Next Article in Journal
Development of PCR-Multiplex Assays for Identification of the Herpotrichiellaceae Family and Agents Causing Chromoblastomycosis
Previous Article in Journal
New and Interesting Pine-Associated Hyphomycetes from China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of Secondary Metabolites by UHPLC-ESI-HRMS/MS in Antifungal Strain Trichoderma harzianum (LBAT-53)

by
Giselle Hernández
1,
Amaia Ponce de la Cal
2,
Yuset Louis
1,
Yamilé Baró Robaina
2,
Yamilet Coll
1,
Iraida Spengler
1,* and
Yaneris Mirabal-Gallardo
3,*
1
Center for Natural Products Research, Faculty of Chemistry, University of Havana, Havana 10400, Cuba
2
Plant Health Research Institute (INISAV), 110 Str. 514, Havana 11600, Cuba
3
Faculty of Engineering, Institute of Applied Chemistry, Universidad Autónoma de Chile, Talca 3460000, Chile
*
Authors to whom correspondence should be addressed.
J. Fungi 2024, 10(8), 547; https://doi.org/10.3390/jof10080547
Submission received: 3 June 2024 / Revised: 21 July 2024 / Accepted: 30 July 2024 / Published: 3 August 2024

Abstract

:
Trichoderma spp. are filamentous fungi generally observed in nature, which are widely marketed as biocontrol agents. The secondary metabolites produced have obtained special attention since they possess attractive chemical structures with a broad spectrum of biological activities. In Cuba, the species of Trichoderma have been commercially applied for the control of several phytopathogens to protect agricultural crops, but few studies have been carried out to detect and characterize the production of metabolites with biological activity. The strain Trichoderma harzianum LBAT-53 was subjected to an antifungal in vitro assay against Fusarium oxysporum f.sp. cubense by dual culture and volatile metabolite assays and fermented in PDB under constant agitation conditions. The ethyl acetate crude extract was obtained by liquid–liquid extraction. The fungal extract was investigated for the composition of secondary metabolites through chemical screening and ultrahigh performance liquid chromatography-tandem mass spectrometry (UHPLC-MS/MS) in negative ionization mode. As a result, LBAT-53 showed antagonistic activity in vitro (Class 2) against the pathogen evaluated in direct confrontation (76.9% of inhibition in 10 days) and by volatile metabolites (<40% in 7 days). Furthermore, seven low-molecular-weight phenolic compounds, including chrysophanol, phomarin, endocrocin, and trichophenol A, among others, were identified using UHPLC-ESI-MS/MS. This study is the first work on the characterization of secondary metabolites produced by the commercially applied strain LBAT-53, which is a promising source of bioactive compounds. These results provide a better understanding of the metabolism of this fungus, which is widely used in Cuba as biopesticides in agriculture pest control.

1. Introduction

As the global population continues to grow, food demand increases and production needs to increase [1]. Several factors have a negative impact on crop production, including pathogens that affect the quality and yield of crops. Fusarium Wilt of Banana, caused by the fungus Fusarium oxysporum f.sp. cubense Schltdl. (Foc), is considered one of the most destructive vascular wilt fungal diseases recorded in banana history. Throughout the world, the yearly losses of banana owing to this disease range from 60% to 90% [2,3]. It is a limiting factor in global production, and the fight against it represents one of the highest production costs [4]. In Cuba, the spread of Fusarium Wilt has caused a marked impact on production costs and especially on the clonal structure of the planted surface [2].
Humans use chemical control mechanisms that affect ecosystems to combat pests and diseases; however, increased resistance of pests to these chemically synthesized products has been observed [5]. In this scenario, the scientific community has focused its research on the use of alternatives that are more environmentally friendly and safer for producers and consumers. The use of biological control agents is a new alternative to fight fungal diseases [6] and has gained great interest in the last years in many pathosystems, including Foc/banana. This has been mainly due to the large input of pesticides, which cause economic, environmental, and safety concerns. Greenhouse and in vitro studies have reported microorganisms that are antagonistic to Foc [7].
Trichoderma spp. are distributed widely in soil and have been developed as a source of biocontrol agents for years [8,9]. These fungi are remarkable for their rapid growth and utilization of diverse substrates. In addition, they use different mechanisms, including antibiosis, mycoparasitism, pathogen competition, plant growth promotion, resistance to biotic and abiotic stresses, and activation of a pathogen defensive system [6,10,11,12]. Their use as a biocontrol agent against phytopathogens that cause losses in important agricultural crops such as Sclerotium rolfsii, Macrophomina phaseolina, Botrytis cinerea, Rhizoctonia spp. and Fusarium spp. is highlighted [13,14,15,16], and even with oomycetes such as Pythium ultimum [17].
Trichoderma spp. have the ability to produce different kinds of chemical substances, like volatile and nonvolatile compounds, polyketides, and siderophores, that can promote biocontrol activities [18]. A wide variety of compounds have been identified during the interaction between T. harzianum and Rhizoctonia solani, such as heptelidic acid, trichoviridine, harzianic acid, gliotoxin, glioviridin, viridin, and viridiol. The ability of the same strain of Trichoderma to secrete several antifungal compounds simultaneously limits the risk of the appearance of microorganisms resistant to these metabolites, a relevant aspect from a practical point of view [19].
Analytical methods including spectrophotometry, high performance liquid chromatography (HPLC), gas chromatography-tandem mass spectrometry (GC-MS/MS), high pressure liquid chromatography mass spectrometry (LC-MS), and, most recently, ultrahigh performance liquid chromatography-tandem mass spectrometry (UHPLC-MS/MS), to facilitate rapid and accurate identification of chemical compounds in complex mixtures, refs. [12,20,21,22,23] have been development for several years for the identification of secondary metabolites.
In Cuba, species of Trichoderma have also been used for several years to improve crop yields; specifically, the strains Trichoderma harzianum Rifai LBAT-34 and LBAT-53, as well as Trichoderma viride Persoon LBAT-TS3, are mass-produced in the Centers for the Reproduction of Entomophages and Entomopathogens (CREE) on solid support and also on liquid and agitated cultures [24]. Currently, these Cuban strains of Trichoderma are used within the national Integrated Pest Management (IPM) programs in crops of economic importance and have been widely used and commercially applied for the control of phytopathogenic fungi and nematodes as antagonists in soils to protect different types of crops, such as beans and bananas [25,26,27,28,29].
The production of Trichoderma for its application in the control of soil phytopathogens is an important line of research in the Plant Health Research Institute (INISAV). Specifically, the T. harzianum strain LBAT-53 is the active ingredient of the commercial product TRICOSAVE 53. For a further evaluation of its biological potential, the study of its antagonism against Foc is required as a pathogen of great importance for the country. Furthermore, the secondary metabolites produced by this strain, which are considered to play a significant and effective role in suppressing plant pathogens and promoting growth, has not yet been characterized.
The aim of this research was to identify the secondary metabolites present in the crude extract of ethyl acetate through chemical screening and UHPLC-MS/MS and evaluate the antifungal activity in vitro of Trichoderma harzianum LBAT-53 against Fusarium oxysporum f.sp. cubense, the causal agent of Fusarium Wilt, also known as Panama disease, a severe fungal disease in banana.

