Next Article in Journal
Uncovering a Novel cyp51A Mutation and Antifungal Resistance in Aspergillus fumigatus through Culture Collection Screening
Previous Article in Journal
Identification and Characterization of a Plant Endophytic Fungus Paraphaosphaeria sp. JRF11 and Its Growth-Promoting Effects
Previous Article in Special Issue
Decolorization of Textile Azo Dye via Solid-State Fermented Wheat Bran by Lasiodiplodia sp. YZH1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Secretome Analysis of Thermothelomyces thermophilus LMBC 162 Cultivated with Tamarindus indica Seeds Reveals CAZymes for Degradation of Lignocellulosic Biomass

by
Alex Graça Contato
1,2,
Tiago Cabral Borelli
3,
Marcos Silveira Buckeridge
4,
Janet Rogers
5,
Steven Hartson
5,
Rolf Alexander Prade
2 and
Maria de Lourdes Teixeira de Moraes Polizeli
1,6,*
1
Departamento de Bioquímica e Imunologia, Faculdade de Medicina de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto 14049-900, SP, Brazil
2
Department of Microbiology and Molecular Genetics, Oklahoma State University, Stillwater, OK 74078, USA
3
Departamento de Biologia Celular e Molecular e Bioagentes Patogênicos, Faculdade de Medicina de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto 14049-901, SP, Brazil
4
Departamento de Botânica, Instituto de Biociências, Universidade de São Paulo, São Paulo 05508-090, SP, Brazil
5
Department of Biochemistry and Molecular Biology, Oklahoma State University, Stillwater, OK 74078, USA
6
Departamento de Biologia, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto 14040-901, SP, Brazil
*
Author to whom correspondence should be addressed.
J. Fungi 2024, 10(2), 121; https://doi.org/10.3390/jof10020121
Submission received: 11 December 2023 / Revised: 6 January 2024 / Accepted: 11 January 2024 / Published: 1 February 2024

Abstract

:
The analysis of the secretome allows us to identify the proteins, especially carbohydrate-active enzymes (CAZymes), secreted by different microorganisms cultivated under different conditions. The CAZymes are divided into five classes containing different protein families. Thermothelomyces thermophilus is a thermophilic ascomycete, a source of many glycoside hydrolases and oxidative enzymes that aid in the breakdown of lignocellulosic materials. The secretome analysis of T. thermophilus LMBC 162 cultivated with submerged fermentation using tamarind seeds as a carbon source revealed 79 proteins distributed between the five diverse classes of CAZymes: 5.55% auxiliary activity (AAs); 2.58% carbohydrate esterases (CEs); 20.58% polysaccharide lyases (PLs); and 71.29% glycoside hydrolases (GHs). In the identified GH families, 54.97% are cellulolytic, 16.27% are hemicellulolytic, and 0.05 are classified as other. Furthermore, 48.74% of CAZymes have carbohydrate-binding modules (CBMs). Observing the relative abundance, it is possible to state that only thirteen proteins comprise 92.19% of the identified proteins secreted and are probably the main proteins responsible for the efficient degradation of the bulk of the biomass: cellulose, hemicellulose, and pectin.

1. Introduction

Plant biomass consists of proteins, lignin, holocellulose (a fraction composed of cellulose fibers wrapped in hemicellulose-pectin), ash, salts, and minerals [1]. The increase in agro-industrial activity has led to the accumulation of many lignocellulosic residues, such as wood and various agro-industrial residues around the world [2,3,4,5]. The economic interest in these residues has increased significantly in recent years since they are renewable and cheap, having the potential to produce and generate chemicals and bioenergy [6,7,8]. The conversion of lignocellulosic biomass into ethanol and other chemical compounds can be performed using a multi-enzyme system acting in synergism [9,10], and it is of fundamental importance to study different microorganisms and understand the secretion of the enzymes of interest that can be applied to these processes [11].
The analysis of the fungal secretome has gained great visibility since, through these studies, it is possible to know the proteins secreted by different microorganisms, especially carbohydrate-active enzymes (CAZymes), grown under different conditions [12,13,14,15]. Based on their protein sequence similarities and three-dimensional folding structure, CAZymes are classified into several hundred different enzyme protein families [16]. These enzymes are involved in many biological processes, and they are responsible for the degradation, synthesis, and modification of carbohydrates [17,18].
Thermophilic fungi are a promising source of new enzymes for cost-effective industrial applications, including abundant thermostable enzymes for biomass degradation and generation of chemicals and biofuels [19,20,21]. Among them, a fungus that is described in the literature as a source of many CAZymes, especially glycoside hydrolases (GHs) and oxidative enzymes that aid in the breakdown of lignocellulosic materials, is the thermophilic ascomycete Thermothelomyces thermophilus (formerly Myceliophthora thermophila) [22,23,24,25]. This filamentous fungus has been shown to be safe for large-scale production processes and can utilize cost-effective sources of plant biomass [26], as waste from the fruit pulp industry, especially tamarind seeds [27].
Tamarind (Tamarindus indica L.) is a fruit plant native to equatorial Africa, India, and Southeast Asia and grows in tropical and subtropical regions, with an ideal average temperature of 25 °C [28]. It consists of pulp and seeds with a hard coating. Seeds constitute 30–40% of the fruit, with a large proportion being an agricultural by-product [29]. According to Gonçalves et al. [30], the tamarind seed composition is 1.82 ± 0.01% ash, 33.07 ± 1.40% lignin, 33.31 ± 3.56% cellulose, and 10.45 ± 1.45% hemicellulose, proving to be a promising source for the detection of CAZymes.
Due to their constitution, tamarind seeds have been utilized for the cultivation of microorganisms to produce microbial enzymes that cleave lignocellulosic biomass or as substrates in assays to test enzymatic activity [31,32]. In addition, these seeds are rich in xyloglucan, which corresponds to about 40% of their dry mass [33]. In this context, this study aimed to report the elucidation of the secretome profile and categorization of CAZymes by function and family of the filamentous fungus T. thermophilus LMBC 162 cultivated by submerged fermentation using tamarind seeds as a carbon source, which is a residue from the fruit pulp industry [28]. Obtaining these data, it was determined by relative abundance which are the main proteins responsible for the degradation of the biomass bulk: cellulose, hemicellulose, and pectin.

2. Methods and Materials

2.1. Maintenance of the Fungus and Culture Medium

The fungus T. thermophilus LMBC 162 used in this work was isolated in Ribeirão Preto, SP, Brazil. Its identification and deposit in GenBank with the accession code MK559967.1 was described by Contato et al. [31]. The maintenance of the thermophilic microorganism was carried out through the inoculation of its spores in potato dextrose agar medium (PDA) (Sigma-Aldrich, Saint Louis, MO, USA), keeping it through successive transfers in glass tubes containing the same medium and incubating at the temperature of 40 °C. Afterward, the tubes were kept under refrigeration for up to 30 days.

2.2. Submerged Cultivation of T. thermophilus LMBC 162 for Protein Secretion Induction

The submerged cultivation was performed according to Contato et al. [30]. A solution with 106–107 spores/mL was prepared. The fungus was grown in test tubes and suspended in sterile distilled water, and its spores were counted in a microscope through a Neubauer chamber. The suspension was inoculated into 125 mL Erlenmeyer flasks with 25 mL of Khanna medium (Khanna’s salt solution [20×]: NH4NO3 (52.72%), KH2PO4 (34.27%), MgSO4·7H2O (9.54%), KCl (2.58%), MnSO4·H2O (0.36%), ZnSO4·H2O (0.18%), Fe2(SO4)3.6H2O (0.17%), CuSO4.5H2O (0.16%), distilled water q.s. (100 mL) (5.0 mL); yeast extract (0.1 g); carbon source (1.0 g); distilled water q.s. 100 mL) [34]. The medium was supplemented with 1% (w/v) of tamarind (Tamarindus indica, Fabaceae) seeds, which were previously pretreated (boiled in water, dried, and ground to 20 mesh) to secure the sanitary quality of the seeds and avoid the growth of other associated fungi. The Erlenmeyer flasks were incubated at 40 °C under static conditions for 72 h, the best conditions for protein induction described by Contato et al. [31] who showed that these conditions were substantially better than in shaken cultivation and with shorter (24 and 48 h) or longer (96 h) times for evaluating enzyme production. The cell biomass was filtered with the aid of a vacuum pump, and the filtrates were used as enzymatic extracts. This was performed in triplicate.

