Open access peer-reviewed chapter

Trichoderma spp.: Approach for Bio-Control Agent

Written By

Lovely Bharti, Kajol Yadav and Ashok Kumar Chaubey

Submitted: 29 September 2023 Reviewed: 04 October 2023 Published: 17 July 2024

DOI: 10.5772/intechopen.1003697

From the Edited Volume

Challenges in Plant Disease Detection and Recent Advancements

Amar Bahadur

Chapter metrics overview

View Full Metrics

Abstract

The novel technologies in all areas of agriculture have improved agricultural production, but some modern practices cause environmental pollution and human hazards. The recent challenge faced by advanced farming has been to achieve higher yields. Thus, there is an immediate need to find eco-friendly solutions. Among the various types of species being used as biocontrol agents, fungi of the genus Trichodermaare a very large group of microorganisms widely used as biocontrol agents against different kinds of plant pathogens. Trichoderma spp. are asexual, free-living organisms that are abundantly present in all types of agricultural soils. Recent studies have shown that Trichoderma can not only prevent diseases but also promote plant growth, improve nutrient utilization efficiency, enhance plant resistance, and improve the agrochemical pollution environment. Trichoderma spp. behaves as a low-cost, effective, and eco-friendly biocontrol agent for different crop species. This chapter provides information on Trichoderma as a biocontrol agent, its biocontrol activity, and plant disease management programs.

Keywords

  • Trichoderma
  • biological control
  • plant growth
  • agriculture
  • plant pathogens

1. Introduction

Plant diseases caused by viruses, bacteria, fungi, or other macroorganisms have a significant impact on the reduction in food production, growth, and development in agriculture, which results in serious economic losses every year. The first option that farmers frequently use to control plant diseases is chemical pesticides. The primary benefit of these pesticides is their immediate use and “solution” to the issue. Chemical pesticides used over an extended period of time can contaminate soil and water, affecting both humans and animals, while occasionally leaving toxic residues that can encourage the growth of particular resistant species. In addition, pesticides have negative impacts on soil microbiomes, soil-dwelling species (such as beneficial insects, like pollinators), and the overall health of terrestrial and aquatic ecosystems [1]. In recent decades, Chemical pesticides, which are the most widely used technique for defending plants against fungal diseases, have placed serious pressure on the agricultural environment [2].

In order to avoid the negative effects of chemical pesticides, researchers are searching for alternate solutions to these issues, such as the use of biocontrol agents (BCAs) for disease control, minimizing the damage caused by plant pathogens, and an integrated approach with other chemicals that is environmentally safe. One of these methods is the use of BCAs, which are based on living microorganisms or their metabolites and products of natural origin that reduce the population of plant pathogens [3]. These biological disease management strategies are effective, long-lasting, eco-friendly, affordable, and safe for human health. At present, integrated pest management strategies and avoidance or regulation of pesticides by using Aspergillus, Gliocladium, Trichoderma, Ampelomyces, Candida, and Coniothyrium are some examples of fungal agents, and other bacterial agents include Pseudomas, Bacillus, and Agrobacterinum [4]. Among these BCAs, Trichoderma spp. is one of the most adaptable and has been used for a long time to control plant pathogenic fungus and lessen the usage of pesticides on economically significant crops. T. harzianum and T. virens prevented the pathogenic fungus from the growth of Ganoderma [5].

Fungi that act as biocontrol agents have become useful tools in the modern agricultural system. They have the capacity to ameliorate abiotic stresses like drought, salinity, extremely high or low temperatures, and heavy metal impacts, as well as the harmful effects of plant pathogens. Trichoderma spp., which are beneficial fungi, have received a lot of attention because of their high reproductive capacity, ability to survive in adverse conditions, prolific production of secondary metabolites, and resistance to plant pathogenic fungi [6]. Additionally, they have been used in biotechnological applications and are crucial to agricultural endeavors because of their capacity to reduce biotic (species and vegetable variety and edaphic microbial interactions) and abiotic stresses (type of soil, water potential, temperature, and pH) and improve plant growth and yield [7]. Colonization with Trichoderma and plant root is the initial stage of successful interaction, signal exchange, and elicitor generation that results in symbiotic relationship between them. Produced enzymes and antibiotics move toward fungal pathogens, successfully breaking down their hyphae to allow entry into the host cell. Both enzymes and antibiotics have antifungal properties and work together to combat fungi [8]. Trichoderma are known to be resistant to Alternaria alternate, Botrytis cinerea, Rhizoctonia solani, Sclerotinia sclerotiorum, Pythium spp., and Fusarium spp. [9, 10], as well as nematode [11]. After successful colonization, they cause induced systemic resistance (ISR) in the plant, which results in the signaling of several hormones and eventually promotes development [8]. Abiotic factors like salt, drought, heavy metal buildup, and severe temperatures have an impact on crop yield all around the world. However, recent studies have shown that Trichoderma induces tolerance against abiotic stresses and improves growth in plants [12] like radish, cucumber, pepper, bottle gourd, periwinkle, bitter gourd, chrysanthemum, lettuce, and tomato [13]. Trichoderma-colonized plants produce substances like auxins (indole-3-acetic acid: IAA), ethylene (ET), gibberellins (GA), plant enzymes, antioxidants, phytoalexins, and phenols that offer long-lasting resistance to abiotic stresses and improve the root system’s ability to branch. The main advantage of priming the plant for certain stress responses is that it provides a faster and stronger response if the stress occurs again.

The purpose of this chapter is to describe the role of Trichoderma species in detail as a biocontrol agent and biofertilizer and provide a brief overview of the commercially available products of Trichoderma species for application.

Advertisement

2. An overview of the genus Trichoderma

Trichoderma is a genus of filamentous, facultative, primarily asexual (the teleomorphic forms are Hypocrea), and anaerobic fungus that are widely distributed around the world and typically colonize decaying wood and other types of organic plant matter. Trichoderma is an essential component in the mycobiome of a wide range of soil ecosystems (such as farmland, prairie, forests, salt marshes, and deserts), including temperate and tropical areas, Antarctica, and tundra [2]. Trichoderma genus is a group of saprotrophic fungi that are widely distributed and frequently inhabit woody plants as endophytes [14]. The systematics and taxonomy of these fungi have changed when Persoon first coined the word “Trichoderma” in 1794 [15]. According to the present genus, it is a member of the eukaryota domain, fungi kingdom, ascomycota division, pezizomycotina subdivision, sordariomycetes class, hypocreales order, and hypocreaceae family. Over 370 Trichoderma species have previously been identified in the genus Hypocrea/Trichoderma, such as T. harzianum, T. viride, T. asperellum, T. hamatum, T. atroviride, T. koningii, T. longibrachiatum, and T. aureoviride [16, 17]. Trichoderma species are effective biocontrol agents in soil ecosystems because of their capacity for rapid growth, ability to use a variety of substrates, and resistance to numerous toxic chemicals, including fungicides (such as azoxystrobin, 3,4-dichloroaniline, and trifloxystrobin), herbicides, and other organic pollutants [18, 19].

Trichoderma species have the potential to act as biological plant protection agents. This was initially described in the early 1930s. According to the findings of researcher Weindling [20], the T. lignorum strain uses necrotrophic mycoparasitism to defend citrus seedlings from the pathogen Rhizoctonia solani. Since then, Trichoderma’s biocontrol properties have been thoroughly studied for the treatment of diseases caused by several types of soil phytopathogens [21]. The mechanisms by which Trichoderma decreases the occurrence of plant diseases include competition for nutrients and space, the synthesis of antifungal metabolites, mycoparasitism, production of lytic enzymes that degrade the cell walls of fungal plant pathogens, as well as the induction of plant resistance [2]. According to Di Marco et al. [22], the T. harzianum, T. hamatum, T. longibrachiatum, T. koningii, T. viride, T. polysporum, and T. asperellum have the most effective biocontrol capabilities. Numerous studies have shown that the majority of Trichoderma spp. can create bioactive compounds and have antagonistic effects on plant pathogenic fungi [23]. These bioactive compounds, such as cell wall-degrading enzymes and secondary metabolites, can effectively increase crop resistance, lower plant illnesses, and stimulate growth of plants [24].

Trichoderma has a wide range of applications in agriculture due to its biocontrol, biostimulation, and biofertilization abilities. Trichoderma is recognized as the genus with the highest potential for biocontrol as it has the most isolated antifungal bioactive chemicals [25]. According to Rush et al. [25], Trichoderma species account for more than 60% of the fungal BCAs. For example, approximately 250 Trichoderma-derived biofungicides are used in India, but compared to biological management, Indian farmers still rely more heavily on synthetic chemical fungicides [26].

2.1 Trichoderma biology

The mycoflora of the genus Trichoderma, which are typically global in distribution and exhibit a wide range of genetic variety, are generally found in areas with access to decomposing plant materials, primarily cellulosic materials [27]. Trichoderma species are considered to be an imperfect fungus that belongs to the order Hypocreales of the Ascomycota. They can be easily separated from natural soil, decaying plant organic matter, and wood. Trichoderma multiplies and grows very quickly in different nutrient sources like Malt Agar (MA), Czapek Dox Agar (CDA), and Potato Dextrose Agar (PDA). It also produces conidia/spores of various shades that are characterized by a green color, especially at the center of a growing spot or in concentric ring-like zones on the agar surface (Figure 1), and some species also produce thick-walled chlamydospores. The most significant characteristic of this genus is its ability to parasitize other pathogenic mycoflora, particularly those that are associated with root rot and wilt diseases [29]. Trichoderma species have been described as endophytic fungi, even though they are typically found as opportunistic plant symbionts in all types of soil, including agricultural soil, orchard soil, and forest soil [30].

Figure 1.

Three different isolated strains of Trichoderma spp. [28].

