Next Article in Journal
Curling Leaf 1, Encoding a MYB-Domain Protein, Regulates Leaf Morphology and Affects Plant Yield in Rice
Next Article in Special Issue
Biochemical and Molecular Characterization of Musa sp. Cultured in Temporary Immersion Bioreactor
Previous Article in Journal
Fruit and Vegetable Production
Previous Article in Special Issue
Optimization of Callus Induction and Shoot Regeneration from Tomato Cotyledon Explants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multifaceted Role of Nanomaterials in Modulating In Vitro Seed Germination, Plant Morphogenesis, Metabolism and Genetic Engineering

1
Department of Botany, University of Rajasthan, Jaipur 302004, Rajasthan, India
2
Amity Institute of Biotechnology, Amity University Rajasthan, Jaipur 303002, Rajasthan, India
*
Author to whom correspondence should be addressed.
Plants 2023, 12(17), 3126; https://doi.org/10.3390/plants12173126
Submission received: 23 May 2023 / Revised: 7 August 2023 / Accepted: 12 August 2023 / Published: 30 August 2023
(This article belongs to the Special Issue In Vitro Morphogenesis of Plants)

Abstract

:
The agricultural practices of breeding, farm management and cultivation have improved production, to a great extent, in order to meet the food demands of a growing population. However, the newer challenges of climate change, global warming, and nutritional quality improvement will have to be addressed under a new scenario. Plant biotechnology has emerged as a reliable tool for enhancing crop yields by protecting plants against insect pests and metabolic engineering through the addition of new genes and, to some extent, nutritional quality improvement. Plant tissue culture techniques have provided ways for the accelerated clonal multiplication of selected varieties with the enhanced production of value-added plant products to increase modern agriculture. The in vitro propagation method has appeared as a pre-eminent approach for the escalated production of healthy plants in relatively shorter durations, also circumventing seasonal effects. However, there are various kinds of factors that directly or indirectly affect the efficiency of in vitro regeneration like the concentration and combination of growth regulators, variety/genotype of the mother plant, explant type, age of seedlings and other nutritional factors, and elicitors. Nanotechnology as one of the latest and most advanced approaches in the material sciences, and can be considered to be very promising for the improvement of crop production. Nanomaterials have various kinds of properties because of their small size, such as an enhanced contact surface area, increased reactivity, stability, chemical composition, etc., which can be employed in plant sciences to alter the potential and performance of plants to improve tissue culture practices. Implementing nanomaterials with in vitro production procedures has been demonstrated to increase the shoot multiplication potential, stress adaptation and yield of plant-based products. However, nanotoxicity and biosafety issues are limitations, but there is evidence that implies the promotion and further exploration of nanoparticles in agriculture production. The incorporation of properly designed nanoparticles with tissue culture programs in a controlled manner can be assumed as a new pathway for sustainable agriculture development. The present review enlists different studies in which treatment with various nanoparticles influenced the growth and biochemical responses of seed germination, as well as the in vitro morphogenesis of many crop species. In addition, many studies suggest that nanoparticles can be useful as elicitors for elevating levels of important secondary metabolites in in vitro cultures. Recent advancements in this field also depict the suitability of nanoparticles as a promising carrier for gene transfer, which show better efficiency than traditional Agrobacterium-mediated delivery. This review comprehensively highlights different in vitro studies that will aid in identifying research gaps and provide future directions for unexplored areas of research in important crop species.

1. Introduction

Owing to their minuscule size, NPs acquire novel and unique properties that differ from their bulkier counterparts [1], giving rise to breakthrough technology with application-based solutions in many sectors of agriculture and plant biotechnology. However, the release of NPs into the environment has raised concern because of their toxic effects on the environment and human health [2]. Moreover, the release of NPs into the environment could result in their entry and accumulation in agricultural soils from bio-solids impregnated with NPs through the application of sewage sludge for agricultural purposes [3]. Thus, the application of NPs in plant tissue cultures is promising, as this technique is used to screen different aspects of plants’ growth and development, as well as to engage in genetic manipulation, bioactive compound production and plant improvement [4]. It has been noted that NPs have a positive impact because of their reduced size, elevated reactivity, mass-to-area ratio and other physico-chemical properties, but the negative effects of NPs have also been noted, which mainly depend on the type of metal, dissolution power and plant species [5,6]. In recent years, the application of NPs has demonstrated a positive effect on callus induction, organogenesis, somatic embryogenesis, genetic transformation and secondary metabolite production. Although there are a number of reviews on the application of nanomaterials in plants [7,8,9,10] and in agriculture [11,12] to mitigate various stresses [13,14], reviews on their application in plant tissue culture are scant [15]. Thus, this review comprises numerous studies that were conducted to explore the in vitro application of NPs on commercially important crops with respect to various aspects like preventing contamination, impact on seed germination, production of metabolites, induction of morphogenesis, biochemical and molecular changes, and genetic engineering.

2. Efficiency of NPs in Eliminating Contamination

The production of healthy plantlets is a prime concern behind the technique of plant tissue culture but microbial contamination is a common problem faced during this procedure. Conventionally, antibiotics are employed to eliminate microbes, but their frequent application can negatively affect plant tissue growth, e.g., antibiotics like carbenicillin and cefotaxime inhibit plant cell growth, organogenesis and embryogenesis [16,17]. Reports suggest that streptomycin and chloramphenicol interact with protein synthesis, rifampicin hinders nucleic acid synthesis and penicillin inhibits cell-wall membrane synthesis [18,19]. There is also the risk of a decreased genetic stability and lower regeneration capability of plants when a high level of antibiotics is used [20]. Nanomaterials are an alternative because of their distinctive features, which have been shown to possess antifungal and antibacterial properties that restrict microbial growth in in vitro cultures resulting in the successful mass propagation of selected species [21]. Silver nanoparticles (AgNPs) have been considered one of the better options, as the anchoring and penetration of Ag ions into microbes alter the cellular signals, via dephosphorylation, of key peptide substrates on tyrosine [22,23]. Another study suggested that Ag+ ions cause a reduction in DNA replication, as well as inactivate the thiol group in proteins, that ultimately reduces microbial growth [24]. Similarly, Min et al. [25] reported that AgNPs restrict the growth of sclerotium-forming phytopathogenic fungi and, hence, can become an alternative to pesticides. AgNPs have been employed to reduce contamination during in vitro cultures of Olea europaea L. [26], Nicotiana tabacum L. [27,28], Gerbera jamesonii Bolus ex Hook.f. [29], Solanum tuberosum L. [21], almond x peach (G x N15) hybrid rootstock [30], Rosa hybrida L. [31], Vitis vinifera L. [32], Vanilla planifolia Jacks. ex Andrews [33], and Phoenix dactylifera L. cv. Sewi and Medjool [34]. In addition, combined treatment with nanosilver and nano-iron particles was reported to have a significant effect on decreasing the contamination rate in Fragaria × ananassa L. cv. Roby Gem [35]. Similarly, biosynthesized silver, chitosan, and selenium NPs were tested for their antimicrobial potential for the in vitro multiplication of three olive cultivars (Koroneiki, Picual, and Manzanillo). Of all the three NPs, AgNPs showed the best antimicrobial properties in all cultivars [36]. However, some studies have also suggested that the concentrations of AgNPs played a crucial role in culture growth as higher concentrations might induce adverse effects on explant response [3,21]. The phytotoxic effect of higher AgNPs has been observed in crop plants of Phaseolus radiatus L. and Sorghum bicolor (L.) Moench [37]. Whereas in tomato and potato plants, it has been reported that lower concentrations of AgNPs with longer exposure time effectively reduced the contamination without hampering explant viability [38].
Titanium dioxide (TiO2) is another NP that has gained attention due to its antimicrobial potential, as it has photocatalytic properties to eliminate contamination from various sources, but its toxicity against microbial growth depends on the intensity and wavelength of light with concentration and particle size [19]. TiO2 reacts with water molecules and forms free radicals like OH, HO2, and H2O2 which in turn results in the oxidation of bacterial cells, suggesting that the photo-activation of TiO2 via UV irradiation retards the bacterial growth [39,40]. It has been evaluated that the addition of TiO2NPs in the Murashige and Skoog (MS) [41] medium enhanced the microbial resistance during the micropropagation of tobacco [27], S. tuberosum [19], and Hordeum vulgare L. [42]. Zinc oxide nanoparticles (ZnONPs) have eliminated nine strains of bacteria (Bacillus megaterium, Cellulomonas uda, C. flarigena, Corynebacterium panrometabolum, Erwinia cypripedii, Klebsiella spp., Pseudomonas spp., Proteus spp., and Staphylococcus spp.) and four fungal species (Aspergillus spp., Candida spp., Fusarium spp., and Penicillium spp.) which increased difficulties during banana micropropagation [43]. Thus, it can be observed that although nanomaterials at higher concentrations have been proven as toxic for plant growth, they can be employed as disinfectant agents for the in vitro multiplication of various economically important crops. The majority of the reports used Ag, TiO2, and Zn-based NPs for the inhibition of microbial growth during in vitro propagation, but new types of NPs should also be assessed. In this regard, various kinds of advanced nanomaterials like graphene, graphite, quantum dots, carbon nanotubes, polymer dendrimers, and atomic clusters will provide enough scope for the study; along with this, evaluations of concentrations, sizes, and types of NPs on various crop species and type of explant are also needed [44].

3. Influence of NPs on Seed Germination

Seed germination is a crucial stage for crop development since young seedlings are more vulnerable to biotic and abiotic stresses [45]. Therefore, lots of efforts to improve the efficiency of seed germination are published from time to time with new technological interventions. Studies to analyze the effect of NPs have been conducted during the last few years, and it was observed that genotype, variety, seed age, and environmental conditions determined the response to NPs [46]. Yasur and Rani [47] and Hatami [48] suggested that the water uptake during seed germination is critical because seeds are relatively dry and requires a substantial amount of water to initiate cellular metabolism and growth. The positive effects of NPs on germination begin with the high capability of NPs to penetrate the seed coat and promote water uptake along with the absorption of nutrients in the seed [49]. Mehrian et al. [50] documented that NP treatment accelerated seed germination from better water uptake by the seeds during the initial days, whereas a decrease in germination efficiency was noted as time passed because of the breakdown of stored nutrients or alternations in permeability properties of the cell membrane. Similarly, Rizwan et al. [51] noted that NPs can penetrate through the seed coat and affect the development processes of embryos through stimulation of the enzymes of metabolic processes. During the radicle appearance stage of seed germination, root apex tissues come in contact with NPs, which then move into the rhizodermis through the apoplast with endocytosis. In the root, they flow towards the plant secretory tissue using symplastic pathways and translocate to other plant organs. However, it has been noted that NPs at a high concentration result in a perforation of the cell wall and penetrate the protoplast and damage the root cell vacuoles. This triggers more production of reactive oxygenspecies (ROS) and it causes a blockage of electron transfer which induces oxidative stress. NPs also up-regulate the genes involved in cell division and carbon/nitrogen metabolism, and the negative effects observed in seedling growth are due to chromosomal aberrations and mitotic abnormalities. This leads to a decrease in cell division of the root meristem, hormonal imbalance, ROS over-production, and increased levels of lipid peroxidation [52]. The increased oxidative stress, in turn, increases hydrogen peroxide (H2O2) contents, activities of malondialdehyde (MDA), catalases (CAT), peroxidases (POD), and superoxide dismutase (SOD), as well as the production of compounds having antioxidant activities like phenolics and flavonoids [53]. Many studies have documented that NPs exert positive or negative influences on seed germination, seedling biomass as well as biochemical and metabolite contents. In the present review, we have taken only those examples where NPs were added into the media and not where seeds were placed on filter paper or water agar media after sonication treatment with NPs.

3.1. AgNPs

In the majority of the studies, NPs’ effect has been evaluated under in vivo conditions [54], but few were tested under in vitro conditions on the culture media. It is also observed that most reports suggested the usage of AgNPs (Table 1), e.g., Lee et al. [37] recorded a negative effect of AgNPs on P. radiates and S. bicolour seedling growth. Similarly, the growth of Physalis peruviana L. seedlings also decreased along with chlorophyll content, but biomass in terms of fresh (FW) and dry weights (DW) was increased. It was also revealed that the seedling growths were not much affected in soil as compared to the agar-based medium. This might be due to changes in the physico-chemical properties of NPs in the soil, as pore water harbours a range of electrolytes that increase the aggregation of AgNPs in soil. These aggregates were larger than the pore size of plant root cells and thus failed to pass through the cells. Greater aggregation may be the principal reason for the reduced phytotoxicity of AgNPs in soil. Thus, the relative germination index is extensively used as an indicator of phytotoxicity, and root growth is one of the sensitive biomarkers for the phytotoxicity assay [55]. Zaka et al. [56] compared AgNPs, gold nanoparticles (AuNPs), and copper nanoparticles (CuNPs) for Eruca sativa Mill. and observed that AgNPs increased seed germination, shoot and root lengths, and seed vigour index, whereas the other two adversely affected these parameters (Table 1). Further evaluation unveiled that all the NPs affected the biochemical milieu of the plants differently (Table 2). In another study, green synthesized AgNPs using Curculigo orchioides Gaertn. were found to exert a positive influence on seedling growth and biomass of Oryza sativa L. cv. Swarna. When the germinated seedlings were biochemically analyzed, an increase in chlorophyll, flavonol contents and enzymes (POD, SOD, CAT, APX, and GR) activities, and a decrease in phenolics, flavonoids, H2O2, and MDA contents were observed. The gene expression analyses revealed that the SOD gene was down-regulated, whereas genes for CAT and ascorbate peroxidase (APX) were up-regulated after AgNP treatment [57]. Similarly, increased seed germination, seed vigour index, shoot and root lengths, and fresh and dry biomass in Pennisetum glaucum (L.) R. Br. after the addition of AgNPs in the medium was reported [58]. The maximum germination was recorded at 40 ppm; at this concentration of AgNPs, mild activities of 2,2-Diphenyl-1-picrylhydrazyl (DPPH), SOD activities and proline content were recorded that significantly increased at higher dose of AgNPs. On the contrary, phenolic contents were higher at optimum germination concentration (40 ppm) and lower at higher concentration, whereas flavonoids were lower at 40 ppm and increased at high levels. AgNPs positively influenced the germination and seedling traits of Brassica oleracea L. var. sabellica ‘Nero di Toscana’ and Raphanus sativus L. var. sativus ‘Ramona’, whereas these traits were decreased in Solanum lycopersicum L. ‘Poranek’. One of the reasons behind decreased growth S. lycopersicum might be due to the presence of AgNPs in plasmodesmata, precluding the transport of nutrients that led to a reduction in plant biomass [59]. Recently, Tomaszewska-Sowa et al. [60] observed the effect of AgNPs and AuNPs on Brassica napus L., and revealed that application of both NPs decreased shoot and root lengths of seedlings irrespective of treatment time. However, total chlorophylls, carotenoids, anthocyanins, free sugars, and H2O2 contents were higher, but no major change in phenolics was found. The seed germination of N. tabacum was carried out using CTAB- and PVP-coated AgNPs, and coating with CTAB showed a positive influence whereas coating with PVP failed to show any positive effect on germination rate and biomass [61]. Similarly, positive influences of AgNPs have been also documented in Brassica juncea (L.) Czern. var. pusajaikisan [62], Hylocereus undatus (Haw.) Britton and Rose [63], and P. vulgaris [64] (Table 1).

3.2. Other Metal and Metal Oxide NPs

Apart from AgNPs, other metal NPs are also used for seedling germination under in vitro conditions; Dehkourdi and Mosavi [69] utilized TiO2NPs and documented a positive influence on seed germination as well as on chlorophyll synthesis in Petroselinum crispum (Mill.) Fuss, whereas Nair et al. [71] observed that the application of copper oxide nanoparticles (CuONPs) on Vigna radiata L. decreased seedling growth in terms of length and biomass. They have also reported that CuONPs decreased chlorophyll and increased proline contents, whereas it increased H2O2 and MDA contents in the root; however, no change in carotenoid, H2O2, and MDA contents in the shoot and increased lignification of root cells were detected (Table 2). The negative effect of CuONPs on seedlings of Cicer arietinum L. was also documented where decreased growth and biomass have been recorded at all the tried concentrations (50–500 mg/L), and elevated H2O2 generation, MDA level, and POD activity along with increased lignifications in roots were observed. Further expression analysis revealed that CuZn-SOD, CAT, and APX genes were up-regulated in roots but no change was found in shoots [66]. Similarly, CuONPs, when used for the treatment of Brassica nigra (L.) K. Koch, delayed the germination of seedlings and decreased plantlet length and biomass significantly [65]. ZnONPs in the media containing seeds of the same plant negatively influenced seedling growth, shoot FW, and reduced stem diameter as the NP amount increased in the media. However, the treatment increased free radical scavenging activity, total antioxidant capacity, total reducing power, phenolics, and flavonoid contents in the shoot and root of the seedling (Table 2) [65]. Moreover, in seeds of Linum usitatissimum L. cv. Barbara, different concentrations of ZnONPs (1, 10, 100, 500, and 1000 mg/L) were tried, and 100 mg/L concentrations proved beneficial in terms of shoot and root lengths as well as seedling biomass, further higher concentrations adversely affected seedling growth [68]. In another study, treatment with multi-walled carbon nanotubes (MWCNTs) showed a positive influence on germination, seedling lengths, as well as biomass in Glycine max (L.) Merr. hybrid S42-T4, H. vulgare hybrid Robust, and Zea mays L. hybrid N79Z 300GT [67]. Unlike the spherical shapes of other NPs, MWCNTs are the allotropes of carbon that are arranged in an elongated, tubular manner with many rolled sheets. Its unique features like functional group, diameter, length, and solubility make its penetration inside the seed coat convenient and it is efficiently translocated in plants [72]. Similar observations have been well documented previously where MWCNTs improve germination, plant growth, and agronomic traits by penetration, and increasing the water and nutrient uptake [73,74].
Table 2. Biochemical changes in seedlings and cultures after NP treatment.
Table 2. Biochemical changes in seedlings and cultures after NP treatment.
PlantNanoparticle (NP) Treatment and Culture TypeBiochemical ChangesReference
Brassica juncea
var. pusa jaikisan
AgNPs, shootsIncreased chlorophyll and decreased MDA, H2O2, and proline content, increased CAT, GPX, and APX activities[62]
Brassica napusAgNPs/AuNPs, shootsIncreased chlorophylls, carotenoids, anthocyanins, free sugars, H2O2 contents, no change in phenolic content[60]
Brassica nigraZnONPs, shoots and roots (seedling), callusIncreased free radical scavenging activity, total antioxidant capacity, total reducing power, phenolic, and flavonoid contents[65]
Brassica nigraCuONPs, seedling and roots (from leaf and stem derived callus)Seedlings increased free radical scavenging activity, total phenolic, and flavonoid content, decreased total antioxidant and reducing potential;
Roots increased free radical scavenging activity, total antioxidant and reducing potential, total phenolic, and flavonoid contents
[5]
Brassica oleracea var. sabellica ‘Nero di Toscana’AgNPs, leavesDecreased chlorophyll, carotenoid, and anthocyanin contents, no change in phenolic, protein contents and SOD activities, increased GPOX activity[59]
Campomanesia rufaAgNPs, shootsNo significant difference in SOD activity[75]
Caralluma tuberculataAgNPs, callusIncreased PAL and free radical scavenging, SOD, POD, CAT, APX activities, total phenolics, and flavonoid contents[76]
Cicer arietinumCuONPs, seedlingIncreased H2O2 generation, MDA content, POD activity, and lignification in roots[66]
Cichorium intybusFe2O3NPs, hairy rootsIncreased hairy root growth, total phenolic, and flavonoid contents[77]
Corylus avellana
cv. Gerd Eshkevar
AgNPs, cell suspensionIncreased CAT, APX, H2O2, PAL activities, decreased SOD and POD activities, and total soluble phenol content[78]
Corylus avellana
cv. Gerd Eshkevar
AgNPs, cell suspensionIncreased MDA, total phenolic, anthocyanin, and flavonoid contents[79]
Cucumis anguriaAgNPs, hairy rootsIncreased total phenolic and flavonoid contents, and antioxidant activities[80]
Eruca sativaAuNPs, CuNPs, and AgNPs, seedlingAuNPs decreased total antioxidant capacity, total phenolic and flavonoid contents, increased DPPH, SOD and POD activities, no change in protein content;
CuNPs decreased total antioxidant capacity, DPPH activity, protein content, increased total phenolic, and flavonoid contents, SOD and POD activities;
AgNP decreased total antioxidant capacity, DPPH activity, decreased total phenolics and flavonoid contents, POD activity, increased SOD activity, no change in protein
[56]
Fragaria × ananassa
cv. Queen Elisa
FeNPs, shootsIncreased chlorophyll a, chlorophyll b, total chlorophyll, carotenoid, total carbohydrates, total protein, and total free proline and iron contents, decreased H2O2 and MDA content, higher SOD and POD activities[81]
Linum usitatissimum
cv. Kerman Shahdad
ZnONPs/TiO2NPs, cell suspensionIncreased PAL and CAD activities, and total phenol content[82]
Linum usitatissimum
cv. Barbara
ZnONPs, seedling and callusIncreased ROS production, membrane lipid peroxidation, protein carbonylation and 8-oxo guanine formation, SOD, POD, radical scavenging activities, total phenolics, and flavonoid contents[68]
Maerua oblongifoliaAgNPs, shootsHigher chlorophyll, total protein and proline contents, and increased activities of antioxidant enzymes[83]
Momordica charantiaAgNPs, cell suspensionIncreased MDA, H2O2, total phenolics and flavonoid contents, and antioxidant activity[84]
Musa paradisiacal
cv. Grand Nain
ZnNPs and ZnONPs, shootsHigher proline, chlorophyll, and antioxidant enzymes activities[43]
Musa spp.AgNPs, shootsIncreased chlorophyll content[85]
Nicotiana benthamianaCH-ZnO, callusIncreased chlorophyll, carotenoid, proline contents and PAL and AO activities, decreased MDA and H2O2 levels[86]
Nicotiana tabacum
cv. Bright Yellow-2
ZnONPs, cell suspensionDecreased dehydrogenase, oxidoreductase SOD, POD and APX activities, increased GR, PAL, protease, caspase-like and acid phosphatases activities, and total phenolic content[87]
Oryza sativa cv. SwarnaAgNPs, seedling leavesIncreased chlorophyll and flavonol contents and POD, SOD, CAT, APX and GR activities, decreased phenolics, flavonoids, H2O2 and MDA contents[57]
Oryza sativa cv. IR64AgNPs, shootDecreased MDA, proline and H2O2 levels[88]
Pennisetum glaucumAgNPs, seedlingIncreased DPPH, proline, SOD, POD, and CAT activities, total phenolics and flavonoid contents[58]
Phoenix dactyliferaMWCNTs, shootsIncreased flavonoid, chlorophylls and carotenoid, nutrient contents, decreased phenolics and tannin contents, SOD, GPOX, and GR activities[89]
Phoenix dactylifera
cv. Hayani
AgNPs, somatic embryosIncreased chlorophyll content[90]
Physalis peruvianaAgNPs, seedling derived shoots and shootsSeedling derived shoots- increased CAT and APX activity, and decreased chlorophyll content, SOD and MDA activities;Shoots- no change in SOD, APX and MDA levels, decreased CAT activity[70]
Raphanus sativus
var. sativus ‘Ramona’
AgNPs, leavesIncreased carotenoid, phenolic contents, and SOD activity, decreased chlorophyll, anthocyanins, protein contents, and GPOX activity[59]
Saccharum spp.
cv. Mex 69-290
AgNPs, leavesIncreased N, Ca, Mg, Fe, Cu, Zn, Mn, and decreased P, K, and B content, higher total phenolics, ROS and lipid peroxidation contents, and antioxidant activity[91]
Simmondsia chinensisMWCNTs, shootsIncreased total tannin content and antioxidant activities, decreased phenolics and flavonoid contents[92]
Solanum lycopersiconFe3O4NPs, shootsIncreased proline content and osmotic potential[93]
Solanum lycopersicumZnONPs, callusIncreased Na, N, P, K, and Zn ionic, protein contents, SOD and GPX activity[4]
Solanum lycopersicum
var. Poranek
AgNPs, leavesIncreased chlorophyll, anthocyanins, phenolics, protein contents and SOD and GPOX activities, decreased carotenoid content[59]
Solanum tuberosumSiO2NPs, leavesIncreased antioxidant enzymes activity and expression of proteins[94]
Solanum tuberosum
cv. White Desiree
AgNPs, shootsIncreased total chlorophyll, carotenoids, proline, total flavonoids, phenolics, lipid peroxidation and H2O2 contents, decreased anthocyanins[95]
Vanilla planifoliaAgNPs, shootsHigher chlorophyll, increased elements like N and B, no change in P, Ca and Mg, and decreased K, Fe, Cu, Zn, Mn, and B contents, higher total phenolics, ROS and lipid peroxidation contents, and antioxidant activity[33]
Vigna radiataCuONPs, seedlingDecreased chlorophyll and increased proline contents, H2O2 and MDA contents in root, no change in carotenoid, H2O2 and MDA contents in shoots[71]
AgNPs: silver nanoparticles; APX: ascorbate peroxidase; AO: ascorbate oxidase; AuNPs: gold nanoparticles; CAT: catalase; CH-ZnO: chitosan-zinc oxide nano-bioformulation; CuNPs: copper nanoparticles; CuONPs: copper oxide nanoparticles; DPPH: 2,2-diphenyl-1-picrylhydrazyl, FeNPs: iron nanoparticles; Fe2O3NPs/Fe3O4NPs: iron oxide nanoparticles; GPX: guaiacol peroxidase; GPOX: glutathione peroxidase; GR: glutathione reductase; H2O2: hydrogen peroxide; MDA: melondialdehyde; MWCNTs: multi-walled carbon nanotubes; PAL: phenylalanine ammonia lyase; POD: peroxidase; ROS: reactive oxygen species; SiO2NPs: silicon dioxide nanoparticles; SOD: superoxide dismutase; TiO2NPs: titanium dioxide nanoparticles; ZnNPs: zinc nanoparticles; ZnONPs: zinc oxide nanoparticles.

