Next Article in Journal
Biochar and Straw Amendments over a Decade Divergently Alter Soil Organic Carbon Accumulation Pathways
Previous Article in Journal
The Effect of Herbicides and Biostimulant Application on the Seed Yield and Seed Quality of Soybean (Glycine max (L.) Merr.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Traditional Strategies and Cutting-Edge Technologies Used for Plant Disease Management: A Comprehensive Overview

by
Hira Akhtar
1,
Muhammad Usman
2,
Rana Binyamin
1,
Akhtar Hameed
1,
Sarmad Frogh Arshad
2,
Hafiz Muhammad Usman Aslam
1,3,
Imran Ahmad Khan
4,
Manzar Abbas
5,
Haitham E. M. Zaki
6,7,
Gabrijel Ondrasek
8,* and
Muhammad Shafiq Shahid
9
1
Institute of Plant Protection, MNS University of Agriculture, Multan 66000, Pakistan
2
Department of Biochemistry and Biotechnology, MNS University of Agriculture, Multan 66000, Pakistan
3
Department of Plant Pathology, San Luis Valley Research Center, Colorado State University, Fort Collins, CO 80523, USA
4
Department of Pharmacy, MNS University of Agriculture, Multan 66000, Pakistan
5
Inner Mongolia Saikexing Institute of Breeding and Reproductive Biotechnology in Domestic Animal, Hohhot 011517, China
6
Horticulture Department, Faculty of Agriculture, Minia University, El-Minia 61517, Egypt
7
Applied Biotechnology Department, University of Technology and Applied Sciences-Sur, Sur 411, Oman
8
Faculty of Agriculture, University of Zagreb, Svetošimunska Cesta 25, 10000 Zagreb, Croatia
9
Department of Plant Sciences, College of Agricultural and Marine Sciences, Sultan Qaboos University, Al-Khoud, Muscat 123, Oman
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(9), 2175; https://doi.org/10.3390/agronomy14092175
Submission received: 28 August 2024 / Revised: 17 September 2024 / Accepted: 18 September 2024 / Published: 23 September 2024

Abstract

:
Agriculture plays a fundamental role in ensuring global food security, yet plant diseases remain a significant threat to crop production. Traditional methods to manage plant diseases have been extensively used, but they face significant drawbacks, such as environmental pollution, health risks and pathogen resistance. Similarly, biopesticides are eco-friendly, but are limited by their specificity and stability issues. This has led to the exploration of novel biotechnological approaches, such as the development of synthetic proteins, which aim to mitigate these drawbacks by offering more targeted and sustainable solutions. Similarly, recent advances in genome editing techniques—such as meganucleases (MegNs), zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs) and clustered regularly interspaced short palindromic repeats (CRISPR)—are precise approaches in disease management, but are limited by technical challenges and regulatory concerns. In this realm, nanotechnology has emerged as a promising frontier that offers novel solutions for plant disease management. This review examines the role of nanoparticles (NPs), including organic NPs, inorganic NPs, polymeric NPs and carbon NPs, in enhancing disease resistance and improving pesticide delivery, and gives an overview of the current state of nanotechnology in managing plant diseases, including its advantages, practical applications and obstacles that must be overcome to fully harness its potential. By understanding these aspects, we can better appreciate the transformative impact of nanotechnology on modern agriculture and can develop sustainable and effective strategies to mitigate plant diseases, ensuring enhanced agricultural productivity.

1. Introduction

Agriculture is the backbone of developing countries as 60% of the total population relies on it for their income [1]. In contrast to underdeveloped countries, it is an unpopular source of income in developed countries. Nowadays, it is a big challenge to fulfill the demand for food for the growing population. As predicted by the United Nations, the expected global population will be around 10 billion by 2050. This growth in population demands a 60% increase in food production to ensure a balanced food supply [2]. Therefore, researchers and scientists have worked on the issues of global food security to improve the production of agricultural products, including both crops and livestock. However, many factors—including high input expenses, drought, cross bordering of plant and animal pests and diseases, decline in soil organic matter, less land for farming, poverty, urban development, high competition of water and land, changes in regulatory landscapes, degradation of natural resources, diminished ecosystems, climate change and loss of biodiversity—affect the issue. An ongoing challenge for farmers is the incorporation of available technology to make conventional production systems more cost-effective [1,3,4,5,6]. Moreover, the contribution of animal pests to overall crop losses is 18%; microbial pathogens and weeds contribute approximately 16% and 34%, respectively. In microbial pathogens, nearly 70–80% of these crop losses are due to fungal pathogens [7]. A review of plant disease control methods is shown in Figure 1.

1.1. Conventional Agricultural Practices

Various conventional agricultural practices have been used for disease management, including field sanitation, legal methods, resistant varieties, cropping systems, soil solarization, bio fumigants, soil amendments, anaerobic soil disinfection, soil steam sterilization, soilless culture and soil fertility. Field sanitation is used to avoid pathogen spread in crop growing areas by managing humidity in the crop canopy, ensuring that it is not excessive, as most plant pathogens benefit from high humidity. Additional measures include the decontamination of equipment using heat treatment, (UV) treatment, and light treatment, as well as removing volunteer plants (such as weeds), ploughing the soil, and implementing appropriate tillage practices. These strategies help to minimize the conditions favorable to pathogen growth and reduce the risk of disease spread [8]. Additionally, legal measures have been adopted to avoid the spread of disease from infected areas to uninfected areas. Long-distance pathogen emerge in previously uninfected areas by means of plant propagation materials, packaging materials, containers and seeds [9]. It has been reported that the trading of ornamental plants was the main cause of phytophthora (P. palmivora and P. syringae, P. ramorum, P. drechsleri and P. nicotianae) in North America and Europe [10,11]. Lab testing and quarantine practices were adopted to avoid such disease incidences [9]. Moreover, the development of resistant plant varieties has also been an efficient approach used to manage plant diseases [12]. Nevertheless, genetic resistance loses its effectiveness over time because of the pressure of selection against phytopathogens, the emergence of new strains of plant pathogens that overcome that resistance or the interaction nematodes and fungi. However, soil-borne pathogens of the tomato plant, such as F. oxysporum f. sp. radicis-lycopersici and F. oxysporum f. sp. lycopersici, have been successfully controlled by using this method [13]. Multiple cropping systems—i.e., mixed cropping (the use of multiple lines in cereals), intercropping and crop rotation—were adopted to prevent pathogen inoculum, as growing the same crop in the same field year after year provides a host crop for pathogens and increases the risk of disease epidemics. In this manner, crop rotation is used to control disease, increase soil fertility and soil moisture and improve the physiochemical properties of soil [10]. Crop rotation involving barley and clover has been used to reduce Rhizoctonia canker and black scurf disease [14]. In addition, solar energy is another method for reducing soil pathogens. It involves spreading a plastic sheet on production beds just after irrigation. Solar radiation is trapped at the soil surface and heats the upper layer [13,15]. Deep tillage and exposure of the soil to solar radiation are recommended during the warmer months of the year to reduce the inoculum of soil-inhabiting plant pathogens. It is an effective approach in killing many important soil-borne pathogens, such as Sclerotinia spp., Fusarium spp., Agrobacterium tumefaciens, nematodes and Streptomyces scabies.
In addition to the above, Brassicaceae crops, such as canola, cabbage and rapeseed, have substances that can control soil-borne pathogens and pests. These crops have the ability to produce glucosinolates, which are sulfur compounds toxic to soil organisms such as P. nicotianae and R. solani, as well as to some nematodes species, such as Meloidogyne; these glucosinolates also act as biofumigants [16,17,18]. It is reported by Baysal-Gurel et al. that R. solani and P. nicotianae can be effectively controlled in woody ornamental plants by using cover crops [19]. In addition to this, soil steam sterilization is another approach that has been used in open fields or in greenhouses to sterilize the soil, ultimately inhibiting soil-borne diseases. Hot water vapors are injected into the soil using conductors and boilers; this technology is applicable for the disinfection of substrates and feedstocks in plant nurseries and greenhouses. It is a better way to control fungal pathogens than using methyl bromide or chloropicrin [20]. Although some biocidal products are effective, they damage the ozone layer and are banned in some countries. These products have more effectively controlled root knot nematode in Florida than treatment with methyl bromide [21]. The use of organic amendment is another conventional way of suppressing the growth of pathogens. Liquids and composts nourished with essential oils, organic acids, phenols and other biological compounds from herbs have also been used occasionally in managing plant diseases [22,23]. Today, the use of plant-based pesticides or phytoplaxicides has become very important for the environment and the safety of food. Organic manures, such as compost and peat moss, control soil-borne pathogens, such as Thielaviopsis basicola, Pythium, Sclerotium, R. solani, Fusarium and Phytophthora [24,25]. Along with improving soil health, organic amendments—which improve the physical properties of the soil and promote the growth of rhizogenic microorganisms—also enhance the activity of beneficial microorganisms in the soil, such as Pseudomonas, Rhizobacteria and Trichoderma. They enhance the production of plant growth regulators, phenols and tannins that have an antagonistic effect on disease-causing microbes [26].

Limitations of Conventional Agricultural Practices

These methods are effective in reducing the incidence of disease, but have a few limitations, as shown in Figure 2. These methods are ineffective on pathogens that have a broad range of host crops, and are effective in controlling only soil-borne pathogens (Rhizoctonia, Sclerotium, Fusarium, Macrophomina, etc.), which are characterized by having a wide range of hosts, normally between 200 and 500 crops [27]. They are also dependent on climatic conditions [15] and are time-consuming, fuel-consuming and laborious approaches, which make their adoption unappealing for controlling diseases [28]. Moreover, they are dependent on the physiochemical properties of the soil, the type of soil, the amount of organic amendments added, soil pH, cation exchange capacity and phytotoxicity [27]. In addition, the use of Brassica cover crops kills pathogenic microorganisms, but they also lead to increased phytotxicity and disease severity [29]. Due to their negative impacts, they are unable to support agriculture and yield production efficiently.
Seeing as there is a need to boost yields for a growing population, which will reach up to 10 billion by the end of 2050 [30], in order to address the concerns about food security, improving production is the ultimate goal, but one which cannot be achieved through conventional approaches. Global crop losses due to plant diseases are estimated to be about 16–20% of potential crop yields annually. Diseases such as wheat rust, rice blast and late blight in potatoes can cause up to 30–50% yield losses in severe cases. The economic impact of plant diseases is substantial. For example, wheat rust diseases alone have caused an estimated loss of USD 60 billion annually worldwide. The impact of fungal diseases on maize can lead to losses of up to UDS 1.5 billion per year in the U.S. alone. Therefore, to reduce pest pressure and to combat food security issues, agriculture has switched to chemical pesticides [31].

1.2. Chemical Pesticides

Chemical pesticides are target-specific and therefore have many subdivisions/classes, such as insecticides, bactericides, herbicides, nematicides and fungicides. Insecticides are used to kill insect pests, fungicides are meant to destroy fungal pathogens and herbicides are used to eradicate weeds, while bactericides are used to kill bacteria and nematicides are used to kill nematodes. Traditional synthetic pesticides are considered one of the most cost-friendly, effective and quick approaches for managing plant diseases [32,33]. They play a significant role in reducing crop losses caused by insect pests, microbes and weeds [34]. It has been reported that there was a 78% reduction in crop losses in fruit crops, 32% in cereals and 54% in vegetables due to the use of chemical pesticides [35]. As a result, a significant decrease in the hunger pattern has been observed since the mid-twentieth century. The prime objective of this pesticide use has been to improve crop production. However, apart from their profitable effects, they also hold many disadvantages [36]. The extensive use of fungicides, with their site-specific mode of action, can lead to the development of resistance in pathogens, environmental pollution, risks to human health and a decline in soil fertility by negatively impacting beneficial organisms, including predators, earthworms, and pollinators. Additionally, they can disrupt microbial diversity by altering soil conditions. However, it is possible to mitigate these effects through bioremediation, a process that involves introducing certain microbial organisms into the soil to help decontaminate and restore its natural balance. These beneficial microbes can break down harmful chemicals, improve soil health and support microbial diversity [32,33,36]. The activity of root colonizing microbes, such as bacteria, mycorrhizal fungi, Rhizogenous antagonist fungi and algae, is also affected in soil that is treated with exogenous pesticides, as these are toxic to the metabolic activity, growth and other factors of beneficial soil microbes [37]. The biochemical reactions of soil, such as nitrogen fixation, ammonification and nitrification, are also disrupted due to the activation or deactivation of specific microorganisms and enzymes [38]. Moreover, they cause retardation in soil organic mineralization, which is responsible for soil quality and its production capacity [39]. Along with their positive impacts, they also have negative impacts. For example, there is urea and ammonia present in pesticides that cause impairment of both environmental and human health. They cause metabolic issues—i.e., diabetes, infertility and endocrine disruptions, neurological disorders, compromised immune system and cancer—in humans [40,41]. Similarly, in recent years, synthetic and natural plant defense elicitors have emerged as promising alternatives or complementary strategies to chemical treatments. These elicitors, which activate plants’ innate immune responses, offer a more sustainable and targeted approach to disease control. For instance, synthetic elicitors like those described by the authors in [42] can mimic natural defense mechanisms, while novel proteins, such as those from Phytophthora parasitica [43], induce both basal immunity and systemic acquired resistance. Additionally, the intricate role of elicitor-receptor molecules in orchestrating plant defense, as discussed in [44], highlights the potential of these molecules in enhancing crop resilience to pathogens. In short, the continuous use of pesticides has had negative consequences on the environment and public health, and to some extent has contributed to a rise in disease incidence. The pros and cons of chemical pesticides are summarized in Figure 3. Therefore, as a sustainable alternative, biopesticides have become more and more popular due to their low toxicity and organic nature, and the fact that they are renewable, ecosystem-friendly and promote food safety [45].

1.3. Biopesticides

In order to reduce the increasing concerns related to chemical pesticide use, researchers have begun to utilize biopesticides [46], which involve the use of pesticides developed from living organisms (i.e., microorganisms and plants or synthesized substances derived from living organisms). Among these, botanicals or plant-based biopesticides have become more popular for the management of plant diseases without inducing toxicity to the food chain and are safer than agrochemicals, promoting safe food [47]. They are also non-toxic to the ecosystem and have an environment-friendly mode of action. The results of various studies have shown no related residual effects when applied in optimized concentrations [40]. Beneficial soil microorganisms are not harmed due to their target specificity [48]. Therefore, it is considered a sustainable pest management approach that has the ability to push agriculture towards sustainability [49]. The use of these natural pesticides is effective as they have not led to the development of resistance among pests [50]. Additionally, they have the ability to decontaminate soils through the addition of certain microbial organisms [51,52]. However, although biopesticides offer benefits, there are some obstacles that have prevented their implementation as an alternative in managing plant diseases. One of the major concerns is that high doses are required for their effectiveness under field conditions. The microorganisms (Trichoderma, Bacillus, Purpureocillium, etc.) used as biocontrol agents have the disadvantages of lack of adaptation and colonization. They should be applied preventively and repeatedly, and are not always available in commercial formulations. Plant-derived pesticides are dependent on the availability of their plant sources and their cultivation. Their formulations are difficult and should be applied at short intervals and at high doses; moreover, they have a shorter shelf life. The efficacy of microbial biopesticides is reduced by environmental factors such as temperature, UV light and desiccation. Moreover, they have issues of high costs and complexity in development [40]. In addition, many considerations are required before adopting biopesticides, including regarding the nature of the host and their ability to disperse [53]. This method has numerous resource constraints when it comes to its implementation [54]. Even today, biopesticides are generally unknown among policy makers, stakeholders and small-scale growers [55]. Due to all these limitations, this method has not been adequate for sustainable agriculture. Figure 4 offers a concise overview of the positive and negative aspects of biopesticides. Therefore, to revolutionize agricultural practices and to mitigate environmental concerns, a shift towards molecular techniques was imperative.

1.4. Molecular Techniques/Approaches

The integration of molecular techniques in crop improvement began in the latter half of the 20th century [56]. Numerous molecular techniques were used/adopted to achieve sustainable and targeted solutions for pest management, i.e., DNA barcoding and genome editing [57]. The mechanism of genome editing involves the use of sequence-specific nucleases (SSNs), which are programmable molecules that have the capacity to alter particular DNA sequences. It was reported that SSNs have been used to make targeted genome changes in multiple crops [58]. In genome editing, four principal mechanisms (meganucleases, zinc finger nucleases, transcription-like effector nucleases and clustered regularly interspaced short palindromic repeats associated protein 9 (Cas-9)) were used to perform targeted nuclease activities, which have opened the door to agricultural advancement [58].

