Next Article in Journal
Beyond Earth: Harnessing Marine Resources for Sustainable Space Colonization
Previous Article in Journal
Antioxidative and Anti-Atopic Dermatitis Effects of Peptides Derived from Hydrolyzed Sebastes schlegelii Tail By-Products
Previous Article in Special Issue
Streptomyces-Fungus Co-Culture Enhances the Production of Borrelidin and Analogs: A Genomic and Metabolomic Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heterologous Expression of Type II PKS Gene Cluster Leads to Diversified Angucyclines in Streptomyces albus J1074

1
Key Laboratory of Marine Drugs, Chinese Ministry of Education, School of Medicine and Pharmacy, Ocean University of China, Qingdao 266003, China
2
Laboratory for Marine Drugs and Bioproducts, Qingdao Marine Science and Technology Center, Qingdao 266237, China
3
Sanya Oceanographic Institute, Ocean University of China, Sanya 572025, China
4
Marine Biomedical Research Institute of Qingdao, Qingdao 266101, China
*
Author to whom correspondence should be addressed.
Mar. Drugs 2024, 22(11), 480; https://doi.org/10.3390/md22110480
Submission received: 27 September 2024 / Revised: 18 October 2024 / Accepted: 18 October 2024 / Published: 22 October 2024

Abstract

:
Heterologous expression has emerged as an effective strategy in activating Streptomyces cryptic gene clusters or improving yield. Eight compounds were successfully obtained by heterologous expression of the type II PKS gene cluster spi derived from marine Streptomyces sp. HDN155000 in the chassis host Streptomyces albus J1074. The structures with absolute configurations were elucidated using extensive MS and NMR spectroscopic methods, as well as theoretical NMR calculations and electronic circular dichroism (ECD) calculations. Interestingly, compound WS009 Z (2) contains a rare thiomethyl group, angumycinone T (4) has a novel oxo-bridge formed between C12a and C4, and angumycinone X (3) showed cytotoxicity toward K562 and NCI-H446/EP cell lines.

1. Introduction

Angucyclines are a significant class of aromatic polyketide compounds biosynthesized by type II polyketide synthases (PKS) via an iterative Claisen condensation to generate Poly-β-ketone backbones [1]. So far, over 400 members of angucyclines have been reported [2] and nearly half of them showed a wide range of biological activities, such as antimicrobial and/or anticancer, glutamate receptor agonist, gastric mucosal protectant, and platelet aggregation-inhibiting actions [3], which continually enhanced the importance of expanding the chemical space and obtaining novel angucycline derivatives for drug discovery and development purposes [4].
Streptomyces, a type of actinobacteria, is a Gram-positive bacterium with high GC genomes, primarily isolated from soil and marine sediments and other habitat samples [5,6,7,8], also known for its ability to produce a complex variety of secondary metabolites [9,10,11]. As genome sequencing technology advances, an increasing number of gene clusters (BGC) for natural products are being discovered [12]. However, only a limited number of natural products have been identified to date [13], making it essential to convert genomic information into valuable compounds for drug discovery and also commercial production. Heterologous expression of gene clusters is an effective strategy to address this problem by enabling the activation of cryptic gene clusters in heterologous host strains and increasing the yield of gene clusters expressed at low levels under laboratory conditions, as well as obtaining new analogues other than wild-type [14,15,16]. However, the biosynthetic capacity of an exogenous gene cluster is greatly affected by the metabolic characteristics of the chassis cells; for example, the competitive consuming of precursors in the native biosynthetic pathways will stop or drastically decrease the production of the desired compound. For this reason, it is critical to select a proper surrogate host, because different hosts influence the yield and abundance of compounds, even possibly producing entirely distinct compounds [17,18]. Thus far, various Streptomyces species including S. coelicolor, S. lividans, S. venezuelae, S. albus, S. fradiae, S. roseosporus, and S. toyocaensis have been specifically engineered and frequently employed for heterologous expression [19].
During our ongoing research into bioactive chemicals from marine-derived microorganisms, six angucycline derivatives, including one novel oxaspirocyclic structure, were obtained by heterologous expression of a type II polyketide biosynthesis gene cluster spi in Streptomyces coelicolor A3(2) [20]. In the current study, to further explore the metabolic capacity of the BGC of spi and acquire more angucycline analogues, this cluster was expressed in another feasible heterologous host, S. albus J1074. A total of six angucycline derivatives and two by-products were obtained from the heterologous strains. Here, we will describe the construction of the recombinant strain, and the characterization of the structures obtained from the culture.

2. Results

2.1. Heterologous Expression of spi BGC in S. albus J1074

A type II PKS gene cluster spi (GenBank accession number OP009365, Figure 1a), which is responsible for angucyclines synthesis, has been previously identified and confirmed from the genome of a marine-derived actinobacteria strain of Streptomyces sp. HDN155000 [20]. To develop the metabolic potential of the spi BGC and obtain more angucycline analogues, we decided to have spi expressed in S. albus J1074, which has been demonstrated as a friendly surrogate host for heterologously expressing actinobacteria sourced genes due to its faster growth, small genome size, easy genetic manipulation procedure, and clear metabolic background [21,22,23]. The empty vector p15A was transferred into S. albus J1074 as the negative control (Figure 1b, trace i), and a series of new peaks were detected by HPLC analysis (Figure 1b, trace ii), demonstrating that the spi gene cluster was successfully activated and that changing heterologous host would greatly develop the metabolic capacity of the spi BGC (Figure S2).

