Next Article in Journal
Assessment of Growth, Yield, and Nutrient Uptake of Mediterranean Tomato Landraces in Response to Salinity Stress
Next Article in Special Issue
The Effects of Accompanying Ryegrass on Bayberry Trees by Change of Soil Property, Rhizosphere Microbial Community Structure, and Metabolites
Previous Article in Journal
Eugenol: In Vitro and In Ovo Assessment to Explore Cytotoxic Effects on Osteosarcoma and Oropharyngeal Cancer Cells
Previous Article in Special Issue
Essential Oils and Antagonistic Microorganisms as Eco-Friendly Alternatives for Coffee Leaf Rust Control
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fertilization of Microbial Composts: A Technology for Improving Stress Resilience in Plants

1
Xianghu Laboratory, Hangzhou 311231, China
2
Institute of Biotechnology, Zhejiang University, Hangzhou 310058, China
3
Department of Bioinformatics and Biotechnology, Government College University, Faisalabad 38000, Pakistan
4
Department of Environmental Sciences, Government College University, Faisalabad 38040, Pakistan
5
Department of Plant Breeding and Genetics, University of Agriculture, Faisalabad 38000, Pakistan
6
MEU Research Unit, Middle East University, Amman 11831, Jordan
7
Institute of Eco-Environmental Protection, Shanghai Academy of Agricultural Sciences, Shanghai 201403, China
8
Department of Botany and Microbiology, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
9
Department of Life Sciences, College of Science and Art Mahyel Aseer, King Khalid University, Abha 62529, Saudi Arabia
10
Department of Biology, College of Science, King Khalid University, Abha 61421, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Plants 2023, 12(20), 3550; https://doi.org/10.3390/plants12203550
Submission received: 28 August 2023 / Revised: 28 September 2023 / Accepted: 9 October 2023 / Published: 12 October 2023
(This article belongs to the Special Issue Pathogenesis and Disease Control in Crops—2nd Edition)

Abstract

:
Microbial compost plays a crucial role in improving soil health, soil fertility, and plant biomass. These biofertilizers, based on microorganisms, offer numerous benefits such as enhanced nutrient acquisition (N, P, and K), production of hydrogen cyanide (HCN), and control of pathogens through induced systematic resistance. Additionally, they promote the production of phytohormones, siderophore, vitamins, protective enzymes, and antibiotics, further contributing to soil sustainability and optimal agricultural productivity. The escalating generation of organic waste from farm operations poses significant threats to the environment and soil fertility. Simultaneously, the excessive utilization of chemical fertilizers to achieve high crop yields results in detrimental impacts on soil structure and fertility. To address these challenges, a sustainable agriculture system that ensures enhanced soil fertility and minimal ecological impact is imperative. Microbial composts, developed by incorporating characterized plant-growth-promoting bacteria or fungal strains into compost derived from agricultural waste, offer a promising solution. These biofertilizers, with selected microbial strains capable of thriving in compost, offer an eco-friendly, cost-effective, and sustainable alternative for agricultural practices. In this review article, we explore the potential of microbial composts as a viable strategy for improving plant growth and environmental safety. By harnessing the benefits of microorganisms in compost, we can pave the way for sustainable agriculture and foster a healthier relationship between soil, plants, and the environment.

1. Introduction

In recent years, global food security has been increasingly threatened by the combination of population increase and the scarcity of limited arable land worldwide [1,2]. Consequently, the need to boost crop productivity has become a significant challenge in order to meet the demands of a continuously increasing global population [3]. To enhance crop productivity, various chemical fertilizers have been extensively employed worldwide; however, this widespread use has led to the deterioration of both human health and environmental ecology with significant severity [4,5]. Biofertilizers offer a promising solution in order to counteract the harms associated with chemical fertilizers. They are an essential component of integrated nutrient management, contributing significantly to crop output and food security in the agricultural sector [6,7]. Biofertilizers consist of one or more beneficial microbes capable of colonizing the interior or exterior of plants after soil application, as seeds, or as direct exposure on the plants. This colonization facilitates the enhancement of nutrient supply to the host plant, effectively promoting its growth [8,9]. These microbe-containing fertilizers improve the soil fertility through different mechanisms including atmospheric nitrogen fixation, solubilization of unavailable nutrients (such as phosphate, zinc, potassium, and iron), and synthesis of phytohormones [10]. Thus, the biofertilizers improve soil fertility and crop yield by harvesting the natural biological system of nutrients and recycling them [11]. The efficiency and efficacy of biofertilizers has remained controversial in terms of their application in the field and their potential to replace the chemical fertilizers, mainly due to the heterogeneous nature of soil, adaptability to harsh ecological conditions, and unsuitable formulations and carrier materials [12]. Thus, if a nutrient-rich carrier material with potential to facilitate microbial growth is used, outcomes from biofertilizers can be substantially enhanced and the use of chemical fertilizers can be completely or partially cut down. Nowadays, as a result of intensive agricultural activities, a huge amount of organic waste is generated by agricultural fields, and the disposal of this poses difficulties [13].
Production of nutrient-rich compost from agricultural waste has been emerging as an alternative way of disposal [14]. A dark brown or black earthy matter, rich in micro- and macronutrients and produced as a result of aerobic decomposition of biodegradable waste is known as compost [15]. Compost not only enhances the soil fertility in terms of micro- and macronutrients, it also improves the soil architecture by improving water- and air-holding capacity for better root growth. Moreover, compost can play an important role in the bioeconomy because its production does not depend on finite inputs [16]. This technique of waste disposal has proved to be very economic, as it enhances plant nutrient uptake, soil organic matter content, crop yield, and soil biophysical parameters [13,17]. The application of compost has been proven beneficial, as it has good impact on environment and soil quality [18].
Composting is the process in which organic materials like food scraps, leaves, twigs, lawn clippings, wood waste, and other organic matter are decayed by the soil microorganisms under controlled ambiance [15,19]. To carry out the decomposition process, either a pit can be dug in the ground, or the process can be conducted in special vessels called compost bins [20]. Certain factors, including temperature, proper aeration, pH, the quality and quantity of feedstock, etc., must be considered before initializing the process of composting to make it more efficient and reliable [15]. Another key factor to be considered is the ratio between carbon and nitrogen (C/N), which plays an essential function in the composting process. The application of compost enhances the physico-chemical properties of the soil and also exerts a positive impact on the soil’s microbial diversity [21]. The widespread adoption of compost in agriculture faces constraints due to its extended time of action and comparatively reduced nutrient supply to crops when compared with chemical fertilizers [17]. Various reports have been published on the composting of agricultural feedstock and its potential application in improving soil fertility and plant growth [22,23]. Moreover, the potential role of different phytobeneficial microbial communities to the process of composting has also been elucidated by several researchers around the world [24,25].
This review examines the feasibility and advantages of producing microbial compost-based biofertilizers using agricultural feedstock enriched with decomposing phytobeneficial microorganisms. Compost serves as an ideal carrier material for potential plant-beneficial microorganisms, presenting a possible alternative to chemical fertilizers. The application of microbial compost in fertilization holds immense promise for bolstering soil sustainability and elevating overall plant productivity. Additionally, we explore the factors influencing the microbial composting process and highlight the future prospects of adopting microbial compost-based biofertilizers.

2. Biofertilizers and Their Advantages

Agricultural productivity is decreasing continuously due to nutrient deficiency in the soils and the growth of obnoxious weeds and pests [26]. Conversely, the production cost has been raised over the last two decades. As a result of these factors, the growth rate in agriculture is falling behind the rapid pace of population growth [27,28]. The population explosion around the globe makes the use of chemicals fertilizers for improving crop production in order to meet food requirements inevitable [29]. Frequent soil tillage, chemical fertilizers, and narrow crop rotations are some of the intensive agricultural techniques that increase production in a short time. However, over time, these techniques have led to a decline in soil organic carbon, soil aggregation strength, and biodiversity, consequently reducing the productivity of field crops. Additionally, they contribute to air and groundwater pollution [30,31]. In this scenario, the biofertilizers have a great potential to improve soil fertility, thus reducing the need for the application of chemical fertilizers. Biofertilizer-mediated agricultural practices also result in crops with improved yield [32,33].
Biofertilizers utilize beneficial microorganisms to optimize plant growth by increasing nutrient supply, effectively enhancing overall nutrient availability, and promoting healthier and more robust plant development [34,35]. Generally, the potential of different beneficial microbes, including nitrogen fixers, phosphorus solubilizers, potassium solubilizers, iron mobilizers, as well as the microbes capable of producing phytohormones, is utilized during biofertilizer synthesis which improves the nutrient profile of the soil [36,37]. Subsequently, the microbe-oriented nutrient recycling through biofertilizers enhances the soil organic matter content and maintains the soil health and sustainability, which is ultimately followed by healthy plant growth. Several bacterial species, known as plant-growth-promoting rhizobacteria (PGPR), such as Azotobacter spp., Escherichia coli, Pseudomonas spp., and Bacillus spp., and arbuscular mycorrhizal fungi (AMF), such as Glomus versiforme, Aspergillus awamori, Glomus macrocarpum, and Sclerocystis coremioides, are often exploited in broad spectrum biofertilizers without causing any negative chemical influence on the soils [38,39,40]. Moreover, co-inoculation of PGPR and AMF also results in enhanced plant growth, nutrient uptake, disease tolerance, and resistance to abiotic stress. For example, co-culture of Rhizobium, Azotobacter, and vesicular arbuscular mycorrhiza (VAM) as biofertilizer enhanced the straw and grain yield when applied to wheat plants along with rock phosphate as a phosphate source [41]. Similarly, a mixed culture comprising Thiobacillus thioxidans, Bacillus subtilis, and Saccharomyces sp. was found capable of converting micronutrients into soluble forms such as Mn, Zn, Fe, etc., and making them available to plants [33]. Trichoderma spp. are also known to improve the tolerance of plants to biotic and abiotic stresses by producing enzymes that can detoxify harmful chemicals and by increasing the production of stress-response proteins in the plant [42]. For example, Trichoderma harzianum inoculation improved tomato plants’ tolerance to chilling stress by enhancing physiological, biochemical, and molecular responses [43]. The biofertilizers that achieve nitrogen fixing, i.e., potassium- and phosphate-solubilizing bacterial strains, significantly improve the growth, production, and qualitative characteristics of food crops [44]. Hence, these microbial-based fertilizers can significantly contribute to the establishment of sustainable agriculture systems, playing a pivotal role in achieving this goal.
Overall, the utilization of biofertilizers has been in practice for a considerable period due to their eco-friendly nature and superior cost-efficiency compared with chemical fertilizers. The biofertilizers can also be used to convert complex organic materials into simple compounds, followed by a change in the color and texture of the soil [45,46]. They have the capability to enhance the crop yield by 25–30% and can help the soil fight against dehydration and other soil-borne illnesses [15]. Therefore, to promote agricultural productivity, it is essential to acknowledge the application of biofertilizers [47,48].

3. Microbial Compost

Agricultural waste is now emerging as a compelling and cost-effective resource that can provide organic matter and essential plant nutrients to the soil, effectively supporting crop production [49]. However, raw organic waste cannot be directly applied because it is unsuitable for land and agricultural crops [50]. Hence, composting stands out as one of the most favorable, straightforward, and cost-effective methods employed to treat this type of wastes [51,52]. Compost can be locally produced on the farm and is an attractive technique of waste disposal. It is can recover valuable plant nutrients and improve the soil’s biophysical characteristics, soil organic matter, and crop yield [53,54]. As the nutrients in compost are slowly released depending upon the microbial biomass, the gradual nutrition availability for the plants is ensured [55]. Thus, the utilization of microbial compost aids in preventing nutrients leaching into groundwater and significantly increases soil productivity in the long term. The soil quality can be enhanced by repeated compost applications, which result in the enriching of microorganisms that are beneficial to the soil, an increased total carbon content, an enhanced cation exchange capacity, and a reduction in the abundance of plant-parasitic nematodes [56]. The use of pesticides and herbicides also decreases due to the improved plant resistance against diseases as a result of compost application [57]. The application of compost, along with certain soil microorganisms such as PGPR and AMF, has been proven to effectively boost soil fertility and health [58].

3.1. The Composting Process

The process of controlled decomposition of organic material like agricultural residues is known as composting [59]. A composting process requires the synergistic action of the natural forces that decompose the organic waste into the organic fertilizer to occur in a safe way [60]. The primary raw materials utilized for composting include agricultural waste such as fruit and vegetable waste, domestic kitchen residues, various crop residues (e.g., stover, cobs, and leaves), and different types of manure, e.g., cattle and poultry [61]. Both aerobic and anaerobic bacteria are involved in the composting process. For better a understanding of the degradation processes that occur in compost, characterization and identification of microorganisms is necessary in the composting process. Common PGPR, like Azotobacter spp., Pseudomonas spp., Escherichia coli, and Bacillus spp., and some AMF, such as Aspergillus awamori, Glomus macrocarpum, and Sclerocystis coremioides, were found to be involved in the decomposition process [62,63]. These microorganisms utilize amino acids, sugars, and lipids present in the feedstock as their energy source [64,65]. Composting offers numerous benefits beyond providing a significant release of various nutrient elements for plants. It contributes to soil conditioning, facilitates efficient manure handling, reduces the risks associated with different pollutants and weed seeds, and promotes pathogen destruction through a high-temperature composting processes [66]. A schematic diagram of the composting process has been presented in Figure 1.

3.2. The Biochemistry of Composting

The compost feedstock is often rich in organic compounds that are rich in components like carbon (C), hydrogen (H), oxygen (O2), and nitrogen [67]. The lignin, sugars, fats, cellulose, and proteins that are important components of agricultural raw material facilitate the decomposition of organic compounds during the process of composting. These components become progressively more oxidized and form molecules with more functional groups that have, however, a lower molecular weight during the aerobic decomposition [68]. The product obtained after the decomposition of organic matter is known as humus. In addition, the biodegradation occurs under both anaerobic and aerobic circumstances, but the composting process proceeds excellently under aerobic conditions [69]. Many microbes can perform their function in the absence of O2, but the process of anaerobic respiration is less energy-efficient and it utilizes chemical species like sulfate, carbon dioxide, nitrates, sulfur, and oxidized metal ions. Hence, to suppress the development of anaerobic conditions, artificial aerobic conditions should be introduced in the compost pile through pulling or pushing the air [70,71].

3.3. Microorganisms Involved in Plant Growth and the Composting Process

The composting process can be characterized as the biodegradation of organic waste into useful products under controlled ambiance. These products are applied to the soil, and they effectively enhance the physico-chemical properties of the soil [72,73]. The biodegradation of organic waste is accelerated by diverse microbial species belonging to different microbial groups, namely, bacteria, actinomycetes, and filamentous fungi. The bacterial species belonging to the genera Pseudomonas, Bacillus, Paenibacilus, and Enterobacter were found to be the most abundant microorganisms, having a very high population density of 3.0 × 108 CFU/g throughout the composting process, followed by actinomycetes (mainly the species of genus Streptomyces, Nocardia, Micromonospora, Thermomonospora Dactylosporangium, and Kibdelosporangium). The members of filamentous fungi that drive the composting process mainly belong to the genus Aspergillus and have a low population density (i.e., 1.2–1.6 × 108 CFU/g) [72].
A group of specific microorganisms which positively influences plant growth are PGPR and AMF [48,74]. A number of PGPR and AMF, along with their potential roles in plant improvement, are presented in Table 1. These microbes belong to many genera, including Xanthomonas, Agrobacterium, Streptomyces, Alcaligenes, Cellulomonas, Arthrobacter, Amorpho sporangium, Bacillus, Pseudomonas sp., Azotobacter, Actinoplanes, Rhizobium, Erwinia, Bradyrhizobium, Enterobacter, Rhizophagus irregularis, and Glomus intraradices [6]. Generally, the microorganisms that are used as BF are potassium solubilizers, phosphorus solubilizers, nitrogen fixers, and phytohormone producers [75]. Nitrogen-fixing microbes convert atmospheric N2 to ammonia [76]. Some microorganisms belonging to the Ectorhizospheric strains and Endosymbiotic rhizobia have been defined as efficient phosphate solubilizers [7,77]. The most potent strains from bacterial genera that solubilize phosphorus are Pseudomonas, Bacillus, Enterobacter, and Rhizobium genera [9,78]. Similarly, numerous PGPR species belonging to different genera, namely, Bradyrhizobium, Pseudomonas, Rhizobium, Agrobacterium, Klebsiella, Enterobacter, Azotobacter, and Bacillus, are best known for their phytohormones production potential. These microbe-oriented phytohormones have an effective role in plant growth stimulation.
The ability of PGPR to solubilize potassium rock by secreting organic acids has also been investigated. A number of bacterial strains, including Burkholderia sp., Bacillus mucilaginosus, Ferrooxidans sp., Paenibacillus sp., Pseudomonas sp., Bacillus edaphicus, and Acidothiobacillus sp., are PGPR that solubilize potassium-bearing minerals, thereby releasing the potassium in the form that is available for plants [79]. Therefore, utilizing PGPR as biofertilizers to enhance agriculture can reduce the dependence on agrochemicals and promote sustainable crop production (Figure 2).
Table 1. Some PGPR and AMF strains involved in plant growth promotion.
Table 1. Some PGPR and AMF strains involved in plant growth promotion.
ChemicalsMicroorganismsBeneficial EffectsReferences
Direct Mechanism
Nitrogen fixationBradyrhizobium japonicum, Glomus macrocarpum, Azotobacter vinelandii, Bacillus, Rhizobium, Beijerinckiaderxii, Klebsiella pneumoniae, Enterobacter cloacae, Citrobacterfreundii, and Pseudomonas putidaThe conversion of atmospheric N2 into plant-utilizable forms triggers improvement in plant development and yield[80,81,82,83,84,85]
Phosphate solubilizationPenicillium brevicompactum, Aspergillus niger, Pseudomonas striata, Enterobacter, Erwinia, Bacillusmegaterium, Ochrobactrumanthropi, Bacilus, Beijerinckia, Burkholderia, Rhizobium, and SerratiaSolubilizing the inorganic phosphorus from insoluble compounds and making them available to the plants[86,87,88,89,90]
Potassium solubilizationAspergillus niger, Aspergillus terreus, Acidothiobacillus sp., Bacillus edaphicus, Ferrooxidans sp., Bacillus mucilaginosus, Pseudomonas sp., Burkholderia sp., and Paenibacillus sp.Solubilizing potassium rock by producing and secreting organic acids and making them available to the plants for growth and development[79,91,92,93]
Zinc mobilizationBeauveria caledonica, Hymenoscyphus ericae, Oidiodendron maius Pennisetum glaucum, Gluconacetobacter diazotrophicus, fluorescent pseudomonads, and Bacillus sp.Solubilizing the insoluble Zn into soluble form and hence having efficient role in plant growth and development[94,95,96,97,98,99]
Production of phytohormonesPaecilomyces formosus, Asprgillus fumigatus, Fusarium proliferatum Azotobacter, Arthrobacter, Azospirillum, Pseudomonas, Bacillus, Acinetobacter, Flavobacterium, Enterobacter, Micrococcus, Agrobacterium, Clostridium, Rhizobium, and XanthomonasPlay an important role as regulators of growth and development of plants[100,101,102,103,104,105]
Siderophore productionAspergillus fumigatus, Glomus etunicatum, Glomus mossae, Trichoderma spp., Pseudomonas fluorescens, Rhodococcus,
Acinetobacte, and Pseudomonas putida
Solubilize and sequester iron from the soil and then provide it to the plant cells[86,106,107,108,109]
Exopolysaccharides productionAzotobacter vinelandii, Bacillus drentensis, Enterobacter cloacae, Rhizobium sp., Agrobacterium sp., and Xanthomonas sp.Plays a pivotal role in increasing the number of soil macropores, aggregating rhizospheric soil particles, and maintaining water potential[110,111]
Indirect Mechanisms
Hydrogen cyanideAlcaligenes, Aeromonas, Rhizobium, Pseudomonas, and Bacillus sp.Powerful inhibitor of many metal enzymes, especially copper-containing cytochrome Coxidases[112,113,114]
ACC deaminase activityGigaspora rosea, Achromobacter, Azospirillum, Pseudomonas, Enterobacter, Bacillus, and RhizobiumPlants were able to tolerate environmental stresses by keeping a normal amount of ethylene in their root zone[115,116,117]
Induced systemic resistanceTrichoderma virens, Pseudomonas, and Bacillus spp.Induced resistance is the state of an enhanced defensive ability developed by plants when appropriately stimulated[118,119,120,121]
Production of vitaminsGlomus aggregatum, Glomus viscosum, Azotobacter vinelandii, Azospirillum brasilense, Azospirillum spp., Pseudomonas fluorescens, Rhizobium leguminosarum, Rhizobium etli, Sinorhizobium meliloti, Mesorhizobium loti, and Bacillus subtilisFacilitate the production of essential compounds for plants and bacteria, induce resistance against pathogens, and directly promote plant growth[122,123,124]
Production of protective enzymesAspergillus niger, Glomus spp., Bacillus, Burkholderia, Enterobacter, Pseudomonas, Serratia, and StaphylococcusMay have a dramatic effect on the cycling of nutrients such as phosphorus, nitrogen, and sulfur[125,126,127]
Production of antibioticsPseudomonas, Bacillus, and AzotobacterPrevent the detrimental effects of pathogens on plants through production of inhibitory substances[128,129]
Volatile organic compound (VOCs)Bacillus amyloliquefaciens, Bacillus subtilis, Pseudomonas fluorescens, Bacillus mojavensis, Trichoderma spp., Trametes gibbosa, and Trametes versicolorVOCs have a significant role in the plant growth promotion and suppression of plant diseases[130,131,132,133]