1.4.1. Meganucleases (MegNs)

MegNs are naturally occurring endodeoxyribonucleases [59] that were discovered in the late 1980s. MegNs are members of the endonuclease family, which is capable of identifying and cleaving lengthy DNA sequences (between 20 and 40 base pairs) [60] found in a variety of microbial organisms, and can also be found in the mitochondria and chloroplast of eukaryotes [51]. In terms of molecular biology, the use of MegNs is adventitious due to their long recognition sites, high specificity, easy delivery, small size and their giving rise to more recombinant DNA by producing a 3′ overhang after DNA cleavage. In addition to this, they have the capacity to reduce the possibility of cytotoxicity [61,62]. Successful applications of MegNs have been seen in Arabidopsis, cotton and corn [63,64,65]. In an investigation, MegNs were developed to create resistance against insects and produce transgenic cotton by cleaving the specifically targeted DNA sequences [66]. It was reported that MegNs caused double strand breaks (DSBs) in embryogenic callus cells that lead to tolerance against two herbicides [67]. However, despite there being no reports on the use of MegNs in rice crops, it can potentially serve as an alternative option due to its low efficacy [63,64,65]. Additionally, the lack of naturally occurring MegNs is a major constraint, as it requires the costly, time-consuming and laborious construction of sequence specific enzymes [68]. Moreover, they can recognize a few specific DNA sequences, and there is a probability of errors due to deletion or addition at the cutting sites [63,69,70]. Figure 5 illustrates the key points of the capabilities and challenges of MegNs.

1.4.2. Zinc Finger Nucleases (ZFNs)

The era of ZFNs started in 1996, and they are recognized as site specific nucleases [71]. ZFNs are synthetic restriction enzymes that are capable of cleaving long stretches of double-stranded DNA sequences [72,73,74]. These are artificially engineered nucleases and the synthesis of their monomers involves the fusion of two domains: a Cys2-His2 zinc finger domain and a non-specific DNA cleavage domain from the DNA restriction enzyme Flavobacterium okeanokoites I (FokI) [75]. Despite their complex modular construction, ZFNs have been utilized intensively in the genetic modification of the Arabidopsis plant [73,74,76,77,78], tobacco (Nicotiana tabacum) [79,80], canola (Brassica napus), soybean (Glycine max) and maize (Zea mays) [75,81,82]. The abscisic acid (ABA) insensitive phenotype in Arabidopsis, herbicide resistance in tobacco and bialaphos resistance in maize were accomplished using ZFN technology [83,84,85]. Moreover, artificial zinc finger proteins (AZPs) have played a significant role in conferring antiviral resistance to plants by restricting the viral replication proteins’ DNA binding sites [86,87]. Chen et al. published a report employing ZFN technology to enhance disease resistance in crop plants. In this study, AZPs were designed to target a conserved sequence motif found in begomoviruses. Through this approach, the researchers achieved resistance against multiple begomoviruses—including Tobacco curly shoot virus (TbSCV) and Tomato yellow leaf curl China virus (TYLCCNV)—by specifically targeting a site within the viral DNA [88]. By preventing viral replication of the proteins’ DNA binding sites, artificial zinc finger proteins, or AZPs, have significantly increased plant resistance to viruses [86,87].
However, because of their engineering complexity and multiplexing difficulties, ZFNs have had a limited impact in inducing disease resistance in crops through the modification of genes that are associated with disease development [83,84,85]. Moreover, challenges such as off-target effects, low efficiency [89] and the existence of target sites that are sparsely distributed have made their use more challenging [90]. Figure 6 illustrates the benefits and drawbacks of ZFNs.

1.4.3. Transcription Activator-like Effector Nucleases (TALENs)

The limitations of ZFNs opened the door for a new class of nucleases: TALENs. Since the discovery of their DNA binding mechanism in 2009, it has become possible to use TALENs for DNA targeting [91,92], which are more efficient, safer and cheaper than ZFNs, and capable of targeting a specific site in the genome [93]. TALENs originate from phyto-pathogenic bacteria Xanthomonas spp. and their homologs from Ralstonia Solanacerum, and have a core DNA binding sites comprising tandem repeats made up of almost 30 to 35 identical amino acids [92,94,95]. Successful applications of this technology have been found in model plants, i.e., Arabidopsis and Brachypodium, and in some important cash crops, such as barley [96], maize [97,98] and rice. In rice, researchers successfully engineered resistance against bacterial blight disease in rice, caused by Xanthomonas oryzae, by changing the promoter region of the OsCWEET 14 gene, which is essential for susceptibility to the pathogen [99]. Similarly, targeting of the FAD2 gene has successfully improved the oil quality in soybean crops [100]. Wheat has acquired heritable resistance to powdery mildew disease after three homologs of MLO were successfully targeted for simultaneous knockout [101]. TALENs have been used to develop improved seeds that have a characteristic fragrance [102] and improved storage potential [103]. Additionally, TALENs-engineered potatoes have the characteristics suitable for cold storage and better processing qualities [104]. Despite the advancements and simplifications of TALEN methods, it remains complex for individuals unfamiliar with molecular biology studies. Moreover, compared to ZFNs, it faces certain constraints due to their larger size, which hinders delivery [105,106]. The construction of TALENs for genome editing via PCR is a difficult task due to the repetition of sequences required for their design. The requirement to create a new TALEN protein for every DNA target site increases the expense and time of their development. Since their cutting effectiveness is dependent on the target sequence, methylated DNA cannot be targeted using them [107]. A review of the pros and cons of TALENs is presented in Figure 7.

1.4.4. Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)

The CRISPR/Cas9 system, sometimes referred to as “short regularly spaced repeats”, was discovered in the 1980s and is regarded as the third generation of genome editing tools (Figure 8). In 2013, it was utilized for modifications of a plants’ genome for the first time. In contrast to the earlier genome editing technology, i.e., ZNFs and TALENs, which relied on artificial proteins, CRISPR/Cas systems recognize DNA binding sites on the basis of DNA/RNA interactions [108]. The TALEN and ZFN tools, especially ZFNs, were overshadowed by the benefits of the CRISPR/Cas system due to their simple design for targeting any genomic sites, straightforward prediction of off-target sites and the potential to modify multiple genomic sites simultaneously [109,110,111]. The utilization of CRISPR for targeted mutation in the genome of the tomato has successfully induced resistance against bacterial speck disease caused by Pseudomonas syringae pv tomato [112]. Similarly, in the case of citrus canker caused by Xanthomonas citri ssp. citri, the LATERAL ORGAN BOUNDARIES 1 (CsLOB1) that promotes the proliferation of bacterium has been identified and successfully knocked out [113]. In another investigation, the connection between CsLOB1 and the effector PthA4 of the bacterium was disrupted by using CAS9 to develop immunity in plants against the bacterium Xcc [114]. Furthermore, it was reported that the apple and many other ornamental and commercially-important plants were seriously threatened by Erwinia amylovora. To reduce the impact of this bacterium, directly purified CRISPR ribonucleoprotein was injected into the protoplast of the apple by targeting three genes (DIPM-2, DIPM-1 and DIPM-4), and researchers observed enhanced resistance against apple fire blight disease [115]. Moreover, CRISPR technology was used to induce resistance against bacterial blight disease caused by Xanthomonas oryzae pv. Oryzae via modifications in the Xa13/cdc region [112]. Similar effects were achieved in the following cases: in the MLO-7/cds region against powdery mildew of grapes caused by Erysiphe necator [115], in the OsERF922/cds region against rice blast caused by Magnaporthe oryzae [116] and in the SlMlo1/cds region against powdery mildew of tomato caused by Oidium neolycopersici [117]. Barley plants resistant to wheat dwarf virus have been successfully developed using CRISPR/Cas9 mutagenesis [118]. CRISPR has not only induced resistance in plants, but has also improved numerous other features, such as crop production [119,120], quality of grain [121,122], resistance to biotic stress [70] and abiotic stress [123] and male sterility via simple [124] and multiplex genome editing [125,126]. Yet, there are some significant constraints that prevent the Cas9 system from being beneficial in the development of disease resistance. Firstly, the direct targeting of the S gene (pathogens exploit S genes to facilitate infection and disease progression) may result in some adaptive costs due to their linkage with other desirable genes. Furthermore, any mutations in the S gene may cause a disruption in its pathways that ultimately damages a large number of other products by causing mutations in the targeted gene. These mutations have the potential to cause phenotypic abnormalities (such as chlorosis, necrosis, leaf deformation, yellowing, etc.) and the shortage of important micronutrients (Fe, Zn, Mn, etc.) [127]. In short, the CAS system is susceptible to off-target mutations [128]. The efficacy of genome editing using CRISPR is directly affected by both the sequence and the location of the target [129,130], as well as by the method of transformation (stable or transient expression methods) [131,132,133]. Moreover, it has caused difficulties/issues for the virologists while trying to control the spread of plant DNA viruses. There is a chance for endogenous genes to be inactivated through mutations or recombination, and the variations in plant virus systems due to genome editing lead to the evolution of viruses resulting from the lack of proofreading mechanisms in RNA replication. When two or more strains of an RNA virus infect the same host plant, recombination can occur, leading to the formation of new viral strains with mixed genetic material [134,135]. Although genetically modified plants developed through this technology have low viral infection rates and remain stable for up to three generations, they are subjected to regulations for genetically modified organisms and may not be accepted in some countries [136]. Figure 6 illustrates the benefits and drawbacks of CRISPR.
Therefore, the need of the hour is to develop such techniques that could increase crop production and protection with high efficacy, a lower dose and that could be environmentally friendly [137]. By keeping in view all the pros and cons of all the previously used practices and technologies (conventional approaches, chemical control, biocontrol, genome editing), researchers can continuously work to come up with emerging technologies that may have better a capacity/potential to overcome all such drawbacks.

2. Nanotechnology Used for Crop Protection

In recent years, nanotechnology has gained the attention of researchers for agricultural production [138]. “Nanotechnology” is the invention of Richard Feynman, who received the Nobel Prize in Physics in 1965. In 1959, during an American Physical Society meeting at Caltech, he introduced the idea of manipulating matter at the atomic level by presenting a paper titled “There’s Plenty of Room at the Bottom”. Nanotechnology represents the cutting edge of material sciences, comprising substances with unique properties in comparison to their larger macroscopic counterparts [139]. It is the use of tiny particles that are ranging from 1 to 100 nanometers in size, known as nanoparticles (NPs), that have a great potential to be used in medical fields, agriculture and industries due their extremely small size and large surface area [140]. For the identification of plant pathogens, it was introduced as a fourth resourceful tool for cellular and molecular biology [141].
Agriculture-based nanotechnology holds great potential for dealing with issues related to food security, including precise farming, waste management, nutrition management and disease control [142]. The earliest example of nanotechnology used in agriculture was reported in 2004 in animal breeding, crop growth and aquaculture [143,144]. In agriculture, NPs with larger specific surface areas function as unique agrochemical carriers that enable site-specific and controlled delivery of nutrients and chemicals [145], by improving their stability and solubility through nanoencapsulations [140]. Moreover, in the agro-environmental sector, numerous potential uses for this technology have been reported in protecting soil [146], improving stress tolerance in plants [147] and in the removal of contaminants [143]. The use of nanocapsules and nanodevices for disease detection and treatment is another emerging area of nanotechnology in agriculture. This includes the use of enzymatic biosensors for targeted sensing, quantum dots for fluorescent labeling to biologically recognize pathogens and in situ sensors for real-time monitoring [148,149,150]. Nanosensors serve as an effective means of identifying nutrient deficiencies, toxicity levels, animal and plant diseases and improving food quality [151]. By using this emerging technology, we can sort out farming issues effectively [152]. It has shown remarkable results for innovation in agriculture via the introduction of the newest and latest methods for disease detection, specified treatments, improved plant capacity for nutrients absorption, tackling disease and facing environmental challenges [141]. Furthermore, genetic modifications and seed treatments have been altered by agricultural nanotechnology. The coating of seeds with NPs promotes root development and early disease resistance and enhances seed germination [153]. Accurate genetic modifications have become possible through nanogenomics that optimize plants’ characteristics for increased adaptability, nutritional value and productivity [153]. In short, it can be concluded that nanotechnology holds the potential to push agriculture towards sustainability [154]. The NPs are classified as inorganic, organic, carbon based and polymeric.

2.1. Inorganic NPs

In the composition of inorganic NPs, carbon atoms are totally absent. These include metal-based and metal oxide-based NPs [155].

2.1.1. Metallic NPs Used for Plant Disease Management

The most common categories included in metal NPs are copper (Cu), cadmium (Cd), gold (Au), Aluminium (Al), Cobalt (Co), Zinc (Zn), Lead (Pb) and Silver (Ag) [156]. On the basis of their size and characteristics, they possess extraordinary properties, such as cylindrical and spherical shapes, amorphous and crystalline structures, small surface area, pore sizes and surface charge densities [157]. Moreover, their numerous remarkable properties have been elaborated on by researchers in the field of agriculture, including their prolonged storage durability, high efficacy, extremely small size and easy transportation and handling when used in an appropriate way. Consequently, these nanomaterials hold the potential to be superior to traditional agrochemicals and to be the preferred choice of farmers [158]. This review will present the overview of previous research focusing on the utilization of metallic nanoparticles (NPs) in crop protection. We summarize the findings and advancements in this field in Table 1, highlighting their effectiveness in improving crop health by minimizing biotic and abiotic stress factors.

2.1.2. Metal Oxide NPs

Small metal oxide particles have unique characteristics because they are extremely small and their surface atoms can easily combine in reactions, making them different from bigger forms [186]. The most common varieties include Aluminum oxide (Al2O3), Cerium oxide (CeO2), silver oxide (Ag2O), copper oxide (CuO), Magnesium oxide (MgO), Iron oxide (Fe2O3), Magnetite (Fe3O4), Silicon dioxide (SiO2), Titanium oxide (TiO2) and Zinc oxide (ZnO) [187]. All of these MONPs possess remarkable properties, which include effective surface qualities, high thermal and chemical stability and flexibility in pore size. These outstanding properties have gained the attention of researchers and have caused them to be considered one of the superior agents for agricultural production. Multiple MONPs have been employed to understand their long-term impact in crop production and plant growth. The results have shown significant effects in absorption, accumulation and movement of the plant, leading to an overall improvement in agricultural output. Additionally, they are highly valuable in the prevention and control of plant diseases and pests due to their increased durability, efficacy and, in particular, their large surface area, which favors their interactions with living plant cells [188]. Furthermore, they are less toxic, highly stable [189], antifungal, antibacterial [190] and antioxidant [191]. A review of the results from numerous research findings that underscore the advantage of metal oxide NPs in crop protection is shown in Table 2.

2.2. Organic NPs

Organic NPs are environment friendly and nontoxic, and are also known as nanocapsules. Organic NPs that have high sensitivity when exposed to light and heat are liposomes, dendrimers, ferritin and micelles [258]. They have potential in various fields due to their properties, such as high reactivity with target substances and susceptibility to many factors, including heat, moisture, the atmosphere and light. They have toxic properties, as well as antibacterial, antifungal and disinfection properties that increase their value in the biomedical field. Furthermore, they possess oxidation, reduction and anti-corrosion properties. They have elasticity, flexibility, ductility, tensile strength, hydrophobicity, settling tendencies, suspension behavior, and diffusion rate make the valuable in diverse fields. Moreover, they are conductors, semiconductors and resistant [187]. Numerous researchers have investigated the antimicrobial potential of organic NPs, and their findings are summarized in Table 3 and Table 4.

2.3. Carbon-Based NPs

The most significant class of NPs is carbon-based nanomaterials (CB-NMs) [271], which include fullerenes, carbon nanotubes (CNTs), graphene, graphene oxide and carbon nanofibers (CNFs). They have been extensively utilized in waste water treatment, crop protection, agriculture, antimicrobial activities, sensor technology and medicine due to distinctive features. Their uniqueness stems from their large surface area as well as electrical and optical [272,273,274,275,276,277]. thermal, mechanical and chemical properties [271]. Furthermore, CB-NMs, particularly, graphene, CNFs and CNTs, exhibit strong penetration capabilities, which enable their easy penetration through the seed coat and movement throughout the plant from roots to leaves and shoots. They serve as carriers of metal, metal oxide and agrochemicals by facilitating their easy translocation within the plant. This translocation ability of CB-NMs is observed due to their negatively charged surfaces, as well as their size [278]. Carbon-based materials have been extensively used in crop protection in multiple studies. A review of those studies and their findings is shown in Table 5, Table 6 and Table 7.

