Next Article in Journal
Is It Possible to Differentiate Pneumocystis jirovecii Pneumonia and Colonization in the Immunocompromised Patients with Pneumonia?
Next Article in Special Issue
Immunosuppressive Sesquiterpenoids from the Edible Mushroom Craterellus odoratus
Previous Article in Journal
Metal Nanoparticles as Novel Antifungal Agents for Sustainable Agriculture: Current Advances and Future Directions
Previous Article in Special Issue
Sesquiterpenoids Specially Produced by Fungi: Structures, Biological Activities, Chemical and Biosynthesis (2015–2020)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Trichoderma reesei Contains a Biosynthetic Gene Cluster That Encodes the Antifungal Agent Ilicicolin H

1
Institute for Organic Chemistry and Biomolekulares Wirkstoffzentrum (BMWZ), Schneiderberg 38, 30167 Hannover, Germany
2
Pharmacognosy Department, Faculty of Pharmacy, Alexandria University, Alexandria 21521, Egypt
*
Author to whom correspondence should be addressed.
J. Fungi 2021, 7(12), 1034; https://doi.org/10.3390/jof7121034
Submission received: 8 November 2021 / Revised: 26 November 2021 / Accepted: 29 November 2021 / Published: 1 December 2021

Abstract

:
The trili biosynthetic gene cluster (BGC) from the well-studied organism Trichoderma reesei was studied by heterologous expression in the fungal host Aspergillus oryzae. Coexpression of triliA and triliB produces two new acyl tetramic acids. Addition of the ring-expanding cytochrome P450 encoded by triliC then yields a known pyridone intermediate to ilicicolin H and a new chain-truncated shunt metabolite. Finally, addition of the intramolecular Diels-Alderase encoded by triliD affords a mixture of 8-epi ilicicolin H and ilicicolin H itself, showing that the T. reesei trili BGC encodes biosynthesis of this potent antifungal agent. Unexpected A. oryzae shunt pathways are responsible for the production of the new compounds, emphasising the role of fungal hosts in catalysing diversification reactions.

1. Introduction

Trichoderma reesei (Hypocrea jecorina) is a filamentous fungus widely used for the industrial production of cellulase [1]. The organism can grow to high cell density and can grow robustly on a variety of low-quality substrates. It has been subject to a wide variety of biotechnological improvements since the discovery of the wild-type QM6A strain in the 1940s, and many molecular tools are available for its genetic manipulation. Many Trichoderma species have been used in the remediation of fungal diseases. For example, Foster and coworkers have shown that some Trichoderma spp. can effectively treat Armillaria mellea (honey fungus) that causes root rot in many tree species [2]. T. reesei is also well-known for the production of a number of secondary metabolites such as sorbicillinol 1 [3,4], and numerous multimeric derivatives such as bisvertinolone 2 [5].
Other members of the Trichoderma genus are well-known producers of many secondary metabolites [6]. In particular, compounds such as acyl tetramic acids (acyl 2-pyrollones) and acyl 2-pyridones, synthesised by hybrid polyketide synthase-nonribosomal peptide synthetases (PKS-NRPS) are commonly produced by various Trichoderma species. These include metabolites such as harzianic acid 3 (T. harzianum) [7,8], and deacetyl-18-deoxy cytochalasin H 4 [9] among the pyrollones; and trichodin A 5 [10] and harzianopyridone 6 [11,12] among the 2-pyridones (Figure 1). The biosynthesis of related pyrollones such as fusarin C 7 in Fusarium spp [13] and acurin A 8 in Aspergillus aculeatus, [14] and 2-pyridones such as tenellin 19 in Beauveria bassiana [15] and ilicicolin H 10 in Talaromyces variabilis and Nectria spp [16,17] are well-understood.
However, despite decades of study, T. reesei itself has not been reported to produce any PKS-NRPS compounds. Since we have a broad interest in fungal PKS-NRPS compounds and we recently made detailed studies on the biosynthesis of sorbicillinoids in T. reesei [3,4], we set out to determine if this organism can also produce PKS-NRPS compounds.

2. Materials and Methods

All chemicals and media ingredients used in this work were purchased from Duchefa Biochemie, Roth, VWR, Fisher scientific, Sigma Aldrich, abcr and Formedium. All media, buffers, solutions and antibiotics were prepared with double-distilled water (ddH2O), where deionized water was further purified by GenPure Pro UV/UF Milipore device from Thermo scientific. All growth media and sterile solutions were autoclaved at 121 °C for 15 min using Systec VX150 or a classic Prestige Medical 2100 autoclave. Antibiotic solutions were sterilized using a sterile syringe filter (0.45 µm pore size, Roth). See ESI for details of all growth media, buffers and solutions, enzymes and antibiotics.

