Next Article in Journal
Cytogenetic, Morphometric, and Ecological Characterization of Festuca indigesta Boiss. in the Southeast of Spain
Next Article in Special Issue
Characterization and Toxicity of Hypoxoside Capped Silver Nanoparticles
Previous Article in Journal
Impact of Nanomaterials on the Regulation of Gene Expression and Metabolomics of Plants under Salt Stress
Previous Article in Special Issue
Role of Synthetic Plant Extracts on the Production of Silver-Derived Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanobionics in Crop Production: An Emerging Approach to Modulate Plant Functionalities

1
Academy of Biology and Biotechnology, Southern Federal University, Stachki 194/1, 344090 Rostov-on-Don, Russia
2
Proteomics Laboratory, Division of Plant Biotechnology, Sher-e-Kashmir University of Agricultural Sciences & Technology of Kashmir, Shalimar, Srinagar 190025, India
3
Key Laboratory of Ministry of Education for Genetics, Breeding and Multiple Utilization of Crops, Center of Legume Crop Genetics and Systems Biology/College of Agriculture, Oil Crops Research Institute, Fujian Agriculture and Forestry University (FAFU), Fuzhou 350002, China
4
Department of Biotechnology, Chonnam National University, Yeosu 59626, Korea
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(5), 692; https://doi.org/10.3390/plants11050692
Submission received: 24 January 2022 / Revised: 23 February 2022 / Accepted: 1 March 2022 / Published: 4 March 2022
(This article belongs to the Special Issue Nanotechnology Advances in Plant Science and Biotechnology)

Abstract

:
The “Zero Hunger” goal is one of the key Sustainable Development Goals (SDGs) of the United Nations. Therefore, improvements in crop production have always been a prime objective to meet the demands of an ever-growing population. In the last decade, studies have acknowledged the role of photosynthesis augmentation and enhancing nutrient use efficiency (NUE) in improving crop production. Recently, the applications of nanobionics in crop production have given hope with their lucrative properties to interact with the biological system. Nanobionics have significantly been effective in modulating the photosynthesis capacity of plants. It is documented that nanobionics could assist plants by acting as an artificial photosynthetic system to improve photosynthetic capacity, electron transfer in the photosystems, and pigment content, and enhance the absorption of light across the UV-visible spectrum. Smart nanocarriers, such as nanobionics, are capable of delivering the active ingredient nanocarrier upon receiving external stimuli. This can markedly improve NUE, reduce wastage, and improve cost effectiveness. Thus, this review emphasizes the application of nanobionics for improving crop yield by the two above-mentioned approaches. Major concerns and future prospects associated with the use of nanobionics are also deliberated concisely.

1. Introduction

Nanotechnology (NT) is witnessing exponential growth in several sectors, and it is believed to have more advancement in the near future. The global market of NT reached up to USD 42.2 billion in the year 2020 despite the global COVID-19 pandemic, and it is estimated to reach USD 70.7 billion by 2026 with a surging Compounded Average Growth Rate (CAGR) of 9.2% (Globe News Wire, 2021). The unique physicochemical features of nanoparticles (NPs) empower their significant application in crop production and food security [1,2,3]. Attaining food security by enhancing agricultural production has been a major goal for scientists and technologists across the globe.
Metal-based NPs are the most studied and found greatly useful in crop production to enhance the growth of crops and protect their health from various biotic stressors [4,5], as well as protect from abiotic stresses such as drought, temperature, humidity, and higher salt concentrations [6,7,8]. The agriculture sector has taken up the application of NPs progressively as a result of several commercial formulations such as Nano-Calcium, Magic Green (PAC International Network Co., Ltd., Köln, Germany), Nano Urea (liquid) fertilizer (Indian Farmers Fertilizer Cooperative Limited, New Delhi, India), NanoRise™ (Aqua-Yield Hub, Sandy, UT, USA), etc., are now available to enhance crop yield and qualitative traits [9,10].
The concept of nanobionics in agriculture deals with the use of NPs and their interaction with a plant’s system to improve or modulate specific functions that can notably boost crop production [11]. Nanobionics features implanting the specific NPs into plants to unleash the obscured potential that may not have been accomplished without the use of NPs. For instance, chloroplasts have a limited capacity to absorb light in the visible range, and they further limit the utilization of solar energy by almost 50% [12]. Ineptly, the photosynthetic capacity is saturated by nearly 10% of total sunlight [13]. In such circumstances, nanobionics can significantly improve photosynthesis by various means [14].
The photocatalytic ability of NPs (e.g., TiO2 NPs) expands the range of absorption of sunlight specifically in the near-infrared range, which is substantially helpful in improving the absorption of solar radiation by leaves, thereby enhancing photosynthesis. The process is governed by modulation of the electron transport chain, photophosphorylation, carboxylation of ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO), and also by preventing the aging of chloroplasts [15,16]. Enhanced expression of RuBisCO activase (rca) mRNA (by 51% compared to the control that was treated with bulk TiO2 NPs) has been proven to be the fundamental principle that helps in improved carbon metabolism by photosynthesis [17]. The application of nanobionics has also been effective on plants under mild heat stress by increasing the water conductance and transpiration as well as modulating the energy dissipation of photosystem-II [18].
Improved nutrient use efficiency (NUE) through controlled, efficient fertilizer uses and site-specific pesticide delivery is another unique application of nanobionics in crop production. It ensures the delivery of the active ingredient only upon receiving the external or environmental stimuli, for example pH, temperature, humidity or even manmade stimuli such as ultrasonic triggers [19,20,21]. Thus, these applications are certainly promising in terms of improving NUE as well as protecting the non-target organism and environment from inevitable exposure of agrochemicals caused by their excessive use [22].
In this review article, the applications of metal NPs as nanobionics and their application in improving crop production by modulating photosynthesis are covered. The review also deals with nanobionics as targeted and stimuli-responsive NPs that can efficiently deliver agrochemicals and control their release. The future prospect of nanobionics is deliberated by focusing on the key areas that have to be dealt in the coming time regarding socioeconomic aspects pertaining to small farmers and developing countries.

2. Nanobionics and Modulation of Photosynthesis

Modulation of photosynthesis is always one of the most advantageous procedures that can considerably increase the rate of crop production [23]. In addition, the notion of manipulating photosynthesis has been a source of debate for years; nonetheless, new findings have asserted that it is practically viable [24]. The efficiency of photosynthesis among C3 and C4 plants vary due to the unfavourable function of RuBisCO [25]. In terms of photosynthesis, among all the plants in the ecosystem (C3 and C4 plants), the majority (85%) of them are C3. However, the C4 plants have a better photosynthetic ability than the C3 [26]. Previously, some studies were conducted using a conventional approach to influence the function of photosynthesis by (i) improving the efficiency of the enzyme RuBisCO [27], (ii) modification of C3 plants to use the C4 photosynthesis process [26,28], (iii) manipulating the chlorophyll antenna and their size within the photosystems [29], (iv) increasing the efficiency of chlorophyll to absorb light in extended wavebands [30], and improving the photoinhibition frequency [31].
NPs as nanobionics have considerably been effective in modulating the photosynthesis capacity of plants. It helps fast-growing plants by acting as an artificial photosynthetic system, thereby improving the process [32,33]. For example, TiO2 NPs have been successfully studied for the potential of enhancing photosynthesis by improving the light absorption of leaves [34,35,36]. They safeguard the chloroplast from photochemical stress that might lead to ageing [37]. Recent studies have reported that TiO2 NPs considerably improve RuBisCO, which is one of the key enzymes helpful in improving the rate of photosynthesis [25,38]. It has also been reported to enrich the chlorophyll content in Brassica napus and Solanum lycopersicum L. [39,40].
Other metal/metal oxide NPs have also been explored as nanobionics that modulate the photosynthesis process, thereby improving the growth and overall features of the plants. In recent years, Au-, ZnO- and Ag-based NPs are being investigated for their capacity to improve photosynthesis. They mainly act on the plants by improving the chlorophyll content of the plants [41], enhancing the relative electron transport rate [42], photochemical quenching [23], and aiding the photosystem functions [43]. However, these functions are again driven by specific doses of NPs, as the higher dosage has been reported to diminish the photosynthesis apparatus [44], rate of photosynthesis [45], photosystems [46], and other associated parameters. For example, Ag NPs in the concentration ranging from 40–60 mg L−1 have successfully been studied for improving the chlorophyll content in the Trigonella foenum-graecum L. (effective dose of 40 mg L−1) [47], Cucumis sativus (50 mg L−1) [48], Eschscholzia californica Cham (25 mg L−1) [49], and Helianthus annuus L. (60 mg L−1). A higher dose (100 mg L−1) caused sharp decline (by 48.3%) in the concentration of chlorophyll [50]. Ag NPs are also reported to enhance the coefficient of photochemical quenching, thereby implying an inhibition of the electron transport chain [51].
In the case of myco-synthesized TiO2 NPs (50–100 nm, spherical), the dose of 25–100 mg L−1 could induce only a slight increase in the total chlorophyll content [52], or a dose-dependent increase [53]. However, it was reported that at 250 mg kg−1 of chemically synthesized TiO2 (15–40 nm, spherical), NPs when coupled with Plant Growth Promoting Rhizobacteria (PGPR) saw a 43.27% increase in total chlorophyll [54]. TiO2 NPs have excellent photocatalytic properties, and therefore 500 mg L−1 of TiO2 (2.0−5.0 nm, spherical) in combination with 210.87 mg L−1 of chloroplast, when coupled together to form a hybrid photosynthesis system, exhibited the highest electron transfer rate from PS-II to PS-I [55].
To further elaborate on these aspects two tables are appended below, in which Table 1 explains the application of metal NPs as nanobionics to improve the photosynthesis process, their beneficial or adverse effect, and mode of action. Table 2 includes the recent findings on metal NPs that improve or affect the plant’s chlorophyll content, their effective dose, and mode of application.
ZnO NPs are most effective NPs at the lower concentration (approximately 10–50 mg L−1) either by seed priming or foliar application. They enhance chlorophyll content by 40–49% and also enhance the photosystem efficiency, stomatal conductance, and CO2 assimilation up to 35%. ZnO nanostructures enabled nano-fertilizer enhanced growth of Lycopersicum esculentum L. plants via improving cellular enzyme activities [71]. It is well noted that Zn is integral part of antioxidative enzymes (superoxide catalase and dismutase), which regulates ROS generation in plant cells and is used by polymerase, kinase, dehydrogenase, and phosphatase enzymes in the photosynthesis process [72]. During the stress conditions, Zn can modulate the cellular redox and antioxidant defence system. In a recent study, it was shown that ZnO NPs with the combination of salicylic acid improved the activities of CAT, POX, and SOD in rice under As stress condition [73]. Based on the cellular importance of Zn, the Zn-based nano-fertilizer could enhanced the effectiveness of micronutrients applied in agriculture [74].
TiO2 NPs, however, require a higher dosage of application. The lower dose of TiO2 either causes a slight improvement in chlorophyll content or this happens in a dose-dependent manner. A hybrid photosynthetic apparatus prepared with chloroplasts coupled with a higher dose of TiO2 NPs has been studied to enhance the photosynthesis by improving the electron transfer rate in the photosystems; however, its practical application is still subject to approval for field application. Ag NPs exhibit effects similar to ZnO NPs that requires lower doses, and Au NPs have worked well with a higher dosage. However, for using Ag NPs and Au NPs, cost effectiveness could be a major issue.
The above studies are conclusive for the use of ZnO NPs that have shown excellent results at lower concentrations and are believed to have practical aspects for field application, in addition to their lower ecotoxicological implications, especially for green synthesized NPs inputs [75].

