Next Article in Journal
Genomic Analysis of Aspergillus Section Terrei Reveals a High Potential in Secondary Metabolite Production and Plant Biomass Degradation
Previous Article in Journal
Functional Characterization of the Transcription Factor Gene CgHox7 in Colletotrichum gloeosporioides, Which Is Responsible for Poplar Anthracnose
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Fungal Biotechnologies in Accomplishing Sustainable Development Goals (SDGs): What Do We Know and What Comes Next?

Department of Horticulture & Life Science, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Author to whom correspondence should be addressed.
J. Fungi 2024, 10(7), 506; https://doi.org/10.3390/jof10070506
Submission received: 26 June 2024 / Revised: 16 July 2024 / Accepted: 18 July 2024 / Published: 22 July 2024
(This article belongs to the Special Issue Innovative Applications and Biomanufacturing of Fungi)

Abstract

:
The present era has witnessed an unprecedented scenario with extreme climate changes, depleting natural resources and rising global food demands and its widespread societal impact. From providing bio-based resources to fulfilling socio-economic necessities, tackling environmental challenges, and ecosystem restoration, microbes exist as integral members of the ecosystem and influence human lives. Microbes demonstrate remarkable potential to adapt and thrive in climatic variations and extreme niches and promote environmental sustainability. It is important to mention that advances in fungal biotechnologies have opened new avenues and significantly contributed to improving human lives through addressing socio-economic challenges. Microbe-based sustainable innovations would likely contribute to the United Nations sustainable development goals (SDGs) by providing affordable energy (use of agro-industrial waste by microbial conversions), reducing economic burdens/affordable living conditions (new opportunities by the creation of bio-based industries for a sustainable living), tackling climatic changes (use of sustainable alternative fuels for reducing carbon footprints), conserving marine life (production of microbe-based bioplastics for safer marine life) and poverty reduction (microbial products), among other microbe-mediated approaches. The article highlights the emerging trends and future directions into how fungal biotechnologies can provide feasible and sustainable solutions to achieve SDGs and address global issues.

1. Introduction

The rising global population, climatic perturbations, and exhausting natural resources are key drivers of ecological imbalance and extinction of plants and animals. In the face of widespread damages and climatic uncertainties, existing biodiversity can support mankind and address the current challenges associated with providing bio-based resources and tackling environmental challenges, ecosystem restoration, and addressing global food demands [1,2,3]. Microorganisms exist as integral members of the ecosystem, demonstrating ubiquitous presence. Due to their significant association with and influence on human lives, microbes demonstrate remarkable potential to adapt and thrive in climatic variations and extreme niches and promote environmental sustainability [4]. Among other biological species, fungi comprise an integral component of our biodiversity and are estimated to include 2.2 and 3.8 million species [5]; however, the vast majority remain unexplored due to limited knowledge/insights about fungal biology and sophisticated technologies. The era of fungal biotechnology started with citric acid production (by controlled fermentation of Aspergillus niger) by Pfizer in 1919 and has expanded to commercial use in food additives and the chemical and pharmaceutical sectors [6,7]. Worldwide companies like Bayer, DuPont, Kerry Group, AB Enzymes, etc., are harnessing fungal resources for economic purposes. Several species of edible fungi are being extensively investigated as attractive resources of ‘high-value’ metabolites including antibiotics, food ingredients/additives, chemicals, industrial enzymes, pigments, etc. [8,9]. While filamentous fungi have been widely explored and harnessed, edible mushrooms (from Ascomycota and Basidiomycota), Saccharomyces cerevisiae, Pichia pastoris, and Yarrowia lipolytica, have been increasingly exploited for commercial use. The advances in fungal biotechnology have opened new avenues and significantly contributed to creating engineered strains with high product yields, bio-functionality, and value addition [10]. For the discovery of new/novel transformative medicines, an insightful discussion suggested that fungi have evolved to create genetically encoded small molecules (GEMs) that can be effective against human targets, and tend to have better pharmacokinetics– brain penetration, oral bioavailability, and less off-target effects, compared to synthetic agents [11] facilitated by advanced high-throughput technologies [12]. With considerable progress in omics biology and their integrated use, a vast repertoire of natural products has been identified and biologically evaluated, attributed to the recent insights on the biosynthetic pathways/mechanisms. Furthermore, optimized production of these compounds can be achieved in cultures via cultivation and metabolic methods including CRISPR-Cas9-mediated gene editing, metabolic engineering, and gene silencing [13]. The publicly available genome resources for fungal species Trichoderma spp., Aspergillus spp., Ganoderma lucidum, Penicillium spp., Rhizopus spp., and others [14,15] have opened new avenues in bridging knowledge gaps in fungal biology and biotechnologies. The advanced molecular predictions have considerably expanded the metabolic pool of fungal high-value metabolites and utilization for creating a bio-based economy and achieving SDGs. Through its policies and reforms, the United Nations SDGs aim to improve people’s livelihood and facilitate sustainable practices (https://sdgs.un.org/, accessed on 20 June 2024), and it is crucial to preserve global biodiversity and bridge the gap between the microbiome and its role in global health [16] (Figure 1).
The rapid developments in fungal biology have facilitated the development of biomass-conversion technologies, and the production of high-value substances as food and feed components. Microbe-based sustainable innovations would likely contribute to United Nations SDGs by providing affordable energy (use of agro-industrial waste by microbial conversions), reducing economic burdens/affordable living conditions (new opportunities by the creation of bio-based industries for sustainable living), tackling climatic changes (use of sustainable alternatives fuels for reducing carbon footprints), conserving marine life (production of microbe-based bioplastics for safer marine life) and poverty reduction (microbial products/microalgae farming), among other initiatives [17,18,19]. These objectives can be achieved via fungal biotechnologies to enhance the production of metabolites, chemicals, and proteins, microbial processing (using microbial enzymes), and advances in biorefineries to develop high-value products. Field and coworkers [20] discussed the potential of mycorrhizal associations as a sustainable approach to achieving food security, conservation, and SDGs [20]. It is important to mention that many mycorrhizal associations of fungi form edible mushrooms, while their collection and consumption are significant for nutrition, traditions, and the global economy [21]. Furthermore, staple cereal crops [22,23,24] and high-value food crops (vanilla flavors) [25] benefit from mycorrhizas, with an important yet overlooked impact on human societies and the ecosystem.
Delving into how the advances in fungal biotechnologies can attain SDGs, state-of-the-art concepts, transformative approaches, achievements, and prospects/directions in the future are discussed in this paper.

2. Fungal Biotechnologies and SDGs—How Far We Have Come

In the face of climate adversities and changing landscapes, human reliance on fossil fuels has impacted productivity and lifestyles and has driven increased emission rates and environmental deterioration [26]. The increased recognition and need to prioritize sustainable practices [27,28] to address and regulate the environmental impact of human activities [29] have been the main goals of SDGs of the United Nations.
The enriched yet less tapped fungal biodiversity can contribute to achieving SDGs, a prospective initiative of the United Nations [30]. Fungal species provide transformative opportunities from petroleum-based to bio-based economy opportunities attributed to converting organic substances into diverse ‘high-value’ products for addressing socio-economic concerns. The utilization of fungal bio-based products is sustainable in securing and enhancing the food supply for a growing population and limiting greenhouse emissions. In addition, the advances in fungal biotechnologies have the potential to tackle global climate change and accomplish SDG reforms (Table 1).

2.1. Fungal ‘High-Value’ Products to Achieve Global Food Security, Tackle Hunger and Malnutrition

Unlocking the road to sustainable food production is challenged by the growing world population, climate fluctuations, food prices, global catastrophes, and agricultural losses due to pathogens [82,83,84]. The development of bio-based products via fungal biotechnologies demonstrates potential in reducing hunger and malnutrition and ensuring food security. Moreover, multiple lifestyle diseases can be tackled by functional foods and nutraceuticals of fungal origin [85,86], following balanced nutrition. Alternative food resources have gained key consensus due to their beneficial health impact and nutritional value. The multi-faceted aspects of food components are improved following microbial synthesis including bio-functionality, quality/nutritional value, peptide synthesis, antimicrobial function, and reduction in antinutritive components, etc. [87,88]. Fungi-based food demonstrates potential as a high-nutritional source for addressing global hunger and malnutrition, besides demonstrating industrial importance (Figure 2).
Among the 2–11 million fungal species in nature, only a fraction (approx. 1.5 lakh species) have been reported; furthermore, only some adhere to the acceptable guidelines of functional food. Since prehistoric times, fungal species have been used to prepare beverages and food products including cheese, bread, food flavors, etc. Only recently, the horizon has expanded to other biotechnological utilities. The prospective pharmaceutical/industrial use of economically viable strains can be attributed to the tractability and transformative potential of fungi, which lead to horizontal gene acquisition and overall plasticity [89]. The edible mushrooms from phyla Ascomycota and Basidiomycota are widely utilized in food preparations across the globe and viewed as exquisite delicacies. The notable fungi namely Aspergillus, Penicillium, and Fusarium sp. are widely recognized for their nutraceutical/pharmacological properties. The fungal mycelium comprises dietary fiber, health-promoting lipids, and vitamins and has health benefits. Some edible fungal species are key sources of probiotics and food flavors [90], while certain filamentous fungi are good protein sources (high protein content) [91]. The food derived from fungal sources has the following major advantages: amino acid profiles, high nutritional and protein content [91,92], and high concentration of fibers, vitamins, and unsaturated fats in the case of edible mushrooms [93]. Research into harnessing the socio-economic benefits of fungi has delved into developing food components comprising nutraceuticals, functional food products, pharmaceuticals, and enzymes [9,94]. Key studies have documented the bioactivities of fungal constituents. Polysaccharides from Morchella esculenta promote antioxidant enzyme function [33], Ganoderma enhance immune functions [95], Tremella relieves epidermal bleeding [96] and Agaricus bisporus restricts the growth of cancer cells [97], among other examples. Health promotion effects are demonstrated by oligosaccharides from copropilous fungi and are developed as a type of functional food [98]. With the advent of white fungal biotechnology, the quality and nutritional value of food products have remarkably improved in the flavor of bread and beverages, single-cell protein (SCP) quality and the yield and shelf life of products [99]. Worldwide, mushrooms are considered to be major aspects of various cuisines and highly nutritional sources of carbohydrates (60%), protein (27–48%), and lipids (2–8%) [100], amino acids (glutamine, valine, leucine, etc.) and vitamins [101,102]. The commercially cultivated mushroom species are represented by Agaricus bisporus, Pleurotus spp., Auricula auricula, Lentinus edodes, and Volvariella volvacea [103,104]. The commercial market for mushrooms has witnessed a tremendous upsurge, with the value for oysters, shiitake, and champignons exceeding USD 50 billion by 2022 [105].
Tian and coworkers [106] showed the beneficial effects of A. bisporus on glucose homeostasis, and the prebiotic effect on glucose homeostasis and regulation of diabetes. In C57BL/6 mice, succinate and propionate produced by Prevotella sp. signals intestinal gluconeogenesis, affects the gut–brain neural circuit, and reduces glucose in hepatic cells. The growing awareness about the nutritional components in multiple fungal species and their increased consumption has raised the demand, and sustainable methods are being employed to meet the increased demands globally. The development of novel strains via genetic engineering studies would be a prospective approach to increase the desired product yield and productivity. Fungal species are ideal resources to develop alternate food components, novel drug molecules, and maintain environmental sustainability.