4. Modulation of In Vitro Morphogenesis by NPs

The ions supplemented into the culture medium are transported via phloem cells [96] and the apoplastic pathway [97]. In a similar manner, as NPs have extremely small sizes, they enter the explants in a similar way, but their effect mainly depends on NP type, concentration, exposure time, and plant species [98]. Many studies have reported the influence of different types of NPs on the in vitro morphogenesis of various crops which are listed in Table 3.

4.1. Metal NPs

In a study on in vitro cultures of L. usitatissimum, BAP was coated with different nanoparticles (AuNPs and AgNPs), and AuNPs proved better than AgNPs for callus formation and somatic embryogenesis [99]. On the contrary, the positive effect of AgNPs on the in vitro regeneration of S. tuberosum cv. White Desiree has been documented where most of the morphological traits showed improvement (Table 3) along with contents of total chlorophyll, carotenoids, proline, total flavonoids, phenolics, lipid peroxidation, and H2O2, whereas anthocyanin content was decreased (Table 2). They have concluded that this beneficial effect is possibly due to the inhibition of ethylene perception by AgNPs [95]. In vitro cultures are known to produce ethylene during the culture period but their excess production inhibits cell division [100]. The adverse effect causes the mortality of cultures, and over-accumulation of ethylene induces senescence, abscission of leaves, and eventual leaf drop [101]. It is well known that Ag+ ions inhibit the physiological actions and production of ethylene because of its properties, easy uptake, and mobility in cells [102].
The number of reports suggested the application of AgNPs for improved regeneration; e.g., Bello-Bello et al. [91] documented positive influence of AgNPs on shoot formation in Saccharum spp. cv. Mex 69-290. They have reported that the better shoot induction might be due to an increase in nutrient elements suchas N, Mg, and Fe which are essential for plant growth. Similarly, a temporary immersion system with fortification of AgNPs has evoked a profuse multiplication in V. planifolia [33]. Mustafa et al. [64] used nano-priming of P. vulgaris seeds using CuNPs and AgNPs, and utilized seedling-derived hypocotyls as an explant for callus formation. They have documented that the callus formation was better when seeds were treated with AgNPs in comparison to control and CuNP treatment. In another study on Musa spp., it was confirmed that the addition of AgNPs into media increased the number of shoots, their length, leaf number, shoot FW/DW, and total chlorophyll content [85]. Also, AgNPs have been used at different stages such as callus formation, shoot induction, shoot multiplication, and rooting in Musa spp. [103] and O. europaea cv. Picual [104]. Three NPs, i.e., AgNPs, selenium (SeNPs), and chitosan (CSNPs), were utilized in different cultivars of O. europaea (Manzanillo, Picual, and Koroneiki), and AgNPs showed positive effect on shoot growth, whereas CSNPs and control media had less effectiveness and SeNPs exerted a negative effect [36]. The toxic effect of Se metal may be due to the replacement of sulfur atoms in sulfur-containing amino acids by Se, which results in changes in protein structure and function; simultaneously, it can cause oxidative stress, cellular damage and disrupt the plant’s metabolism [105,106]. The addition of nano-iron instead of traditional iron along with silver nitrate nanoparticles (AgNO3NPs) increased the shoot regeneration, leaf number/shoot, shoot FW and DW in F. ananassa cv. Ruby Gem [35]. El-Kosary et al. [34] had observed that callus formation was optimum at a higher concentration of AgNPs (500 μg/L) but a lower concentration (125 μg/L) was favorable for globular embryo formation as well as for the multiplication of embryos in P. dactylifera cv. Medjool and Sewi. The positive effect of AgNPs on somatic embryogenesis has been well documented in the same plant [90]. Whereas a negative effect of AgNPs was recorded in Campomanesia rufa (O. Berg) Nied, as a decreased shoot number was found as compared to control, no change in shoot biomass was noted [75]. The efficacy of silver thiosulfate, silver nitrate, and AgNPs was analyzed on the micropropagation of Citrus australasica F. Muell. and it was reported that AgNPs were less effective as compared to silver thiosulfate in terms of leaf abscission as well as shoot number and length, but were better than AgNO3 [101] (Table 3).
In vitro cultures of rice cells, when treated with MWCNTs, exhibited decreased cell density at higher concentration of NPs, and it was suggested that this might be due to self-defense response [107]. Later on, Taha et al. [89] reported a positive effect of MWCNTs on somatic embryo germination and elongation in P. dactylifera. In accordance, the positive influence of MWCNTs on the nodal culture of Simmondsia chinensis (Link) Schneider has been well documented [92]. In in vitro cultures of O. sativa ssp. indica cv. KDML105, a comparison of activated charcoal and nanocarbon on callusing and plant regeneration was carried out, which confirmed that the addition of nanocarbon proved better for callus induction frequency, its size, callus FW/DW, and ratio of no. of seedlings to calli [108].
Table 3. Effect of various NPs on in vitro morphogenesis in crops.
Table 3. Effect of various NPs on in vitro morphogenesis in crops.
PlantExplantNanoparticle (NP) TreatmentCallus/Number of Shoots or SEs/Explant
(% Response)
Shoot Length (cm)Root Induction MediaNumber of Roots/Explant (% Response)Root Length (cm)EffectReference
Alternanthera sessilisNodeGFAgNPs (2.0 mg/L) 153.6 ± 2.3 (100%)-- --In vitro cultures are genetically uniform with
mother plant
[109]
Brassica napus
cv. Hayola 401
HypocotylZnONPs (10 mg/L)Callus-
300 mg (FW),
29 mg (DW) (88%)
----Improved callus FW and DW[110]
Brassica nigraStemZnONPs (1 mg/L)Callus-
11.95 ± 1.7 g (FW)
0.70 ± 0.2 g (DW)
----Induced roots and few shoots from callus, decreased FW and DW of callus[65]
Brassica nigraLeaf and stemCuONPs (1 mg/L)Callus (leaf)-
8.5 ± 1.9 g (FW),
0.3 ± 0.01 g (DW);
Callus (stem)-
8.7 ± 1.8 g (FW),
0.4 ± 0.06 g (DW)
----Decreased FW and DW of callus[5]
Campomanesia rufaNodeAgNPs (1.54 mg/L)171.1---Less shoots in presence of NPs, but no significance difference in fresh mass of the shoots[75]
Cicer arietinumEmbryo axes (EA) and
embryo axes with adjacent part of cotyledon (EXC)
IONPs (15 mg/L)EA- 51.6 ± 0.9 (86%), EXC- 53.0 ± 1.5 (88%) (var. Punjab-Noor 09);
EA- 47.4 ± 0.4 (79%), EXC- 45.7 ± 2.5 (76%) (var. Bittle-98)
EA- 9.9 ± 0.3, EXC- 11.8 ± 0.5 (var. Pujab-Noor 09);
EA- 7.5 ± 0.3, EXC-
8.5 ± 0.4 (var. Bittle-98)
IONPs (15 mg/L)EA- 45.0 ± 1.2 (75%),
EXC- 49.8 ± 0.9 (83%)
(var. Pujab-Noor 09); EA-
41.5 ± 1.5 (69%),
EXC-
47.0 ± 2.6 (78%) (var. Bittle-98)
-Higher iron content[111]
Citrus australasicaNodeAgNPs (40 μM)17.43.53 ± 0.02---Less shoot regeneration[101]
Daucus carota
cv. Berlicum
HypocotylFe3O4NPs (4.02 mg/L)-----Decreased SEs formation, mitotic index of cell culture[112]
Fragaria × ananassa
cv. Queen Elisa
Runner tipsFeNPs (0.8 ppm)4 (Branch number)2.80 ± 0.03 (Branch length)--3.40 ± 0.20Increased biomass, higher percentage of relative water content (RWC), and membrane stability index (MSI)[81]
Fragaria × ananassa
cv. Ruby Gem
Runner tipsAgNO3NPs (10 mg/L)11.00 4.17---Increasedpercentage of open buds, shoot regeneration, leaf number, shoots FW and DW[35]
Fragaria × ananassaLeafShoot induction- Explants were treated with 200 mg/L AgNPs solution (20 min), shoot multiplication- AgNPs (0.20 mg/L)Regeneration- 21.00 (64.44%), multiplication- 12.67 (100%)3.93AgNPs (0.50 mg/L) 6.673.40Increased regeneration and rooting response, and biomass of plants[113]
Hordeum vulgare cv. NosratMature embryosTiO2NPs (60 µg/mL)-Callus diameter-
21 mm2
---Increased number and size of callus[42]
Linum usitatissimum
‘Blue di Riga’
Stem of in vitro seedlingCNPs (1×10−3 g/L)Callus- 83%,
0.5 ± 0.1 g (FW), indirect embryogenesis- 25%
Callus diameter- 8.5 ± 0.3 mm---Reduce callus formation, embryogenesis, and organogenesis[114]
Linum usitatissimumStem of in vitro seedlingBAP (1 mg/L) (coated with AgNPs or AuNPs)Callus regeneration zone- 1.40 ± 0.65, rhizogenesis (50%), embryogenesis (50%) (AgNPs);
Callus regeneration zone- 3.40 ± 1.22, rhizogenesis (30%), embryogenesis (70%) (AuNPs)
Callus length and width-
5.38 ± 1.30 and 5.00 ± 2.14 mm (AgNPs),
8.38 ± 1.60 and 5.38 ± 1.06 (AuNPs)
---Increased callus formation and embryogenesis[99]
Linum usitatissimumStem of in vitro seedlingExperiment A-
Fe3O4NPs (1.5 mg/L),
Experiment B-
Fe3O4NPs (1 mg/L)
Experiment A-
100% somatic embryogenesis,
Experiment B-
100% rhizogenesis
Callus length and width-
1.4 ± 0.38 and 1.11 ± 0.26 cm (Exp. A), 1.17 ± 0.55 and 0.98 ± 0.32 cm (Exp. B)
---Increased callus size and embryogenesis, NPs induced genotoxicity incallus cultures[115]
Maerua oblongifoliaNodeAgNPs (20 mg/L)16.67 ± 0.5710.43 ± 0.45---Increased shoot number and length, leaf number, shoot FW and DW[83]
Mentha longifoliaNodeCuNPs (0.5 mg/L)-6.83 ± 0.74- --Increased regeneration and shoot formation[116]
Musa paradisiacal cv. Grand NainShoot tip (Suckers)ZnNPs/ZnONPs (100 mg/L) Callus- 92%, shoot- 2.5 (92%)-ZnNPs/ZnONPs (100 mg/L)6.57 (89%)2.93Reduced contamination and increased callus formation, shoot regeneration, shoots and roots FW, and rooting[43]
Musa paradisiacalShoot tip (Suckers)FCNTs (100 µg/mL)12.55.2---Increased shoot formation[117]
Musa spp.In vitro shoot tipAgNPs (1 ppm)8.402.45AgNPs (3 ppm)7.107.70Increased number of shoot, its length, leaf number, shoot FW and DW[85]
Musa spp.Pseudo-stemCallus formation- AgNPs (8 ppm), shoot regeneration- AgNPs (4 ppm), multiplication- AgNPs (6 ppm) Callus- 97.78%,
multiplication coefficient- 4.22 (100%)
3.44AgNPs (4 ppm)5.22 (98.33%)4.26Increased callus formation, shoot induction and multiplication as well as rooting response[103]
Olea europaea
cv. Picual
NodeAgNPs (5 mg/L)1.725.44---Increased bud sprouting, shoot length, shoot number, and number of
leaves/shoot
[104]
Olea europaeaNodeAgNPs (10 mg/L)4.3 ± 0.17
(cv. Manzanillo);
4.0 ± 0.00
(cv. Picual);
5.0 ± 0.00
(cv. Koroneiki)
7.0 ± 0.00 (cv. Manzanillo);
8.0 ± 0.57
(cv. Picual);
10.0 ± 0.00
(cv. Koroneiki)
---Higher number of shoots, shoot length, leaf number, and
multiplication rate
[36]
Oryza sativa
cv. KDML105
SeedTiO2NPs (25 mg/L)2.80 ± 0.03
(56.46 ± 0.82%)
----Better regeneration[118]
Oryza sativaSeedCallus- CuONPs (10 mg/L), regeneration- CuONPs (20 mg/L) (var. Basmati 2000, Basmati 370, Basmati 385); Callus- CuONPs (10 mg/L), regeneration- CuONPs (15 mg/L) (var. Super Basmati)Callus- 74%, regeneration- 80%
(var. Basmati 2000);
Callus- 86%, regeneration- 42%
(var. Basmati 370); Callus- 90%, regeneration- 92%
(var. Basmati 385); Callus- 94%, regeneration- 65%
(var. Super Basmati)
----Increased callogenesis and regeneration[119]
Oryza sativa ssp. indica cv. RD49SeedTiO2NPs (20 mg/L)Callus- 97.73 ± 0.17%,
regeneration- 67%
----Better regeneration[120]
Oryza sativa ssp. indicaSeedCallus- TiO2NPs (50 mg/L), regeneration- TiO2NPs (40 mg/L)Callus- 94.67 ± 1.01%, regeneration- 3.11 (61.89 ± 1.13%) (cv. Suphanburi1); Callus- 93.25 ± 1.02%, regeneration- 3.06 (60.25 ± 1.13%) (cv. Suphanburi90)----Better regeneration[121]
Oryza sativa
cv. KDML105
SeedCallus- NCNPs (5 mg/L), regeneration- NCNPs (20 mg/L)Callus- 94.70 ± 0.86%,
regeneration-
3.16 ± 0.04
(62.75 ± 0.89%)
----Increased callus frequency, FW and DW, ratio of no. of seedlings to no. of
regenerated calli
[108]
Oryza sativa
cv. IR64
SeedsCallus- AgNPs (10 mg/L), regeneration- AgNPs (5 mg/L)Callus- 82.4 ± 5.2%,
regeneration- 61 ± 6.3%
-AgNPs
(10 mg/L)
11.2 ± 0.64.9 ± 0.3Increased regeneration and rooting[88]
Panicum virgatumSeed/
internode
ZnONPs (20 and 30 mg/L)Callus induction-
90% (seed), 96% (internode),
shoot regeneration- 23.10 ± 2.1 (76%) (seed),
24.00 ± 0.01 (80%) (internode)
----Enhanced plant growth and development[122]
Phaseolus vulgarisHypocotylAgNPs (50 mg/mL)Callus- 97%----Increased callus formation, FW and DW[64]
Phoenix dactyliferaLeafletsCallus—MWCNTs (0.05 mg/L), SE formation and elongation—MWCNTs (0.1 mg/L)Callus- 3.80 g, SE- 24.04.3MWCNTs (0.1 mg/L) 5.36.0Increased embryogenesis and elongation of shoots[89]
Phoenix dactyliferaImmature inflorescencesCallus establishment- AgNPs (500 μg/L), callus differentiation- AgNPs (125 μg/L), SE formation- AgNPs (125 μg/L) (cv. Medjool); Callus establishment- AgNPs (500 μg/L), callus differentiation- AgNPs (500 μg/L), SE formation- AgNPs (125 μg/L)(cv. Sewi)Callus- 76.66%,
globular SE- 16.00,
direct SE- 68.33%,
germination- 0.81, multiplication- 1.00 (cv. Medjool);
Callus- 73.33%
globular SE- 17.33,
direct SE- 68.33%, germination- 0.92, multiplication- 1.00 (cv. Sewi)
----Increased SE formation[34]
Phoenix dactylifera cv. HayaniShoot tipAgNPs (1 mL/L)Callus- 4.60 g (FW), SE initiation- 9.39, SE development- 35.301.80---Increased SE length, no. and length of leaves, shoots were genetically uniform[90]
Rhizoma polygonatiTuberFe3O4NPs (0.4 mg/L)4-Fe3O4NPs (0.5 mg/L)9.5-Increased number of shoots and roots[123]
Rubus adenotrichosIn vitro shoot---SWCNTs-COOH (4 µg/mL)8.60 ± 5.750.6Increased rooting of shoots and growth of plants[124]
Saccharum spp.cv. Mex 69-290In vitro ShootAgNPs (50 mg/L)47.28 ± 1.695.55 ± 0.24---Improved regeneration and shoot length in temporary immersion bioreactors[91]
Simmondsia chinensisNodeMWCNTs (0.002 g/L)16.001.36---Improved regeneration[92]
Solanum lycopersiconHypocotyl (For callus),
cotyledonary nodes (For regeneration)
Fe3O4NPs (3 mg/L)Callus- 64.26 ± 0.38%, shoot-
8.2 ± 0.09 (cv. Nora);
Callus- 83.28 ± 0.94%, shoot-
10.8 ± 0.09 (cv. PS-10); Callus- 74.48 ± 0.39%, shoot-
9.7 ± 0.09 (cv. Peto);
Callus- 56.32 ± 0.47%, shoot-
6.6 ± 0.12 (cv. Roma)
----Better callus and shoot formationshowing resistance to salinity stress[93]
Solanum lycopersicum
cv. Edkawy
CotyledonZnONPs (15 mg/L)Regeneration-
83.34 ± 0.23%
(cv. Edkawy),
64.58 ± 0.15% (cv. Anna Aasa), 78.16 ± 0.23% (cv. Australische Rosen), 67.7 ± 0.47% (cv. Sankt Ignatius), 87.64 ± 0.58% (cv. Sandpoint)
----Improved salinity stress and regeneration frequency[4]
Solanum tuberosumLeafSiO2NPs (50 mg/L)Callus-
1.1 g (FW),
0.07 g (DW)
(cv. Proventa);
Callus-
1 g (FW),
0.05 g (DW)
(cv. Sante)
9 (cv. Proventa);
8 (cv. Sante)
-6 (cv. Proventa);
4 (cv. Sante)
10 (cv. Proventa); 8 (cv. Sante) Increased resistance to salinity stress in terms of various morphological traits[94]
Solanum tuberosum cv. White DesireeNodeAgNPs (2 mg/L)-7.8--12Increased shoot and root DW, root length and leaf area, decreased shoot length[95]
Solanum tuberosum cv. SpuntaSproutCSNPs (250 mg/L)90.97 ± 1.4112.40 ± 0.38---Production of potato virus Y (PVY) free plants[125]
Triticum aestivumMature embryoCallus- 1X-3X of all NPs (ZnO/CuO/γ-Fe3O4), embryogenic callus- 1x (CuO/γ-Fe3O4), SE formation- 3X γ-Fe3O4NPs
(genotypeKırik); Callus- 2XZnONPs/3XZnONPs/3XCuONPs, embryogenic callus-3XZnONPs, regeneration- 3X ZnONPs
(genotype ES-26)
Callus- 100%, embryogenic callus- 97.5%, SEs- 1.69, plantlet- 9.00
(genotypeKırik); Callus- 97.50%, embryogenic callus- 41.38%, SE- 1.70,
plantlet- 6.75
(genotype ES-26)
----Genotype Kırik: higher callus, SE and plantlet formation;
Genotype ES-26: same frequency of callus but less SE and plantlet formation
[126]
Triticum aestivumMature embryoCuNPs (0.015 mg/L) + AgNPs (4 mg/L)Callus- 90.00%, embryogenic callus- 84.67%, regeneration- 71.67% (genotype AS-2002); Callus- 95.00%, embryogenic callus- 78.00%, regeneration- 68.33% (genotype Wafaq-2001)----Increased callus and regeneration frequency[127]
Vanilla planifoliaIn vitro shootAgNPs (50 mg/L)14.89 ± 0.404.71 ± 0.23---Increased regeneration, shoot length, and biomass[33]
Vigna unguiculata cv. ÜlkemPlumule of embryoMgONPs (555 mg/L)10.00 (82.50%)1.45MgONPs (370 mg/L)0.75 (22.50%)0.72Increased shoot number, frequency, and rooting response[128]
AgNPs: silver nanoparticles; AgNO3: silver nitrate; AgNO3NPs: silver nitrate nanoparticles; AuNPs: gold nanoparticles; CH-ZnO: chitosan–zinc oxide nano-bioformulation; CNPs: carbon nanoparticles; CSNPs: chitosan nanoparticles; CuNPs: copper nanoparticles; CuONPs: copper oxide nanoparticles; DW: dry weight; FCNTs: functionalized carbon nanotubes; FeNPs: iron nanoparticles; Fe3O4NPs: iron oxide nanoparticles; FW: fresh weight; GFAgNPs: Gracilariafoliifera coated silver nanoparticles; IONPs: iron oxide nanoparticles; MgONPs: magnesium oxide nanoparticles; MWCNTs: multi-walled carbon nanotubes; NCNPs: nanocarbon nanoparticles; SE(s): somatic embryo(s); SiO2NPs: silicon dioxide nanoparticles; SWCNTs: single-walled carbon nanotubes; TiO2NPs: titanium dioxide nanoparticles; ZnNPs: zinc nanoparticles; ZnONPs: zinc oxide nanoparticles.