3. Nanobionics to Improve Nutrient Use Efficiency

Even though the soil is nutrient-rich, crops are often unable to effectively utilize those nutrients, because modern agriculture completely relies upon the fertilizers that can avail the macro and micronutrients such as N, K, Mg, P, S, Ca, etc., and Cu, Mn, Fe, B, Cl, Mo, and Zn, etc., respectively [76,77]. However, these commercial and chemical fertilizers have limits, and they can only provide 30 to 50% of total crop yield nutrition [78]. NUE is very much dependent on the soil characteristics (physicochemical properties), vaporization of chemical fertilizers and their leaching and also on the physicochemical characteristics of the fertilizers [79]. The cost of production in agriculture surges with the lowered NUE, and that ultimately affects the lower-income group farmers as well as stakeholders. Fertilizers with lower NUE also affect the environment and ecosystems, as their utilization is more compared to fertilizers with higher NUE [80]. The best agriculture practices must include fertilizers with greater NUE so that they can be cost effective and have minimal environmental impacts.
Nanobionics have emerged as a great means that can improve the NUE by employing a smart delivery system (Figure 1). Such nanocarriers are capable of performing controlled release of the ingredient triggered by various environmental conditions, and can load tiny amounts of fertilizers, thereby widening their distribution [81]. The release of nutrients depends on the types of carriers, choice of fertilizers and design specific to the stimuli.
The release even can be pre-planned as per the requirements and executed through some external stimuli. These stimuli can be pH, temperature, humidity, molecular recognition, light, and ultrasonic shock waves [82,83,84]. Such controlled release of ingredients acts as a function of time or interaction with an appropriate environment, as per the strategy of the nano-encapsulation (Figure 2).
Owing to the promising applications of stimuli-responsive NPs for agricultural application, in this context, few studies have reported the controlled release of active ingredients along with extended-release duration, which are vital to their field application. Polymeric ZnO NPs that were intended to work with agricultural soils at specific pH and organic matter exhibited a slow-release pattern of Zn in the soil with time. This demonstrated that the release of Zn was in accordance with the requirement of Zea mays L., thereby preventing excess Zn bio-availability in the soil and therefore ecotoxicity [85].
Engineered SiO2 NPs (SiO2: NPs as the core, chitosan: semi-permeable coating, and Sodium Alginate and Kaolin: outermost superabsorbent coating) ((K-SA-CS-SiO2 NPs) loaded with NPK have exhibited sustained release of nutrients for six months at room temperature. They have also helped in fighting drought and salinity stress [86]. SiO2 NPs with polyacrylate emulsion used for the formulation of controlled release urea formulation (CRUF) along with polyacrylate/silica hybrid were found to be significant. The 1.0 wt.% SiO2 used with polyacrylate emulsion reduced the solubility of CRU 38.3 wt.% to 2.2 wt.%. The release time was also extended from 8 to 27 days [87].
Apart from metal and slow-release type NPs, the pH-responsive NPs composed of poly[(meth)acrylic acid] and poly[N,N dimethylaminoethyl(meth)acrylate] are recognized to be functional in a range of physiological conditions. They have been reported to be promising for agricultural application due to their nondegradable all-carbon backbones [88]. Other NPs are also being studied for assessing the potential material for nano-carriers composed of chitosan, polyacrylic acids and zeolites [14,89,90].
A wide range of commercial preparations of nano-fertilizers are available around the globe, such as nitrogen (IFFCO Nano Urea, IFFCO, New Delhi, India), phosphorus (TAG Nano Phos, SK Organic Farms, Chennai, India), potassium (NanoMax Potash, JU Agri Sciences, New Delhi, India), potassium and phosphorus (Fosvit K30, Kimitec Group, Almería, Spain), zinc (Geolife Nano Zn, Geolife Agritech India Ltd., Mumbai, India, Silvertech Kimya Sanayi ve Ticaret Ltd., Istanbul, Turkey and Nano Zinc Chelate Fertilizer, AFME Trading Group, Canary Wharf, UK), and iron and magnesium (Nubiotek®Hyper Fe+Mg, Bioteksa, Chihuahua, Mexico) [91].

4. Foliar Application of Nanoparticles

4.1. Adhesion to the Foliar Surface

Foliar application of NPs has been documented to be very effective and efficient for crop production. It can reduce the leaching of agrochemicals, thereby reducing the impacts on the environment and ecosystem [58]. Controlled and targeted delivery ensures localized delivery of the active ingredients on the crops for effective utilization and improved efficiency [92]. Cuticle and epidermal tissues of the foliage have distinct features and environmental conditions [93]. In the kingdom Plantae, amphistomatic, epistomatic lamina and hypostomatic are differentiations that are controlled by environmental conditions such as temperature and humidity [94]. The design of NPs should be focused to target the types of foliage morphology and chemical characteristics. Usually, the leaf cuticles have an abundance of wax layers that enable low surface energy, which results in increased contact angle of the water, hence preventing wetting [95]. Therefore, to ensure the sufficient adhesion of the NPs after a foliar application, the design should be preferred that can consider their shape, size and surface chemistry [96]. The most preferred shape of the carriers are spherical; however, rod- and platelet-shaped carriers can be given preference for the foliar application pertaining to their better contact areas [96].
A report on a poly-lactic acid (PLA) carrier of 450 nm loaded with abamectin (an insecticide) was studied for improving the hydrophobicity and positive surface groups. NPs with functional groups such as acetyl, amine or carboxylic acid were greatly helpful in adhesion to the cucumber leaves, and the order of adhesion was observed to be amine > acetyl > carboxylic acid. Particles modified with the amine functional group showed the strongest adhesion, which retained itself even after multiple irrigations [97]. Nanoionics surfaces modified with tannic acid have also shown enhanced adhesion to foliage tissues for delivery of insecticide and fungicides [97,98]. The adhesion is greatly contributed by H-bonding brought about by OH groups present on the phenol rings, and 50% greater adhesion was achieved by the modified nanobionics compared to the unmodified surface. The modification with polyphenolic groups greatly increases the photostability of the active ingredient without disturbing the release event [99]. Other surface modifications such as carboxymethylcellulose with ethylenediamine [100], graphene oxide nanosheets with polydopamine [101], and styrene-methacrylic acid are also successful in improving the adhesion of NPs to the plants [102].

4.2. Uptake and Translocation within the Plant System

Plant uptake of NPs from the soil is achieved by roots that distribute and translocate up to the leaves. However, in the case of foliar application, the leaf to leaf translocation in distribution is common [103]. Translocation through the roots is a normal hydroponic phenomenon attributed to the plant vascular system; however, the uptake via leaves is important to foliar application and particulate delivery [104]. Uptake after the foliar application is greatly influenced by physicochemical properties of the NPs such as size, shape, and surface chemistry [105]. Studies have elaborated the understanding of plants’ mechanisms and distribution limits and their fate in the plant system after the uptake [106]. Internalization of NPs by leaves usually follows many routes. They enter through stomata via the symplastic or apoplastic transport system of the plants [107,108]. The apertures of stomata that are usually in micrometres allows spontaneous uptake of the NPs, and leaf mesophylls play a vital role in the transport and distribution. Smaller NPs can penetrate the junctions of leave cuticles [109]. As NPs are transferred from soil into plants, they are mobilized by apoplastic and symplastic mechanisms. It is reported that the apoplastic pathway supports radial distribution, which leads NPs to be directed toward the root core cylinder and vascular systems, as well as an upward tendency towards the aerial sections of the plant [110,111]. The apoplastic pathway is critical for the transport of NPs, whereas the Casparian strip prevents NPs from migrating radially in the endodermis of roots. The Casparian strip can be avoided by switching from the apoplastic to the symplastic pathway [112,113]. This pathway is more orderly and well-regulated than the other pathways that allow NPs to transit through the plant body. Once the NPs have reached the cytoplasm of the cell, the plasmodesmata aid in cell-to-cell migration [114].
A study using Au NPs of various sizes (3, 10 or 50 nm) and surface chemistry (treated with polyvinylpyrrolidone (PVP) and citrate) was performed to explore the role of physicochemical properties of the carrier NPs on interactions and translocation in the plant system. The study inferred that distribution in the plant was greatly affected by size and surface chemistry. The PVP-modified NPs showed better adhesion compared to citrate, which could not penetrate the cuticle for almost 15 days. However, the PVP-Au NPs were internalized and translocated up to the mesophyll cells. The smaller Au NPs (3, 10 nm) after foliar application were more persistent with respect to Au NPs of size 50 nm [96].
Another study that included the liposomes of 100 nm prepared from hydrogenated soybean L-α-phosphatidylcholine carrying Fe and Mg ions were greatly helpful in overcoming the acute nutrient deficiency of the tomato. The study reported up to one-third of the applied liposome penetrated the leaves and delivered the ions quickly compared to conventional treatment that penetrates 1% of the total free molecules applied in the same fashion. The liposomes were readily distributed to the plant system and detected in the roots after just 24 h of exposure [115].
The poly(ε-caprolactone) nanocapsules of 300 nm that carried atrazine were studied for leaf uptake after foliar application on mustard leaves. The nanocapsules remained active on the surface for 24 h after the application. After 48 h, they translocated to vessel elements and after 96 h of the treatment, they were found in the gaps between the mesophyll cells. The atrazine encapsulated when delivered in a site-specific manner demonstrated excellent herbicide activity even after being diluted 10 times [116]. Findings from the studies affirm that NPs that are used as a carrier must pass through several physiological barriers until they are taken up by the plants and translocated. Interfacial interactions between the foliar tissue and the nanobionics play a decisive role.
The impact of shape, geometry, and dimension of nanobionics are the factors that alter their uptake and translocation in the plant system, and they should be further systematically studied on different plants for their beneficial or adverse effects.