2.2. Harnessing Pharmaceutical Metabolites from Fungi in Healthcare

SDGs established by the United Nations aim to attain sustainable growth and holistic upliftment of human lives by 2030, utilize alternative bio-based resources, and address global issues. The transition from a fossil-based to a bio-based economy requires the integration of advanced biotechnologies with bioeconomy [107]. The diverse yet interesting group of known fungi inhabits different ecological niches and contributes to multi-faceted roles in the environment, ranging from symbionts, and decomposers to pathogens.
Fungal species produce a plethora of diverse, ‘high-value’ compounds including therapeutics, food components, biofuels, chemicals, vegan leather, organic acids, industrial materials, etc., that can be effectively utilized for sustainable living. Fungi, as the major drivers of bio-based economy, demonstrate diverse fermentation capabilities (industrial value) attributed to their active metabolism (ecological relevance) and adaptation to wider niches (industrial applications). The commercially important high-value products, namely antibiotics and drugs, can be utilized to treat human ailments and positively impact human health and well-being. To date, thousands of pharmacologically active metabolites have been purified and characterized using fungi- demonstrating potent efficacies in treating multiple disorders [108,109].
The landmark discoveries of penicillin and cephalosporin C from fungi opened new avenues and revolutionized fungi-mediated drug discovery. Constituting both classes of traditional drugs and recent landmarks, fungi-derived drugs have been effective in treating the following chronic diseases: autoimmune disorders (immunosuppressants), hypercholesterolemia (statins), and chronic infections (antifungal and antibiotics) [110]. The representative examples include cephalosporins (antibiotic), penicillin V (antibiotic), fusidic acid (antibiotic), griseofulvin (antifungal), retapamulin and enfumafungin (antifungal), among other notable examples. The translational success of these drugs can be attributed to their validation in clinical trials (drugs for drug-resistant depression and cancer). Subsequently, fungal-derived immunosuppressants, such as cyclosporin A (from Tolypocladium inflatum), block the calcineurin pathway (hampering T-cell activation in humans) and have been pivotal in organ transplantations [111]. Another drug (isolated from Penicillium brevicompactum) named mycophenolic acid hampers inosine monophosphate dehydrogenase and biosynthesis of guanine, which restricts the proliferation of lymphocyte (in organ transplantations) [112]. The synthetic compound, fingolimod (inspired by fungi-derived myriocin), is produced by Novartis and achieved blockbuster success as an immunosuppressant for multiple sclerosis, generating USD 1 billion in 2012 [113]. Tiwari et al. [2] extensively discussed and highlighted the potential of plant-associated endophytes to produce potent antimicrobials and counter drug-resistant microbes, an emerging medical concern in the present era. The antimicrobials, namely hypericin, cryptocandin, leucinostatin A, colletotric acid, munumbicins, and their derivatives demonstrated clinical efficacies in treating drug-resistant pathogens; however, assessment and further trials are imperative to establish their therapeutic potential and drug development. In obstetric medicine, bromocriptine (a synthetic form of ergocryptine), is a dopamine agonist and restricts prolactin release from the pituitary gland [114,115]. It is used for the treatment of hyperprolactinemia-related conditions. In therapeutic advances for treating blood cholesterol levels, fungal-derived drugs have proved pivotal in achieving key success. The discovery of mevastatin (compactin) from Penicillium citrinum by Akira Endo, a Japanese scientist, ushered in a new era [34]. Lovastatin (the statin drug), isolated from Monascus ruber (documented as monacolin K) [116] and subsequently from Aspergillus terreus (documented as mevinolin) [117], was quite successful in lowering blood cholesterol. Lovastatin was successfully marketed as a cholesterol-lowering drug in 1987, followed by mevastatin [118]. Statins comprise one of the highest-marketed drugs worldwide, generating sales of USD 25 billion in 2005. Furthermore, several compounds of fungal origin and their derivatives are currently in clinical trials for multiple diseases and include Halimide (synthetic derivative Plinabulin) in phase III trials for cancer, Hypothemycin (synthetic derivative E6201) in phase I trials for solid tumors/melanoma, Wortmannin (synthetic derivative PX-866) in phase II clinical trials for prostate cancer, Cordycepin (synthetic derivative NUC-7738) in phase I trials for lymphoma/solid tumors and Radicicol (synthetic derivative Ganetespib) in phase III trials for lung cancer, among other therapeutics. Gomes and coworkers [93] have extensively discussed and highlighted the importance of marine-derived fungal metabolites for cancer treatment, including leptosins, gliotoxin, shearinine, meleagrin, neoechinulin A, and bostrycin, etc. [119]. The biosynthetic gene clusters (BGCs) in fungal genomes synthesize bioactive, high-value metabolites and can be investigated/engineered for obtaining higher yields of the targeted metabolites.

2.3. Novel Fungal Cell Factories for the Production of Bioactive Metabolites

Fungal-derived metabolites exhibit enormous diversity and interesting bioactivities, namely antimicrobial, hypoglycemic, antiviral, antitumor, immunosuppressant bioactivities, etc. The increasing evidence from studies highlights the potent efficacies of fungal-derived bioactive metabolites as key therapeutics. In addition, functional food/nutraceuticals from fungi have been documented to promote human health and well-being [9,120] and multiple fungal species are powerful resources used to generate ‘high-value’ substances of socio-economic relevance.
Filamentous fungi are widely recognized as efficient producers of natural products, industrial substances, enzymes, proteins, organic acids, etc., and are employed as novel tools for targeted morphology engineering [121]. In addition, fungal biomass is also important in textile industries and as a food component. Discussing industrial relevance, fungi produce key enzymes including phytases, proteases, catalases, and glucoamylases and others with wider usage [122]. Fungal enzymes are also utilized in biofuel production to convert lignocellulosic biomass to fermentable sugars, generating an economic return of over EUR 4 billion [8]. For large-scale cultivation (both solid-state and submerged fermentations), understanding and reprogramming fungal morphogenesis and growth are crucial. Further efforts are needed in process design to optimize fungal morphology for producing a targeted product. Multiple investigations/research in this direction have speculated that septal secretion in fungi may have industrial value and optimization of fungal morphology would improve septal junctions by genetic manipulation studies, in addition to prospective yet less-explored intercalary secretion pathways [123].
Fungal secondary metabolism and its exploration are promising, with studies suggesting that more than 60% of medicines comprise natural products [124]. While efforts are being made for the bio-prospection of fungal resources, new techniques for the activation of silent gene clusters (BGCs) in the laboratory and pilot fermentation studies [125] have been employed and enhanced production via targeted genome manipulation has been achieved [126]. The advances in synthetic biology and a deeper elucidation of the filamentous life cycle for fungal genome engineering facilitate targeted strain development [123].
Aspergillus is fast emerging as a model for genome manipulation, attributed to the technological advances in whole genome sequencing. Engineering initiatives started in the 1950s, ranging from manipulating fungal morphologies and mutagenesis to achieving high product titers. For instance, strains of A. oryzae were subjected to nitrous acid and UV mutagenesis, resulting in less viscous broth and higher production of glucoamylase [127]. Subsequently, mutagenesis of Trichoderma reesei with diethyl sulfite led to a highly branched and short chimeric strain showing enhanced cellulase production [128]. Through the efforts in genome sequencing, an increased understanding of the candidate genes/metabolic pathway has been achieved for strain improvement [14]. In addition, attempts have been made for single nucleotide polymorphism (SNP) identification in fungal genomes for better growth of fungal strains; however, studies are limited. The signaling pathways govern morphological regulation in fungi, and engineering attempts have been made to target components in the cascade for enhanced biotechnological utilities. The key signaling pathways in filamentous fungi, protein kinase A (PKA)/cyclic adenosine monophosphate (cAMP) signaling, calcium ion responses, and mitogen-activated protein kinase (MAPK) are the prime focus of targeted fungal engineering for fungal growth and morphological improvements [123].
In this direction, synthetic biology has made significant advances to create designer chimeras possessing minimized genomes, less complexity, and improved attributes, respectively. A reduced genome of S. cerevisiae was created, and a significant portion (14%) of chromosome 3 was deleted (tRNA, transposans, and wild-type base pairs) [129]. The genome editing of S. cerevisiae chromosome 16 and fusion experiments resulted in chimeras with reduced genome size [130,131]. In A. niger, targeting the fungal genome for minimization was achieved by deletion/inactivation of certain genes/chromosome sections by the CRISPR-Cas9 tool [132]. The genomes of economically viable fungal species have been engineered by the CRISPR-Cas9 editing system and are as follows: available T. reesei [133], M. thermophila [134], A. oryzae [135] and P. chrysogenum [136], among other notable ones, and are exploited industrially. The concept of engineering fungal genomes for size reduction relies on the deletion of non-essential genes and focuses on targeted genes/pathways for morphological changes/growth phenotype and the creation of desired fungal chimeras.
The natural product discovery pipeline has been greatly expedited, which is attributed to the advances in the synthetic biology toolkit. In addition to CRISPR-Cas, advanced synthetic tools are promising in creating designer fungal cell factories, improving the morphological feature and high titer of the desired metabolite. The heterologous expression of key biosynthetic genes (for natural products) has been achieved in A. niger [38], A. nidulans [137], and P. chrysogenum [138], etc., and synthetic fungal chimeras with new/novel attributes have been created by domain swapping [139] and fungal media optimization. Synthetic biology toolkits have expanded the horizons to facilitate polycistronic gene expression in filamentous fungi [140], and next-stage morphological engineering via controlled gene expression of multiple genes using a single promoter, offers interesting insights [141]. Successful attempts in engineering and optimization of tuneable gene switches in filamentous fungi [142] offer precise details of the strain’s morphological characteristics and gene function. It is imperative that advances in fungal imaging have provided precise information about fungal morphology; X-ray microtomography has elucidated the three-dimensional morphology of P. chrysogenum and A. niger [143] and defined new prospects in precise quantification of hyphal number distribution in the pellet, providing future directions in understanding how pellet morphologies affect the titer of the product. These technological developments, optimized in filamentous fungi and other fungal systems (in progress), will result in novel fungal cell factories, including minimizing genomes, higher product titers, and optimized fungal morphologies, in the future.