4.2. Metal Oxide NPs

Metal oxide NPs are another type of NP that has proven results on many crops (Table 3). In O. sativa ssp. indica cv. KDML105, comparisons of two NPs (ZnO and TiO2) suggested that ZnONPs showed toxic effects whereas TiO2NPs enhanced the regeneration frequency [118]. Later on, Chutipaijit and Sutjaritvorakul [121] also documented a positive influence of TiO2NPs on the indirect regeneration of rice cultivars Suphanburi1 and Suphanburi90. Zafar et al. [65] observed the negative influence of ZnONPs on B. nigra stem explant as only a few shoots and roots were emerged from the callus. Later on, when the CuONP fortification of NPs in media was performed, it caused root emergence from callus from both leaf and stem explants of B. nigra; also, the biochemical potency of the roots which were emerged from the callus was different as compared to the seedlings, and this is due to lower concentrations of NPs used for leaf and stem explants as compared to seeds [5]. Similar to rice cultivar KDML105, TiO2NPs displayed a beneficial effect as compared to ZnONPs on callus induction and plant regeneration in another cultivar RD49 [120]. Later on, the negative effect of ZnONPs on N. tabacum cv. Bright Yellow-2 (BY-2) cells with respect to viability, packed cell volume, and FWs were also reported. There was a significant decrease in mitotic index and changes in cell structure such as endoplasmic reticulum, mitochondrial dysfunction, and Golgi apparatus, along with an increase in ROS and reactive nitrogen species (RNS) [87]. Further, to evaluate the programmed cell death, an increase in plasma membrane integrity, and activities of protease, caspase-like, and acid phosphatases were observed along with nuclear cell morphology and DNA fragmentation, suggesting the phytotoxic effect of ZnONPs. Some studies suggested that metal and metal oxide NPs showed similar effects on cultures, e.g., zinc nanoparticles (ZnNPs) and ZnONPs both increased regeneration and rooting responses in Musa paradisiacal L. Upon analysis, it was confirmed that treatment with both types of NPs elevated total proline and chlorophyll contents as well as increased the activities of antioxidant enzymes in shoots [43]. The comparison between ZnO bulk and NPs on B. napus cv. Hayola 401 showed that the application of ZnONPs improved the formation of calli in terms of FW and DW in comparison to bulk ZnO [110]. Irum et al. [111] documented that the callus of C. arietinum var. Punjab-Noor 09 and Bittle-98 showed good callus proliferation from embryo axes and embryo axes with the adjacent parts of cotyledon explant on media containingiron oxide nanoparticles (IONPs) in comparison to control. However, transferring this callus on regeneration medium showed only an increase in callus size which failed to undergo redifferentiation.
The effect of iron oxide nanoparticles (Fe3O4NPs) on hypocotyls of Daucus carota L. cv. Berlicum revealed that its lower concentrations facilitated somatic embryo formation, but higher concentrations ceased its differentiation [112]. Li et al. [123] compared Fe3O4NPs and micro-cube on the in vitro morphogenic response of Rhizoma polygonati Odorati, and confirmed the positive influence of the former than later on shoot and root formation. In L. usitatissimum cultures, augmentation of the medium with Fe3O4NPs during the culture initiation phase induced somatic embryogenesis, but its addition after callus formation led to rhizogenesis [115]. An interesting study was carried out on two genotypes of Triticum aestivum L. Kırik and ES-26, where Fe, Cu, and Zn of MS media were replaced by NP versions of elements (ZnO, CuO, and γ-Fe3O4) in the same concentration (1X) and two and three times higher amount (2X and 3X). The results suggested that the overall response in terms of callus, somatic embryo, and plantlet formation was less for the ES-26 genotype in comparison to the control and Kırik genotypes, suggesting variation in NPs’ effect between the genotypes [126]. In the same way, Malik et al. [127] compared CuSO4 and AgNO3 with their NP counterpart and evaluated their effect on the in vitro response of two genotypes of T. aestivum cv. AS-2002 and Wafaq-2001. They have suggested that the combined application of both the NPs was more beneficial than their individual usage, also the response was better for cv. AS-2002 than Wafaq-2001. The majority of studies suggested that the enhancement in growths of plants in response to NPs might be due to the up-regulation or down-regulation of different hormonal pathways, especially the cytokinin, which evoked culture growth [110].

4.3. Role of Green NPs

Rapid progress in the field of nanotechnology has enabled the synthesis of NPs of different types, sizes, and morphologies; and, NPs generated using plants are reported to have a less toxic and more stable effects [129]. CuONPs synthesized utilizing Azadirachta indica leaf extracts were used to evaluate its effect on four varieties of O. sativa (Super Basmati, Basmati 2000, Basmati 370, and Basmati 385). They have suggested that the augmentation of NPs in the media improved callogenesis and organogenesis; however, variation has been detected between the varieties [119]. In another study, AgNPs coated with marine red alga Gracilaria foliifera (GFAgNPs) showed growth-stimulating properties in comparison to the traditionally used hormones like 6-benzylaminopurine (BAP) and kinetin (Kn) on Alternanthera sessilis L. regeneration [109]. In line with this, AgNPs synthesized using Parthenium hysterophorus extract when augmented into the media have promoted callus formation, shoot regeneration, and rhizogenesis in O. sativa cv. IR64, and also suggested the inhibitory effect of AgNPs on ethylene perception [88]. Likewise, AgNPs synthesized using leaf extract of Ochradenus arabicus increased the shoot number, length, FW, DW, chlorophyll, total protein, and total proline contents as well as activities of enzymes like SOD and CAT in cultures of Maerua oblongifolia (Forssk.) A. Rich [83]. Green synthesis of ZnONPs using Cymbopogon citrates extract enhanced callus induction and regeneration from seed and internode explants of Panicum virgatum L. [122]. They have observed that ZnONPs have a positive influence on both the explant, but a 20 mg/L concentration was suitable for the seed explant whereas, for the internode explant, 30 mg/L was better. In addition, the application of manganese oxide nanoparticles (MgONPs) synthesized using walnut shell extract increased the regeneration response in Vigna unguiculata L. Walp cv. Ülkem [128] (Table 3).

5. Ramifications of NPs on In Vitro Cultures

5.1. NPs to Mitigate Stress and Virus Resistance

The application of nanomaterials can help in improving tolerance against biotic and abiotic stresses via in vitro cultures that help in the micropropagation of stress-resistant crops. In five cultivars of S. lycopersicum, the salinity stress provided using NaCl was used to evaluate the effect of ZnONPs, and results showed that the treatment significantly improved callus growth and regeneration frequency (Table 3), as well as the content of different elements, protein, and antioxidant enzymes (Table 2), which justifies the resistance towards salinity-induced stress [4]. In another study to improve salinity stress in potato cv. Sante and Proventa, Gowayed et al. [94] utilized silicon dioxide nanoparticles (SiO2NPs) and observed better morphological traits when SiO2NPs along with NaCl were used, signifying the improved plant growth under salinity stress. A comparative study was carried out in four cultivars of S. lycopersicon in which the effect of NPs like Fe3O4 and ZnO was seen and it was observed that the aforesaid treatments significantly eliminated the salinity stress in callus and shoot formation, and the maximum response for cv. PS-10 and least for cv. Roma was recorded [93]. Drought stress has been a serious concern for the crops and thus in F. ananassa, a combination of iron nanoparticles (FeNPs, 0.8 ppm) + salicylic acid (SA, 0.01 mM) proved beneficial as an increased biomass of plantlets with a higher percentage of relative water content (RWC) and membrane stability index (MSI) was observed, which confirmed that the cultures withstand drought stress [81]. Another concern for crop productivity is infection with virus, and potato virus Y (PVY) causes serious loss in the yield and quality of potatoes. Recently, to overcome this, Elsahhar et al. [125] evaluated the role of CSNPs and suggested that their treatment proved effective in producing virus-free plants.

5.2. NPs’ Influence on Induction of Somaclonal Variation

The addition of NPs in the culture media is known to affect the plant cells at biochemical and/or molecular levels; thus, there are chances that it might produce somaclonal variations. Somaclonal variation is one of the consequences of in vitro cultures which is associated with any changes in chromosome number, chromosome structure, DNA sequence, DNA methylation, mitotic crossing over, and activation of transposable elements [130]. However, it has advantages as well as disadvantages, and the main advantages are the development of useful characteristics like plant size, flower colour, leaf variegation, fruit ripening, resistance to biotic and abiotic stresses, and elevated secondary metabolites production [131]. Sometimes, the higher concentrations of NPs cause toxicity in plants which affects the mitotic index, DNA integrity, and alters the protein and DNA expression [54]. Some studies reported the change in ploidy levels, e.g., L. usitatissimum calli grown on media containing carbon NPs (CNPs) showed an increased number of tetraploid cells and level of DNA methylation [114]. In another study, Kokina et al. [132] observed a high rate of somaclonal variation in calli and regenerated shoots of L. usitatissimum when the medium consisted of AuNPs and less variation in the presence of AgNPs. Hence, it has been suggested that the regenerated plants need to be investigated if the study aims to induce true-to-type plants.

5.3. NPs as an Elicitor for In Vitro Production of Secondary Metabolites

Earlier studies have suggested that NPs act as signal components and modify the physiological and metabolic responses of plants. This has opened an alternative strategy for the production of targeted secondary metabolites in plant cell cultures using NPs as an elicitor [133]. The effectiveness of NPs is due to their small size, as they can easily attach to the plant cell walls, destroy them or change their permeability, and thus significantly affect the cellular metabolism [134]. This is due to the dual role played by NPs, first by acting as an efficient nutrient and second by acting as an elicitor, which enhances secondary metabolite production [135]. Recently, the exploitation of NPs for the production of economically and commercially important secondary metabolites from crop species has been well documented (Table 4). Al-Oubaidi and Mohammed-Ameen [136] found that AgNO3NPs increased callus formation in Calendula officinalis L. at a 0.3 mg/L concentration, but its higher concentration (1.2 mg/L) favored the synthesis of various essential oils. Hairy root cultures are one of the promising ways for secondary metabolite production and in Datura metel L. hairy roots, the addition of AgNPs proved beneficial for the enhancement of biomass as well as atropine content in comparison to AgNO3 and biotic elicitors (Bacillus cereus and Staphylococcus aureus) [137]. In Cucumis anguria L. hairy roots, AgNPs significantly elevated the content of different metabolites in comparison to its bulk counterpart (AgNO3) [80]. In a cell suspension of Corylus avellana L. cv. Gerd Eshkevar, augmentation of media with AgNPs increased the taxol [78] as well as taxane contents [79]. AgNPs obtained via green synthesis using an extract of Bacillus marisflavi increased the contents of different groups of compounds like hydroxybenzoic acids, hydroxycinnamic acids, and flavonols in the cell suspension of Momordica charantia L. [84]. A comparison between AuNPs and AgNPs has been carried out for the shoot culture of Lavandula angustifolia Mill. cv. Munstead where shoots grown in the presence of both NPs significantly affected the composition of essential oil. Their addition in media decreased the content of low-molecular-weight compounds (α- and β-pinene, camphene, δ-3-carene, p-cymene, 1,8-cineole, trans-pinocarveol, and camphoriborneol), which were replaced with high-molecular-weight compounds (τ- and α-cadinol 9-cedranone, cadalene, α-bisabolol, cis-14-nor-muurol-5-en-4-one, and (E,E)-farnesol) [138]. In callus cultures of Allium sativum L., the contents of allicin, di-allyldisulfide, and vinyldithiin have been elicited using AgNPs and NaCl [139].
The addition of CuNPs and cobalt nanoparticles (CoNPs) in shoot cultures of Mentha longifolia L. revealed that they have positively influenced linalool synthesis and negatively affected linalyl acetate content. Their results also confirmed that CuNPs were better for regeneration, but maximum essential oil synthesis was observed in the presence of CoNPs [116]. Contrarily, CuONPs significantly enhanced the contents of glucosinolates, phenolic compounds, hydroxy-benzoic acids, hydroxycinnamic acids, and flavonols in hairy roots of B. rapa spp. pekinensis. They have also confirmed the up-regulation of different pathway genes such as MYB34, MYB122, MYB28, MYB29, PAL, CHI, and FLS after the exposure of hairy roots with NPs [140]. Al-Oubaidi and Al-Khafagi [141] compared the effectiveness of MgONPs and CuONPs on Punica granatum L. callus cultures for metabolite synthesis, and reported that the contents of metabolites varied according to concentration and type of NPs augmented in the media. They have observed that the level of gallic acid, tannic acid, ellagic acid, chlorogenic acid, acacetin, cinnamic acid, and geniste in was increased in the presence of MgONPs, whereas CuONPs elevated the levels of brevifolincarboxylic acid, catechin, rutin, coumaric acid, ferulic acid, benzoic acid, and kaempferol. Al-Khafagi and Al-Oubaidi [142] studied the contents of similar metabolites of the same species after treatment using NPs on shoot tip culture. However, their results differed as compared to an earlier report as the majority of the compounds were elicited after MgONP treatment except for catechin and kaempferol which were increased with CuONPs. Similarly, TiO2NPs stimulated higher lignin content in the cell suspension of L. usitatissimum cv. Kerman Shahdad as compared to ZnONPs [82]. In another cultivar (Barbara) of L. usitatissimum, different lignans (secoisolariciresinoldiglucoside and lariciresinoldiglucoside) and neolignans (dehydrodiconiferyl alcohol glucoside and guaiacylglycerol-β-coniferyl alcohol ether glucoside) were detected in seedlings and calli, and it was noted that a higher concentration (500 mg/L) of ZnONPs proved beneficial in elevating the contents in seedlings, whereas a lower concentration (10 mg/L) favoured metabolite synthesis in callus [68]. The effectiveness of the chitosan–zinc oxide nano-bioformulation (CH-ZnO) combination proved better in comparison to ZnONPs not only for callus biomass but also for tannin content, whereas nicotine content was better when ZnONPs were used in media containing leaf-derived callus of Nicotiana benthamiana Domin [86]. In a recent study, CuO and ZnO NPs were synthesized using Nigella sativa L. extract, and their effect on three varieties of V. radiate (var. NCM-13, MgAT-7, and MgAT-4) was assessed. Their results confirmed that both the NPs affected different metabolites in all three varieties as increased phenolics were recorded in the presence of CuONPs, whereas increased glycoside was detected in the presence of ZnONPs [135]. Recently, Abu-Al hayl and Al-Oubaidi [143] carried out an experiment using SiO2NPs on the callus of Tagetes erecta L. where amounts of gallic acid, syringic acid, ellagic acid, quercetin, quercetagetin, lutein, and kaempferol increased significantly at a higher concentration in comparison to the lower concentration of NPs (Table 4).
Table 4. Effects of various NPs on elicitation of metabolites in in vitro cultures of different crops.
Table 4. Effects of various NPs on elicitation of metabolites in in vitro cultures of different crops.
PlantMetabolite(s)Culture TypeNanoparticle (NP) Treatment and TimeRemarksReference
Allium sativumAllicin, di-allyldisulfide and vinyldithiinCallusAgNPs (2 mg/L),
time- 4 w
Increased content of all metabolites[139]
Brassica rapa spp. pekinensisGlucosinolates (gluconasturtiin, glucobrassicin, 4-methoxyglucobrassicin, neoglucobrassicin, 4-hydroxyglucobrassicin, glucoallysin, glucobrassicanapin, sinigrin, progoitrin, and gluconapin), phenolic compounds (flavonols, hydroxybenzoic and hydroxycinnamic acids), hydroxy-benzoic acids (vanillin, p-hydroxybenzoic, protocatechuic, syringic, gentisic acids), hydroxycinnamic acids (chlorogenic, p-coumaric, ferulic, and t-cinnamic acids) and flavonols (myricetin, quercetin, catechin, kaempferol, rutin, naringenin and hesperidin)Hairy rootsCuONPs (100 mg/L),
time- 48 h
Increased content of all metabolites and expression of genes[140]
Calendula officinalisα-Pinene, β-pinene, ρ-cymene, α-thujene, calendulaglycoside, α-cadinene, cadinol, t-muurolol, 1,8-cineole and limoneneCallusAgNO3NPs (1.2 mg/L),
time- 4 w
Increased contents[136]
Capsicum annum and
C. frutescens
CapsaicinCell suspensionAgNO3NPs (3 mg/L),
time- 6 d
Increased content[144]
Corylus avellana cv. Gerd EshkevarTaxolCell suspensionAgNPs (5 ppm),
time- 1 w
Increased content, decreased cell viability[78]
Corylus avellana cv. Gerd EshkevarTaxol and baccatin IIICell suspensionAgNPs (5 ppm),
time- 24 h
Increased contents[79]
Cucumis anguriaHydroxybenzoic acids (p-Hydroxybenzoic acid, gallic acid, protocatechuic acid, syringic acid, gentisic acid, salicylic acid, vanillic acid, β-resorcylic acid, hydroxycinnamic acids (Caffeic acid, p-coumaric acid, o-coumaric acid, ferulic acid, chlorogenic acid, t-cinnamic acid), flavonols (Myricetin, quercetin, kaempferol, catechin, rutin, naringenin, biochanin A), phenolics (vanillin, veratric acid, homogentisic acid, hesperidin)Hairy rootAgNPs (1 mg/L),
time- 21 d
Increased biomass and content[80]
Datura metelAtropineHairy rootsNanosilver (conc.- NM),
time- 48 h
Increased biomass and content[137]
Lavandula
angustifolia
cv. Munstead
Essential oilsShootsAgNPs and AuNPs (10 mg/dm3),
time- NM
Decreased content of low-molecular-weight compounds (e.g., α- and β-pinene, camphene, δ-3-carene, p-cymene, 1,8-cineole, trans-pinocarveol, camphoriborneol), and increased content of high-molecular-weight compounds (τ- and α-cadinol 9-cedranone, cadalene, α-bisabolol, cis-14-nor-muurol-5-en-4-one, (E,E)-farnesol)[138]
Linum
usitatissimum
cv. Kerman Shahdad
LignanCell suspensionTiO2NPs (150 mg/L),
time- 72 h
Increased content[82]
Linum
usitatissimum
cv. Barbara
Lignans (secoisolariciresinoldiglucoside, lariciresinoldiglucoside) and neolignans (dehydrodiconiferyl alcohol glucoside and guaiacylglycerol-β-coniferyl alcohol ether
glucoside)
Seedlings and callusZnONPs (500 mg/L)/(10 mg/L),
time- 30 d
Higher ZnONPs increased contents in seedling; lower ZnONPs increased contents in callus[68]
Mentha longifoliaEssential oils (Linalool and linalyl acetate)ShootsCoNPs (0.8 mg/L),
time- 30 d
Increased linalool and decreased linalyl acetate contents[116]
Momordica
charantia
Hydroxybenzoic acids (p-Hydroxybenzoic acid, gallic acid, protocatechuic acid, syringic acid, gentisic acid, salicylic acid, vanillic acid, β-resorcylic acid), hydroxycinnamic acids (Caffeic acid, p-coumaric acid, o-coumaric acid, ferulic acid, chlorogenic acid, t-cinnamic acid), flavonols (Myricetin, quercetin, kaempferol, catechin, rutin, naringenin, biochanin A)Cell suspensionAgNPs (5 mg/L),
time- 48 h
Increased contents[84]
Nicotiana
benthamiana
Tannin and nicotineCallusCH-ZnO (400 ppm)/ZnONPs (200 ppm),
time- 7 d
CH-ZnOincreased tannin content;
ZnONPs increased nicotine content
[86]
Olea europaeaOleuropein, OH-tyrosol, ligustroside and oleaceinShoot tipsAgNO3NPs (1 and 2 mg/L),
time- 30 days
Increased contents[145]
Oryza sativa
cv. Swarna
CarotenoidsSeedling leavesAgNPs (40 ppm),
time- 14 d
Increased content[57]
Punica granatumTannins (Gallic acid, tannic acid, ellagic acid, brevifolincarboxylic acid), phenols (chlorogenic acid, catechin, rutin, coumaric acid, ferulic acid, benzoic acid, acacetin, cinnamic acid, genistein, kaempferol)Callus, shoot tipMgONPs (2.5–10 mg/L)/CuONPs (5–20 mg/L),
time- 21 d
Callus: MgONPs, gallic acid, tannic acid, ellagic acid, chlorogenic acid, acacetin, cinnamic acid, genistein;
CuONPs: brevifolincarboxylic acid, catechin, rutin, coumaric acid, ferulic acid, benzoic acid, kaempferol;
Shoot tip- MgONPs- gallic acid, tannic acid, ellagic acid, brevifolincarboxylic acid, chlorogenic acid, rutin, coumaric acid, ferulic acid, benzoic acid, acacetin, cinnamic acid, genistein;
CuONPs: catechin, kaempferol
[141,142]
Tagetes erectaGallic acid, syringic acid, ellagic acid, quercetin, quercetagetin, lutein and kaempferolCallusSiO2NPs (200 mg/L),
time- 30 d
Increased contents[143]
Vigna radiata
var. NCM-13, MgAT-7, and MgAT-4
Phenolic and glycosidesCallus and shootsCuONPs (0.5 mg/L)/
ZnONPs (0.5 mg/L),
time- NM
Overall, callus synthesized more metabolite than shoots, CuONPs-increased phenolics,
ZnONPs increased glycoside content
[135]
AgNPs: silver nanoparticles; AgNO3NPs: silver nitrate nanoparticles; AuNPs: gold nanoparticles; CH-ZnO: chitosan–zinc oxide nano-bioformulation; CoNPs: cobalt nanoparticles; CuONPs: copper oxide nanoparticles; MgONPs: magnesium oxide nanoparticles; TiO2NPs: titanium dioxide nanoparticles; ZnONPs: zinc oxide nanoparticles. NM: not mentioned.
NPs, when acting as elicitors, either bind directly to elicitor-binding sites or produce endogenous messenger molecules that will bind to the sites and initiate the responses. This is due to NPs’ interaction with some of the plant cell-wall and membrane components [146]. Initially, NPs involve an active exchange of ions like Na+/K+/Cl effluxes and Ca2+/H+ influxes through the plasma membrane into the cytosol. Among all these, Ca2+ influxes are considered asthe most important event due to their involvement in various physiological and cellular pathways, and it plays a pivotal role in the first steps of the elicitation mechanism [147]. In another hypothesis, a cascade of events has been described where Ca2+ flux movements and ROS produced by oxidative burst act as messengers that led to the up-regulation/phosphorylation of mitogen-activated protein kinase (MAPK) [146] or affected ATPase activity and increase the cytoplasmic acidity, leading to metabolite synthesis [148]. The structure of the plant cell wall is consistent with the size of the NPs for entry into the cell where the ROS accumulation can be triggered [149]. This ROS will interfere with the plasma membrane and affect the permeability of the cells; thus, as a result, more NPs enter into cells, causing more stress and stimulating the production of stress-induced secondary metabolites [21]. Other studies suggested that NADPH and other oxidases also become activated through Ca2+ movements and they are responsible for the generation of ROS in plant cells [150,151]. This ROS generation results in the activation of cGMP-dependent protein kinase and the phosphorylation of MAPKs, which results in transcriptional reprogramming events of genes of secondary metabolite pathways [152]. Kohan-Baghkheirati and Geisler-Lee [153] stated that the G-proteins (Guanine nucleotide-binding proteins) can also activate the metabolite accumulation through de novo biosynthesis of stress-signaling compounds such as SA, jasmonic acid (JA), and methyl jasmonic acid (MeJA). Similarly, it is also suggested that the expression of genes involved in oxidative stress and the accumulation of ROS acts as a signal for metabolite synthesis [154,155].