5. Major Concern and Future Prospects

The application of NPs in crop production is promising yet in the nascent phase. Both engineered and green synthesized NPs have been explored extensively for improving crop production and other agricultural advances. Considering the application of nanobionics in augmenting photosynthesis and enhancing NUE, it is certainly believed that NT has considerable potential to revamp conventional agriculture practices. However, several other applications of NPs have equal potential as nanobionics. For example, NPs can be used to provide abiotic and biotic stress tolerance, site-specific delivery of agrochemicals, and nano-sensors for precision agriculture. Nano-enabled products are now being developed and debated in the scientific community for their agricultural application. Nearly 9425 nano-based products have been developed so far, which was contributed by 64 countries and 2658 industries, among which 229 products are for agriculture developed by 73 industries and 16 countries (https://product.statnano.com/; accessed on 10 January 2021). In the coming time, the focus of NP research and application must be addressed on the development of more nano-enabled products that can address conventional agricultural issues in an intelligent way.
However, with mere development of products, other aspects pertaining to the application of NPs in agriculture must be deliberated in the future: (a) guidelines for manufacturing, handling, and application, (b) crop-specific dose of application, (c) availability in the environment and their ecotoxicological concerns, and (d) toxicological aspects in food chains, including humans.
With the growing application of NPs their release in soil and water is gradually becoming inevitable. Deliberate application of NPs on the crops may lead to accumulation in the soil, thereby uptake by vital soil microorganisms and lower organisms is useful to agriculture. This may affect total organic content, nutrient cycling, biogeochemistry of the crop field and the overall ecosystem [117,118]. Metal-based NPs are most likely to release their constituents in the soil as elemental or ionic forms that can be used as nutrients (in the case of NPs of Cu, Mg, Mn, Se, Zn, etc.); however, the NPs of Ag, As, Pt, Ti, etc., that are equally promising in crop production, may accumulate after continuous and repeated applications and affect the environment negatively [119,120,121]. This can affect the germination of seeds as well as root and shoot development, thereby affecting the biomass and yield of the crops. NPs such as Ag NPs have been analysed in the soil, surface water, and sediment [59,121,122,123]. The accumulated NPs are transported via interfaces of soil particles and affect the beneficial soil microbes undesirably [118,123].
On the one hand, NPs have potential applications in improving crop production, and on the other hand, their repeated application causes a growing threat to agricultural soil and crops to hamper crop growth [6]. Therefore, further studies that can define and ensure the crop/disease/soil specific dosage, mechanism of internalization, and translocation and distribution in the plant system are prerequisites and challenges for this decade [8].
The ecotoxicological and toxicological aspect is an important concern that should also be dealt with in utmost tenaciousness. Once NPs enter the environment they undergo several modifications, such as dissolution, binding with soil/sediment particles, sulfidation aggregation/agglomeration, etc., and these modifications are key for assessing the accumulation in soil and biota, precipitation and also toxicity [124]. Other studies have also reported several NPs for the trophic transfer from green algae to fish by a two-stage food chain [125], in river snails (Cipangopaludina chinensis), and Chinese muddy loaches (Misgurnusmizolepis) [126]. These studies could be alarming as well as trendsetters, followed by which further studies should delineate more sensitive biotas. Interaction of NPs with key protein and biomarkers (also from humans), genotoxicity, toxicodynamics and toxicokinetics are the key areas where more data generation is certainly required.
The feasibility of acceptance of NP-based formulations is one of the biggest challenges identified ahead in this decade. No matter how great the potential of NP-formulations, if they are not cost effective, it will be hard to find a market among the developing and underdeveloped countries where the livelihoods of the majority of the farmers are dependent on limited land resources and smallholders. Precision farming and the use of nanobionics may imply sensors and gadgets that would be far beyond the reach of smallholders.
To sum up, the sustainable development that the world is striving for should be inclusive. If such development can improve crop production, at the same time it is required to protect the natural resources (land, water and forests), biota, and also human beings. To seek and achieve the “Zero Hunger SDG 2030”, research planning should be ahead of the times to protect from the aftermath and adverse consequences, unlike the implementation of the use of chemical pesticides that have caused extensive damage to the environment and continue to persist in damage to environmental health. Application of NPs by the nanobionics approach would certainly shape and build the future of crop production, but understanding all the aspects of their practical application is necessary before they unleash their harmful impacts on the environment and human health.

6. Conclusions

Nanobionics can modulate photosynthesis by improving the light absorption up to the near-infrared region, improving the number of thylakoids and chlorophyll content that overall boosts the biomass and crop production. Nanocarriers that respond to external stimuli such as pH, temperature, humidity, pathogens, or even manmade stimuli to deliver agrochemicals in a controlled release are promising for improving the NUE and minimizing wastage. These two approaches are considerable for cost-effective crop production and sustainable agriculture. They can together improve conventional agriculture practices to a smarter and affordable agriculture. Thus, to achieve the “Zero Hunger” goal of UN SDG by 2030 there is an urgent need for a paradigm shift in conventional agriculture practices that can effectively be brought about using emerging and smart NPs such as nanobionics.
However, the increasing application of NPs in agriculture, where they directly interact with the environment, poses threats to aquatic and terrestrial species as well as humans. Deliberate application may lead to the accumulation of NPs or their constituent elements in the soil [118], and that could affect physicochemical characteristics as well as fertility of the soil. Such accumulation can also affect the plant health brought about by NP-induced oxidative stress, cell death and genotoxicity, thereby affecting the biomass of the plant, chlorophyll content, and resulting in diminished yield. Moreover, beneficial microbes and lower soil animals (e.g., nematodes and earthworms) who ensure better nutrient cycling and maintain soil health are also affected.
Colloidal microstructures or aggregates of the soil particles are responsible for carrying the NPs from the soil to water bodies, where they are taken up by lower trophic aquatic species and then biomagnified. NPs affect the biological system by interfering in the biochemical processes at the molecular as well as tissue organ levels. The ROS-induced oxidative stress, followed by DNA damage, apoptosis, and altered protein functions, are key events by which NPs exert toxicological effects.

