Skip to main content
  • Original Article
  • Published:

Production of vineomycin A1 and chaetoglobosin A by Streptomyces sp. PAL114

Abstract

An actinobacteria strain PAL114, isolated from a Saharan soil in Algeria, produces bioactive compounds. Morphological and chemical studies indicated that this strain belongs to the genus Streptomyces. Analysis of the 16S rRNA gene sequence showed a similarity level of 99.8 % with S. griseoflavus LMG 19344T, the most closely related species. Two bioactive compounds, named P44 and P40, were extracted by dichloromethane from the cell-free supernatant broth and were purified by HPLC. Minimum inhibitory concentrations (MIC) of the compounds were determined against pathogenic and toxigenic microorganisms, most of which are multiresistant to antibiotics. The P40 fraction showed a strong activity especially against Candida albicans, Bacillus subtilis, and Staphylococcus aureus and has lower MIC values than those of P44 against most microorganisms tested. Chemical structures of compounds were determined based on spectroscopic and spectrometric analyses (UV-visible, mass, 1H, and 13C NMR spectra). The compounds P44 and P40 were identified as vineomycin A1 and chaetoglobosin A, respectively. Vineomycin A1 is known to be produced by some Streptomyces species. However, chaetoglobosin A is known to be produced only by fungi belonging to the genera Chaetomium, Penicillium, and Calonectria. This is the first time that chaetoglobosin A, known for its antimicrobial, anticancer, and cytotoxic effects, is reported in prokaryotes.

Introduction

Actinobacteria are Gram-positive bacteria with a percentage of guanine-cytosine higher than 55 %, and most of them produce mycelia. They are particularly interesting for their high capacity to produce secondary metabolites with diverse chemical structures (Watve et al. 2001). Among these compounds, there are, for example, antivirals, antiparasitics, immunostimulants, and immunosuppressants. However, actinobacteria are especially known for the production of antibiotics (Takahashi and Omura 2003; Solanki et al. 2008). Indeed, over 45 % of bioactive molecules of microbial origin are produced by actinobacteria (Solecka et al. 2012). Among these molecules, about 80 % are produced by species of the genus Streptomyces, the most common genus in the environment (Demain 2006; Jose et al. 2011). The antibiotics secreted by this genus may have antibacterial or antifungal activities, or cytostatic and antitumor properties, such as adriamycin and anthramycin (Butler 2004). Many of these molecules have found an important therapeutic application (Jose and Jebakumar 2013).

Our previous works showed the richness of Algeria Saharan soil with actinobacteria producers of bioactive compounds (Sabaou et al. 1998). Therefore, many strains proved to be new species producing new bioactive molecules (Lamari et al. 2002a, b; Zitouni et al. 2004a, b, 2005; Badji et al. 2006, 2007; Merrouche et al. 2010; Boubetra et al. 2012).

A previous work performed on a Saharan actinobacteria strain named PAL114 revealed the production of two bioactive molecules identified as saquayamycins A and C, known as anticancer agents (Aouiche et al. 2014). In the present article, which is the continuation of a previously published work, the taxonomy of actinomycete strain and the production, purification, structure elucidation, and bioactivity of two other bioactive molecules produced by the same strain are described.

Materials and methods

Identification of actinobacteria strain PAL114

The actinobacteria strain PAL114 was isolated from a Saharan soil collected from Ghardaïa province, south Algeria (Aouiche et al. 2014).

The morphological and cultural characteristics were observed by naked-eye examination of 14 day-old cultures grown on various ISP (International Streptomyces Project) media: yeast extract–malt extract agar (ISP2), oatmeal agar (ISP3), inorganic salts–starch agar (ISP4), and glycerol–asparagine agar (ISP5) (Shirling and Gottlieb 1966), and also on Bennett medium (Waksman 1961). The micromorphology and sporulation were observed by light microscopy.

For chemotaxonomic analyses, biomass was obtained from a culture grown in shaken ISP2 medium (Shirling and Gottlieb 1966) and incubated at 30 °C for 4 days. Analysis of diaminopimelic acid isomers and whole-cell sugar pattern were carried out using the method of Becker et al. (1964) and Lechevalier and Lechevalier (1970). Phospholipids were analysed according to the procedures developed by Minnikin et al. (1977).

