance in Pseudomonas aeruginosa. Gene,
1999, 238, 417-425.
Lang, S. Biological amphiphiles (microbial biosurfactants). Curr.
Opin. Colloid Inter. Sci., 2002, 7, 12-20.
Desai, J.D.; Banat, I.M. Microbial production of surfactants and
their commercial potential. Microbiol. Mol. Biol. R., 1997, 61, 4764.
Kosaric, N. Biosurfactants and their application for soil bioremediation. Food Technol. Biotechnol., 2001, 39, 295-304.
Das, P.; Mukherjee, S.; Sen, R. Improved bioavailability and biodegradation of a model polyaromatic hydrocarbon by a biosurfactant producing bacterium of marine origin. Chemosphere, 2008,
72(9), 1229-1234.
Kumar, A.S.; Mody, K.; Jha, B. Evaluation of biosurfactant/bioemulsifier production by a marine bacterium. Bull. Environ.
Contam. Toxicol., 2007, 79, 617-621.
Gandhimathi, R.; Seghal Kiran, G.; Hema, T.A.; Selvin, J.; Rajeetha Raviji, T.; Shanmughapriya, S. Production and characterization of lipopeptide biosurfactant by a sponge-associated marine
actinomycetes Nocardiopsis alba MSA10. Bioprocess. Biosyst.
Eng., 2009, 32, 825-835.
Gnanamani, A.; Kavitha, V.; Radhakrishnan, N.; Suseela Rajakumar, G.; Sekaran, G.; Mandal, A.B. Microbial products (biosurfactant and extracellular chromate reductase) of marine microorganism are the potential agents reduce the oxidative stress induced
by toxic heavy metals. Colloids Surf. B Biointerfaces, 2010, 79(2),
334-339.
Pacwa-Płociniczak, M., Płaza, G.A., Piotrowska-Seget, Z.;
Cameotra, S.S. Environmental applications of biosurfactants: recent advances. Intern. J. Mol. Sci., 2011, 12, 633-654.
Santos-Gandelman et al.
[130]
[131]
[132]
[133]
[134]
[135]
[136]
[137]
[138]
[139]
[140]
[141]
[142]
[143]
[144]
[145]
[146]
[147]
[148]
[149]
[150]
Khopade, A.; Ren, B.; Liu, X.Y.; Mahadik, K.; Zhang, L.; Kokare,
C. Production and characterization of biosurfactant from marine
Streptomyces species B3. J. Colloid Interface Sci., 2012, 367(1),
311-318.
Nakano, M.; Iehata, S.; Tanaka, R.; Maeda, H. Extracellular neutral
lipids produced by the marine bacteria Marinobacter sp. Biocontrol
Sci., 2012, 17(2), 69-75.
Juwarkar, A.A.; Dubey, K.V.; Nair, A.; Singh, S.K. Bioremediation
of multi-metal contaminated soil using biosurfactant-a novel approach. Indian J. Microbiol., 2008, 48(1), 142-146.
Das, P.; Mukherjee, S.; Sen, R. Biosurfactant of marine origin
exhibiting heavy metal remediation properties. Bioresour. Technol.,
2009, 100, 4887-4890.
Santos-Gandelman, J.F.; Santos, O.C.S.; Pontes, P.V.M.; Andrade,
C.L.; Korenblum, E.; Muricy, G.; Giambiagi-deMarval, M.; Laport,
M.S. Community structure and characterization of spongeassociated bacteria from Brazilian coast. Mar. Biotechnol., 2013.
15, 668-676.
Santos-Gandelman, J.F.; Cruz, K.; Crane, S.; Muricy, G.; Giambiagi-deMarval M.; Barkay, T.; Laport, M.S. Potential application in
mercury bioremediation of a marine sponge-isolated Bacillus
cereus strain Pj1. Curr. Microbiol., 2014, DOI 10.1007/s00284014-0597-5.
Gandhimathi R, Seghal Kiran G, Hema TA, Selvin J, Rajeetha
Raviji T, Shanmughapriya S. Production and characterization of
lipopeptide biosurfactant by a sponge-associated marine actinomycetes Nocardiopsis alba MSA10. Bioprocess Biosyst. Eng., 2009,
32(6), 825-835.