4. The Formulation of the Microbial Compost-Based Biofertilizers

Enriching compost with nutrients, beneficial bacteria, and AMF is a key strategy for improving its nutritional value and enhancing its positive effects on plant growth [134]. By establishing a mutualistic relationship, soil microorganisms and plants work together to facilitate nutrient uptake without disrupting the overall physico-chemical or biological equilibrium of the system. Endophytic PGPR, for example, resides within the plant roots, positively impacting the plants through the secretion of phytohormones, nitrogen fixation, enhanced phosphorus uptake, and the solubilization of inorganic phosphates. This cooperative partnership promotes healthier plant growth and fosters a sustainable ecosystem [135,136]. In general, supplementation of compost with different types of nutrients as well as its bioaugmentation with potential PGPR and AMF strains can lead to value addition in BF technology. Thus, such compost-based biofertilizers need to be added in future long-term farming strategies in order to make the farm yield more sustainable and cost-effective. For this purpose, the compost can be developed through natural processes followed by adding the inoculum of known PGPR or AMF strain(s) at an optimal population density. However, the survival of PGPR or AMF strain(s) in the compost and their synergistic effect, along with natural microflora of compost, are vital factors, which need to be assessed before developing compost-based biofertilizers.

5. Feedstock for Microbial Compost-Based Biofertilizers

The number of global agricultural products is rising steadily despite the global decrease in cultivated land area, due to the increased demand to produce food to feed an increasing population [137,138]. The United Nation’s Food and Agriculture Organization (FAO) predicted that the need for food will rise rapidly as the population of the world will surpass eight billion by 2030 [139]. Agricultural waste and residue will also increase as the food production increases [140]. The physiochemical properties and composition of the final compost are mainly dependent on the agricultural residues or waste which are used as a feedstock. Many essential characteristics of the compost, including the C-to-N ratio, available micronutrients (e.g., Fe, Mg, Mo) and macronutrients (e.g., N, P, K, Ca) for the plants, as well as the texture and structure of the compost, depend on the method of composting and the type of feedstock used [66]. The feedstock used in composting varies remarkably from area to area and season to season depending on the type of agricultural produce. Various classes of agriculture feedstocks being used for composting include cereal residues, animal manure, fruit and vegetable waste, and grassland residues [140,141].

5.1. Cereal and Grassland Residues

Cereal residues encompass the materials remaining in croplands after crop harvest, comprising stems, leaves, straw, and seedpods. Additionally, process residues like husk, roots, bagasse, and seeds are the by-products remaining after the processing of cereal crops [142]. The harvesting of cereals generates a substantial amount of agricultural residues, which are frequently underutilized as livestock feedstock or for other purposes. For instance, straw is one of the primary agricultural residues left after harvesting several cereal crops [143]. Due to its low nutritional value, straw is not considered a suitable livestock feedstock, leading to the common practice of leaving it in the fields to enhance the soil’s carbon stock. Alternatively, straw is used for livestock housing and bedding, as well as in various horticultural applications [144]. Despite being used for livestock housing, straw requires periodic disposal. However, it has been found to be a suitable substrate for anaerobic digestion. To achieve the highest compost yields, mechanical, thermochemical, or biological pre-treatment of the straw is necessary to reduce digestion retention times compared with other substrates. Major agricultural residues like rice straw, wheat straw, corn stover, and sugarcane bagasse are abundant and readily available for compost production, offering valuable opportunities for sustainable waste management and resource utilization in agricultural systems [145].
About 70% of the world’s agricultural land and 26% of the world’s total land area is comprised of grasslands [146,147]. In terms of purpose and harvesting type, grassland is usually classified as permanent grassland and temporary grassland. The European agricultural sector is very much dependent on temporary grasslands; they are the most significant origin of feed for livestock production. Since the use of grass silage as animal feed is decreasing, grass is receiving more attention as an alternative feedstock for composting [61].

5.2. Fruit and Vegetable Waste

Fruits and vegetables are essential natural sources of minerals, dietary fibers, and vitamins in both developing and developed countries. The fruit and vegetable production is continuously improving in many developing countries in, for example, South America and Asia [148]. The largest source of vegetables, accounting for more than 61% of the global output, is Asia. Similarly, according to an estimate, about 30% of the world’s fruit is supplied by India, Brazil, and China [149]. Items that are extracted from vegetables and fruits during processing, cleaning, cooking, and packaging are considered vegetable and fruit solid waste. Many of these by-products are either abandoned or landfilled. The major fruit and vegetable waste products which can be used for preparing compost include twigs, pomace, skins, rinds, cores, pulp, stems, pits, leaves, seeds, peels, and spoiled fruits and vegetables [150].

5.3. Animal Manure

Animal manure and slurry cause environmental pollution due to the improper recycling of them [151]. The composting of animal manure is an efficient strategy to manage the environmental risks related to animal manure. The composting process converts the animal manure into high-quality products that can be utilized for numerous agricultural purposes [152]. Various factors like low porosity, high humidity, low C/N ratio, and high pH affect the manure composting process. Thus, for the production of quality compost, efficient management techniques for manure composting is a necessary pre-requisite. In order to obtain a quality product in less time and at a lower cost, various different aeration techniques, bulking agents, substrate-conditioning feed stocks, and other available options have been used for the composting of manure [153,154]. The bulking agent added during composting of animal manure improves the mineralization rate of the substrate by optimizing its properties like porosity, humidity content, pH, particle density, C/N ratio, etc. Generally, for the composting of nitrogen-rich waste (such as animal manure), agricultural and forestry by-products are used as bulking agents. The most commonly used materials are cotton waste, hay, and cereal straw [155]. The high carbon content of these bulking agents increases the low C/N value of animal manures [156].

6. Plant Growth Promotions Mediated by Microbial Compost-Based Biofertilizers

Microorganisms are an integral part of soil nutrient cycling in soil–plant–microbe systems. Plant growth and development is facilitated by the soil microbial community in several different ways [157]. Generally, microorganisms promote plant growth by improving nutrient mobilization and availability through a number of mechanisms such as nitrogen fixation, organic compounds mineralization, mineral nutrients solubilization, phytohormone production, etc. [158,159]. Some rhizospheric microbes improve plants’ physiology and resistance against phytopathogens indirectly by producing substances that act synergistically with the native immune system of a plant to neutralize the harmful effects of pathogens [160,161]. Through several activities, the microorganisms either enhance the growth of plants or suppress the disease. The major microbial direct and indirect mechanisms involved in plant growth promotion are presented in Table 2 and briefly described below.

6.1. Direct Mechanisms of Microorganisms’ Influence on Plants

6.1.1. Nitrogen Fixation

Nitrogen (N) is an essential plant nutrient. Significant energy is needed to break the triple bonds between the two N-atoms of atmospheric molecular nitrogen (N2), because the plants are unable to use the atmospheric molecular nitrogen (N2) directly [205,206]. Plants normally use inorganic forms of nitrogen like NO3− and NH4+ as well as low-molecular-weight dissolved organic nitrogen (DON), especially amino acids. Plants cannot utilize the atmospheric dinitrogen (N2) unless some diazotrophic microorganisms reduce it to more usable form like ammonia (NH3) [207]. In biological nitrogen fixation (BNF), nitrogenase enzyme present in diazotrophic microorganisms catalyzes and converts the atmospheric N2 to NH3. BNF provides more than 2 × 1013 g nitrogen annually throughout the world. Symbiotic fixation adds about 80% of this amount, while the remaining is contributed by either free-living or associative nitrogen-fixing systems [208,209]. Bacteria and fungi are the only organisms capable of converting atmospheric N2 into a form that plants can utilize. Symbiotic and associative symbiotic interactions with plant roots are accomplished by different AMF species including Glomus macrocarpum, Glomus hoi, and Glomus mosseae [82,83,84] and several bacterial species like Pseudomonas spp., Alcaligenes spp., Azotobacter spp., Azospirillum spp., Enterobacter spp., Arthrobacter spp., Acinetobacter spp., Bacillus spp., Bradyrhizobium japonicum, Serratia (Brady) rhizobium spp., and Burkholderia spp. [210,211]. Compost consisting of nitrogen-fixing microbial strains showed improved nitrogen metabolism. It has been reported that the species of genus Gordonia exhibits nitrogenase reductase activity and was found to be involved in the nitrogen metabolic reaction of compost [212]. Thus, adding nitrogen-fixing PGPR to compost can enhance the growth of plants by improving the nitrogen availability to plants.

6.1.2. Phosphate Solubilization

Phosphorus (P) is the second most crucial nutrient for plant growth, and its scarcity can severely limit crop production to a critical level [213]. To address P deficiency in soils, agricultural fields are commonly treated with various types of chemical fertilizers. However, a major portion of P fertilizers is fixed in soils and becomes unavailable for plants. For example, tropical and subtropical soils come under the classification of phosphorous-deficient soils due to their high pH [214]. One of the economical ways to deal with P deficiency in soils is to focus on rock phosphate sources in the soil and make them available for plants. Various microorganisms have the ability to solubilize fixed forms of P such as rock phosphate. Rhizobacterial species reported to solubilize phosphorous include B. megatheriumand, Enterobacter, Erwinia, O. anthropi TRS-2, and Pseudomonas striata [86]. Such bacteria make the soil rich with soluble inorganic phosphate by decomposing phosphate-rich organic compounds. Enterobacter, Serratia, Azotobacter, Beijerinckia, Bacillus, Pseudomonas, Burkholderia, Microbacterium, Erwinia, Rhizobium, and Flavobacterium are amongst the most important bacterial genera that have been reported to solubilize phosphate [36]. Rivera-Cruz et al. (2008) revealed that mixing of P-solubilizing microbial strains with compost improved plant physiology and the physical and microbiological characters of the soil [215]. These modified biofertilizers have been reported to enhance the available P content of the soil [216,217].

6.1.3. Potassium Mobilization

Potassium (K) is an essential macronutrient and is required for plant growth. The concentration of soluble potassium is usually low in soil because more than 90 percent of the potassium exists in the form of silicate minerals and insoluble rock. Potassium deficiency is also one of the major constraints for an optimal crop production [218,219]. The plants suffering from potassium deficiency often suffer from poor root development, low seed production, slower growth, and smaller yields [220]. In order to cope with potassium deficiency without using mineral fertilizers, one of the possible ways is to focus on the alternative endemic sources of potassium present in unavailable rock forms in the soils [221]. Several soil microbes have been isolated and characterized for their potential to solubilize potassium rock and to make it available to plants through the production and secretion of organic acids. Several PGPR, including Paenibacillus sp., Pseudomonas sp., Bacillus edaphicus, Acidothiobacillus sp., Burkholderia sp., Bacillus mucilaginosus, and Ferrooxidans sp., have been reported to release potassium in an accessible form by solubilizing potassium-bearing minerals in soils [79]. Similarly, AMF such as Aspergillus terreus and Aspergillus niger have demonstrated the capability to solubilize potassium and produce acids by utilizing feldspar and potassium aluminum silicate as insoluble K sources [93]. Hence, by exploiting the potential of potassium-solubilizing microorganisms, an eco-friendly crop production can be supported by reducing the use of agrochemicals. Grigiz MGZ (2006) concluded that the combination of compost and K-mobilizing bacterial strains (Bacillus sp. UBFBa4 and UBFBa7) improves the growth of wheat by increasing the uptake of K. The compost generally caused an increase in the population density of K-mobilizers, which convert the insoluble K into plant-useable forms by lowering the pH through organic acid production [222].

6.1.4. Zinc Solubilization

Zinc (Zn) is an important nutrient required by the plant tissues in small amounts for growth and gene regulation [223]. Increasing the zinc concentration in crop plants is a major challenge However, due to the relatively low solubility and decreased availability of Zn in agricultural soils, there is a prevailing universal Zn deficiency in crops [223]. A prediction reveals that about 50% of the total Asian agricultural soils are deficient in zinc [224]. The deficiency of Zn is because of several reasons depending on the soil condition; for example, the solubility of Zn reduces with an increase in pH, high magnesium-to-calcium ratio, high availability of P and Fe, and variability in organic matter and bicarbonate content [97,225]. Using Zn fertilizers in the fields can overcome Zn deficiency, but generally, chemical fertilizers are expensive and have unfavorable effects on the environment. Thus, there should be an eco-friendly and cost-effective approach to overcoming the Zn deficiency. Thus, PGPR like Bacillus sp., Gluconacetobacter diazotrophicus, fluorescent pseudomonads, and Pennisetum have been reported to increase Zn solubility in soil [96,97,98,99]. Similarly, some AMF like Beauveria caledonica, Hymenoscyphus ericae, and Oidiodendron maius have been found to be Zn solubilizers [94,95]. Compost combined with Zn-solubilizing microbial strains was found to improve the crop productivity by increasing Zn mineralization and uptake in plants. Shanmugam PM (2000) reported that the application of green manure compost amended with Zn-solubilizing PGPR enhances the growth and physiology of rice plants. Thus, modification of compost by inoculating Zn-solubilizing microorganisms is a viable strategy for improving the bioavailability of this micronutrient to plants to achieve sustainable agricultural production [226].

6.1.5. Production of Phytohormones

Plant growth hormones are organic compounds utilized by plants as messengers to interact with and respond to their environment [227]. These hormones are very efficient if manufactured in a minute quantity; however, they may result in plant growth inhibition if found in a greater amount [228]. These plant growth hormones can elicit physiological actions that influence growth and fruit ripening, as they are produced in one part of the plant and transported to other parts [229]. In addition to plants, several PGPR and AMF have demonstrated the ability to produce growth hormones, i.e., Azospirillum, Enterobacter, Pseudomonas, Bacillus, Azotobacter, Paecilomyces formosus, Asprgillus fumigatus, and Fusarium proliferatum. Amongst the phytohormones, indole-3-acitic acid (IAA) plays a significant role in the growth and division of cells and increases the growth of lateral roots in plants [103,104,228,230]. IAA functions as a signal molecule in plant development, contributing to processes such as organogenesis. Moreover, it plays a crucial role in gene regulation and cell division, elongation, and expansion. Phytohormones are manufactured by the bacteria that solubilize phosphate [231,232]. Various bacterial strains with the potential to produce IAA in pure cultures and soils, along with their interactions with plants roots, have been widely studied [230,233].
IAA has a role in organogenesis, gene regulation, cell division, cell elongation, and cell expansion [234,235]. It is required not only for phytostimulation but is also used by various microorganisms for their interaction with plants due to its significance in bacterial colonization of plant roots. It directly influences bacterial physiology, as it functions as a signaling molecule for bacteria as well [236]. Sometimes, IAA production by the microbial strains is triggered in the presence of tryptophan, which serves as a precurser for its production [237]. For example, a free-living nitrogen-fixing bacterium, Azospirillum, is distinguished in terms of its production of a significant concentration of IAA to stimulate plant growth [238]. Active IAA-producing PGPR are often isolated and characterized from the plant rhizospheres and re-inoculated with plants for plant benefits [239]. For example, when eucalyptus cuttings were grown on a substrate that was inoculated with IAA-producing rhizobacteria, a notable development in root proliferation and root dry matter was observed [240]. It has been documented that the integrated use of organic manure compost and phytohormones-producing biofertilizers improves the growth parameters (such as leaf number and plant height) of onions [241]. Both compost and biofertilizers synergistically produce plant growth regulators (i.e., IAA and gibberellins) and influence the plant growth in more favorable ways.

6.1.6. Siderophore Production

Siderophores are low-molecular-weight substances which have chelation properties and are primarily produced and employed by microorganisms to fulfil the nutritional requirements for iron [242]. Siderophores have the ability to form complexes with ferric ion through their ligands, improve its solubilization, and enable its removal from natural complexes or from minerals [243]. Iron hydroxide, being poorly soluble, remains unavailable to biological systems and, at physiological pH under aerobic conditions, the unstable ferrous (Fe2+) form of iron is oxidized and converted into ferric (Fe3+), which is a relatively more stable form [244,245]. When siderophores are released into an environment, they solubilize the iron by forming a ferric–siderophore complex, which can move through the diffusion process and reach the cell surface [246]. After the ferric–siderophore complexes are formed, different Gram-positive and Gram-negative bacteria recognize these complexes, and their transport systems start working. Several PGPR and AMF, including Pseudomonas fluorescens, Rhodococcus, Acinetobacte, Pseudomonas putida, Aspergillus fumigatus, Glomus etunicatum, Glomus mossae, Trichoderma spp., and Aspergillus niger, have been reported to harbor the potential for production of siderophores [86,106,107,108]. Microbial siderophores are categorized into four major classes based on the types of ligands and fundamental features of functional groups involved in iron complexation. These classes include carboxylates, pyoverdines, phenol catecholates, and hydroxamates [247]. The application of biofertilizers in 50 and 75% compost enhances the plant growth significantly. The biofertilizers added to compost exhibit siderophore-producing activity along with other plant-growth-promoting characters [248]. The compost with siderophore-producing microbial strains has a better ability to facilitate plant Fe uptake and also increase the microbial density of antagonistic bacteria involved in plant growth promotion.

6.1.7. Exopolysaccharides Production

Exopolysaccharides (EPSs) are biodegradable polymers usually with a high molecular weight. Various bacteria, algae, and plants have been reported to produce these extrapolymeric substances [249]. EPSs play a pivotal role in increasing the number of soil macropores, aggregating rhizospheric soil particles, and maintaining water potential, thus helping plants ameliorate different stress conditions including salinity, drought, or water logging [250]. The EPS-producing rhizobacterial species include Bacillus drentensis, Azotobacter vinelandii, Enterobacter cloacae, Rhizobium sp., Agrobacterium sp., and Xanthomonas sp. [251]. The combination of compost and EPS-producing PGPR can improve soil fertility by improving soil porosity and structure and can also provide plants with the ability to tolerate various stress conditions.