5.4. NPs’ Uptake, Biochemical and Molecular Attributes in Plant Cell

The pore size of plant cell walls is usually in the range of a few nanometers which acts as a barrier to foreign materials [156]. However, the diameters of NPs are usually smaller in comparison to the diameters of the cell-wall pores; hence, they can easily penetrate and reach the plasma membrane. On the other hand, if the size of NPs is higher than the pore size, they enter the cell either by changing the size of existing pores or by inducing new larger pores in the cell wall [7]. After crossing the cell wall, NPs reach the cell membrane and are then internalized towards cytosol or other organelles either by endocytosis, specific membrane-bound transporter proteins (aquaporins), or through the induction of new pores using ion-carrier substances [157,158]. NPs help in the regulation of processes like cell signaling and the regulation of the plasma membrane; they bind with different cytoplasmic organelles and interfere with the metabolic processes at the site [159,160]. The translocation of NPs also depends on the concentration and the nature of the plant species, but usually, the passages of uptake and transportation are via the xylem [161], and it was also found that NPs followed the stomatal pathway in the leaf [162]. Further, they can be transported from one cell to another either by the apoplastic or symplastic pathway or via plasmodesmata [163,164]. Nair et al. [165] suggested that NPs, after crossing the membrane, stick with the membrane and interfere with different organelles. It has been reported in many studies that the uptake of NPs is closely associated with the absorption of moisture and nutrients from the media [166,167]. In addition, few in vitro studies have reported the uptake of NPs via clathrin-independent and -dependent pathways in N. tabaccum [168] or the endosomal pathway in Catharanthus roseus (L.) G. Don [169]. Kokina et al. [99] documented that the plant cells take up plant growth regulators (PGRs) for differentiation and redifferentiation, and the metal NPs are transported along with PGRs. It is also suggested that NP internalization becomes different in the case of cell-suspension cultures, where endocytosis in the vacuole occurs from the apoplast through vesicles formed from the plasma membrane [170,171], e.g., carbon nanotubes (CNTs) entered the cell wall through endocytosis and moved towards the cell membrane in tobacco cell suspension [172]. In comparison, NPs enter through parenchymatous intercellular spaces which assist the diffusion of liquid solution to cotyledon in seeds [173,174]. Physiologically, they affect the plant metabolism by delivering micronutrients [175], along with this, they regulate various gene functions [176] and interfere with different oxidative processes [177]. Further, they take part in electron transfer in plants, thus increasing the activity of many enzymes and influencing plant mineral nutrition [178,179]. As the size of NPs accumulates in intracellular spaces, their higher concentration renders toxicity [37,180,181]. However, it is noted that the toxic effects of NPs are dose-dependent [182], and a negative response is due to injury in the cell wall and membrane [183].
Plants induce various responses to combat stress such as the production of various ROS like singlet oxygen, superoxide, hydrogen peroxide, and hydroxyl radical, which are the main oxidative outbursts in plant cells after stress induction [177]. Depending on their concentration, ROS can work positively or negatively. At low concentration, they act as secondary messengers in intracellular signaling that induce several responses in plant cells including stress tolerance [62,184], whereas a higher concentration of ROS causes damage to biomolecules by apoptosis or necrosis, a disruption of the metabolic pathway through inactivation of the enzyme thatresults in plant cell death [185,186]. ROS are involved in many stress adaptations in plants [187,188], and therefore robust defense mechanisms have been developed by plants, viz, enzymatic (SOD, POD, CAT, APX, etc.), non-enzymatic (ascorbate, glutathione, carotenoids, tocopherols, phenolics, etc.), and antioxidant production [189]. Plants also activate the MAPK pathway that boosts the plant antioxidant elements to come in contact with ROS [190]; also, H2O2 and MDA are the measures to evaluate the stress in plants as they modulate the unstable ROS [189,191]. Proline is also a well-accepted stress marker as it has ameliorative properties suggesting its involvement in mitigating oxidative stress [192]. Its accumulation is reported to act as an antioxidant for neutralizing the toxic effects of ROS and it is also known to maintain the structure of proteins and membranes of cells [193]. Similarly, the activities of enzymes are also evaluated and the most commonly analyzed is phenylalanine ammonia-lyase (PAL). It is the first enzyme of the phenylpropanoid pathway that synthesizes many compounds which are the major protectants against stress, and hence its evaluation can be correlated with the effect of stress on plants [194]. Pigments like chlorophyll and carotenoids are attributed as precursors of abscisic acid (ABA) that modulates stress responses [195]. The loss in chlorophyll causes a surplus of electrons to combine with molecular oxygen and eventually form ROS [196]. The chlorophyll donates an electron to a series of molecular intermediates called an electron transport chain [197], whereas carotenoids are structural components of the photosynthetic antenna and reaction center complexes that protect photosynthetic organelles against harmful photo-oxidative processes [198].
The enzymes require metallic ions as a co-factor to complete the function during photosynthesis [199], and due to the physical properties of NPs, they dissociate quickly in the cytosol and aid enzymes at the cellular level to facilitate photosynthesis [200]. The addition of NPs increases the activities of enzymes like POD, CAT, and nitrate reductase, that also favours regeneration by affecting important physiological and biochemical processes [201]. Parida and Das [202] documented that the treatment of cultures with NPs induced a better chlorophyll a/b ratio indicating the activeness of PS-I and PS-II, which might be beneficial for regeneration. Various biochemical parameters are being analyzed to observe the stress imposed on plant cells or tissues after the application of various NPs (Table 2). It has been suggested that Fe and Zn stimulates the antioxidant enzyme activity in plants, and helps in the reduction of the free radical effect [203]; FeNPs are also reported to increase the gene expression of enzymes involved in photosynthesis and thus assist in enhancing the process [204]. Likewise, ZnONPs improved plant growth by affecting the electron transfer chain and increasing enzymatic antioxidants, reducing ion leakage, and improving the Hill reaction [205]. Zn also plays a vital role as a co-factor for several enzymes comprising superoxide, catalase, and dismutase, which inhibit ROS stress [206]. In seedling and callus of L. usitatissimum cv. Barbara, the gradual increase in ROS production was observed as the concentration of ZnONPs increased and it elevated the formation of membrane lipid peroxidation, protein carbonylation, and 8-oxo guanine [68]. Similarly, in B. nigra callus and seedling ZnONP treatment, increased antioxidant activity, phenolic, and flavonoid contents [65] were observed, whereas in cell suspension cultures of cultivar Kerman Shahdad of L. usitatissimum, increased PAL and CAD activities, and levels of total phenols [82] were observed. Likewise, in in vitro cultures of different crops like M. paradisiacal cv. Grand Nain [43], different cultivars of S. lycopersicum [4], N. tabacum cv. BY-2 [87], etc., augmentation of media with ZnONPs changes the biochemical parameters. In N. benthamiana, a combination of chitosan–zinc oxide nano-bioformulation (CH-ZnO) increased chlorophyll, carotenoid, proline contents, and enzyme (PAL and AO) activities, but decreased MDA and H2O2contents [86] (Table 2). In addition, iron oxide nanoparticles (Fe2O3NPs) changed the activities of different enzymes and antioxidant compounds in Cichorium intybus L. [77] and S. lycopersicon [93], and FeNPs in F. ananassa [81].
Cu is another important metal element, but at a higher level, it induces toxicity due to its binding with sulfhydryl groups in proteins which eventually inhibit enzyme activity [207]. Similarly, higher concentrations of CuONPs adversely affected the growth as Cu ions released from NPs are impermeable to the plasma membrane, thus causing a deficiency of essential nutrients [208]. CuONPs inside the cell taken up by lysosomes increase the release of Cu ions that ultimately produce intracellular ROS [209,210]. It also induces oxidative stress by catalyzing the formation of OH radicals from the non-enzymatic chemical reactions between superoxide and H2O2 [211]. Alternatively, the positive effect of CuONPs at an optimum concentration on callus induction can be explained as Cu being an essential nutrient in plant growth and acting as a co-factor in many metalloproteins. Cu also acts as a structural element in regulatory proteins and is involved in important physiological processes like the electron transport chain, hormone signaling, and cell wall metabolism [212]. Studies revealed that the application of CuONPs counteracts stress by changing various biochemical reactions in V. radiate [71], C. arietinum [66], and B. nigra [5] (Table 2). Ti is another essential element that also increases the nutrient absorption of metals like Ca, Mg, Zn, and P [213]. It has been reported that TiO2NPs at a proper concentration promotes plant growth by assisting water absorption inplant cells and inducing cellular metabolism [214], by activating photosynthetic complexes, Rubisco carboxylase activity and nitrogen metabolism in the plant cell [215,216]. Mandeh et al. [42] reported that TiO2NPs facilitated plant growth as they play a role similar to PGRs like cytokinin and gibberellic acid (GA3), whereas SiO2NPs had increased the levels of GA3 in the cells, having a plant hormone-like property and play a vital role in cell division, and consequently increased the elongation [217]. TiO2NPs showed an influence on cultures of L. usitatissimum cv. Kerman Shahdad for variation in enzyme activities [82]. Gowayed et al. [94] studied S. tuberosum cv. Sante and Proventa under SiO2NPs influence and observed that it increased the number of protein bands in both cultivars compared to control and NaCl treatment. This increase in bands indicated that SiO2NPs activated genes which are important proteins associated with salt-stress resistance. Elevated activities of antioxidant enzymes (GPX and SOD) were also observed (Table 2).
The role of ethylene in in vitro plant regeneration has been well documented [218], and Ag+ ions are known to inhibit ethylene action by replacing Cu+2 ions with Ag+ and blocking ethylene receptor (ETR1) [219]. The beneficial effect of Ag can also be attributed to enhance polyamine biosynthesis rather than reduce ethylene production [220], and an increase in auxin efflux independent of ethylene response that affects plant growth [221]. The supplementation of AgNPs can enhance the plant cell’s nutrient and water uptake from culture media by mutilating the cell wall [222]. Another hypothesis suggested that AgNPs modify the structural components of cellular membranes, and macromolecules, influence cell division and defense systems, and interfere with the physiological and biochemical processes of plants by altering the gene expression [223]. But, the higher concentration of AgNPs cause lipid peroxidation because of the ROS generation, inhibition of ethylene production, and restriction in the electron transport chain of mitochondria and chloroplast, which all lead to oxidative burst, rise in ROS concentration, and eventual cell death [200,224]. Whereas at low concentrations, it modulates the redox status of plants, because of its efficient catalytic activity in redox reactions by acting as electron relay centers [225] and its ability to support electron exchange with Fe2+ and Co3+ [226]. Vannini et al. [227] observed that AgNPs cause changes in proteins involved in redox regulation and sulfur metabolism; they also alter some proteins related to the endoplasmic reticulum and vacuole. In B. juncea var. pusa jaikisan, they have increased chlorophyll content and the activities of enzymes, but MDA, proline, and H2O2 content were decreased [189]. Similarly, decreased MDA, proline, and H2O2 levels were also observed after AgNP treatment in cultures of O. sativa cv. IR64 [88]. In addition, variation in response has been well documented between crop species, as in some crops the contents were found to be decreased, e.g., B. oleracea var. sabellica ‘Nero di Toscana’ [59]. On the contrary, increased metabolites and enzyme activities have been reported in M. charantia [84], Caralluma tuberculata N.E.Br. [76], and Maerua oblongifolia (Forssk.) A. Rich [83]. Jamshidi et al. [78] had observed that AgNPs in the cell suspension culture of C. avellana cv. Gerd Eshkevar showed a positive influence on the contents of ascorbate peroxidase (APX), CAT, H2O2, and PAL, but they decreased SOD and POD activities, and total soluble phenol content. Recently, in B. napus shoots, they increased metabolites and H2O2 levels but no change in phenolics was found [60]. The examples of different NPs on biochemical changes in seedlings and different cultures of important crops are given in Table 2.
NPs after internalization evoked changes at molecular levels as metal NPs are known to induce systemic stress, and to overcome this stress, they alter the expression of genes [228]. CNT treatment is known to affect the expression of the water channel proteins (aquaporins) [229] that are considered to be crucial for the process of seed germination and plant growth [230]. The expression of water channel genes (aquaporin, LeAqp2) is reported to be activated in response to MWCNTs in tomato seedlings [73,229] and tobacco cells [231]. Villagarcia et al. [73] revealed that MWCNTs affect the expression of genes regulating cell division and cell wall extension in treated cells, resulting in faster growth than the unexposed control cells. However, a few in vitro studies that were carried out on gene expression analysis, e.g., Nair and Chung [66] analyzed the effect of CuONPs on seedlings of C. arietinum and correlated the biochemical changes with oxidative stress response genes such as SOD and CAT, but there was no significant change in the expression of APX. Likewise, on seedlings of B. rapa ssp. rapa, AgNPs up-regulated the expression of different genes related to antioxidant defense (catalase, CAT; peroxidase, POD; glutathione S-transferase, GST), biotic and abiotic stresses (pathogenesis-related gene 1, PR1; lipoxygenase 2, LOX2), carotenoids (β-cyclase, CYB; zeaxanthinepoxidase-1, ZEP1), anthocyanins (production of anthocyanin pigment 1, PAP1; anthocyanin synthase, ANS; phenylalanine ammonia-lyase, PAL), and glucosinolates (BrMYB28; BrMYB29; BrMYB34; BrMYB51; sulfotransferase, St5C; and superroot1, SUR1). Further, over-expression of the Geranyl diphosphate synthase gene (GPPS gene), a key gene involved in the thymoquinonebiosynthesis pathway, has been observed in N. sativa after TiO2 and SiO2NPs [232]. Manickavasagam et al. [88] depicted AgNPs in media containing O. sativa L. cv. IR64 seeds showed up-regulation of ethylene (ERF063), ABA (OsRab16), auxin (OslAA1), cytokinin (RR2), and gibberellic acid (PBZ1) responsive genes, justifying the stress induced by NPs treatment. In addition, treatment of M. charantia with selenium nanoparticles (SeNPs) showed a variation in methylation-susceptible loci (MSL) between the control and treated group which suggested an epigenetic modification in response to NPs. The results also revealed that there was significant up-regulation of transcription factor WRKY1, and genes like PAL and 4-coumarate:CoA ligase (4CL) [233].