2.4. Fungi-Based Bioremediation for Environmental Subsistence

The present era has witnessed an increased interest in microbial biodegradation of toxic contaminants for ecosystem restoration. Microbe-assisted bioremediation comprises a cost-effective and eco-friendly approach for the transformation of recalcitrant pollutants into environmentally degradable substances. In addition to other microbial species, fungi-mediated bioremediation is a safe and renewable strategy for mitigating contaminants/polluted locations [144]. Fungi play a critical role as degraders and symbionts, colonizing diverse environmental niches and possessing consistent morphology and multi-faceted metabolic potential. A combination of biostimulation, bioaugmentation, natural attenuation, or individual approaches can be used [145] as per the requirement and efficiency of the microbial strain. Mycoremediation has been a method of choice in environmental cleanups, with multiple efficient fungal species documented for their potency in mitigating heavy metal contamination, pollutants, greenhouse gases, industrial chemicals, etc. [146,147].
Mycorrhizal fungi play a key role in the ecosystem by promoting plant access to nutrients and water in soil and plant tolerance to pathogens. In addition, fungal species in mycorrhizal associations contribute to bioremediation, conservation, and ecosystem well-being [148]. Mycorrhizal associations confer salt tolerance to the plant and promote plant growth and overall health. Bioremediation achieved in the capacities of microbial degradations minimizes the amount and harmful impact of diverse contaminants, while microbial processes aid in the mitigation of pollutants in contaminated sites. Microbe-assisted chemical and physical processes cause disintegration and structural changes in the pollutants and accelerate metabolism. In addition, microbes facilitate energy-dependent chemical reactions for the dissemination of contaminants and electron transfer [149] via oxidation and reduction reactions. In nature, microbes acquire carbon from contaminants for growth and degrade them into simple substances. Quite interestingly, mycoremediation is effective in the removal of heavy metals and radioactive agents to be further decomposed [150]. During pesticide degradation, fungal species obtain nitrogen, carbon, or energy for growth. Molds, e.g., Botrytis and Aspergillus, decompose sugar polymers, celluloses, starches, pectins, oils, chitin, oil components, etc. Subsequently, an environmental hazard, e.g., endosulphan, is effectively degraded by Trichoderma harzianum, Cladosporium oxysporum, Aspergillus spp., and Mucor thermohyalospora. Moreover, fungi-mediated degradation of pesticides into non-toxic substances occurs via processes namely hydroxylation, dehydrogenation, esterification, and deoxygenation [151]. Other fungal strains are capable of bioremediating different contaminants, including textile wastewater detoxification by Zygomycetes and Aspergillus, polychlorinated biphenyl (PCB) degradation by Fusarium solani, Penicillum chrysogenum and Penicillum digitatum, biosorption of pentachlorophenol by Rhizopus oryzae CDBB-H-1877, cellulose degradation by brown rot fungi, xenobiotics degradation by Agaricus bisporus, Pleurotus ostreatus, Pleurotus pulmonarius, etc., heptachlor and heptachlor epoxide remediation by Phanerochaete ostreatus and effluents from textile industries by Ascomycetes and Basidiomycetes fungi, among other distinct examples [151].
Besides the remediation of contaminants present in the environment, the restoration of polluted sites has been achieved via naturally occurring microbes. The representative examples Penicillium, and Aspergillus were effective alleviators of contaminants like textile dyes, chemicals, pesticides, industrial effluents, organic pollutants, etc. [152,153]. In addition, the substantial removal of petroleum hydrocarbons and diesel contaminants in soil has been successful by short-term inoculation of Phanerochaete chrysosporium and Aspergillus niger, which facilitated bioremediation [154,155]. Literature studies have shown that white rot fungi disintegrate harmful substances, namely phenols, effluent, pesticides, heavy metals, polychlorinated biphenyls, etc., and alleviate the adverse impact on soil. Studies have also established the significant potential of fungal enzymes (lipases, catalases, amylases, proteases, peroxidases, etc.) in organic waste management [156], highlighting their industrial value [157,158]. Advanced technologies have immensely contributed to addressing limitations with fungi-mediated bioremediation. In recent times, immobilized fungi in bioreactors (fluidized bed reactors and rotating biological reactors) have been adopted for bioremediation [159,160]. For the treatment of wastewater sludge from sewage plants, it is mixed with microbial inoculum in a broad-scale bioreactor and considered a sustainable approach [161,162]. Furthermore, advanced practices for PAH mitigation include Trichoderma longibrachiatum-based biobarriers on nylon sponges, where high efficiency of PAH removal was achieved [163]. An upcoming approach utilizes yeast expression systems to generate cytochrome P450 monooxygenases that can tackle hydrocarbons and aid in mitigation [164].

2.5. Addressing Climate Changes via Fungal Biotechnologies

For addressing climate change, it is imperative to achieve net-zero emissions by the mid-century to limit temperature rise within 1.5 °C, while adopting measures to sequester, capture, and store excess atmospheric carbon [165]. In a recent report by the World Economic Forum, fungi can play a crucial role in addressing climate change [166]. Fungal species inhabiting natural environments assist forests in absorbing carbon and tackling the potential impacts of climatic fluctuations. While fungi occupy diverse ecological niches and mushrooms are present in shady and damp places, mycorrhizal fungi (ectomycorrhizal fungi) assist trees and forests to absorb CO2 faster and reduce the rate of carbon flow/return in the atmosphere. However, the rapid deforestation every year threatens the beneficial interactions, and promoting the regrowth of forests would reduce global emissions by 30%, as per the guidelines of the COP26 summit in Glasgow [166]. Since little information is available on the role of fungal networks in combating climate change, the Society for the Protection of Underground Networks (SPUN) has devised a project to understand the role of mycorrhizal fungi in areas of climate science and map the ‘Wood Wide Web’. Thousands of fungal samples are collected to map fungal networks and utilized by SPUN (via machine learning) to create these networks and their function as carbon sinks. This information could be used to identify high-priority zones for more carbon storage and tackle extreme conditions.
The growing evidence from the literature suggests that fungi can contribute to farming practices and agriculture. The inoculation of seeding soil with beneficial fungi promotes soil attributes, enriches soil fertility, and decreases atmospheric CO2 levels, crucial to environmental functioning [167,168], while pathogenic macrofungi exploit plants and animals to absorb nutrients and also contribute to biodiversity in the ecosystems. The interconnected mycelial network with the host is crucial and improves nutrient acquisition, transport, and enzyme secretion [169]. These fungi-mediated processes are essential for sustaining biodiversity and ecosystem well-being. The translational success of arbuscular mycorrhizal fungi (AMF) as potential biofertilizers has a major impact on the global market with a value of USD 2 billion. In addition, fungi are key players and perform essential functions in the ecosystem. Globally distributed, fungi carry out processes including bioconversion, energy flow processes, and nutrient cycling and act as symbionts, pathogens, and decomposers in nature [170,171]. According to a study in nature, the biodiversity of fungi determines plant biodiversity, productivity, and variation in the ecosystem and approximately 90% of plants form integral associations with fungi. In addition to other functions, fungi perform mycoremediation (as earlier discussed), degrade chemicals heavy metals, crude oils, etc., absorb heavy metals and radioactive components, and maintain ecological subsistence. However, excessive human activities and pollution levels are impacting fungal diversity/population and signaling climate change. Adequate and urgent efforts are required to stop/minimize deforestation, restore ectomycorrhizal forests, and switch from fossil fuel to renewable energy sources (as in America).

3. Achievements and Prospects in the Present Decade: What Do We Know and What Comes Next?

Recently, the United Nations General Assembly (UNGA) Science Summit stated that “understanding the world of microbes is imperative either to curb dangerous effects or to harness their power for healthier life, for sustainable energy sources, for biodiversity, for tackling climate change and for solving hunger problems”, which is one of the key objectives of the United Nations SDGs. The microbes in the environment are integrally associated and impact human lives. The increased recognition of the favorable impact of beneficial microbes on humans and the environment has contributed to their potential applications in healthcare, agriculture, and ecosystem restoration.
Widely exploited as a source of ‘high-value’ food ingredients (food flavors, pigments, nutritional substances, etc.), the present era has witnessed the utilization of fungi-based biofertilizers to boost crop health and productivity. Moreover, water quality and sanitation have been remarkably improved by microbe-assisted remediation of contaminated water bodies. Other achievements in microbial biotechnologies in achieving SDGs have been biofuel production as a direct source of affordable and clean energy, industrial production of high-value metabolites, potent drugs from fungi approved and marketed for disease treatment with others in different stages of drug development, environmental cleanup via bioremediation of contaminants and plastics and conferring stress tolerance to plants in times of global climate adversities.
Cutting-edge research has focused on deciphering and highlighting the prospects of beneficial microbes in different socio-economic contexts. With the beginning of the transition towards a bio-based economy and the efficient utilization of fungal resources, answers to the following pertinent questions are required: which species has valuable/useful traits, and how can self-sustainability be achieved by fungal production [1]? A better understanding can be achieved with these answers on the road towards a sustainable future. In light of the current findings, it is important to investigate/screen the vast repertoire of fungal species and validate the bioactivities, which are necessary to define the safety profiles for socio-economic applications.
Advanced biotechnological tools have revolutionized the exploration of natural resources. The phylogenomics-guided exploration of specific traits has been inferred from the relationship between microbial species. In addition, progress in analytical equipment and omics-assisted identification of species have contributed to bridging the knowledge gaps in metabolite biosynthesis and evaluation of their bioactive potential [12,172]. Metabolomics studies have attempted to understand the fungal metabolic networks and their dynamics, providing critical insights into the taxonomic identification, fungal stress response, metabolite discovery, metabolic engineering, and plant–fungal interactions. A deeper knowledge of complex fungal interactions and their environmental responses has been attained via metabolomics [172,173]. Omics biology has also contributed to research on edible fungi (cointegrated with other methods), delving into processes including stress resistance, growth and development, and its pharmaceutical value [174,175], providing in-depth information. With the advent of modern genome editing tools, like CRISPR-Cas, the production of high-value metabolites can be optimized by fungal genome engineering, heterologous expression, and gene disruption [13], among others. Molecular analysis of the fungal genome provides a framework to screen beneficial traits, metabolite discovery, and efficient production under laboratory conditions. The key to strengthening fungal resources and biotechnologies to achieve the sustainable goals involves obtaining extensive knowledge of fungal biology (a global effort is needed), building a global network, and providing a knowledge base platform for fungal identification, classification and collection of fungal species [176].

4. The Road Ahead: Future Directions in a Fungal Bio-Based Economy

The enriched yet less tapped fungal biodiversity can contribute to realizing SDGs, a prospective initiative of the United Nations for a better world. Fungal biology and biotechnologies provide transformative opportunities from petroleum-based to bio-based economies attributed to converting organic substances into diverse ‘high-value’ products for socio-economic sustainability. Fungi have been associated with land plants during their evolutionary course, and harnessing the power of ancient players would benefit natural habitats and biodiversity [20]. The utilization of fungal bio-based products is sustainable in securing and enhancing the food supply for a growing population and limiting greenhouse emissions. The development of alternate food products includes Quorn (meat substitute) (https://mycorena.com, accessed on 20 June 2024) [177], filamentous fungi-based biomaterials [178], biorefinery applications (second-generation biofuels) [179], biodegradation of plastics [180], and other notable examples. In addition, the advances in fungal biotechnologies have the potential to tackle climate change and contribute to the United Nations SDGs. The road to sustainable development is not yet reachable, and fungal resources represent prime resources in addressing sustainable livelihood and development in a global context.