2.2. Isolation and Purification of Compounds from Recombinant Strain

To specifically identify the compounds, we processed a larger-scale fermentation (20 L) of the strain J1074::spi, and ethyl acetate extraction of the fermentation product yielded 11.3 g crude extract. Subsequent stepwise separation by column chromatography (CC) with SiliaSphere C18 (ODS), medium pressure liquid chromatography (MPLC) and HPLC of the crude extract yielded a total of eight compounds, including four undescribed angucyclinone analogues: angumycinone Z (1) (7.1 mg), WS009 Z (2) (5.6 mg), angumycinone X (3) (6.4 mg), and angumycinone T (4) (4.7 mg). Compounds 58 were identified by comparing their 1D NMR and MS data to previously published literature (Figure 2).
Compound 1 was separated as a pale brown powder. Its molecular formula, C19H18O8, was identified by HR-ESIMS m/z 373.0916 [M − H], indicating eleven degrees of unsaturation. The signals indicated in the 1H NMR data illustrate that there are one methyl (δH 1.93, s), three methylenes (δH 2.89, 2.06; δH 1.98, 1.60; δH 2.28, 1.74), one vinyl methine (δH 5.71, s), three aromatic methines (δH 7.34, d, J = 8.2 Hz; δH 7.76, t, J = 8.0 Hz; δH 7.60, d, J = 7.5 Hz), and five exchangeable protons (δH 11.10, s; δH 6.32, s; δH 6.03, s; δH 7.01, s; δH 5.66, s) (Table 1). The 13C NMR spectrum combined with the HSQC spectrum reveals the existence of one methyl, three sp3 methylenes, four protonated sp2 carbons, and eleven quaternary carbons (Table 2). These data imply that compound 1 has a tetracyclic benz[a]anthracene skeleton belonging to the angucycline analogue [24,25], which possibly resulted from the expression of the spi BGC. The planar structure of 1 was subsequently identified through a comprehensive analysis of 2D NMR data (Figure 3). Firstly, the 1H-1H COSY correlations from H-9 (δH 7.34, d, J = 8.2 Hz) via H-10 (δH 7.76, t, J = 8.0 Hz) through H-11 (δH 7.60, d, J = 7.5 Hz) indicate the presence of a 1,2,3-trisubstituted benzene ring (ring D) (Figure 3). The HMBC spectrum correlations from 8-OH (δH 11.10) to C-8 (δC 159.9, s), C-9 (δC 123.6, d), and C-7a (δC 116.3, s) showed a phenolic OH at C-8. The HMBC correlations from Me-13 to C-4 (δC 41.8, t), C-3 (δC 158.2, s), C-1 (δC 192.8, s), and C-2 (δC 122.9 d), from H-2 (δH 5.71, s) to C-4 and C-12b (δC 76.2, s), as well as from H-4 (δH 2.06, d, J = 17.7 Hz) to C-4a (δC 75.5, s) and C-12b (δC 76.2, s) identified the presence of the A ring. The B ring was determined via the HMBC correlations from the active exchangeable protons 12a-OH (δH 7.01) to C-12a (δC 76.4, s), C-12 (δC 194.1, s), and C-12b (δC 76.2, s), from 12b-OH (δH 5.66) to C-4a (δC 75.5, s), C-12a (δC 76.4, s), and C-12b, from 4a-OH (δH 6.32) to C-4a, and C-5 (δC 28.2, t), and from 6a-OH (δH 6.03) to C-6a (δC 77.6, s) and C-12a (δC 76.4, s). The HMBC correlations from H-11 to C-12 and C-7 (δC 200.5, s) confirmed the connection of rings B and D via ring C. Thus, the planar structure of 1 was identified (Figure 3).
The absolute configuration of 1 was determined by NOESY spectrum analysis and ECD calculations. Analyzing the NOESY spectrum data revealed a correlation between 4a-OH (δH 6.32) and 12a-OH (δH 7.01), as well as a correlation between 4a-OH and 12b-OH (δH 5.66), while there was no correlation with 6a-OH (δH 6.03), which indicated that 4a-OH, 12a-OH, and 12b-OH were all in the same plane and 6a-OH was on the opposite side. Thus, the relative configuration of 1 was confirmed to be (6aS*, 4aR*, 12aR*, 12bR*)-1 or (6aR*, 4aS*, 12aS*, 12bS*)-1. ECD calculation demonstrated that the latter was more compatible with the measured values, based on which the absolute configuration of compound 1 was confirmed as (6aR, 4aS, 12aS, 12bS)-1 (Figure 4).
Compound 2 is an even brown amorphous powder with a molecular formula of C20H20O7S given based on HRESIMS data. One-dimensional NMR and HSQC spectra showed that there were two methyl (δC 23.7 and 13.3), three sp3 methylenes (δC 17.9, 28.8, and 41.6), four protonated sp2 carbons (δC 119.0, 122.7, 124.4, and 136.0), and eleven quaternary carbons including four non-protonated sp3 carbons (δC 61.2, 75.7, 75.7, and 78.4), four non-protonated sp2 carbons (δC 115.4, 131.5, 159.4, and 160.0), and three carbonyls (δC 191.0, 193.5, and 195.5) (Table 2). These signals all indicate that compound 2 has a backbone highly similar to 1. The difference lies in the high-field chemical shift of C-6 (δC 17.9, t), C-6a (δC 61.2, s), and C-7 (δC 195.5, s), recombined with the HMBC correlation from -SCH3 (δH 1.67) to C-6a (Figure 3) and the HR-ESIMS data, which determined that the hydroxyl signal (6a-OH) in compound 1 was replaced by the thiomethyl group (δC 13.3, q) in 2. However, the NOESY data could not determine the relative configuration of the thiomethyl group and three hydroxyl groups. Compounds 2 and 1 are proposed to have originated from the same epoxide precursor (discussed later in Scheme 1), which went through a ring opening process probably promoted by hydrolysis or methane thiol addition, hence the absolute configuration of 2 was tentatively assigned as 6aR, 4aS, 12aS, 12bS.