2.4. Polymeric NPs

Polymeric NPs play a substantial role as drug nanocarriers in plants, functioning either as pesticides or plant growth regulators [295]. Research conducted over the past decade has found that the most important polymeric NPs in the agricultural sector are alginate [296,297,298,299], chitosan [300,301] and zein [302,303,304]. Meanwhile, the realm of synthetic polymers PLGA (poly lactic-co-glycolic acid) has been crucial in developing new NP-based materials [295]. Researchers have focused on polymeric NPs due to their potential to control the release of active ingredients and to protect them from unfavorable environmental conditions. Their stability and ability to deliver the active ingredients more precisely and accurately to the targeted areas of plants have generated significant interest in their use within agriculture. Additionally, the biodegradable and biocompatible properties of polymeric NPs result in low ecological toxicity. Furthermore, the encapsulation of large numbers of active ingredients and their slow release minimizes the environmental impact. Therefore, they need to be combined with active compounds such as herbicides, pesticides and antibiotics [143]. Polymeric NPs in agriculture also enable the efficient delivery of drugs, enhanced adhesion in soil, uptake by plants, thermal and photostability and reduce soil leaching [305]. Their coatings as nanocarriers enhance the life span of drugs due to their outstanding properties, such as easy water dispersal and their bioavailability for hydrophobic compounds. Biodegradable and biocompatible polymers can be considered as an alternative to inorganic NPs in order to minimize the issues of ecological toxicity [295]. This review focuses on the applications of polymeric NPs in crop protection shown in Table 8, highlighting their efficacy in delivering and stabilizing active molecules.
In addition to conventional nanomaterials, recent developments in self-assembled nano-bioprotectants have shown great promise for sustainable crop protection. These nanomaterials, which can self-assemble into functional structures, offer precise delivery and the controlled release of active compounds, enhancing their efficacy and reducing their environmental impact. For instance, Ref. [306] demonstrated the high efficiency of a self-assembled, multicomponent nano-bioprotectant in managing potato late blight. Similarly, Ref. [307] reported the successful use of self-assembled nanoparticles of a prodrug conjugate based on pyrimethanil for efficient disease control. Furthermore, Ref. [308] developed a plant protein-based self-assembling core-shell nanocarrier that not only controls plant viruses effectively but also promotes plant growth and induces resistance. These advancements highlight the potential of self-assembled nanomaterials in enhancing crop protection while addressing sustainability concerns.
Nanoparticles have been increasingly utilized to improve the efficiency of both traditional and frontier plant disease control strategies. Their unique properties allow for better delivery, stability, and controlled release of agrochemicals, contributing to enhanced disease resistance. For example, the authors of [309] discuss the role of nanoparticle-mediated strategies in enhancing plant disease resistance, illustrating their ability to work synergistically with traditional methods. Additionally, advanced applications, such as the use of exosome/liposome-like nanoparticles as carriers for CRISPR-based genome editing in plants, represent a frontier strategy in disease control [310]. These nanoparticles improve the precision and effectiveness of gene editing technologies, opening new avenues for crop improvement and disease management.
Table 8. Applications of carbon NPs in plant disease management.
Table 8. Applications of carbon NPs in plant disease management.
Polymeric
NPs
NPs TypeConcentrationDisease & Pathogen ManagementReferences
ChitosanMoringa chitosan NPs200 mg/LRice blast (Magnoparthae oryzae)[311]
Chitosan in acetic acid distilled water solution4 g/LDry rot and wilt (Fusarium sp.)[312]
Chitosan biopolymer2.5 mg/mLPowdery mildew of cucumber[313,314]
Chitosan100 and 200 µg/mLBacterial wilt of potato and tomato[315]
Bioengineered chitosan iron nanocompositesIn vitro: 250 μg m/L
In vivo: 250 μg m/L
Bacterial leaf blight of rice (Xanthomonas oryzae pv. oryzae)[172]
Chitosan300 mg/L and 400 mg/LBean yellow mosaic virus of faba bean [316]
Chitosan composite film having chitosan, calcium, auxiliaries, ferulic acid and dextrin0.71–1.42 g/LSoft rot of kiwi (B. dothidea and Phomopsis sp.)[317]
Copper chitosan NPs0.10, 0.20, and 0.30 mg/mLFusarium wilt of banana (Fusarium oxysporum f. sp. cubense)[318]
Chitosan0.1–2.0 g/LRoot rot of fenugreek (Fusarium solani)[319]
Chitosan500 mg/LPowdery mildew of Rosa roxburghii[320]
Chitosan0.2 g/L and 0.4 g/LBlue mold of apple (Penicillium expansum)[321]
Chitosan/dextran NPs100 µg m/LAlfalafa mosaic virus on Nicotiana glutinosa plant[322]
Nickle chitosan nanocojugate0.04 mg/mL Fusarium rot of wheat[323]
Fluoroalkenyl-Grafted Chitosan Oligosaccharide Derivative1 mg/mLRoot knot nematode (Meloidogyne incongita)[324]
Chitosan along with botanicals (Argemone mexicana L., Achyranthes aspera L., and Ricinus communis L.)2500, 2000, 1500, 1000, and 500 ppmMeloidogyne incongita in carrot[325]
Zein NPsNatamycin-loaded zein-casein NPs (N-Z/C NPs)20 and 80 µg/mBrown rot of peach (Monilinia fructicola)[326]
Carvacrol-loaded zein NPs135 μg/mL and 270 μg/mL Bacterial canker (Pseudomonas syringae)
Fusarium wilt (Fusarium oxysporum)
[327]
Natamycin-loaded zein NPs stabilized by carboxymethyl chitosan10 mg/LPostharvest gray mold, rot and mildew of strawberry[313]
Satureja montana Essential Oil in combination with zein NPs1 mg/mLBacterial spot of tomato (Xanthomonas sp.)[314]
Rotenone loaded zein NPs16 μg m/L
48 μg m/L
Pseudomonas syringae Fusarium oxysporum[328]
PLGA NPsCTAB-PLGA Curcumin NP52.57 μg/mL and 44.67 μg/mL and 15 μg/mL Pythium ultimum var. ultimum[329]
Poly (lactic-co-glycolic acid) NPs(PLGA NPs)1.25–0.07 μg mLGray mold disease (Botrytis cinerea)[330]
Alginate NPsAlginate oligosaccharide (AOS) combined with Meyerozyma guilliermondii5 g/LBlue mold decay (Penicillium expansum)[331]
Alginate oligosaccharide (AOS)50 mg/LGray mold of kiwi fruit (Botrytis cinerea)[332]
Alginate polysaccharide1 g/LBayoud disease of date palm (Fusarium oxysporum f. sp. albedins)[333]
Alginate2 g/LVerticillium wilt of olive (Verticillium dahliae)[334]
Nano Cu-Cu2O/Alginate17.8 mg Cu/L.Rice blast (Pyricularia oryzae)[335]

2.5. Limitations of Nanotechnology

Nanotechnology is considered a groundbreaking technology that holds the potential to revolutionize numerous industries, including agriculture. Despite its promising applications, it also carries a number of possible risks and drawbacks [336]—as illustrated in Figure 9—that should be addressed. One of the major concerns is their nano-scale size, which enables them to be easily transported by air and water, thereby causing contamination. Once released in the environment, their accumulation in the air, water and soil leads to the development of ecological threats. For instance, a disruption of the balance of macro- and microorganisms leads to a decrease in soil fertility [337,338,339]. Nanopesticides can enhance the efficiency of pesticides by providing controlled release to the targeted pathogens. However, several studies have shown that only 0.1% of nanopesticides reach the intended target, while the remaining 99.9% disperse in the surrounding environment, which results in the loss of biodiversity, soil and water pollution and increased resistance among plant pests and pathogens. Studies have shown that nanopesticides are harmful to bees, which are important pollinators for the spread of pollen [36]. Additionally, their accumulation in plants and animals have adverse impacts on human health [340], leading to cardiovascular issues, respiratory diseases and neurological damage as well [341]. Direct exposure to nanopesticides through the skin facilitates their entry into systemic circulation, which causes systemic toxicity [342]. The interaction of NPs with cellular components, such as nucleic acid, lipids, proteins etc., causes inflammation, toxicity and oxidative stress [343]. As a result, reactive oxygen species (ROS) are produced in the body that damage the cellular membranes [344]. Furthermore, substantial investment is required for the research and development of nanotechnology, making it expensive. It is unaffordable for small scale growers, resulting in an unequal distribution of the benefits of nanotechnology in society [336]. Finally, last but not least, there is a knowledge gap regarding the environmental and ecosystem impacts of nanotechnology, which justifies significant investment to mitigate the potential hazards these tiny particles represent to humans and wildlife [345].

3. Conclusions

In conclusion, the integration of nanotechnology into plant disease management represents a promising pathway for sustainable agriculture. As traditional methods face numerous increasing challenges, such as pesticide resistance and environmental concerns, nanotechnology offers innovative solutions for enhanced disease resistance, controlled drug delivery and reduced pest resistance. NPs bring the promising benefit of targeted delivery to enhance the plant defense mechanism.
However, there are certain obstacles to the adoption of nanotechnology. Concerns regarding NPs toxicity, regulatory frameworks and environmental persistence demands thorough investigation and careful consideration. Collaborative efforts between researchers, policymakers and industry stakeholders are imperative to address these challenges and to ensure the safe application of nanotechnology in agriculture.
Despite these challenges, the potential benefits of nanotechnology in plant disease management are vast. By utilizing nanoscale materials and technologies, farmers can mitigate the impacts of plant disease by promoting sustainable agriculture. In addition, genetic engineering technologies also play a crucial role in developing disease-resistant crops, offering complementary and powerful tools to enhance crop resilience. Continued research, coupled with protective measures to address regulatory concerns and safety measures, will be essential in understanding the full potential of nanotechnology to revolutionize modern agriculture and ensure global food security for future generations.

Future Directions

Nanotechnology can offer vast and promising applications in plant disease management. One of its exciting domains of exploration is the development of multi-functional NPs or nanocomposites that can detect, treat and deliver drugs and nutrients simultaneously to plants to improve plant health status. Additionally, the development of advanced nanocarrier systems designed for the targeted delivery of treatments to specific plant pathogens holds great potential. This system would enable precise application, reducing the impact on non-target organisms and decreasing the reliance on chemical pesticides. Furthermore, the development of nanosensors for the real-time monitoring of plant health and early disease detection is a critical area for future research directions. Such technology would allow for timely interventions, improving management practices and potentially preventing large scale disease outbreaks. These future directions highlight the transformative potential of nanotechnology in creating more efficient, sustainable and effective plant disease management strategies.
Challenges: The challenges in genome editing include handling the off-target effects, improving delivery methods for CRISPR components, and addressing ethical issues related to safety and use. Regulatory concerns focus on ensuring the safety and effectiveness of genome-edited products and managing public perception. These concerns can differ by region and impact how genome editing technologies are developed, approved, and adopted.

Author Contributions

H.A.: Investigation, Software, Validation, Visualization, Writing—original draft; M.U.: Data curation, Investigation, Software, Visualization, Writing—original draft; R.B.: Conceptualization, Resources, Software, Validation, Writing—original draft; A.H.: Conceptualization, Validation, Visualization, Writing—original draft; S.F.A.: Conceptualization, Resources, Software, Validation, Visualization, Writing—review & editing; H.M.U.A.: Formal analysis, Investigation, Software, Validation, Writing—review & editing; I.A.K.: Data curation, Investigation, Validation, Writing—review & editing; M.A.: Data curation, Formal analysis, Investigation, Software, Validation, Writing—review & editing; H.E.M.Z.: Data curation, Formal analysis, Investigation, Validation, Writing—review & editing; G.O.: Resources, Validation, Visualization, Writing—review & editing; M.S.S.: Conceptualization, Investigation, Supervision, Visualization, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partly performed under the European Union program Next GenerationEU.

Institutional Review Board Statement

Not applicable for studies not involving humans or animals.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data created were used in this article.

Acknowledgments

The authors acknowledge Canva for providing the tools used in the creation of the figures in this manuscript. Author Zaki thanks and acknowledges the Department of Research and Consultation at the University of Technology and Applied Sciences-Sur, Oman, for their ongoing support and facilities.