6.2. Indirect Mechanisms of Microorganisms’ Influence on Plants

6.2.1. Hydrogen Cyanide Production

Hydrogen cyanide (HCN) is a poisonous chemical which is produced by the soil microbes and has a low molecular weight and antifungal properties [112]. HCN is produced through the interaction of glycine with the HCN synthetase enzyme, which is located on the plasma membrane of specific rhizobacteria [252]. Until now, many microbial genera such as Rhizobium, Aeromonas, Alcaligenes, etc., have been reported to harbor the potential for HCN production in the rhizospheric soil and plant root nodules [253]. Many studies have revealed the disease-controlling effects of HCN, e.g., inhibition of “root-knot” and black rot in tomato and tobacco [254,255]. However, a few studies have also indicated the harmful effects of HCN on the metabolism and growth of root cells in different plants. For example, HCN produced by Pseudomonas in the rhizosphere was found to hinder the primary growth of roots in Arabidopsis because of the suppression of an auxin-responsive gene [255,256]. The use of HCN-synthesizing PGPR strains in conjunction with compost can be a viable approach to protecting plants from soil-borne phytopathogens and to improving plant growth and physiology with greater efficiency.

6.2.2. ACC Deaminase Production

Plants require a suitable quantity of ethylene for their growth and development. However, its high concentration can influence the plant cellular processes and can hinder growth [257]. The level of ethylene in plants is also regulated by the soil microbes. For example, the ethylene concentration in the root region of Arabidopsis thaliana was found to be managed by soil microbes using their 1-amino-cyclopropane-1-carboxylic acid (ACC) deaminase, which stops ACC from acting in ethylene biosynthesis pathways [258]. Some PGPR and AMF strains, including Azospirillum, Enterobacter, Bacillus, Achromobacter, Rhizobium, Pseudomonas, and Gigaspora rosea, have the potential to produce ACC deaminase, an enzyme responsible for the cleavage of ACC into ammonia and α-ketobutyrate [115,116,117]. For example, Ghosh et al. (2003) achieved a higher root length in Brassica compestris through inoculation with three bacillus spp. (Bacillus circulans DUC1, Bacillus firmus DUC2, and Bacillus globisporus DUC3) carrying ACC deaminase activity. In another study, the shoot dry biomass was enhanced in Brassica napus after the inoculation with Pseudomonas asplenii harboring the ACC deaminase gene [259]. Therefore, PGPR with ACC deaminase activity boost plant biomass in a stressed environment like pathogenicity, high temperature, contaminants drought, salinity, waterlogging, etc. [260]. Plants showed an increase in root length and biomass when treated with compost containing ACC deaminase, producing plant-beneficial bacteria [261].

6.2.3. Induced Systemic Resistance

Many non-pathogenic rhizobacteria and beneficial fungi have the ability to overcome diseases by producing a defense mechanism known as “Induced Systemic Resistance” (ISR) [262]. ISR is a state of increased basal resistance of the plant, which relies upon two signaling molecules (i.e., salicylic acid and jasmonic acid) [263]. Pathogens show variability in their sensitivity against signaling molecules that elicit resistance. The same strain shows antagonism against many phytopathogens. Bacillus spp. and Pseudomonas are the most examined rhizobacteria that activate ISR [119,264]. The treatment of crop plants with resistance-inducing and antagonistic PGPR and AMF leads to more effective biocontrol strategies for making the cropping systems better [264]. Several PGPR and AMF, such as Trichoderma virens, Pseudomonas, and Bacillus spp., have been used to induce systemic resistance in plants [120,121,265]. Thus, the fortification of compost by using ISR-inducing microbes can suppress numerous diseases in plants caused by soil-inhabiting pathogens. This blending of antagonistic microbes with compost could be an efficient strategy for improving plants’ basal resistance response against harmful microbial communities.

6.2.4. Production of Vitamins

Microbes that produce vitamins are widespread in the rhizosphere. Vitamins are either obtained from root exudates or synthesized from rhizospheric bacteria and fungi [266,267]. In general, the vitamins secreted by plant roots are not exclusively synthesized by the plants themselves. Instead, they are produced by microorganisms and then taken up by the plants. Subsequently, the plants release these vitamins along with root exudates [123]. The composition of many vitamins like biotin, pyrroloquinoline quinone, niacin, thiamine, and pantothenic acid plays an important role in microbial chemotaxis [268]. Due to the frequent presence of microbial-produced vitamins in the rhizosphere and the adaptation of plants to be able to uptake microbially-synthesized vitamins, such vitamins are considered to play an important role in rhizosphere interactions [123,124]. The synthesis of thiamine and many other vitamins has been reported to be caused by numerous potential microbes such as Glomus aggregatum, Glomus viscosum, Azospirillum brasilense, Rhizobium leguminosarum, Azotobacter vinelandii, Azospirillum spp., Bacillus subtilis, Sinorhizobium meliloti, Rhizobium etli, Mesorhizobium loti, and Pseudomonas fluorescens [122,123,124]. The primary function of vitamins includes the induction of defenses against pathogens, being a cofactor in different metabolic pathways, the promotion of plant growth, and the production of essential compounds for plant growth [118]. The microbes harboring the potential to synthesize vitamins and compost may promote the proliferation and colonization of plant-beneficial microbes in the root zone that will subsequently enhance plant growth through the aforementioned mechanisms.

6.2.5. Production of Protective Enzymes

Plant microorganisms possess the capacity to produce certain proteins that promote and enhance plant development and growth. Root symbionts, such as ectomycorrhizal fungi and other rhizosphere microorganisms, play a vital role in facilitating these processes [269]. The enzymes are utilized for defense mechanisms, but some play role in cell wall degradation, like hydrolytic enzymes. The synthesis of enzymes comprises cellulase, chitinase, protease, b-1, 3-glucanase, and lipase, but some fungal cells may be lost during their production [270]. Phytase, peroxidase, phenylalanine ammonia-lyase, and polyphenol oxidase are examples of the most exceptional defense enzymes [271]. Many PGPR and AMF can produce phytase enzymes that have the ability to mineralize the phytates, including beneficial species that produce protective enzymes, which belong to the Burkholderia, Staphylococcus, Enterobacter, Bacillus, Pseudomonas and Serratia, Aspergillus niger, and Glomus spp. [125,255]. In the rhizosphere, phosphatases, arylsulfatases, proteases, and other extracellular enzymes tend to exhibit higher activity compared with the bulk soil. This heightened enzymatic activity significantly impacts the cycling of nutrients, such as phosphorus, nitrogen, and sulfur [272]. Pseudomonas, Bacillus, and Serratia can provide numerous extracellular lytic enzymes including b-1, 3 glucanases, chitinases, cellulases, laminarinase, and proteases [273]. Protective enzymes producing microbes when mixed with compost may increase the soluble sugar level for plant uptake and utilization along with other nutrients (i.e., N and P). Thus, such modified forms of compost will help plants grow in stressed environments.

6.2.6. Production of Antibiotics

The most significant antagonistic activity for combating phytopathogens is the generation of diverse antibiotics by microbes [274]. Several PGPR have been reported to produce different types of antibiotics including oligomycin A, viscosinamide, xanthobaccin, pyoluteorin, pyrrolnitrin, and 2,4 diacetyl phloroglucinol (2,4-DAPG) [255]. P. fluorescens BL915 strains produce the pyrrolnitrin that can stop the destruction of Rhizoctonia solani during the damping off of the cotton plant [275]. Pseudomonads can synthesize the 2,4-Dthat, which is a very active and extensively studied antibiotic, and can cause damage to the membrane of Pythium spp. [276]. Moreover, Pseudomonads produce the phenazine exhibiting redox activity, through which it can suppress the pathogens of plants such as Gaeumannomyces graminis and F. oxysporum [262]. The antibiotics generated by the majority of Bacillus spp. (circulin, colistin, and polymyxin) are very effective against Gram-negative and Gram-positive bacteria, and against many other pathogenic fungi as well [277]. In another study, different bacterial strains such as Pseudomonas, Bacillus, and Azotobacter exhibited a broad-spectrum antifungal activity on Muller–Hinton medium against different fungal pathogens such as Aspergillus, Fusarium, and Rhizoctonia bataticola [278]. Compost and antibiotic-producing bacteria can synergistically relieve plants from devastating microbes. Compost may facilitate the growth of beneficial microorganisms, which, in turn promote plant growth by retarding the growth of plant pathogens via their antimicrobial activity.

6.2.7. Volatile Organic Compounds

Soil microorganisms generate different types of metabolites during their metabolic activities [279]. The production of various volatile organic compounds (VOCs) by microbial metabolites has garnered significant attention from the global scientific community. These VOCs are the outcomes of the primary and secondary metabolism of soil microorganisms [133]. During primary metabolism, soil microbes generate VOCs as by-products while breaking down food to extract the necessary nutrients required for cell maintenance. In contrast, during secondary metabolism, VOCs are produced by microbes as a response to resource competition in nutrient-poor environments. Recently, several bacterial and fungal strains including Pseudomonas fluorescens, Bacillus mojavensis, and Trichoderma sp. have been reported to produce different types of VOCs [130,132,133]. Microbial VOCs have been extensively documented for their roles as antibacterial, antifungal, and antinematode agents, as well as for promoting plant growth. Additionally, these VOCs function as signaling molecules for cell-to-cell communication between different species. For instance, Bacillus subtilis and Bacillus amyloliquefaciens produce 2,3-butanediol and acetoin, which play a significant role in promoting plant growth [280]. Other plant-growth-promoting VOCs, like 2-pentylfuran, 13-tetradecadien-1-ol, 2-butanone, and 2-methyl-n-1-tridecene, have been identified in various bacterial strains [281]. VOCs can also trigger induced systemic resistance in plants when faced with pathogen challenges. Notably, volatile compounds such as HCN, dimethyl sulfide, and inorganic volatiles have been found to inhibit growth or have phytotoxic effects [130]. On the other hand, fungi-based VOCs act as location cues for host selection in fungivorous arthropods. For instance, 1-octen-3-ol produced by the wood-rotting white rot fungus Trametes gibbosa attracts fungus-eating beetles (Coleoptera) [131]. Another white rot species, Trametes versicolor, emits sesquiterpenes like d-cadinene, b-guaiene, isoledene, and g-patchoulene, which attract fungivorous beetles in behavioral experiments [131]. Moreover, plants can be protected from biotic stress by treating them with compost supplemented with VOC-producing microbes. The compost acts as a carrier, facilitating the colonization of plant-growth-promoting microbes in the rhizospheric region, subsequently protecting the plant from pathogen attacks by producing VOCs.

7. Factors Affecting the Composting Process

The composting process is influenced by several parameters that need to be monitored throughout to ensure the production of high-quality compost. The mechanism of composting is affected by various physico-chemical characteristics of the agricultural feedstocks, including porosity, temperature, oxygen levels, C/N ratio, pH value, proper aeration, and moisture content [15,156,282]. Some of the factors affecting the composting process are described herein.

7.1. Temperature and pH Value

Temperature is a key factor which significantly affects the growth and activity of the functional microbial communities in the composting of organic mass (OM). The OM transformation into compost takes place between thermophilic and mesophilic environments. The thermophilic conditions trigger the decomposition of organic waste. The catabolic activities of aerobes help in establishing the ambiance favorable for the decomposition process by shooting up the temperature to 65–70 °C within a few days [283,284]. Hassen et al. (2001) described that the decrease in temperature during composting is associated with the decrease in bacterial population [285]. The pH for the optimal decomposition of the organic waste should also be determined. Conventional composting requires a pH value in the range of 6 to 8 to decompose biomass optimally [286].

7.2. Carbon/Nitrogen Ratio

The carbon and nitrogen (C/N) ratio has a major impact on organic waste decomposition during composting. Every organism has a fixed C/N ratio at tissue level. This ratio has a crucial role in deciding the pathway for microbe-mediated decomposition of OM. Carbon serves as an energy source for microorganisms, and hence, it is the basic structural unit of life, whereas nitrogen is an important constituent of proteins, nucleic acids, and other cellular components [287]. Hence, microbes have to be supplemented with carbon-rich materials and nitrogenous sources in a balanced manner. The optimum C/N ratio for rapid composting is 25–35:1. As the carbon content exceeds the optimum amount, the decomposition process becomes progressive [288]. On the other hand, an increase in nitrogen concentration results in unpleasant odor [15]. The regular turning of windrow piles helps in maintaining the C/N ratio by removing the residual nitrogen in ammoniacal form [289].

7.3. Quality and Quantity of Incoming Raw Material

The chemical profile of the raw material is very important in the composting process. The low-end product recovery is achieved using organic-content-depleted biomass [156]. Therefore, periodic chemical analysis is required to ensure the quality of the feedstock. The composition of biomass varies as the source varies. Hence, the organic-fraction-rich raw material from fish, vegetable, and fruit markets should be utilized for the composting process [290].

7.4. Proper Aeration and Turning

As the biochemical transformation is facilitated by aerobes, proper aeration is needed to ensure oxygen availability. Proper aeration is linked with the periodic turning of the feedstock piles [291]. Turning helps in the smooth running of aerobic compositing processes by evenly distributing the oxygen to every component of the compost pile. Therefore, turning biomass heaps is important in order to maintain adequate aeration [15]. It has been reported that the aerobic composting process is stimulated when a huge amount of oxygen is provided to the project site [292].

7.5. Moisture Content

Moisture level is another important factor for the bioconversion process. The optimal range of moisture content is 50 to 55% for organic waste composting [13]. The presence of adequate moisture is crucial for the biochemical activities of microbes facilitating the degradation process. As the moisture exceeds the optimal amount, the environment turns anaerobic as the air spaces are occupied by water, thus reducing oxygen availability to aerobes, whereas a low moisture level will result in the killing of microorganisms [156]. Periodic turning of compost piles not only adjusts the moisture levels but also chops down the large organic matter cakes, and hence, provides microbes with more surface area to attack the material, resulting in a better decomposition process [293].

8. Conclusions and Future Directions

In conclusion, as the agricultural industry faces challenges posed by escalating production costs and a growing global population, it is evident that a paradigm shift is needed in the fertilizer sector. The era of the green revolution, once dominated by chemical fertilizers, is gradually fading due to their exorbitant costs and environmental implications. In response to these concerns, the exploration of alternative technologies, such as compost and rhizospheric microorganisms, for sustainable agricultural practices has gained momentum, offering a profitable and eco-friendly solution. Globally, extensive research has been undertaken to investigate the potential of bacterial, cyanobacterial, and fungal biofertilizers for various crops. The application of these biofertilizers has demonstrated positive effects on soil fertility and crop yields in arable lands. Beyond improving nutrient availability, these microorganisms contribute to plant growth through the production of beneficial plant hormones, induction of stress resistance, and biocontrol of plant pathogens. While their performance may vary under different field conditions, the use of microbial fertilizers is anticipated to surge in the foreseeable future due to their economic and ecological benefits.
Looking ahead, the technology of microbial fertilizers can be further enhanced by incorporating sustainable materials as carriers. Utilizing compost as a carrier material, given its sustainability and inherent fertilizing properties, has the potential to elevate biofertilizer technology to the next level. Moreover, in-depth research is required to identify compatible microbial strains capable of thriving in compost and demonstrating optimal physiological performance therein. The application of genome engineering holds potential as a tool for developing superior microbial strains, further advancing the efficacy of biofertilizers. The concept of compost-based biofertilizers represents a novel approach, fostering a unique setting for plant–microbe interactions by harnessing the combined effects of compost and characterized microbial strains. Compost, when combined with a microbial consortium, offers an alternative nutrient source for agricultural plants, while also exhibiting antagonistic effects against soil-borne phytopathogens and inducing tolerance against biotic and abiotic stresses. Furthermore, compost-based biofertilizers not only positively impact soil fertility, but also uphold their safety for use, owing to their eco-friendly nature. Overall, the continuous exploration and development of compost-based biofertilizers, alongside advancements in microbial strain selection and genome engineering, promise to revolutionize modern agriculture, making it more sustainable, economically viable, and environmentally responsible. Embracing these innovations will pave the way for a greener, more productive, and resilient agricultural future.