2.2 Morphological characteristics

Trichoderma species are mostly identified based on their morphology, although this method is not precise enough to distinguish between species’ differences in diversity. According to [31], most Trichoderma cultures develop quickly in the ideal range of 25–30°C but not at 35°C, yet some species grow well at 35°C. This plays a key role in differentiating species with similar morphology. T. harzianum can be distinguished from species like T. atroviride and T. aggressivum that have a similar morphology by growing it at 35°C. After 96 hours, T. harzianum develops well and sporulates at 35°C, whereas T. aggressivum and T. atroviride cannot have colonies with a radius higher than 5 mm after 96 hours. Mycelia development and color features can be more easily identified in a rich media like Potato Dextrose Agar (PDA). Trichoderma uses a number of compounds, including carbon and nitrogen sources, for its sporulation. Sporulates of Trichoderma also produce powder masses with branched conidiophore bearing bright green, or occasionally colorless, grayish, or brownish, conidia [32], which is characteristic of genera such as Myrothecium, Clonostachys, and Aspergillus as well as Penicillium, respectively. Most of the time, their surfaces are smooth, but other species, such T. viride, have rough conidia [31]. The conidiophore is poorly defined, but it is mostly branched and contains unicellular conidia and phialides at the tip of the branched hyphal system that are not visible on one-week-old media. Phialides of Trichoderma spp. are often sterile hyphae creeping septate, forming a flat, firm, and tuft conidiophores erect originating from a short branch, whereas conidia have double-layered walls that have an electron-dense rough outer layer (epispore) and a moderately electron-dense inner layer. Conidia color morphology varies from species to species, but it is normally green or can be gray, white, and yellow. Colonies of Trichoderma spp. can develop slowly or swiftly, depending on the species. Their aerial mycelium is frequently restricted, floccose to arachnoid, and reverse colorless to dull yellow. Various isolates have a distinct aroma that is similar to coconut. Conidiation may be variously vibrant, loosely tufted, or produce compact pustules that initially appear white but eventually turn green (rarely brown). The conidiophores of T. harzianum with phialides and phialospores are depicted in Figure 2. Shah et al. [33] claim that the light green conidia of T. viride are globose to subglobose, whereas those of T. harzianum are globose. T. pseudokoningii showed little, light green conidia.

Figure 2.

Phialides and phialospores of (a) Trichoderma harzianum and (b) Trichoderma viride (taken from the Université de Bretagne Occidentale website).

Advertisement

3. Biocontrol mechanisms of Trichoderma spp. against phytopathogens

Plant diseases are caused by the interaction among various components, including the host, pathogens, and environment, that is, the disease triangle. Bioagents are the organisms that prevent disease by interacting with different disease triangles. The interaction of pathogens and bioagents allows for modification of the soil environment to establish favorable conditions for effective biocontrol techniques against plant disease. Biocontrol agents involve a variety of processes in achieving disease control. However, conclusive evidences for the involvement of a specific factor in biological control are identified by the strong correlation between the factor’s appearance and biological control. The study of various strategies for managing phytopathogens and plant diseases is the most significant and fascinating aspect of Trichoderma. Trichoderma uses competition, parasitism, antibiosis, induction of defense responses in host plants, and/or a combination of these strategies to control a variety of organisms.

3.1 Mycoparasitism of Trichoderma

A complex mechanism known as mycoparasitism or hyperparasitism allows an antagonistic fungus (mycoparasite) to parasitize on another fungus (host) and ultimately result in the death of pathogen cells [2]. The majority of the Trichoderma fungus is classified as a necrotrophic mycoparasite. Trichoderma parasitize a variety of mycoparasites, especially soilborne pathogens. Over 75 Trichoderma species have been identified with high mycoparasite potential and have the ability to lyse and attack plant pathogenic fungi like Alternaria alternata, Botrytis cinerea, Rhizoctonia solani, Sclerotinia sclerotiorum, Pythium spp., and Fusarium spp. [10]. Trichoderma necrotrophs have a mycoparasitic effect on fungi, which involves chemotaxis and prey sensing, attachment to the host, and physical attack through vigorous branching and coiling around the hyphae of the host. Moreover, Trichoderma can also create penetration structures that are similar to pathogen appressoria, or appressoria-like structures [34]. The act of mycoparasitism consists of three fundamental phases. This role can be performed in the rhizosphere of plants, an ecosystem where Trichoderma has successfully colonized and where the biological control of potential pathogens is crucial to prevent plant diseases. First, Trichoderma identifies its host (or potential plant pathogenic fungus), in which the formation of oligochitins has been considered as a sensor molecule [35]. Similar to this, it is understood that during the preceding stage, a number of gene-encoding proteases and oligopeptide transporters are expressed prior to interaction with the fungal host. Second, hydrophobin-like proteins might be useful when Trichoderma comes into contact with the plant pathogenic fungus, which results in the production of papillae or appressoria-like structures. The third step takes place when Trichoderma coils around the pathogen hyphae and begins to degrade it by producing cell wall-degrading enzymes (CWDE), such as cellulases and hemicellulases, chitinases, proteases, xylanase, pectinase, lipase, amylase, arabinase, protease, 1,3-glucanases, and lytic enzymes among other secondary metabolites (Table 1 and Figure 3), which are essential for the mycoparasitism/biocontrol process [41]. These enzymes degrade the pathogen cell walls, which are composed of chitin and glucan polysaccharides. Production and regulation of lytic enzymes, which include the chitinolytic enzymes β-N-acetyl glucosaminidase, endochitinase, and chitobiosidase, are essential for promoting cell wall degradation, mycelial autolysis, chitin assimilation, and fungal parasitism and inhibiting spore germination of other plant pathogenic fungi, which are produced by T. harzianum, T. atroviride P. karst, and T. asperellum [42]. For plant protection, Trichoderma produces several types of volatile metabolites, such as 6-n-pentyl-2H-pyran-2-one (6-PAP) [43]. The enzymes β-1,3- and β-1,6-glucanases determine the hyperparasitic capability of Trichoderma in response to Phytophthora sp. and Pythium species. Endo- and exoproteases are proteolytic enzymes of Trichoderma that are responsible for enzyme secretion for the control of Botrytis cinerea, Rhizoctonia solani, and Fusarium culmorum. Ceratin cellulase enzymes, such as exo-β-1,4-glucanases, endo-β-1,4-glucanases, and β-glucosidases, synthesized by antagonists, also play a key role in hyperparasitism. Chemical attack and degradation of the pathogen’s cell wall by hydrolytic enzymes and antifungal compounds produced by Trichoderma is the final stage of the mycoparasitic interaction, ultimately causing host death (Figure 4) [44].

SecondaryMetabolites effectReference
PeptaibolsInhibit the activity of β-1,3 glucan synthase[36]
Gliotoxin, Gliovirin, Koninginins, Trichothecenes, and 6-Pentyl-Α-Pyrone (6-PAP)Antifungal[37]
Trichodermin, Suzukacillin, and AlamethicinAssist in destroying the cell walls of other pathogenic fungi[38]
Harzianic acid, Tricholin, Massoilactone, Viridin, Glisoprenins, and Heptelidic acidAntibiotic[37]
Jasmonic acid and TerpenesRepel insects from consuming plant leaves[39]

Table 1.

Secondary metabolites produced by Trichoderma spp.

Figure 3.

Example of SMs produced by Trichoderma spp. chemical structure of 1 6-pentyl-α-pyrone, 2 koninginin A, 3 viridin, 4 harzianopyridone, 5 harzianic acid, 6 harzianolide, 7 T39butenolide, 8 dehydro-harzianolide, 9 cerinolactone, 10 gliotoxin, 11 coprogen, 12 aspinolide C, 13 trichodiene, 14 harzianum A [40].

Figure 4.

Mycoparasitism by Trichoderma. (A) Interaction between colonies of Trichoderma and a fungal prey. (C) Main factors and structures used by Trichoderma during antagonistic interaction. (B and D) Host invasion. Transmission electron micrographs of T. Atroviride parasitizing Rhizoctonia solani. (B) T. Atroviride (T) penetrates R. solani (R). (D) T. atroviride (T) grows inside R. solani (R) hypha [45].

Numerous Trichoderma genes that code for proteases and oligopeptide transporters have been identified during interactions between different Trichoderma species and the host. Additionally, it has been proposed that pathogenic host sensing is mediated by class IV G-protein-coupled receptors (GPCRs), which operate as sensors for oligopeptides and other chemicals released from phytopathogen cell walls by the activity of protease enzymes [10]. The conserved signaling cascade of G-proteins, which consists of the Gα, Gβ, and Gγ subunits, is used for further signal transduction. Mutants of T. atroviride that lack the Gα subunit are completely incapable of mycoparasitism, have decreased chitinolytic activity, and are unable to produce the antifungal chemical 6-pentyl-α-pyrone [10]. Furthermore, during mycoparasitism in Trichoderma, mitogen-activated protein kinases (MAPKs), particularly pathogenicity MAPK (TmkA and Tmk1), are implicated in signal transduction pathways [46]. The antagonistic interactions between S. rolfsii, R. solani, and Pythium ultimatum and T. virens are reduced when TmkA is removed.

Trichoderma’s mycoparasitism extends beyond the host’s pathogenic hyphae, and these fungi may use the pathogen’s conidia as a source of further biocontrol. The Trichoderma TkZ3A0 strain’s hyphae were able to parasitize phialides with macroconidia of the F. culmorum phytopathogen (isolated from winter wheat with severe fusariosis symptoms) and exhibited clear chemotaxis and adhesion to their structures. This strain has a high degree of similarity to the T. koningiopsis species. Even in the initial phases of the interaction, macroconidia’s shape changed and their ability to germinate was inhibited (mycostasis) [47].