Compound 3 is a yellowish solid powder. HR-ESIMS analysis established the chemical formula as C19H16O7, with twelve degrees of unsaturation. The 1H and 13C NMR spectra revealed that 3 contains one methyl, three methylene, four methenyl, and 11 quaternary carbon signals (including three carbonyl and six aromatic quaternary carbons). The 1H and 13C NMR spectra showed that 3 was highly similar to angumycinone A [26], with the only difference being the low-field chemical shift at C-12a (δC 67.5), from which it can be deduced that there is hydroxylation at this position, thus determining the planar structure of compound 3. Nevertheless, the relative configuration was unable to be defined by the NOESY data. The stereo-structure was determined based on NMR calculations, DP4+ analysis, and ECD calculations, primarily NMR calculations and DP4+ probability analyses of the four possible configurations (6aS*, 12aR*, 12bS*, 4aR*)-3a, (6aS*, 12aR*, 12bR*, 4aS*)-3b, (6aS*, 12aR*, 12bS*, 4aS*)-3c, and (6aS*, 12aR*, 12bR*, 4aR*)-3d at the B3LYP/6-311+G(d,p) level using PCM model in DMSO. The calculated results for 3d (R2 = 0.9958) were more consistent with the experimental data than others (Figure S3). Furthermore, the DP4+ probability calculations identified 3d with a 99.90% probability (Table S7). The absolute configuration of 3 was then determined to be 6aS,12aR,12bR,4aR through ECD calculation at the B3LYP/6-31+G(d,p)//B3LYP/6-31G(d) level was compared with the experimental ECD curve (Figure 5).
Compound 4 is a brown amorphous powder with molecular formula C19H16O7 deduced from HR-ESIMS, indicating 12 degrees of unsaturation. The 1H NMR spectrum of 4 displayed one methyl (δH 2.33, d, J = 1.2 Hz), two methylenes (δH 1.98; δH 2.24, 2.14), five methines (δH 3.98, d, J = 1.3 Hz; δH 3.14, s; δH 7.35, dd, J = 8.6, 1.0 Hz; δH 7.76, dd, J = 8.3, 7.6 Hz; δH 7.34, d, J = 7.5, 1.0 Hz), a vinyl methine (δH 5.83, s), and three exchangeable protons (δH 5.59, s; δH 7.05, s; δH 11.72, s) (Table 1). The 13C NMR and HSQC spectrum showed that 4 contains one methyl (δC 26.3, q), two methylene (δC 26.8 and 32.6), six methines (δC 52.7, 85.1, 118.3, 123.7, 124.6, and 137.7), and ten quaternary carbons, including three non-protonated sp3 carbons (δC 78.0, 81.6 and 92.5), four non-protonated sp2 carbons (δC 114.9, 135.3, 161.1, and 162.4), and three carbonyls (δC 192.1, 195.6, and 201.6) (Table 2). The planar structure of 4 was then identified via a detailed analysis of 2D NMR (Figure 3). The A ring was determined by the HMBC correlations from H-2 (δH 5.83, s) to C-1 (δC 195.6), C-4 (δC 85.1), and C-12b (δC 52.7), from 13-CH3 (δH 2.33, d) to C-2 (δC 124.6), C-3 (δC 162.4), C-4, and from H-4 (δH 3.98, d, J = 1.3 Hz) to C-4a (δC 81.6) and C-12b. The 1H-1H COSY correlations from H-9 (δH 7.35, dd, J = 8.6, 1.0 Hz) via H-10 (δH 7.76, dd, J = 8.3, 7.6 Hz) through H-11(δH 7.34, dd, J = 7.5, 1.0 Hz) suggested the presence of the 1,2,3-trisubstituted benzene ring (ring D). HMBC correlation analyses from H-9 to C-7a (δC 114.9) and C-7 (δC 201.6), from H-10 to C-11a (δC 135.3) and C-8 (δC 161.1), and from H-11 to C-12 (δC 192.1) extend ring D to ring C. Further analysis of the HMBC correlations from H-5 (δH 1.98, m) to C-6a (δC 78.0) and C-12b and from H-6 (δH 2.14) to C-12a (δC 92.5) determined the presence of ring B, closing the angular skeleton of compound 4. HMBC correlations from the active exchangeable protons 4a-OH (δH 5.59) to C-4a, C-4 and C-5 (δC 32.6), from 6a-OH (δH 7.05, s) to C-6a, C-7 and C-6, together with the fact that there are 12 degrees of unsaturation, suggests that another unsaturation should be the presence of an epoxide between C-12a and C-4 (δC 85.1). Thus, the planar structure of 4 was determined.
The stereo configuration of compound 4 was determined by NOESY spectral analysis and ECD calculations. The NOESY spectrum analysis revealed that H-12b (δH 3.14) correlates with 4a-OH and 6a-OH, indicating that H-12b, 4a-OH, and 6a-OH are in the same plane, resulting in only two possible structures: (4R*, 4aR*, 6aS*, 12aS*, 12bS*)-4 and (4S*, 4aS*, 6aR*, 12aR*, 12bR*)-4. Finally, the absolute configuration of 4 was determined to be 4S,4aS,6aR,12aR based on the results of ECD calculations (Figure 6). Angucyclines carrying epoxide modifications on the skeletons were frequently discovered in nature [27], however, this is the first example of angucycline derivative with the epoxidation-bridged C-12a and C-4.
Based on the comparison of NMR and MS data with those reported in the literature, the six known compounds isolated in this study were identified as WP 3688-2 (5) [28], gephyromycin (6) [29,30], SEK43 (7) [31], and SEK15 (8) [32].
All new angucyclines were evaluated for cytotoxicity against L-02, MDA-MB-231, K562, ASPC-1, and NCI-H446/EP cell lines in vitro. Adriamycin was used as the positive control. Compound 3 showed cytotoxicity toward five cancer cell lines and especially strongly inhibited K562 and NCI-H446/EP cell lines with IC50 values of 11.72 and 8.92 μM, respectively. It is worth mentioning that both compounds 3 and 4 possess an oxygen bridge and two hydroxyl groups in the structure; however, compound 4 did not display obvious cytotoxicity, which indicated that the oxidative modifications on rings A and B are very important for their mode of actions.