Author Contributions

Conceptualization, A.R. (Anuj Ranjan), V.D.R., and T.M.; methodology, A.R. (Anuj Ranjan), V.D.R., S.S., and S.S.M.; formal analysis, V.D.R., A.K., and E.V.P.; resources, T.M. and S.S.M.; data curation, V.D.R., S.M.Z., and T.M.; writing—original draft preparation, A.R. (Anuj Ranjan); writing—review and editing, A.R. (Anuj Ranjan), V.D.R., and A.K.; critical review, G.C., A.R. (Ali Raza), V.D.R., S.S.M., S.S., and S.M.Z.; supervision, T.M. and S.S.M.; financial assistance, A.R. (Ali Raza) and G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The research was supported by the “Strategic Academic Leadership Program of the Southern Federal University (“Priority 2030”)” and by the “Cooperative Research Program for Agriculture Science and Technology Development (Project No. PJ01665001)”, Rural Development Administration, Republic of Korea. The authors (T.M., S.S., S.S.M.) would like to thank the Centres for collective use of Southern Federal University “Modern Microscopy” and “High Technology” for performing chemical-analytical experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bakshi, M.; Abhilash, P.C. Nanotechnology for soil remediation: Revitalizing the tarnished resource. Nano-Mater. Photocatal. Degrad. Environ. Pollut. Chall. Possibilities 2020, 11, 345–370. [Google Scholar] [CrossRef]
  2. Bhusare, S.; Kadam, S. Applications of Nanotechnology in Fruits and Vegetables. Food Agric. Spectr. J. 2021, 2, 90–95. [Google Scholar]
  3. Usman, M.; Farooq, M.; Wakeel, A.; Nawaz, A.; Cheema, S.A.; ur Rehman, H.; Ashraf, I.; Sanaullah, M. Nanotechnology in agriculture: Current status, challenges and future opportunities. Sci. Total Environ. 2020, 721, 137778. [Google Scholar] [CrossRef]
  4. Tourinho, P.S.; Van Gestel, C.A.M.; Lofts, S.; Svendsen, C.; Soares, A.M.V.M.; Loureiro, S. Metal-based nanoparticles in soil: Fate, behavior, and effects on soil invertebrates. Environ. Toxicol. Chem. 2012, 31, 1679–1692. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, H. Metal based nanoparticles in agricultural system: Behavior, transport, and interaction with plants. Chem. Speciat. Bioavailab. 2018, 30, 123–134. [Google Scholar] [CrossRef] [Green Version]
  6. Ranjan, A.; Rajput, V.D.; Minkina, T.; Bauer, T.; Chauhan, A.; Jindal, T. Nanoparticles induced stress and toxicity in plants. Environ. Nanotechnol. Monit. Manag. 2021, 15, 100457. [Google Scholar] [CrossRef]
  7. Liu, R.; Lal, R. Potentials of engineered nanoparticles as fertilizers for increasing agronomic productions. Sci. Total Environ. 2015, 514, 131–139. [Google Scholar] [CrossRef]
  8. Rajput, V.D.; Minkina, T.; Kumari, A.; Singh, V.K.; Verma, K.K.; Mandzhieva, S.; Sushkova, S.; Srivastava, S.; Keswani, C. Coping with the Challenges of Abiotic Stress in Plants: New Dimensions in the Field Application of Nanoparticles. Plants 2021, 10, 1221. [Google Scholar] [CrossRef]
  9. Kumar, S.; Nehra, M.; Dilbaghi, N.; Marrazza, G.; Hassan, A.A.; Kim, K.-H. Nano-based smart pesticide formulations: Emerging opportunities for agriculture. J. Control. Release 2019, 294, 131–153. [Google Scholar] [CrossRef]
  10. Sabourin, V. Commercial opportunities and market demand for nanotechnologies in agribusiness sector. J. Technol. Manag. Innov. 2015, 10, 40–51. [Google Scholar] [CrossRef] [Green Version]
  11. Giraldo, J.P.; Landry, M.P.; Faltermeier, S.M.; McNicholas, T.P.; Iverson, N.M.; Boghossian, A.A.; Reuel, N.F.; Hilmer, A.J.; Sen, F.; Brew, J.A.; et al. Erratum: Plant nanobionics approach to augment photosynthesis and biochemical sensing (Nature Materials (2014) 13 (400-408)). Nat. Mater. 2014, 13, 530. [Google Scholar] [CrossRef] [Green Version]
  12. Terry, N. Limiting factors in photosynthesis: I. Use of iron stress to control photochemical capacity in vivo. Plant Physiol. 1980, 65, 114–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhu, X.-G.; Long, S.P.; Ort, D.R. Improving photosynthetic efficiency for greater yield. Annu. Rev. Plant Biol. 2010, 61, 235–261. [Google Scholar] [CrossRef] [Green Version]
  14. Marchiol, L. Nanotechnology in agriculture: New opportunities and perspectives. New Vis. Plant Sci. 2018, 9, 161. [Google Scholar]
  15. Middepogu, A.; Hou, J.; Gao, X.; Lin, D. Effect and mechanism of TiO2 nanoparticles on the photosynthesis of Chlorella pyrenoidosa. Ecotoxicol. Environ. Saf. 2018, 161, 497–506. [Google Scholar] [CrossRef]
  16. Azmat, R.; Altaf, I.; Moin, S. The reflection of the photocatalytic properties of TiO2 nanoparticles on photosynthetic activity of Spinacia oleracea plants. Pak. J. Bot 2020, 52, 1229–1234. [Google Scholar] [CrossRef]
  17. Linglan, M.; Chao, L.; Chunxiang, Q.; Sitao, Y.; Jie, L.; Fengqing, G.; Fashui, H. Rubisco activase mRNA expression in spinach: Modulation by nanoanatase treatment. Biol. Trace Elem. Res. 2008, 122, 168–178. [Google Scholar] [CrossRef]
  18. Qi, M.; Liu, Y.; Li, T. Nano-TiO2 improve the photosynthesis of tomato leaves under mild heat stress. Biol. Trace Elem. Res. 2013, 156, 323–328. [Google Scholar] [CrossRef]
  19. Mustafa, I.F.; Hussein, M.Z. Synthesis and technology of nanoemulsion-based pesticide formulation. Nanomaterials 2020, 10, 1608. [Google Scholar] [CrossRef]
  20. Habibi, N.; Kamaly, N.; Memic, A.; Shafiee, H. Self-assembled peptide-based nanostructures: Smart nanomaterials toward targeted drug delivery. Nano Today 2016, 11, 41–60. [Google Scholar] [CrossRef] [Green Version]
  21. Rameshaiah, G.N.; Pallavi, J.; Shabnam, S. Nano fertilizers and nano sensors–an attempt for developing smart agriculture. Int. J. Eng. Res. Gen. Sci. 2015, 3, 314–320. [Google Scholar]
  22. Petosa, A.R.; Rajput, F.; Selvam, O.; Öhl, C.; Tufenkji, N. Assessing the transport potential of polymeric nanocapsules developed for crop protection. Water Res. 2017, 111, 10–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, Y.; Liu, N.; Wang, W.; Sun, J.; Zhu, L. Photosynthesis and related metabolic mechanism of promoted rice (Oryza sativa L.) growth by TiO2 nanoparticles. Front. Environ. Sci. Eng. 2020, 14, 1–12. [Google Scholar] [CrossRef]
  24. Lew, T.T.S.; Koman, V.B.; Gordiichuk, P.; Park, M.; Strano, M.S. The emergence of plant nanobionics and living plants as technology. Adv. Mater. Technol. 2020, 5, 1900657. [Google Scholar] [CrossRef]
  25. Qu, Y.; Sakoda, K.; Fukayama, H.; Kondo, E.; Suzuki, Y.; Makino, A.; Terashima, I.; Yamori, W. Overexpression of both Rubisco and Rubisco activase rescues rice photosynthesis and biomass under heat stress. Plant Cell Environ. 2021, 44, 2308–2320. [Google Scholar] [CrossRef]
  26. Kajala, K.; Covshoff, S.; Karki, S.; Woodfield, H.; Tolley, B.J.; Dionora, M.J.A.; Mogul, R.T.; Mabilangan, A.E.; Danila, F.R.; Hibberd, J.M. Strategies for engineering a two-celled C4 photosynthetic pathway into rice. J. Exp. Bot. 2011, 62, 3001–3010. [Google Scholar] [CrossRef]
  27. Amthor, J.S. Improvement of Crop Plants for Industrial End Uses; Ranalli, P., Ed.; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
  28. Hibberd, J.M.; Sheehy, J.E.; Langdale, J.A. Using C4 photosynthesis to increase the yield of rice—Rationale and feasibility. Curr. Opin. Plant Biol. 2008, 11, 228–231. [Google Scholar] [CrossRef]
  29. Melis, A. Solar energy conversion efficiencies in photosynthesis: Minimizing the chlorophyll antennae to maximize efficiency. Plant Sci. 2009, 177, 272–280. [Google Scholar] [CrossRef]
  30. Qiang, W.; Qi-De, Z.; Xin-Guang, Z.H.U.; Cong-Ming, L.U.; Ting-Yun, K.; Cheng-Quan, L.I. PSII Photochemistry and Xanthophyll Cycle in Two Superhigh yield Rice Hybrids, Liangyoupeijiu and Hua an 3 During Photoinhibition and Subsequent Restoration. J. Integr. Plant Biol. 2002, 44, 1297–1302. [Google Scholar]
  31. Evans, J.R. Improving photosynthesis. Plant Physiol. 2013, 162, 1780–1793. [Google Scholar] [CrossRef] [Green Version]
  32. Noji, T.; Kamidaki, C.; Kawakami, K.; Shen, J.-R.; Kajino, T.; Fukushima, Y.; Sekitoh, T.; Itoh, S. Photosynthetic oxygen evolution in mesoporous silica material: Adsorption of photosystem II reaction center complex into 23 nm nanopores in SBA. Langmuir 2011, 27, 705–713. [Google Scholar] [CrossRef] [PubMed]
  33. Singh, M.R.; Bell, A.T. Design of an artificial photosynthetic system for production of alcohols in high concentration from CO2. Energy Environ. Sci. 2016, 9, 193–199. [Google Scholar] [CrossRef] [Green Version]