3.2 Competitive role of Trichoderma

Starvation is the most prevalent cause of mortality for all living organisms. Microorganisms typically die from starvation because there are not enough nutrients in their local surroundings, which include soil and plant surfaces. Trichoderma successfully established themselves in the environment, both pathogens and biocontrol agents are competing hard with one another for resources like food and space. Competition is a process that occurs between microorganisms when there is a scarcity of micro- and macronutrients, and the competition for those nutrients results in the natural management of fungal communities and the development of phytopathogens [48]. Competition for micro- and macronutrients, which are carbon, nitrogen, and iron, is essential for the interactions between beneficial and harmful fungi and is associated with biocontrol systems. Trichoderma is usually considered as a strong competitor against soilborne fungal pathogens. It exhibits higher capacity to mobilize and quickly absorb the nutrients needed for the growth of the pathogen fungi, leading to nutritional deficit and preventing the growth and reproduction of the pathogenic fungi [49]. Rapid germination rate of Trichoderma has produced a wider accessible area and a greater supply of nutrients for growth. The most effective utilization of nutrients depends upon the ability of Trichoderma spp. to obtain energy from the metabolism of carbohydrates such as cellulose, chitin, glucan, and glucose, which are frequently present in the mycelial environment [50]. Trichoderma has a strong capacity for environmental adaptation [51]. It can seize nutrients and space close to the plant rhizosphere by its rapid growth and reproduction, consume oxygen in the air, and inhibit the growth of plant pathogenic fungi [52]. Trichoderma can effectively suppress the growth of plant pathogenic fungi since its growth rate is much faster than that of plant pathogenic fungi. After 24 hours in the soil, Trichoderma can quickly attach to plant roots for reproduction, and the hyphae quickly wrap plant roots to create a protective layer that prevents pathogen invasion and removes surrounding pathogens. According to Risoli et al. [53], T. harzianum grew 2.0–4.2 times more quickly than B. cinerea. Trichoderma mycelium competed with Fusarium graminearum by adhering, twining, interpenetration, and other mechanisms, causing Fusarium graminearum’s mycelium to become distorted and subsequently disappear [54]. Through rapid growth and reproduction, Trichoderma can weaken and exclude the gray mold pathogen in the same habitat by capturing water and nutrients, occupying space, using oxygen, and other things [55]. Trichoderma accomplishes this by biosynthesizing and releasing organic acids including gluconic, citric, and fumaric acids to cause a decrease in soil pH. Additionally, these organic acids help to dissolve micronutrients and mineral cations such phosphates, iron, manganese, and magnesium.

Siderophores, which are formed under iron-deficiency stress and have a low molecular weight (less than 10 kDa) chelator molecule with a high affinity for iron (Fe), are considered to be of the greatest significance in relation to the competitive phenomenon in Trichoderma [56]. The Fe ions serve as cofactors for numerous enzymes and are crucial for the healthy growth and development of both microbes and plants. Generally, the three classes of microbial siderophores—hydroxamate, catecholate, and carboxylate—are separated based on the chemical structure and the position of their coordination sites with iron [57]. The majority of the times, fungi produce hydroxamate-type siderophores, like coprogens, ferrichromes, and fusarinines, which have the structural component N5-acyl-N5-hydroxyornithine [58]. Iron is mostly found in soil in the form of Fe3+ under neutral pH conditions and in the presence of oxygen. Fe tends to produce insoluble iron oxides in the aerobic environment, making it unavailable to plants. Siderophores have the ability to interact with insoluble iron (Fe3+) and subsequently change it into soluble Fe2+, which is readily absorbed by plants and microbes (Figure 4).

In the absence of iron supplies from an associated niche, Trichoderma may inhibit the growth and activity of target soil pathogens. The interaction of the pathogens Fusarium acuminatum, Alternaria alternata, and Alternaria infectoria with T. harzianum was studied in vitro, and the results revealed that the studied fungi died from nutrient starvation. Further, the iron competition was suggested as one of the essential components in T. asperellum’s antagonistic relationship with F. oxysporum f. sp. lycopersici and tomato plants’ defense against wilt disease. Trichoderma has been shown to successfully compete with phytopathogens from the genera Colletotrichum sp., Botrytis sp., Verticillium sp., and Phytophthora sp. for simple and complex carbon substrates [59]. Further, the powerful inhibitory impact of Trichoderma on the pathogens B. cinerea, F. graminearum, and Macrophomina phaseolina was mediated by enzymes that create competition for nutrients and space [60].

Therefore, the coordination of many strategies, including the competition for nutrients, which is considered to be among the most essential, is required for the control management of some pathogens (such as B. cinerea) by using Trichoderma [61].

3.3 Induced systemic resistance of Trichoderma

One of the most significant mechanisms of Trichoderma’s biocontrol actions is induced systemic resistance. It is well-known that Trichoderma spp. are able to induce systemic and local resistance in plants against plant pathogens through alterations in their physiological and biochemical processes, as a result of complex interactions between microbial biomolecules and plant receptors [62]. While preventing the pathogenic fungus from proliferating and growing, it can also encourage crops to produce self-defense mechanisms for obtaining local or systemic disease resistance. This causes the synthesis of regulatory plant proteins that may contribute in the defense mechanism by recognizing the microbial biomolecules, as well as the formation of the defense-enhancing chemical salicylic acid throughout the entire plant as a systematically induced defense mechanism [63]. Trichoderma-induced plant disease resistance is achieved by two methods: first, regulate the elicitors or effectors that initiate the plant disease resistance response, and second, the cell wall-degrading enzymes (such as celluloses, chitinases, and glucanases) produced by Trichoderma release oligosaccharides that can induce plant resistance [64]. Elicitors play a key role in the initiation of Trichoderma-induced systemic resistance (TISR) in plants. Currently, there are more than 10 Trichoderma elicitors known to promote plant resistance, including Sm1, QID74 hydrophobic protein, MRSP1, chitin-degrading enzyme, xylanase, cellulase, endopolygalacturonase, sucrase, and antibacterial peptides. These substances primarily come from five Trichoderma species: T. asperellum, T. viride, T. atroviride, T. harzianum, and T. virens[65, 66, 67, 68].

The elicitors are divided into two categories: (1) race-specific elicitors that only activate gene-to-gene defense in particular host cultivars and (2) general elicitors released from pathogenic and nonpathogenic strains that activate non-race-specific defense in both host and nonhost plants. Numerous classes of elicitors have been identified, such as oligosaccharides (glucans, chitins, and oligogalacturonides), proteins and peptides (elicitins and endoxylanase), glycopeptides and glycoproteins (e.g., glycopeptide fragments of invertase), glycolipids (e.g., lipopolysaccharides), and lipophilic substances (e.g. fatty acids). In plants, the activation of signal transduction pathways by elicitors causes physical, biochemical, and molecular changes such as ion flow across the membrane, the production of reactive oxygen species (ROS), the construction of a physical barrier that inhibits the spread of phytopathogens (callose deposition and reinforcement of the plant cell wall), and the synthesis of various defense compounds (such as phytoalexins, volatile organic compounds, enzymes, and phytohormones) [69].

Numerous plant species, including monocotyledonous and dicotyledonous, have an enhanced immune response when nonpathogenic Trichoderma fungi are present. The recognition of conserved domains, such as the pathogen-associated molecular pattern (PAMP) or the microbe-associated molecular pattern (MAMP), is the main component of the plant defense response. Both MAMP-triggered immunity (MTI)/PAMP-triggered immunity (PTI) and effector-triggered immunity (ETI) are induced by these domains in plants [47].

According to Saravanakumar et al. [70], Trichoderma-coated maize seeds significantly increased in peroxidase (POD) and phenylalanine ammonia lyase (PAL) activity, and the plants were resistant to curvularia leaf spot of maize.

3.4 Antibiosis effect of Trichoderma

The biological control mechanism referred to as antibiosis involves the production and excretion of secondary metabolites, including compounds of a different chemical nature with cytotoxic activity, which can reduce or inhibit the growth of phytopathogens by secreting antagonistic substances [71]. Trichoderma species have the ability to create antimicrobial secondary metabolites, such as trichomycin, gelatinomycin, chlorotrichomycin, and antibacterial peptides, which may inhibit the development of several types of pathogens. This inhibiting process is known as antibiosis. Antibiosis is one of the primary mechanisms by which Trichoderma and other biological pest controllers, such as plant growth-promoting bacteria (PGPB), work [72, 73]. More than 180 secondary metabolites of Trichoderma have been classified into different categories of chemical compounds based on their roles in competition, iron-chelating metabolites, inducers of plant resistance, plant growth-promoting metabolites, and antibiotics [40]. These substances can be divided into peptaibols, water-soluble substances, and volatile or nonantibiotic substances. These substances have a variety of roles in the biological control of various pathogens and are also able to function as antibacterial agents, encourage plant growth, and serve as valuable building blocks for the creation of agricultural antibiotics [74]. Some secondary metabolites have the ability to modify the metabolism and growth of plants. Some Trichoderma strains, including T. viride, T. harzianum, and T. koningii, have been shown to be able to produce and secrete 6-PAP, a volatile metabolite that can inhibit the growth of colonies to varying degrees, and some of them can inhibit the growth of colonies by more than 80% [75, 76, 77]. These volatile metabolites are crucial to the biocontrol of a variety of pathogenic species, including Fusarium oxysporum, Botrytis cinerea, and Rhizoctonia solani [78]. For example, T. virens species produce trichodermamides, whereas T. koningii produce Koninginins, both of which have antibacterial and antifungal activity. Additionally, in T. harzianum and T. virens substances like azaphilones, viridins, nitrogen heterocyclic compounds (such harzianopyridone and harzianic acid), and volatile terpenes have been characterized and are used in the biocontrol of pathogenic fungus [79]. Different Trichoderma spp. have also been studied for their ability to produce hydrolytic enzymes and proteases, including exo- and endochitinases, chitinases, xylanases, glucanases, lipases, and endo- and exopeptidases, among others, with antifungal activity [80]. The most prevalent volatile organic compound (VOC) from T. atroviride is 6-pentyl-2H-pyran-2-one (6-PP), which, together with other VOCs produced by the fungus, promotes plant growth and controls sugar transport in Arabidopsis roots [81].

Advertisement

4. Trichoderma as biocontrol agent against soilborne pathogen

The rapid use of chemical fertilizers considerably contributes to the current condition of environmental deterioration through the use of fossil fuels, the production and release of carbon dioxide, and the contamination of water supplies. Globally, environmental deterioration is currently a major concern. The most effective way to prevent environmental deterioration is by using biological agents [82]. Biocontrol is the use of biological populations to control the proliferation of pests. A few of the biocontrol mechanisms by which BCAs can prevent the growth of soilborne pathogens include the capability to grow faster than soilborne pathogens for nutrients and space, the production of numerous potent plant-degrading enzymes like lytic enzymes and proteolytic enzymes, and the production of more than 200 antibiotics that are extremely toxic to all macro- and microorganisms. It is believed that the ability to produce various antibiotics will enhance biological control by preventing a variety of microbial competitors, some of which are likely plant diseases [28].

Antibiotics such as 2,4-diacetylphloroglucinol, produced by Pseudomonas fluorescens F113 against Pythium species, which causes damping-off disease, and gliotoxin, produced by Trichoderma virens against Rhizoctonia solani, which causes plant root rot, have been reported to be involved in the suppression of plant pathogens. Furthermore, BCA enzymes can also hydrolyze chitin, proteins, cellulose, and hemicellulose. As a result, plant pathogens are quickly inhibited. Several types of BCAs can produce enzymes that are beneficial against particular plant diseases [83].