Author Contributions

T.A.: Conceptualization, Investigation, Validation, Resources, and Writing—original draft. M.N.: Validation, Investigation, and Writing—review and editing. M.S., H.A.M. and S.H.: Investigation and Writing—original draft. Y.Q., H.M.A. and L.X.: Writing—review and editing, Resources, and Validation. S.N. and A.F.E.-K.: Writing—review and editing, Resources, and Validation. B.L., Y.Y. and X.Q.: Supervision, Writing—review and editing, Resources, Validation. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by the Shanghai Agriculture Applied Technology Development Program, China (Grant No.X2021-02-08-00-12-F00760), Hangzhou Science and Technology Development Plan Project (202003A05) and Xianghu laboratory special research funds.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the large groups program under grant number RGP2/232/44.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahanty, T.; Bhattacharjee, S.; Goswami, M.; Bhattacharyya, P.; Das, B.; Ghosh, A.; Tribedi, P. Biofertilizers: A potential approach for sustainable agriculture development. Environ. Sci. Pollut. Res. 2017, 24, 3315–3335. [Google Scholar] [CrossRef]
  2. Suryanto, S.; Trinugroho, I.; Susilowati, F.; Aboyitungiye, J.B.; Hapsari, Y. The Impact of Climate Change, Economic Growth, and Population Growth on Food Security in Central Java Indonesia. Nat. Environ. Pollut. Technol. 2023, 22, 1017–1022. [Google Scholar] [CrossRef]
  3. Ahmed, T.; Noman, M.; Gardea-Torresdey, J.L.; White, J.C.; Li, B. Dynamic interplay between nano-enabled agrochemicals and the plant-associated microbiome. Trends Plant Sci. 2023. [Google Scholar] [CrossRef] [PubMed]
  4. Soleimani, H.; Mansouri, B.; Kiani, A.; Omer, A.K.; Tazik, M.; Ebrahimzadeh, G.; Sharafi, K. Ecological risk assessment and heavy metals accumulation in agriculture soils irrigated with treated wastewater effluent, river water, and well water combined with chemical fertilizers. Heliyon 2023, 9, e14580. [Google Scholar] [CrossRef]
  5. Khatun, J.; Intekhab, A.; Dhak, D. Effect of uncontrolled fertilization and heavy metal toxicity associated with arsenic (As), lead (Pb) and cadmium (Cd), and possible remediation. Toxicology 2022, 477, 153274. [Google Scholar] [CrossRef]
  6. Mohammadi, K.; Sohrabi, Y. Bacterial biofertilizers for sustainable crop production: A review. ARPN J. Agric. Biol. Sci. 2012, 7, 307–316. [Google Scholar]
  7. Shahid, M.; Hameed, S.; Imran, A.; Ali, S.; van Elsas, J.D. Root colonization and growth promotion of sunflower (Helianthus annuus L.) by phosphate solubilizing Enterobacter sp. Fs-11. World J. Microbiol. Biotechnol. 2012, 28, 2749–2758. [Google Scholar] [CrossRef] [PubMed]
  8. Malusá, E.; Sas-Paszt, L.; Ciesielska, J. Technologies for beneficial microorganisms inocula used as biofertilizers. Sci. World J. 2012, 2012, 491206. [Google Scholar] [CrossRef]
  9. Shahid, M.; Hameed, S.; Tariq, M.; Zafar, M.; Ali, A.; Ahmad, N. Characterization of mineral phosphate-solubilizing bacteria for enhanced sunflower growth and yield-attributing traits. Ann. Microbiol. 2015, 65, 1525–1536. [Google Scholar] [CrossRef]
  10. Mazid, M.; Khan, T.A. Future of bio-fertilizers in Indian agriculture: An overview. Int. J. Agric. Food Res. 2015, 3, 10–23. [Google Scholar] [CrossRef]
  11. Pandey, J.; Singh, A. Opportunities and constraints in organic farming: An Indian perspective. J. Sci. Res. 2012, 56, 47–72. [Google Scholar]
  12. Wang, Y.; Liu, Z.; Hao, X.; Wang, Z.; Wang, Z.; Liu, S.; Tao, C.; Wang, D.; Wang, B.; Shen, Z. Biodiversity of the beneficial soil-borne fungi steered by Trichoderma-amended biofertilizers stimulates plant production. NPJ Biofilms Microbiomes. 2023, 9, 46. [Google Scholar] [CrossRef]
  13. Ahmad, R.; Jilani, G.; Arshad, M.; Zahir, Z.A.; Khalid, A. Bio-conversion of organic wastes for their recycling in agriculture: An overview of perspectives and prospects. Ann. Microbiol. 2007, 57, 471–479. [Google Scholar] [CrossRef]
  14. Moretti, S.M.L.; Bertoncini, E.I.; Abreu-Junior, C.H. Composting sewage sludge with green waste from tree pruning. Sci. Agric. 2015, 72, 432–439. [Google Scholar] [CrossRef]
  15. Baruah, S.; Srivastava, S. Production and Assesment of Compost as a Bio-Fertilizer Using Kitchen Waste from NIT Rourkela Hostels and Study of Comparative Plant Growth. Ph.D. Thesis, National Institute of Technology, Rourkela, India, 2015. [Google Scholar]
  16. Ingrao, C.; Bacenetti, J.; Bezama, A.; Blok, V.; Geldermann, J.; Goglio, P.; Koukios, E.G.; Lindner, M.; Nemecek, T.; Siracusa, V. Agricultural and forest biomass for food, materials and energy: Bio-economy as the cornerstone to cleaner production and more sustainable consumption patterns for accelerating the transition towards equitable, sustainable, post fossil-carbon societies. J. Clean. Prod. 2016, 117, 4–6. [Google Scholar] [CrossRef]
  17. Singh, A.; Manna, M. Solid Waste Management Through Vermi-and Phosphate-Enriched Composting Techniques for Sustainable Agriculture. In Water Quality Management; Springer: Berlin/Heidelberg, Germany, 2018; pp. 399–406. [Google Scholar]
  18. Oyetunji, O.; Bolan, N.; Hancock, G. A comprehensive review on enhancing nutrient use efficiency and productivity of broadacre (arable) crops with the combined utilization of compost and fertilizers. J. Environ. Manag. 2022, 317, 115395. [Google Scholar] [CrossRef] [PubMed]
  19. Qian, S.; Zhou, X.; Fu, Y.; Song, B.; Yan, H.; Chen, Z.; Sun, Q.; Ye, H.; Qin, L.; Lai, C. Biochar-compost as a new option for soil improvement: Application in various problem soils. Sci. Total Environ. 2023, 870, 162024. [Google Scholar] [CrossRef]
  20. El-Desuki, M.; Hafez, M.M.; Mahmoud, A.R.; El-Al, A.; Faten, S. Effect of organic and bio fertilizers on the plant growth, green pod yield, quality of pea. Int. J. Acad. Res. 2010, 2, 87–92. [Google Scholar]
  21. Tumuhairwe, J.B.; Tenywa, J.S. Bacterial community changes during composting of municipal crop waste using low technology methods as revealed by 16S rRNA. Afr. J. Environ. Sci. Technol. 2018, 12, 209–221. [Google Scholar]
  22. Sánchez, Ó.J.; Ospina, D.A.; Montoya, S. Compost supplementation with nutrients and microorganisms in composting process. Waste Manag. 2017, 69, 136–153. [Google Scholar] [CrossRef] [PubMed]
  23. Li, Z.; Lu, H.; Ren, L.; He, L. Experimental and modeling approaches for food waste composting: A review. Chemosphere 2013, 93, 1247–1257. [Google Scholar] [CrossRef] [PubMed]
  24. Osman, A.I.; Fawzy, S.; Farghali, M.; El-Azazy, M.; Elgarahy, A.M.; Fahim, R.A.; Maksoud, M.A.; Ajlan, A.A.; Yousry, M.; Saleem, Y. Biochar for agronomy, animal farming, anaerobic digestion, composting, water treatment, soil remediation, construction, energy storage, and carbon sequestration: A review. Environ. Chem. Lett. 2022, 20, 2385–2485. [Google Scholar] [CrossRef]
  25. Mahapatra, S.; Yadav, R.; Ramakrishna, W. Bacillus subtilis impact on plant growth, soil health and environment: Dr. Jekyll and Mr. Hyde. J. Appl. Microbiol. 2022, 132, 3543–3562. [Google Scholar] [CrossRef]
  26. Rehman, A.D.; Ayyub, M.; Zahir, Z.A. The trend towards suppressing most troublesome weeds for sustainable production of wheat. In Food and Agriculture on Social, Economic and Environmental Linkages; Iksad Publications: Ankara, Türkiye, 2023; p. 74. [Google Scholar]
  27. Jat, S.L.; Suby, S.; Parihar, C.M.; Gambhir, G.; Kumar, N.; Rakshit, S. Microbiome for sustainable agriculture: A review with special reference to the corn production system. Arch. Microbiol. 2021, 203, 2771–2793. [Google Scholar] [CrossRef] [PubMed]
  28. Sambo, U.; Sule, B. Impact of Climate Change on Food Security in Northern Nigeria. Green Low-Carbon Econ. 2023. [Google Scholar] [CrossRef]
  29. Babu, S.; Rana, D.; Yadav, G.; Singh, R.; Yadav, S. A review on recycling of sunflower residue for sustaining soil health. Int. J. Agron. 2014, 2014, 601049. [Google Scholar] [CrossRef]
  30. Chen, J.-H. The combined use of chemical and organic fertilizers and/or biofertilizer for crop growth and soil fertility. In Proceedings of the International Workshop on Sustained Management of the Soil-Rhizosphere System for Efficient Crop Production and Fertilizer Use, Bangkok, Thailand, 16–20 October 2006; p. 20. [Google Scholar]
  31. Zhang, J.; Hamza, A.; Xie, Z.; Hussain, S.; Brestic, M.; Tahir, M.A.; Ulhassan, Z.; Yu, M.; Allakhverdiev, S.I.; Shabala, S. Arsenic transport and interaction with plant metabolism: Clues for improving agricultural productivity and food safety. Environ. Pollut. 2021, 290, 117987. [Google Scholar] [CrossRef] [PubMed]
  32. Son, T.; Thu, V.; Duong, V.; Hiraoka, H. Effect of Organic and Bio-Fertilizers on Soybean and Rice Cropping System; Japan International Research Center for Agricultural Sciences: Tsukuba, Ibaraki, Japan, 2007. [Google Scholar]
  33. Bhattacharjee, R.; Dey, U. Biofertilizer, a way towards organic agriculture: A review. Afr. J. Microbiol. Res. 2014, 8, 2332–2343. [Google Scholar]
  34. de Moraes, A.C.P.; Ribeiro, L.d.S.; de Camargo, E.R.; Lacava, P.T. The potential of nanomaterials associated with plant growth-promoting bacteria in agriculture. 3 Biotech 2021, 11, 318. [Google Scholar] [CrossRef]
  35. Mitter, E.K.; Tosi, M.; Obregón, D.; Dunfield, K.E.; Germida, J.J. Rethinking crop nutrition in times of modern microbiology: Innovative biofertilizer technologies. Front. Sustain. Food Syst. 2021, 5, 606815. [Google Scholar] [CrossRef]
  36. Bhattacharyya, P.; Jha, D. Plant growth-promoting rhizobacteria (PGPR): Emergence in agriculture. World J. Microbiol. Biotechnol. 2012, 28, 1327–1350. [Google Scholar] [CrossRef]
  37. Mourouzidou, S.; Ntinas, G.K.; Tsaballa, A.; Monokrousos, N. Introducing the Power of Plant Growth Promoting Microorganisms in Soilless Systems: A Promising Alternative for Sustainable Agriculture. Sustainability 2023, 15, 5959. [Google Scholar] [CrossRef]
  38. Chunhaleuchanon, S. Screening of rhizobacteria for their plant growth promoting activities. KMITL Sci. Technol. J. 2008, 8, 18–23. [Google Scholar]
  39. Jelin, J.; Dhanarajan, M.; Mariappan, V. Assessment of compost as a bio-fertilizer for the growth of paddy. J. Environ. Biol. 2013, 34, 975. [Google Scholar]
  40. Sadhana, B. Arbuscular Mycorrhizal Fungi (AMF) as a Biofertilizer-a Review. Int. J. Curr. Microbiol. App. Sci 2014, 3, 384–400. [Google Scholar]
  41. Kachroo, D.; Razdan, R. Growth, nutrient uptake and yield of wheat (Triticum aestivum) as influenced by biofertilizers and nitrogen. Indian J. Agron. 2006, 51, 37–39. [Google Scholar]
  42. Fazeli-Nasab, B.; Shahraki-Mojahed, L.; Piri, R.; Sobhanizadeh, A. Trichoderma: Improving growth and tolerance to biotic and abiotic stresses in plants. In Trends of Applied Microbiology for Sustainable Economy; Elsevier: Amsterdam, The Netherlands, 2022; pp. 525–564. [Google Scholar]
  43. Ghorbanpour, A.; Salimi, A.; Ghanbary, M.A.T.; Pirdashti, H.; Dehestani, A. The effect of Trichoderma harzianum in mitigating low temperature stress in tomato (Solanum lycopersicum L.) plants. Sci. Hortic. 2018, 230, 134–141. [Google Scholar] [CrossRef]
  44. Youssef, M.; Eissa, M. Biofertilizers and their role in management of plant parasitic nematodes. A review. E3 J. Biotechnol. Pharm. Res 2014, 5, 1–6. [Google Scholar]
  45. Adesemoye, A.O.; Kloepper, J.W. Plant–microbes interactions in enhanced fertilizer-use efficiency. Appl. Microbiol. Biotechnol. 2009, 85, 1–12. [Google Scholar] [CrossRef]
  46. Mahmood, F.; Khan, I.; Ashraf, U.; Shahzad, T.; Hussain, S.; Shahid, M.; Abid, M.; Ullah, S. Effects of organic and inorganic manures on maize and their residual impact on soil physico-chemical properties. J. Soil Sci. Plant Nutr. 2017, 17, 22–32. [Google Scholar] [CrossRef]
  47. Vakili, M.; Rafatullah, M.; Ibrahim, M.H.; Salamatinia, B.; Gholami, Z.; Zwain, H.M. A review on composting of oil palm biomass. Environ. Dev. Sustain. 2015, 17, 691–709. [Google Scholar] [CrossRef]
  48. Tariq, M.; Noman, M.; Ahmed, T.; Hameed, A.; Manzoor, N.; Zafar, M. Antagonistic features displayed by Plant Growth Promoting Rhizobacteria (PGPR): A Review. J. Plant Sci. Phytopathol. 2017, 1, 38–43. [Google Scholar]
  49. Tariq, M.; Jameel, F.; Ijaz, U.; Abdullah, M.; Rashid, K. Biofertilizer microorganisms accompanying pathogenic attributes: A potential threat. Physiol. Mol. Biol. Plants 2022, 28, 77–90. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, Y.; Liu, Y.; Feng, W.; Zeng, S. Waste Haven Transfer and Poverty-Environment Trap: Evidence from EU. Green Low-Carbon Econ. 2023, 1, 41–49. [Google Scholar] [CrossRef]
  51. Mącik, M.; Gryta, A.; Frąc, M. Biofertilizers in agriculture: An overview on concepts, strategies and effects on soil microorganisms. Adv. Agron. 2020, 162, 31–87. [Google Scholar]
  52. Lima, J.Z.; da Silva, E.F.; Patinha, C.; Rodrigues, V.G.S. Sorption and post-sorption performances of Cd, Pb and Zn onto peat, compost and biochar. J. Environ. Manag. 2022, 321, 115968. [Google Scholar] [CrossRef] [PubMed]
  53. Viaene, J.; Van Lancker, J.; Vandecasteele, B.; Willekens, K.; Bijttebier, J.; Ruysschaert, G.; De Neve, S.; Reubens, B. Opportunities and barriers to on-farm composting and compost application: A case study from northwestern Europe. Waste Manag. 2016, 48, 181–192. [Google Scholar] [CrossRef] [PubMed]
  54. Nkonya, E.; Kato, E.; Kabore, C. Impact of Farmer-Managed Natural Regeneration on Resilience and Welfare in Mali. Green Low-Carbon Econ. 2023. [Google Scholar] [CrossRef]
  55. DeLuca, T.H.; Gundale, M.J.; MacKenzie, M.D.; Jones, D.L. Biochar effects on soil nutrient transformations. Biochar Environ. Manag. Sci. Technol. Implement. 2015, 2, 421–454. [Google Scholar]
  56. Stirling, G.R. Biological control of plant-parasitic nematodes. In Diseases of Nematodes; CRC Press: Boca Raton, FL, USA, 2017; pp. 103–150. [Google Scholar]
  57. Yadav, S.; Babu, S.; Yadav, M.; Singh, K.; Yadav, G.; Pal, S. A review of organic farming for sustainable agriculture in Northern India. Int. J. Agron. 2013, 2013, 718145. [Google Scholar] [CrossRef]
  58. Rengalakshmi, R.; Manjula, M.; Prabavathy, V.; Jegan, S.; Selvamukilan, B. Building Bioeconomy in Agriculture: Harnessing Soil Microbes for Sustaining Ecosystem Services. In Towards a Sustainable Bioeconomy: Principles, Challenges and Perspectives; Springer: Berlin/Heidelberg, Germany, 2018; pp. 261–277. [Google Scholar]
  59. Qu, F.; Wu, D.; Li, D.; Zhao, Y.; Zhang, R.; Qi, H.; Chen, X. Effect of Fenton pretreatment combined with bacterial inoculation on humification characteristics of dissolved organic matter during rice straw composting. Bioresour. Technol. 2022, 344, 126198. [Google Scholar] [CrossRef] [PubMed]
  60. Harindintwali, J.D.; Zhou, J.; Yu, X. Lignocellulosic crop residue composting by cellulolytic nitrogen-fixing bacteria: A novel tool for environmental sustainability. Sci. Total Environ. 2020, 715, 136912. [Google Scholar] [CrossRef] [PubMed]
  61. Meyer, A.; Ehimen, E.; Holm-Nielsen, J. Future European biogas: Animal manure, straw and grass potentials for a sustainable European biogas production. Biomass Bioenergy 2017, 111, 154–164. [Google Scholar] [CrossRef]
  62. Avery, L.M.; Booth, P.; Campbell, C.; Tompkins, D.; Hough, R.L. Prevalence and survival of potential pathogens in source-segregated green waste compost. Sci. Total Environ. 2012, 431, 128–138. [Google Scholar] [CrossRef] [PubMed]
  63. Akhtar, M.J.; Asghar, H.N.; Shahzad, K.; Arshad, M. Role of plant growth promoting rhizobacteria applied in combination with compost and mineral fertilizers to improve growth and yield of wheat (Triticum aestivum L.). Pak. J. Bot. 2009, 41, 381–390. [Google Scholar]
  64. Gil, M.; Carballo, M.; Calvo, L. Fertilization of maize with compost from cattle manure supplemented with additional mineral nutrients. Waste Manag. 2008, 28, 1432–1440. [Google Scholar] [CrossRef]
  65. Shilev, S.; Naydenov, M.; Vancheva, V.; Aladjadjiyan, A. Composting of food and agricultural wastes. In Utilization of By-Products and Treatment of Waste in the Food Industry; Springer: Boston, MA, USA, 2007; pp. 283–301. [Google Scholar]
  66. Chatterjee, N.; Flury, M.; Hinman, C.; Cogger, C.G. Chemical and Physical Characteristics of Compost Leachates. A Review; Washington State University: Pullman, WA, USA, 2013. [Google Scholar]
  67. Guo, X.-x.; Liu, H.-t.; Zhang, J. The role of biochar in organic waste composting and soil improvement: A review. Waste Manag. 2020, 102, 884–899. [Google Scholar] [CrossRef]
  68. Meunchang, S.; Panichsakpatana, S.; Weaver, R.W. Inoculation of sugar mill by-products compost with N 2-fixing bacteria. Plant Soil 2005, 271, 219–225. [Google Scholar] [CrossRef]
  69. Zanella, A.; Ponge, J.-F.; Guercini, S.; Rumor, C.; Nold, F.; Sambo, P.; Gobbi, V.; Schimmer, C.; Chaabane, C.; Mouchard, M.-L. Humusica 2, article 16: Techno humus systems and recycling of waste. Appl. Soil Ecol. 2018, 122, 220–236. [Google Scholar] [CrossRef]
  70. Mata-Alvarez, J.; Mace, S.; Llabres, P. Anaerobic digestion of organic solid wastes. An overview of research achievements and perspectives. Bioresour. Technol. 2000, 74, 3–16. [Google Scholar] [CrossRef]