6. NPs as a Tool for Genetic Engineering in Crops

Genetic engineering has proven useful in the face of climate change and the growing global population by bestowing desirable genetic traits and enhancing crop productivity. The delivery of genetic materials such as DNA and small interfering RNA (Si-RNA) is important for the development of pest, pathogen, and stress-resistant strains of crops by altering the gene expression [234,235]. The bottleneck in genetic transformation is the plant cell wall, which causes obstacles such as targeting the delivery system, transportation through the cell membrane, uptake and degradation in endolysosomes, and intracellular trafficking of DNA to the nucleus [236]. Similarly, the traditional delivery methods also have some demerits like viral gene vectors have a narrow host range, allowing only a limited size of genetic material to be delivered, and they also face the possibility of inducing viral symptoms. Other methods are microinjection, Agrobacterium-mediated transformation, and microprojectile bombardment. All these methods had either very low efficiency (0.01–20%) or were mainly applied for dicotyledons [11]. Another concern behind Agrobacterium-mediated transformation is the usage of antibiotics such as carbenicillin, cefotaxime, rifampicin, and timentin for the removal of bacteria after co-cultivation, which affected the regeneration potential and genetic stability of the regenerated plantlets [15]. An alternative method for genetic engineering like genome editing using CRISPR/Cas9 is comparatively precise and can manipulate the genome, but it also relies on an Agrobacterium-mediated pathway and thus has drawbacks like undesirable off-target effects and insertional mutations in the genome [236]. To circumvent these obstacles, a technique that enables specific horizontal gene transfer is required that allows the delivery of genes into abroad range of plant species without the need of external force to induce desirable traits in commercially important crops [237].
NPs having an extremely small size and easy uptake into plant cells are a potential vehicle for passive gene transfer in different tissues like seeds, leaves, calli, roots, etc. [8,238]. The charge and shape of NP greatly influence the cell membrane translocation, and thus these properties are central to nanocarrier optimization. Another benefit of NP-mediated delivery is that it has high DNA-binding ability and thus has high transformation efficiency without genome integration [239,240]. It has been commonly reported that the internalization is faster and more efficient for cationic NPs as they easily bind with negatively charged cell membranes in comparison to anionic NPs [241]. The conjugation of DNA with NPs and transformation in the cell cytoplasm integrates DNA into the target genome, and develops the transgenic plants with desired traits [242]. Nonetheless, fewer reports are available on the usage of nanomaterials as carriers to deliver biomolecules into the in vitro cultures of crops as compared to research available on morphogenesis. The first study on gene transfer using NP has been documented by Torney et al. [243] where gold-capped mesoporous silica nanoparticles (MSNs) were delivered to N. tabacum. Liu et al. [244] synthesized starch NPs and coated them with poly-L-lysine and fluorescent material Ru(bpy)32+·6H2O. To deliver the DNA, they conjugated NPs with pEGAD plasmid DNA and successfully transformed the suspension culture of Dioscrea zigiberensis G H Wright. Furthermore, poly-L-lysine-coated ZnS nanoparticles with an average size of 3–5 nm efficiently delivered the β-glucuronidase (GUS)-encoding plasmid into young tobacco leaves using the ultrasonic treatment. The efficiency of gene transfection of the treated tobacco plant under various conditions indicated that the highest efficiency is achieved when an ultrasonic treatment with intensity of 60 W for 20 min is applied. These results indicated that the optimum condition for the ultrasonic treatment to achieve the highest gene transfection efficiency depends on the plant type (protoplast, cells, leaves, roots, etc.) as well as nanocarriers and their size [245].
The carbon-based NPs have also been proven as an efficient system; Vijayakumar et al. [246] found that the carbon-supported AuNPs delivered GUS genes more efficiently as compared to the gold particles using a gene gun into N. tabacum, O. sativa and Leucaena leucocephala (Lam.) de Wit. Similarly, positive results have been obtained for fluorescein isothiocyanate (FITC)-tagged SWCNTs and complexes of FITC-tagged DNA molecules with MWCNTs in the suspension cultures of N. tabacum BY-2 cells [172]. In addition, the genetic transformation of N. tabacum protoplasts with a plasmid construct pGreen 0029 having a yfp reporter was carried out using SWCNTs and MWCNTs. It was found that SWCNTs were able to transform both protoplasts and walled plant cells, whereas MWCNTs could only transform the protoplasts because of the presence of a cellulose wall which hindered NP penetration [247]. Later on, FITC has been delivered into B. napus var. Jet Neuf and D. carota var. Konservnaja 63 protoplast using SWCNTs [248] and magnetic AuNPs [249], suggesting successful delivery of the molecules.
An interesting study was carried out to transfer pCambia 1301 having the GUS gene into B. juncea cv. pusa jaikisan, where a better transformation efficiency was achieved with calcium phosphate nanoparticles (CaPNPs, 80.7%), followed by A. tumefaciens (54.4%) and naked DNA (8%) [250]. Similarly, the utilization of CaPNPs to deliver the pBI121-harboring GFP gene into tobacco cells was reported by Ardekani et al. [251]. In corroboration with these, usage of CaPNPs in C. intybus to deliver the HMGR gene showed a positive transformation as higher chlorophyll, proteins, and esculin contents, as well as higher HMGR activity, were detected [252]. For the efficient and stable transformation of Jatropha curcas L. callus [253] and cell suspension [254], a complex of CdSe fluorescent quantum dots (QDs) with L-cysteine and chitosan–DNA (CS-DNA) NP conjugate have been reported. In many reports, MSNs have been documented as an efficient nanocarrier; e.g., Martin-Ortigosa et al. [255] documented the usage of MSNs in different ways like gold plating for MSNs, CaCl2/spermidine DNA coating, and NPs with gold microparticles and gold nanorods to enhance the NP-mediated DNA delivery using the biolistic method in onion, maize, and tobacco. In another study on the same plants, protein-loaded Au-MSNs can be subsequently coated with plasmid DNA and introduced into plant tissues through particle bombardment by which both protein and DNA can be transferred efficiently [256]. For instance, MSNs have been well documented as a carrier to deliver Cre recombinase protein into the Z. mays [257], cryIAb gene in S. lycopersicum var. falat [258], and GUS gene in N. tabacum [259].
Furthermore, layered double hydroxide (LDH) nanosheets have shown positive results in transforming N. tabacum with fluorescent dyes such as tetramethyl rhodamine isothiocyanate (TRITC), FITC, and DNA molecules [260]. New polymeric dimethylaminoethyl methacrylate (DMAEM)-based polymer NPs have been reported to carry the yfp gene in N. tabacum protoplast with the help of polyethylene glycol (PEG) [261]. Similarly, Zhao et al. [262] used magnetic Fe3O4NPs and documented the stable transformation of the BTΔα-CPTI gene in Gossypium hirsutum Linn. pollen, which remained integrated into the genome, which transcribed, expressed and produced an insect-resistant transgenic progeny of cotton plants. Later on, S. tuberosum (cv. lady and spunta)-resistant varieties were generated for pathogenic fungi like Alternaria alternate and Rhizoctonia solani, with the help of two thionin genes delivered using NPs [263]. Gil-Humanes et al. [264] also successfully transformed Wheat dwarf virus (WDV)-derived replicons along with CRISPR/Cas9 to induce targeted mutagenesis in T. aestivum cv. Bobwhite. Recently, green synthesized FeNPs using the leaf extract of Camellia sinensis were utilized for the successful and stable transformation of pBIN.35s-mgfp5-ER carrying GFP gene to the Abelmoschus esculentus [265] and UidA gene with the help of chitosan NPs to the in vitro plants of S. tuberosum [266].

7. Conclusions and Future Perspectives

The world population is expected to reach 9.6 billion by 2050 and to feed this ever-increasing population, there will be increased pressure on land, which is not extendable. A higher usage of fertilizers causes soil damage and environmental pollution. Recent advancements in the field of nanotechnology have demonstrated the potential to revolutionize agricultural production. However, the concern raised behind the in vivo usage of NPs is their release in the environment which might result in the accumulation of different NPs in each trophic level of the ecological pyramid; therefore, we should use NPs judicially.
An alternative strategy to test the NPs is to use them in a tissue culture system which is a powerful tool for the screening of plantlets; it also provides a unique opportunity for studying many aspects of plant growth and development under well-defined and controlled environmental conditions. NPs have been shown to enhance plant germination/production, improve plant resistance to abiotic and biotic stress, assist efficient nutrient exploitation, and promote plant growth, with reduced environmental impact compared to traditional approaches with bulk material. The main application of NPs under in vitro conditions is to increase the crop yield, and ability of plantlets to cope with stressful conditions by which it will adapt to the conditions before transplanting into the field. However, the reports reviewed in the present manuscript suggested that the type and concentration of NPs, as well as the crop species and even cultivar, showed variation in responses. Therefore, the effect of different types and concentration ranges of NPs on plant tissue should be optimized accordingly to determine the optimum dose, which usually ranges between 1–50 mg/L. This optimization will be useful in precision agriculture for individual crops. Further, the usage of NPs as an elicitor has been explored which showed promising results for many commercially valuable metabolites; hence, nano-elicitors may be exploited for the commercial production of secondary metabolites at the bioreactor level.
Although many reports are available on NPs’ effect on plants under in vivo conditions, to gain a clear understanding of the underlying mechanisms behind the role of NPs in plant morphogenesis, these can be further astudied via in vitro routes. The advantage of NPs is that they are promising materials for biomolecule delivery, owing to their ability to traverse plant cells without external force, their tunability for diverse cargo conjugation, and broad host range applicability. These qualities make them a promising tool for the genetic engineering of plants for an easier delivery of genes and without injury to the plant cells. They can also be used for targeted gene delivery to the nucleus, chloroplast, and mitochondria to achieve transgenesis in plants. Reports on molecular mechanisms of elicitation using NPs are scant; hence, systematic omics-based analyses (e.g., genomics, transcriptomics, proteomics, and metabolomics) are necessary.

Author Contributions

Conceptualization, S.K., A.P. and P.D.; Writing—original draft preparation, A.P., N.M. and P.D.; Writing—review and editing, A.P., S.H., S.L.K. and S.K.; Supervision, S.K. All authors have read and agreed to the published version of the manuscript.