Author Contributions

P.T. conceptualized the manuscript. P.T. contributed to the literature collection and writing of the manuscript. K.-I.P. made critical suggestions for improvement. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors express gratitude to their organizations.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Antonelli, A.; Smith, R.J.; Simmonds, M.S.J. Unlocking the properties of plants and fungi for sustainable development. Nat. Plants 2019, 5, 1100–1102. [Google Scholar] [CrossRef] [PubMed]
  2. Tiwari, P.; Thakkar, S.; Dufossé, L. Antimicrobials from Endophytes as novel therapeutics to counter drug-resistant pathogens. Crit. Rev. Biotechnol. 2024, 1–27. [Google Scholar] [CrossRef] [PubMed]
  3. Tiwari, P.; Mishra, B.N.; Sangwan, N.S. β-glucosidases from the fungus Trichoderma: Efficient cellulose machinery in biotechnological application. Biomed. Res. Int. 2013, 2013, 203735. [Google Scholar] [CrossRef] [PubMed]
  4. Abhilash, P.C.; Dubey, R.K.; Tripathi, V.; Gupta, V.K.; Singh, H.B. Plant growth-promoting microorganisms for environmental sustainability. Trends Biotechnol. 2016, 34, 847–850. [Google Scholar] [CrossRef] [PubMed]
  5. Hawksworth, D.L.; Lücking, R. Fungal diversity revisited: 2.2 to 3.8 million species. Microbiol. Spectr. 2017, 5, 1–14. [Google Scholar] [CrossRef] [PubMed]
  6. Cairns, T.C.; Nai, C.; Meyer, V. How a fungus shapes biotechnology: 100 years of Aspergillus niger research. Fungal Biol. Biotechnol. 2018, 5, 13. [Google Scholar] [CrossRef] [PubMed]
  7. Takahashi, J.A.; Barbosa, B.V.R.; Martins, B.d.A.; Guirlanda, C.P.; Moura, M.A.F. Use of the versatility of fungal metabolism to meet modern demands for healthy aging, functional foods, and sustainability. J. Fungi 2020, 6, 223. [Google Scholar] [CrossRef] [PubMed]
  8. Meyer, V.; Andersen, M.R.; Brakhage, A.A.; Braus, G.H.; Caddick, M.X.; Cairns, T.C.; de Vries, R.P.; Haarmann, T.; Hansen, K.; Hertz-Fowler, C.; et al. Current challenges of research on flamentous fungi in relation to human welfare and a sustainable bio-economy: A white paper. Fungal Biol. Biotechnol. 2016, 3, 6. [Google Scholar] [CrossRef] [PubMed]
  9. Tiwari, P.; Dufossé, L. Focus and insights into the synthetic biology-mediated chassis of economically important fungi for the production of high-value metabolites. Microorganisms 2023, 11, 1141. [Google Scholar] [CrossRef]
  10. Skellam, E. Strategies for engineering natural product biosynthesis in Fungi. Trends Biotechnol. 2019, 37, 416–427. [Google Scholar] [CrossRef]
  11. Fungal genomes scoured for drugs. Nat. Biotechnol. 2022, 40, 628. [CrossRef]
  12. Pavarini, D.P.; da Silva, D.B.; Carollo, C.A.; Portella, A.P.F.; Latansio-Aidar, S.R.; Cavalin, P.O.; Oliveira, V.C.; Rosado, B.H.P.; Aidar, M.P.M.; Bolzani, V.S.; et al. Application of MALDI-MS analysis of Rainforest chemodiversity: A keystone for biodiversity conservation and sustainable use. J. Mass Spectrom. 2012, 47, 1482–1485. [Google Scholar] [CrossRef]
  13. Liang, M.-H.; Wang, L.; Wang, Q.; Zhu, J.; Jiang, J.G. High-value bioproducts from microalgae: Strategies and progress. Crit. Rev. Food Sci. Nutr. 2019, 59, 2423–2441. [Google Scholar] [CrossRef]
  14. Grigoriev, I.V.; Nikitin, R.; Haridas, S.; Kuo, A.; Ohm, R.; Otillar, R.; Riley, R.; Salamov, A.; Zhao, X.; Korzeniewski, F.; et al. MycoCosm portal: Gearing up for 1000 fungal genomes. Nucleic Acids Res. 2014, 42, D699–D704. [Google Scholar] [CrossRef] [PubMed]
  15. Stajich, J.E.; Harris, T.; Brunk, B.P.; Brestelli, J.; Fischer, S.; Harb, O.S.; Kissinger, J.C.; Li, W.; Nayak, V.; Pinney, D.F.; et al. FungiDB: An integrated functional genomics database for fungi. Nucleic Acids Res. 2012, 40, D675–D681. [Google Scholar] [CrossRef] [PubMed]
  16. Cavicchioli, R.; Ripple, W.J.; Timmis, K.N.; Azam, F.; Bakken, L.R.; Baylis, M.; Behrenfeld, M.J.; Boetius, A.; Boyd, P.W.; Classen, A.T.; et al. Scientists’ warning to humanity: Microorganisms and climate change. Nat. Rev. Microbiol. 2019, 17, 569–586. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, Z.; Wang, K.; Chen, Y.; Tan, T.; Nielsen, J. Third-generation biorefineries as the means to produce fuels and chemicals from CO2. Nat. Catal. 2020, 3, 274–288. [Google Scholar] [CrossRef]
  18. Rosenboom, J.-G.; Langer, R.; Traverso, G. Bioplastics for a circular economy. Nat. Rev. Mater. 2022, 7, 117–137. [Google Scholar] [CrossRef]
  19. Diaz-Rodriguez, A.M.; Gastelum, L.A.S.; Pablos, C.M.F.; Parra-Cota, F.I.; Santoyo, G.; Puente, M.L.; Bhattacharya, D.; Mukherjee, J.; Santos-Villalobos, S.d.L. The current and future role of microbial culture collections in food security worldwide. Front. Sustain. Food Syst. 2021, 4, 101. [Google Scholar] [CrossRef]
  20. Field, K.J.; Daniell, T.; Johnson, D.; Helgason, T. Mycorrhizal mediation of sustainable development goals. Plants People Planet 2021, 3, 430–432. [Google Scholar] [CrossRef]
  21. Pérez-Moreno, J.; Guerin-Laguette, A.; Rinaldi, A.C.; Yu, F.; Verbeken, A.; Hernández-Santiago, F.; Martínez-Reyes, M. Edible mycorrhizal fungi of the world: What is their role in forest sustainability, food security, biocultural conservation and climate change? Plants People Planet 2021, 3, 471–490. [Google Scholar] [CrossRef]
  22. Frew, A. Contrasting effects of commercial and native arbuscular mycorrhizal fungal inoculants on plant biomass allocation, nutrients, and phenolics. Plants People Planet 2020, 3, 536–540. [Google Scholar] [CrossRef]
  23. Thirkell, T.J.; Campbell, M.; Driver, J.; Pastok, D.; Merry, B.; Field, K.J. Cultivar-dependent increases in mycorrhizal nutrient acquisition by barley in response to elevated CO2. Plants People Planet 2020, 3, 553–566. [Google Scholar] [CrossRef]
  24. Watts-Williams, S.J.; Gilbert, S.E. Arbuscular mycorrhizal fungi affect the concentration and distribution of nutrients in the grain differently in barley compared with wheat. Plants People Planet 2020, 3, 567–577. [Google Scholar] [CrossRef]
  25. Johnson, L.J.A.N.; Gónzalez-Chávez, M.C.A.; Carrillo-González, R.; Porras-Alfaro, A.; Mueller, G.M. Vanilla aerial and terrestrial roots host rich communities of orchid mycorrhizal and ectomycorrhizal fungi. Plants People Planet 2020, 3, 541–552. [Google Scholar] [CrossRef]
  26. IPCC. Climate change 2021: The physical science basis. In Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Masson-Delmotte, P.V., Zhai, A., Pirani, S.L., Connors, C., Péan, S., Berger, N., Caud, Y., Chen, L., Goldfarb, M.I., Gomis, M., Eds.; Cambridge University Press: Cambridge, UK, 2021. [Google Scholar]
  27. Brito, I.; Carvalho, M.; Goss, M.J. Managing the functional diversity of arbuscular mycorrhizal fungi for the sustainable intensification of crop production. Plants People Planet 2021, 3, 491–505. [Google Scholar] [CrossRef]
  28. Cobb, A.B.; Duell, E.B.; Haase, K.B.; Miller, R.M.; Wu, Y.Q.; Wilson, G.W.T. Utilizing mycorrhizal responses to guide selective breeding for agricultural sustainability. Plants People Planet 2021, 3, 578–587. [Google Scholar] [CrossRef]
  29. Verbruggen, E.; Struyf, E.; Vicca, S. Can arbuscular mycorrhizal fungi speed up carbon sequestration by enhanced weathering? Plants People Planet 2021, 3, 445–453. [Google Scholar] [CrossRef]
  30. Vuong, P.; Chong, S.; Kaur, P. The little things that matter: How bioprospecting microbial biodiversity can build towards the realization of United Nations Sustainable Development Goals. NPJ Biodivers. 2022, 1, 4. [Google Scholar] [CrossRef]
  31. Liu, L.; Chen, Z.; Liu, W.; Ke, X.; Tian, X.; Chu, J. Cephalosporin C biosynthesis and fermentation in Acremonium chrysogenum. Appl. Microbiol. Biotechnol. 2022, 106, 6413–6426. [Google Scholar] [CrossRef]
  32. Drugs.com. Lentinan. Available online: https://www.drugs.com/npp/lentinan.html (accessed on 21 June 2024).
  33. Browne, A.G.P.; Fisher, M.C.; Henk, D.A. Species-specific PCR to describe local-scale distributions of four cryptic species in the Penicillium chrysogenum complex. Fungal Ecol. 2013, 6, 419–429. [Google Scholar] [CrossRef]
  34. Stossel, T.P. The discovery of statins. Cell 2008, 134, 903–905. [Google Scholar] [CrossRef]
  35. van den Berg, M.A. Impact of the Penicillium chrysogenum genome on industrial production of metabolites. Appl. Microbiol. Biotechnol. 2011, 92, 45–53. [Google Scholar] [CrossRef]
  36. Anand, S.; Singh, K.S.; Aggarwal, D. Expanding Avenues for Probiotic Yeast. In Microbial Cell Factories, 1st ed.; Sharma, D., Saharan, B.S., Eds.; CRC Press: Boca Raton, FL, USA, 2018; pp. 125–141. [Google Scholar]
  37. Boruta, T.; Bizukojc, M. Production of lovastatin and itaconic acid by Aspergillus terreus: A comparative perspective. World J. Microbiol. Biotechnol. 2017, 33, 34. [Google Scholar] [CrossRef] [PubMed]
  38. Richter, L.; Wanka, F.; Boecker, S.; Storm, D.; Kurt, T.; Vural, O.; Süßmuth, R.; Meyer, V. Engineering of Aspergillus niger for the production of secondary metabolites. Fungal Biol. Biotechnol. 2014, 1, 4. [Google Scholar] [CrossRef] [PubMed]
  39. Kraševec, N.; Milunovic, T.; Lasnik, M.A.; Lukancic, I.; Komel, R.; Porekar, V.G. Human granulocyte colony-stimulating factor (G-CSF) produced in the filamentous fungus Aspergillus niger. Acta Chim. Slov. 2014, 61, 709–717. [Google Scholar]
  40. Kuivanen, J.; Wang, Y.J.; Richard, P. Engineering Aspergillus niger for galactaric acid production: Elimination of galactaric acid catabolism by using RNA sequencing and CRISPR/Cas9. Microb. Cell Fact. 2016, 15, 210–219. [Google Scholar] [CrossRef] [PubMed]
  41. Mir, R.A.; Kaushik, P.S.; Chowdery, R.A.; Anuradha, M. Elicitation of forskolin in cultures of Rhizactonia bataticola-a phytochemical synthesizing endophytic fungi. Int. Pharm. Pharmaceut. Sci. 2015, 7, 185–189. [Google Scholar]
  42. Maehara, S.; Simanjuntak, P.; Ohashi, K.; Shibuya, H. Composition of endophytic fungi living in Cinchona ledgeriana (Rubiaceae). J. Nat. Med. 2010, 64, 227–230. [Google Scholar] [CrossRef]
  43. Kaul, S.; Ahmed, M.; Zargar, K.; Sharma, P.; Dhar, M.K. Prospecting endophytic fungal assemblage of Digitalis lanata Ehrh. (foxglove) as a novel source of digoxin: A cardiac glycoside. 3 Biotech 2013, 3, 335–340. [Google Scholar] [CrossRef]
  44. Nanou, K.; Roukas, T. Waste cooking oil: A new substrate for carotene production by Blakeslea trispora in submerged fermentation. Bioresour. Technol. 2016, 203, 198–203. [Google Scholar] [CrossRef] [PubMed]
  45. Mantzouridou, F.T.; Naziri, E. Scale translation from shaken to diffused bubble aerated systems for lycopene production by Blakeslea trispora under stimulated conditions. Appl. Microbiol. Biotechnol. 2017, 101, 1845–1856. [Google Scholar] [CrossRef] [PubMed]
  46. Chaudhary, V.; Katyal, P.; Poonia, A.K.; Kaur, J.; Puniya, A.K.; Panwar, H. Natural pigment from Monascus: The production and therapeutic significance. J. Appl. Microbiol. 2021, 133, 18–38. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, Y.; Xiao, W.; Wang, Y.; Liu, H.; Li, X.; Yuan, Y. Lycopene overproduction in Saccharomyces cerevisiae through combining pathway engineering with host engineering. Microb. Cell Fact. 2016, 15, 113. [Google Scholar] [CrossRef] [PubMed]