2.1. Microbiological Methods

Bacterial and fungal strains and all microbiological methods used in this work are summarized in the Electronic Supplementary Information (ESI).

2.2. Construction of Heterologous Expression Vectors

The genomic DNA (gDNA) of T. reesei QM6a ∆tmus53 ∆pyr4 [18] was used as a template to amplify genes for yeast homologous recombination (without removal of introns). The vector pEYA was used to assemble the PKS-NRPS (TRIREDRAFT_58285, triliA) gene (~12 kb) in four different DNA fragments of around 3 kb each using yeast homologous recombination. This was followed by LR recombination between the entry plasmid pEYA·triliA (MSIII139) and the expression vector pTYGS·arg to yield vector pTYGS·argB·triliA (MSIII144). The enoyl reductase gene (TRIREDRAFT_58289, triliB) was also amplified from the gDNA of T. reesei using primers with overhang to PgpdA and TgpdA. Using yeast homologous recombination, the final vector pTYGS·argB·triliA·triliB (MSIII152) was constructed which contains the PKS-NRPS (triliA)- and ER (triliB)-encoding genes from the specified cluster. A vector for the expression of the first three genes of the cluster; PKS/NRPS (TRIREDRAFT_58285, triliA), ER (TRIREDRAFT_58289, triliB), P450RE (TRIREDRAFT_58953, triliC) was constructed using yeast homologous recombination using gDNA of T. reesei QM6a·∆tmus53·∆Pyr4 as a template for the amplification of the genes. A vector for the expression of the first four genes of the cluster; PKS/NRPS (TRIREDRAFT_58285, triliA), ER (TRIREDRAFT_58289, triliB), P450RE (TRIREDRAFT_58953, triliC) and Diels-Alderase (triliD) was constructed via yeast homologous recombination using gDNA of T. reesei QM6a·∆tmus53·∆Pyr4 as template for the amplification of the genes.

2.3. Chemical Analysis

All the chemicals and materials were purchased from one of the following companies: Bio-Rad (München, Germany), New England Biolabs (Beverly, MA, USA), Roth (Karlsruhe, Germany), Sigma Aldrich (Steinheim, Germany), and Thermo Fisher Scientific (Waltham, MA, USA).

2.3.1. Liquid Chromatography–Mass Spectrometry (LC–MS)

Analytical LC–MS data of the organic extracts were obtained by either a Waters mass-directed autopurification system (Waters, Wilmslow, UK) including a Waters 2767 autosampler, a Waters 2545 pump and Phenomenex Kinetex column (2.6 μm, C18, 100 Å, 4.6 mm × 100 mm; Phenomenex) equipped with a Phenomenex Security Guard precolumn (Luna, C5, 300 Å). Detection was performed by a diode array detector from 210–600 nm (DAD; Waters 2998 or Waters 996), an electron light-scattering detector (ELSD; Waters 2424) and an electrospray ionisation mass detector in the range of 100–1000 m/z (Waters SQD-2). Gradient was run over 15 min starting at 10% acetonitrile/ 90% HPLC grade water (+ 0.05% formic acid) and ramping to 90% acetonitrile/ 10% HPLC grade water (+ 0.05% formic acid). Flowrate was 1 mL/min and 20 μL of the sample was injected. Data were displayed using the software MassLynx. The LC–MS program is summarized in the ESI.
Single compounds were purified from the raw organic extracts by a Waters mass-directed autopurification system. It comprises a Waters 2767 autosampler, a Waters 2545 binary gradient module system and a Phenomenex Kinetex Axia column (5 μm, C18, 100 Å, 21.2 mm × 250 mm) equipped with a Phenomenex Security Guard precolumn (Luna, C5, 300 Å). A water/acetonitrile gradient was run over 15 min with a flowrate of 20 mL/min and a post-column flow split of 100:1. The minority flow was applied for simultaneous analysis by a diode array detector (Waters 2998) in the range 210 to 1600 nm, an evaporative light-scattering detector (ELSD; Waters 2424) and a Waters SQD-2 mass detector, operating in ES+ and ES− modes between 100 and 1000 m/z. Selected peaks were collected into test tubes and solvent was evaporated under reduced pressure. Compounds were dissolved in methanol (1 mg/mL). High-Resolution Mass Spectrometry was performed on a Q-Tof Premier mass spectrometer (Waters) coupled to an Acquity UPLC–domain (Waters). Electron spray Ionisation (EsI) mass spectroscopy was measured in positive or negative mode depending on the compound.