Trichoderma has a significant application value and potential for biological management of plant diseases. The use of Trichoderma to control plant diseases has been studied all around the world. Trichoderma spp. are effective biocontrol agents that are widely used against soilborne diseases. Some of the important gene types that are essential to the function of biocontrol include protease, chitinase, glucanase, tubulins, cell adhesion proteins, and stress-tolerant genes. These genes play a role in the degradation of cell walls, hyphal growth, stress tolerance, and parasitic activity. For example, chitinase breaks down glycosidic linkages, while xylanase breaks down hemicellulose and so forth. Trichoderma fights against other plant pathogenic fungi and promotes the growth of the plant and its roots. For plant pathogenic diseases, it uses a variety of management strategies, such as antibiosis, mycoparasitism, induced host cell resistance, and competition for nutrients and space. The antagonist’s proliferative potential in the soil is a key factor in the efficient biocontrol of soilborne pathogens. Trichoderma species are recognized as prospective biological control agents due to their capacity to either increase resistance or restrict the growth of a number of phytopathogenic fungus. Trichoderma’s induction of systemic resistance is another important method of pathogen resistance. Through mycoparasitism, competition, or the production of antibiotics, plants can defend themselves directly [84].

As a biocontrol agent, T. viride and T. harzianum are known to have various degrees of inhibitory effects on pathogens including Fusarium oxysporum, Rhizoctonia solani, Pythium aphanidermatum, Fusarium culmorum, Gaeumannomyces graminis var. tritici, Sclerotium rolfsii, Phytophthora cactorum, Botrytis cinerea, Pseudocercospora spp., Colletotrichum spp., and Alternaria species [85, 86, 87, 88, 89, 90]. Trichomonas has been widely used in the biological management of tobacco root rot, tomato, potato, rice, wheat, and other plant diseases (Table 2).

DiseaseCropPathogenBiocontrol strainReferences
Root rot diseaseSoybean (Glycine max (L.) Merr. cv)Pythium arrhenomanes f. sp. adzukiT. viride[91]
Corn (Zea mays)Fusarium oxysporum f. sp. adzuki
Cocoyam (Xanthosoma sagittifolium)Pythium myriotylumT. asperellum[92]
Pepper plants (Capsicum annuum)Rhizoctonia solaniT. harzianum[93]
Eggplant (Solanum melongena L.)Macrophomina phaseolinaT. harzianum, T. polysporum, T. viride[94]
Damping-offPepper (Capsicum annuum)Phytophthora capsiciT. harzianum[95]
Cucumber (Cucumis sativus)Pythium sp.T. harzianum[96]
Cotton (Gossyphtm hirsutum)Rhizoctonia solaniT. harzianum[97]
Cotton (Gossyphtm hirsutum)Pythium aphanidermatum, Pythium ultimum, Rhizopus oryzaeT. virens[98]
WiltTomato (Solanum lycopersicum)Fusarium oxysporum f. sp. Lycopersici (FOL)T. asperellum[99]
Melon (Cucumis melo)F. oxysporumT. harzianum T-78[100]
Fruit rotChili (Capsicum annuum)Alternaria tenuisT. harzianum[101]
Tomato (Solanum lycopersicum)Rhizoctonia solaniT. viride, T. virens,
T. harzianum
[102]
Brown spotTobacco (Nicotiana tabacum)Alternaria alternataT. harzianum[103]
Brown root rotPeanut (Arachis hypogaea)Fusarium solaniT. harzianum[104]
Anthracnose gray moldStrawberry (Fragaria ananassa)Colletotrichum acutatum, Botrytis cinereaT. hamatum, T. atroviride, T. longibrachiatum[105]
Head blightWheat and other small grain cereals (Triticum aestivum)Fusarium graminearum, Fusarium culmorumT. gamsii[106]
Sheath blightRice (Oryza sativa)Rhizoctonia solaniT. harzianum[107]
Blossom blightAlfalfa (Medicago sativa)Sclerotinia sclerotiorumT. atroviride[108]
Web blightBean (Phaseolus vulgaris)Sclerotinia sclerotiorumT. viride[102]
Collar rotTomato (Solanum lycopersicum)Sclerotium rolfsiiT. virens, T. harzianum[102]

Table 2.

Various diseases controlled by Trichoderma spp.

Advertisement

5. Trichoderma as biofertilizers

Severe environmental conditions are one of the factors that have decreased crop growth and productivity. Crop productivity would suffer as a result of abiotic stresses caused by climate changes, such as temperature rise, increased carbon dioxide levels, protracted drought, and hurricanes [109]. Additionally, these circumstances have an impact on the proliferation and distribution of plant pathogens and pests, adding to the biotic stresses imposed on crops [110]. Genetically uniform improved crop are frequently vulnerable to invasion caused by non-native pests and pathogens. Irrigation practices in dry environments affect soil nutrients and increase salinization. Salinization has been associated to decreased soil microbial activity and also has an impact on physical properties of soil, such as producing compaction. Compacted soil with low oxygen content affects root growth, which limits nutrient and water intake and reduces the production of plants. As a result, reports indicate that the average production of important crops is facing losses of up to 50%. In order to maintain agricultural production and productivity under various environmental challenges, Trichoderma spp. application as biofertilizer offers a sustainable and environmentally friendly alternative. In addition, they also play an essential role as a biocontrol agent, protecting plants from a variety of disease attacks. Trichoderma spp. interacts with plants to promote root growth, assist in nutrient uptake, and enhance the plant’s resistance to abiotic stresses, all of which increase the production of agricultural products [47]. According to previous studies, Trichoderma was able to colonize roots and affect the transcriptome, proteome, and epigenome of plants, improving their capacity to tolerate external stresses. In the presence and absence of abiotic stress factors, Trichoderma spp. treatment has positive effects on plants (Table 3).

Strain as
biofertilizer
CropsApplication modeBeneficial outcome
T. azevedoiLettuceSimple exposureIncreases carotenoids and chlorophyll with reduction in the white mold attack to about 78.83%
T. afroharzianumTomatoSeed inoculation or treatmentHelps in the secretion of phytohormones like homeostasis, antioxidant activity, phenylpropanoid biosynthesis, and glutathione metabolism
T. harzianum,Chinese cabbageIrrigationIncreased yield by 37%; Increased enzyme activity in the soils and providing more inorganic nitrogen and phosphorus content to the soil
T. asperellum,
T. hamatum,
T. atroviride
T. brevicompactum,TomatoSeedling drenchingImproved growth and yield due to the production of indole-3 acetic acid
T. gamsii,
T. harzianum,
T. harzianum Rifai,TomatoSeed treatmentImproves phosphorus uptake
T. asperellum,
T. brevicompactum,TomatoSeed drenchingImproves phosphorus solubilization
T. gamsii,
T. harzianum
T. erinaceumRiceSeed treatmentImproves germination, vigor, and yield
T. harzianum T22TomatoSeed treatmentImproves soil fertility, level of minerals and antioxidants, nutrient uptake, and yield
T. harzianum T22TomatoSoil amendment as compostIncrease in yield to about 12.9%
T. virideSugar canePowder as fertilizersImproves nutrient uptake
T. asperellum T34CucumberSeedling drenchingEnhanced nutrient uptake
T. harzianumAll cropsCompostEnhances residue decomposition resulting in availability of soil nutrients
T. simmonsiiBell pepperSeedling drenchingImproves yield to about 67%
T. reeseiChickpeaSeed treatmentEnhanced mineral uptake
T. harzianumMustardSoil inoculationImproved nitrogen absorption and increased yield to about 108 and 203%
T. harzianumChiliSoil inoculationIncreased yield
T. harzianumBarleySeed inoculation17% increase in yield
T. viridaeWheatSoil and seed inoculation75.8% increase in yield with improved nutrient absorption
T. viridaePotatoSoil inoculationIncreased yield with an average of 16.25 tubers/plant
T. viridaeRed beet cabbageSeed inoculation29% increase in yield
T. harzianumOnion seedlingsSeedling inoculationEnhanced growth and yield
T. asperellumMaizeSoil granulesImproves yield

Table 3.

Trichoderma spp. as biofertilizers function in increasing plant production and growth [111].

5.1 Plant growth enhancement by Trichoderma spp.

Trichoderma frequently coexists with the root ecosystems of the host plants. Therefore, Trichoderma is generally referred to as a genus of symbiotic, opportunistic, and nonpathogenic microorganisms that colonize the roots and regulate plant pathogens. Trichoderma spp. not only inhibit the growth of pathogens but also increase plant growth, serve as a biofertilizer, and stimulate plant defense mechanisms. According to Hyakumachi and Kubota [112], plant growth-promoting fungi (PGPF) are fungi that can play a role in the growth of plants. The main effects of PGPF are frequently seen in the production, quality, and growth of the crop. Recent studies have shown that Trichoderma spp. can make a great PGPF. It has been shown that some Trichoderma strains may penetrate the epidermis and create a robust and persistent colonization of root surfaces. Trichoderma species have positive impacts on plants, such as promoting growth, enhancing seed germination and viability, improving root structure and condition, and increasing photosynthetic efficiency, blooming, and yield quality [113]. Trichoderma has been shown to enhance the growth of lettuce, pepper plants, and tomatoes. The production of phytohormones and phytoregulators is the most significant stimulating factor at almost all stages of plant growth and development by creating a favorable environment and production of a large amount of secondary metabolites [47]. Trichoderma spp. produce secondary metabolites that regulate plant growth, such as koningin A (Trichoderma koningii) and 6-pentyl-alpha-pyrone (T. harzianum). Other significant activities carried out by Trichoderma spp. include increased solubilization of phosphate minerals (such as Fe, Mg, and Mn), reduction in soil pH, and formation of gluconic and citric acids as well as micronutrients.