Funding

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A.P. and S.H. would like to acknowledge the Ministry of Human Resource Development, Department of Higher Education, Government of India, for providing a post-doctoral fellowship under RUSA 2.0, Thematic Project III, sanctioned to Sumita Kachhwaha (F.35-532/PP/2021/1203). N.M. is thankful to University Grants Commision (UGC), New Delhi, for providing senior research fellowship [F.No.16-6(DEC.2018)/2019(NET/CSIR)].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Duhan, J.S.; Kumar, R.; Kumar, N.; Kaur, P.; Nehra, K.; Duhan, S. Nanotechnology: The new perspective in precision agriculture. Biotechnol. Rep. 2017, 15, 11–23. [Google Scholar] [CrossRef] [PubMed]
  2. Hu, X.; Cook, S.; Wang, P.; Hwang, H.M. In vitro evaluation of cytotoxicity of engineered metal oxide nanoparticles. Sci. Total Environ. 2009, 407, 3070–3072. [Google Scholar] [CrossRef] [PubMed]
  3. Rico, C.M.; Majumdar, S.; Duarte-Gardea, M.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Interaction of nanoparticles with edible plants and their possible implications in the food chain. J. Agric. Food Chem. 2011, 59, 3485–3498. [Google Scholar] [CrossRef]
  4. Alharby, H.F.; Metwali, E.M.R.; Fuller, M.P.; Aldhebiani, A.Y. Impact of application of zinc oxide nanoparticles on callus induction, plant regeneration, element content and antioxidant enzyme activity in tomato (Solanum lycopersicum Mill.) under salt stress. Arch. Biol. Sci. 2016, 68, 723–735. [Google Scholar] [CrossRef]
  5. Zafar, H.; Alim, A.; Zia, M. CuO nanoparticles inhibited root growth from Brassica nigra seedlings but induced root from stem and leaf explants. Appl. Biochem. Biotechnol. 2017, 181, 365–378. [Google Scholar] [CrossRef]
  6. Zia, M.; Gul, S.; Akhtar, J.; Ul Haq, I.; Abbasi, B.H.; Hussain, A.; Naz, S.; Chaudhary, M.F. Green synthesis of silver nanoparticles from grape and tomato juices and evaluation of biological activities. IET Nanobiotechnol. 2017, 11, 193–199. [Google Scholar] [CrossRef]
  7. Aslani, F.; Bagheri, S.; Julkapli, N.M.; Juraimi, A.S.; Hashemi, F.S.G.; Baghdadi, A. Effects of engineered nanomaterials on plants growth: An overview. Sci. World J. 2014, 2014, 641759. [Google Scholar] [CrossRef]
  8. Wang, P.; Lombi, E.; Zhao, F.J.; Kopittke, P.M. Nanotechnology: A new opportunity in plant sciences. Trends Plant Sci. 2016, 21, 699–712. [Google Scholar] [CrossRef]
  9. Sanzari, I.; Leone, A.; Ambrosone, A. Nanotechnology in plant science: To make a long story short. Front. Bioeng. Biotechnol. 2019, 7, 120. [Google Scholar] [CrossRef]
  10. Kralova, K.; Jampilek, J. Responses of medicinal and aromatic plants to engineered nanoparticles. Appl. Sci. 2021, 11, 1813. [Google Scholar] [CrossRef]
  11. Ghormade, V.; Deshpande, M.V.; Paknikar, K.M. Perspectives for nano-biotechnology enabled protection and nutrition of plants. Biotechnol. Adv. 2011, 29, 792–803. [Google Scholar] [CrossRef] [PubMed]
  12. Xin, X.; Judy, J.D.; Sumerlin, B.B.; He, Z. Nano-enabled agriculture: From nanoparticles to smart nanodelivery systems. Environ. Chem. 2020, 17, 413–425. [Google Scholar] [CrossRef]
  13. Mujtaba, M.; Wang, D.; Carvalho, L.B.; Oliveira, J.L.; Pereira, A.D.E.S.; Sharif, R.; Jogaiah, S.; Paidi, M.K.; Wang, L.; Ali, Q.; et al. Nanocarrier-mediated delivery of miRNA, RNAi, and CRISPR-Cas for plant protection: Current trends and future directions. ACS Agric. Sci. Technol. 2021, 1, 417–435. [Google Scholar] [CrossRef]
  14. Khalid, M.F.; Khan, R.I.; Jawaid, M.Z.; Shafqat, W.; Hussain, S.; Ahmed, T.; Rizwan, M.; Ercisli, S.; Pop, O.L.; Marc, R.A. Nanoparticles: The plant saviour under abiotic stresses. Nanomaterials 2022, 12, 3915. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, D.H.; Gopal, J.; Sivanesan, I. Nanomaterials in plant tissue culture: The disclosed and undisclosed. RSC Adv. 2017, 7, 36492–36505. [Google Scholar] [CrossRef]
  16. Nauerby, B.; Billing, K.; Wyndaele, R. Influence of the antibiotic timentin on plant regeneration compared to carbenicillin and cefotaxime in concentrations suitable for elimination of Agrobacterium tumefaciens. Plant Sci. 1997, 123, 169–177. [Google Scholar] [CrossRef]
  17. Wiebke, B.; Ferreira, F.; Pasquali, G.; Bodanese-Zanettini, M.H.; Droste, A. Influence of antibiotics on embryogenic tissue and Agrobacterium tumefaciens suppression in soybean genetic transformation. Bragantia 2006, 65, 543–551. [Google Scholar] [CrossRef]
  18. Abdi, G.; Salehi, H.; Khosh-Khui, M. Nano silver: A novel nanomaterial for removal of bacterial contaminants in valerian (Valeriana officinalis L.) tissue culture. Acta Physiol. Plant. 2008, 30, 709–714. [Google Scholar] [CrossRef]
  19. Safavi, K. Effect of titanium dioxide nanoparticles in plant tissue culture media for enhance resistance to bacterial activity. Bull. Environ. Pharmacol. Life Sci. 2014, 3, 163–166. [Google Scholar]
  20. da Silva, T.G.A.; Duong, T.; Michi, T.; Seiichi, F. The effect of antibiotics on the in vitro growth response of chrysanthemum and tobacco stem transverse thin cell layers (tTCLs). Sci. Hortic. 2003, 97, 397–410. [Google Scholar] [CrossRef]
  21. Safavi, K. Evaluation of using nanomaterial in tissue culture media and biological activity. In Proceedings of the 2nd International Conference on Ecological, Environmental and Biological Sciences (EEBS’2012), Bali, Indonesia, 13–14 October 2012; pp. 5–8. [Google Scholar]
  22. Braydich-Stolle, L.; Hussain, S.; Schlager, J.J.; Hofmann, M.C. In vitro cytotoxicity of nanoparticles in mammalian germline stem cells. Toxicol. Sci. 2005, 88, 412–419. [Google Scholar] [CrossRef] [PubMed]
  23. Shrivastava, S.; Bera, T.; Roy, A.; Singh, G.; Ramachandrarao, P.; Debabrata, D. Characterization of enhanced antibacterial effects of novel silver nanoparticles. Nanotechnology 2007, 18, 225103. [Google Scholar] [CrossRef]
  24. Jeong, S.H.; Hwang, Y.H.; Yi, S.C. Antibacterial properties of padded PP/PE nonwovens incorporating nano-sized silver. J. Mater. Sci. 2005, 40, 5413–5418. [Google Scholar] [CrossRef]
  25. Min, J.S.; Kim, K.S.; Kim, S.W.; Jung, J.H.; Lamsal, K.; Kim, S.B.; Jung, M.; Lee, Y.S. Effects of colloidal silver nanoparticles on sclerotium-forming phytopathogenic fungi. Plant Pathol. J. 2009, 25, 376–380. [Google Scholar] [CrossRef]
  26. Rostami, A.A.; Shahsavar, A. Nano-silver particles eliminate the in vitro contaminations of Olive ‘Mission’ explants. Asian J. Plant Sci. 2009, 8, 505–509. [Google Scholar] [CrossRef]
  27. Safavi, K.; Mortazaeinezahad, F.; Esfahanizadeh, M.; Asgari, M.J. In vitro antibacterial activity of nanomaterial for using in tobacco plants tissue culture. World Acad. Sci. Eng. Technol. 2011, 55, 372–373. [Google Scholar]
  28. Bansod, S.; Bawskar, M.; Rai, M. In vitro effect of biogenic silver nanoparticles on sterilization of tobacco leaf explants and for higher yield of protoplasts. IET Nanobiotechnol. 2015, 9, 239–245. [Google Scholar] [CrossRef]
  29. Fakhrfeshani, M.; Bagheri, A.; Sharifi, A. Disinfecting effects of nano silver fluids in gerbera (Gerbera jamesonii) capitulum tissue culture. J. Biol. Environ. Sci. 2012, 6, 121–127. [Google Scholar]
  30. Arab, M.M.; Yadollahi, M.M.H.; Bagheri, S. Effects of antimicrobial activity of silver nanoparticles on in vitro establishment of G × N15 (hybrid of almond × peach) rootstock. J. Genet. Eng. Biotechnol. 2014, 12, 103–110. [Google Scholar] [CrossRef]
  31. Shokri, S.; Babaei, A.; Ahmadian, M.; Hessami, S.; Arab, M.M. The effects of different concentrations of nano-silver on elimination of bacterial contaminations and phenolic exudation of Rosae (Rosa hybrida L.) in vitro culture. Acta Hortic. 2014, 3, 50–54. [Google Scholar]
  32. Gouran, A.; Jirani, M.; Mozafari, A.A.; Saba, M.K.; Ghaderi, N.; Zaheri, S. Effect of silver nanoparticles on grapevine leaf explants sterilization at in vitro conditions. In Proceedings of the 2nd National Conference on Nanotechnology from Theory to Application, Isfahan, Iran, 20 February 2014; pp. 1–6. [Google Scholar]
  33. Spinoso-Castillo, J.L.; Chavez-Santoscoy, R.A.; Bogdanchikova, N.; Pérez-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]
  34. El-Kosary, S.; Abd-Allatif, A.M.; Stino, R.G.; Hassan, M.M.; Kinawy, A.A. Effect of silver nanoparticles on micropropagation of date palm (Phoenix dactylifera L. cv. sewi and medjool). Plant Arch. 2020, 20, 9701–9706. [Google Scholar]
  35. Abbas, H.K.; Abdulhussein, M.A.A. Improving shoot multiplication of strawberry (Fragaria ananassa L. Cv. Roby Gem) in vitro by using AgNPs and iron nanoparticles. Nat. Volatiles Essent. 2021, 8, 2521–2530. [Google Scholar]
  36. Darwesh, O.M.; Hassan, S.A.M.; Abdallatif, A.M. Enhancing in vitro multiplication of some olive cultivars using silver, selenium and chitosan nanoparticles. Res. Sq. 2021, 8, 995940. [Google Scholar] [CrossRef]
  37. Lee, W.M.; Kwak, J.I.; An, Y.J. Effect of silver nanoparticles in crop plants Phaseolus radiatus and Sorghum bicolor: Media effect on phytotoxicity. Chemosphere 2012, 86, 491–499. [Google Scholar] [CrossRef]
  38. Mahna, N.; Vahed, S.Z.; Khani, S. Plant in vitro culture goes nano: Nanosilver mediated decontamination of ex vitro explants. J. Nanomed. Nanotechnol. 2013, 4, 161. [Google Scholar] [CrossRef]
  39. Cho, M.; Chung, H.; Choi, W.; Yoon, J. Different inactivation behaviors of MS-2 phage and Escherichia coli in TiO2 photocatalytic disinfection. Appl. Eniron. Microbiol. 2005, 71, 270–275. [Google Scholar] [CrossRef]
  40. Shiraishi, K.; Koscki, H.; Tsurumoto, T.; Baba, K.; Naito, M.; Nakayama, K.; Shindo, H. Antimicrobial metal implant with a TiO2-conferred photocatalytic bactericidal effect against Staphylococcus aureus. Surf. Interface Anal. 2008, 41, 17–21. [Google Scholar] [CrossRef]
  41. Murashige, T.; Skoog, F. A revised medium for rapid growth and bioassays with tobacco tissue cultures. Plant Physiol. 1962, 15, 473–479. [Google Scholar] [CrossRef]
  42. Mandeh, M.; Omidi, M.; Rahaie, M. In vitro influences of TiO2 nanoparticles on barley (Hordeum vulgare L.) tissue culture. Biol. Trace Elem. Res. 2012, 150, 376–380. [Google Scholar] [CrossRef]
  43. Helaly, M.N.; El-Metwally, M.A.; El-Hoseiny, H.; Omar, S.A.; El-Sheery, N.I. Effect of nanoparticles on biological contamination of in vitro cultures and organogenic regeneration of banana. Aust. J. Crop. Sci. 2014, 8, 612–624. [Google Scholar]
  44. Álvarez, S.P.; Tapia, M.A.M.; Vega, M.E.G.; Ardisana, E.F.H.; Medina, J.A.C.; Zamora, G.L.F.; Bustamante, D.V. Nanotechnology and plant tissue culture. In Plant Nanobionics, Nanotechnology in the Life Sciences; Prasad, R., Ed.; Springer: Cham, Germany, 2019; pp. 333–370. [Google Scholar]
  45. Waterworth, W.M.; Bray, C.M.; West, C.E. The importance of safeguarding genome integrity in germination and seed longevity. J. Exp. Bot. 2015, 66, 3549–3558. [Google Scholar] [CrossRef] [PubMed]
  46. Barrena, R.; Casals, E.; Colón, J.; Font, X.; Sánchez, A.; Puntes, V. Evaluation of the ecotoxicity of model nanoparticles. Chemosphere 2009, 75, 850–857. [Google Scholar] [CrossRef] [PubMed]
  47. Yasur, J.; Rani, P.U. Environmental effects of nanosilver: Impact on castor seed germination, seedling growth, and plant physiology. Environ. Sci. Pollut. Res. Int. 2013, 20, 8636–8648. [Google Scholar] [CrossRef]
  48. Hatami, M. Stimulatory and inhibitory effects of nanoparticulates on seed germination and seedling vigor indices. In Nanoscience and Plant-Soil Systems; Ghorbanpour, M., Manika, K., Varma, A., Eds.; Springer International Publishing: New York, NY, USA, 2017; pp. 357–385. [Google Scholar]
  49. Zheng, L.; Mingyu, S.; Xiao, W.; Chao, L.; Chunxiang, Q.; Liang, C.; Huang, H.; Xiaoqing, L.; Hong, F. Effect of nano anatase on spectral characteristics and distribution of LHCLL on the thylakoid membranes of spinach. Biol. Trace Elem. Res. 2007, 120, 273–280. [Google Scholar]
  50. Mehrian, S.K.; Heidari, R.; Rahmani, F.; Najafi, S. Effect of chemical synthesis silver nanoparticles on germianation indices and seedling growth in seven varietes of Lycopersicon esculentum Mill (tomato) plants. J. Clust. Sci. 2016, 27, 327–340. [Google Scholar] [CrossRef]
  51. Rizwan, M.; Ali, S.; Qayyum, M.F.; Ok, Y.S.; Adrees, M.; Ibrahim, M.; Zia-ur-Rehman, M.; Farid, M.; Abbas, F. Effect of metal and metal oxide nanoparticles on growth and physiology of globally important food crops: A critical review. J. Hazard. Mater. 2017, 322, 2–16. [Google Scholar] [CrossRef] [PubMed]
  52. Szőllősi, R.; Molnár, Á.; Kondak, S.; Kolbert, Z. Dual effect of nanomaterials on germination and seedling growth: Stimulation vs. phytotoxcity. Plants 2020, 9, 1745. [Google Scholar] [CrossRef]
  53. Mazumdar, H.; Ahmed, G.U. Phytotoxicity effect of silver nanoparticles on Oryza sativa. Int. J. Chem. Tech. Res. 2011, 3, 1494–1500. [Google Scholar]
  54. Tripathi, D.K.; Singh, S.; Singh, S.; Pandey, R.; Singh, V.P.; Sharma, N.C.; Chauhan, D.K. An overview on manufactured nanoparticles in plants: Uptake, translocation, accumulation and phytotoxicity. Plant Physiol. Biochem. 2017, 110, 2–12. [Google Scholar] [CrossRef]
  55. Ali, A.; Phull, A.R.; Zia, M.; Shah, A.M.A.; Malik, R.N.; Haq, I.U. Phytotoxicity of river Chenab sediments: In vitro morphological and biochemical response of Brassica napus L. Environ. Nanotechnol. Monit. Manag. 2015, 4, 74–84. [Google Scholar] [CrossRef]
  56. Zaka, M.; Abbasi, B.H.; Rahman, L.; Shah, A.; Zia, M. Synthesis and characterisation of metal nanoparticles and their effects on seed germination and seedling growth in commercially important Eruca sativa. IET Nanobiotechnol. 2016, 10, 134–140. [Google Scholar] [CrossRef] [PubMed]
  57. Dutta Gupta, S.; Agarwal, A.; Pradhan, S. Phytostimulatory effect of silver nanoparticles (AgNPs) on rice seedling growth: An insight from antioxidative enzyme activities and gene expression patterns. Ecotoxicol. Environ. Saf. 2018, 161, 624–633. [Google Scholar] [CrossRef]
  58. Khan, I.; Raza, M.A.; Awan, S.A.; Khalid, M.H.B.; Raja, N.I.; Min, S.; Zhang, A.; Naeem, M.; Meraj, T.A.; Iqbal, N.; et al. In vitro effect of metallic silver nanoparticles (AgNPs): A novel approach toward the feasible production of biomass and natural antioxidants in pearl millet (Pennisetum glaucum L.). Appl. Ecol. Environ. Res. 2019, 17, 12877–12892. [Google Scholar] [CrossRef]
  59. Tymoszuk, A. Silver nanoparticles effects on in vitro germination, growth, and biochemical activity of tomato, radish, and kale seedlings. Materials 2021, 14, 5340. [Google Scholar] [CrossRef] [PubMed]
  60. Tomaszewska-Sowa, M.; Lisiecki, K.; Pańka, D. Response of rapeseed (Brassica napus L.) to silver and gold nanoparticles as a function of concentration and length of exposure. Agronomy 2022, 12, 2885. [Google Scholar] [CrossRef]
  61. Biba, R.; Matić, D.; Lyons, D.M.; Štefanić, P.P.; Cvjetko, P.; Tkalec, M.; Balen, B. Coating-dependent effects of silver nanoparticles on tobacco seed germination and early growth. Int. J. Mol. Sci. 2020, 21, 3441. [Google Scholar] [CrossRef]
  62. Sharma, P.; Bhatt, D.; Zaidi, M.G.H.; Saradhi, P.P.; Khanna, P.K.; Arora, S. Silver nanoparticle-mediated enhancement in growth and antioxidant status of Brassica juncea. Appl. Biochem. Biotechnol. 2012, 167, 2225–2233. [Google Scholar] [CrossRef]
  63. Timoteo, C.O.; Paiva, R.; dos Reis, M.V.; da Silva, D.P.C.; da Silva, R.B.; de Oliveira, J.E. Silver nanoparticles on dragon fruit in vitro germination and growth. Plant Cell Cult. Micropropag. 2018, 14, 18–25. [Google Scholar]
  64. Mustafa, H.S.; Oraibi, A.G.; Ibrahim, K.M.; Ibrahim, N.K. Influence of silver and copper nanoparticles on physiological characteristics of Phaseolus vulgaris L. in vitro and in vivo. Int. J. Curr. Microbiol. App. Sci. 2017, 6, 834–843. [Google Scholar] [CrossRef]
  65. Zafar, H.; Ali, A.; Ali, J.S.; Haq, I.U.; Zia, M. Effect of ZnO nanoparticles on Brassica nigra seedlings and stem explants: Growth dynamics and antioxidative response. Front. Plant Sci. 2016, 7, 535. [Google Scholar] [CrossRef] [PubMed]
  66. Nair, P.M.G.; Chung, I.M. Changes in the growth, redox status and expression of oxidative stress related genes in chickpea (Cicer arietinum L.) in response to copper oxide nanoparticle exposure. J. Plant Growth Regul. 2015, 34, 350–361. [Google Scholar] [CrossRef]
  67. Lahiani, M.H.; Dervishi, E.; Chen, J.; Nima, Z.; Gaume, A.; Biris, A.S.; Khodakovskaya, M.V. Impact of carbon nanotube exposure to seeds of valuable crops. ACS Appl. Mater. Interface 2013, 5, 7965–7973. [Google Scholar] [CrossRef]
  68. Zaeem, A.; Drouet, S.; Anjum, S.; Khurshid, 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] [PubMed]
  69. 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]
  70. Timoteo, C.O.; Paiva, R.; dos Reis, M.V.; Claro, P.I.C.; Ferraz, L.M.; Marconcini, J.M.; de Oliveira, J.E. In vitro growth of Physalis peruviana L. affected by silver nanoparticles. 3 Biotech 2019, 9, 145. [Google Scholar] [CrossRef] [PubMed]
  71. Nair, P.M.G.; Kim, S.H.; Chung, I.M. Copper oxide nanoparticle toxicity in mung bean (Vigna radiata L.) seedlings: Physiological and molecular level responses of in vitro grown plants. Acta Physiol. Plant. 2014, 36, 2947–2958. [Google Scholar] [CrossRef]
  72. Husen, A.; Siddiqi, K.S. Carbon and fullerene nanomaterials in plant system. J. Nanobiotechnol. 2014, 12, 16–27. [Google Scholar] [CrossRef]
  73. Villagarcia, H.; Dervishi, E.; de Silva, K.; Biris, A.S.; Khodakovskaya, M.V. Surface chemistry of carbon nanotubes impacts the growth and expression of water channel protein in tomato plants. Small 2012, 8, 2328–2334. [Google Scholar] [CrossRef]
  74. Dhingra, P.; Sharma, S.; Singh, K.H.; Kushwaha, H.S.; Barupal, J.K.; Haq, S.; Kothari, S.L.; Kachhwaha, S. Seed priming with carbon nanotubes and silicon dioxide nanoparticles influence agronomic traits of Indian mustard (Brassica juncea) in field experiments. J. King Saud Univ. Sci. 2022, 34, 102067. [Google Scholar] [CrossRef]
  75. Timoteo, C.O.; 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]
  76. Ali, A.; Mohammad, S.; Khan, M.A.; Raja, N.I.; Arif, M.; Kamil, A.; Mashwani, Z.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]
  77. 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]
  78. Jamshidi, M.; Ghanati, F.; Rezaei, A.; Bemani, E. Change of antioxidant enzymes activity of hazel (Corylus avellana L.) cells by AgNPs. Cytotechnology 2016, 68, 525–530. [Google Scholar] [CrossRef]
  79. Jamshidi, M.; Ghanati, F. Taxanes content and cytotoxicity of hazel cells extract after elicitation with silver nanoparticles. Plant Physiol. Biochem. 2017, 110, 178–184. [Google Scholar] [CrossRef]
  80. Chung, I.M.; Rajakumar, G.; Thiruvengadam, M. Effect of silver nanoparticles on phenolic compounds production and biological activities in hairy root cultures of Cucumis anguria. Acta Biol. Hung. 2018, 69, 97–109. [Google Scholar] [CrossRef]
  81. Mozafari, A.; Havas, F.; Ghaderi, N. Application of iron nanoparticles and salicylic acid in in vitro culture of strawberries (Fragaria × ananassa Duch.) to cope with drought stress. Plant Cell Tissue Organ Cult. 2018, 132, 511–523. [Google Scholar] [CrossRef]
  82. Karimzadeh, F.; Haddad, R.; Garoosi, G.H.; 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]
  83. Shaikhaldein, H.O.; Al-Qurainy, F.; Nadeem, M.; Khan, S.; Tarroum, M.; Salih, A.M. Biosynthesis and characterization of silver nanoparticles using Ochradenus arabicus and their physiological effect on Maerua oblongifolia raised in vitro. Sci. Rep. 2020, 10, 17569. [Google Scholar] [CrossRef]
  84. 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 2018, 8, 412. [Google Scholar] [CrossRef] [PubMed]
  85. Do, D.G.; Dang, T.K.T.; Nguyen, T.H.T.; Nguyen, T.D.; Tran, T.T.; Duong, D.H. Effects of nano silver on the growth of banana (Musa spp.) cultured in vitro. J. Viet. Environ. 2018, 10, 92–98. [Google Scholar] [CrossRef]
  86. 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. 3 Biotech 2020, 10, 477–492. [Google Scholar] [CrossRef] [PubMed]
  87. Balážová, Ľ.; Baláž, M.; Babula, P. Zinc oxide nanoparticles damage tobacco BY-2 cells by oxidative stress followed by processes of autophagy and programmed cell death. Nanomaterials 2020, 10, 1066. [Google Scholar] [CrossRef] [PubMed]
  88. Manickavasagam, M.; Pavan, G.; Vasudevan, V. A comprehensive study of the hormetic influence of biosynthesized AgNPs on regenerating rice calli of indica cv. IR64. Sci. Rep. 2019, 9, 8821. [Google Scholar] [CrossRef] [PubMed]
  89. Taha, R.A.; Hassan, M.M.; Ibrahim, E.A.; Baker, N.H.A.; Shaaban, E.A. Carbon nanotubes impact on date palm in vitro cultures. Plant Cell Tissue Organ Cult. 2016, 127, 525–534. [Google Scholar] [CrossRef]
  90. Elsayh, S.A.A. Impact of silver nanoparticles on enhancing in vitro proliferation of embryogenic callus and somatic embryos regeneration of date palm cv. Hayani. Int. J. Environ. Agric. Biotech. 2021, 6, 40–52. [Google Scholar]
  91. Bello-Bello, J.J.; Chavez-Santoscoy, R.A.; Lecona-Guzman, C.A.; Bogdanchikova, N.; Salinas-Ruız, J.; Gomez-Merino, F.C.; Pestryakov, A. Hormetic response by silver nanoparticles on in vitro multiplication of sugarcane (Saccharum spp. cv. Mex 69-290) using a temporary immersion system. Dose Response 2017, 15, 1559325817744945. [Google Scholar] [CrossRef]
  92. Gaafar, A.A.; Taha, R.A.; Abou-Baker, N.H.; Shaaban, E.A.; Salama, Z.A. Evaluation of regeneration, active ingredients and antioxidant activities in jojoba tissue cultures as affected by carbon nanotubes. Biosci. Res. 2018, 15, 2283–2392. [Google Scholar]