2. Materials and Methods

2.1. Fungal Strains

The Trichoderma harzianum strain LBAT-53 from INISAV Microbial Culture Collection (Havana, Cuba) was used in this study.
The phytopathogen Fusarium oxysporum f.sp. cubense (Foc) PalPR7 race 1 was collected from the municipality of Los Palacios, Pinar del Río, Cuba, and is deposited in the Microbial Culture Collection of INISAV.
All the fungal strains were cultured on potato dextrose agar (PDA) plates for 5 days in darkness at 30 ± 1 °C.

2.2. Fermentation Process

To obtain liquid cultures, the strain LBAT-53 was inoculated in 2 L Erlenmeyer flasks with 1 L of sterile potato dextrose broth (PDB). The inoculation was carried out from 5 mm diameter plugs obtained from actively growing margins of PDA cultures. The cultures were placed under constant agitation, 150 rpm for 7 days at 30 ± 1 °C. They were subsequently filtered under vacuum through filter paper (Whatman No. 4; Brentford, UK) to eliminate the mycelium and obtain the supernatant, which were centrifuged at 8000 rpm and filtered again on a 0.2 µm nitrocellulose membrane.

2.3. Antifungal In Vitro Assay

The antifungal activity of T. harzianum LBAT-53 against F. oxysporum f.sp. cubense was performed by a dual culture technique and volatile metabolite assay.
First, culture disks from a 7-day-old culture dish of both fungi (5 mm diameter from growing edge colonies) were placed on the opposite ends of Petri dishes with PDA (90 mm) and incubated at 30 °C in the dark for 10 days. As control, a disk of mycelium of the phytopathogen was placed on separate plates. The 5-grade Class Scale described by Bell et al. [30] was used to classify the Trichoderma strain.
The evaluation of volatile metabolites was carried out according to the methodology described by Dennis and Webster [31]. Culture disks of both fungi (5 mm) were inoculated in the center of Petri dishes with PDA (90 mm), and the lids were removed. The bottoms containing an antagonist and a pathogen were placed together and sealed using Parafilm® M (Darmstadt, Germany)and incubated at 30 ± 1 °C under dark conditions. The pathogen was in the upper plate in order to avoid any interference by antagonistic spores in the plate inoculated with Foc. As control, a bottom containing the pathogen was used, overlapping with another containing only PDA medium (Section S1).
All experiments were performed in triplicate. The radial growth of the pathogen was measured every 24 h with a graduated ruler, and the percentage of radial growth inhibition (PRGI) was calculated according to Hernández et al. [32].

2.4. Extraction of Secondary Metabolites

The fermentation broth of the Trichoderma strain LBAT-53 was filtered with a Whatman filter paper No. 2 (Brentford, UK) by gravity to separate the insoluble components present in the fermentation broth. The pH of the supernatants was adjusted to 2.0 with H3PO4 and then extracted with ethyl acetate (3 × 25 mL), concentrated in a rotary evaporator (Büchi R-300, BÜCHI Labortechnik GmbH, Essen, Germany) coupled to a vacuum pump (Büchi V-300, BÜCHI Labortechnik GmbH, Essen, Germany) to yield the crude metabolite, which was re-suspended in 2 mL of methanol, and 10 μL was injected for each analysis.

2.5. Chemical Screening

The phytochemical screening for the ethyl acetate crude extract of LBAT-53 was analyzed after extraction with ethanol. The different chemical groups were characterized with reference to the Rondina and Coussio [33] technique, with slight modifications (Section S2).

2.6. Cleanup of the Sample

An amount of 2 mL of MeOH/H2O 8:2 (v/v) solution of the extract (1 mg/mL) was subjected to solid-phase extraction using SPE cartridges Chromabond® C18 (loading, 200 mg/3 mL; particle size, 54 μm, Macherey-Nagel, Düren, Germany) eluted with MeOH/H2O 8:2 (v/v). After drying, 1 mg was dissolved in 1 mL of MeOH/H2O 8:2 (v/v) solution (solution A), and an aliquot (10 μL) was diluted with MeOH/H2O 8:2 (v/v) up to a final volume of 1 mL, and was filtered through a 0.22 μm membrane of a nylon filter. The solution (150 μL) was diluted with MeOH/H2O 8:2 (v/v) up to a final concentration of 1 ppm and submitted to UHPLC-ESI-HRMS/MS.

2.7. UHPLC-ESI-HRMS/MS Conditions and Data Analysis

The negative ion high-resolution ESI mass spectra were obtained from an Orbitrap Elite mass spectrometer (Thermo Fisher Scientific, Bremen, Germany) equipped with a heated ESI electrospray ion source (spray voltage negative ion mode, 3.5 kV; source heater temperature, 150 °C; capillary temperature, 325 °C; FTMS resolution, 30,000). Nitrogen was used as sheath and auxiliary gas. The MS system was coupled online to an ultrahigh performance liquid chromatography (UHPLC) system (Dionex UltiMate 3000, Thermo Fisher Scientific), equipped with a RP C18 column (1.8 µm; 100 × 1.0 mm; ACQUITY UPLC HSS T3 C18; Waters, column temperature, 45 °C), and a photodiode array detector (PDA, Thermo Fisher Scientific). For the UHPLC, a gradient system was used starting from H2O (A; Milli-Q, Merck Millipore, Burlington, MA, USA):CH3CN (B; Chromasolv LC-MS, Riedel-de Haen, Honeywell, Charlotte, NC, USA) 95:5 (each of them containing 0.1% (v/v) formic acid (eluent additive for LC-MS, Honeywell Fluka, Charlotte, NC, USA), isocratic for 1 min) raised to 50:50 within 3 min and in further 10 min to 10:90 to finally 5:95 within 1 min to then hold on 5:95 for further 3 min, the flow rate at 150 µL min−1. The wavelength range of the PDA measurements was 190–400 nm used for detection. The CID tandem mass spectra (buffer gas: helium; FTMS resolution, 15,000) were recorded in data-dependent acquisition mode (DDA) using normalized collision energies (NCE) of 35%. The instrument was externally calibrated by the Pierce® LTQ Velos ESI negative ion calibration solution (product number 88324, Thermo Fisher Scientific, Rockford, IL, 61105 USA). The data were evaluated by the Xcalibur software 2.2 SP1 (Thermo Fisher Scientific).