  34. Gao, F.; Liu, C.; Qu, C.; Zheng, L.; Yang, F.; Su, M.; Hong, F. Was improvement of spinach growth by nano-TiO2 treatment related to the changes of Rubisco activase? Biometals 2008, 21, 211–217. [Google Scholar] [CrossRef] [PubMed]
  35. Thomas, A.G.; Syres, K.L. Adsorption of organic molecules on rutile TiO2 and anatase TiO2 single crystal surfaces. Chem. Soc. Rev. 2012, 41, 4207–4217. [Google Scholar] [CrossRef]
  36. Ji, Y.; Luo, Y. Direct Donation of Protons from H2O to CO2 in Artificial Photosynthesis on the Anatase TiO2 (101) Surface. J. Phys. Chem. C 2019, 123, 3019–3023. [Google Scholar] [CrossRef]
  37. Hong, F.; Yang, F.; Liu, C.; Gao, Q.; Wan, Z.; Gu, F.; Wu, C.; Ma, Z.; Zhou, J.; Yang, P. Influences of nano-TiO2 on the chloroplast aging of spinach under light. Biol. Trace Elem. Res. 2005, 104, 249–260. [Google Scholar] [CrossRef]
  38. Suganami, M.; Suzuki, Y.; Tazoe, Y.; Yamori, W.; Makino, A. Co-overproducing Rubisco and Rubisco activase enhances photosynthesis in the optimal temperature range in rice. Plant Physiol. 2021, 185, 108–119. [Google Scholar] [CrossRef]
  39. Raliya, R.; Nair, R.; Chavalmane, S.; Wang, W.-N.; Biswas, P. Mechanistic evaluation of translocation and physiological impact of titanium dioxide and zinc oxide nanoparticles on the tomato (Solanum lycopersicum L.) plant. Metallomics 2015, 7, 1584–1594. [Google Scholar] [CrossRef]
  40. Lei, Z.; Mingyu, S.; Chao, L.; Liang, C.; Hao, H.; Xiao, W.; Xiaoqing, L.; Fan, Y.; Fengqing, G.; Fashui, H. Effects of nanoanatase TiO2 on photosynthesis of spinach chloroplasts under different light illumination. Biol. Trace Elem. Res. 2007, 119, 68–76. [Google Scholar] [CrossRef]
  41. Torres, R.; Diz, V.E.; Lagorio, M.G. Effects of gold nanoparticles on the photophysical and photosynthetic parameters of leaves and chloroplasts. Photochem. Photobiol. Sci. 2018, 17, 505–516. [Google Scholar] [CrossRef]
  42. Li, X.; Sun, H.; Mao, X.; Lao, Y.; Chen, F. Enhanced photosynthesis of carotenoids in microalgae driven by light-harvesting gold nanoparticles. ACS Sustain. Chem. Eng. 2020, 8, 7600–7608. [Google Scholar] [CrossRef]
  43. Shoyhet, H.; Pavlopoulos, N.G.; Amirav, L.; Adir, N. Functionalized gold-nanoparticles enhance photosystem II driven photocurrent in a hybrid nano-bio-photoelectrochemical cell. J. Mater. Chem. A 2021, 9, 17231–17241. [Google Scholar] [CrossRef]
  44. Venzhik, Y.V.; Moshkov, I.E.; Dykman, L.A. Influence of Nanoparticles of Metals and Their Oxides on the Photosynthetic Apparatus of Plants. Biol. Bull. 2021, 48, 140–155. [Google Scholar] [CrossRef]
  45. Dias, M.C.; Santos, C.; Pinto, G.; Silva, A.M.S.; Silva, S. Titanium dioxide nanoparticles impaired both photochemical and non-photochemical phases of photosynthesis in wheat. Protoplasma 2019, 256, 69–78. [Google Scholar] [CrossRef]
  46. Vardar, F.; Yanık, F. Oxidative stress response to aluminum oxide (Al2O3) nanoparticles in Triticum aestivum. Biologia 2018, 73, 1–7. [Google Scholar]
  47. Sadak, M.S. Impact of silver nanoparticles on plant growth, some biochemical aspects, and yield of fenugreek plant (Trigonella foenum-graecum). Bull. Natl. Res. Cent. 2019, 43, 1–6. [Google Scholar] [CrossRef]
  48. Zhang, H.; Chen, S.; Jia, X.; Huang, Y.; Ji, R.; Zhao, L. Comparation of the phytotoxicity between chemically and green synthesized silver nanoparticles. Sci. Total Environ. 2021, 752, 142264. [Google Scholar] [CrossRef]
  49. Hajian, M.H.; Ghorbanpour, M.; Abtahi, F.; Hadian, J. Differential effects of biogenic and chemically synthesized silver-nanoparticles application on physiological traits, antioxidative status and californidine content in California poppy (Eschscholzia californica Cham). Environ. Pollut. 2022, 292, 118300. [Google Scholar] [CrossRef]
  50. Batool, S.U.; Javed, B.; Zehra, S.S.; Mashwani, Z.-R.; Raja, N.I.; Khan, T.; ALHaithloul, H.A.S.; Alghanem, S.M.; Al-Mushhin, A.A.M.; Hashem, M. Exogenous Applications of Bio-fabricated Silver Nanoparticles to Improve Biochemical, Antioxidant, Fatty Acid and Secondary Metabolite Contents of Sunflower. Nanomaterials 2021, 11, 1750. [Google Scholar] [CrossRef]
  51. Biba, R.; Tkalec, M.; Cvjetko, P.; Peharec Štefanić, P.; Šikić, S.; Pavoković, D.; Balen, B. Silver nanoparticles affect germination and photosynthesis in tobacco seedlings. Acta Bot. Croat. 2021, 80, 1–11. [Google Scholar] [CrossRef]
  52. Sonawane, H.; Arya, S.; Math, S.; Shelke, D. Myco-synthesized silver and titanium oxide nanoparticles as seed priming agents to promote seed germination and seedling growth of Solanum lycopersicum: A comparative study. Int. Nano Lett. 2021, 11, 371–379. [Google Scholar] [CrossRef]
  53. Mustafa, H.; Ilyas, N.; Akhtar, N.; Raja, N.I.; Zainab, T.; Shah, T.; Ahmad, A.; Ahmad, P. Biosynthesis and characterization of titanium dioxide nanoparticles and its effects along with calcium phosphate on physicochemical attributes of wheat under drought stress. Ecotoxicol. Environ. Saf. 2021, 223, 112519. [Google Scholar] [CrossRef] [PubMed]
  54. Zand, A.D.; Mikaeili Tabrizi, A.; Vaezi Heir, A. Application of titanium dioxide nanoparticles to promote phytoremediation of Cd-polluted soil: Contribution of PGPR inoculation. Bioremediat. J. 2020, 24, 171–189. [Google Scholar] [CrossRef]
  55. Pan, X.; Li, D.; Fang, Y.; Liang, Z.; Zhang, H.; Zhang, J.Z.; Lei, B.; Song, S. Enhanced photogenerated electron transfer in a semiartificial photosynthesis system based on highly dispersed titanium oxide nanoparticles. J. Phys. Chem. Lett. 2020, 11, 1822–1827. [Google Scholar] [CrossRef] [PubMed]
  56. Rai-Kalal, P.; Jajoo, A. Priming with zinc oxide nanoparticles improve germination and photosynthetic performance in wheat. Plant Physiol. Biochem. 2021, 160, 341–351. [Google Scholar] [CrossRef]
  57. Azeez, L.; Lateef, A.; Adetoro, R.O.; Adeleke, A.E. Responses of Moringa oleifera to alteration in soil properties induced by calcium nanoparticles (CaNPs) on mineral absorption, physiological indices and photosynthetic indicators. Beni-Suef Univ. J. Basic Appl. Sci. 2021, 10, 115. [Google Scholar] [CrossRef]
  58. Rizwan, M.; Ali, S.; ur Rehman, M.Z.; Malik, S.; Adrees, M.; Qayyum, M.F.; Alamri, S.A.; Alyemeni, M.N.; Ahmad, P. Effect of foliar applications of silicon and titanium dioxide nanoparticles on growth, oxidative stress, and cadmium accumulation by rice (Oryza sativa). Acta Physiol. Plant 2019, 41, 112. [Google Scholar]
  59. Li, X.; Ke, M.; Zhang, M.; Peijnenburg, W.; Fan, X.; Xu, J.; Zhang, Z.; Lu, T.; Fu, Z.; Qian, H. The interactive effects of diclofop-methyl and silver nanoparticles on Arabidopsis thaliana: Growth, photosynthesis and antioxidant system. Environ. Pollut. 2018, 232, 212–219. [Google Scholar] [CrossRef]
  60. Wang, X.; Yang, X.; Chen, S.; Li, Q.; Wang, W.; Hou, C.; Gao, X.; Wang, L.; Wang, S. Zinc oxide nanoparticles affect biomass accumulation and photosynthesis in Arabidopsis. Front. Plant Sci. 2016, 6, 1243. [Google Scholar] [CrossRef] [Green Version]
  61. Da Costa, M.V.J.; Sharma, P.K. Effect of copper oxide nanoparticles on growth, morphology, photosynthesis, and antioxidant response in Oryza sativa. Photosynthetica 2016, 54, 110–119. [Google Scholar] [CrossRef]
  62. Barhoumi, L.; Oukarroum, A.; Taher, L.B.; Smiri, L.S.; Abdelmelek, H.; Dewez, D. Effects of superparamagnetic iron oxide nanoparticles on photosynthesis and growth of the aquatic plant Lemna gibba. Arch. Environ. Contam. Toxicol. 2015, 68, 510–520. [Google Scholar] [CrossRef] [PubMed]
  63. Khan, I.; Awan, S.A.; Raza, M.A.; Rizwan, M.; Tariq, R.; Ali, S.; Huang, L. Silver nanoparticles improved the plant growth and reduced the sodium and chlorine accumulation in pearl millet: A life cycle study. Environ. Sci. Pollut. Res. 2021, 28, 13712–13724. [Google Scholar] [CrossRef] [PubMed]
  64. Jiang, M.; Dai, S.; Wang, B.; Xie, Z.; Li, J.; Wang, L.; Li, S.; Tan, Y.; Tian, B.; Shu, Q. Gold nanoparticles synthesized using melatonin suppress cadmium uptake and alleviate its toxicity in rice. Environ. Sci. Nano 2021, 8, 1042–1056. [Google Scholar] [CrossRef]
  65. Das, S.; Debnath, N.; Pradhan, S.; Goswami, A. Enhancement of photon absorption in the light-harvesting complex of isolated chloroplast in the presence of plasmonic gold nanosol—a nanobionic approach towards photosynthesis and plant primary growth augmentation. Gold Bull. 2017, 50, 247–257. [Google Scholar] [CrossRef]