2.3.2. Nuclear Magnetic Resonance (NMR) Analysis

Bruker Ascend 400 MHz, Bruker DRX 500 MHz or a Bruker Ascend 600 MHz Spectrometer (Bruker) were used for NMR measurements of the samples. Raw data were then analyzed using the software Bruker TopSpin 3.5. Chemical shifts are expressed in parts per million (ppm) in comparison to the Tetramethylsilane (TMS) standard and are referenced to the deuterated solvent.

2.3.3. Extraction of Fungal Cultures

For A. oryzae transformants, a small cell sample was dried by Büchner filtration or gravity filtration and used for gDNA analysis. Mycelia were ground using a hand blender and removed from the culture supernatant by Büchner filtration. Supernatant was acidified with 2 M HCL to pH 3–4 and extracted twice with ethyl acetate. Combined organic layers were dried over MgSO4 and solvents were removed under vacuum. Organic residue was dissolved in methanol or acetonitrile to a concentration of 5–10 mg/mL (analytical) or 50 mg/mL (preparative) and filtered over glass wool.

3. Results

The genome of T. reesei QM6A, the wild-type isolate, was sequenced in 2008 [19]. We examined this genome sequence using a combination of automated AntiSMASH [20] and manual screening, with a particular focus on PKS and PKS-NRPS encoding genes. Bioinformatic analysis of all PKS and PKS-NRPS genes in T. reesei QM6a revealed the presence of 11 PKS and two hybrid PKS-NRPS encoding genes (see ESI for details). Of these, only the two PKS genes responsible for the sorbicillins and related compounds have been identified and confirmed (TRIREDRAFT_73618 and 73621) [4,21,22]. The HR-PKS TRIREDRAFT_65116 shows a very high similarity to the recently investigated PKS from Trichoderma virens, which is encoded in the biosynthetic gene cluster (BGC) that directs the production of virensols and trichoxide [23]. However, the other PKS and PKS-NRPS BGCs could not be linked to the production of any specific compound.
Although the T. reesei genome contains two hybrid PKS-NRPS encoding genes, no PKS-NRPS secondary metabolites have been reported from this fungus and these BGCs are likely to be silent under laboratory conditions. The two PKS-NRPS themselves show similarity (>60%) to proteins involved in fusarin C 8 biosynthesis in Fusarium moniliforme [13] and illicicolin H 10 biosynthesis in Penicillium variabile (= Talaromyces variabilis) [16]. Ilicicolin H 10 is an interesting antifungal agent that targets the cytochrome bc1 respiration complex in the 3–5 nM range [24], and it is conceivable that ilicicolin H 10 production may be responsible for the observed antifungal effects of Trichoderma species in general [2] and T. reesei in particular [25]. We therefore selected the putative T. reesei ilicicolin H BGC (trili) for further investigation.
Based on the NCBI prediction, the trili BGC (Figure 2) consists of a hybrid PKS/NRPS (triliA), an ER (triliB), and a putative cytochrome P450 (triliC). Reannotation of the cluster by FungiSMASH and Clinker [26] and Artemis comparison to a closely related BGC in T. citrinoviride resulted in the identification of a putative S-adenosylmethionine (SAM)-dependent Diels-Alderase in the cluster (triliD), that shows high similarity to iccD from the illicicolin-H BGC in Talaromyces variabilis [16]. Finally, triliE encodes a putative TIM-like epimerase homologous to IccE, known to epimerise epi-ilicicolin H in T. variabilis. Multigene analysis using cblaster [27] identified highly similar ili BGCs in other strains of T. reesei (RUT C-30) and other Trichoderma species, including T. parareesei, T. virens Gv29-8, T. citrinoviride TUCIM6016 and T. longibrachiatum ATCC 18648 (see ESI).
In an initial heterologous expression experiment, triliA and triliB were cloned into the vector pTYGS·argB [28] using rapid recombination of genome-derived PCR fragments in yeast and in vitro (Gateway) recombination. The PKS-NRPS-encoding gene triliA was placed downstream of the inducible A. oryzae amyB promoter (PamyB), while triliB was placed under the control of the Aspergillus gpdA promoter (PgpdA). The resulting vector pTYGS·argB·triliA·triliB was transformed into the fungal host Aspergillus oryzae NSAR1 [29]. Comparison of the organic extracts of transformants with those of untransformed A. oryzae NSAR1 revealed the production of new compounds by LC–MS analysis eluting at 6.7 min, 8.0 min and 11.0 min (Figure 3).
Compound 11 (RT 6.7 min, 30 mg) was isolated by preparative LC–MS, and the structure was elucidated using 1D and 2D NMR. HRMS analysis of 11 confirmed a molecular formula of C23H28NO6 ([M+H]+ calculated 414.1917, found 414.1917). Compound 12 (RT 8.0 min, 4 mg) was isolated by preparative LC–MS and the structure was elucidated using NMR. HRMS analysis of 12 confirmed a molecular formula of C27H32NO6 ([M+H]+ calculated 466.2230, found 466.2237).
NMR analysis of these compounds (see ESI for full details) showed that 11 is a linear acyl tetramic acid with a C12 sidechain terminating in an unusual (at this position) carboxylic acid. Minor compound 12 was shown to be a related acyl tetramic acid featuring a decalin, with an unusual 3-propenoic acid substituent. Although the minor compound 13 at 11.0 min could not be fully identified by NMR, mass spectroscopic and UV analysis suggested that it is likely to be a third related tetramic acid with a C16 sidechain, first reported by Tang and coworkers during biosynthetic studies of ilicicolin H (Scheme 1) [16].
In the next round of experiments, triliC, encoding a cytochrome P450 oxidase, was added to the A. oryzae strains, driven by Padh. TriliC is related to ring-expanding cytochrome P450 oxidases first characterised from the tenellin pathway [15]. Coexpression of triliA, triliB and triliC resulted in production of two new compounds in comparison to the previous transformants and the untransformed A. oryzae (Figure 4), eluting at 6.7 (14), and 10.8 (15) min. Full NMR analysis of the RT 6.7 min compound showed it to be the pyridone homolog of 11. Similarly, the 10.8 min compound was shown to be the pyridone homolog of 13.
Finally, a gene homologous to Diels-Alderase iccD (triliD) was added to the expression system downstream of Peno and transformed into A. oryzae. This experiment again produced 11 (RT 6.7 min). Additionally, two new peaks in comparison to the previous transformants and the untransformed A. oryzae were also observed by LC–MS (Figure 5) at RT 10.76 and 11.0 min. The minor compound at 11.0 min was again assigned as 13 after analysis of partial data. NMR and HRMS analysis (see ESI for full details) showed the peak at 10.76 min consists of an almost coeleuting mixture of the known 8-epi ilicicolin H 16 and ilicicolin H 10 itself.