  71. Tambone, F.; Scaglia, B.; D’Imporzano, G.; Schievano, A.; Orzi, V.; Salati, S.; Adani, F. Assessing amendment and fertilizing properties of digestates from anaerobic digestion through a comparative study with digested sludge and compost. Chemosphere 2010, 81, 577–583. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, C.-Y.; Mei, H.-C.; Cheng, C.-Y.; Lin, J.-H.; Chung, Y.-C. Enhancing the conversion of organic waste into biofertilizer with thermophilic bacteria. Environ. Eng. Sci. 2012, 29, 726–730. [Google Scholar] [CrossRef]
  73. Lin, C.; Cheruiyot, N.K.; Hoang, H.-G.; Le, T.-H.; Tran, H.-T.; Bui, X.-T. Benzophenone biodegradation and characterization of malodorous gas emissions during co-composting of food waste with sawdust and mature compost. Environ. Technol. Innov. 2021, 21, 101351. [Google Scholar] [CrossRef]
  74. Igiehon, N.O.; Babalola, O.O. Biofertilizers and sustainable agriculture: Exploring arbuscular mycorrhizal fungi. Appl. Microbiol. Biotechnol. 2017, 101, 4871–4881. [Google Scholar] [CrossRef] [PubMed]
  75. Gothwal, R.; Nigam, V.; Mohan, M.; Sasmal, D.; Ghosh, P. Screening of nitrogen fixers from rhizospheric bacterial isolates associated with important desert plants. Appl. Ecol. Eviron. Res. 2008, 6, 101–109. [Google Scholar] [CrossRef]
  76. Bakulin, M.; Grudtsyna, A.; Pletneva, A.Y. Biological fixation of nitrogen and growth of bacteria of the genus Azotobacter in liquid media in the presence of perfluorocarbons. Appl. Biochem. Microbiol. 2007, 43, 399. [Google Scholar] [CrossRef]
  77. Igual, J.; Valverde, A.; Cervantes, E.; Velázquez, E. Phosphate-solubilizing bacteria as inoculants for agriculture: Use of updated molecular techniques in their study. Agronomie 2001, 21, 561–568. [Google Scholar] [CrossRef]
  78. Satyaprakash, M.; Nikitha, T.; Reddi, E.; Sadhana, B.; Vani, S.S. Phosphorous and phosphate solubilising bacteria and their role in plant nutrition. Int. J. Curr. Microbiol. App. Sci 2017, 6, 2133–2144. [Google Scholar]
  79. Liu, D.; Lian, B.; Dong, H. Isolation of Paenibacillus sp. and assessment of its potential for enhancing mineral weathering. Geomicrobiol. J. 2012, 29, 413–421. [Google Scholar] [CrossRef]
  80. James, E.K.; Gyaneshwar, P.; Barraquio, W.L.; Mathan, N.; Ladha, J.K. Endophytic diazotrophs associated with rice. In The Quest for Nitrogen Fixation in Rice; International Rice Research Institute: Makati City, Philippines, 2000; pp. 119–140. [Google Scholar]
  81. Meunchang, S.; Panichsakpatana, S.; Weaver, R. Tomato growth in soil amended with sugar mill by-products compost. Plant Soil 2006, 280, 171–176. [Google Scholar] [CrossRef]
  82. Meng, L.; Zhang, A.; Wang, F.; Han, X.; Wang, D.; Li, S. Arbuscular mycorrhizal fungi and rhizobium facilitate nitrogen uptake and transfer in soybean/maize intercropping system. Front. Plant Sci. 2015, 6, 339. [Google Scholar] [CrossRef]
  83. Bulgarelli, R.G.; Marcos, F.C.C.; Ribeiro, R.V.; de Andrade, S.A.L. Mycorrhizae enhance nitrogen fixation and photosynthesis in phosphorus-starved soybean (Glycine max L. Merrill). Environ. Exp. Bot. 2017, 140, 26–33. [Google Scholar] [CrossRef]
  84. Hodge, A.; Fitter, A.H. Substantial nitrogen acquisition by arbuscular mycorrhizal fungi from organic material has implications for N cycling. Proc. Natl. Acad. Sci. 2010, 107, 13754–13759. [Google Scholar] [CrossRef] [PubMed]
  85. Mehmood, S.; Khatoon, Z.; Amna; Ahmad, I.; Muneer, M.A.; Kamran, M.A.; Ali, J.; Ali, B.; Chaudhary, H.J.; Munis, M.F.H. Bacillus sp. PM31 harboring various plant growth-promoting activities regulates Fusarium dry rot and wilt tolerance in potato. Arch. Agron. Soil Sci. 2023, 69, 197–211. [Google Scholar] [CrossRef]
  86. Chakraborty, U.; Chakraborty, B.; Basnet, M.; Chakraborty, A. Evaluation of Ochrobactrum anthropi TRS-2 and its talc based formulation for enhancement of growth of tea plants and management of brown root rot disease. J. Appl. Microbiol. 2009, 107, 625–634. [Google Scholar] [CrossRef] [PubMed]
  87. Rodriguez, H.; Gonzalez, T.; Goire, I.; Bashan, Y. Gluconic acid production and phosphate solubilization by the plant growth-promoting bacterium Azospirillum spp. Naturwissenschaften 2004, 91, 552–555. [Google Scholar] [CrossRef] [PubMed]
  88. Bashan, Y.; Kamnev, A.A.; de-Bashan, L.E. A proposal for isolating and testing phosphate-solubilizing bacteria that enhance plant growth. Biol. Fertil. Soils 2013, 49, 1–2. [Google Scholar] [CrossRef]
  89. Altaf, M.M.; Imran, M.; Abulreesh, H.H.; Khan, M.S.; Ahmad, I. Diversity and Applications of Penicillium spp. in Plant-Growth Promotion. In New and Future Developments in Microbial Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2017; pp. 261–276. [Google Scholar]
  90. Rossati, K.F.; Figueiredo, C.C.d.; Mendes, G.d.O. Aspergillus niger Enhances the Efficiency of Sewage Sludge Biochar as a Sustainable Phosphorus Source. Sustainability 2023, 15, 6940. [Google Scholar] [CrossRef]
  91. Teotia, P.; Kumar, V.; Kumar, M.; Shrivastava, N.; Varma, A. Rhizosphere Microbes: Potassium Solubilization and Crop Productivity–Present and Future Aspects. In Potassium Solubilizing Microorganisms for Sustainable Agriculture; Springer: Berlin/Heidelberg, Germany, 2016; pp. 315–325. [Google Scholar]
  92. Ahmad, M.; Nadeem, S.M.; Naveed, M.; Zahir, Z.A. Potassium-solubilizing bacteria and their application in agriculture. In Potassium Solubilizing Microorganisms for Sustainable Agriculture; Springer: Berlin/Heidelberg, Germany, 2016; pp. 293–313. [Google Scholar]
  93. Prajapati, K.; Sharma, M.; Modi, H. Isolation of two potassium solubilizing fungi from ceramic industry soils. Life Sci. Leafl. 2012, 5, 71–75. [Google Scholar]
  94. Fomina, M.; Alexander, I.J.; Hillier, S.; Gadd, G. Zinc phosphate and pyromorphite solubilization by soil plant-symbiotic fungi. Geomicrobiol. J. 2004, 21, 351–366. [Google Scholar] [CrossRef]
  95. Martino, E.; Perotto, S.; Parsons, R.; Gadd, G.M. Solubilization of insoluble inorganic zinc compounds by ericoid mycorrhizal fungi derived from heavy metal polluted sites. Soil Biol. Biochem. 2003, 35, 133–141. [Google Scholar] [CrossRef]
  96. Kumar, V.; Sarma, M.; Saharan, K.; Srivastava, R.; Kumar, L.; Sahai, V.; Bisaria, V.; Sharma, A. Effect of formulated root endophytic fungus Piriformospora indica and plant growth promoting rhizobacteria fluorescent pseudomonads R62 and R81 on Vigna mungo. World J. Microbiol. Biotechnol. 2012, 28, 595–603. [Google Scholar] [CrossRef] [PubMed]
  97. Li, M.; Ahammed, G.J.; Li, C.; Bao, X.; Yu, J.; Huang, C.; Yin, H.; Zhou, J. Brassinosteroid ameliorates zinc oxide nanoparticles-induced oxidative stress by improving antioxidant potential and redox homeostasis in tomato seedling. Front. Plant Sci. 2016, 7, 615. [Google Scholar] [CrossRef] [PubMed]
  98. Rokhbakhsh-Zamin, F.; Sachdev, D.; Kazemi-Pour, N.; Engineer, A.; Pardesi, K.R.; Zinjarde, S.; Dhakephalkar, P.K.; Chopade, B.A. Characterization of plant-growth-promoting traits of Acinetobacter species isolated from rhizosphere of Pennisetum glaucum. J. Microbiol. Biotechnol. 2011, 21, 556–566. [Google Scholar] [CrossRef] [PubMed]
  99. Sharma, S.K.; Sharma, M.P.; Ramesh, A.; Joshi, O.P. Characterization of zinc-solubilizing Bacillus isolates and their potential to influence zinc assimilation in soybean seeds. J. Microbiol. Biotechnol. 2012, 22, 352–359. [Google Scholar] [CrossRef] [PubMed]
  100. Tsavkelova, E.; Klimova, S.Y.; Cherdyntseva, T.; Netrusov, A. Microbial producers of plant growth stimulators and their practical use: A review. Appl. Biochem. Microbiol. 2006, 42, 117–126. [Google Scholar] [CrossRef]
  101. Bottini, R.; Cassán, F.; Piccoli, P. Gibberellin production by bacteria and its involvement in plant growth promotion and yield increase. Appl. Microbiol. Biotechnol. 2004, 65, 497–503. [Google Scholar] [CrossRef]
  102. Akhtar, M.; Siddiqui, Z. Effects of phosphate solubilizing microorganisms and Rhizobium sp. on the growth, nodulation, yield and root-rot disease complex of chickpea under field condition. Afr. J. Biotechnol. 2009, 8, 3489–3496. [Google Scholar]
  103. Waqas, M.; Khan, A.L.; Shahzad, R.; Ullah, I.; Khan, A.R.; Lee, I.-J. Mutualistic fungal endophytes produce phytohormones and organic acids that promote japonica rice plant growth under prolonged heat stress. J. Zhejiang Univ.-Sci. B 2015, 16, 1011–1018. [Google Scholar] [CrossRef]
  104. Bilal, L.; Asaf, S.; Hamayun, M.; Gul, H.; Iqbal, A.; Ullah, I.; Lee, I.-J.; Hussain, A. Plant growth promoting endophytic fungi Asprgillus fumigatus TS1 and Fusarium proliferatum BRL1 produce gibberellins and regulates plant endogenous hormones. Symbiosis 2018, 76, 117–127. [Google Scholar] [CrossRef]
  105. Kumar, V.; Prasher, I. Phosphate solubilization and indole-3-acetic acid (IAA) produced by Colletotrichum gloeosporioides and Aspergillus fumigatus strains isolated from the rhizosphere of Dillenia indica L. Folia Microbiol. 2023, 68, 219–229. [Google Scholar] [CrossRef] [PubMed]
  106. Haas, H. Fungal siderophore metabolism with a focus on Aspergillus fumigatus. Nat. Prod. Rep. 2014, 31, 1266–1276. [Google Scholar] [CrossRef]
  107. Patel, D.; Patel, S.; Thakar, P.; Saraf, M. Siderophore Producing Aspergillus spp as Bioinoculant for Enhanced Growth of Mung Bean. Int. J. Adv. Agric. Sci. Technol. 2017, 6, 111–120. [Google Scholar] [CrossRef]
  108. Li, Y.; Wang, Z.; Liu, X.; Song, Z.; Li, R.; Shao, C.; Yin, Y. Siderophore biosynthesis but not reductive iron assimilation is essential for the dimorphic fungus Nomuraea rileyi conidiation, dimorphism transition, resistance to oxidative stress, pigmented microsclerotium formation, and virulence. Front. Microbiol. 2016, 7, 931. [Google Scholar] [CrossRef] [PubMed]
  109. Happacher, I.; Aguiar, M.; Yap, A.; Decristoforo, C.; Haas, H. Fungal siderophore metabolism with a focus on Aspergillus fumigatus: Impact on biotic interactions and potential translational applications. Essays Biochem. 2023, 67, 829–842. [Google Scholar] [PubMed]
  110. Hindersah, R. Exopolysaccharide-Producing Azotobacter for Bioremediation of Heavy Metal-Contaminated Soil. In Advances in Agricultural and Industrial Microbiology; Volume-2: Applications of Microbes for Sustainable Agriculture and In-Silico Strategies; Springer: Singapore, 2022; pp. 103–117. [Google Scholar]
  111. Çam, S.; Bicek, S. The effects of temperature, salt, and phosphate on biofilm and exopolysaccharide production by Azotobacter spp. Arch. Microbiol. 2023, 205, 87. [Google Scholar] [CrossRef]
  112. Sehrawat, A.; Sindhu, S.S.; Glick, B.R. Hydrogen cyanide production by soil bacteria: Biological control of pests and promotion of plant growth in sustainable agriculture. Pedosphere 2022, 32, 15–38. [Google Scholar] [CrossRef]
  113. El-Rahman, A.; Shaheen, H.A.; El-Aziz, A.; Rabab, M.; Ibrahim, D.S. Influence of hydrogen cyanide-producing rhizobacteria in controlling the crown gall and root-knot nematode, Meloidogyne incognita. Egypt. J. Biol. Pest Control 2019, 29, 41. [Google Scholar] [CrossRef]
  114. Sneka, M.; Shanmitha, A.; Priyadharshini, A.; Thilakavathi, G.; Joselin, J.; Sarenya, R.; Nachiar, T.; Kaleeswari, G.; Pushpakanth, P.; Tamilselvi, S. Isolation and characterization of Rhizobium from green gram (Vigna radiata). Curr. Agric. Res. J. 2022, 10, 277–289. [Google Scholar] [CrossRef]
  115. Gamalero, E.; Berta, G.; Massa, N.; Glick, B.R.; Lingua, G. Synergistic interactions between the ACC deaminase-producing bacterium Pseudomonas putida UW4 and the AM fungus Gigaspora rosea positively affect cucumber plant growth. FEMS Microbiol. Ecol. 2008, 64, 459–467. [Google Scholar] [CrossRef]
  116. Duan, J.; Müller, K.M.; Charles, T.C.; Vesely, S.; Glick, B.R. 1-aminocyclopropane-1-carboxylate (ACC) deaminase genes in rhizobia from southern Saskatchewan. Microb. Ecol. 2009, 57, 423–436. [Google Scholar] [CrossRef] [PubMed]
  117. Ghosh, S.; Penterman, J.N.; Little, R.D.; Chavez, R.; Glick, B.R. Three newly isolated plant growth-promoting bacilli facilitate the seedling growth of canola, Brassica campestris. Plant Physiol. Biochem. 2003, 41, 277–281. [Google Scholar] [CrossRef]
  118. Pastori, G.M.; Kiddle, G.; Antoniw, J.; Bernard, S.; Veljovic-Jovanovic, S.; Verrier, P.J.; Noctor, G.; Foyer, C.H. Leaf vitamin C contents modulate plant defense transcripts and regulate genes that control development through hormone signaling. Plant Cell 2003, 15, 939–951. [Google Scholar] [CrossRef] [PubMed]
  119. Van Wees, S.C.; Van der Ent, S.; Pieterse, C.M. Plant immune responses triggered by beneficial microbes. Curr. Opin. Plant Biol. 2008, 11, 443–448. [Google Scholar] [CrossRef] [PubMed]
  120. Djonović, S.; Vargas, W.A.; Kolomiets, M.V.; Horndeski, M.; Wiest, A.; Kenerley, C.M. A proteinaceous elicitor Sm1 from the beneficial fungus Trichoderma virens is required for induced systemic resistance in maize. Plant Physiol. 2007, 145, 875–889. [Google Scholar] [CrossRef] [PubMed]
  121. Naznin, H.A.; Kiyohara, D.; Kimura, M.; Miyazawa, M.; Shimizu, M.; Hyakumachi, M. Systemic resistance induced by volatile organic compounds emitted by plant growth-promoting fungi in Arabidopsis thaliana. PLoS ONE 2014, 9, e86882. [Google Scholar] [CrossRef]
  122. Karunakaran, R.; Ebert, K.; Harvey, S.; Leonard, M.; Ramachandran, V.; Poole, P. Thiamine is synthesized by a salvage pathway in Rhizobium leguminosarum bv. viciae strain 3841. J. Bacteriol. 2006, 188, 6661–6668. [Google Scholar] [CrossRef]
  123. Palacios, O.A.; Bashan, Y.; de-Bashan, L.E. Proven and potential involvement of vitamins in interactions of plants with plant growth-promoting bacteria—An overview. Biol. Fertil. Soils 2014, 50, 415–432. [Google Scholar] [CrossRef]
  124. Bona, E.; Lingua, G.; Manassero, P.; Cantamessa, S.; Marsano, F.; Todeschini, V.; Copetta, A.; D’Agostino, G.; Massa, N.; Avidano, L. AM fungi and PGP Pseudomonads increase flowering, fruit production, and vitamin content in strawberry grown at low nitrogen and phosphorus levels. Mycorrhiza 2015, 25, 181–193. [Google Scholar] [CrossRef]
  125. Pawar, P.M.A.; Derba-Maceluch, M.; Chong, S.L.; Gómez, L.D.; Miedes, E.; Banasiak, A.; Ratke, C.; Gaertner, C.; Mouille, G.; McQueen-Mason, S.J. Expression of fungal acetyl xylan esterase in Arabidopsis thaliana improves saccharification of stem lignocellulose. Plant Biotechnol. J. 2016, 14, 387–397. [Google Scholar] [CrossRef]
  126. Talaat, N.B.; Shawky, B.T. Protective effects of arbuscular mycorrhizal fungi on wheat (Triticum aestivum L.) plants exposed to salinity. Environ. Exp. Bot. 2014, 98, 20–31. [Google Scholar] [CrossRef]
  127. Albalawi, M.A.; Abdelaziz, A.M.; Attia, M.S.; Saied, E.; Elganzory, H.H.; Hashem, A.H. Mycosynthesis of silica nanoparticles using Aspergillus niger: Control of Alternaria solani causing early blight disease, induction of innate immunity and reducing of oxidative stress in eggplant. Antioxidants 2022, 11, 2323. [Google Scholar] [CrossRef]
  128. Verma, J.P.; Yadav, J.; Tiwari, K.N.; Kumar, A. Effect of indigenous Mesorhizobium spp. and plant growth promoting rhizobacteria on yields and nutrients uptake of chickpea (Cicer arietinum L.) under sustainable agriculture. Ecol. Eng. 2013, 51, 282–286. [Google Scholar] [CrossRef]
  129. Chaiharn, M.; Chunhaleuchanon, S.; Lumyong, S. Screening siderophore producing bacteria as potential biological control agent for fungal rice pathogens in Thailand. World J. Microbiol. Biotechnol. 2009, 25, 1919–1928. [Google Scholar] [CrossRef]
  130. Rath, M.; Mitchell, T.; Gold, S. Volatiles produced by Bacillus mojavensis RRC101 act as plant growth modulators and are strongly culture-dependent. Microbiol. Res. 2018, 208, 76–84. [Google Scholar] [CrossRef] [PubMed]
  131. Morath, S.U.; Hung, R.; Bennett, J.W. Fungal volatile organic compounds: A review with emphasis on their biotechnological potential. Fungal Biol. Rev. 2012, 26, 73–83. [Google Scholar] [CrossRef]
  132. Jalali, F.; Zafari, D.; Salari, H. Volatile organic compounds of some Trichoderma spp. increase growth and induce salt tolerance in Arabidopsis thaliana. Fungal Ecol. 2017, 29, 67–75. [Google Scholar] [CrossRef]
  133. Raza, W.; Ling, N.; Liu, D.; Wei, Z.; Huang, Q.; Shen, Q. Volatile organic compounds produced by Pseudomonas fluorescens WR-1 restrict the growth and virulence traits of Ralstonia solanacearum. Microbiol. Res. 2016, 192, 103–113. [Google Scholar] [CrossRef] [PubMed]
  134. Bhantana, P.; Rana, M.S.; Sun, X.-c.; Moussa, M.G.; Saleem, M.H.; Syaifudin, M.; Shah, A.; Poudel, A.; Pun, A.B.; Bhat, M.A. Arbuscular mycorrhizal fungi and its major role in plant growth, zinc nutrition, phosphorous regulation and phytoremediation. Symbiosis 2021, 84, 19–37. [Google Scholar] [CrossRef]
  135. Paul, E. 1-Soil Microbiology, Ecology, and Biochemistry in Perspective; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  136. Iqbal, S.; Hameed, S.; Shahid, M.; Hussain, K.; Ahmad, N.; Niaz, M. In vitro characterization of bacterial endophytes from tomato (Solanum lycopersicum L.) for phytobeneficial traits. Appl. Ecol. Environ. Res. 2018, 16, 1037–1051. [Google Scholar] [CrossRef]
  137. Tian, X.; Engel, B.A.; Qian, H.; Hua, E.; Sun, S.; Wang, Y. Will reaching the maximum achievable yield potential meet future global food demand? J. Clean. Prod. 2021, 294, 126285. [Google Scholar] [CrossRef]
  138. Wang, Z.; Yin, Y.; Wang, Y.; Tian, X.; Ying, H.; Zhang, Q.; Xue, Y.; Oenema, O.; Li, S.; Zhou, F. Integrating crop redistribution and improved management towards meeting China’s food demand with lower environmental costs. Nat. Food 2022, 3, 1031–1039. [Google Scholar] [CrossRef] [PubMed]