  66. Wahid, I.; Rani, P.; Kumari, S.; Ahmad, R.; Hussain, S.J.; Alamri, S.; Tripathy, N.; Khan, M.I.R. Biosynthesized gold nanoparticles maintained nitrogen metabolism, nitric oxide synthesis, ions balance, and stabilizes the defense systems to improve salt stress tolerance in wheat. Chemosphere 2022, 287, 132142. [Google Scholar] [CrossRef]
  67. Dai, S.; Wang, B.; Song, Y.; Xie, Z.; Li, C.; Li, S.; Huang, Y.; Jiang, M. Astaxanthin and its gold nanoparticles mitigate cadmium toxicity in rice by inhibiting cadmium translocation and uptake. Sci. Total Environ. 2021, 786, 147496. [Google Scholar] [CrossRef]
  68. Faizan, M.; Bhat, J.A.; Chen, C.; Alyemeni, M.N.; Wijaya, L.; Ahmad, P.; Yu, F. Zinc oxide nanoparticles (ZnO-NPs) induce salt tolerance by improving the antioxidant system and photosynthetic machinery in tomato. Plant Physiol. Biochem. 2021, 161, 122–130. [Google Scholar] [CrossRef]
  69. Wu, F.; Fang, Q.; Yan, S.; Pan, L.; Tang, X.; Ye, W. Effects of zinc oxide nanoparticles on arsenic stress in rice (Oryza sativa L.): Germination, early growth, and arsenic uptake. Environ. Sci. Pollut. Res. 2020, 27, 26974–26981. [Google Scholar] [CrossRef]
  70. Salam, A.; Khan, A.R.; Liu, L.; Yang, S.; Azhar, W.; Ulhassan, Z.; Zeeshan, M.; Wu, J.; Fan, X.; Gan, Y. Seed priming with zinc oxide nanoparticles downplayed ultrastructural damage and improved photosynthetic apparatus in maize under cobalt stress. J. Hazard. Mater. 2022, 423, 127021. [Google Scholar] [CrossRef]
  71. Jabeen, N.; Maqbool, Q.; Bibi, T.; Nazar, M.; Hussain, S.Z.; Hussain, T.; Jan, T.; Ahmad, I.; Maaza, M.; Anwaar, S. Optimised synthesis of ZnO-nano-fertiliser through green chemistry: Boosted growth dynamics of economically important L. esculentum. IET Nanobiotechnol. 2018, 12, 405–411. [Google Scholar] [CrossRef]
  72. Rajput, V.D.; Minkina, T.; Fedorenko, A.; Chernikova, N.; Hassan, T.; Mandzhieva, S.; Sushkova, S.; Lysenko, V.; Soldatov, M.A.; Burachevskaya, M. Effects of zinc oxide nanoparticles on physiological and anatomical indices in spring barley tissues. Nanomaterials 2021, 11, 1722. [Google Scholar] [CrossRef] [PubMed]
  73. Panpatte, D.G.; Jhala, Y.K. Nanotechnology for Agriculture: Crop Production & Protection; Springer: Berlin/Heidelberg, Germany, 2019; ISBN 9813293748. [Google Scholar]
  74. Verma, K.K.; Song, X.-P.; Joshi, A.; Tian, D.-D.; Rajput, V.D.; Singh, M.; Arora, J.; Minkina, T.; Li, Y.-R. Recent Trends in Nano-Fertilizers for Sustainable Agriculture under Climate Change for Global Food Security. Nanomaterials 2022, 12, 173. [Google Scholar] [CrossRef] [PubMed]
  75. Anwaar, S.; Maqbool, Q.; Jabeen, N.; Nazar, M.; Abbas, F.; Nawaz, B.; Hussain, T.; Hussain, S.Z. Addendum: The Effect of Green Synthesized CuO Nanoparticles on Callogenesis and Regeneration of Oryza sativa L. Front. Plant Sci. 2016, 7, 1330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Broadley, M.; Brown, P.; Cakmak, I.; Rengel, Z.; Zhao, F. Function of nutrients: Micronutrients. In Marschner’s Mineral Nutrition of Higher Plants; Elsevier: Amsterdam, The Netherlands, 2012; pp. 191–248. [Google Scholar]
  77. Raza, A.; Tabassum, J.; Zahid, Z.; Charagh, S.; Bashir, S.; Barmukh, R.; Khan, R.S.A.; Barbosa Jr, F.; Zhang, C.; Chen, H.; et al. Advances in “Omics” Approaches for Improving Toxic Metals/Metalloids Tolerance in Plants. Front. Plant Sci. 2022, 12, 794373. [Google Scholar] [CrossRef]
  78. Snapp, S.S.; Swinton, S.M.; Labarta, R.; Mutch, D.; Black, J.R.; Leep, R.; Nyiraneza, J.; O’neil, K. Evaluating cover crops for benefits, costs and performance within cropping system niches. Agron. J. 2005, 97, 322–332. [Google Scholar] [CrossRef]
  79. Salim, N.; Raza, A. Nutrient use efficiency (NUE) for sustainable wheat production: A review. J. Plant Nutr. 2020, 43, 297–315. [Google Scholar] [CrossRef]
  80. Eriksen, A.B.; Kjeldby, M. A comparative study of urea hydrolysis rate and ammonia volatilization from urea and urea calcium nitrate. Fertil. Res. 1987, 11, 9–24. [Google Scholar] [CrossRef]
  81. Worrall, E.A.; Hamid, A.; Mody, K.T.; Mitter, N.; Pappu, H.R. Nanotechnology for plant disease management. Agronomy 2018, 8, 285. [Google Scholar] [CrossRef] [Green Version]
  82. Yang, C.; Wang, Q.; Xiang, Y.; Yuan, R.; Chai, Y. Target-induced strand release and thionine-decorated gold nanoparticle amplification labels for sensitive electrochemical aptamer-based sensing of small molecules. Sens. Actuators B Chem. 2014, 197, 149–154. [Google Scholar] [CrossRef]
  83. Taurozzi, J.S.; Hackley, V.A.; Wiesner, M.R. Ultrasonic dispersion of nanoparticles for environmental, health and safety assessment—Issues and recommendations. Nanotoxicology 2011, 5, 711–729. [Google Scholar] [CrossRef]
  84. Yoon, H.Y.; Lee, J.G.; Esposti, L.D.; Iafisco, M.; Kim, P.J.; Shin, S.G.; Jeon, J.-R.; Adamiano, A. Synergistic release of crop nutrients and stimulants from hydroxyapatite nanoparticles functionalized with humic substances: Toward a multifunctional nanofertilizer. ACS Omega 2020, 5, 6598–6610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Martins, N.C.T.; Avellan, A.; Rodrigues, S.; Salvador, D.; Rodrigues, S.M.; Trindade, T. Composites of Biopolymers and ZnO NPs for Controlled Release of Zinc in Agricultural Soils and Timed Delivery for Maize. ACS Appl. Nano Mater. 2020, 3, 2134–2148. [Google Scholar] [CrossRef]
  86. Mushtaq, A.; Jamil, N.; Rizwan, S.; Mandokhel, F.; Riaz, M.; Hornyak, G.L.; Malghani, M.N.; Shahwani, M.N. Engineered Silica Nanoparticles and silica nanoparticles containing Controlled Release Fertilizer for drought and saline areas. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Quetta, Pakistan, 2018; Volume 414, p. 12029. [Google Scholar]
  87. Shen, Y.; Zhou, J.; Du, C. Development of a polyacrylate/silica nanoparticle hybrid emulsion for delaying nutrient release in coated controlled-release urea. Coatings 2019, 9, 88. [Google Scholar] [CrossRef] [Green Version]
  88. Atta, S.; Bera, M.; Chattopadhyay, T.; Paul, A.; Ikbal, M.; Maiti, M.K.; Singh, N.D.P. Nano-pesticide formulation based on fluorescent organic photoresponsive nanoparticles: For controlled release of 2, 4-D and real time monitoring of morphological changes induced by 2, 4-D in plant systems. RSC Adv. 2015, 5, 86990–86996. [Google Scholar] [CrossRef]
  89. Ditta, A.; Arshad, M. Applications and perspectives of using nanomaterials for sustainable plant nutrition. Nanotechnol. Rev. 2016, 5, 209–229. [Google Scholar] [CrossRef] [Green Version]
  90. Dimkpa, C.O.; Bindraban, P.S. Nanofertilizers: New products for the industry? J. Agric. Food Chem. 2017, 66, 6462–6473. [Google Scholar] [CrossRef]
  91. Rajput, V.D.; Singh, A.; Minkina, T.; Rawat, S.; Mandzhieva, S.; Sushkova, S.; Shuvaeva, V.; Nazarenko, O.; Rajput, P.; Verma, K.K. Nano-Enabled Products: Challenges and Opportunities for Sustainable Agriculture. Plants 2021, 10, 2727. [Google Scholar] [CrossRef]
  92. Campos, E.V.R.; de Oliveira, J.L.; da Silva, C.M.G.; Pascoli, M.; Pasquoto, T.; Lima, R.; Abhilash, P.C.; Fraceto, L.F. Polymeric and solid lipid nanoparticles for sustained release of carbendazim and tebuconazole in agricultural applications. Sci. Rep. 2015, 5, 1–14. [Google Scholar] [CrossRef] [Green Version]
  93. Li, C.; Wang, P.; van der Ent, A.; Cheng, M.; Jiang, H.; Lund Read, T.; Lombi, E.; Tang, C.; de Jonge, M.D.; Menzies, N.W. Absorption of foliar-applied Zn in sunflower (Helianthus annuus): Importance of the cuticle, stomata and trichomes. Ann. Bot. 2019, 123, 57–68. [Google Scholar] [CrossRef] [Green Version]
  94. Wang, C.J.; Liu, Z.Q. Foliar uptake of pesticides—present status and future challenge. Pestic. Biochem. Physiol. 2007, 87, 18. [Google Scholar] [CrossRef]
  95. Fagerström, A.; Kocherbitov, V.; Westbye, P.; Bergström, K.; Mamontova, V.; Engblom, J. Characterization of a plant leaf cuticle model wax, phase behaviour of model wax—Water systems. Thermochim. Acta 2013, 571, 42–52. [Google Scholar] [CrossRef]
  96. Avellan, A.; Yun, J.; Zhang, Y.; Spielman-Sun, E.; Unrine, J.M.; Thieme, J.; Li, J.; Lombi, E.; Bland, G.; Lowry, G.V. Nanoparticle size and coating chemistry control foliar uptake pathways, translocation, and leaf-to-rhizosphere transport in wheat. ACS Nano 2019, 13, 5291–5305. [Google Scholar] [CrossRef] [PubMed]