4. Discussion

PKS-NRPS compounds have been reported before from the genus Trichoderma, but T. reesei QM6A itself has not been reported to produce these compounds. Genomic analysis, however, reveals two typical fungal PKS-NRPS BGCs. Heterologous expression experiments reveal that the trili BGC is fully functional, but apparently silent in T. reesei QM6A under laboratory conditions. Initial expression of the PKS-NRPS (triliA) and trans-ER (triliB) genes in A. oryzae led to production of two new acyl tetramic acids (11 and 12) and a third compound 13 that is likely to be a precursor of illicicolin H, recently described by Tang and coworkers [16] from Nectria sp. B13 and T. variabilis and Gao and coworkers [17] in Neonectria sp. DH2. Decalin 12 appears to have been oxidised to a carboxylic acid at its terminal polyketide methyl group. Addition of the ring-expanding P450 oxidase then produces the corresponding 2-pyridones 14 and 15. Compound 15 is identical to an ilicicolin H precursor previously identified by Tang and coworkers [16], but 14 is new. Finally, addition of the Diels-Alderase encoded by triliD produces a mixture of 8-epi ilicicolin H 16 and ilicicolin H 10 itself.
The T. variabilis icc [16] and Neonectra DH2 ili [17] BGCs were previously investigated by expression in A. nidulans. Our results are highly similar to those obtained from these studies, except that A. oryzae as a host appears to catalyse some unexpected steps. For example, in A. oryzae the initial tetramic acid precursor 13 appears to be oxidatively degraded to remove the terminal four polyketide carbons and leaves a carboxylic acid in the structure of 11. Previous heterologous expression studies on A. oryzae also indicate that this host has a propensity for the oxygenation of polyenes [30,31] and oxidative cleavage of olefins [32]. Likewise, oxygenation of the terminal methyl of 12 to a carboxylic acid is newly observed in A. oryzae. This is similar to the BueE-catalysed oxidation recently observed by Piggott and coworkers in the case of the fungal polyketide decalin burnettiene A in Aspergillus burnettii [33]. It also appears that a spontaneous (or A. oryzae-catalysed) Diels–Alder reaction can occur to produce a decalin that is then oxidised to a carboxylic acid at the terminal methyl to generate new compound 12. In previous studies on ilicicolin H biosynthesis, using A. nidulans as the host, a specific Diels–Alder catalyst was required for this step. We also observed the chain-truncated pyridone 14—again a new natural product. In A. nidulans it appears that 8-epi ilicicolin H 16 is a stable compound, requiring expression of iccE for the conversion to ilicicolin H. However, in A. oryzae, observable amounts of ilicicolin H 10 are produced in the absence of this epimerase catalyst.
Our results therefore illustrate that T. reesei QM6A has a fully functional PKS-NRPS BGC that encodes the biosynthesis of ilicicolin H, a potent antifungal agent. The ability of T. reesei QM6A to make this compound has never before been observed [34], despite nearly 70 years of laboratory-based investigations. It is likely that this BGC can be activated under unknown environmental conditions in its native habitat and may contribute to its utility as an agricultural antifungal agent, but it appears to be silent under laboratory conditions. Finally, our results also show that different fungal heterologous hosts can introduce unexpected shunt pathways which can diversify known secondary metabolites and produce new-to-nature compounds.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/jof7121034/s1. All experimental details including growth media, buffers, enzyme and antibiotic solutions, bacterial and fungal strains, vectors and oligonucleotides, bioinformatic procedures and results, structure elucidation details, tabulated NMR data, NMR spectra, and HRMS data.