  48. Procópio, D.P.; Lee, J.W.; Shin, J.; Tramontina, R.; Ávila, P.F.; Brenelli, L.B.; Squina, F.M.; Damasio, A.; Rabelo, S.C.; Goldbeck, R.; et al. Metabolic engineering of Saccharomyces cerevisiae for second-generation ethanol production from xylo-oligosaccharides and acetate. Sci. Rep. 2023, 13, 19182. [Google Scholar] [CrossRef]
  49. Runguphan, W.; Keasling, J.D. Metabolic engineering of Saccharomyces cerevisiae for production of fatty acid-derived biofuels and chemicals. Metab. Eng. 2014, 21, 103–113. [Google Scholar] [CrossRef] [PubMed]
  50. Paramasivan, K.; Mutturi, S. Progress in terpene synthesis strategies through the engineering of Saccharomyces cerevisiae. Crit. Rev. Biotechnol. 2017, 37, 974–989. [Google Scholar] [CrossRef] [PubMed]
  51. Fu, L.; Wang, Y.; Wang, J.; Yang, Y.; Hao, L. Evaluation of the antioxidant activity of extracellular polysaccharides from Morchella esculenta. Food Funct. 2013, 4, 871–879. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, L.; Li, C.; Ren, L.; Guo, H.; Li, Y. Production of pork sausages using Pleaurotus eryngii with different treatments as replacements for pork back fat. J. Food Sci. 2019, 84, 3091–3098. [Google Scholar] [CrossRef]
  53. Finnigan, T.J.A.; Wall, B.T.; Wilde, P.J.; Stephens, F.B.; Taylor, S.L.; Freedman, M.R. Mycoprotein: The future of nutritious nonmeat protein, a symposium review. Curr. Dev. Nutr. 2019, 3, nzz021. [Google Scholar] [CrossRef]
  54. Barzee, T.J.; Cao, L.; Pan, Z.; Zhang, R. Fungi for future foods. J. Future Foods 2021, 1, 25–37. [Google Scholar] [CrossRef]
  55. Patricia, M. Cheese: Exploring Taste and Tradition; Gibbs Smith: Layton, UT, USA, 2010; p. 12. ISBN 9781423606512. [Google Scholar]
  56. Caron, T.; Piver, M.L.; Péron, A.C.; Lieben, P.; Lavigne, R.; Brunel, S.; Roueyre, D.; Place, M.; Bonnarme, P.; Giraud, T.; et al. Strong effect of Penicillium roqueforti populations on volatile and metabolic compounds responsible for aromas, flavor and texture in blue cheeses. Int. J. Food Microbiol. 2021, 354, 109174. [Google Scholar] [CrossRef] [PubMed]
  57. Jo, C.; Zhang, J.; Tam, J.M.; Church, G.M.; Khalil, A.S.; Segrè, D.; Tang, T.C. Unlocking the magic in mycelium: Using synthetic biology to optimize filamentous fungi for biomanufacturing and sustainability. Mater. Today Bio 2023, 19, 100560. [Google Scholar] [CrossRef]
  58. Chai, K.F.; Ng, K.R.; Samarasiri, M.; Chen, W.N. Precision fermentation to advance fungal food fermentations. Curr. Opin. Food Sci. 2022, 47, 100881. [Google Scholar] [CrossRef]
  59. Bamforth, C.W.; Cook, D.J. Food, Fermentation, and Micro-Organisms: Second Edition; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2019. [Google Scholar]
  60. Allwood, J.G.; Wakeling, L.T.; Bean, D.C. Fermentation and the microbial community of Japanese koji and miso: A review. J. Food Sci. 2021, 86, 2194–2207. [Google Scholar] [CrossRef] [PubMed]
  61. Papagianni, M. Advances in citric acid fermentation by Aspergillus niger: Biochemical aspects, membrane transport and modeling. Biotechnol. Adv. 2007, 25, 244–263. [Google Scholar] [CrossRef] [PubMed]
  62. Larroude, M.; Celinska, E.; Back, A.; Thomas, S.; Nicaud, J.M.; Ledesma-Amaro, R. A synthetic biology approach to transform Yarrowia lipolytica into a competitive biotechnological producer of b-carotene. Biotechnol. Bioeng. 2018, 115, 464–472. [Google Scholar] [CrossRef] [PubMed]
  63. Pollmann, H.; Breitenbach, J.; Sandmann, G. Engineering of the carotenoid pathway in Xanthophyllomyces dendrorhous leading to the synthesis of zeaxanthin. Appl. Microbiol. Biotechnol. 2017, 101, 103–111. [Google Scholar] [CrossRef] [PubMed]
  64. Sakamoto, T.; Sakuradani, E.; Okuda, T.; Kikukawa, H.; Ando, A.; Kishino, S.; Izumi, Y.; Bamba, T.; Shima, J.; Ogawa, J. Metabolic engineering of oleaginous fungus Mortierella alpina for high production of oleic and linoleic acids. Bioresour. Technol. 2017, 247, 1610–1615. [Google Scholar] [CrossRef]
  65. Lu, X.; Brennan, M.A.; Serventi, L.; Liu, J.; Guan, W.; Brennan, C.S. Addition of mushroom powder to pasta enhances the antioxidant content and modulates the predictive glycaemic response of pasta. Food Chem. 2018, 264, 199–209. [Google Scholar] [CrossRef]
  66. Becker, J.; Tehrani, H.H.; Ernst, P.; Blank, L.M.; Wierckx, N. An optimized Ustilago maydis for Itaconic acid production at maximal theoretical yield. J. Fungi 2020, 7, 20. [Google Scholar] [CrossRef]
  67. Rosa, J.C.C.; Colombo, L.T.; Alvim, M.C.T.; Avonce, N.; Dijck, P.V.; Passos, F.M.L. Metabolic engineering of Kluyveromyces lactis for L-ascorbic acid (vitamin C) biosynthesis. Microb. Cell Fact. 2013, 12, 59. [Google Scholar] [CrossRef]
  68. Fitz, E.; Gamauf, C.; Seiboth, B.; Wanka, F. Deletion of the small GTPase rac1 in Trichoderma reesei provokes hyperbranching and impacts growth and cellulase production. Fungal Biol. Biotechnol. 2019, 6, 16. [Google Scholar] [CrossRef] [PubMed]
  69. Mycotex: Textile Made from Mushroom Mycelium. Available online: https://materialdistrict.com/article/mycotex-textile-mushroom-mycelium/ (accessed on 21 June 2024).
  70. Haneef, M.; Ceseracciu, L.; Canale, C.; Bayer, I.S.; Heredia-Guerrero, J.A.; Athanassiou, A. Advanced materials from fungal mycelium: Fabrication and tuning of physical properties. Sci. Rep. 2017, 7, 41292. [Google Scholar] [CrossRef]
  71. Somacal, S.; Pinto, V.S.; Vendruscolo, R.G.; Somacal, S.; Wagner, R.; Ballus, C.A.; Kuhn, R.C.; Mazutti, M.A.; Menezes, C.R. Maximization of microbial oil containing polyunsaturated fatty acid production by Umbelopsis (Mortierella) isabellina. Biocatal. Agric. Biotechnol. 2020, 30, 101831. [Google Scholar] [CrossRef]
  72. Mattern, D.J.; Valiante, V.; Horn, F.; Petzke, L.; Brakhage, A.A. Rewiring of the Austinoid biosynthetic pathway in filamentous fungi. ACS Chem. Biol. 2017, 12, 2927–2933. [Google Scholar] [CrossRef] [PubMed]
  73. Mascarin, G.M.; Jaronski, S.T. The production and uses of Beauveria bassiana as a microbial insecticide. World J. Microbiol. Biotechnol. 2016, 32, 177. [Google Scholar] [CrossRef]
  74. Mascarin, G.M.; Lopes, R.B.; Delalibera, Í., Jr.; Fernandes, É.K.K.; Luz, C.; Faria, M. Current status and perspectives of fungal entomopathogens used for microbial control of arthropod pests in Brazil. J. Invertebr. Pathol. 2019, 165, 46–53. [Google Scholar] [CrossRef]
  75. Djonovic, S.; Pozo, M.J.; Dangott, L.J.; Howell, C.R.; Kenerley, C.M. Sm1, a proteinaceous elicitor by the biocontrol fungus Trichoderma virens induces plant defense responses and systemic resistance. Mol. Plant-Microbe Interact. 2006, 19, 838–853. [Google Scholar] [CrossRef]
  76. Garnica-Vergara, A.; Barrera-Ortiz, S.; Muñoz-Parra, E.; Raya-González, J.; Méndez-Bravo, A.; Macías-Rodríguez, L.; Ruiz-Herrera, L.F.; López-Bucio, J. The volatile 6-pentyl-2H-pyran-2-one from Trichoderma atroviride regulates Arabidopsis thaliana root morphogenesis via auxin signaling and ETHYLENE INSENSITIVE 2 functioning. New Phytol. 2015, 209, 1496–1512. [Google Scholar] [CrossRef]
  77. Vinale, F.; Marra, R.; Scala, F.; Ghisalberti, E.L.; Lorito, M.; Sivasithamparam, K. Major secondary metabolites produced by two commercial Trichoderma strains active against different phytopathogens. Lett. Appl. Microbiol. 2006, 43, 143–148. [Google Scholar] [CrossRef] [PubMed]
  78. El-Saadony, M.T.; Saad, A.M.; Soliman, S.M.; Salem, H.M.; Ahmed, A.I.; Mahmood, M.; El-Tahan, A.M.; Ebrahim, A.A.M.; El-Mageed, T.A.A.; Negm, S.H.; et al. Plant growth-promoting microorganisms as biocontrol agents of plant diseases: Mechanisms, challenges and future perspectives. Front. Plant Sci. 2022, 13, 923880. [Google Scholar] [CrossRef] [PubMed]
  79. Marumo, S.; Katayama, M.; Komori, E.; Ozaki, Y.; Natsume, M.; Kondo, S. Microbial production of Abscisic acid by Botrytis cinerea. Agri. Biol. Chem. 1982, 46, 1967–1968. [Google Scholar] [CrossRef]
  80. Shanthiyaa, V.; Saravanakumar, D.; Rajendran, L.; Karthikeyan, G.; Prabakar, K.; Raguchander, T. Use of Chaetomium globosum for biocontrol of potato late blight disease. Crop Prot. 2013, 52, 33–38. [Google Scholar] [CrossRef]
  81. Kusari, S.; Hertweck, C.; Spiteller, M. Chemical ecology of endophytic fungi: Origins of secondary metabolites. Chem. Biol. 2012, 19, 792–798. [Google Scholar] [CrossRef] [PubMed]
  82. Khoury, C.K.; Bjorkman, A.D.; Dempewolf, H.; Ramirez-Villegas, J.; Guarino, L.; Jarvis, A.; Struik, P.C. Increasing homogeneity in global food supplies and the implications for food security. Proc. Natl. Acad. Sci. USA 2014, 111, 4001–4006. [Google Scholar] [CrossRef] [PubMed]
  83. Oaklander, M. Many Foods Subsidized By the Government Are Unhealthy. 2016. Available online: http://time.com/4393109/food-subsidies-obesity (accessed on 5 March 2024).
  84. Tiwari, P. Sustainable agriculture and nanotechnologies for food and nutraceutical production—An update. In Plant and Nanoparticles; Chen, J., Ed.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 315–338. [Google Scholar]
  85. Tiwari, P.; Bae, H. Endophytic fungi: Key insights, emerging prospects, and challenges in natural product drug discovery. Microorganisms 2022, 10, 360. [Google Scholar] [CrossRef]
  86. Tiwari, P.; Seogchan, K.; Bae, H. Plant-endophyte associations: Rich yet under-explored sources of novel bioactive molecules and applications. Microbiol. Res. 2023, 266, 127241. [Google Scholar] [CrossRef]
  87. Mallikarjuna, N.; Yellamma, K. Genetic and metabolic engineering of microorganisms for the production of various food products. Recent Dev. Appl. Microbiol. Biochem. 2019, 60, 167–182. [Google Scholar]
  88. Tiwari, P.; Khare, T.; Shriram, V.; Bae, H.; Kumar, V. Plant synthetic biology for producing potent phyto-antimicrobials to combat antimicrobial resistance. Biotechnol. Adv. 2021, 48, 107729. [Google Scholar] [CrossRef]
  89. Roth, M.G.; Westrick, N.M.; Baldwin, T.T. Fungal biotechnology: From yesterday to tomorrow. Front. Fungal Biol. 2023, 4, 1135263. [Google Scholar] [CrossRef] [PubMed]
  90. Hashempour-Baltork, F.; Khosravi-Darani, K.; Hosseini, H.; Farshi, P.; Reihani, S.F.S. mycoproteins as safe meat substitutes. J. Clean. Prod. 2020, 253, 119958. [Google Scholar] [CrossRef]
  91. Derbyshire, E.J.; Delange, J. Fungal protein—What is it and what is the health evidence? A systematic review focusing on mycoprotein. Front. Sustain. Food Syst. 2021, 5, 581682. [Google Scholar] [CrossRef]
  92. Souza Filho, P.F.; Andersson, D.; Ferreira, J.A.; Taherzadeh, M.J. Mycoprotein: Environmental impact and health aspects. World J. Microbiol. Biotechnol. 2019, 35, 147. [Google Scholar] [CrossRef] [PubMed]
  93. Vega Oliveros, C. Comparación de la Producción de Metabolitos Secundarios Bioactivos con dos Fuentes de Carbono en la Fermentación Líquida de Una Especie de Pleurotus y su uso Potencial en un Alimento de Tipo Funcional. Ph.D. Thesis, Universidad Nacional de Colombia, Bogotá, DC, Colombia, 2016. [Google Scholar]
  94. Tiwari, P. Endophytes: Types, Potential Uses and Mechanisms of Action; Nova Publishers: Hauppauge, NY, USA, 2022; ISBN 979-8-88697-205-4. [Google Scholar]