  139. Carle, J.; Holmgren, P. Wood from planted forests: A global outlook 2005–2030. For. Prod. J. 2008, 58, 6. [Google Scholar]
  140. Paudel, S.R.; Banjara, S.P.; Choi, O.K.; Park, K.Y.; Kim, Y.M.; Lee, J.W. Pretreatment of agricultural biomass for anaerobic digestion: Current state and challenges. Bioresour. Technol. 2017, 245, 1194–1205. [Google Scholar] [CrossRef] [PubMed]
  141. Notarnicola, B.; Hayashi, K.; Curran, M.A.; Huisingh, D. Progress in working towards a more sustainable agri-food industry. J. Clean. Prod. 2012, 28, 1–8. [Google Scholar] [CrossRef]
  142. Gupta, A.P.; Upadhyay, P.; Sen, T.; Dutta, J. Agricultural Waste as a Resource: The Lesser Travelled Road to Sustainability. In Agriculture Waste Management and Bioresource: The Circular Economy Perspective; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2023; pp. 1–20. [Google Scholar]
  143. Haas, E.; Carozzi, M.; Massad, R.S.; Butterbach-Bahl, K.; Scheer, C. Long term impact of residue management on soil organic carbon stocks and nitrous oxide emissions from European croplands. Sci. Total Environ. 2022, 836, 154932. [Google Scholar] [CrossRef]
  144. Kim, S.; Dale, B.E. Global potential bioethanol production from wasted crops and crop residues. Biomass Bioenergy 2004, 26, 361–375. [Google Scholar] [CrossRef]
  145. Meyer, A.; Ehimen, E.; Holm-Nielsen, J. The potential of animal manure, straw and grass for European biogas production in 2030. In Proceedings of the 24th European Biomass Conference and Exhibition, Amsterdam, The Netherlands, 6–9 June 2016. [Google Scholar]
  146. Rodriguez, C.; Alaswad, A.; Benyounis, K.; Olabi, A. Pretreatment techniques used in biogas production from grass. Renew. Sustain. Energy Rev. 2017, 68, 1193–1204. [Google Scholar] [CrossRef]
  147. Tsapekos, P.; Kougias, P.G.; Angelidaki, I. Anaerobic mono-and co-digestion of mechanically pretreated meadow grass for biogas production. Energy Fuels 2015, 29, 4005–4010. [Google Scholar] [CrossRef]
  148. Sarkar, T.; Salauddin, M.; Roy, S.; Chakraborty, R.; Rebezov, M.; Shariati, M.A.; Thiruvengadam, M.; Rengasamy, K.R.R. Underutilized green leafy vegetables: Frontier in fortified food development and nutrition. Crit. Rev. Food Sci. Nutr. 2022, 1–55. [Google Scholar] [CrossRef]
  149. Barbosa-Cánovas, G.; Fernández-Molina, J.; Alzamora, S.; Tapia, M.; López-Malo, A.; Chanes, J. General considerations for preservation of fruits and vegetables. In Handling and Preservation of Fruits and Vegetables by Combined Methods for Rural Areas; Food and Agriculture Organization of the United Nations: Rome, Italy, 2003. [Google Scholar]
  150. Plazzotta, S.; Manzocco, L.; Nicoli, M.C. Fruit and vegetable waste management and the challenge of fresh-cut salad. Trends Food Sci. Technol. 2017, 63, 51–59. [Google Scholar] [CrossRef]
  151. Su, G.; Ong, H.C.; Zulkifli, N.W.M.; Ibrahim, S.; Chen, W.H.; Chong, C.T.; Ok, Y.S. Valorization of animal manure via pyrolysis for bioenergy: A review. J. Clean. Prod. 2022, 343, 130965. [Google Scholar] [CrossRef]
  152. Srivastava, R.K.; Shetti, N.P.; Reddy, K.R.; Nadagouda, M.N.; Badawi, M.; Bonilla-Petriciolet, A.; Aminabhavi, T.M. Valorization of biowastes for clean energy production, environmental depollution and soil fertility. J. Environ. Manag. 2023, 332, 117410. [Google Scholar] [CrossRef]
  153. Michel Jr, F.C.; Pecchia, J.A.; Rigot, J.; Keener, H.M. Mass and nutrient losses during the composting of dairy manure amended with sawdust or straw. Compos. Sci. Util. 2004, 12, 323–334. [Google Scholar] [CrossRef]
  154. Solano, M.; Iriarte, F.; Ciria, P.; Negro, M. SE—Structure and environment: Performance characteristics of three aeration systems in the composting of sheep manure and straw. J. Agric. Eng. Res. 2001, 79, 317–329. [Google Scholar] [CrossRef]
  155. Höhn, J.; Lehtonen, E.; Rasi, S.; Rintala, J. A Geographical Information System (GIS) based methodology for determination of potential biomasses and sites for biogas plants in southern Finland. Appl. Energy 2014, 113, 1–10. [Google Scholar] [CrossRef]
  156. Bernal, M.P.; Alburquerque, J.; Moral, R. Composting of animal manures and chemical criteria for compost maturity assessment. A review. Bioresour. Technol. 2009, 100, 5444–5453. [Google Scholar] [CrossRef] [PubMed]
  157. Ahemad, M.; Kibret, M. Mechanisms and applications of plant growth promoting rhizobacteria: Current perspective. J. King Saud Univ.-Sci. 2014, 26, 1–20. [Google Scholar] [CrossRef]
  158. Kumar, A.; Maurya, B.; Raghuwanshi, R.; Meena, V.S.; Islam, M.T. Co-inoculation with Enterobacter and Rhizobacteria on Yield and Nutrient Uptake by Wheat (Triticum aestivum L.) in the Alluvial Soil Under Indo-Gangetic Plain of India. J. Plant Growth Regul. 2017, 36, 608–617. [Google Scholar] [CrossRef]
  159. Verma, R.; Maurya, B.; Meena, V.; Dotaniya, M.; Deewan, P.; Jajoria, M. Enhancing Production Potential of Cabbage and Improves Soil Fertility Status of Indo-Gangetic Plain through Application of Bio-organics and Mineral Fertilizer. Int. J. Curr. Microbiol. App. Sci 2017, 6, 301–309. [Google Scholar]
  160. Verma, J.P.; Jaiswal, D.K.; Meena, V.S.; Kumar, A.; Meena, R. Issues and challenges about sustainable agriculture production for management of natural resources to sustain soil fertility and health. J. Clean. Prod. 2015, 107, 793–794. [Google Scholar] [CrossRef]
  161. Nath, D.; Maurya, B.R.; Meena, V.S. Documentation of five potassium-and phosphorus-solubilizing bacteria for their K and P-solubilization ability from various minerals. Biocatal. Agric. Biotechnol. 2017, 10, 174–181. [Google Scholar] [CrossRef]
  162. Afzal, A.; Bano, A. Rhizobium and phosphate solubilizing bacteria improve the yield and phosphorus uptake in wheat (Triticum aestivum). Int. J. Agric. Biol. 2008, 10, 85–88. [Google Scholar]
  163. Indiragandhi, P.; Anandham, R.; Madhaiyan, M.; Sa, T. Characterization of plant growth–promoting traits of bacteria isolated from larval guts of diamondback moth Plutella xylostella (Lepidoptera: Plutellidae). Curr. Microbiol. 2008, 56, 327–333. [Google Scholar] [CrossRef] [PubMed]
  164. Selvakumar, G.; Mohan, M.; Kundu, S.; Gupta, A.; Joshi, P.; Nazim, S.; Gupta, H. Cold tolerance and plant growth promotion potential of Serratia marcescens strain SRM (MTCC 8708) isolated from flowers of summer squash (Cucurbita pepo). Lett. Appl. Microbiol. 2008, 46, 171–175. [Google Scholar] [CrossRef] [PubMed]
  165. Dutta, D.; Bandyopadhyay, P. Performance of chickpea (Cicer arietinum L.) to application of phosphorus and bio-fertilizer in laterite soil. Arch. Agron. Soil Sci. 2009, 55, 147–155. [Google Scholar] [CrossRef]
  166. Leaungvutiviroj, C.; Ruangphisarn, P.; Hansanimitkul, P.; Shinkawa, H.; Sasaki, K. Development of a new biofertilizer with a high capacity for N2 fixation, phosphate and potassium solubilization and auxin production. Biosci. Biotechnol. Biochem. 2010, 74, 1098–1101. [Google Scholar] [CrossRef]
  167. Mia, M.B.; Shamsuddin, Z.; Wahab, Z.; Marziah, M. Rhizobacteria as bioenhancer and biofertilizer for growth and yield of banana (Musa spp. cv.‘Berangan’). Sci. Hortic. 2010, 126, 80–87. [Google Scholar] [CrossRef]
  168. Tani, A.; Akita, M.; Murase, H.; Kimbara, K. Culturable bacteria in hydroponic cultures of moss Racomitrium japonicum and their potential as biofertilizers for moss production. J. Biosci. Bioeng. 2011, 112, 32–39. [Google Scholar] [CrossRef]
  169. Zayed, M.S. Improvement of growth and nutritional quality of Moringa oleifera using different biofertilizers. Ann. Agric. Sci. 2012, 57, 53–62. [Google Scholar] [CrossRef]
  170. Akhani, A.; Darzi, M.T.; Hadi, M.H.S. Effects of biofertilizer and plant density on yield components and seed yield of coriander (Coriandrum sativum). Int. J. Agric. Crop Sci. 2012, 4, 1205–1211. [Google Scholar]
  171. Ibiene, A.; Agogbua, J.; Okonko, I.; Nwachi, G. Plant growth promoting rhizobacteria (PGPR) as biofertilizer: Effect on growth of Lycopersicum esculentus. J. Am. Sci. 2012, 8, 318–324. [Google Scholar]
  172. Pešaković, M.; Karaklajić-Stajić, Ž.; Milenković, S.; Mitrović, O. Biofertilizer affecting yield related characteristics of strawberry (Fragaria× ananassa Duch.) and soil micro-organisms. Sci. Hortic. 2013, 150, 238–243. [Google Scholar] [CrossRef]
  173. Ghilavizadeh, A.; Darzi, M.T.; Hadi, M.H.S. Effects of biofertilizer and plant density on essential oil content and yield traits of ajowan (Carum copticum). Middle-East J. Sci. Res. 2013, 14, 1508–1512. [Google Scholar]
  174. Kumar, S.; Bauddh, K.; Barman, S.; Singh, R.P. Amendments of microbial biofertilizers and organic substances reduces requirement of urea and DAP with enhanced nutrient availability and productivity of wheat (Triticum aestivum L.). Ecol. Eng. 2014, 71, 432–437. [Google Scholar] [CrossRef]
  175. Abd-Alla, M.H.; El-Enany, A.-W.E.; Nafady, N.A.; Khalaf, D.M.; Morsy, F.M. Synergistic interaction of Rhizobium leguminosarum bv. viciae and arbuscular mycorrhizal fungi as a plant growth promoting biofertilizers for faba bean (Vicia faba L.) in alkaline soil. Microbiol. Res. 2014, 169, 49–58. [Google Scholar] [CrossRef]
  176. Shen, Z.; Ruan, Y.; Wang, B.; Zhong, S.; Su, L.; Li, R.; Shen, Q. Effect of biofertilizer for suppressing Fusarium wilt disease of banana as well as enhancing microbial and chemical properties of soil under greenhouse trial. Appl. Soil Ecol. 2015, 93, 111–119. [Google Scholar] [CrossRef]
  177. Araujo, J.; Díaz-Alcántara, C.-A.; Velázquez, E.; Urbano, B.; González-Andrés, F. Bradyrhizobium yuanmingense related strains form nitrogen-fixing symbiosis with Cajanus cajan L. in Dominican Republic and are efficient biofertilizers to replace N fertilization. Sci. Hortic. 2015, 192, 421–428. [Google Scholar] [CrossRef]
  178. Shen, Z.; Ruan, Y.; Chao, X.; Zhang, J.; Li, R.; Shen, Q. Rhizosphere microbial community manipulated by 2 years of consecutive biofertilizer application associated with banana Fusarium wilt disease suppression. Biol. Fertil. Soils 2015, 51, 553–562. [Google Scholar] [CrossRef]
  179. Ochoa-Velasco, C.E.; Valadez-Blanco, R.; Salas-Coronado, R.; Sustaita-Rivera, F.; Hernández-Carlos, B.; García-Ortega, S.; Santos-Sánchez, N.F. Effect of nitrogen fertilization and Bacillus licheniformis biofertilizer addition on the antioxidants compounds and antioxidant activity of greenhouse cultivated tomato fruits (Solanum lycopersicum L. var. Sheva). Sci. Hortic. 2016, 201, 338–345. [Google Scholar] [CrossRef]
  180. Kantachote, D.; Nunkaew, T.; Kantha, T.; Chaiprapat, S. Biofertilizers from Rhodopseudomonas palustris strains to enhance rice yields and reduce methane emissions. Appl. Soil Ecol. 2016, 100, 154–161. [Google Scholar] [CrossRef]
  181. Ge, B.; Liu, B.; Nwet, T.T.; Zhao, W.; Shi, L.; Zhang, K. Bacillus methylotrophicus Strain NKG-1, isolated from Changbai Mountain, China, has potential applications as a biofertilizer or biocontrol Agent. PLoS ONE 2016, 11, e0166079. [Google Scholar] [CrossRef] [PubMed]
  182. Tamreihao, K.; Ningthoujam, D.S.; Nimaichand, S.; Singh, E.S.; Reena, P.; Singh, S.H.; Nongthomba, U. Biocontrol and plant growth promoting activities of a Streptomyces corchorusii strain UCR3-16 and preparation of powder formulation for application as biofertilizer agents for rice plant. Microbiol. Res. 2016, 192, 260–270. [Google Scholar] [CrossRef] [PubMed]
  183. Shang, C.; Chen, A.; Chen, G.; Li, H.; Guan, S.; He, J. Microbial Biofertilizer Decreases Nicotine Content by Improving Soil Nitrogen Supply. Appl. Biochem. Biotechnol. 2017, 181, 1–14. [Google Scholar] [CrossRef]
  184. Mukhtar, S.; Shahid, I.; Mehnaz, S.; Malik, K.A. Assessment of two carrier materials for phosphate solubilizing biofertilizers and their effect on growth of wheat (Triticum aestivum L.). Microbiol. Res. 2017, 205, 107–117. [Google Scholar] [CrossRef] [PubMed]
  185. Kumar, A.; Usmani, Z.; Kumar, V. Biochar and flyash inoculated with plant growth promoting rhizobacteria act as potential biofertilizer for luxuriant growth and yield of tomato plant. J. Environ. Manag. 2017, 190, 20–27. [Google Scholar]
  186. Goyal, P.; Swaroop, N.; Thomas, T. Effect of Organic Inorganic Bio Fertilizer and Seed Inoculation on Soil Properties, Growth and Yield of Maize (Zea mays L.) Var. Hybrid MM-2255. Int. J. Curr. Microbiol. App. Sci 2017, 6, 95–100. [Google Scholar] [CrossRef]
  187. Sapre, S.; Gontia-Mishra, I.; Tiwari, S. Klebsiella sp. confers enhanced tolerance to salinity and plant growth promotion in oat seedlings (Avena sativa). Microbiol. Res. 2018, 206, 25–32. [Google Scholar] [CrossRef]
  188. Colla, G.; Rouphael, Y.; Cardarelli, M.; Tullio, M.; Rivera, C.M.; Rea, E. Alleviation of salt stress by arbuscular mycorrhizal in zucchini plants grown at low and high phosphorus concentration. Biol. Fertil. Soils 2008, 44, 501–509. [Google Scholar] [CrossRef]
  189. Wu, Q.; Zou, Y. Mycorrhizal influence on nutrient uptake of citrus exposed to drought stress. Philipp. Agric. Sci. 2009, 92, 33–38. [Google Scholar]
  190. Chandanie, W.; Kubota, M.; Hyakumachi, M. Interactions between the arbuscular mycorrhizal fungus Glomus mosseae and plant growth-promoting fungi and their significance for enhancing plant growth and suppressing damping-off of cucumber (Cucumis sativus L.). Appl. Soil Ecol. 2009, 41, 336–341. [Google Scholar] [CrossRef]
  191. Djebali, N.; Turki, S.; Zid, M.; Hajlaoui, M. Growth and development responses of some legume species inoculated with a mycorrhiza-based biofertilizer. Agr. Biol. J. N. Am. 2010, 1, 748–754. [Google Scholar] [CrossRef]
  192. Ogbo, F.C. Conversion of cassava wastes for biofertilizer production using phosphate solubilizing fungi. Bioresour. Technol. 2010, 101, 4120–4124. [Google Scholar] [CrossRef]
  193. Latef, A.A.H.A.; Chaoxing, H. Effect of arbuscular mycorrhizal fungi on growth, mineral nutrition, antioxidant enzymes activity and fruit yield of tomato grown under salinity stress. Sci. Hortic. 2011, 127, 228–233. [Google Scholar] [CrossRef]
  194. Srivastava, R.; Mehta, C.; Sharma, A. Fusarium pallidoroseum–A new biofertilizer responsible for enhancing plant growth in different crops. Int. Res. J. Microbiol. 2011, 2, 192–199. [Google Scholar]
  195. Malviya, J.; Singh, K.; Joshi, V. Effect of phosphate solubilizing fungi on growth and nutrient uptake of ground nut (Arachis hypogaea) plants. Adv. Biores 2011, 2, 110–113. [Google Scholar]
  196. Boonlue, S.; Surapat, W.; Pukahuta, C.; Suwanarit, P.; Suwanarit, A.; Morinaga, T. Diversity and efficiency of arbuscular mycorrhizal fungi in soils from organic chili (Capsicum frutescens) farms. Mycoscience 2012, 53, 10–16. [Google Scholar] [CrossRef]
  197. Andrade, M.M.M.; Stamford, N.P.; Santos, C.E.R.; Freitas, A.D.S.; Sousa, C.A.; Junior, M.A.L. Effects of biofertilizer with diazotrophic bacteria and mycorrhizal fungi in soil attribute, cowpea nodulation yield and nutrient uptake in field conditions. Sci. Hortic. 2013, 162, 374–379. [Google Scholar] [CrossRef]
  198. Saxena, J.; Saini, A.; Ravi, I.; Chandra, S.; Garg, V. Consortium of phosphate-solubilizing bacteria and fungi for promotion of growth and yield of chickpea (Cicer arietinum). J. Crop Improv. 2015, 29, 353–369. [Google Scholar] [CrossRef]
  199. Bhattacharjee, S.; Sharma, G. Effect of Arbuscular mycorrhizal fungi (AM fungi) and Rhizobium on the nutrient uptake of pigeon pea plant. Int. J. Adv. Res. 2015, 3, 833–836. [Google Scholar]
  200. Bai, B.; Suri, V.K.; Kumar, A.; Choudhary, A.K. Influence of Dual Inoculation of AM Fungi and Rhizobium on Growth Indices, Production Economics, and Nutrient Use Efficiencies in Garden Pea (Pisum sativum L.). Commun. Soil Sci. Plant Anal. 2016, 47, 941–954. [Google Scholar] [CrossRef]
  201. Bona, E.; Cantamessa, S.; Massa, N.; Manassero, P.; Marsano, F.; Copetta, A.; Lingua, G.; D’Agostino, G.; Gamalero, E.; Berta, G. Arbuscular mycorrhizal fungi and plant growth-promoting pseudomonads improve yield, quality and nutritional value of tomato: A field study. Mycorrhiza 2017, 27, 1–11. [Google Scholar] [CrossRef] [PubMed]
  202. Tarraf, W.; Ruta, C.; Tagarelli, A.; De Cillis, F.; De Mastro, G. Influence of arbuscular mycorrhizae on plant growth, essential oil production and phosphorus uptake of Salvia officinalis L. Ind. Crops Prod. 2017, 102, 144–153. [Google Scholar] [CrossRef]
  203. Poorabdollah, A.; Gholamalizadeh-Ahangar, A.; Sabbagh, S.; Sirousmehr, A. Bio-fertilizers and Systemic Acquired Resistance in Fusarium Infected Wheat. J. Agr. Sci. Tech. 2018, 19, 453–464. [Google Scholar]
  204. Mathur, S.; Sharma, M.P.; Jajoo, A. Improved photosynthetic efficacy of maize (Zea mays) plants with arbuscular mycorrhizal fungi (AMF) under high temperature stress. J. Photochem. Photobiol. B Biol. 2018, 180, 149–154. [Google Scholar] [CrossRef]
  205. Hoffman, B.M.; Lukoyanov, D.; Yang, Z.-Y.; Dean, D.R.; Seefeldt, L.C. Mechanism of nitrogen fixation by nitrogenase: The next stage. Chem. Rev. 2014, 114, 4041–4062. [Google Scholar] [CrossRef]
  206. Yang, H.; Li, Y.; Cao, Y.; Shi, W.; Xie, E.; Mu, N.; Du, G.; Shen, Y.; Tang, D.; Cheng, Z. Nitrogen nutrition contributes to plant fertility by affecting meiosis initiation. Nat. Commun. 2022, 13, 485. [Google Scholar] [CrossRef]