2.8. Statistical Analysis

The experimental design was completely random, and the data obtained were processed using a simple classification Analysis of Variance (ANOVA) with its corresponding significance test (p ˂ 0.05) according to Tukey’s test using the statistical software InfoStat version 2008 for Windows.

3. Results

3.1. Antifungal In Vitro Assay

The results obtained in dual culture showed that the T. harzianum LBAT-53 strain has a higher growth rate than F. oxysporum f.sp. cubense PalPR7. After 10 days, it was placed in Class 2 on the Bell et al. scale [30], showing a high antagonistic effect, with the growth of the pathogen stopping upon contact with the antagonist mycelium (Figure 1).
In the interaction with the pathogenic strain PalPR7 at 48 h, prior to physical contact, a growth inhibitory effect greater than 10% was observed due to the effect of the LBAT-53 strain, with significant differences with the control treatment (Table 1). At 7 days, the inhibitory effect was greater than 66%, with significant differences compared with the control, and after 10 days, the percentage of inhibition of pathogen growth was 76.90%.
Regarding the release of volatile metabolites, the strain LBAT-53 inhibited the growth of F. oxysporum f.sp. cubense PalPR7. At 72 h, the percentage of inhibition in the mycelium of F. oxysporum f.sp. cubense presented significant differences with the control. After 7 days, growth inhibition was greater than 40% due to the release of volatile metabolites (Table 1), inducing changes in the coloration and shape of the pathogen colony (Figure 2).

3.2. Chemical Screening

The chemical screening of a crude extract of T. harzianum LBAT-53 revealed the presence of amine compounds, saponins, and glycosylated triterpenes–steroids; moderate amounts of free flavonoids and quinones; and higher amounts of phenols and reducing sugars. On the other hand, no free triterpenes–steroids, alkaloids, cardenolides, glycosylated flavonoids, and proanthocyanidins/catechins were detected (Table 2). A detailed description of the procedure to obtain each fraction of increasing polarity and the assays corresponding to the different groups of metabolites detected is given in the Supplementary Information (Section S2).

3.3. UHPLC-ESI-MS/MS

The extract of T. harzianum was also analyzed using UHPLC-ESI-MS/MS in negative ion mode. Figure 3A corresponds to the chromatogram of the total ions obtained in a UHPLC equipped with a Watters RP C18 column (1.8 µm; 100 × 1.0 mm), and Figure 3B to the expanded chromatogram of the metabolites eluted with a Rt of 4–9.5 min, where, considering the characteristics of the reversed phase, in this region, the metabolites of medium polarity should appear. The spectrum of total ions in the extended m/z 140–900 region afforded characteristic deprotonated ions corresponding to the different metabolites present in the crude extract (Figure 3C).

3.3.1. Anthraquinone Identification

The presences of the different anthraquinones within the extract are visualized in the extracted ion chromatograms (EICs) at different retention times (Figure 4A–C). The assignment of the structures is based on their elemental composition determined by high-resolution mass spectrometry (Table 3). For the data evaluation, the target m/z values were extracted from the total ion chromatogram to obtain the corresponding extracted ion chromatograms for each compound. Due to the resolving power of the Orbitrap detector, a differentiation of isobaric ions was possible, as shown in the EIC of Figure 4A, where the anthraquinone peak at m/z 253.0515 is clearly separated from other accompanying ion at the same nominal mass. In the second-order spectrum of the pseudomolecular ion with Rt = 8.32 min, common losses of H2O and CO were observed. The compound was identified as chrysophanol according to the fragmentation pattern, and the second peak, eluting at 8.50 min, was tentatively identified as phomarin (1,6-dihydroxy-3-methyl-9,10-anthracenedione), a positional isomer of chrysophanol (Figure 5).
Two additional anthraquinones—1,8-dihydroxy-3-(hydroxymethyl)anthracene-9,10-dione (m/z = 269.0457 with Rt = 6.49 min and endocrocin (m/z = 313.0353 with Rt = 6.95 min)—were identified (Figure 6) based on combined data available on the mass spectra with molecular weight and characteristic fragment ions that were previously identified by Laub et al., 2020 [34].

3.3.2. Other Phenolic Compounds

Caffeic acid (m/z 179.0564), trichophenol A (m/z 299.2593), and isorhamnetin (m/z 315.0545) were identified on the ethyl acetate crude extract of T. harzianum (LBAT-53). The structures of each compound are shown in Figure 7.