  95. Li, R.; Zhang, J.; Zhang, T.H. Immunomodulatory activities of polysaccharides from Ganoderma on immune effector cells. Food Chem. 2020, 340, 127933. [Google Scholar]
  96. Wen, L.; Gao, Q.; Ma, C.W.; Ge, Y.; You, L.; Liu, R.H.; Fu, X.; Liu, D. Effect of polysaccharides from Tremella fuciformis on UV-induced photoaging. J. Func. Foods. 2016, 20, 400–441. [Google Scholar] [CrossRef]
  97. Jeong, S.C.; Koyyalamudi, S.R.; Jeong, Y.T.; Song, C.H.; Pang, G. Macrophage immunomodulating and antitumor activities of polysaccharides isolated from Agaricus bisporus white button mushrooms. J. Med. Food 2012, 15, 58–65. [Google Scholar] [CrossRef]
  98. Ojwach, J.; Adetunji, A.I.; Mutanda, T.; Mukaratirwa, S. Oligosaccharides production from coprophilous fungi: An emerging functional food with potential health-promoting properties. Biotechnol. Rep. 2022, 33, e00702. [Google Scholar] [CrossRef] [PubMed]
  99. Challa, S.; Dutta, T.; Neelapu, N.R.R. Fungal White Biotechnology Applications for Food Security: Opportunities and Challenges. In Recent Advancement in White Biotechnology through Fungi; Yadav, A., Singh, S., Mishra, S., Gupta, A., Eds.; Fungal Biology; Springer: Cham, Switzerland, 2019. [Google Scholar] [CrossRef]
  100. Valverde, M.E.; Hernandez-Perez, T.; Paredes-Lopez, O. Edible mushrooms: Improving human health and promoting quality life. Int. J. Microbiol. 2015, 2015, 376387. [Google Scholar] [CrossRef]
  101. Mattila, P.; Konko, K.; Eurola, M.; Pihlava, J.M.; Astola, J.; Vahteristo, L.; Hietaniemi, V.; Kumpulainen, J.; Valtonen, M.; Piironen, V. Contents of vitamins, mineral elements, and some phenolic compounds in cultivated mushrooms. J. Agric. Food Chem. 2001, 49, 2343–2348. [Google Scholar] [CrossRef]
  102. Kalac, P. A review of chemical composition and nutritional value of wild-growing and cultivated mushrooms. J. Sci. Food Agric. 2013, 93, 209. [Google Scholar] [CrossRef] [PubMed]
  103. Aida, F.M.; Shuhaimi, M.; Yazid, M.; Maaruf, A.G. Mushroom as a potential source of prebiotics: A review. Trends Food Sci. Technol. 2009, 20, 567–575. [Google Scholar] [CrossRef]
  104. Cheung, P.C. Mini-review on edible mushrooms as source of dietary fiber: Preparation and health benefits. Food Sci. Hum. Wellness 2013, 2, 162–166. [Google Scholar] [CrossRef]
  105. Available online: https://www.grandviewresearch.com/industry-analysis/mushroom-market (accessed on 20 June 2024).
  106. Tian, Y.; Nichols, R.G.; Roy, P.; Gui, W.; Smith, P.B.; Zhang, J.; Lin, Y.; Weaver, V.; Cai, J.; Patterson, A.D.; et al. Prebiotic effects of white button mushroom (Agaricus bisporus) feeding on succinate and intestinal gluconeogenesis in C57BL/6 mice. J. Funct. Foods 2018, 45, 223–232. [Google Scholar] [CrossRef]
  107. Singh, I.; Thakur, P. Impact of fungi on the world economy and its sustainability: Current status and potentials. In Fungal Resources for Sustainable Economy; Singh, I., Ed.; Springer Nature: Singapore, 2023. [Google Scholar] [CrossRef]
  108. Aly, A.H.; Debbab, A.; Proksch, P. Fifty years of drug discovery from fungi. Fungal Divers. 2011, 50, 3–19. [Google Scholar] [CrossRef]
  109. Singh, V.K.; Tiwari, R.; Kumar, A.; Gupta, R.; Kumar, R. Therapeutic potential of fungal endophyte-derived bioactive compound in Protozoan diseases. In Endophytic Fungi Fungal Biology; Singh, B.P., Abdel-Azeem, A.M., Gautam, V., Singh, G., Singh, S.K., Eds.; Springer: Cham, Switzerland, 2024. [Google Scholar] [CrossRef]
  110. Prescott, T.A.K.; Hill, R.; Mas-Claret, E.; Gaya, E.; Burns, E. Fungal drug discovery for chronic disease: History, new discoveries and new approaches. Biomolecules 2023, 13, 986. [Google Scholar] [CrossRef] [PubMed]
  111. Bushley, K.E.; Raja, R.; Jaiswal, P.; Cumbie, J.S.; Nonogaki, M.; Boyd, A.E.; Owensby, C.A.; Knaus, B.J.; Elser, J.; Miller, D.; et al. The genome of Tolypocladium inflatum: Evolution, organization, and expression of the cyclosporin biosynthetic gene cluster. PLoS Genet. 2013, 9, e1003496. [Google Scholar] [CrossRef] [PubMed]
  112. Freedman, R.; Yu, R.; Sarkis, A.W.; Hedstrom, L. A structural determinant of mycophenolic acid resistance in eukaryotic inosine 5′-monophosphate dehydrogenases. Protein Sci. 2020, 29, 686–694. [Google Scholar] [CrossRef] [PubMed]
  113. Chew, W.S.; Wang, W.; Herr, D.R. To fingolimod and beyond: The rich pipeline of drug candidates that target S1P signaling. Pharmacol. Res. 2016, 113, 521–532. [Google Scholar] [CrossRef]
  114. Haarmann, T.; Rolke, Y.; Giesbert, S.; Tudzynski, P. Ergot: From witchcraft to biotechnology. Mol. Plant Pathol. 2009, 10, 563–577. [Google Scholar] [CrossRef]
  115. Ozery, M.; Wadhwa, R. StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  116. Endo, A. Monacolin K, a new hypocholesterolemic agent produced by a Monascus species. J. Antibiot. 1979, 32, 852–854. [Google Scholar] [CrossRef]
  117. Alberts, A.W.; Chen, J.; Kuron, G.; Hunt, V.; Huff, J.; Hoffman, C.; Rothrock, J.; Lopez, M.; Joshua, H.; Harris, E.; et al. Mevinolin: A highly potent competitive inhibitor of hydroxymethylglutaryl-coenzyme A reductase and a cholesterol-lowering agent. Proc. Natl. Acad. Sci. USA 1980, 77, 3957–3961. [Google Scholar] [CrossRef] [PubMed]
  118. Subhan, M.; Faryal, R.; Macreadie, I. Exploitation of Aspergillus terreus for the production of natural statins. J. Fungi 2016, 2, 13. [Google Scholar] [CrossRef] [PubMed]
  119. Gomes, N.G.M.; Lefranc, F.; Kijjoa, A.; Kiss, R. Can some marine-derived fungal metabolites become actual anticancer agents? Mar. Drugs 2015, 13, 3950–3991. [Google Scholar] [CrossRef] [PubMed]
  120. Tiwari, P.; Bose, S.K.; Park, K.-I.; Dufosse, L.; Fouillaud, M. Plant-microbe interactions under the extreme habitats and their potential applications. Microorganisms 2024, 12, 448. [Google Scholar] [CrossRef] [PubMed]
  121. Meyer, V.; Cairns, T.; Barthel, L.; King, R.; Kunz, P.; Schmideder, S.; Müller, H.; Briesen, H.; Dinius, A.; Krull, R. Understanding and controlling filamentous growth of fungal cell factories: Novel tools and opportunities for targeted morphology engineering. Fungal Biol. Biotechnol. 2021, 8, 8. [Google Scholar] [CrossRef] [PubMed]
  122. El-Gendi, H.; Saleh, A.K.; Badierah, R.; Redwan, E.M.; El-Maradny, Y.A.; El-Fakharany, E.M. A Comprehensive insight into fungal enzymes: Structure, classification, and their role in mankind’s challenges. J. Fungi 2021, 8, 23. [Google Scholar] [CrossRef] [PubMed]
  123. Cairns, T.C.; Zheng, X.; Zheng, P.; Sun, J.; Meyer, V. Moulding the mould: Understanding and reprogramming filamentous fungal growth and morphogenesis for next generation cell factories. Biotechnol. Biofuels 2019, 12, 77. [Google Scholar] [CrossRef] [PubMed]
  124. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the 30 years from 1981 to 2010. J. Nat. Prod. 2012, 75, 311–335. [Google Scholar] [CrossRef]
  125. Yuan, Y.; Cheng, S.; Bian, G.; Yan, P.; Ma, Z.; Dai, W.; Chen, R.; Fu, S.; Huang, H.; Chi, H.; et al. Efficient exploration of terpenoid biosynthetic gene clusters in filamentous fungi. Nat. Catal. 2022, 5, 277–287. [Google Scholar] [CrossRef]
  126. Wang, Q.; Zhong, C.; Xiao, H. Genetic Engineering of filamentous fungi for efficient protein expression and secretion. Front. Bioeng. Biotechnol. 2020, 8, 293. [Google Scholar] [CrossRef]
  127. Booking, S.P.; Wiebe, M.G.; Robson, G.D.; Hansen, K.; Christiansen, L.H.; Trinci, A.P.J. Effect of branch frequency in Aspergillus oryzae on protein secretion and culture viscosity. Biotechnol. Bioeng. 1999, 65, 638–648. [Google Scholar] [CrossRef]
  128. He, R.; Li, C.; Ma, L.; Zhang, D.; Chen, S. Effect of highly branched hyphal morphology on the enhanced production of cellulase in Trichoderma reesei DES-15. Biotech 2016, 6, 214. [Google Scholar] [CrossRef]
  129. Dymond, J.S.; Richardson, S.M.; Coombes, C.E.; Babatz, T.; Muller, H.; Annaluru, N.; Blake, W.J.; Schwerzmann, J.W.; Dai, J.; Lindstrom, D.L.; et al. Synthetic chromosome arms function in yeast and generate phenotypic diversity by design. Nature 2011, 477, 471–476. [Google Scholar] [CrossRef]
  130. Shao, Y.; Lu, N.; Wu, Z.; Cai, C.; Wang, S.; Zhang, L.-L.; Zhou, F.; Xiao, S.; Liu, L.; Zeng, X.; et al. Creating a functional single-chromosome yeast. Nature 2018, 560, 331–335. [Google Scholar] [CrossRef] [PubMed]
  131. Luo, J.; Sun, X.; Cormack, B.P.; Boeke, J.D. Karyotype engineering by chromosome fusion leads to reproductive isolation in yeast. Nature 2018, 560, 392–396. [Google Scholar] [CrossRef]
  132. Zheng, X.; Zheng, P.; Zhang, K.; Cairns, T.C.; Meyer, V.; Sun, J.; Ma, Y. 5S rRNA promoter for guide RNA expression enabled highly efficient CRISPR/Cas9 genome editing in Aspergillus niger. ACS Synthetic Biol. 2018, 8, 1568–1574. [Google Scholar] [CrossRef] [PubMed]
  133. Liu, R.; Chen, L.; Jiang, Y.; Zhou, Z.; Zou, G. Effcient genome editing in flamentous fungus Trichoderma reesei using the CRISPR/Cas9 system. Cell Discov. 2015, 1, 15007. [Google Scholar] [CrossRef] [PubMed]
  134. Liu, Q.; Gao, R.; Li, J.; Lin, L.; Zhao, J.; Sun, W.; Tian, C. Development of a genome editing CRISPR/Cas9 system in thermophilic fungal Myceliophthora species and its application to hyper-cellulase production strain engineering. Biotechnol. Biofuels 2017, 10, 1. [Google Scholar] [CrossRef]
  135. Katayama, T.; Tanaka, Y.; Okabe, T.; Nakamura, H.; Fujii, W.; Kitamoto, K.; Maruyama, J.I. Development of a genome editing technique using the CRISPR/Cas9 system in the industrial flamentous fungus Aspergillus oryzae. Biotechnol. Lett. 2016, 38, 637–642. [Google Scholar] [CrossRef]
  136. Pohl, C.; Kiel, J.A.K.W.; Driessen, A.J.M.; Bovenberg, R.A.L.; Nygard, Y. CRISPR/Cas9 based genome editing of Penicillium chrysogenum. ACS Synth. Biol. 2016, 5, 754–764. [Google Scholar] [CrossRef] [PubMed]
  137. Dijk, J.W.A.; Wang, C.C.C. Heterologous expression of fungal secondary metabolite pathways in the Aspergillus nidulans host system. Methods Enzymol. 2016, 575, 127–142. [Google Scholar] [PubMed]
  138. McLean, K.J.; Hans, M.; Meijrink, B.; van Scheppingen, W.B.; Vollebregt, A.; Tee, K.L.; van der Laan, J.-M.; Leys, D.; Munro, A.W.; Berg, M.A.v.D. Single-step fermentative production of the cholesterol-lowering drug pravastatin via reprogramming of Penicillium chrysogenum. Proc. Natl. Acad. Sci. USA 2015, 112, 2847–2852. [Google Scholar] [CrossRef] [PubMed]
  139. Steiniger, C.; Hofmann, S.; Mainz, A.; Kaiser, M.; Voigt, K.; Meyer, V.; Süssmuth, R.D. Harnessing fungal nonribosomal cyclodepsipeptide synthetases for mechanistic insights and tailored engineering. Chem. Sci. 2017, 8, 7834–7843. [Google Scholar] [CrossRef] [PubMed]
  140. Geib, E.; Brock, M. ATNT: An enhanced system for expression of polycistronic secondary metabolite gene clusters in Aspergillus niger. Fungal Biol. Biotechnol. 2017, 4, 13. [Google Scholar] [CrossRef]
  141. Jorgensen, T.R.; Goosen, T.; Hondel, C.; Ram, A.F.J.; Iversen, J.J.L. Transcriptomic comparison of Aspergillus niger growing on two different sugars reveals coordinated regulation of the secretory pathway. BMC Genomics 2009, 10, 44. [Google Scholar] [CrossRef] [PubMed]