The main advantage of Trichoderma spp. for plant growth is the development of roots. This theory is confirmed by recent studies [114] that have shown Trichoderma spp. produces or controls plant hormones like auxin, harzianic acid, and harzionalide, which are responsible for promoting root development. According to Yedidia et al. [115], it was found that plants inoculated with T. harzianum on the 28th day developed significantly greater roots. Additionally, this study found that the content of Cu, P, Fe, Zn, Mn, and Na increased in the inoculated root (Figure 5). At the same time, it was found that the concentration of Mn, Zn, and P had increased by 70%, 25%, and 30%, respectively, in the plant’s shoot. The study supports the theory made by Contreras-Cornejo et al. [116] that a high root surface area caused by Trichoderma spp. inoculation permits the root to explore a larger area of soil. This makes it possible for plants to absorb more macronutrients and micronutrients from the soil, which is advantageous for them when fighting for minerals with other organisms or when there is a lack of minerals. The identification of Trichoderma spp. metabolites could be helpful in the application of novel biofertilizers in agriculture as an alternative to synthetic/chemical fertilizers. To increase crop yield, using Trichoderma spp. as a biofertilizer in combination with fungi as inoculants may be beneficial. Additionally, it reduces pollution caused by the overuse of synthetic/chemical fertilizer in the agricultural sector (Figure 6).

Figure 5.

Transformation of the insoluble form of iron (Fe3+) into a soluble and easily assimilating form (Fe2+) by siderophores produced by Trichoderma fungi [117].

Figure 6.

Trichoderma spp. increase the nutrients uptake by root improvement [118].

5.2 Plant root colonization by Trichoderma Spp.

Many rhizosphere Trichoderma species may colonize the root surfaces of monocotyledonous and dicotyledonous plants, which may have a significant effect on the plant’s metabolism. Trichoderma species recognize the host plant and adhere and penetrate plant roots to colonize it. Trichoderma spp. colonization of plant roots enhances plant defense by causing the production of 1,3-glucanase, peroxidases, phenylalanine, chitinases, and hydroperoxidase, which activated signaling of the plant’s biosynthetic pathways and led to the accumulation of low-molecular weight phytoalexins. Since it colonized such a diverse range of host plants, Trichoderma most likely developed efficient techniques to get over plant defenses. Trichoderma harzianum and plant interacted physically in a symbiotic way where Trichoderma stimulate the increased activity of metabolites such as peroxidase and chitinase, thus protecting the plant from disease in exchange for it providing a nutritional niche for the fungus. Yedidia et al. [119] used an electron microscope to investigate the physical interaction between T. harzianum T-203 and a cucumber plant and discovered that the fungus pierced the root and spread across the epidermis and outer cortex, which prompted the production of more peroxidase and chitinase. Consequently, the interaction seems to be a symbiotic one in which Trichoderma lived in the nutritional niche offered by the plant and the plant was protected from disease.

Various studies have confirmed Trichoderma’s capacity to colonize the root surface in addition to the rhizosphere soil. Previously, the root hair and elongation zone was the main area where Trichoderma colonized the roots. However, recent research has shown that these fungi have colonized the root cap border cells (RBCs) zone. RBCs play a vital role in the interaction between plants and soil microbes. During root extension, the RBCs are discharged into the soil environment from the cap cells’ outer layer. These cells are thought to be a valuable source of nutrients and biologically active substances for microorganisms, along with the related root exudates. Additionally, RBCs have an impact on the structure and content of the microbial rhizosphere population and serve as attractants for microbes, promoting interaction with plants and root colonization. RBCs are also considered to be an important factor in protecting plants from a variety of biotic and abiotic stresses [120]. The ability of Trichoderma to colonize RBCs, as shown for the wheat roots injected with DEMTkZ3A0 strain conidia, suggests a major role of these fungi in the indirect defense mechanism against phytopathogens through the creation of a mantle-like structure.

Advertisement

6. Commercially available Trichoderma-based bioproducts

Trichoderma spp. are considered as one of the most extensively studied fungal agents as microbial biocontrol agents (MBCAs) in agriculture and are sold as biopesticides, biofertilizers, growth promoters, and naturally occurring resistance-inducing agents. Since these compounds are developed to protect plants from adverse environmental conditions, it is important to establish set standards for their production and maintenance because changes in pH, temperature, and other environmental factors may have an impact on the formulation’s activity. Additionally, it should not be phytotoxic, be economical, dissolve quickly in water, be easy and affordable to obtain carrier materials for, and be compatible with agrochemicals. The distribution of the different Trichoderma spp.-based products for important agricultural diseases is listed in Table 4.

Bioproducts nameSpecies strainCompany, Country
Topshield, RootshieldT. harzianum T-22Bioworks, Geneva, N.Y.
T35T. harzianumMakhteshim-Agan Chemicals, Israel
Harzian 20, Harzian 10T. harzianumNatural Plant Protection, Noguerres, France
F-stopT. harzianumEastman Kodak Co., United States TGT Inc., New York
SupraavitT. harzianumBonegaard and Reitzel, Denmark
Solsain, Hors-solsain, PlantsainTrichoderma spp.Prestabiol, Montpellier, France
ANTI-FUNGUSTrichoderma spp.Grondontsmettingen De Ceuster, Belgium
TyTrichoderma spp.Mycontrol, Israel
GlioGard and SoilGardT. virens (Gliocladium virens)Grace-Sierra Co., Maryland, USA
Bip TT. viridePoland
Promot Plus WP Promot PlusDDTrichoderma spp., T. koningii, T. harzianumTan Quy, Vietnam
TRiB1Trichoderma spp.National Institute of Plant, Vietnam
TRICÔ-DHCTTrichoderma spp.Can Tho University, Vietnam
Vi – DKTrichoderma spp.Pesticide Corp., Vietnam
NLU-TriTrichoderma virensHo Chi Minh University of Agriculture and Forestry, Vietnam
Biobus 1.00WPTrichoderma virideNam Bac, Vietnam
Bio – Humaxin Sen Vàng 6SC, Fulhumaxin 5.15SCTrichoderma spp.An Hung Tuong, Vietnam
BioSpark TrichodermaT. parceramosum, T. pseudokoningii, and Ultraviolet irradiated strain of T. harzianumBioSpark Corporation, Philippines
Biocure FT. virideT. Stanes and Company Limited, European Union, available in India
Trichoderma viride powderT. virideOrganic Dews, Bio Organic, India
Bio-Shield, BioveerT. virideAmbika Biotech, India
Mycofungicyd, TrichoderminT. viride 16, T. lignorumBizar-agro LTD, Ukraine
ICB Nutrisolo SC e WPT. viride, T. harzianum, T. koningii and Trichoderma spp.ICB BIOAGRITEC Ltd., Brazil
Trichosav-34T. harzianum A-34Institute for Research in Plant Protection (INISAV), Cuba
Trichosav-55T. harzianum A-55Institute for Research in Plant Protection (INISAV), Cuba
Antagon TVT. virideGreen Tech Agroproducts, Tamil Nadu, India
TrichostarT. harzianumGreen Tech Agroproducts, Tamil Nadu, India
GliostarT. virensGBPUAT, Pantnagar, India
BiodermaT. viride, T. harzianumBiotech International Ltd., India
Bio FitT. virideAjay Biotech (India) Ltd., India
EcofitT. virideHoechst Schering Afgro Evo Ltd., India
TrichoguardT. virideAnu Biotech Int. Ltd. Faridabad, India
BioconT. virideTocklai Experimental Station Tea Research Association, Jorhat (Assam), India

Table 4.

Description of commercially available bioactive products of Trichoderma spp. worldwide [111].

Advertisement

7. Conclusion

Currently, the primary strategy for controlling plant diseases is chemical control, which is accomplished by spraying pesticides and fungicides. Although chemical control has a positive effect and is beneficial in boosting agricultural production, it also has negative effects on people and the environment and increases pathogens’ resistance. Scientists and their studies have demonstrated that Trichoderma is nonpathogenic to plants and exhibits various biocontrol features, which makes it one of the most efficient organisms against several types of plant pathogens, and reduces the use of harmful pesticides in the agricultural sector. Trichoderma spp., which are primarily found in soil and the rhizosphere region, are antagonistic to the majority of plant pathogens. A biocontrol program is only established when the biocontrol agent is able to successfully control the relationship between the host plant and pathogen. Trichoderma has a well-established ability to control this interaction. It has also been shown that the fungi improve plant defense mechanisms. Furthermore, Trichoderma provides a variety of control strategies for various plant pathogens, which gives it an edge over other phytopathogen control strategies. Mechanisms that are usually involved are mycoparasitism, antibiotics, competition for nutrients, and stimulation of systemic resistance in plants. Another fascinating feature is that Trichoderma species produce a number of secondary metabolites that significantly affect plant growth or prevent pathogen growth along with promoting localized and systemic resistance as well as stress tolerance in plants. Currently, Trichoderma species are being used in the sustainable disease management system for managing plant diseases. These methods of managing plant diseases may be safer for the agroecosystem, overall environment, the health of the soil, and human health. They would also be the proper move in sustaining agricultural output to satisfy the demands of growing populations. As a result, farmers have the choice of using it to enhance agricultural yields and produce higher quality. Trichoderma spp. can be used for waste/organic material decomposition, contaminated area purification, and decreasing diseases and enhancing plant growth.