4. Discussion

Trichoderma species are the most studied and used fungi for the control of plant disease [35]. Diverse studies revealed the different mechanisms that Trichoderma has as a biocontrol agent [8,9,27,36,37,38,39]. In 2016, Khaledi and Taheri investigated the biocontrol mechanisms of 11 Trichoderma isolates against M. phaseolina in dual culture tests [40]. The results showed that all isolates inhibited the mycelial growth of the pathogen from 20.2 to 58.7%. Likewise, in Cuba, in a study conducted by Martínez-Coca et al. (2018) [41], most of the Trichoderma strains evaluated inhibited the growth of Fusarium spp. isolates by over 40%. When specifically evaluating the LBAT-53 strain, Sierra et al. (2007) obtained the highest inhibition percentages for the pathogens Fusarium subglutinans and Rhizoctonia solani, with values of 48% and 58%, respectively [42]. The PRGI reported in this study are higher than the values reported by these authors.
In this study, T. harzianum (LBAT-53) displayed a faster growth compared with F. oxysporum and inhibited the mycelial growth of the pathogen in more than 66% at 7 days. In dual culture assays with Foc race 1 strains, several authors have obtained inhibition values of 38.82% and 53.46% with the Trichoderma sp. TB1 isolate [43]. Hernández-Melchor et al. [44] evaluated 15 native Trichoderma isolates against 5 Fusarium oxysporum f.sp. cubense RACE 1 isolates, and the best 8 strains showed inhibition percentages in the range of 3% to 54%; the values were lower than those obtained in this study. These results could be explained because of the highest competition for nutrients and space of LBAT-53. Furthermore, the excretion of metabolites with significant fungistatic action over the pathogen was also corroborated with inhibition values greater than 40%. Recently, Lakhdari et al. [45] corroborated the presence of volatile metabolites with antifungal activity present in the ethyl acetate and n-butanol extracts of T. harzianum.
Secondary metabolites in nature perform specialized functions at very low concentrations. They facilitate symbiosis with microorganisms, insects, plants, and higher animals, and are considered mediators of chemical communication between soil inhabitants in different ecological niches [8,37]. Since they are produced in very low concentrations, analytical methods for detecting analytes present at trace levels in a complex matrix and screening large numbers of samples have been developed. Among others, ultrahigh performance liquid chromatography-tandem mass spectrometry (UHPLC-MS/MS) is a better choice due to higher sensitivity, higher resolution, and a shorter run-time. However, conventional chemical assays remain the choice for preliminary chemical screening because they are inexpensive and simple and require fewer resources [46]. These techniques combine metabolite solubility with solvent polarity and pH variation during the extraction procedure. The crude extract is separated into fractions that are subjected to qualitative assays of color development and/or solid precipitation, depending on the type of metabolite.
In this study, the chemical screening allowed for detecting the presence of phenolic compounds. These kinds of secondary metabolites were identified in different species of Trichoderma (T. polypore, T. polyalthiae, and T. gamsii) [47]. In the medium-high polarity fraction corresponding to the CHCl3/EtOH extract (Fraction D), the Shinoda assay was positive, suggesting a moderate presence of flavonoids such as flavonones, flavones, flavanonols, chalcones, and flavonols. Ni et al. [48] identified flavonoids, for example, dihydromyricetin, isorhamnetin, and 4-hydroxy-5,7-dimethoxyflavanone, present in the fermentation broth of T. asperellum TJ01. Ninhidrine addition to the respective fraction A and an aqueous fraction F produced a slightly violet color, indicating the presence of amine compounds. The Börntrager assay for quinones detection was positive, and the color of a solution indicated the presence of anthraquinones. This family of compounds was also detected by Dennis et al. [31]. They identified chrysophanol, phomarin, emodin, and other anthraquinones from the T. harzianum strain Th-R16 [49]. In the scientific literature consulted, it was found that several investigations have shown that Trichoderma species are capable of producing these metabolites [50]. Anthraquinones are very important for their antifungal, antimicrobial, and antioxidant action, among others [50,51,52]. They have been considered among the most abundant fungal natural products [53]. Based on results of chemical screening, the strain LBAT-53 was characterized for the first time with a highly sensitive method like UPLC-ESI-MS/MS in negative ionization mode. Four anthraquinones were detected and corroborated their structures according to the fragmentation patterns observed. In the case of chrysophanol, the antifungal activity that this molecule presents against Blumeria graminis f.sp. Hordei, Podosphaera xanthii, Candida albicans, Cryptococcus neoformans, Trichophyton mentagrophytes, and Aspergillus fumigatus is well documented [54]. When tested against Botrytis cinerea and Rhizoctonia solani, it also showed inhibition of the growth of these pathogens [55]. Liu et al. [55] demonstrated that chrysophanol is involved in the stimulation of plant growth, photosynthesis, and the induction of host defense responses during T. harzianum colonization on cabbage. The anthraquinone phomarin, which is a positional isomer of chrysophanol, detected in the crude acetate extract of the Cuban strain LBAT-53, was also identified in Trichoderma species [56].
As an example, the fragmentation patterns of chrysophanol and endocrocin were examined by Ms2 analysis (Figure 8 and Figure 9). Characteristic product ions of chrysophanol corresponded to loss of a neutral molecular fragment, such as CO2 resulting in a product ion at m/z = 209, H2O [M-H-18] at m/z = 235, and CO resulting in the base peak at m/z = 225 was detected. In Figure 8 is shown the fragmentation pathway proposed for this molecule.
Characteristic losses in the second-order spectra of endocrocin were detected under negative ion electrospray (Figure 9). The loss of a neutral molecular fragment was also a characteristic of methylated anthraquinones [57]. The main product ion corresponds to [M-H-CH3-CO] (base peak) obtained from ion at m/z = 298 [M-H-CH3], which corresponds to the loss of the methyl group; this ion is further decomposed by the loss of CO.
The isocoumarin trichophenol A, detected in LBAT-53, was isolated and reported by Liu et al. in 2020 from the marine-alga-endophytic fungus Trichoderma citrinoviride A-WH-20-3 [58]. The flavonol isorhamnetin also was identified in this study. In a research conducted by Unver [59] in 2024, the antifungal activity was demonstrated with a significant inhibitory effect (MIC = 1.875 mg/mL) against Candida species. This bioactive metabolite was reported by Ni et al. [48] recently by liquid chromatography coupled to mass spectrometry with a triple quadrupole analyzer. The authors studied the metabolic changes in Trichoderma asperellum TJ01 at different fermentation times. Additionally, they detected at 72 h that the culture had a higher proportion of upregulated flavonoids.
In a study conducted by Zakaba and Pavela in 2013 [60], the authors tested 21 phenolic compounds against filamentous fungi and demonstrated that caffeic acid had an inhibitory effect on the mycelial growth of F. oxysporum at a basic concentration of 1000 μg/mL with 33.33 ± 0.09% of inhibition.
In general, the secondary metabolites of Trichoderma confer the biocontrol activity of the strain either by directly inhibiting pathogens (direct antagonism) of the host or by inducing host plant resistance [49]. They are essential for fungal development and actively determine interactions with other organisms. Compounds possessing a phenolic ring system have been found to exhibit a number of pharmacological properties; for example, some of the phenolic compounds, such as phenolic acids, flavonoids, catechins, anthocyanins, tannins, anthraquinones, and naphthoquinones, which are lipophilic in nature, can inhibit the activity of the ABC transporters [61]. Polyphenols can also bind directly to proteins, interfering with the tertiary structure of proteins and thus effectively inhibiting the function of ABC transporters, which could affect the pathogenicity mechanisms of the fungus that causes the disease in the plant. The structure of phenolic compounds is such that they can diffuse through the microbial membranes and can penetrate into the cell, where they can interfere in the metabolic pathways [61]. According to these results, the anthraquinones detected in this research could be involved directly in the biocontrol function of T. harzianum against the pathogen tested, and a synergistic effect with other phenolic compounds identified also could be observed.

5. Conclusions

This investigation represents a first approach to the chemical study of the Trichoderma harzianum (LBAT-53) strain from the Microbial Culture Collection of INISAV. Chemical screening and a highly sensitive and high-resolution UHPLC-ESI-MS/MS technique were used as qualitative methods to identify secondary metabolites produced by this important strain, which is used as a biocontrol agent. In vitro antifungal activity showed that LBAT-53 has an antagonistic effect against Fusarium oxysporum f.sp. cubense PalPR7. The results support the potential use of Trichoderma harzianum as an alternative for the control of this pathogen, which affects plantains and banana crops worldwide.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/jof10080547/s1, Section S1: Schematic overview of the volatile metabolite assay; Section S2: Chemical screening procedure.

Author Contributions

Conceptualization, G.H., A.P.d.l.C., Y.L., Y.C., Y.B.R. and I.S.; methodology, G.H., A.P.d.l.C., Y.L., Y.B.R. and I.S.; software, G.H., I.S. and A.P.d.l.C.; validation, Y.B.R. and I.S.; formal analysis, I.S.; investigation, G.H., A.P.d.l.C., Y.L., Y.C., I.S. and Y.B.R.; resources, A.P.d.l.C. and Y.M.-G.; writing—original draft preparation, G.H. and A.P.d.l.C.; writing—review and editing G.H., A.P.d.l.C., Y.L., Y.C., Y.B.R. and I.S.; supervision, Y.B.R. and I.S.; project administration, A.P.d.l.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the research project “Fundamentos para la obtención de un bioproducto a partir de Trichoderma para el control de fitopatógenos de suelo con capacidad bioestimulante”, financed by the Cuban Ministry of Agriculture, under research grant number PSSAV PS223MY003112, Cuban Plant Health Sectorial Program, grant number PS131LH0030112.