2.3. Plausible Biosynthetic Pathways for Compounds 18 Produced by the Mutant Strain

The plausible biogenetic pathway of compounds 18 was determined based on the bioinformatics analyses (Table S2) and literature surveys. First, the polyketide chain was generated by an iterative Claisen condensation reaction of a starting unit malonyl-CoA and nine extension unit malonyl-CoA under the action of the minimal PKS SpiA, SpiB and SpiC, and was subsequently folded by cyclase SpiE and SpiD to form UWM6 [33,34], a characteristic intermediate of angucyclines; however, if it is spontaneously folded, the by-products 7 and 8 will be produced. Thereafter, UWM6 can achieve the oxidation of C-12 catalyzed by monooxygenase SpiH2 [35] and subsequently undergo reduction at C-6, dehydration and enoyl reduction at C-5 and C-6, and oxidation at C-12b to produce 5, which could be converted into compounds 1 and 2 via a 6a, 12a-epoxide intermediate. The methanethiol residue of 2 was probably derived from methionine which was incorporated via methane thiol addition to the epoxide. In alternative routes, UWM6 may undergo multi-step catalysis including oxidation, dehydration and oxo-bridge generation to give 3, 4 and 6 in the presence of relevant endogenous enzymes of the heterologous host.

3. Materials and Methods

3.1. General Experimental Procedures

High-quality actinobacteria genomes were obtained using methods previously described in the literature [36,37]. Polymerase chain reaction (PCR) was performed using 2 × Phanta Flash Master Mix (P510, Vazyme, Nanjing, China). PCR analyses were conducted with a 2 × Hieff® PCR Master Mix (with dye, Yeasen, Shanghai, China). Shanghai Sangon DNA Technologies (Shanghai, China) offered custom oligonucleotide synthesis and fragment sequencing services.
NMR spectra were measured on an Agilent 500 MHz DD2 spectrometer with tetramethylsilane (TMS) as an internal standard. Optical rotations in MeOH were recorded on a JASCO-1020 digital polarimeter. A Thermo Scientific LTQ Orbitrap XL mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) was used to collect HRESIMS data. UV-vis spectra were acquired on the UFLC system (Shimadzu, Tokyo, Japan). The JASCO J-715 spectropolarimeter was used to record the ECD spectra. A UFLC system (Shimadzu, Tokyo, Japan) with a C18 column (Shimadzu, 4.6 mm × 150 mm, 5 μm, 1 mL/min) coupled to an LCMS-2020 mass spectrometer (Shimadzu, Tokyo, Japan) was used to perform LC-MS analyses. Column chromatography (CC) was performed with SiliaSphere C18 (Octadecylsilyl, ODS) monomeric (SiliCycle Inc., Quebec, QC, Canada, 50 μm, 120 A), and Sephadex LH-20 (GE Healthcare, Uppsala, Sweden). MPLC was performed using a C18 column (Welch Materials Inc., Ultimate® XB-C18, 21.2 mm × 250 mm, 5 μm, 10 mL/min).
Prediction and analysis of gene clusters using the online tool antiSMASH. DNA and protein sequence homology analysis using the online tool NCBI.

3.2. Materials and Culture Conditions

Streptomyces sp. HDN155000 was obtained from a marine sediment sample collected from the South China Sea at coordinates 125°28.550′ E and 29°1.618′ N. The strain was identified via genome sequencing and submitted to GenBank (accession number MN822699). S. albus J1074 was grown for 6 days at 28 °C on MS plates with 2% mannitol, 2% soya bean flour, and 1.5% agar. For genome extraction, the strain was inoculated into 250mL Erlenmeyer flasks containing 50 mL of TSB liquid medium (1.7% tryptone, 0.3% peptone, 0.25% glucose, 0.25% KH2PO4, and 0.3% NaCl) for 4 days.
Escherichia coli strains DH10B as the general host for cloning and E. coli ET12567/pUZ8002 as the donor in intergeneric conjugation were cultured on Luria–Bertani plates or liquid medium (1% tryptone, 0.5% yeast extract, and 1% NaCl) at 37 °C.

3.3. S. albus J1074 Intergeneric Conjugation

The intergeneric conjugation between E. coli and S. albus J1074 was performed as previously described with some modifications [38]. The culture of the donor E. coli ET12567/pUZ8002 containing the heterologous vector p15A::spi was incubated with appropriate antibiotics LB liquid culture medium (containing 50 mg/L kanamycin, 15 mg/L chloramphenicol, and 50 mg/L apramycin,) until the optical density at 600 nm (OD600) of 0.6–0.8. Bacteriophages were collected at 9000 rpm, rinsed three times with 1 mL 2× YT medium [39] to remove antibiotics, and then resuspended in 0.2 mL of 2× YT (containing 10 mM MgCl2) as the donor cells. S. albus J1074 spores were washed three times and suspended in 2× YT broth at a concentration of 109 per mL. Subsequently, spores obtained in the previous step were heated at 50 °C for 10 min to serve as recipients. Donor and recipient cells were mixed evenly on MS plates (containing 30 mM MgCl2 and 30 mM CaCl2), and cultivated for 18 h at 30 °C. After incubation, plates were covered with 1 mL of water with apramycin (50 μg/mL) and nalidixic acid (50 μg/mL) and then kept at 30 °C for 4–6 days until exconjugants appeared. The conjugation frequency reached 7.8 × 10−5.

3.4. Fermentation and LC/LC-MS Analysis

Heterologous expression strains and negative control strains were cultured for 7 days at 28 °C in 500 mL Erlenmeyer flasks with 100 mL of M1 medium (2 g/L peptone, 4 g/L yeast extract, 10 g/L starch), after which the culture products were collected and extracted three times with ethyl acetate. The organic phase was evaporated, and the residue was re-dissolved in 150 μL MeOH, which was analyzed using HPLC (C18 column, Shimadzu, 4.6 mm × 150 mm, 5 μm, 1 mL/min), revealing a considerable change in metabolite synthesis in the heterologous expression strain J1074::spi (Figure 1).
For the isolation of secondary metabolites, we scaled up cultures with the 500 mL Erlenmeyer flask (total 20 L) according to the conditions used for small-scale analyses. The fermentation supernatant was extracted with ethyl acetate three times and then evaporated to dryness, obtaining a total of 11.3 g of crude product.

3.5. Extraction, Isolation, and Purification

The crude extract was eluted using a step gradient of MeOH-H2O using metabolite detection on a C18 column, yielding seven subfractions (Fr.1-Fr.7, 30% to 100%). Fr.2 was purified by semi-preparative HPLC (33:67 MeOH-H2O, 3 mL/min) to generate compounds 1 (7.1 mg, tR = 15.1 min) and 6 (3.1 mg, tR = 16.7 min). Fr.3 was purified by a semi-preparative C18 HPLC column (47:53 MeOH-H2O), producing compound 3 (6.4 mg, tR = 14.7 min). Fr.4 was separated using preparative C18 HPLC column (58:42 MeOH-H2O) to yield two sub-fractions (Fr.4.1 and Fr.4.2). Then, Fr.4.1 and Fr.4.2 were purified by semi-preparative HPLC (53:47, 60:40, MeOH-H2O, 3 mL/min) to provide compounds 4 (4.7 mg, tR = 18.1 min), 5 (3.4 mg, tR = 20.7 min), 7 (3.8 mg, tR = 16.9 min), and 8 (3.0 mg, tR = 19.4 min), respectively. Fr.5 was applied to a Sephadex LH-20 column and eluted with methanol, obtaining compound 2 (5.6 mg).
Angumycinone Z (1): pale brown powder, [ α ] D 25 + 19.5 (c 0.03, CH3OH); UV (DAD) λmax 208 nm, 250 nm, 327 nm, 386 nm; CD (MeOH) λmax (∆ε) 210 (−19.06), 243 (+33.91), 283 (−4.88), 302 (+3.13), 336 (−3.39), 376 (+5.21); 1H and 13C NMR data, see Table 1 and Table 2; negative ion HRESIMS m/z 373.0916 [M − H] (calcd. for C19H17O7, 373.0928).
WS009 Z (2): even brown amorphous powder, [ α ] D 25 + 55.8 (c 0.03, CH3OH); UV (DAD) λmax 249, 302, 352; 1H and 13C NMR data, see Table 1 and Table 2; positive ion HRESIMS m/z 405.0996 [M + H]+ (calcd. for C20H21O7S+, 405.1003).
Angumycinone X (3): yellowish, solid powder, [ α ] D 25 + 79.1 (c 0.03, CH3OH); UV (DAD) λmax 206, 238, 309, 364; CD (MeOH) λmax (∆ε) 211 (−12.77), 242 (+25.31), 281 (−3.28), 300 (+2.63), 334 (−2.96), 372 (+4.21); 1H and 13C NMR data, see Table 1 and Table 2; negative ion HRESIMS m/z 355.0821 [M − H] (calcd. for C19H15O7, 355.0823).
Angumycinone T (4): brown amorphous powder, [ α ] D 25 + 31.0 (c 0.03, CH3OH); UV (DAD) λmax 206, 236, 298, 350; CD (MeOH) λmax (∆ε) 207 (+3.30), 227 (+19.67), 243 (−11.30), 274 (+14.31), 332 (−2.67), 370 (+2.94); 1H and 13C NMR data, see Table 1 and Table 2; negative ion HRESIMS m/z 355.0815[M − H] (calcd. for C19H15O7, 355.0823)