References

  1. 1. Alizadeh M, Vasebi Y, Safaie N. Microbial antagonists against plant pathogens in Iran: A review. Open Agriculture. 2020;5:404-440
  2. 2. Ghorbanpour M, Omidvari M, Abbaszadeh-Dahaji P, Omidvar R, Kariman K. Mechanisms underlying the protective effects of beneficial fungi against plant diseases. Biological Control. 2018;117:147-157
  3. 3. Grasswitz TR. Integrated pest management (IPM) for small-scale farms in developed economies: Challenges and opportunities. Insects. 2019;10:179
  4. 4. Gilardi G, Manker DC, Garibaddi A, Gullino ML. Efficacy of the biocontrol agents Bacillus subtilis and Ampebmyces quisqualis applied in combination with fungicides against powdery mildew of zucchini. Journal of Plant Disease and Protection. 2008;115:208-213
  5. 5. Naher L, Yusuf UK, Siddiquee S, Ferdous J, Rahman MA. Effect of media on growth and antagonistic activity of selected Trichoderma strains against Ganoderma. African Journal of Microbiology Research. 2012;6:7449-7453
  6. 6. Contreras-Cornejo HA, Macías-Rodríguez L, del-Val E, Larsen J. Ecological functions of Trichoderma spp. and their secondary metabolites in the rhizosphere: Interactions with plants. FEMS Microbiology Ecology. 2016;92(4):fiw036. DOI: 10.1093/femsec/fiw036
  7. 7. Hermosa R, Viterbo A, Chet I, Monte E. Plant beneficial effects of Trichoderma and of its genes. Microbiology. 2012;158:17-25
  8. 8. Shoresh M, Harman GE, Mastouri F. Induced systemic resistance and plant responses to fungal biocontrol agents. Annual Review of Phytopathology. 2010;48:21-43
  9. 9. Harman GE, Howell CR, Viterbo A, Chet I, Lorito M. Trichoderma plant symbionts species-opportunistic, avirulent plant symbionts. Nature Reviews. Microbiology. 2004;2:43-56
  10. 10. Druzhinina IS, Seidl-Seiboth V, Herrera-Estrella A, Horwitz BA, Kenerley CM, Monte E, et al. Trichoderma: The genomics of opportunistic success. Nature Reviews. Microbiology. 2011;9:749-759. DOI: 10.1038/nrmicro2637
  11. 11. Sharon E, Bar-Eyal M, Chet I, Herrera-Estrella A, Kleifeld O, Spiegel Y. Biological control of root-knot nematode Meloidogyne javanica by Trichoderma harzianum. Phytopathology. 2001;91(7):687-693
  12. 12. Yasmeen R, Siddiqui ZS. Physiological responses of crop plants against Trichoderma harzianum in saline environment. Acta Botanica Croatica. 2017;76(2):154-162
  13. 13. Brotman Y, Landau U, Cuadros-Inostroza L, Takayuki T, Fernie AR, Chet I, et al. Trichoderma-plant root colonization: Escaping early plant defense responses and activation of the antioxidant machinery for saline stress tolerance. PLoS Pathogens. 2013;9(3):e1003221. DOI: 10.1371/journal.ppat.1003221
  14. 14. Kamala T, Devi SI, Sharma KC, Kennedy K. Phylogeny and taxonomical investigation of Trichoderma spp. from Indian region of Indo-Burma biodiversity hot spot region with special reference to Manipur. BioMed Research International. 2015;2015:285261. DOI: 10.1155/2015/285261
  15. 15. Błaszczyk L, Siwulski M, Sobieralski K, Lisiecka J, J˛edryczka M. Trichoderma spp.—Application and prospects for use in organic farming and industry. Journal of Plant Protection Research. 2014;54(4):309-317
  16. 16. Sánchez-Montesinos B, Santos M, Moreno-Gavíra A, Marín-Rodulfo T, Gea FJ, Diánez F. Biological control of fungal diseases by Trichoderma aggressivum f. europaeum and its compatibility with fungicides. Journal of Fungi. 2021;7:598. DOI: 10.3390/jof7080598
  17. 17. Sun J, Karuppiah V, Li Y, Pandian S, Kumaran S, Chen J. Role of cytochrome P450 genes of Trichoderma atroviride T23 on the resistance and degradation of dichlorvos. Chemosphere. 2022;290:133173. DOI: 10.1016/j.chemosphere.2021.133173
  18. 18. De Padua JC, dela Cruz TEE. Isolation and characterization of nickel-tolerant Trichoderma strains from marine and terrestrial environments. Journal of Fungi. 2021;7:591. DOI: 10.3390/jof7080591
  19. 19. Escudero-Leyva E, Alfaro-Vargas P, Muñoz-Arrieta R, Charpentier-Alfaro C, Granados-Montero MM, Valverde-Madrigal KS, et al. Tolerance and biological removal of fungicides by Trichoderma species isolated from the endosphere of wild Rubiaceae plants. Frontiers in Agronomy. 2022;3:772170. DOI: 10.3389/fagro.2021.772170
  20. 20. Weindling R. Trichoderma lignorum as a parasite of other soil fungi. Phytopathology. 1932;22:837-845
  21. 21. Kumar MA, Sharma P. Morphological characterization of biocontrol isolates of Trichoderma to study the correlation between morphological characters and biocontrol efficacy. International Letters of Natural Science. 2016;55:57-67
  22. 22. Di Marco S, Metruccio EG, Moretti S, Nocentini M, Carella G, Pacetti A, et al. Activity of Trichoderma asperellum strain ICC 012 and Trichoderma gamsii strain ICC 080 toward diseases of esca complex and associated pathogens. Frontiers in Microbiology. 2022;12:813410. DOI: 10.3389/fmicb.2021.813410
  23. 23. Druzhinina IS, Chenthamara K, Zhang J, Atanasova L, Yang D, Miao Y, et al. Massive lateral transfer of genes encoding plant cell wall-degrading enzymes to the mycoparasitic fungus Trichoderma from its plant- associated hosts. PLoS Genetics. 2018;14:e1007322. DOI: 10.1371/journal.pgen.1007322
  24. 24. Kubicek CP, Steindorff AS, Chenthamara K, Manganiello G, Henrissat B, Zhang J, et al. Evolution and comparative genomics of the most common Trichoderma species. BMC Genomics. 2019;20:485. DOI: 10.1186/s12864-019-5680-7
  25. 25. Rush TA, Shrestha HK, Gopalakrishnan Meena M, Spangler MK, Ellis JC, Labbé JL, et al. Bioprospecting Trichoderma: A systematic roadmap to screen genomes and natural products for biocontrol applications. Frontiers in Fungal Biology. 2021;2:716511. DOI: 10.3389/ffunb.2021.716511
  26. 26. Singh HB, Singh BN, Singh SP, Singh SR, Sarma BK. Biological control of plant diseases: Status and prospects. In: Recent Advances in Biopesticides: Biotechnological Applications. Vol. 322. New Delhi, India: New India Pub.; 2009
  27. 27. Jaklitsch WM. European species of Hypocrea part I: The green-spored species. Studies in Mycology. 2009;63:1-91. DOI: 10.3114/sim.2009.63.01
  28. 28. Zin NA, Badaluddin NA. Biological functions of Trichoderma spp. for agriculture applications. Annals of Agricultural Science. 2020;65:168-178. DOI: 10.1016/j.aoas.2020.09.003
  29. 29. Santoro PH, Cavaguchi SA, Alexandre TM, Zorzetti J, Neves PMOJ. In vitro sensitivity of antagonistic Trichoderma atroviride to herbicides. Brazilian Archives of Biology and Technology. 2014;57:238-243
  30. 30. Kim CS, Park MS, Kim SC, Maekawa N, Yu SH. Identification of Trichoderma, a competitor of shiitake mushroom (Lentinula edodes), and competition between Lentinula edodes and Trichoderma species in Korea. Plant Pathology Journal. 2012;28:137-148
  31. 31. Hatvani L, Kredics L, Allaga H, Manczinger L, Vágvölgyi C, Kuti K, et al. First report of Trichoderma aggressivum f. aggressivum green mold on Agaricus bisporus in Europe. Plant Disease. 2017;101:1052. DOI: 10.1094/PDIS-12-16-1783-PDN
  32. 32. Chaudhari PJ, Shrivastava P, Khadse AC. Substrate evaluation for mass cultivation of Trichoderma viride. Asiatic Journal of Biotechnology Resources. 2011;2:441-446
  33. 33. Shah S, Nasreen S, Sheikh PA. Cultural and morphological characterization of Trichoderma spp. associated with green mold disease of Pleurotus spp. in Kashmir. Research Journal of Microbiology. 2012;7:139-144
  34. 34. Moreno-Ruiz D, Lichius A, Turra D, Di Pietro A, Zeilinger S. Chemotropism assays for plant symbiosis and mycoparasitism related compound screening in Trichoderma atroviride. Frontiers in Microbiology. 2020;11:601251. DOI: 10.3389/fmicb.2020.601251
  35. 35. Viterbo A, Horwitz BA. Mycoparasitism. In: Borkovich KA, Ebbole DJ, editors. Cellular and Molecular Biology of Filamentous Fungi. Vol. 42. Washington: American Society for Microbiology; 2010. pp. 676-693
  36. 36. Nawrocka J, Małolepsza U. Diversity in plant systemic resistance induced by Trichoderma. Biological Control. 2013;67(2):149-156. DOI: 10.1016/j.biocontrol.2013.07.005
  37. 37. Kolli SK, Adusumilli N. Trichoderma—Its paramount role in agriculture. In: Singh J, Gehlot P, editors. New and Future Developments in Microbial Biotechnology and Bioengineering. Amsterdam, The Netherlands: Elsevier; 2020. pp. 69-83. DOI: 10.1016/B978-0-12-821007-9.00007-3
  38. 38. Manibhushanrao K, Sreenivasaprasad S, Baby U, Joe Y. Susceptibility of rice sheath blight pathogen to mycoparasites. Current Science. 1989;58(9):515-518
  39. 39. Contreras-Cornejo HA, Macías-Rodríguez L, del-Val E, Larsen J. The root endophytic fungus Trichoderma atroviride induces foliar herbivory resistance in maize plants. Applied Soil Ecology. 2018;124:45-53. DOI: 10.1016/j.apsoil.2017.10.004
  40. 40. Vinale F, Sivasithamparam K. Beneficial effects of Trichoderma secondary metabolites on crops. Phytotherapy Research. 2020;34:2835-2842
  41. 41. Strakowska J, Błaszczyk L, Chełkowski J. The significance of cellulolytic enzymes produced by Trichoderma in opportunistic lifestyle of this fungus. Journal of Basic Microbiology. 2014;54(Suppl.1):S2-S13. DOI: 10.1002/jobm.201300821