Data Availability Statement

The original contributions presented in the study are included in the article and Supplementary Materials, further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors would like to thank to Enrique Ponce Grijuela from INISAV for excellent technical support and Fondo de Innovación para la Competitividad FIC del Gobierno Regional del Maule, IDEA DE I+D 2023, Subdirección de Investigación Aplicada, FONDEF, ANID.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. FAO Global Agriculture towards 2050. Available online: https://www.fao.org/fileadmin/templates/wsfs/docs/Issues_papers/HLEF2050_Global_Agriculture.pdf (accessed on 14 November 2023).
  2. González, E.; Rodríguez, S. ‘INIVIT PB-12’. Nuevo cultivar de plátano (Musa spp.) tipo burro en entidades productivas del país. Agric. Trop. 2018, 4, 120–126. [Google Scholar]
  3. Fusarium Wilt. 2019. Available online: https://fusariumwilt.org/index.php/es/el-problema/ (accessed on 24 January 2024).
  4. Niwas, R.; Chand, G.; Nath Gupta, R. Fusarium Wilt: A Destructive Disease of Banana and Their Sustainable Management. In Fusarium: An Overview of the Genus; IntechOpen: London, UK, 2022. [Google Scholar] [CrossRef]
  5. Rupp, S.; Weber, R.W.; Rieger, D.; Detzel, P.; Hahn, M. Spread of Botrytis cinerea Strains with Multiple Fungicide Resistance in German Horticulture. Front. Microbiol. 2017, 7, 2075. [Google Scholar] [CrossRef] [PubMed]
  6. Rodrigo, S.; García, C.; Santamaria, O. Metabolites Produced by Fungi against Fungal Phytopathogens: Review, Implementation and Perspectives. Plants 2022, 11, 81. [Google Scholar] [CrossRef] [PubMed]
  7. Bubici, G.; Kaushal, M.; Prigigallo, M.I.; Gomez-Lama Cabanás, C.; Mercado-Blanco, J. Biological control agents against Fusarium Wilt of banana. Front. Microbiol. 2019, 10, 616. [Google Scholar] [CrossRef] [PubMed]
  8. Zin, N.A.; Badaluddin, N.A. Biological functions of Trichoderma spp. for agriculture applications. Ann. Agric. Sci. 2020, 65, 168–178. [Google Scholar] [CrossRef]
  9. Bhandari, S.; Pandey, K.R.; Joshi, Y.R.; Lamichhane, S.K. An overview of multifaceted role of Trichoderma spp. for sustainable agriculture. Arch. Agric. Environ. Sci. 2021, 6, 72–79. [Google Scholar] [CrossRef]
  10. Sood, M.; Kapoor, D.; Kumar, V.; Sheteiwy, M.D.; Ramakrishnan, M.; Landi, M.; Araniti, F.; Sharma, A. Trichoderma: The “Secrets” of a Multitalented Biocontrol Agent. Plants 2020, 9, 762. [Google Scholar] [CrossRef] [PubMed]
  11. Küçük, C.; Kivanc, M. Isolation of Trichoderma spp. and Determination of Their Antifungal, Biochemical and Physiological Features. Turk. J. Biol. 2003, 27, 247–253. [Google Scholar]
  12. Infante, D.; González, N.; Reyes, Y. Mecanismos deacción de Trichoderma frente a hongos fitopatógenos. Rev. Prot. Veg. 2009, 24, 14–21. [Google Scholar]
  13. Sánchez-García, B.M.; Espinosa Huerta, E.; Villordo-Pineda, E.; Rodríguez-Guerra, R.; Mora Avilés, M.A. Identificación molecular y evaluación antagónica in vitro de cepas nativas de Trichoderma spp. sobre hongos fitopatógenos de raíz en frijol (Phaseolus vulgaris L.) cv. Montcalm. Agrociencia 2017, 51, 63–79. [Google Scholar]
  14. Sridharan, A.P.; Sugitha, T.; Karthikeyan, G.; Sivakumar, U. Comprehensive profiling of the VOCs of Trichoderma longibrachiatum EF5 while interacting with Sclerotium rolfsii and Macrophomina phaseolina. Microbiol. Res. 2020, 236, 126436–126449. [Google Scholar] [CrossRef]
  15. Gutiérrez-Rivas, R.E. Evaluación In Vitro e In Vivo de la Capacidad Antagónica de Trichoderma harzianum para el Control de Rhizoctonia spp. Bachelor’s Thesis, Universidad Nacional de Tumbes, Tumbes, Peru, 2022. [Google Scholar]
  16. Ttacca, B.L.; Yataco, R.J.Y.; Farfán, A.A.; Calderón, L.M.; Cantoral, J.C.A. Antibiosis y micoparasitismo de hongos endófitos sobre el agente causal del moho gris del arándano (Botrytis cinerea). Bioagro 2022, 34, 209–222. [Google Scholar] [CrossRef]
  17. Karlsson, M.; Atanasova, L.; Jensen, D.F.; Zeilinger, S. Necrotrophic mycoparasites and their genomes. Microbiol. Spec. 2017, 5, 1–21. [Google Scholar] [CrossRef]
  18. Joo, J.H.; Hussein, K.A. Biological Control and Plant Growth Promotion Properties of Volatile Organic Compound-Producing Antagonistic Trichoderma spp. Front. Plant Sci. 2022, 13, 897668. [Google Scholar] [CrossRef] [PubMed]
  19. Osorio-Hernández, E.O.; Castillo, F.D.H.; Herrera, R.R.; Fuentes, S.E.V.; Drouaillet, B.E.; Santillán, J.A.L. Actividad antagónica de Trichoderma spp. sobre Rhizoctonia solani in vitro. Investig. Cienc. Univ. Autón. Aguascalientes 2016, 65, 5–11. [Google Scholar] [CrossRef]
  20. Saldan, N.C.; Almeida, R.T.; Avíncola, A.; Porto, C.; Galuch, M.B.; Magon, T.F.; Oliveira, C.C. Development of an analytical method for identification of Aspergillus flavus based on chemical markers using HPLC-MS. Food Chem. 2018, 241, 113–121. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, A.J.; Choi, J.N.; Kim, J.; Park, S.B.; Yeo, S.H.; Choi, J.H.; Lee, C.H. GC-MS based metabolite profiling of rice koji fermentation by various fungi. Biosci. Biotechnol. Biochem. 2010, 74, 2267–2272. [Google Scholar] [CrossRef] [PubMed]
  22. Salvatore, M.M.; Nicoletti, R.; Salvatore, F.; Naviglio, D.; Andolfi, A. GC–MS approaches for the screening of metabolites produced by marine-derived Aspergillus. Mar. Chem. 2018, 206, 19–33. [Google Scholar] [CrossRef]
  23. García, T.H.; da Rocha, C.Q.; Dias, M.J.; Pino, L.L.; del Barrio, G.; Roque, A.; Pérez, C.E.; Dos Santos, L.C.; Spengler, I.; Vilegas, W. Comparison of the qualitative chemical composition of extracts from Ageratina havanensis collected in two different phenological stages by FIA-ESI-IT-MSn and UPLC/ESI-MSn:Antiviral activity. Nat. Prod. Commun. 2017, 12, 31–34. [Google Scholar] [CrossRef]
  24. Fernández, O.; Elósegui, O. Alternativas de producción de Trichoderma en Cuba. Fitosanidad 2006, 10, 147. [Google Scholar]
  25. Stefanova, M. Introducción y eficacia técnica del biocontrol de fitopatógenos con Trichoderma spp. en Cuba. Fitosanidad 2007, 11, 75–79. [Google Scholar]
  26. Calero, A.; Quintero, E.; Pérez, Y. Utilización de diferentes bioproductos en la producción de frijol común (Phaseolus vulgaris L.). Agrotec. Cuba 2017, 41, 17–24. [Google Scholar]
  27. Pérez Torres, E.J.; Bernal Cabrera, A.; Milanés Virreles, P.; Leiva Mora, M.; Sierra Reyes, Y.; Cupull Santana, R. Actividad antagónica de Trichoderma harzianum Rifai sobre el agente causal del tizón del arroz (Pyricularia grisea Sacc.). Centro Agrícola 2017, 44, 13–19. [Google Scholar]