For the physiological study, 18 tests were used (Locci 1989), including the production of melanoid pigments on ISP6 and ISP7 media, the assimilation of nine carbohydrates as sole carbon source, the degradation of xanthine, the production of nitrate reductase, the sensitivities to sodium chloride (7 % w/v), sodium azide (0.01 % w/v), phenol (0.1 % w/v), and penicillin (10 UI), and the growth at 45 °C.

For molecular analysis, DNA was extracted using the procedure recommended by Liu et al. (2000). PCR amplification of the 16S rRNA gene sequence of strain PAL114 was performed using two primers: 27f (5′-AGAGTTTGATCCTGGCTCAG-3′) and 1492r (5′- GGTTACCTTGTTACGACTT-3′). The 16S rRNA gene sequence was amplified by PCR using an Invitrogen Kit and sequenced. The sequences obtained were compared with sequences present in the public sequence databases and with the EzTaxon-e server (Kim et al. 2012).

Phylogenetic analyses were conducted using MEGA version 5.0 (Tamura et al. 2011). The 16S rRNA gene sequence of strain PAL114 was aligned using the CLUSTAL W program (Thompson et al. 1994) against corresponding nucleotide sequences of representatives of the Streptomyces genus retrieved from GenBank. Evolutionary distance matrices were generated as described by Jukes and Cantor (1969) and a phylogenetic tree was established using the neighbor-joining method of Saitou and Nei (1987). The topology of the tree was evaluated by bootstrap analysis (Felsenstein 1985) using 1,000 resamplings.

Production and purification of antimicrobial compounds

Production of bioactive compounds were conducted in ISP2 broth (malt extract 10 g, yeast extract 4 g, glucose 4 g, distilled water 1,000 mL, pH 7.2). A seed culture was prepared with the same medium and used to inoculate 80 Erlenmeyer flasks of 500 mL, each containing 100 mL of medium. The cultures were incubated on a rotary shaker (250 rpm) at 30 °C. The extraction of active compounds was performed on the fifth day, previously determined as being the day of optimum production. The ISP2 culture broth was centrifuged to eliminate cells. The cell-free supernatant was extracted with an equal volume of dichloromethane. The organic extract was concentrated to dryness by rotary evaporator under a vacuum at a temperature lower than 40 °C. The resulting dry extract was recuperated in 10 mL of methanol and bioassayed against Candida albicans M3, Aspergillus carbonarius M333, and Bacillus subtilis ATCC 6633 by paper disc diffusion method.

Preparative chromatography with silica gel plates (Merck Art. 5735, Kiesselgel 60HF 254–366; 20 × 20 cm) was employed for the partial purification of bioactive product. TLC plates were developed in the solvent system ethyl acetate–methanol (100:15 v/v). The developed plates were air dried overnight to remove all traces of solvents. The separated compounds were visualized under UV at 254 nm (absorbance) and 365 nm (fluorescence), and the active spot was detected by bioautography (Betina 1973). The retention factor (Rf) of the active spot was measured.

The final purification of bioactive compound was carried out by Waters reverse phase HPLC using an XBridge C18 (5 μm) column (200 × 10 mm, Waters) with a continuous linear gradient solvent system from 20 to 100 % methanol in water, a flow rate of 2 mL/min, and UV detection at 220 and 254 nm. All peak fractions were collected and tested against Candida albicans M3, Aspergillus carbonarius M333, and Bacillus subtilis ATCC 6633. The active fractions were re-injected into the HPLC system under the same conditions as previously, until their final purification.