Author Contributions

M.L.S. and R.J.C. designed the study. Experimental work was performed by M.L.S., assisted by M.A. The paper was written by R.J.C. and M.L.S., R.J.C. was responsible for providing physical resources. Funding was gained by M.L.S. (DAAD GERLS) and R.J.C. All authors have read and agreed to the published version of the manuscript.

Funding

M.L.S. thanks the German Academic Exchange Service (DAAD) and the Egyptian Ministry of Higher Education and Scientific Research (MHESR) for funding (GERLS programme, 2017, 57311832). This work was supported by DFG [CO 1328/2-1, INST 187/621-1, INST 187/686-1)]. The APC was funded by the Open Access Fund of the Leibniz Universität Hannover.

Acknowledgments

The authors would like to thank Eric Kuhnert and Kevin Becker for the helpful discussions and Astrid Mach-Aigner for donating T. reesei strain.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. Beier, S.; Hinterdobler, W.; Monroy, A.A.; Bazafkan, H.; Schmoll, M. The Kinase USK1 Regulates Cellulase Gene Expression and Secondary Metabolite Biosynthesis in Trichoderma reesei. Front. Microbiol. 2020, 11, 974. [Google Scholar] [CrossRef] [PubMed]
  2. Rees, H.J.; Bashir, N.; Drakulic, J.; Cromey, M.G.; Bailey, A.M.; Foster, G.D. Identification of native endophytic Trichoderma spp. for investigation of in vitro antagonism towards Armillaria mellea using synthetic- and plant-based substrates. J. Appl. Microbiol. 2021, 131, 392–403. [Google Scholar] [CrossRef] [PubMed]
  3. Al Fahad, A.; Abood, A.; Fisch, K.M.; Osipow, A.; Davison, J.; Avramović, M.; Butts, C.P.; Piel, J.; Simpson, T.J.; Cox, R.J. Oxidative dearomatisation: The key step of sorbicillinoid biosynthesis. Chem. Sci. 2014, 5, 523–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kahlert, L.; Bassiony, E.F.; Cox, R.; Skellam, E.J. Diels–Alder Reactions during the Biosynthesis of Sorbicillinoids. Angew. Chem. Int. Ed. 2020, 59, 5816–5822. [Google Scholar] [CrossRef] [Green Version]
  5. Harned, A.M.; Volp, K.A. The sorbicillinoid family of natural products: Isolation, biosynthesis, and synthetic studies. Nat. Prod. Rep. 2011, 28, 1790–1810. [Google Scholar] [CrossRef]
  6. Shenouda, M.L.; Cox, R.J. Molecular methods unravel the biosynthetic potential of Trichoderma species. RSC Adv. 2021, 11, 3622–3635. [Google Scholar] [CrossRef]
  7. Sawa, R.; Mori, Y.; Iinuma, H.; Naganawa, H.; Hamada, M.; Yoshida, S.; Furutani, H.; Kajimura, Y.; Fuwa, T.; Takeuchi, T. Harzianic acid, a new antimicrobial antibiotic from a fungus. J. Antibiot. 1994, 47, 731–732. [Google Scholar] [CrossRef] [Green Version]
  8. Xie, L.; Zang, X.; Cheng, W.; Zhang, Z.; Zhou, J.; Chen, M.; Tang, Y. Harzianic Acid from Trichoderma afroharzianum Is a Natural Product Inhibitor of Acetohydroxyacid Synthase. J. Am. Chem. Soc. 2021, 143, 9575–9584. [Google Scholar] [CrossRef]
  9. Chen, H.; Daletos, G.; Okoye, F.; Lai, D.; Dai, H.; Proksch, P. A New Cytotoxic Cytochalasin from the Endophytic Fungus Trichoderma harzianum. Nat. Prod. Commun. 2015, 10, 585–587. [Google Scholar] [CrossRef] [Green Version]
  10. Wu, B.; Oesker, V.; Wiese, J.; Schmaljohann, R.; Imhoff, J.F. Two New Antibiotic Pyridones Produced by a Marine Fungus, Trichoderma sp. Strain MF106. Mar. Drugs 2014, 12, 1208–1219. [Google Scholar] [CrossRef] [Green Version]