3.6. NMR and ECD Calculations

Conformation searches based on molecular mechanics with MMFF force fields were performed for stereoisomers to obtain stable conformers. The Gaussian 16 programme package [40] used the density functional theory (DFT) approach at the B3LYP/6-31G(d) level to optimize all conformers. Gauge Independent Atomic Orbital (GIAO) calculations of their 1H and 13C NMR chemical shifts were performed utilizing density functional theory (DFT) at the mPW1PW91/6-311+G(d,p) level with the PCM model in DMSO [41]. The calculated NMR data of these conformers were averaged using the Boltzmann distribution theory and their respective Gibbs free energy. The 1H and 13C NMR chemical shifts for TMS were computed using the same methodology and served as a reference. Following calculation, the experimental and calculated data were compared using linear correlation coefficients (R2) and the improved probability DP4+ technique [42].
All stable conformers were further optimized at the B3LYP/6-31G(d)-GD3BJ level with the Gaussian 16 programme package. The ECD was computed using time-dependent density functional theory (TDDFT) at a B3LYP/6-311+G(d,p) level with the IEFPCM model. The calculated ECD curves were all generated using the SpecDis 1.71 computer package and the calculated ECD data of all conformers were Boltzmann averaged using Gibbs free energy [43].

3.7. Cytotoxicity Assay

The cytotoxic assay involved five human cancer cell lines: K562 (using the MTT method), ASPC-1, NCI-H446/EP, MDA-MB-231, and L-02 (using the SRB method). Adriamycin (ADM) was employed as a positive control. The detailed procedures for biological testing were conducted as previously stated [20,44]. The cancer cell lines are purchased from the National Collection of Authenticated Cell Cultures of China (Shanghai).

4. Conclusions

Secondary metabolites from actinobacteria have long been a valuable source of natural medicines [45]. However, the success rate for identifying actinobacteria-derived novel compounds is limited by gene cluster silencing or low expression under laboratory conditions, especially for micro-organisms from special habitats such as marine actinobacteria. In this study, the type II PKS gene cluster spi derived from Streptomyces sp. HDN155000 was successfully heterologously expressed in S. albus J1074, and a total of eight compounds were obtained through further fermentation, isolation, and purification. Based on the literature and our previous study, we proposed a plausible biosynthetic pathway leading to those new compounds [4,34]. Generally, compounds 36 were directly produced from UWM6 by a multi-step enzymatic reaction involving oxidation, dehydration, enoyl reduction, etc. Compounds 1 and 2 were derived from compound 5 via dehydrogenation, hydrolysis, methane thiol addition, and other processes. Angucyclines are a chemical family that has been extensively explored in actinobacteria since the first structure was described in 1965 in Streptomyces rimosus [2]. More than 400 cases have been recorded, with a range of skeleton modifications; however, compound 4 is the first example of oxidized angucyclines with an oxo-bridge between C-12a and C-4 [3]. Compound 3 showed cytotoxic to the K562 and NCI-H446/EP cell lines, providing an alternative lead compound for the study of human chronic myeloid leukemia and lung cancer. The above data enrich our understanding of the type II PKS pathways and related aromatic products, and also indicate heterologous expression as a promising strategy for maximizing the metabolic capability of Streptomyces.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/md22110480/s1, Figure S1: Restriction to validate heterologous expression plasmid p15A-spi and PCR analysis for confirming the heterologous strain J1074::spi; Figure S2: Comparative analysis of metabolites from two heterologous hosts, S. albus J1074 and S. coelicolor A3; Figure S3: 13C NMR calculation results of four possible isomers of 3; Figures S4–S10: NMR and HRESIMS spectra of compound 1; Figures S11–S16: NMR and HRESIMS spectra of compound 2; Figures S17–S22: NMR and HRESIMS spectra of compound 3; Figures S23–S29: NMR and HRESIMS spectra of compound 4; Table S1: The primers used in this study; Table S2: Deduced function of 25 genes in spi; Table S3–S6: Calculated 13C NMR results for 3a, 3b, 3c, 3d; Table S7: DP4+ analysis results of 3a (Isomer 1), 3b (Isomer 2), 3c (Isomer 3) and 3d (Isomer 4).

Author Contributions

The contributions of the respective authors are as follows: X.Z. drafted the work and performed isolation and structural elucidation of the extract. X.Z. and F.Z. performed the NMR and ECD Calculations of the compounds. Biological evaluations were performed by C.L., J.L., X.X., T.Z., Q.C. and D.L. checked the procedures of this work. G.Z. designed the project and contributed to the critical reading of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (81991522, 32370061), Qingdao Marine Science and Technology Center (2022QNLM030003-1, 2022QNLM030003-2), NSFC-Shandong Joint Fund (U1906212), Hainan Provincial Joint Project of Sanya Yazhou Bay Science and Technology City (2021CXLH0012), Taishan Scholar Youth Ex-pert Program in Shandong Province (tsqn 202103153), Major Basic Research Programs of Natural Science Foundation of Shandong Province (ZR2021ZD28).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data given in this research are available in this article and the Supplementary Materials.