  142. Wanka, F.; Cairns, T.; Boecker, S.; Berens, C.; Happel, A.; Zheng, X.; Sun, J.; Krappmann, S.; Meyer, V. Tet-On, or Tet-Of, that is the question: Advanced conditional gene expression in Aspergillus. Fungal Genet. Biol. 2016, 89, 72–83. [Google Scholar] [CrossRef] [PubMed]
  143. Schmideder, S.; Barthel, L.; Friedrich, T.; Thalhammer, M.; Kovačević, T.; Niessen, L.; Meyer, V.; Briesen, H. An X-ray microtomography-based method for detailed analysis of the three-dimensional morphology of fungal pellets. Biotechnol. Bioeng. 2019, 116, 1355–1365. [Google Scholar] [CrossRef] [PubMed]
  144. Tomer, A.; Singh, R.; Singh, S.K.; Dwivedi, S.A.; Reddy, C.U.; Keloth, M.R.A.; Rachel, R. Role of fungi in bioremediation and environmental sustainability. In Mycoremediation and Environmental Sustainability; Prasad, R., Nayak, S.C., Kharwar, R.N., Dubey, N.K., Eds.; Fungal Biology; Springer: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  145. Bisht, J.; Harsh, N.S.; Palni, L.M.; Agnihotri, V.; Kumar, A. Biodegradation of chlorinated organic pesticides endosulfan and chlorpyrifos in soil extract broth using fungi. Remediat. J. 2019, 29, 63–77. [Google Scholar] [CrossRef]
  146. Tiwari, P.; Muhammad, A.; Bae, H. Endophyte-mediated bioremediation—An efficient biological strategy in ecological subsistence and agriculture. In Endophytic and Arbuscular Mycorrhizal Fungi and Their Role in Sustainable Agriculture; Erwin, D.J., Ed.; Nova Publishers: Hauppauge, NY, USA, 2023; ISBN 979-8-88697-766-0. [Google Scholar]
  147. Tiwari, P.; Bae, H. Trends in harnessing plant Endophytic microbiome for heavy metal mitigation in plants: A perspective. Plants 2023, 12, 1515. [Google Scholar] [CrossRef]
  148. Alaux, P.-L.; Zhang, Y.; Gilbert, L.; Johnson, D. Can common mycorrhizal fungal networks be managed to enhance ecosystem functionality? Plant People Planet 2021, 3, 433–444. [Google Scholar] [CrossRef]
  149. Friesen, M.L.; Porter, S.S.; Stark, S.C.; von Wettberg, E.J.; Sachs, J.L.; Martinez-Romero, E. Microbially mediated plant functional traits. Annu. Rev. Ecol. Evol. Syst. 2011, 42, 23–46. [Google Scholar] [CrossRef]
  150. Ceci, A.; Pinzari, F.; Russo, F.; Persiani, A.M.; Gadd, G.M. Roles of saprotrophic fungi in biodegradation or transformation of organic and inorganic pollutants in co-contaminated sites. Appl. Microbiol. Biotechnol. 2019, 103, 53–68. [Google Scholar] [CrossRef] [PubMed]
  151. Verma, S.; Srivastava, J. Mycoremediation: A novel approach to rescue soil from heavy metal contamination. In Mycoremediation and Environmental Sustainability; Prasad, R., Nayak, S.C., Kharwar, R.N., Dubey, N.K., Eds.; Fungal Biology; Springer: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  152. Prasad, R. (Ed.) Mycoremediation and Environmental Sustainability, Volume 1; Springer: Cham, Switzerland, 2017. [Google Scholar]
  153. Prasad, R. (Ed.) Mycoremediation and Environmental Sustainability, Volume 2; Springer: Cham, Switzerland, 2018. [Google Scholar]
  154. Redman, R.S.; Kim, Y.O.; Woodward, C.J.; Greer, C.; Espino, L.; Doty, S.L.; Rodriguez, R.J. Increased fitness of rice plants to abiotic stress via habitat adapted symbiosis: A strategy for mitigating impacts of climate change. PLoS ONE 2011, 6, e14823. [Google Scholar] [CrossRef] [PubMed]
  155. Echeveria, L.; Gilmore, S.; Harrison, S.; Heinz, K.; Chang, A.; Nunz-Conti, G.; Cosi, F.; Singh, P.; Bond, T. Versatile Bio-Organism Detection Using Microspheres for Future Biodegradation and Bioremediation Studies. In Laser Resonators, Microresonators, and Beam Control XXII; Kudryashov, A.V., Paxton, A.H., Ilchenko, V.S., Armani, A.M., Eds.; SPIE Proceedings; SPIE: Bellingham, WC, USA, 2020; Volume 11266. [Google Scholar]
  156. Claus, H. Microbial degradation of 2,4,6-trinitrotoluene in vitro and in natural environments. In Biological Remediation of Explosive Residues; Singh, S.N., Ed.; Springer: Cham, Switzerland, 2014; pp. 15–38. [Google Scholar]
  157. Kumar, M.; Prasad, R.; Goyal, P.; Teotia, P.; Tuteja, N.; Varma, A.; Kumar, V. Environmental biodegradation of xenobiotics: Role of potential microfora. In Xenobiotics in the Soil Environment: Monitoring, Toxicity and Management; Hashmi, M.Z., Kumar, V., Varma, A., Eds.; Springer: Cham, Switzerland, 2017; pp. 319–334. [Google Scholar]
  158. Quintella, C.M.; Mata, A.M.; Lima, L.C. Overview of bioremediation with technology assessment and emphasis on fungal bioremediation of oil contaminated soils. J. Environ. Manag. 2019, 241, 156–166. [Google Scholar] [CrossRef] [PubMed]
  159. Lien, P.J.; Ho, H.J.; Lee, T.H.; Lai, W.L.; Kao, C.M. Effects of Aquifer Heterogeneity and Geochemical Variation on Petroleum-Hydrocarbon Biodegradation at a Gasoline Spill Site; Advanced materials research 1079; Trans Tech Publications Ltd.: Wollerau, Switzerland, 2015; pp. 584–588. [Google Scholar]
  160. Roccuzzo, S.; Beckerman, A.P.; Trögl, J. New perspectives on the bioremediation of endocrine disrupting compounds from wastewater using algae-, bacteria- and fungi-based technologies. Int. J. Environ. Sci. Technol. 2020, 18, 89–106. [Google Scholar] [CrossRef]
  161. Yadav, A.N.; Mishra, S.; Singh, S.; Gupta, A. (Eds.) Recent Advancement in White Biotechnology through Fungi; Springer: Cham, Switzerland, 2019. [Google Scholar]
  162. Singh, S.; Kumar, V.; Dhanjal, D.S.; Datta, S.; Bhatia, D.; Dhiman, J.; Samuela, J.; Prasad, R.; Singh, J. A sustainable paradigm of sewage sludge biochar: Valorization, opportunities, challenges and future prospects. J. Clean. Prod. 2020, 269, 122259. [Google Scholar] [CrossRef]
  163. Tyagi, M.; da Fonseca, M.M.; de Carvalho, C.C. Bioaugmentation and biostimulation strategies to improve the effectiveness of bioremediation processes. Biodegradation 2011, 22, 231–241. [Google Scholar] [CrossRef]
  164. Li, Y.; Fu, K.; Gao, S.; Wu, Q.; Fan, L.; Li, Y.; Chen, J. Increased virulence of transgenic Trichoderma koningi strains to the Asian corn borer larvae by over expressing heterologous chit42 gene with chitin-binding domains. J. Environ. Sci. Health 2013, 48, 376–383. [Google Scholar] [CrossRef]
  165. The Role of Fungi in Fighting Climate Change; and Why They Are At Risk 2021. Available online: https://www.cnbctv18.com/environment/explained-the-role-of-fungi-in-fighting-climate-change-and-why-they-are-at-risk-11664192.htm (accessed on 20 June 2024).
  166. Available online: https://www.weforum.org/agenda/2022/07/fungi-forests-carbon-climate (accessed on 20 June 2024).
  167. Tiwari, P.; Bajpai, M.; Singh, L.K.; Mishra, S.; Yadav, A.N. Phytohormones producing fungal communities: Metabolic engineering for abiotic stress tolerance in plants. In Agriculturally Important Fungi for Sustainable Agriculture; Gupta, V.K., Tuohy, M., Eds.; Springer: Cham, Switzerland, 2020; pp. 171–197. [Google Scholar]
  168. Tiwari, P.; Bajpai, M.; Singh, L.K.; Yadav, A.; Bae, H. Portraying fungal mechanisms in stress tolerance: Perspective for sustainable agriculture. In Springer-Nature Book on Recent Trends in Mycological Research, Vol 1: Agricultural and Medical Perspective; Springer: Berlin/Heidelberg, Germany, 2021; pp. 269–292. [Google Scholar]
  169. Niego, A.G.T.; Rapior, S.; Thongklang, N.; Raspé, O.; Hyde, K.D.; Mortimer, P. Reviewing the contributions of macrofungi to forest ecosystem processes and services. Fungal Biol. Rev. 2023, 44, 100294. [Google Scholar] [CrossRef]
  170. Boddy, L. Fungi, Ecosystems, and Global Change. In The Fungi, 3rd ed.; Academic Press: New York, NY, USA, 2016; pp. 361–400. [Google Scholar]
  171. Muhammad Adil, T.P.; Chen, J.-T.; Kanwal, S. Major bioactive metabolites and antimicrobial potential of Orchidaceae Fungal endophytes. In Advances in Orchid Biology, Biotechnology, and Omics; Tiwari, P., Chen, J., Eds.; Springer Publishers: Berlin/Heidelberg, Germany, 2023. [Google Scholar]
  172. Li, G.; Jian, T.; Liu, X.; Lv, Q.; Zhang, G.; Ling, J. Application of metabolomics in fungal research. Molecules 2022, 27, 7365. [Google Scholar] [CrossRef]
  173. Ijoma, G.N.; Heri, S.M.; Matambo, T.S.; Tekere, M. Trends and applications of omics technologies to functional characterization of enzymes and protein metabolites produced by fungi. J. Fungi 2021, 7, 700. [Google Scholar] [CrossRef] [PubMed]
  174. Caesar, L.K.; Butun, F.A.; Robey, M.T.; Ayon, N.J.; Gupta, R.; Dainko, D.; Bok, J.W.; Nickles, G.; Stankey, R.J.; Johnson, D.; et al. Correlative metabologenomics of 110 fungi reveals metabolite–gene cluster pairs. Nat. Chem. Biol. 2023, 19, 846–854. [Google Scholar] [CrossRef] [PubMed]
  175. Cao, L.; Zhang, Q.; Miao, R.; Lin, J.; Feng, R.; Ni, Y.; Li, W.; Yang, D.; Zhao, X. Application of omics technology in the research on edible fungi. Curr. Res. Food Sci. 2022, 6, 100430. [Google Scholar] [CrossRef]
  176. Lange, L. Fungal enzymes and yeasts for conversion of plant biomass to bioenergy and high-value products. Microbiol. Spectr. 2017, 5, 1029–1048. [Google Scholar] [CrossRef]
  177. Mycorena. Creating Green Protein with No Plants. Available online: https://mycorena.com/ (accessed on 20 June 2024).
  178. Attias, N.; Danai, O.; Abitbol, T.; Tarazi, E.; Ezov, N.; Pereman, I.; Grobman, Y.J. Mycelium bio-composites in industrial design and architecture: Comparative review and experimental design. J. Clean. Prod. 2020, 246, 119037. [Google Scholar] [CrossRef]
  179. Lange, L. The importance of fungi and mycology for addressing major global challenges. IMA Fungus 2014, 5, 463–471. [Google Scholar] [CrossRef]
  180. Novoa, C.; Dhoke, G.V.; Mate, D.M.; Martinez, R.; Haarmann, T.; Schreiter, M.; Eidner, J.; Schwerdtfeger, R.; Lorenz, P.; Davari, M.D.; et al. Know Volution of a fungal laccase toward alkaline pH. ChemBioChem 2019, 20, 1458–1466. [Google Scholar] [CrossRef]
Figure 1. Sustainable development goals (SDGs) and role of fungal biotechnologies.
Figure 1. Sustainable development goals (SDGs) and role of fungal biotechnologies.
Jof 10 00506 g001
Figure 2. Development of a bio-based economy via production and utilization of high-value products from fungi [9].
Figure 2. Development of a bio-based economy via production and utilization of high-value products from fungi [9].
Jof 10 00506 g002
Table 1. Representative examples of ‘high-value’ products from fungi (natural and engineered strains) and their potential to achieve SDGs.
Table 1. Representative examples of ‘high-value’ products from fungi (natural and engineered strains) and their potential to achieve SDGs.
Fungal SpeciesHigh-Value ProductBiotechnological/
Economic Utilities
References
Fungal high-value metabolites in medicinal applications
Acremonium chrysogenumβ-lactam antibiotics
(cephalosporins)
Pharmaceutical value[31]
Lentiana edodesLentinanAs chemotherapy adjuvant in healthcare[32]
Penicillium rubens
P. solitum
P. chrysogenum
Penicillin
Mevastatin
β-lactam antibiotics
(penicillins)
Pharmaceutical value
Statins are widely used in lowering blood
cholesterol levels
Antibiotics in healthcare
[33,34,35]
Saccharomyces boulardiiProbioticsHealth supplements[36]
Aspergillus terreus
A. niger
Secondary metabolites
(lovastatin)
Secondary metabolites
(enniatins)
Human granulocyte
colony-stimulating factor
(G-CSF)
Galactaric acid
Pharmaceutical value