  28. Pérez, L.; Porras, A. Impacto potencial del cambio climático sobre las plagas de bananos y plátanos en Cuba. Fitosanidad 2015, 19, 201–221. [Google Scholar]
  29. Samaniego, L.M.; Harouna, M.; Odalys Corbea, O.; Rondón, A.J.; Placeres, I. Aislamiento, identificación y evaluación de cepas autóctonas de Trichoderma spp. antagonistas de patógenos del suelo. Rev. Prot. Veg. 2018, 33, 1–11. [Google Scholar]
  30. Bell, D.K.; Wells, H.D.; Markham, C.R. In vitro antagonism of Trichoderma species against six fungal plant pathogens. Phytopathology 1982, 72, 379–382. [Google Scholar] [CrossRef]
  31. Dennis, C.; Webster, J. Antagonistic properties of species-groups of Trichoderma. II. Production of volatile antibiotics. Trans. Br. Mycol. Soc. 1971, 57, 41–48. [Google Scholar] [CrossRef]
  32. Hernández, G.; Ramos, B.; Sultani, H.N.; Ortiz, Y.; Spengler, I.; Castañeda, R.F.; Rivera, D.G.; Arnold, N.; Westermann, B.; Mirabal, Y. Cultural Characterization and Antagonistic Activity of Cladobotryum virescens against Some Phytopathogenic Fungi and Oomycetes. Agronomy 2023, 13, 389. [Google Scholar] [CrossRef]
  33. Rondina, R.V.D.; Coussio, J.D. Estudio fitoquímico de plantas medicinales argentinas (1). Rev. Investig. Agropec. (Ser. 2. Biol. Prod. Veg.) 1969, 6, 352. [Google Scholar]
  34. Laub, A.; Sendatzki, A.K.; Palfner, G.; Wessjohann, L.; Schmidt, J.; Arnold, N. HPTLC-DESI-HRMS based Profiling of Anthraquinones in Complex Mixtures –A Proof-Of-Concept Study using Crude Extracts of Chilean Mushrooms. Foods 2020, 9, 156. [Google Scholar] [CrossRef]
  35. Khan, M.R.; Mohiddin, F.A. Trichoderma: Its multifarious utility in crop improvement. In Crop Improvement trough Microbial Biotechnology; Elsevier: Amsterdam, The Netherlands, 2018; pp. 263–291. [Google Scholar] [CrossRef]
  36. Pérez, E.; Bernal, A.; Milanés, P.; Sierra, Y.; Leiva, M.; Marín, S.; Monteagudo, O. Eficiencia de Trichoderma harzianum (cepa a-34) y sus filtrados en el control de tres enfermedades fúngicas foliares en arroz. Bioagro 2018, 30, 17–26. [Google Scholar]
  37. Ali, M.; Ali, H.; Haq, I.; Shah, I.; Naeem, M.; Samad, A.; Sher, Z.; Asmat, T.M. Evaluation of Trichoderma as a biological control against different pathogenic bacteria and fungi. Pure Appl. Biol. 2020, 9, 1223–1229. [Google Scholar] [CrossRef]
  38. Babychan, M.; Simon, S. Efficacy of Trichoderma spp. against Fusarium oxysporum f. sp. lycopersici. (FOL) infecting pre-and post-seedling of tomato. J. Pharmacogn. Phytochem. 2017, 6, 616–619. [Google Scholar]
  39. Ferreira, F.V.; Musumeci, M.A. Trichoderma as biological control agent: Scope and prospects to improve efficacy. World J. Microbiol. Biotechnol. 2021, 37, 90. [Google Scholar] [CrossRef] [PubMed]
  40. Khaledi, N.; Taheri, P. Biocontrol mechanisms of Trichoderma harzianum against soybean charcoal rot caused by Macrophomina phaseolina. J. Plant Prot. Res. 2016, 56, 21–31. [Google Scholar] [CrossRef]
  41. Martínez-Coca, B.; Infante, D.; Caraballo, W.; Duarte-Leal, Y.; Echevarría-Hernández, A. Antagonism of strains of Trichoderma asperellum Samuels, Lieckfeldt & Nirenberg against isolates of Fusarium spp. from chickpea. Rev. Prot. Veg. 2018, 3, 1–13. [Google Scholar]
  42. Sierra, A.; Hernández, A.A.; Díaz, J.A.; Carr, A. Control in vitro con Trichoderma spp. de hongos fitopatógenos de vitroplantas de piña (Ananas comosus L.) Merr. Control in vitro with Trichoderma spp. of phytopathogenus of pineapple vitroplants (Pineapples comosus (L). Merr. Centro Agrícola 2007, 34, 19–23. [Google Scholar]
  43. Caballero, A.J.; Pocasangre, L.E.; Casanoves, F.; Avelin, J.; Tapia, A.C.; Ortiz, J.L. Uso de aislamientos endofíticos de Trichoderma spp., para el biocontrol del Fusarium oxysporum f. sp. cubense (Mal de Panamá) raza 1 en vitroplantas de banano del cultivar Gros Michel (AAA) en condiciones de invernadero. Univ. Rev. Cient. 2013, 4, 71–82. [Google Scholar]
  44. Hernández-Melchor, D.J.; Ferrera-Cerrato, R.; López-Pérez, P.A.; Ferrera-Rodríguez, M.R.; García-Ávila, C.d.l.J.; Alarcón, A. Qualitative and quantitative enzimatic profile of native Trichoderma strains and biocontrol potential against Fusarium oxysporum f.sp. cubense RACE 1. J. Microbiol. Biotech. Food Sci. 2022, 11, e3264. [Google Scholar] [CrossRef]
  45. Lakhdari, W.; Benyahia, I.; Bouhenna, M.M.; Bendif, H.; Khelafi, H.; Bachir, H.; Ladjal, A.; Hammi, H.; Mouhoubi, D.; Khelil, H.; et al. Exploration and Evaluation of Secondary Metabolites from Trichoderma harzianum: GC-MS Analysis, Phytochemical Profiling, Antifungal and Antioxidant Activity Assessment. Molecules 2023, 28, 5025. [Google Scholar] [CrossRef]
  46. Camba, W.E.; Petroche, D.J.; Cortez, L.A.; Mariscal, W.E. Tamizaje fitoquímico, fenoles totales y actividad antioxidante de Citrus aurantium. RECIAMUC 2022, 6, 470–479. [Google Scholar] [CrossRef]
  47. Zhao, D.L.; Zhang, X.F.; Huang, R.H.; Wang, D.; Wang, X.Q.; Li, Y.Q.; Zheng, C.J.; Zhang, P.; Zhang, C.S. Antifungal nafuredin and epithiodiketopiperazine derivatives from the mangrove-derived fungus Trichoderma harzianum D13. Front. Microbiol. 2020, 11, 1495. [Google Scholar] [CrossRef]
  48. Ni, M.; Wu, Q.; Wang, G.S.; Liu, Q.Q.; Yu, M.X.; Tang, J. Analysis of metabolic changes in Trichoderma asperellum TJ01 at different fermentation time-points by LC-QQQ-MS. J. Environ. Sci. Health Part B 2019, 54, 20–26. [Google Scholar] [CrossRef]
  49. Liu, S.Y.; Lo, C.T.; Shibu, M.A.; Leu, Y.L.; Jen, B.Y.; Peng, K.C. Study on the Anthraquinones Separated from the Cultivation of Trichoderma harzianum Strain Th-R16 and Their Biological Activity. J. Agric. Food Chem. 2009, 57, 7288–7292. [Google Scholar] [CrossRef] [PubMed]
  50. De Mattos-Shipley, K.M.; Simpson, T.J. The ‘emodin family’ of fungal natural products–amalgamating a century of research with recent genomics-based advances. Nat. Prod. Rep. 2023, 40, 174–201. [Google Scholar] [CrossRef]
  51. Malik, M.S.; Alsantali, R.I.; Jassas, R.S.; Alsimaree, A.A.; Syed, R.; Alsharif, M.A.; Kalpana, K.; Morad, M.; Althagafi, I.I.; Ahmed, S.A. Journey of anthraquinones as anticancer agents a systematic review of recent literature. RSC Adv. 2021, 11, 35806–35827. [Google Scholar] [CrossRef]
  52. Sebak, M.; Molham, F.; Greco, C.; Tammam, M.A.; Sobehd, M.; El-Demerdash, A. Chemical diversity, medicinal potentialities, biosynthesis, and pharmacokinetics of anthraquinones and their congeners derived from marine fungi: A comprehensive update. RSC Adv. 2022, 12, 24887. [Google Scholar] [CrossRef]
  53. Fouillaud, M.; Venkatachalam, M.; Girard-Valenciennes, E.; Caro, Y.; Dufossé, L. Anthraquinones and Derivatives from Marine-Derived Fungi: Structural Diversity and Selected Biological Activities. Mar. Drugs 2016, 14, 1–64. [Google Scholar] [CrossRef] [PubMed]
