Next Article in Journal
Potential of Induced Pluripotent Stem Cells for Use in Gene Therapy: History, Molecular Bases, and Medical Perspectives
Next Article in Special Issue
Transcriptional Cascade in the Regulation of Flowering in the Bamboo Orchid Arundina graminifolia
Previous Article in Journal
Macrophages and Stem Cells—Two to Tango for Tissue Repair?
Previous Article in Special Issue
Modulation of the Antioxidant Defense System by Exogenous l-Glutamic Acid Application Enhances Salt Tolerance in Lentil (Lens culinaris Medik.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biostimulants Application: A Low Input Cropping Management Tool for Sustainable Farming of Vegetables

by
Mohamad Hesam Shahrajabian
*,
Christina Chaski
,
Nikolaos Polyzos
and
Spyridon A. Petropoulos
*
Department of Agriculture Crop Production and Rural Environment, University of Thessaly, 38446 Volos, Greece
*
Authors to whom correspondence should be addressed.
Biomolecules 2021, 11(5), 698; https://doi.org/10.3390/biom11050698
Submission received: 23 March 2021 / Revised: 4 May 2021 / Accepted: 6 May 2021 / Published: 7 May 2021

Abstract

:
Biostimulants, are a diverse class of compounds including substances or microorganism which have positive impacts on plant growth, yield and chemical composition as well as boosting effects to biotic and abiotic stress tolerance. The major plant biostimulants are hydrolysates of plant or animal protein and other compounds that contain nitrogen, humic substances, extracts of seaweeds, biopolymers, compounds of microbial origin, phosphite, and silicon, among others. The mechanisms involved in the protective effects of biostimulants are varied depending on the compound and/or crop and mostly related with improved physiological processes and plant morphology aspects such as the enhanced root formation and elongation, increased nutrient uptake, improvement in seed germination rates and better crop establishment, increased cation exchange, decreased leaching, detoxification of heavy metals, mechanisms involved in stomatal conductance and plant transpiration or the stimulation of plant immune systems against stressors. The aim of this review was to provide an overview of the application of plant biostimulants on different crops within the framework of sustainable crop management, aiming to gather critical information regarding their positive effects on plant growth and yield, as well as on the quality of the final product. Moreover, the main limitations of such practice as well as the future prospects of biostimulants research will be presented.

1. Introduction

The growing need for food production through sustainable cultivation practices, without reducing crop yield and producer income, is a major objective due to increased environmental pollution and the gradual degradation of cultivated soils [1]. In the context of global climate change and food security, there is a need for cultivating crops under unfavorable conditions, particularly in dry and semi-dry areas, as well as for the sustainable use of valuable and finite natural resources through the protection of biodiversity [2,3,4]. Various farming systems have been suggested throughout the last decades with biostimulants being a novel and sustainable approach towards crop production, especially under biotic and abiotic stressors [4,5]. The expected market growth in the biostimulant sector at a compound annual growth rate of 11.24% and up to USD 4.9 billion by 2025 [6]. Therefore, there is increasing interest in the farming sector for new biostimulant products and a lot of research is carried out in this gradually evolving section of the industry. There are several commercial products available which are currently applied on various crops within the context of sustainable and organic farming [7].
Various compounds with bioactive properties can be utilized as biostimulants to boost plant growth and development under normal and stressful conditions [8,9], while among the distinctive characteristics a biostimulatory product must improve nutrients use efficiency, tolerance to abiotic stressors, quality of the final product and nutrients availability in soil [10]. So far, six distinct categories of biostimulants are recognized, including microbial inoculants, humic substances, such as humic and fulvic acids, protein hydrolysates and amino acids, biopolymers, inorganic compounds. and seaweed extracts, all of which are commercially available with wide applications in agriculture [11,12]. Biostimulant application can be considered as an effective and sustainable nutritional crop supplementation and may alleviate the environmental problems associated with excessive fertilization [5,13]. In intensive cropping sectors such as in horticulture and floriculture, the biostimulants can also increase nutrient use efficiency, partly substitute the chemical fertilizer inputs and ameliorate the yield and quality of crops [14,15,16]. However, biostimulants are not only considered as important substitutes to mineral fertilizers, but also notable in organic farming systems within sustainable crop production management [17]. Increased root and shoot growth, improved resistance against stressors, better root growth potential, and reduction in nitrogen fertilizer inputs are some of the most noteworthy impacts of biostimulant application in sustainable agriculture system [1]. Most of these impacts could be attributed to their auxin-like effect, as well as to the improvement in nitrogen uptake and metabolism, the regulation of K/Na ratio, and the proline accumulation which serves as an osmoprotectant against salinity stress [18,19,20]. Moreover, biostimulatory compounds may also have a positive impact on soil biology and they can be recognized as a good strategy for recovering semiarid areas and degraded ecosystems [21,22,23]. However, the variable composition of raw materials used for the production of biostimulant products make the task of revealing the mechanisms of action more difficult and long-term studies and standardization processes are needed [24]. The major biostimulant impacts on crops are shown in Figure 1.
Considering the numerous literature reports during the last decade related to biostimulants and their effects on various crops, this review aims to present the most up-to-date key results for biostimulant practical applications on crops and the new tools available for the unraveling of mechanisms behind the observed effects. In the present review, all relevant reports in English language were collected. The literature search was performed by using the keywords of plant biostimulants, seaweed extract, leafy vegetables, phenolic compounds, arbuscular mycorrhizal fungi, and biofertilizers in main indexing systems including PubMed/MEDLINE, Scopus, the search engine of Google Scholar, as well as the Institute for Scientific Information Web of Science.

2. Biostimulant Categories

Biostimulants are classified into two distinct classes based on their origin. Therefore, one category includes all those products that have biological origin being obtained from pathogens or from the plant itself and the second category includes all the products that do not have biological origin such as physical factors and chemicals [25]. Moreover, biotic biostimulants may have a defined composition and contain molecules of known structure or being more complex including several molecules with different structures [26]. Another classification approach divides biostimulant products in microbial, which are obtained from arbuscular mycorrhizal fungi and plant growth promoting bacteria, and non-microbial biostimulants which include plant micro-algae extracts, humic substances and biopolymers such as chitosan [27,28,29,30,31]. In particular, the microbial biostimulants may promote plant growth both directly and indirectly; biofertilization, stimulation of root growth, tolerance to plant stressors and rhizoremediation are a few examples of direct effects on plant growth promotion [28,31], while controlling plant pathogens and enhancing the enzymatic activity of plants may indirectly induce plant growth [29,30]. Finally, many researchers divide the non-microbial biostimulants in phytohormonal and non-phytohormonal (those biostimulants that include protein-containing compounds) [32].
Among the various compounds with biostimulatory activity, protein hydrolysates are in the spotlight of scientific research due to their promising properties. Such compounds are actually a mixture of amino acids and soluble peptides, which are mainly produced after enzymatic, thermal and chemical processes and derived from animal or plant origin proteins [33,34]. Their positive effects are associated with the up-regulation of metabolites involved in plant growth processes and the elicitation of hormone-like activities which altogether affect plant growth and productivity [20,35]. The most important benefits of protein hydrolysates are presented in Figure 2.
Similarly, seaweed extracts are also a widely known category of biostimulants with a steadily increasing penetration into the farming sector during the last decades. These compounds have found applications in various crops since they may induce tolerance against abiotic stressors and boost crop performance while they may also improve the shelf-life of various crop products [36]. They are largely prepared from brown seaweeds, such as Ascophyllum nodosum, Ecklonia maxima, and Macrocystis pyrifera and they consist of promoting hormones or trace elements such as Fe, Cu, Zn, and Mn [37,38]. Other compounds such as phloroglucinol and eckol are active biomolecules obtained from the brown seaweed Ecklonia maxima, which is one of the most common species of Kelp which is widely utilized as liquid fertilizer [39]. Moreover, the use of extracts from seaweeds that are industrially processed for other purposes (e.g., the production of carrageenan from Kappaphycus alvarezii) may reduce the carbon footprint of industrial sector and increase at the same time the added value of seaweeds [40].
Trichoderma–based biostimulants are another important category of microbial biostimulants that have found applications in crop production since they may improve plant nutrient status and tolerance against environmental stressors stress via the boost of root growth, the increased nutrient uptake and the production of auxins and secondary metabolites (e.g., peptides, volatile organic compounds) [41,42,43,44,45]. However, several other fungi have shown biostimulatory activity on crops with beneficial effects on plant growth and yield and response to oxidative stress [46].
On the other hand, the number of plant growth promoting bacteria (PGPB) used in various formulation is quite low when considering their great biodiversity [47]. PGPB and symbiotic microorganisms may act through various mechanisms related to hormone release or changes in hormonal balance within plants, the improvement in nutrients availability, the biosynthesis of volatile organic compounds, and the increased tolerance to abiotic stressors through the induction of systemic tolerance [17,48,49,50]. The main negative effects of abiotic stressors (e.g., salinity, drought) on plants are related to changes in endogenous hormones balance (e.g., ethylene production, increase of absicic acid and decrease of cytokinins levels) which results to reduced shoot and root growth as a means to plant homeostasis regulation [51]. This is the key point where PGPB come into play since they promote the production of indole acetic acid which in turn alters root architecture and induce root development resulting to a larger root area and to more root tips. In more detail, these additional (exogenous) phytormones along with the already existing hormones in plant tissues (endogenous) regulate cell proliferation (especially in the roots) which facilitate the uptake of water and minerals from required to support plant growth [52]. Roots and shoots may communicate through hormonal signaling and actually roots may regulate the development and growth of the aerial parts of the plant by transferring endogenous hormones via the xylem to the shoots which act as hormonal sinks [53]. However, apart from the endogenous hormones which plant may produce itself, several other phytormones have been detected in the root-soil environment related to the soil microbiome which may enter the plant through the transpiration flow (e.g., xylem) and regulate plant growth depending on their balance [54]. Soil detected phytormones may be produced from plant roots acting as signals for root functioning, or by soil microbiome (bacteria and fungi) [55]. The overall balance of these ex planta hormones is regulated by biosynthesis and uptake from roots, as well as by production, uptake and degradation of hormones from soil microbes [54], and interacts with in plant hormones, thus regulating plant growth and development [51]. Moreover, arbuscular mycorrhizal fungi (AMF) and rhizosphere microflora combinations seem to be effective not only in improving crop productivity but also in preserving soil health and fertility [56].
Humic-like substances such as humic and fulvic acids may also exhibit biostimulatory activity, since various reports have suggested improved crop performance attributed mostly to auxin- and cytokinin-like effects [57,58]. They are derived from organic matter decomposition and metabolic products of soil microbes and they contribute to plant growth through the improvement of soil physic-chemical properties and the increased availability of nutrients in rhizosphere [7]. The main effects of humic substances are in general the improvement in root growth and morphology, the increase in the uptake of nutrients and their use efficiency, the better crop performance, and finally the increase in fruit quality and in tolerance against abiotic stressors [59,60]. The actual mechanisms of action seem to be the result of synergistic between the various bioactive compounds that raw material include, although the effects may differ depending on the crop, the soil type and soil microbes present in the rhizosphere [61]. In addition, humic and fulvic acids may promote plant growth through hormone-like effects, since the breakdown of these substances releases auxins and other pre-cursors [58,62,63,64]. Moreover, Canellas et al. [65] suggested a hormone-like activity of humic substances fraction on tomato plants through the release of auxin-like biomolecules. Considering that humic-like substances and humic acids in particular can be obtained from various raw materials such as natural organic matter, plant tissues and bio-waste, they present a variable composition with heterogenous effects, depending on their molecular weight [65,66].
Phosphite (Phi) and biopolymers such as chitosan were also reported to possess biostimulant properties with several applications on horticultural crops [67,68]. Regarding Phi, it is widely used as fungicide against various pathogens or as a supplement of P nutrition in crops; however, its application is also associated with plant growth promoting effects which are attributed to promoted root growth and better uptake and assimilation of nutrients from plants [67,69]. On the other hand, chitosan is a biopolymer produced after the deacetylation of chitin and is in the research focus during the last decades due to the interesting effects on crops [68]. It is commercially produced from seafood shells and its main application is related to plant defense against pathogens, since it may induce the production of protective molecules against pathogens [70]. Biostimulant activities have also been reported being mainly associated with increased photosynthetic activity, tolerance to drought, salinity, and extreme temperatures stress and activity of antioxidant enzymes [71,72]. However, considering that chitosan is a biopolymer that comprises compounds of different deacetylation and polymerization degree there is a great variability in the composition of the commercially available products which may also result in variable effects on crops [73].
Apart from these well-established categories, there is significant interest from the biostimulant sector for waste and by-products which exhibit important biological activities and they could be considered as a new category among the existing ones creating alternative pathways in for by-products management [74,75]. In this context, the production of dissolved organic matter (DOM) through anaerobic digestion has shown promising results for the design of new biostimulatory products that may improve plant health through an auxin-like mode of action [66]. According to Messias et al. [76], shale water, which is generated after the pyrolysis of pyrobituminous shale rock, has also shown important biostimulant effects in horticultural crops and could be used as a yield enhancer and biofortifying agent. Finally other compounds such as melatonin and vitamins have shown biostimulatory activities, especially under abiotic stress conditions, and apart from the induction of secondary metabolites biosynthesis they also improve the quality and the functional properties of the final products [77,78,79].
Considering the novel status of the biostimulants sector, as well as the fact that various substances and organisms can be classified as biostimulants by definition, biosafety criteria are important for choosing new microorganisms as biostimulants, and biosafety measures should be addressed according to bioassays rather than on taxonomy and based on environmental and human safety indices (EHSI; [80]) [81]. Moreover, any negative effects associated with unintended effects on reactive nitrogen losses need more attention [82]. The main research topics for the biostimulant characterization of new substances and compounds characterization include, (1) evaluation of the biostimulant composition; (2) standardization of the production methods; (3) characterization of plant responses especially in combination with environmental conditions; (4) identification of crop-specific responses to biostimulants products; and (5) fine-tuning of application timing and doses [83]. The principal classification of plant biostimulants are shown in Table 1.