Acknowledgments

We thank Bin Wang (Institute of Microbiology, Chinese Academy of Sciences) for providing the heterologous expression strain.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rohr, J.; Thiericke, R. Angucycline group antibiotics. Nat. Prod. Rep. 1992, 9, 103–137. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, J.; Duan, Y.; Zhu, X.; Yan, X. Novel angucycline/angucyclinone family of natural products discovered between 2010 and 2020. Sheng Wu Gong Cheng Xue Bao 2021, 37, 2147–2165. [Google Scholar] [PubMed]
  3. Hibi, G.; Shiraishi, T.; Umemura, T.; Nemoto, K.; Ogura, Y.; Nishiyama, M.; Kuzuyama, T. Discovery of type II polyketide synthase-like enzymes for the biosynthesis of cispentacin. Nat Commun. 2023, 14, 8065. [Google Scholar] [CrossRef] [PubMed]
  4. Metsä-Ketelä, M.; Palmu, K.; Kunnari, T.; Ylihonko, K.; Mäntsälä, P. Engineering anthracycline biosynthesis toward angucyclines. Antimicrob. Agents Chemother. 2003, 47, 1291–1296. [Google Scholar] [CrossRef]
  5. Watve, M.G.; Tickoo, R.; Jog, M.M.; Bhole, B.D. How many antibiotics are produced by the genus Streptomyces? Arch. Microbiol. 2001, 176, 386–390. [Google Scholar] [CrossRef]
  6. van Bergeijk, D.A.; Terlouw, B.R.; Medema, M.H.; van Wezel, G.P. Ecology and genomics of Actinobacteria: New concepts for natural product discovery. Nat. Rev. Microbiol. 2020, 18, 546–558. [Google Scholar] [CrossRef]
  7. Liu, Y.; Yue, S.J.; Wang, W.; Hu, H.B.; Zhang, X.H. Elucidation of the biosynthesis of griseoluteic acid in Streptomyces griseoluteus P510. J. Nat. Prod. 2024, 87, 1540–1547. [Google Scholar] [CrossRef]
  8. Nett, M.; Ikeda, H.; Moore, B.S. Genomic basis for natural product biosynthetic diversity in the actinomycetes. Nat. Prod. Rep. 2009, 26, 1362–1384. [Google Scholar] [CrossRef]
  9. Lopatkin, A.J.; Bening, S.C.; Manson, A.L.; Stokes, J.M.; Kohanski, M.A.; Badran, A.H.; Earl, A.M.; Cheney, N.J.; Yang, J.H.; Collins, J.J. Clinically relevant mutations in core metabolic genes confer antibiotic resistance. Science 2021, 371, eaba0862. [Google Scholar] [CrossRef]
  10. Platon, V.M.; Dragoi, B.; Marin, L. Erythromycin formulations-A journey to advanced drug delivery. Pharmaceutics 2022, 14, 2180. [Google Scholar] [CrossRef]
  11. Chopra, I.; Roberts, M. Tetracycline antibiotics: Mode of action, applications, molecular biology, and epidemiology of bacterial resistance. Microbiol. Mol. Biol. Rev. MMBR 2001, 65, 232–260. [Google Scholar] [CrossRef] [PubMed]
  12. Katz, M.; Hover, B.M.; Brady, S.F. Culture-independent discovery of natural products from soil metagenomes. J. Ind. Microbiol. Biotechnol. 2016, 43, 129–141. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, J.J.; Tang, X.; Moore, B.S. Genetic platforms for heterologous expression of microbial natural products. Nat. Prod. Rep. 2019, 36, 1313–1332. [Google Scholar] [CrossRef]
  14. Liu, Z.; Sun, W.; Hu, Z.; Wang, W.; Zhang, H. Marine Streptomyces-derived novel alkaloids discovered in the past decade. Mar. Drugs 2024, 22, 51. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, Z.; He, J.; Wei, X.; Ju, J.; Ma, J. Exploration and genome mining of natural products from marine Streptomyces. Appl. Microbiol. Biotechnol. 2020, 104, 67–76. [Google Scholar] [CrossRef]
  16. Poli, A.; Finore, I.; Romano, I.; Gioiello, A.; Lama, L.; Nicolaus, B. Microbial diversity in extreme marine habitats and their biomolecules. Microorganisms 2017, 5, 25. [Google Scholar] [CrossRef]
  17. Rodríguez Estévez, M.; Myronovskyi, M.; Gummerlich, N.; Nadmid, S.; Luzhetskyy, A. Heterologous expression of the nybomycin gene cluster from the marine strain Streptomyces albus subsp. chlorinus NRRL B-24108. Mar. Drugs 2018, 16, 435. [Google Scholar] [CrossRef]
  18. Liu, S.H.; Wang, W.; Wang, K.B.; Zhang, B.; Li, W.; Shi, J.; Jiao, R.H.; Tan, R.X.; Ge, H.M. Heterologous expression of a cryptic giant Type I PKS gene cluster leads to the production of ansaseomycin. Org. Lett. 2019, 21, 3785–3788. [Google Scholar] [CrossRef]
  19. Klumbys, E.; Xu, W.; Koduru, L.; Heng, E.; Wei, Y.; Wong, F.T.; Zhao, H.; Ang, E.L. Discovery, characterization, and engineering of an advantageous Streptomyces host for heterologous expression of natural product biosynthetic gene clusters. Microb. Cell Fact. 2024, 23, 149. [Google Scholar] [CrossRef]
  20. Xu, X.; Chang, Y.; Chen, Y.; Zhou, L.; Zhang, F.; Ma, C.; Che, Q.; Zhu, T.; Pfeifer, B.A.; Zhang, G.; et al. Biosynthesis of Atypical Angucyclines Unveils New Ring Rearrangement Reactions Catalyzed by Flavoprotein Monooxygenases. Org. Lett. 2024, 26, 7489–7494. [Google Scholar] [CrossRef]
  21. Hoz, J.F.; Méndez, C.; Salas, J.A.; Olano, C. Novel Bioactive Paulomycin Derivatives Produced by Streptomyces albus J1074. Molecules 2017, 22, 1758. [Google Scholar] [CrossRef]
  22. Yuan, H.; Zheng, Y.; Yan, X.; Wang, H.; Zhang, Y.; Ma, J.; Fu, J. Direct cloning of a herpesvirus genome for rapid generation of infectious BAC clones. J. Adv. Res. 2023, 43, 97–107. [Google Scholar] [CrossRef] [PubMed]