Conflicts of Interest

Author Manzar Abbas was employed by the company Inner Mongolia Saikexing Institute of Breeding and Reproductive Biotechnology in Domestic Animal. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Elizabath, A.; Babychan, M.; Mathew, A.M.; Syriac, G.M. Application of nanotechnology in agriculture. Int. J. Pure Appl. Biosci. 2019, 7, 131–139. [Google Scholar] [CrossRef]
  2. Ma, W.; Hong, S.; Reed, W.R.; Duan, J.; Luu, P. Yield effects of agricultural cooperative membership in developing countries: A meta-analysis. Ann. Public Coop. Econ. 2023, 94, 761–780. [Google Scholar] [CrossRef]
  3. Asfaw, S.; Pallante, G.; Palma, A. Distributional impacts of soil erosion on agricultural productivity and welfare in Malawi. Ecol. Econ. 2020, 177, 106764. [Google Scholar] [CrossRef]
  4. Challinor, A.J.; Watson, J.; Lobell, D.B.; Howden, S.M.; Smith, D.R.; Chhetri, N. A meta-analysis of crop yield under climate change and adaptation. Nat. Clim. Chang. 2014, 4, 287–291. [Google Scholar] [CrossRef]
  5. Lachaud, M.A.; Bravo-Ureta, B.E.; Ludena, C.E. Economic effects of climate change on agricultural production and productivity in Latin America and the Caribbean (LAC). Agric. Econ. 2022, 53, 321–332. [Google Scholar] [CrossRef]
  6. Varma, V.; Bebber, D.P. Climate change impacts on banana yields around the world. Nat. Clim. Chang. 2019, 9, 752–757. [Google Scholar] [CrossRef]
  7. Moore, D.; Robson, G.D.; Trinci, A.P. 21st Century Guidebook to Fungi; Cambridge University Press: Cambridge, UK, 2020. [Google Scholar]
  8. Panth, M.; Hassler, S.C.; Baysal-Gurel, F. Methods for management of soilborne diseases in crop production. Agriculture 2020, 10, 16. [Google Scholar] [CrossRef]
  9. Crooks, J.A. Lag times and exotic species: The ecology and management of biological invasions in slow-motion1. Ecoscience 2005, 12, 316–329. [Google Scholar] [CrossRef]
  10. Goss, E.M.; Carbone, I.; Grünwald, N.J. Ancient isolation and independent evolution of the three clonal lineages of the exotic sudden oak death pathogen Phytophthora ramorum. Mol. Ecol. 2009, 18, 1161–1174. [Google Scholar] [CrossRef]
  11. Moralejo, E.P.; Pérez-Sierra, A.M.; Álvarez, L.A.; Belbahri, L.; Lefort, F.; Descals, E. Multiple alien Phytophthora taxa discovered on diseased ornamental plants in Spain. Plant Pathol. 2009, 58, 100–110. [Google Scholar] [CrossRef]
  12. Katan, J. Diseases caused by soilborne pathogens: Biology, management and challenges. J. Plant Pathol. 2017, 99, 305–315. [Google Scholar]
  13. Yadav, B.; Gurjar, M.K.; Sheshma, R.; Chopdar, R. Management Strategies for Seed and Soil Borne Diseases in Vegetable Production. Agric. Mag. 2022, 1, 35–40. [Google Scholar]
  14. Larkin, R.P.; Griffin, T.S.; Honeycutt, C.W. Rotation and cover crop effects on soilborne potato diseases, tuber yield, and soil microbial communities. Plant Dis. 2010, 94, 1491–1502. [Google Scholar] [CrossRef] [PubMed]
  15. Baysal-Gurel, F.; Gardener, B.M.; Miller, S.A. Soil Borne Disease Management in Organic Vegetable Production. 2012. Available online: https://eorganic.org/node/7581 (accessed on 27 August 2024).
  16. Baysal-Gurel, F.; Liyanapathiranage, P.; Mullican, J. Biofumigation: Opportunities and challenges for control of soilborne diseases in nursery production. Plant Health Prog. 2018, 19, 332–337. [Google Scholar] [CrossRef]
  17. Larkin, R.P.; Griffin, T.S. Control of soilborne potato diseases using Brassica green manures. Crop Prot. 2007, 26, 1067–1077. [Google Scholar] [CrossRef]
  18. Larkin, R.P.; Honeycutt, C.W. Effects of different 3-year cropping systems on soil microbial communities and Rhizoctonia diseases of potato. Phytopathology 2006, 96, 68–79. [Google Scholar] [CrossRef] [PubMed]
  19. Baysal-Gurel, F.; Liyanapathiranage, P.; Addesso, K.M. Effect of Brassica crop-based biofumigation on soilborne disease suppression in woody ornamentals. Can. J. Plant Pathol. 2020, 42, 94–106. [Google Scholar] [CrossRef]
  20. Tanaka, S.; Kobayashi, T.; Iwasaki, K.; Yamane, S.; Maeda, K.; Sakurai, K. Properties and metabolic diversity of microbial communities in soils treated with steam sterilization compared with methyl bromide and chloropicrin fumigations. Soil. Sci. Plant Nutr. 2003, 49, 603–610. [Google Scholar] [CrossRef]
  21. Kokalis-Burelle, N.; Butler, D.M.; Holzinger, J.; Rosskopf, E.N. Evaluation of steam for meloidogyne arenaria control in production of in-ground floriculture crops in florida. J. Nematol. 2016, 48, 183–192. [Google Scholar] [CrossRef]
  22. El-Sharouny, E.E. Effect of different soil amendments on the microbial count correlated with resistance of apple plants towards pathogenic Rhizoctonia solani AG-5. Biotechnol. Biotechnol. Equip. 2015, 29, 463–469. [Google Scholar] [CrossRef]
  23. Kirkegaard, J.A.; Sarwar, M.; Wong, P.T.; Mead, A.; Howe, G.; Newell, M. Field studies on the biofumigation of take-all by Brassica break crops. Aust. J. Agric. Res. 2000, 51, 445–456. [Google Scholar] [CrossRef]
  24. Bonanomi, G.; Antignani, V.; Pane, C.; Scala, F. Suppression of soilborne fungal diseases with organic amendments. J. Plant Pathol. 2007, 89, 311–324. [Google Scholar]
  25. Shafique, H.A.; Sultana, V.; Ehteshamul-Haque, S.; Athar, M. Management of soil-borne diseases of organic vegetables. J. Plant Prot. Res. 2016, 56, 221–230. [Google Scholar] [CrossRef]
  26. Welke, S.E. The effect of compost extract on the yield of strawberries and the severity of Botrytis cinerea. J. Sustain. Agric. 2005, 25, 57–68. [Google Scholar] [CrossRef]
  27. Sullivan, D.M.; Miller, R.O. Compost Quality Attributes, Measurements, and Variability. In Compost Utilization In Horticultural Cropping Systems; CRC Press: Boca Raton, FL, USA, 2001. [Google Scholar]
  28. Luvisi, A.; Panattoni, A.; Materazzi, A. Heat treatments for sustainable control of soil viruses. Agron. Sustain. Dev. 2015, 35, 657–666. [Google Scholar] [CrossRef]
  29. Kwerepe, B.; Labuschagne, N. Biofumigation and solarization as integrated pest management (IPM) components for control of root knot nematode (Meloidogyne incognita (Kofoid & White) Chitwoodi) on bambara groundnut (Vigna subterranea (L.) Verdc.). UNISWA J. Agric. 2003, 11, 56–63. [Google Scholar]
  30. Abeyratne, R. Regulation of Air Transport; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar]
  31. Sarkar, S.; Gil, J.D.; Keeley, J.; Jansen, K. The Use of Pesticides in Developing Countries and Their Impact on Health and the Right to Food; European Union: Brussels, Belgium, 2021. [Google Scholar]
  32. 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]
  33. Malandrakis, A.A.; Kavroulakis, N.; Chrysikopoulos, C.V. Use of copper, silver and zinc nanoparticles against foliar and soil-borne plant pathogens. Sci. Total Environ. 2019, 670, 292–299. [Google Scholar] [CrossRef]
  34. Aktar, W.; Sengupta, D.; Chowdhury, A. Impact of pesticides use in agriculture: Their benefits and hazards. Interdiscip. Toxicol. 2009, 2, 1–12. [Google Scholar] [CrossRef]
  35. Lamichhane, J.R. Pesticide use and risk reduction in European farming systems with IPM: An introduction to the special issue. Crop Prot. 2017, 97, 1–6. [Google Scholar] [CrossRef]
  36. Khan, B.A.; Nadeem, M.A.; Nawaz, H.; Amin, M.M.; Abbasi, G.H.; Nadeem, M.; Ali, M.; Ameen, M.; Javaid, M.M.; Maqbool, R.; et al. Pesticides: Impacts on agriculture productivity, environment, and management strategies. In Emerging Contaminants and Plants: Interactions, Adaptations and Remediation Technologies; Springer: Berlin/Heidelberg, Germany, 2023; pp. 109–134. [Google Scholar]
  37. Ahemad, M.; Khan, M. Toxicological assessment of selective pesticides towards plant growth promoting activities of phosphate solubilizing Pseudomonas aeruginosa. Acta Microbiol. Et. Immunol. Hung. 2011, 58, 169–187. [Google Scholar] [CrossRef]
  38. Hussain, J.; Ullah, R.; Rehman, N.U.; Khan, A.L.; Muhammad, Z.; Khan, F.U.; Hussain, S.T.; Anwar, S. Endogenous transitional metal and proximate analysis of selected medicinal plants from Pakistan. J. Med. Plants Res. 2010, 4, 267–270. [Google Scholar]
  39. Sebiomo, A.; Ogundero, V.W.; Bankole, S.A. Effect of four herbicides on microbial population, soil organic matter and dehydrogenase activity. Afr. J. Biotechnol. 2011, 10, 770–778. [Google Scholar]
  40. Essiedu, J.A.; Adepoju, F.O.; Ivantsova, M.N. Benefits and limitations in using biopesticides: A review. In AIP Conference Proceedings; AIP Publishing: New York, NY, USA, 2020. [Google Scholar]
  41. Thakur, N.J.P. Increased soil-microbial-eco-physiological interactions and microbial food safety in tomato under organic strategies. In Probiotics and Plant Health; Springer: Berlin/Heidelberg, Germany, 2017; pp. 215–232. [Google Scholar]
  42. Bektas, Y.; Eulgem, T. Synthetic plant defense elicitors. Front. Plant Sci. 2015, 5, 804. [Google Scholar] [CrossRef]
  43. Chang, Y.H.; Yan, H.Z.; Liou, R.F. A novel elicitor protein from P hytophthora parasitica induces plant basal immunity and systemic acquired resistance. Mol. Plant Pathol. 2015, 16, 123–136. [Google Scholar] [CrossRef] [PubMed]
  44. Abdul Malik, N.A.; Kumar, I.S.; Nadarajah, K. Elicitor and receptor molecules: Orchestrators of plant defense and immunity. Int. J. Mol. Sci. 2020, 21, 963. [Google Scholar] [CrossRef]
  45. Abdullah, S.; Zahoor, I. Biopesticides: A Green Substitute to Chemical Pesticide. Int. J. Chem. Biochem. Sci. 2023, 24, 141–156. [Google Scholar]
  46. Koul, O.; Walia, S.; Dhaliwal, G.S. Essential oils as green pesticides: Potential and constraints. Biopestic. Int. 2008, 4, 63–84. [Google Scholar]
  47. Munjanja, B.; Chaparadza, A.; Majoni, S. Biopesticide residue in foodstuffs. In Biopesticides Handbook; CRC Press: Boca Raton, FL, USA, 2015; pp. 71–92. [Google Scholar]
  48. Shiberu, E.G.T. Assessment of selected botanical extracts against Liriomyza species (Diptera: Agromyzidae) on tomato under glasshouse condition. Int. J. Fauna Biol. Stud. 2016, 3, 87–90. [Google Scholar]
  49. Nawaz, M.; Mabubu, J.I.; Hua, H. Current status and advancement of biopesticides: Microbial and botanical pesticides. J. Entomol. Zool. Stud. 2016, 4, 241–246. [Google Scholar]
  50. Tadele, S.; Emana, G. Entomopathogenic effect of Beauveria bassiana (Bals.) and Metarrhizium anisopliae (Metschn.) on Tuta absoluta (Meyrick) (Lepidoptera: Gelechiidae) larvae under laboratory and glasshouse conditions in Ethiopia. J. Plant Pathol. Microbiol. 2017, 8, 411–414. [Google Scholar]
  51. Khalil, A.M. The genome editing revolution. J. Genet. Eng. Biotechnol. 2020, 18, 68. [Google Scholar] [CrossRef] [PubMed]
  52. Javaid, M.K.; Ashiq, M.; Tahir, M. Potential of biological agents in decontamination of agricultural soil. Scientifica 2016, 2016, 1598325. [Google Scholar] [CrossRef] [PubMed]
  53. Gerson, U. Pest control by mites (Acari): Present and future. Acarologia 2014, 54, 371–394. [Google Scholar] [CrossRef]
  54. Stoneman, B. Challenges to commercialization of biopesticides. In Proceedings Microbial Biocontrol of Arthropods, Weeds and Plant Pathogens: Risks, Benefits and Challenges; National Conservation Training Center: Shepherdstown, WV, USA, 2010; pp. 123–127. [Google Scholar]
  55. Kumar, S.; Singh, A. Biopesticides: Present status and the future prospects. J. Fertil. Pestic. 2015, 6, 129. [Google Scholar] [CrossRef]
  56. Moose, S.P.; Mumm, R.H. Molecular plant breeding as the foundation for 21st century crop improvement. Plant Physiol. 2008, 147, 969–977. [Google Scholar] [CrossRef]
  57. Ramarasu, A.; Asokan, R.; Pavithra, B.S.; Sridhar, V. Innovative molecular approaches for pest management. In Genetic Methods and Tools for Managing Crop Pests; Springer: Berlin/Heidelberg, Germany, 2022; pp. 27–43. [Google Scholar]
  58. Nekrasov, V.; Staskawicz, B.; Weigel, D.; Jones, J.D.; Kamoun, S. Targeted mutagenesis in the model plant Nicotiana benthamiana using Cas9 RNA-guided endonuclease. Nat. Biotechnol. 2013, 31, 691–693. [Google Scholar] [CrossRef]
  59. Stoddard, B.L. Homing endonucleases: From microbial genetic invaders to reagents for targeted DNA modification. Structure 2011, 19, 7–15. [Google Scholar] [CrossRef]
  60. Gallagher, R.R.; Li, Z.; Lewis, A.O.; Isaacs, F.J. Rapid editing and evolution of bacterial genomes using libraries of synthetic DNA. Nat. Protoc. 2014, 9, 2301–2316. [Google Scholar] [CrossRef]
  61. Mandip, K.; Steer, C.J. A new era of gene editing for the treatment of human diseases. Swiss Med. Wkly. 2019, 149, w20021. [Google Scholar]
  62. Devarajan, A. Optically Controlled CRISPR-Cas9 and Cre Recombinase for Spatiotemporal Gene Editing: A Review. ACS Synth. Biol. 2023, 13, 25–44. [Google Scholar] [PubMed]
  63. Daboussi, F.; Stoddard, T.J.; Zhang, F. Engineering meganuclease for precise plant genome modification. In Advances in New Technology for Targeted Modification of Plant Genomes; Springer: Berlin/Heidelberg, Germany, 2015; pp. 21–38. [Google Scholar]
  64. Viana, V.E.; Pegoraro, C.; Busanello, C.; Costa de Oliveira, A. Mutagenesis in rice: The basis for breeding a new super plant. Front. Plant Sci. 2019, 10, 1326. [Google Scholar]
  65. Zhu, C.; Bortesi, L.; Baysal, C.; Twyman, R.M.; Fischer, R.; Capell, T.; Schillberg, S.; Christou, P. Characteristics of genome editing mutations in cereal crops. Trends Plant Sci. 2017, 22, 38–52. [Google Scholar] [PubMed]
  66. Singh, A.; Srivastava, A.; Saidulu, D.; Gupta, A.K. Advancements of sequencing batch reactor for industrial wastewater treatment: Major focus on modifications, critical operational parameters, and future perspectives. J. Environ. Manag. 2022, 317, 115305. [Google Scholar] [CrossRef] [PubMed]
  67. D’Halluin, K.; Vanderstraeten, C.; Van Hulle, J.; Rosolowska, J.; Van Den Brande, I.; Pennewaert, A.; D’Hont, K.; Bossut, M.; Jantz, D.; Ruiter, R.; et al. Targeted molecular trait stacking in cotton through targeted double-strand break induction. Plant Biotechnol. J. 2013, 11, 933–941. [Google Scholar]
  68. Prieto, J.; Redondo, P.; Padro, D.; Arnould, S.; Epinat, J.C.; Pâques, F.; Blanco, F.J.; Montoya, G. The C-terminal loop of the homing endonuclease I-CreI is essential for site recognition, DNA binding and cleavage. Nucleic Acids Res. 2007, 35, 3262–3271. [Google Scholar]