  11. Dickinson, J.M.; Hanson, J.R.; Hitchcock, P.B.; Claydon, N. Structure and biosynthesis of harzianopyridone, an antifungal metabolite of Trichoderma harzianum. J. Chem. Soc. Perkin Trans. 1 1989, 1, 1885–1887. [Google Scholar] [CrossRef]
  12. Bat-Erdene, U.; Kanayama, D.; Tan, D.; Turner, W.C.; Houk, K.N.; Ohashi, M.; Tang, Y. Iterative Catalysis in the Biosynthesis of Mitochondrial Complex II Inhibitors Harzianopyridone and Atpenin B. J. Am. Chem. Soc. 2020, 142, 8550–8554. [Google Scholar] [CrossRef]
  13. Song, Z.; Cox, R.J.; Lazarus, C.M.; Simpson, T.J. Fusarin C Biosynthesis in Fusarium moniliforme and Fusarium venenatum. ChemBioChem 2004, 5, 1196–1203. [Google Scholar] [CrossRef]
  14. Wolff, P.B.; Nielsen, M.L.; Slot, J.C.; Andersen, L.N.; Petersen, L.M.; Isbrandt, T.; Holm, D.K.; Mortensen, U.H.; Nødvig, C.S.; Larsen, T.O.; et al. Acurin A, a novel hybrid compound, biosynthesized by individually translated PKS- and NRPS-encoding genes in Aspergillus aculeatus. Fungal Genet. Biol. 2020, 139, 103378. [Google Scholar] [CrossRef]
  15. Halo, L.M.; Heneghan, M.N.; Yakasai, A.A.; Song, Z.; Williams, K.; Bailey, A.; Cox, R.; Lazarus, C.M.; Simpson, T.J. Late Stage Oxidations during the Biosynthesis of the 2-Pyridone Tenellin in the Entomopathogenic FungusBeauveria bassiana. J. Am. Chem. Soc. 2008, 130, 17988–17996. [Google Scholar] [CrossRef]
  16. Zhang, Z.; Jamieson, C.S.; Zhao, Y.-L.; Li, D.; Ohashi, M.; Houk, K.N.; Tang, Y. Enzyme-Catalyzed Inverse-Electron Demand Diels–Alder Reaction in the Biosynthesis of Antifungal Ilicicolin H. J. Am. Chem. Soc. 2019, 141, 5659–5663. [Google Scholar] [CrossRef]
  17. Lin, X.; Yuan, S.; Chen, S.; Chen, B.; Xu, H.; Liu, L.; Li, H.; Gao, Z. Heterologous Expression of Ilicicolin H Biosynthetic Gene Cluster and Production of a New Potent Antifungal Reagent, Ilicicolin J. Molecules 2019, 24, 2267. [Google Scholar] [CrossRef] [Green Version]
  18. Derntl, C.; Kiesenhofer, D.P.; Mach, R.L.; Mach-Aigner, A.R. Novel Strategies for Genomic Manipulation of Trichoderma reesei with the Purpose of Strain Engineering. Appl. Environ. Microbiol. 2015, 81, 6314–6323. [Google Scholar] [CrossRef] [Green Version]
  19. Chia-Ling, C.; Huang, C.-H.; Chen, C.-L.; Tung, S.-Y.; Wang, T.-F. Trichoderma reesei complete genome sequence, repeat-induced point mutation, and partitioning of CAZyme gene clusters. Biotechnol. Biofuels 2017, 10, 170. [Google Scholar] [CrossRef]
  20. Blin, K.; Shaw, S.; Kloosterman, A.M.; Charlop-Powers, Z.; van Wezel, G.P.; Medema, M.H.; Weber, T. antiSMASH 6.0: Improving cluster detection and comparison capabilities. Nucleic Acids Res. 2021, 49, W29–W35. [Google Scholar] [CrossRef]
  21. Derntl, C.; Rassinger, A.; Srebotnik, E.; Mach, R.L.; Mach-Aigner, A.R. Identification of the Main Regulator Responsible for Synthesis of the Typical Yellow Pigment Produced by Trichoderma reesei. Appl. Environ. Microbiol. 2016, 82, 6247–6257. [Google Scholar] [CrossRef] [Green Version]
  22. Derntl, C.; Guzman-Chavez, F.; Mello-De-Sousa, T.M.; Busse, H.-J.; Driessen, A.J.M.; Mach, R.; Mach-Aigner, A.R. In Vivo Study of the Sorbicillinoid Gene Cluster in Trichoderma reesei. Front. Microbiol. 2017, 8, 2037. [Google Scholar] [CrossRef]
  23. Liu, L.; Tang, M.-C.; Tang, Y. Fungal Highly Reducing Polyketide Synthases Biosynthesize Salicylaldehydes That Are Precursors to Epoxycyclohexenol Natural Products. J. Am. Chem. Soc. 2019, 141, 19538–19541. [Google Scholar] [CrossRef]