  93. Aazami, M.A.; Rasouli, F.; Ebrahimzadeh, A. Oxidative damage, antioxidant mechanism and gene expression in tomato responding to salinity stress under in vitro conditions and application of iron and zinc oxide nanoparticles on callus induction and plant regeneration. BMC Plant Biol. 2021, 21, 597. [Google Scholar] [CrossRef]
  94. Gowayed, S.M.H.; Al-Zahrani, H.S.M.; Metwali, E.M.R. Improving the salinity tolerance in potato (Solanum tuberosum) by exogenous application of silicon dioxide nanoparticles. Int. J. Agric. Biol. 2017, 19, 183–194. [Google Scholar]
  95. Homaee, M.B.; Ehsanpour, A.A. Physiological and biochemical responses of potato (Solanum tuberosum) to silver nanoparticles and silver nitrate treatments under in vitro conditions. Ind. J. Plant Physiol. 2015, 20, 353–359. [Google Scholar] [CrossRef]
  96. Koontz, H.V.; Berle, K.L. Silver uptake, distribution, and effect on calcium, phosphorus, and sulfur uptake. Plant Physiol. 1980, 65, 336–339. [Google Scholar] [CrossRef] [PubMed]
  97. Bell, P.F.; Chaney, R.L.; Angle, J.S. Free metal activity and total metal concentrations as indices of micro nutrient availability to barley [Hordeum vulgare L. Klages]. Plant Soil. 1991, 130, 51–62. [Google Scholar] [CrossRef]
  98. Zuverza-Mena, N.; Martínez-Fernández, D.; Du, W.; Hernandez-Viezcas, J.A.; Bonilla-Bird, N.; López-Moreno, M.L.; Komárek, M.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Exposure of engineered nanomaterials to plants: Insights into the physiological and biochemical responses-A review. Plant Physiol. Biochem. 2017, 110, 236–264. [Google Scholar] [CrossRef]
  99. Kokina, I.; Gerbreders, V.; Sledevskis, E.; Bulanovs, A. Penetration of nanoparticles in flax (Linum usitatissimum L.) calli and regenerants. J. Biotechnol. 2013, 165, 127–132. [Google Scholar] [CrossRef]
  100. Iqbal, N.; Khan, N.A.; Ferrante, A.; Trivellini, A.; Francini, A.; Khan, M.I.R. Ethylene role in plant growth, development and senescence: Interaction with other phytohormones. Front. Plant Sci. 2017, 8, 475. [Google Scholar] [CrossRef]
  101. Mahmoud, L.M.; Grosser, J.W.; Dutt, M. Silver compounds regulate leaf drop and improve in vitro regeneration from mature tissues of Australian finger lime (Citrus australasica). Plant Cell Tissue Organ Cult. 2020, 141, 455–464. [Google Scholar] [CrossRef]
  102. Sarmast, M.K.; Salehi, H. Silver nanoparticles: An influential element in plant nanobiotechnology. Mol. Biotechnol. 2016, 58, 441–449. [Google Scholar] [CrossRef]
  103. Huong, B.T.T.; Xuan, T.D.; Trung, K.H.; Ha, T.T.T.; Duong, V.X.; Khanh, T.D.; Gioi, D.H. Influences of silver nanoparticles in vitro morphogenesis of specialty king banana (Musa ssp.) in Vietnam. Plant Cell Biotechnol. Mol. Biol. 2021, 22, 163–175. [Google Scholar]
  104. Hegazi, E.S.S.; Yousef, A.; Abd-Allatif, A.M.A.; Mahmoud, T.S.M.; Mostafa, M.K.M. Effect of silver nanoparticles, medium composition and growth regulators on in vitro propagation of picual olive cultivar. Egypt. J. Chem. 2021, 64, 6961–6969. [Google Scholar] [CrossRef]
  105. Terry, N.; Zayed, A.M.; De Souza, M.P.; Tarun, A.S. Selenium in higher plants. Ann. Rev. Plant Physiol. Plant Mol. Biol. 2000, 51, 401–432. [Google Scholar] [CrossRef]
  106. Hasanuzzaman, M.; Hossain, M.A.; Fujita, M. Selenium-induced upregulation of the antioxidant defense and methylglyoxal detoxification system reduces salinity induced damage in rapeseed seedlings. Biol. Trace Elem. Res. 2011, 43, 1704–1721. [Google Scholar] [CrossRef] [PubMed]
  107. Tan, X.M.; Fugetsu, B. Multi-walled carbon nanotubes interact with cultured rice cells: Evidence of a self-defense response. J. Biomed. Nanotechnol. 2007, 3, 285–288. [Google Scholar] [CrossRef]
  108. Chutipaijit, S.; Sutjaritvorakul, T. Application of activated charcoal and nanocarbon to callus induction and plant regeneration in aromatic rice (Oryza sativa L.). Chem. Soc. Rev. 2018, 30, 1–8. [Google Scholar] [CrossRef]
  109. Venkatachalam, P.; Malar, S.; Thiyagarajan, M.; Indiraarulselvi, P.; Geetha, N. Effect of phycochemical coated silver nanocomplexes as novel growth-stimulating compounds for plant regeneration of Alternanthera sessilis L. J. Appl. Phycol. 2017, 29, 1095–1106. [Google Scholar] [CrossRef]
  110. Kouhi, S.M.; Lahouti, M. Application of ZnO nanoparticles for inducing callus in tissue culture of rapeseed. Int. J. Nanosci. Nanotechnol. 2018, 14, 133–141. [Google Scholar]
  111. Irum, S.; Jabeen, N.; Ahmad, K.S.; Shafique, S.; Khan, T.F.; Gul, H.; Anwaar, S.; Shah, N.I.; Mehmood, A.; Hussain, S.Z. Biogenic iron oxide nanoparticles enhance callogenesis and regeneration pattern of recalcitrant Cicer arietinum L. PLoS ONE 2020, 15, e0242829. [Google Scholar] [CrossRef]
  112. Giorgetti, L.; Ruffini Castiglione, M.; Bernabini, M.; Geri, C. Nanoparticles effects on growth and differentiation in cell culture of carrot (Daucus carota L.). Agrochimica 2010, 6, 1–9. [Google Scholar]
  113. Tung, H.T.; Thuong, T.T.; Cuong, D.M.; Luan, V.Q.; Hien, V.T.; Hieu, T.; Nhut, D.T. Silver nanoparticles improved explant disinfection, in vitro growth, runner formation and limited ethylene accumulation during micropropagation of strawberry (Fragaria × ananassa). Plant Cell Tissue Organ Cult. 2021, 145, 393–403. [Google Scholar] [CrossRef]
  114. Kokina, I.; Sïedevskis, Ç.; Gerbreders, V.; Grauda, D.; Jermaïonoka, M.; Valaine, K.; Gavarâne, I.; Pigiòka, I.; Filipoviès, M.; Rashal, I. Reaction of flax (Linum usitatissimum L.) calli culture to supplement of medium by carbon nanoparticles. Proc. Latv. Acad. Sci. Sect. B 2012, 66, 200–209. [Google Scholar]
  115. Kokina, I.; Mickeviča, I.; Jahundoviča, I.; Ogurcovs, A.; Krasovska, M.; Jermaļonoka, M.; Mihailova, I.; Tamanis, E.; Gerbreders, V. Plant explants grown on medium supplemented with Fe3O4 nanoparticles have a significant increase in embryogenesis. J. Nanomater. 2017, 2017, 4587147. [Google Scholar] [CrossRef]
  116. 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]
  117. Chetia, I.; Chaliha, A.K.; Gogoi, M.B.; Verma, G. Establishment of efficient in-vitroregeneration protocol in ‘Malbhog’ Banana (Musa paradisiaca) using MWCNTs and plant growth regulators. Int. J. Curr. Microbiol. App. Sci. 2020, 9, 2930–2937. [Google Scholar] [CrossRef]
  118. Chutipaijit, S. Establishment of condition and nanoparticle factors influencing plant regeneration from aromatic rice (Oryza sativa). Int. J. Agric. Biol. 2015, 17, 1049–1054. [Google Scholar]
  119. Anwaar, S.; Maqbool, Q.; Jabeen, N.; Nazar, M.; Abbas, F.; Nawaz, B.; Hussain, T.; Hussain, S.Z. The effect of green synthesized CuO nanoparticles on callogenesis and regeneration of Oryza sativa L. Front. Plant Sci. 2016, 7, 1330. [Google Scholar] [CrossRef]
  120. Chutipaijit, S.; Sutjaritvorakul, T. Application of nanomaterials in plant regeneration of rice (Oryza sativa L.). Mater. Today Proc. 2017, 4, 6140–6145. [Google Scholar] [CrossRef]
  121. Chutipaijit, S.; Sutjaritvorakul, T. Titanium dioxide (TiO2) nanoparticles induced callus induction and plant regeneration of indica rice cultivars (suphanburi1 and suphanburi90). Dig. J. Nanomater. Biostruct. 2018, 13, 1003–1010. [Google Scholar]
  122. Shafique, S.; Jabeen, N.; Ahmad, K.S.; Irum, S.; Anwaar, S.; Ahmad, N.; Alam, S.; Ilyas, M.; Khan, T.F.; Hussain, S.Z. Green fabricated zinc oxide nanoformulated media enhanced callus induction and regeneration dynamics of Panicum virgatum L. PLoS ONE 2020, 15, e0230464. [Google Scholar] [CrossRef]
  123. Li, X.; Zhang, F.; He, N.; Zhang, B.; Liu, X.; Tan, X. In vitro influence of Fe3O4 nanoparticles on Rhizoma polygonati micropropagation. J. Nanosci. Nanotechnol. 2017, 17, 2047–2053. [Google Scholar] [CrossRef]
  124. Flores, D.; Chacón, R.; Alvarado, L.; Schmidt, A.; Alvarado, C.; Chaves, J. Effect of using two different types of carbon nanotubes for blackberry (Rubus adenotrichos) in vitro plant rooting, growth and histology. Am. J. Plant Sci. 2014, 5, 3510–3518. [Google Scholar] [CrossRef]
  125. Elsahhar, S.; Shahba, M.; Elsayed, T.; Mikhail, M.; Galal, A. Effect of chitosan nanoparticles (CS-NPs) on in vitro regeneration response and production of potato virus Y (PVY)-free plants of potato. Agronomy 2022, 12, 2901. [Google Scholar] [CrossRef]
  126. Nalci, O.B.; Nadaroglu, H.; Pour, A.H.; Gungor, A.A.; Haliloglu, K. Effects of ZnO, CuO and γ-Fe3O4 nanoparticles on mature embryo culture of wheat (Triticum aestivum L.). Plant Cell Tissue Organ Cult. 2019, 136, 269–277. [Google Scholar] [CrossRef]
  127. Malik, W.A.; Mahmood, I.; Razzaq, A.; Afzal, M.; Shah, G.A.; Iqbal, A.; Zain, M.; Ditta, A.; Asad, S.A.; Ahmad, I.; et al. Exploring potential of copper and silver nano particles to establish efficient callogenesis and regeneration system for wheat (Triticum aestivum L.). GM Crop. Food 2021, 12, 564–585. [Google Scholar] [CrossRef] [PubMed]
  128. Koçak, R.; Okcu, M.; Haliloglu, K.; Alayli, A.; Nadaroglu, H. Determination of the effect of magnesium oxide nanoparticles (MgO-NP) on in vitro culture of cowpea (Vigna unguiculata L. Walp). Res. Sq. 2021, 13, 1088469. [Google Scholar] [CrossRef]
  129. Iravani, S. Green synthesis of metal nanoparticles using plants. Green Chem. 2011, 13, 2638–2650. [Google Scholar] [CrossRef]
  130. Pathak, A.; Joshi, A. Somaclonal variation. In Plant Biotechnology: Principles, Techniques and Applications; Prasad, B.D., Sahni, S., Kumar, P., Siddiqui, W.M., Eds.; Apple Academic Press (CRC Press): New York, NY, USA, 2018; Volume 1, pp. 185–213. [Google Scholar]
  131. Jeong, B.R.; Sivanesan, I. Direct adventitious shoot regeneration, in vitro flowering, fruiting, secondary metabolite content and antioxidant activity of Scrophularia takesimensis Nakai. Plant Cell Tissue Organ Cult. 2015, 123, 607–618. [Google Scholar] [CrossRef]
  132. Kokina, I.; Mickevica, I.; Jermalonoka, M.; Bankovska, L.; Gerbreders, V.; Ogurcovs, A.; Jahundovica, I. Case study of somaclonal variation in resistance genes Mlo and Pme3 in flaxseed (Linum usitatissimum L.) induced by nanoparticles. Int. J. Genomics 2017, 2017, 1676874. [Google Scholar] [CrossRef]
  133. Hatami, M.; Naghdi, B.H.; Ghorbanpour, M. Nano-elicitation of secondary pharmaceutical metabolites in plant cells: A review. J. Med. Plant. 2019, 18, 6–36. [Google Scholar] [CrossRef]
  134. Raei, M.; Angaji, A.A.; Omidi, M.; Khodayari, M. Effect of abiotic elicitors on tissue culture of Aloe vera. Int. J. Biosci. 2014, 5, 74–81. [Google Scholar]
  135. Iqbal, Z.; Javad, S.; Naz, S.; Shah, A.A.; Shah, A.N.; Paray, B.A.; Gulnaz, A.; Abdelsalam, N.R. Elicitation of the in vitro cultures of selected varieties of Vigna radiata L. with zinc oxide and copper oxide nanoparticles for enhanced phytochemicals production. Front. Plant Sci. 2022, 13, 908532. [Google Scholar] [CrossRef]
  136. Al-Oubaidi, H.K.M.; Mohammed-Ameen, A.S. The effect of (AgNO3) NPs on increasing of secondary metabolites of Calendula officinalis L. in vitro. Int. J. Pharm. Pract. 2014, 5, 267–272. [Google Scholar]
  137. 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]
  138. Wesołowska, 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]
  139. Al-Taie, A.A.G.; Aboohanah, M.A.; Issa, F.H. Effect of silver nanoparticles in stimulating some active compounds in garlic callus under salt stress, in vitro. IOP Conf. Ser. Earth Environ. Sci. 2021, 923, 012023. [Google Scholar] [CrossRef]
  140. 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]
  141. Al-Oubaidi, H.K.M.; Al-Khafagi, M.F.J. In vitroincreasing medical compounds (tannins and phenols) of Punica granatum L. in callus using MgO NPs and CuO NPs. J. Pharm. Sci. Res. 2018, 10, 1085–1088. [Google Scholar]
  142. Al-Khafagi, M.F.J.; Al-Oubaidi, H.K.M. Effect of MgO and CuO nanoparticles on increasing tannin and phenol compounds of Punica granatum L. using shoot tip in vitro. Int. J. Drug Deliv. Technol. 2020, 10, 499–504. [Google Scholar] [CrossRef]
  143. Hayl, L.A.A.-A.; Al-Oubaidi, H.K.M. Effect of SiO2 NPs on increase of active compounds in leave callus of Tagetes erecta L. (Marigold) plant in vitro. J. Pharm. Negat. 2022, 13, 86–92. [Google Scholar]
  144. Bhat, P.; Bhat, A. Silver nanoparticles for the enhancement of accumulation of capsaicin in suspension culture of Capsicum sp. J. Exp. Sci. 2016, 7, 1–6. [Google Scholar]
  145. Al-Sowaidi, W.M.M.; Al-Oubaidi, H.K.M. Increasing (glycosides compounds) of Olea europaea L. from shoot tips using AgNO3 nano particle in vitro. Int. J. Phytopharmacol. 2015, 6, 31–35. [Google Scholar]
  146. Anjum, S.; Anjum, I.; Hano, C.; Kousar, S. Advances in nanomaterials as novel elicitors of pharmacologically active plant spcialized metabolites: Current status and future outlooks. RSC Adv. 2019, 9, 40404–40423. [Google Scholar] [CrossRef] [PubMed]
  147. Bolwell, G.P.; Wojtaszek, P. Mechanisms for the generation of reactive oxygen species in plant defence–A broad perspective. Physiol. Mol. Plant Pathol. 1997, 51, 347–366. [Google Scholar] [CrossRef]
  148. Khan, A.K.; Kousar, S.; Tungmunnithum, D.; Hano, C.; Abbasi, B.H.; Anjum, S. Nano-elicitation as an effective and emerging strategy for in vitro production of industrially important flavonoids. Appl. Sci. 2021, 11, 1694. [Google Scholar] [CrossRef]
  149. Navarro, E.; Baun, A.; Behra, R.; Hartmann, N.B.; Filser, J.; Miao, A.J.; Quigg, A.; Santschi, P.H.; Sigg, L. Environmental behavior and ecotoxicity of engineered nanoparticles to algae, plants, and fungi. Ecotoxicology 2008, 17, 372–386. [Google Scholar] [CrossRef] [PubMed]
  150. Khan, M.N.; Mobin, M.; Abbas, Z.K.; AlMutairi, K.A.; Siddiqui, Z.H. Role of nanomaterials in plants under challenging environments. Plant Physiol. Biochem. 2017, 110, 194–209. [Google Scholar] [CrossRef] [PubMed]
  151. Berni, R.; Luyckx, M.; Xu, X.; Legay, S.; Sergeant, K.; Hausman, J.F.; Lutts, S.; Cai, G.; Guerriero, G. Reactive oxygen species and heavy metal stress in plants: Impact on the cell wall and secondary metabolism. Environ. Exp. Bot. 2018, 161, 98–106. [Google Scholar] [CrossRef]
  152. Phukan, U.J.; Jeena, G.S.; Shukla, R.K. WRKY transcription factors: Molecular regulation and stress responses in plants. Front. Plant Sci. 2016, 7, 760. [Google Scholar] [CrossRef]
  153. Kohan-Baghkheirati, E.; Geisler-Lee, J. Gene expression, protein function and pathways of Arabidopsis thaliana responding to silver nanoparticles in comparison to silver ions, cold, salt, drought, and heat. Nanomaterials 2015, 5, 436–467. [Google Scholar] [CrossRef]
  154. Marslin, G.; Sheeba, C.J.; Franklin, G. Nanoparticles alter secondary metabolism in plants via ROS burst. Front. Plant Sci. 2017, 8, 832. [Google Scholar] [CrossRef]
  155. Rivero-Montejo, S.J.; Hernandez, M.V.; Pacheco, I.T. Nanoparticles as novel elicitors to improve bioactive compounds in plants. Agriculture 2021, 11, 134. [Google Scholar] [CrossRef]
  156. Carpita, N.; Sabularse, D.; Montezinos, D.; Delmer, D.P. Determination of the pore size of cell walls of living plant cells. Science 1979, 205, 1144–1147. [Google Scholar] [CrossRef]
  157. Maine, M.A.; Duarte, M.V.; Suñé, N.L. Cadmium uptake by floating macrophytes. Water Res. 2001, 35, 2629–2634. [Google Scholar] [CrossRef] [PubMed]
  158. Kurepa, J.; Paunesku, T.; Vogt, S.; Arora, H.; Rabatic, B.M.; Lu, J.; Wanzer, M.B.; Woloschak, G.E.; Smalle, J.A. Uptake and distribution of ultrasmall anatase TiO2 alizarin red S nanoconjugates in Arabidopsis thaliana. Nano Lett. 2010, 10, 2296–2302. [Google Scholar] [CrossRef] [PubMed]
  159. Zhang, L.; Feng, C.; Chen, Z.; Liu, L.; Jiang, K.; Li, Q.; Fan, S. Super aligned carbon nanotube grid for high resolution transmission electron microscopy of nanomaterials. Nano Lett. 2008, 8, 2564–2569. [Google Scholar] [CrossRef]
  160. Etxeberria, E.; Gonzalez, P.; Pozueta, J. Evidence for two endocytic transport pathways in plant cells. Plant Sci. 2009, 177, 341–348. [Google Scholar] [CrossRef]
  161. Pola, M.; Tamara, L.C.; Andrew, T.H. Toxicity, uptake, and translocation of engineered nanomaterials in vascular plants. Environ. Sci. Technol. Lett. 2012, 46, 9224–9239. [Google Scholar]
  162. Birbaum, K.; Brogioli, R.; Schellenberg, M.; Martinoia, E.; Stark, W.J.; Günther, D.; Limbach, L.K. No evidence for cerium dioxide nanoparticle translocation in maize plants. Environ. Sci. Technol. 2010, 44, 8718–8723. [Google Scholar] [CrossRef]
  163. Hauck, T.S.; Ghazani, A.A.; Chan, W.C.W. Assessing the effect of surface chemistry on gold nanorod uptake, toxicity, and gene expression in mammalian cells. Small 2008, 4, 153–159. [Google Scholar] [CrossRef]
  164. Deng, Y.Q.; White, J.C.; Xing, B.S. Interactions between engineered nanomaterials and agricultural crops: Implications for food safety. J. Zhejiang Univ. Sci. A 2014, 15, 552–572. [Google Scholar] [CrossRef]
  165. Nair, R.; Varghese, H.S.; Nair, B.G.; Maekawa, T.; Yoshida, Y.; Kumar, D.S. Nanoparticulate material delivery to plants. Plant Sci. 2010, 179, 154–163. [Google Scholar] [CrossRef]
  166. Lin, S.; Reppert, J.; Hu, Q.; Hudson, J.S.; Reid, M.L.; Ratnikova, T.A.; Rao, A.M.; Luo, H.; Ke, P.C. Uptake, translocation, and transmission of carbon nanomaterials in rice plants. Small 2009, 5, 1128–1132. [Google Scholar] [CrossRef]
  167. Lee, C.W.; Mahendra, S.; Zodrow, K.; Li, D.; Tsai, Y.C.; Braam, J.; Alvarez, P.J.J. Developmental phytotoxicity of metal oxide nanoparticles to Arabidopsis thaliana. Environ. Toxicol. Chem. 2010, 29, 669–675. [Google Scholar] [CrossRef]
  168. Onelli, E.; Prescianotto-Baschong, C.; Caccianiga, M.; Moscatelli, A. Clathrin dependent and independent endocytic pathways in tobacco protoplasts revealed by labelling with charged nanogold. J. Exp. Bot. 2008, 59, 3051–3068. [Google Scholar] [CrossRef]
  169. Serag, M.F.; Kaji, N.; Gaillard, C.; Okamoto, Y.; Terasaka, K.; Jabasini, M.; Tokeshi, M.; Mizukami, H.; Bianco, A.; Baba, Y. Trafficking and subcellular localization of multiwalled carbon nanotubes in plant cells. ACS Nano 2010, 5, 493–499. [Google Scholar] [CrossRef]
  170. Whiteside, M.D.; Treseder, K.K.; Atsatt, P.R. The brighter side of soils: Quantum dots track organic nitrogen through fungi and plants. Ecology 2009, 90, 100–108. [Google Scholar] [CrossRef] [PubMed]
  171. Chen, R.; Ratnikova, T.A.; Stone, M.B.; Lin, S.; Lard, M.; Huang, G.; Hudson, J.S.; Ke, P.C. Differential uptake of carbo nanoparticles by plant and mammalian cells. Small 2010, 6, 612–617. [Google Scholar] [CrossRef] [PubMed]
  172. Liu, Q.; Chen, B.; Wang, Q.; Shi, X.; Xiao, Z.; Lin, J.; Fang, X. Carbon nanotubes as molecular transporters for walled plant cells. Nano Lett. 2009, 9, 1007–1010. [Google Scholar] [CrossRef] [PubMed]
  173. Ma, X.; Geiser-Lee, J.; Deng, Y.; Kolmakov, A. Interactions between engineered nanoparticles (ENPs) and plants: Phytotoxicity, uptake and accumulation. Sci. Total Environ. 2010, 408, 3053–3061. [Google Scholar] [CrossRef]
  174. Ma, Y.; Kuang, L.; He, X.; Bai, W.; Ding, Y.; Zhang, Z.; Zhao, Y.; Chai, Z. Effects of rare earth oxide nanoparticles on root elongation of plants. Chemosphere 2010, 78, 273–279. [Google Scholar] [CrossRef]
  175. Liu, R.; Lal, R. Potentials of engineered nanoparticles as fertilizers for increasing agronomic productions. Sci. Total Environ. 2015, 514, 131–139. [Google Scholar] [CrossRef]
  176. Nair, P.M.G.; Chung, I.M. Impact of copper oxide nanoparticles exposure on Arabidopsis thaliana growth, root system development, root lignificaion, and molecular level changes. Environ. Sci. Pollut. Res. 2014, 21, 12709–12722. [Google Scholar] [CrossRef]
  177. Hossain, Z.; Mustafa, G.; Komatsu, S. Plant responses to nanoparticle stress. Int. J. Mol. Sci. 2015, 16, 26644–26653. [Google Scholar] [CrossRef]
  178. Kole, C.; Kole, P.; Randunu, K.M.; Choudhary, P.; Podila, R.; Ke, P.C. 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, 9, 37. [Google Scholar] [CrossRef] [PubMed]
  179. Razzaq, A.; Ammara, R.; Jhanzab, H.M.; Mahmood, T.; Hafeez, A.; Hussain, S. A novel nanomaterial to enhance growth and yield of wheat. J. Nanosci. Technol. 2016, 2, 55–58. [Google Scholar]
  180. Choi, O.; Deng, K.K.; Kim, N.J.; Ross, L.; Surampalli, R.Y.; Hu, Z. The inhibitory effects of silver nanoparticles, silver ions, and silver chloride colloids on microbial growth. Water Res. 2008, 42, 3066–3074. [Google Scholar] [CrossRef] [PubMed]
  181. Nguyen, N.H.A.; Padil, V.V.T.; Slaveykova, V.I.; Černık, M.; Sevců, A. Green synthesis of metal and metal oxide nanoparticles and their effect on the unicellular alga Chlamydomonas reinhardtii. Nanoscale Res. Lett. 2018, 13, 159. [Google Scholar] [CrossRef] [PubMed]
  182. Stampoulis, D.; Sinha, S.K.; White, J.C. Assay-dependent phytotoxicity of nanoparticles to plants. Environ. Sci. Technol. 2009, 43, 9473–9479. [Google Scholar] [CrossRef]
  183. Mirzajani, F.; Askari, H.; Hamzelou, S.; Farzaneh, M.; Ghassempour, A. Effect of silver nanoparticles on Oryza sativa L. and its rhizosphere bacteria. Ecotoxicol. Environ. Saf. 2013, 88, 48–54. [Google Scholar] [CrossRef] [PubMed]
  184. Dat, J.; Vandenabeele, S.; Vranová, E.; Van Montagu, M.; Inzé, D.; Van Breusegem, F. Dual action of the active oxygen species during plant stress responses. Cell. Mol. Life Sci. 2000, 57, 779–795. [Google Scholar] [CrossRef]
  185. Bethke, P.C.; Jones, R.L. Cell death of barley aleurone protoplasts is mediated by reactive oxygen species. Plant J. 2001, 25, 19–29. [Google Scholar]
  186. Faisal, M.; Saquib, Q.; Alatar, A.A.; Al-Khedhairy, A.A.; Hegazy, A.K.; Musarrat, J. Phytotoxic hazards of NiO-nanoparticles in tomato: A study on mechanism of cell death. J. Hazard. Mater. 2013, 250–251, 318–332. [Google Scholar] [CrossRef] [PubMed]
  187. Niu, L.; Liao, W. Hydrogen peroxide signaling in plant development and abiotic responses: Crosstalk with nitric oxide and calcium. Front. Plant Sci. 2016, 7, 230. [Google Scholar] [CrossRef] [PubMed]