  54. Prateeksha; Aslam, M.Y.; Singh, B.N.; Sudheer, S.; Kharwar, R.N.; Siddiqui, S.; Abdel-Azeem, A.M.; Fernandes Fraceto, L.; Dashora, K.; Gupta, V.K. Chrysophanol: A Natural Anthraquinone with Multifaceted Biotherapeutic Potential. Biomolecules 2019, 9, 68. [Google Scholar] [CrossRef]
  55. Liu, S.Y.; Liao, C.K.; Lo, C.T.; Yang, H.H.; Lin, K.C.; Peng, K.C. Chrysophanol is involved in the biofertilization and biocontrol activities of Trichoderma. Physiol. Mol. Plant Pathol. 2016, 96, 1–7. [Google Scholar] [CrossRef]
  56. Guo, R.; Li, G.; Zhang, Z.; Peng, X. Structures and Biological Activities of Secondary Metabolites from Trichoderma harzianum. Mar. Drugs 2022, 20, 701. [Google Scholar] [CrossRef]
  57. Schmidt, J. Negative ion electrospray high-resolution tandem mass spectrometry of polyphenols. J. Mass. Spectrom. 2015, 51, 33–43. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, J.; Leng, L.; Liu, Y.; Gao, H.; Yang, W.; Chen, S.; Liu, A. Identification and quantification of target metabolites combined with transcriptome of two rheum species focused on anthraquinone and flavonoids biosynthesis. Sci. Rep. 2020, 10, 20241. [Google Scholar] [CrossRef] [PubMed]
  59. Unver, T. Isorhamnetin as a promising natural bioactive flavonoid: In vitro assessment of its antifungal property. Int. J. Agric. Environ. Food Sci. 2024, 8, 54–61. [Google Scholar] [CrossRef]
  60. Zabka, M.; Pavela, R. Antifungal efficacy of some natural phenolic compounds against significant pathogenic and toxinogenic filamentous fungi. Chemosphere 2013, 93, 1051–1056. [Google Scholar] [CrossRef]
  61. Ansari, M.A.; Anurag, A.; Fatima, Z.; Hameed, S. Natural phenolic compounds: A potential antifungal agent. Microb. Pathog. Strateg. Combat. Sci. Thenol. Educ. 2013, 1, 1189–1195. [Google Scholar]
Figure 1. Dual culture of the strain T. harzianum LBAT-53 and F. oxysporum f.sp. cubense PALPR7 after 10 days. (A) Dual culture. (B) Control of F. oxysporum f.sp. cubense PALPR7.
Figure 1. Dual culture of the strain T. harzianum LBAT-53 and F. oxysporum f.sp. cubense PALPR7 after 10 days. (A) Dual culture. (B) Control of F. oxysporum f.sp. cubense PALPR7.
Jof 10 00547 g001
Figure 2. Effect of volatile metabolites of Trichoderma harzianum LBAT-53 on F. oxysporum f.sp. cubense PalPR7. (A) Confrontation. (B) Control of F. oxysporum f.sp. cubense PalPR7.
Figure 2. Effect of volatile metabolites of Trichoderma harzianum LBAT-53 on F. oxysporum f.sp. cubense PalPR7. (A) Confrontation. (B) Control of F. oxysporum f.sp. cubense PalPR7.
Jof 10 00547 g002
Figure 3. (A) Total ion chromatogram (TIC) obtained for ethyl acetate from T. harzianum LBAT-53 by reversed-phase UHPLC-ESI-MS/MS, (B) expanded ion chromatogram of metabolites with a Rt = 4–9.5 min, and (C) total ion spectrum of the ethyl acetate extract obtained from the culture broth of the strain LBAT-53 in negative ionization mode.
Figure 3. (A) Total ion chromatogram (TIC) obtained for ethyl acetate from T. harzianum LBAT-53 by reversed-phase UHPLC-ESI-MS/MS, (B) expanded ion chromatogram of metabolites with a Rt = 4–9.5 min, and (C) total ion spectrum of the ethyl acetate extract obtained from the culture broth of the strain LBAT-53 in negative ionization mode.
Jof 10 00547 g003
Figure 4. Extracted ion chromatograms (AC) obtained from the ethyl acetate crude extract of T. harzianum (LBAT-53) based on the theoretical masses of anthraquinones studied by Laub et al. [34].
Figure 4. Extracted ion chromatograms (AC) obtained from the ethyl acetate crude extract of T. harzianum (LBAT-53) based on the theoretical masses of anthraquinones studied by Laub et al. [34].
Jof 10 00547 g004
Figure 5. Structures of anthraquinones chrysophanol and phomarin identified in ethyl acetate crude extract of T. harzianum LBAT-53 by UHPLC-ESI-MS/MS in negative ion mode.
Figure 5. Structures of anthraquinones chrysophanol and phomarin identified in ethyl acetate crude extract of T. harzianum LBAT-53 by UHPLC-ESI-MS/MS in negative ion mode.
Jof 10 00547 g005
Figure 6. Structures of anthraquinones detected.
Figure 6. Structures of anthraquinones detected.
Jof 10 00547 g006
Figure 7. Caffeic acid (a), trichophenol A, (b) and isorhamnetin (c) detected in the crude extract of ethyl acetate from T. harzianum (LBAT-53).
Figure 7. Caffeic acid (a), trichophenol A, (b) and isorhamnetin (c) detected in the crude extract of ethyl acetate from T. harzianum (LBAT-53).
Jof 10 00547 g007
Figure 8. Fragmentation pathway proposed for chrysophanol, [M−H] ion at m/z 253 according to the MS/MS spectra.
Figure 8. Fragmentation pathway proposed for chrysophanol, [M−H] ion at m/z 253 according to the MS/MS spectra.
Jof 10 00547 g008
Figure 9. Second-order spectra of endocrocin.
Figure 9. Second-order spectra of endocrocin.
Jof 10 00547 g009
Table 1. Effect of T. harzianum LBAT-53 on the radial growth of F. oxysporum f.sp. cubense PalPR7.
Table 1. Effect of T. harzianum LBAT-53 on the radial growth of F. oxysporum f.sp. cubense PalPR7.
TreatmentPercentage of Radial Growth Inhibition(%)
Dual CultureVolatile Metabolites
24 h48 h72 h7 d10 d24 h48 h72 h7 d8 d
Control PalPR70 a0 a0 a0 a0 a0 a0 a0 a0 a0 a
LBAT-53 vs. PalPR716.67 a10.17 b14.94 ab66.09 cd76.90 cd20.83 ab8.53 abc14.14 cd40.65 cd44.44 e
Means with different letters, in the same column, indicate significant differences (ANOVA and Tukey test, p ≤ 0.05).
Table 2. Chemical screening of tested extract from Trichoderma harzianum LBAT-53.
Table 2. Chemical screening of tested extract from Trichoderma harzianum LBAT-53.
Family CompoundsLBAT-53
Fraction A
Phenols+++
Amine compounds+
Fraction B
Triterpenes/steroids-
Quinones++
Fraction C1
Alkaloids-
Fraction C2
Triterpenes/steroids-
Alkaloids-
Cardenolides-
Fraction D
Flavonoids++
Cardenolides-
Alkaloids-
Proanthocyanidins/catechins-
Triterpenes/steroids+
Fraction E
Proanthocyanidins/catechins-
Flavonoids-
Reducing sugars+++
Fraction F
Saponins+
Amine compounds+++
Legend: (-): Negative test (absence of turbidity or precipitation). (+): Weak positive test. (++): Positive test (moderate amount). (+++): Test strongly positive.
Table 3. Anthraquinones: elemental composition and exact mass.
Table 3. Anthraquinones: elemental composition and exact mass.
Ionic SpeciesRt (min)Elemental Composition[M−H]
(m/z Theoretical)
[M−H]
(m/z Experimental)
RDBΔ ppm
[M−H]8.32C15H9O4253.0579253.051511.50.66
[M−H]8.50C15H9O4253.0579253.051511.50.66
[M−H]6.39C15H9O5269.0455269.045911.50.57
[M−H]6.92C16H9O7313.0354313.035012.5−1.09
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

