Next Article in Journal
Is Yessotoxin the Main Phycotoxin in Croatian Waters?
Next Article in Special Issue
Production of Metabolites as Bacterial Responses to the Marine Environment
Previous Article in Journal
Inhibition of Nitric Oxide (NO) Production in Lipopolysaccharide (LPS)-Activated Murine Macrophage RAW 264.7 Cells by the Norsesterterpene Peroxide, Epimuqubilin A
Previous Article in Special Issue
3-O-Methylfunicone, a Selective Inhibitor of Mammalian Y-Family DNA Polymerases from an Australian Sea Salt Fungal Strain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Development of Novel Drugs from Marine Surface Associated Microorganisms

School of Biotechnology and Biomolecular Sciences and Centre for Marine Bio-Innovation, University of New South Wales, Sydney 2052, Australia
*
Author to whom correspondence should be addressed.
Mar. Drugs 2010, 8(3), 438-459; https://doi.org/10.3390/md8030438
Submission received: 18 January 2010 / Revised: 3 February 2010 / Accepted: 22 February 2010 / Published: 1 March 2010
(This article belongs to the Special Issue Bioactive Compounds from Marine Microbes)

Abstract

:
While the oceans cover more than 70% of the Earth’s surface, marine derived microbial natural products have been largely unexplored. The marine environment is a habitat for many unique microorganisms, which produce biologically active compounds (“bioactives”) to adapt to particular environmental conditions. For example, marine surface associated microorganisms have proven to be a rich source for novel bioactives because of the necessity to evolve allelochemicals capable of protecting the producer from the fierce competition that exists between microorganisms on the surfaces of marine eukaryotes. Chemically driven interactions are also important for the establishment of cross-relationships between microbes and their eukaryotic hosts, in which organisms producing antimicrobial compounds (“antimicrobials”), may protect the host surface against over colonisation in return for a nutrient rich environment. As is the case for bioactive discovery in general, progress in the detection and characterization of marine microbial bioactives has been limited by a number of obstacles, such as unsuitable culture conditions, laborious purification processes, and a lack of de-replication. However many of these limitations are now being overcome due to improved microbial cultivation techniques, microbial (meta-) genomic analysis and novel sensitive analytical tools for structural elucidation. Here we discuss how these technical advances, together with a better understanding of microbial and chemical ecology, will inevitably translate into an increase in the discovery and development of novel drugs from marine microbial sources in the future.

1. Introduction

The number of natural products, discovered from various living organisms including plants, animals and microbes, to date exceeds 1 million [1], with the majority (40–60%) derived from terrestrial plants. Of these natural products, 20–25% possess various bioactive properties including antibacterial, antifungal, antiprotozoal, antinematode, anticancer, antiviral and anti-inflammatory activities.
Plants and plant extracts have been used for the treatment of human diseases for millennia, and their use has been recorded in the most ancient archaeological sources. In contrast, the exploration of microorganisms as producers of therapeutical agents only began in the 20th century [2]. However, despite this relatively short history, nearly 10% of all currently known biologically active natural products are of microbial origin. These include the majority of antibiotics, clearly demonstrating the potential of microorganisms as an emerging source for the production of biologically active products. Indeed, by the 20th century microbially derived bioactives had become the foundation of modern pharmaceuticals [3]. For example, the production of antimicrobials is observed in 30–80% of actinomycete and fungal strains screened in various studies [4,5]. Moreover, mathematical models predict that the number of undiscovered antibiotics from actinomycetes could be in the order of 107 [6].
An emerging source of new bioactives may result from the many recent studies of microbial diversity in the marine environment, particularly those microbes associated with marine plants and animals [716]. Several studies have demonstrated that “living surfaces” represent an environment rich in epibiotic microorgansims that produce bioactives [1720]. Nevertheless, the vast biotechnological potential of marine epibiotic microorgansims remains mostly unexplored [21]. This review discusses the importance of exploring new sources potentially rich in bioactives, and highlights the significance of considering the chemical ecology of marine microorganism-host associations for the targeted isolation of bioactive producing microorganisms.

2. Exploring the Under-Explored–Marine Microorganisms as a Source of New Drugs

In the past, it was presumed that the marine environment was a “desert” with scarcity of life forms [22]. However, it is now clear that the oceans are thriving with tremendous diversity of living microorganisms, with cell counts of 106–109 cells per milliliter [7,8] and levels of species diversity and richness predicted to exceed many of the Earths rainforests [2326]. This microbial diversity is presumed to translate into metabolic diversity resulting in the potential for new bioactives to be discovered. Indeed, in the past decades we have witnessed an increase in the number of marine natural products a large proportion of which are of microbial origin [27] (MarinLit database, University of Canterbury: http://www.chem.canterbury.ac.nz/marinlit/marinlit.shtml). For example, in 2007 there was a significant increase (38%), compared to the preceding year, in the number of new marine microbially derived compounds [27]. In addition to structural variety, bioactives obtained from marine microorganisms are known for their broad range of biological effects, which include antimicrobial, antiprotozoan, antiparasitic, and antitumour activities [2835], as well as antifouling activities that prevent the surface-settlement of various marine organisms [3638]. Many of these compounds are noted for their high potency, which could be related to the need to overcome the dilution of allelochemicals in the seawater [26,39].
Use of bioactive producing marine eukaryotes in large-scale production faces many difficulties mainly due to the fact, that, in many cases, the eukaryotic organism is killed in the process of obtaining the bioactive, and many of these eukaryotes cannot be cultured in laboratory, but need to be hand-picked by SCUBA [40]. It also raises the issue of sustainability of these organisms in the nature. In contrast, many bioactive producing marine microorganisms can be easily cultured and manipulated in bioreactors and, therefore, represent the best renewable source of biologically active compounds [41].

2.1. Marine Surface Associated Microoorganisms

The marine environment is a complex ecosystem with an enormous diversity of different life forms often existing in close associations. Among these, microorganism-eukaryote associations have gained significant attention in the past decade [21].
The surfaces of all marine eukaryotes are covered with microbes that live attached in diverse communities often embedded in a matrix, forming a biofilm. The microbial consortia living on various eukaryotes differ significantly from each other and from the microorganisms living in the surrounding seawater [916,42,43]. For example, a DGGE based comparison of the microbial community composition of the coral Montastraea franksi and the surrounding seawater revealed almost no overlap [42]. Likewise, Longford et al., [10] found only two (out of a total of one hundred) bacterial species that were common to three different marine sessile eukaryotes. Moreover, host specificity has also been illustrated by studies that have shown the presence of unique stable communities living on geographically distant individuals belonging to the same species [14,44].
In contrast to free living planktonic microorganisms, which often encounter fluctuations in environmental conditions, requiring quick short-term adaptive responses, surface associated microorganisms supposedly have developed more specialised and stable adaptations, specific to the microenvironment created by a particular host. The high level of specificity of microbial communities on various marine eukaryotes highlights the existence of close cross-relationships between microbial epibionts and their eukaryotic hosts. In fact, some epibiotic microorganisms have been shown to be essential for the normal life and development of the eukaryote, for example, being involved in the development of host morphology [4549]. Host specific bacteria have also been shown to be vertically transferred from the parental eukaryotic organism to its offspring, which indicates the importance of these bacteria for the host. Such inheritance of members of the microbial community has been reported to occur in sponges [16,50,51], bivalves [52] and ascidians [53]. While the exact nature of the relationship between the microorganisms and their hosts remains unclear, it has been hypothesized that the microbial partners construct chemical microenvironments with the eukaryotic host and live in syntrophy, participating in cycling of nutrients, as well as preventing predation of the host via the production of bioactive molecules [50].
Marine surface associated bacteria are often metabolically linked with their host. For example, epiphytes belonging to the Roseobacter lineage are known to degrade the algal osmolyte dimethylsulfoniopropionate (DMSP) yielding the climate-relevant gas dimethylsulfide (DMS) and are regarded as major players in sulphur cycling in the ocean [54,55]. In contrast, some marine eukaryotes heavily rely on the metabolites produced by their microbial symbionts to survive. For example, some marine sponges use the carbon produced by their associated photosynthetic cyanobacteria [56] and may even rely on their autotrophic cyanobacterial symbionts to provide more than 50% of their energy requirements, which allows them to grow in low-nutrient environments [57].
Close metabolic associations between microorganisms and their host can make it difficult to reveal which partner organism is responsible for the production of a particular metabolite. As a result, many bioactive products, previously ascribed to the eukaryotes, have later been found to be produced by associated microorganisms [5866]. For example, the microbial origin of the cytotoxic compound bryostatin was demonstrated by the identification of polyketide synthase genes, involved in its biosynthesis, in the genome of the bryozoan bacterial symbiont “Candidatus Endobugula sertula” [67]. Moreover, it was proposed that microbially derived bryostatin, found on the larvae of bryozoan Bugula neritina, serves to defend the larvae against potential predators [68].
The interactions between the epibiotic microorganisms and their host, in which microorganisms are thought to acquire nutrients from the eukaryote, while the host benefits from the wide range of bioactives produced by its associated microorganism, seems to be widespread in the marine environment [69]. For example, the gamma-proteobacterium Pseudoalteromonas tunicata, known for the production of several bioactive compounds, is proposed to play a role in defending the host against surface colonisation by producing antimicrobial, antilarval and antiprotozoan compounds [7073]. Likewise, the surfaces of the healthy embryos of the lobster Homarus americanus are covered almost exclusively by a single gram-negative bacterium, that produces an antifungal compound highly effective against the fungus Lagenidium callinectes, a common pathogen of many crustaceans [74].
The production of antimicrobials by epiphytic microorganisms could also give the producers a distinct advantage in competition with other surface-dwelling microbes. This is especially important given the fierce competition that exists on the surfaces of marine living organisms, that are relatively rich in nutrients compared to seawater, and, therefore, are attractive for numerous microorganisms [21].
The fact that many marine microorganism-host associations are based on metabolic or chemical interactions may explain the abundance of bioactive producing bacteria on living surfaces [1719,75,76]. Thus, a better understanding of the ecological challenges and the underlying mechanisms involved in such interactions should accelerate the search for novel bioactives.