2.3. Protein Quantification

The proteins obtained in crude extract after cultivation were quantified using Bradford method [35]. Reactions were added with 160 µL of Bradford’s reagent and 40 µL of the enzymatic extracts and incubated for 5 min at room temperature. The absorbance was measured on a spectrophotometer (Shimadzu, Kyoto, Japan) at a wavelength of 595 nm, using bovine albumin as standard. The results were expressed in μg of protein/mL.

2.4. Sample Processing

The supernatant of T. thermophilus cultivated in tamarind seeds under submerged cultivation was collected by filtration after 72 h, concentrated by ultra-filtration (10,000 MWCO, PES membrane, Vivaspin, Littleton, CO, USA), rinsed twice with 5 mL of sodium acetate buffer 50 mM pH 5.0, and the proteins were separated using SDS-PAGE electrophoresis [36].

2.5. Characterization of the T. thermophilus LMBC 162 Using Liquid Chromatography-Tandem Mass Spectrometry (LC-MS/MS)

For secretome peptide mapping, concentrated T. thermophilus LMBC 162 was analyzed by reducing SDS-PAGE on 12% separation gels. For secretome LC–MS/MS analysis, 15–20 μg of identified secretome proteins were loaded onto an SDS-PAGE gel. Preparate PAGE gel electrophoresis was used to separate the protein secretomes from the complex carbohydrate and phenolic species accumulated in the supernatant. Proteins were briefly electrophoresed into the PAGE separating gel, with the electrophoresis being terminated after the bromophenol blue tracking dye had migrated 2–3 cm into the separating gel, stained with Coomassie blue, and the entire protein banding profile excised, processed for LC–MS/MS [37]. Isolated gel bands were reduced with 10 mM Tris (2-carboxyethyl) phosphine for 1 h at room temperature, alkylated by 10 mM using 2-iodoacetamide for 1 h at room temperature in the dark, and digested overnight at 37 °C with 8 μg/mL trypsin (Promega V5072, Madison, WI, USA) (relation proteins in bands: trypsin of 10 µg:mL trypsin/LysC) using 25 mM ammonium bicarbonate buffer, pH 8.0. Peptides were extracted from the gel segments with three sequential extractions at room temperature using 0.3 mL, 0.2 mL, and 0.2 mL of 0.5% trifluoracetic acid, respectively. After intermittent mixing at the vortex, they were dried in SpeedVac (ThermoFisher Scientific, Waltham, MA, USA). Finally, samples were desalted by solid phase extraction using C18 pipet tips, following the manufacturer’s recommendations (Agilent P/N A57003100, Agilent Technologies, Santa Clara, CA, USA). The desalted peptides were redissolved in 0.1% aqueous formic acid and injected onto a 75-micron × 50 cm capillary HPLC column packed with 2-micron C18 particles (Thermo P/N 164942, ThermoFisher Scientific, Waltham, MA, USA). Peptides were separated using a 60 min gradient of formic acid/acetonitrile with a flow rate of 250 nL/min and ionized in a Nanospray Flex (ThermoFisher Scientific, Waltham, MA, USA) ion source using stainless-steel emitters connected to a quadrupole-Orbitrap mass spectrometer (Fusion model, ThermoFisher Scientific, Waltham, MA, USA). Peptide ions were analyzed using a “high-low” “top-speed” data-dependent MS/MS strategy, wherein peptide precursors were analyzed at high resolution in the Orbitrap sector, selected for MS/MS using the quadrupole sector, fragmented by HCD in the ion routing multipole, followed by analysis of the fragment ions in the ion trap sector. MS/MS parameters used in the experiments are ions spray voltage (1900 W); capillary temperature (300 °C); mass range in full MS mode (375–1575 m/z); resolution setting for full mass MS scan, AGC target value, maximum injection time (120,000 nominal resolution, 4 × 105 ions, 50 ms, respectively); number of peptides selected to be fragmented in each duty cycle (data-dependent acquisition limited only by cyclic rate, set at 5 s); value of normalized collision energy (32% HCD energy); resolution settings for MS/MS acquisition, AGC target value, maximum injection time (MS/MS analysis in the ion trap sector using a rapid scan rate, 5 × 104 ions, dynamic injection timing limits wherein the system maximizes the injection times available for relative to the stated cycle time and maximizing sensitivity, respectively); charges of precursor ions excluded (below +2 or above +6 were excluded); and dynamic exclusion time (dynamic exclusion was set to 45 s).
Each sample was analyzed twice by LC-MS/MS, and the two RAW data files were specified as a single sample for database searching using MaxQuant (version v2.0.1.0, Max-Planck-Institute of Biochemistry, Planegg, Germany) [38]. Spectra were searched against a database of 18,464 T. thermophilus protein sequences downloaded from NCBI on 27 May 2022, using Thermothelomyces as a genus search term. Searches were annotated using Python version v3.11 (Python Software Foundation, Wilmington, DE, USA) to annotate NCBI T. thermophilus ID’s by transferring annotations from related curated proteins at Uniprot (https://www.uniprot.org/ accessed on 14 December 2022). Sequences with a false discovery rate (FDR or q-value) greater than 0.00 were removed from the analysis. Finally, we identified conserved CAZy domains using Hidden Markov Models (HMM) profiles available on the dbCAN2 web platform (https://bcb.unl.edu/dbCAN2/index.php accessed on 14 December 2022). Only domains with e-values > 10−17 and coverage > 0.35 were considered.