  69. Majid, A.; Parray, G.A.; Wani, S.H.; Kordostami, M.; Sofi, N.R.; Waza, S.A.; Shikari, A.B.; Gulzar, S. Genome editing and its necessity in agriculture. Int. J. Curr. Microbiol. Appl. Sci. 2017, 6, 5435–5443. [Google Scholar] [CrossRef]
  70. Yin, K.; Qiu, J.-L. Genome editing for plant disease resistance: Applications and perspectives. Philos. Trans. R. Soc. B 2019, 374, 20180322. [Google Scholar] [CrossRef]
  71. Kim, Y.G.; Cha, J.; Chandrasegaran, S. Hybrid restriction enzymes: Zinc finger fusions to Fok I cleavage domain. Proc. Natl. Acad. Sci. USA 1996, 93, 1156–1160. [Google Scholar] [CrossRef]
  72. Carroll, D. Genome engineering with zinc-finger nucleases. Genetics 2011, 188, 773–782. [Google Scholar] [CrossRef]
  73. Osakabe, K.; Osakabe, Y.; Toki, S. Site-directed mutagenesis in Arabidopsis using custom-designed zinc finger nucleases. Proc. Natl. Acad. Sci. USA 2010, 107, 12034–12039. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, X.; Zhang, J.; Zhu, K.Y. Chitosan/double-stranded RNA nanoparticle-mediated RNA interference to silence chitin synthase genes through larval feeding in the African malaria mosquito (Anopheles gambiae). Insect Mol. Biol. 2010, 19, 683–693. [Google Scholar] [CrossRef] [PubMed]
  75. Curtin, S.J.; Zhang, F.; Sander, J.D.; Haun, W.J.; Starker, C.; Baltes, N.J.; Reyon, D.; Dahlborg, E.J.; Goodwin, M.J.; Coffman, A.P.; et al. Targeted mutagenesis of duplicated genes in soybean with zinc-finger nucleases. Plant Physiol. 2011, 156, 466–473. [Google Scholar] [CrossRef] [PubMed]
  76. Even-Faitelson, L.; Samach, A.; Melamed-Bessudo, C.; Avivi-Ragolsky, N.; Levy, A.A. Localized egg-cell expression of effector proteins for targeted modification of the Arabidopsis genome. Plant J. 2011, 68, 929–937. [Google Scholar] [CrossRef] [PubMed]
  77. Petolino, J.F.; Worden, A.; Curlee, K.; Connell, J.; Strange Moynahan, T.L.; Larsen, C.; Russell, S. Zinc finger nuclease-mediated transgene deletion. Plant Mol. Biol. 2010, 73, 617–628. [Google Scholar] [CrossRef]
  78. Qi, Y.; Li, X.; Zhang, Y.; Starker, C.G.; Baltes, N.J.; Zhang, F.; Sander, J.D.; Reyon, D.; Joung, J.K.; Voytas, D.F. Targeted deletion and inversion of tandemly arrayed genes in Arabidopsis thaliana using zinc finger nucleases. G3 Genes Genomes Genet. 2013, 3, 1707–1715. [Google Scholar] [CrossRef]
  79. Townsend, J.A.; Wright, D.A.; Winfrey, R.J.; Fu, F.; Maeder, M.L.; Joung, J.K.; Voytas, D.F. High-frequency modification of plant genes using engineered zinc-finger nucleases. Nature 2009, 459, 442–445. [Google Scholar] [CrossRef]
  80. Wright, D.A.; Townsend, J.A.; Winfrey, R.J., Jr.; Irwin, P.A.; Rajagopal, J.; Lonosky, P.M.; Hall, B.D.; Jondle, M.D.; Voytas, D.F. High-frequency homologous recombination in plants mediated by zinc-finger nucleases. Plant J. 2005, 44, 693–705. [Google Scholar] [CrossRef]
  81. Ainley, W.M.; Sastry-Dent, L.; Welter, M.E.; Murray, M.G.; Zeitler, B.; Amora, R.; Corbin, D.R.; Miles, R.R.; Arnold, N.L.; Strange, T.L.; et al. Trait stacking via targeted genome editing. Plant Biotechnol. J. 2013, 11, 1126–1134. [Google Scholar] [CrossRef]
  82. Shukla, V.K.; Doyon, Y.; Miller, J.C.; DeKelver, R.C.; Moehle, E.A.; Worden, S.E.; Mitchell, J.C.; Arnold, N.L.; Gopalan, S.; Meng, X.; et al. Precise genome modification in the crop species Zea mays using zinc-finger nucleases. Nature 2009, 459, 437–441. [Google Scholar] [CrossRef]
  83. de Galarreta, M.R.; Lujambio, A. Therapeutic editing of hepatocyte genome in vivo. J. Hepatol. 2017, 67, 818–828. [Google Scholar] [CrossRef] [PubMed]
  84. Jaganathan, D.; Ramasamy, K.; Sellamuthu, G.; Jayabalan, S.; Venkataraman, G. CRISPR for crop improvement: An update review. Front. Plant Sci. 2018, 9, 985. [Google Scholar] [CrossRef] [PubMed]
  85. Khandagale, K.; Nadaf, A. Genome editing for targeted improvement of plants. Plant Biotechnol. Rep. 2016, 10, 327–343. [Google Scholar] [CrossRef]
  86. Sera, T. Inhibition of virus DNA replication by artificial zinc finger proteins. J. Virol. 2005, 79, 2614–2619. [Google Scholar] [CrossRef]
  87. Takenaka, K.; Koshino-Kimura, Y.; Aoyama, Y.; Sera, T. Inhibition of tomato yellow leaf curl virus replication by artificial zinc-finger proteins. Nucleic Acids Symp. Ser. 2007, 429–430. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, W.; Qian, Y.; Wu, X.; Sun, Y.; Wu, X.; Cheng, X. Inhibiting replication of begomoviruses using artificial zinc finger nucleases that target viral-conserved nucleotide motif. Virus Genes 2014, 48, 494–501. [Google Scholar] [CrossRef]
  89. Nemudryi, A.A.; Valetdinova, K.R.; Medvedev, S.P.; Zakian, S.Á. TALEN and CRISPR/Cas genome editing systems: Tools of discovery. Acta Naturae (Англoязычная Версия) 2014, 6, 19–40. [Google Scholar] [CrossRef] [PubMed]
  90. Maeder, M.L.; Thibodeau-Beganny, S.; Osiak, A.; Wright, D.A.; Anthony, R.M.; Eichtinger, M.; Jiang, T.; Foley, J.E.; Winfrey, R.J.; Townsend, J.A.; et al. Rapid “open-source” engineering of customized zinc-finger nucleases for highly efficient gene modification. Mol. Cell 2008, 31, 294–301. [Google Scholar] [CrossRef]
  91. Boch, J.; Scholze, H.; Schornack, S.; Landgraf, A.; Hahn, S.; Kay, S.; Lahaye, T.; Nickstadt, A.; Bonas, U. Breaking the code of DNA binding specificity of TAL-type III effectors. Science 2009, 326, 1509–1512. [Google Scholar] [CrossRef]
  92. Moscou, M.J.; Bogdanove, A.J. A simple cipher governs DNA recognition by TAL effectors. Science 2009, 326, 1501. [Google Scholar] [CrossRef]
  93. Christian, M.L.; Demorest, Z.L.; Starker, C.G.; Osborn, M.J.; Nyquist, M.D.; Zhang, Y.; Carlson, D.F.; Bradley, P.; Bogdanove, A.J.; Voytas, D.F. Targeting G with TAL effectors: A comparison of activities of TALENs constructed with NN and NK repeat variable di-residues. PLoS ONE 2012, 7, e45383. [Google Scholar] [CrossRef] [PubMed]
  94. de Lange, O.; Schreiber, T.; Schandry, N.; Radeck, J.; Braun, K.H.; Koszinowski, J.; Heuer, H.; Strauß, A.; Lahaye, T. Breaking the DNA-binding code of Ralstonia solanacearum TAL effectors provides new possibilities to generate plant resistance genes against bacterial wilt disease. New Phytol. 2013, 199, 773–786. [Google Scholar] [CrossRef] [PubMed]
  95. Jankele, R.; Svoboda, P. TAL effectors: Tools for DNA targeting. Brief. Funct. Genom. 2014, 13, 409–419. [Google Scholar] [CrossRef] [PubMed]
  96. Wendt, T.; Holm, P.B.; Starker, C.G.; Christian, M.; Voytas, D.F.; Brinch-Pedersen, H.; Holme, I.B. TAL effector nucleases induce mutations at a pre-selected location in the genome of primary barley transformants. Plant Mol. Biol. 2013, 83, 279–285. [Google Scholar] [CrossRef]
  97. Char, S.N.; Unger-Wallace, E.; Frame, B.; Briggs, S.A.; Main, M.; Spalding, M.H.; Vollbrecht, E.; Wang, K.; Yang, B. Heritable site-specific mutagenesis using TALEN s in maize. Plant Biotechnol. J. 2015, 13, 1002–1010. [Google Scholar] [CrossRef]
  98. Liang, Z.; Zhang, K.; Chen, K.; Gao, C. Targeted mutagenesis in Zea mays using TALENs and the CRISPR/Cas system. J. Genet. Genom. 2014, 41, 63–68. [Google Scholar] [CrossRef]
  99. Li, T.; Liu, B.; Spalding, M.H.; Weeks, D.P.; Yang, B. High-efficiency TALEN-based gene editing produces disease-resistant rice. Nat. Biotechnol. 2012, 30, 390–392. [Google Scholar] [CrossRef]
  100. Haun, W.; Coffman, A.; Clasen, B.M.; Demorest, Z.L.; Lowy, A.; Ray, E.; Retterath, A.; Stoddard, T.; Juillerat, A.; Cedrone, F.; et al. Improved soybean oil quality by targeted mutagenesis of the fatty acid desaturase 2 gene family. Plant Biotechnol. J. 2014, 12, 934–940. [Google Scholar] [CrossRef]
  101. Wang, Y.; Cheng, X.; Shan, Q.; Zhang, Y.; Liu, J.; Gao, C.; Qiu, J.L. Simultaneous editing of three homoeoalleles in hexaploid bread wheat confers heritable resistance to powdery mildew. Nat. Biotechnol. 2014, 32, 947–951. [Google Scholar] [CrossRef]
  102. Shan, Q.; Zhang, Y.; Chen, K.; Zhang, K.; Gao, C. Creation of fragrant rice by targeted knockout of the Os BADH 2 gene using TALEN technology. Plant Biotechnol. J. 2015, 13, 791–800. [Google Scholar] [CrossRef]
  103. Ma, L.; Zhu, F.; Li, Z.; Zhang, J.; Li, X.; Dong, J.; Wang, T. TALEN-based mutagenesis of lipoxygenase LOX3 enhances the storage tolerance of rice (Oryza sativa) seeds. PLoS ONE 2015, 10, e0143877. [Google Scholar] [CrossRef]
  104. Clasen, B.M.; Stoddard, T.J.; Luo, S.; Demorest, Z.L.; Li, J.; Cedrone, F.; Tibebu, R.; Davison, S.; Ray, E.E.; Daulhac, A.; et al. Improving cold storage and processing traits in potato through targeted gene knockout. Plant Biotechnol. J. 2016, 14, 169–176. [Google Scholar] [CrossRef] [PubMed]
  105. Gaj, T.; Sirk, S.J.; Shui, S.L.; Liu, J. Genome-editing technologies: Principles and applications. Cold Spring Harb. Perspect. Biol. 2016, 8, a023754. [Google Scholar] [CrossRef]
  106. Khan, S.H. Genome-editing technologies: Concept; pros, and cons of various genome-editing techniques and bioethical concerns for clinical application. Mol. Ther. Nucleic Acids 2019, 16, 326–334. [Google Scholar] [CrossRef] [PubMed]
  107. Malzahn, A.; Lowder, L.; Qi, Y. Plant genome editing with TALEN and CRISPR. Cell Biosci. 2017, 7, 21. [Google Scholar] [CrossRef]
  108. Huo, Z.; Tu, J.; Xu, A.; Li, Y.; Wang, D.; Liu, M.; Zhou, R.; Zhu, D.; Lin, Y.; Gingold, J.A.; et al. Generation of a heterozygous p53 R249S mutant human embryonic stem cell line by TALEN-mediated genome editing. Stem Cell Res. 2019, 34, 101360. [Google Scholar] [CrossRef]
  109. Hille, F.; Richter, H.; Wong, S.P.; Bratovič, M.; Ressel, S.; Charpentier, E. The biology of CRISPR-Cas: Backward and forward. Cell 2018, 172, 1239–1259. [Google Scholar] [CrossRef]
  110. Koonin, E.V.; Makarova, K.S.; Zhang, F. Diversity, classification and evolution of CRISPR-Cas systems. Curr. Opin. Microbiol. 2017, 37, 67–78. [Google Scholar] [CrossRef] [PubMed]
  111. Lino, C.A.; Harper, J.C.; Carney, J.P.; Timlin, J.A. Delivering CRISPR: A review of the challenges and approaches. Drug Deliv. 2018, 25, 1234–1257. [Google Scholar] [CrossRef]
  112. Ortigosa, A.; Gimenez-Ibanez, S.; Leonhardt, N.; Solano, R. Design of a Bacterial Speck Resistant Tomato by CRISPR /Cas9-mediated Editing of Sl JAZ 2. Plant Biotechnol. J. 2019, 17, 665–673. [Google Scholar] [CrossRef]
  113. Hu, Y.; Zhang, J.; Jia, H.; Sosso, D.; Li, T.; Frommer, W.B.; Yang, B.; White, F.F.; Wang, N.; Jones, J.B. Lateral organ boundaries 1 is a disease susceptibility gene for citrus bacterial canker disease. Proc. Natl. Acad. Sci. USA 2014, 111, E521–E529. [Google Scholar] [CrossRef] [PubMed]
  114. Peng, A.; Chen, S.; Lei, T.; Xu, L.; He, Y.; Wu, L.; Yao, L.; Zou, X. Engineering canker-resistant plants through CRISPR/Cas9-targeted editing of the susceptibility gene Cs LOB 1 promoter in citrus. Plant Biotechnol. J. 2017, 15, 1509–1519. [Google Scholar] [CrossRef] [PubMed]
  115. Malnoy, M.; Viola, R.; Jung, M.H.; Koo, O.J.; Kim, S.; Kim, J.S.; Velasco, R.; Nagamangala Kanchiswamy, C. DNA-free genetically edited grapevine and apple protoplast using CRISPR/Cas9 ribonucleoproteins. Front. Plant Sci. 2016, 7, 1904. [Google Scholar] [CrossRef] [PubMed]
  116. Wang, F.; Wang, C.; Liu, P.; Lei, C.; Hao, W.; Gao, Y.; Liu, Y.G.; Zhao, K. Enhanced rice blast resistance by CRISPR/Cas9-targeted mutagenesis of the ERF transcription factor gene OsERF922. PLoS ONE 2016, 11, e0154027. [Google Scholar] [CrossRef] [PubMed]
  117. Nekrasov, V.; Wang, C.; Win, J.; Lanz, C.; Weigel, D.; Kamoun, S. Rapid generation of a transgene-free powdery mildew resistant tomato by genome deletion. Sci. Rep. 2017, 7, 482. [Google Scholar] [CrossRef]
  118. Kis, A.; Hamar, É.; Tholt, G.; Bán, R.; Havelda, Z. Creating highly efficient resistance against wheat dwarf virus in barley by employing CRISPR/Cas9 system. Plant Biotechnol. J. 2019, 17, 1004. [Google Scholar] [CrossRef]
  119. Bao, A.; Burritt, D.J.; Chen, H.; Zhou, X.; Cao, D.; Tran, L.S. The CRISPR/Cas9 system and its applications in crop genome editing. Crit. Rev. Biotechnol. 2019, 39, 321–336. [Google Scholar] [CrossRef]
  120. Hussain, B.; Lucas, S.J.; Budak, H. CRISPR/Cas9 in plants: At play in the genome and at work for crop improvement. Brief. Funct. Genom. 2018, 17, 319–328. [Google Scholar] [CrossRef]
  121. Chao, S.; Cai, Y.; Feng, B.; Jiao, G.; Sheng, Z.; Luo, J.; Tang, S.; Wang, J.; Hu, P.; Wei, X. Editing of rice isoamylase gene ISA1 provides insights into its function in starch formation. Rice Sci. 2019, 26, 77–87. [Google Scholar]
  122. Fiaz, S.; Ahmad, S.; Noor, M.A.; Wang, X.; Younas, A.; Riaz, A.; Riaz, A.; Ali, F. Applications of the CRISPR/Cas9 system for rice grain quality improvement: Perspectives and opportunities. Int. J. Mol. Sci. 2019, 20, 888. [Google Scholar] [CrossRef]
  123. Zafar, S.A.; Zaidi, S.S.; Gaba, Y.; Singla-Pareek, S.L.; Dhankher, O.P.; Li, X.; Mansoor, S.; Pareek, A. Engineering abiotic stress tolerance via CRISPR/Cas-mediated genome editing. J. Exp. Bot. 2020, 71, 470–479. [Google Scholar] [CrossRef] [PubMed]
  124. Barman, H.N.; Sheng, Z.; Fiaz, S.; Zhong, M.; Wu, Y.; Cai, Y.; Wang, W.; Jiao, G.; Tang, S.; Wei, X.; et al. Generation of a new thermo-sensitive genic male sterile rice line by targeted mutagenesis of TMS5 gene through CRISPR/Cas9 system. BMC Plant Biol. 2019, 19, 109. [Google Scholar] [CrossRef] [PubMed]
  125. Li, S.; Shen, L.; Hu, P.; Liu, Q.; Zhu, X.; Qian, Q.; Wang, K.; Wang, Y. Developing disease-resistant thermosensitive male sterile rice by multiplex gene editing. J. Integr. Plant Biol. 2019, 61, 1201–1205. [Google Scholar] [CrossRef] [PubMed]
  126. Shen, L.; Hua, Y.; Fu, Y.; Li, J.; Liu, Q.; Jiao, X.; Xin, G.; Wang, J.; Wang, X.; Yan, C.; et al. Rapid generation of genetic diversity by multiplex CRISPR/Cas9 genome editing in rice. Sci. China Life Sci. 2017, 60, 506–515. [Google Scholar] [CrossRef] [PubMed]
  127. van Schie, C.C.; Takken, F.L. Susceptibility genes 101: How to be a good host. Annu. Rev. Phytopathol. 2014, 52, 551–581. [Google Scholar] [CrossRef]
  128. Langner, T.; Kamoun, S.; Belhaj, K. CRISPR crops: Plant genome editing toward disease resistance. Annu. Rev. Phytopathol. 2018, 56, 479–512. [Google Scholar] [CrossRef]
  129. Slaymaker, I.M.; Gao, L.; Zetsche, B.; Scott, D.A.; Yan, W.X.; Zhang, F. Rationally engineered Cas9 nucleases with improved specificity. Science 2016, 351, 84–88. [Google Scholar] [CrossRef]
  130. Tycko, J.; Myer, V.E.; Hsu, P.D. Methods for optimizing CRISPR-Cas9 genome editing specificity. Mol. Cell 2016, 63, 355–370. [Google Scholar] [CrossRef]