  42. 42. Saravanakumar K, Li Y, Yu C, Wang Q , Wang M, Sun J, et al. Effect of Trichoderma harzianum on maize rhizosphere microbiome and biocontrol of Fusarium stalk rot. Scientific Reports. 2017;7:1771
  43. 43. Jeleń H, Błaszczyk L, Chełkowski J, Rogowicz K, Strakowska J. Formation of 6-n-pentyl-2H-pyran-2-one (6-PAP) and other volatiles by different Trichoderma species. Mycological Progress. 2013;13(3):589-600. DOI: 10.1007/s11557-013-0942-2
  44. 44. Seidl-Seiboth V, Ihrmark K, Druzhinina IS, Karlsson M. Molecular evolution of Trichoderma chitinases. In: Gupta VK, Schmoll M, Herrera-Estrella A, Upadhyay RS, Druzhinina I, Tuohy MG, editors. Biotechnology and Biology of Trichoderma. Oxford, UK: Elsevier; 2014. pp. 67-78
  45. 45. Guzmán-Guzmán P, Porras-Troncoso MD, Olmedo-Monfil V, Herrera-Estrella A. Trichoderma species: Versatile plant symbionts. Review. Phytopathology. 2019;109:6-16. DOI: doi.org/10.1094/phyto-07-18-0218-rvw
  46. 46. Schmoll M. The information highways of a biotechnological workhorse—Signal transduction in Hypocrea jecorina. BMC Genomics. 2008;9:430
  47. 47. Jaroszuk-Ściseł J, Tyśkiewicz R, Nowak A, Ozimek E, Majewska M, Hanaka A, et al. Phytohormones (auxin, gibberellin) and ACC deaminase in vitro synthesized by the mycoparasitic Trichoderma DEMTkZ3A0 strain and changes in the level of auxin and plant resistance markers in wheat seedlings inoculated with this strain conidia. International Journal of Molecular Sciences. 2019;20:4923
  48. 48. Poveda J, Baptista P. Filamentous fungi as biocontrol agents in olive (Olea europaea L.) diseases: Mycorrhizal and endophytic fungi. Crop Protection. 2021;146:105672. DOI: 10.1016/j.cropro.2021.105672
  49. 49. Halifu S, Deng X, Song X, Song R, Liang X. Inhibitory mechanism of Trichoderma virens ZT05 on Rhizoctonia solani. Planning Theory. 2020;9:912. DOI: 10.3390/plants9070912
  50. 50. Mahmood A, Kataoka R. Potential of biopriming in enhancing crop productivity and stress tolerance. In: Advances in Seed Priming. Berlin/Heidelberg, Germany: Springer; 2018. pp. 127-145
  51. 51. Pescador L, Fernandez I, Pozo MJ, Romero-Puertas MC, Pieterse C, Martínez-Medina A. Nitric oxide signalling in roots is required for MYB72-dependent systemic resistance induced by Trichoderma volatile compounds in Arabidopsis. Journal of Experimental Botany. 2022;73:584-595. DOI: 10.1093/jxb/erab294
  52. 52. Panchalingam H, Powell D, Adra C, Foster K, Tomlin R, Quigley BL, et al. Assessing the various antagonistic mechanisms of Trichoderma strains against the brown root rot pathogen Pyrrhoderma noxium infecting heritage fig trees. Journal of Fungi. 2022;8:1105. DOI: 10.3390/jof8101105
  53. 53. Risoli S, Cotrozzi L, Sarrocco S, Nuzzaci M, Pellegrini E, Vitti A. Trichoderma-induced resistance to Botrytis cinerea in Solanum species: A meta-analysis. Planning Theory. 2022;11:180. DOI: 10.3390/plants11020180
  54. 54. Dugassa A, Alemu T, Woldeha-wariat Y. In-vitro compatibility assay of indigenous Trichoderma and Pseudomonas species and their antagonistic activities against black root rot disease (Fusarium solani) of faba bean (Vicia faba L.). BMC Microbiology. 2021;21:115. DOI: 10.1186/s12866-021-02181-7
  55. 55. Herrera-Téllez VI, Cruz-Olmedo AK, Plasencia J, Gavilanes-Ruíz M, Arce-Cervantes O, Hernández-León S, et al. The protective effect of Trichoderma asperellum on tomato plants against Fusarium oxysporum and Botrytis cinerea diseases involves inhibition of reactive oxygen species production. International Journal of Molecular Sciences. 2019;20:2007. DOI: 10.3390/ijms20082007
  56. 56. Zhao L, Wang Y, Kong S. Effects of Trichoderma asperellum and its siderophores on endogenous auxin in Arabidopsis thaliana under iron-deficiency stress. International Microbiology. 2020;23:501-509
  57. 57. Ghosh SK, Banerjee S, Sengupta C. Siderophore production by antagonistic fungi. Journal of Biopesticides. 2017;10:105-112
  58. 58. Chowdappa S, Jagannath S, Konappa N, Udayashankar AC, Jogaiah S. Detection and characterization of antibacterial siderophores secreted by endophytic fungi from cymbidium aloifolium. Biomolecules. 2020;10:1412
  59. 59. Oszust K, Cybulska J, Frąc M. How do Trichoderma genus fungi win a nutritional competition battle against soft fruit pathogens? A report on niche overlap nutritional potentiates. International Journal of Molecular Sciences. 2020;21:4235
  60. 60. Saravanakumar K, Wang MH. Isolation and molecular identification of Trichoderma species from wetland soil and their antagonistic activity against phytopathogens. Physiological and Molecular Plant Pathology. 2020;109:101458
  61. 61. Bargaz A, Lyamlouli K, Chtouki M, Zeroual Y, Dhiba D. Soil microbial resources for improving fertilizers efficiency in an integrated plant nutrient management system. Frontiers in Microbiology. 2018;9:1606
  62. 62. Nusaibah SA, Musa H. A review report on the mechanism of Trichoderma spp. as biological control agent of the basal stem Rot (BSR) disease of Elaeis guineensis. In: Shah MM, Sharif U, Buhari TR, editors. Trichoderma the Most Widely Used Fungicide. London: IntechOpen; 2019
  63. 63. Kachroo A, Robin GP. Systemic signaling during plant defense. Current Opinion in Plant Biology. 2013;16:527
  64. 64. Gomes EV, Costa M, de Paula RG, de Azevedo RR, da Silva FL, Noronha EF, et al. The Cerato-Platanin protein Epl-1 from Trichoderma harzianum is involved in mycoparasitism, plant resistance induction and self-cell wall protection. Scientific Reports. 2015;5:17998. DOI: 10.1038/srep17998
  65. 65. Ngo MT, Nguyen MV, Han JW, Park MS, Kim H, Choi GJ. In vitro and in vivo antifungal activity of Sorbicillinoids produced by Trichoderma longibrachiatum. Journal of Fungi. 2021;7:428. DOI: 10.3390/jof7060428
  66. 66. Matas-Baca MÁ, Urías García C, Pérez-Álvarez S, Flores-Córdova MA, Escobedo-Bonilla CM, Magallanes-Tapia MA, et al. Morphological and molecular characterization of a new autochthonous Trichoderma sp. isolate and its biocontrol efficacy against Alternaria sp. Saudi Journal of Biological Sciences. 2022;29:2620-2625. DOI: 10.1016/j.sjbs.2021.12.052
  67. 67. Zaid R, Koren R, Kligun E, Gupta R, Leibman-Markus M, Mukherjee PK, et al. Gliotoxin, an immunosuppressive fungal metabolite, primes plant immunity: Evidence from Trichoderma virens-tomato interaction. MBio. 2022;13:e0038922. DOI: 10.1128/mbio.00389-22
  68. 68. Zhu N, Zhou JJ, Zhang SW, Xu BL. Mechanisms of Trichoderma longibrachiatum T6 fermentation against Valsa mali through inhibiting its growth and reproduction, pathogenicity and gene expression. Journal of Fungi. 2022;8:113. DOI: 10.3390/jof8020113
  69. 69. Bhaskar R, Xavier LSE, Udayakumaran G, Kumar DS, Venkatesh R, Nagella P. Biotic elicitors: A boon for the in-vitro production of plant secondary metabolites. Plant Cell Tissue and Organ Culture. 2021;147:1-18
  70. 70. Saravanakumar K, Fan L, Fu K, Yu C, Wang M, Xia H, et al. Cellulase from Trichoderma harzianum interacts with roots and triggers induced systemic resistance to foliar disease in maize. Scientific Reports. 2016;6:35543. DOI: 10.1038/srep35543
  71. 71. El-Hasan A, Walker F, Klaiber I, Schöne J, Pfannstiel J, Voegele RT. New approaches to manage Asian soybean rust (Phakopsora pachyrhizi) using Trichoderma spp. or their antifungal secondary metabolites. Meta. 2022;12:507. DOI: 10.3390/metabo12060507
  72. 72. Mukhopadhyay R, Kumar D. Trichoderma: A beneficial antifungal agent and insights into its mechanism of biocontrol potential. Egyptian Journal of Biological Pest Control. 2020;30(1):133. DOI: 10.1186/s41938-020-00333-x
  73. 73. Alfiky A, Weisskopf L. Deciphering Trichoderma-plant-pathogen interactions for better development of biocontrol applications. Journal of Fungi. 2021;7:61. DOI: 10.3390/jof7010061
  74. 74. Nawrocka J, Małolepsza U, Szymczak K, Szczech M. Involvement of metabolic components, volatile compounds, PR proteins, and mechanical strengthening in multilayer protection of cucumber plants against Rhizoctonia solani activated by Trichoderma atroviride TRS25. Protoplasma. 2018;255:359-373. DOI: 10.1007/s00709-017-1157-1
  75. 75. Thambugala KM, Daranagama DA, Phillips A, Kannangara SD, Promputtha I. Fungi vs. fungi in biocontrol: An overview of fungal antagonists applied against fungal plant pathogens. Frontiers in Cellular and Infection Microbiology. 2020;10:604923. DOI: 10.3389/fcimb.2020.604923
  76. 76. Kong WL, Ni H, Wang WY, Wu XQ. Antifungal effects of volatile organic compounds produced by Trichoderma koningiopsis T2 against Verticillium dahliae. Frontiers in Microbiology. 2022;13:1013468. DOI: 10.3389/fmicb.2022.1013468
  77. 77. Li M, Ren Y, He C, Yao J, Wei M, He X. Complementary effects of dark septate endophytes and Trichoderma strains on growth and active ingredient accumulation of Astragalus mongholicus under drought stress. Journal of Fungi. 2022;8:920. DOI: 10.3390/jof8090920
  78. 78. Rincón AM, Benítez T, Codón AC, Moreno-Mateos MA. Biotechnological aspects of Trichoderma spp. In: Rai M, Bridge PD, editors. Applied Mycology. London, UK: CAB International; 2009. pp. 216-223
  79. 79. Sood M, Kapoor D, Kumar V, Sheteiwy MS, Ramakrishnan M, Landi M, et al. Trichoderma: The “secrets” of a multitalented biocontrol agent. Plants. 2020;9:762