  24. Singh, S.B.; Liu, W.; Li, X.; Chen, T.; Shafiee, A.; Card, D.; Abruzzo, G.; Flattery, A.; Gill, C.; Thompson, J.R.; et al. Antifungal Spectrum, In Vivo Efficacy, and Structure–Activity Relationship of Ilicicolin H. ACS Med. Chem. Lett. 2012, 3, 814–817. [Google Scholar] [CrossRef] [Green Version]
  25. Damodaran, T.; Rajan, S.; Muthukumar, M.; Gopal, R.; Yadav, K.; Kumar, S.; Ahmad, I.; Kumari, N.; Mishra, V.K.; Jha, S.K. Biological Management of Banana Fusarium Wilt Caused by Fusarium Oxysporum f. sp. cubense Tropical Race 4 Using Antagonistic Fungal Isolate CSR-T-3 (Trichoderma Reesei). Front. Microbiol. 2020, 11, 595845. [Google Scholar] [CrossRef]
  26. Gilchrist, C.L.M.; Chooi, Y.-H. Clinker & clustermap.js: Automatic generation of gene cluster comparison figures. Bioinformatics 2021, 37, 2473–2475. [Google Scholar] [CrossRef]
  27. Gilchrist, C.L.M.; Booth, T.J.; van Wersch, B.; van Grieken, L.; Medema, M.H.; Chooi, Y.-H. cblaster: A remote search tool for rapid identification and visualization of homologous gene clusters. Bioinform. Adv. 2021, 1, vbab016. [Google Scholar] [CrossRef]
  28. Pahirulzaman, K.A.K.; Williams, K.; Lazarus, C.M. A Toolkit for Heterologous Expression of Metabolic Pathways in Aspergillus oryzae. Methods Enzymol. 2012, 517, 241–260. [Google Scholar] [CrossRef]
  29. Jin, F.J.; Maruyama, J.-I.; Juvvadi, P.; Arioka, M.; Kitamoto, K. Development of a novel quadruple auxotrophic host transformation system byargBgene disruption usingadeAgene and exploiting adenine auxotrophy inAspergillus oryzae. FEMS Microbiol. Lett. 2004, 239, 79–85. [Google Scholar] [CrossRef] [Green Version]
  30. Song, Z.; Bakeer, W.; Marshall, J.W.; Yakasai, A.A.; Khalid, R.M.; Collemare, J.; Skellam, E.; Tharreau, D.; Lebrun, M.-H.; Lazarus, C.M.; et al. Heterologous expression of the avirulence gene ACE1 from the fungal rice pathogen Magnaporthe oryzae. Chem. Sci. 2015, 6, 4837–4845. [Google Scholar] [CrossRef] [Green Version]
  31. Heneghan, M.N.; Yakasai, A.A.; Williams, K.; Kadir, K.A.; Wasil, Z.; Bakeer, W.; Fisch, K.M.; Bailey, A.M.; Simpson, T.J.; Cox, R.J.; et al. The programming role of trans-acting enoyl reductases during the biosynthesis of highly reduced fungal polyketides. Chem. Sci. 2011, 2, 972–979. [Google Scholar] [CrossRef]
  32. Fujii, R.; Ugai, T.; Ichinose, H.; Hatakeyama, M.; Kosaki, T.; Gomi, K.; Fujii, I.; Minami, A.; Oikawa, H. Reconstitution of biosynthetic machinery of fungal polyketides: Unexpected oxidations of biosynthetic intermediates by expression host. Biosci. Biotechnol. Biochem. 2016, 80, 426–431. [Google Scholar] [CrossRef] [Green Version]
  33. Roux, I.; Bowles, S.; Kalaitzis, J.A.; Vuong, D.; Lacey, E.; Chooi, Y.-H.; Piggott, A.M. Characterisation and Heterologous Biosynthesis of Burnettiene A, a New Polyene-Decalin Polyketide from Aspergillus Burnettii. Org. Biomol. Chem. 2021, 19, 9506–9513. [Google Scholar] [CrossRef] [PubMed]
  34. Rush, T.A.; Shrestha, H.K.; Meena, M.G.; Spangler, M.K.; Ellis, J.C.; Labbé, J.L.; Abraham, P.E. Bioprospecting Trichoderma: A Systematic Roadmap to Screen Genomes and Natural Products for Biocontrol Applications. Front. Fungal Biol. 2021, 2, 716511. [Google Scholar] [CrossRef]
Figure 1. Typical compounds found in various Trichoderma species (16) and related fungal metabolites (710).
Figure 1. Typical compounds found in various Trichoderma species (16) and related fungal metabolites (710).