Hernández, G.; Ponce de la Cal, A.; Louis, Y.; Baró Robaina, Y.; Coll, Y.; Spengler, I.; Mirabal-Gallardo, Y. Identification of Secondary Metabolites by UHPLC-ESI-HRMS/MS in Antifungal Strain Trichoderma harzianum (LBAT-53). J. Fungi 2024, 10, 547. https://doi.org/10.3390/jof10080547

AMA Style

Hernández G, Ponce de la Cal A, Louis Y, Baró Robaina Y, Coll Y, Spengler I, Mirabal-Gallardo Y. Identification of Secondary Metabolites by UHPLC-ESI-HRMS/MS in Antifungal Strain Trichoderma harzianum (LBAT-53). Journal of Fungi. 2024; 10(8):547. https://doi.org/10.3390/jof10080547

Chicago/Turabian Style

Hernández, Giselle, Amaia Ponce de la Cal, Yuset Louis, Yamilé Baró Robaina, Yamilet Coll, Iraida Spengler, and Yaneris Mirabal-Gallardo. 2024. "Identification of Secondary Metabolites by UHPLC-ESI-HRMS/MS in Antifungal Strain Trichoderma harzianum (LBAT-53)" Journal of Fungi 10, no. 8: 547. https://doi.org/10.3390/jof10080547

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