3. Results and Discussion

Analysis of Secretome Protein Composition

To characterize the secretome of T. thermophilus LMBC 162, the supernatants of cultures were collected and analyzed using LC-MS/MS searching against a database of Thermothelomyces sequences downloaded from the NCBI. The identified proteins were annotated by searching the T. thermophilus sequences against curated protein sequences available in the Uniprot/Swiss-Prot database. Our analysis identified 79 proteins in the T. thermophilus LMBC 162 secretome (all non-anchored extracellular proteins). Taking into account the quantification of these proteins through their relative abundance referenced by the IBAQ value (sum of all the peptides intensities divided by the number of observable peptides of a protein), we found five diverse classes of CAZymes: 5.55% auxiliary activity (AAs); 2.58% carbohydrate esterases (CEs); 20.58% polysaccharide lyases (PLs); and 71.29% glycoside hydrolases (GHs), which were 54.97% cellulolytic GHs, 16.27% hemicellulolytic GHs, and 0.05 classified as other GHs. Furthermore, 48.74% of CAZymes have carbohydrate-binding modules (CBMs) (Figure 1). These values are consistent with others shown in the literature for other filamentous fungi [24,39].
Auxiliary activity (AA) enzymes. The AAs are families of catalytic proteins that are potentially involved in plant cell degradation through an ability to help the original glycoside hydrolase, polysaccharide lyase, and carbohydrate esterase enzymes to gain access to the carbohydrates comprising the plant cell wall [40]. Among the 17 auxiliary activity enzymes, we observed eight (8) lytic polysaccharide monooxygenases (LPMOs), with the majority being from the AA9 CAZy domain (Table 1). Analyzing quantitatively by the relative abundance (IBAQ value), we identified that the AAs correspond to 5.55% of the proteins detected in the secretome analysis. Other AA CAZy domains found are AA3, AA5, AA7, AA8, AA12, and AA13. These results corroborate studies that, when performing the secretome profile of this microorganism with other cultivated sources, also verified the presence of these oxidative enzymes [22,24].
Carbohydrate esterases (CEs). CE catalyzes the de-O- or de-N-acylation by removing the ester decorations from carbohydrates. They represent biocatalysts important for bioconversion of plant biomass and saccharification of plant cell wall polysaccharide fractions that have not gone through an alkaline pretreatment or process that would destroy the ester linkages [41]. In this study, seven (7) CEs were found, corresponding to 2.58% of the proteins of the T. thermophilus LMBC 162 secretome cultivated using submerged fermentation with tamarind seeds (Table 2). These values are consistent with those shown by Rocha et al. [42], who found a relative abundance of 2% CEs in the Trichoderma harzianum secretome when cultivated on sugarcane bagasse, and with the study by Machado et al. [39] who obtained 3.4% of esterases in the Trametes versicolor secretome cultivated on microcrystalline cellulose. Among the CEs identified in this work, two were highlighted with 2.29% relative abundance, that is, 88.75% of the CEs. They are an acetylesterase CE16, an enzyme that catalyzes the conversion of acetate esters and water into alcohols and acetate [43], and a pectinesterase CE8 (accession number G2QLD0) that catalyzes the de-esterification of pectin into pectate and methanol [44]. They correspond to 1.19% and 1.10% of relative abundance, respectively. Other CE CAZy domains found are CE1, CE3, CE5, and CE12.
Polysaccharide lyases (PLs). PLs are a group of enzymes that cleave uronic acid-containing polysaccharide chains via a β-elimination mechanism to generate an unsaturated hexenuronic acid residue and a new reducing end [45]. In this work, six (6) PLs were visualized, which correspond to 20.58% of the secretome. However, there is a huge emphasis on a particular protein from the PL1 family, accession number (G2QH79) in the UniProt/Swiss-Prot database and hypothetical molecular weight of (34 kDa) (Table 3). This PL alone corresponds to 19.95% of relative abundance, which is 96.94% of all identified PLs. These values are higher than the 9% seen by Verma et al. [46] in the secretome of the fungal phytopathogen Ascochyta rabiei and the 7% of those seen by Rubio et al. [47] for Aspergillus nidulans. However, this is consistent with the study by dos Santos et al. [24] who used another strain of Myceliophthora thermophila and cultivated on lignocellulosic residues. Other PL CAZy domains found are PL3 and PL4. The classes of CEs and PLs are mainly responsible for the degradation of pectin, one of the main components of the cell wall of plants [48].
Cellulolytic glycoside hydrolases (GHs). The GHs are enzymes that catalyze the hydrolysis of the glycosidic linkage of glycosides, leading to the formation of a sugar hemiacetal or hemiketal and the corresponding free aglycon [49]. The GHs that cleave sugars from cellulose are named cellulolytic glycoside hydrolases, and those that cleave sugars from hemicellulose are named hemicellulolytic glycoside hydrolases. Of the CAZymes found in the secretome of T. thermophilus LMBC 162, forty-nine (49) are GHs, corresponding to 71.29% in relative abundance. Among them, the majority are those that breakdown cellulose. They account for 54.97% of the CAZymes produced in the secretome (Table 4). One cellobiohydrolase GH7 and a glucoside hydrolase from the GH7 CAZy domain are the main proteins found. They can be seen with accession numbers G2Q665 and G2QNN8 in the UniProt/Swiss-Prot database. The hypothetical molecular weight of each is 56 kDa and 49 kDa, respectively. Other GHs CAZy domains found are GH6, GH15, GH31, and GH45. The value of cellulolytic GHs found is like those seen by Machado et al. [39] who found 48.1% for Phanerochaete chrysosporium and 48.0% for T. versicolor.
Hemicellulolytic glycoside hydrolases (GHs). Regarding the hemicellulolytic GHs found (Table 5), they correspond to 16.27% in relative abundance, once again being equivalent to values reported by others [24,39]. The hemicellulolytic GHs belong to the CAZy domains: GH2, GH3, GH10, GH11, GH16, GH26, GH27, GH43, GH47, GH55, GH62, GH74, GH76, GH79, GH92, GH93, GH125, GH131, and GH135. The hemicellulolytic GHs with the highest relative abundance are a xylanase GH10 with a hypothetical molecular weight of (45 kDa) and accession number (G2QJ91) in the Uniprot/Swiss-Prot database, with 2.11% of relative abundance; a xylanase GH11 with a hypothetical molecular weight of (24 kDa) and accession number (G2Q4M3), with a relative abundance of 2.59%; a xyloglucanase GH74 of hypothetical molecular weight of (79 kDa) and accession number (G2QHR7), with 5.39% of relative abundance; and an exo-α-L-1,5-arabinanase GH93 with a hypothetical molecular weight of (42 kDa) and accession number (G2Q5Q6), with a relative abundance of 2.30%. The GH74 found in this work showed a hypothetical molecular weight corresponding to the xyloglucanase found by Berezina et al. [50], who expressed a GH74 xyloglucanase from M. thermophila in Pichia pastoris. Another relevant factor is that this secretome was performed in tamarind seeds, which are rich in xyloglucan [32], thus proving why the GH74 was produced with the highest relative abundance. Regarding the GH93 family, it is known to hydrolyze linear α-1,5-L-arabinan [51].
In the analysis of T. thermophilus LMBC 162 secreted proteins belonging to the hemicellulolytic GH family, two uncharacterized proteins were determined. The GH131 family are β-glucanases that exhibit activity for a wide range of β-glucan polysaccharides, including laminarin, curdlan, lichenan, and cellulosic derivatives [52], while the GH135 family has disclosed fungal glycoside hydrolases with the ability to degrade the fungal heteropolysaccharide galactosaminogalactan [53].
Other relevant GHs found in T. thermophilus LMBC 162 secretome are the GH3, GH10, GH16, GH43, and GH55 CAZy domains. GH3 removes single glycoside residues from the non-reducing ends of their substrates and currently groups together exo-acting β-D-glucosidases, α-L-arabinofuranosidases, β-D-xylopyranosidases, N-acetyl-β-D-glucosaminidases, and N-acetyl-β-D-glucosaminide phosphorylases [54]; GH10 are endo-β-1,4-xylanases [55]; GH16 are active on β-1,4 or β-1,3 glycosidic bonds in various glucans and galactans, which include lichenases (EC 3.2.1.73) [56]. GH43 are α-L-arabinofuranosidases, endo-α-L-arabinanases (or endo-processive arabinanases), and β-D-xylosidases [57]; and GH55 consists exclusively of β-1,3-glucanases, including both exo-and endo-enzymes [58].
Other glycoside hydrolases (GHs). Table 6 shows GHs identified that breakdown other components. Only two GHs were detected: a trehalase GH37, which corresponds to 0.01% of the secretome, and an α-L-fucosidase GH95, corresponding to 0.04% of the secretome. The enzymes from family GH37 have been shown to hydrolyze the α-1,1 bound trehalose (α-D-glucopyranosyl-(1→1)-α-D-glucopyranoside) into two molecules of D-glucose [59], while the α-L-fucosidase GH95 hydrolyzes α-Fuc-1,2-Gal linkages attached at the non-reducing ends of oligosaccharides [60].
Carbohydrate binding modules (CBMs). CBMs are protein domains found in CAZymes, whose main role is to recognize and bind specifically to carbohydrates. The consequences of this event result in different functions, such as increased hydrolysis of insoluble substrates, bringing the catalytic domain closer to the substrate, polysaccharide structure disruption, and cell surface protein anchoring [61]. In the secretome of T. thermophilus LMBC 162, sixteen (16) CAZymes have CBMs (Table 7). In sum, CBMs are present in 48.74% of the proteins found in the T. thermophilus secretome profile, a value like those shown in the literature for other microorganisms [39].
Observing the iBAQ/total iBAQ values as a percentage, it is possible to state that only thirteen (13) proteins comprise 92.19% of the identified proteins secreted, and they are probably the main proteins responsible for the degradation of the bulk of the biomass: cellulose, hemicellulose, and pectin (Table 8). The most abundant protein in the secretome of T. thermophilus LMBC 162 (31.43%) was a cellobiohydrolase, like the secretome of Trichoderma reesei RUT C30, where a cellobiohydrolase is the most abundant protein [12]. However, the presence of other enzymes, such as β-xylanase, lytic polysaccharide monooxygenase, and pectinesterase, was reported. Nevertheless, one of the limitations of shotgun proteomics is incomplete sequence coverage when using only one protease. Therefore, there is a possibility that other proteins, such as small proteins due to the few theoretical peptides produced in digestion, were not detected [62].