  23. Fu, S.; An, Z.; Wu, L.; Xiang, Z.; Deng, Z.; Liu, R.; Liu, T. Evaluation and optimization of analytical procedure and sample preparation for polar Streptomyces albus J1074 metabolome profiling. Synth. Syst. Biotechnol. 2022, 7, 949–957. [Google Scholar] [CrossRef]
  24. Cao, M.; Zheng, C.; Yang, D.; Kalkreuter, E.; Adhikari, A.; Liu, Y.C.; Rateb, M.E.; Shen, B. Cryptic Sulfur incorporation in thioangucycline biosynthesis. Angew. Chem. Int. Ed. Engl. 2021, 60, 7140–7147. [Google Scholar] [CrossRef]
  25. Kharel, M.K.; Pahari, P.; Shepherd, M.D.; Tibrewal, N.; Nybo, S.E.; Shaaban, K.A.; Rohr, J. Angucyclines: Biosynthesis, mode-of-action, new natural products, and synthesis. Nat. Prod. Rep. 2012, 29, 264–325. [Google Scholar] [CrossRef]
  26. Su, H.; Shao, H.; Zhang, K.; Li, G. Antibacterial metabolites from the Actinomycete Streptomyces sp. P294. J. Microbiol. 2016, 54, 131–135. [Google Scholar] [CrossRef] [PubMed]
  27. Jackson, D.R.; Yu, X.; Wang, G.; Patel, A.B.; Calveras, J.; Barajas, J.F.; Sasaki, E.; Metsä-Ketelä, M.; Liu, H.W.; Rohr, J.; et al. Insights into complex oxidation during BE-7585A biosynthesis: Structural determination and analysis of the polyketide monooxygenase BexE. ACS Chem. Biol. 2016, 11, 1137–1147. [Google Scholar] [CrossRef] [PubMed]
  28. Gould, S.J.; Cheng, X.C. New benz[a]anthraquinone secondary metabolites from Streptomyces phaeochromogenes. J. Org. Chem. 1994, 59, 400–405. [Google Scholar] [CrossRef]
  29. Gao, H.M.; Xie, P.F.; Zhang, X.L.; Yang, Q. Isolation, phylogenetic and gephyromycin metabolites characterization of new exopolysaccharides-bearing antarctic actinobacterium from feces of emperor penguin. Mar. Drugs 2021, 19, 458. [Google Scholar] [CrossRef]
  30. Bringmann, G.; Lang, G.; Maksimenka, K.; Hamm, A.; Gulder, T.A.; Dieter, A.; Bull, A.T.; Stach, J.E.; Kocher, N.; Müller, W.E. Fiedler HP. Gephyromycin, the first bridged angucyclinone, from Streptomyces griseus strain NTK 14. Phytochemistry 2005, 66, 1366–1373. [Google Scholar] [CrossRef]
  31. Gao, Y.; Zhao, Y.; Zhou, J.; Yang, M.; Lin, L.; Wang, W.; Tao, M.; Deng, Z.; Jiang, M. Unexpected role of a short-chain dehydrogenase/reductase family protein in Type II polyketide biosynthesis. Angew. Chem. Int. Ed. Engl. 2022, 61, e202110445. [Google Scholar] [CrossRef]
  32. Ridley, C.P.; Khosla, C. Synthesis and biological activity of novel pyranopyrones derived from engineered aromatic polyketides. ACS Chem. Biol. 2007, 2, 104–108. [Google Scholar] [CrossRef]
  33. Xu, D.; Metz, J.; Harmody, D.; Peterson, T.; Winder, P.; Guzmán, E.A.; Russo, R.; McCarthy, P.J.; Wright, A.E.; Wang, G. Brominated and sulfur-containing angucyclines derived from a single pathway: Identification of nocardiopsistins D-F. Org. Lett. 2022, 24, 7900–7904. [Google Scholar] [CrossRef]
  34. Chen, Y.H.; Wang, C.C.; Greenwell, L.; Rix, U.; Hoffmeister, D.; Vining, L.C.; Rohr, J.; Yang, K.Q. Functional analyses of oxygenases in jadomycin biosynthesis and identification of JadH as a bifunctional oxygenase/dehydrase. J. Biol. Chem. 2005, 280, 22508–22514. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, C.; Zhang, L.; Zhang, W.; Huang, C.; Zhu, Y.; Jiang, X.; Liu, W.; Zhao, M.; De, B.C.; Zhang, C. Biochemical and structural insights of multifunctional flavin-dependent monooxygenase FlsO1-catalyzed unexpected xanthone formation. Nat. Commun. 2022, 13, 5386. [Google Scholar] [CrossRef] [PubMed]
  36. Monciardini, P.; Iorio, M.; Maffioli, S.; Sosio, M.; Donadio, S. Discovering new bioactive molecules from microbial sources. Microb. Biotechnol. 2014, 7, 209–220. [Google Scholar] [CrossRef] [PubMed]
  37. Enghiad, B.; Huang, C.; Guo, F.; Jiang, G.; Wang, B.; Tabatabaei, S.K.; Martin, T.A.; Zhao, H. Cas12a-assisted precise targeted cloning using in vivo Cre-lox recombination. Nat. Commun. 2021, 12, 1171. [Google Scholar] [CrossRef] [PubMed]
  38. Hou, Y.H.; Li, F.C.; Wang, S.J.; Qin, S.; Wang, Q.F. Intergeneric conjugation in holomycin-producing marine Streptomyces sp. strain M095. Microbiol. Res. 2008, 163, 96–104. [Google Scholar] [CrossRef]
  39. Du, L.; Liu, R.H.; Ying, L.; Zhao, G.R. An efficient intergeneric conjugation of DNA from Escherichia coli to mycelia of the lincomycin-producer Streptomyces lincolnensis. Int. J. Mol. Sci. 2012, 13, 4797–4806. [Google Scholar] [CrossRef]
  40. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  41. Xu, X.; Zhang, F.; Zhou, L.; Chang, Y.; Che, Q.; Zhu, T.; Li, D.; Zhang, G. Overexpression of global regulator SCrp leads to the discovery of new angucyclines in Streptomyces sp. XS-16. Mar. Drugs 2023, 21, 240. [Google Scholar] [CrossRef]
  42. Marcarino, M.O.; Cicetti, S.; Zanardi, M.M.; Sarotti, A.M. A critical review on the use of DP4+ in the structural elucidation of natural products: The good, the bad and the ugly. A practical guide. Nat. Prod. Rep. 2022, 39, 58–76. [Google Scholar] [CrossRef] [PubMed]