Lawrance, A.; Balakrishnan, M.; Joseph, T.C.; Sukumaran,
D.P.; Valsalan, V.N.; Gopal, D.; Ramalingam, K. Functional and
molecular characterization of a lipopeptide surfactant from the marine sponge-associated eubacteria Bacillus licheniformis NIOTAMKV06 of Andaman and Nicobar Islands, India. Mar. Pollut.
Bull., 2014, 82, 76-85.
Matz, C.; Webb, J.S.; Schupp, P.J.; Phang, S.Y.; Penesyan, A.;
Egan, S.; Steinberg, P.; Kjelleberg, S. Marine biofilm bacteria
evade eukaryotic predation by targeted chemical defense. PLoS
One, 2008, 3(7):e2744.
Muthusamy, A.K.; Kanapathi, T.K.A.; Karuppiah, P. Production
and characterization of exopolysaccharides (eps) from biofilm
forming marine bacterium. Brazilian Arch. Biol. Technol., 2011,
54(2), 259-265.
Lee, J.W.; Nam, J.H.; Kim, Y.H.; Lee, K.H.; Lee, D.H. Bacterial
communities in the initial stage of marine biofilm formation on artificial surfaces. J. Microbiol., 2008, 46(2), 174-182.
Davey, M.E.; O'toole, G.A. Microbial biofilms: from ecology to
molecular genetics. Microbiol. Mol. Biol. Rev., 2000, 64(4), 847867.
Whitchurch, C.B.; Tolker-Nielsen, T.; Ragas, P.C.; Mattick, J.S.
Extracellular DNA required for bacterial biofilm formation. Science, 2002, 95(5559), 1487.
Allesen-Holm, M.; Barken, K.B.; Yang, L.; Klausen, M.; Webb,
J.S.; Kjelleberg, S.; Molin, S.; Givskov, M.; Tolker-Nielsen, T. A
characterization of DNA release in Pseudomonas aeruginosa cultures and biofilms. Mol. Microbiol., 2006, 59(4), 1114-1128.
Stoodley, P.; Sauer, K.; Davies, D.G.; Costerton, J. W. Biofilms as
complex differentiated communities. Annu. Rev. Microbiol., 2002,
56, 187-209.
Dang, H.; Lovell, C.R. Bacterial primary colonization and early
succession on surfaces in marine waters as determined by amplified
rRNA gene restriction analysis and sequence analysis of 16S rRNA
genes. Appl. Environ. Microbiol., 2000, 66, 467-475.
Singh, R.; Paul, D.; Jain, R.K. Biofilms: implications in bioremediation. Trends Microbiol., 2006, 14(9), 389-397.
Vu, B.; Chen, M.; Crawford, R.J.; Ivanova, E.P. Bacterial extracellular polysaccharides involved in biofilm formation. Molecules,
2009, 14, 2535-2554.
Tribelli, P.M.; Martino, C.D.; Lopez, N.I.; Iustman, L.J.R. Biofilm
lifestyle enhances diesel bioremediation and biosurfactant production in the Antarctic polyhydroxyalkanoate producer Pseudomonas
extremaustralis. Biodegradation, 2012, 23, 645-651.
Decho, A.W. Microbial biofilms in intertidal systems: an overview.
Cont. Shelf Res., 2000, 20, 1257-1273.
Akinson, B. Immobilized biomass-a basis for process development
in wastewater treatment. In Biological Fluidized Bed Treatment of
Biotechnological Potential of Sponge-Associated Bacteria
[151]
[152]
[153]
[154]
[155]
[156]
[157]
[158]
[159]
Water and Wastewater. Cooper, P.E.; Akinson, B., Eds, Ellis Horwood: New York, USA, 1981, pp. 22-34.
Paul, D.; Pandey, G.; Pandey, J.; Jain, R.K. Accessing microbial
diversity for bioremediation and environmental restoration. Trends
Biotechnol., 2005, 23, 135-142.
Pandey, G. and Jain, R.K. Bacterial chemotaxis toward environmental pollutants: role in bioremediation. Appl. Environ. Microbiol., 2002, 68, 5789-5795.
Barkay, T.; Schaefer, J. Metal and radionuclide bioremediation:
issues, considerations and potentials. Curr. Opin. Microbiol., 2001,
4, 318-323.
Almaguer-Cantú, V.; Morales-Ramos, L.H.; Balderas-Rentería, I.
Biosorption of lead (II) and cadmium (II) using Escherichia coli
genetically engineered with mice metallothionen I. Water Sci.
Technol., 2011, 63, 1607-1613.