4. Conclusions

The secretome analysis of T. thermophilus LMBC 162 cultivated by submerged fermentation with tamarind seeds, an abundant residue from the fruit pulp industry, reveals seventy-nine (79) CAZymes diversified into the five classes of CAZy database: 5.55% AAs; 1.48% CBMs; 2.58% CEs, 20.58% PLs; and 70.55% GHs, which are 54.97% cellulolytic GHs, 15.51% hemicellulolytic GHs, and 0.05 classified as other GHs. Between them, sixteen (16) CAZymes have CBMs, protein domains found in CAZymes, whose main role is to recognize and bind specifically to carbohydrates. In sum, CBMs are present in 48.74% of the proteins found in the T. thermophilus secretome profile, a value like those shown in the literature for other microorganisms. Observing the relative abundance, it is possible to state that only thirteen (13) proteins comprise 92.19% of the identified proteins secreted, and they are probably the main proteins responsible for the degradation of the bulk of the biomass: cellulose, hemicellulose, and pectin. The findings of this work allow us to say that tamarind seeds are a residue option for the identification and production of lignocellulosic CAZymes.

Author Contributions

A.G.C.: Methodology, Investigation, Data Curation, Formal Analysis, Writing—original draft. T.C.B.: Methodology, Investigation, Data Curation, Formal Analysis. M.S.B.: Funding acquisition, Conceptualization. J.R.: Methodology, Investigation, Data Curation, Formal Analysis, Writing—review and editing. S.H.: Methodology, Investigation, Data Curation, Formal Analysis, Writing—review and editing. R.A.P.: Methodology, Investigation, Data Curation, Formal Analysis, Funding acquisition, Supervision, Writing—review and editing. M.d.L.T.d.M.P.: Funding acquisition, Conceptualization, Supervision, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) for the doctorate grants awarded to A.G. Contato (Processes nº 2017/25862-6 and 2021/07066-3) and to T.C. Borelli (Process nº 2021/08235-3); and for financial support from National Institute of Science and Technology of Bioethanol (Process FAPESP 2014/50884-5), and Process 2018/07522-6. Conselho Nacional de Desenvolvimento Científico (CNPq), process 465319/2014-9. M.L.T.M. Polizeli (process 301963/2017-7) and M.S. Buckeridge are Research Fellows of CNPq.

Institutional Review Board Statement

The research meets all ethical guidelines, including adherence to the legal requirements of the study countries.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Acknowledgments