Enhanced production of high-value metabolites

High protein titre for medicinal applications


Efficiently produce galactaric acid for industrial applications
[37,38,39,40]
Rhizoctonia bataticolaForskolinAnti-HIV, anti-tumor, therapeutic application[41]
Phomopsis sp.QuinineAntimalarial, used in malaria treatment[42]
Alternaria sp.DigoxinCardiotonic, therapeutic application[43]
Fungal species in food industries/food applications
Blakeslea trisporaCarotene
Lycopene
Food pigments for application in food sector[44,45]
Monascus ankaMonascus pigmentsFood pigments as natural food colorants[46]
S. cerevisiaeLycopene (carotenoid)
Ethanol
Production of fatty
acid-derived biofuels
Terpene production
Food pigment for use in food sector
Biofuel production
Industrial applications

Genetic engineering for enhanced terpene production
[47,48,49,50]
Morchella esculentaPolysaccharidesAs food (nutritional) supplement[51]
Pleaurotus eryngiiPork sausage (food component)Used as food component[52]
Fusarium venenatum
Fusarium sp.
Quorn (meat substitute)

Dairy-free cream cheese
Nutritional food (high amino acid and fiber, fungal
protein)
Food industries
[53,54]
Penicillium camemberti
P. roquefortii
Production of cheese
Blue cheese
Food industries
---
[55,56]
Mushroom myceliumPlant-based baconAlternative food product[57]
Aspergillus sp.
A. oryzae
A. sojae
A. niger
Fermented meat
Soy sauce
Miso
Jiuqu
Citric acid
Enzymes
Alternative meat source, high protein content
Traditional fermented food
---
Food industries
[58,59,60,61]
Yarrowia lipolyticaβ-carotenoidHigh metabolite yield for food sector application[62]
Xanthophyllomyces dendrorhousZeaxanthinFood pigment usage in food industry[63]
Mortierella alpinaLinoleic and oleic acidsFood industry[64]
L. edodesPasta (functional food)Nutritional supplements[65]
Fungal metabolites for industrial applications
Ustilago maydisItaconic acidBio-based building block in the polymer industry, pharmaceutical value[66]
Kluyveromyces lactisL-ascorbic acid (vitamin C)Enhanced production for industrial applications[67]
Trichoderma reeseiEnzyme (cellulase)Enhanced production for industrial applications[68]
Schizophyllum communeTextilesIndustrial application[69]
Ganoderma lucidumComposite material, construction materialBiomaterials to reduce environmental pollution[70]
Umbelopsis isabellinaConstituents of biodiesel (polyunstaturated fatty acids)Biofuel production, energy source[71]
Fungal metabolites for agricultural applications
A. nidulansInsecticides (austinoids)Production of austenoid derivatives including
7-hydroxydehydroaustin,
1,2-dihydro-7-hydroxydehydroaustin,
1,2-dehydro-precalidodehydroaustin, calidodehydroaustin, etc.
[72]
Beauveria bassianaMycoinsecticidesIntegrated pest management, biocontrol of arthropod pests[73,74]
Trichoderma spp.
T. harzianum T22
T. harzianum TC39
Auxin-like metabolites,
proteinaceous compounds
Azaphilone, harzianolide, 1-hydroxy-3-methylanthraquinone and harzianopyridone
Regulate plant growth and development,
agricultural applications
Biocontrol agents, suppress the growth of plant pathogens
[75,76,77]
Gliocladium virensAntifungal compounds gliovirin, viridiol, valinotrocin, viridin, gliotoxin, and heptelidic acidProtect agricutural crops from multiple pathogens, bicontrol functions[78]
Botrytis cinereaAbscisic acidPhytohormone regulates abiotic stresses, application in agriculture[79]
Chaetomium globosum Cg-7,
C. globosum Cg-6
C. globosum Cg-5
ChaetoglobosinReduce post-harvest diseases in multiple fruits[80]
Eupenicillium parvumAzadirachtin A and BFor the control of insects, biocontrol functions[81]
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

Tiwari, P.; Park, K.-I. Advanced Fungal Biotechnologies in Accomplishing Sustainable Development Goals (SDGs): What Do We Know and What Comes Next? J. Fungi 2024, 10, 506. https://doi.org/10.3390/jof10070506

AMA Style

Tiwari P, Park K-I. Advanced Fungal Biotechnologies in Accomplishing Sustainable Development Goals (SDGs): What Do We Know and What Comes Next? Journal of Fungi. 2024; 10(7):506. https://doi.org/10.3390/jof10070506

Chicago/Turabian Style

Tiwari, Pragya, and Kyeung-Il Park. 2024. "Advanced Fungal Biotechnologies in Accomplishing Sustainable Development Goals (SDGs): What Do We Know and What Comes Next?" Journal of Fungi 10, no. 7: 506. https://doi.org/10.3390/jof10070506

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