  188. Smirnoff, N.; Arnaud, D. Hydrogen peroxide metabolism and functions in plants. New Phytol. 2019, 221, 1197–1214. [Google Scholar] [CrossRef] [PubMed]
  189. Sharma, P.; Jha, A.B.; Dubey, R.S.; Pessarakli, M. Reactive oxygen species, oxidative damage, and antioxidative defense mechanism in plants under stressful conditions. J. Bot. 2012, 2012, 217037. [Google Scholar] [CrossRef]
  190. Sinha, A.K.; Jaggi, M.; Raghuram, B.; Tuteja, N. Mitogen-activated protein kinase signaling in plants under abiotic stress. Plant Signal. Behav. 2011, 6, 196–203. [Google Scholar] [CrossRef]
  191. Gechev, T.S.; Hille, J. Hydrogen peroxide as a signal controlling plant programmed cell death. J. Cell. Biol. 2005, 168, 17–20. [Google Scholar] [CrossRef]
  192. Joshi, P.K.; Saxena, S.C.; Arora, S. Characterization of Brassica juncea antioxidant potential under salinity stress. Acta Physiol. Plant. 2011, 33, 811–822. [Google Scholar] [CrossRef]
  193. Hayat, S.; Hayat, Q.; Alyemeni, M.N.; Wani, A.S.; Pichtel, J.; Ahmad, A. Role of proline under changing environments. Plant Signal. Behav. 2012, 7, 1456–1466. [Google Scholar] [CrossRef]
  194. Dixon, R.A.; Paiva, N.L. Stress-induced phenylpropanoid metabolism. Plant Cell 1995, 7, 1085–1097. [Google Scholar] [CrossRef]
  195. Felemban, A.; Braguy, J.; Zurbriggen, M.D.; Al-babili, S. Apocarotenoids involved in plant development and stress response. Front. Plant Sci. 2019, 10, 1168. [Google Scholar] [CrossRef]
  196. Thiruvengadam, M.; Gurunathan, S.; Chung, I.M. Physiological, metabolic, and transcriptional effects of biologically-synthesized silver nanopartciles in turnip (Brassica rapa ssp. rapa L.). Protoplasma 2015, 252, 1031–1046. [Google Scholar] [CrossRef] [PubMed]
  197. Carter, G.A.; Jones, J.H.; Mitchell, R.J.; Brewer, C.H. Detection of solar-excited chlorophyll a fluorescence and leaf photosynthetic capacity using a fraunhofer line radiometer. Remote Sens. Environ. 1996, 55, 89–92. [Google Scholar] [CrossRef]
  198. Bartley, G.E. Plant carotenoids: Pigments for photoprotection, visual attraction, and human health. Plant Cell 1995, 7, 1027–1038. [Google Scholar]
  199. Sillanpää, M.; Ncibi, C. Biofuels and bioenergy. In A Sustainable Bioeconomy; Sillanpää, M., Ncibi, C., Eds.; Springer: Cham, Germany, 2017; pp. 79–139. [Google Scholar]
  200. Yang, J.; Cao, W.; Rui, Y. Interaction between nanoparticles and plants: Phytotoxicity and defense mechanisms. J. Plant Interact. 2017, 12, 158–169. [Google Scholar] [CrossRef]
  201. Husen, A.; Siddiqi, K.S. Phytosynthesis of nanoparticles: Concept, controversy and application. Nanoscale Res. Lett. 2014, 9, 1–24. [Google Scholar] [CrossRef]
  202. Parida, A.K.; Das, A.B. Salt tolerance and salinity effects on plants: A review. Ecotoxicol. Environ. Saf. 2005, 60, 324–349. [Google Scholar] [CrossRef] [PubMed]
  203. Sperotto, R.A.; Boff, T.; Duarte, G.L.; Fett, J.P. Increased senescence-associated gene expression and lipid peroxidation induced by iron deficiency in rice roots. Plant Cell Rep. 2008, 27, 183–195. [Google Scholar] [CrossRef]
  204. Rawat, M.; Nayan, R.; Negi, B.; Zaidi, M.; Arora, S. Physio-biochemical basis of iron-sulfide nanoparticle induced growth and seed yield enhancement in B. juncea. Plant Physiol. Biochem. 2017, 118, 274–284. [Google Scholar] [CrossRef]
  205. Hezaveh, T.A.; Pourakbar, L.; Rahmani, F.; Alipour, H. Interactive effects of salinity and ZnO nanoparticles on physiological and molecular parameters of rapeseed (Brassica napus L.). Commun. Soil Sci. Plant Anal. 2019, 50, 698–715. [Google Scholar] [CrossRef]
  206. Winkel, B.S.J. When an enzyme isn’t just an enzyme anymore. J. Exp. Bot. 2017, 68, 1387–1389. [Google Scholar] [CrossRef]
  207. Van Assche, F.; Clijsters, H. Effects of metalson enzyme activity in plants. Plant Cell Environ. 1990, 13, 195–206. [Google Scholar] [CrossRef]
  208. Meharg, A.A. Integrated tolerance mechanisms: Constitutive and adaptive plant responses to elevated metal concentrations in the environment. Plant Cell Environ. 1994, 17, 989–993. [Google Scholar] [CrossRef]
  209. Bandara, J.; Guasaquillo, I.; Bowen, P.; Soare, L.; Jardim, W.F.; Kiwi, J. Photocatalytic storing of O2 as H2O2 mediated by high surface area CuO evidence for a reductive–oxidative interfacial mechanism. Langmuir 2005, 21, 8554–8559. [Google Scholar] [CrossRef] [PubMed]
  210. Guo, B.; Zebda, R.; Drake, S.J.; Sayes, C.M. Synergistic effect of co-exposure to carbon black and Fe2O3 nanoparticles on oxidative stress in cultured lung epithelial cells. Part. Fiber Toxicol. 2009, 6, 4. [Google Scholar] [CrossRef] [PubMed]
  211. Cheloni, G.; Marti, E.; Slaveykova, V.I. Interactive effects of copper oxide nanoparticles and light to green alga Chlamydomonas reinhardtii. Aquat. Toxicol. 2016, 170, 120–128. [Google Scholar] [CrossRef]
  212. Yruela, I. Copper in plants. Braz. J. Plant Physiol. 2005, 17, 145–156. [Google Scholar] [CrossRef]
  213. Pais, I. The biological importance of titanium. J. Plant Nutr. 1983, 6, 3–131. [Google Scholar] [CrossRef]
  214. Zheng, L.; Hong, F.; Lu, S.; Liu, C. Effect of nano-TiO2 on strength of naturally aged seeds and growth of spinach. Biol. Trace Elem. Res. 2005, 105, 83–91. [Google Scholar] [CrossRef]
  215. Yang, F.; Hong, F.; You, W.; Liu, C.; Gao, F.; Wu, C.; Yang, P. Influences of nano-anatase TiO2 on the nitrogen metabolism of growing spinach. Biol. Trace Elem. Res. 2006, 110, 179–190. [Google Scholar] [CrossRef]
  216. Zheng, L.; Mingyu, S.; Xiao, W.; Chao, L.; Chunxiang, Q.; Liang, C.; Hao, H.; Xiaoqing, L.; Fashui, H. Antioxidant stress is promoted by nano-anatase in spinach chloroplasts under UV-B radiation. Biol. Trace Elem. Res. 2008, 121, 69–79. [Google Scholar]
  217. Soundararajan, P.; Sivanesan, I.; Jana, S.; Jeong, B.R. Influence of silicon supplementation on the growth and tolerance to high temperature in Salvia splendens. Hortic. Environ. Biotechnol. 2014, 55, 271–279. [Google Scholar] [CrossRef]
  218. Kumar, V.; Ramakrishna, A.; Ravishankar, G.A. Influence of different ethylene inhibitors on somatic embryogenesis and secondary embryogenesis from Coffea canephora P ex Fr. Plant Cell Tissue Organ Cult. 2007, 43, 602–607. [Google Scholar] [CrossRef]
  219. Zhao, X.C.; Qu, X.; Mathews, D.E.; Schaller, G.E. Effect of ethylene-pathway mutations upon expression of the ethylene receptor ETR1 from Arabidopsis. Plant Physiol. 2002, 130, 1983–1991. [Google Scholar] [CrossRef] [PubMed]
  220. Pua, E.C.; Deng, X.; Koh, A.T.C. Genotypic variability of de novo shoot morphogenesis Brassica oleracea in vitro in response to ethylene inhibitors and putrescine. J. Plant Physiol. 1999, 155, 598–605. [Google Scholar] [CrossRef]
  221. Strader, L.C.; Beisner, E.R.; Bartel, B. Silver ions increase auxin efflux independently of effects on ethylene response. Plant Cell 2009, 21, 3585–3590. [Google Scholar] [CrossRef]
  222. Ali, H.; Khan, M.A.; Ullah, N.; Khan, R.S. Impacts of hormonal elicitors and photoperiod regimes on elicitation of bioactive secondary volatiles in cell cultures of Ajuga bracteosa. J. Photochem. Photobiol. B Biol. 2018, 183, 242–250. [Google Scholar] [CrossRef]
  223. Khan, I.; Awan, S.A.; Raza, M.A.; Rizwan, M.; Tariq, R.; Ali, S.; Huang, L. Silver nanoparticles improved plant growth and reduced the sodium and chlorine accumulation in pearl millet: A life cycle study. Environ. Sci. Pollut. Res. Int. 2021, 28, 13712–13724. [Google Scholar] [CrossRef]
  224. Cvjetko, P.; Milosic, A.; Domijan, A.M.; Vinkovic, I.; Tolic, S.; Peharec, P. Toxicity of silver ions and differently coated silver nanoparticles in Allium cepa roots. Ecotoxicol. Environ. Saf. 2017, 137, 18–28. [Google Scholar] [CrossRef]
  225. Mallick, K.; Witcomb, M.; Scurrella, M. Silver nanoparticle catalysed redox reaction: An electron relay effect. Mater. Chem. Phys. 2006, 97, 283–287. [Google Scholar] [CrossRef]
  226. Mukherjee, M.; Mahapatra, A. Catalytic effect of silver nanoparticle on electron transfer reaction: Reduction of [Co(NH3)5Cl](NO3)2 by iron(II). Colloids Surf. A Physicochem. Eng. Asp. 2009, 350, 1–7. [Google Scholar] [CrossRef]
  227. Vannini, C.; Domingo, G.; Onelli, E.; Prinsi, B.; Marsoni, M.; Espen, L.; Bracale, M. Morphological and proteomic responses of Eruca sativa exposed to silver nanoparticles or silver nitrate. PLoS ONE 2013, 8, 68752. [Google Scholar] [CrossRef] [PubMed]
  228. Vankova, R.; Landa, P.; Podlipna, R.; Dobrev, P.I.; Prerostova, S.; Langhansova, L.; Gaudinova, A.; Motkova, K.; Knirsch, V.; Vanek, T. ZnO nanoparticle effects on hormonal pools in Arabidopsis thaliana. Sci. Total Environ. 2017, 59, 535–542. [Google Scholar] [CrossRef] [PubMed]
  229. Khodakovskaya, M.V.; de Silva, K.; Nedosekin, D.A.; Dervishi, E.; Biris, A.S.; Shashkov, E.V.; Galanzha, E.I.; Zharov, V.P. Complex genetic, photothermal, and photoacoustic analysis of nanoparticle-plant interactions. Proc. Natl. Acad. Sci. USA 2011, 108, 1028–1033. [Google Scholar] [CrossRef]
  230. Vander Willigen, C.; Postaire, O.; Tournaire-Roux, C.; Boursiac, Y.; Maurel, C. Expression and inhibition of aquaporins in germinating Arabidopsis seeds. Plant Cell Physiol. 2006, 47, 1241–1250. [Google Scholar] [CrossRef] [PubMed]
  231. Khodakovskaya, M.V.; de Silva, K.; Biris, A.S.; Dervishi, E.; Villagarcia, H. Carbon nanotubes induce growth enhancement of tobacco cells. ACS Nano 2012, 6, 2128–2135. [Google Scholar] [CrossRef]
  232. 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]
  233. Rajaee Behbahani, S.; 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 bitter melon (Momordica charantia); an in vitro experiment. PLoS ONE 2020, 15, e0235556. [Google Scholar] [CrossRef]
  234. Gelvin, S.B. Agrobacterium-mediated plant transformation: The “Gene-Jockeying” tool: The biology behind. Microbiol. Mol. Biol. Rev. 2003, 67, 16–37. [Google Scholar] [CrossRef]
  235. Price, D.R.G.; Gatehouse, J.A. RNAi-mediated crop protection against insects. Trends Biotechnol. 2008, 26, 393–399. [Google Scholar] [CrossRef]
  236. Izuegbunam, C.L.; Wijewantha, N.; Wone, B.; Ariyarathne, M.A.; Seredab, G.; Wone, B.M.W. A nano-biomimetic transformation system enables in planta expression of a reporter gene in mature plants and seeds. Nanoscale Adv. 2021, 3, 32–40. [Google Scholar] [CrossRef]
  237. Abdallah, N.A.; Prakash, C.S.; McHughen, A.G. Genome editing for crop improvement: Challenges and opportunities. GM Crop. Food 2015, 6, 183–205. [Google Scholar] [CrossRef] [PubMed]
  238. Schwab, F.; Zhai, G.; Kern, M.; Turner, A.; Schnoor, J.L.; Wiesner, M.R. Barriers, pathways and processes for uptake, translocation and accumulation of nanomaterials in plants-Critical review. Nanotoxicology 2016, 10, 257–278. [Google Scholar] [CrossRef] [PubMed]
  239. Demirer, G.S.; Zhang, H.; Matos, J.L.; Goh, N.S.; Cunningham, F.J.; Sung, Y.; Chang, R.; Aditham, A.J.; Chio, L.; Cho, M.J.; et al. High aspect ratio nanomaterials enable delivery of functional genetic material without DNA integration in mature plants. Nat. Nanotechnol. 2019, 14, 456–464. [Google Scholar] [CrossRef] [PubMed]
  240. Wang, P.; Zhao, F.J.; Kopittke, P.M.J. Engineering crops without genome integration using nanotechnology. Trends Plant Sci. 2019, 24, 574–577. [Google Scholar] [CrossRef] [PubMed]
  241. Albanese, A.; Tang, P.S.; Chan, W.C. The effect of nanoparticle size, shape, and surface chemistry on biological systems. Annu. Rev. Biomed. Eng. 2012, 14, 1–16. [Google Scholar] [CrossRef]
  242. Niazian, M.; MolaahmadNalousi, A.; Azadi, P.; Ma’mani, L.; Chandler, S.F. Perspectives on new opportunities for nano-enabled strategies for gene delivery to plants using nanoporous materials. Planta 2021, 254, 83. [Google Scholar] [CrossRef]
  243. Torney, F.; Trewyn, B.G.; Lin, V.S.; Wang, K. Mesoporous silica nanoparticles deliver DNA and chemicals into plants. Nat. Nanotechnol. 2007, 2, 295–300. [Google Scholar] [CrossRef]
  244. Liu, J.; Wang, F.; Wang, L.; Xiao, S.; Tong, C.; Tang, D.; Liu, X. Preparation of fluorescence starch-nanoparticle and its application as plant transgenic vehicle. J. Cent. S. Univ. Technol. 2008, 15, 768–773. [Google Scholar] [CrossRef]
  245. Fu, Y.Q.; Li, L.H.; Wang, P.W.; Qu, J.; Fu, Y.P.; Wang, H.; Sun, J.R.; Lü, C.L. Delivering DNA into plant cell by gene carriers of ZnS nanoparticles. Chem. Res. Chin. Univ. 2012, 28, 672–676. [Google Scholar]
  246. Vijayakumar, P.S.; Abhilash, O.U.; Khan, B.M.; Prasad, B.L.V. Nanogold-loaded sharp-edged carbon bullets as plant-gene carriers. Adv. Funct. Mater. 2010, 20, 2416–2423. [Google Scholar] [CrossRef]
  247. Burlaka, O.M.; Pirko, Y.V.; Yemets, A.I.; Blume, Y.B. Plant genetic transformation using carbon nanotubes for DNA delivery. Cytol. Genet. 2015, 49, 349–357. [Google Scholar] [CrossRef]
  248. Hao, Y.; Yang, X.; Shi, Y.; Xing, J.; Marowitch, J.; Chen, J.; Chen, J. FITC delivery into plant cells using magnetic single-walled carbon nanotubes. J. Nanosci. Nanotechnol. 2012, 12, 6287–6293. [Google Scholar] [CrossRef] [PubMed]
  249. Hao, Y.; Yang, X.; Shi, Y.; Song, S.; Xing, J.; Marowitch, J.; Chen, J.; Chen, J. Magnetic gold nanoparticles as a vehicle for fluorescein isothiocyanate and DNA delivery into plant cells. Botany 2013, 91, 457–466. [Google Scholar] [CrossRef]
  250. Naqvi, S.; Maitra, A.N.; Abdin, M.Z.; Akmal, M.; Arors, I.; Samim, M. Calcium phosphate nanoparticle mediated genetic transformation in plants. J. Mater. Chem. 2012, 22, 3500–3507. [Google Scholar] [CrossRef]
  251. Ardekani, M.R.S.; Abdeni, M.Z.; Nasrullaah, N.; Samim, M. Calcium phosphate nanoparticles a novel non-viral gene delivery system for genetic transformation of tobacco. Int. J. Pharm. Sci. 2014, 6, 605–609. [Google Scholar]
  252. Rafsanjani, M.S.O.; Kiran, U.; Ali, A.; Abdin, M.Z. Transformation efficiency of calcium phosphate nanoparticles for genetic manipulation of Cichorium intybus L. Ind. J. Biotechnol. 2016, 15, 145–152. [Google Scholar]
  253. Wang, Q.; Chen, J.; Zhang, H.; Lu, M.; Qiu, D.; Wen, Y.; Kong, Q. Synthesis of water soluble quantum dots for monitoring carrier-DNA nanoparticles in plant cells. J. Nanosci. Nanotechnol. 2011, 11, 2208–2214. [Google Scholar] [CrossRef]
  254. Wang, Q.; Chen, J.N.; Zhan, P.; Zhang, L.; Kong, Q.Q. Establishment of a suspension cell system for transformation of Jatropha curcas using nanoparticles. Adv. Mater. Res. 2013, 608–609, 314–319. [Google Scholar] [CrossRef]
  255. Martin-Ortigosa, S.; Valenstein, J.S.; Sun, W.; Moeller, L.; Fang, N.; Trewyn, B.G.; Wang, K. Parameters affecting the efficient delivery of mesoporous silica nanoparticle materials and gold nanorods into plant tissues by the biolistic method. Small 2012, 8, 413–422. [Google Scholar] [CrossRef]
  256. Martin-Ortigosa, S.; Valenstein, J.S.; Lin, V.S.Y.; Trewyn, B.G.; Wang, K. Gold functionalized mesoporous silica nanoparticle mediated protein and DNA codelivery to plant cells via the biolistic method. Adv. Funct. Mater. 2012, 22, 3576–3582. [Google Scholar] [CrossRef]
  257. Martin-Ortigosa, S.; Peterson, D.J.; Valenstein, J.S.; Lin, V.S.Y.; Trewyn, B.G.; Lyznik, L.A.; Wang, K. Mesoporous silica nanoparticle-mediated intracellular cre protein delivery for maize genome editing via loxP site excision. Plant Physiol. 2014, 164, 537–547. [Google Scholar] [CrossRef]
  258. Hajiahmadi, Z.; Shirzadian-Khorramabad, R.; Kazemzad, M.; Sohani, M.M.; Khajehali, J. A novel, simple, and stable mesoporous silica nanoparticle-based gene transformation approach in Solanum lycopersicum. 3 Biotech 2020, 10, 370. [Google Scholar] [CrossRef] [PubMed]
  259. Zolghadrnasab, M.; Mousavi, A.; Farmany, A.; Arpanaei, A. Ultrasound-mediated gene delivery into suspended plant cells using polyethyleneimine-coated mesoporous silica nanoparticles. Ultrason. Sonochem. 2021, 73, 105507. [Google Scholar] [CrossRef] [PubMed]
  260. Bao, W.; Wang, J.; Wang, Q.; O’Hare, D.; Wan, Y. Layered double hydroxide nanotransporter for molecule delivery to intact plant cells. Sci. Rep. 2016, 6, 26738. [Google Scholar] [CrossRef]
  261. Finiuk, N.; Buziashvili, A.; Burlaka, O.; Zaichenko, A.; Mitina, N.; Miagkota, O.; Lobachevska, O.; Stoika, R.; Blume, Y.; Yemets, A. Investigation of novel oligoelectrolyte polymer carriers for their capacity of DNA delivery into plant cells. Plant Cell Tissue Organ Cult. 2017, 131, 27–39. [Google Scholar] [CrossRef]
  262. Zhao, X.; Meng, Z.; Wang, Y.; Chen, W.; Sun, C.; Cui, B.; Cui, H. Pollen magnetofection for genetic modification with magnetic nanoparticles as gene carriers. Nat. Plants. 2017, 3, 956–964. [Google Scholar] [CrossRef] [PubMed]
  263. Abdel-Razik, A.B.; Hammad, I.A.; Tawfik, E. Transformation of thionin genes using chitosan nanoparticle into potato plant to be resistant to fungal infection. IOSRJ Biotechnol. Biochem. 2017, 3, 01–13. [Google Scholar] [CrossRef]
  264. Gil-Humanes, J.; Wang, Y.; Liang, Z.; Shan, Q.; Ozuna, C.V.; Sánchez-León, S.; Baltes, N.J.; Starker, C.; Barro, F.; Gao, C.; et al. High-efficiency gene targeting in hexaploid wheat using DNA replicons and CRISPR/Cas9. Plant J. 2017, 89, 1251–1262. [Google Scholar] [CrossRef]
  265. Farooq, N.; Ather, L.; Shafiq, M.; Nawaz-ul-Rehman, M.S.; Haseeb, M.; Anjum, T.; Abbas, Q.; Hussain, Q.; Ali, N.; Abbas, S.A.A.A. Magnetofection approach for the transformation of okra using green iron nanoparticles. Sci. Rep. 2022, 12, 16568. [Google Scholar] [CrossRef] [PubMed]
  266. Tawfik, E.; Ahmed, M.F. Chitosan nanoparticles as a new technique in gene transformation into different plants tissues. Nat. Resour. Human Health 2022, 2, 215–221. [Google Scholar] [CrossRef]
Table 1. Effects of various NP on seed germination ofdifferent crops under in vitro conditions.
Table 1. Effects of various NP on seed germination ofdifferent crops under in vitro conditions.
PlantNanoparticle (NP) TreatmentParametersReference
Brassica juncea
var. pusa jaikisan
AgNPsEnhancement in the growth of seedlings in terms of shoot FW, shoot and root length, and vigor index[62]
Brassica napusAgNPs/AuNPsDeceased shoot and root lengths, as well as shoot FW and DW[60]
Brassica nigraZnONPsIncreased shoot length and shoot DW, decreased root length, shoot FW, root FW and DW[65]
Brassica nigraCuONPsDelayed seed germination, decreased plantlet length, and their FWs and DWs[5]
Brassica oleracea
var. sabellica ‘Nero di Toscana’
AgNPsIncreased germination response, shoot and root lengths, as well as biomass[59]
Cicer arietinumCuONPsDecreased shoot and root lengths, FWs and DWs of shoot and root, increased lignifications in root cells[66]
Eruca sativaAuNPs, CuNPs and AgNPsAgNP-increased seed germination, shoot and root lengths, and seed vigourindex;
AuNP-and CuNP-decreased seed germination, shoot and root lengths, and seed vigour index
[56]
Glycine max
hybrid S42-T4
MWCNTsEarly and better germination, increased shoot, root and leaf lengths, shoot and root FWs and DWs[67]
Hylocereus undatusAgNPsIncreased germination, shoot number, shoot, and root lengths, cladode size, and FW[63]
Hordeum vulgare hybrid RobustMWCNTsEarly and better germination, increased shoot, root, and leaf lengths, shoot and root FWs and DWs[67]
Linum usitatissimum cv. BarbaraZnONPsIncreased shoot and root length, as well as their FWs and DWs[68]
Nicotiana tabacumAgNPsIncreased germination and dry biomass[61]
Oryza sativa
cv. Swarna
AgNPsIncreased shoot and root length, FWs and DWs of shoot and root[57]
Pennisetum glaucumAgNPsIncreased germination, seed vigour index, shoot and root lengths, and fresh and dry biomass[58]
Petroselinum
crispum
TiO2NPsIncreased germination, shoot and root lengths, and their FWs[69]
Phaseolus radiatusAgNPsAdverse effect on seedling growth[37]
Phaseolus vulgarisAgNPsIncreased seed germination, shoot and root length, their FWs and DWs, number of axillary buds, adventitious buds and leaves[64]
Physalis peruvianaAgNPsDecreased shoot and root lengths, chlorophyll content, but increased FW and DW[70]
Raphanus sativus
var. sativus ‘Ramona’
AgNPsIncreased germination response, shoot and root lengths, and seedling biomass[59]
Solanum lycopersicum var. PoranekAgNPsDecreased germination response, shoot and root lengths, and seedling biomass[59]
Sorghum bicolourAgNPsAdverse effect on seedling growth[37]
Vigna radiataCuONPsDecreased shoot and root lengths and their FWs, increased lignifications in root cells[71]
Zea mays
hybrid N79Z 300GT
MWCNTsEarly and better germination, increased shoot, root and leaf lengths, shoot and root FW and DW[67]
AgNPs: silver nanoparticles; AuNPs: gold nanoparticles; CuNPs: copper nanoparticles; CuONPs: copper oxide nanoparticles; DW: dry weight; FW: fresh weight; MWCNTs: multi-walled carbon nanotubes; TiO2NPs: titanium dioxide nanoparticles; ZnONPs: zinc oxide nanoparticles.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pathak, A.; Haq, S.; Meena, N.; Dwivedi, P.; Kothari, S.L.; Kachhwaha, S. Multifaceted Role of Nanomaterials in Modulating In Vitro Seed Germination, Plant Morphogenesis, Metabolism and Genetic Engineering. Plants 2023, 12, 3126. https://doi.org/10.3390/plants12173126

AMA Style

Pathak A, Haq S, Meena N, Dwivedi P, Kothari SL, Kachhwaha S. Multifaceted Role of Nanomaterials in Modulating In Vitro Seed Germination, Plant Morphogenesis, Metabolism and Genetic Engineering. Plants. 2023; 12(17):3126. https://doi.org/10.3390/plants12173126

Chicago/Turabian Style

Pathak, Ashutosh, Shamshadul Haq, Neelam Meena, Pratibha Dwivedi, Shanker Lal Kothari, and Sumita Kachhwaha. 2023. "Multifaceted Role of Nanomaterials in Modulating In Vitro Seed Germination, Plant Morphogenesis, Metabolism and Genetic Engineering" Plants 12, no. 17: 3126. https://doi.org/10.3390/plants12173126

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