3. Practical Applications of Biostimulants and Biostimulatory Products on Horticultural Crops

Various biostimulant products have been studied in numerous research reports. Ascophyllum nodosum extracts are among the most commonly studied biostimulants with varied effects on several crops such as the yield and nutritional quality of spinach [16,103,104,105], the nutritional status and shelf-life of lettuce [106], increased the drought tolerance in tomato plants [107], improved plant growth and yield in carrot and strawberry [108,109,110,111], or alleviated the water stress effects on common bean [15,112]. The mechanisms behind these beneficial effects of A. nodosum extracts are still under investigation, although various studies postulated hormonal effects on plant growth through the up- or down-regulation of auxin-responsive genes [113]. This argument is supported by the composition of A. nodosum extracts which contain several hormones (e.g., abscisic acid, auxins, brassinosteroids, cytokinins, ethylene, gibberellins, and strigolactones) [113], although the opposition suggests that the low hormones content along with the low application doses of biostimulants cannot justify these positive effects on plant growth [90,114]. However, the recent study of Dookie et al. [115] came to confirm the hormonal effects of seaweed extracts, since the foliar application of A. nodosum and Sargassum sp. extracts on tomato plants up-regulated the expression of six flowering genes. Other recent studies highlight the protective effects of seaweed extracts against oxidative stress in plants subjected to environmental stress, thus reducing electrolyte leakage and lipid peroxidation [116].
Ecklonia maxima is another brown microalgae the extracts of which have found several application in crop production via various biostimulatory products. The effects of these extracts have shown positive results on crop yield and leaf color of lettuce plants [117], on mung bean germination and plant growth [118], and on growth and nutritional quality of spinach [119]. The application of E. maxima extracts on potato plants also had varied effects including the increase of marketable yield [120] the improved tolerance to abiotic stress and total assimilation area [121] whereas contrasting effects on quality were reported with no effects on dry matter, protein, total sugars and vitamin C content [122] or increasing trends on total and true proteins content being observed [123]. The detailed analysis of the extracts identified new plant growth biostimulants, namely eckol and phloroglucinol [119,124,125], while other plant growth regulators such as abscicic acid, gibberellins and brassinosteroids were also detected in commercial products indicating a hormone-like activity [126]. Apart from these two algae species, several other macro- and microalgae extracts have been incorporated in commercial formulation that are currently used in various horticultural crops [127]. However, despite the scientific evidence regarding the hormonal effects of seaweed extracts, several factors associated with the variability in experimental set-ups, the plethora of seaweed-based products and their species-specific effects, the lack of information regarding the analytical composition of such products and the variable composition of raw material throughout the year make the definition of mechanisms of action difficult [113].
Another category of biostimulants widely used in horticulture is protein hydrolysates and nitrogen-containing compounds. There are several commercial products available derived from plant or animal proteins with various applications in horticultural crops during the last few years [128,129,130]. For example, one of the first studies was conducted on common bean with protein hydrolysates derived from tomato plant residues and reported a significant increase in nitrogen assimilation of bean [131]. Other crop residues were also promising sources of protein hydrolysates and have found practical applications in vegetable crops, e.g., tomato grown in organic [132,133] and conventional farming systems [134], or common bean plants grown under water-stress conditions [135]. Other application of protein hydrolysates refer to plants subjected to stress conditions such as in the study of Koleška et al. [136] who tested the effectiveness of biostimulants in alleviating macronutrient deficiency effects on tomato plants, or Casadesús et al. [137] and Ertani et al. [138] who studied the hormonal effects of plant biostimulants on water-stressed tomato plants. It is suggested that foliar or root applications of protein hydrolysates may improve root development, C and N assimilation and nutrients uptake from plants [33,139,140], regulate the metabolic processes through a multi-level signaling that involves auxin-like activities [141,142,143], as well as to increase the effectiveness of plant defense mechanisms against abiotic stressors in a sustainable manner [144]. However, apart from increased tolerance to stressors the application of protein hydrolysates may enhance quality parameters in fruit and leafy vegetables [35,105,145].
On the other hand, nitrogen containing products have also shown promising results in alleviating stress effects on horticultural crops, as in the case of spinach [16] and common bean [15]. Gelatin is another compound included in animal-derived protein hydrolysates which may improve vegetable crops performance through the up-regulation of nitrogen assimilation by plants [146]. The application of amino acids has also shown positive effects on plant growth, photosynthetic processes and nutritional quality of lettuce [147,148], nutritional quality and physiological parameters of common bean [149,150], plant growth and nutritional status of fennel [151], physiological parameters and chlorophyll content of broccoli [152], and plant growth and fruit quality of tomato grown under iron deficiency [19] or macronutrients deprivation regimes [153].
Humic substances (HS) include humic and fulvic acids which are present in soil organic matter as well in aquatic environments and the atmosphere and differ with each other in their molecular weight [88]. Their application in pepper plants resulted to an increase in plant growth and to accelerated fruit development without significant difference in terms of fruit yield being observed from the untreated plants [154]. In tomato plants, the exogenous application of HS in combination with chelated FeEDDHA increased iron uptake, while it also improved phosphorus content in leaves [155]. Moreover, HS application increased early yield in potato crop when plants were grown under low temperatures and water availability [156], while the incorporation of HS in growing medium increased seed germination and seedling growth in tomato and okra [157,158]. Similar positive effects of HS application were observed on garlic through the stimulation of N and S uptake [159,160], or on onion plants without however significant correlations between the yield and nutrients uptake [161]. The increased yield and quality of potato tubers after the incorporation of HS in soil was attributed to the better availability use efficiency of nutrients due to reduced leaching, as well as to increased water holding capacity of soil [162,163]. Humic substances may also alleviate negative effects of high salinity, as reported by Shalaby and El-Messairy [164] in melon plants. In contrast, Ibrahim and Ramadan [165] reported inconsistent results for the foliar application of zinc combined with humic acid and chitosan on common bean plants, while Hartz and Bottoms [166] suggested no significant effect of HS application on dry matter accumulation or fruit yield in lettuce and tomato, respectively. However, these contradicting reports could be associated with application time and doses, as already suggested by Bettoni et al. [167] for onion crop, or the application methods in the case of mung bean [168]. According to De Hita et al. [60], root application showed more consistent effects than foliar application of HS, since although hormone-like activities in both methods of application the foliar spraying has transient impact and has to be repeated during the growing period.
Biostimulants based on plant growth promoting microorganisms include microbial inocula from bacteria and fungi of various genera and have also found practical applications in horticultural crops, either alone or in combination with each other [169]. In the case of fungi, Trichoderma-based products are the most widely used in horticultural crops [170,171,172,173], although other fungi such as Glomus sp. are also applied in sustainable horticulture [15,27,41,174,175,176]. Regarding the plant growth promoting rhizobacteria (PGPR), the strategy to choose the appropriate ones includes six steps: (i) Determination of the target crop and commercial strategy; (ii) selection of growth media for the isolation of microbial candidates; (iii) screening for traits giving considerable agronomical advantages; (iv) screening for traits belonged to product development; (v) characterization of the mode of action of PGPR; and (vi) evaluation of plant growth efficacy [177]. However, considerable variability is observed in the obtained results and considering that mechanisms of action are not fully understood and the importance of soil physicochemical parameters it is advisable to analyze soil characteristics and then apply those PGPRs that best suit the conditions [48]. The application of PGPR is associated mostly with stress alleviation effects in various crops such as common bean, potato, and lettuce which consequently results to better plant growth and yield [48,178,179,180]. The plant growth promoting effects are related to hormonal regulation in plants, since soil microbes produce various phytohormones which enter the plant from roots via the transpiration flow and reach the shoot sinks where they can induce alterations in shoot and leaf morphology and physiology [181].
Phosphite is a novel biostimulant which may function as a phosphate source affecting plant growth and performance, as well as a biocide against various pathogens and abiotic stress reliever [67,97,182]. It is usually applied with foliar spraying or through the nutrient solution with drip irrigation in the form of potassium phosphite or phosphorus acid, resulting to beneficial effects on plant growth and yield in several vegetable crops [67,69]. The most profound effects were observed in potato crop, where the foliar spraying or potato seed treatment with potassium Phi-improved plant growth and yield and earliness of tuber maturity through the induction of defense mechanisms and the increase of mycorrhizal colonization [94,96,97]. However, there are also studies where negative effects were reported for Phi application on tomato and pepper due to phytotoxicity effects, especially when plants are subjected to P-deficient conditions [182]. According to the same study, the positive or negative effects of Phi are highly associated with the P availability of plants, since Phi per se is not an effective form for P supplementation of plants and positive effects are due to pathogens control [182].
Biopolymers, such as chitosan, have been widely used in horticultural crops cultivation for many years mostly for pathogen control purposes [85]. The modes of application include foliar spraying, direct incorporation in soil or coating of vegetable products [68,71]. In particular, chitosan application was found beneficial for lettuce and tomato plants growth [71,183] and increased phytochemicals defensive metabolites content in spinach leaves [72]. However, there is evidence that bulk chitosan is associated with root growth inhibition when applied in non-optimal concentrations, therefore, alternative forms have been suggested including chitosan micro- and nanoparticles which are safer for agricultural use [184,185].
Silicon (Si) has many biostimulant activities such as enhancing growth and development of horticultural crops, especially under abiotic stress conditions; Si mechanisms of action are involved in oxidative damage, water relations, photosynthesis, ion uptake, hormones, and acts mostly via silica deposition in tissues providing mechanical strength [10]. Although Si effects are mostly visible under stressful conditions, its application may also have beneficial effects on crops grown under optimal conditions since it improves photosynthetic activity and plant growth [186]. The main application form is the foliar spraying, soil incorporation or fertigation of silicic acid and silicates [187], although new forms have also been suggested, such as silicon nanoparticles, for better uptake of Si from plants compared to the bulk form [188]. There are several examples of beneficial effects of Si on vegetable crops, such as tomato [189,190,191], cucumber [192], pepper [193], and squash [194], where Si application alleviated the negative effects of abiotic stressors on plant growth.
A great variety of biostimulant products with different active compounds have been suggested for application in the agricultural sector, including phenolic acids [195], triglycerides [196], titanium [2,197], or zeatin from Moringa oleifera leaves [198]. Moreover, chitosan is another important biopolymer with biostimulant activities which has been used to alleviate water stress negative effects and increase shelf life on horticultural crops such as basil [92], lettuce [71], spinach [72], tomato [183], and pepper [184], among others.
Apart from single product effects, there are several reports where the combination of biostimulants resulted in beneficial effects on horticultural crops which usually are better than those of single biostimulants [145]. For example, microalgae combined with humic acids improved the growth and yield of onion [161], plant growth promoting bacteria acted synergistically with humic acids to improve the growth of tomato [199,200] and potato plants [201], A. nodosum extracts combined with humic acids enhanced plant growth and shelf life of lettuce and spinach [202], or the interaction of plant growth promoting rhizhobacteria with seaweed extracts from E. maxima which increased plant growth and photosynthetic pigments content in Amaranthus hybridus plants [203]. In this context, Rouphael and Colla [204] suggested complex synergistic and additive effects of microbial and non-microbial biostimulants with mechanisms of action that have to be unraveled at a molecular level aiming to design the new generation of biostimulant products.
The main effects of biostimulants on vegetable crops are summarized in Table 2.

4. Future Remarks and Conclusions

Sustainable farming of vegetables is the focal point of research within the last decade considering the ongoing climate change and the increasing incidences of weather extremities as well as the pressure on crops from other abiotic and biotic stressors e.g., water and salinity stress or pathogens infestations. Moreover, food security and the efficient use of natural resources are in a tug-of-war with increasing food demands on the one side and replenishment of natural resources and their efficient use on the other side. Conventional cultivation of vegetables under these conditions becomes more and more difficult and farmers throughout the world start to adapt sustainable cultivation practices seeking always new methods. Biostimulants application is proven a useful tool towards this aim, allowing vegetable producers to cultivate under unfavorable conditions without adverse effects on crop yield. Moreover, the great variety of biostimulant products means there are commercially available products suitable for various conditions and crops. The present review gathered the most up-to-date information regarding the classification of biostimulatory agents and their main mechanisms of action, as well as their practical applications on vegetable crops. Although there are several cases where biostimulant application resulted to beneficial effects on plant growth and yield, more studies are needed to fine-tune application practices, since it seems there are product and crop specificities to be addressed and negative or no effects are also reported. These variable effects reported in the literature are usually due to the variable composition of biostimulants which are natural matrices that include various compounds from different classes and different activities, as well as to uncertainties in application times, methods and doses. Finally, the crop factor is also important since the genotype has a great effect on the response to biostimulant products, especially under stressful conditions.
Considering the above, the production and application of is an evolving process and new biostimulant products are needed. However, this should be realized under a new approach focusing on the synergistic effects of various biostimulatory agents instead of single-product application. Moreover, studying the molecular mechanisms behind the observed activities will help to reveal those physiological and plant metabolism pathways involved in this process and provide farmers with tailor-made products suitable for variable conditions. The application of biostimulants is not just a promising and environmentally-friendly practice, but it may also lead to increased use efficiency of natural resources through water deficit irrigation regimes and the reduced input of agrochemicals (e.g., mineral fertilizers and chemical for pests and pathogen control). It can also increase the sustainability of agricultural and horticultural production systems as well as improve the quality and quantity of food for the ever-growing world’s population.