Identification of bioactive compounds

These analyses were made with the pure bioactive compounds. The UV spectra were determined with a Shimadzu UV1605 spectrophotometer. The mass spectra were recorded on an LCQ ion-trap mass spectrometer (Finnigan MAT, San Jose, CA, USA) with nanospray ion electro-spray ionization (ESI) source (positive and negative ion modes). 1H and 13C NMR spectroscopy were used for the characterization of two bioactive molecules. NMR sample was prepared by dissolving 3 mg of each compound in 600 μL of CD3OD. All spectra were recorded on a Bruker Avance 500 spectrometer equipped with a cryoprobe. All chemical shifts for 1H and 13C are relative to TMS using 1H (residual) or 13C chemical shifts of the solvent as a secondary standard. The temperature was set at 298 K. All the 1H and 13C signals were assigned on the basis of chemical shifts, spin-spin coupling constants, splitting patterns and signal intensities, and by using 1H-1H COSY45, 1H-1H TOCSY, 1H-13C HSQC, and 1H-13C HMBC experiments. Gradient-enhanced 1H COSY45 was realised including 36 scans for per increment. The mixing time for TOCSY was 60 ms and 96 scans per increment were accumulated. 1H-13C correlation spectra using a gradient-enhanced HSQC sequence (delay was optimised for 1JCH of 145 Hz) was obtained with 200 scans per increment. Gradient-enhanced HMBC experiment was performed allowing 62.5 ms for long-range coupling evolution (340 scans were accumulated). Typically, 2,048 t2 data points were collected for 256 t1 increments.

Determination of minimum inhibitory concentrations

Minimum inhibitory concentrations (MIC) of pure bioactive molecules were carried out using conventional agar dilution (Oki et al. 1990). Thirteen target microorganisms, the majority of which are pathogenic or toxigenic to humans and multiresistant to antibiotics (Table 1), were used. Three bacteria (Bacillus subtilis ATCC 6633, Staphylococcus aureus S1, and Escherichia coli E195), five filamentous fungi (Aspergillus carbonarius M333, A. flavus AF3, A. ochraceus AO1, Fusarium culmorum FC200, and Penicillium glabrum PG1), and five yeasts (Candida albicans M1, M3, IPA200, and IPA988, and Saccharomyces cerevisiae ATCC 4226) were inoculated onto Mueller Hinton medium for bacteria and Sabouraud medium for fungi, containing different concentrations of active compounds (1, 2, 5, 10, 20, 30, 50, 75, and 100 μg/mL). After a growth period of 24–48 h at 37 °C for bacteria and 48–72 h at 28 °C for fungi, the plates were examined for growth and the lowest bioactive compound concentration that inhibited the growth of each organism was determined. Mueller Hinton and Sabouraud media, without active compound and inoculated with target organisms, were used as control treatments.

Table 1 Resistance pattern of target fungi and bacteria to antibiotics

Results and discussion

Identification of strain PAL114

The strain PAL114 showed good growth on ISP2, ISP3, ISP4, ISP5, and Bennett media. The aerial and substrate mycelia were light to medium grey and light brown, respectively. A diffusible pigment with light brown color was produced on ISP2, ISP3, and Bennett media. The strain PAL114 formed a very well-developed aerial mycelium with long spiraled chains containing between 10 and 50 spores per chain. The spores were carried by sporophores and were oval and 1–1.5 × 0.6–1 μm in size. The substrate mycelium was nonfragmented.

The chemotaxonomic study of strain PAL114 showed the presence of LL-diaminopimelic acid isomer and glycin in the cell wall. The whole-cell hydrolysates contained non-characteristic sugars (ribose, glucose, and galactose), typical of cell wall type IC (Lechevalier and Lechevalier 1970). The phospholipid profile contained phosphatidylethanolamine, corresponding to phospholipid type PII (Lechevalier et al. 1977). Based on the morphological and chemical characteristics, strain PAL114 was identified to the genus Streptomyces (Holt et al. 1994).

The strain PAL114 used arabinose, fructose, inositol, mannitol, melibiose, rhamnose, and xylose as carbon source, but not sucrose and raffinose. Xanthine was degraded and nitrate reduced. The strain grew at 45 °C in the presence of penicillin (10 UI), phenol (0.1 % w/v), sodium azide (0.01 % w/v), and NaCl (7 % w/v). The melanoid pigments were not produced on ISP6 and ISP7 media.

The alignment of the 16S rRNA gene sequence (1471 nucleotides) of strain PAL114 with those of Streptomyces reference species available in the GenBank database, can be seen in the neighbor-joining dendrogram (Fig. 1). The similarity level was 99.8 % with S. griseoflavus LMG 19344T (Holt et al. 1994), the most closely related species. The morphological and physiological properties of strain PAL114 are similar to those of the type strain of S. griseoflavus except for the tests of nitrate reduction and growth at 45 °C.