Mass spectrometry analyses and primary database searches were conducted in the Center for Genomics and Proteomics at Oklahoma State University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yamaguchi, A.; Mimura, N.; Shirai, M.; Sato, O. Cascade utilization of biomass: Strategy for conversion of cellulose, hemicellulose, and lignin into useful chemicals. ACS Sustain. Chem. Eng. 2019, 7, 10445–10451. [Google Scholar] [CrossRef]
  2. Ravindra, K.; Singh, T.; Mor, S. Emissions of air pollutants from primary crop residue burning in India and their mitigation strategies for cleaner emissions. J. Clean. Prod. 2019, 208, 261–273. [Google Scholar] [CrossRef]
  3. Yadav, M.; Paritosh, K.; Pareek, N.; Vivekanand, V. Coupled treatment of lignocellulosic agricultural residues for augmented biomethanation. J. Clean. Prod. 2019, 213, 75–88. [Google Scholar] [CrossRef]
  4. Singh, T.A.; Sharma, M.; Sharma, M.; Sharma, G.D.; Passari, A.K.; Bhasin, S. Valorization of agro-industrial residues for production of commercial biorefinery products. Fuel 2022, 322, 124284. [Google Scholar] [CrossRef]
  5. Ortega, F.; Versino, F.; López, O.V.; García, M.A. Biobased composites from agro-industrial wastes and by-products. Emergent Mater. 2022, 5, 873–921. [Google Scholar] [CrossRef]
  6. Bechara, R.; Gomez, A.; Saint-Antonin, V.; Schweitzer, J.M.; Maréchal, F.; Ensinas, A. Review of design works for the conversion of sugarcane to first and second-generation ethanol and electricity. Renew. Sust. Energy Rev. 2018, 91, 152–164. [Google Scholar] [CrossRef]
  7. Klein, B.C.; Sampaio, I.L.M.; Mantelatto, P.E.; Maciel Filho, R.; Bonomi, A. Beyond ethanol, sugar, and electricity: A critical review of product diversification in Brazilian sugarcane mills. Biofuel Bioprod. Biorefining 2019, 13, 809–821. [Google Scholar] [CrossRef]
  8. Magalhães, A.I., Jr.; de Carvalho, J.C.; Pereira, G.V.M.; Karp, S.G.; Câmara, M.C.; Medina, J.D.C.; Soccol, C.R. Lignocellulosic biomass from agro-industrial residues in South America: Current developments and perspectives. Biofuel Bioprod. Biorefining 2019, 13, 1505–1519. [Google Scholar] [CrossRef]
  9. Patel, A.; Shah, A.R. Integrated lignocellulosic biorefinery: Gateway for production of second generation ethanol and value added products. J. Bioresour. Bioprod. 2021, 6, 108–128. [Google Scholar] [CrossRef]
  10. Chukwuma, O.B.; Rafatullah, M.; Tajarudin, H.A.; Ismail, N. Lignocellulolytic enzymes in biotechnological and industrial processes: A review. Sustainability 2020, 12, 7282. [Google Scholar] [CrossRef]
  11. dos Santos, J.R.; Moreira, L.R.S.; Ferreira Filho, E.X. Cellulose-degrading enzymes: Key players in biorefinery development. Biologia 2023, 78, 1759–1772. [Google Scholar] [CrossRef]
  12. Borin, G.P.; Sanchez, C.C.; de Souza, A.P.; de Santana, E.S.; de Souza, A.T.; Leme, A.F.P.; Squina, F.M.; Buckeridge, M.S.; Goldman, G.H.; Oliveira, J.V.D.C. Comparative secretome analysis of Trichoderma reesei and Aspergillus niger during growth on sugarcane biomass. PLoS ONE 2015, 10, e0129275. [Google Scholar] [CrossRef]
  13. Basotra, N.; Kaur, B.; Di Falco, M.; Tsang, A.; Chadha, B.S. Mycothermus thermophilus (Syn. Scytalidium thermophilum): Repertoire of a diverse array of efficient cellulases and hemicellulases in the secretome revealed. Bioresour. Technol. 2016, 222, 413–421. [Google Scholar] [CrossRef] [PubMed]
  14. Raheja, Y.; Kaur, B.; di Falco, M.; Tsang, A.; Chadha, B.S. Secretome analysis of Talaromyces emersonii reveals distinct CAZymes profile and enhanced cellulase production through response surface methodology. Ind. Crop. Prod. 2020, 152, 112554. [Google Scholar] [CrossRef]
  15. Wijayawardene, N.N.; Boonyuen, N.; Ranaweera, C.B.; de Zoysa, H.K.; Padmathilake, R.E.; Nifla, F.; Dai, D.Q.; Liu, X.; Suwannarachi, N.; Kumla, J.; et al. OMICS and other advanced technologies in mycological applications. J. Fungi 2023, 9, 688. [Google Scholar] [CrossRef]
  16. Barrett, K.; Jensen, K.; Meyer, A.S.; Frisvad, J.C.; Lange, L. Fungal secretome profile categorization of CAZymes by function and family corresponds to fungal phylogeny and taxonomy: Example Aspergillus and Penicillium. Sci. Rep. 2020, 10, 5158. [Google Scholar] [CrossRef] [PubMed]
  17. Nin-Hill, A.; Piniello, B.; Rovira, C. In silico modelling of the function of disease-related CAZymes. Essays Biochem. 2023, 67, 355–372. [Google Scholar] [CrossRef] [PubMed]
  18. Salazar-Cerezo, S.; de Vries, R.P.; Garrigues, S. Strategies for the development of industrial fungal producing strains. J. Fungi 2023, 9, 834. [Google Scholar] [CrossRef]
  19. Perez, C.L.; Casciatori, F.P.; Thomeo, J.C. Strategies for scaling-up packed-bed bioreactors for solid-state fermentation: The case of cellulolytic enzymes production by a thermophilic fungus. Chem. Eng. J. 2019, 361, 1142–1151. [Google Scholar] [CrossRef]
  20. Liu, E.; Segato, F.; Wilkins, M.R. Fed-batch production of Thermothelomyces thermophilus lignin peroxidase using a recombinant Aspergillus nidulans strain in stirred-tank bioreactor. Bioresour. Technol. 2021, 325, 124700. [Google Scholar] [CrossRef]
  21. Yang, J.; Yue, H.R.; Pan, L.Y.; Feng, J.X.; Zhao, S.; Suwannarangsee, S.; Champreda, V.; Liu, C.G.; Zhao, X.Q. Fungal strain improvement for efficient cellulase production and lignocellulosic biorefinery: Current status and future prospects. Bioresour. Technol. 2023, 385, 129449. [Google Scholar] [CrossRef] [PubMed]
  22. Berka, R.M.; Grigoriev, I.V.; Otillar, R.; Salamov, A.; Grimwood, J.; Reid, I.; Ishmael, N.; John, T.; Darmond, C.; Moisan, M.C.; et al. Comparative genomic analysis of the thermophilic biomass-degrading fungi Myceliophthora thermophila and Thielavia terrestris. Nat. Biotechnol. 2011, 29, 922–927. [Google Scholar] [CrossRef] [PubMed]
  23. Marin-Felix, Y.; Stchigel, A.M.; Miller, A.N.; Guarro, J.; Cano-Lira, J.F. A re-evaluation of the genus Myceliophthora (Sordariales, Ascomycota): Its segregation into four genera and description of Corynascus fumimontanus sp. nov. Mycologia 2015, 107, 619–632. [Google Scholar] [CrossRef] [PubMed]
  24. dos Santos, H.B.; Bezerra, T.M.S.; Pradella, J.C.G.; Delabona, P.; Lima, D.; Gomes, E.; Hartson, S.D.; Rogers, J.; Couger, B.; Prade, R. Myceliophthora thermophila M77 utilizes hydrolytic and oxidative mechanisms to deconstruct biomass. AMB Express 2016, 6, 103. [Google Scholar] [CrossRef] [PubMed]
  25. Chorozian, K.; Karnaouri, A.; Karantonis, A.; Souli, M.; Topakas, E. Characterization of a dual cellulolytic/xylanolytic AA9 lytic polysaccharide monooxygenase from Thermothelomyces thermophilus and its utilization toward nanocellulose production in a multi-step bioprocess. ACS Sustain. Chem. Eng. 2022, 10, 8919–8929. [Google Scholar] [CrossRef]
  26. Balabanova, L.; Seitkalieva, A.; Yugay, Y.; Rusapetova, T.; Slepchenko, L.; Podvolotskaya, A.; Yatsunskaya, M.; Vasyutkina, E.; Son, O.; Tekutyeva, L.; et al. Engineered fungus Thermothelomyces thermophilus producing plant storage proteins. J. Fungi 2022, 8, 119. [Google Scholar] [CrossRef]
  27. Contato, A.G.; Nogueira, K.M.V.; Buckeridge, M.S.; Silva, R.N.; Polizeli, M.L.T.M. Trichoderma longibrachiatum and Thermothelomyces thermophilus co-culture: Improvement the saccharification profile of different sugarcane bagasse varieties. Biotechnol. Lett. 2023, 45, 1093–1102. [Google Scholar] [CrossRef]
  28. Ramesh, T.; Rajalaksmi, N.; Dhatathrevan, K.S. Activated carbons derived from tamarind seeds for hydrogen storage. J. Energy Storage 2015, 4, 89–95. [Google Scholar] [CrossRef]
  29. Reis, P.M.C.L.; Dariva, C.; Vieira, G.A.B.; Hense, H. Extraction and evaluation of antioxidant potential of the extracts obtained from tamarind seeds (Tamarindus indica), sweet variety. J. Food Eng. 2016, 173, 116–123. [Google Scholar] [CrossRef]
  30. Gonçalves, G.R.; Gandolfi, O.R.; Bonomo, R.C.F.; Fontan, R.D.C.I.; Veloso, C.M. Synthesis of activated carbon from hydrothermally carbonized tamarind seeds for lipase immobilization: Characterization and application in aroma ester synthesis. J. Chem. Technol. Biotechnol. 2021, 96, 3316–3329. [Google Scholar] [CrossRef]
  31. Contato, A.G.; de Oliveira, T.B.; Aranha, G.M.; de Freitas, E.N.; Vici, A.C.; Nogueira, K.M.V.; de Lucas, R.C.; Scarcella, A.S.A.; Buckeridge, M.S.; Silva, R.N.; et al. Prospection of fungal lignocellulolytic enzymes produced from jatoba (Hymenaea courbaril) and tamarind (Tamarindus indica) seeds: Scaling for bioreactor and saccharification profile of sugarcane bagasse. Microorganisms 2021, 9, 533. [Google Scholar] [CrossRef] [PubMed]
  32. Calderaro, F.; Keser, M.; Akeroyd, M.; Bevers, L.E.; Eijsink, V.G.H.; Várnai, A.; van den Berg, M.A. Characterization of an AA9 LPMO from Thielavia australiensis, TausLPMO9B, under industrially relevant lignocellulose saccharification conditions. Biotechnol. Biofuels 2020, 13, 1–17. [Google Scholar] [CrossRef]
  33. Buckeridge, M.S.; Dietrich, S.M.C. Galactomannan from Brazilian legume seeds. Rev. Bras. Bot. 1990, 13, 109–112. [Google Scholar]
  34. Khanna, P.; Sundari, S.S.; Kumar, N.J. Production, isolation, and partial purification of xylanases from an Aspergillus sp. World J. Microbiol. Biotechnol. 1995, 11, 242–243. [Google Scholar] [CrossRef] [PubMed]
  35. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef] [PubMed]
  36. Weber, K.; Osborn, M. The reliability of molecular weight determinations by dodecyl sulfate-polyacrylamide gel electrophoresis. J. Biol. Chem. 1969, 244, 4406–4412. [Google Scholar] [CrossRef]