Author Contributions

M.H.S.: Writing—original draft preparation; C.C.: Writing—original draft preparation; N.P.: Writing—original draft preparation; S.A.P.: Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been co-financed by the European Regional Development Fund of the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH–CREATE–INNOVATE (project code: T2EDK-05281). The APC was funded by the European Regional Development Fund of the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH–CREATE–INNOVATE (project code: T2EDK-05281).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Russo, R.O.; Berlyn, G.P. The Use of Organic Biostimulants to Help Low Input Sustainable Agriculture. J. Sustain. Agric. 1991, 1, 19–42. [Google Scholar] [CrossRef]
  2. Szparaga, A.; Kuboń, M.; Kocira, S.; Czerwińska, E.; Pawłowska, A.; Hara, P.; Kobus, Z.; Kwaśniewski, D. Towards Sustainable Agriculture—Agronomic and Economic Effects of Biostimulant Use in Common Bean Cultivation. Sustainability 2019, 11, 4575. [Google Scholar] [CrossRef] [Green Version]
  3. Postel, S.L. Entering an era of water scarcity: The challenges ahead. Ecol. Appl. 2000, 10, 941–948. [Google Scholar] [CrossRef]
  4. Del Buono, D. Can biostimulants be used to mitigate the effect of anthropogenic climate change on agriculture? It is time to respond. Sci. Total. Environ. 2021, 751, 141763. [Google Scholar] [CrossRef]
  5. Bulgari, R.; Cocetta, G.; Trivellini, A.; Vernieri, P.; Ferrante, A. Biostimulants and crop responses: A review. Biol. Agric. Hortic. 2014, 31, 1–17. [Google Scholar] [CrossRef]
  6. Caradonia, F.; Battaglia, V.; Righi, L.; Pascali, G.; La Torre, A. Plant Biostimulant Regulatory Framework: Prospects in Europe and Current Situation at International Level. J. Plant Growth Regul. 2019, 38, 438–448. [Google Scholar] [CrossRef]
  7. De Pascale, S.; Rouphael, Y.; Colla, G. Plant biostimulants: Innovative tool for enhancing plant nutrition in organic farming. Eur. J. Hortic. Sci. 2018, 82, 277–285. [Google Scholar] [CrossRef]
  8. Askari-Khorasgani, O.; Hatterman-Valenti, H.; Pardo, F.B.F.; Pessarakli, M. Plant and symbiont metabolic regulation and biostimulants application improve symbiotic performance and cold acclimation. J. Plant Nutr. 2019, 42, 2151–2163. [Google Scholar] [CrossRef]
  9. Massa, D.; Lenzi, A.; Montoneri, E.; Ginepro, M.; Prisa, D.; Burchi, G. Plant response to biowaste soluble hydrolysates in hibiscus grown under limiting nutrient availability. J. Plant Nutr. 2017, 41, 396–409. [Google Scholar] [CrossRef]
  10. Rouphael, Y.; Colla, G. Editorial: Biostimulants in Agriculture. Front. Plant Sci. 2020, 11, 40. [Google Scholar] [CrossRef] [Green Version]
  11. Colla, G.; Rouphael, Y. Biostimulants in horticulture. Sci. Hortic. 2015, 196, 1–2. [Google Scholar] [CrossRef]
  12. Du Jardin, P. Plant biostimulants: Definition, concept, main categories and regulation. Sci. Hortic. 2015, 196, 3–14. [Google Scholar] [CrossRef] [Green Version]
  13. Halpern, M.; Bar-Tal, A.; Ofek, M.; Minz, D.; Muller, T.; Yermiyahu, U. The Use of Biostimulants for Enhancing Nutrient Uptake. Adv. Agron. 2015, 130, 141–174. [Google Scholar] [CrossRef]
  14. Petropoulos, S.A.; Taofiq, O.; Fernandes, Â.; Tzortzakis, N.; Ciric, A.; Sokovic, M.; Barros, L.; Ferreira, I.C. Bioactive properties of greenhouse-cultivated green beans (Phaseolus vulgaris L.) under biostimulants and water-stress effect. J. Sci. Food Agric. 2019, 99, 6049–6059. [Google Scholar] [CrossRef] [PubMed]
  15. Petropoulos, S.A.; Fernandes, Â.; Plexida, S.; Chrysargyris, A.; Tzortzakis, N.; Barreira, J.C.M.; Barros, L.; Ferreira, I.C.F.R. Biostimulants Application Alleviates Water Stress Effects on Yield and Chemical Composition of Greenhouse Green Bean (Phaseolus vulgaris L.). Agronomy 2020, 10, 181. [Google Scholar] [CrossRef] [Green Version]
  16. Pereira, C.; Dias, M.I.; Petropoulos, S.A.; Plexida, S.; Chrysargyris, A.; Tzortzakis, N.; Calhelha, R.C.; Ivanov, M.; Stojković, D.; Soković, M.; et al. The Effects of Biostimulants, Biofertilizers and Water-Stress on Nutritional Value and Chemical Composition of Two Spinach Genotypes (Spinacia oleracea L.). Molecules 2019, 24, 4494. [Google Scholar] [CrossRef] [Green Version]
  17. Bertrand, C.; Gonzalez-Coloma, A.; Prigent-Combaret, C. Plant metabolomics to the benefit of crop protection and growth stimulation. Adv. Bot. Res. 2021, 98, 107–132. [Google Scholar] [CrossRef]
  18. Colla, G.; Svecová, E.; Cardarelli, M.; Rouphael, Y.; Reynaud, H.; Canaguier, R.; Planques, B. Effectiveness of A Plant-Derived Protein Hydrolysate to Improve Crop Performances under Different Growing Conditions. Acta Hortic. 2013, 175–179. [Google Scholar] [CrossRef]
  19. Cerdán, M.; Sánchez-Sánchez, A.; Jordá, J.D.; Juárez, M.; Sánchez-Andreu, J. Effect of commercial amino acids on iron nutrition of tomato plants grown under lime-induced iron deficiency. J. Plant Nutr. Soil Sci. 2013, 176, 859–866. [Google Scholar] [CrossRef]
  20. Ertani, A.; Cavani, L.; Pizzeghello, D.; Brandellero, E.; Altissimo, A.; Ciavatta, C.; Nardi, S. Biostimulant activity of two protein hydrolyzates in the growth and nitrogen metabolism of maize seedlings. J. Plant Nutr. Soil Sci. 2009, 172, 237–244. [Google Scholar] [CrossRef]
  21. Tejada, M.; Benítez, C.; Gómez, I.; Parrado, J. Use of biostimulants on soil restoration: Effects on soil biochemical properties and microbial community. Appl. Soil Ecol. 2011, 49, 11–17. [Google Scholar] [CrossRef]
  22. Karapouloutidou, S.; Gasparatos, D. Effects of Biostimulant and Organic Amendment on Soil Properties and Nutrient Status of Lactuca sativa in a Calcareous Saline-Sodic Soil. Agriculture 2019, 9, 164. [Google Scholar] [CrossRef] [Green Version]
  23. Calvo, P.; Nelson, L.; Kloepper, J.W. Agricultural uses of plant biostimulants. Plant Soil 2014, 383, 3–41. [Google Scholar] [CrossRef] [Green Version]
  24. Yakhin, O.I.; Lubyanov, A.A.; Yakhin, I.A.; Brown, P.H. Biostimulants in Plant Science: A Global Perspective. Front. Plant Sci. 2017, 7, 2049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Rafiee, H.; Badi, H.N.; Mehrafarin, A.; Qaderi, A.; Zarinpanjeh, N.; Sekara, A.; Zand, E. Application of Plant Biostimulants as New Approach to Improve the Biological Responses of Medicinal Plants- A Critical Review. J. Med. Plants 2016, 15, 1–39. [Google Scholar]
  26. Vasconsuelo, A.; Boland, R. Molecular aspects of the early stages of elicitation of secondary metabolites in plants. Plant Sci. 2007, 172, 861–875. [Google Scholar] [CrossRef]
  27. Rouphael, Y.; Colla, G. Toward a Sustainable Agriculture Through Plant Biostimulants: From Experimental Data to Practical Applications. Agronomy 2020, 10, 1461. [Google Scholar] [CrossRef]
  28. Lugtenberg, B.; Kamilova, F. Plant-Growth-Promoting Rhizobacteria. Annu. Rev. Microbiol. 2009, 63, 541–556. [Google Scholar] [CrossRef] [Green Version]
  29. Pérez-Montaño, F.; Alías-Villegas, C.; Bellogín, R.; del Cerro, P.; Espuny, M.; Jiménez-Guerrero, I.; López-Baena, F.; Ollero, F.; Cubo, T. Plant growth promotion in cereal and leguminous agricultural important plants: From microorganism capacities to crop production. Microbiol. Res. 2014, 169, 325–336. [Google Scholar] [CrossRef] [Green Version]
  30. 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] [Green Version]
  31. De Vries, F.T.; Griffiths, R.I.; Knight, C.G.; Nicolitch, O.; Williams, A. Harnessing rhizosphere microbiomes for drought-resilient crop production. Science 2020, 368, 270–274. [Google Scholar] [CrossRef] [PubMed]
  32. Geelen, D.; Xu, L. The Chemical Biology of Plant Biostimulants; Wiley: Ghent, Belgium, 2020. [Google Scholar]
  33. Carillo, P.; Colla, G.; Fusco, G.M.; Dell’Aversana, E.; El-Nakhel, C.; Giordano, M.; Pannico, A.; Cozzolino, E.; Mori, M.; Reynaud, H.; et al. Morphological and Physiological Responses Induced by Protein Hydrolysate-Based Biostimulant and Nitrogen Rates in Greenhouse Spinach. Agronomy 2019, 9, 450. [Google Scholar] [CrossRef] [Green Version]
  34. Colla, G.; Nardi, S.; Cardarelli, M.; Ertani, A.; Lucini, L.; Canaguier, R.; Rouphael, Y. Protein hydrolysates as biostimulants in horticulture. Sci. Hortic. 2015, 196, 28–38. [Google Scholar] [CrossRef]
  35. Colla, G.; Hoagland, L.; Ruzzi, M.; Cardarelli, M.; Bonini, P.; Canaguier, R.; Rouphael, Y. Biostimulant Action of Protein Hydrolysates: Unraveling Their Effects on Plant Physiology and Microbiome. Front. Plant Sci. 2017, 8, 2202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Battacharyya, D.; Babgohari, M.Z.; Rathor, P.; Prithiviraj, B. Seaweed extracts as biostimulants in horticulture. Sci. Hortic. 2015, 196, 39–48. [Google Scholar] [CrossRef]
  37. Gupta, V.; Kumar, M.; Brahmbhatt, H.; Reddy, C.; Seth, A.; Jha, B. Simultaneous determination of different endogenetic plant growth regulators in common green seaweeds using dispersive liquid–liquid microextraction method. Plant Physiol. Biochem. 2011, 49, 1259–1263. [Google Scholar] [CrossRef] [PubMed]
  38. Sivasankari, S.; Venkatesalu, V.; Anantharaj, M.; Chandrasekaran, M. Effect of seaweed extracts on the growth and biochemical constituents of Vigna sinensis. Bioresour. Technol. 2006, 97, 1745–1751. [Google Scholar] [CrossRef]
  39. Rengasamy, K.R.; Kulkarni, M.G.; Pendota, S.C.; Van Staden, J. Enhancing growth, phytochemical constituents and aphid resistance capacity in cabbage with foliar application of eckol—A biologically active phenolic molecule from brown seaweed. New Biotechnol. 2016, 33, 273–279. [Google Scholar] [CrossRef]
  40. Ghosh, A.; Anand, K.V.; Seth, A. Life cycle impact assessment of seaweed based biostimulant production from onshore cultivated Kappaphycus alvarezii (Doty) Doty ex Silva—Is it environmentally sustainable? Algal Res. 2015, 12, 513–521. [Google Scholar] [CrossRef]
  41. Colla, G.; Rouphael, Y.; Di Mattia, E.; El-Nakhel, C.; Cardarelli, M. Co-inoculation of Glomus intraradices and Trichoderma atroviride acts as a biostimulant to promote growth, yield and nutrient uptake of vegetable crops. J. Sci. Food Agric. 2015, 95, 1706–1715. [Google Scholar] [CrossRef]
  42. Fiorentino, N.; Ventorino, V.; Woo, S.L.; Pepe, O.; De Rosa, A.; Gioia, L.; Romano, I.; Lombardi, N.; Napolitano, M.; Colla, G.; et al. Trichoderma-Based Biostimulants Modulate Rhizosphere Microbial Populations and Improve N Uptake Efficiency, Yield, and Nutritional Quality of Leafy Vegetables. Front. Plant Sci. 2018, 9, 743. [Google Scholar] [CrossRef] [Green Version]
  43. Ahmad, P.; Hashem, A.; Abd-Allah, E.F.; Alqarawi, A.A.; John, R.; Egamberdieva, D.; Gucel, S. Role of Trichoderma harzianum in mitigating NaCl stress in Indian mustard (Brassica juncea L) through antioxidative defense system. Front. Plant Sci. 2015, 6, 868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. López-Bucio, J.; Pelagio-Flores, R.; Herrera-Estrella, A. Trichoderma as biostimulant: Exploiting the multilevel properties of a plant beneficial fungus. Sci. Hortic. 2015, 196, 109–123. [Google Scholar] [CrossRef]
  45. Vinale, F.; Nigro, M.; Sivasithamparam, K.; Flematti, G.; Ghisalberti, E.L.; Ruocco, M.; Varlese, R.; Marra, R.; Lanzuise, S.; Eid, A.; et al. Harzianic acid: A novel siderophore from Trichoderma harzianum. FEMS Microbiol. Lett. 2013, 347, 123–129. [Google Scholar] [CrossRef] [Green Version]
  46. Drobek, M.; Frąc, M.; Cybulska, J. Plant Biostimulants: Importance of the Quality and Yield of Horticultural Crops and the Improvement of Plant Tolerance to Abiotic Stress—A Review. Agronomy 2019, 9, 335. [Google Scholar] [CrossRef] [Green Version]
  47. Olivares, F.L.; Busato, J.G.; De Paula, A.M.; Lima, L.D.S.; Aguiar, N.O.; Canellas, L.P. Plant growth promoting bacteria and humic substances: Crop promotion and mechanisms of action. Chem. Biol. Technol. Agric. 2017, 4, 30. [Google Scholar] [CrossRef] [Green Version]
  48. Ruzzi, M.; Aroca, R. Plant growth-promoting rhizobacteria act as biostimulants in horticulture. Sci. Hortic. 2015, 196, 124–134. [Google Scholar] [CrossRef]
  49. 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] [Green Version]
  50. Glick, R.B. Bacteria with ACC deaminase can promote plant growth and help to feed the world. Microbiol. Res. 2014, 169, 30–39. [Google Scholar] [CrossRef] [PubMed]
  51. Yang, J.; Kloepper, J.W.; Ryu, C.-M. Rhizosphere bacteria help plants tolerate abiotic stress. Trends Plant Sci. 2009, 14, 1–4. [Google Scholar] [CrossRef]
  52. Wong, W.S.; Zhong, H.T.; Cross, A.T.; Wan, J.; Yong, H. Plant Biostimulants in Vermicomposts: Characteristics and Plausible Mechanisms. In The Chemical Biology of Plant Biostimulants; Geelen, D., Xu, L., Eds.; Wiley: Hoboken, NJ, USA, 2020. [Google Scholar]
  53. Jackson, M.B. Are Plant Hormones Involved in Root to Shoot Communication? In Advances in Botanical Research; Callow, J.A., Ed.; Academic Press: Cambridge, MA, USA, 1993; Volume 19, pp. 103–187. ISBN 0065-2296. [Google Scholar]
  54. Lu, Y.; Wang, E.; Tang, Z.; Rui, J.; Li, Y.; Tang, Z.; Dong, W.; Liu, X.; George, T.S.; Song, A.; et al. Roots and microbiome jointly drive the distributions of 17 phytohormones in the plant soil continuum in a phytohormone-specific manner. Plant Soil 2021, 1–13. [Google Scholar] [CrossRef]
  55. Rosier, A.; Medeiros, F.H.V.; Bais, H.P. Defining plant growth promoting rhizobacteria molecular and biochemical networks in beneficial plant-microbe interactions. Plant Soil 2018, 428, 35–55. [Google Scholar] [CrossRef] [Green Version]
  56. Giovannini, L.; Palla, M.; Agnolucci, M.; Avio, L.; Sbrana, C.; Turrini, A.; Giovannetti, M. Arbuscular Mycorrhizal Fungi and Associated Microbiota as Plant Biostimulants: Research Strategies for the Selection of the Best Performing Inocula. Agronomy 2020, 10, 106. [Google Scholar] [CrossRef] [Green Version]
  57. Puglisi, I.; Barone, V.; Sidella, S.; Coppa, M.; Broccanello, C.; Gennari, M.; Baglieri, A. Biostimulant activity of humic-like substances from agro-industrial waste on Chlorella vulgaris and Scenedesmus quadricauda. Eur. J. Phycol. 2018, 53, 433–442. [Google Scholar] [CrossRef]
  58. Pizzeghello, D.; Francioso, O.; Ertani, A.; Muscolo, A.; Nardi, S. Isopentenyladenosine and cytokinin-like activity of different humic substances. J. Geochem. Explor. 2013, 129, 70–75. [Google Scholar] [CrossRef]
  59. Vujinović, T.; Zanin, L.; Venuti, S.; Contin, M.; Ceccon, P.; Tomasi, N.; Pinton, R.; Cesco, S.; De Nobili, M. Biostimulant Action of Dissolved Humic Substances from a Conventionally and an Organically Managed Soil on Nitrate Acquisition in Maize Plants. Front. Plant Sci. 2020, 10, 1–14. [Google Scholar] [CrossRef]
  60. De Hita Mejía, D.; Fuentes, M.; García, A.; Olaetxea, M.; Baigorri, R.; Zamarreño, A.M.; Berbara, R.; Garcia-Mina, J.M. Humic substances: A valuable agronomic tool for improving crop adaptation to saline water irrigation. Water Sci. Technol. Water Supply 2019, 19, 1735–1740. [Google Scholar] [CrossRef]
  61. Olaetxea, M.; De Hita, D.; Garcia, C.A.; Fuentes, M.; Baigorri, R.; Mora, V.; Garnica, M.; Urrutia, O.; Erro, J.; Zamarreño, A.M.; et al. Hypothetical framework integrating the main mechanisms involved in the promoting action of rhizospheric humic substances on plant root- and shoot- growth. Appl. Soil Ecol. 2018, 123, 521–537. [Google Scholar] [CrossRef]
  62. Jindo, K.; Martim, S.A.; Navarro, E.C.; Pérez-Alfocea, F.; Hernandez, T.; Garcia, C.; Aguiar, N.O.; Canellas, L.P. Root growth promotion by humic acids from composted and non-composted urban organic wastes. Plant Soil 2012, 353, 209–220. [Google Scholar] [CrossRef]
  63. Garcia, A.C.; Olaetxea, M.; Santos, L.A.; Mora, V.; Baigorri, R.; Fuentes, M.; Zamarreño, A.M.; Berbara, R.L.; Garcia-Mina, J.M. Involvement of hormone- and ROS- signaling pathways in the beneficial action of humic substances on plants growing under normal and stressing conditions. BioMed. Res. Int. 2016, 2016, 3747501. [Google Scholar] [CrossRef]
  64. Conselvan, G.B.; Fuentes, D.; Merchant, A.; Peggion, C.; Francioso, O.; Carletti, P. Effects of humic substances and indole-3-acetic acid on Arabidopsis sugar and amino acid metabolic profile. Plant Soil 2018, 426, 17–32. [Google Scholar] [CrossRef]
  65. Canellas, L.P.; Olivares, F.L.; Aguiar, N.O.; Jones, D.L.; Nebbioso, A.; Mazzei, P.; Piccolo, A. Humic and fulvic acids as biostimulants in horticulture. Sci. Hortic. 2015, 196, 15–27. [Google Scholar] [CrossRef]
  66. Scaglia, B.; Pognani, M.; Adani, F. The anaerobic digestion process capability to produce biostimulant: The case study of the dissolved organic matter (DOM) vs. auxin-like property. Sci. Total. Environ. 2017, 589, 36–45. [Google Scholar] [CrossRef]
  67. Gómez-Merino, F.C.; Trejo-Téllez, L.I. Biostimulant activity of phosphite in horticulture. Sci. Hortic. 2015, 196, 82–90. [Google Scholar] [CrossRef] [Green Version]
  68. Pichyangkura, R.; Chadchawan, S. Biostimulant activity of chitosan in horticulture. Sci. Hortic. 2015, 196, 49–65. [Google Scholar] [CrossRef]
  69. Rickard, D.A. Review of phosphorus acid and its salts as fertilizer materials. J. Plant Nutr. 2000, 23, 161–180. [Google Scholar] [CrossRef]
  70. Goñi, O.; Quille, P.; O’Connell, S. Production of chitosan oligosaccharides for inclusion in a plant biostimulant. Pure Appl. Chem. 2016, 88, 881–889. [Google Scholar] [CrossRef]
  71. Xu, C.; Mou, B. Chitosan as Soil Amendment Affects Lettuce Growth, Photochemical Efficiency, and Gas Exchange. HortTechnology 2018, 28, 476–480. [Google Scholar] [CrossRef] [Green Version]
  72. Singh, S. Enhancing phytochemical levels, enzymatic and antioxidant activity of spinach leaves by chitosan treatment and an insight into the metabolic pathway using DART-MS technique. Food Chem. 2016, 199, 176–184. [Google Scholar] [CrossRef]
  73. Malerba, M.; Cerana, R. Recent Advances of Chitosan Applications in Plants. Polymers 2018, 10, 118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ugolini, L.; Cinti, S.; Righetti, L.; Stefan, A.; Matteo, R.; D’Avino, L.; Lazzeri, L. Production of an enzymatic protein hydrolyzate from defatted sunflower seed meal for potential application as a plant biostimulant. Ind. Crop. Prod. 2015, 75, 15–23. [Google Scholar] [CrossRef]
  75. Pane, C.; Palese, A.M.; Spaccini, R.; Piccolo, A.; Celano, G.; Zaccardelli, M. Enhancing sustainability of a processing tomato cultivation system by using bioactive compost teas. Sci. Hortic. 2016, 202, 117–124. [Google Scholar] [CrossRef]
  76. Messias, R.D.S.; Silveira, C.A.P.; Galli, V.; Pillon, C.N.; Rombaldi, C.V. Multimineral and Organic Composition of a Liquid By-Product from the Pyrobituminous Shale Pyrolysis Process and its Potential Use in Agriculture. J. Plant Nutr. 2015, 38, 959–972. [Google Scholar] [CrossRef]
  77. Debnath, B.; Sikdar, A.; Islam, S.; Hasan, K.; Li, M.; Qiu, D. Physiological and Molecular Responses to Acid Rain Stress in Plants and the Impact of Melatonin, Glutathione and Silicon in the Amendment of Plant Acid Rain Stress. Molecules 2021, 26, 862. [Google Scholar] [CrossRef]
  78. Hanson, A.D.; Beaudoin, G.A.; Mccarty, D.R.; Gregory, J.F. Does Abiotic Stress Cause Functional B Vitamin Deficiency in Plants? Plant Physiol. 2016, 172, 2082–2097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Teklić, T.; Parađiković, N.; Špoljarević, M.; Zeljković, S.; Lončarić, Z.; Lisjak, M. Linking abiotic stress, plant metabolites, biostimulants and functional food. Ann. Appl. Biol. 2021, 178, 169–191. [Google Scholar] [CrossRef]
  80. Vílchez, J.I.; Navas, A.; González-López, J.; Arcos, S.C.; Manzanera, M. Biosafety test for plant growth-promoting bacteria: Proposed Environmental and Human Safety Index (EHSI) Protocol. Front. Microbiol. 2016, 6, 1–14. [Google Scholar] [CrossRef]
  81. Barros-Rodríguez, A.; Rangseekaew, P.; Lasudee, K.; Pathom-Aree, W.; Manzanera, M. Regulatory risks associated with bacteria as biostimulants and biofertilizers in the frame of the European Regulation (EU) 2019/1009. Sci. Total Environ. 2020, 740, 140239. [Google Scholar] [CrossRef] [PubMed]
  82. Souza, E.F.; Rosen, C.J.; Venterea, R.T. Contrasting effects of inhibitors and biostimulants on agronomic performance and reactive nitrogen losses during irrigated potato production. Field Crop. Res. 2019, 240, 143–153. [Google Scholar] [CrossRef]
  83. Toscano, S.; Romano, D.; Massa, D.; Bulgari, R.; Franzoni, G.; Ferrante, A. Biostimulant applications in low input horticultural cultivation systems. Italus Hortus 2018, 27–36. [Google Scholar] [CrossRef]
  84. Ahmad, R.; Lim, C.J.; Kwon, S.-Y. Glycine betaine: A versatile compound with great potential for gene pyramiding to improve crop plant performance against environmental stresses. Plant Biotechnol. Rep. 2013, 7, 49–57. [Google Scholar] [CrossRef]
  85. Xu, L.; Geelen, D. Developing Biostimulants from Agro-Food and Industrial By-Products. Front. Plant Sci. 2018, 9, 1567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Colla, G.; Rouphael, Y.; Canaguier, R.; Svecova, E.; Cardarelli, M. Biostimulant action of a plant-derived protein hydrolysate produced through enzymatic hydrolysis. Front. Plant Sci. 2014, 5, 448. [Google Scholar] [CrossRef] [Green Version]
  87. Vioque, J.; Sánchez-Vioque, R.; Clemente, A.; Pedroche, J.; Millán, F. Partially hydrolyzed rapeseed protein isolates with improved functional properties. J. Am. Oil Chem. Soc. 2000, 77, 447–450. [Google Scholar] [CrossRef]
  88. Nardi, S.; Pizzeghello, D.; Schiavon, M.; Ertani, A. Plant biostimulants: Physiological responses induced by protein hydrolyzed-based products and humic substances in plant metabolism. Sci. Agricola 2016, 73, 18–23. [Google Scholar] [CrossRef] [Green Version]
  89. Khan, W.; Rayirath, U.P.; Subramanian, S.; Jithesh, M.N.; Rayorath, P.; Hodges, D.M.; Critchley, A.T.; Craigie, J.S.; Norrie, J.; Prithiviraj, B. Seaweed Extracts as Biostimulants of Plant Growth and Development. J. Plant Growth Regul. 2009, 28, 386–399. [Google Scholar] [CrossRef]
  90. Craigie, J.S. Seaweed extract stimuli in plant science and agriculture. Environ. Boil. Fishes 2011, 23, 371–393. [Google Scholar] [CrossRef]
  91. González, A.; Castro, J.; Vera, J.; Moenne, A. Seaweed Oligosaccharides Stimulate Plant Growth by Enhancing Carbon and Nitrogen Assimilation, Basal Metabolism, and Cell Division. J. Plant Growth Regul. 2012, 32, 443–448. [Google Scholar] [CrossRef] [Green Version]
  92. Ghasemi Pirbalouti, A.; Malekpoor, F.; Salimi, A.; Golparvar, A. Exogenous application of chitosan on biochemical and phys-iological characteristics, phenolic content and antioxidant activity of two species of basil (Ocimum ciliatum and Ocimum basilicum) under reduced irrigation. Sci. Hortic. 2017, 217, 114–122. [Google Scholar] [CrossRef]
  93. Varadarajan, D.K.; Karthikeyan, A.S.; Matilda, P.D.; Raghothama, K.G. Phosphite, an Analog of Phosphate, Suppresses the Coordinated Expression of Genes under Phosphate Starvation. Plant Physiol. 2002, 129, 1232–1240. [Google Scholar] [CrossRef] [Green Version]
  94. Tambascio, C.; Covacevich, F.; Lobato, M.C.; de Lasa, C.; Caldiz, D.; Dosio, G.; Andreu, A. The Application of K Phosphites to Seed Tubers Enhanced Emergence, Early Growth and Mycorrhizal Colonization in Potato (Solanum tuberosum). Am. J. Plant Sci. 2014, 5, 132–137. [Google Scholar] [CrossRef]
  95. Olivieri, F.; Feldman, M.; Machinandiarena, M.; Lobato, M.; Caldiz, D.; Daleo, G.; Andreu, A. Phosphite applications induce molecular modifications in potato tuber periderm and cortex that enhance resistance to pathogens. Crop. Prot. 2012, 32, 1–6. [Google Scholar] [CrossRef]
  96. Lobato, M.C.; Machinandiarena, M.F.; Tambascio, C.; Dosio, G.A.A.; Caldiz, D.O.; Daleo, G.R.; Andreu, A.B.; Olivieri, F.P. Effect of foliar applications of phosphite on post-harvest potato tubers. Eur. J. Plant Pathol. 2011, 130, 155–163. [Google Scholar] [CrossRef]
  97. Oyarburo, N.S.; Machinandiarena, M.F.; Feldman, M.L.; Daleo, G.R.; Andreu, A.B.; Olivieri, F.P. Potassium phosphite increases tolerance to UV-B in potato. Plant Physiol. Biochem. 2015, 88, 1–8. [Google Scholar] [CrossRef]
  98. Moor, U.; Põldma, P.; Tõnutare, T.; Karp, K.; Starast, M.; Vool, E. Effect of phosphite fertilization on growth, yield and fruit composition of strawberries. Sci. Hortic. 2009, 119, 264–269. [Google Scholar] [CrossRef]
  99. Estrada-Ortiz, E.; Trejo-Téllez, L.I.; Gómez-Merino, F.C.; Núñez-Escobar, R.; Sandoval-Villa, M. The effects of phosphite on strawberry yield and fruit quality. J. Soil Sci. Plant Nutr. 2013, 13, 612–620. [Google Scholar] [CrossRef] [Green Version]
  100. Lovatt, C.J. Properly Timing Foliar-applied Fertilizers Increases Efficacy: A Review and Update on Timing Foliar Nutrient Applications to Citrus and Avocado. HortTechnology 2013, 23, 536–541. [Google Scholar] [CrossRef] [Green Version]
  101. Aremu, A.O.; Stirk, W.A.; Kulkarni, M.G.; Tarkowská, D.; Turečková, V.; Gruz, J.; Šubrtová, M.; Pěnčík, A.; Novák, O.; Dolezal, K.; et al. Evidence of phytohormones and phenolic acids variability in garden-waste-derived vermicompost leachate, a well-known plant growth stimulant. Plant Growth Regul. 2014, 75, 483–492. [Google Scholar] [CrossRef]
  102. Zhang, H.; Tan, S.N.; Teo, C.H.; Yew, Y.R.; Ge, L.; Chen, X.; Yong, J.W.H. Analysis of phytohormones in vermicompost using a novel combinative sample preparation strategy of ultrasound-assisted extraction and solid-phase extraction coupled with liquid chromatography–tandem mass spectrometry. Talanta 2015, 139, 189–197. [Google Scholar] [CrossRef] [PubMed]
  103. Fan, D.; Hodges, D.M.; Critchley, A.T.; Prithiviraj, B. A Commercial Extract of Brown Macroalga (Ascophyllum nodosum) Affects Yield and the Nutritional Quality of Spinach In Vitro. Commun. Soil Sci. Plant Anal. 2013, 44, 1873–1884. [Google Scholar] [CrossRef]
  104. Fan, D.; Hodges, D.M.; Zhang, J.; Kirby, C.W.; Ji, X.; Locke, S.J.; Critchley, A.T.; Prithiviraj, B. Commercial extract of the brown seaweed Ascophyllum nodosum enhances phenolic antioxidant content of spinach (Spinacia oleracea L.) which protects Caenorhabditis elegans against oxidative and thermal stress. Food Chem. 2011, 124, 195–202. [Google Scholar] [CrossRef] [Green Version]
  105. Rouphael, Y.; Giordano, M.; Cardarelli, M.; Cozzolino, E.; Mori, M.; Kyriacou, M.C.; Bonini, P.; Colla, G. Plant- and Seaweed-Based Extracts Increase Yield but Differentially Modulate Nutritional Quality of Greenhouse Spinach through Biostimulant Action. Agronomy 2018, 8, 126. [Google Scholar] [CrossRef] [Green Version]
  106. Chrysargyris, A.; Xylia, P.; Anastasiou, M.; Pantelides, I.; Tzortzakis, N. Effects of Ascophyllum nodosum seaweed extracts on lettuce growth, physiology and fresh-cut salad storage under potassium deficiency. J. Sci. Food Agric. 2018, 98, 5861–5872. [Google Scholar] [CrossRef] [PubMed]
  107. Goñi, O.; Quille, P.; O’Connell, S. Ascophyllum nodosum extract biostimulants and their role in enhancing tolerance to drought stress in tomato plants. Plant Physiol. Biochem. 2018, 126, 63–73. [Google Scholar] [CrossRef]
  108. Alam, M.Z.; Braun, G.; Norrie, J.; Hodges, D.M. Ascophyllum extract application can promote plant growth and root yield in carrot associated with increased root-zone soil microbial activity. Can. J. Plant Sci. 2014, 94, 337–348. [Google Scholar] [CrossRef] [Green Version]
  109. Alam, M.Z.; Braun, G.; Norrie, J.; Hodges, D.M. Effect of Ascophyllum extract application on plant growth, fruit yield and soil microbial communities of strawberry. Can. J. Plant Sci. 2013, 93, 23–36. [Google Scholar] [CrossRef]
  110. Weber, N.; Schmitzer, V.; Jakopic, J.; Stampar, F. First fruit in season: Seaweed extract and silicon advance organic strawberry (Fragaria×ananassa Duch.) fruit formation and yield. Sci. Hortic. 2018, 242, 103–109. [Google Scholar] [CrossRef]
  111. Spinelli, F.; Fiori, G.; Noferini, M.; Sprocatti, M.; Costa, G. A novel type of seaweed extract as a natural alternative to the use of iron chelates in strawberry production. Sci. Hortic. 2010, 125, 263–269. [Google Scholar] [CrossRef]
  112. Galvão, Í.M.; dos Santos, O.F.; de Souza, M.L.C.; de Jesus Guimarães, J.; Kühn, I.E.; Broetto, F. Biostimulants action in common bean crop submitted to water deficit. In Toward a Sustainable Agriculture Through Plant Biostimulants: From Data to Practical Applications; Rouphael, Y., Colla, G., Eds.; MDPI: Basel, Switzerland, 2019; Volume 225, p. 105762. [Google Scholar]
  113. De Saeger, J.; Van Praet, S.; Vereecke, D.; Park, J.; Jacques, S.; Han, T.; Depuydt, S. Toward the molecular understanding of the action mechanism of Ascophyllum nodosum extracts on plants. Environ. Boil. Fishes 2019, 32, 573–597. [Google Scholar] [CrossRef] [Green Version]
  114. Rayorath, P.; Jithesh, M.N.; Farid, A.; Khan, W.; Palanisamy, R.; Hankins, S.D.; Critchley, A.T.; Prithiviraj, B. Rapid bioassays to evaluate the plant growth promoting activity of Ascophyllum nodosum (L.) Le Jol. using a model plant, Arabidopsis thaliana (L.) Heynh. Environ. Boil. Fishes 2007, 20, 423–429. [Google Scholar] [CrossRef]
  115. Dookie, M.; Ali, O.; Ramsubhag, A.; Jayaraman, J. Flowering gene regulation in tomato plants treated with brown seaweed extracts. Sci. Hortic. 2021, 276, 109715. [Google Scholar] [CrossRef]
  116. Shukla, P.S.; Prithiviraj, B. Ascophyllum nodosum Biostimulant Improves the Growth of Zea mays Grown Under Phosphorus Impoverished Conditions. Front. Plant Sci. 2021, 11, 1–17. [Google Scholar] [CrossRef] [PubMed]
  117. Di Mola, I.; Cozzolino, E.; Ottaiano, L.; Giordano, M.; Rouphael, Y.; El Nakhel, C.; Leone, V.; Mori, M. Effect of seaweed (Ecklonia maxima) extract and legume-derived protein hydrolysate biostimulants on baby leaf lettuce grown on optimal doses of nitrogen under greenhouse conditions. Aust. J. Crop. Sci. 2020, 14, 1456–1464. [Google Scholar] [CrossRef]
  118. Herrera, D.D.F.; Muñoz-Ochoa, M.; Hernández-Herrera, R.M.; Hernández-Carmona, G. Biostimulant activity of individual and blended seaweed extracts on the germination and growth of the mung bean. Environ. Boil. Fishes 2018, 31, 2025–2037. [Google Scholar] [CrossRef]
  119. Kulkarni, M.G.; Rengasamy, K.R.; Pendota, S.C.; Gruz, J.; Plačková, L.; Novák, O.; Doležal, K.; Van Staden, J. Bioactive molecules derived from smoke and seaweed Ecklonia maxima showing phytohormone-like activity in Spinacia oleracea L. New Biotechnol. 2019, 48, 83–89. [Google Scholar] [CrossRef]
  120. Zarzecka, K.; Gugała, M.; Sikorska, A.; Grzywacz, K.; Niewęgłowski, M. Marketable yield of potato and its quantitative parameters after application of herbicides and biostimulants. Agriculture 2020, 10, 49. [Google Scholar] [CrossRef] [Green Version]
  121. Wadas, W.; Dziugieł, T. Changes in assimilation area and chlorophyll content of very early potato (Solanum tuberosum L.) cultivars as influenced by biostimulants. Agronomy 2020, 10, 387. [Google Scholar] [CrossRef] [Green Version]
  122. Wadas, W.; Dziugieł, T. Quality of new potatoes (Solanum tuberosum L.) in response to plant biostimulants application. Agriculture 2020, 10, 265. [Google Scholar] [CrossRef]
  123. Zarzecka, K.; Gugała, M.; Mystkowska, I.; Sikorska, A. Total and True Protein Content in Potato Tubers Depending on Herbicides and Biostimulants. Agronomy 2020, 10, 1106. [Google Scholar] [CrossRef]
  124. Rengasamy, K.R.; Kulkarni, M.G.; Papenfus, H.B.; Van Staden, J. Quantification of plant growth biostimulants, phloroglucinol and eckol, in four commercial seaweed liquid fertilizers and some by-products. Algal Res. 2016, 20, 57–60. [Google Scholar] [CrossRef]
  125. Rengasamy, K.R.R.; Kulkarni, M.G.; Stirk, W.A.; Van Staden, J. Eckol—A new plant growth stimulant from the brown seaweed Ecklonia maxima. Environ. Boil. Fishes 2015, 27, 581–587. [Google Scholar] [CrossRef]
  126. Stirk, W.A.; Tarkowská, D.; Turečová, V.; Strnad, M.; Van Staden, J. Abscisic acid, gibberellins and brassinosteroids in Kelpak®, a commercial seaweed extract made from Ecklonia maxima. Environ. Boil. Fishes 2013, 26, 561–567. [Google Scholar] [CrossRef]
  127. Ronga, D.; Biazzi, E.; Parati, K.; Carminati, D.; Carminati, E.; Tava, A. Microalgal Biostimulants and Biofertilisers in Crop Productions. Agronomy 2019, 9, 192. [Google Scholar] [CrossRef] [Green Version]
  128. Consentino, B.B.; Virga, G.; La Placa, G.G.; Sabatino, L.; Rouphael, Y.; Ntatsi, G.; Iapichino, G.; La Bella, S.; Mauro, R.P.; D’Anna, F.; et al. Celery (Apium graveolens L.) Performances as Subjected to Different Sources of Protein Hydrolysates. Plants 2020, 9, 1633. [Google Scholar] [CrossRef]
  129. García-Santiago, J.C.; Cavazos, C.J.L.; González-Fuentes, J.A.; Zermeño-González, A.; Alvarado, E.R.; Duarte, A.R.; Preciado-Rangel, P.; Troyo-Diéguez, E.; Ramos, F.M.P.; Valdez-Aguilar, L.A.; et al. Effects of fish-derived protein hydrolysate, animal-based organic fertilisers and irrigation method on the growth and quality of grape tomatoes. Biol. Agric. Hortic. 2021, 37, 107–124. [Google Scholar] [CrossRef]
  130. Madende, M.; Hayes, M. Fish By-Product Use as Biostimulants: An Overview of the Current State of the Art, Including Relevant Legislation and Regulations within the EU and USA. Molecules 2020, 25, 1122. [Google Scholar] [CrossRef] [Green Version]
  131. Baglieri, A.; Cadili, V.; Monterumici, C.M.; Gennari, M.; Tabasso, S.; Montoneri, E.; Nardi, S.; Negre, M. Fertilization of bean plants with tomato plants hydrolysates. Effect on biomass production, chlorophyll content and N assimilation. Sci. Hortic. 2014, 176, 194–199. [Google Scholar] [CrossRef]
  132. Chehade, L.A.; Al Chami, Z.; De Pascali, S.A.; Cavoski, I.; Fanizzi, F.P. Biostimulants from food processing by-products: Agronomic, quality and metabolic impacts on organic tomato (Solanum lycopersicum L.). J. Sci. Food Agric. 2018, 98, 1426–1436. [Google Scholar] [CrossRef]
  133. Caruso, G.; De Pascale, S.; Cozzolino, E.; Cuciniello, A.; Cenvinzo, V.; Bonini, P.; Colla, G.; Rouphael, Y. Yield and Nutritional Quality of Vesuvian Piennolo Tomato PDO as Affected by Farming System and Biostimulant Application. Agronomy 2019, 9, 505. [Google Scholar] [CrossRef] [Green Version]
  134. Rouphael, Y.; Colla, G.; Giordano, M.; El-Nakhel, C.; Kyriacou, M.C.; De Pascale, S. Foliar applications of a legume-derived protein hydrolysate elicit dose-dependent increases of growth, leaf mineral composition, yield and fruit quality in two greenhouse tomato cultivars. Sci. Hortic. 2017, 226, 353–360. [Google Scholar] [CrossRef]
  135. Sitohy, M.Z.; Desoky, E.-S.M.; Osman, A.; Rady, M.M. Pumpkin seed protein hydrolysate treatment alleviates salt stress effects on Phaseolus vulgaris by elevating antioxidant capacity and recovering ion homeostasis. Sci. Hortic. 2020, 271, 109495. [Google Scholar] [CrossRef]
  136. Koleška, I.; Hasanagić, D.; Todorović, V.; Murtić, S.; Klokić, I.; Parađiković, N.; Kukavica, B. Biostimulant prevents yield loss and reduces oxidative damage in tomato plants grown on reduced NPK nutrition. J. Plant Interact. 2017, 12, 209–218. [Google Scholar] [CrossRef] [Green Version]
  137. Casadesús, A.; Polo, J.; Munné-Bosch, S. Hormonal Effects of an Enzymatically Hydrolyzed Animal Protein-Based Biostimulant (Pepton) in Water-Stressed Tomato Plants. Front. Plant Sci. 2019, 10, 758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Ertani, A.; Schiavon, M.; Nardi, S. Transcriptome-Wide Identification of Differentially Expressed Genes in Solanum lycopersicon L. in Response to an Alfalfa-Protein Hydrolysate Using Microarrays. Front. Plant Sci. 2017, 8, 1159. [Google Scholar] [CrossRef] [Green Version]
  139. Sestili, F.; Rouphael, Y.; Cardarelli, M.; Pucci, A.; Bonini, P.; Canaguier, R.; Colla, G. Protein Hydrolysate Stimulates Growth in Tomato Coupled With N-Dependent Gene Expression Involved in N Assimilation. Front. Plant Sci. 2018, 9, 1233. [Google Scholar] [CrossRef] [Green Version]
  140. Ceccarelli, A.; Miras-Moreno, B.; Buffagni, V.; Senizza, B.; Pii, Y.; Cardarelli, M.; Rouphael, Y.; Colla, G.; Lucini, L. Foliar Application of Different Vegetal-Derived Protein Hydrolysates Distinctively Modulates Tomato Root Development and Metabolism. Plants 2021, 10, 326. [Google Scholar] [CrossRef]
  141. Paul, K.; Sorrentino, M.; Lucini, L.; Rouphael, Y.; Cardarelli, M.; Bonini, P.; Reynaud, H.; Canaguier, R.; Trtílek, M.; Panzarová, K.; et al. Understanding the Biostimulant Action of Vegetal-Derived Protein Hydrolysates by High-Throughput Plant Phenotyping and Metabolomics: A Case Study on Tomato. Front. Plant Sci. 2019, 10, 1–17. [Google Scholar] [CrossRef]
  142. Paul, K.; Sorrentino, M.; Lucini, L.; Rouphael, Y.; Cardarelli, M.; Bonini, P.; Miras Moreno, M.B.; Reynaud, H.; Canaguier, R.; Trtílek, M.; et al. A Combined Phenotypic and Metabolomic Approach for Elucidating the Biostimulant Action of a Plant-Derived Protein Hydrolysate on Tomato Grown Under Limited Water Availability. Front. Plant Sci. 2019, 10, 493. [Google Scholar] [CrossRef]
  143. Lucini, L.; Miras-Moreno, B.; Rouphael, Y.; Cardarelli, M.; Colla, G. Combining Molecular Weight Fractionation and Metabolomics to Elucidate the Bioactivity of Vegetal Protein Hydrolysates in Tomato Plants. Front. Plant Sci. 2020, 11, 976. [Google Scholar] [CrossRef] [PubMed]
  144. Hamedani, S.R.; Rouphael, Y.; Colla, G.; Colantoni, A.; Cardarelli, M. Biostimulants as a Tool for Improving Environmental Sustainability of Greenhouse Vegetable Crops. Sustainability 2020, 12, 5101. [Google Scholar] [CrossRef]
  145. Giordano, M.; El-Nakhel, C.; Caruso, G.; Cozzolino, E.; De Pascale, S.; Kyriacou, M.C.; Colla, G.; Rouphael, Y. Stand-Alone and Combinatorial Effects of Plant-based Biostimulants on the Production and Leaf Quality of Perennial Wall Rocket. Plants 2020, 9, 922. [Google Scholar] [CrossRef]
  146. Wilson, H.T.; Amirkhani, M.; Taylor, A.G. Evaluation of Gelatin as a Biostimulant Seed Treatment to Improve Plant Performance. Front. Plant Sci. 2018, 9, 1006. [Google Scholar] [CrossRef] [Green Version]
  147. Khan, S.; Yu, H.; Li, Q.; Gao, Y.; Sallam, B.N.; Wang, H.; Liu, P.; Jiang, W. Exogenous Application of Amino Acids Improves the Growth and Yield of Lettuce by Enhancing Photosynthetic Assimilation and Nutrient Availability. Agronomy 2019, 9, 266. [Google Scholar] [CrossRef] [Green Version]
  148. Tsouvaltzis, P.; Kasampalis, D.S.; Aktsoglou, D.-C.; Barbayiannis, N.; Siomos, A.S. Effect of Reduced Nitrogen and Supplemented Amino Acids Nutrient Solution on the Nutritional Quality of Baby Green and Red Lettuce Grown in a Floating System. Agronomy 2020, 10, 922. [Google Scholar] [CrossRef]
  149. Abdelhamid, M.T.; Sadak, M.S.; Schmidhalter, U. Effect of Foliar Application of Aminoacids on Plant Yield and Physiological Parameters in Bean Plants Irrigated with Seawater. Acta Biológica Colombiana 2014, 20, 140–152. [Google Scholar] [CrossRef]
  150. Kocira, S.; Szparaga, A.; Hara, P.; Treder, K.; Findura, P.; Bartoš, P.; Filip, M. Biochemical and economical effect of application biostimulants containing seaweed extracts and amino acids as an element of agroecological management of bean cultivation. Sci. Rep. 2020, 10, 1–16. [Google Scholar] [CrossRef]
  151. Mostafa, G.G. Improving the Growth of Fennel Plant Grown under Salinity Stress using some Biostimulants. Am. J. Plant Physiol. 2015, 10, 77–83. [Google Scholar] [CrossRef]
  152. Kałużewicz, A.; Krzesiński, W.; Spiżewski, T.; Zaworska, A. Effect of Biostimulants on Several Physiological Characteristics and Chlorophyll Content in Broccoli under Drought Stress and Re-watering. Not. Bot. Horti Agrobot. Cluj-Napoca 2017, 45, 197–202. [Google Scholar] [CrossRef] [Green Version]
  153. Klokić, I.; Koleška, I.; Hasanagić, D.; Murtić, S.; Bosančić, B.; Todorović, V. Biostimulants’ influence on tomato fruit characteristics at conventional and low-input NPK regime. Acta Agric. Scand. Sect. B Plant Soil Sci. 2020, 70, 233–240. [Google Scholar] [CrossRef]
  154. Azcona, I.; Pascual, I.; Aguirreolea, J.; Fuentes, M.; García-Mina, J.M.; Sánchez-Díaz, M. Growth and development of pepper are affected by humic substances derived from composted sludge. J. Plant Nutr. Soil Sci. 2011, 174, 916–924. [Google Scholar] [CrossRef]
  155. Sánchez, A.S.; Juárez, M.; Sánchez-Andreu, J.; Jordá, J.; Bermúdez, D. Use of Humic Substances and Amino Acids to Enhance Iron Availability for Tomato Plants from Applications of the Chelate FeEDDHA. J. Plant Nutr. 2005, 28, 1877–1886. [Google Scholar] [CrossRef]
  156. Dziugieł, T.; Wadas, W. Possibility of increasing early crop potato yield with foliar application of seaweed extracts and humic acids. J. Central Eur. Agric. 2020, 21, 300–310. [Google Scholar] [CrossRef]
  157. Türkmen, Ö.; Dursun, A.; Turan, M.; Erdinç, Ç. Calcium and humic acid affect seed germination, growth, and nutrient content of tomato (Lycopersicon esculentum L.) seedlings under saline soil conditions. Acta Agric. Scand. Sect. B Plant Soil Sci. 2004, 54, 168–174. [Google Scholar] [CrossRef]
  158. Paksoy, M.; Türkmen, Ö.; Dursun, A. Effects of potassium and humic acid on emergence, growth and nutrient contents of okra (Abelmoschus esculentus L.) seedling under saline soil conditions. Afr. J. Biotechnol. 2010, 9, 5343–5346. [Google Scholar]
  159. Pizzeghello, D.; Schiavon, M.; Francioso, O.; Vecchia, F.D.; Ertani, A.; Nardi, S. Bioactivity of Size-Fractionated and Unfractionated Humic Substances from Two Forest Soils and Comparative Effects on N and S Metabolism, Nutrition, and Root Anatomy of Allium sativum L. Front. Plant Sci. 2020, 11, 1203. [Google Scholar] [CrossRef]
  160. Balmori, D.M.; Domínguez, C.Y.A.; Carreras, C.R.; Rebatos, S.M.; Farías, L.B.P.; Izquierdo, F.G.; Berbara, R.L.L.; García, A.C. Foliar application of humic liquid extract from vermicompost improves garlic (Allium sativum L.) production and fruit quality. Int. J. Recycl. Org. Waste Agric. 2019, 8, 103–112. [Google Scholar] [CrossRef] [Green Version]
  161. Gemin, L.G.; Mógor, Á.F.; De Oliveira Amatussi, J.; Mógor, G. Microalgae associated to humic acid as a novel biostimulant improving onion growth and yield. Sci. Hortic. 2019, 256, 108560. [Google Scholar] [CrossRef]
  162. Selim, E.; Mosa, A.; El-Ghamry, A. Evaluation of humic substances fertigation through surface and subsurface drip irrigation systems on potato grown under Egyptian sandy soil conditions. Agric. Water Manag. 2009, 96, 1218–1222. [Google Scholar] [CrossRef]
  163. Selladurai, R.; Purakayastha, T.J. Effect of Humic Acid Multinutrient Fertilizers on Yield and Nutrient Use Efficiency of Potato. J. Plant Nutr. 2015, 39, 949–956. [Google Scholar] [CrossRef]
  164. Shalaby, O.A.E.-S.; El-Messairy, M.M. Humic acid and boron treatment to mitigate salt stress on the melon plant. Acta Agric. Slov. 2018, 111, 349–356. [Google Scholar] [CrossRef] [Green Version]
  165. Ibrahim, E.A.; Ramadan, W.A. Effect of zinc foliar spray alone and combined with humic acid or/and chitosan on growth, nutrient elements content and yield of dry bean (Phaseolus vulgaris L.) plants sown at different dates. Sci. Hortic. 2015, 184, 101–105. [Google Scholar] [CrossRef]
  166. Hartz, T.K.; Bottoms, T.G. Humic Substances Generally Ineffective in Improving Vegetable Crop Nutrient Uptake or Productivity. HortScience 2010, 45, 906–910. [Google Scholar] [CrossRef] [Green Version]
  167. Bettoni, M.M.; Mogor, Á.F.; Pauletti, V.; Goicoechea, N.; Aranjuelo, I.; Garmendia, I. Nutritional quality and yield of onion as affected by different application methods and doses of humic substances. J. Food Compos. Anal. 2016, 51, 37–44. [Google Scholar] [CrossRef] [Green Version]
  168. Waqas, M.; Ahmad, B.; Arif, M.; Munsif, F.; Khan, A.L.; Amin, M.; Kang, S.-M.; Kim, Y.-H.; Lee, I.-J. Evaluation of Humic Acid Application Methods for Yield and Yield Components of Mungbean. Am. J. Plant Sci. 2014, 5, 2269–2276. [Google Scholar] [CrossRef] [Green Version]
  169. Sani, M.N.H.; Islam, M.N.; Uddain, J.; Chowdhury, M.S.N.; Subramaniam, S. Synergistic effect of microbial and nonmicrobial biostimulants on growth, yield, and nutritional quality of organic tomato. Crop Sci. 2020, 60, 2102–2114. [Google Scholar] [CrossRef]
  170. Bonini, P.; Rouphael, Y.; Miras-Moreno, B.; Lee, B.; Cardarelli, M.; Erice, G.; Cirino, V.; Lucini, L.; Colla, G. A Microbial-Based Biostimulant Enhances Sweet Pepper Performance by Metabolic Reprogramming of Phytohormone Profile and Secondary Metabolism. Front. Plant Sci. 2020, 11, 1–13. [Google Scholar] [CrossRef] [PubMed]
  171. Rouphael, Y.; Carillo, P.; Colla, G.; Fiorentino, N.; Sabatino, L.; El-Nakhel, C.; Giordano, M.; Pannico, A.; Cirillo, V.; Shabani, E.; et al. Appraisal of Combined Applications of Trichoderma virens and a Biopolymer-Based Biostimulant on Lettuce Agronomical, Physiological, and Qualitative Properties under Variable N Regimes. Agronomy 2020, 10, 196. [Google Scholar] [CrossRef] [Green Version]
  172. Patkowska, E.; Mielniczuk, E.; Jamiołkowska, A.; Skwaryło-Bednarz, B.; BłaŻewicz-Woźniak, M. The Influence of Trichoderma harzianum Rifai T-22 and Other Biostimulants on Rhizosphere Beneficial Microorganisms of Carrot. Agronomy 2020, 10, 1637. [Google Scholar] [CrossRef]
  173. Pereira, T.D.S.; Macêdo, A.G.; Da Silva, J.; Pinheiro, J.B.; De Paula, A.M.; Biscaia, D.; Busato, J.G. Water-extractable fraction of vermicomposts enriched with Trichoderma enhances the growth of bell pepper and tomato as well as their tolerance against Meloidogyne incognita. Sci. Hortic. 2020, 272, 109536. [Google Scholar] [CrossRef]
  174. Rouphael, Y.; Franken, P.; Schneider, C.; Schwarz, D.; Giovannetti, M.; Agnolucci, M.; De Pascale, S.; Bonini, P.; Colla, G. Arbuscular mycorrhizal fungi act as biostimulants in horticultural crops. Sci. Hortic. 2015, 196, 91–108. [Google Scholar] [CrossRef]
  175. Latef, A.A.H.A.; Chaoxing, H. Does Inoculation with Glomus mosseae Improve Salt Tolerance in Pepper Plants? J. Plant Growth Regul. 2014, 33, 644–653. [Google Scholar] [CrossRef]
  176. Petropoulos, S.A. Practical Applications of Plant Biostimulants in Greenhouse Vegetable Crop Production. Agronomy 2020, 10, 1569. [Google Scholar] [CrossRef]
  177. Vasseur-Coronado, M.; du Boulois, H.D.; Pertot, I.; Puopolo, G. Selection of plant growth promoting rhizobacteria sharing suitable features to be commercially developed as biostimulant products. Microbiol. Res. 2021, 245, 126672. [Google Scholar] [CrossRef]
  178. Sarma, R.K.; Saikia, R. Alleviation of drought stress in mung bean by strain Pseudomonas aeruginosa GGRJ21. Plant Soil 2014, 377, 111–126. [Google Scholar] [CrossRef]
  179. Korir, H.; Mungai, N.W.; Thuita, M.; Hamba, Y.; Masso, C.; Tola, Y.H. Co-inoculation Effect of Rhizobia and Plant Growth Promoting Rhizobacteria on Common Bean Growth in a Low Phosphorus Soil. Front. Plant Sci. 2017, 8, 141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Vurukonda, S.S.K.P.; Vardharajula, S.; Shrivastava, M.; SkZ, A. Enhancement of drought stress tolerance in crops by plant growth promoting rhizobacteria. Microbiol. Res. 2016, 184, 13–24. [Google Scholar] [CrossRef]
  181. Yong, J.W.H.; Letham, D.S.; Wong, S.C.; Farquhar, G.D. Rhizobium-induced elevation in xylem cytokinin delivery in pigeonpea induces changes in shoot development and leaf physiology. Funct. Plant Biol. 2014, 41, 1323–1335. [Google Scholar] [CrossRef]
  182. Thao, H.T.B.; Yamakawa, T. Phosphite (phosphorous acid): Fungicide, fertilizer or bio-stimulator? Soil Sci. Plant Nutr. 2009, 55, 228–234. [Google Scholar] [CrossRef]
  183. Reyes-Pérez, J.J.; Enríquez-Acosta, E.A.; Ramírez-Arrebato, M.Á.; Rodríguez-Pedroso, A.T.; Falcón-Rodríguez, A. Effect of humic acids, mycorrhiza, and chitosan on growth indicators of two tomato cultivars (Solanum lycopersicum L.). Terra Latinoam. 2020, 38, 653–666. [Google Scholar] [CrossRef]
  184. Asgari-Targhi, G.; Iranbakhsh, A.; Ardebili, Z.O. Potential benefits and phytotoxicity of bulk and nano-chitosan on the growth, morphogenesis, physiology, and micropropagation of Capsicum annuum. Plant Physiol. Biochem. 2018, 127, 393–402. [Google Scholar] [CrossRef]
  185. Iglesias, M.J.; Colman, S.L.; Terrile, M.C.; París, R.; Martín-Saldaña, S.; Chevalier, A.A.; Álvarez, V.A.; Casalongué, C.A. Enhanced Properties of Chitosan Microparticles over Bulk Chitosan on the Modulation of the Auxin Signaling Pathway with Beneficial Impacts on Root Architecture in Plants. J. Agric. Food Chem. 2019, 67, 6911–6920. [Google Scholar] [CrossRef]
  186. Luyckx, M.; Hausman, J.-F.; Lutts, S.; Guerriero, G. Silicon and Plants: Current Knowledge and Technological Perspectives. Front. Plant Sci. 2017, 8, 411. [Google Scholar] [CrossRef] [Green Version]
  187. Laane, H.-M. The Effects of Foliar Sprays with Different Silicon Compounds. Plants 2018, 7, 45. [Google Scholar] [CrossRef] [Green Version]
  188. Rastogi, A.; Tripathi, D.K.; Yadav, S.; Chauhan, D.K.; Živčák, M.; Ghorbanpour, M.; El-Sheery, N.I.; Brestic, M. Application of silicon nanoparticles in agriculture. 3 Biotech 2019, 9, 1–11. [Google Scholar] [CrossRef] [Green Version]
  189. Shi, Y.; Zhang, Y.; Yao, H.; Wu, J.; Sun, H.; Gong, H. Silicon improves seed germination and alleviates oxidative stress of bud seedlings in tomato under water deficit stress. Plant Physiol. Biochem. 2014, 78, 27–36. [Google Scholar] [CrossRef] [PubMed]
  190. Romero-Aranda, M.R.; Jurado, O.; Cuartero, J. Silicon alleviates the deleterious salt effect on tomato plant growth by improving plant water status. J. Plant Physiol. 2006, 163, 847–855. [Google Scholar] [CrossRef] [PubMed]
  191. Shi, Y.; Zhang, Y.; Han, W.; Feng, R.; Hu, Y.; Guo, J.; Gong, H. Silicon enhances water stress tolerance by improving root hydraulic conductance in Solanum lycopersicum L. Front. Plant Sci. 2016, 7, 1–15. [Google Scholar] [CrossRef]
  192. Zhu, Y.-X.; Xu, X.-B.; Hu, Y.-H.; Han, W.-H.; Yin, J.-L.; Li, H.-L.; Gong, H.-J. Silicon improves salt tolerance by increasing root water uptake in Cucumis sativus L. Plant Cell Rep. 2015, 34, 1629–1646. [Google Scholar] [CrossRef] [PubMed]
  193. Tantawy, A.S.; Salama, Y.; El-Nemr, M.A.; Abdel-Mawgoud, A. Nano silicon application improves salinity tolerance of sweet pepper plants. Int. J. ChemTech Res. 2015, 8, 11–17. [Google Scholar]
  194. Siddiqui, M.H.; Al-Whaibi, M.H.; Faisal, M.; Al Sahli, A.A. Nano-silicon dioxide mitigates the adverse effects of salt stress on Cucurbita pepo L. Environ. Toxicol. Chem. 2014, 33, 2429–2437. [Google Scholar] [CrossRef]
  195. Przybysz, A.; Gawrońska, H.; Gajc-Wolska, J. Biological mode of action of a nitrophenolates-based biostimulant: Case study. Front. Plant Sci. 2014, 5, 1–15. [Google Scholar] [CrossRef]
  196. Elmarie, V.D.W.; Johan, C.P.; Van Der Watt, E.; Pretorius, J.C. A triglyceride from Lupinus albus L. seed with bio-stimulatory properties. Afr. J. Biotechnol. 2013, 12, 5431–5443. [Google Scholar] [CrossRef] [Green Version]
  197. Wadas, W.; Kalinowski, K. Effect of Tytanit on the dry matter and macroelement contents in potato tuber. J. Central Eur. Agric. 2018, 19, 557–570. [Google Scholar] [CrossRef] [Green Version]
  198. Rady, M.M.; Varma, C.B.; Howladar, S.M. Common bean (Phaseolus vulgaris L.) seedlings overcome NaCl stress as a result of presoaking in Moringa oleifera leaf extract. Sci. Hortic. 2013, 162, 63–70. [Google Scholar] [CrossRef]
  199. Galambos, N.; Compant, S.; Moretto, M.; Sicher, C.; Puopolo, G.; Wäckers, F.; Sessitsch, A.; Pertot, I.; Perazzolli, M. Humic Acid Enhances the Growth of Tomato Promoted by Endophytic Bacterial Strains Through the Activation of Hormone-, Growth-, and Transcription-Related Processes. Front. Plant Sci. 2020, 11, 1–18. [Google Scholar] [CrossRef]
  200. Olivares, F.L.; Aguiar, N.O.; Rosa, R.C.C.; Canellas, L.P. Substrate biofortification in combination with foliar sprays of plant growth promoting bacteria and humic substances boosts production of organic tomatoes. Sci. Hortic. 2015, 183, 100–108. [Google Scholar] [CrossRef]
  201. Ekin, Z. Integrated Use of Humic Acid and Plant Growth Promoting Rhizobacteria to Ensure Higher Potato Productivity in Sustainable Agriculture. Sustainability 2019, 11, 3417. [Google Scholar] [CrossRef] [Green Version]
  202. Sandepogu, M.; Shukla, P.S.; Asiedu, S.; Yurgel, S.; Prithiviraj, B. Combination of Ascophyllum nodosum Extract and Humic Acid Improve Early Growth and Reduces Post-Harvest Loss of Lettuce and Spinach. Agriculture 2019, 9, 240. [Google Scholar] [CrossRef] [Green Version]
  203. Ngoroyemoto, N.; Kulkarni, M.G.; Stirk, W.A.; Gupta, S.; Finnie, J.F.; van Staden, J. Interactions Between Microorganisms and a Seaweed-Derived Biostimulant on the Growth and Biochemical Composition of Amaranthus hybridus L. Nat. Prod. Commun. 2020, 15, 1–11. [Google Scholar] [CrossRef]
  204. Rouphael, Y.; Colla, G. Synergistic Biostimulatory Action: Designing the Next Generation of Plant Biostimulants for Sustainable Agriculture. Front. Plant Sci. 2018, 9, 1655. [Google Scholar] [CrossRef] [Green Version]
  205. Shehata, S.A.; AbdelGawad, K.F.; Elmogy, M. Quality and Shelf-life of Onion Bulbs Influenced by Bio-stimulants. Int. J. Veg. Sci. 2017, 23, 362–371. [Google Scholar] [CrossRef]
  206. Semida, W.M.; El-Mageed, T.A.A.; Hemida, K.; Rady, M.M. Natural bee-honey based biostimulants confer salt tolerance in onion via modulation of the antioxidant defence system. J. Hortic. Sci. Biotechnol. 2019, 94, 632–642. [Google Scholar] [CrossRef]
  207. Anbarasi, D.; Haripriya, K. Response of aggregatum onion (Allium cepa L. var. aggregatum Don.) to organic inputs, bio fertilizers and biostimulants. Plant Arch. 2020, 20, 759–762. [Google Scholar]
  208. Agung, I.G.A.M.S.; Diara, I.W. Biostimulants Enhanced Seedling Root Growth and Bulb Yields of True Seed Shallots (Allium cepa var aggregatum L.). Int. J. Environ. Agric. Biotechnol. 2019, 4, 598–601. [Google Scholar] [CrossRef]
  209. Ngoroyemoto, N.; Gupta, S.; Kulkarni, M.; Finnie, J.; Van Staden, J. Effect of organic biostimulants on the growth and biochemical composition of Amaranthus hybridus L. S. Afr. J. Bot. 2019, 124, 87–93. [Google Scholar] [CrossRef]
  210. Arthur, G.D.; Aremu, A.O.; Kulkarni, M.G.; Okem, A.; Stirk, W.A.; Davies, T.C.; Van Staden, J. Can the use of natural biostimulants be a potential means of phytoremediating contaminated soils from goldmines in South Africa? Int. J. Phytoremediation 2015, 18, 427–434. [Google Scholar] [CrossRef] [PubMed]
  211. Kałużewicz, A.; Gąsecka, M.; Spiżewski, T. Influence of biostimulants on phenolic content in broccoli heads directly after harvest and after storage. Folia Hortic. 2017, 29, 221–230. [Google Scholar] [CrossRef] [Green Version]
  212. Navarro-León, E.; López-Moreno, F.J.; Rios, J.J.; Blasco, B.; Ruiz, J.M. Assaying the use of sodium thiosulphate as a biostimulant and its effect on cadmium accumulation and tolerance in Brassica oleracea plants. Ecotoxicol. Environ. Saf. 2020, 200, 110760. [Google Scholar] [CrossRef]
  213. Sadeghi, H.; Taban, A. Crushed maize seeds enhance soil biological activity and salt tolerance in caper (Capparis spinosa L.). Ind. Crop. Prod. 2021, 160, 113103. [Google Scholar] [CrossRef]
  214. Barrajón-Catalán, E.; Álvarez-Martínez, F.J.; Borrás, F.; Pérez, D.; Herrero, N.; Ruiz, J.J.; Micol, V. Metabolomic analysis of the effects of a commercial complex biostimulant on pepper crops. Food Chem. 2020, 310, 125818. [Google Scholar] [CrossRef]
  215. Dzung, P.D.; Van Phu, D.; Du, B.D.; Ngoc, L.S.; Duy, N.N.; Hiet, H.D.; Nghia, D.H.; Thang, N.T.; Van Le, B.; Hien, N.Q. Effect of foliar application of oligochitosan with different molecular weight on growth promotion and fruit yield enhancement of chili plant. Plant Prod. Sci. 2017, 20, 389–395. [Google Scholar] [CrossRef]
  216. Shirkhodaei, M.; Darzi, M.T.; Haj, M.; Hadi, M.H.S. Influence of Vermicompost and Biostimulant on the growth and biomass of coriander (Coriandrum sativum L.). Int. J. Adv. Biol. Biomed. Res. 2014, 2, 706–714. [Google Scholar]
  217. Pokluda, R.; Sękara, A.; Jezdinský, A.; Kalisz, A.; Neugebauerová, J.; Grabowska, A. The physiological status and stress biomarker concentration of Coriandrum sativum L. plants subjected to chilling are modified by biostimulant application. Biol. Agric. Hortic. 2016, 32, 258–268. [Google Scholar] [CrossRef]
  218. Negro, D.; Montesano, V.; Sonnante, G.; Rubino, P.; De Lisi, A.; Sarli, G. Fertilization strategies on cultivars of globe artichoke: Effects on yield and quality performance. J. Plant Nutr. 2015, 39, 279–287. [Google Scholar] [CrossRef]
  219. Wszelaczyńska, E.; Szczepanek, M.; Pobereżny, J.; Kazula, M.J. Effect of biostimulant application and long-term storage on the nutritional value of carrot. Hortic. Bras. 2019, 37, 451–457. [Google Scholar] [CrossRef] [Green Version]
  220. Lucini, L.; Rouphael, Y.; Cardarelli, M.; Canaguier, R.; Kumar, P.; Colla, G. The effect of a plant-derived biostimulant on metabolic profiling and crop performance of lettuce grown under saline conditions. Sci. Hortic. 2015, 182, 124–133. [Google Scholar] [CrossRef]
  221. Hernandez, O.L.; Calderín, A.; Huelva, R.; Martínez-Balmori, D.; Guridi, F.; Aguiar, N.O.; Olivares, F.L.; Canellas, L.P. Humic substances from vermicompost enhance urban lettuce production. Agron. Sustain. Dev. 2015, 35, 225–232. [Google Scholar] [CrossRef] [Green Version]
  222. Torres, P.; Novaes, P.; Ferreira, L.G.; Santos, J.P.; Mazepa, E.; Duarte, M.E.R.; Noseda, M.D.; Chow, F.; Dos Santos, D.Y. Effects of extracts and isolated molecules of two species of Gracilaria (Gracilariales, Rhodophyta) on early growth of lettuce. Algal Res. 2018, 32, 142–149. [Google Scholar] [CrossRef]
  223. Vetrano, F.; Miceli, C.; Angileri, V.; Frangipane, B.; Moncada, A.; Miceli, A. Effect of Bacterial Inoculum and Fertigation Management on Nursery and Field Production of Lettuce Plants. Agronomy 2020, 10, 1477. [Google Scholar] [CrossRef]
  224. Alkuwayti, M.; El-Sherif, F.; Yap, Y.-K.; Khattab, S. Foliar application of Moringa oleifera leaves extract altered stress-responsive gene expression and enhanced bioactive compounds composition in Ocimum basilicum. S. Afr. J. Bot. 2020, 129, 291–298. [Google Scholar] [CrossRef]
  225. Taha, R.; Alharby, H.; Bamagoos, A.; Medani, R.; Rady, M. Elevating tolerance of drought stress in Ocimum basilicum using pollen grains extract; a natural biostimulant by regulation of plant performance and antioxidant defense system. S. Afr. J. Bot. 2020, 128, 42–53. [Google Scholar] [CrossRef]
  226. Rady, M.; Desoky, E.-S.; Elrys, A.; Boghdady, M. Can licorice root extract be used as an effective natural biostimulant for salt-stressed common bean plants? S. Afr. J. Bot. 2019, 121, 294–305. [Google Scholar] [CrossRef]
  227. Rady, M.M.; Mohamed, G.F. Modulation of salt stress effects on the growth, physio-chemical attributes and yields of Phaseolus vulgaris L. plants by the combined application of salicylic acid and Moringa oleifera leaf extract. Sci. Hortic. 2015, 193, 105–113. [Google Scholar] [CrossRef]
  228. Elzaawely, A.A.; Ahmed, M.E.; Maswada, H.F.; Xuan, T.D. Enhancing growth, yield, biochemical, and hormonal contents of snap bean (Phaseolus vulgaris L.) sprayed with moringa leaf extract. Arch. Agron. Soil Sci. 2016, 63, 687–699. [Google Scholar] [CrossRef]
  229. Elzaawely, A.A.; Ahmed, M.E.; Maswada, H.F.; Al-Araby, A.A.; Xuan, T.D. Growth traits, physiological parameters and hormonal status of snap bean (Phaseolus vulgaris L.) sprayed with garlic cloves extract. Arch. Agron. Soil Sci. 2018, 64, 1068–1082. [Google Scholar] [CrossRef]
  230. Merwad, A.-R.M.A. Using Moringa oleifera extract as biostimulant enhancing the growth, yield and nutrients accumulation of pea plants. J. Plant Nutr. 2017, 41, 425–431. [Google Scholar] [CrossRef]
  231. Desoky, E.-S.M.; ElSayed, A.I.; Merwad, A.-R.M.; Rady, M.M. Stimulating antioxidant defenses, antioxidant gene expression, and salt tolerance in Pisum sativum seedling by pretreatment using licorice root extract (LRE) as an organic biostimulant. Plant Physiol. Biochem. 2019, 142, 292–302. [Google Scholar] [CrossRef] [PubMed]
  232. Rady, M.M.; Rehman, H. Supplementing organic biostimulants into growing media enhances growth and nutrient uptake of tomato transplants. Sci. Hortic. 2016, 203, 192–198. [Google Scholar] [CrossRef]
  233. Chanthini, K.M.-P.; Senthil-Nathan, S.; Stanley-Raja, V.; Thanigaivel, A.; Karthi, S.; Sivanesh, H.; Sundar, N.S.; Palanikani, R.; Soranam, R. Chaetomorpha antennina (Bory) Kützing derived seaweed liquid fertilizers as prospective bio-stimulant for Lycopersicon esculentum (Mill). Biocatal. Agric. Biotechnol. 2019, 20, 101190. [Google Scholar] [CrossRef]
  234. Colman, S.L.; Salcedo, M.F.; Mansilla, A.Y.; Iglesias, M.J.; Fiol, D.F.; Martín-Saldaña, S.; Alvarez, V.A.; Chevalier, A.A.; Casalongué, C.A. Chitosan microparticles improve tomato seedling biomass and modulate hormonal, redox and defense pathways. Plant Physiol. Biochem. 2019, 143, 203–211. [Google Scholar] [CrossRef] [PubMed]
  235. Maach, M.; Boudouasar, K.; Akodad, M.; Skalli, A.; Moumen, A.; Baghour, M. Application of biostimulants improves yield and fruit quality in tomato. Int. J. Veg. Sci. 2020, 1–6. [Google Scholar] [CrossRef]
  236. Supraja, K.V.; Behera, B.; Balasubramanian, P. Efficacy of microalgal extracts as biostimulants through seed treatment and foliar spray for tomato cultivation. Ind. Crops Prod. 2020, 151, 112453. [Google Scholar] [CrossRef]
  237. Rahou, Y.A.; Ait-El-Mokhtar, M.; Anli, M.; Boutasknit, A.; Ben-Laouane, R.; Douira, A.; Benkirane, R.; El Modafar, C.; Meddich, A. Use of mycorrhizal fungi and compost for improving the growth and yield of tomato and its resistance to Verticillium dahliae. Arch. Phytopathol. Plant Prot. 2020, 1–26. [Google Scholar] [CrossRef]
  238. Dong, C.; Wang, G.; Du, M.; Niu, C.; Zhang, P.; Zhang, X.; Ma, D.; Ma, F.; Bao, Z. Biostimulants promote plant vigor of tomato and strawberry after transplanting. Sci. Hortic. 2020, 267, 109355. [Google Scholar] [CrossRef]
  239. Rachidi, F.; Benhima, R.; Sbabou, L.; El Arroussi, H. Microalgae polysaccharides bio-stimulating effect on tomato plants: Growth and metabolic distribution. Biotechnol. Rep. 2020, 25, e00426. [Google Scholar] [CrossRef] [PubMed]
  240. Pohl, A.; Grabowska, A.; Kalisz, A.; Sękara, A. Biostimulant Application Enhances Fruit Setting in Eggplant—An Insight into the Biology of Flowering. Agronomy 2019, 9, 482. [Google Scholar] [CrossRef] [Green Version]
  241. Pohl, A.; Grabowska, A.; Kalisz, A.; Sekara, A. The eggplant yield and fruit composition as affected by genetic factor and biostimulant application. Not. Bot. Horti Agrobot. Cluj-Napoca 2019, 47, 929–938. [Google Scholar] [CrossRef] [Green Version]
  242. Ali, M.; Cheng, Z.-H.; Hayat, S.; Ahmad, H.; Ghani, M.I.; Liu, T. Foliar spraying of aqueous garlic bulb extract stimulates growth and antioxidant enzyme activity in eggplant (Solanum melongena L.). J. Integr. Agric. 2019, 18, 1001–1013. [Google Scholar] [CrossRef]
  243. Abbas, S.M. The influence of biostimulants on the growth and on the biochemical composition of Vicia faba CV. Giza 3 beans. Rom. Biotechnol. Lett. 2013, 18, 8061–8068. [Google Scholar]
Figure 1. The most important biostimulant effects on crops.
Figure 1. The most important biostimulant effects on crops.
Biomolecules 11 00698 g001
Figure 2. Protein hydrolysates biostimulatory effects.
Figure 2. Protein hydrolysates biostimulatory effects.
Biomolecules 11 00698 g002
Table 1. Classification of plant biostimulants.
Table 1. Classification of plant biostimulants.
Plant BiostimulantsKey PointsReferences
Protein hydrolysates (PHs) and other N-containing compounds (amino acids)a. Mixtures of peptides and amino acids which are produced via enzymatic, chemical or thermal hydrolysis of animal- or plant-derived proteins. [34,84]
b. Effective in increasing yield and quality of various crop products.[85]
c. Categorization based on proteins, sources and the hydrolysis system; PHs boost both primary and secondary plant metabolism biochemical and physiological procedures.[86,87]
d. Effective in alleviating negative abiotic stress effects.[24]
Humic substancesa. Include fulvic acids and humic acids which they differ in color, molecular weight, carbon content and the degree of polymerization. [88]
b. They could increase cationic exchange capacity (CEC) of the soil and interact with root membrane transporters. [65]
Seaweed extractsa. Extracts from brown seaweeds, e.g., Ascophyllum, Fucus, and Laminaria genera. [89]
b. They are rich in polysaccharides, polyphenols and compounds with hormonal activity that affect plant growth and development.[90,91]
Biopolymers (Chitosans and other polymers)a. Chitosans are naturally occurring components in fungi nematodes, insects and crustaceans.[68]
b. They regulate plant-defense mechanisms related to phytoalexins biosynthesis, reactive oxygen species, and pathogenesis-related proteins making plants more resistant to biotic and abiotic stressors.[92]
Microbial biostimulants (Mycorrhizal and non-mycorrhizal fungi, Rhizobium, Trichoderma, and Plant Growth-Promoting Rhizobacteria (PGPR)) a. Symbiotic fungi, especially arbuscular mycorrhizal fungi (AMF) within the genus Glomus. [14,16]
b. Trichoderma genus[44]
c. Beneficial bacteria with plant growth promoting properties also known as PGPBs (Bacillus, Rhizobium, Pseudomonas, Azospirillum, Azotobacter, and many others).[48]
Phosphite (Phi)a. A phosphate (H2PO4) analog which affects various plant growth and development processes.[93]
b. Several beneficial effects have been reported in various vegetable crops.[69,94,95,96,97]
c. Biostimulatory impacts on fruit such as avocado, banana, citrus, peach, raspberry and strawberry.[69,98,99,100]
Silicona. Effective against abiotic and biotic stressors.[11]
Vermicompostsa. Hormonal activity of vermicompost leachates due to content in trace elements of hormones such as cytokinins, indolo-acetic acid (IAA), eighteen gibberellins (GAs) and brasinosteroids.[101]
b. Phytohormones from three different classes, including cytokinins, auxins and gibberellins provide plant growth promoting activities in vermicompost[102]
Table 2. Selected biostimulants effects on various vegetable crops.
Table 2. Selected biostimulants effects on various vegetable crops.
PlantCommon NameKey PointsEffectsReferences
Allium cepa L.Oniona. Biostimulants containing humic acids, organic substances, amino acids, carbon and boron or algae extractsImproved plant growth and yield, and shelf life of bulbs[205]
b. Application of diluted bee-honey extract (DHE)Increased photosynthetic parameters, biomass production and yield, and antioxidants activity[206]
Allium cepa var. aggregatum L. Shallota. Application of seaweed extracts, vermicompost and mixture of animal wasteImproved yield and bulb traits[207]
b. Soaking of seeds in PGPB biostimulantsIncreased germination percentage, plant growth and bulb parameters [208]
Allium sativum L.Garlica. Foliar application of liquid humic substances obtained from vermicompost extractsImproved yield and quality parameters of bulbs[160]
Amaranthus hybridus L.Amarantha. Foliar application of vermicompost leachate, smoke-water, karrikinolide, eckol and KelpakIncreased growth, higher chlorophylls, carotenoids and proteins content [209]
b. Combination of plant growth-promoting rhizobacteria (PGPRs), and Ecklonia maxima extractsImproved plant growth and photosynthetic pigment content, stress relief[203]
Brassica juncea L.Mustard greena. Foliar application of vermicompost leachate, smoke-water, indole-3-butyric acid and Ecklonia maxima extracts on seedlings grown in soils from goldmines Increased phytoremediative activities though the accumulation of heavy metals [210]
Brassica oleraceae L.Broccolia. Combination of foliar spraying with Ascophyllum nodosum extracts and watering with amino acids on broccoli plants subjected to water stress and re-wateringIncreased photosynthetic parameters under water stress conditions[152]
b. Combination of foliar spraying with Ascophyllum nodosum extracts and watering with amino acids on broccoli plantsIncreased total phenolic compounds, sinapic acid and quercetin content[211]
Brassica oleraceae L.Cabbagea. Foliar application of eckol from Ecklonia maxima extracts Increased root and shoot length, photosynthetic pigments and proteins, proline and iridoid glycosides; inhibition of infestation from aphids[39]
b. Thiosulfate application through the nutrient solution in cabbage plants subjected to Cd toxicityImproved phytoremediative properties of Cd without biostimulant effects on cabbage plants [212]
Capparis spinosa L. Capera. Incorporation of crushed maize seeds in growing medium of caper plants subjected to salinity stressIncreased activity of soil enzymes, Na exclusion from plant tissues[213]
Capsicum annuum L. Peppera. Application of a lipo-complex biostimulant containing mainly polysaccharides, polypeptides and vitamins Increased phenylalanine and metabolites associated with fruit ripening (organic acids, monosaccharides, carotenoids)[214]
Capsicum frutescens L.Chilli peppera. Foliar application of oligochitosan Increased plant growth, chlorophyll content and fruit weight[215]
Coriandrum sativum L.Coriandera. Seed inoculation with Azotobacter chroococcum and Azospirillum lipoferumIncreased biomass production[216]
b. Foliar spraying with biostimulants (Asahi SL or Goemar Goteo) on plants subjected to chilling stressIncreased photosynthetic parameters, L-ascorbic acid and total phenolic compounds content and total antioxidant activity [217]
Cynara scolymus L.Globe artichokea. Application of A. nodosum extracts and trace elements. Increased number and weight of heads [218]
Daucus carota subsp. sativus Carrota. Foliar application of Kelpak SL and Asahi SLKelpak SL improved nutritional value and increased storage life of carrots[219]
Lactuca sativa L. Lettucea. Foliar and root application of protein hydrolysates in lettuce plants grown under salinity conditionsMitigation of oxidative stress, increased osmolytes and glucosinolates content [220]
b. Foliar application of liquid humic substances obtained from vermicompostImproved earliness of plants, increased the number of leaves per plant and total yield[221]
c. Application of crude seaweed extracts (Gracilaria caudate and Gracilaria domingensis) on lettuce seedlings Increased root growth[222]
d. Inoculation of growth substrate with Bacillus spp.Positive effects on plant growth nitrate content [223]
Manihot esculenta CrantzCassavaa. Foliar application of Moringa oleifera leaves extracts. The plant height and leaf number of cassava plant were increased because of foliar application of MLE. Improved plant growth and decreased incidence of Zonocerus variegatus attacks[219]
Nasturtium officinaleWatercressa. Foliar spraying of algal biostimulant on watercress plants grown in Cd contaminated soilIncreased plant growth and reduced Cd accumulation in plant tissues[208]
Ocimum basilicum L.Basila. Foliar application of Moringa oleifera leaves extracts Increased growth and yield, and estragole and eucalyptol contents [224]
b. Foliar spraying with palm pollen grains extract alleviated the negative effects of deficit irrigation on basil plants. Improved plant growth and essential oils content and antioxidant enzyme activities; maintained relative water content, electrolyte leakage and water use efficiency; improved leaf and stem anatomy[225]
Phaseolus vulgaris L.Common beana. Foliar application of protein hydrolysates from pumkin seeds on Phaseolus vulgaris plants grown under saline conditionsMaintained plant growth, yield and anatomical features; mitigated negative effects of salt stress on macronutrients, photosynthetic pigments, relative water content and stability of cell membranes [135]
b. Foliar and soil application of Nomoren, EKOprop, Veramin Ca on Phaseouls vulgaris plants grown under normal irrigation and water stress conditionsPositive effects on yield, nutritional parameters, chemical composition and bioactive properties of fresh pods and seeds [14,15]
c. Seed soaking and foliar spraying with licorice root extract on common bean plants subjected to salinity stressImproved growth, yield and physicochemical parameters[226]
d. Seed soaking and foliar spraying with salicylic acid and Moringa oleifera leaves extracts on common bean plants subjected to salinity stressImproved growth, yield and physicochemical parameters[227]
Phaseolus vulgaris L.Snap beana. Foliar spraying with Moringa oleifera leaves extracts Improved plant growth and yield components, increased total phenolic compounds and minerals content in pods[228]
b. Foliar spraying with garlic cloves extractsImproved plant growth parameters, yield and chemical composition of pods[229]
Pisum sativum L.Peaa. Foliar spraying with Moringa oleifera leaves extractsIncreased biomass production, pod and seed yield, proteins and nutrients content in seeds [230]
b. Seed soaking in licorice root extractIncreased seedling growth, photosynthetic activity and antioxidant enzymes activity[231]
Solanum lycopersicum L.Tomatoa. Incorporation of humic acids and/or crushed maize grainImproved shoot and root growth, increased relative water content and membrane stability of transplants, improved macronutrients uptake [232]
b. Seed pretreatment with liquid extracts of Chaetomorpha antennina green seaweedIncreased germination percentage and vegetative growth, improved biochemical profile [233]
c. Foliar spraying with Chitosan microparticlesImproved seed germination and seedling vigor, modulation of antioxidant enzymes activities[234]
d. Foliar treatment with saffron extractImprovement in morphological and biochemical parameters, antifungal effects against Phytophthora infestans [234]
e. Foliar application of humic (Megafol) and amino acids (Viva) biostimulantsImproved plant growth under normal fertilization rates and minimized yield loses under nutrients deprivation[153]
f. Foliar application of Tecamin Brix® and/or Tecamin Flower ® in tomato plants grown in saline conditions. Improved yield and fruit quality[235]
g. Deed treatment and foliar spraying of microalgal extractsImproved germination and seedling growth rates[236]
h. Soil application of compost and arbuscular mycorrhizal fungiImproved plant growth and photosynthetic parameters, reduced incidence of Verticillium dahliae infestations[237]
i. Soil application of biostimulants containing plant extracts, Ascophyllum nodosum extracts or animal derived protein hydrolysates in tomato plants after transplantationReduced transplantation shock through the increase of root and shoot development[238]
j. Fertigation with microalgae polysaccharidesImproved vegetative growth, increased nutrients, protein and sterols content in leaves[239]
k. Foliar application of brown seaweed extracts from A. nodosum and Sargassum sp.Induced flower formation and fruit setting[115]
Solanum melongena L. Eggplanta. Foliar application of A. nodosum extractsImproved flower and fruit set, fruit yield and chemical composition[240,241]
b. Foliar application of aqueous garlic bulbSingle application increased plant growth, photosynthetic parameters and antioxidant enzymes activity[242]
Solanum tuberosum L. Potatoa. Combined application of Ecklonia maxima extracts and Asahi SL with herbicidesIncreased content of true and total proteins, increased marketable yield and yield parameters[120,123]
b. Soil spraying of biostimulant containing N-fixing microbes (NFM0 combined or not with an amino acid blendUnintended impacts on nitrogen losses [82]
c. Potato seed pretreatment and foliage spraying with phosphite Induced structural and biochemical changes in tuber periderm and cortex, increased tolerance to UV-B, enhanced sprouts emergence and early growth [94,95,97]
d. Foliar application of biostimulants containing A. nodosum extracts, E. maxima extracts or humic and fulvic acidsIncreased yield under drought stress, increased marketable yield[156]
Spinacia oleracea L.Spinacha. Foliar spraying of smoke-water and Ecklonia maxima extracts, Increased growth and biochemical parameters (antioxidant enzymes activity and sinapic acid content)[119]
b. Application of various biostimulants (Megafol, Aminovert, Veramin Ca, Twin Antistress and irrigation treatments on spinach plants grown under normal and water stress conditionsImproved nutritional value and bioactive properties[16]
Vicia faba L.Broad beana. Foliar spraying with Bacillus licheniformis, yeast (5 g/L), extracts form algae and humic acid (20 g/L), increased pigments, carotenoids concentrations and total carbohydrates. Improved photosynthesis and nutrients uptake, induced endogenous hormones and protein biosynthesis[243]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shahrajabian, M.H.; Chaski, C.; Polyzos, N.; Petropoulos, S.A. Biostimulants Application: A Low Input Cropping Management Tool for Sustainable Farming of Vegetables. Biomolecules 2021, 11, 698. https://doi.org/10.3390/biom11050698

AMA Style

Shahrajabian MH, Chaski C, Polyzos N, Petropoulos SA. Biostimulants Application: A Low Input Cropping Management Tool for Sustainable Farming of Vegetables. Biomolecules. 2021; 11(5):698. https://doi.org/10.3390/biom11050698

Chicago/Turabian Style

Shahrajabian, Mohamad Hesam, Christina Chaski, Nikolaos Polyzos, and Spyridon A. Petropoulos. 2021. "Biostimulants Application: A Low Input Cropping Management Tool for Sustainable Farming of Vegetables" Biomolecules 11, no. 5: 698. https://doi.org/10.3390/biom11050698

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