Fig. 1
figure 1

Neighbor-joining tree based on 16S rRNA gene sequences showing the relation between strain PAL114 and type species of the genus Streptomyces. The numbers at the nodes indicate the levels of bootstrap support (≥50 %) based on neighbour-joining analyses of 1,000 resampled data sets. Bar, 0.001 nt substitution per nt position

Purification of bioactive compounds

The dichloromethane extract was chromatographed by TLC and developed in an ethyl acetate–methanol system (100–15 v/v). After migration, we observed one active bioautographic compound, which showed a strong activity against Candida albicans M3 and Bacillus subtilis ATCC 6633. This compound (Rf = 0.9), was selected and analyzed by HPLC. Two active fractions (P44 and P40) were purified with 80 % methanol in water. The fraction P40 was eluted at a retention time of 39.68 min and P44 at 44.08 min. A quantity of 5 mg was obtained for each molecule from 8 L of culture filtrate.

Identification of bioactive compounds

The UV-VIS spectrum of the bioactive compound P44 showed maxima at 218, 318, and 438 nm.

The mass spectrum of the compound was obtained in negative mode. It yielded a pseudo-molecular ion [M - H] = 933. Thus, the molecular weight of antimicrobial compound is M = 934.

The 1H and 13C chemical shifts of P44 compound are given in Table 2 and structure in Fig. 2. The HSQC and HMBC spectra show 49 carbon signals for P44 molecule. It was possible to discern five ketone groups (δ c 182.30 to 205.28), four hydroxyl groups (δ c 71.13 to 157.40), nine ether functions (δ c 67.34 to 95.13), 13 sp2–hybridized carbons (δ c from 114.00 to 144.09), and 13 sp3-hybridized carbons (δ c 14.04 to 42.61) for the P44 molecule. The hydrogens of the hydroxyl group are not observed due to rapid exchange with MeOD. The 2D 1H-1H and 1H-13C experiments, and especially the long range 1H-13C couplings observed in the HMBC spectrum (see Fig. 2b), permitted us to established the connectivity between all the groups of the molecule.

Table 2 1H and 13C NMR data assignments of P44 compound in CD3OD at 298K. See Fig. 2b for numbering of hydrogen and carbon atoms
Fig. 2
figure 2

Structure of bioactive compound P44 (a) and HMBC and COSY correlations (b)

The structure of the P44 molecule was determined by NMR and mass spectrometry to be vineomycin A1. It is an antibiotic belonging to the aquayamycin group, class of angucycline, family of anthracycline (Maskey et al. 2003). It is known to be active against Gram-positive bacteria and sarcoma-180 solid tumors in mice (Imamura et al. 1982; Omura 2011). Also, vineomycin A1 is a suitable treatment for hypertrophic scar tissue and keloid disease. It is known to be a potent inhibitor of prolyl 4-hydroxylase, an essential enzyme involved in the post translational modification of collagen causing several diseases such as fibrosis, arterio-sclerosis, and scleroderma (Cunliffe and Franklin 1986; Chen 2007). Furthermore, cytostatic activity in vivo has been reported for vineomycins (Rohr and Thiericke 1992). Vineomycins are known to be produced by actinobacteria, especially Streptomyces species such as S. matensis and S. albogriseolus (Ono et al. 1974; Chen 2007; Omura 2011), but have never been reported in S. griseoflavus, to which our strain was closely related.

For the P40 compound, the UV-VIS spectrum showed maxima at 220 and 276 nm.

The mass spectrum of the compound was obtained in negative mode. It yielded a pseudo-molecular ion [M - H] = 527. Thus, the molecular weight of antimicrobial compound is M = 528. The 1H and 13C chemical shifts of the P40 compound are given in Table 3 and structure in Fig. 3. The HSQC and HMBC spectra show 32 carbon signals for P40. From these data, it was possible to discern two ketone groups (δ c 197.3 and 200.5), one amide group (δ c 174.3), one hydroxyl group (δ c 81.5), one epoxy group (δ c 58.0 and 62.1), 14 sp2–hybridized carbons (δ c from 108.0 to 136.6), and 14 sp3-hybridized carbons (δ c 12.1 to 81.5). The hydrogens of the hydroxyl group are not observed due to rapid exchange with MeOD. The 2D 1H-1H and 1H-13C experiments, and especially the long range 1H-13C couplings observed in the HMBC spectrum (see Fig. 3b), permitted us to establish the connectivity between all the groups of the P40 molecule.