  37. Voruganti, S.; Kline, J.T.; Balch, M.J.; Rogers, J.; Matts, R.L.; Hartson, S.D. Proteomic profiling of Hsp90 inhibitors. In Chaperones; Humana Press: New York, NY, USA, 2018; pp. 139–162. [Google Scholar]
  38. Cox, J.; Mann, M. MaxQuant enables high peptide identification rates, individualized ppb-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 2008, 26, 1367–1372. [Google Scholar] [CrossRef]
  39. Machado, A.S.; Valadares, F.; Silva, T.F.; Milagres, A.M.; Segato, F.; Ferraz, A. The secretome of Phanerochaete chrysosporium and Trametes versicolor grown in microcrystalline cellulose and use of the enzymes for hydrolysis of lignocellulosic materials. Front. Bioeng. Biotechnol. 2020, 8, 826. [Google Scholar] [CrossRef]
  40. Levasseur, A.; Drula, E.; Lombard, V.; Coutinho, P.M.; Henrissat, B. Expansion of the enzymatic repertoire of the CAZy database to integrate auxiliary redox enzymes. Biotechnol. Biofuels 2013, 6, 41. [Google Scholar] [CrossRef]
  41. Biely, P. Microbial carbohydrate esterases deacetylating plant polysaccharides. Biotechnol. Adv. 2012, 30, 1575–1588. [Google Scholar] [CrossRef]
  42. Rocha, V.A.L.; Maeda, R.N.; Pereira, N., Jr.; Kern, M.F.; Elias, L.; Simister, R.; Steele-King, C.; Gómez, L.D.; McQueen-Mason, S.J. Characterization of the cellulolytic secretome of Trichoderma harzianum during growth on sugarcane bagasse and analysis of the activity boosting effects of swollenin. Biotechnol. Progress 2016, 32, 327–336. [Google Scholar] [CrossRef] [PubMed]
  43. Kanelli, M.; Topakas, E. Acylation of soluble polysaccharides in a biphasic system catalyzed by a CE2 acetylesterase. Carbohydr. Polym. 2017, 163, 208–215. [Google Scholar] [CrossRef] [PubMed]
  44. Satapathy, S.; Rout, J.R.; Kerry, R.G.; Thatoi, H.; Sahoo, S.L. Biochemical prospects of various microbial pectinase and pectin: An approachable concept in pharmaceutical bioprocessing. Front. Nutr. 2020, 7, 117. [Google Scholar] [CrossRef] [PubMed]
  45. Garron, M.L.; Cygler, M. Uronic polysaccharide degrading enzymes. Curr. Opin. Struct. Biol. 2014, 28, 87–95. [Google Scholar] [CrossRef]
  46. Verma, S.; Gazara, R.K.; Nizam, S.; Parween, S.; Chattopadhyay, D.; Verma, P.K. Draft genome sequencing and secretome analysis of fungal phytopathogen Ascochyta rabiei provides insight into the necrotrophic effector repertoire. Sci. Rep. 2016, 6, 24638. [Google Scholar] [CrossRef]
  47. Rubio, M.V.; Zubieta, M.P.; Cairo, J.P.L.F.; Calzado, F.; Leme, A.F.P.; Squina, F.M.; Prade, R.A.; Damásio, A.R.L. Mapping N-linked glycosylation of carbohydrate-active enzymes in the secretome of Aspergillus nidulans grown on lignocellulose. Biotechnol. Biofuels 2016, 9, 168. [Google Scholar] [CrossRef] [PubMed]
  48. Bonnin, E.; Pelloux, J. Pectin degrading enzymes. In Pectin: Technological and Physiological Properties; Kontogiorgos, V., Ed.; Springer International Publishing: Cham, Switzerland, 2020; pp. 37–60. [Google Scholar]
  49. Prieto, A.; de Eugenio, L.; Méndez-Líter, J.A.; Nieto-Domínguez, M.; Murgiondo, C.; Barriuso, J.; Bejarano-Muñoz, L.; Martínez, M.J. Fungal glycosyl hydrolases for sustainable plant biomass valorization: Talaromyces amestolkiae as a model fungus. Int. Microbiol. 2021, 24, 545–558. [Google Scholar] [CrossRef]
  50. Berezina, O.V.; Herlet, J.; Rykov, S.V.; Kornberger, P.; Zavyalov, A.; Kozlov, D.; Sakhibgaraeva, L.; Krestyanova, I.; Schwarz, W.H.; Zverlov, Z.Z.; et al. Thermostable multifunctional GH74 xyloglucanase from Myceliophthora thermophila: High-level expression in Pichia pastoris and characterization of the recombinant protein. Appl. Microbiol. Biotechnol. 2017, 101, 5653–5666. [Google Scholar] [CrossRef]
  51. Velasco, J.; Oliva, B.; Gonçalves, A.L.; Lima, A.S.; Ferreira, G.; França, B.A.; Mulinari, E.J.; Gonçalves, T.A.; Squina, F.M.; Kadowaki, M.A.S.; et al. Functional characterization of a novel thermophilic exo-arabinanase from Thermothielavioides terrestris. Appl. Microbiol. Biotechnol. 2020, 104, 8309–8326. [Google Scholar] [CrossRef]
  52. Lafond, M.; Navarro, D.; Haon, M.; Couturier, M.; Berrin, J.G. Characterization of a broad-specificity β-glucanase acting on β-(1,3)-, β-(1,4)-, and β-(1,6)-glucans that defines a new glycoside hydrolase family. Appl. Environ. Microbiol. 2012, 78, 8540–8546. [Google Scholar] [CrossRef]
  53. Bamford, N.C.; Snarr, B.D.; Gravelat, F.N.; Little, D.J.; Lee, M.J.; Zacharias, C.A.; Chabot, J.C.; Geller, A.M.; Baptista, S.D.; Baker, P.; et al. Sph3 is a glycoside hydrolase required for the biosynthesis of galactosaminogalactan in Aspergillus fumigatus. J. Biol. Chem. 2015, 290, 27438–27450. [Google Scholar] [CrossRef] [PubMed]
  54. Harvey, A.J.; Hrmova, M.; De Gori, R.; Varghese, J.N.; Fincher, G.B. Comparative modeling of the three-dimensional structures of family 3 glycoside hydrolases. Proteins 2000, 41, 257–269. [Google Scholar] [CrossRef] [PubMed]
  55. Velasco, J.; Oliva, B.; Mulinari, E.J.; Quintero, L.P.; Lima, A.S.; Gonçalves, A.L.; Gonçalves, T.A.; Damasio, A.; Squina, F.M.; Milagres, A.M.F.; et al. Heterologous expression and functional characterization of a GH10 endoxylanase from Aspergillus fumigatus var. niveus with potential biotechnological application. Biotechnol. Rep. 2019, 24, e00382. [Google Scholar] [CrossRef]
  56. Liberato, M.V.; Prates, E.T.; Gonçalves, T.A.; Bernardes, A.; Vilela, N.; Fattori, J.; Ematsu, G.C.; Chinaglia, M.; Gomes, E.R.M.; Figueira, A.C.M.; et al. Insights into the dual cleavage activity of the GH16 laminarinase enzyme class on β-1,3 and β-1,4 glycosidic bonds. J. Biol. Chem. 2021, 296, 100385. [Google Scholar] [CrossRef]
  57. Mewis, K.; Lenfant, N.; Lombard, V.; Henrissat, B. Dividing the large glycoside hydrolase family 43 into subfamilies: A motivation for detailed enzyme characterization. Appl. Environ. Microbiol. 2016, 82, 1686–1692. [Google Scholar] [CrossRef]
  58. Bianchetti, C.M.; Takasuka, T.E.; Deutsch, S.; Udell, H.S.; Yik, E.J.; Bergeman, L.F.; Fox, B.G. Active site and laminarin binding in glycoside hydrolase family 55. J. Biol. Chem. 2015, 290, 11819–11832. [Google Scholar] [CrossRef]
  59. Sakaguchi, M. Diverse and common features of trehalases and their contributions to microbial trehalose metabolism. Appl. Microbiol. Biotechnol. 2020, 104, 1837–1847. [Google Scholar] [CrossRef]
  60. Silchenko, A.S.; Rubtsov, N.K.; Zueva, A.O.; Kusaykin, M.I.; Rasin, A.B.; Ermakova, S.P. Fucoidan-active α-L-fucosidases of the GH29 and GH95 families from a fucoidan degrading cluster of the marine bacterium Wenyingzhuangia fucanilytica. Arch. Biochem. Biophys. 2022, 728, 109373. [Google Scholar] [CrossRef]
  61. Guillén, D.; Sánchez, S.; Rodríguez-Sanoja, R. Carbohydrate-binding domains: Multiplicity of biological roles. Appl. Microbiol. Biotechnol. 2010, 85, 1241–1249. [Google Scholar] [CrossRef]
  62. McDonald, W.H.; Yates, J.R., III. Shotgun proteomics and biomarker discovery. Dis. Markers 2002, 18, 99–105. [Google Scholar] [CrossRef]
Figure 1. CAZymes from secretome analysis of T. thermophilus LMBC 162. Graph units are based on relative abundance according to data from the LC-MS/MS. Auxiliary activity (AA), carbohydrate-binding module (CBM), carbohydrate esterases (CEs), polysaccharide lyases (PLs), and glycoside hydrolases (GHs).
Figure 1. CAZymes from secretome analysis of T. thermophilus LMBC 162. Graph units are based on relative abundance according to data from the LC-MS/MS. Auxiliary activity (AA), carbohydrate-binding module (CBM), carbohydrate esterases (CEs), polysaccharide lyases (PLs), and glycoside hydrolases (GHs).
Jof 10 00121 g001
Table 1. LC-MS/MS secretome analysis for auxiliary activity (AA) enzymes.
Table 1. LC-MS/MS secretome analysis for auxiliary activity (AA) enzymes.
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy
Domain
BLAST
E-Value b
iBAQ ciBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
glucose–methanol–choline GMC oxidoreductaseG2PZJ2XP_003660923.169AA39.5 × 10−153116990000.35
cellobiose dehydrogenaseG2QGP4XP_003663851.168AA3—CBM12.2 × 10−762311400.73
cellobiose dehydrogenaseG2QNS9XP_003666548.157AA37 × 10−1972103600.01
glyoxal oxidaseG2Q335XP_003658743.1105AA54.8 × 10−142860490.01
FAD linked oxidaseG2QG48XP_003663758.154AA75.9 × 10−928182800.02
FAD linked oxidaseG2Q654XP_003660778.169AA71.8 × 10−45981500.01
cellobiose dehydrogenaseG2QFY4XP_003664543.184AA86 × 10−659840500.03
cellobiose dehydrogenaseA9XK88XP_003663382.189AA8—CBM12.5 × 10−718469800.02
lytic polysaccharide monooxygenase GG2Q4M0XP_003659754.132AA9—CBM11.7 × 10−67658750001.93
lytic polysaccharide monooxygenase BG2QCJ3XP_003663414.132AA9—CBM11.4 × 10−74467610001.36
lytic polysaccharide monooxygenase HG2Q9T3XP_003661787.135AA9—CBM14.3 × 10−71251880000.74
lytic polysaccharide monooxygenase JG2Q7A5XP_003661261.126AA91.4 × 10−7462028000.18
lytic polysaccharide monooxygenase FG2QK49XP_003665200.124AA92.7 × 10−6930435000.09
lytic polysaccharide monooxygenase IG2Q9F7XP_003661661.131AA9—CBM12 × 10−714453100.01
lytic polysaccharide monooxygenase AG2QI82XP_003665516.124AA93.8 × 10−721533300.01
pyrroloquinoline quinone-dependent oxidoreductaseG2QES6XP_003664200.154AA121 × 10−1715421000.02
lytic polysaccharide monooxygenaseG2QH80XP_003663985.120AA132.5 × 10−6510548000.03
a Hypothetical molecular weight of the proteins, b BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, c the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 2. LC-MS/MS secretome analysis for carbohydrate esterases (CEs).
Table 2. LC-MS/MS secretome analysis for carbohydrate esterases (CEs).
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy
Domain
BLAST
E-Value b
iBAQ ciBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