  131. Scheben, A.; Wolter, F.; Batley, J.; Puchta, H.; Edwards, D. Towards CRISPR/Cas crops–bringing together genomics and genome editing. New Phytol. 2017, 216, 682–698. [Google Scholar] [CrossRef]
  132. Tsai, S.Q.; Joung, J.K. Defining and improving the genome-wide specificities of CRISPR–Cas9 nucleases. Nat. Rev. Genet. 2016, 17, 300–312. [Google Scholar] [CrossRef]
  133. Yin, K.; Gao, C.; Qiu, J.L. Progress and prospects in plant genome editing. Nat. Plants 2017, 3, 17107. [Google Scholar] [CrossRef] [PubMed]
  134. Mehta, D.; Stürchler, A.; Anjanappa, R.B.; Zaidi, S.S.; Hirsch-Hoffmann, M.; Gruissem, W.; Vanderschuren, H. Linking CRISPR-Cas9 interference in cassava to the evolution of editing-resistant geminiviruses. Genome Biol. 2019, 20, 80. [Google Scholar] [CrossRef] [PubMed]
  135. Cao, Y.; Zhou, H.; Zhou, X.; Li, F. Control of plant viruses by CRISPR/Cas system-mediated adaptive immunity. Front. Microbiol. 2020, 11, 593700. [Google Scholar] [CrossRef] [PubMed]
  136. Tyagi, S.; Kumar, R.; Kumar, V.; Won, S.Y.; Shukla, P. Engineering disease resistant plants through CRISPR-Cas9 technology. GM Crop. Food 2021, 12, 125–144. [Google Scholar] [CrossRef]
  137. Delgado, J.A.; Short Jr, N.M.; Roberts, D.P.; Vandenberg, B. Big data analysis for sustainable agriculture on a geospatial cloud framework. Front. Sustain. Food Syst. 2019, 3, 54. [Google Scholar] [CrossRef]
  138. Parisi, C.; Vigani, M.; Rodríguez-Cerezo, E. Agricultural nanotechnologies: What are the current possibilities? Nano Today 2015, 10, 124–127. [Google Scholar] [CrossRef]
  139. Hulla, J.E.; Sahu, S.C.; Hayes, A.W. Nanotechnology: History and future. Hum. Exp. Toxicol. 2015, 34, 1318–1321. [Google Scholar] [CrossRef]
  140. Ghidan, A.Y.; Al-Antary, T.M.; Salem, N.M.; Awwad, A.M. Facile green synthetic route to the zinc oxide (ZnONPs) nanoparticles: Effect on green peach aphid and antibacterial activity. J. Agric. Sci. 2017, 9, 131–138. [Google Scholar] [CrossRef]
  141. Sharon, M.; Choudhary, A.K.; Kumar, R. Nanotechnology in agricultural diseases and food safety. J. Phytol. 2010, 2, 83–92. [Google Scholar]
  142. Guleria, G.; Thakur, S.; Shandilya, M.; Sharma, S.; Thakur, S.; Kalia, S. Nanotechnology for sustainable agro-food systems: The need and role of nanoparticles in protecting plants and improving crop productivity. Plant Physiol. Biochem. 2023, 194, 533–549. [Google Scholar] [CrossRef]
  143. Khan, S.; Naushad, M.; Al-Gheethi, A.; Iqbal, J. Engineered nanoparticles for removal of pollutants from wastewater: Current status and future prospects of nanotechnology for remediation strategies. J. Environ. Chem. Eng. 2021, 9, 106160. [Google Scholar] [CrossRef]
  144. Zhang, P.; Lynch, I.; Handy, R.D.; White, J.C. A brief history of nanotechnology in agriculture and current status. In Nano-Enabled Sustainable and Precision Agriculture; Elsevier: Amsterdam, The Netherlands, 2023; pp. 3–14. [Google Scholar]
  145. Shang, Y.; Hasan, M.K.; Ahammed, G.J.; Li, M.; Yin, H.; Zhou, J. Applications of nanotechnology in plant growth and crop protection: A review. Molecules 2019, 24, 2558. [Google Scholar] [CrossRef]
  146. Liu, W.; Li, Y.; Feng, Y.; Qiao, J.; Zhao, H.; Xie, J.; Fang, Y.; Shen, S.; Liang, S. The effectiveness of nanobiochar for reducing phytotoxicity and improving soil remediation in cadmium-contaminated soil. Sci. Rep. 2020, 10, 858. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, P.; Guo, Z.; Ullah, S.; Melagraki, G.; Afantitis, A.; Lynch, I. Nanotechnology and artificial intelligence to enable sustainable and precision agriculture. Nat. Plants 2021, 7, 864–876. [Google Scholar] [CrossRef] [PubMed]
  148. Daniel, A.; Laës-Huon, A.; Barus, C.; Beaton, A.D.; Blandfort, D.; Guigues, N.; Knockaert, M.; Munaron, D.; Salter, I.; Woodward, E.M.; et al. Toward a harmonization for using in situ nutrient sensors in the marine environment. Front. Mar. Sci. 2020, 6, 773. [Google Scholar] [CrossRef]
  149. Sargazi, S.; Fatima, I.; Kiani, M.H.; Mohammadzadeh, V.; Arshad, R.; Bilal, M.; Rahdar, A.; Díez-Pascual, A.M.; Behzadmehr, R. Fluorescent-based nanosensors for selective detection of a wide range of biological macromolecules: A comprehensive review. Int. J. Biol. Macromol. 2022, 206, 115–147. [Google Scholar] [CrossRef]
  150. Tovar-Lopez, F.J. Recent progress in micro- and nanotechnology-enabled sensors for biomedical and environmental challenges. Sensors 2023, 23, 5406. [Google Scholar] [CrossRef]
  151. Shawon, Z.B.; Hoque, M.E.; Chowdhury, S.R. Nanosensors and nanobiosensors: Agricultural and food technology aspects. In Nanofabrication for Smart Nanosensor Applications; Elsevier: Amsterdam, The Netherlands, 2020; pp. 135–161. [Google Scholar]
  152. Manjunatha, S.B.; Biradar, D.P.; Aladakatti, Y.R. Nanotechnology and its applications in agriculture: A review. J. Farm. Sci. 2016, 29, 1–13. [Google Scholar]
  153. do Espirito Santo Pereira, A.; Caixeta Oliveira, H.; Fernandes Fraceto, L.; Santaella, C. Nanotechnology potential in seed priming for sustainable agriculture. Nanomaterials 2021, 11, 267. [Google Scholar] [CrossRef]
  154. Pramanik, P.; Krishnan, P.; Maity, A.; Mridha, N.; Mukherjee, A.; Rai, V. Application of nanotechnology in agriculture. Environ. Nanotechnol. 2020, 4, 317–348. [Google Scholar]
  155. Zahoor, M.; Nazir, N.; Iftikhar, M.; Naz, S.; Zekker, I.; Burlakovs, J.; Uddin, F.; Kamran, A.W.; Kallistova, A.; Pimenov, N.; et al. A review on silver nanoparticles: Classification, various methods of synthesis, and their potential roles in biomedical applications and water treatment. Water 2021, 13, 2216. [Google Scholar] [CrossRef]
  156. Pandiyaraj, V.; Murmu, A.; Pandy, S.K.; Sevanan, M.; Arjunan, S. Metal nanoparticles and its application on phenolic and heavy metal pollutants. Phys. Sci. Rev. 2023, 8, 2879–2897. [Google Scholar] [CrossRef]
  157. Amer, M.; Awwad, A. Green synthesis of copper nanoparticles by Citrus limon fruits extract, characterization and antibacterial activity. Chem. Int. 2020, 7, 1–8. [Google Scholar]
  158. Acharya, A.; Pal, P.K. Agriculture nanotechnology: Translating research outcome to field applications by influencing environmental sustainability. NanoImpact 2020, 19, 100232. [Google Scholar] [CrossRef]
  159. Das, A.; Dutta, P. Antifungal activity of biogenically synthesized silver and gold nanoparticles against sheath blight of rice. J. Nanosci. Nanotechnol. 2021, 21, 3547–3555. [Google Scholar] [CrossRef]
  160. Ponmurugan, P. Biosynthesis of silver and gold nanoparticles using Trichoderma atroviride for the biological control of Phomopsis canker disease in tea plants. IET Nanobiotechnology 2017, 11, 261–267. [Google Scholar] [CrossRef]
  161. Thakur, R.K.; Prasad, P. Synthesis of gold nanoparticles and assessment of in vitro toxicity against plant pathogens. Indian Phytopathol. 2022, 75, 101–108. [Google Scholar] [CrossRef]
  162. Kaman, P.; Dutta, P.; Bhattacharyya, A. Synthesis of gold nanoparticles from Metarhizium anisopliae for management of blast disease of rice and its effect on soil biological index and physicochemical properties. Res. Sq. 2022, 1–21. [Google Scholar] [CrossRef]
  163. Patra, P.; Mitra, S.; Debnath, N.; Goswami, A. Biochemical-, biophysical-, and microarray-based antifungal evaluation of the buffer-mediated synthesized nano zinc oxide: An in vivo and in vitro toxicity study. Langmuir 2012, 28, 16966–16978. [Google Scholar] [CrossRef]
  164. Wagner, G.; Korenkov, V.; Judy, J.D.; Bertsch, P.M. Nanoparticles composed of Zn and ZnO inhibit Peronospora tabacina spore germination in vitro and P. tabacina infectivity on tobacco leaves. Nanomaterials 2016, 6, 50. [Google Scholar] [CrossRef]
  165. Akpinar, I.; Unal, M.; Sar, T. Potential antifungal effects of silver nanoparticles (AgNPs) of different sizes against phytopathogenic Fusarium oxysporum f. sp. radicis-lycopersici (FORL) strains. SN Appl. Sci. 2021, 3, 506. [Google Scholar] [CrossRef]
  166. Mishra, S.; Singh, B.R.; Singh, A.; Keswani, C.; Naqvi, A.H.; Singh, H.B. Biofabricated silver nanoparticles act as a strong fungicide against Bipolaris sorokiniana causing spot blotch disease in wheat. PLoS ONE 2014, 9, e97881. [Google Scholar] [CrossRef]
  167. Jahan, Q.S.; Sultana, Z.; Ud-Daula, M.A.; Ashikuzzaman, M.; Reja, M.S.; Rahman, M.M.; Khaton, A.; Tang, M.A.; Rahman, M.S.; Faruquee, H.M.; et al. Optimization of green silver nanoparticles as nanofungicides for management of rice bakanae disease. Heliyon 2024, 10, e27579. [Google Scholar] [CrossRef] [PubMed]
  168. Ansari, M.; Ahmed, S.; Khan, M.T.; Hamad, N.A.; Ali, H.M.; Abbasi, A.; Mubeen, I.; Intisar, A.; Hasan, M.E.; Jasim, I.K. Evaluation of in vitro and in vivo antifungal activity of green synthesized silver nanoparticles against early blight in tomato. Horticulturae 2023, 9, 369. [Google Scholar] [CrossRef]
  169. Bibi, S.; Raza, M.; Shahbaz, M.; Ajmal, M.; Mehak, A.; Fatima, N.; Abasi, F.; Seelan, J.S.; Raja, N.I.; Yongchao, B.; et al. Biosynthesized silver nanoparticles enhanced wheat resistance to Bipolaris sorokiniana. Plant Physiol. Biochem. 2023, 203, 108067. [Google Scholar] [CrossRef]
  170. Qureshi, A.K.; Farooq, U.; Shakeel, Q.; Ali, S.; Ashiq, S.; Shahzad, S.; Tariq, M.; Seleiman, M.F.; Jamal, A.; Saeed, M.F.; et al. The Green Synthesis of Silver Nanoparticles from Avena fatua Extract: Antifungal Activity against Fusarium oxysporum f. sp. lycopersici. Pathogens 2023, 12, 1247. [Google Scholar] [CrossRef]
  171. Ghasemi, S.; Harighi, B.; Ashengroph, M. Biosynthesis of silver nanoparticles using Pseudomonas canadensis, and its antivirulence effects against Pseudomonas tolaasii, mushroom brown blotch agent. Sci. Rep. 2023, 13, 3668. [Google Scholar] [CrossRef]
  172. Ahmed, T.; Noman, M.; Jiang, H.; Shahid, M.; Ma, C.; Wu, Z.; Nazir, M.M.; Ali, M.A.; White, J.C.; Chen, J.; et al. Bioengineered chitosan-iron nanocomposite controls bacterial leaf blight disease by modulating plant defense response and nutritional status of rice (Oryza sativa L.). Nano Today 2022, 45, 101547. [Google Scholar] [CrossRef]
  173. Iqbal, M.; Raja, N.I.; Khan, S.A.; Ali, A.; Hanif, A.; Hussain, M.; Anwar, T.; Qureshi, H.; Saeed, M.; Rauf, A.; et al. Evaluation of Green Synthesized Silver Nanoparticles against Bacterial Pathogenic Strains of Plants. Pak. J. Bot. 2023, 55, 1967–1972. [Google Scholar] [CrossRef]
  174. Al-Otibi, F.; Perveen, K.; Al-Saif, N.A.; Alharbi, R.I.; Bokhari, N.A.; Albasher, G.; Al-Otaibi, R.M.; Al-Mosa, M.A. Biosynthesis of silver nanoparticles using Malva parviflora and their antifungal activity. Saudi J. Biol. Sci. 2021, 28, 2229–2235. [Google Scholar] [CrossRef]
  175. Yassin, M.A.; Elgorban, A.M.; El-Samawaty, A.E.; Almunqedhi, B.M. Biosynthesis of silver nanoparticles using Penicillium verrucosum and analysis of their antifungal activity. Saudi J. Biol. Sci. 2021, 28, 2123–2127. [Google Scholar] [CrossRef] [PubMed]
  176. Namburi, K.R.; Kora, A.J.; Chetukuri, A.; Kota, V.S. Biogenic silver nanoparticles as an antibacterial agent against bacterial leaf blight causing rice phytopathogen Xanthomonas oryzae pv. oryzae. Bioprocess Biosyst. Eng. 2021, 44, 1975–1988. [Google Scholar] [CrossRef] [PubMed]
  177. Ajaz, S.; Ahmed, T.; Shahid, M.; Noman, M.; Shah, A.A.; Mehmood, M.A.; Abbas, A.; Cheema, A.I.; Iqbal, M.Z.; Li, B. Bioinspired green synthesis of silver nanoparticles by using a native Bacillus sp. strain AW1-2: Characterization and antifungal activity against Colletotrichum falcatum Went. Enzym. Microb. Technol. 2021, 144, 109745. [Google Scholar] [CrossRef]
  178. Ibrahim, E.; Zhang, M.; Zhang, Y.; Hossain, A.; Qiu, W.; Chen, Y.; Wang, Y.; Wu, W.; Sun, G.; Li, B. Green-synthesization of silver nanoparticles using endophytic bacteria isolated from garlic and its antifungal activity against wheat Fusarium head blight pathogen Fusarium graminearum. Nanomaterials 2020, 10, 219. [Google Scholar] [CrossRef]
  179. Dorjee, L.; Gogoi, R.; Kamil, D.; Kumar, R.; Verma, A. Copper nanoparticles hold promise in the effective management of maize diseases without impairing environmental health. Phytoparasitica 2023, 51, 593–619. [Google Scholar] [CrossRef]
  180. Truong, H.T.; Nguyen, L.C.; Le, L.Q. Synthesis and antifungal activity of copper nanoparticles against Fusarium oxysporum pathogen of plants. Mater. Res. Express 2023, 10, 065001. [Google Scholar] [CrossRef]
  181. Iliger, K.S.; Sofi, T.A.; Bhat, N.A.; Ahanger, F.A.; Sekhar, J.C.; Elhendi, A.Z.; Al-Huqail, A.A.; Khan, F. Copper nanoparticles: Green synthesis and managing fruit rot disease of chilli caused by Colletotrichum capsici. Saudi J. Biol. Sci. 2021, 28, 1477–1486. [Google Scholar] [CrossRef] [PubMed]
  182. Natesan, K.; Ponmurugan, P.; Gnanamangai, B.M.; Manigandan, V.; Joy, S.P.J.; Jayakumar, C.; Amsaveni, G. Biosynthesis of silica and copper nanoparticles from Trichoderma, Streptomyces and Pseudomonas spp. evaluated against collar canker and red root-rot disease of tea plants. Arch. Phytopathol. Plant Prot. 2021, 54, 56–85. [Google Scholar] [CrossRef]
  183. Noman, M.; Ahmed, T.; White, J.C.; Nazir, M.M.; Azizullah Li, D.; Song, F. Bacillus altitudinis-Stabilized Multifarious Copper Nanoparticles Prevent Bacterial Fruit Blotch in Watermelon (Citrullus lanatus L.): Direct Pathogen Inhibition, In Planta Particles Accumulation, and Host Stomatal Immunity Modulation. Small 2023, 19, 2207136. [Google Scholar] [CrossRef]
  184. Majumdar, T.D.; Singh, M.; Thapa, M.; Dutta, M.; Mukherjee, A.; Ghosh, C.K. Size-dependent antibacterial activity of copper nanoparticles against Xanthomonas oryzae pv. oryzae–A synthetic and mechanistic approach. Colloid Interface Sci. Commun. 2019, 32, 100190. [Google Scholar] [CrossRef]
  185. Soliman, A.M.M.; Abdallah, E.A.M.; Hafez, E.E.; Kadous, E.A.; Kassem, F.A. Nematicidal Activity of Chemical and Green Biosynthesis of Copper Nanoparticles Against Root-Knot Nematode, Meloidogyne Incognita. Alex. Sci. Exch. J. 2022, 43, 583–591. [Google Scholar] [CrossRef]
  186. Liu, Q.; Zhang, A.; Wang, R.; Zhang, Q.; Cui, D. A review on metal-and metal oxide-based nanozymes: Properties, mechanisms, and applications. Nano Micro Lett. 2021, 13, 154. [Google Scholar] [CrossRef] [PubMed]