Table 3 1H and 13C NMR data assignments of the P40 compound in CD3OD at 298 K. See Fig. 3b for numbering of hydrogen and carbon atoms
Fig. 3
figure 3

Structure of bioactive compound P40 (a) and HMBC and COSY correlations (b)

The structure of the P40 compound was determined by NMR and mass spectrometry to be chaetoglobosin A. This compound is known for its antibacterial, antifungal, phytotoxic (June et al. 1998; Larsen et al. 2005; Zhang et al. 2013), anticancer, and cytotoxic effects towards mammal cells (Fisvad et al. 2004; Fogle et al. 2008), and nematicidal effects (Hu et al. 2012). It has also been reported that this compound increases fibrinolytic activity in bovine animals (Shinohara et al. 2000). This compound is a cytochalasin derivative (Ohtsubo et al. 1978; June et al. 1998). Furthermore, it is considered a mycotoxin, and it is known to be produced only by fungi such as Chaetomium globosum (Hu et al. 2012), Penicillium discolor, P. expansum, P. marinum (Fisvad et al. 2004), and Calonectria morganii (Von Wallbrunn et al. 2001). It has never been described in prokaryotes. This is the first time that this compound is reported in the genus Streptomyces, belonging to the class of Actinobacteria. All experiments were made three times at different cultivation times to confirm fully the production of chaetoglobosin A by strain PAL114.

Minimum inhibitory concentrations

Minimum inhibitory concentrations (MIC) of bioactive molecules purified by HPLC are summarized in Table 4.

Table 4 Minimum inhibitory concentrations (MIC) of the bioactive compounds P44 and P40 secreted by the strain PAL114 against several fungi and bacteria

For vineomycin A1 (P44), MIC values were between 50 and 75 μg/mL for yeasts, and 75 and 100 μg/mL for filamentous fungi. For bacteria, Bacillus subtilis ATCC 6633 was the most sensitive (50 μg/mL). All other bacteria tested were resistant.

For chaetoglobosin A (P40), MIC values were between 30 and 75 μg/mL for yeasts, 50 and 75 μg/mL for filamentous fungi, and 20 and 30 μg/mL for Gram-positive bacteria. All Gram-negative bacteria tested were resistant. The most sensitive microorganisms were Bacillus subtilis ATCC 6633 (20 μg/mL), Staphylococcus aureus S1 (30 μg/mL), and Candida albicans M3 (30 μg/mL).

In conclusion, the strain PAL114, closely related to S. griseoflavus, showed an antibacterial and antifungal activity against pathogenic and toxigenic microorganisms, most of which are resistant to several antibiotics. The bioactive compounds produced by the strain proved to be the vineomycin A1 and chaetoglobosin A.

Vineomycin A1 belongs to the same group, class, and family as saquayamycins (Aouiche et al. 2014); it differs from these compounds only by the presence of a saccharide derivative bonded to the carbon 40. Chaetoglobosin A is known to be produced only by a small number of fungal species. This is the first time that chaetoglobosin A, which, in addition to its antimicrobial activity, is also considered a mycotoxin in the scientific literature, is produced by prokaryotes such as Actinobacteria.

Moreover, strain PAL114 represents a sample which confirms the potential of actinobacteria to produce a large variety of bioactive molecules.