3. The Challenges of Microbial Bioactive Natural Product Development from Marine Epibiotic Microorganisms

The general procedure for the isolation of natural products from marine epibiotic microorganisms includes several essential steps. The process begins with the isolation of microorganisms from the environment. Often, in the past, the isolation of microorganisms has been a random process. However there is now a growing recognition that the source of microbial samples can be important for increasing the success rate of bioactive discovery [17,20]. Thus, as discussed above, due to the various and often chemically mediated interactions that occur between microorganisms and their host and between members of the epibiotic community, isolation of microorganisms from marine living surfaces can significantly increase the chances of obtaining bioactive producing strains. After growing the microorganisms in the laboratory on nutritional media, the screening of individual isolates for biological activity is performed, for example, based on the inhibition of growth of microorganisms surrounding the test organism in the case of antimicrobials. The phylogenetic and phenotypic identification of the bioactive producing organism is then performed as the first de-replication to ensure that the organism has not been previously used for the particular activity and, subsequently, to maximise the possibility of finding a novel bioactive compound. The extraction and purification of biologically active compounds are then performed, followed by chemical structure elucidation. At this stage a second de-replication can be done to exclude already known compounds. Once novel compounds are identified, the various growth conditions of the producer organism can be assessed to optimize their production. Finally, compounds are assessed for use in the treatment of certain diseases [77,78] and in a variety of industrial settings (Figure 1).

3.1. Improving the Culturability and Production of Bioactives from Marine Microorganisms

The limited ability to culture the majority of environmental strains represents a major bottleneck in classical culture-based screening programs for microbial derived bioactives, including those from marine surfaces. It is estimated that the majority (98–99%) of microorganisms cannot be cultured by traditional techniques [7981]. Nevertheless, marine living surfaces may provide an advantage as, in some cases, a higher percentage of eukaryote associated microorganisms can be readily cultured [82].
Being able to grow the organism in vitro provides great advantages, such as better access to its physiology [83,84]. This may allow the manipulation of different growth parameters to achieve the maximum yield of various products and for their large-scale production via fermentation. Among strategies to improve the culturability of microorganisms, those that attempt to grow the organisms under conditions that mimic the physical and chemical parameters of their natural environment, have been the most successful. For example, by using specialised environmental chambers, Kaeberlein et al., [85] could successfully culture up to 40% of of all microbial cells present in a marine environmental sample.
The close associations often present between marine eukaryotic organisms and their microbial epibionts, clearly impose conditions that are difficult to replicate with standard laboratory procedures. It has, therefore, been proposed that the development of suitable culturing techniques for such organisms should involve conditions that reflect the microenvironment created by their host [86]. This approach has proven successful for the isolation of the sponge associated bacterium Oscillatoria spongeliae [87], for which hyperosmotic medium, resembling the osmolarity of the sponge mesophyle, was used to cultivate the organism.
Cultivation conditions, such as temperature, aeration, pH of the media, incubation time and media composition, can affect the production of the desired metabolite, and, therefore, must be taken into account and fine-tuned [8894]. Usually this requires a producer strain to be grown in the conditions optimal for the production of the active compound. These conditions can differ significantly from the optimal growth conditions of the strain. In some cases, the producer organism is grown under a variety of conditions in parallel and the differences in metabolic spectra are assessed [9597]. Marine surface associated microorganisms may also require conditions that resemble their native environment in order to produce the maximum amount of bioactives. For example, several studies have shown an increase in the production of antimicrobial compounds when the surface associated bacteria were grown, in vitro, to form surface attached biofilms [41,98,99]. In addition, Okazaki et al., [100] have shown that marine isolate SS-228 was able to produce the antibiotic compound only when the growth medium was supplemented with powdered Laminaria seaweeds, common in the habitat from which strain SS-228 was obtained. It is now becoming clear that knowledge of a microorgansims’ natural habitat, including the specifics of the host organism, can improve production of microbial derived bioactives.
Sequence information obtained via the sequencing of the environmental DNA (“metagenome”) can greatly assist in understanding the metabolic potential of the organisms present in the environment, and thus guide the development of specialised cultivation conditions. For example Tyson et al., [101] successfully cultured Leptospirillum ferrodiazotrophum by developing a selective isolation strategy. The predicted nitrogen fixing capability of this organism, based on the sequence information of an acid mine drainage biofilm community, underlined the development of that strategy.
Over the past decade genomics has emerged as an alternative to directly culturing microorganisms for the isolation of new bioactives. In particular, functional metagenomics was first developed to tackle the biotechnological potential of unculturable microorganisms. In this approach the DNA obtained from the environment (“environmental DNA”) is inserted into a host organism, such as E. coli, and a functional screen of libraries is performed to detect the desired activity in the clones [102107]. Some of the successes of this approach were the discovery of terragine A [108], bioactive N-acyl-tyrosine derivatives [109] as well as indirubin [110] from the soil metagenomes.
Functional (meta-)genomics can provide an insight into the genes and gene clusters involved in the production of certain metabolites, and, thus, provide information about the possible biosynthetic pathway leading to that metabolite [111]. Such an approach has been used by Burke et al., [112] to propose the biosynthetic pathway of the antifungal compound tambjamine produced by the marine bacterium P. tunicata. Recently the same approach was successful in identifying two positive clones from a P. tunicata genome library, with different modes of action against the nematode Caenorhabditis elegans (Ballestriero et al., unpublished data). In addition, a functional (meta-) genomics approach provides an advantage for further purification of the bioactive compound produced by a clone, since the “extra” metabolite can be relatively easily pinpointed by using a reference non-bioactive-producing clone [113].
However, despite some success, currently the hit rate of using metagenomic functional screening to obtain bioactive producing clones generally remains low, in the order of 1 in 10,000 [109,114], or even as low as 1 in 730,000 [115] clones screened. This low hit rate is mainly a result of the limited ability of host expression strains to express compounds of foreign origin [78,116,117]. Therefore it is possible that with improvements of such strains the hit rate for positive clones in metagenomics functional screens will greatly increase. For example, recent studies have demonstrated that the use of a variety of host expression strains can assist in the expression of the desired metabolite [118121]. Specifically, these strains are often chosen based on possible similarities with the producer bacteria (if known), such as, for example, the similarities in codon usage, as well as presence of specific machinery necessary for the production of particular metabolites [119,120].
Alternatively, shotgun sequencing of environmental DNA and subsequent data analysis has the potential to identify genes encoding new structures of known compound classes, e.g., polyketide synthases (PKS) and non-ribosomal peptide synthetases (NRPS) usually involved in production of bioactive secondary metabolites [122]. For example, recent analysis of the genome of P. tunicata has revealed the presence of nine NRPS [123]. Two products of these NRPS with predicted biological activity have been recently identified in the laboratory via heterologous expression, and their presence was confirmed in the original strain of P. tunicata by varying its conditions of growth [124].
Moreover, the availability of sequence data from a variety of microorganisms has further highlighted the importance of developing culturing conditions that would be suitable for the production of bioactives. There are now several examples of genes involved in the biosynthesis of bioactives found in non-bioactive producing organisms, suggesting that, given suitable growth conditions, these organisms have the potential to produce bioactive metabolites. For example, the genome of the myxobacterium Stigmatella aurantiaca DW4/3-1 showed the presence of multiple polyketide synthase/non-ribosomal peptide synthetase gene clusters, which had not been previously observed in this organism [125]. In another example, the previously unknown potential of the production of the antitumour compound terrequinone A in the fungus Aspergillus nidulans was revealed [126].