  187. Ealia, S.A.M.; Saravanakumar, M.P. A review on the classification, characterisation, synthesis of nanoparticles and their application. IOP Conf. Ser. Mater. Sci. Eng. 2017, 263, 032019. [Google Scholar] [CrossRef]
  188. Maity, D.; Gupta, U.; Saha, S. Biosynthesized metal oxide nanoparticles for sustainable agriculture: Next-generation nanotechnology for crop production, protection and management. Nanoscale 2022, 14, 13950–13989. [Google Scholar] [CrossRef]
  189. Cheira, M.F. Performance of poly sulfonamide/nano-silica composite for adsorption of thorium ions from sulfate solution. SN Appl. Sci. 2020, 2, 398. [Google Scholar] [CrossRef]
  190. Sharma, S.; Virk, K.; Sharma, K.; Bose, S.K.; Kumar, V.; Sharma, V.; Focarete, M.L.; Kalia, S. Preparation of gum acacia-poly (acrylamide-IPN-acrylic acid) based nanocomposite hydrogels via polymerization methods for antimicrobial applications. J. Mol. Struct. 2020, 1215, 128298. [Google Scholar] [CrossRef]
  191. Singh, K.R.; Nayak, V.; Sarkar, T.; Singh, R.P. Cerium oxide nanoparticles: Properties, biosynthesis and biomedical application. RSC Adv. 2020, 10, 27194–27214. [Google Scholar] [CrossRef] [PubMed]
  192. Dulta, K.; Koşarsoy Ağçeli, G.; Chauhan, P.; Jasrotia, R.; Chauhan, P.K. Ecofriendly synthesis of zinc oxide nanoparticles by Carica papaya leaf extract and their applications. J. Clust. Sci. 2022, 33, 603–617. [Google Scholar] [CrossRef]
  193. Helmy, K.G.; Partila, A.M.; Salah, M. Gamma radiation and polyvinyl pyrrolidone mediated synthesis of zinc oxide/zinc sulfide nanoparticles and evaluation of their antifungal effect on pre and post harvested orange and pomegranate fruits. Biocatal. Agric. Biotechnol. 2020, 29, 101728. [Google Scholar] [CrossRef]
  194. Zhu, W.; Hu, C.; Ren, Y.; Lu, Y.; Song, Y.; Ji, Y.; Han, C.; He, J. Green synthesis of zinc oxide nanoparticles using Cinnamomum camphora (L.) Presl leaf extracts and its antifungal activity. J. Environ. Chem. Eng. 2021, 9, 106659. [Google Scholar] [CrossRef]
  195. Khan, R.A.; Tang, Y.; Naz, I.; Alam, S.S.; Wang, W.; Ahmad, M.; Najeeb, S.; Rao, C.; Li, Y.; Xie, B.; et al. Management of Ralstonia solanacearum in tomato using ZnO nanoparticles synthesized through Matricaria chamomilla. Plant Dis. 2021, 105, 3224–3230. [Google Scholar] [CrossRef] [PubMed]
  196. Ali, M.; Wang, X.; Haroon, U.; Chaudhary, H.J.; Kamal, A.; Ali, Q.; Saleem, M.H.; Usman, K.; Alatawi, A.; Ali, S.; et al. Antifungal activity of Zinc nitrate derived nano Zno fungicide synthesized from Trachyspermum ammi to control fruit rot disease of grapefruit. Ecotoxicol. Environ. Saf. 2022, 233, 113311. [Google Scholar] [CrossRef]
  197. Ahmad, H.; Venugopal, K.; Rajagopal, K.; De Britto, S.; Nandini, B.; Pushpalatha, H.G.; Konappa, N.; Udayashankar, A.C.; Geetha, N.; Jogaiah, S. Green synthesis and characterization of zinc oxide nanoparticles using Eucalyptus globules and their fungicidal ability against pathogenic fungi of apple orchards. Biomolecules 2020, 10, 425. [Google Scholar] [CrossRef] [PubMed]
  198. Chaithanya, G.; Kumar, A.; Vijay, D.; Singh, P.K.; Hussain, Z.; Basu, S.; Lal, S.K. Efficacy of nanoparticles against purple blotch (Alternaria porri) of onion. Indian Phytopathol. 2023, 76, 845–852. [Google Scholar] [CrossRef]
  199. Rehman, F.U.; Paker, N.P.; Khan, M.; Naeem, M.; Munis, M.F.H.; Rehman, S.U.; Chaudhary, H.J. Bio-fabrication of zinc oxide nanoparticles from Picea smithiana and their potential antimicrobial activities against Xanthomonas campestris pv. Vesicatoria and Ralstonia solanacearum causing bacterial leaf spot and bacterial wilt in tomato. World J. Microbiol. Biotechnol. 2023, 39, 176. [Google Scholar] [CrossRef]
  200. Jomeyazdian, A.; Pirnia, M.; Alaei, H.; Taheri, A.; Sarani, S. Control of Fusarium wilt disease of tomato and improvement of some growth factors through green synthesized zinc oxide nanoparticles. Eur. J. Plant Pathol. 2024, 169, 333–345. [Google Scholar] [CrossRef]
  201. Karmous, I.; Vaidya, S.; Dimkpa, C.; Zuverza-Mena, N.; da Silva, W.; Barroso, K.A.; Milagres, J.; Bharadwaj, A.; Abdelraheem, W.; White, J.C.; et al. Biologically synthesized zinc and copper oxide nanoparticles using Cannabis sativa L. enhance soybean (Glycine max) defense against Fusarium virguliforme. Pestic. Biochem. Physiol. 2023, 194, 105486. [Google Scholar] [CrossRef]
  202. Sardar, M.; Ahmed, W.; Al Ayoubi, S.; Nisa, S.; Bibi, Y.; Sabir, M.; Khan, M.M.; Ahmed, W.; Qayyum, A. Fungicidal synergistic effect of biogenically synthesized zinc oxide and copper oxide nanoparticles against Alternaria citri causing citrus black rot disease. Saudi J. Biol. Sci. 2022, 29, 88–95. [Google Scholar] [CrossRef] [PubMed]
  203. Parveen, A.; Siddiqui, Z.A. Effect of silver oxide nanoparticles on growth, activities of defense enzymes and fungal and bacterial diseases of tomato. Gesunde Pflanz. 2023, 75, 405–414. [Google Scholar] [CrossRef]
  204. Ogunyemi, S.O.; Luo, J.; Abdallah, Y.; Yu, S.; Wang, X.; Alkhalifah, D.H.; Hozzein, W.N.; Wang, F.; Bi, J.A.; Yan, C.; et al. Copper oxide nanoparticles: An effective suppression tool against bacterial leaf blight of rice and its impacts on plants. Pest. Manag. Sci. 2024, 80, 1279–1288. [Google Scholar] [CrossRef]
  205. Tiwari, V.; Bambharoliya, K.S.; Bhatt, M.D.; Nath, M.; Arora, S.; Dobriyal, A.K.; Bhatt, D. Application of green synthesized copper oxide nanoparticles for effective mitigation of Fusarium wilt disease in roots of Cicer arietinum. Physiol. Mol. Plant Pathol. 2024, 131, 102244. [Google Scholar] [CrossRef]
  206. Kamel, S.M.; Elgobashy, S.F.; Omara, R.I.; Derbalah, A.S.; Abdelfatah, M.; El-Shaer, A.; Al-Askar, A.A.; Abdelkhalek, A.; Abd-Elsalam, K.A.; Essa, T.; et al. Antifungal activity of copper oxide nanoparticles against root rot disease in cucumber. J. Fungi 2022, 8, 911. [Google Scholar] [CrossRef] [PubMed]
  207. Sawake, M.M.; Moharil, M.P.; Ingle, Y.V.; Jadhav, P.V.; Ingle, A.P.; Khelurkar, V.C.; Paithankar, D.H.; Bathe, G.A.; Gade, A.K. Management of Phytophthora parasitica causing gummosis in citrus using biogenic copper oxide nanoparticles. J. Appl. Microbiol. 2022, 132, 3142–3154. [Google Scholar] [CrossRef] [PubMed]
  208. Ashraf, H.; Anjum, T.; Riaz, S.; Batool, T.; Naseem, S.; Ahmad, I.S. Sustainable synthesis of microwave assisted IONPs by using spinacia oleracea: Enhances resistance against fungal wilt infection by inducing ROS and modulating defense system in tomato plants. J. Nanobiotechnol. 2021, 20, 8. [Google Scholar]
  209. Gaba, S.; Rai, A.K.; Varma, A.; Prasad, R.; Goel, A. Biocontrol potential of mycogenic copper oxide nanoparticles against Alternaria brassicae. Front. Chem. 2022, 10, 966396. [Google Scholar] [CrossRef] [PubMed]
  210. Khan, A.U.; Khan, M.; Khan, A.A.; Parveen, A.; Ansari, S.; Alam, M. Effect of Phyto-Assisted Synthesis of Magnesium Oxide Nanoparticles (MgO-NPs) on Bacteria and the Root-Knot Nematode. Bioinorg. Chem. Appl. 2022, 2022, 3973841. [Google Scholar] [CrossRef]
  211. El-Shetehy, M.; Moradi, A.; Maceroni, M.; Reinhardt, D.; Petri-Fink, A.; Rothen-Rutishauser, B.; Mauch, F.; Schwab, F. Silica nanoparticles enhance disease resistance in Arabidopsis plants. Nat. Nanotechnol. 2021, 16, 344–353. [Google Scholar] [CrossRef]
  212. Zaheer, S.; Shehzad, J.; Chaudhari, S.K.; Hasan, M.; Mustafa, G. Morphological and biochemical responses of Vigna radiata L. seedlings towards green synthesized SiO2 NPs. Silicon 2023, 15, 5925–5936. [Google Scholar] [CrossRef]
  213. Abdallah, Y.; Nehela, Y.; Ogunyemi, S.O.; Ijaz, M.; Ahmed, T.; Elashmony, R.; Alkhalifah, D.H.; Hozzein, W.N.; Xu, L.; Yan, C.; et al. Bio-functionalized nickel-silica nanoparticles suppress bacterial leaf blight disease in rice (Oryza sativa L.). Front. Plant Sci. 2023, 14, 1216782. [Google Scholar] [CrossRef]
  214. Goswami, P.; Sharma, M.; Srivastava, N.; Mathur, J. Assessment of the fungicidal efficacy of biogenic SiO2 NPs in Eruca sativa against fusarium wilt. J. Nat. Pestic. Res. 2022, 2, 100011. [Google Scholar] [CrossRef]
  215. Awad-Allah, E.F.; Shams, A.H.; Helaly, A.A. Suppression of bacterial leaf spot by green synthesized silica nanoparticles and antagonistic yeast improves growth, productivity and quality of sweet pepper. Plants 2021, 10, 1689. [Google Scholar] [CrossRef] [PubMed]
  216. Abdelrhim, A.S.; Mazrou, Y.S.; Nehela, Y.; Atallah, O.O.; El-Ashmony, R.M.; Dawood, M.F. Silicon dioxide nanoparticles induce innate immune responses and activate antioxidant machinery in wheat against Rhizoctonia solani. Plants 2021, 10, 2758. [Google Scholar] [CrossRef]
  217. Elamawi, R.M.; Tahoon, A.M.; Elsharnoby, D.E.; El-Shafey, R.A. Bio-production of silica nanoparticles from rice husk and their impact on rice bakanae disease and grain yield. Arch. Phytopathol. Plant Prot. 2020, 53, 459–478. [Google Scholar] [CrossRef]
  218. Khan, M.R.; Siddiqui, Z.A. Efficacy of titanium dioxide nanoparticles in the management of disease complex of beetroot (Beta vulgaris L.) caused by Pectobacterium betavasculorum, Rhizoctonia solani, and Meloidogyne incognita. Gesunde Pflanz. 2021, 73, 445–464. [Google Scholar] [CrossRef]
  219. Satti, S.H.; Raja, N.I.; Javed, B.; Akram, A.; Mashwani, Z.U.; Ahmad, M.S.; Ikram, M. Titanium dioxide nanoparticles elicited agro-morphological and physicochemical modifications in wheat plants to control Bipolaris sorokiniana. PLoS ONE 2021, 16, e0246880. [Google Scholar] [CrossRef] [PubMed]
  220. Khan, M.; Siddiqui, Z.A.; Parveen, A.; Khan, A.A.; Moon, I.S.; Alam, M. Elucidating the role of silicon dioxide and titanium dioxide nanoparticles in mitigating the disease of the eggplant caused by Phomopsis vexans, Ralstonia solanacearum, and root-knot nematode Meloidogyne incognita. Nanotechnol. Rev. 2022, 11, 1606–1619. [Google Scholar] [CrossRef]
  221. Hossain, A.; Abdallah, Y.; Ali, M.A.; Masum, M.M.; Li, B.; Sun, G.; Meng, Y.; Wang, Y.; An, Q. Lemon-fruit-based green synthesis of zinc oxide nanoparticles and titanium dioxide nanoparticles against soft rot bacterial pathogen Dickeya dadantii. Biomolecules 2019, 9, 863. [Google Scholar] [CrossRef]
  222. El-Gazzar, N.; Ismail, A.M. The potential use of Titanium, Silver and Selenium nanoparticles in controlling leaf blight of tomato caused by Alternaria alternata. Biocatal. Agric. Biotechnol. 2020, 27, 101708. [Google Scholar] [CrossRef]
  223. Irshad, M.A.; Nawaz, R.; ur Rehman, M.Z.; Imran, M.; Ahmad, J.; Ahmad, S.; Inam, A.; Razzaq, A.; Rizwan, M.; Ali, S. Synthesis and characterization of titanium dioxide nanoparticles by chemical and green methods and their antifungal activities against wheat rust. Chemosphere 2020, 258, 127352. [Google Scholar] [CrossRef]
  224. Satti, S.H.; Raja, N.I.; Ikram, M.; Oraby, H.F.; Mashwani, Z.U.; Mohamed, A.H.; Singh, A.; Omar, A.A. Plant-based titanium dioxide nanoparticles trigger biochemical and proteome modifications in Triticum aestivum L. under biotic stress of Puccinia striiformis. Molecules 2022, 27, 4274. [Google Scholar] [CrossRef]
  225. Acuña-Fuentes, N.L.; Vargas-Hernandez, M.; Rivero-Montejo, S.D.; Rivas-Ramirez, L.K.; Macias-Bobadilla, I.; Palos-Barba, V.; Rivera-Muñoz, E.M.; Guevara-Gonzalez, R.G.; Torres-Pacheco, I. Antiviral activity of TiO2 NPs against tobacco mosaic virus in chili pepper (Capsicum annuum L.). Agriculture 2022, 12, 2101. [Google Scholar] [CrossRef]
  226. Monclou-Salcedo, S.A.; Correa-Torres, S.N.; Kopytko, M.I.; Santoyo-Muñóz, C.; Vesga-Guzmán, D.M.; Castellares-Lozano, R.; López-Amaris, M.; Saavedra-Mancera, A.D.; Herrera-Barros, A.P. Evaluación antifúngica de nanopartículas de TiO2 para inhibición de Fusarium solani en Palma Africana. Int. J. Agric. Nat. Resour. 2020, 47, 126–133. [Google Scholar]
  227. Adisa, I.O.; Reddy Pullagurala, V.L.; Rawat, S.; Hernandez-Viezcas, J.A.; Dimkpa, C.O.; Elmer, W.H.; White, J.C.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Role of cerium compounds in Fusarium wilt suppression and growth enhancement in tomato (Solanum lycopersicum). J. Agric. Food Chem. 2018, 66, 5959–5970. [Google Scholar] [CrossRef]
  228. Shahbaz, M.; Fatima, N.; Mashwani, Z.U.; Akram, A.; Haq, E.U.; Mehak, A.; Abasi, F.; Ajmal, M.; Yousaf, T.; Raja, N.I.; et al. Effect of phytosynthesized selenium and cerium oxide nanoparticles on wheat (Triticum aestivum L.) against stripe rust disease. Molecules 2022, 27, 8149. [Google Scholar] [CrossRef] [PubMed]
  229. Alotaibi, M.O.; Alotaibi, N.M.; Ghoneim, A.M.; ul Ain, N.; Irshad, M.A.; Nawaz, R.; Abbas, T.; Abbas, A.; Rizwan, M.; Ali, S. Effect of green synthesized cerium oxide nanoparticles on fungal disease of wheat plants: A field study. Chemosphere 2023, 339, 139731. [Google Scholar] [CrossRef]
  230. Elbasuney, S.; El-Sayyad, G.S.; Abdelaziz, A.M.; Rizk, S.H.; Tolba, M.M.; Attia, M.S. Stable Colloidal Iron Oxide Nanoparticles: A New Green Nanofertilizer and Therapeutic Nutrient for Eggplant Immune Response against Fusarium Wilt Disease. J. Clust. Sci. 2024, 35, 983–997. [Google Scholar] [CrossRef]
  231. Niazi, F.; Ali, M.; Haroon, U.; Farhana Kamal, A.; Rashid, T.; Anwar, F.; Nawab, R.; Chaudhary, H.J.; Munis, M.F. Effect of green Fe2O3 nanoparticles in controlling Fusarium fruit rot disease of loquat in Pakistan. Braz. J. Microbiol. 2023, 54, 1341–1350. [Google Scholar] [CrossRef]
  232. Akbar, M.; Haroon, U.; Ali, M.; Tahir, K.; Chaudhary, H.J.; Munis, M.F. Mycosynthesized Fe2O3 nanoparticles diminish brown rot of apple whilst maintaining composition and pertinent organoleptic properties. J. Appl. Microbiol. 2022, 132, 3735–3745. [Google Scholar] [CrossRef]
  233. Umar, H.; Aliyu, M.R.; Ozsahin, D.U. Iron oxide nanoparticles synthesized using Mentha spicata extract and evaluation of its antibacterial, cytotoxicity and antimigratory potential on highly metastatic human breast cells. Biomed. Phys. Eng. Express 2024, 10, 035019. [Google Scholar] [CrossRef]
  234. Alam, T.; Khan, R.A.; Ali, A.; Sher, H.; Ullah, Z.; Ali, M. Biogenic synthesis of iron oxide nanoparticles via Skimmia laureola and their antibacterial efficacy against bacterial wilt pathogen Ralstonia solanacearum. Mater. Sci. Eng. C 2019, 98, 101–108. [Google Scholar] [CrossRef]
  235. Ali, M.; Haroon, U.; Khizar, M.; Chaudhary, H.J.; Hussain Munis, M.F. Scanning electron microscopy of bio-fabricated Fe2O3 nanoparticles and their application to control brown rot of citrus. Microsc. Res. Tech. 2021, 84, 101–110. [Google Scholar] [CrossRef] [PubMed]