References

  • Aouiche A, Bijani C, Zitouni A, Mathieu F, Sabaou N (2014) Antimicrobial activity of saquayamycins produced by Streptomyces sp. PAL114 isolated from a Saharan soil. J Mycol Med 24:17–23

    Article  Google Scholar 

  • Badji B, Zitouni A, Mathieu F, Lebrihi A, Sabaou N (2006) Antimicrobial compounds produced by Actinomadura sp. AC104 isolated from an Algerian Saharan soil. Can J Microbiol 52:373–382

    Article  CAS  PubMed  Google Scholar 

  • Badji B, Mostefaoui A, Sabaou N, Lebrihi A, Mathieu F, Seguin E, Tillequin F (2007) Isolation and partial characterization of antimicrobial compounds from a new strain Nonomuraea sp. NM94. J Ind Microbiol 34:403–412

    Article  CAS  Google Scholar 

  • Becker B, Lechevalier MP, Gordon RE, Lechevalier HA (1964) Rapid differentiation between Nocardia and Streptomyces by paper chromatography of whole-cell hydrolysates. Appl Microbiol 12:421–423

    CAS  PubMed Central  PubMed  Google Scholar 

  • Betina V (1973) Bioautography in paper and thin layer chromatography and its scope in the antibiotic field. J Chromatogr 78:41–51

    Article  CAS  PubMed  Google Scholar 

  • Boubetra D, Sabaou N, Zitouni A, Bijani C, Lebrihi A, Mathieu F (2012) Taxonomy and chemical characterization of new antibiotics produced by Saccharothrix SA198 isolated from a Saharan soil. Microbiol Res 168:223–230

    Article  PubMed  Google Scholar 

  • Butler MS (2004) The role of natural product chemistry in drug discovery. J Nat Prod 67:2141–2215

    Article  CAS  PubMed  Google Scholar 

  • Chen CL (2007) Methodologies for the synthesis of functionalized naphthols and progress toward the total synthesis of vineomycinone B2 methyl ester and actinophyllic acid. Ph.D. dissertation, University of Texas, Austin

  • Cunliffe CJ, Franklin TJ (1986) Inhibition of prolyl 4-hydroxylase by hydroxyanthraquinones. Biochem J 239:311–315

    CAS  PubMed Central  PubMed  Google Scholar 

  • Demain AL (2006) From natural products discovery to commercialization: a success story. J Ind Microbiol Biotechnol 33:486–495

    Article  CAS  PubMed  Google Scholar 

  • Felsenstein J (1985) Confidence limits on phylogenies: an approach using the bootstrap. Evolution 39:783–791

    Article  Google Scholar 

  • Fisvad JC, Smedsgaard J, Larsen TO, Samson RA (2004) Mycotoxins, drugs and other extrolites produced by species in Penicillium subgenus Penicillium. Stud Mycol 49:201–241

    Google Scholar 

  • Fogle MR, Douglas DR, Jumper CA, Straus DC (2008) Growth and mycotoxin production by Chaetomium globosum is favored in a neutral pH. Int J Mol Sci 9:2357–2365

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Holt JG, Kreig NR, Sneath PHA, Staley JT, Williams ST (1994) Bergey’s manual of determinative bacteriology, 9th edn. Williams and Wilkins Co., Baltimore

    Google Scholar 

  • Hu Y, Zhang W, Zhang P, Ruan W, Zhu X (2012) Nematicidal activity of chaetoglobosin A poduced by Chaetomium globosum NK102 against Meloidogyne incognita. J Agric Food Chem 61:41–46

    Article  PubMed  Google Scholar 

  • Imamura N, Kakinuma K, Ikekawa N, Tanaka H, Omura S (1982) Biosynthesis of vineomycins A1 and B2. J Antibiot 35:602–608

    Article  CAS  PubMed  Google Scholar 

  • Jose PA, Jebakumar SRD (2013) Non-streptomycete actinomycetes nourish the current microbial antibiotic drug discovery. Front Microbiol. doi:10.3389/fmicb.2013.00240

    Google Scholar 

  • Jose PA, Santhi VS, Jebakumar SRD (2011) Phylogenetic-affiliation, antimicrobial potential and PKS gene sequence analysis of moderately halophilic Streptomyces sp. inhabiting an Indian saltpan. J Basic Microbiol 51:348–356

    Article  PubMed  Google Scholar 

  • Jukes TH, Cantor CR (1969) Evolution of protein molecules. In: Munro HN (ed) Mammalian protein metabolism, vol 3. Academic, New York, pp 21–132

  • June N, Yeon SW, Paek NS, Kim TH, Kim YH, Kim CJ, Kim KW (1998) Isolation and structural determination of anti-Helicobacter pylori compound from fungus 60686. San’oeb misaengmul haghoeji 26:137–142