  207. Schwabedissen, S.; Reed, S.; Lohse, K.; Magnuson, T. Climatic and Grazing Controls on Biological Soil Crust Nitrogen Fixation in Semi-arid Ecosystems. In Proceedings of AGU Fall Meeting Abstracts; American Geophysical Union: Washington, DC, USA, 2014. [Google Scholar]
  208. Sheffer, E.; Batterman, S.A.; Levin, S.A.; Hedin, L.O. Biome-scale nitrogen fixation strategies selected by climatic constraints on nitrogen cycle. Nat. Plants 2015, 1, 15182. [Google Scholar] [CrossRef]
  209. Pankievicz, V.; Amaral, F.P.; Santos, K.F.; Agtuca, B.; Xu, Y.; Schueller, M.J.; Arisi, A.C.M.; Steffens, M.; Souza, E.M.; Pedrosa, F.O. Robust biological nitrogen fixation in a model grass–bacterial association. Plant J. 2015, 81, 907–919. [Google Scholar] [CrossRef]
  210. Tilak, K.; Ranganayaki, N.; Pal, K.; De, R.; Saxena, A.; Nautiyal, C.S.; Mittal, S.; Tripathi, A.; Johri, B. Diversity of plant growth and soil health supporting bacteria. Curr. Sci. 2005, 89, 136–150. [Google Scholar]
  211. Egamberdiyeva, D. Plant-growth-promoting rhizobacteria isolated from a Calcisol in a semi-arid region of Uzbekistan: Biochemical characterization and effectiveness. J. Plant Nutr. Soil Sci. 2005, 168, 94–99. [Google Scholar] [CrossRef]
  212. Jurelevicius, D.; Korenblum, E.; Casella, R.; Vital, R.L.; Seldin, L. Polyphasic analysis of the bacterial community in the rhizosphere and roots of Cyperus rotundus L. grown in a petroleum-contaminated soil. J. Microbiol. Biotechnol. 2010, 20, 862–870. [Google Scholar] [CrossRef] [PubMed]
  213. Paz-Ares, J.; Puga, M.I.; Rojas-Triana, M.; Martinez-Hevia, I.; Diaz, S.; Poza-Carrión, C.; Miñambres, M.; Leyva, A. Plant adaptation to low phosphorus availability: Core signaling, crosstalks and applied implications. Mol. Plant 2022, 15, 104–124. [Google Scholar] [CrossRef] [PubMed]
  214. Kamilova, F.; Kravchenko, L.V.; Shaposhnikov, A.I.; Azarova, T.; Makarova, N.; Lugtenberg, B. Organic acids, sugars, and L-tryptophane in exudates of vegetables growing on stonewool and their effects on activities of rhizosphere bacteria. Mol. Plant-Microbe Interact. 2006, 19, 250–256. [Google Scholar] [CrossRef] [PubMed]
  215. del Carmen Rivera-Cruz, M.; Narcía, A.T.; Ballona, G.C.; Kohler, J.; Caravaca, F.; Roldan, A. Poultry manure and banana waste are effective biofertilizer carriers for promoting plant growth and soil sustainability in banana crops. Soil Biol. Biochem. 2008, 40, 3092–3095. [Google Scholar]
  216. Jain, R.; Saxena, J.; Sharma, V. Effect of phosphate-solubilizing fungi Aspergillus awamori S29 on mungbean (Vigna radiata cv. RMG 492) growth. Folia Microbiol. 2012, 57, 533–541. [Google Scholar] [CrossRef]
  217. Jain, R.; Saxena, J.; Sharma, V. The evaluation of free and encapsulated Aspergillus awamori for phosphate solubilization in fermentation and soil–plant system. Appl. Soil Ecol. 2010, 46, 90–94. [Google Scholar] [CrossRef]
  218. Kamali, S.; Singh, A. Jasmonates as emerging regulators of plants response to variable nutrient environment. Crit. Rev. Plant Sci. 2022, 41, 271–285. [Google Scholar] [CrossRef]
  219. Ankit, A.; Singh, A.; Kumar, S.; Singh, A. Morphophysiological and transcriptome analysis reveal that reprogramming of metabolism, phytohormones and root development pathways governs the potassium (K+) deficiency response in two contrasting chickpea cultivars. Front. Plant Sci. 2023, 13, 1054821. [Google Scholar] [CrossRef]
  220. Parmar, P.; Sindhu, S. Potassium solubilization by rhizosphere bacteria: Influence of nutritional and environmental conditions. J. Microbiol. Res. 2013, 3, 25–31. [Google Scholar]
  221. Kumar, P.; Dubey, R. Plant growth promoting rhizobacteria for biocontrol of phytopathogens and yield enhancement of Phaseolus vulgaris L. J. Curr. Pers. Appl. Microbiol. 2012, 1, 6–38. [Google Scholar]
  222. Girgis, M. Response of wheat to inoculation with phosphate and potassium mobilizers and organic amendment. Ann. Agric. Sci.-Cairo- 2006, 51, 85. [Google Scholar]
  223. Gontia-Mishra, I.; Sapre, S.; Sharma, A.; Tiwari, S. Alleviation of mercury toxicity in wheat by the interaction of mercury-tolerant plant growth-promoting rhizobacteria. J. Plant Growth Regul. 2016, 35, 1000–1012. [Google Scholar] [CrossRef]
  224. Ramesh, A.; Sharma, S.K.; Sharma, M.P.; Yadav, N.; Joshi, O.P. Inoculation of zinc solubilizing Bacillus aryabhattai strains for improved growth, mobilization and biofortification of zinc in soybean and wheat cultivated in Vertisols of central India. Appl. Soil Ecol. 2014, 73, 87–96. [Google Scholar] [CrossRef]
  225. Wissuwa, M.; Ismail, A.M.; Yanagihara, S. Effects of zinc deficiency on rice growth and genetic factors contributing to tolerance. Plant Physiol. 2006, 142, 731–741. [Google Scholar] [CrossRef]
  226. Shanmugam, P.; Veeraputhran, R. Effect of organic manure, biofertilizers, inorganic nitrogen and zinc on growth and yield of rabi rice (Oryza sativa L.). Madras Agric. J. 2000, 87, 90–93. [Google Scholar]
  227. Katel, S.; Yadav, S.P.S.; Sharma, B. Impacts of plant growth regulators in strawberry plant: A review. Heliyon 2022, 8, e11959. [Google Scholar] [CrossRef]
  228. Araújo, F.F.; Henning, A.A.; Hungria, M. Phytohormones and antibiotics produced by Bacillus subtilis and their effects on seed pathogenic fungi and on soybean root development. World J. Microbiol. Biotechnol. 2005, 21, 1639–1645. [Google Scholar] [CrossRef]
  229. Davies, P.J. The plant hormones: Their nature, occurrence, and functions. In Plant Hormones; Springer: Berlin/Heidelberg, Germany, 2010; pp. 1–15. [Google Scholar]
  230. Venieraki, A.; Dimou, M.; Vezyri, E.; Kefalogianni, I.; Argyris, N.; Liara, G.; Pergalis, P.; Chatzipavlidis, I.; Katinakis, P. Characterization of nitrogen-fixing bacteria isolated from field-grown barley, oat, and wheat. J. Microbiol. 2011, 49, 525–534. [Google Scholar] [CrossRef]
  231. Ryu, R.J.; Patten, C.L. Aromatic amino acid-dependent expression of indole-3-pyruvate decarboxylase is regulated by TyrR in Enterobacter cloacae UW5. J. Bacteriol. 2008, 190, 7200–7208. [Google Scholar] [CrossRef]
  232. Chen, Y.; Rekha, P.; Arun, A.; Shen, F.; Lai, W.-A.; Young, C. Phosphate solubilizing bacteria from subtropical soil and their tricalcium phosphate solubilizing abilities. Appl. Soil Ecol. 2006, 34, 33–41. [Google Scholar] [CrossRef]
  233. Akram, M.S.; Shahid, M.; Tariq, M.; Azeem, M.; Javed, M.T.; Saleem, S.; Riaz, S. Deciphering Staphylococcus sciuri SAT-17 mediated anti-oxidative defense mechanisms and growth modulations in salt stressed maize (Zea mays L.). Front. Microbiol. 2016, 7, 867. [Google Scholar] [CrossRef] [PubMed]
  234. Teale, W.D.; Paponov, I.A.; Palme, K. Auxin in action: Signalling, transport and the control of plant growth and development. Nat. Rev. Mol. Cell Biol. 2006, 7, 847. [Google Scholar] [CrossRef] [PubMed]
  235. Schruff, M.C.; Spielman, M.; Tiwari, S.; Adams, S.; Fenby, N.; Scott, R.J. The AUXIN RESPONSE FACTOR 2 gene of Arabidopsis links auxin signalling, cell division, and the size of seeds and other organs. Development 2006, 133, 251–261. [Google Scholar] [CrossRef] [PubMed]
  236. Spaepen, S.; Vanderleyden, J.; Remans, R. Indole-3-acetic acid in microbial and microorganism-plant signaling. FEMS Microbiol. Rev. 2007, 31, 425–448. [Google Scholar] [CrossRef]
  237. Ahmad, F.; Ahmad, I.; Khan, M.S. Indole acetic acid production by the indigenous isolates of Azotobacter and fluorescent Pseudomonas in the presence and absence of tryptophan. Turk. J. Biol. 2005, 29, 29–34. [Google Scholar]
  238. Spaepen, S.; Bossuyt, S.; Engelen, K.; Marchal, K.; Vanderleyden, J. Phenotypical and molecular responses of Arabidopsis thaliana roots as a result of inoculation with the auxin-producing bacterium Azospirillum brasilense. N. Phytol. 2014, 201, 850–861. [Google Scholar] [CrossRef]
  239. Khalid, A.; Arshad, M.; Zahir, Z. Screening plant growth-promoting rhizobacteria for improving growth and yield of wheat. J. Appl. Microbiol. 2004, 96, 473–480. [Google Scholar] [CrossRef]
  240. Teixeira, D.A.; Alfenas, A.C.; Mafia, R.G.; Ferreira, E.M.; de Siqueira, L.; Maffia, L.A.; Mounteer, A.H. Rhizobacterial promotion of eucalypt rooting and growth. Braz. J. Microbiol. 2007, 38, 118–123. [Google Scholar] [CrossRef]
  241. Lee, J.; Jo, N.-Y.; Shim, S.-Y.; Linh, L.T.Y.; Kim, S.-R.; Lee, M.-G.; Hwang, S.-G. Effects of Hanwoo (Korean cattle) manure as organic fertilizer on plant growth, feed quality, and soil bacterial community. Front. Plant Sci. 2023, 14, 1135947. [Google Scholar] [CrossRef]
  242. Yi, S.; Li, F.; Wu, C.; Wei, M.; Tian, J.; Ge, F. Synergistic leaching of heavy metal-polycyclic aromatic hydrocarbon in co-contaminated soil by hydroxamate siderophore: Role of cation-π and chelation. J. Hazard. Mater. 2022, 424, 127514. [Google Scholar] [CrossRef] [PubMed]
  243. Zhou, D.; Huang, X.-F.; Chaparro, J.M.; Badri, D.V.; Manter, D.K.; Vivanco, J.M.; Guo, J. Root and bacterial secretions regulate the interaction between plants and PGPR leading to distinct plant growth promotion effects. Plant Soil 2016, 401, 259–272. [Google Scholar] [CrossRef]
  244. Krewulak, K.D.; Vogel, H.J. Structural biology of bacterial iron uptake. Biochim. Biophys. Acta (BBA)-Biomembr. 2008, 1778, 1781–1804. [Google Scholar] [CrossRef] [PubMed]
  245. Osorio, H.; Martínez, V.; Nieto, P.A.; Holmes, D.S.; Quatrini, R. Microbial iron management mechanisms in extremely acidic environments: Comparative genomics evidence for diversity and versatility. BMC Microbiol. 2008, 8, 203. [Google Scholar] [CrossRef] [PubMed]
  246. Andrews, S.C.; Robinson, A.K.; Rodríguez-Quiñones, F. Bacterial iron homeostasis. FEMS Microbiol. Rev. 2003, 27, 215–237. [Google Scholar] [CrossRef]
  247. Crowley, D.E. Microbial siderophores in the plant rhizosphere. In Iron Nutrition in Plants and Rhizospheric Microorganisms; Springer: Berlin/Heidelberg, Germany, 2006; pp. 169–198. [Google Scholar]
  248. Taie, H.A.; El-Mergawi, R.; Radwan, S. Isoflavonoids, flavonoids, phenolic acids profiles and antioxidant activity of soybean seeds as affected by organic and bioorganic fertilization. Am.-Eurasian J. Agric. Environ. Sci. 2008, 4, 207–213. [Google Scholar]
  249. Sanalibaba, P.; Çakmak, G. Exopolysaccharides production by lactic acid bacteria. Appl. Microbiol. Open Access 2016, 2, 1000115. [Google Scholar] [CrossRef]
  250. Grover, M.; Ali, S.Z.; Sandhya, V.; Rasul, A.; Venkateswarlu, B. Role of microorganisms in adaptation of agriculture crops to abiotic stresses. World J. Microbiol. Biotechnol. 2011, 27, 1231–1240. [Google Scholar] [CrossRef]
  251. Mahmood, S.; Daur, I.; Al-Solaimani, S.G.; Ahmad, S.; Madkour, M.H.; Yasir, M.; Hirt, H.; Ali, S.; Ali, Z. Plant growth promoting rhizobacteria and silicon synergistically enhance salinity tolerance of mung bean. Front. Plant Sci. 2016, 7, 876. [Google Scholar] [CrossRef]
  252. Blumer, C.; Haas, D. Mechanism, regulation, and ecological role of bacterial cyanide biosynthesis. Arch. Microbiol. 2000, 173, 170–177. [Google Scholar] [CrossRef]
  253. Dhole, A.; Shelat, H. Non-Rhizobial Endophytes Associated with Nodules of Vigna radiata L. and Their Combined Activity with Rhizobium sp. Curr. Microbiol. 2022, 79, 103. [Google Scholar] [CrossRef] [PubMed]
  254. Siddiqui, Z.A. PGPR: Prospective biocontrol agents of plant pathogens. In PGPR: Biocontrol and Biofertilization; Springer: Berlin/Heidelberg, Germany, 2005; pp. 111–142. [Google Scholar]
  255. Martínez-Viveros, O.; Jorquera, M.; Crowley, D.; Gajardo, G.; Mora, M. Mechanisms and practical considerations involved in plant growth promotion by rhizobacteria. J. Soil Sci. Plant Nutr. 2010, 10, 293–319. [Google Scholar] [CrossRef]
  256. Rudrappa, T.; Splaine, R.E.; Biedrzycki, M.L.; Bais, H.P. Cyanogenic pseudomonads influence multitrophic interactions in the rhizosphere. PLoS ONE 2008, 3, e2073. [Google Scholar] [CrossRef] [PubMed]
  257. Xiang, L.; Harindintwali, J.D.; Wang, F.; Redmile-Gordon, M.; Chang, S.X.; Fu, Y.; He, C.; Muhoza, B.; Brahushi, F.; Bolan, N. Integrating biochar, bacteria, and plants for sustainable remediation of soils contaminated with organic pollutants. Environ. Sci. Technol. 2022, 56, 16546–16566. [Google Scholar] [CrossRef]
  258. Desbrosses, G.; Contesto, C.; Varoquaux, F.; Galland, M.; Touraine, B. PGPR-Arabidopsis interactions is a useful system to study signaling pathways involved in plant developmental control. Plant Signal. Behav. 2009, 4, 319–321. [Google Scholar] [CrossRef]
  259. Reed, M.L.; Glick, B.R. Growth of canola (Brassica napus) in the presence of plant growth-promoting bacteria and either copper or polycyclic aromatic hydrocarbons. Can. J. Microbiol. 2005, 51, 1061–1069. [Google Scholar] [CrossRef]
  260. Saleem, M.; Arshad, M.; Hussain, S.; Bhatti, A.S. Perspective of plant growth promoting rhizobacteria (PGPR) containing ACC deaminase in stress agriculture. J. Ind. Microbiol. Biotechnol. 2007, 34, 635–648. [Google Scholar] [CrossRef]
  261. Hameeda, B.; Rupela, O.; Reddy, G.; Satyavani, K. Application of plant growth-promoting bacteria associated with composts and macrofauna for growth promotion of Pearl millet (Pennisetum glaucum L.). Biol. Fertil. Soils 2006, 43, 221–227. [Google Scholar] [CrossRef]
  262. López, A.C.; Giorgio, E.M.; Vereschuk, M.L.; Zapata, P.D.; Luna, M.F.; Alvarenga, A.E. Ilex paraguariensis Hosts Root-Trichoderma spp. with Plant-Growth-Promoting Traits: Characterization as Biological Control Agents and Biofertilizers. Current Microbiol. 2023, 80, 120. [Google Scholar] [CrossRef]
  263. Silveira, A.B.d.; Ambrosini, A.; Passaglia, L.M.P. Plant growth-promoting rhizobacteria (PGPR): Their potential as antagonists and biocontrol agents. Genet. Mol. Biol. 2012, 35, 1044–1051. [Google Scholar]
  264. Kloepper, J.W.; Ryu, C.-M.; Zhang, S. Induced systemic resistance and promotion of plant growth by Bacillus spp. Phytopathology 2004, 94, 1259–1266. [Google Scholar] [CrossRef] [PubMed]
  265. Liu, K.; Garrett, C.; Fadamiro, H.; Kloepper, J.W. Induction of systemic resistance in Chinese cabbage against black rot by plant growth-promoting rhizobacteria. Biol. Control 2016, 99, 8–13. [Google Scholar] [CrossRef]
  266. Xu, Y.; Chen, Z.; Li, X.; Tan, J.; Liu, F.; Wu, J. Mycorrhizal fungi alter root exudation to cultivate a beneficial microbiome for plant growth. Funct. Ecol. 2023, 37, 664–675. [Google Scholar] [CrossRef]
  267. Basu, S.; Priyadarshini, P.; Prasad, R.; Kumar, G. Effects of microbial signaling in plant growth and development. In Beneficial Microorganisms in Agriculture; Springer: Berlin/Heidelberg, Germany, 2022; pp. 329–348. [Google Scholar]
  268. Dakora, F.D.; Matiru, V.N.; Kanu, A.S. Rhizosphere ecology of lumichrome and riboflavin, two bacterial signal molecules eliciting developmental changes in plants. Front. Plant Sci. 2015, 6, 700. [Google Scholar] [CrossRef] [PubMed]
  269. Mitra, D.; Mondal, R.; Khoshru, B.; Senapati, A.; Radha, T.; Mahakur, B.; Uniyal, N.; Myo, E.M.; Boutaj, H.; Sierra, B.E.G. Actinobacteria-enhanced plant growth, nutrient acquisition, and crop protection: Advances in soil, plant, and microbial multifactorial interactions. Pedosphere 2022, 32, 149–170. [Google Scholar] [CrossRef]
  270. Muleta, D.; Assefa, F.; Granhall, U. In vitro antagonism of rhizobacteria isolated from Coffea arabica L. against emerging fungal coffee pathogens. Eng. Life Sci. 2007, 7, 577–586. [Google Scholar] [CrossRef]
  271. Latha, P.; Anand, T.; Ragupathi, N.; Prakasam, V.; Samiyappan, R. Antimicrobial activity of plant extracts and induction of systemic resistance in tomato plants by mixtures of PGPR strains and Zimmu leaf extract against Alternaria solani. Biol. Control 2009, 50, 85–93. [Google Scholar] [CrossRef]
  272. Pinton, R.; Varanini, Z.; Nannipieri, P. The rhizosphere as a site of biochemical interactions among soil components, plants, and microorganisms. In The Rhizosphere; CRC Press: Boca Raton, FL, USA, 2001. [Google Scholar]
  273. Prashar, P.; Kapoor, N.; Sachdeva, S. Rhizosphere: Its structure, bacterial diversity and significance. Rev. Environ. Sci. Bio/Technol. 2014, 13, 63–77. [Google Scholar] [CrossRef]
  274. Mikaelyan, A.; Babayan, B.; Vartanyan, A.; Tokmajyan, H. Tartaric acid synthetic derivatives effect on phytopathogenic bacteria. Agron. Res. 2022, 20, 644–659. [Google Scholar]
  275. Loper, J.E.; Gross, H. Genomic analysis of antifungal metabolite production by Pseudomonas fluorescens Pf-5. Eur. J. Plant Pathol. 2007, 119, 265–278. [Google Scholar] [CrossRef]
  276. de Souza, J.T.; Arnould, C.; Deulvot, C.; Lemanceau, P.; Gianinazzi-Pearson, V.; Raaijmakers, J.M. Effect of 2, 4-diacetylphloroglucinol on Pythium: Cellular responses and variation in sensitivity among propagules and species. Phytopathology 2003, 93, 966–975. [Google Scholar] [CrossRef]
  277. Maksimov, I.; Abizgil’Dina, R.; Pusenkova, L. Plant growth promoting rhizobacteria as alternative to chemical crop protectors from pathogens (review). Appl. Biochem. Microbiol. 2011, 47, 333–345. [Google Scholar] [CrossRef]
  278. Ahmad, R.; Arshad, M.; Khalid, A.; Zahir, Z.A. Effectiveness of organic-/bio-fertilizer supplemented with chemical fertilizers for improving soil water retention, aggregate stability, growth and nutrient uptake of maize (Zea mays L.). J. Sustain. Agric. 2008, 31, 57–77. [Google Scholar] [CrossRef]
  279. Koza, N.A.; Adedayo, A.A.; Babalola, O.O.; Kappo, A.P. Microorganisms in plant growth and development: Roles in abiotic stress tolerance and secondary metabolites secretion. Microorganisms 2022, 10, 1528. [Google Scholar] [CrossRef]