  43. Bruhn, T.; Schaumlöffel, A.; Hemberger, Y.; Bringmann, G. SpecDis: Quantifying the comparison of calculated and experimental electronic circular dichroism spectra. Chirality 2013, 25, 243–249. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, F.; Ma, C.; Che, Q.; Zhu, T.; Zhang, G.; Li, D. Extending the structural diversity of labdane diterpenoids from marine-derived fungus Talaromyces sp. HDN151403 using heterologous expression. Mar. Drugs 2023, 21, 628. [Google Scholar] [CrossRef] [PubMed]
  45. Lewis, K. The science of antibiotic discovery. Cell 2020, 181, 29–45. [Google Scholar] [CrossRef]
Figure 1. (a) The spi gene cluster in comparison to homologous gene clusters that encode saquayamycin, landomycin, frigocyclinone, and ossamine (oss). (b) ⅰ: HPLC analysis of heterologous expression strain J1074::spi; ⅱ: HPLC analysis of negative control host strain S. albus J1074::p15A.
Figure 1. (a) The spi gene cluster in comparison to homologous gene clusters that encode saquayamycin, landomycin, frigocyclinone, and ossamine (oss). (b) ⅰ: HPLC analysis of heterologous expression strain J1074::spi; ⅱ: HPLC analysis of negative control host strain S. albus J1074::p15A.
Marinedrugs 22 00480 g001
Figure 2. Chemical structures of 18.
Figure 2. Chemical structures of 18.
Marinedrugs 22 00480 g002
Figure 3. 1H-1H COSY, key HMBC and NOESY correlations of 14.
Figure 3. 1H-1H COSY, key HMBC and NOESY correlations of 14.
Marinedrugs 22 00480 g003
Figure 4. Calculated and experimental ECD spectra of 1.
Figure 4. Calculated and experimental ECD spectra of 1.
Marinedrugs 22 00480 g004
Figure 5. Calculated and experimental ECD spectra of 3.
Figure 5. Calculated and experimental ECD spectra of 3.
Marinedrugs 22 00480 g005
Figure 6. Calculated and experimental ECD spectra of 4.
Figure 6. Calculated and experimental ECD spectra of 4.
Marinedrugs 22 00480 g006
Scheme 1. Proposed biosynthetic pathways of compounds 18.
Scheme 1. Proposed biosynthetic pathways of compounds 18.
Marinedrugs 22 00480 sch001
Table 1. 1H Spectroscopic Data for 14 in DMSO-d6 (500 MHZ).
Table 1. 1H Spectroscopic Data for 14 in DMSO-d6 (500 MHZ).
No.1234
δH (J in Hz)δH (J in Hz)δH (J in Hz)δH (J in Hz)
25.71, s5.70, s5.97, s5.83, s
42.89, d (18.0)2.94, d (18.0)3.06, d (18.0)3.98, d (1.3)
2.06, d (18.0)2.13, d (18.0)2.02, d (18.0)-
51.98, m2.07, m1.76, m1.98, m
1.60, m1.63, m1.35, m-
62.28, m2.36, m2.53, m2.24, m
1.74, m1.84, m2.38, m2.14, m
97.34, d (8.0)7.35, d (8.0)7.31, m7.35, dd (8.0, 1.0)
107.76, t (8.0)7.73, t (8.0)7.69, dd (8.2, 7.6)7.76, dd (8.0, 7.6)
117.60, d (8.0)7.58, d (8.0)7.34, m7.34, dd (7.6, 1.0)
12b---3.14, s
4a-OH6.32, s-2.54, s5.59, s
6a-OH6.03, s--7.05, s
12a-OH7.01, s---
12b-OH5.66, s---
8-OH11.10, s11.17, s11.06, s11.72, s
131.93, s1.96, s1.81, s2.33, d (1.2)
S-CH3-1.67, s--
Table 2. 13C Spectroscopic Data for 14 in DMSO-d6 (125 MHZ).
Table 2. 13C Spectroscopic Data for 14 in DMSO-d6 (125 MHZ).
No.1234
δC, TypeδC, TypeδC, TypeδC, Type
1192.8, C193.5, C196.7, C195.6, C
2122.9, CH122.7, CH122.6, CH124.6, CH
3158.2, C159.4, C156.1, C162.4, C
441.8, CH241.6, CH238.2, CH285.1, CH
4a75.5, C75.7, C65.1, C81.6, C
528.2, CH228.8, CH225.8, CH232.6, CH2
621.6, CH217.9, CH220.1, CH226.8, CH2
6a77.6, C61.2, C74.2, C78.0, C
7200.5, C195.5, C194.5, C201.6, C
7a116.3, C115.4, C114.7, C114.9, C
8159.9, C160.0, C159.5, C161.1, C
9123.6, CH124.4, CH123.9, CH123.7, CH
10136.2, CH136.0, CH136.4, CH137.7, CH
11118.9, CH119.0, CH118.9, CH118.3, CH
11a133.0, C131.5, C132.7, C135.3, C
12194.1, C191.0, C189.6, C192.1, C
12a76.4, C78.4, C67.5, C92.5, C
12b76.2, C75.7, C73.7, C52.7, CH
1323.6, CH323.7, CH323.6, CH326.3, CH3
S-CH3-13.3, CH3--
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, X.; Zhang, F.; Li, C.; Li, J.; Xu, X.; Zhu, T.; Che, Q.; Li, D.; Zhang, G. Heterologous Expression of Type II PKS Gene Cluster Leads to Diversified Angucyclines in Streptomyces albus J1074. Mar. Drugs 2024, 22, 480. https://doi.org/10.3390/md22110480

AMA Style

Zhang X, Zhang F, Li C, Li J, Xu X, Zhu T, Che Q, Li D, Zhang G. Heterologous Expression of Type II PKS Gene Cluster Leads to Diversified Angucyclines in Streptomyces albus J1074. Marine Drugs. 2024; 22(11):480. https://doi.org/10.3390/md22110480

Chicago/Turabian Style

Zhang, Xiaoting, Falei Zhang, Chen Li, Jiayi Li, Xiao Xu, Tianjiao Zhu, Qian Che, Deihai Li, and Guojian Zhang. 2024. "Heterologous Expression of Type II PKS Gene Cluster Leads to Diversified Angucyclines in Streptomyces albus J1074" Marine Drugs 22, no. 11: 480. https://doi.org/10.3390/md22110480

APA Style

Zhang, X., Zhang, F., Li, C., Li, J., Xu, X., Zhu, T., Che, Q., Li, D., & Zhang, G. (2024). Heterologous Expression of Type II PKS Gene Cluster Leads to Diversified Angucyclines in Streptomyces albus J1074. Marine Drugs, 22(11), 480. https://doi.org/10.3390/md22110480

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop