Next Article in Journal
Melt Electrospinning of Polymers: Blends, Nanocomposites, Additives and Applications
Next Article in Special Issue
Nano-Silver Particles Reduce Contaminations in Tissue Culture but Decrease Regeneration Rate and Slows Down Growth and Development of Aldrovanda vesiculosa Explants
Previous Article in Journal
Research on Microstructure of Soft Clay under Various Artificial Ground Freezing Conditions Based on NMR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Responses of Medicinal and Aromatic Plants to Engineered Nanoparticles †

1
Institute of Chemistry, Faculty of Natural Sciences, Comenius University, Ilkovicova 6, 842 15 Bratislava, Slovakia
2
Department of Analytical Chemistry, Faculty of Natural Sciences, Comenius University, Ilkovicova 6, 842 15 Bratislava, Slovakia
3
Regional Centre of Advanced Technologies and Materials, Czech Advanced Technology and Research Institute, Palacky University, Slechtitelu 27, 783 71 Olomouc, Czech Republic
*
Author to whom correspondence should be addressed.
This article is sincerely dedicated to Dr. Elena Masarovičová, Professor of Plant Physiology at the Department of Soil Science, Faculty of Natural Sciences, Comenius University in Bratislava, Slovakia.
Appl. Sci. 2021, 11(4), 1813; https://doi.org/10.3390/app11041813
Submission received: 21 January 2021 / Revised: 8 February 2021 / Accepted: 15 February 2021 / Published: 18 February 2021
(This article belongs to the Special Issue Interaction between Nanoparticles and Plants)

Abstract

:
Medicinal and aromatic plants have been used by mankind since ancient times. This is primarily due to their healing effects associated with their specific secondary metabolites (some of which are also used as drugs in modern medicine), or their structures, served as a basis for the development of new effective synthetic drugs. One way to increase the production of these secondary metabolites is to use nanoparticles that act as elicitors. However, depending on the specific particle size, composition, concentration, and route of application, nanoparticles may have several other benefits on medicinal and aromatic plants (e.g., increased plant growth, improved photosynthesis, and overall performance). On the other hand, particularly at applications of high concentrations, they are able to damage plants mechanically, adversely affect morphological and biochemical characteristics of plants, and show cytotoxic and genotoxic effects. This paper provides a comprehensive overview of the beneficial and adverse effects of metal-, metalloid-, and carbon-based nanoparticles on the germination, growth, and biochemical characteristics of a wide range of medicinal and aromatic plants, including the corresponding mechanisms of action. The positive impact of nanopriming and application of nanosized fertilizers on medicinal and aromatic plants is emphasized. Special attention is paid to the effects of various nanoparticles on the production of valuable secondary metabolites in these plants cultivated in hydroponic systems, soil, hairy root, or in vitro cultures. The beneficial impact of nanoparticles on the alleviation of abiotic stresses in medicinal and aromatic plants is also discussed.

1. Introduction

The 20th century was characterized by a rapid increase in the use of synthetic (often toxic) substances in various sectors of the economy, including agriculture, food, and in the pharmaceutical industry. Moreover, huge amounts of harmful solvents were often used to prepare required drugs. Due to the growing environmental contamination, which is also reflected in adverse effects on human health, “green” solutions are indispensable today. However, unlike the “Green Revolution” of the mid-20th century, which significantly increased the amounts of synthetic pesticides and fertilizers to achieve high yields [1], increased attention is now being paid to the widest possible use of biologically/pharmacologically active compounds of natural origins [2,3,4]. Medicinal plants, and in particular their essential oils (EOs), have been used in folk medicine since prehistoric times. For example, the use of medicinal plants is mentioned in the Bible. Moreover, recipes with medicinal plants can be found on a Sumerian tablet dating back to 3000 BC as well as Ebers papyrus (1500 BC). Several important works dealing with medicinal plants and corresponding medicines prepared from these plants, which arose during antiquity and the Middle Ages, were also preserved. The Herball, or Generall Historie of Plantes, published by the English botanist John Gerarde in 1597, includes approximately 600 medicinal plants [5,6,7,8,9,10,11,12]. Some fragrant medicinal plants, such as Myrtus communis L., Ocimum basilicum L., Rosmarinus officinalis L., and Salvia fruticosa Mill, are used as ritual plants in the main monotheistic religion [13].
Medicinal plants provide valuable sources of compounds that contain healing effects [5,6]. The phytotherapeutic effects of medicinal plants are caused by specific secondary metabolites, i.e., biologically active compounds formed in the processes of secondary metabolism. Unlike primary metabolism, which covers the growth and development of an individual, and can be characterized as indispensable, universal, uniform, and conservative, secondary metabolism covers the interaction of an individual with the environment, and although it is dispensable for growth and development, it is indispensable for survival in the environment and is considered unique, diverse, and adaptive [14]. For the pharmaceutical industry, secondary metabolites of medicinal plants that exhibit healing effects are particularly useful as lead compounds for the design of effective drugs [15,16,17]. Recently, there has been growing interest in obtaining increased amounts of valuable secondary metabolites of medicinal plants through hairy root cell suspension cultures. The advantage of these “green methods” compared to synthetic methods is the elimination of harmful solvents.
On the other hand, aromatic plants produce and secrete aromatic substances, and some of which exhibit therapeutic properties, such as antimicrobial and antioxidant activities, which are often used for culinary purposes, as well as in the food and liqueur industries [15,16,17,18,19,20,21,22,23,24]. Similar to medicinal plants, several culinary herbs and spices have been used as medicines over the centuries. For example, in Egypt, under the pharaohs, garlic was consumed by workers and slaves who built large pyramids “in order to increase their stamina and strength and protect them from disease” [25]. The use of spices can, to some extent, prevent food spoilage, especially in hot climates [26,27]. Today, it is possible to produce herbal-enriched dairy products, and functional dairy products with improved nutritional quality and medicinal values [28]. Laws and regulations on medicinal and aromatic plants in Europe, including the European Union, were summarized by Steinhoff [29] and Barbieri [30].
Over the last 20 years, we have experienced an unprecedented boom in nanotechnology associated with the application of nanosized particles/formulations in agriculture, food, and medicine. According to the U.S. National Nanotechnology Initiative [31], nanotechnology represents “the understanding and control of matter at dimensions of roughly 1 to 100 nanometers, where unique phenomena enable novel applications”. The European Commission in 2011 defined the term “nanomaterial” as “a natural, incidental or manufactured material containing particles, in an unbound state or as an aggregate or as an agglomerate and where, for 50% or more of the particles in the number size distribution, one or more external dimensions is in the size range 1–100 nm“ [32]. The uniqueness of nanomaterials is related to the fact that, despite the same chemical composition as the corresponding bulk materials, they may not only have different physical and chemical properties, but also different effects on living organisms [3,33,34,35,36,37,38,39,40,41,42].
Nanoparticles (NPs) can generally be prepared by mechanical division from top to bottom (physical method) or by “growth” from bottom to top (chemical method). Top-down methods, most often grinding using stabilizing agents, were the first to be used to prepare NPs. They usually lead to particles with diameters > 10 nm and relatively large variability, which is also reflected in problematic reproducibility. The bottom-up methods are much more elegant and, nowadays, a more widespread option for preparing NPs, which allows better control over the size of the NPs. A large number of different stabilizers (donor ligands, polymers, and detergents) are used to control the growth of NPs and to protect them from aggregation. Electrochemical reduction of metal salts and controlled decomposition of metastable organometallic compounds are often used to prepare metal-based NPs. In the initial state of nucleation, the metal salt is reduced to form metal atoms that collide with other ions, metal atoms, or clusters, and irreversibly form “nuclei” of a stable metal core. The diameter of the cores should be below 1 nm depending on the strength of the metal bonds, the magnitude of the redox potential of the metal salt, and on the reducing agent used. Nowadays, when green synthesis of NPs is preferred, it refers to various plant or microbial extracts containing numerous biologically active compounds for the reduction of metal ions, acting as capping and stabilizing agents (increasingly being used for the synthesis of metal-based NPs). These bioactive compounds that are bound to the surface of NPs significantly improve the interaction of NPs with living cells and improve their biological efficiency [43,44,45,46].
Controlled release of active substances, the need for lower doses to achieve the same effect, and lower toxicity to non-target organisms are the main advantages of nanomaterials compared to their bulk counterparts, which contributes to reducing environmental pollution and costs [3,37,47,48,49,50,51].
The impact of nanoparticles (NPs) on plants can be different and depends on the chemical composition of NPs, the applied concentration, environmental conditions (e.g., pH, temperature, soil acidity, etc.), as well as on the treated plant species [52,53,54,55,56,57]. NPs used as fertilizers stimulate plant growth and yield and can positively affect the nutritional quality of crops [50,52,54,55,56], while nanosized herbicides suppress the unwanted weeds [48,49,53,56,58,59,60] and NPs used in plant protection contribute to healthy development and growth of plants by killing harmful insects, fungi, and other pathogens [3,34,35,48,49,53]. NP-induced oxidative stress can be used to produce valuable secondary metabolites, not only in field culture, but also in vitro using tissue or hairy root (HR) culture [51,61,62,63,64,65,66,67]. Reducing the adverse effects of abiotic stresses on plants using NPs offers an effective way to achieve higher yields and crops [50,51,68,69,70,71].
The increasing use of NPs in industry and agriculture is associated with their gradual accumulation in environmental matrices, where they could migrate, persist, or be transformed, accumulate in algae and plants, and enter the food chain. As a result, some toxic NPs can adversely affect non-target organisms and pose a serious threat to human health. To eliminate these risks, environmental monitoring is required for NPs, providing research on their fate in environmental matrices and contents in edible parts of plants. This is especially important in terms of medicinal plants that are consumed for their medicinal properties [36,49,52,53,72,73,74,75].
This review article summarizes the current findings on the effects of metal-, carbon-, and silicon-based NPs as well as NPs derived from organic materials on medicinal and aromatic plants. Increased attention is paid to the production of important secondary metabolites exhibiting medicinal properties under in vitro conditions suitable for use in the mass production of these biologically active compounds, to be an alternative to synthetically produced ones. The beneficial effects of NP on medicinal plants exposed to some abiotic stresses are also discussed.

2. Effects of Metal-, Metalloid-, and Carbon-Based NPs on Vascular Plants

Metal-, carbon-, silicon-, and selenium-based compounds, applied in bulk or in nanoform, affect growth and development of vascular plants. In general, metal ions and metal NPs, including some essential metals (e.g., Cu, Zn, Fe. Ni, etc.) applied at higher concentrations, are phytotoxic, generate oxidative stress, adversely affect germination, photosynthetic processes, inhibit activities of enzymes and biosynthesis of photosynthetic pigments and non-enzymatic antioxidants, and accumulate high metal amounts in their plant organs, which is reflected in impaired plant growth and development [55,56,57,76,77,78,79,80,81,82,83]. On the other hand, some medicinal plants (e.g., Hypericum perforatum L. or Matricaria recutita L.) were found to be tolerant metal excess and accumulate in their shoots/leaves high concentrations of toxic metal, such as cadmium [83,84]. Therefore, to ensure food safety, cultivation of medicinal plants for commercial purposes under natural conditions must be performed on soils that are not contaminated with toxic metals [85].
Moreover, bulk and nanoscale metals and metal oxides can also be cytotoxic and genotoxic, cause cell-cycle arrest and apoptotic induction, as well as serious damage to DNA [57,86,87,88,89,90]. On the other hand, their low doses may stimulate plant growth and can be used as effective fertilizers for achieving higher yields [55,56,91,92,93,94,95,96]. Therefore, for individual plant species, it is necessary to known which concentrations have plant growth simulating effects and those with plant growth inhibiting impact.
Carbon-based nanomaterials, such as multi-walled carbon nanotubes (MWCNTs), can penetrate seed coats and, when used in low concentrations, improve water absorption, water transport, seed germination, nitrogenase, photosystem and antioxidant activities, activate water channel proteins, and stimulate nutritional absorption [97,98]. The entry of carbon-based nanomaterials into roots in an apoplastic way is thought to cause elongation of cell membrane pores [99] and the increased ability of plants to absorb water and essential nutrients when exposed to carbon nanotubes leads to stimulation of plant growth [98,100,101]. They can also affect the plant phenotype and composition of soil microbiota [102]. On the other hand, at higher concentrations, CNTs may have an adverse effect on plants due to the reactive oxygen species (ROS) generation, leading to cell death [103].
Silicon is a metalloid, which is beneficial for plant growth and development, particularly under abiotic stress conditions (e.g., high salinity, water deficit, high temperature, presence of toxic metals) because it can alleviate damages in stressed plants; it can activate some defense mechanisms and regulate some physiological processes related to defense mechanisms in plants [50,54,104,105]. Savvas and Ntatsi [106] stated that mitigation of abiotic stresses in vascular plants by Si is associated with SiO2 deposition inside the plant tissues. Se acts as antioxidant and protects plants from oxidative damages and, depending on applied concentration; it can have both beneficial and adverse impact on plants, and modulate negative effects of abiotic stresses on plants, e.g., [107,108].
NPs affecting medicinal and aromatic plants discussed in this review are presented in Figure 1. Beneficial and adverse effects of NPs on medicinal and aromatic plants are shown in Figure 2.

3. Improved Production of Secondary Metabolites in Presence of NPs

As mentioned above, NPs exhibit similar effects to their bulk counterparts; however due to their specific properties connected with their nanoscale sizes, their ultimate effects on plants can be amplified or attenuated [80,109,110,111,112,113,114]. By green production of valuable secondary metabolites of medicinal plants elicited by various NPs, using in vitro cultures (Table 1) or HRs (Table 2), higher amounts of these pharmacologically important compounds can be achieved exceeding those observed with corresponding bulk elicitors. These secondary metabolites can be subsequently utilized by the pharmaceutical industry [115,116,117,118,119,120].
Elicitors are compounds that are capable of causing stress to the plant, leading to increased synthesis and accumulation of secondary metabolites, or to the induction of new secondary metabolites [116,167]. In general, secondary metabolites of plants play a role in their communication and adaptation. Plant secondary metabolism utilizes building blocks and biosynthetic enzymes derived from primary metabolic processes, and it produces large number of specialized compounds that are essential for plant survival in its environment [168,169,170]. Many of the secondary metabolites are also biologically active compounds, which are therefore desirable to produce in large quantities directly from medicinal plants [115,117,119,171]. It has been found that many NPs can be used as effective elicitors, the use of which makes it possible to avoid various other, mostly harmful/ecotoxic chemicals [115,117,118,172]. Stress often induces the production of secondary metabolites in the plant tissue culture system [116]. The accumulation of valuable secondary metabolites in micropropagated medicinal plants induced by abiotic/biotic stress using NPs is studied intensively [113,115,117,172,173]. Hairy roots are also important for the large-scale production of industrially important secondary metabolites. These are differentiated cultures of the roots of higher plants that have been injured and subsequently infected with Agrobacterium rhizogenes. Such transformed roots are capable to grow rapidly on the basal medium, show genetic and biosynthetic stability, and can produce increased amounts of secondary metabolites [119,120,171,173,174,175,176,177].

4. Effect of Metal-Based NPs on Medicinal and Aromatic Plants

4.1. Silver-Based NPs

A comprehensive overview focused on phytotoxicity assessment of AgNPs and their interaction with plants, including whole plant and organs, as well as interactions on the cellular and molecular levels, was presented by Tkalec et al. [178]. Coating of AgNPs can affect their toxicity.

4.1.1. Impact on Plant Growth

Beneficial effect of AgNP seed germination characteristics of Thymus kotschyanus was reported by Khalaki et al. [179]. Moreover, treatment with 40 mg/kg AgNPs during the germination process effectively improved germination and growth of Ocimum basilicum L. plants and their resistance to salinity stress [180]. By application of 0.2 mg AgNPs (8–21 nm) per fenugreek seedling, the root and shoot length of seedlings was approximately doubled and wet weight of treated plants was even 3-fold higher compared to the control [115]. Increasing concentrations of AgNPs (27.5 ± 4.8 nm) up to 10 mg/L showed beneficial impact on growth and development of Lavandula angustifolia propagated in vitro, which was reflected in improved formation of shoots, increased plant weight, and longer roots compared to control. Increasing AgNPs concentrations reduced the number of secretory trichomes; the greatest diameter of trichomes on the adaxial surface of the leaf blade was observed at exposure to 2 mg/L AgNPs and the smallest one at 5 mg/L AgNPs. The largest trichome diameter on the abaxial surface was formed at treatment with 1 mg/L and the lowest one using 5 mg/L AgNPs [181].
Spraying of borago (Borago officinalis L.) plants at the seed growth stage with AgNPs using doses 20, 40, and 60 ppm resulted in improved seed yield compared to the control; this beneficial impact increased with increasing AgNPs concentration, although content of polyphenols in treated plants showed a decrease [182]. Spraying of shoots of borago plants at the onset of the flowering stage (65 days after cultivation) and maintained until flowering (98 days after cultivation) with solutions containing 0.2, 0.4, and 0.6 mM AgNPs resulted in considerable increase of leaf numbers, plant, and inflorescence dry weights, and petal abscission [183].
For vanilla (Vanilla planifolia) micropropagated in a semi-solid Murashige and Skoog (MS) medium treatment, 25 mg/L and 50 mg/L AgNPs with a mean diameter of 35 ± 15 nm was found to be the most favorable for shoot multiplication. The greatest lengths, number of shoots, and number of leaves were obtained with 25 mg/L and 50 mg/L AgNPs, and such treatment effectively reduced the contamination during micropropagation of V. planifolia [184].
Ag-SiO2 core-shell NPs (101.8 ± 8.9 nm) applied at 900 mg/L for 3 d to the HR cultures of Artemisia annua promoted their dry weight, which was 71% higher than that of the control [164]. AgNPs and AgNO3 did not affect the in vitro multiplication of Campomanesia rufa at doses 0.385 and 0.77 mg/L, respectively, while treatments with 15.4 mg/L showed toxic effects reflected in the reduced number of shoots [185].

4.1.2. Elicitor Effect

AgNPs also acted as elicitor in HR cultures of Brassica rapa, causing a pronounced increase in the contents of glucosinolates and their transcripts, as well as in total phenolic and flavonoid concentrations and transcripts, compared to non-elicited HRs [186]. After treatment of in vitro grown shoots of Isatis constricta with 2 mg/L AgNPs, lasting 5 days, the indigo and tryptanthrin content were higher by 15 and 71% than in the control; however, at prolonged treatment (10 or 15 days), their content showed a decrease. On the other hand, treatments with AgNPs were found to decrease indirubin production [123].
Foliar treatment of 40 mg/L AgNPs, besides stimulating growth of fenugreek (Trigonella foenum-graecum), plants increased the total content of photosynthetic pigments by 25%, while content of indole acetic acid was even doubled compared to the control. Moreover, seeds yielded from treated plants showed pronouncedly higher content of phenolics, flavonoids, and tannins, as well as a higher percentage of proteins and carbohydrates than those of control plants [187]. Diosgenin content in fenugreek seedlings treated with 0.2 mg AgNPs (8–21 nm) per seedling considerably increased compared to the control (214.06 ± 17.07 vs. 164.49 ± 7.67 μg/mL) [115]. Pronounced enhancement of secondary metabolites, including phenols, tannins, and alkaloids, was found in borago plants sprayed with solutions containing 0.2, 0.4, and 0.6 mM AgNPs [183].
Salvia officinalis plants foliarly treated with AgNPs contained enhanced levels of phenolic compounds; the content of rosmarinic acid (RA) and salvianolic acid A and B was positively correlated with the activity of phenylalanine ammonia-lyase (PAL) and RA synthase, but not with tyrosine aminotransferase. After treatment of plants with 100 mg/L AgNPs, even 8-fold higher RA content was observed than in control plants [128]. Rosmarinus officinalis L. plants sprayed with solution containing 25, 50, 100 ppm AgNPs showed beneficial impact on the contents of secondary metabolites. Application of 200 ppm AgNPs for 12 days resulted in enhanced carnosic acid (CA) content by <11% compared to the control. In AgNP-treated plants, positive correlation was observed between TFC and CA content. Translocation of AgNPs to rosemary roots was confirmed as well [188]. The composition of EO extracted from Lavandula angustifolia propagated on MS medium containing AgNPs (27.5 ± 4.8 nm; 10 and 50 mg/L) was considerably altered, showing a decrease in compounds of lower molecular weight, which were replaced by those with higher molecular weight [130].
AgNPs more effectively elicited production of cichoric acid in cell suspension cultures of Echinacea purpurea than AgNO3, which reached even 9.54 mg/g dry weight (d.w.) in cell suspension culture after 48 h of treatment with 2 mg/L AgNPs and 8 mg/g d.w. in leaf suspension culture after 72 h of treatment [122]. Moreover, enhanced production of secondary metabolites, TPC, TFC, and protein content, and improved antioxidant and superoxide dismutase (SOD) activities were observed in Artemisia absinthium grown in vitro on MS medium treated with AgNPs [131]. Due to oxidative stress generated by AgNPs, artemisinin content in HRs increased and its production in 20-day-old HR cultures achieved up to 13.3 mg/L, being 3.9-fold higher than that of the control [164]. By application of 45 mg/L AgNPs to callus culture of Stevia rebaudiana, the amount of stevioside was enhanced, up to 32.34 mg/g d.w. of callus, suggesting that by using appropriate concentrations of AgNPs as an elicitor enhanced production of this secondary metabolite can be obtained [126]. AgNPs effectively elicited production of both lignans and neolignans in cell suspension culture of Linum usitatissimum [125].
Cell suspension cultures (CSC) of Momordica charantia elicited with AgNPs showed pronouncedly higher levels of Ag, malondialdehyde (MDA), and H2O2 compared to non-elicited ones, and in their extracts, considerably enhanced total phenolic content (TPC) and total flavonoid content (TFC) compared to the control were observed. In CSC elicited by 5 mg/L AgNPs. the amounts of flavonols, hydroxybenzoic, and hydroxycinnamic acids greatly exceeded the amounts observed in the control [121].

4.1.3. Effect on Antioxidant Enzymes

Significantly enhanced callus induction of Citrus reticulata L., as well as high TPC, TFC, and antioxidant activity were observed using co-treatment with 20 ppm AgNPs and 1 mg/L thidiazuron; considerable increase of SOD, peroxidase (POD), and catalase (CAT) levels were recorded on MS medium supplemented with 30 ppm AgNPs and thidiazuron [189]. Hormetic effects of Argovit on in vitro regeneration of vanilla using a temporary immersion bioreactor system were reported by Spinoso-Castillo et al. [127]. Whereas doses 25 and 50 mg/L of Argovit increased Vanilla planifolia growth, application of 100 and 200 mg/L showed strong inhibitory effect; a dose-dependent effect on generation of ROS formation, and increasing TPC, antioxidant capacity, and lipid peroxidation was observed. AgNPs in combination with plant growth regulators added to MS medium pronouncedly enhanced callus biomass of Caralluma tuberculata. Stimulated production of phenolics, flavonoids, and increased PAL, SOD, POD, CAT, ascorbate peroxidase (APX) activities, as well as antioxidant activity were observed in the callus cultures treated with 90 µg/L AgNPs [124]. AgNPs green synthesized from leaf extracts of Swertia chirata applied on regenerating shoot cultures of S. chirata were found to generate ROS. However, due to induced antioxidant defense by antioxidant enzymes, ROS content was balanced and shoot regeneration increased, suggesting that AgNPs applied to plant tissue cultures can act as growth stimulant for improved shoot regeneration [190].
Calendula officinalis L. plants exposed to 50 ppm AgNPs and magnetic fields (3 mT), either alone or in combination, pronouncedly increased phenolic content, activities of several enzymes, such as PAL, polyphenol oxidase (PPO), ascorbate peroxidase (APX), guaiacol peroxidase (GPX), glutathione reductase (GR), and radical scavenging in plants, suggesting positive effect of AgNPs on the defense mechanism of C. officinalis plants, combined treatment being the most effective [191]. Treatment of Calendula officinalis with AgNPs also reduced its DPPH (2,2-diphenyl-1-picrylhydrazyl) scavenging activity and increased formation of ROS and membrane peroxidation [192]. In the extracts of Calendula officinalis plants cultivated in hydroponic solution containing 0.4 mM AgNPs (30–50 nm), the extracts contained lower contents of chlorophyll (Chl), carotenoids, and anthocyanins than the control. Only content of flavonoids showed an increase by 14.6%; however, at higher AgNP concentrations, their content decreased as well, and saponin content increased by ca. 77% [192].
It was reported that AgNPs pronouncedly suppressed oxidative stress triggered by naphthalene toxicity in Moringa oleifera called the “Miracle Tree” [193]. AgNPs applied foliarly to the Salvia officinalis plants, penetrated via the leaf epidermis into the parenchyma cells, and after application of 50 and 1000 mg/L AgNPs, reduced levels of assimilation pigments and increased activities of antioxidants (enzymatic and non-enzymatic) were able to cope with the oxidation stress, compared to the control plants [128]. The cetyltrimethylammonium bromide (CTAB) coated AgNPs were found to be toxic to A. cepa roots, causing robust inhibition of roots and pronounced oxidative damage, whereas much bigger negatively charged citrate coated AgNPs aggregating to larger particles were least toxic. However, it could be mentioned that the toxicity of Ag+ ions was higher than that of both coated AgNPs [194].

4.1.4. Attenuation of Abiotic Stress

AgNPs enhanced the tolerance to drought stress in germination and early growth stages of Thymus daenensis Celak and Thymus vulgaris L.; they improved germination characteristics and root length up to 200 mM salinity, although they exhibited positive impact on vigor and shoot length only up to 100 mM salinity as compared to the control [195].

4.1.5. Cytotoxic and Genotoxic Effects of AgNPs

AgNPs with particle sizes >100 nm were found to penetrate the root tip cells of Allium cepa and adversely affected stages of cell division resulting in formation chromatin bridge, stickiness, disturbed metaphase, multiple chromosomal breaks, and cell disintegration, whereby at application of 100 ppm AgNPs mitotic index (MI) decreased from 60.30% (control) to 27.62% [196]. Mitigation of genotoxic and cytotoxic effects of AgNPs in onion root tips reflected by a decrease of MI, elevated chromosome aberrations, and tail DNA caused by oxidative stress can be attenuated by using antioxidants (e.g., vitamin E), pretreatment being more effective than combined treatment with AgNPs [197]. Cytotoxic and genotoxic impact of Ag nanorods on A. cepa roots depended on their concentration, a dose of 15 µM being more toxic than doses of 5 and 10 µM, respectively [198]. Although polyvinylpyrrolidone (PVP)-coated as well as uncoated AgNPs were found to be genotoxic, cytotoxic, and induce oxidative stress in A. cepa roots, even at low concentrations; under exposure to visible-light, their aggregation state will be altered, resulting in reduced toxicity [199]. Due to application of 25 and 50 mg/L AgNPs to V. planifolia, only a small MI decrease was observed and genotoxic effects were minimal, with < 5% of the total chromatic aberrations and 3 micronuclei in 3000 cells, despite the long-time exposure to AgNPs [200].
In general, considering the beneficial impact of AgNPs on enhanced production of secondary metabolites, they could be recommended to be used, particularly for production of stevioside in callus culture of Stevia rebaudiana L. [126] and for enhanced production of rosmarinic acid by foliar spraying of Salvia officinalis L. plants [128] (Table 1).

4.2. Copper-Based NPs

4.2.1. Impact on Plant Growth and Photosynthesis

Capsicum annuum plants grown in soil amended with 500 mg/kg CuNPs, which were evaluated at the vegetative stage, i.e., 45 days post transplantation, showed pronouncedly increased length and dry weight of roots, while plant tissues accumulated lower Cu amounts compared to treatment with bulk Cu particles. On the other hand, when treated plants were evaluated in the reproductive stage, i.e., 90 days post transplantation corresponding to a full life cycle, treatment with 500 mg/kg CuNPs resulted in increased photosynthesis, stomatal conductance, and Cu leaf content compared to that observed with application of bulk Cu particles. Treatment with 62.5 mg/kg CuNPs was found to increase transpiration, but total sugar, carotenoid, Chl, and vitamin C contents in the fruit, and the root and leaf enzyme activity, were not affected 90 days post transplantation in plants exposed to 62.5, 125, or 500 mg/kg CuNPs. The results of this experiment suggested that a dose of 500 mg /kg CuNPs, ensuring higher Cu bioaccumulation in the root, and leaf tissue over long-term exposure, is suitable to be used as a nanofertilizer [201]. CuO NPs phytosynthesized using Punica granatum peel extract foliarly sprayed on Capsicum annuum plants increased the height of plants, as well as growth and the Chl content of leaves; the fruit practically did not contain residues of sprayed NPs [202]. Application of 125 mg/kg CuO NPs to Capsicum annuum plants significantly reduced leaf P content by 41%, while exposure to 500 mg/kg CuO NPs decreased Zn content in leaves and fruits by 55% and 47%, respectively, compared to the control, thereby affecting the nutritional quality of bell pepper [203].
Foliar spraying of Mentha piperita L plants performed three times of the interval of 15 days up to flowering stages with 1.0 g/L CuNPs was found to enhance Chl content and EO percentage of 35% and 20%, respectively, as compared to control [133]. Investigation of the clonal microreproduction of Mentha longifolia plant showed that addition of colloidal solutions of CuNPs (0.5 mg/L) and CoNPs (0.8 mg/L) to MS nutrient medium resulted in an increase of microplant height and growth index on 45–48.4%, the quantity of internodes on 29.4–33.9%, quantity of shoots on 55.6–66.2%, and reproduction coefficient on 30–40% [204]. Application of 5 µM CuNPs (20–40 nm) for plant regeneration of Ocimum basilicum via somatic embryogenesis resulted in 5.6-fold higher percentage of explants produced somatic embryos (84 vs. 15%) and 4.35-fold higher average number of regenerated plantlets/explant (18.7 vs. 4.3) as compared to treatment with 0.1 µM CuSO4 5 H2O [205]. Exposure of Origanum vulgare plants for 60 days to CuNPs at doses 0–200 mg Cu/kg increased root and shoot growth, and water content, but reduced shoot biomass, as well as starch, total sugar, and reducing sugar in leaves; whereas shoot length, MDA, and Chl levels were not affected. However, CuNPs affected root and shoot contents of Ca, Fe, Mg, and Mn, suggesting that nutrition value of oregano could be modified [206]. Phytotoxic effect of 200, 400, and 800 mg/L CuNPs on hydroponically cultivated coriander (Coriandrum sativum) plants was reflected in reduced biomass and root length, lower levels of Chls (Chla and Chlb) compared to control plants, and increased levels of H2O2 compared to untreated plants [207].

4.2.2. Elicitor Effect

Foliar application of 0.5 g/L CuNPs effectively enhanced the levels of camphene, α-phellandrene and limonene in Mentha piperita L. plants, while a dose of 1.0 g/L increased the concentrations of menthol, menthone, and menthofuran [133]. Allium fistulosum plants grown 80 days at greenhouse conditions in soil spiked with 150 mg/kg CuO NPs showed higher contents of Cu, Ca, and Fe in roots, and higher Ca and Mg content in bulb than plants exposed to bulk CuO or CuSO4. Moreover, plants of Chinese scallion plants cultivated in soil supplemented with 75–600 mg/kg CuO NPs showed higher levels of allicin in leaves (56–187%) than the control, suggesting that CuO NPs have potential to be used as nanofertilizer for onion production [208].
Crystalline monoclinic cubic CuO NPs with mean size 47 nm applied to MS media for direct organogenesis of S. rebaudiana from nodal segments at a dose of 10 mg/L pronouncedly increased shoot organogenesis (88.5%), shoot length, mean number of shoots per explant, and fresh weight, as well as the content of glycosides, rebaudioside A, and stevioside [209]. The NPs of bimetallic alloys of Cu and Au (Cu:Au = 3:1) applied at a dose of 30 µg/L together with 0.5 mg/L naphthalene acetic acid were found to stimulate biomass accumulation on the 27th day of log phases, increased the radical scavenging activity, and enhanced TPC and TFC in submerge adventitious root cultures of S. rebaudiana (Bert.) [210]. Uncapped and polyethylene glycol (PEG) or PVP capped CuO NPs applied to the growth medium of in vitro cultured S. rebaudiana plants were able to increase the production of commercially important secondary metabolites. Regarding shoot organogenesis, increased growth parameters and higher content of steviol glycosides and non-enzymatic antioxidants (TPC, TFC) in shoots were observed at treatment with polymer-capped CuO NPs compared to uncapped ones [211].
Using cell suspension cultures of Gymnema sylvestre (Retz.) R. Br and CuO NPs as elicitor (3 mg/L for 48 h), the yields of gymnemic acid II, TPC, and TFC were considerably enhanced, and the cultures exhibited significant antioxidant, antidiabetic, anti-inflammatory, and anticancer activities [136]. Bacopa monnieri (L.) Pennell herb treated in vivo with Cu-based NPs (CuNP/CuO NPs) applied at doses 0–100 mg/L showed biphasic dependence of total contents of saponins, alkaloids, phenolics, and flavonoids on the concentration of used NPs. While at lower Cu/CuO NPs concentrations, the content of secondary metabolites increased; at higher concentrations, their strong reduction was observed. Besides enhancement of secondary metabolism, the CuNPs applied at sub-toxic doses also improved the antioxidant capacity in Bacopa monnieri herb via ROS-mediated defense response [212].

4.2.3. Genotoxicity

CdS and CuO NPs induced growth inhibition and caused cytological aberrations in both mitotic and meiotic cells in Coriandrum sativum L, the effect being like that of conventional mutagen, ethyl methanesulfonate, which was used as control [213]. Damaged root plasma membrane and altered genome of hydroponically cultivated coriander plants exposed to higher CuNP concentrations suggested that genotoxic impact of CuNPs was observed as well [207].
Considering the beneficial impact CuO NPs on the production of gymnemic acid II in Gymnema sylvestre (Retz.) R. Br cell suspension, which is 9.0-fold higher than that of the control [136] (Table 1), this method can be recommended for recovery of this anti-sweet compound.

4.3. Iron-Based NPs

4.3.1. Impact on Plant Growth and Assimilation Pigments

Chitosan-coated Fe3O4 NPs with sizes 3–22 nm applied at doses 200 and 400 mg/kg showed beneficial impacts on seed germination and growth of Capsicum annuum L. seedlings, whereby their efficiency was better than that of bare Fe3O4 NPs [214]. Priming of roots of 28-day-old chilli and marigold plants for 3 h in the suspension of FeS2 NPs (100 µg/mL) prior to transplantation had beneficial impact on chilli yields and mean flower number of marigold, which was approximately doubled as compared to control (27 vs. 13) [215]. Introduction of 0.3 mg/L FeNPs (27.0 ± 0.51 nm) instead of FeSO4 into a MS nutrient medium increased root length, root activity (evaluated as the reduction of triphenyltetrazolium), and leaf Chl in Capsicum annuum plants by 118, 58 and 5% related to control. A 10-fold higher FeNPs concentration increased Chl content even by 27% and beneficial impact on the above-mentioned characteristics were observed also at a combination of FeNPs with CuNPs or ZnNPs. It could be mentioned that the applied FeNPs concentration of 0.3 mg/L was 18.7-fold lower than that contained in standard MS medium [216].
Treatment with Fe2O3 NPs and a chelated Fe-EDTA fertilizer increased Fe content of Arachis hypogaea plants cultivated in pots compared to control; Fe2O3 NPs increased root length, plant height, and biomass of peanut plants. The Fe2O3 NPs promoted the plant growth via regulating contents of phytohormones and antioxidant enzyme activities. Adsorption of Fe2O3 NPs onto sandy soil improved availability of Fe to the plants suggesting that Fe2O3 NPs can be used as fertilizer instead of traditional Fe fertilizers [91]. Zingiber officinale Rosc. plants cultivated hydroponically in Hoagland solution containing 100 ppm Fe2O3 NPs exhibited higher protein levels (1.699 μg/mL) compared to plants treated with Fe-EDTA (1.108 μg/mL) and plants cultivated without Fe3+ addition to nutrient solution (0.208 μg/mL). Similarly, in plants exposed to Fe2O3 NPs, the total Chl and carotenoid content was 1.18- and 1.28-fold higher than in plants treated with Fe-EDTA, and 2.23- and 2.25-fold higher than in control plants. Fe content in rhizome of Fe2O3 NPs-treated plants increased as well, suggesting that these NPs can be used to mitigate chlorosis [217].
Comparison of yield, growth, and Fe uptake of Cuminum cyminum L. plants grown on the calcareous soils and sprayed twice during the season, before and after flowering (with a 10 day interval), using three various Fe-containing fertilizers, showed that the highest weight of 1000 grains and plant growth was observed at application of Fe-nano-chelate using a dose 1 g/L, while treatment with Fe-siderophore was found to be the most suitable for seed fortification with Fe [218].

4.3.2. Elicitor Effect

Treatment of cell suspension cultures of Dracocephalum polychaetum Bornm with 100 ppm Fe2O3 NPs, static magnetic field (SMF) of 30 mT, as well as combined treatment with Fe2O3 NPs and SMF pronouncedly increased intracellular content of secondary metabolites, such as RA, naringin, apigenin, thymol, carvacrol, quercetin, and rutin; secretion of these secondary metabolites from cells exposed to the double stress (SMF and NPs) to the culture media was observed as well. Moreover, the applied treatments resulted in increased TPC, TFC, anthocyanins, lignin, and MDA contents and increased activity of PPO and PAL [143]. Exposures of HRs derived from 4-week-old leaves of Dracocephalum kotschyi Boiss and inoculated with Agrobacterium rhizogenes strain ATCC15834 to 75 mg/L Fe2O3 NPs for 24 h resulted in improved biomass accumulation and activities of antioxidant enzymes, enhanced TPC and TFC, and contents of some valuable secondary metabolites, such as RA and xanthomicrol, which reached 9.7- and 11.87-fold higher values than in the control [162].
Moreover, inoculation with Agrobacterium rhizogenes of HRs derived from cotyledon explants of Hyoscyamus reticulatus L, which were elicited with 900 and 450 mg/L FeNPs using exposure time of 24 and 48 h, respectively, achieved approximately fivefold higher content of tropane alkaloids, hyoscyamine, and scopolamine, and considerably higher activities of antioxidant enzymes compared to control [64]. Fe2O3 NPs acted as highly effective elicitor, increasing the growth and production of secondary metabolites of Cichorium intybus L. in HR culture induced by Agrobacterium rhizogenes strain 15834 [219]. Treatment of S. rebaudiana cultivated in vitro with 45 µg/L FeNPs resulted in beneficial impact on plant growth characteristics and production of stevioside, and rebaudioside A reached 4.2 ± 0.058 and 4.9 ± 0.068 mg/g (d.w.). At exposure to 135 µg/L FeNPs, the levels of TPC and TFC reached 3.2 ± 0.042 and 1.6 ± 0.022 mg/g (d.w.), respectively, although adverse effect on growth characteristics was observed. Exposure to 90 µg/L FeNPs resulted in increased activities of SOD, POD, CAT, and APX [220].
In cell suspension culture of Hypericum perforatum L. elicited with 100 ppb Fe2O3 NPs for 72 h, enhanced concentrations of hypericin (195.62 ± 10.00 vs. 16.272 ± 4.62 μg/g (d.w.)) and hyperforin (11.181 ± 0.17 vs. 2.072 ± 0.77 μg/g (d.w.)) compared to the control were observed, whereas the dry weight was comparable with control. It could be assumed that the Fe2O3 NPs may induce jasmonate production, playing a role in hypericin and hyperforin production, or can be involved in signal transduction process regulating jasmonate production genes in cells [141]. Simultaneous foliar application of 0.01% MgO NPs and 0.03% Fe2O3 NPs on Hibiscus sabdariffa plants exhibited synergistic and beneficial impact on the activity of antioxidant enzymes and anthocyanin and flavonoids levels, and reduced H2O2 concentration in plants. On the other hand, at combined application of 0.03% MgO NPs and 0.03% Fe2O3, ROS production increased, and the antioxidant defense system was not able to attenuate toxic effects of produced H2O2 and, therefore, the observed characteristics were comparable to those of control plants [221].
Foliar treatment of Ocimum basilicum L plants grown in soil with Fe nanocomplexes Fe(His)3 and Fe(Arg)3 resulted in an increase of sesquiterpenes and a decrease of the content of oxygenated monoterpenes in the composition of secondary metabolites of EO, and enhanced antioxidant and antimicrobial activities of plants [222]. FeNPs applied foliarly to Mentha piperita L. at a dose of 1.5 g/L, three times of the interval of 15 days up to flowering stages, considerably enhanced contents of Chl and EO. Application of 0.5 g/L NPs increased the EO content and oil production by 60 and 50%, and menthone and menthol contents by 65% and 30%, respectively, compared to the control. On the other hand, at a threefold higher dose of FeNPs, menthofuran concentration was 2.5-fold higher than in the control [133].
Spraying of basil plants with 0.2% Fe-urea nanocomplex increased the content of epi-α-cadinol and trans-α-bergamotene to 27.09 ± 2.5% and 14.93 ± 1.77%, respectively, although n-decane content was reduced strongly even by 99.1% and the concentration of RA achieved 5.81 ± 0.18 mg/g [223].

4.3.3. Attenuation of Abiotic Stress

Spraying of salt-stressed (100 mM of NaC1) Dracocephalum moldavica L. plants with Fe2O3 NPs considerably increased the leaf area, content of secondary metabolites (TPC, TFC, anthocyanins), as well as activities of antioxidant enzymes (GPX, APX, CAT, GR), suggesting improved antioxidant defense. Moreover, under salinity stress, the Fe2O3 NPs can serve as a source of Fe for this medicinal plant [224]. Application of Fe3O4 NPs was also able to greatly attenuate Cd and Pb induced toxicity in coriander plants grown in soil artificially contaminated with these toxic metals [225]. Treatment of Mentha piperita plants under salt stress with 10–30 µM Fe2O3NPs improved the fresh and dry biomass of leaves, contents of essential elements such as P, K, Fe, Zn, and Ca, although did not affect Na content. Moreover, Fe2O3 NPs pronouncedly reduced lipid peroxidation and proline contents and decreased activities of CAT, SOD and GPX in salt-stressed plants [226]. In Melissa officinalis L. plants exposed to drought stress (80%, 60%, and 40% field capacity (FC)) considerable increase in MDA, H2O2, as well as proline content and electrolyte leakage was observed with decreasing FC, while TPC, activities of antioxidant enzymes, Chl content and relative water content were reduced. Although increasing of water deficit from 80% to 60% FC resulted in improved yield of EO, further reduction to 40% already showed an opposite effect. However, application of FeNPs (5–20 µM) had beneficial impact on drought-stressed plants, which was reflected in further increase of EO yield, increased proline levels and mitigation of oxidative stress by activation of antioxidant defense system of plants and, thus, FeNPs protected the plants against oxidative stress damages [142].

4.3.4. Genotoxicity

Ghosh et al. [227] investigated the impact of FeNPs with different surface chemistries on the uptake, root morphology, DNA damage, oxidative stress, and cell death in Allium cepa roots after 24 h exposure. It was shown that FeNPs with low zeta potential, conductivity, and high polydispersity index, which were adsorbed on root surfaces, damaged the root tip, epidermal, and root hairs due to their colloidal destabilization and smaller size. High uptake of NPs was responsible for considerable DNA damage, chromosome/nuclear aberrations, and micronuclei formation. On the other hand, FeNPs, which were characterized with higher dissolution and substantial uptake, induced genotoxicity. Both types of FeNPs generated strong oxidation stress and in treated A. cepa roots ROS-mediated apoptotic and necrotic cell death was observed.
Elicitation of the production of hyoscyamine and scopolamine in HR of Hyoscyamus reticulatus L. by Fe2O3 NPs allows achieving ca. 5-fold higher amounts than in the control. This is a promising method that can be considered to produce tropane alkaloids [65] (Table 2). Moreover, 5.4- and 12-fold higher amounts of hypericin and hyperforin, compared to the control observed in cell suspension culture of Hypericum perforatum L., using 100 mg/L Fe2O3 NPs as an elicitor, suggest this method is very encouraging [141] (Table 1).

4.4. Zinc-Based NPs

4.4.1. Impact on Germination and Plant Growth

ZnO nanorods with a mean aspect ratio of 7, coated in gel and mixed with the seeds of Solanum melongena L. and Capsicum annum L., applied at doses 50, 100, and 150 mg/L, accelerated germination, which was reflected in reduced germination time and improved mean germination rate compared to control plants; application of 100 mg/L ZnO NPs ensured maximal transplant length as well as fresh and dry weight [228]. Amelioration of the seed germination rate and seedling vigor of Capsicum annuum L. was observed at treatment with ZnO NPs suspensions at 100, 200, and 500 ppm, respectively. On the other hand, application of these ZnO NPs concentrations resulted in the inhibition of radical growth and stimulated accumulation of phenolic compounds [229]. Spraying of Capsicum chinense Jacq. plants during the main stages of phenological development with 1000 ZnO NPs stimulated plant height, stem diameter, and Chl content, and increased fruit yield and biomass accumulation in treated plants compared to control and treatment with ZnSO4, whereas the double dose of ZnO NPs adversely affected plant growth [230].
Germination of seeds of Abelmoschus esculentus (L.) Moench was positively affected by treatment with ZnNPs (50–500 ppm), and in 30-day-old treated plants, an increase in Chl, protein, proline, and carbohydrate contents were observed, along with pronounced reduction of the carbohydrate content of leaves [231]. Spraying of leaves of 10-day old chickpea (Cicer arietinum L. var. HC-1) seedlings, with solution containing 1.5 ppm ZnO NPs, stimulated shoot dry weight, but when a dose of 10 ppm was applied, root inhibition was observed. Positive impact of ZnO NPs on plant biomass related to reduced ROS generation caused less lipid peroxidation, which also resulted in lower activity of antioxidant enzymes, such as SOD and POD, compared to the control [232]. Phytotoxic effects of the ZnO NPs and Zn microparticles (MPs) (10–2000 mg/L) on Fagopyrum esculentum was reflected in biomass reduction, increased Zn bioaccumulation in plants, and higher translocation factor of Zn being observed for treatment with ZnO NPs. Moreover, reduced glutathione level and CAT estimated at application of ZnO NPs indicated oxidative stress due to ROS generation [233]. In 6-week-old callus produced from leaf explants of S. rebaudiana Bertoni grown in MS medium supplemented with plant growth regulators addition of 0.1 mg/L ZnO NPs significantly enhanced fresh and dry weight of callus [117].
ZnO NPs (40 nm) positively affected growth of the Echinacea leaf callus and showed beneficial impact on the anticancer activity of plant extracts, in contrast to the control treatment with ZnO MPs. Moreover, application of ZnO NPs increased flavonoid contents of Echinacea purpurea [234]. A dose of 2000 ppm ZnO NPs increased dry weight of Tanacetum parthenium (L.) Schultz Bip plants from 28.09 g/day (control) to 32.54 g/day, while using 1000 ppm ZnO NPs, the EO content achieved 0.9% v/w compared to 0.56% v/w observed in the control plants; an increase in Zn absorption at application of ZnO NPs was observed as well, while Fe absorption was reduced [235].
ZnO NPs at doses 0.5–2.5 mg/L applied to hormone-free MS medium inhibited root growth of ginseng root culture, deceased the lateral root number, and decreased percentage of lateral root formation was observed at exposure to 2.5 mg/L ZnO NPs. In treated plants thinner vascular tissue or cortex area and more black points in root cap area were observed compared to control. On the other hand, exposure to ZnO NPs resulted in thicker endodermis [236].
Foliar application of ZnO NPs at doses 20, 60, and 100 ppm to Momordica charantia L. plants grown in pots did not considerably affect shoot weight and anthocyanin content, but pronouncedly increased Chla content, as well as contents of secondary metabolites (phenols, flavonoids, and carotenoids), along with carbohydrate and proline content in a dose-dependent manner [132]. Growth characteristics in the callus of Punica granatum cv. Hegazy were adversely affected by 200 µg/mL ZnO NPs reflected in a stronger decrease in callus dry weight than at application of bulk ZnO (57.53% vs. 37.78%) and more elevated increment in Zn concentration over the control compared with that of ZnO bulk (335.24% vs. 247%) [237].

4.4.2. Elicitor Effect

Improved production of tropane alkaloids was observed using HRs obtained from two-week-old cotyledon explants of Hyoscyamus reticulatus treated with 100 mg/L of ZnO NPs after 48 and 72 h, when the levels of hyoscyamine and scopolamine reached 37.63% and 37%, respectively. The highest expression of h6h gene in HRs was observed at treatment with 100 mg/L of ZnO NPs after 24 h [67]. Zn-Ag NPs (molar ratio of 19:1) exhibiting photocatalytic and elicitor activity were able to improve the light absorption and increase oxidative stress in Withania somnifera plants, resulting in increased photosynthesis and transpiration rates along with an increase in some carbohydrates and enhanced content of withanolide, which was associated with upregulation of genes involved in photosynthesis, Calvin cycle, carbohydrate metabolism, and withanolide biosynthesis. Moreover, cyclic electron flux managing the electron flow around photosystem (PS) I was observed [63].
Spraying of Capsicum chinense Jacq. plants with 1000 ZnO NPs pronouncedly increased TPC, TFC, and content of secondary metabolites, such as dihydrocapsaicin in fruits, suggesting ameliorated nutraceutical quality of fruits [230]. Hypericum perforatum cell suspension cultures elicited with 100 ppm ZnO NPs stimulated production of hypericin and hyperforin, up to 7.87 and 217.45 μg/g d.w., providing 3 and 13-fold higher amounts compared to the control [141].
ZnO NPs capped with PEG and PVP enhanced TPC, TFC, content of steviol glycosides in in shoots of S. rebaudiana Bertoni cultured in vitro more than upcapped ones [211]. By application of 2 mg/L ZnO NPs, 91% induction of in vitro roots formation in regenerants of S. rebaudiana, and increased contents of secondary metabolites of rebaudioside A (2.96% vs. 4.42%) and stevioside (1.01% vs. 1.28%) in leaves, compared to the control, were observed. With application of increasing ZnO NPs, the content of secondary metabolites gradually decreased to 1.22% and 0.21% for rebaudioside A and stevioside, respectively, corresponding to the values observed under stress induced by 2000 mg/L ZnO NPs [135]. In contrast to the beneficial impact of 1 mg/L ZnO NPs on shoot formation of micropropagated S. rebaudiana and steviol glycosides levels, which were practically doubled compared to control, further increase of ZnO NPs was accompanied with a decrease in formation of secondary metabolites and antioxidant activities, and at application of 1000 mg/L ZnO NPs, a strong phytotoxic effect was observed [137]. Toxic impact of ZnNPs (<100 nm) applied at concentrations 400–1000 mg/L accompanied with pronounced reduction of stevioside was also observed in the nodal explants of Stevia cultured in vitro on MS medium [238]. Fertilization of Stevia rebaudiana with 75 mg/L of green synthesized ZnO NPs through roots enhanced the concentrations of TPC and TFC and Zn content up to 406.8% with respect to the control, although the biosynthetic pathway of steviol glycosides was not affected; treatment did not show adverse impact on plant growth [139].
Seedlings of flax (Linum usitatissimum) cultivated on MS in the presence of 10 mg/L ZnO NPs showed improved growth compared to control, while application of 500 mg/L ZnO NPs increased the production of secoisolariciresinol diglucoside, lariciresinol diglucoside, dehydrodiconiferyl alcohol glucoside and guaiacylglycerol-β-coniferyl alcohol ether glucoside. In another experiment, in which the stem explants were cultured on MS media supplemented with plant growth regulator and 25 mg/L ZnO NPs, enhanced rooting effect and effective accumulation of metabolites was observed [138].
In Thymus kotschyanus and T. daenensis micropropagated in the presence of 150 mg/L ZnO NPs thymol and carvacrol contents of 22.8 mg/L and 0.68 mg/L, respectively, were reported by [172]. Treatment of Glycyrrhiza glabra L. seedlings with 1 and 10 µM ZnO NPs increased the contents of glycyrrhizin, TPC, TFC, anthocyanins and proline, whereas by application of bulk ZnO at a dose of 10 µM content of tannins was enhanced [239].

4.4.3. Impact on Antioxidant Enzymes

Growth of fresh and dry weight of HRs treated with 100 mg/L ZnO NPs showed a decrease compared to the control, whereas activities of CAT, GPX, and APX were considerably higher than in the control [67]. Significant increase of CAT activity was observed at application 60 and 100 ppm of ZnO NPs to Momordica charantia L. plants grown in pots, but the activities of GPX and AXP were not affected [132].
In cell suspension cultures of Linum usitatissimum L. exposed to ZnO NPs, the highest PAL activity was observed at treatment with 30 mg/L for 48 h, whereas at application of 60 mg/L ZnO NPs, the highest activity of cinnamyl alcohol dehydrogenase was achieved. On the other hand, using both above-mentioned ZnO NP concentrations, the highest TPC and lignan amounts could be obtained [140]. In seedlings of flax (Linum usitatissimum) cultivated on MS medium and in stem-derived callus, the presence of ZnO NPs (1–1000 mg/L and 10–50 mg/L, respectively) caused the dose-dependent ROS generation. A differential induction of SOD (in seedlings) vs. POD (in vitro cultures) enzymes was observed. At concentrations > 500 mg/L (seedlings) and >25 mg/L ZnONPs (callus) oxidative damage was observed [138].
Spherical green synthesized ZnO NPs with mean size of 41 nm had positive impact on root and shoot length of Borago officinalis plants, but did not affect significantly soluble sugar and anthocyanin levels, and CAT activity in the plants. Due to enhanced ROS generation by ZnO NPs, the contents of proline and Chl decreased and oxidative stress in plants was accompanied with increased H2O2 content, lipid peroxidation, phenol content, and APX, GPX, PAL activities [240].
ZnO NPs capped with PEG and PVP enhanced total antioxidant capacity, total reducing power and DPPH free radical scavenging activity in shoots of S. rebaudiana Bertoni cultured in vitro more than uncapped ZnO NPs, concentration 1 mg/L being highly effective [211]. Addition of 100 mg/L ZnO NPs resulted in effective increase in TPC, TFC, total antioxidant capacity and DPPH free radical scavenging activity in 6-week-old callus produced from leaf explants of S. rebaudiana Bertoni grown in MS medium; the highest total reducing power was observed at application 50 mg/L ZnO NPs [117].

4.4.4. Cytotoxic and Genotoxic Effects

Cytotoxic and genotoxic impact of ZnO NPs (5 and 50 µg/mL) causing oxidative stress in roots of Allium cepa was reflected in lack of membrane integrity, reduced metabolic activity, ROS generation, chromosome aberration, DNA damage, cell-cycle arrest, and cell death in onion roots; considerably higher Zn levels were estimated in the cytoplasmic and nuclear fractions of A. cepa roots. It was confirmed that, for the toxicity of ZnO NPs suspension beside Zn2+ ions, predominantly ZnO NPs were responsible [241]. Similar effects were observed by Ghosh et al. [242] who treated A. cepa roots with ZnO NPs having a size of 85 nm, which deregulated the components of ROS antioxidant machinery causing DNA strand breaks and cell-cycle arrest at the G2/M checkpoint. Mitochondrial and chromosomal damage induced by oxidative stress in A. cepa roots treated with Zn2+ ions, bulk ZnO, and ZnO NPs (25 nm) was also reported by Ahmed et al. [243]. ZnO NPs attachment and internalization, and a considerable increase in levels of thiobarbituric acid reactive substances and antioxidant enzymes in treated roots, were observed as well. Summarizing the adverse effects of ZnO NPs on onion roots, they can be considered as a clastogenic/genotoxic and cytotoxic agent [244]. Treatment of A. cepa roots with 100 mg/L ZnO NPs reduced MI compared to the control (53.2% vs. 63.6%), which further decreased to 50% under UV irradiation. At application of 0.1 mg/L with ZnO NPs lagged metaphase and anaphase with multiple chromatin bridges in the root tip cells were recorded, although at doses 10 mg/L and 100 mg/L, respectively, both stickiness and diagonal anaphase were predominant. It was observed that, following exposure to ZnO NPs, the generation of ROS and MDA was accelerated [245].
Elicitation of the production of hyoscyamine and scopolamine in HR of Hyoscyamus reticulatus L. by ZnO NPs allows achieving 4.61- and 3.20-fold higher amounts than in the control. This is a promising method that can be considered to produce tropane alkaloids [67] (Table 2). Moreover, 3.80- and 13.36-fold higher amounts of hypericin and hyperforin compared to the control observed in cell suspension culture of Hypericum perforatum L. using 100 mg/L ZnO NPs as an elicitor suggest this method as very encouraging [141] (Table 1).

4.5. TiO2-Based NPs

TiO2 NPs may have a positive effect on photosynthetic efficiency by increasing electron transport between the PSII and PSI, positively affecting Rubisco activity in the Calvin and Benson cycle [246].

4.5.1. Impact on Germination, Plant Growth and Assimilation Pigments

TiO2 NPs (10–25 nm) applied at doses 20–40 mg/L were found to accelerate the germination characteristics of Alyssum homolocarpum, Carum copticum, and Nigella sativa seeds, and increase their vigor, while most effective stimulation of germination of Salvia mirzayanii seeds was obtained with 80 mg/L TiO2 NPs [247]. TiO2 NPs (43 nm) coated in gel, which were mixed with the seeds of pepper using doses 50, 100, and 150 mg/L, accelerated germination and enhanced lengths, fresh and dry weights of seedlings compared to control, concentration of 100 mg/L TiO2 NPs being the most effective [228]. Nanopriming of seeds with TiO2 NPs (anatase) (10–40 mg/mL) considerably enhanced germination as well as morphological characteristics and Chl content of Petroselinum crispum seedlings grown in vitro in MS medium, the dose of 30 mg/L being the most effective [248]. Exposure to 40 ppm TiO2 NPs improved fennel seed germination time by 31.8% in comparison to the untreated control, while treatment with 40 ppm TiO2 NPs improved dry weight compared to the untreated control [249]. On the other hand, Khater and Osman [250] reported that spraying of fennel plant with 6 ppm TiO2 NPs enhanced Chl synthesis, which was reflected in improved photosynthesis.
Foliar treatment of Mentha piperita plants with TiO2 NPs using a dose of 150 mg/L increased fresh and dry weights by 48% and 62.6%, and nitrate reductase and carbonic anhydrase activities 150 days after planting by 17.7% and 19.1%, respectively, compared to control plants; an increase of nitrogen content, stomatal conductance, and net photosynthetic rate by 12.6%, 8.5% and 23.8%, respectively, compared to the control plants, was observed as well. Diameter and density of peltate glandular trichomes, which are important from the aspect of EO production, were increase by 77.8% and 62.5%, respectively, at treatment with 100 mg/mL TiO2 NPs [144]. Mentha piperita L. and Salvia officinalis L. plants exposed to TiO2 NPs either via the leaves (using TiO2 NPs suspension) or via the root system (in soil) resulted in reduced Chl contents in M. piperita and S. officinalis plants treated through root system, but not after application of NPs suspension on leaves of sage plants, likely due to the abundance of trichomes, providing a natural barrier for the NP accumulation to the symplast. In the soil-exposed plants translocation of uptaken TiO2 NPs from roots to shoots resulted in reduced Chl levels. The reduction of Chl correlated mainly with the loss of Mg (sage) and Mn (peppermint) in treated plants [251].
Foliar application of 0.04% TiO2 NPs improved the contents of total Chl, carotenoids, and soluble proteins, as well as grain yield of safflower plants to greater extent than application of 0.02% TiO2 NPs and bulk TiO2 [252].
In contrast to smaller TiO2 NPs of 50 nm, larger TiO2 NPs of size 68 nm had beneficial effect on germination of Abelmoschus esculentus seeds. However, the small-sized TiO2 NPs applied at a dose of 200 mg/kg pronouncedly increased Chla and total Chl content compared to the control; a dose of 800 mg/kg considerably increased fresh weight of plants compared to the control. On the other hand, both types of NPs reduced APX in roots and leaves, whereas GR activity in roots decreased, but increased in the leaves. A considerable raise of SOD activity in roots and leaves of plants exposed to TiO2 NPs signalized their phytotoxic impact. Moreover, treatment with TiO2 NPs caused a size-dependent and dose-dependent decrease in essential minerals Ca, Mg, and Fe contents in fruits, as well as adversely altered the proximate compositions (except carbohydrate) of fruits [253].
Treatment with TiO2 NPs at 50−200 mg/L (83.7 nm) moderately increased the root and shoot fresh biomass of Coriandrum sativum L. plants cultivated in hydroponic solution. At application of 100–400 mg/L TiO2 NPs, enhanced activities of SOD, CAT, and APX mitigated the oxidative stress; pronouncedly reduced root fresh biomass and water content was observed using a dose of 400 mg/L. At exposure to 50 mg/L TiO2 NPs stimulated accumulation of Ca, Mg, Fe, Mn, Zn, and B in shoots, resulting in improved nutrient quality of edible plan parts was observed; internalization or translocation of TiO2 NPs to shoots was not detected [254]. At studying of the transgenerational effects TiO2 NPs in Ocimum basilicum plants, it was found that hydrophobic or hydrophilic TiO2 NPs exhibited beneficial impact on plants growth, but more negative effects on photosynthesis compared to pristine TiO2 NPs. Treatments with hydrophobic and hydrophilic TiO2 NPs resulted in the reduction of Chlb and total Chl by 52% and 30%, respectively, while total sugar and reducing sugar contents achieved 186% and 145% of contents observed in unexposed plants in both cycles [255].
Spraying of Vetiveria zizanioides plants with 90 mg/L TiO2 NPs, which was performed 300 d after transplantation, increased the Chl content and maximum photochemical efficiency of PS II and yield of EO by 27.2%, 23.5%, and 55.1%, respectively, compared to the control [145].

4.5.2. Elicitor Effect

Spraying of S. rebaudiana Bertoni plants with 400 ppm TiO2 NPs showed pronounced beneficial effect on fresh and dry weights of shoots, while application of 200 ppm mostly enhanced stevioside glycoside and reduced MDA levels [147]. In addition, application of 100 mg/L TiO2 NPs also increased EO content and contents of important secondary metabolites. Moreover, TiO2 NPs ameliorated agronomic traits and increased antioxidant enzyme activities in D. moldavica L. plants grown under salt stress, and considerably reduced H2O2 levels [149].
The beneficial impact of 150 mg/L TiO2 NPs on foliarly treated M. piperita plants was reflected in an increase of EO yield up to 105.1% and application of 150 mg/L TiO2 NPs increased the yields of menthol, menthone, and menthyl acetate by 124.1%, 169.4%, and 130.5%, respectively, over the control, whereby contents of these secondary metabolites were enhanced as well [144]. Concentration of 0.06% TiO2 NPs was reported as most favorable for foliar spraying of Cuminum cyminum L. to achieve the highest EO yield of cumin [256]. Treatment of black cumin plants in the early flowering stage with 100 mg/L TiO2 NPs promoted the geranyl diphosphate synthase (GPPS) expression 24 h after treatment more than application of a half dose, which resulted in improved production of thymoquinone [257]. TiO2 NPs applied foliarly at a dose 100 mg/L acted as an elicitor by improving secondary metabolism in S. officinalis plants resulting in enhanced biosynthesis of natural antioxidants and at application of 100 mg/L TiO2 NPs highest dry weights of roots and shoots were observed as well [146]. Spraying of rosemary leaves with TiO2 NPs (7-times, using doses 20–400 ppm) pronouncedly affected quantity of EO, mostly increasing the content of secondary metabolites, although at concentrations >200 ppm some of these compounds (e.g., β-pinene, 3-pinanol, myrcene, 1,8-cineol) showed a decrease [258].
Ebadollahi et al. [154] compared the effect of TiO2/perlite nanocomposites (15.50–24.61 nm) applied at doses 25–200 mg/L on production of secondary metabolites in the calli obtained from in vitro, as well as from field grown plants of Hypericum perforatum. They found that, at application of 25 mg/L TiO2/perlite NCs, the enhancement of alkaloids production in the calli obtained from in vitro grown plants related to control was almost by one order higher than in calli obtained from field grown plants. Exposure of Tanacetum parthenium L. grown in greenhouse to TiO2 NPs affected the expression of genes involved in biosynthesis pathway of parthenolide and β-caryophyllene; TiO2 NPs upregulated the costunolide synthase (COST) and feverfew (T. parthenium) germacrene A synthase (TpGAS) but downregulated the feverfew (T. parthenium) (E)-β-caryophyllene synthase (TpCarS) with increasing time from 6 to 24 h. Exposure to salinity increased the expression of these three genes with increasing time, suggesting that TiO2 NPs and salinity can be used as elicitors for achieving more effective production of secondary metabolites [259].
Khusimol content and yield of Vetiveria zizanioides plants sprayed with 90 mg/L TiO2 NPs 300 d after transplantation increased by 24.5% and 93.2%, respectively, compared to control [145]. In hydroponically cultivated Nigella arvensis plants treated for 21 days with TiO2 NPs increased amounts of secondary metabolites, such as glaucine and quercetin, as well as higher TPC levels were observed compared to the control, although the concentrations of the elicitor providing the highest levels of individual bioactive components differed each from other [148].

4.5.3. Attenuation of Abiotic Stress

Treatment of Verbascum sinuatum plants exposed to artificial drought conditions (−0.6 MPa) with 20 ppm TiO2 was found to mitigate negative impact of water deficit on growth characteristics and contents of assimilation pigments by stimulation of antioxidant defense systems, resulting in improved drought tolerance [260]. Foliar application of 10 ppm TiO2 NPs increased shoot dry mass and EO content of Dracocephalum moldavica L. plants and in plants exposed to drought stress. Plants grown under water deficit, which were treated with 10 ppm TiO2 NPs, had more proline and considerably lower content of H2O2 and MDA compared to untreated plants, suggesting that TiO2 NPs can mitigate water deficit-induced oxidative damages. On the other hand, assimilation pigments in leaves were pronouncedly reduced in Dracocephalum moldavica plants treated with 10 ppm TiO2 NPs [261]. Whereas spraying with TiO2 NPs positively affected growth of Thymus vulgaris cultivated under drought stress, it did not affect the components of EO [262].
TiO2 NPs (2 and 5 ppm) increased tolerance of Glycyrrhiza glabra L. plants to cold stress, reduced MDA and H2O2 levels in cold-stressed (4 °C) plants, resulting in lesser oxidative damage, and increased the contents of phenolics, total protein, and osmolytes [263]. Treatments of sensitive and tolerant chickpea genotypes with TiO2 NPs did not induce oxidative damage in plans, which were able to mitigate membrane damage and electrolyte leaching, and pronouncedly reduce MDA content also under cold stress conditions via improved redox status of the genotypes, with a concentration of 5 ppm TiO2 NPs being the most effective [264].
Everyday spraying with TiO2 NPs at doses 25 and 50 mg/L was found to prevent Crocus sativus L. plants cultivated hydroponically against harmful effects on morphological characteristics of plants caused by UV irradiation (30 and 45 min daily for one month). Combined treatment reduction of dissolved sugars, increased levels of total anthocyanins and MDA in leaves, as well as higher radical scavenging activity were observed; a raise in TPC and TFC in saffron stigmas was estimated as well, which could contribute to enhanced nutritive value of saffron [265].
TiO2 NPs (anatase; 10–25 nm) were able to increase concentrations of assimilation pigment in leaves of flax under normal and water deficit conditions, reduce the levels of H2O2 and MDA, resulting in lower degree of lipid peroxidation, particularly at a dose of 10 mg/L, and enhance seed oil and protein contents in treated plants [266]. In cell suspension cultures of flax exposed to 50, 100, and 150 mg/L TiO2 NPs, increased activity of cinnamyl alcohol dehydrogenase and maximum TPC content was observed at application of 150 mg/L TiO2 NPs; TiO2 NPs at tested concentrations increased the total lignan content as well [140].

4.5.4. Genotoxicity

Using A. cepa root tip test, it was shown that TiO2 NPs (>50 nm) applied at 0.1 and 100 g/L reduced the MI from 63.6% (control) to 59.5% and 53.5%, respectively, and at additional UV stress further decrease to 58.0% and 51.4%, respectively, was observed. Whereas at treatment with 0.1 mg/L TiO2 NPs (anaphase) multiple chromatin bridges, at exposure to 10 mg/L and 100 mg/L TiO2 NPs distributed and fragmented chromosomes can be observed. The chromosomal aberration of A. cepa root cells can be associated with accelerated generation of ROS and higher MDA levels at the presence of TiO2 NPs [245]. Similar results were observed also by Pakrashi et al. [267] or Ahmed et al. [268] who reported that due to oxidative stress induced by of TiO2 NPs the ROS levels as well as SOD and CAT activities of A. cepa roots increased by 10%, 20.8%, and 12.4%, respectively, and the levels of O2 in roots showed a dose-dependent increase. TiO2 applied at a dose 1000 mg/L caused DNA damage in A. cepa root meristem cells, which, expressed as %DNA tail, achieved 56.54 ± 3.82% and 37.57 ± 3.60% for TiO2 NPs of 50 and 21 nm, respectively [269].
Considering 2.4-fold enhancement of glaucine in Nigella arvensis L. plants cultivated in hydroponic system in the presence of 2500 mg/L TiO2 NPs, this method seems to be suitable to achieve enhanced levels of this aporphine alkaloid in N. arvensis plants [147] (Table 1).

4.6. NPs of Other Metals

4.6.1. MgO NPs and Mn2O3 NPs

Elicitor Effect
MgO NPs immobilized on the surface of nanoperlite using Melissa officinalis extract as a capping agent applied at doses 25, 50, and 150 mg/L of MgO/perlite NCs, showing the size 30 nm elicited in vitro biosynthesis of RA in M. officinalis plant organ cultures [270]. Mn2O3 NPs (ca. 30 nm) supplemented to MS culture medium with tissue culture of Atropa belladonna at doses 25–200 mg/L affected growth characteristics, leaf relative water content, Chl levels, H2O2, and MDA contents, and electrolyte leakage by altering the protein content and activation of antioxidant and defense enzymes. Mn2O3 NPs also affected biosynthesis of secondary metabolites, including TPC, TFC, and alkaloids in a dose-dependent manner. Application of 25 mg/L Mn2O3 NPs was found to promote not only plant growth, but also the production of alkaloids in shoot tip cultures of Atropa belladonna, which reached 1.84- and 2.92-times greater levels as that of positive (22.3 mg/L MnSO4 · 4H2O) and negative (medium without Mn) controls, respectively. Extracts prepared from plantlets showed antioxidant properties [151].
Attenuation of Abiotic Stress
Mn compounds are able to stimulate crop growth and mitigate abiotic stress, whereby at application of nanosized Mn particles, higher efficiency in attenuation of abiotic stresses and less toxic effects can be achieved compared to bulk Mn particles. In addition, plant Mn deficiency can be overcome by additional supply of Mn, resulting in induction of manganese superoxide dismutase at the transcriptional level, to suppress ROS production and promote Mn-dependent proteins for preservation of cell integrity [271].
Genotoxicity
Chromosomal aberrations and a decrease in MI compared to the control, with a pronounced raise in H2O2 and O2 production and lipid peroxidation, was observed in Allium cepa roots treated with 12.5–100 µg/mL MgO MPs and MgO NPs, the toxicity of NPs being higher [272].

4.6.2. CoNPs

Elicitor Effect
Artemisia annua cell suspension culture exposed to 5 mg/L CoNPs for 24 h achieved artemisinin content of 113.35 mg/g (d.w.), being 2.25-fold higher than that of the control. Increased artemisinin production was assumed to be associated with decreased expression levels of SQS and DBR2 genes playing essential roles in the regulation of artemisinin pathway [134].

4.6.3. NiNPs and NiO NPs

Impact on Plant Growth
Exposure of Coriandrum sativum L. plants to NiNPs with mean particle size 20 nm, applied at doses 20, 40, and 80 ppm for 22 days, resulted in reduced water content, root, and shoot growth, decreased levels of photosynthetic pigments, lower antioxidant activity and increased NiNPs content in plants, suggesting toxic effect of NiNPs on this medicinal plant [273]. In vitro grown seedlings of Abelmoschus esculentus treated with NiO NPs (100–1000 mg/L) showed impaired plant growth, lower Chl content, pronouncedly altered levels of anthocyanin contents, TPC, and TFC and increased ROS and MDA levels, suggesting phytotoxicity of NiO NPs [274].
Elicitor Effect
Nigella arvensis plants treated with 50 mg/L NiO NPs resulted in 2.2- and 1.8-fold higher quercetin content in the shoots and roots compared to the control plants. On the other hand, following exposure of black cumin plants to 1000 mg/L NiO NPs up to 3.2-fold higher glaucine content in shoots, 2.9-fold higher kaempferol content in roots, and considerably higher levels of TPC and TFC were observed compared to the control plants, suggesting that NiO NPs can be used as an elicitor for effective in vitro production of secondary metabolites [148].
Genotoxicity
Treatment of Allium sativum L., Allium schoenoprasum L., Allium porrum L., and Allium fistulosum L. growing root tips with NiO NPs, even at a dose of 10 mg/L induced genotoxicity in plants. NiO NPs caused perturbation of biochemical homeostasis and disrupted normal physiology of the cells, which was reflected in enhanced lipid peroxidation, MDA levels, and activities of antioxidant enzymes (CAT, SOD, and GPX) [275].

4.6.4. Cr2O3 NPs

Genotoxicity
Chromosomal aberrations and cytogenetic effects in root tip cells of A. cepa due to exposure to Cr2O3 NPs (0.01–100 µg/mL) were reported by Kumar et al. [276] and in A. cepa roots treated with 25–100 ppm Bi2O3 stickiness chromosome laggards, disturbed anaphase-telophase, and anaphase bridges were observed in anaphase-telophase cells, pro-metaphase, and c-metaphase in other cells, as well as a pronounced rise in DNA damage [277].

4.6.5. CeO2 NPs

Coriandrum sativum L. plants germinated and cultivated for 30 days in soil amended with 125 mg/kg CeO2 NPs were characterized with longer roots, pronouncedly increased CAT activity in shoots, and APX activity in roots, and changed the chemical environment of carbohydrates in shoots, suggesting impact of CeO2 NPs on the nutritional properties of cilantro [278]. Green synthesized organometallic CeO2 nanostructures with mean size 38 ± 5 nm applied at a dose of 4 mg/L increased callus induction of Berberis lycium Royle to 90%, whereas using a dose of 20 mg/L direct shoot regeneration was improved to 79% [279]. On the other hand, CeO2 NPs (≈20.28 nm) and bulk CeO2 (≈4.24 µm) particles applied at doses 12.5–100 ppm for 4 h exhibited cytotoxic and genotoxic effects on the root meristem cells of A. cepa reflected in decreased MI values, considerably increased chromosomal aberrations including chromosome laggards, disturbed anaphase-telophase, stickiness, bridges, and caused DNA damage [280].

4.6.6. Al2O3 NPs

Due to oxidative stress induced in roots of A. cepa, which were exposed to Al2O3 NPs—ROS, SOD, and CAT activities were enhanced by 30%, 17.1%, and 13.2%, respectively, as compared to the control and the levels of O2, which showed a dose dependent increase [268].
Genotoxicity
Exposure of A. cepa roots to Al2O3 NPs showed chromosomal aberration, whereby application of low Al2O3 NPs concentration (0.1 mg/L) disturbed metaphase. At treatment with 10 and 100 mg/L Al2O3 NPs—abnormal anaphase and sticky metaphase were observed; reduction of MI and generation of ROS was observed as well [245]. Genotoxic effects and oxidative damage on A. cepa caused by Al2O3 NPs were pronouncedly higher than those of bulk Al2O3, and a considerable raise of GPX accompanied with the depletion in CAT activity was observed in plants treated with both Al2O3 NPs and bulk Al2O3, respectively [281].

4.6.7. AuNPs

Elicitor Effect
AuNPs (24.2 ± 2.4 nm) applied at concentrations of 50 and 10 mg/L, respectively, elicited biosynthesis of secondary metabolites in Lavandula angustifolia in vitro cultures, whereby lower molecular weight compounds of EO (e.g., α- and β-pinene, camphene, δ-3-carene, p-cymene, 1,8-cineole) were replaced by compounds showing higher molecular weight, e.g., cadalene, α-bisabolol, and (E,E)-farnesol [130]. Spraying of shoots of 6-year-old ginseng plants with AuNPs, performed 3-fold, increased the content of ginsenosides Rg1, Re, Rf, and Rb1; using the steaming process only, the levels of Rd and Rg3 were enhanced [282].
Considering 3.2-fold enhancement of glaucine in Nigella arvensis L. plants cultivated in hydroponics in the presence of 1000 mg/L NiO NPs, this method seems to be suitable to achieve enhanced levels of this aporphine alkaloid in N. arvensis plants [136] (Table 1).

5. Effect of Metalloid-Based NPs on Medicinal Plants

5.1. Silicon and Silica NPs

Silicon has beneficial impact on plants, particularly under stress conditions, and is considered a non-essential beneficial plant nutrient. It ameliorates not only the vigor of plants, but also their resistance to exogenous stresses [50,283,284].

5.1.1. Impact on Germination and Plant Growth

SiO2 NPs applied at a dose of 400 mg/L pronouncedly improved seed germination, enhanced the length and fresh/dry weight of plant organs of Nigella sativa L., and increased levels of photosynthetic pigments, total protein, total amino acid, and proline [285]. Irrigation of hawthorn (Crataegus aronia L.) seedlings with SiO2 NPs at doses 10, 50, and 100 mg/L for 45 days before exposure to drought stress resulted in improved plant biomass, xylem water potential, and MDA content, especially under drought conditions, and had positive influence on photosynthetic pigments; relative water content and membrane electrolyte leakage were not affected and carbohydrate and proline content showed a decrease under all water regimes, especially under water stress [286]. Plants grown on MS medium from Matricaria recutita seeds treated for 1 h with SiO2 NPs (4 or 6 g/L) were characterized with improved growth characteristics, higher relative water content, TPC, and TFC, but reduced levels of H2O2 and MDA [153].

5.1.2. Elicitor Effect

Pretreatment of Calendula officinalis L. seeds with 200 mg/L SiNPs resulted in maximal increase of quercetin content in plants grown under 50% FC [287]. Treatment of Tanacetum parthenium L. plants with 25 mM SiO2 NPs increased the expression of TpGAS, COST, and TpCarS genes, which are involved in the biosynthesis pathway of the secondary metabolites, parthenolide, and β-caryophyllene, when the time increased from 6 to 24 h, and a similar effect on the expression of genes exhibited salinity stress [259]. Mentha piperita L. plants treated with 50 and 100 mg/L SiNPs considerably enhanced the density and diameter of the peltate glandular trichomes, Chl content, net photosynthetic rate, TPC, and EO content, as well as content of secondary metabolites at 150 days after plantation [152]. Dracocephalum kotschyi HRs elicited by 100 mg/L SiO2 NPs after 48 h exposure time showed considerable increase in biomass compared to the control and pronounced enhancement of the expression of pal and ras genes, as well as a strong increase in several secondary metabolites was observed as well [163]. SiO2 NPs and TiO2 promoted the expression of GPPS, a key gene involved in thymoquinone biosynthesis pathway in Nigella sativa L. plants, TiO2 NPs being more effective, and a dose of 100 mg/L raised the gene expression more than that of 50 mg/L. The tested NPs acting as elicitors increased the biosynthesis of thymoquinone in plants via upregulation of related metabolic pathway genes [257]. Continuous production of metabolites from plant cultures using nanoharvesting, in which NPs are designed to bind and carry biomolecules out of living cells, was described by Khan et al. [288] who treated Solidago nemoralis HR cultures in MS medium with mesoporous SiO2NPs (165 nm diameter) functionalized with both TiO2 (425 mg/g particles) for coordination binding sites, and NH2 (145 mg/g particles) to stimulate cellular internalization in order to obtain increased production of polyphenolic flavonoids. Moreover, observed post-nanoharvesting growth indicating the viability of the roots after nanoharvesting, and thus their ability to synthesize henceforward the flavonoids, suggested that application of NPs can facilitate continuous isolation of many biologically active compounds from living and functioning plant cultures.

5.1.3. Attenuation of Abiotic Stress

SiNPs effectively reduced adverse effect of saline stress on leaf dry and fresh weight and Chl content in Ocimum basilicum plants and increased proline levels at exposure to SiNPs suggested induction of the tolerance in basil plants [289]. Exposure of Allium cepa roots to non-irradiated suspensions of SiO2 NPs and those irradiated with UV-A light did not lead to the adsorption of SiO2 NPs on the root surface and no considerable adverse impact was observed even at high exposure concentrations. Consequently, it can be supposed that cell wall shielded the cell membrane from direct contact with NPs. It could be mentioned that similar results were obtained also with application of TiO2 NPs showing photocatalytic potential [290]. Priming of Calendula officinalis L. seeds with 200 mg/L SiNPs greatly enhanced antioxidant activity and TFC of plants exposed to drought stress of 25% FC [287].
Production of rosmarinic acid elicited by SiO2 NPs in HR of Dracocephalum kotschyi achieving up to 8.26-fold higher yield compared to control can be considered as very promising [163]. However, also beneficial impact of foliar treatment of Mentha piperita plants with 100 mg/L SiNPs resulting in nearly double enhancement of menthol, menthone and menthyl-acetate in plants seems to be convenient for application also in the field [152] (Table 1).

5.2. Selenium NPs

Selenium is a member of the chalcogen family. It is classified by some researchers as metalloid showing intermediate properties between metals and nonmetals [291,292]. It resembles chemically to sulfur and shares a similar pathway for uptake and translocation in plants.

5.2.1. Impact on Plant Growth and Elicitor Effect

Impact of SeNPs and nitric oxide on growth, metabolism, antioxidant machinery, gene expression, and flowering Cichorium intybus L. plants was discussed by Abedi et al. [293]. Stimulating effect on Capsicum annuum grown in vitro in MS medium was observed after addition of 0.5 and 1.0 mg/L SeNPs; considerable induction in nitrate reductase activity was observed as well [294]. The low doses of 0.5 and 1 mg/L of SeNPs showed growth-stimulating effects of growth of Capsicum annuum grown in vitro in MS medium, while higher SeNPs concentrations (10 and 30 mg/L) were found to be phytotoxic due to DNA hypermethylation, upregulation of the bZIP1 transcription factor as well as upregulation of the expression of the WRKY1 transcription factor and inhibition in the differentiation of xylem tissues [294]. Exposure of Ocimum basilicum plants to CdSSe quantum dots (QDs) and Cd2+ ions resulted in an initial adverse impact on plant growth, but after 3 and 6 weeks, the roots of treated plants were comparable with those of the control group. Although the Ocimum basilicum plants exposed to CdSSe QDs accumulated Cd predominantly in roots, when they were applied at doses 25 and 50 mg/kg of plant material, higher translocation of Cd to shoots was observed compared with application of Cd2+ ions during 3 and 6 weeks. This can be related to the functional groups on the surface of QDs, which make easier their movement by the cationic transporters or permit free movement of QDs through the cells, whereas due to interaction of Cd2+ ions with CO32− and PO43− species inside the plant, insoluble salts are formed, which can be immobilized in the apoplastic or symplastic compartments [295].

5.2.2. Elicitor Effect

Treatment of Capsicum annuum grown in vitro in MS medium with SeNPs enhanced the activity of phenylalanine ammonia-lyase and concentrations of soluble phenols as well [294]. Similar results were obtained at in vitro treatment of Momordica charantia with SeNPs [296].

6. Effects of Carbon-Based NPs on Medicinal and Aromatic Plants

In addition to biochar, which is widely used to improve soil quality, nanoscale carbonaceous materials, such as carbon dots (CDs), graphene (GR), graphene quantum dots (GR QDs), graphene oxide (GO), reduced GO (rGO), fullerenes, single-walled (SWCNTs), and multi-walled carbon nanotubes (MWCNTs) on plants are also increasingly used in agriculture [40,297,298,299]. CDs are quasi-spherical carbon-based NPs (>10 nm) with a core that can be either amorphous or nanocrystalline. GR is a two-dimensional allotropic form of carbon, formed by individual layers of sp2-bonded carbon atoms arranged in a hexagonal lattice with a distance of 0.142 nm between adjacent carbon hexagon atoms. GO is achieved by introducing carbonyl, hydroxy, and epoxide groups on planar surfaces and edges of graphene sheets, while rGO is obtained by oxidation of GR, followed by reduction and exfoliation. Carbon nanotubes (CNTs) are cylindrical structures with open or closed ends and, depending on the number of concentric layers of rolled graphene sheets, are classified as SWCNTs (outer diameter 0.8–2 nm) and MWCNTs (outer diameter 5–20 nm) with a length ranging from 100 nm to several centimeters. Fullerenes are composed entirely of carbon and exist in the form of hollow balls (buckyballs), ellipsoids, or tubes (buckytubes or CNTs), e.g., [40,300,301].

6.1. Carbon Dots

Impact on Photosynthesis
Li et al. [81] comprehensively overviewed the interactions of CDs on plant growth, internal physiological processes, and further external factors affecting plant growth. Amine functionalized CDs, effectively conjugated over the surface of the chloroplast after absorbing light or photons can transfer electrons towards chloroplasts, thereby positively affecting photosynthetic electron transport resulting in faster conversion of light energy to the electrical energy and finally to the chemical energy [302]. Li et al. [303] prepared far-red (FR) CDs that are able to efficiently convert UV-A light to 625–800 nm FR emission, i.e., wave lengths suitable to be directly absorbed by photosynthesizing organisms. In an in vitro experiment, coating of chloroplasts of Lactuca sativa L. var. longifolia with FR CDs resulted in a hybrid photosynthetic system, ensuring improved efficiency of electron transport between photosystem PSII and PSI, leading to increased ATP production. Moreover, in an in vivo experiment, the FR CDs-treated lettuce achieved higher electron transfer rate by 28.00% compared to the control, resulting in increased fresh and dry weights by 51.14% and 24.60%, respectively, suggesting that FR CDs could be used for improved conversion efficiency from solar energy to chemical energy.

6.2. Graphene Quantum Dots and Graphene Nanosheets

Treatment of coriander and garlic seeds with 0.2 mg/mL of GR QDs for 3 h before planting resulted in improved growth rate of plant organs of both plants, including fruits, suggesting that GR QDs could be used as plant growth regulators [304]. Spraying of graphene nanosheets (GNS) on Capsicum annuum L. and Solanum melongena L. plants during two seasons at doses 0.1, 0.2 and 0.3 GNS/g induced metabolic regulation of the leaves physiological status and resulted in an increased number of branches and number of fruits per plant, as well as fruit yield. GNS were localized on plastids, cell walls, and intercellular spaces of both plants, and GNS situated inside chloroplasts were found to activate assimilation pigments resulting in promotion of fructose, sucrose, and starch. Application of GNS triggered activity of antioxidant enzymes CAT, APX, GPX, and glutathione-S-transferase, and induced SOD and antioxidant molecules, reducing the levels of OH and O2 radicals and, thus, prevented the lipid peroxidation and electrolyte leaching. Moreover, increased PAL activity due to GNS treatment stimulated the formation of valuable secondary metabolites [305].

6.3. Graphene Oxide

6.3.1. Impact on Germination and Plant Growth

Exposure of Salvia mirzayanii seed to GO retarded and reduced the germination as well as growth characteristics of plants due to lower water uptake and oxidative stress, the effects being greater in demucilaged compared to intact (mucilaginous) seeds. On the other hand, the dependence of growth characteristics on the applied concentration of GO/polyaniline (PANI) NC was bi-phasic, showing growth stimulation at low composite doses and growth inhibition at high composite doses. In shoots of S. mirzayanii plants exposed to GO, and high concentration of GO/PANI, elevated H2O2 levels were estimated. GO/PANI NC also showed beneficial impact on germination and ensured a regular porosity pattern in roots resulting in ameliorated water uptake and, therefore, they could be utilized in drought-prone ecosystems [306].

6.3.2. Attenuation of Abiotic Stress

The addition of 800 µg/L nanosized GO to Plantago major L. leaf-derived calli grown on the 1/2 MS medium, in which drought stress conditions were induced by PEG, caused a 78.5% decrease in relative growth rate and 48.2% reduction of osmotic potential value, while dry matter and H2O2 contents increased by 35.1% and 54.2%, respectively. Under normal water availability in the presence of 800 µg/L nanosized GO, a pronounced increase of phenolic and flavonoid contents by 40.9% and 35.3%, respectively, compared to the control, was observed, while proline content showed a considerable decrease by 26.9% [307].

6.4. SWCNTs

Attenuation of Abiotic Stress
Exposure of Hyoscyamus niger seeds under drought stress (0.5–1.5 MPa) to low concentrations of SWCNTs for 14 days resulted in improved water uptake and upregulation of mechanisms involved in starch hydrolysis, while oxidative stress and electrolyte leakage were suppressed. In the presence of SWCNTs, plant defense system was activated, resulting in increased activities of antioxidant enzymes, such as APX, CAT, POD, and SOD as well as improved biosynthesis of phenolics and proline, which finally contributed to the mitigation of drought stress [308].

6.5. MWCNTs

6.5.1. Impact on Germination and Plant Growth

Beneficial impact of MWCNTs on growth rate, germination, and morphological characteristics of Salvia sclarea and Salvia macrosiphon was also reported by Mehrjardi et al. [309].

6.5.2. Elicitor Effect

In vitro cultivation of Thymus daenensis seeds in MS medium containing MWCNTs, which were found to pass across the plant cell wall and enter the cellular cytoplasm, affected early germination of plants, and stimulated the production of plant biomass up to a dose of 250 µg/mL, while further increasing of MWCNTs concentration reduced the biomass. Moreover, MWCNTs upregulated the activities of enzymes and antioxidants and elicited biosynthesis of some secondary metabolites and antioxidants, the optimum dose of MWCNTs being 250 µg/mL [157]. Treatment of Satureja khuzistanica Jamzad plants grown in vitro with 250 mg/L MWCNTs resulted in RA content of 140.49 mg/g (d.w.), while at in vivo exposure of S. khuzistanica to MWCNTs for 24 h, the RA content of 7.13 mg RA/g (d.w.) was achieved [310]. Stimulation of callus induction, biosynthesis of secondary metabolites, and improved antioxidant capacity in medicinal plant S. khuzestanica grown in vitro by MWCNTs was also reported previously by Ghorbanpour and Hadian [158].
Spraying of the leaves of two-month-old Salvia verticillata L with MWCNTs (0–1000 mg/L) resulted in their absorption via the epidermal cells layer into the parenchymal cells of the exposed leaves; the treatment reduced levels of assimilation pigments and caused oxidative stress in a dose-dependent manner. However, at application of 50 and 1000 mg/L MWCNTs, an (approximate) 4-fold increase of RA compared to the control was observed, whereby the activity and gene expression patterns of RA synthase correlated with the RA accumulation. While lower ROS levels at application of lower MWCNT concentrations contributed to improved production of secondary metabolites, treatment with higher concentrations was phytotoxic due to increased oxidative stress [155]. The γ-irradiation of Salvia nemorosa callus, using a dose 70 Gy, resulted in a high yielding cell line able to produce up to 18.53, 5.21, 1.9, and 7.59 mg/g d.w. of RA, salvianolic acid B, ferulic acid, and cinnamic acid, respectively. The cell suspension culture prepared of irradiated callus and elicited with 100 mg/L MWCNT-COOH (outer diameter of 20–30 nm) achieved fresh and dry biomass of 268.47 g/L and 22.17 g/L, respectively. Both treatment with γ-irradiation and elicitation using MWCNT-COOH pronouncedly improved the antioxidant activity of cultures. These findings showed that improved production of secondary metabolites in S. nemorosa achieved by combination of cell line selection via γ-irradiation with elicitation with MWCNT-COOH could be utilized for large-scale production of phenolic compounds [159].
MWCNTs increased the production of biomass production in callus culture of Catharanthus roseus cultivated in the dark, promoted the biosynthesis of total produced alkaloids, whereby strong improvement in the production of vinblastine and vincristine alkaloids was observed, not only with callus cultivation at dark, but also in light conditions [311].

6.5.3. Attenuation of Abiotic Stress

The addition of 50 mg/L of MWCNTs functionalized with carboxylic acid groups (MWCNT-COOH) in hydroponic solution was able to improve the adverse impact of salt stress (50 and 100 mM NaCl) on hydroponically cultivated O. basilicum L. plants, resulting in increased contents of assimilation pigments and the levels of non-enzymatic antioxidants and enhanced activities of antioxidant enzymes, including APX, CAT, and GPX. Whereas application of 25 mg/L MWCNT-COOH was found to be beneficial for plant growth and EO content and compound profile; treatment with a 4-fold higher dose was found to be phytotoxic. Hence, MWCNT-COOH could be considered as plant growth promoting and stress protecting agents, which have potential to be used in agriculture [312].

6.5.4. Genotoxicity

Uptake of MWCNTs in Allium cepa root cells modified cellular morphology, adversely affected membrane integrity and mitochondrial function, and caused considerable DNA damage, micronucleus formation, and chromosome aberration, as well as formation of internucleosomal fragments, which is indicative of apoptotic cell death. Accumulation of cells in the sub-G0 phase of the cell cycle, pronounced increases in CpG methylation, and in the levels of 5-methyl-deoxycytidine verified the cyto-genotoxic effect of MWCNT [313].
The use of MWCNT-COOH as an elicitor of rosmarinic acid and salvianolic acid B in a cell suspension culture prepared from γ-irradiated (70 Gy) Salvia nemorosa L callus, which allows to reach >13- and 14-fold higher concentrations of these secondary metabolites than in the control, can be recommended for obtaining these bioactive acids on a larger scale [159] (Table 1).

7. Effects of Organic Material-Derived NPs on Medicinal and Aromatic Plants

Chitosan/tripolyphosphate NPs supplemented to culture media modified the root architecture and differentiation in micropropagated Capsicum annuum, and at toxic doses 5–20 mg/L caused interruption of plant growth and development. On the other hand, at a dose of 1 mg/L, it acted as an effective elicitor to trigger organogenesis via micropropagation, causing considerable increases in contents of secondary metabolites, including soluble phenols, proline, and alkaloid [314]. Exposure of seeds of Capsicum annuum to aqueous suspensions of nanoscale chitin (0.001–0.05% w/v) or hydropriming reduced mean germination time at 25 °C on blotter paper to 4.9–5.3 days compared to 5.4–6.7 days observed with untreated seeds, seeds treated with 1% acetic acid, or 1% Captan fungicide. Application of 0.05% chitin NPs or hydropriming ameliorated seedling emergence at 19 to 30 °C and chitin NPs was able to reduce fungal growth as well [315].
Root meristems of A. cepa obtained from seeds exposed to nanoplastics (50 nm) at doses 0.01–1 g/L showed reduction of MI and caused induction and of cytogenetic anomalies and micronuclei suggesting their cytotoxicity and genotoxicity. This adverse impact of nanoplastics is caused by the mechanical surface contact of nanoplastics with root external layers and the internalization of nanoplastics in different cellular compartments. Consequently, nanoplastics internalized into crop plants could enter in different trophic levels of the food chain and, so, they represent a hazard for the health of the human population [316]. Polystyrene NPs applied at doses 25–400 mg/L considerably inhibited growth of A. cepa roots, induced the production of OH and O2 radicals, and increased DPPH scavenging activity and lipid peroxidation, causing a decrease in MI compared to the control. They also caused various chromosomal and nuclear aberrations and downregulated the expression of the plant cyclin-dependent kinase (CDKA) encoding gene: cdc2, an important cell cycle regulator. The above results suggest the cytotoxic and genotoxic potential of polystyrene NPs [317].

8. Conclusions

Aromatic and medicinal plants grown in the field require low energy input for cultivation, can contribute to increasing biodiversity in agroecosystems, and their cultivation can also be used to restore degraded areas. Secondary metabolites of aromatic and medicinal plants represent a rich source of phytochemicals that can be used not only for medicinal purposes or as nutraceuticals, but also as “lead compounds” for the design of new drugs. Medicinal plants should generally not be grown in contaminated soil, and chemicals used in the growth or protection of crops should be kept to a minimum. On the other hand, it is desirable to ensure favorable conditions for their germination and plant growth. In this context, the use of nanoparticles, which allow a controlled release of the active ingredient and the use of a smaller amount of active ingredient to achieve the same (or even better) biological effect than in the case of bulk, has a great perspective. The use of seed nanopriming can significantly improve plant germination and growth and can be considered a suitable approach for the revegetation of medicinal plants in harsh environmental conditions. In agricultural soils with low nitrogen content, mesoporous SiO2 NPs loaded with urea can contribute to achieving the desired levels of this macronutrient. On the other hand, in calcareous soils characterized by a lack of micronutrients (especially Zn and Fe), mineral fertilization of medicinal and aromatic plants using nanoscale particles of these essential nutrients or their oxides can be favorable. In particular, foliar spraying with nanosized metal fertilizers allows maximum use of plant nutrients and contributes to the fortification of plants with essential micronutrients. In general, low-concentration nanoparticles applied in such treatment do not contribute to environmental contamination and do not pose a risk to non-target organisms. Medicinal plants, especially their essential oils containing secondary metabolites, show many biological activities, e.g., anti-inflammatory, antimicrobial, antiviral, antidiabetic, anticancer, etc. Oxidative stress induced by NPs stimulates the production of these valuable compounds, not only in medicinal and aromatic plants grown in the field, but also with higher efficiency in tissue and hair root cultures of these plants. Therefore, these “green” in vitro methods, which do not use harmful solvents, can be promising for the efficient production of these beneficial bioactive metabolites. Utilizing the ability of nanoparticles to mitigate the adverse impact of abiotic stresses on medicinal and aromatic plants can be particularly useful in changing climatic conditions, where longer periods of drought and elevated temperatures can be expected in the near future. Increased plant tolerance to salts due to the application of nanoparticles will facilitate the cultivation of medicinal and aromatic plants in soil with higher salinity without adversely affecting their performance. Medicinal and aromatic plants have been associated with the human population since ancient times; their popularity as alternatives to synthetic drugs is currently growing, and the use of appropriate nanoparticles can significantly contribute to their better yield, production of higher concentrations of valuable metabolites, and better nutritional value.

Author Contributions

K.K. and J.J. composition and writing. Both authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Slovak Research and Development Agency (projects APVV-17-0373 and APVV-17-0318), and by the Slovak Grant Agency for Science (VEGA 1/0787/18).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

APX (ascorbate peroxidase); CA (carnosic acid); CAT (catalase); CDKA (cyclin-dependent kinase); CDs (carbon dots); Chl (chlorophyll); CNTs (carbon nanotubes); CSC (cell suspension culture); CTAB (cetyltrimethylammonium bromide); DPPH (2,2-diphenyl-1-picrylhydrazyl); EDTA (ethylenediaminetetraacetic acid); EO (essential oil); FC (field capacity); FITC (fluorescein isothiocyanate); GNS (graphene nanosheets); GO (graphene oxide); GPPS (geranyl diphosphate synthase); GPX (guaiacol peroxidase); GR (glutathione reductase); GR QDs (graphene quantum dots); HR (hairy root); MDA (malondialdehyde); MI (mitotic index); MIC (minimum inhibitory concentration); MP (microparticle); (MS (Murashige and Skoog); MWCNTs (multi-walled carbon nanotubes); NC (nanocomposite); NP (nanoparticle); PAL (phenylalanine ammonia-lyase); PANI (polyaniline); PEG (polyethylene glycol); POD (peroxidase); PPO (polyphenol oxidase); PS (photosystem); PVP (polyvinylpyrrolidone); RA (rosmarinic acid); rGO (reduced GO); ROS (reactive oxygen species); QDs (quantum dots); SMF (static magnetic field); SM (secondary metabolite); SOD (superoxide dismutase); SWCNTs (single-walled carbon nanotubes); TFC (total flavonoid content); TPC (total phenolic content).

References

  1. Pimentel, D. Green revolution agriculture and chemical hazards. Sci. Total Environ. 1996, 188, S86–S98. [Google Scholar] [CrossRef]
  2. Villaverde, J.J.; Sandin-Espana, P.; Sevilla-Moran, B.; Lopez-Goti, C.; Alonso-Prados, J.L. Biopesticides from natural products: Current development, legislative framework, and future trends. BioResources 2016, 11, 5618–5640. [Google Scholar] [CrossRef] [Green Version]
  3. Jampilek, J.; Kralova, K. Nanobiopesticides in agriculture: State of the art and future opportunities. In Nano-Biopesticides Today and Future Perspectives; Koul, O., Ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2019; pp. 397–447. [Google Scholar]
  4. Salmeron-Manzano, E.; Garrido-Cardenas, J.A.; Manzano-Agugliaro, F. Worldwide research trends on medicinal plants. Int. J. Environ. Res. Public Health 2020, 17, 3376. [Google Scholar] [CrossRef]
  5. Masarovicova, E.; Kralova, K. Medicinal plants—past, nowadays, future. Acta Horticult. 2007, 749, 19–27. [Google Scholar] [CrossRef]
  6. Masarovicova, E.; Kralova, K.; Vykoukova, I.; Krissakova, Z. Responses of medicinal plants to abiotic stresses. In Handbook of Plant and Crop Stress, 4th ed.; Pessarakli, M., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 713–753. [Google Scholar]
  7. Bryan, C.P. The Papyrus Ebers; D. Appleton and Co.: New York, NY, USA, 1936. [Google Scholar]
  8. Duke, J.A. Medicinal Plants of the Bible; Conch Magazine Ltd.: New York, NY, USA, 1983. [Google Scholar]
  9. Duke, J.A.; Telatnik, M.A.; Duke, P.K. Herbs of the Bible: 2000 Years of Plant. Medicine; Wendell Whitman Co.: Warsaw, IN, USA, 2007. [Google Scholar]
  10. Gunther, R.T. The Greek Herbal of Dioscorides; Oxford University Press: Oxford, UK, 1934. [Google Scholar]
  11. Gerard, J. The Herbal or General History of Plants, the 2nd Edition of the Famous 1597 Work; Thomas Johnson, T., Ed.; Dover Publications: Mineola, NY, USA, 1975. [Google Scholar]
  12. Janick, J. Herbals: The connection between horticulture and medicine. HortTechnology 2003, 13, 229–238. [Google Scholar] [CrossRef] [Green Version]
  13. Dafni, A.; Petanidou, T.; Vallianatou, I.; Kozhuharova, E.; Blanche, C.; Pacini, E.; Peyman, M.; Stevanovic, Z.D.; Franchi, G.G.; Benitez, G. Myrtle, basil, rosemary, and three-lobed sage as ritual plants in the monotheistic religions: An historical-ethnobotanical comparison. Econ. Bot. 2020, 74, 330–355. [Google Scholar] [CrossRef]
  14. Hartmann, T. From waste products to ecochemicals: Fifty years research of plant secondary metabolism. Phytochemistry 2007, 68, 2831–2846. [Google Scholar] [CrossRef]
  15. Dolab, J.G.; Lima, B.; Spaczynska, E.; Kos, J.; Cano, N.H.; Feresin, G.; Tapla, A.; Garibotto, F.; Petenatti, E.; Olivella, M.; et al. Antimicrobial activity of Annona emarginata (Schltdl.) H. Rainer and most active isolated compound against clinically important bacteria. Molecules 2018, 23, 1187. [Google Scholar] [CrossRef] [Green Version]
  16. Zhang, H.J.; Jampilek, J. Anti-infective drug discovery based on diversified plant natural compounds. Curr. Org. Chem. 2017, 21, 1775–1776. [Google Scholar] [CrossRef]
  17. Jampilek, J. Design of antimalarial agents based on natural products. Curr. Org. Chem. 2017, 21, 1824–1846. [Google Scholar] [CrossRef]
  18. Tyliszczak, B.; Drabczyk, A.; Kudkacik-Kramarczyk, S.; Rudnicka, K.; Gatkowska, J.; Sobczak-Kupiec, A.; Jampilek, J. In vitro biosafety of pro-ecological chitosan based hydrogels modified with natural substances. J. Biomed. Mater. Res. A 2019, 107, 2501–2511. [Google Scholar] [CrossRef]
  19. Do Nascimento, L.D.; de Moraes, A.A.B.; da Costa, K.S.; Galucio, J.M.P.; Taube, P.S.; Costa, C.M.L.; Cruz, J.N.; Andrade, E.H.D.; de Faria, L.J.G. Bioactive natural compounds and antioxidant activity of essential oils from spice plants: New findings and potential applications. Biomolecules 2020, 10, 988. [Google Scholar] [CrossRef]
  20. Nieto, G. A review on applications and uses of Thymus in the food industry. Plants 2020, 9, 961. [Google Scholar] [CrossRef]
  21. Amiri, A.; Mottaghipisheh, J.; Jamshidi-Kia, F.; Saeidi, K.; Vitalini, S.; Iriti, M. Antimicorbial potency of major functional foods’ essential oils in liquid and vapor phases: A short review. Appl. Sci. 2020, 10, 8103. [Google Scholar] [CrossRef]
  22. Bhavaniramya, S.; Vishnupriya, S.; Saleh Al-Aboody, M.; Vijayakumar, R.; Baskaran, D. Role of essential oils in food safety: Antimicrobial and antioxidant applications. Grain Oil Sci. Technol. 2019, 2, 49–55. [Google Scholar] [CrossRef]
  23. Zhao, Y.; Wu, Y.; Wang, M. Bioactive substances of plant origin. In Handbook of Food Chemistry; Cheung, P., Mehta, B., Eds.; Springer: Berlin/Heidelberg, Germany, 2015; pp. 967–1008. [Google Scholar]
  24. Inoue, M.; Hayashi, S.; Craker, L.E. Role of medicinal and aromatic plants: Past, present, and future. In Pharmacognosy–Medicinal Plants; Perveen, S., Al-Taweel, A., Eds.; IntechOpen: Rijeka, Croatia, 2019; Available online: https://www.intechopen.com/books/pharmacognosy-medicinal-plants/role-of-medicinal-and-aromatic-plants-past-present-and-future (accessed on 15 December 2020).
  25. Porritt, G.M.; Garlic, F. Christopher’s Herbal Legacy. Available online: https://www.herballegacy.com/Porritt_History.html (accessed on 15 December 2020).
  26. Tapsell, L.C.; Hemphill, I.; Cobiac, L.; Patch, C.S.; Sullivan, D.R.; Fenech, D.R.; Roodenrys, S.; Keogh, J.B.; Clifton, P.M.; Williams, P.G.; et al. Health benefits of herbs and spices: The past, the present, the future. Med. J. Aust. 2006, 185, S1–S24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Sherman, P.W.; Hash, G.A. Why vegetable recipes are not very spicy. Evol. Hum. Behav. 2001, 22, 147–163. [Google Scholar] [CrossRef]
  28. El-Sayed, S.M.; Youssef, A.M. Potential application of herbs and spices and their effects in functional dairy products. Heliyon 2019, 5, e01989. [Google Scholar] [CrossRef] [Green Version]
  29. Steinhoff, B. Laws and regulation on medicinal and aromatic plants in Europe. Acta Hortic. 2005, 678, 13–22. [Google Scholar] [CrossRef]
  30. Barbieri, C. Medicinal and aromatic plants legislation in the European Union, Italy and several of its regions. Nat. Prod. Res. 2013, 27, 1576–1582. [Google Scholar] [CrossRef] [PubMed]
  31. U.S. National Nanotechnology Initiative. The National Nanotechnology Initiative Strategic Plan. December. Available online: https://www.nano.gov/sites/default/files/pub_resource/nni_strategic_plan_2004.pdf (accessed on 15 December 2020).
  32. European Commission. 2011. Recommendation of 18 October 2011on the Definition of Nanomaterial (2011/696/EU). Official Journal of European Union L 275/38. Available online: https://ec.europa.eu/research/industrial_technologies/pdf/policy/commission-recommendation-on-the-definition-of-nanomater-18102011_en.pdf (accessed on 15 December 2020).
  33. Jampilek, J.; Kralova, K. Nano-antimicrobials: Activity, benefits and weaknesses. In Nanostructures in Therapeutic Medicine, Vol. 2—Nanostructures for Antimicrobial Therapy; Ficai, A., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, the Netherlands, 2017; pp. 23–54. [Google Scholar]
  34. Jampilek, J.; Kralova, K. Impact of nanoparticles on toxigenic fungi. In Nanomycotoxicology, Treating Mycotoxins in the Nano Way; Rai, M., Abd-Elsalam, K.A., Eds.; Academic Press, Elsevier, 2020; pp. 309–348. [Google Scholar]
  35. Jampilek, J.; Kralova, K. Nanocomposites: Synergistic nanotools for management of mycotoxigenic fungi. In Nanomycotoxicology, Treating Mycotoxins in the Nano Way; Rai, M., Abd-Elsalam, K.A., Eds.; Academic Press: Cambridge, MA, USA, 2020. [Google Scholar]
  36. Jampilek, J.; Kralova, K.; Fedor, P. Bioactivity of nanoformulated synthetic and natural insecticides and their impact on environment. In Nanopesticides; Fraceto, L., de Castro, V.S.S., Grillo, R., Ávila, D., Caixeta Oliveira, H., Lima, R., Eds.; Springer: Cham, Switzerland, 2020; pp. 165–225. [Google Scholar]
  37. Tarrahi, R.; Mahjouri, S.; Khataee, A. A review on in vivo and in vitro nanotoxicological studies in plants: A headlight for future targets. Environ. Ecotoxicol. Environ. Saf. 2021, 208, 111697. [Google Scholar] [CrossRef] [PubMed]
  38. Jampilek, J.; Kos, J.; Kralova, K. Potential of nanomaterial applications in dietary supplements and foods for special medical purposes. Nanomaterials 2019, 9, 296. [Google Scholar] [CrossRef] [Green Version]
  39. Jampilek, J.; Kralova, K. Potential of nanonutraceuticals in increasing immunity. Nanomaterials 2020, 10, 2224. [Google Scholar] [CrossRef]
  40. Placha, D.; Jampilek, J. Graphenic materials for biomedical applications. Nanomaterials 2019, 9, 1758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Placha, D.; Jampilek, J. Chronic inflammatory diseases, anti-inflammatory agents and their delivery nanosystems. Pharmaceutics 2021, 13, 64. [Google Scholar] [CrossRef]
  42. Vaculikova, E.; Grunwaldova, V.; Kral, V.; Dohnal, J.; Jampilek, J. Preparation of candesartan and atorvastatin nanoparticles by solvent evaporation. Molecules 2012, 17, 13221. [Google Scholar] [CrossRef] [PubMed]
  43. Makhlouf, A.S.H.; Barhoum, A. Fundamentals of Nanoparticles: Classifications, Synthesis Methods, Properties and Characterization; Elsevier: Amsterdam, the Netherlands, 2018. [Google Scholar]
  44. Luque, R.; Varma, R.S. Sustainable Preparation of Metal. Nanoparticles: Methods and Applications; Royal Society of Chemistry: Cambridge, UK, 2012. [Google Scholar]
  45. Sharon, M. History of Nanotechnology: From Prehistoric to Modern Times; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  46. Shukla, A.K.; Iravani, S. Green Synthesis, Characterization and Applications of Nanoparticles; Elsevier: Amsterdam, the Netherlands, 2019. [Google Scholar]
  47. Jampilek, J.; Kralova, K. Application of nanobioformulations for controlled release and targeted biodistribution of drugs. In Nanobiomaterials: Applications in Drug Delivery; Sharma, A.K., Keservani, R.K., Kesharwani, R.K., Eds.; CRC Press: Warentown, NJ, USA, 2018; pp. 131–208. [Google Scholar]
  48. Jampilek, J.; Kralova, K. Nanopesticides: Preparation, targeting and controlled release. In New Pesticides and Soil Sensors in Nanotechnology in the Agri-Food Industry; Grumezescu, A.M., Ed.; Elsevier: London, UK, 2017; Volume 10, pp. 81–127. [Google Scholar]
  49. Jampilek, J.; Kralova, K. Benefits and potential risks of nanotechnology applications in crop protection. In Nanobiotechnology Applications in Plant Protection; Abd-Elsalam, K., Prasad, R., Eds.; Springer: Cham, Switzerland, 2018; pp. 189–246. [Google Scholar]
  50. Jampilek, J.; Kralova, K. Beneficial effects of metal- and metalloid-based nanoparticles on crop production. In Nanotechnology for Agriculture; Panpatte, D.G., Jhala, Y.K., Eds.; Springer Nature: Singapore, 2019; pp. 161–219. [Google Scholar]
  51. Jampilek, J.; Kralova, K. Nanoparticles for improving and augmenting plant functions. In Advances in Nano-Fertilizers and Nano-Pesticides in Agriculture; Jogaiah, S., Singh, H.B., Fraceto, L.F., Lima, R., Eds.; Elsevier: Amsterdam, the Netherlands, 2021; pp. 171–227. [Google Scholar]
  52. Jampilek, J.; Kralova, K. Impact of nanoparticles on photosynthesizing organisms and their use in hybrid structures with some components of the photosynthetic apparatus. In Plant Nanobionics; Prasad, R., Ed.; Springer Nature: Cham, Switzerland, 2019; pp. 255–331. [Google Scholar]
  53. Jampilek, J.; Kralova, K. Application of nanotechnology in agriculture and food industry, its prospects and risks. Ecol. Chem. Eng. S 2015, 22, 321–361. [Google Scholar] [CrossRef] [Green Version]
  54. Jampilek, J.; Kralova, K. Nanomaterials for delivery of nutrients and growth-promoting compounds to plants. In Nanotechnology: An Agricultural Paradigm; Prasad, R., Kumar, M., Kumar, V., Eds.; Springer-Verlag: Berlin/Heidelberg, Germany, 2017; pp. 177–226. [Google Scholar]
  55. Masarovicova, E.; Kralova, K. Metal nanoparticles and plants. Ecol. Chem. Eng. S 2013, 20, 9–22. [Google Scholar]
  56. Masarovicova, E.; Kralova, K.; Zinjarde, S.S. Metal nanoparticles in plants. Formation and action. In Handbook of Plant and Crop Physiology, 3rd ed.; Pessarakli, M., Ed.; CRC: Boca Raton, FL, USA; Taylor and Francis: Boca Raton, FL, USA, 2014; pp. 683–731. [Google Scholar]
  57. Kralova, K.; Masarovicova, E.; Jampilek, J. Plant responses to stress induced by toxic metals and their nanoforms. In Handbook of Plant and Crop Stress, 4th ed.; Pessarakli, M., Ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 479–522. [Google Scholar]
  58. Oliveira, H.C.; Stolf-Moreira, R.; Martinez, C.B.R.; Grillo, R.; de Jesus, M.B.; Fraceto, L.F. Nanoencapsulation enhances the post-emergence herbicidal activity of atrazine against nustard plants. PLoS ONE 2015, 10, e0132971. [Google Scholar] [CrossRef]
  59. Bombo, A.B.; Pereira, A.E.S.; Lusa, M.G.; Oliveira, E.D.; de Oliveira, J.L.; Campos, E.V.R.; de Jesus, M.B.; Oliveira, H.C.; Fraceto, L.F.; Mayer, J.L.S. A mechanistic view of interactions of a nanoherbicide with target organism. J. Agric. Food Chem. 2019, 67, 4453–4462. [Google Scholar] [CrossRef] [Green Version]
  60. Balah, M.A.; Pudake, R.N. Use nanotools for weed control and exploration of weed plants in nanotechnology. In Nanoscience for Sustainable Agriculture; Pudake, R.N., Chauhan, N., Kole, C., Eds.; Springer Nature: Basel, Switzerland, 2019; pp. 207–231. [Google Scholar]
  61. Patel, K.V.; Nath, M.; Bhatt, M.D.; Dobriyal, A.K.; Bhatt, D. Nanofomulation of zinc oxide and chitosan zinc sustain oxidative stress and alter secondary metabolite profile in tobacco. 3Biotech 2020, 10, 477. [Google Scholar] [CrossRef]
  62. Ma, Y.J.; Xia, J.; Wang, Y.; Wang, J.W. Stimulation of tanshinone production in Salvia miltiorrhiza hairy roots by beta-cyclodextrin-coated silver nanoparticles. Sustain. Chem. Pharm. 2020, 18, 100271. [Google Scholar] [CrossRef]
  63. Singh, R.; Singh, D.P.; Gupta, P.; Jain, P.; Sanchita; Mishra, T.; Kumar, A.; Dhawan, S.S.; Shirke, P.A. Nanoparticles alter the withanolide biosynthesis and carbohydrate metabolism in Withania somnifera (Dunal). Ind. Crops Prod. 2019, 127, 94–109. [Google Scholar] [CrossRef]
  64. Kralova, K.; Masarovicova, E.; Jampilek, J. Risks and benefits of metal-based nanoparticles for vascular plants. In Handbook of Pant and Crop Physiology, 4th ed.; Pessarakli, M., Ed.; Taylor & Francis: Abingdon, UK, 2021; in press. [Google Scholar]
  65. Moharrami, F.; Hosseini, B.; Sharafi, A.; Farjaminezhad, M. Enhanced production of hyoscyamine and scopolamine from genetically transformed root culture of Hyoscyamus reticulatus L. elicited by iron oxide nanoparticles. In Vitro Cell. Dev. Biol. Plant 2017, 3, 104–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Anjum, S.; Anjum, I.; Hano, C.; Kousar, S. Advances in nanomaterials as novel elicitors of pharmacologically active plant specialized metabolites: Current status and future outlooks. RSC Adv. 2019, 9, 40404–40423. [Google Scholar] [CrossRef] [Green Version]
  67. Asl, K.R.; Hosseini, B.; Sharafi, A.; Palazon, J. Influence of nano-zinc oxide on tropane alkaloid production, h6h gene transcription and antioxidant enzyme activity in Hyoscyamus reticulatus L. hairy roots. Eng. Life Sci. 2019, 19, 73–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Linh, T.M.; Mai, N.C.; Hoe, P.T.; Lien, L.Q.; Ban, N.K.; Hien, L.T.T.; Chau, N.H.; Van, N.T. Metal-based nanoparticles enhance drought tolerance in soybean. J. Nanomater. 2020, 2020, 4056563. [Google Scholar] [CrossRef]
  69. Adrees, M.; Khan, Z.S.; Ali, S.; Hafeez, M.; Khalid, S.; Rehman, M.Z.U.; Hussain, A.; Hussain, K.; Chatha, S.A.S.; Rizwan, M. Simultaneous mitigation of cadmium and drought stress in wheat by soil application of iron nanoparticles. Chemosphere 2020, 238, 124681. [Google Scholar] [CrossRef]
  70. Morales-Espinoza, M.C.; Cadenas-Pliego, G.; Perez-Alvarez, M.; Hernandez-Fuentes, A.D.; de la Fuente, M.C.; Benavides-Mendoza, A.; Valdes-Reyna, J.; Juarez-Maldonado, A. Se nanoparticles induce changes in the growth, antioxidant responses, and fruit quality of tomato developed under NaCl stress. Molecules 2019, 24, 3030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Xiao, L.; Guo, H.Y.; Wang, S.X.; Li, J.L.; Wang, Y.Q.; Xing, B.S. Carbon dots alleviate the toxicity of cadmium ions (Cd2+) toward wheat seedlings. Environ. Sci. Nano 2019, 6, 1493–1506. [Google Scholar] [CrossRef]
  72. Yang, J.; Cao, W.; Rui, Y. Interactions between nanoparticles and plants: Phytotoxicity and defense mechanisms. J. Plant. Interact. 2017, 12, 158–169. [Google Scholar] [CrossRef]
  73. Goswami, P.; Yadav, S.; Mathur, J. Positive and negative effects of nanoparticles on plants and their applications in agriculture. Plant. Sci. Today 2019, 6, 232–234. [Google Scholar] [CrossRef]
  74. Rajput, V.; Minkina, T.; Mazarji, M.; Shende, S.; Sushkova, S.; Mandzhieva, S.; Burachevskaya, M.; Chaplygin, V.; Singh, A.; Jatav, H. Accumulation of nanoparticles in the soil-plant systems and their effects on human health. Ann. Agric. Sci. 2020, 65, 137–143. [Google Scholar] [CrossRef]
  75. Rajput, V.; Minkina, T.; Sushkova, S.; Behal, A.; Maksimov, A.; Blicharska, E.; Ghazaryan, K.; Movsesyan, H.; Barsova, N. ZnO and CuO nanoparticles: A threat to soil organisms, plants, and human health. Environ. Geochem. Health 2020, 42, 147–158. [Google Scholar] [CrossRef] [PubMed]
  76. Masarovicova, E.; Kralova, K. Essential elements and toxic metals in some crops, medicinal plants, and trees. In Phytoremediation: Management of Environmental Contaminants; Ansari, A.A., Gill, S.S., Gill, R., Lanza, G.R., Newman, L., Eds.; Springer International Publishing AG: Cham, Switzerland, 2017; Volume 5, pp. 183–255. [Google Scholar]
  77. Masarovicova, E.; Kralova, K.; Sersen, F. Plant responses to toxic metal stress. In Handbook of Plant and Crop Stress, 3rd ed.; Pessarakli, M., Ed.; CRC Press: Boca Raton, FL, USA, 2010; pp. 595–634. [Google Scholar]
  78. Masarovicova, E.; Kralova, K. Plant-heavy metal interaction: Phytoremediation, biofortification and nanoparticles. In Advances in Selected Plant. Physiology Aspects; Montanaro, G., Dichio, B., Eds.; InTech: Rijeka, Croatia, 2012; pp. 75–102. [Google Scholar]
  79. Mittal, D.; Kaur, G.; Singh, P.; Yadav, K.; Ali, S.A. Nanoparticle-based sustainable agriculture and food science: Recent advances and future outlook. Front. Nanotechnol. 2020, 2, 579954. [Google Scholar] [CrossRef]
  80. Usman, M.; Farooq, M.; Wakeel, A.; Nawaz, A.; Cheema, S.A.; Rehman, H.U.; Ashraf, I.; Sanaullah, M. Nanotechnology in agriculture: Current status, challenges and future opportunities. Sci. Total Environ. 2020, 721, 137778. [Google Scholar] [CrossRef]
  81. Li, Y.D.; Xu, X.K.; Wu, Y.; Zhuang, J.L.; Zhang, X.J.; Zhang, H.R.; Lei, B.F.; Hu, C.F.; Liu, Y.L. A review on the effects of carbon dots in plant systems. Mater. Chem. Front. 2020, 4, 437–448. [Google Scholar]
  82. Faraz, A.; Faizan, M.; Sami, F.; Siddiqui, H.; Pichtel, J.; Hayat, S. Nanoparticles: Biosynthesis, translocation and role in plant metabolism. IET Nanobiotechnol. 2019, 13, 345–352. [Google Scholar] [CrossRef]
  83. Szollosi, R.; Molnar, A.; Kondak, S.; Kolbert, Z. Dual effect of nanomaterials on germination and seedling growth: Stimulation vs. phytotoxicity. Plants 2020, 9, 1745. [Google Scholar] [CrossRef]
  84. Masarovicova, E.; Kralova, K.; Kummerova, M. Principles of classification of medicinal plants as hyperaccumulators or excluders. Acta Physiol. Plant. 2010, 32, 823–829. [Google Scholar] [CrossRef]
  85. Salamon, I.; Kralova, K.; Masarovicova, E. Accumulation of cadmium in chamomile plants cultivated in Eastern Slovakia regions. Acta Hort (ISHS) 2007, 749, 217–222. [Google Scholar] [CrossRef]
  86. Mehrian, S.K.; De Lima, R. Nanoparticles cyto and genotoxicity in plants: Mechanisms and abnormalities. Environ. Nanotechnol. Monit. Manag. 2016, 6, 184–193. [Google Scholar]
  87. Koca, F.D.; Duman, F. Genotoxic and cytotoxic activity of green synthesized TiO2 nanoparticles. Appl. Nanosci. 2019, 9, 815–823. [Google Scholar] [CrossRef]
  88. Liman, R.; Basbug, B.; Ali, M.M.; Acikbas, Y.; Cigerci, I.H. Cytotoxic and genotoxic assessment of tungsten oxide nanoparticles in Allium cepa cells by Allium ana-telophase and comet assays. J. Appl. Genet. 2021, 62, 85–92. [Google Scholar] [CrossRef]
  89. Youssef, M.S.; Elamawi, R.M. Evaluation of phytotoxicity, cytotoxicity, and genotoxicity of ZnO nanoparticles in Vicia faba. Environ. Sci. Pollut. Res. 2020, 27, 18972–18984. [Google Scholar] [CrossRef]
  90. Heikal, Y.M.; Sutan, N.A.; Rizwan, M.; Elsayed, A. Green synthesized silver nanoparticles induced cytogenotoxic and genotoxic changes in Allium cepa L. varies with nanoparticles doses and duration of exposure. Chemosphere 2020, 243, 125430. [Google Scholar] [CrossRef]
  91. Rui, M.M.; Ma, C.X.; Hao, Y.; Guo, J.; Rui, Y.K.; Tang, X.L.; Zhao, Q.; Fan, X.; Zhang, Z.T.; Hou, T.Q.; et al. Iron oxide nanoparticles as a potential iron fertilizer for peanut (Arachis hypogaea). Front. Plant. Sci. 2016, 7, 815. [Google Scholar] [CrossRef] [Green Version]
  92. Raliya, R.; Saharan, V.; Dimkpa, C.; Biswas, P. Nanofertilizer for precision and sustainable agriculture: Current state and future perspectives. J. Agric. Food Chem. 2018, 66, 6487–6503. [Google Scholar] [CrossRef]
  93. El-Ramady, H.; Abdalla, N.; Alshaal, T.; El-Henawy, A.; Elmahrouk, M.; Bayoumi, Y.; Shalaby, T.; Amer, M.; Shehata, S.; Fari, M.; et al. Plant nano-nutrition: Perspectives and challenges. In Nanotechnology, Food Security and Water Treatment; Book Series: Environmental Chemistry for a Sustainable, World; Gothandam, K.M., Ranjan, S., Dasgupta, N., Ramalingam, C., Lichtfouse, E., Eds.; Springer: Cham, Switzerland, 2018; Volume 11, pp. 129–161. [Google Scholar]
  94. Ju, M.; Navarreto-Lugo, M.; Wickramasinghe, S.; Milbrandt, N.B.; McWhorter, A.; Samia, A.C.S. Exploring the chelation-based plant strategy for iron oxide nanoparticle uptake in garden cress (Lepidium sativum) using magnetic particle spectrometry. Nanoscale 2019, 11, 18582–18594. [Google Scholar] [CrossRef] [PubMed]
  95. Abbasifar, A.; Shahrabadi, F.; ValizadehKaji, B. Effects of green synthesized zinc and copper nano-fertilizers on the morphological and biochemical attributes of basil plant. J. Plant. Nutr. 2020, 43, 1104–1118. [Google Scholar] [CrossRef]
  96. Adisa, I.O.; Pullagurala, V.L.R.; Peralta-Videa, J.R.; Dimkpa, C.O.; Elmer, W.H.; Gardea-Torresdey, J.L.; White, J.C. Recent advances in nano-enabled fertilizers and pesticides: A critical review of mechanisms of action. Environ. Sci. Nano 2019, 6, 2002–2030. [Google Scholar] [CrossRef]
  97. Verma, S.K.; Das, A.K.; Gantait, S.; Kumar, V.; Gurel, E. Applications of carbon nanomaterials in the plant system: A perspective view on the pros and cons. Sci. Total Environ. 2017, 667, 485–499. [Google Scholar] [CrossRef]
  98. Lahiani, M.H.; Dervishi, E.; Chen, J.H.; Nima, Z.; Gaume, A.; Biris, A.S.; Khodakovskaya, M.V. Impact of carbon nanotube exposure to seeds of valuable crops. ACS Appl. Mater. Interfaces 2013, 5, 7965–7973. [Google Scholar] [CrossRef]
  99. Tripathi, S.; Kapri, S.; Datta, A.; Bhattacharyya, S. Influence of the morphology of carbon nanostructures on the stimulated growth of gram plant. RSC Adv. 2016, 6, 43864–43873. [Google Scholar] [CrossRef]
  100. Mathew, S.; Tiwari, D.K.; Tripathi, D. Interaction of carbon nanotubes with plant system: A review. Carbon Lett. 2020, in press. [Google Scholar] [CrossRef]
  101. Patel, D.K.; Kim, H.B.; Dutta, S.D.; Ganguly, K.; Lim, K.T. Carbon nanotubes-nased manomaterials and their agricultural and biotechnological applications. Materials 2020, 13, 1679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Khodakovskaya, M.V.; Kim, B.S.; Kim, J.N.; Alimohammadi, M.; Dervishi, E.; Mustafa, T.; Cernigla, C.E. Carbon nanotubes as plant growth regulators: Effects on tomato growth, reproductive system, and soil microbial community. Small 2013, 9, 115–123. [Google Scholar] [CrossRef]
  103. Vithanage, M.; Seneviratne, M.; Ahmad, M.; Sarkar, B.; Ok, Y.S. Contrasting effects of engineered carbon nanotubes on plants: A review. Environ. Geochem. Health 2017, 39, 1421–1439. [Google Scholar] [CrossRef] [PubMed]
  104. Liang, Y.; Sun, W.; Zhu, Y.G.; Christie, P. Mechanisms of silicon-mediated alleviation of abiotic stresses in higher plants: A review. Environ. Pollut. 2007, 147, 422–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ma, J.F. Role of silicon in enhancing the resistance of plants to biotic and abiotic stresses. Soil Sci. Plant. Nutr. 2004, 50, 11–18. [Google Scholar] [CrossRef]
  106. Savvas, D.; Ntatsi, G. Biostimulant activity of silicon in horticulture. Sci. Hortic. 2015, 196, 66–81. [Google Scholar] [CrossRef]
  107. Kaur, N.; Sharma, S.; Kaur, S.; Nayyar, H. Selenium in agriculture: A nutrient or contaminant for crops? Arch. Agron. Soil Sci. 2014, 60, 1593–1624. [Google Scholar] [CrossRef]
  108. Hasanuzzaman, M.; Bhuyan, M.H.M.B.; Raza, A.; Hawrylak-Nowak, B.; Matraszek-Gawron, R.; Al Mahmud, J.; Nahar, K.; Fujita, M. Selenium in plants: Boon or bane? Environ. Exp. Bot. 2020, 178, 104170. [Google Scholar] [CrossRef]
  109. Use of Nanofertilizers in Agriculture: Advantages and Safety Concerns. Dextra International. 2020. Available online: http://www.dextrainternational.com/use-of-nanofertilizers-in-agriculture-advantages-and-safety-concerns (accessed on 15 December 2020).
  110. Kopittke, P.M.; Lombi, E.; Wang, P.; Schjoerring, J.K.; Husted, S. Nanomaterials as fertilizers for improving plant mineral nutrition and environmental outcomes. Environ. Sci. Nano 2019, 6, 3513–3524. [Google Scholar] [CrossRef]
  111. Santiago, E.F.; Pontes, M.S.; Arruda, G.J.; Caires, A.R.L.; Colbeck, I.; Maldonado-Rodriguez, R.; Grillo, R. Understanding the interaction of nanopesticides with plants. In Nanopesticides; Fraceto, L.F., de Castro, V.L., Grillo, R., Avila, D., Caixeta Oliveira, H., Lima, R., Eds.; Springer: Cham, Switzerland, 2020; pp. 69–109. [Google Scholar]
  112. Bundschuh, M.; Filser, J.; Luderwald, S.; McKee, M.S.; Metreveli, G.; Schaumann, G.E.; Schulz, R.; Wagner, S. Nanoparticles in the environment: Where do we come from, where do we go to? Environ. Sci. Eur. 2018, 30, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Paramo, L.A.; Feregrino-Perez, A.A.; Guevara, R.; Mendoza, S.; Esquivel, K. Nanoparticles in agroindustry: Applications, toxicity, challenges, and trends. Nanomaterials 2020, 10, 1654. [Google Scholar] [CrossRef]
  114. Rastogi, A.; Zivcak, M.; Sytar, O.; Kalaji, H.M.; He, X.; Mbarki, S.; Brestic, M. Impact of metal and metal oxide nanoparticles on plant: A critical review. Front. Chem. 2017, 5, 78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Jasim, B.; Thomas, R.; Mathew, J.; Radhakrishnan, E.K. Plant growth and diosgenin enhancement effect of silver nanoparticles in Fenugreek (Trigonella foenum-graecum L.). Saudi Pharm. J. 2017, 25, 443–447. [Google Scholar] [CrossRef]
  116. Naik, P.M.; Al-Khayri, J.M. Abiotic and biotic elicitors–role in secondary metabolites production through in vitro culture of medicinal plants. In Abiotic and Biotic Stress in Plants—Recent Advances and Future Perspectives; Shanker, A., Shanker, C., Eds.; IntechOpen: Rijeka, Croatia, 2016; pp. 247–278. [Google Scholar]
  117. Javed, R.; Yucesan, B.; Zia, M.; Gurel, E. Elicitation of secondary metabolites in callus cultures of Stevia rebaudiana Bertoni grown under ZnO and CuO nanoparticles stress. Sugar Tech. 2018, 20, 194–201. [Google Scholar] [CrossRef]
  118. Kavianifar, S.; Ghodrati, K.; Naghdi Badi, H.; Etminan, A. Effects of nano elicitors on callus induction and mucilage production in tissue culture of Linum usitatissimum L. J. Med. Plant. 2018, 17, 45–54. [Google Scholar]
  119. Abraham, J.; Thomas, T.D. Hairy root culture for the production of useful secondary metabolites. In Biotechnology and Production of Anti-Cancer Compounds; Malik, S., Ed.; Springer: Cham, Switzerland, 2017; pp. 201–230. [Google Scholar]
  120. Dhiman, N.; Patial, V.; Bhattacharya, A. The current status and future applications of hairy root cultures. In Biotechnological Approaches for Medicinal and Aromatic Plants; Kumar, N., Ed.; Springer: Singapore, 2018; pp. 87–155. [Google Scholar]
  121. Chung, I.M.; Rekha, K.; Rajakumar, G.; Thiruvengadam, M. Elicitation of silver nanoparticles enhanced the secondary metabolites and pharmacological activities in cell suspension cultures of bitter gourd. 3 Biotech 2019, 8, 412. [Google Scholar] [CrossRef]
  122. Ramezannezhad, R.; Aghdasi, M.; Fatemi, M. Enhanced production of cichoric acid in cell suspension culture of Echinacea purpurea by silver nanoparticle elicitation. Plant. Cell Tissue Organ. Cult. 2019, 139, 261–273. [Google Scholar] [CrossRef]
  123. Karakas, O. Effect of silver nanoparticles on production of indole alkaloids in Isatis constricta. Iran. J. Sci. Technol. Trans. A Sci. 2020, 44, 621–627. [Google Scholar] [CrossRef]
  124. Ali, A.; Mohammad, S.; Khan, M.A.; Raja, N.I.; Arif, M.; Kamil, A.; Mashwani, Z.U.R. Silver nanoparticles elicited in vitro callus cultures for accumulation of biomass and secondary metabolites in Caralluma tuberculata. Artif. Cells Nanomed. Biotechnol. 2019, 47, 715–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Zahir, A.; Nadeem, M.; Ahmad, W.; Giglioli-Guivarc’h, N.; Hano, C.; Abbasi, B.H. Chemogenic silver nanoparticles enhance lignans and neolignans in cell suspension cultures of Linum usitatissimum L. Plant. Cell Tissue Organ. Cult. 2019, 136, 589–596. [Google Scholar] [CrossRef]
  126. Golkar, P.; Moradi, M.; Garousi, G.A. Elicitation of Stevia glycosides using salicylic acid and silver nanoparticles under callus culture. Sugar Tech. 2019, 21, 569–577. [Google Scholar] [CrossRef]
  127. Spinoso-Castillo, J.L.; Chavez-Santoscoy, R.A.; Bogdanchikova, N.; Perez-Sato, J.A.; Morales-Ramos, V.; Bello-Bello, J.J. Antimicrobial and hormetic effects of silver nanoparticles on in vitro regeneration of vanilla (Vanilla planifolia Jacks. ex Andrews) using a temporary immersion system. Plant. Cell Tissue Organ. Cult. 2017, 129, 195–207. [Google Scholar] [CrossRef]
  128. Farida, S.H.M.; Karamian, R.; Albrectsen, B.R. Silver nanoparticle pollutants activate oxidative stress responses and rosmarinic acid accumulation in sage. Physiol. Plant 2020, in press. [Google Scholar] [CrossRef]
  129. Fazal, H.; Abbasi, B.H.; Ahmad, N.; Ali, M.; Ali, S.S.; Khan, A.; Wei, D.Q. Sustainable production of biomass and industrially important secondary metabolites in cell cultures of selfheal (Prunella vulgaris L.) elicited by silver and gold nanoparticles. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2553–2561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Wesolowska, A.; Jadczak, P.; Kulpa, D.; Przewodowski, W. Gas chromatography-mass spectrometry (GC-MS) analysis of essential oils from AgNPs and AuNPs elicited Lavandula angustifolia in vitro cultures. Molecules 2019, 24, 606. [Google Scholar] [CrossRef] [Green Version]
  131. Hussain, M.; Raja, N.I.; Mashwani, Z.U.R.; Iqbal, M.; Sabir, S.; Yasmeen, F. In vitro seed germination and biochemical profiling of Artemisia absinthium exposed to various metallic nanoparticles. 3 Biotech 2017, 7, 101. [Google Scholar] [CrossRef] [Green Version]
  132. Sharifi-Rad, R.; Esmaeilzadeh, B.S.; Samzadeh-Kermani, A.; Gholami, M. The effect of non-biological elicitors on physiological and biochemical properties of medicinal plant Momordica charantia L. Iran. J. Sci. Technol. Trans. A Sci. 2020, 44, 1315–1326. [Google Scholar] [CrossRef]
  133. Lafmejani, Z.N.; Jafari, A.A.; Moradi, P.; Moghadam, A.L. Impact of foliar application of iron-chelate and iron nano particles on some morpho-physiological traits and essential oil composition of peppermint (Mentha piperita L.). J. Essent. Oil-Bear. Plants 2018, 21, 1374–1384. [Google Scholar] [CrossRef]
  134. Ghasemi, B.; Hosseini, R.; Dehghan Nayeri, F. Effects of cobalt nanoparticles on artemisinin production and gene expression in Artemisia annua. Turk. J. Bot. 2015, 39, 769–777. [Google Scholar] [CrossRef]
  135. Ahmad, M.A.; Javed, R.; Adeel, M.; Rizwan, M.; Ao, Q.; Yang, Y.S. Engineered ZnO and CuO nanoparticles ameliorate morphological and biochemical response in tissue culture regenerants of candyleaf (Stevia rebaudiana). Molecules 2020, 25, 1356. [Google Scholar] [CrossRef] [Green Version]
  136. Chung, I.M.; Rajakumar, G.; Subramanian, U.; Venkidasamy, B.; Thiruvengadam, M. Impact of copper oxide nanoparticles on enhancement of bioactive compounds using cell suspension cultures of Gymnema sylvestre (Retz.) R. Br. Appl. Sci. 2019, 9, 2165. [Google Scholar] [CrossRef] [Green Version]
  137. Javed, R.; Usman, M.; Yucesan, B.; Zia, M.; Gurel, E. Effect of zinc oxide (ZnO) nanoparticles on physiology and steviol glycosides production in micropropagated shoots of Stevia rebaudiana Bertoni. Plant. Physiol. Biochem. 2017, 110, 94–99. [Google Scholar] [CrossRef]
  138. Zaeem, A.; Drouet, S.; Anjum, S.; Khursihd, R.; Younas, M.; Blondeau, J.P.; Tungmunnithum, D.; Giglioli-Guivarc’h, N.; Hano, C.; Abbasi, B.H. Effects of biogenic zinc oxide nanoparticles on growth and oxidative stress response in flax seedlings vs. in vitro cultures: A comparative analysis. Biomolecules 2020, 10, 918. [Google Scholar] [CrossRef]
  139. Velazquez-Gamboa, M.C.; Rodriguez, H.L.; Abud-Archila, M.; Gutierrez-Miceli, F.A.; Gonzalez-Mendoza, D.; Valdez-Salas, B.; Gonzalez-Terreros, E.; Lujan-Hidalgo, M.C. Agronomic biofortification of Stevia rebaudiana with zinc oxide (ZnO) phytonanoparticles and antioxidant compounds. Sugar Tech. 2020, in press. [Google Scholar] [CrossRef]
  140. Karimzadeh, F.; Haddad, R.; Garoosi, G.; Khademian, R. Effects of nanoparticles on activity of lignan biosynthesis enzymes in cell suspension culture of Linum usitatissimum L. Russ. J. Plant. Physiol. 2019, 66, 756–762. [Google Scholar] [CrossRef]
  141. Sharafi, E.; Fotokian, M.H.; Davoodi, D.; Hasanloo, H.H. Improvement of hypericin and hyperforin production using zinc and iron nano-oxides as elicitors in cell suspension culture of St John’s wort (Hypericum perforatum L.). JMPB 2013, 2, 177–184. [Google Scholar]
  142. Mohasseli, V.; Farbood, F.; Moradi, A. Antioxidant defense and metabolic responses of lemon balm (Melissa officinalis L.) to Fe-nano-particles under reduced irrigation regimes. Ind. Crops Prod. 2020, 149, 112338. [Google Scholar] [CrossRef]
  143. Taghizadeh, M.; Nasibi, F.; Kalantari, K.M.; Ghanati, F. Evaluation of secondary metabolites and antioxidant activity in Dracocephalum polychaetum Bornm. cell suspension culture under magnetite nanoparticles and static magnetic field elicitation. Plant. Cell Tissue Organ. Cult. 2019, 136, 489–498. [Google Scholar] [CrossRef]
  144. Ahmad, B.; Shabbir, A.; Jaleel, H.; Masroor, M.; Khan, A.; Sadiq, Y. Efficacy of titanium dioxide nanoparticles in modulating photosynthesis, peltate glandular trichomes and essential oil production and quality in Mentha piperita L. Curr. Plant. Biol. 2018, 13, 6–15. [Google Scholar] [CrossRef]
  145. Shabbir, A.; Khan, M.M.A.; Ahmad, B.; Sadiq, Y.; Jaleel, H.; Uddin, M. Efficacy of TiO2 nanoparticles in enhancing the photosynthesis, essential oil and khusimol biosynthesis in Vetiveria zizanioides L. Nash. Photosynthetica 2019, 57, 599–606. [Google Scholar] [CrossRef] [Green Version]
  146. Ghorbanpour, M. Major essential oil constituents, total phenolics and flavonoids content and antioxidant activity of Salvia officinalis plant in response to nano-titanium dioxide. Indian J. Plant. Physiol. 2015, 20, 249–256. [Google Scholar] [CrossRef]
  147. Rezaizad, M.; Hashemi-Moghaddam, H.; Abbaspour, H.; Gerami, M.; Mueller, A. Photocatalytic effect of TiO2 nanoparticles on morphological and photochemical properties of stevia plant (Stevia rebaudiana Bertoni). Sugar Tech. 2019, 21, 1024–1030. [Google Scholar] [CrossRef]
  148. Modarresi, M.; Chahardoli, A.; Karimi, N.; Chahardoli, S. Variations of glaucine, quercetin and kaempferol contents in Nigella arvensis against Al2O3, NiO, and TiO2 nanoparticles. Heliyon 2020, 6, e04265. [Google Scholar] [CrossRef]
  149. Gohari, G.; Mohammadi, A.; Akbari, A.; Panahirad, S.; Dadpour, M.R.; Fotopoulos, V.; Kimura, S. Titanium dioxide nanoparticles (TiO2 NPs) promote growth and ameliorate salinity stress effects on essential oil profile and biochemical attributes of Dracocephalum moldavica. Sci. Rep. 2020, 10, 912. [Google Scholar] [CrossRef] [PubMed]
  150. Kamalizadeh, M.; Bihamta, M.; Zarei, A. Drought stress and TiO2 nanoparticles affect the composition of different active compounds in the Moldavian dragonhead plant. Acta Physiol. Plant. 2019, 41, 21. [Google Scholar] [CrossRef]
  151. Tian, H.; Ghorbanpour, M.; Kariman, K. Manganese oxide nanoparticle-induced changes in growth, redox reactions and elicitation of antioxidant metabolites in deadly nightshade (Atropa belladonna L.). Ind. Crops Prod. 2018, 126, 403–414. [Google Scholar] [CrossRef]
  152. Ahmad, B.; Khan, M.M.A.; Jaleel, H.; Shabbir, A.; Sadiq, Y.; Uddin, M. Silicon nanoparticles mediated increase in glandular trichomes and regulation of photosynthetic and quality attributes in Mentha piperita L. J. Plant. Growth Regul. 2020, 39, 346–357. [Google Scholar] [CrossRef]
  153. Torabzadeh, D.; Hassabpour, H.; Asgarpanah, J.; Rezayian, M. Nanoparticles induced antioxidative compounds in Matricaria chamomilla. Iran. J. Plant. Physiol. 2019, 9, 2955–2961. [Google Scholar]
  154. Ebadollahi, R.; Jafarirad, S.; Kosari-Nasab, M.; Mahjouri, S. Effect of explant source, perlite nanoparticles and TiO2/perlite nanocomposites on phytochemical composition of metabolites in callus cultures of Hypericum perforatum. Sci. Rep. 2019, 9, 12998. [Google Scholar] [CrossRef]
  155. Rahmani, N.; Radjabian, T.; Soltani, B.M. Impacts of foliar exposure to multi-walled carbon nanotubes on physiological and molecular traits of Salvia verticillata L., as a medicinal plant. Plant. Physiol. Biochem. 2020, 150, 27–38. [Google Scholar] [CrossRef] [PubMed]
  156. Ghasempour, M.; Iranbakhsh, A.; Ebadi, M.; Ardebili, Z.O. Multi-walled carbon nanotubes improved growth, anatomy, physiology, secondary metabolism, and callus performance in Catharanthus roseus: An in vitro study. 3 Biotech 2019, 9, 404. [Google Scholar] [CrossRef]
  157. Samadi, S.; Saharkhiz, M.J.; Azizi, M.; Samiei, L.; Ghorbanpour, M. Multi-walled carbon nanotubes stimulate growth, redox reactions and biosynthesis of antioxidant metabolites in Thymus daenensis celak. in vitro. Chemosphere 2020, 249, 126069. [Google Scholar] [CrossRef]
  158. Ghorbanpour, M.; Hadian, J. Multi-walled carbon nanotubes stimulate callus induction, secondary metabolites biosynthesis and antioxidant capacity in medicinal plant Satureja khuzestanica grown in vitro. Carbon 2015, 94, 749–759. [Google Scholar] [CrossRef]
  159. Heydari, H.R.; Chamani, E.; Esmaielpour, B. Cell line selection through gamma irradiation combined with multi-walled carbon nanotubes elicitation enhanced phenolic compounds accumulation in Salvia nemorosa cell culture. Plant. Cell Tissue Organ. Cult. 2020, 142, 353–367. [Google Scholar] [CrossRef]
  160. Ahmadi, S.Z.; Ghorbanpour, M.; Aghaee, A.; Hadian, J. Deciphering morpho-physiological and phytochemical attributes of Tanacetum parthenium L. plants exposed to C60 fullerene and salicylic acid. Chemosphere 2020, 259, 127406. [Google Scholar] [CrossRef]
  161. Kole, C.; Kole, P.; Randunu, K.M.; Choudhary, P.; Podila, R.; Ke, P.C.; Rao, A.M.; Marcus, R.K. Nanobiotechnology can boost crop production and quality: First evidence from increased plant biomass, fruit yield and phytomedicine content in bitter melon (Momordica charantia). BMC Biotechnol. 2013, 13, 37. [Google Scholar] [CrossRef] [Green Version]
  162. Nourozi, E.; Hosseini, B.; Maleki, R.; Mandoulakani, B.A. Iron oxide nanoparticles: A novel elicitor to enhance anticancer flavonoid production and gene expression in Dracocephalum kotschyi hairy-root cultures. J. Sci. Food Agric. 2019, 99, 6418–6430. [Google Scholar] [CrossRef] [PubMed]
  163. Nourozi, E.; Hosseini, B.; Maleki, R.; Mandoulakani, B.A. Pharmaceutical important phenolic compounds overproduction and gene expression analysis in Dracocephalum kotschyi hairy roots elicited by SiO2 nanoparticles. Ind. Crops Prod. 2019, 133, 435–446. [Google Scholar] [CrossRef]
  164. Zhang, B.; Zheng, L.P.; Li, W.Y.; Wang, J.W. Stimulation of artemisinin production in Artemisia annua hairy roots by Ag-SiO2 core-shell nanoparticles. Curr. Nanosci. 2013, 9, 363–370. [Google Scholar] [CrossRef]
  165. Shakeran, Z.; Keyhanfar, M.; Asghari, G.; Ghanadian, M. Improvement of atropine production by different biotic and abiotic elicitors in hairy root cultures of Datura metel. Turk. J. Biol. 2015, 39, 111–118. [Google Scholar] [CrossRef]
  166. Chung, I.M.; Rekha, K.; Rajakumar, G.; Thiruvengadam, M. Production of bioactive compounds and gene expression alterations in hairy root cultures of chinese cabbage elicited by copper oxide nanoparticles. Plant. Cell Tissue Organ. Cult. 2018, 134, 95–106. [Google Scholar] [CrossRef]
  167. Juarez-Maldonado, A.; Ortega-Ortiz, H.; Morales-Diaz, A.B.; Gonzalez-Morales, S.; Morelos-Moreno, A.; Cabrera-De la Fuente, M.; Sandoval-Rangel, A.; Cadenas-Pliego, G.; Benavides-Mendoza, A. Nanoparticles and nanomaterials as plant biostimulants. Int. J. Mol. Sci. 2019, 20, 162. [Google Scholar] [CrossRef] [Green Version]
  168. Wink, M. Plant breeding: Importance of plant secondary metabolites for protection against pathogens and herbivores. Theoret. Appl. Genet. 1988, 75, 225–233. [Google Scholar] [CrossRef]
  169. Isah, T. Stress and defense responses in plant secondary metabolites production. Biol. Res. 2019, 52, 39. [Google Scholar] [CrossRef] [Green Version]
  170. Delgoda, R.; Murray, J.E. Evolutionary Perspectives on the Role of Plant Secondary Metabolites. In Pharmacognosy: Fundamentals, Applications and Strategies; Badal, S., Delgoda, R., Eds.; Academic Press & Elsevier: London, UK, 2017; pp. 93–100. [Google Scholar]
  171. Mottaki, Z.; Rezayian, M.; Niknam, V.; Ebrahimzadeh, H.; Mirmasoumi, M. Using hairy roots for production of secondary metabolites in Artemisia. Plant. Biotechnol. Rep. 2019, 13, 263–271. [Google Scholar] [CrossRef]
  172. Mosavat, N.; Golkar, P.; Yousefifard, M.; Javed, R. Modulation of callus growth and secondary metabolites in different Thymus species and Zataria multiflora micropropagated under ZnO nanoparticles stress. Biotechnol. Appl. Biochem. 2019, 66, 316–322. [Google Scholar] [CrossRef]
  173. Ramirez-Estrada, K.; Vidal-Limon, H.; Hidalgo, D.; Moyano, E.; Golenioswki, M.; Cusidó, R.M.; Palazon, J. Elicitation, an effective strategy for the biotechnological production of bioactive high-added value compounds in plant cell factories. Molecules 2016, 21, 182. [Google Scholar] [CrossRef]
  174. Pistelli, L.; Giovannini, A.; Ruffoni, B.; Bertoli, A.; Pistelli, L. Hairy root cultures for secondary metabolites production. Adv. Exp. Med. Biol. 2010, 698, 167–184. [Google Scholar]
  175. Sharma, P.; Radh, H.; Shrivastava, N. Hairy root cultures: A suitable biological system for studying secondary metabolic pathways in plants. Eng. Life Sci. 2013, 13, 62–75. [Google Scholar] [CrossRef]
  176. Häkkinen, S.T.; Moyano, E.; Cusido, R.M.; Oksman-Caldentey, K.M. Exploring the Metabolic stability of engineered hairy roots after 16 years maintenance. Front. Plant. Sci. 2016, 7, 1486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Hedayati, A.; Hosseini, B.; Palazon, J.; Maleki, R. Improved tropane alkaloid production and changes in gene expression in hairy root cultures of two Hyoscyamus species elicited by silicon dioxide nanoparticles. Plant. Physiol. Biochem. 2020, 155, 416–428. [Google Scholar] [CrossRef]
  178. Tkalec, M.; Stefanic, P.P.; Balen, B. Phytotoxicity of silver nanoparticles and defence mechanisms. Compr. Anal. Chem. 2019, 84, 145–198. [Google Scholar]
  179. Khalaki, M.A.; Ghorbani, A.; Moameri, M. Effects of silica and silver nanoparticles on seed germination traits of Thymus kotschyanus in laboratory conditions. J. Rangel. Sci. 2016, 6, 221–231. [Google Scholar]
  180. Darvishzadeh, F.; Nejatzadeh, F.; Iranbakhsh, A.R. Effects of silver nanoparticles on salinity tolerance in basil plant (Ocimum basilicum L.) during germination in vitro. New Cell. Nilecular Biotechnol. J. 2015, 5, 63–70. [Google Scholar]
  181. Jadczak, P.; Kulpa, D.; Bihun, M.; Przewodowski, W. Positive effect of AgNPs and AuNPs in in vitro cultures of Lavandula angustifolia Mill. Plant. Cell Tissue Organ. Cult. 2019, 139, 191–197. [Google Scholar] [CrossRef] [Green Version]
  182. Sahandi, S.M.; Sorooshzadeh, A.; Rezazadeh, H.S.; Naghdibadi, H.A. Effect of nano silver and silver nitrate on seed yield of borage. J. Med. Plant. Res. 2011, 5, 706–710. [Google Scholar]
  183. Seifsahandi, M.; Sorooshzadeh, A. Comparison between the influences of silver nanoparticles and silver nitrate on the growth and phytochemical properties of borage (Borago officinalis L). Curr. Nanosci. 2013, 9, 241–247. [Google Scholar] [CrossRef]
  184. Pastelin-Solano, M.C.; Ramirez-Mosqueda, M.A.; Bogdanchikova, N.; Castro-Conzalez, C.G.; Bello-Bello, J.J. Silver nanoparticles affect the micropropagation of vanilla (Vanilla planifolia Jacks. ex Andrews). Agrociencia 2020, 54, 1–13. [Google Scholar]
  185. Timoteo, C.D.; Paiva, R.; dos Reis, M.V.; Claro, P.I.C.; da Silva, D.P.C.; Marconcini, J.M.; de Oliveira, J.E. Silver nanoparticles in the micropropagation of Campomanesia rufa (O. Berg) Nied. Plant. Cell Tissue Organ. Cult. 2019, 137, 359–368. [Google Scholar] [CrossRef]
  186. Chung, I.M.; Rekha, K.; Rajakumar, G.; Thiruvengadam, M. Influence of silver nanoparticles on the enhancement and transcriptional changes of glucosinolates and phenolic compounds in genetically transformed root cultures of Brassica rapa ssp rapa. Bioprocess. Biosyst. Eng. 2018, 41, 1665–1677. [Google Scholar] [CrossRef] [PubMed]
  187. Sadak, M.S. Impact of silver nanoparticles on plant growth, some biochemical aspects, and yield of fenugreek plant (Trigonella foenum-graecum). Bull. Natl. Res. Cent. 2019, 43, 38. [Google Scholar] [CrossRef]
  188. Soltanabad, M.H.; Bagherieh-Najjar, M.B.; Mianabadi, M. Carnosic acid content increased by silver nanoparticle treatment in rosemary (Rosmarinus officinalis L.). Appl. Biochem. Biotechnol. 2020, 191, 482–495. [Google Scholar] [CrossRef]
  189. Anum, F.; Raja, N.I.; Hussain, M.; Iqbal, M.; Chaudhari, S.K.; Ehsan, M.; Javaid, U.; Zafar, N.U. Effect of green synthesised silver nanoparticles on morphogenic and biochemical variations in callus cultures of kinnow mandarin (Citrus reticulata L.). IET Nanobiotechnol. 2019, 13, 541–545. [Google Scholar] [CrossRef]
  190. Saha, N.; Gupta, S.D. Promotion of shoot regeneration of Swertia chirata by biosynthesized silver nanoparticles and their involvement in ethylene interceptions and activation of antioxidant activity. Plant. Cell Tissue Organ. Cult. 2018, 134, 289–300. [Google Scholar] [CrossRef]
  191. Jalilzadeh, E.; Jamei, E.; Sarghein, S.H. Magnetic field and silver nanoparticles induced changes on phenolic compound and oxidative status of marigold seedlings. J. Plant. Physiol. Breed. 2018, 8, 75–88. [Google Scholar]
  192. Ghanati, F.; Bakhtiarian, S. Effect of methyl jasmonate and silver nanoparticles on production of secondary metabolites by Calendula officinalis L (Asteraceae). Trop. J. Pharm. Res. 2014, 13, 1783–1789. [Google Scholar] [CrossRef] [Green Version]
  193. Lawal, A.T.; Azeez, L.; Sulaiman, W.K. Silver nanoparticles (AgNPs) alleviate naphthalene-triggered oxidative stress and physiological deficiencies in Moringa oleifera. Chem. Ecol. 2020, in press. [Google Scholar] [CrossRef]
  194. Cvjetko, P.; Milosic, A.; Domijan, A.M.; Vrcek, I.V.; Tolic, S.; Stefanic, P.P.; Letofsky-Papst, I.; Tkalec, M.; Balen, B. Toxicity of silver ions and differently coated silver nanoparticles in Allium cepa roots. Ecotox. Environ. Saf. 2017, 137, 18–28. [Google Scholar] [CrossRef]
  195. Ghavam, M. Effect of silver nanoparticles on tolerance to drought stress in Thymus daenensis Celak and Thymus vulgaris L. in germination and early growth stages. Environ. Stress. Crop. Sci. 2019, 12, 555–566. [Google Scholar]
  196. Kumari, M.; Mukherjee, A.; Chandrasekaran, N. Genotoxicity of silver nanoparticles in Allium cepa. Sci. Total Environ. 2009, 407, 5243–5246. [Google Scholar] [CrossRef]
  197. Hafez, R.M.; Fouad, A.S. Mitigation of genotoxic and cytotoxic effects of silver nanoparticles on onion root tips using some antioxidant scavengers. Egypt. J. Bot. 2020, 60, 133–145. [Google Scholar] [CrossRef]
  198. Durairaj, K.; Roy, B.; Chandrasekaran, N.; Krishnan, S.P.; Mukherjee, A. Silver nanorods induced oxidative stress and chromosomal aberrations in the Allium cepa model. IET Nanobiotechnol. 2020, 14, 161–166. [Google Scholar] [CrossRef] [PubMed]
  199. Souza, I.R.; Silva, L.R.; Fernandes, L.S.P.; Salgado, L.D.; Silva, H.C.D.; Firak, D.S.; Bach, L.; Santos-Filho, R.; Voigt, C.L.; Barros, A.C. Visible-light reduced silver nanoparticles’ toxicity in Allium cepa test system. Environ. Pollut. 2020, 257, 113551. [Google Scholar] [CrossRef]
  200. Bello-Bello, J.J.; Spinoso-Castillo, J.L.; Arano-Avalos, S.; Martinez-Estrada, E.; Arellano-Garcia, M.E.; Pestryakov, A.; Toledano-Magana, Y.; Garcia-Ramos, J.C.; Bogdanchikova, N. Cytotoxic, genotoxic, and polymorphism effects on Vanilla planifolia Jacks ex Andrews after long-term exposure to Argovit® silver nanoparticles. Nanomaterials 2018, 8, 754. [Google Scholar] [CrossRef] [Green Version]
  201. Rawat, S.; Adisa, I.O.; Wang, Y.; Sun, Y.P.; Fadil, A.S.; Niu, G.H.; Sharma, N.; Hernandez-Viezcas, J.A.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Differential physiological and biochemical impacts of nano vs micron Cu at two phenological growth stages in bell pepper (Capsicum annuum) plant. Nanoimpact 2019, 14, 100161. [Google Scholar] [CrossRef]
  202. Ghidan, A.Y.; Al-Antary, T.M.; Awwad, A.M.; Ayad, J.Y. Physiological effect of some nanomaterials on pepper (Capsicum annuum L.) plants. Fresen. Environ. Bull. 2018, 27, 7872–7878. [Google Scholar]
  203. Rawat, S.; Pullagurala, V.L.R.; Hernandez-Molina, M.; Sun, Y.P.; Niu, G.H.; Hernandez-Viezcas, J.A.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Impacts of copper oxide nanoparticles on bell pepper (Capsicum annum L.) plants: A full life cycle study. Environ. Sci. Nano 2018, 5, 83–95. [Google Scholar] [CrossRef]
  204. Talankova-Sereda, T.E.; Liapina, K.V.; Shkopinskij, E.A.; Ustinov, A.I.; Kovalyova, A.V.; Dulnev, P.G.; Kucenko, N.I. The influence of Cu and Co nanoparticles on growth characteristics and biochemical structure of Mentha longifolia in vitro. Nanosci. Nanoeng. 2016, 4, 31–39. [Google Scholar] [CrossRef]
  205. Ibrahim, A.S.; Fahmy, A.H.; Ahmed, S.S. Copper nanoparticles elevate regeneration capacity of (Ocimum basilicum L.) plant via somatic embryogenesis. Plant. Cell Tissue Organ. Cult. 2019, 136, 41–50. [Google Scholar] [CrossRef]
  206. Du, W.C.; Tan, W.J.; Yin, Y.; Ji, R.; Peralta-Videa, J.R.; Guo, H.Y.; Gardea-Torresdey, J.L. Differential effects of copper nanoparticles/microparticles in agronomic and physiological parameters of oregano (Origanum vulgare). Sci. Total Environ. 2018, 618, 306–312. [Google Scholar] [CrossRef] [PubMed]
  207. AlQuraidi, A.O.; Mosa, K.A.; Ramamoorthy, K. Phytotoxic and genotoxic effects of copper nanoparticles in coriander (Coriandrum sativum-Apiaceae). Plants 2019, 8, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Wang, Y.; Deng, C.Y.; Cota-Ruiz, K.; Peralta-Videa, J.R.; Sun, Y.P.; Rawat, S.; Tan, W.J.; Reyes, A.; Hernandez-Viezcas, J.A.; Niu, G.H.; et al. Improvement of nutrient elements and allicin content in green onion (Allium fistulosum) plants exposed to CuO nanoparticles. Sci. Total Environ. 2020, 725, 138387. [Google Scholar] [CrossRef]
  209. Javed, R.; Mohamed, A.; Yucesan, B.; Gurel, E.; Kausar, R.; Zia, M. CuO nanoparticles significantly influence in vitro culture, steviol glycosides, and antioxidant activities of Stevia rebaudiana Bertoni. Plant. Cell Tissue Organ. Cult. 2017, 131, 611–620. [Google Scholar] [CrossRef]
  210. Ghazal, B.; Saif, S.; Farid, K.; Khan, A.; Rehman, S.; Reshma, A.; Fazal, H.; Ali, M.; Ahmad, A.; Rahman, L.; et al. Stimulation of secondary metabolites by copper and gold nanoparticles in submerge adventitious root cultures of Stevia rebaudiana (Bert.). IET Nanobiotechnol. 2018, 12, 569–573. [Google Scholar] [CrossRef]
  211. Javed, R.; Zia, M.; Yucesan, B.; Gurel, E. Abiotic stress of ZnO-PEG, ZnO-PVP, CuO-PEG and CuO-PVP nanoparticles enhance growth, sweetener compounds and antioxidant activities in shoots of Stevia rebaudiana Bertoni. IET Nanobiotechnol. 2017, 11, 898–902. [Google Scholar] [CrossRef]
  212. Lala, S. Enhancement of secondary metabolites in Bacopa monnieri (L.) Pennell plants treated with copper-based nanoparticles in vivo. IET Nanobiotechnol. 2020, 14, 78–85. [Google Scholar] [CrossRef]
  213. Pramanik, A.; Datta, A.K.; Das, D.; Kumbhakar, D.V.; Ghosh, B.; Mandal, A.; Gupta, S.; Saha, A.; Sengupta, S. Assessment of nanotoxicity (cadmium sulphide and copper oxide) using cytogenetical parameters in Coriandrum sativum L. (Apiaceae). Cytol. Genet. 2018, 52, 299–308. [Google Scholar] [CrossRef]
  214. Bahrami, M.K.; Movafeghi, A.; Mahdavinia, G.R.; Hassanpouraghdam, M.B.; Gohari, G. Effects of bare and chitosan-coated Fe3O4 magnetic nanoparticles on seed germination and seedling growth of Capsicum annuum L. Biointerface Res. Appl. Chem. 2018, 8, 3552–3559. [Google Scholar]
  215. Jangir, H.; Das, C.K.; Kumar, J.; Mahapatra, S.S.; Srivastava, G.; Bhardwaj, A.; Das, M. Nano pyrite (FeS2) root priming enhances chilli and marigold production in nutrients-deficient soil: A nano strategy for fertiliser tuning. Appl. Nanosci. 2019, 9, 327–340. [Google Scholar] [CrossRef]
  216. Nechitailo, G.S.; Bogoslovskaya, O.A.; Ol’khovskaya, I.P.; Glushchenko, N.N. Influence of iron, zinc, and copper nanoparticles on some growth indices of pepper plants. Nanotechnol. Russ. 2018, 13, 161–167. [Google Scholar] [CrossRef]
  217. Siva, G.V.; Benita, L.F.J. Iron oxide nanoparticles promotes agronomic traits of ginger (Zingiber officinale Rosc.). Int. J. Adv. Res. Biol. Sci. 2016, 3, 230–237. [Google Scholar]
  218. Sabet, H.; Mortazaeinezhad, F. Yield, growth and Fe uptake of cumin (Cuminum cyminum L.) affected by Fe-nano, Fe-chelated and Fe-siderophore fertilization in the calcareous soils. J. Trace Elem. Med. Biol. 2018, 50, 154–160. [Google Scholar] [CrossRef] [PubMed]
  219. Mohebodini, M.; Fathi, R.; Mehri, N. Optimization of hairy root induction in chicory (Cichorium intybus L.) and effects of nanoparticles on secondary metabolites accumulation. Iran. J. Genet. Plant. Breed. 2017, 6, 60–68. [Google Scholar]
  220. Khan, M.A.; Ali, A.; Mohammad, S.; Ali, H.; Khan, T.; Mashwani, Z.U.R.; Jan, A.; Ahmad, P. Iron nano modulated growth and biosynthesis of steviol glycosides in Stevia rebaudiana. Plant. Cell Tissue Organ. Cult. 2020, 143, 121–130. [Google Scholar] [CrossRef]
  221. Kiapour, H.; Moaveni, P.; Sani, B.; Rajabzadeh, F.; Mozafari, H. Investigating the effect of magnesium and iron oxide nanoparticles on the levels of enzymatic and non-enzymatic antioxidants in roselle. JMPB 2020, 9, 19–31. [Google Scholar]
  222. Tavallali, V.; Kiani, M.; Hojati, S. Iron nano-complexes and iron chelate improve biological activities of sweet basil (Ocimum basilicum L.). Plant. Physiol. Biochem. 2019, 144, 445–454. [Google Scholar] [CrossRef]
  223. Tavallali, V.; Rowshan, V.; Gholami, H.; Hojati, S. Iron-urea nano-complex improves bioactive compounds in essential oils of Ocimum basilicum L. Sci. Hortic. 2020, 265, 109222. [Google Scholar] [CrossRef]
  224. Moradbeygi, H.; Jamei, R.; Heidari, R.; Darvishzadeh, R. Investigating the enzymatic and non-enzymatic antioxidant defense by applying iron oxide nanoparticles in Dracocephalum moldavica L. plant under salinity stress. Sci. Hortic. 2020, 272, 109537. [Google Scholar] [CrossRef]
  225. Fahad; Balouch, A.; Agheem, M.H.; Memon, S.A.; Baloch, A.R.; Tunio, A.; Abdullah; Pato, A.H.; Jagirani, M.S.; Panah, P.; et al. Efficient mitigation of cadmium and lead toxicity in coriander plant utilizing magnetite (Fe3O4) nanofertilizer as growth regulator and antimicrobial agent. Int. J. Environ. Anal. Chem. 2020, in press. [Google Scholar] [CrossRef]
  226. Askary, M.; Talebi, S.M.; Amini, F.; Bangan, A.D.B. Effects of iron nanoparticles on Mentha piperita L. under salinity stress. Biologia 2017, 63, 65–75. [Google Scholar] [CrossRef]
  227. Ghosh, I.; Mukherjee, A.; Mukherjee, A. In planta genotoxicity of nZVI: Influence of colloidal stability on uptake, DNA damage, oxidative stress and cell death. Mutagenesis 2017, 32, 371–387. [Google Scholar] [CrossRef]
  228. Younes, N.A.; Hassan, H.S.; Elkady, M.F.; Hamed, A.M.; Dawood, M.F.A. Impact of synthesized metal oxide nanomaterials on seedlings production of three Solanaceae crops. Heliyon 2020, 6, e03188. [Google Scholar] [CrossRef] [Green Version]
  229. Garcia-Lopez, J.I.; Zavala-Garcia, F.; Olivares-Saenz, E.; Lira-Saldivar, R.H.; Diaz Barriga-Castro, E.; Ruiz-Torres, N.A.; Ramos-Cortez, E.; Vazquez-Alvarado, R.; Nino-Medina, G. Zinc oxide nanoparticles boosts phenolic compounds and antioxidant activity of Capsicum annuum L. during germination. Agronomy 2018, 8, 215. [Google Scholar] [CrossRef] [Green Version]
  230. Garcia-Lopez, J.I.; Nino-Medina, G.; Olivares-Saenz, E.; Lira-Saldivar, R.H.; Barriga-Castro, E.D.; Vazquez-Alvarado, R.; Rodriguez-Salinas, P.A.; Zavala-Garcia, F. Foliar application of zinc oxide nanoparticles and zinc sulfate boosts the content of bioactive compounds in habanero peppers. Plants 2019, 8, 254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Gokak, I.B.; Taranath, T.C. Morphological and biochemical responses of Abelmoschus esculantus (L.) Moench to zinc nanoparticles. Adv. Nat. Sci. Nanosci. 2015, 6, 025017. [Google Scholar] [CrossRef] [Green Version]
  232. Burman, U.; Saini, M.; Kumar, P. Effect of zinc oxide nanoparticles on growth and antioxidant system of chickpea seedlings. Toxicol. Environ. Chem. 2013, 95, 605–612. [Google Scholar] [CrossRef]
  233. Lee, S.; Kim, S.; Kim, S.; Lee, I. Assessment of phytotoxicity of ZnO NPs on a medicinal plant, Fagopyrum esculentum. Environ. Sci. Pollut. Res. 2013, 20, 848–854. [Google Scholar] [CrossRef] [Green Version]
  234. Karimi, N.; Behbahani, M.; Dini, G.; Razmjou, A. Enhancing the secondary metabolite and anticancer activity of Echinacea purpurea callus extracts by treatment with biosynthesized ZnO nanoparticles. Adv. Nat. Sci. Nanosci. 2018, 9, 045009. [Google Scholar] [CrossRef]
  235. Shahhoseini, R.; Azizi, M.; Asili, J.; Moshtaghi, N.; Samiei, L. Effects of zinc oxide nanoelicitors on yield, secondary metabolites, zinc and iron absorption of Feverfew (Tanacetum parthenium (L.) Schultz Bip.). Acta Physiol. Plant. 2020, 42, 52. [Google Scholar] [CrossRef]
  236. Linh, N.T.N.; Tung, H.T.; Hien, V.T.; Luan, V.Q.; Huy, N.P.; Loc, N.H.; Nhut, D.T. Effect of metal nanoparticles on the growth of Ngoc Linh ginseng (Panax vietnamensis Ha et Grushv.) lateral roots cultured in vitro. HU JOS 2017, 126, 47–56. [Google Scholar]
  237. Radi, A.A.; Farghaly, F.A.; Al-Kahtany, F.A.; Hamada, A.M. Zinc oxide nanoparticles-mediated changes in ultrastructure and macromolecules of pomegranate callus cells. Plant. Cell Tissue Organ. Cult. 2018, 135, 247–261. [Google Scholar] [CrossRef]
  238. Desai, C.V.; Desai, H.B.; Suthar, K.P.; Singh, D.; Patel, R.M.; Taslim, A. Phytotoxicity of zinc-nanoparticles and its influence on stevioside production in Stevia rebaudiana Bertoni. Appl. Biol. Res. 2015, 17, 1–7. [Google Scholar] [CrossRef]
  239. Oloumi, H.; Soltaninejad, R.; Baghizadeh, A. The comparative effects of nano and bulk size particles of CuO and ZnO on glycyrrhizin and phenolic compounds contents in Glycyrrhiza glabra L. seedlings. Indian J. Plant. Physiol. 2015, 20, 157–161. [Google Scholar] [CrossRef]
  240. Nadernejad, N. Investigation of toxicity of zinc oxide nanoparticles synthesized by olive extract on growth and pigments in Borago officinalis. J. Plant. Process. Funct. 2019, 8, 303–315. [Google Scholar]
  241. Sun, Z.Q.; Xiong, T.T.; Zhang, T.; Wang, N.F.; Chen, D.; Li, S.S. Influences of zinc oxide nanoparticles on Allium cepa root cells and the primary cause of phytotoxicity. Ecotoxicology 2019, 28, 175–188. [Google Scholar] [CrossRef]
  242. Ghosh, M.; Jana, A.; Sinha, S.; Jothiramajayam, M.; Nag, A.; Chakraborty, A.; Mukherjee, A.; Mukherjee, A. Effects of ZnO nanoparticles in plants: Cytotoxicity, genotoxicity, deregulation of antioxidant defenses, and cell-cycle arrest. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2016, 807, 25–32. [Google Scholar] [CrossRef]
  243. Ahmed, B.; Dwivedi, S.; Abdin, M.Z.; Azam, A.; Al-Shaeri, M.; Khan, M.S.; Saquib, Q.; Al-Khedhairy, A.A.; Musarrat, J. Mitochondrial and chromosomal damage induced by oxidative stress in Zn2+ ions, ZnO-bulk and ZnO NPs treated Allium cepa roots. Sci. Rep. 2017, 7, 40685. [Google Scholar] [CrossRef]
  244. Kumari, M.; Khan, S.S.; Pakrashi, S.; Mukherjee, A.; Chandrasekaran, N. Cytogenetic and genotoxic effects of zinc oxide nanoparticles on root cells of Allium cepa. J. Hazard. Mater. 2011, 190, 613–621. [Google Scholar] [CrossRef]
  245. Debnath, P.; Mondal, A.; Sen, K.; Mishra, D.; Mondal, N.K. Genotoxicity study of nano Al2O3, TiO2 and ZnO along with UV-B exposure: An Allium cepa root tip assay. Sci. Total Environ. 2020, 713, 136592. [Google Scholar] [CrossRef] [PubMed]
  246. Tighe-Neira, R.; Carmora, E.; Recio, G.; Nunes-Nesi, A.; Reyes-Diaz, M.; Alberdi, M.; Rengel, Z.; Inostroza-Blancheteau, C. Metallic nanoparticles influence the structure and function of the photosynthetic apparatus in plants. Plant. Physiol. Biochem. 2018, 130, 408–417. [Google Scholar] [CrossRef] [PubMed]
  247. Hatami, M.; Ghorbanpour, M.; Salehiarjomand, H. Nano-anatase TiO2 modulates the germination behavior and seedling vigority of some commercially important medicinal and aromatic plants. J. Biol. Environ. Sci. 2014, 8, 5359. [Google Scholar]
  248. Dehkourdi, E.H.; Mosavi, M. Effect of anatase nanoparticles (TiO2) on parsley seed germination (Petroselinum crispum) in vitro. Biol. Trace Elem. Res. 2013, 155, 283–286. [Google Scholar] [CrossRef] [PubMed]
  249. Feizi, H.; Kamali, M.; Jafari, L.; Moghaddam, P.R. Phytotoxicity and stimulatory impacts of nanosized and bulk titanium dioxide on fennel (Foeniculum vulgare Mill). Chemosphere 2013, 91, 506–511. [Google Scholar] [CrossRef]
  250. Khater, M.S.; Osman, Y.A.H. Influence of TiO2 nanoparticles on growth, chemical constituents and toxicity of fennel plant. AJNSA 2015, 48, 178–186. [Google Scholar]
  251. Motyka, O.; Strbova, K.; Zinicovscaia, I. Chlorophyll content in two medicinal plant species following nano-TiO2 exposure. Bull. Environ. Contam. Toxicol. 2020, 104, 373–379. [Google Scholar] [CrossRef]
  252. Morteza, E.; Moaveni, P.; Morteza, T.; Saemi, H.; Joorabloo, A. Effects of TiO2 (nano and bulk) foliar application on physiological traits and grain yield of Safflower (Carthamus tinctorius L.). Biol. Forum 2015, 7, 1725–1731. [Google Scholar]
  253. Ogunkunle, C.O.; Adegboye, E.F.; Okoro, H.K.; Vishwakarma, V.; Alagarsamy, K.; Fatoba, P.O. Effect of nanosized anatase TiO2 on germination, stress defense enzymes, and fruit nutritional quality of Abelmoschus esculentus (L.) Moench (okra). Arab. J. Geosci. 2020, 13, 120. [Google Scholar] [CrossRef]
  254. Hu, J.; Wu, X.Y.; Wu, F.; Chen, W.X.; White, J.C.; Yang, Y.; Wang, B.; Xing, B.S.; Tao, S.; Wang, X.L. Potential application of titanium dioxide nanoparticles to improve the nutritional quality of coriander (Coriandrum sativum L.). J. Hazard. Mater. 2020, 389, 121837. [Google Scholar] [CrossRef] [PubMed]
  255. Tan, W.J.; Du, W.C.; Darrouzet-Nardi, A.J.; Hernandez-Viezcas, J.A.; Ye, Y.Q.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Effects of the exposure of TiO 2 nanoparticles on basil (Ocimum basilicum) for two generations. Sci. Total Environ. 2018, 636, 240–248. [Google Scholar] [CrossRef]
  256. Morteza, E.; Morteza, T.; Saemi, H.; Joorabloo, A. Effect of nano titanium dioxide spraying on essential oil yield and traits of cumin (Cuminum cyminum L.). Biol. Environ. Sci. 2016, 8, 101–108. [Google Scholar]
  257. Kahila, M.M.H.; Najy, A.M.; Rahaie, M.; Mir-Derikvand, M. Effect of nanoparticle treatment on expression of a key gene involved in thymoquinone biosynthetic pathway in Nigella sativa L. Nat. Prod. Res. 2018, 32, 1858–1862. [Google Scholar] [CrossRef]
  258. Golami, A.; Abbaspour, H.; Hashemi-Moghaddam, H.; Gerami, M. Photocatalytic effect of TiO2 nanoparticles on essential oil of Rosmarinus officinalis. J. Biochem. Technol. 2018, 9, 50–56. [Google Scholar]
  259. Khajavi, M.; Rahaie, M.; Ebrahimi, A. The effect of TiO2 and SiO2 nanoparticles and salinity stress on expression of genes involved in parthenolide biosynthesis in Feverfew (Tanacetum parthenium L.). Caryologia 2019, 72, 3–14. [Google Scholar]
  260. Karamian, R.; Ghasemlou, F.; Amiri, H. Physiological evaluation of drought stress tolerance and recovery in Verbascum sinuatum plants treated with methyl jasmonate, salicylic acid and titanium dioxide nanoparticles. Plant. Biosyst. 2020, 154, 277–287. [Google Scholar] [CrossRef]
  261. Mohammadi, H.; Esmailpour, M.; Gheranpaye, A. Effects of TiO2 nanoparticles and water-deficit stress on morpho-physiological characteristics of dragonhead (Dracocephalum moldavica L.) plants. Acta Agric. Slov. 2016, 107, 385–396. [Google Scholar] [CrossRef] [Green Version]
  262. Fazeli-Nasab, B.; Sirousmehr, A.R.; Azad, H. Effect of titanium dioxide nanoparticles on essential oil quantity and quality in Thymus vulgaris under water deficit. JMPB 2018, 7, 125–133. [Google Scholar]
  263. Ghabel, V.K.; Karamian, R. Effects of TiO2 nanoparticles and spermine on antioxidant responses of Glycyrrhiza glabra L. to cold stress. Acta Bot. Croat. 2020, 79, 137–147. [Google Scholar] [CrossRef]
  264. Mohammadi, R.; Maali-Amiri, R.; Abbasi, A. Effect of TiO2 nanoparticles on chickpea response to cold stress. Biol. Trace Elem. Res. 2013, 152, 403–410. [Google Scholar] [CrossRef]
  265. Rikabad, M.M.; Pourakbar, L.; Moghaddam, S.S.; Popovic-Djordjevic, J. Agrobiological, chemical and antioxidant properties of saffron (Crocus sativus L.) exposed to TiO2 nanoparticles and ultraviolet-B stress. Ind. Crops Prod. 2019, 137, 137–143. [Google Scholar] [CrossRef]
  266. Aghdam, M.T.B.; Mohammadi, H.; Ghorbanpour, M. Effects of nanoparticulate anatase titanium dioxide on physiological and biochemical performance of Linum usitatissimum (Linaceae) under well-watered and drought stress conditions. Braz. J. Bot. 2016, 39, 139–146. [Google Scholar] [CrossRef]
  267. Pakrashi, S.; Jain, N.; Dalai, S.; Jayakumar, J.; Chandrasekaran, P.T.; Raichur, A.M.; Chandrasekaran, N.; Mukherjee, A. In vivo genotoxicity assessment of titanium dioxide nanoparticles by Allium cepa root tip assay at high exposure concentrations. PLoS ONE 2014, 9, e87789. [Google Scholar] [CrossRef]
  268. Ahmed, B.; Shahid, M.; Khan, M.S.; Musarrat, J. Chromosomal aberrations, cell suppression and oxidative stress generation induced by metal oxide nanoparticles in onion (Allium cepa) bulb. Metallomics 2018, 10, 1315–1327. [Google Scholar] [CrossRef] [PubMed]
  269. Demir, E.; Kaya, N.; Kaya, B. Genotoxic effects of zinc oxide and titanium dioxide nanoparticles on root meristem cells of Allium cepa by comet assay. Turk. J. Biol. 2014, 38, 31–39. [Google Scholar] [CrossRef]
  270. Rezaei, Z.; Jafarirad, S.; Kosari-Nasab, M. Modulation of secondary metabolite profiles by biologically synthesized MgO/perlite nanocomposites in Melissa officinalis plant organ cultures. J. Hazard. Mater. 2019, 380, 120878. [Google Scholar] [CrossRef] [PubMed]
  271. Ye, Y.Q.; Medina-Velo, I.A.; Cota-Ruiz, K.; Moreno-Olivas, F.; Gardea-Torresdey, J.L. Can abiotic stresses in plants be alleviated by manganese nanoparticles or compounds? Ecotoxicol. Environ. Saf. 2019, 184, 109671. [Google Scholar] [CrossRef]
  272. Mangalampalli, B.; Dumala, N.; Grover, P. Allium cepa root tip assay in assessment of toxicity o magnesium oxide nanoparticles and microparticles. J. Environ. Sci. 2018, 66, 125–137. [Google Scholar] [CrossRef] [PubMed]
  273. Miri, A.H.; Shakib, E.S.; Ebrahimi, O.; Sharifi-Rad, J. Impacts of nickel nanoparticles on grow characteristics, photosynthetic pigment content and antioxidant activity of Coriandrum sativum L. Orient. J. Chem. 2017, 33, 1297–1303. [Google Scholar] [CrossRef] [Green Version]
  274. Baskar, V.; Safia, N.; Preethy, K.S.; Dhivya, S.; Thiruvengadam, M.; Sathishkumar, R. A comparative study of phytotoxic effects of metal oxide (CuO, ZnO and NiO) nanoparticles on in-vitro grown Abelmoschus esculentus. Plant. Biosyst. 2020, in press. [Google Scholar] [CrossRef]
  275. Manna, I.; Bandyopadhyay, M. Engineered nickel oxide nanoparticle causes substantial physicochemical perturbation in plants. Front. Chem. 2017, 5, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Kumar, D.; Rajeshwari, A.; Jadon, P.S.; Chaudhuri, G.; Mukherjee, A.; Chandrasekaran, N.; Mukherjee, A. Cytogenetic studies of chromium (III) oxide nanoparticles on Allium cepa root tip cells. J. Environ. Sci. 2015, 38, 150–157. [Google Scholar] [CrossRef]
  277. Liman, R. Genotoxic effects of Bismuth (III) oxide nanoparticles by Allium and Comet assay. Chemosphere 2013, 93, 269–273. [Google Scholar] [CrossRef] [PubMed]
  278. Morales, M.I.; Rico, C.M.; Hernandez-Viezcas, J.A.; Nunez, J.C.; Barrios, A.C.; Tafoy, A.; Flores-Marges, J.P.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Toxicity assessment of cerium oxide nanoparticles in cilantro (Coriandrum sativum L.) plants grown in organic soil. J. Agric. Food Chem. 2013, 61, 6224–6230. [Google Scholar] [CrossRef]
  279. Maqbool, Q.; Sidorowicz, A.; Nazar, M.; Jabeen, N.; Anwaar, S.; Ahmad, I. Organometallic cerium oxide nanostructures acted as nanofertilizer during callogenesis and organogenesis of recalcitrant plant Berberis lycium Royle. Mater. Today Commun. 2019, 20, 100544. [Google Scholar] [CrossRef]
  280. Liman, R.; Acikbas, Y.; Cigerci, I.H. Cytotoxicity and genotoxicity of cerium oxide micro and nanoparticles by Allium and Comet tests. Ecotoxicol. Environ. Saf. 2019, 168, 408–414. [Google Scholar] [CrossRef]
  281. De, A.; Chakrabarti, M.; Ghosh, I.; Mukherjee, A. Evaluation of genotoxicity and oxidative stress of aluminium oxide nanoparticles and its bulk form in Allium cepa. Nucleus 2016, 59, 219–225. [Google Scholar] [CrossRef]
  282. Kang, H.; Hwang, Y.G.; Lee, T.G.; Jin, C.R.; Cho, C.H.; Jeong, H.Y.; Kim, D.O. Use of gold nanoparticle fertilizer enhances the ginsenoside contents and anti-inflammatory effects of red ginseng. J. Microbiol. Biotechnol. 2016, 26, 26–1668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Epstein, E. Silicon: Its manifold roles in plants. Ann. Appl. Biol. 2009, 155, 155–160. [Google Scholar] [CrossRef]
  284. 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] [PubMed] [Green Version]
  285. Sharifi-Rad, J.; Sharifi-Rad, M.; da Silva, J.A.T. Morphological, physiological and biochemical responses of crops (Zea mays L., Phaseolus vulgaris L.), medicinal plants (Hyssopus officinalis L., Nigella sativa L.), and weeds (Amaranthus retroflexus L., Taraxacum officinale F. H. Wigg) exposed to SiO2 nanoparticles. J. Agric. Sci. Technol. 2016, 18, 1027–1040. [Google Scholar]
  286. Ashkavand, P.; Tabari, M.; Zarafshar, M.; Tomaskova, I.; Struve, D. Effect of SiO2 nanoparticles on drought resistance in hawthorn seedlings. For. Res. Pap. 2015, 76, 350–359. [Google Scholar] [CrossRef] [Green Version]
  287. Rahimi, S.; Hatami, M.; Ghorbanpour, M. Effect of seed priming with nanosilicon on morpho-physiological characteristics, quercetin content and antioxidant capacity in Calendula officinalis L. under drought stress conditions. J. Med. Plants 2020, 18, 186–203. [Google Scholar]
  288. Khan, M.A.; Wallace, W.T.; Sambi, J.; Rogers, D.T.; Littleton, J.M.; Rankin, S.E.; Knutson, B.L. Nanoharvesting of bioactive materials from living plant cultures using engineered silica nanoparticles. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 106, 110190. [Google Scholar] [CrossRef]
  289. Kalteh, M.; Alipour, Z.T.; Ashraf, S.; Aliabadi, M.M.; Nosratabadi, A.F. Effect of silica nanoparticles on basil (Ocimum basilicum) under salinity stress. J. Chem. Health Risks 2014, 4, 49–55. [Google Scholar]
  290. Koce, J.D.; Drobne, D.; Klancnik, K.; Makovec, D.; Novak, S.; Hocevar, M. Oxidative potential of ultraviolet-A irradiated or nonirradiated suspensions of titanium dioxide or silicon dioxide nanoparticles on Allium cepa roots. Environ. Toxicol. Chem. 2014, 33, 858–867. [Google Scholar] [CrossRef]
  291. Brasted, R.C. Selenium. Available online: https://www.britannica.com/science/selenium (accessed on 14 January 2021).
  292. Kurokawa, S.; Berry, M.J. Selenium. Role of the essential metalloid in health. Met. Ions Life Sci. 2013, 13, 499–534. [Google Scholar]
  293. Abedi, S.; Iranbakhsh, A.; Ardebili, Z.O.; Ebadi, M. Nitric oxide and selenium nanoparticles confer changes in growth, metabolism, antioxidant machinery, gene expression, and flowering in chicory (Cichorium intybus L.): Potential benefits and risk assessment. Environ. Sci. Pollut. Res. 2020, in press. [Google Scholar] [CrossRef]
  294. Sotoodehnia-Korani, S.; Iranbakhsh, A.; Ebadi, M.; Majd, A.; Ardebili, Z.O. Selenium nanoparticles induced variations in growth, morphology, anatomy, biochemistry, gene expression, and epigenetic DNA methylation in Capsicum annuum; an in vitro study. Environ. Pollut. 2020, 265, 114727. [Google Scholar] [CrossRef]
  295. Alamo-Nole, L.; Su, Y.F. Translocation of cadmium in Ocimum basilicum at low concentration of CdSSe nanoparticles. Appl. Mater. Today 2017, 9, 314–318. [Google Scholar] [CrossRef]
  296. Behbahani, S.R.; Iranbakhsh, A.; Ebadi, M.; Majd, A.; Ardebili, Z.O. Red elemental selenium nanoparticles mediated substantial variations in growth, tissue differentiation, metabolism, gene transcription, epigenetic cytosine DNA methylation, and callogenesis in bittermelon (Momordica charantia); an in vitro experiment. PLoS ONE 2020, 15, e0235556. [Google Scholar]
  297. Zaytseva, O.; Neumann, G. Carbon nanomaterials: Production, impact on plant development, agricultural and environmental applications. Chem. Biol. Technol. Agric. 2016, 3, 17. [Google Scholar] [CrossRef] [Green Version]
  298. Mukherjee, A.; Majumdar, S.; Servin, A.D.; Pagano, L.; Dhankher, O.P.; White, J.C. Carbon nanomaterials in agriculture: A critical review. Front. Plant. Sci. 2016, 7, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Mittal, J.; Osheen, S.; Gupta, A.; Kumar, R. Carbon nanomaterials in agriculture. In Nanoscience for Sustainable Agriculture; Pudake, R., Chauhan, N., Kole, C., Eds.; Springer: Cham, Switzerland, 2019; pp. 153–170. [Google Scholar]
  300. Crista, M.A.; da Silva, J.C.G.E.; da Silva, L.P. Evaluation of different bottom-up routes for the fabrication of carbon dots. Nanomaterials 2020, 10, 1316. [Google Scholar] [CrossRef]
  301. Zarzycki, P.K. Pure and Functionalized Carbon Based Nanomaterials: Analytical, Biomedical, Civil. and Environmental Engineering Applications; CRC Press: Boca Raton, FL, USA, 2020. [Google Scholar]
  302. Chandra, S.; Pradhan, S.; Mitra, S.; Patra, P.; Bhattacharya, A.; Pramanik, P.; Goswami, A. High throughput electron transfer from carbon dots to chloroplast: A rationale of enhanced photosynthesis. Nanoscale 2014, 6, 3647–3655. [Google Scholar] [CrossRef] [PubMed]
  303. Li, D.N.; Li, W.; Zhang, H.R.; Zhang, X.J.; Zhuang, J.L.; Liu, Y.L.; Hu, C.F.; Lei, B.F. Far-red carbon dots as efficient light-harvesting agents for enhanced photosynthesis. ACS Appl. Mater. Interfaces 2020, 12, 21009–21019. [Google Scholar] [CrossRef] [PubMed]
  304. Chakravarty, D.; Erande, M.B.; Late, D.J. Graphene quantum dots as enhanced plant growth regulators: Effects on coriander and garlic plants. J. Sci. Food Agric. 2015, 95, 2772–2778. [Google Scholar] [CrossRef]
  305. Younes, N.A.; Dawood, M.F.A.; Wardany, A.A. Biosafety assessment of graphene nanosheets on leaf ultrastructure, physiological and yield traits of Capsicum annuum L. and Solanum melongena L. Chemosphere 2019, 228, 318–327. [Google Scholar] [CrossRef]
  306. Hatami, M.; Hosseini, S.M.; Ghorbanpour, M.; Kariman, K. Physiological and antioxidative responses to GO/PANI nanocomposite in intact and demucilaged seeds and young seedlings of Salvia mirzayanii. Chemosphere 2019, 233, 920–935. [Google Scholar] [CrossRef]
  307. Ghorbanpour, M.; Farahani, A.H.K.; Hadian, J. Potential toxicity of nano-graphene oxide on callus cell of Plantago major L. under polyethylene glycol-induced dehydration. Ecotoxicol. Environ. Saf. 2018, 148, 910–922. [Google Scholar] [CrossRef]
  308. Hatami, M.; Hadian, J.; Ghorbanpour, M. Mechanisms underlying toxicity and stimulatory role of single-walled carbon nanotubes in Hyoscyamus niger during drought stress simulated by polyethylene glycol. J. Hazard. Mater. 2017, 324, 306–320. [Google Scholar] [CrossRef] [PubMed]
  309. Mehrjardi, H.A.; Jonoubi, P.; Majd, A.; Hosseini, R.H.; Abarghouei, H.B.; Hosseinizadeh, S.A. Study of MWCNTs effects on growth rate, germination and morphological characteristics of two Salvia species seedlings, Salvia sclarea and Salvia macrosiphon. J. Curr. Res. Sci. 2015, 3, 110–118. [Google Scholar]
  310. Fatemi, F.; Abdollahi, M.R.; Mirzaie-Asl, A.; Dastan, D.; Papadopoulou, K. Phytochemical, antioxidant, enzyme activity and antifungal properties of Satureja khuzistanica in vitro and in vivo explants stimulated by some chemical elicitors. Pharm. Biol. 2020, 58, 286–296. [Google Scholar] [CrossRef] [Green Version]
  311. McGehee, D.L.; Alimohammadi, M.; Khodakovskaya, M.V. Carbon-based nanomaterials as stimulators of production of pharmaceutically active alkaloids in cell culture of Catharanthus roseus. Nanotechnology 2019, 30, 275102. [Google Scholar] [CrossRef] [PubMed]
  312. Gohari, G.; Safai, F.; Panahirad, S.; Akbari, A.; Rasouli, F.; Dadpour, M.R.; Fotopoulos, V. Modified multiwall carbon nanotubes display either phytotoxic or growth promoting and stress protecting activity in Ocimum basilicum L. in a concentration-dependent manner. Chemosphere 2020, 249, 126171. [Google Scholar] [CrossRef]
  313. Ghosh, M.; Bhadra, S.; Adegoke, A.; Bandyopadhyay, M.; Mukherjee, A. MWCNT uptake in Allium cepa root cells induces cytotoxic and genotoxic responses and results in DNA hyper-methylation. Mutat. Res. Fund. Mol. Mech. Mut. 2015, 774, 49–58. [Google Scholar] [CrossRef]
  314. 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]
  315. Samarah, N.H.; Wang, H.; Welbaum, G.E. Pepper (Capsicum annuum) seed germination and vigour following nanochitin, chitosan or hydropriming treatments. Seed Sci. Technol. 2016, 44, 609–623. [Google Scholar] [CrossRef]
  316. Giorgetti, L.; Spano, C.; Muccifora, S.; Bottega, S.; Barbieri, F.; Bellani, L.; Castiglione, M.R. Exploring the interaction between polystyrene nanoplastics and Allium cepa during germination: Internalization in root cells, induction of toxicity and oxidative stress. Plant. Physiol. Biochem. 2020, 149, 170–177. [Google Scholar] [CrossRef] [PubMed]
  317. Maity, S.; Chatterjee, A.; Guchhait, R.; De, S.; Pramanick, K. Cytogenotoxic potential of a hazardous material, polystyrene microparticles on Allium cepa L. J. Hazard. Mater. 2020, 385, 121560. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Types of nanoparticles (NPs) affecting medicinal and aromatic plants discussed in this paper.
Figure 1. Types of nanoparticles (NPs) affecting medicinal and aromatic plants discussed in this paper.
Applsci 11 01813 g001
Figure 2. Beneficial and adverse effects of NPs on medicinal and aromatic plants.
Figure 2. Beneficial and adverse effects of NPs on medicinal and aromatic plants.
Applsci 11 01813 g002
Table 1. Enhanced production of secondary metabolites (SM) of medicinal plants elicited by various NPs using in vitro cultures and plants grown in soil or hydroponic solutions compared to corresponding bulk elicitors. (CSC = cell suspension culture, TPC = total phenolic content, TFC = total flavonoid content, MS = Murashige and Skoog medium, EO = essential oil)
Table 1. Enhanced production of secondary metabolites (SM) of medicinal plants elicited by various NPs using in vitro cultures and plants grown in soil or hydroponic solutions compared to corresponding bulk elicitors. (CSC = cell suspension culture, TPC = total phenolic content, TFC = total flavonoid content, MS = Murashige and Skoog medium, EO = essential oil)
NPs of ElicitorPlantCultivation ModeElicitor DoseSMMultiple of SM Content in ControlRefs.
AgMomordica charantia L.CSC5 mg/LTPC
TFC
flavonoids
hydroxybenzoic acid
hydroxycinnamic acid
1.40
1.56
1.52
1.23
1.15
[121]
AgEchinacea purpurea L.CSC, root derived callus (48 h)
CSC, leaf derived callus (72 h)
2 mg/Lcichoric acid
cichoric acid
≈2.22
≈1.80
[122]
AgIsatis constricta L.leaves of shoots regenerated in MS (5 days)2 mg/Ltryptanthrin1.71[123]
AgCaralluma tuberculata L.callus culture90 µg/LTPC
TFC
≈3.75
≈3.60
[124]
AgTrigonella foenum-graecum L2% agar (5 days) and then cultivation in sterile soil200 μg/plantdiosgenin1.30[115]
AgLinum usitatissimum Lcell suspension culture, 20 days30 μg/Lsecoisolariciresinol diglucoside
lariciresinol diglucoside
dehydrodiconiferyl alcohol glucoside
guaiacylglycerol-coniferyl alcohol ether glucoside
10.0

2.8
5.0

1.75
[125]
AgStevia rebaudiana L.callus culture45 mg/L, after 2 d 30 mg/L, 45 mg/L, after 6 dstevioside
stevioside
rebaudioside A
≈4.32
≈3.26
≈1.70
[126]
AgVanilla planifolia L.in vitro cultivation in MS25 mg/L
50 mg/L
TPC≈1.83
≈1.83
[127]
AgSalvia officinalis L.foliar spaying100 mg/Lrosmarinic acid≈11.0[128]
AgAu (1:3)Prunella vulgaris Lcell culture (+ 2 mg/L NAA), 14 days30 μg/LTPC
TFC
1.54
1.39
[129]
AuLavandula angustifolia L.in vitro culture50 mg/L
10 mg/L
(E,E)-farnesol
bisabolol oxide A
cadale
necaryophyllene oxide
1.91
1.15
1.76
1.23
[130]
CuArtemisia absinthium L.seeds inoculated on MS medium30 µg/mLTPC
TFC
≈1.66
≈1.07
[131]
ZnMomordica charantia L.foliar spraying20 ppmcarotenoids
anthocyanins
TPC
TFC
2.55
1.11
1.19
1.45
[132]
Zn-Ag (0.95:0.05)
(0.75:0.25)
Withania somnifera L. Dunalin vitro culture (MS medium; 1 month)20 mg/Lwithanolide≈14.08
≈5.60
[63]
NiWithania somnifera L. Dunalin vitro culture (MS medium; 1 month)20 mg/Lwithanolide≈7.90[63]
FeMentha piperita L.foliar application, 3 times0.5 g/L
0.5 g/L
1.5 g/L
menthone
menthol
menthofuran
1.65
1.30
2.50
[133]
CoArtemisia annua L.cell suspension culture (24 h)5 mg/Lartemisinin2.25[134]
CdSe QDsWithania somnifera L. Dunalin vitro culture (MS medium; 1 month)20 mg/Lwithanolide≈3.75[63]
CuOStevia rebaudianaleaf regenerants tissue culture20 mg/Lrebaudioside A
stevioside
1.50
1.94
[135]
CuOGymnema sylvestre (Retz.) R. Brcell suspension culture3 mg/Lgymnemic acid II
TPC
TFC
9.0
≈2.56
≈1.79
[136]
ZnOStevia rebaudiana L. Bertoni tissue culture grown shoots0.1 mg/L
1.0 mg/L
rebaudioside A
stevioside
TPC
stevioside
TPC
1.35
1.53
1.42
1.60
1.63
[137]
ZnOStevia rebaudiana L.leaf regenerants tissue culture2 mg/Lrebaudioside A
stevioside
1.49
1.27
[135]
ZnOLinum usitatissimum L.in vitro culture (MS medium)500 mg/Lsecoisolariciresinol diglucoside
lariciresinol diglucoside
dehydrodiconiferyl alcohol glucoside
guaiacylglycerol-β-coniferyl alcohol ether
glucoside
1.28

1.35
1.60
1.54
[138]
ZnOStevia rebaudiana L.hydroponic cultivation75 mg/LTPC
TFC
1.61
1.88
[139]
ZnOLinum usitatissimum L.cell suspension culture60 mg/Ltotal lignan≈1.59[140]
ZnOHypericum perforatum L.cell suspension culture100 ppbhypericin
hyperforin
3.8013.36[141]
Fe2O3Melissa officinalis L.plant irrigation at 60% FC30 µMEO per plant≈1.60[142]
Fe2O3Hypericum perforatum L.cell suspension culture100 ppbhypericin
hyperforin
5.40
12.02
[141]
Fe3O4Dracocephalum polychaetum Bornm.cell suspension culture100 ppmnaringin
rutin
quercetin
apigenin
rosmarinic acidt
hymol carvacrol
2.02
3.15
6.29
4.81
3.18
5.07
3.22
[143]
TiO2Linum usitatissimum L.cell suspension culture150 mg/Ltotal lignan1.50[140]
TiO2Mentha piperita L.foliar spraying150 mg/Lmenthol
menthone
menthyl acetate
1.09
1.32
1.11
[144]
TiO2Vetiveria zizanioides Lfoliar spraying90 mg/L measurement 300 DAT)khusimol1.24[145]
TiO2Salvia officinalis L.spraying of the 4-month old plants200 mg/LTPC
TFC
p-cymene
1,8-cineol
cis-thujonecamphor
1.63
1.72
1.61
2.23
1.88
1.31
[146]
TiO2Stevia rebaudiana L. Bertonicultivation in soil, spraying 3-fold in 3 weeks60 mg/L
200 mg/L
stevioside≈1.67
≈1.77
[147]
TiO2Nigella arvensis L.hydroponic cultivation, 21 d2500 mg/L
50 mg/L
50 mg/L
2500 mg/L
1000 mg/L
glaucine (shoots)
glaucine (roots)
quercetin (shoots)
quercetin (roots)
TPC
2.4
1.7
1.5
1.3
2.2
[148]
TiO2Dracocephalum moldavica L.irrigation of hydroponically grown plants100 mg/Lgeraniol
geranial (E-citral)Z-citral
1.41
1.05
1.07
[149]
TiO2Dracocephalum moldavica L.spraying of plants grown in pots30 ppm
30 ppm
100 ppm
rosmarinic acid
ellagitan
ninchlorogenic acid
caffeic acid
1.23
1.40
1.22
1.41
[150]
NiONigella arvensis L.hydroponic cultivation, 21 d1000 mg/L
2500 mg/L
50 mg/L
50 mg/L
100 mg/L
glaucine (shoots)
glaucine (roots)
quercetin (shoots)
quercetin (roots)
TFC
3.20
2.60
2.2
1.2
2.5
[148]
Al2O3Nigella arvensis L.hydroponic cultivation, 21 d2500 mg/L
2500 mg/L
2500 mg/L
1000 mg/L
glaucine (shoots)
glaucine (roots)
quercetin(shoots)
quercetin(roots)
1.37
1.60
1.22
1.43
[148]
Mn2O3Atropa belladonna Lin vitro culture on MS medium100 mg/L
25 mg/L
TPC
TFC
3.05
4.49
2.92
[151]
SiMentha piperita L.foliar spraying100 mg/Lmenthol/per plant
menthone/per plant
menthyl-acetate/per plant
1.811.751.83[152]
SiO2Matricaria recutita L.seed treatment (1 h), then in vitro culture on MS4 g/L
6 g/L
6 g/L
TPC
TPC
TFC
2.50 *
4.40 *
10.58 *
[153]
perliteHypericum perforatum L.callus cultures from in vitro grown plants
field grown plants
25 mg/L
50 mg/L
50 mg/L
alkaloids14.24
12.69
1.75
[154]
perlite-TiO2Hypericum perforatum L.callus cultures from in vitro grown plants
field grown plants
25 mg/L
50 mg/L
25 mg/L
alkaloids12.35
11.80
1.34
[154]
MWCNTsSalvia verticillata Lspraying of plants50 mg/L
50 mg/L
1000 mg/L
250 mg/L
TPC
rosmarinic acid
rosmarinic acid
caffeic acid
1.20
3.40
3.99
2.88
[155]
MWCNTsCatharanthus roseus L.in vitro cultivation (MS medium)50 mg/L
100 mg/L
150 mg/L
50 mg/L
alkaloids
TPC
≈1.86
≈1.63
≈1.86
≈1.36
[156]
MWCNTsThymus daenensis Celak.in vitro culture (MS medium)250 mg/LTPC
TFC
2.10
1.09
[157]
MWCNTsSatureja khuzistanicaLeaf segments cultured in B5 basal medium100 µg/mLTPC
TFC
rosmarinic acid
caffeic acid
1.96
2.61
2.60
1.76
[158]
MWCNT-
COOH
Salvia nemorosa L.cell suspension culture + 70 Gy γ-irradiation100 mg/Lrosmarinic acid
salvianolic acid B
ferulic acid
cinnamic acid
13.0 *
14.2 *
20.0 *
3.0 *
[159]
C60 fullereneTanacetum parthenium L., Pharmasaat genotypefoliar spraying (harvest at full flowering stage)250 mg/Lparthenolide≈8.2[160]
fullerenol [C60(OH)20]Momordica charantiaseed treatment10.88 nMcucurbitacin-B
lycopene
charantin
insulin
1.74
1.09
1.05
1.91
[161]
* compared to wild plant, ≈ evaluated from graphs.
Table 2. Enhanced production of secondary metabolites (SM) of medicinal plants elicited by various NPs using hairy root cultures compared to corresponding bulk elicitors.
Table 2. Enhanced production of secondary metabolites (SM) of medicinal plants elicited by various NPs using hairy root cultures compared to corresponding bulk elicitors.
NPsMedicinal PlantDose (Length of Treatment)SMMultiple of Control SM ContentRefs.
ZnOHyoscyamus reticulatus L.100 mg/L (48 h)
100 mg/L (72 h)
hyoscyamine
scopolamine
4.61
3.20
[67]
Fe2O3Hyoscyamus reticulatus L.900 mg/L (24 h)
450 mg/L (48 h)
hyoscyamine
scopolamine
≈5.0
≈5.0
[65]
Fe2O3Dracocephalum kotschyi Boiss75 mg/L (24 h)rosmarinic acid
xanthomicrol
cirsimaritin
isokaempferide
9.7
11.87
3.85
2.27
[162]
SiO2Dracocephalum kotschyi100 mg/Lrosmarinic acid
xanthomicrol
cirsimaritin
isokaempferide
8.26
13.00
13.42
10.00
[163]
Ag-SiO2Artemisia annua L.900 mg/L (3 days)artemisinin3.90[164]
AgDatura metel L.20 ppm
(12 h)
(24 h)
(48 h)
atropine
1.147
1.117
2.420
[165]
CuOBrassica rapa L. spp. pekinensis50, 100, 250 mg/L for 24, 48, 72 hgluconapin
glucobrassicanapin
4-methoxyglucobrassicin
neoglucobrassicin
4-hydroxyglucobrassicin
≈1.38
≈1.34
≈1.51
≈1.54
≈1.82
[166]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kralova, K.; Jampilek, J. Responses of Medicinal and Aromatic Plants to Engineered Nanoparticles. Appl. Sci. 2021, 11, 1813. https://doi.org/10.3390/app11041813

AMA Style

Kralova K, Jampilek J. Responses of Medicinal and Aromatic Plants to Engineered Nanoparticles. Applied Sciences. 2021; 11(4):1813. https://doi.org/10.3390/app11041813

Chicago/Turabian Style

Kralova, Katarina, and Josef Jampilek. 2021. "Responses of Medicinal and Aromatic Plants to Engineered Nanoparticles" Applied Sciences 11, no. 4: 1813. https://doi.org/10.3390/app11041813

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