  280. Idris, E.E.; Iglesias, D.J.; Talon, M.; Borriss, R. Tryptophan-dependent production of indole-3-acetic acid (IAA) affects level of plant growth promotion by Bacillus amyloliquefaciens FZB42. Mol. Plant-Microbe Interact. 2007, 20, 619–626. [Google Scholar] [CrossRef] [PubMed]
  281. Tahir, H.A.; Gu, Q.; Wu, H.; Raza, W.; Hanif, A.; Wu, L.; Colman, M.V.; Gao, X. Plant growth promotion by volatile organic compounds produced by Bacillus subtilis SYST2. Front. Microbiol. 2017, 8, 171. [Google Scholar] [CrossRef] [PubMed]
  282. Wakchaure, V.N.; Zhou, J.; Hoffmann, S.; List, B. Catalytic Asymmetric Reductive Amination of α-Branched Ketones. Angew. Chem. 2010, 122, 4716–4718. [Google Scholar] [CrossRef]
  283. Asses, N.; Farhat, A.; Cherif, S.; Hamdi, M.; Bouallagui, H. Comparative study of sewage sludge co-composting with olive mill wastes or green residues: Process monitoring and agriculture value of the resulting composts. Process Saf. Environ. Prot. 2018, 114, 25–35. [Google Scholar] [CrossRef]
  284. Zhang, J.; Chen, G.; Sun, H.; Zhou, S.; Zou, G. Straw biochar hastens organic matter degradation and produces nutrient-rich compost. Bioresour. Technol. 2016, 200, 876–883. [Google Scholar] [CrossRef]
  285. Hassen, A.; Belguith, K.; Jedidi, N.; Cherif, A.; Cherif, M.; Boudabous, A. Microbial characterization during composting of municipal solid waste. Bioresour. Technol. 2001, 80, 217–225. [Google Scholar] [CrossRef]
  286. Manios, T. The composting potential of different organic solid wastes: Experience from the island of Crete. Environ. Int. 2004, 29, 1079–1089. [Google Scholar] [CrossRef] [PubMed]
  287. Kavitha, R.; Subramanian, P. Bioactive compost-a value added compost with microbial inoculants and organic additives. J. Appl. Sci. 2007, 7, 2514–2518. [Google Scholar] [CrossRef]
  288. Adhikari, B.K.; Barrington, S.; Martinez, J.; King, S. Effectiveness of three bulking agents for food waste composting. Waste Manag. 2009, 29, 197–203. [Google Scholar] [CrossRef] [PubMed]
  289. Vázquez, M.; de la Varga, D.; Plana, R.; Soto, M. Nitrogen losses and chemical parameters during co-composting of solid wastes and liquid pig manure. Environ. Technol. 2017, 39, 2017–2029. [Google Scholar] [CrossRef] [PubMed]
  290. Das, S.; Bhattacharyya, B.K. Selective organic fraction of municipal solid waste degradation under controlled composting conditions. J. Environ. Waste Manag. 2017, 4, 156–163. [Google Scholar]
  291. Gao, M.; Liang, F.; Yu, A.; Li, B.; Yang, L. Evaluation of stability and maturity during forced-aeration composting of chicken manure and sawdust at different C/N ratios. Chemosphere 2010, 78, 614–619. [Google Scholar] [CrossRef] [PubMed]
  292. Sanchez-Monedero, M.; Cayuela, M.; Roig, A.; Jindo, K.; Mondini, C.; Bolan, N. Role of biochar as an additive in organic waste composting. Bioresour. Technol. 2017, 247, 1155–1164. [Google Scholar] [CrossRef]
  293. Zang, B.; Li, S.; Michel Jr, F.; Li, G.; Luo, Y.; Zhang, D.; Li, Y. Effects of mix ratio, moisture content and aeration rate on sulfur odor emissions during pig manure composting. Waste Manag. 2016, 56, 498–505. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the composting process.
Figure 1. Schematic representation of the composting process.
Plants 12 03550 g001
Figure 2. Schematic representation of the compost-based biofertilizers. Mycorrhizal fungal filaments and plant-growth-promoting rhizobacteria in the soil act as to support the development of the plant root system and are more effective at water and nutrient absorption than the roots themselves. PGPR and AMF also explore the soil and reach places unattainable to roots and increase nutrient uptake by plants from the soil. Their effects on plant growth through two different mechanisms, such as direct mechanism and indirect mechanism, are illustrated.
Figure 2. Schematic representation of the compost-based biofertilizers. Mycorrhizal fungal filaments and plant-growth-promoting rhizobacteria in the soil act as to support the development of the plant root system and are more effective at water and nutrient absorption than the roots themselves. PGPR and AMF also explore the soil and reach places unattainable to roots and increase nutrient uptake by plants from the soil. Their effects on plant growth through two different mechanisms, such as direct mechanism and indirect mechanism, are illustrated.
Plants 12 03550 g002
Table 2. Effects of inoculation with PGPR- and AMF-based biofertilizers on the plant, physiological performance under different plant growth promotion mechanisms.
Table 2. Effects of inoculation with PGPR- and AMF-based biofertilizers on the plant, physiological performance under different plant growth promotion mechanisms.
BiofertilizersPlantsExperiment SitesResults/Discovery MadePlant Growth MechanismReference
PGPR-Based Biofertilizers
Bradyrhizobium sp.Wheat
(Triticum aestivum)
Pot experimentResults indicate that the inoculation with phosphate biofertilizer significantly enhances grain yield of wheatIAA production,
phosphate solubilization, siderophore production, and HCN production
[162]
Acinetobacter sp.Canola
(Brassica napus)
and Tomato
(Solanum lycopersicum)
Pot experimentResult showed that strain inoculation increased the root length and dry biomass of the canola test plantsIAA production,
phosphate solubilization,
ACC deaminase, and antifungal activity
[163]
Serratia mercescensWheat
(Triticum aestivum)
Pot experimentResult revealed that there was a significant increase in root and shoot lengths, and a significant increase in the root and shoot biomass was also observedIAA production,
siderophore production,
and HCN production
[164]
Pseudomonas striataChickpea
(Cicer arietinum L.)
Field experimentIt was concluded that there was significant seed grain or crop yield through inoculation with Pseudomonas striataNitrogen fixation and phosphate solubilization[165]
Azotobacter tropicalis,
Burkhoderia unamae, and Bacillus subtilis
Corn
(Zea mays)
Pot experimentResult showed that inoculation significantly enhanced the growth or yield and also increased the fresh and dry weight of cornAuxin production,
nitrogen fixation,
phosphate solubilization, and potassium solubilization
[166]
Azospirillum brasilense and Bacillus sphaericusBanana
(Musa spp. cv. ‘Berangan’)
Pot and field experimentResult showed that PGPR inoculation greatly increased the bunch yieldNitrogen fixation[167]
Pseudomonas, Rhodococcus, and Duganella sp.Moss
(Racomitrium japonicum)
Hydroponic experimentIt was concluded that isolate utilization should promote the moss growth and has potential to be utilized as biofertilizers for moss productionAuxin production, phosphate solubilization, siderophore production, HCN production, and antifungal activity[168]
Rhodobacter capsulatusRice
(Oryza sativa L.)
Field experimentResults revealed that both biological and grain yields in all the Rhodobacter capsulatus inoculated treatments were significantly higher than those in the uninoculated corresponding treatments in both fieldsNitrogen fixation[168]
Azotobacter chroococcum and Bacillus circulansHorseradish Tree (Moringa oleifera)Pot experimentResult showed that he highest records of shoot and root lengths, and shoot and root dry weights, were obtained with soil inoculation with mixed culturesIAA production,
cytokinin production, nitrogen fixation,
phosphate solubilization, and potassium solubilization
[169]
Azotobacter chroococcum and Azospirillum lipoferumCoriander
(Coriandrum Sativum)
Field experimentResults showed that the highest plant height, umbel number per plant, weight of 1000 seeds,
dry weight of plant, and seed yield were obtained by using the biofertilizer twice
Nitrogen fixation and phosphate solubilization[170]
Azotobacter sp., Nitrobacter sp., and Nitrosomonas sp.Tomato (Lycoperscum esculentus)Pot experimentResult indicated that combined biofertilizers are recommended for excellent growth performance of plantsIAA and siderophore production[171]
Klebsiella planticola, Azotobacter, and BacillusStrawberry (Fragaria × ananassa Duch)Pot experimentA significant effect on yield per plant was noted.Nitrogen fixation and phytohormones production[172]
Azotobacter chroococcum and Azospirillum lipoferumAjowan
(Carum copticum)
Field experimentIt was noted that biological yield, seed yield, essential oil content, and essential oil yield were obtained by using the biofertilizerNitrogen fixation[173]
Azotobacter chroococcum and Bacillus subtilisWheat
(Triticum aestivum L.)
Field experimentOrganic matrix can be more effective for achieving enhanced productivity of wheat in sub-tropical agro-climatic conditionsNitrogen fixation,
phosphate solubilization, and potassium solubilization
[174]
Rhizobium leguminosarumFaba bean
(Vicia faba L.)
Pot experimentEfficient bioinoculant development to enhance the tolerance of faba bean plants to alkalinity stress and thereby improve the fitness of plantsNitrogenase activity[175]
Acidobacteria and BacteroidetesBanana
(Cavendish)
Pot experimentSuppressed the strength of Fusarium wilt disease through improving soil chemical condition and manipulating the composition of soil microbial communityNitrogen fixation,
phosphate solubilization, and potassium solubilization
[176]
Bradyrhizobium yuanmingenseFabales
(Cajanus cajan L.)
Field experimentIt was evaluated that their potential as biofertilizers includes being able to replace mineral N fertilizationNitrogenase activity[177]
Acidobacteria, Actinobacteria and ProteobacteriaBanana
(Cavendish)
Field experimentResults showed that these bacterial strains are associated with banana Fusarium wilt disease suppressionAntagonistic activity[178]
Bacillus licheniformisTomato
(Lycoperscum esculentus)
Pot experimentThe combined effect of the nitrogen fertilization dose and biofertilizer addition during tomato cultivation on the content of antioxidant compounds in tomato fruits was shownNitrogenase activity[179]
Rhodopseudomonas palustrisRice
(Oryza sativa L.)
Field experimentIt was concluded that there should be an increase in rice yields in both the organic and saline-flooded paddy fields and concurrently, a reduction in CH4 emissionsNitrogenase activity[180]
Bacillus methylotrophicusTomato
(Lycoperscum esculentus)
Greenhouse experimentResults suggest that NKG-1 has potential for commercial application as a biofertilizer or biocontrol agentAntagonistic activity[181]
Streptomyces
Corchorusii
Rice
(Oryza sativa L.)
Pot experimentSignificant enhancement in shoot length and weight of shoot and root, and total grain yield, and weight of grains in rice plantsIAA production,
siderophore production, phosphate solubilization, ACC deaminase activity, and antagonistic activity
[182]
Bacillus thuringiensisTobacco
(Nicotiana)
Pot experimentResults revealed that there was reduction in nicotine content and improvement in tobacco quality, which may provide some useful information for tobacco cultivationNitrogenase activity[183]
Bacillus endophyticus, Bacillus sphaericus, Enterobacter aerogenes, Bacillus safensis, and Bacillus megateriumWheat (Triticum aestivum L.)Pot experiment and field experimentResults revealed that bacterial isolates with plant beneficial traits such as P solubilization are more promising candidates as biofertilizer when used with carrier materialsPhosphate solubilization[184]
Bacillus sp. and Burkholderia sp.Tomato (Lycoperscum esculentus)Pot experimentThe inoculum can tremendously enhance the productivity of tomato, soil fertility, and can also act as a sustainable substitute for chemical fertilizersIAA production,
siderophore production, and phosphate solubilization
[185]
AzotobacterMaize
(Zea mays L.)
Field experimentSeed Inoculation has a significant effect on soil properties, growth, and yield of maizeNitrogen fixation,
phosphate solubilization, and potassium solubilization
[186]
Klebsiella sp.Oat seedlings
(Avena sativa)
Field experimentDemonstrates that PGPR play an imperative function in stimulating salt tolerance in plants and can be used as biofertilizer to enhance growth of crops in saline areasAntagonistic activity[187]
AMF-Based Biofertilizer
Rhizophagus irregularisPumpkin
(Cucurbita pepo)
Pot experimentResults demonstrate that crop inoculation improved the detrimental effect of salinity on plant growth due to improved nutritional contentsPotassium solubilization[188]
Glomus versiformeTrifoliate orange
(Poncirus trifoliate)
Pot experimentAMF inoculation increased the fresh and dry weight and leaf area of seedlings under drought stressPotassium solubilization and phosphorous solubilization[189]
Glomus mosseaeCucumber (Cucumis sativus L.)Pot experimentResults revealed the significance of enhancing plant growth and suppressing damping off of cucumber achieved through inoculation with AMFAntibiotics production[190]
Glomus fasciculatum and Gigaspora sp.Bean
(Phaseolus) and Pea
(Pisum sativum)
Pot experimentInoculation led to an increase in shoots and increase in root dry weightsNitrogen fixation[191]
Aspergillus fumigatus and Aspergillus nigerPigeon pea (Cajanus cajan L.)Pot experimentTreated plants showed improved nutrient uptake in contrast to non-treated plantsPhosphorous solubilization[192]
Funneliformis mosseaeTomato
(Solanum lycopersicum)
Pot experimentInoculation confers resistance against salt stress to tomato plants via antioxidant enzymes (POD, CAT, etc.) productionHigher leaf concentration of phosphorous, potassium[193]
Fusarium pallidoroseumTomato
(Solanum lycopersicum),
Wheat (Triticum), and
Maize (Zea mays)
Pot experimentEnhanced proline content, acid and alkaline phosphomonoesterase activity, and peroxidase activity was observed in plants inoculated with Fusarium pallidoroseum. The fungus treatment also improved the shoot dry weight and shoot lengthPhosphorous solubilization and zinc solubilization[194]
Aspergillus niger and Penicillium notatumNut
(Arachis hypogaea)
Pot experimentPlants showed better nutrient uptake when treated with plant beneficial microbial inoculumNitrogen fixation and phosphorous solubilization[195]
Acaulospora spp., Entrophospora sp., Glomus spp., and ScutellosporaChili
(Capsicum frutescens L.)
Pot experimentHighest shoot, root length, and yield were observed in inoculated plantsNitrogen fixation, phosphorous solubilization, and potassium solubilization[196]
Glomus etunicatum and Gigaspora
albida
Cowpeas
(Vigna unguiculata)
Pot experiment and field experimentThe treatment resulted in improved grain and shoot biomass production and increased the available P and K levels in soilPhosphorous solubilization and potassium solubilization[197]
Acaulospora laevis, Glomus geosporum, Glomus mosseae,
and Scutellospora armeniaca
Faba bean
(Vicia faba L.)
Pot experimentThe faba bean inoculated with mixed culture of Rhizobium and AMF showed better growth under alkaline environmentNitrogen fixation and phosphorous solubilization[175]
Aspergillus nigerChickpea
(Cicer arietinum)
Pot experimentThe treated plants showed better growth and nutrient profile as compared with non-treated plantsPhosphorous solubilization[198]
Glomus fasciculatumPigeon pea
(Cajanus cajan L.)
Pot experimentImproved nutrient uptake and biomass production was observed in inoculated pigeon pea plantsNitrogen fixation, phosphorous solubilization, and potassium solubilization[199]
Glomus mosseaeGarden Pea (Pisum sativum L.)Pot experimentThe plants showed enhanced plant height, leaf area index, and dry matter accumulation when treated with Glomus mosseaeNitrogen fixation and phosphorous solubilization[200]
Glomus aggregatum, Glomus viscosumTomato
(Solanum lycopersicum L.)
Pot experimentThe inoculation improved the yield, quality, and nutritional value of tomato plantsNutrient solubilization[201]
Glomus viscosum(Salvia officinalis L.)Pot experimentBetter plant growth and biomass production was found in plants treated with fungiPhosphorous solubilization[202]
Glomus intraradicesWheat
(Triticum)
Pot experimentEnhanced nutrient uptake was observed in inoculated plantsNutrient solubilization[203]
Rhizophagus intraradices, Funneliformis mosseae, and F. geosporumMaize
(Zea mays)
Pot experimentMaize treated with AMF showed improved photosynthetic efficacy under high-temperature stressNutrient solubilization[204]
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

Ahmed, T.; Noman, M.; Qi, Y.; Shahid, M.; Hussain, S.; Masood, H.A.; Xu, L.; Ali, H.M.; Negm, S.; El-Kott, A.F.; et al. Fertilization of Microbial Composts: A Technology for Improving Stress Resilience in Plants. Plants 2023, 12, 3550. https://doi.org/10.3390/plants12203550

AMA Style

Ahmed T, Noman M, Qi Y, Shahid M, Hussain S, Masood HA, Xu L, Ali HM, Negm S, El-Kott AF, et al. Fertilization of Microbial Composts: A Technology for Improving Stress Resilience in Plants. Plants. 2023; 12(20):3550. https://doi.org/10.3390/plants12203550

Chicago/Turabian Style

Ahmed, Temoor, Muhammad Noman, Yetong Qi, Muhammad Shahid, Sabir Hussain, Hafiza Ayesha Masood, Lihui Xu, Hayssam M. Ali, Sally Negm, Attalla F. El-Kott, and et al. 2023. "Fertilization of Microbial Composts: A Technology for Improving Stress Resilience in Plants" Plants 12, no. 20: 3550. https://doi.org/10.3390/plants12203550

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