acetylxylan esteraseG2QD29XP_003663492.134CE11.5 × 10−2356475000.17
lipaseG2QGB0XP_003664615.126CE34.5 × 10−541351800.01
acetylxylan esteraseG2QJ94XP_003665705.132CE5—CBM11.3 × 10−4120335000.06
pectinesteraseG2QLD0XP_003666007.136CE81.1 × 10−75402240001.19
pectinesteraseG2QMM2XP_003666447.135CE87 × 10−797600700.02
rhamnogalacturonan acetylesteraseG2QMH3XP_003666398.128CE126.2 × 10−4511789000.03
acetylesteraseG2QJ27XP_003664847.132CE166.1 × 10−101374350001.10
a Hypothetical molecular weight of the proteins, b BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, c the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 3. LC-MS/MS secretome analysis for polysaccharide lyases (PLs).
Table 3. LC-MS/MS secretome analysis for polysaccharide lyases (PLs).
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy DomainBLAST
E-Value b
iBAQ ciBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
pectate lyaseG2QH79XP_003663984.134PL18.3 × 10−8267613000019.95
pectate lyaseG2Q1K5XP_003660241.135PL11.3 × 10−94100960000.30
pectate lyaseG2QMM3XP_003666448.133PL11.9 × 10−4514545000.04
pectate lyaseG2QG50XP_003663760.140PL16.5 × 10−863464200.01
pectate lyaseG2QG74XP_003664579.126PL32.3 × 10−8229398000.09
rhamnogalacturonase BG2QFG7XP_003664441.158PL44.4 × 10−21165253000.19
a Hypothetical molecular weight of the proteins, b BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, c the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 4. LC-MS/MS secretome analysis for glycoside hydrolases (GHs) that breakdown cellulose.
Table 4. LC-MS/MS secretome analysis for glycoside hydrolases (GHs) that breakdown cellulose.
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy DomainBLAST E-Value biBAQ ciBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
cellobiohydrolaseG2QA39XP_003661032.151GH6—CBM18.2 × 10−971671000004.93
cellobiohydrolaseG2QFW6XP_003664525.142GH62.7 × 10−9154198000.16
cellobiohydrolaseG2Q665XP_003660789.156GH7—CBM15.5 × 10−198106610000031.43
glycoside hydrolaseG2QNN8XP_003666507.149GH72 × 10−17546477000013.71
endoglucanaseG2QCS4XP_003663441.149GH7—CBM11 × 10−1381422700004.20
endoglucanaseG2QGA1XP_003664606.149GH73.2 × 10−15391354000.27
glycoside hydrolaseG2Q359XP_003658767.149GH71.1 × 10−1901379800.01
glucan 1,4-α-glucosidaseG2QPS0XP_003666828.167GH15—CBM202.4 × 10−7421910000.06
glycoside hydrolase G2QAE3XP_003661084.1103GH311.2 × 10−148280720.01
endoglucanaseG2Q0Y0XP_003659323.124GH451.1 × 10−9263933000.19
a Hypothetical molecular weight of the proteins, b BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, c the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 5. LC-MS/MS secretome analysis for glycoside hydrolases (GHs) that breakdown hemicellulose.
Table 5. LC-MS/MS secretome analysis for glycoside hydrolases (GHs) that breakdown hemicellulose.
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy
Domain
BLAST
E-Value b
iBAQ ciBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
β-galactosidaseG2QGS8XP_003664680.196GH25.7 × 10−12010178000.03
β-glucosidaseG2QDN2XP_003663588.178GH37.1 × 10−6116280000.05
β-glucosidaseG2QCQ3XP_003663420.195GH32.8 × 10−618803400.03
xylan 1,4-β-xylosidaseG2QKP9XP_003665776.190GH32.4 × 10−593659200.01
xylanaseG2QJ91XP_003665702.145GH10—CBM11.1 × 10−99713710002.11
xylanase G2QGN6XP_003663843.142GH101.3 × 10−9119892000.06
xylanase G2QG07XP_003664565.136GH102.7 × 10−1065286100.02
xylanaseG2Q4M3XP_003659757.124GH112 × 10−53876380002.59
glycoside hydrolaseG2QLD1XP_003666008.131GH162.1 × 10−5384434000.25
glycoside hydrolase G2QHP5XP_003664150.141GH162.5 × 10−7120316000.06
glycoside hydrolaseG2Q2L1XP_003658675.144GH162.9 × 10−71373640.01
mannan endo-1,4-β-mannosidase AG2Q4H7XP_003658915.153GH26—CBM352.3 × 10−3589563000.26
α-galactosidaseG2QNU8XP_003667369.145GH271.1 × 10−5635283000.10
α-L-arabinofuranosidase 1G2QFK1XP_003663668.135GH431.4 × 10−97234540000.69
arabinanaseG2QCC8XP_003662548.161GH432 × 10−132184530000.54
arabinan endo-1,5-α-L-arabinosidaseG2QFK0XP_003663667.135GH431.5 × 10−12545693000.13
arabinanaseG2QDD9XP_003663549.149GH43—CBM351.9 × 10−8442574000.13
arabinanaseG2QQ09XP_003666917.155GH431.9 × 10−11524701000.07
arabinanaseG2QHQ9XP_003664164.139GH435.4 × 10−10014949000.04
α-1,2-mannosidaseG2QHL4XP_003664119.158GH471.5 × 10−1364463100.01
exo-β-1,3-glucanaseG2QCT8XP_003663454.184GH55025343000.07
exo-β-1,3-glucanaseG2PZK7XP_003660938.195GH551.6 × 10−25910302000.03
exo-β-1,3-glucanaseG2QF48XP_003664322.182GH556.9 × 10−2922412000.01
exo-β-1,3-glucanaseG2QIM4XP_003665591.182GH554.7 × 10−217275660.01
α-L-arabinofuranosidaseG2QJQ6XP_003665058.135GH622.1 × 10−13163013000.19
α-L-arabinofuranosidaseG2QLV4XP_003666179.140GH62—CBM11.7 × 10−1252952600.01
1,3-β-glucanosyltransferaseG2QN92XP_003667210.152GH729.3 × 10−1227456800.02
1,3-β-glucanosyltransferaseG2QAD1XP_003661926.157GH721.8 × 10−1286038900.02
xyloglucanase G2QHR7XP_003664172.179GH741.2 × 10−221830600005.39
α-mannanaseG2QL30XP_003665907.146GH761.6 × 10−96396720.01
glycoside hydrolaseG2QJT2XP_003665083.151GH797.1 × 10−6414110000.04
α-mannosidaseG2Q1N1XP_003660267.190GH923.4 × 10−1601355700.01
exo-α-L-1,5-arabinanaseG2Q5Q6XP_003660737.142GH932.8 × 10−108780890002.30
glycoside hydrolaseG2QGJ6XP_003664651.155GH1253.4 × 10−1451065800.01
glycoside hydrolaseG2QNK3XP_003667321.137GH131—CBM11.2 × 10−111256490000.76
uncharacterized protein MYCTH_2295704 G2Q5V8XP_003659079.132GH1313.6 × 10−7634392000.10
uncharacterized protein MYCTH_2301831G2QAC0XP_003661915.133GH1351.4 × 10−4334263000.10
a Hypothetical molecular weight of the proteins, b BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, c the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 6. LC-MS/MS secretome analysis for glycoside hydrolases (GHs) that breakdown other components.
Table 6. LC-MS/MS secretome analysis for glycoside hydrolases (GHs) that breakdown other components.
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy DomainBLAST
E-Value b
iBAQ ciBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
trehalaseG2PZS2XP_003658392.178GH371.4 × 10−1591548600.01
α-L-fucosidaseG2QDI5XP_003662742.191GH957.4 × 10−20918731000.04
a Hypothetical molecular weight of the proteins, b BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, c the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 7. LC-MS/MS secretome analysis for carbohydrate binding modules (CBMs).
Table 7. LC-MS/MS secretome analysis for carbohydrate binding modules (CBMs).
MS/MS View:
Identified Proteins
Accession NumberCAZy
Domain
BLAST
E-Value a
iBAQ biBAQ/Total iBAQ (%)
UniProt/Swiss-Prot DatabaseNCBI Database
cellobiose dehydrogenaseG2QGP4XP_003663851.1AA3—CBM12.2 × 10−762311400.73
cellobiose dehydrogenaseA9XK88XP_003663382.1AA8—CBM12.5 × 10−718469800.02
lytic polysaccharide monooxygenase GG2Q4M0XP_003659754.1AA9—CBM11.7 × 10−67658750001.93
lytic polysaccharide monooxygenase BG2QCJ3XP_003663414.1AA9—CBM11.4 × 10−74467610001.36
lytic polysaccharide monooxygenase HG2Q9T3XP_003661787.1AA9—CBM14.3 × 10−71251880000.74
lytic polysaccharide monooxygenase IG2Q9F7XP_003661661.1AA9—CBM12 × 10−714453100.01
acetylxylan esteraseG2QJ94XP_003665705.1CE5—CBM11.3 × 10−4120335000.06
cellobiohydrolaseG2QA39XP_003661032.1GH6—CBM18.2 × 10−971671000004.93
cellobiohydrolaseG2Q665XP_003660789.1GH7—CBM15.5 × 10−198106610000031.43
endoglucanaseG2QCS4XP_003663441.1GH7—CBM11 × 10−1381422700004.20
glucan 1,4-α-glucosidaseG2QPS0XP_003666828.1GH15—CBM202.4 × 10−7421910000.06
xylanaseG2QJ91XP_003665702.1GH10—CBM11.1 × 10−99713710002.11
mannan endo-1,4-β-mannosidase AG2Q4H7XP_003658915.1GH26—CBM352.3 × 10−3589563000.26
arabinanaseG2QDD9XP_003663549.1GH43—CBM351.9 × 10−8442574000.13
α-L-arabinofuranosidaseG2QLV4XP_003666179.1GH62—CBM11.7 × 10−1252952600.01
glycoside hydrolaseG2QNK3XP_003667321.1GH131—CBM11.2 × 10−111256490000.76
a BLAST E-value is the number of expected hits of similar quality (score) that could be found just by chance, b the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
Table 8. Proteins comprising 92.19% of the identified secreted proteins and classified according to their relative abundance (IBAQ/Total IBAQ) and with which part of the biomass they degrade: cellulose, hemicellulose, or pectin.
Table 8. Proteins comprising 92.19% of the identified secreted proteins and classified according to their relative abundance (IBAQ/Total IBAQ) and with which part of the biomass they degrade: cellulose, hemicellulose, or pectin.
MS/MS View:
Identified Proteins
Accession NumberMolecular Weight (kDa) aCAZy
Domain
iBAQ/Total iBAQ (%) bDegraded Biomass
UniProt/Swiss-Prot DatabaseNCBI Database
cellobiohydrolaseG2Q665XP_003660789.156GH7—CBM131.43cellulose
polysaccharide lyaseG2QH79XP_003663984.134PL119.95pectin
glycoside hydrolaseG2QNN8XP_003666507.149GH713.71cellulose
xyloglucanaseG2QHR7XP_003664172.179GH745.39hemicellulose
cellobiohydrolaseG2QA39XP_003661032.151GH6—CBM14.93cellulose
endoglucanaseG2QCS4XP_003663441.149GH7—CBM14.20cellulose
xylanaseG2Q4M3XP_003659757.124GH112.59hemicellulose
exo-α-L-1,5-arabinanaseG2Q5Q6XP_003660737.142GH932.30hemicellulose
xylanaseG2QJ91XP_003663843.145GH10—CBM12.11hemicellulose
lytic polysaccharide monooxygenase GG2Q4M0XP_003659754.132AA9—CBM11.93cellulose
lytic polysaccharide monooxygenase BG2QCJ3XP_003663414.132AA9—CBM11.36cellulose
pectinesteraseG2QLD0XP_003666007.136CE81.19pectin
acetylesteraseG2QJ27XP_003664847.132CE161.10pectin
a Hypothetical molecular weight of the proteins, b the iBAQ corresponds to the sum of all the peptide intensities divided by the number of observable peptides of a protein.
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