  97. Yu, M.; Yao, J.; Liang, J.; Zeng, Z.; Cui, B.; Zhao, X.; Sun, C.; Wang, Y.; Liu, G.; Cui, H. Development of functionalized abamectin poly (lactic acid) nanoparticles with regulatable adhesion to enhance foliar retention. RSC Adv. 2017, 7, 11271–11280. [Google Scholar] [CrossRef] [Green Version]
  98. Guo, J.; Tardy, B.L.; Christofferson, A.J.; Dai, Y.; Richardson, J.J.; Zhu, W.; Hu, M.; Ju, Y.; Cui, J.; Dagastine, R.R. Modular assembly of superstructures from polyphenol-functionalized building blocks. Nat. Nanotechnol. 2016, 11, 1105–1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Yu, M.; Sun, C.; Xue, Y.; Liu, C.; Qiu, D.; Cui, B.; Zhang, Y.; Cui, H.; Zeng, Z. Tannic acid-based nanopesticides coating with highly improved foliage adhesion to enhance foliar retention. RSC Adv. 2019, 9, 27096–27104. [Google Scholar] [CrossRef] [Green Version]
  100. Xiao, D.; Liang, W.; Li, Z.; Cheng, J.; Du, Y.; Zhao, J. High foliar affinity cellulose for the preparation of efficient and safe fipronil formulation. J. Hazard. Mater. 2020, 384, 121408. [Google Scholar] [CrossRef]
  101. Tong, Y.; Shao, L.; Li, X.; Lu, J.; Sun, H.; Xiang, S.; Zhang, Z.; Wu, Y.; Wu, X. Adhesive and stimulus-responsive polydopamine-coated graphene oxide system for pesticide-loss control. J. Agric. Food Chem. 2018, 66, 2616–2622. [Google Scholar] [CrossRef]
  102. Liang, J.; Yu, M.; Guo, L.; Cui, B.; Zhao, X.; Sun, C.; Wang, Y.; Liu, G.; Cui, H.; Zeng, Z. Bioinspired development of P (St–MAA)—Avermectin nanoparticles with high affinity for foliage to enhance folia retention. J. Agric. Food Chem. 2017, 66, 6578–6584. [Google Scholar] [CrossRef]
  103. Tripathi, D.K.; Tripathi, A.; Singh, S.; Singh, Y.; Vishwakarma, K.; Yadav, G.; Sharma, S.; Singh, V.K.; Mishra, R.K.; Upadhyay, R.G. Uptake, accumulation and toxicity of silver nanoparticle in autotrophic plants, and heterotrophic microbes: A concentric review. Front. Microbiol. 2017, 8, 7. [Google Scholar] [CrossRef]
  104. Mahil, E.; Kumar, B. Foliar application of nanofertilizers in agricultural crops—A review. J. Farm Sci. 2019, 32, 239–249. [Google Scholar]
  105. Hong, J.; Wang, C.; Wagner, D.C.; Gardea-Torresdey, J.L.; He, F.; Rico, C.M. Foliar application of nanoparticles: Mechanisms of absorption, transfer, and multiple impacts. Environ. Sci. Nano 2021, 8, 1196–1210. [Google Scholar] [CrossRef]
  106. Su, Y.; Ashworth, V.; Kim, C.; Adeleye, A.S.; Rolshausen, P.; Roper, C.; White, J.; Jassby, D. Delivery, uptake, fate, and transport of engineered nanoparticles in plants: A critical review and data analysis. Environ. Sci. Nano 2019, 6, 2311–2331. [Google Scholar] [CrossRef]
  107. He, J.; Zhang, L.; He, S.Y.; Ryser, E.T.; Li, H.; Zhang, W. Stomata facilitate foliar sorption of silver nanoparticles by Arabidopsis thaliana. Environ. Pollut. 2022, 292, 118448. [Google Scholar] [CrossRef] [PubMed]
  108. Larue, C.; Castillo-Michel, H.; Sobanska, S.; Cécillon, L.; Bureau, S.; Barthès, V.; Ouerdane, L.; Carrière, M.; Sarret, G. Foliar exposure of the crop Lactuca sativa to silver nanoparticles: Evidence for internalization and changes in Ag speciation. J. Hazard. Mater. 2014, 264, 98–106. [Google Scholar] [CrossRef]
  109. Avellan, A.; Yun, J.; Morais, B.P.; Clement, E.T.; Rodrigues, S.M.; Lowry, G.V. Critical Review: Role of Inorganic Nanoparticle Properties on Their Foliar Uptake and in Planta Translocation. Environ. Sci. Technol. 2021, 55, 13417–13431. [Google Scholar] [CrossRef]
  110. Rajput, V.D.; Minkina, T.M.; Behal, A.; Sushkova, S.N.; Mandzhieva, S.; Singh, R.; Gorovtsov, A.; Tsitsuashvili, V.S.; Purvis, W.O.; Ghazaryan, K.A. Effects of zinc-oxide nanoparticles on soil, plants, animals and soil organisms: A review. Environ. Nanotechnol. Monit. Manag. 2018, 9, 76–84. [Google Scholar] [CrossRef]
  111. Rajput, V.D.; Minkina, T.; Fedorenko, A.; Mandzhieva, S.; Sushkova, S.; Lysenko, V.; Duplii, N.; Azarov, A.; Chokheli, V. Destructive effect of copper oxide nanoparticles on ultrastructure of chloroplast, plastoglobules and starch grains in spring barley (Hordeum sativum distichum). Int. J. Agric. Biol. 2018, 21, 171–174. [Google Scholar]
  112. Kumari, A.; Kumari, P.; Rajput, V.D.; Sushkova, S.N.; Minkina, T. Metal (loid) nanosorbents in restoration of polluted soils: Geochemical, ecotoxicological, and remediation perspectives. Environ. Geochem. Health 2022, 44, 235–246. [Google Scholar] [CrossRef]
  113. Mandal, D. Nanofertilizer and its application in horticulture. J. Appl. Hortic. 2021, 23, 70–77. [Google Scholar] [CrossRef]
  114. Rajput, V.D.; Minkina, T.; Sushkova, S.; Mandzhieva, S.; Fedorenko, A.; Lysenko, V.; Bederska-Błaszczyk, M.; Olchowik, J.; Tsitsuashvili, V.; Chaplygin, V. Structural and Ultrastructural Changes in Nanoparticle Exposed Plants. In Nanoscience for Sustainable Agriculture; Springer: Berlin/Heidelberg, Germany, 2019; pp. 281–295. [Google Scholar]
  115. Karny, A.; Zinger, A.; Kajal, A.; Shainsky-Roitman, J.; Schroeder, A. Therapeutic nanoparticles penetrate leaves and deliver nutrients to agricultural crops. Sci. Rep. 2018, 8, 110. [Google Scholar] [CrossRef] [Green Version]
  116. Bombo, A.B.; Pereira, A.E.S.; Lusa, M.G.; de Medeiros Oliveira, E.; de Oliveira, J.L.; Campos, E.V.R.; de Jesus, M.B.; Oliveira, H.C.; Fraceto, L.F.; Mayer, J.L.S. A mechanistic view of interactions of a nanoherbicide with target organism. J. Agric. Food Chem. 2019, 67, 4453–4462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Yan, A.; Chen, Z. Impacts of silver nanoparticles on plants: A focus on the phytotoxicity and underlying mechanism. Int. J. Mol. Sci. 2019, 20, 1003. [Google Scholar] [CrossRef] [PubMed]
  118. Rajput, V.; Minkina, T.; Sushkova, S.; Behal, A.; Maksimov, A.; Blicharska, E.; Ghazaryan, K.; Movsesyan, H.; Barsova, N. ZnO and CuO nanoparticles: A threat to soil organisms, plants, and human health. Environ. Geochem. Health 2020, 42, 147–158. [Google Scholar] [CrossRef] [PubMed]
  119. Zhang, R.; Zhang, H.; Tu, C.; Hu, X.; Li, L.; Luo, Y.; Christie, P. Phytotoxicity of ZnO nanoparticles and the released Zn (II) ion to corn (Zea mays L.) and cucumber (Cucumis sativus L.) during germination. Environ. Sci. Pollut. Res. 2015, 22, 11109–11117. [Google Scholar] [CrossRef] [PubMed]
  120. Marimón-Bolívar, W.; Tejeda-Benítez, L.; Herrera, A.P. Removal of mercury (II) from water using magnetic nanoparticles coated with amino organic ligands and yam peel biomass. Environ. Nanotechnol. Monit. Manag. 2018, 10, 486–493. [Google Scholar] [CrossRef]
  121. Wimmer, A.; Beyerl, J.; Schuster, M. Copper drinking water pipes as a previously undocumented source of silver-based nanoparticles. Environ. Sci. Technol. 2019, 53, 13293–13301. [Google Scholar] [CrossRef]
  122. Wimmer, A.; Markus, A.A.; Schuster, M. Silver nanoparticle levels in river water: Real environmental measurements and modeling approaches—a comparative study. Environ. Sci. Technol. Lett. 2019, 6, 353–358. [Google Scholar] [CrossRef]
  123. Rajput, V.D.; Minkina, T.; Sushkova, S.; Chokheli, V.; Soldatov, M. Toxicity assessment of metal oxide nanoparticles on terrestrial plants. In Comprehensive Analytical Chemistry; Elsevier: Amsterdam, The Netherlands, 2019; Volume 87, pp. 189–207. ISBN 0166-526X. [Google Scholar]
  124. Tangaa, S.R.; Selck, H.; Winther-Nielsen, M.; Khan, F.R. Trophic transfer of metal-based nanoparticles in aquatic environments: A review and recommendations for future research focus. Environ. Sci. Nano 2016, 3, 966–981. [Google Scholar] [CrossRef] [Green Version]
  125. Xin, X.; Huang, G.; Zhang, B.; Zhou, Y. Trophic transfer potential of nTiO2, nZnO, and triclosan in an algae-algae eating fish food chain. Aquat. Toxicol. 2021, 235, 105824. [Google Scholar] [CrossRef]
  126. Kim, J.I.; Park, H.G.; Chang, K.H.; Nam, D.H.; Yeo, M.K. Trophic transfer of nano-TiO2 in a paddy microcosm: A comparison of single-dose versus sequential multi-dose exposures. Environ. Pollut. 2016, 212, 316–324. [Google Scholar] [CrossRef]
Figure 1. Merits of different types of formulation of nanobionics involved in enhancing the nutrient use efficiency of agrochemicals.