3.2. De-Replication

In the early years of antibiotic discovery the selection of antibiotic producer strains was based on morphology rather than genotypes, resulting in redundancy in many natural product extract libraries [127,128]. The initial success in the discovery of many antibiotic compounds from natural sources included thousands of compounds being described within a few decades. However, the lack of a systematic approach often resulted in the frequent re-discovery of known compounds. Therefore, it is important to put considerable effort into the de-replication process for early detection of both the known producer organisms as well as the known bioactive compounds.
Currently, 16S rRNA gene sequencing [129] and phenotypic characterisation can serve for the identification of producer organisms, and, hence, may reveal whether a given microorganism or its close relatives have been previously known to produce certain bioactives, in order to focus the efforts on organisms with yet uncharacterised activity.
Concerning the early detection of known bioactive compounds, advances in the development of new chemical analysis techniques, coupled with a database evaluation, can serve as tools for rapid detection of known compounds requiring only small quantities of sample and/or minimal efforts in sample preparation [130132]. Hence, they may greatly assist in preventing the waste of resources, which would otherwise be necessary for scale-up and characterisation of the bioactive compounds.
The exploration of relatively unexplored environments can also assist in finding novel microorganisms and chemical structures, and, hence, minimise the re-discovery of known compounds. The oceans have proven to be a habitat for many unique microbes [133], such as, for example, the recently discovered marine genera Salinispora and Marinophilus [133,134]. In addition, some of the members of these groups were found to produce structurally novel bioactive metabolites, for example, salinosporamides, a family of compounds with cytotoxic activity, were successfully isolated from Salinnispora tropica [135]. Likewise, the structurally novel compounds marineosins [136] and largazole [137] have been recently isolated from marine bacteria belonging to the actinomycete and cyanobacterial groups respectively. Recently discovered marine bioactive natural products including compounds with unique structures are reviewed in [138]. This supports the idea that the unique chemical and physical parameters of the marine environment can lead to the evolution of life forms that could also produce metabolites with novel chemical scaffolds [139,140].

3.3. Purification of Bioactives from Crude Extracts

Despite their potential, full characterisation of marine microbially derived bioactives, as well as the development of extraction and purification strategies can be a long and laborious process requiring a great deal of manual work, with little room for automation [78,141].
Purification and structure elucidation of mass limited sample material is considered a major bottleneck. This is usually because the compound of interest often represents less than 1% of the crude extract, which, in most cases, is a mixture of hundreds of different compounds. Therefore, every extract has its unique combination of “contaminants” necessitating a specific approach; as a result, development of the purification strategies remains largely experimental [142].
Furthermore, obtaining adequate quantities of bioactive compounds, necessary for structure elucidation and evaluation, usually requires extensive optimization of conditions and scale-up [143].
To facilitate the chemical characterisation, analytical methods are constantly being developed and improved, one of the major lines of improvement being the possibility of using small quantities of sample and easy sample preparation. For example, the recently developed Ultra High Performance Liquid Chromatography (UPLC) coupled to high resolution mass spectrometers (MS), as well as capillary probe nuclear magnetic resonance spectrometers have greatly assisted the process of natural product discovery from mass limited samples [130,144,145]. Likewise, the newly developed Desorption Electrospray Ionisation Mass Spectrometry (DESI-MS) technique allows for the rapid detection of the compounds requiring minimal effort to be spent on sample preparation [131,132]. These techniques may soon eliminate, or, by some estimates, have already eliminated the purification and chemical characterisation step as a major drawback in natural product discovery [146150].

4. Is There an Alternative to Natural Product Discovery and Development?

The challenges of natural product research have resulted in a search for an alternative to bioactive product development.
Combinatorial biosynthesis has emerged as one of the alternatives and is defined as “the application of genetic engineering to modify biosynthetic pathways to natural products in order to produce new and altered structures using nature’s biosynthetic machinery” [151]. It involves the use of genes from different biosynthetic pathways, in various combinations, in order to generate libraries of hybrid structures. However, in practice, this approach is rather problematic. Firstly, it involves the construction of various mutant organisms and, therefore, is very labour-intensive and costly. Secondly, it often relies on the low substrate specificity of enzymes in the biosynthetic pathways, which is not always the case as many enzymes are rather specific [151].
High-throughput screening (HTS) of synthetic chemical libraries is also regarded as an alternative to bioactive discovery and the development of combinatorial chemistry has allowed for smaller, more drug-like libraries to be screened against defined macromolecular targets. Furthermore, an increase in the availability of genomic data has provided more potential targets for these screens [142,152157]. However, the first libraries of chemically synthesised compounds provided more quantity than quality; some produced more than million compounds, but were a disappointment, as they yielded very low numbers of, or no, active compounds [142,158]. For example, GlaxoSmithKline ( http://www.gsk.com) have recently disclosed the results of a six-year campaign to discover broad-spectrum antibiotics that was abandoned because of the limited chemical diversity of their synthetic screening libraries [157]. These approaches obviously failed to fulfil initial expectations [159164], and are unlikely to substitute the benefits of natural product development.
In contrast to chemical libraries, bioactives of natural product origin provide a diversity and a structural complexity with densely packed functional groups allowing maximum selectivity and interaction with the target [77,165,166]. Such complexity makes the chemical synthesis of these compounds extremely difficult [146,172176]. Nevertheless there has been some success with the total synthesis of natural products from marine microorganisms, such as, for example, dinoflagellate toxins azaspiracids and brevetoxins, and the antibiotic compound uncialamycin produced by marine Streptomyces [167171]. It has been suggested that the success of natural compounds is due to the fact that they have undergone natural selection and, therefore, are best suited to perform their activities [142,160,177]. Thus, further research on bioactive natural products may provide a source of new chemical structures that can guide the design of novel chemical compounds [178,179], as well as reveal yet unknown modes of action [180].
The majority of antibiotics currently used in clinical practice are of natural product origin [77,161,181]. For example, 70 out of the 90 antibiotics marketed in the years 1982–2002 originated from natural products [161]. Notably, the quinolones or fluoroqinones, one of the most successful classes of synthetic antibiotics, are also based on the structure of the natural product quinine [182]. In fact, chemical modifications based on a natural product scaffold is a widely used approach in modifying the chemical and physical properties of the molecule thus making it useful for a particular pharmacological application [183,184]. For example, Jenkins et al., [185] have recently synthesised four new chemical scaffolds useful for drug development based on novel structures of a number of bioactive natural products such as the histrionicotoxins isolated from the skin of the Colombian poison dart frog, Dendrobates histrionicus [185]. Likewise, the chemically synthesised analogs of epothilones—compounds produced by myxobacterium Sorangium cellitlosum [186], have shown an increased potency against tumour cells compared to the original natural product [187,188].

5. Conclusions

Despite the challenges, the search and development of natural products remain an indispensable and unparalleled source of biologically active compounds. Thus, research into the diversity of bioactive natural products justifies the resources invested due to the lack of equivalent alternatives in synthetic compounds [142,178]. Microorganisms are currently accepted as the best renewable source for bioactives, and the exploration of yet underexplored sources, such as the marine living-surface habitat, has a great potential to deliver novel bioactive producing microbes useful for further drug development. Moreover, a systematic approach that takes into consideration unique ecological relationships in the marine environment, such as those discussed in this review, can greatly assist in maximizing the output of obtaining novel bioactive producing organisms and, thus, may prevent the frequent re-discovery of known compounds and the waste of resources that would be necessary for largescale high-throughput screens.

Acknowledgements

The authors thank Tilmann Harder for helpful discussions and review of the manuscript. The authors also thank Australian Research Council Linkage Grant # LP04554677 and the Centre for Marine Bio-Innovation for research funding support.