    Google Scholar 

  • Kim OS, Cho YJ, Lee K, Yoon SH, Kim M, Na H, Park SC, Jeon YS, Lee JH et al (2012) Introducing EzTaxon-e: a prokaryotic 16S rRNA Gene sequence database with phylotypes that represent uncultured species. Int J Syst Evol Microbiol 62:716–721

    Article  CAS  PubMed  Google Scholar 

  • Lamari L, Zitouni A, Boudjella H, Badji B, Sabaou N, Lebrihi A, Lefebvre G, Seguin E, Tillequin F (2002a) New dithiolopyrrolone antibiotics from Saccharothrix sp. SA 233. I. Taxonomy, fermentation, isolation and biological activities. J Antibiot 55:696–701

    Article  CAS  PubMed  Google Scholar 

  • Lamari L, Zitouni A, Dob T, Sabaou N, Lebrihi A, Germain P, Seguin E, Tillequin F (2002b) New dithiolopyrrolone antibiotics from Saccharothrix sp. SA 233. II. Physicochemical properties and structure elucidation. J Antibiot 55:702–707

    Article  CAS  PubMed  Google Scholar 

  • Larsen TO, Smedsgaard J, Nielsen KF, Hansen ME, Frisvad JC (2005) Phenotypic taxonomy and metabolite profiling in microbial drug discovery. Nat Prod Rep 22:672–693

    Article  CAS  PubMed  Google Scholar 

  • Lechevalier MP, Lechevalier HA (1970) Chemical composition as a criterion in the classification of aerobic actinomycetes. Int J Syst Bacteriol 20:435–443

    Article  CAS  Google Scholar 

  • Lechevalier MP, De Bievre C, Lechevalier HA (1977) Chemotaxonomy of aerobic actinomycetes: phospholipid composition. Biochem Syst Ecol 5:249–260

    Article  CAS  Google Scholar 

  • Liu D, Coloe S, Baird R, Pedersen J (2000) Rapid mini-preparation of fungal DNA for PCR. J Clin Microbiol 38:471

    CAS  PubMed Central  PubMed  Google Scholar 

  • Locci R (1989) Streptomyces and related genera. In: Williams ST, Sharpe ME, Holt JG (eds) Bergey’s manual of systematic bacteriology, vol 4. Williams and Wilkins Co., Baltimore, pp 2451–2492

    Google Scholar 

  • Maskey RP, Helmke E, Laatsch H (2003) Himalomycin A and B: isolation and structure elucidation of new fridamycin type antibiotics from a marine Streptomyces isolate. J Antibiot 56:942–949

    Article  CAS  PubMed  Google Scholar 

  • Merrouche R, Bouras N, Coppel Y, Mathieu F, Monje MC, Sabaou N, Lebrihi A (2010) Dithiolopyrrolone antibiotic formation induced by adding valeric acid to the culture broth of Saccharothrix algeriensis. J Nat Prod 73:1164–1166

    Article  CAS  PubMed  Google Scholar 

  • Minnikin DE, Patel PV, Alshamaony L, Goodfellow M (1977) Polar lipid composition in the classification of Nocardia and related bacteria. Int J Syst Bacteriol 27:104–117

    Article  CAS  Google Scholar 

  • Ohtsubo K, Saito M, Sekita S, Yoshihira K, Natori S (1978) Acute toxic effects of chaetoglobosin A, a new cytochalasan compound produced by Chaetomium globosum, on mice and rats. Jpn J Exp Med 48:105–110

    CAS  PubMed  Google Scholar 

  • Oki T, Tenmyo O, Tomatsu K, Kamei H (1990) Pradimicins A, B and C: new antifungal antibiotics. II. In vitro and in vivo biological activities. J Antibiot 43:763–770

    Article  CAS  PubMed  Google Scholar 

  • Omura S (2011) Microbial metabolites: 45 years of wandering, wondering and discovering. Tetrahedron 67:6420–6459

    Article  CAS  Google Scholar 

  • Ono H, Harada S, Kishi T (1974) Maridomycin, a new macrolide antibiotic. VII. J Antibiot 27:442–448