Borrock, D.; Fein, J.B.; Kulpa, C.F. Proton and Cd adsorption onto
natural bacterial consortia: testing universal adsorption behavior.
Geochim. Cosmochim. Acta, 2004, 68, 3231-3238.
Fang, L.; Cai, P.; Li, P.; Wu, H.; Liang, W.; Rong, X.; Chen, W.;
Huang, Q. Micro calorimetric and potentiometric titration studies
on the adsorption of copper by P. putida and B. thuringiensis and
their composites with minerals. J. Hazard. Mater., 2010, 181,
1031-1038.
Hawari, A.H.; Mulligan, C.N. Biosorption of lead (II), cadminum
(II), copper (II) and nickel (II) by anaerobic granular biomass.
Bioresour. Technol., 2005, 97, 692-700.
Joo, J.H.; Hassan, S.H.A.; Oh, S.E. Comparative study of biosorption of Zn2+ by Pseudomonas aeruginosa and Bacillus cereus. Int.
Biodeterior. Biodegrad., 2010, 64, 734-741.
Ozdemir, S.; Kilinc, E.; Poli, A.; Nicolaus, B.; Guven, K. Biosorption of Cd, Cu, Ni, Mn and Zn from aqueous solutions by thermophilic bacteria, Geobacillus toebii sub. sp. decanicus and Geobacillus thermoleovorans sub.sp. stromboliensis: Equilibrium, kinetic
and thermodynamic studies. Chem. Eng. J., 2009, 152, 195-206.
Received: March 27, 2014
View publication stats
Revised: March 28, 2014
Accepted: June 20, 2014
Current Pharmaceutical Biotechnology, 2014, Vol. 15, No. 2
[160]
[161]
[162]
[163]
[164]
[165]
[166]
[167]
[168]
[169]
155
Pokrovsky, O.S.; Feurtet-Mazel, A.; Martinez, R.E.; Morin, S.;
Baudrimont, M.; Duong, T.; Coste, M. Experimental study of cadmium interaction with periphytic biofilms. Appl. Geochem., 2010,
25, 418-427.
Quintelas, C.; Rocha, Z.; Silva, B.; Fonseca, B.; Figueiredo, H.;
Tavares, T. Biosorptive performance of an Escherichia coli biofilm
supported on zeolite NaY for the removal of Cr(VI), Cd(II), Fe(III)
and Ni(II). Chem. Eng. J., 2009, 152, 110-115.
Pérez Silva, R.M.; Abalos-Rodríguez, A.; Gómez-Montes, J.M.;
Cantero-Moreno, D. Biosorption of chromium, copper, manganese
and zinc by Pseudomonas aeruginosa AT18 isolated from a site
contaminated with petroleum. Bioresour. Technol., 2009, 100,
1533-1538.
Dash, H.R.; Mangwani, N.; Das, S. Characterization and potential
application in mercury bioremediation of highly mercury-resistant
marine bacterium Bacillus thuringiensis PW-05. Environ. Sci. Pollut. Res., 2014, 21, 2642-2653.
Zhang, W.; Chen, L.; Liu, D. Characterization of a marine-isolated
mercury-resistant Pseudomonas putida strain SP1 and its potential
application in marine mercury reduction. Appl. Microbiol. Biotechnol., 2012, 93, 1305-1314.
Vogel, G. The inner lives of sponges. Science, 2008, 320(5879),
1028-1030.
Webster, N.S.; Blackall, L.L. What do we really know about
sponge-microbial symbioses? ISME J., 2009, 3(1), 1-3.
Qian, P.Y.; Chen, L., Xu, Y. Mini-review: Molecular mechanisms
of antifouling compounds. Biofouling, 2013, 29(4), 381-400.
Dash, S.; Jin, C.; Lee, O.O.; Xu, Y.; Qian, P.Y. Antibacterial and
antilarval-settlement potential and metabolite profiles of novel
sponge-associated marine bacteria. J. Ind. Microbiol. Biotechnol.,
2009, 36(8), 1047-1056.
Dash, S.; Nogata, Y.; Zhou, X.J.; Zhang, Y.; Xu, Y.; Guo, X.;
Zhang, X.; Qian, P.Y. Poly-ethers from Winogradskyella poriferorum: Antifouling potential, time-course study of production and
natural abundance. Bioresour. Technol., 2011, 102, 7532-7537.
Download