References

  1. Berdy, J. Bioactive microbial metabolites. A personal view. J Antibiot (Tokyo) 2005, 58, 1–26. [Google Scholar]
  2. Monaghan, RL; Tkacz, JS. Bioactive microbial products: Focus upon mechanism of action. Annu Rev Microbiol 1990, 44, 271–301. [Google Scholar]
  3. Capon, RJ. Marine bioprospecting–Trawling for treasure and pleasure. European J Org Chem 2001, 633–645. [Google Scholar]
  4. Basilio, A; Gonzalez, I; Vicente, MF; Gorrochategui, J; Cabello, A; Gonzalez, A; Genilloud, O. Patterns of antimicrobial activities from soil actinomycetes isolated under different conditions of pH and salinity. J Appl Microbiol 2003, 95, 814–823. [Google Scholar]
  5. Pelaez, F; Genilloud, O. Discovering new drugs from microbial natural products. In Microorganisms for Health Care, Food and Enzyme Production; Barredo, JL, Ed.; Research Signpost: Trivendrum, India, 2003; pp. 1–23. [Google Scholar]
  6. Watve, MG; Tickoo, R; Jog, MM; Bhole, BD. How many antibiotics are produced by the genus Streptomyces. Arch Microbiol 2001, 176, 386–390. [Google Scholar]
  7. Rheinheimer, G. Aquatic microbiology; Wiley: Chichester; New York, NY, USA, 1992. [Google Scholar]
  8. Fenical, W; Jensen, PR. Developing a new resource for drug discovery: Marine actinomycete bacteria. Nat Chem Biol 2006, 2, 666–673. [Google Scholar]
  9. Perez-Matos, AE; Rosado, W; Govind, NS. Bacterial diversity associated with the Caribbean tunicate Ecteinascidia turbinata. Antonie van Leeuwenhoek 2007, 92, 155–164. [Google Scholar]
  10. Longford, SR; Tujula, NA; Crocetti, GR; Holmes, AJ; Holmstrom, C; Kjelleberg, S; Steinberg, PD; Taylor, MW. Comparisons of diversity of bacterial communities associated with three sessile marine eukaryotes. Aquat Microb Ecol 2007, 48, 217–229. [Google Scholar]
  11. Santiago-Vazquez, LZ; Bruck, TB; Bruck, WM; Duque-Alarcon, AP; McCarthy, PJ; Kerr, RG. The diversity of the bacterial communities associated with the azooxanthellate hexacoral Cirrhipathes lutkeni. ISME J 2007, 1, 654–659. [Google Scholar]
  12. Rohwer, F; Seguritan, V; Azam, F; Knowlton, N. Diversity and distribution of coral-associated bacteria. Mar Ecol Prog Ser 2002, 243, 1–10. [Google Scholar]
  13. Martinez-Garcia, M; Diaz-Valdes, M; Wanner, G; Ramos-Espla, A; Anton, J. Microbial community associated with the colonial ascidian Cystodytes dellechiajei. Environ Microbiol 2007, 9, 521–534. [Google Scholar]
  14. Webster, NS; Bourne, D. Bacterial community structure associated with the Antarctic soft coral Alcyonium antarcticum. FEMS Microbiol Ecol 2007, 59, 81–94. [Google Scholar]
  15. Bourne, DG; Munn, CB. Diversity of bacteria associated with the coral Pocillopora damicornis from the Great Barrier Reef. Environ Microbiol 2005, 7, 1162–1174. [Google Scholar]
  16. Enticknap, JJ; Kelly, M; Peraud, O; Hill, RT. 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. [Google Scholar]
  17. Penesyan, A; Marshall-Jones, Z; Holmstrom, C; Kjelleberg, S; Egan, S. Antimicrobial activity observed among cultured marine epiphytic bacteria reflects their potential as a source of new drugs: Research article. FEMS Microbiol Ecol 2009, 69, 113–124. [Google Scholar]
  18. 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. [Google Scholar]
  19. Muscholl-Silberhorn, A; Thiel, V; Imhoff, J. Abundance and bioactivity of cultured sponge-sssociated bacteria from the Mediterranean sea. Microb Ecol 2008, 55, 94–106. [Google Scholar]
  20. Burgess, JG; Jordan, EM; Bregu, M; Mearns-Spragg, A; Boyd, KG. Microbial antagonism: A neglected avenue of natural products research. J Biotechnol 1999, 70, 27–32. [Google Scholar]
  21. Egan, S; Thomas, T; Kjelleberg, S. Unlocking the diversity and biotechnological potential of marine surface associated microbial communities. Curr Opin Microbiol 2008, 11, 219–225. [Google Scholar]
  22. Zobell, CE. Marine microbiology: A Monograph on Hydrobacteriology; Chronica Botanica Co.: Waltham, MA, USA, 1946. [Google Scholar]
  23. Sogin, ML; Morrison, HG; Huber, JA; Welch, DM; Huse, SM; Neal, PR; Arrieta, JM; Herndl, GJ. Microbial diversity in the deep sea and the underexplored “rare biosphere”. Proc Natl Acad Sci USA 2006, 103, 12115–12120. [Google Scholar]
  24. Stach, JEM; Bull, AT. Estimating and comparing the diversity of marine actinobacteria. Antonie van Leeuwenhoek 2005, 87, 3–9. [Google Scholar]
  25. Holler, U; Wright, AD; Matthee, GF; Konig, GM; Draeger, S; Aust, HJ; Schulz, B. Fungi from marine sponges: Diversity, biological activity and secondary metabolites. Mycol Res 2000, 104, 1354–1365. [Google Scholar]
  26. Haefner, B. Drugs from the deep: Marine natural products as drug candidates. Drug Discov Today 2003, 8, 536–544. [Google Scholar]
  27. Blunt, JW; Copp, BR; Hu, WP; Munro, MHG; Northcote, PT; Prinsep, MR. Marine natural products. Nat Prod Rep 2009, 26, 170–244. [Google Scholar]
  28. Andrighetti-Frohner, CR; Antonio, RV; Creczynski-Pasa, TB; Barardi, CRM; Simoes, CMO. Cytotoxicity and potential antiviral evaluation of violacein produced by Chromobacterium violaceum. Mem Inst Oswaldo Cruz 2003, 98, 843–848. [Google Scholar]
  29. Lichstein, HC; Van de Sand, VF. Violacein, an antibiotic pigment produced by Chromobacterium violaceum. J Infect Dis 1945, 76, 47–51. [Google Scholar]
  30. Matz, C; Deines, P; Boenigk, J; Arndt, H; Eberl, L; Kjelleberg, S; Jurgens, K. Impact of violacein-producing bacteria on survival and feeding of bacterivorous nanoflagellates. Appl Environ Microbiol 2004, 70, 1593–1599. [Google Scholar]
  31. James, SG; Holmstrom, C; Kjelleberg, S. Purification and characterization of a novel antibacterial protein from the marine bacterium D2. Appl Environ Microbiol 1996, 62, 2783–2788. [Google Scholar]
  32. Franks, A; Haywood, P; Holmstrom, C; Egan, S; Kjelleberg, S; Kumar, N. Isolation and structure elucidation of a novel yellow pigment from the marine bacterium Pseudoalteromonas tunicata. Molecules 2005, 10, 1286–1291. [Google Scholar]
  33. Feling, RH; Buchanan, GO; Mincer, TJ; Kauffman, CA; Jensen, PR; Fenical, W. Salinosporamide A: A highly cytotoxic proteasome inhibitor from a novel microbial source, a marine bacterium of the new genus Salinospora. Angewandte Chemie–International Edition 2003, 42, 355–357. [Google Scholar]
  34. Ratnayake, R; Lacey, E; Tennant, S; Gill, JH; Capon, RJ. Kibdelones: Novel anticancer polyketides from a rare Australian actinomycete. Chemistry–A European Journal 2007, 13, 1610–1619. [Google Scholar]
  35. Fremlin, LJ; Piggott, AM; Lacey, E; Capon, RJ. Cottoquinazoline A and cotteslosins A and B, metabolites from an Australian marine-derived strain of Aspergillus versicolor. J Nat Prod 2009, 72, 666–670. [Google Scholar]
  36. Xu, Y; He, H; Schulz, S; Liu, X; Fusetani, N; Xiong, H; Xiao, X; Qian, PY. Potent antifouling compounds produced by marine. Streptomyces Bioresource Technol 2010, 101, 1331–1336. [Google Scholar]
  37. Dash, S; Jin, C; Lee, OO; Xu, Y; Qian, PY. Antibacterial and antilarval-settlement potential and metabolite profiles of novel sponge-associated marine bacteria. J Ind Microbiol Biotechnol 2009, 36, 1047–1056. [Google Scholar]
  38. Xiong, H; Qi, S; Xu, Y; Miao, L; Qian, P.-Y. Antibiotic and antifouling compound production by the marine-derived fungus Cladosporium sp. F14. J Hydro-environ Res 2009, 2, 264–270. [Google Scholar]
  39. Zhang, L; An, R; Wang, J; Sun, N; Zhang, S; Hu, J; Kuai, J. Exploring novel bioactive compounds from marine microbes. Curr Opin Microbiol 2005, 8, 276–281. [Google Scholar]
  40. Molinski, TF; Dalisay, DS; Lievens, SL; Saludes, JP. Drug development from marine natural products. Nat Rev Drug Discov 2009, 8, 69–85. [Google Scholar]