Figure 1. Merits of different types of formulation of nanobionics involved in enhancing the nutrient use efficiency of agrochemicals.
Plants 11 00692 g001
Figure 2. Illustration of the use of nanobionics for soil and foliar applications to improve the crop production.
Figure 2. Illustration of the use of nanobionics for soil and foliar applications to improve the crop production.
Plants 11 00692 g002
Table 1. Application of various nanoparticles as nanobionics and their impact on photosynthesis.
Table 1. Application of various nanoparticles as nanobionics and their impact on photosynthesis.
Type of NPsSize (nm)ShapePlantsEffect on PhotosynthesisReferences
ZnO NPs (10 mg L−1) 20–30 nmsphericalTriticum aestivum L.Enhanced photosynthesis by enhancing the photosynthetic pigments chlorophyll a, chlorophyll b, and total chlorophyll content
Increased the process of water splitting complex at donor side of PS-II (Fv/Fo)
[56]
25, 50,
75, 100 and 150 μM Ag NPs and AgNO3
50 nmnanospheresNicotiana tabacum L.AgNP and AgNO3 enhanced the chlorophyll content that significantly lowered values of relative electron transport rate
Coefficient of photochemical quenching, implying an inhibition of the electron transport chain
[51]
100 mg kg−1 Ca(NO3)2 in the soil and 50, 75, and 100 mg kg−1 CaNPs in the soil80 nmparticle meshworkMoringa oleifera L.Enhanced physiological parameters of the plant
Photosynthetic ability was significantly enhanced at 50 mg CaNPs kg−1 soil
[57]
SiO2 and TiO2
NP foliar spray 0, 5, 10, 15, and 20 mg L−1
Oryza sativa L.Allowed plants to nurture in contaminated soil with Cd metals
It reduced the uptake of Cd and protects the antioxidant system
[58]
TiO2 (Aeroxide® P25: A80:R20, 21 nm)-NPs
5, 50, and 150 mg L−1
60 nm sphericalTriticum aestivum L.Reduced the chlorophyll content and hampered the efficiency of PS-II
Did not affect the RuBisCo or total sugar content (TSS)
[45]
Al2O3
NPs
5, 25, and 50 mg L−1
Suspension
~150 nm Triticum aestivum L.Caused oxidative stress and affected the pigments of the photosystem
Reduced roots and shoots in all treatments
[46]
Diclofop-methyl (an herbicide) and citrate-coated Ag NPs
Diclofop-methyl (0.5, 1.0 and 1.5 mg L−1) and Ag NPs (0.1, 0.5 and 1.0 mg L−1) on MS media
63–67 nm Arabidopsis thaliana L.Led to the oxidation of chlorophyll and the apparatuses of photosystem
It also affected the growth of the plant
Diclofop-methyl affected the Ag+ released from Ag NPs solutions
[59]
Au NPs10–14 nm Robinia pseudoacacia L.Induced the reabsorption of photoemission from PS-II, improved photosynthesis [41]
ZnO NPs
0, 50, 100, 200, 250, and 300 mgL−1
<50 nm Arabidopsis thaliana L.Zn NPs prevented the biogenesis of chlorophyll and affected PS-I thereby reducing the rate of photosynthesis [60]
CuO NP
1000 mg L−1
<50 nmsphericalOryza sativa L.Reduced the number of thylakoids in the granum Stopped the expression of key proteins of the PS-I
Higher concentration destroyed PS-II
[61]
Fe3O4, Co0.2Zn0.8Fe2O4,
and Co0.5Zn0.5Fe2O4,
0, 12.5, 25, 50, 100, 200 or 400 mg L−1
5–10 nmsphericalLemnagibba gibba L.Induced the production of ROS that destroys chlorophyll and components of PS-II, thereby completely stopping photosynthesis[62]
Table 2. Mostly studied nanoparticles and their impact on chlorophyll content, photosynthesis modulation and response to stress conditions.
Table 2. Mostly studied nanoparticles and their impact on chlorophyll content, photosynthesis modulation and response to stress conditions.
Types of NPsEffective ConcentrationSize and ShapePlantsImpact of Chlorophyll ContentReferences
Ag NPs40 mg L−1 Trigonella foenum-graecum L.Significantly enhanced chlorophyll a and b along with carotenoids, and other pigments[47]
50 mg L−1
(green Ag NP)
20 nmCucumis sativus L.Significantly enhanced total chlorophyll a and b; however, carotenoid content was decreased[48]
10, 25, and 50 mg L−1
(green Ag NP)
<90 nm, sphericalEschscholzia californica ChamChlorophyll a content by 39.7%, 35.1% and 38.7%, respectively,
25 mg L−1 bAg NPs led to a higher increase by 45.6% in total chlorophyll content over the control
Decreased by 48.3% followed at 100 mg L−1
[49]
20 mM 50–100 nmPennisetum glaucum L.(enhanced chlorophyll a by 10% and chlorophyll b by 24%) [63]
60 mg L−1
(green Ag NPs)
100 nm, sphericalHelianthus annuus L. Significantly increased chlorophyll content[50]
Au NPs200 μM Au NPs (Melatonin in the form of Au NPs)40 nm, sphericalOryza sativa L.Treated on the hydroponic system
Restored chlorophyll biosynthesis (85.1%) that was affected by Cd (51.6% decrease)
Cd uptake significantly decreased by 33.0% and 46.2%
[64]
0.1–1.0 mg mL−1
(aspartate-capped AuNPs, BSA-capped AuNPs and sodium citrate-capped AuNPs)
20–45 nm, SphericalVigna radiata L.Improved photon absorption in the light-harvesting molecular complexes
The highest concentration of sodium citrate-capped AuNPs significantly improved the total chlorophyll content
[65]
300 mg L−1
(green Au NPs)
On salt stress condition
27 nm, sphericalTriticum aestivum L.Chlorophyll contents were decreased by 49.11% in salt stress condition
77.05%, 60.16% and 72.34%, respectively, when supplemented with AuNP with NaCl
[66]
100 mg L−1 Astaxanthin in the form of Au NPs
(Ast-Au NPs)
(Cd stressed)
52.5 ± 4.3 nm, sphericalTriticum aestivum L.The presence of Cd in the culture medium diminished the total chlorophyll content; however, treatment with 100 μg/mL Ast-Au NPs restored the normalcy[67]
TiO2
NPs
25, 50 and 100 mg L−150–100 nm, sphericalSolanum lycopersicum L.A slight increase in total chlorophyll content with 25, 50 and 100 µg/mL;
however, reduced with the application of concentration 200 and 400 µg/mL
[52]
100–500 mg kg−1 (Cd-contaminated soil)15–40 nmTrifolium repens L.Slightly increased Chl a and Chl b contents
Compared to control 43.27% increase in total chlorophyll content was achieved at 250 mg kg−1 TiO2 NPs along with PGPR application
[54]
500 mg L−1 TiO2 in conjunction with 210.87 mg L−1 chloroplast (a hybrid semiartificial photosynthesis system)2.0−5.0 nm, crystallineLactuca sativa L.The highest electron-transfer rate from PS-II to PS-I during the photosynthetic process[55]
20 and 40 mg L−1 per kg of the soil
(green TiO2 NPs)
sphericalTriticum aestivum L.A dose-dependent increase in chlorophyll content in chlorophyll content[53]
ZnO NPs 10 mg L−1 of ZnO NPs by seed priming for 18 h20–30 nm, sphericalZea mays L.Compared to control Chl a had a 48% increase and Chl b had a 50% increase reported
Total Chl content was increased by 49% as compared with the control
[56]
50 mg L−1 ZnO NPs,
foliar application
(salt-stressed)
Solanum lycopersicum L.Total chlorophyll concentration was enhanced by 34% with respect to control
The rate of photosynthesis, stomatal conductance, and internal CO2 concentration was also enhanced by 34%, 32%, and 35%, respectively
[68]
20 mg L−1 ZnO NP by seed priming
(arsenic-stressed)
20–30 nm, sphericalOryza sativa L.Total chlorophyll concentration was increased by 40.1%
The higher dose (100 and 200 mg L−1) did not affect the chlorophyll concentration significantly
[69]
500 mg L−1 ZnO NPs by seed priming for 24 h
(cobalt-stressed)
20 nm, crystallineZea mays L.ZnO NPs chlorophyll content, photosynthetic efficiency and biomass accumulation
Plants were also protected from cellular ultra-cellular structure damage under Co stress
[70]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ranjan, A.; Rajput, V.D.; Kumari, A.; Mandzhieva, S.S.; Sushkova, S.; Prazdnova, E.V.; Zargar, S.M.; Raza, A.; Minkina, T.; Chung, G. Nanobionics in Crop Production: An Emerging Approach to Modulate Plant Functionalities. Plants 2022, 11, 692. https://doi.org/10.3390/plants11050692

AMA Style

Ranjan A, Rajput VD, Kumari A, Mandzhieva SS, Sushkova S, Prazdnova EV, Zargar SM, Raza A, Minkina T, Chung G. Nanobionics in Crop Production: An Emerging Approach to Modulate Plant Functionalities. Plants. 2022; 11(5):692. https://doi.org/10.3390/plants11050692

Chicago/Turabian Style

Ranjan, Anuj, Vishnu D. Rajput, Arpna Kumari, Saglara S. Mandzhieva, Svetlana Sushkova, Evgenya V. Prazdnova, Sajad Majeed Zargar, Ali Raza, Tatiana Minkina, and Gyuhwa Chung. 2022. "Nanobionics in Crop Production: An Emerging Approach to Modulate Plant Functionalities" Plants 11, no. 5: 692. https://doi.org/10.3390/plants11050692

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