Contato, A.G.; Borelli, T.C.; Buckeridge, M.S.; Rogers, J.; Hartson, S.; Prade, R.A.; Polizeli, M.d.L.T.d.M. Secretome Analysis of Thermothelomyces thermophilus LMBC 162 Cultivated with Tamarindus indica Seeds Reveals CAZymes for Degradation of Lignocellulosic Biomass. J. Fungi 2024, 10, 121. https://doi.org/10.3390/jof10020121

AMA Style

Contato AG, Borelli TC, Buckeridge MS, Rogers J, Hartson S, Prade RA, Polizeli MdLTdM. Secretome Analysis of Thermothelomyces thermophilus LMBC 162 Cultivated with Tamarindus indica Seeds Reveals CAZymes for Degradation of Lignocellulosic Biomass. Journal of Fungi. 2024; 10(2):121. https://doi.org/10.3390/jof10020121

Chicago/Turabian Style

Contato, Alex Graça, Tiago Cabral Borelli, Marcos Silveira Buckeridge, Janet Rogers, Steven Hartson, Rolf Alexander Prade, and Maria de Lourdes Teixeira de Moraes Polizeli. 2024. "Secretome Analysis of Thermothelomyces thermophilus LMBC 162 Cultivated with Tamarindus indica Seeds Reveals CAZymes for Degradation of Lignocellulosic Biomass" Journal of Fungi 10, no. 2: 121. https://doi.org/10.3390/jof10020121

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