    Article  CAS  PubMed  Google Scholar 

  • Rohr J, Thiericke R (1992) Angucycline group antibiotics. Nat Prod Rep 9:103–137

    Article  CAS  PubMed  Google Scholar 

  • Sabaou N, Boudjella H, Bennadji A, Mostefaoui A, Zitouni A, Lamari L, Bennadji H, Lefebvre G, Germain P (1998) Les sols des oasis du Sahara algérien, source d’actinomycètes rares producteurs d’antibiotiques. Sécheresse 9:147–153

    Google Scholar 

  • Saitou N, Nei M (1987) The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol 4:406–425

    CAS  PubMed  Google Scholar 

  • Shinohara C, Chikanishi T, Nakashima S, Hashimoto A, Hamanaka A, Endo A, Hasumi K (2000) Enhancement of fibrinolytic activity of vascular endothelial cells by chaetoglobosin A, crinipellin B, geodin and triticone B. J Antibiot 53:262–268

    Article  CAS  PubMed  Google Scholar 

  • Shirling EB, Gottlieb D (1966) Methods for characterization of Streptomyces species. Int J Syst Bacteriol 13:313–340

    Article  Google Scholar 

  • Solanki R, Khanna M, Lal R (2008) Bioactive compounds from marine actinomycetes. Indian J Microbiol 48:410–431

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Solecka J, Zajko J, Postek M, Rajnisz A (2012) Biologically active secondary metabolites from actinomycetes. Cent Eur J Biol 7:373–390

    CAS  Google Scholar 

  • Takahashi Y, Omura S (2003) Isolation of new actinomycete strains for the screening of new bioactive compounds. J Gen Appl Microbiol 49:141–154

    Article  CAS  PubMed  Google Scholar 

  • Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S (2011) MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28:2731–2739

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighing, position-specific gap penalties and weight matrix choice. Nucleic Acids Res 22:4673–4680

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Von Wallbrunn C, Luftmann H, Bergander K, Meinhardt F (2001) Phytotoxic chaetoglobosins are produced by the plant pathogen Calonectria morganii (anamorph Cylindrocladium scoparium). J Gen Appl Microbiol 47:33–38

    Article  Google Scholar 

  • Waksman SA (1961) The Actinomycetes. Classification, identification and descriptions of genera and species, vol 2. Williams and Wilkins Co, Baltimore

    Google Scholar 

  • Watve MG, Tickoo R, Jog MM, Bhole BD (2001) How many antibiotics are produced by the genus Streptomyces? Arch Microbiol 176:386–390

    Article  CAS  PubMed  Google Scholar 

  • Zhang G, Zhang Y, Qin J, Qu X, Liu J, Li X, Pan H (2013) Antifungal metabolites produced by chaetomium globosum No.04, an endophytic fungus isolated from Ginkgo biloba. Indian J Microbiol 53:175–180

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Zitouni A, Lamari L, Boudjella H, Badji B, Sabaou N, Gaouar A, Mathieu F, Lebrihi A, Labeda DP (2004a) Saccharothrix algeriensis sp. nov., isolated from Saharan soil. Int J Syst Evol Microbiol 54:1377–1381

    Article  CAS  PubMed  Google Scholar 

  • Zitouni A, Boudjella H, Mathieu F, Sabaou N, Lebrihi A (2004b) Mutactimycin PR, a new anthracycline antibiotic from Saccharothrix sp. SA 103. I. Taxonomy, fermentation, isolation and biological activities. J Antibiot 57:367–372

    Article  CAS  PubMed  Google Scholar 

  • Zitouni A, Boudjella H, Lamari L, Badji B, Mathieu F, Lebrihi A, Sabaou N (2005) Nocardiopsis and Saccharothrix genera in Saharan soils in Algeria: isolation, biological activities and partial characterization of antibiotics. Res Microbiol 156:984–993

    Article  CAS  PubMed  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Nasserdine Sabaou or Florence Mathieu.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aouiche, A., Meklat, A., Bijani, C. et al. Production of vineomycin A1 and chaetoglobosin A by Streptomyces sp. PAL114. Ann Microbiol 65, 1351–1359 (2015). https://doi.org/10.1007/s13213-014-0973-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13213-014-0973-1

Keywords