  41. Sarkar, S; Saha, M; Roy, D; Jaisankar, P; Das, S; Gauri Roy, L; Gachhui, R; Sen, T; Mukherjee, J. Enhanced production of antimicrobial compounds by three salt-tolerant actinobacterial strains isolated from the Sundarbans in a niche-mimic bioreactor. Mar Biotechnol 2008, 10, 518–526. [Google Scholar]
  42. Rohwer, F; Breitbart, M; Jara, J; Azam, F; Knowlton, N. Diversity of bacteria associated with the Caribbean coral Montastraea franksi. Coral Reef 2001, 20, 85–91. [Google Scholar]
  43. Dobretsov, S; Dahms, HU; Tsoi, MY; Qian, PY. Chemical control of epibiosis by Hong Kong sponges: The effect of sponge extracts on micro- and macrofouling communities. Mar Ecol Prog Ser 2005, 297, 119–129. [Google Scholar]
  44. Taylor, MW; Schupp, PJ; Dahllof, I; Kjelleberg, S; Steinberg, PD. Host specificity in marine sponge-associated bacteria, and potential implications for marine microbial diversity. Environ Microbiol 2004, 6, 121–130. [Google Scholar]
  45. Provasoli, L; Pintner, IJ. Effect of media and inoculum on morphology of. Ulva J Phycol 1977, 13, 56–56. [Google Scholar]
  46. Provasoli, L; Pintner, IJ. Bacteria induced polymorphism in an axenic laboratory strain of Ulva lactuca (Chlorophceae). J Phycol 1980, 16, 196–201. [Google Scholar]
  47. Nakanishi, K; Nishijima, M; Nishimura, M; Kuwano, K; Saga, N. Bacteria that induce morphogenesis in Ulva pertusa (Chlorophyta) grown under axenic conditions. J Phycol 1996, 32, 479–482. [Google Scholar]
  48. Nakanishi, K; Nishijima, M; Nomoto, AM; Yamazaki, A; Saga, N. Requisite morphologic interaction for attachment between Ulva pertusa (Chlorophyta) and symbiotic bacteria. Mar Biotechnol 1999, 1, 107–111. [Google Scholar]
  49. Matsuo, Y; Suzuki, M; Kasai, H; Shizuri, Y; Harayama, S. Isolation and phylogenetic characterization of bacteria capable of inducing differentiation in the green alga Monostroma oxyspermum. Environ Microbiol 2003, 5, 25–35. [Google Scholar]
  50. Sharp, KH; Eam, B; John Faulkner, D; Haygood, MG. Vertical transmission of diverse microbes in the tropical sponge Corticium sp. Appl Environ Microbiol 2007, 73, 622–629. [Google Scholar]
  51. Schmitt, S; Weisz, JB; Lindquist, N; Hentschel, U. Vertical transmission of a phylogenetically complex microbial consortium in the viviparous sponge Ircinia felix. Appl Environ Microbiol 2007, 73, 2067–2078. [Google Scholar]
  52. Cary, SC. Verticial transmission of a chemoautotrophic symbiont in the protobranch bivalve Solemya reidi. Mol Mar Biol Biotechnol 1994, 3, 121–130. [Google Scholar]
  53. Hirose, E; Fukuda, T. Vertical transmission of photosymbionts in the colonial ascidian Didemnum molle: The larval tunic prevents symbionts from attaching to the anterior part of larvae. Zool Sci 2006, 23, 669–674. [Google Scholar]
  54. Buchan, A; Gonzalez, JM; Moran, MA. Overview of the marine Roseobacter lineage. Appl Environ Microbiol 2005, 71, 5665–5677. [Google Scholar]
  55. Wagner-Dobler, I; Biebl, H. Environmental Biology of the Marine Roseobacter Lineage. Annu Rev Microbiol 2006, 60, 255–280. [Google Scholar]
  56. Wilkinson, CR. Nutrient translocation from symbiotic cyanobacteria to coral reef sponges. In Biologie des Spongiaires; Levi, C, Boury-Esnault, N, Eds.; Colloques Internationionaux du Centre National de la Recherche Scientifique: Paris France, 1979; pp. 373–380. [Google Scholar]
  57. Wilkinson, CR. Net primary productivity in coral reef sponges. Science 1983, 219, 410–412. [Google Scholar]
  58. Stierle, AC; Cardellina Ii, JH; Singleton, FL. A marine Micrococcus produces metabolites ascribed to the sponge. Tedania ignis Experientia 1988, 44, 1021. [Google Scholar]
  59. Kobayashi, J; Ishibashi, M. Bioactive metabolites of symbiotic marine microorganisms. Chem Rev 1993, 93, 1753–1769. [Google Scholar]
  60. Unson, MD; Faulkner, DJ. Cyanobacterial symbiont synthesis of chlorinated metabolites from Dysidea herbacea (Porifera). Experientia 1993, 44, 1021–1022. [Google Scholar]
  61. Unson, MD; Holland, ND; Faulkner, DJ. A brominated secondary metabolite synthesized by the cyanobacterial symbiont of a marine sponge and accumulation of the crystalline metabolite in the sponge tissue. Mar Biol 1994, 119, 1–12. [Google Scholar]
  62. Oclarit, JM; 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. [Google Scholar]
  63. Bewley, CA; Holland, ND; Faulkner, DJ. Two classes of metabolites from Theonella swinhoei are localized in distinct populations of bacterial symbionts. Experientia 1996, 52, 716–722. [Google Scholar]
  64. Schmidt, EW; Obraztsova, AY; Davidson, SK; Faulkner, DJ; Haygood, MG. Identification of the antifungal peptide-containing symbiont of the marine sponge Theonella swinhoei as a novel delta-proteobacterium, “Candidatus Entotheonella palauensis”. Mar Biol 2000, 136, 969–977. [Google Scholar]
  65. Schmidt, EW. From chemical structure to environmental biosynthetic pathways: Navigating marine invertebrate-bacteria associations. Trends Biotechnol 2005, 23, 437–440. [Google Scholar]
  66. Konig, GM; Kehraus, S; Seibert, SF; Abdel-Lateff, A; Muller, D. Natural products from marine organisms and their associated microbes. Chembiochem 2006, 7, 229–238. [Google Scholar]
  67. Sudek, S; Lopanik, NB; Waggoner, LE; Hildebrand, M; Anderson, C; Liu, H; Patel, A; Sherman, DH; Haygood, MG. Identification of the putative bryostatin polyketide synthase gene cluster from “Candidatus Endobugula sertula”, the uncultivated microbial symbiont of the marine bryozoan Bugula neritina. J Nat Prod 2007, 70, 67–74. [Google Scholar]
  68. Lopanik, N; Gustafson, KR; Lindquist, N. Structure of bryostatin 20: A symbiont-produced chemical defense for larvae of the host bryozoan Bugula neritina. J Nat Prod 2004, 67, 1412–1414. [Google Scholar]
  69. Harder, T. Marine epibiosis: Concepts, ecological consequences and host defence. In Marine and Industrial Biofouling; Costerton, JW, Ed.; Springer-Verlag: Berlin, Germany, 2009; pp. 219–231. [Google Scholar]
  70. Egan, S; James, S; Holmstrom, C; Kjelleberg, S. Inhibition of algal spore germination by the marine bacterium Pseudoalteromonas tunicata. FEMS Microbiol Ecol 2001, 35, 67–73. [Google Scholar]
  71. Egan, S; James, S; Holmstrom, C; Kjelleberg, S. Correlation between pigmentation and antifouling compounds produced by Pseudoalteromonas tunicata. Environ Microbiol 2002, 4, 433–442. [Google Scholar]
  72. Franks, A; Egan, S; Holmstrom, C; James, S; Lappin-Scott, H; Kjelleberg, S. Inhibition of fungal colonization by Pseudoalteromonas tunicata provides a competitive advantage during surface colonization. Appl Environ Microbiol 2006, 72, 6079–6087. [Google Scholar]
  73. Holmstrom, C; James, S; Neilan, BA; White, DC; Kjelleberg, S. Pseudoalteromonas tunicata sp. nov., a bacterium that produces antifouling agents. Int J Syst Bacteriol 1998, 48, 1205–1212. [Google Scholar]
  74. Gil-Turnes, MS; Fenical, W. Embryos of Homarus americanus are protected by epibiotic bacteria. Biol Bull 1992, 182, 105–108. [Google Scholar]
  75. Lemos, ML; Toranzo, AE; Barja, JL. Antibiotic activity of epiphytic bacteria isolated from intertidal seaweeds. Microb Ecol 1985, 11, 149–163. [Google Scholar]
  76. Wilson, GS; Raftos, DA; Corrigan, SL; Nair, SV. Diversity and antimicrobial activities of surface-attached marine bacteria from Sydney Harbour, Australia. Microbiol Res 2009. [Google Scholar]
  77. Singh, SB; Barrett, JF. Empirical antibacterial drug discovery – Foundation in natural products. Biochem Pharmacol 2006, 71, 1006–1015. [Google Scholar]
  78. Pelaez, F. The historical delivery of antibiotics from microbial natural products – Can history repeat. Biochem Pharmacol 2006, 71, 981–990. [Google Scholar]