  80. 80. Khan RAA, Najeeb S, Mao Z, Ling J, Yang Y, Li Y, et al. Bioactive secondary metabolites from Trichoderma spp. against phytopathogenic bacteria and root-knot nematode. Microorganisms. 2020;8:401. DOI: 10.3390/microorganisms8030401
  81. 81. Esparza-Reynoso S, Ruíz-Herrera LF, Pelagio-Flores R, Macías-Rodríguez LI, Martínez-Trujillo M, López-Coria M, et al. Trichoderma atroviride-emitted volatiles improve growth of Arabidopsis seedlings through modulation of sucrose transport and metabolism. Plant, Cell & Environment. 2021;44:1961-1976
  82. 82. Subedi S, Dhunaganaa B, Adhikarib A, Regmia V, Dhungana S. Effect of organic manures on growth and yield of cowpea in Chitwan, Nepal. Plant Physiological and Soil Chemistry. 2022;2(2):4-57
  83. 83. Pandit MA, Kumar J, Gulati S, Bhandari N, Mehta P, Katyal R, et al. Major biological control strategies for plant pathogens. Pathogens. 2022;11(2):273
  84. 84. Odebode AC. Control of postharvest pathogens of fruits by culture filtrate from antagonistic fungi. Journal of Plant Protector Research. 2006;46(1):1-5
  85. 85. Álvarez-García S, Mayo-Prieto S, Gutiérrez S, Casquero PA. Self-inhibitory activity of Trichoderma soluble metabolites and their antifungal effects on Fusarium oxysporum. Journal of Fungi. 2020;6:176. DOI: 10.3390/jof6030176
  86. 86. Zhang S, Gan Y, Liu J, Zhou J, Xu B. Optimization of the fermentation media and parameters for the bio-control potential of Trichoderma longibrachiatum T6 against nematodes. Frontiers in Microbiology. 2020;11:574601
  87. 87. Al-Askar AA, Saber W, Ghoneem KM, Hafez EE, Ibrahim AA. Crude citric acid of Trichoderma asperellum: Tomato growth promotor and suppressor of Fusarium oxysporum f. sp. lycopersici. Plants. 2021;10:222. DOI: 10.3390/plants10020222
  88. 88. Degani O, Dor S. Trichoderma biological control to protect sensitive maize hybrids against late wilt disease in the field. Journal of Fungi. 2021;7:315. DOI: 10.3390/jof7040315
  89. 89. Zhang C, Wang W, Hu Y, Peng Z, Ren S, Xue M, et al. A novel salt-tolerant strain Trichoderma atroviride HN082102.1 isolated from marine habitat alleviates salt stress and diminishes cucumber root rot caused by Fusarium oxysporum. BMC Microbiology. 2022;22:67. DOI: 10.1186/s12866-022-02479-0
  90. 90. Zhang Y, Xiao J, Yang K, Wang Y, Tian Y, Liang Z. Transcriptomic and metabonomic insights into the biocontrol mechanism of Trichoderma asperellum M45a against watermelon Fusarium wilt. PLoS One. 2022;17:e0272702. DOI: 10.1371/journal.pone.0272702
  91. 91. John RP, Tyagi RD, Prévost D, Brar SK, Pouleur S, Surampalli RY. Mycoparasitic Trichoderma viride as a biocontrol agent against Fusarium oxysporum f. sp. adzuki and Pythium arrhenomanes and as a growth promoter of soybean. Journal of Crop Protection. 2010;29:1452-1459
  92. 92. Mbarga JB, Ten Hoopen GM, Kuaté J, Adiobo A, Ngonkeu MEL, Ambang Z, et al. Trichoderma asperellum: A potential biocontrol agent for Pythium myriotylum, causal agent of cocoyam (Xanthosoma sagittifolium) root rot disease in Cameroon. Crop Protection. 2012;36:18-22
  93. 93. Ahmed AS, Ezziyyani M, Sánchez CP, Candela ME. Effect of chitin on biological control activity of bacillus spp. and Trichoderma harzianum against root rot disease in pepper (Capsicum annuum) plants. European Journal of Plant Pathology. 2003;109:633-637
  94. 94. Ramezani H. Biological control of root-rot of eggplant caused by Macrophomina phaseolina. American-Eurasian Journal of Agricultural & Environmental Sciences. 2008;4:218-220
  95. 95. Ezziyyani M, Requena ME, Egea-Gilabert C, Candela ME. Biological control of Phytophthora root rot of pepper using Trichoderma harzianum and Streptomyces rochei in combination. International Journal of Environmental Sciences. 2007;155:342-349
  96. 96. Paulitz TC, Ahmad JS, Baker R. Integration of Pythium Nunn and Trichoderma harzianum isolate T-95 for the biological control of Pythium damping-off of cucumber. Plant and Soil. 1990;121:243-250
  97. 97. Lewis JA, Papavizas GC. Application of Trichoderma and Gliocladium in alginate pellets for control of Rhizoctonia damping-off. Plant Pathology. 1987;36:438-446
  98. 98. Howell CR. Cotton seedling preemergence damping-off incited by Rhizopus oryzae and Pythium spp. and its biological control with Trichoderma spp. Phytopathology. 2002;92:177-180
  99. 99. El Komy MH, Saleh AA, Eranthodi A, Molan YY. Characterization of novel Trichoderma asperellum isolates to select effective biocontrol agents against tomato Fusarium wilt. Plant Pathology Journal. 2015;31:50-60
  100. 100. Bernal-Vicente A, Ros M, Pascual JA. Increased effectiveness of the Trichoderma harzianum isolate T-78 against Fusarium wilt on melon plants under nursery conditions. Journal of the Science of Food and Agriculture. 2009;89:827-833
  101. 101. Begum MF, Rahman MA, Alam MF. Biological control of Alternaria fruit rot of chili by Trichoderma species under field conditions. Mycology. 2010;38:113-117
  102. 102. Amin F, Razdan VK. Potential of Trichoderma species as biocontrol agents of soil borne fungal propagules. Journal of Phytology. 2010;2:38-41
  103. 103. Gveroska B, Ziberoski J. Trichoderma harzianum as a biocontrol agent against Alternaria alternata on tobacco. ATI-Applied Technology Innovations. 2012;7:67-76
  104. 104. Rojo FG, Reynoso MM, Ferez M, Chulze SN, Torres AM. Biological control by Trichoderma species of Fusarium solani causing peanut brown root rot under field conditions. Crop Protection. 2007;26:549-555
  105. 105. Freeman S, Minz D, Kolesnik I, Barbul O, Zveibil A, Maymon M, et al. Trichoderma biocontrol of Colletotrichum acutatum and Botrytis cinerea and survival in strawberry. European Journal of Plant Pathology. 2004;110:361-370
  106. 106. Matarese F, Sarrocco S, Gruber S, Seidl-Seiboth V, Vannacci G. Biocontrol of Fusarium head blight: Interactions between Trichoderma and mycotoxigenic Fusarium. Microbiology. 2012;158:98-106
  107. 107. Naeimi S, Okhovvat SM, Javan-Nikkhah M, Vágvölgyi C, Khosravi V, Kredics L. Biological control of Rhizoctonia solani AG1-1A, the causal agent of rice sheath blight with Trichodermastrains. Phytopathologia Mediterranea. 2010;49:287-300
  108. 108. Li GQ , Huang HC, Acharya SN, Erickson RS. Effectiveness of Coniothyrium minitans and Trichoderma atroviride in suppression of sclerotinia blossom blight of alfalfa. Plant Pathology. 2005;54:204-211
  109. 109. Zhang H, Zhu J, Gong Z, Zhu JK. Abiotic stress responses in plants. Nature Reviews Genetics. 2022;23(2):104-119. DOI: 10.1038/s41576-021-00413-0
  110. 110. Kashyap PL, Rai P, Srivastava AK, Kumar S. Trichoderma for climate resilient agriculture. World Journal of Microbiology and Biotechnology. 2017;33:155. DOI: 10.1007/s11274-017-2319-1
  111. 111. Abdul-Halim AM, Shivanand P, Krishnamoorthy S, Taha H. A review on the biological properties of Trichoderma spp. as a prospective biocontrol agent and biofertilizer. Journal of Applied Biology & Biotechnology. 2023;11(5):34-46. DOI: 10.7324/JABB.2023.11504
  112. 112. Hyakumachi M, Kubota M. Fungi as plant growth promoter and disease suppressor. In: Arora DK, editor. Fungal Biotechnology in Agricultural, Food and Environmental Application. New York: Marcel Dekker; 2003. pp. 101-110
  113. 113. Halifu S, Deng X, Song X, Song R. Effects of two Trichoderma strains on plant growth, rhizosphere soil nutrients, and fungal community of Pinus sylvestris var. mongolica annual seedlings. Forests. 2019;10:758. DOI: 10.3390/f10090758
  114. 114. Vinale F, Nigro M, Sivasithamparam K, Flematti G, Ghisalberti E, Ruocco M, et al. Harzianic acid: A novel siderophore from Trichoderma harzianum. FEMS Microbiology Letters. 2013;347:123-129
  115. 115. Yedidia I, Srivastva AK, Kapulnik Y, Chet I. Effect of Trichoderma harzianum on microelement concentrations and increased growth of cucumber plants. Plant and Soil. 2001;235:235-242
  116. 116. Contreras-Cornejo XA, Macías-Rodríguez L, del-Val E, Larsen J. Ecological functions of Trichoderma spp. and their secondary metabolites in the rhizosphere: Interactions with plants. FEMS Microbiology Ecology. 2016;92:1-17
  117. 117. Tyskiewicz R, Nowak A, Ozimek E, Jaroszuk-Sciseł J. Trichoderma: The current status of its application in agriculture for the biocontrol of fungal phytopathogens and stimulation of plant growth. International Journal of Molecular Sciences. 2022;23:2329. DOI: 10.3390/ijms23042329
  118. 118. Siemering G, Ruark M, Geven A. The Value of Trichoderma for Crop Production. Cooperative Extension: University of Wisconsin–Extension; 2016
  119. 119. Yedidia I, Benhamou N, Chet I. Induction of defence responses in cucumber plants (Cucumis sativus L.) by the biocontrol agent Trichoderma harzianum. Applied Environmental Microbiology. 1999;65:10061-11070
  120. 120. Liu Q , Li K, Guo X, Ma L, Guo Y, Liu Z. Developmental characteristics of grapevine seedlings root border cells and their response to ρ-hydroxybenzoic acid. Plant and Soil. 2019;443:199-218

Written By

Lovely Bharti, Kajol Yadav and Ashok Kumar Chaubey

Submitted: 29 September 2023 Reviewed: 04 October 2023 Published: 17 July 2024