Jof 07 01034 g001
Figure 2. Annotation of T. reesei ilicicolin H BGC (trili) and comparison to the Talaromyces variabilis icc BGC. Ili proteins are from the Neonectria sp. DH2 ilicicolin H biosynthetic pathway [17]; Icc proteins are from the T. variabilis ilicicolin H biosynthetic pathway [16].
Figure 2. Annotation of T. reesei ilicicolin H BGC (trili) and comparison to the Talaromyces variabilis icc BGC. Ili proteins are from the Neonectria sp. DH2 ilicicolin H biosynthetic pathway [17]; Icc proteins are from the T. variabilis ilicicolin H biosynthetic pathway [16].
Jof 07 01034 g002
Figure 3. LC–MS traces (DAD 210–600 nm) of transformed A. oryzae with illiA and illiB. (A), the extract of the wild-type A. oryzae NSAR1. (BE), extracts of four transformants of A. oryzae NSAR1 with triliA and triliB showing the production of different PKS-NRPS-related compounds at RT 6.7, 8.0 and 11.0 min.
Figure 3. LC–MS traces (DAD 210–600 nm) of transformed A. oryzae with illiA and illiB. (A), the extract of the wild-type A. oryzae NSAR1. (BE), extracts of four transformants of A. oryzae NSAR1 with triliA and triliB showing the production of different PKS-NRPS-related compounds at RT 6.7, 8.0 and 11.0 min.
Jof 07 01034 g003
Figure 4. LC–MS traces (DAD 210–600 nm) of transformed A. oryzae NSAR1 with triliA, triliB and triliC. (A), the extract of untransformed A. oryzae NSAR1; (BE), extracts of four transformants of A. oryzae with triliA, triliB and triliC. New compounds were observed at RT 6.7 and 10.8 min.
Figure 4. LC–MS traces (DAD 210–600 nm) of transformed A. oryzae NSAR1 with triliA, triliB and triliC. (A), the extract of untransformed A. oryzae NSAR1; (BE), extracts of four transformants of A. oryzae with triliA, triliB and triliC. New compounds were observed at RT 6.7 and 10.8 min.
Jof 07 01034 g004
Figure 5. LC–MS traces (DAD 210–600 nm) of transformed A. oryzae with illiA, illiB, illiC and illiD. (A), the extract of the wild-type A. oryzae NSAR1; (BC), Extracts of two representative transformants of A. oryzae with illiA, illiB, illiC and illiD.
Figure 5. LC–MS traces (DAD 210–600 nm) of transformed A. oryzae with illiA, illiB, illiC and illiD. (A), the extract of the wild-type A. oryzae NSAR1; (BC), Extracts of two representative transformants of A. oryzae with illiA, illiB, illiC and illiD.
Jof 07 01034 g005
Scheme 1. Overall predicted pathway showing production of new compounds 11, 12 and 14 in A. oryzae NSAR1.
Scheme 1. Overall predicted pathway showing production of new compounds 11, 12 and 14 in A. oryzae NSAR1.
Jof 07 01034 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shenouda, M.L.; Ambilika, M.; Cox, R.J. Trichoderma reesei Contains a Biosynthetic Gene Cluster That Encodes the Antifungal Agent Ilicicolin H. J. Fungi 2021, 7, 1034. https://doi.org/10.3390/jof7121034

AMA Style

Shenouda ML, Ambilika M, Cox RJ. Trichoderma reesei Contains a Biosynthetic Gene Cluster That Encodes the Antifungal Agent Ilicicolin H. Journal of Fungi. 2021; 7(12):1034. https://doi.org/10.3390/jof7121034

Chicago/Turabian Style

Shenouda, Mary L., Maria Ambilika, and Russell J. Cox. 2021. "Trichoderma reesei Contains a Biosynthetic Gene Cluster That Encodes the Antifungal Agent Ilicicolin H" Journal of Fungi 7, no. 12: 1034. https://doi.org/10.3390/jof7121034

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