  79. Staley, JT; Konopka, A. Measurement of in situ activities of nonphotosynthetic microorganisms in aquatic and terrestrial habitats. Annu Rev Microbiol 1985, 39, 321–346. [Google Scholar]
  80. Pace, NR. A molecular view of microbial diversity and the biosphere. Science 1997, 276, 734–740. [Google Scholar]
  81. Ward, DM; Weller, R; Bateson, MM. 16S rRNA sequences reveal numerous uncultured microorganisms in a natural community. Nature 1990, 345, 63–65. [Google Scholar]
  82. Jensen, PR; Kauffman, CA; Fenical, W. High recovery of culturable bacteria from the surfaces of marine algae. Marine Biology 1996, 126, 1–7. [Google Scholar]
  83. DeLong, EF. The microbial ocean from genomes to biomes. Nature 2009, 459, 200–206. [Google Scholar]
  84. Tripp, HJ; Kitner, JB; Schwalbach, MS; Dacey, JWH; Wilhelm, LJ; Giovannoni, SJ. SAR11 marine bacteria require exogenous reduced sulphur for growth. Nature 2008, 452, 741–744. [Google Scholar]
  85. Kaeberlein, T; Lewis, K; Epstein, SS. Isolating “uncultivabte” microorganisms in pure culture in a simulated natural environment. Science 2002, 296, 1127–1129. [Google Scholar]
  86. Osinga, R; Armstrong, E; Grant Burgess, J; Hoffmann, F; Reitner, J; Schumann-Kindel, G. Sponge-microbe associations and their importance for sponge bioprocess engineering. Hydrobiologia 2001, 461, 55–62. [Google Scholar]
  87. Hinde, R; Pironet, F; Borowitzka, MA. Isolation of Oscillatoria spongeliae, the filamentous cyanobacterial symbiont of the marine sponge Dysidea herbacea. Mar Biol 1994, 119, 99–104. [Google Scholar]
  88. Betina, V. Bioactive secondary metabolites of microorganisms. Prog Ind Microbiol 1994, 30. [Google Scholar]
  89. Aharonowitz, Y. Nitrogen metabolite regulation of antibiotic biosynthesis. Annu Rev Microbiol 1980, 34, 209–233. [Google Scholar]
  90. Barberel, SI; Walker, JRL. The effect of aeration upon the secondary metabolism of microorganisms. Biotechnol Genet Eng Rev 2000, 17, 281–323. [Google Scholar]
  91. Pfefferle, C; Theobald, U; Gurtler, H; Fiedler, H. Improved secondary metabolite production in the genus Streptosporangium by optimization of the fermentation conditions. J Biotechnol 2001, 23, 135–142. [Google Scholar]
  92. Schimana, J; Gebhardt, K; Holtzel, A; Schmid, DG; Sussmuth, R; Muller, J; Pukall, R; Fiedler, HP. Arylomycins A and B new biaryl-bridged lipopeptide antibiotics produced by Streptomyces sp. Tu 6075I. Taxonomy, fermentation, isolation and biological activities. J Antibiot (Tokyo) 2002 55, 565–570.
  93. Saitoh, K; Tenmyo, O; Yamamoto, S; Furumai, T; Oki, T. Pradimicin S, a new pradimicin analog I. Taxonomy, fermentation and biological activities. J Antibiot (Tokyo) 1993, 46, 580–588. [Google Scholar]
  94. Shimada, N; Hasegawa, S; Harada, T. Oxetanocin, a novel nucleoside from bacteria. J Antibiot (Tokyo) 1986, 39, 1623–1625. [Google Scholar]
  95. Bills, GF; Platas, G; Fillola, A; Jimenez, MR; Collado, J; Vicente, F; Martin, J; Gonzalez, A; Bur-Zimmermann, J; Tormo, JR; Pelaez, F. Enhancement of antibiotic and secondary metabolite detection from filamentous fungi by growth on nutritional arrays. J Appl Microbiol 2008, 104, 1644–1658. [Google Scholar]
  96. Minas, W; Bailey, JE; Duetz, W. Streptomycetes in micro-cultures: Growth, production of secondary metabolites, and storage and retrieval in the 96-well format. Anton Leeuwenhoek Int J Gen M 2000, 78, 297–305. [Google Scholar]
  97. Knight, V; Sanglier, JJ; DiTullio, D; Braccili, S; Bonner, P; Waters, J; Hughes, D; Zhang, L. Diversifying microbial natural products for drug discovery. Appl Microbiol Biotechnol 2003, 62, 446–458. [Google Scholar]
  98. Matz, C; Webb, JS; Schupp, PJ; Phang, SY; Penesyan, A; Egan, S; Steinberg, P; Kjelleberg, S. Marine biofilm bacteria evade eukaryotic predation by targeted chemical defense. PLoS ONE 2008, 3, e2744. [Google Scholar]
  99. Yan, L; Boyd, KG; Grant Burgess, J. Surface attachment induced production of antimicrobial compounds by marine epiphytic bacteria using modified roller bottle cultivation. Mar Biotechnol 2002, 4, 356–366. [Google Scholar]
  100. Okazaki, T; Kitahara, T; Okami, Y. Studies on marine microorganisms. IV. A new antibiotic SS-228 Y produced by Chainia isolated from shallow sea mud. J Antibiot (Tokyo) 1975, 28, 176–184. [Google Scholar]
  101. Tyson, GW; Lo, I; Baker, BJ; Allen, EE; Hugenholtz, P; Banfield, JF. Genome-directed isolation of the key nitrogen fixer Leptospirillum ferrodiazotrophum sp. nov. from an acidophilic microbial community. Appl Environ Microbiol 2005, 71, 6319–6324. [Google Scholar]
  102. Handelsman, J. Metagenomics: Application of genomics to uncultured microorganisms. Microbiol Mol Biol Rev 2004, 68, 669–685. [Google Scholar]
  103. Handelsman, J. How to find new antibiotics. Scientist 2005, 19, 20–21. [Google Scholar]
  104. Langer, M; Gabor, EM; Liebeton, K; Meurer, G; Niehaus, F; Schulze, R; Eck, J; Lorenz, P. Metagenomics: An inexhaustible access to nature’s diversity. Biotechnol J 2006, 1, 815–821. [Google Scholar]
  105. Sleator, RD; Shortall, C; Hill, C. Metagenomics. Lett Appl Microbiol 2008, 47, 361–366. [Google Scholar]
  106. Osburne, MS; Grossman, TH; August, PR; MacNeil, IA. Tapping into microbial diversity for natural products drug discovery. ASM News 2000, 66, 411–417. [Google Scholar]
  107. Daniel, R. The soil metagenome–A rich resource for the discovery of novel natural products. Curr Opin Biotechnol 2004, 15, 199–204. [Google Scholar]
  108. Wang, GYS; Graziani, E; Waters, B; Pan, W; Li, X; McDermott, J; Meurer, G; Saxena, G; Andersen, RJ; Davies, J. Novel natural products from soil DNA libraries in a streptomycete host. Org Lett 2000, 2, 2401–2404. [Google Scholar]
  109. Brady, SF; Clardy, J. Long-chain N-acyl amino acid antibiotics isolated from heterologously expressed environmental DNA [20]. J Am Chem Soc 2000, 122, 12903–12904. [Google Scholar]
  110. MacNeil, IA; Tiong, CL; Minor, C; August, PR; Grossman, TH; Loiacono, KA; Lynch, BA; Phillips, T; Narula, S; Sundaramoorthi, R; Tyler, A; Aldredge, T; Long, H; Gilman, M; Holt, D; Osburne, MS. Expression and isolation of antimicrobial small molecules from soil DNA libraries. J Mol Microbiol Biotechnol 2001, 3, 301–308. [Google Scholar]
  111. Uria, A; Piel, J. Cultivation-independent approaches to investigate the chemistry of marine symbiotic bacteria. Phytochem Rev 2009, 1–14. [Google Scholar]
  112. Burke, C; Thomas, T; Egan, S; Kjelleberg, S. The use of functional genomics for the identification of a gene cluster encoding for the biosynthesis of an antifungal tambjamine in the marine bacterium Pseudoalteromonas tunicata: Brief report. Environ Microbiol 2007, 9, 814–818. [Google Scholar]
  113. Lefevre, F; Robe, P; Jarrin, C; Ginolhac, A; Zago, C; Auriol, D; Vogel, TM; Simonet, P; Nalin, R. Drugs from hidden bugs: Their discovery via untapped resources. Res Microbiol 2008, 159, 153–161. [Google Scholar]
  114. Brady, SF; Chao, CJ; Clardy, J. Long-chain N-acyltyrosine synthases from environmental DNA. Appl Environ Microbiol 2004, 70, 6865–6870. [Google Scholar]
  115. Henne, A; Schmitz, RA; Bomeke, M; Gottschalk, G; Daniel, R. Screening of environmental DNA libraries for the presence of genes conferring lipolytic activity on Escherichia coli. Appl Environ Microbiol 2000, 66, 3113–3116. [Google Scholar]
  116. Schloss, PD; Handelsman, J. Biotechnological prospects from metagenomics. Curr Opin Biotechnol 2003, 14, 303–310. [Google Scholar]
  117. Handelsman, J. Sorting out metagenomes. Nat Biotechnol 2005, 23, 38–39. [Google Scholar]
  118. Sarovich, DS; Pemberton, JM. pPSX: A novel vector for the cloning and heterologous expression of antitumor antibiotic gene clusters. Plasmid 2007, 57, 306–313. [Google Scholar]
  119. Díaz, M; Ferreras, E; Moreno, R; Yepes, A; Berenguer, J; Santamaría, R. High-level overproduction of Thermus enzymes in Streptomyces lividans. Appl Microbiol Biotechnol 2008, 79, 1001–1008. [Google Scholar]
  120. Butzin, NC; Owen, HA; Collins, MLP. A new system for heterologous expression of membrane proteins: Rhodospirillum rubrum. Protein Expr Purif.
  121. Li, C; Hazzard, C; Florova, G; Reynolds, KA. High titer production of tetracenomycins by heterologous expression of the pathway in a Streptomyces cinnamonensis industrial monensin producer strain. Metab Eng 2009, 11, 319–327. [Google Scholar]
  122. Foerstner, KU; Doerks, T; Creevey, CJ; Doerks, A; Bork, P. A computational screen for type I polyketide synthases in metagenomics shotgun data. PLoS ONE 2008, 3. [Google Scholar]
  123. Thomas, T; Evans, FF; Schleheck, D; Mai-Prochnow, A; Burke, C; Penesyan, A; Dalisay, DS; Stelzer-Braid, S; Saunders, N; Johnson, J; Ferriera, S; Kjelleberg, S; Egan, S. Analysis of the Pseudoalteromonas tunicata genome reveals properties of a surface-associated life style in the marine environment. PLoS ONE 2008, 3, e3252. [Google Scholar]
  124. Blasiak, LC; Clardy, J. Discovery of 3-formyl-tyrosine metabolites from Pseudoalteromonas tunicata through heterologous expression. J Am Chem Soc 2009. [Google Scholar]
  125. Silakowski, B; Kunze, B; Muller, R. Multiple hybrid polyketide synthase/non-ribosomal peptide synthetase gene clusters in the myxobacterium Stigmatella aurantiaca. Gene 2001, 275, 233–240. [Google Scholar]
  126. Bok, JW; Hoffmeister, D; Maggio-Hall, LA; Murillo, R; Glasner, JD; Keller, NP. Genomic mining for Aspergillus natural products. Chem Biol 2006, 13, 31–37. [Google Scholar]
  127. Firn, RD; Jones, CG. The evolution of secondary metabolism–a unifying model. Mol Microbiol 2000, 37, 989–994. [Google Scholar]
  128. Handelsman, J; Rondon, MR; Brady, SF; Clardy, J; Goodman, RM. Molecular biological access to the chemistry of unknown soil microbes: A new frontier for natural products. Chem Biol 1998, 5, R245–249. [Google Scholar]
  129. Lane, DJ. 16S/23S rRNA sequencing. In Nucleic acid techniques in bacterial systematics; Stackebrandt, E, Goodfellow, M, Eds.; John WIley & Sons: New York, NY, USA, 1991; pp. 115–147. [Google Scholar]
  130. Mitova, MI; Murphy, AC; Lang, G; Blunt, JW; Cole, ALJ; Ellis, G; Munro, MHG. Evolving trends in the dereplication of natural product extracts. 2. The isolation of chrysaibol, an antibiotic peptaibol from a New Zealand sample of the mycoparasitic fungus S. epedonium chrysospermum. J Nat Prod 2008, 71, 1600–1603. [Google Scholar]
  131. Nielen, MWF; Hooijerink, H; Claassen, FC; van Engelen, MC; van Beek, TA. Desorption electrospray ionisation mass spectrometry: A rapid screening tool for veterinary drug preparations and forensic samples from hormone crime investigations. Anal Chim Acta 2009, 637, 92–100. [Google Scholar]
  132. Williams, JP; Scrivens, JH. Rapid accurate mass desorption electrospray ionisation tandem mass spectrometry of pharmaceutical samples. Rapid Commun Mass Spectrom 2005, 19, 3643–3650. [Google Scholar]
  133. Jensen, PR; Mincer, TJ; Williams, PG; Fenical, W. Marine actinomycete diversity and natural product discovery. Anton Leeuwenhoek 2005, 87, 43–48. [Google Scholar]
  134. Maldonado, LA; Fenical, W; Jensen, PR; Kauffman, CA; Mincer, TJ; Ward, AC; Bull, AT; Goodfellow, M. Salinispora arenicola gen. nov., sp. nov. and Salinispora tropica sp. nov., obligate marine actinomycetes belonging to the family Micromonosporaceae. Int J Syst Evol Microbiol 2005, 55, 1759–1766. [Google Scholar]
  135. Williams, PG; Buchanan, GO; Feling, RH; Kauffman, CA; Jensen, PR; Fenical, W. New cytotoxic salinosporamides from the marine actinomycete Salinispora tropica. J Org Chem 2005, 70, 6196–6203. [Google Scholar]
  136. Boonlarppradab, C; Kauffman, CA; Jensen, PR; Fenical, W. Marineosins A and B, cytotoxic spiroaminals from a sarine-serived Actinomycete. Org Lett 2008, 10, 5505–5508. [Google Scholar]
  137. Taori, K; Paul, VJ; Luesch, H. Structure and activity of largazole, a potent anti-proliferative agent from the Floridian marine cyanobacterium Symploca sp. J Am Chem Soc 2008, 130, 1806–1807. [Google Scholar]
  138. Hill, RA. Marine natural products. Annu Rep Prog Chem Sect B: Org Chem 2009, 105, 150–166. [Google Scholar]
  139. Jensen, PR; Fenical, W. Strategies for the discovery of secondary metabolites from marine bacteria: Ecological perspectives. Annu Rev Microbiol 1994, 48, 559–584. [Google Scholar]
  140. Williams, PG. Panning for chemical gold: Marine bacteria as a source of new therapeutics. Trends Biotechnol 2009, 27, 45–52. [Google Scholar]
  141. Monaghan, RL; Polishook, JD; Pecore, VJ; Bills, GF; Nallin-Omstead, M; Streicher, SL. Discovery of novel secondary metabolites from fungi–Is it really a random walk through a random forest. Can J Bot 1995, 73, S925–931. [Google Scholar]
  142. Koehn, FE; Carter, GT. The evolving role of natural products in drug discovery. Nat Rev Drug Discov 2005, 4, 206–220. [Google Scholar]
  143. Strobel, GA. Rainforest endophytes and bioactive products. Crit Rev Biotechnol 2002, 22, 315–333. [Google Scholar]
  144. Frenich, AG; Vidal, JLM; Romero-Gonzalez, R; Aguilera-Luiz, MdM. Simple and high-throughput method for the multimycotoxin analysis in cereals and related foods by ultra-high performance liquid chromatography/tandem mass spectrometry. Food Chem 2009, 117, 705–712. [Google Scholar]
  145. Rowe, B; Schmidt, JJ; Smith, LA; Ahmed, SA. Rapid product analysis and increased sensitivity for quantitative determinations of botulinum neurotoxin proteolytic activity. Anal Biochem 2010, 396, 188–193. [Google Scholar]
  146. Bull, AT; Stach, JEM. Marine actinobacteria: New opportunities for natural product search and discovery. Trends Microbiol 2007, 15, 491–499. [Google Scholar]
  147. Cardellina, JH. A place for natural products. Screening 2006, 7, 28–30. [Google Scholar]
  148. Koehn, FE. High impact technologies for natural products screening. Prog Drug Res 2008, 65, 176–210. [Google Scholar]
  149. Lam, KS. New aspects of natural products in drug discovery. Trends Microbiol 2007, 15, 279–289. [Google Scholar]
  150. Bugni, TS; Richards, B; Bhoite, L; Cimbora, D; Harper, MK; Ireland, CM. Marine natural product libraries for high-throughput screening and rapid drug discovery. J Nat Prod 2008, 71, 1095–1098. [Google Scholar]
  151. Floss, HG. Combinatorial biosynthesis-Potential and problems. J Biotechnol 2006, 124, 242–257. [Google Scholar]
  152. Alksne, LE; Dunman, PM. Target-based antimicrobial drug discovery. Methods Mol Biol 2007, 431, 271–283. [Google Scholar]
  153. Alvarez, J; Vicente, M. Using genomics to identify new targets and counteract resistance to antibiotics. Expert Opin Ther Pat 2007, 17, 667–674. [Google Scholar]
  154. Dougherty, TJ; Barrett, JF; Pucci, MJ. Microbial genomics and novel antibiotic discovery: New technology to search for new drugs. Curr Pharm Des 2002, 8, 1119–1135. [Google Scholar]
  155. Dougherty, TJ; Miller, PF. Microbial genomics and drug discovery: Exploring innovative routes of drug discovery in the postgenomic era. IDrugs 2006, 9, 420–422. [Google Scholar]
  156. Monaghan, RL; Barrett, JF. Antibacterial drug discovery–Then, now and the genomics future. Biochem Pharmacol 2006, 71, 901–909. [Google Scholar]
  157. Payne, DJ; Gwynn, MN; Holmes, DJ; Pompliano, DL. Drugs for bad bugs: Confronting the challenges of antibacterial discovery. Nat Rev Drug Discov 2007, 6, 29–40. [Google Scholar]
  158. Walsh, C. Where will new antibiotics come from. Nat Rev Microbiol 2003, 1, 65–70. [Google Scholar]
  159. Butler, MS. The role of natural product chemistry in drug discovery. J Nat Prod 2004, 67, 2141–2153. [Google Scholar]
  160. Muller-Kuhrt, L. Putting nature back into drug discovery. Nat Biotechnol 2003, 21, 602. [Google Scholar]
  161. Newman, DJ; Cragg, GM; Snader, KM. Natural products as a source of new drugs over the period 1981–2002. J. Nat. Prod 2003, 66, 1002–1037. [Google Scholar]
  162. Overbye, KM; Barrett, JF. Antibiotics: Where did we go wrong. Drug Discov Today 2005, 10, 45–52. [Google Scholar]
  163. Norrby, SR; Nord, CE; Finch, R. Lack of development of new antimicrobial drugs: A potential serious threat to public health. Lancet Infect Dis 2005, 5, 115–119. [Google Scholar]
  164. Projan, SJ. Why is big pharma getting out of antibacterial drug discovery. Curr Opin Microbiol 2003, 6, 427–430. [Google Scholar]
  165. Larsson, J; Gottfries, J; Muresan, S; Backlund, A. ChemGPS-NP: Tuned for navigation in biologically relevant chemical space. J Nat Prod 2007, 70, 789–794. [Google Scholar]
  166. Larsen, TO; Smedsgaard, J; Nielsen, KF; Hansen, ME; Frisvad, JC. Phenotypic taxonomy and metabolite profiling in microbial drug discovery. ChemInform 2006, 37. [Google Scholar]
  167. Nicolaou, KC; Frederick, MO; Petrovic, G; Cole, KP; Loizidou, EZ. Total synthesis and confirmation of the revised structures of azaspiracid-2 and azaspiracid-3. Angew Chem Int Ed 2006, 45, 2609–2615. [Google Scholar]
  168. Nicolaou, KC; Zhang, H; Chen, JS; Crawford, JJ; Pasunoori, L. Total synthesis and stereochemistry of uncialamycin. Angew Chem Int Ed 2007, 46, 4704–4707. [Google Scholar]
  169. Tillmann, U; Elbrächter, M; Krock, B; John, U; Cembella, A. Azadinium spinosum gen. et sp. nov. (Dinophyceae) identified as a primary producer of azaspiracid toxins. Eur J Phycol 2009, 44, 63–79. [Google Scholar]
  170. Crimmins, MT; Zuccarello, JL; Ellis, JM; McDougall, PJ; Haile, PA; Parrish, JD; Emmitte, KA. Total synthesis of Brevetoxin A. Org Lett 2009, 11, 489–492. [Google Scholar]
  171. Haywood, AJ; Scholin, CA; Marin Iii, R; Steidinger, KA; Heil, C; Ray, J. Molecular detection of the brevetoxin-producing dinoflagellate Karenia brevis and closely related species using rRNA-targeted probes and a semiautomated sandwich hybridization assay. J Phycol 2007, 43, 1271–1286. [Google Scholar]
  172. Feher, M; Schmidt, JM. Property distributions: Differences between drugs, natural products, and molecules from combinatorial chemistry. J Chem Inf Comput Sci 2003, 43, 218–227. [Google Scholar]
  173. Henkel, T; Brunne, RM; Muller, H; Reichel, F. Statistical investigation into the structural complementarity of natural products and synthetic compounds. Angew Chem Int Ed 1999, 38, 643–647. [Google Scholar]
  174. Verdine, G. The combinatorial chemistry of nature. Nature 1996, 384, 11–13. [Google Scholar]
  175. Tulp, M; Bohlin, L. Unconventional natural sources for future drug discovery. Drug Discov Today 2004, 9, 450–458. [Google Scholar]
  176. Newman, D; Cragg, G; Kingston, D. Natural products as pharmaceuticals and sources for lead structures. In The Practice of Medicinal Chemistry, 2nd ed; Wermuth, CG, Ed.; Academic Press: London, UK, 2003; pp. 91–110. [Google Scholar]
  177. Nisbet, LJ; Moore, M. Will natural products remain an important source of drug research for the future. Curr Opin Biotechnol 1997, 8, 708–712. [Google Scholar]
  178. Nicolaou, KC; Chen, JS; Edmonds, DJ; Estrada, AA. Recent advances in the chemistry and biology of naturally occurring antibiotics. Angew Chem Int Ed 2009, 48, 660–719. [Google Scholar]
  179. Breinbauer, R; Vetter, IR; Waldmann, H. From protein domains to drug candidates--natural products as guiding principles in compound library design and synthesis. Ernst Schering Res Found Workshop 2003, 167–188. [Google Scholar]
  180. Urizar, NL; Liverman, AB; Dodds, DT; Silva, FV; Ordentlich, P; Yan, YZ; Gonzalez, FJ; Heyman, RA; Mangelsdorf, DJ; Moore, DD. A natural product that lowers cholesterol as an antagonist ligand for FXR. Science 2002, 296, 1703–1706. [Google Scholar]
  181. Von Nussbaum, F; Brands, M; Hinzen, B; Weigand, S; Habich, D. Antibacterial natural products in medicinal chemistry–Exodus or revival. Angew Chem Int Ed 2006, 45, 5072–5129. [Google Scholar]
  182. Demain, AL; Sanchez, S. Microbial drug discovery: 80 Years of progress. J Antibiot (Tokyo) 2009, 62, 5–16. [Google Scholar]
  183. Walters, WP; Murcko, A; Murcko, MA. Recognizing molecules with drug-like properties. Curr Opin Chem Biol 1999, 3, 384–387. [Google Scholar]
  184. Leeson, PD; Davis, AM; Steele, J. Drug-like properties: Guiding principles for design–Or chemical prejudice. Drug Discov Today 2004, 1, 189–195. [Google Scholar]
  185. Jenkins, ID; Lacrampe, F; Ripper, J; Alcaraz, L; Van Le, P; Nikolakopoulos, G; De Leone, PA; White, RH; Quinn, RJ. Synthesis of four novel natural product inspired scaffolds for drug discovery. J Org Chem 2009, 74, 1304–1313. [Google Scholar]
  186. Bollag, DM; McQueney, PA; Zhu, J; Hensens, O; Koupal, L; Liesch, J; Goetz, M; Lazarides, E; Woods, CM. Epothilones, a new class of microtubule-stabilizing agents with a taxol-like mechanism of action. Cancer Res 1995, 55, 2325–2333. [Google Scholar]
  187. Nicolaou, K; Scarpelli, R; Bollbuck, B; Werschkun, B; Pereira, M; Wartmann, M; Altmann, KH; Zaharevitz, D; Gussio, R; Giannakakou, P. Chemical synthesis and biological properties of pyridine epothilones. Chem Biol 2000, 7, 593–599. [Google Scholar]
  188. Nicolaou, KC; Pratt, BA; Arseniyadis, S; Wartmann, M; O’Brate, A; Giannakakou, P. Molecular design and chemical synthesis of a highly potent epothilone. ChemMedChem 2006, 1, 41–44. [Google Scholar]
Figure 1. General procedure for the discovery of biologically active natural compounds, such as antimicrobials, of microbial origin. The procedure starts with the isolation of microorganisms from the environment, for example, from the surfaces of marine eukaryotes, followed by their antimicrobial activity screening and the identification of the producer organism. The bioactive compound is then purified and the chemical structure elucidated. Production optimization can be performed to maximize the yield of the desired compound for further in vivo trials and product development. Clip art images provided by Open Clip Art Library ( www.openclipart.org) are used in the figure.
Figure 1. General procedure for the discovery of biologically active natural compounds, such as antimicrobials, of microbial origin. The procedure starts with the isolation of microorganisms from the environment, for example, from the surfaces of marine eukaryotes, followed by their antimicrobial activity screening and the identification of the producer organism. The bioactive compound is then purified and the chemical structure elucidated. Production optimization can be performed to maximize the yield of the desired compound for further in vivo trials and product development. Clip art images provided by Open Clip Art Library ( www.openclipart.org) are used in the figure.
Marinedrugs 08 00438f1

Share and Cite

MDPI and ACS Style

Penesyan, A.; Kjelleberg, S.; Egan, S. Development of Novel Drugs from Marine Surface Associated Microorganisms. Mar. Drugs 2010, 8, 438-459. https://doi.org/10.3390/md8030438

AMA Style

Penesyan A, Kjelleberg S, Egan S. Development of Novel Drugs from Marine Surface Associated Microorganisms. Marine Drugs. 2010; 8(3):438-459. https://doi.org/10.3390/md8030438

Chicago/Turabian Style

Penesyan, Anahit, Staffan Kjelleberg, and Suhelen Egan. 2010. "Development of Novel Drugs from Marine Surface Associated Microorganisms" Marine Drugs 8, no. 3: 438-459. https://doi.org/10.3390/md8030438

Article Metrics

Back to TopTop