  236. Mogazy, A.M.; Mohamed, H.I.; El-Mahdy, O.M. Calcium and iron nanoparticles: A positive modulator of innate immune responses in strawberry against Botrytis cinerea. Process Biochem. 2022, 115, 128–145. [Google Scholar] [CrossRef]
  237. Anwaar, S.; Ijaz, D.E.; Anwar, T.; Qureshi, H.; Nazish, M.; Alrefaei, A.F.; Almutairi, M.H.; Alharbi, S.N. Boosting Solanum tuberosum resistance to Alternaria solani through green synthesized ferric oxide (Fe2O3) nanoparticles. Sci. Rep. 2024, 14, 2375. [Google Scholar] [CrossRef]
  238. Zubair, M.S.; Munis, M.F.; Alsudays, I.M.; Alamer, K.H.; Haroon, U.; Kamal, A.; Ali, M.; Ahmed, J.; Ahmad, Z.; Attia, H. First report of fruit rot of cherry and its control using Fe2O3 nanoparticles synthesized in Calotropis procera. Molecules 2022, 27, 4461. [Google Scholar] [CrossRef]
  239. Ismail, A.; Kabary, H.; Samy, A. Synthesis of α-Al2O3 Nanoparticles from Pepsi Cans Wastes and Its Fungicidal Effect on Some Mycotoxins Producing Fungal Isolates. Res. Sq. 2021. [Google Scholar] [CrossRef]
  240. Suryavanshi, P.; Pandit, R.; Gade, A.; Derita, M.; Zachino, S.; Rai, M. Colletotrichum sp.-mediated synthesis of sulphur and aluminium oxide nanoparticles and its in vitro activity against selected food-borne pathogens. LWT Food Sci. Technol. 2017, 81, 188–194. [Google Scholar] [CrossRef]
  241. Ogunyemi, S.O.; Xu, X.; Xu, L.; Abdallah, Y.; Rizwan, M.; Lv, L.; Ahmed, T.; Ali, H.M.; Khan, F.; Yan, C.; et al. Cobalt oxide nanoparticles: An effective growth promoter of Arabidopsis plants and nano-pesticide against bacterial leaf blight pathogen in rice. Ecotoxicol. Environ. Saf. 2023, 257, 114935. [Google Scholar] [CrossRef] [PubMed]
  242. Ashraf, H.; Batool, T.; Anjum, T.; Illyas, A.; Li, G.; Naseem, S.; Riaz, S. Antifungal Potential of Green Synthesized Magnetite Nanoparticles Black Coffee–Magnetite Nanoparticles against Wilt Infection by Ameliorating Enzymatic Activity and Gene Expression in Solanum lycopersicum L. Front. Microbiol. 2022, 13, 754292. [Google Scholar]
  243. Elenany, A.M.; Atia, M.M.; Abbas, E.E.; Moustafa, M.; Alshaharni, M.O.; Negm, S.; Elnahal, A.S. Nanoparticles and Chemical Inducers: A Sustainable Shield against Onion White Rot. Biology 2024, 13, 219. [Google Scholar] [CrossRef]
  244. El-Ganainy, S.M.; El-Bakery, A.M.; Hafez, H.M.; Ismail, A.M.; El-Abdeen, A.Z.; Ata, A.A.; Elraheem, O.A.; El Kady, Y.M.; Hamouda, A.F.; El-Beltagi, H.S.; et al. Humic acid-coated Fe3O4 nanoparticles confer resistance to Acremonium wilt disease and improve physiological and morphological attributes of grain Sorghum. Polymers 2022, 14, 3099. [Google Scholar] [CrossRef]
  245. Ahmed, T.; Noman, M.; Luo, J.; Muhammad, S.; Shahid, M.; Ali, M.A.; Zhang, M.; Li, B. Bioengineered chitosan-magnesium nanocomposite: A novel agricultural antimicrobial agent against Acidovorax oryzae and Rhizoctonia solani for sustainable rice production. Int. J. Biol. Macromol. 2021, 168, 834–845. [Google Scholar] [CrossRef] [PubMed]
  246. Abdallah, Y.; Ogunyemi, S.O.; Abdelazez, A.; Zhang, M.; Hong, X.; Ibrahim, E.; Hossain, A.; Fouad, H.; Li, B.; Chen, J. The green synthesis of MgO nano-flowers using Rosmarinus officinalis L.(Rosemary) and the antibacterial activities against Xanthomonas oryzae pv. oryzae. BioMed Res. Int. 2019, 2019, 5620989. [Google Scholar] [CrossRef] [PubMed]
  247. Ogunyemi, S.O.; Zhang, M.; Abdallah, Y.; Ahmed, T.; Qiu, W.; Ali, M.A.; Yan, C.; Yang, Y.; Chen, J.; Li, B. The bio-synthesis of three metal oxide nanoparticles (ZnO, MnO2, and MgO) and their antibacterial activity against the bacterial leaf blight pathogen. Front. Microbiol. 2020, 11, 588326. [Google Scholar] [CrossRef]
  248. Rabea, A.; Naeem, E.; Balabel, N.M.; Daigham, G.E. Management of potato brown rot disease using chemically synthesized CuO-NPs and MgO-NPs. Bot. Stud. 2023, 64, 20. [Google Scholar] [CrossRef] [PubMed]
  249. Ahamad, L.; Azmat, A.L.; Masudulla, K.H.; Farid, O.; Mahboob, A.L. Exploring the nano-fungicidal efficacy of green synthesized magnesium oxide nanoparticles (MgO NPs) on the development, physiology, and infection of carrot (Daucus carota L.) with Alternaria leaf blight (ALB): Molecular docking. J. Integr. Agric. 2023, 22, 3069–3080. [Google Scholar] [CrossRef]
  250. Abdel-Aziz, M.M.; Emam, T.M.; Elsherbiny, E.A. Bioactivity of magnesium oxide nanoparticles synthesized from cell filtrate of endobacterium Burkholderia rinojensis against Fusarium oxysporum. Mater. Sci. Eng. C 2020, 109, 110617. [Google Scholar] [CrossRef]
  251. Abdelfattah, N.A.; Yousef, M.A.; Badawy, A.A.; Salem, S.S. Influence of biosynthesized magnesium oxide nanoparticles on growth and physiological aspects of cowpea (Vigna unguiculata L.) plant, cowpea beetle, and cytotoxicity. Biotechnol. J. 2023, 18, 2300301. [Google Scholar] [CrossRef]
  252. Wang, Z.L.; Zhang, X.; Fan, G.J.; Que, Y.; Xue, F.; Liu, Y.H. Toxicity effects and mechanisms of MgO nanoparticles on the oomycete pathogen Phytophthora infestans and its host Solanum tuberosum. Toxics 2022, 10, 553. [Google Scholar] [CrossRef]
  253. Ismail, A.M. Efficacy of copper oxide and magnesium oxide nanoparticles on controlling black scurf disease on potato. Egypt. J. Phytopathol. 2021, 49, 116–130. [Google Scholar] [CrossRef]
  254. Ismail, A.M.; El-Gawad, A.; Mona, E. Antifungal activity of MgO and ZnO nanoparticles against powdery mildew of pepper under greenhouse conditions. Egypt. J. Agric. Res. 2021, 99, 421–434. [Google Scholar]
  255. Liao, J.; Yuan, Z.; Wang, X.; Chen, T.; Qian, K.; Cui, Y.; Rong, A.; Zheng, C.; Liu, Y.; Wang, D.; et al. Magnesium oxide nanoparticles reduce clubroot by regulating plant defense response and rhizosphere microbial community of tumorous stem mustard (Brassica juncea var. tumida). Front. Microbiol. 2024, 15, 1370427. [Google Scholar] [CrossRef] [PubMed]
  256. Hantoosh, M.N.; Hussein, H. Bioactivity of Magnesium Oxide Nanoparticles Synthesized by Alcoholic Extract of Walnut Tree Bark Juglans regia against Thielaviopsis paradoxa and Thielaviopsis punctulata in vitro. IOP Conf. Ser. Earth Environ. Sci. 2023, 1252, 012021. [Google Scholar] [CrossRef]
  257. Liao, Y.Y.; Huang, Y.; Carvalho, R.; Choudhary, M.; Da Silva, S.; Colee, J.; Huerta, A.; Vallad, G.E.; Freeman, J.H.; Jones, J.B.; et al. Magnesium oxide nanomaterial, an alternative for commercial copper bactericides: Field-scale tomato bacterial spot disease management and total and bioavailable metal accumulation in soil. Environ. Sci. Technol. 2021, 55, 13561–13570. [Google Scholar] [CrossRef] [PubMed]
  258. El-Sayed, M.E. Nanoadsorbents for water and wastewater remediation. Sci. Total Environ. 2020, 739, 139903. [Google Scholar] [CrossRef] [PubMed]
  259. Feng, S.; Wang, J.; Zhang, L.; Chen, Q.; Yue, W.; Ke, N.; Xie, H. Coumarin-containing light-responsive carboxymethyl chitosan micelles as nanocarriers for controlled release of pesticide. Polymers 2020, 12, 2268. [Google Scholar] [CrossRef]
  260. Mustafa, I.F.; Hussein, M.Z.; Idris, A.S.; Hilmi, N.H.; Ramli, N.R.; Fakurazi, S. The effect of surfactant type on the physico-chemical properties of hexaconazole/dazomet-micelle nanodelivery system and its biofungicidal activity against Ganoderma boninense. Colloids Surf. A Physicochem. Eng. Asp. 2022, 640, 128402. [Google Scholar] [CrossRef]
  261. Su, W.; Qin, Y.; Wu, J.; Meng, G.; Yang, S.; Cui, L.; Liu, Z.; Guo, X. Linear Supramolecular Block Copolymer Micelles for ROS-Responsive Release of Antimicrobial Pesticides. ACS Appl. Nano Mater. 2023, 6, 12736–12743. [Google Scholar] [CrossRef]
  262. Luong, J.H.; Tran, C.; Ton-That, D. A paradox over electric vehicles, mining of lithium for car batteries. Energies 2022, 15, 7997. [Google Scholar] [CrossRef]
  263. Pérez-de-Luque, A.; Cifuentes, Z.; Beckstead, J.A.; Sillero, J.C.; Ávila, C.; Rubio, J.; Ryan, R.O. Effect of amphotericin B nanodisks on plant fungal diseases. Pest. Manag. Sci. 2012, 68, 67–74. [Google Scholar] [CrossRef]
  264. Xu, Y.; Wei, Y.; Jiang, S.; Xu, F.; Wang, H.; Shao, X. Preparation and characterization of tea tree oil solid liposomes to control brown rot and improve quality in peach fruit. Lwt 2022, 162, 113442. [Google Scholar] [CrossRef]
  265. Chen, X.; Qiu, L.; Liu, Q.; He, Y. Preparation of an environmentally friendly nano-insecticide through encapsulation in polymeric liposomes and its insecticidal activities against the fall armyworm, Spodoptera frugiperda. Insects 2022, 13, 625. [Google Scholar] [CrossRef] [PubMed]
  266. Li, M.; Li, J.; Meng, Y.; Wang, Y.; Gao, M.; Dong, J.; Cao, Z.; Zhang, L.; Ma, S. Preparation of nanoliposomes containing the extracts of Eleocharis dulcis corm-peels and ascertaining their aphidicidal activity against Megoura crassicauda and Acyrthosiphon pisum. Ind. Crop. Prod. 2024, 207, 117746. [Google Scholar] [CrossRef]
  267. Liu, X.; He, B.; Xu, Z.; Yin, M.; Yang, W.; Zhang, H.; Cao, J.; Shen, J. A functionalized fluorescent dendrimer as a pesticide nanocarrier: Application in pest control. Nanoscale 2015, 7, 445–449. [Google Scholar] [CrossRef]
  268. Thanh, V.M.; Bui, L.M.; Bach, L.G.; Nguyen, N.T.; Thi, H.L.; Hoang Thi, T.T. Origanum majorana L. essential oil-associated polymeric nano dendrimer for antifungal activity against Phytophthora infestans. Materials 2019, 12, 1446. [Google Scholar] [CrossRef]
  269. Labidi, S.; Sandhu, R.K.; Beaulieu, C.; Beaudoin, N. Increased ferritin and iron accumulation in tubers of thaxtomin A-habituated potato var. Yukon Gold somaclones with enhanced resistance to common scab. J. Plant Pathol. 2023, 105, 107–119. [Google Scholar] [CrossRef]
  270. Maleki, R.; Abdollahi, H.; Piri, S. Variation of active iron and ferritin content in pear cultivars with different levels of pathogen resistance following inoculation with Erwinia amylovora. J. Plant Pathol. 2022, 104, 281–293. [Google Scholar] [CrossRef]
  271. Khot, L.R.; Sankaran, S.; Maja, J.M.; Ehsani, R.; Schuster, E.W. Applications of nanomaterials in agricultural production and crop protection: A review. Crop Prot. 2012, 35, 64–70. [Google Scholar] [CrossRef]
  272. Khare, P.; Talreja, N.; Deva, D.; Sharma, A.; Verma, N. Carbon nanofibers containing metal-doped porous carbon beads for environmental remediation applications. Chem. Eng. J. 2013, 229, 72–81. [Google Scholar] [CrossRef]
  273. Kumar, D.; Talreja, N. Nickel nanoparticles-doped rhodamine grafted carbon nanofibers as colorimetric probe: Naked eye detection and highly sensitive measurement of aqueous Cr3+ and Pb2+. Korean J. Chem. Eng. 2019, 36, 126–135. [Google Scholar] [CrossRef]
  274. Saraswat, R.; Talreja, N.; Deva, D.; Sankararamakrishnan, N.; Sharma, A.; Verma, N. Development of novel in situ nickel-doped, phenolic resin-based micro–nano-activated carbon adsorbents for the removal of vitamin B-12. Chem. Eng. J. 2012, 197, 250–260. [Google Scholar] [CrossRef]
  275. Talreja, N.; Jung, S.; Kim, T. Phenol-formaldehyde-resin-based activated carbons with controlled pore size distribution for high-performance supercapacitors. Chem. Eng. J. 2020, 379, 122332. [Google Scholar] [CrossRef]
  276. Talreja, N.; Kumar, D.; Verma, N. Removal of hexavalent chromium from water using Fe-grown carbon nanofibers containing porous carbon microbeads. J. Water Process Eng. 2014, 3, 34–45. [Google Scholar] [CrossRef]
  277. Talreja, N.; Verma, N.; Kumar, D. Carbon bead-supported ethylene diamine-functionalized carbon nanofibers: An efficient adsorbent for salicylic acid. CLEAN–Soil. Air Water 2016, 44, 1461–1470. [Google Scholar] [CrossRef]
  278. Omar, R.A.; Talreja, N.; Chauhan, D.; Mangalaraja, R.V.; Ashfaq, M. Nano metal-carbon–based materials: Emerging platform for the growth and protection of crops. In Nanotechnology-Based Sustainable Alternatives for the Management of Plant Diseases; Elsevier: Amsterdam, The Netherlands, 2022; pp. 341–354. [Google Scholar]
  279. González-García, Y.; Cadenas-Pliego, G.; Alpuche-Solís, Á.G.; Cabrera, R.I.; Juárez-Maldonado, A. Carbon nanotubes decrease the negative impact of Alternaria solani in tomato crop. Nanomaterials 2021, 11, 1080. [Google Scholar] [CrossRef] [PubMed]
  280. Hao, Y.; Fang, P.; Ma, C.; White, J.C.; Xiang, Z.; Wang, H.; Zhang, Z.; Rui, Y.; Xing, B. Engineered nanomaterials inhibit Podosphaera pannosa infection on rose leaves by regulating phytohormones. Environ. Res. 2019, 170, 1–6. [Google Scholar] [CrossRef] [PubMed]
  281. Lipșa, F.D.; Ursu, E.L.; Ursu, C.; Ulea, E.; Cazacu, A. Evaluation of the antifungal activity of gold–chitosan and carbon nanoparticles on Fusarium oxysporum. Agronomy 2020, 10, 1143. [Google Scholar] [CrossRef]
  282. Adeel, M.; Farooq, T.; White, J.C.; Hao, Y.; He, Z.; Rui, Y. Carbon-based nanomaterials suppress tobacco mosaic virus (TMV) infection and induce resistance in Nicotiana benthamiana. J. Hazard. Mater. 2021, 404, 124167. [Google Scholar] [CrossRef]
  283. El-Ganainy, S.M.; Mosa, M.A.; Ismail, A.M.; Khalil, A.E. Lignin-loaded carbon nanoparticles as a promising control agent against Fusarium verticillioides in maize: Physiological and biochemical analyses. Polymers 2023, 15, 1193. [Google Scholar] [CrossRef]
  284. Al-Zaban, M.I.; Alhag, S.K.; Dablool, A.S.; Ahmed, A.E.; Alghamdi, S.; Ali, B.; Al-Saeed, F.A.; Saleem, M.H.; Poczai, P. Manufactured nano-objects confer viral protection against cucurbit chlorotic yellows virus (CCYV) infecting nicotiana benthamiana. Microorganisms 2022, 10, 1837. [Google Scholar] [CrossRef]
  285. Baka, Z.A.; El-Zahed, M.M. Biocontrol of chocolate spot disease of broad bean (Vicia faba L.) caused by Botrytis fabae using biosynthesized reduced graphene oxide/silver nanocomposite. Physiol. Mol. Plant Pathol. 2023, 127, 102116. [Google Scholar] [CrossRef]
  286. El-Abeid, S.E.; Ahmed, Y.; Daròs, J.A.; Mohamed, M.A. Reduced graphene oxide nanosheet-decorated copper oxide nanoparticles: A potent antifungal nanocomposite against fusarium root rot and wilt diseases of tomato and pepper plants. Nanomaterials 2020, 10, 1001. [Google Scholar] [CrossRef] [PubMed]
  287. Bytešníková, Z.; Pečenka, J.; Tekielska, D.; Kiss, T.; Švec, P.; Ridošková, A.; Bezdička, P.; Pekárková, J.; Eichmeier, A.; Pokluda, R.; et al. Reduced graphene oxide-based nanometal-composite containing copper and silver nanoparticles protect tomato and pepper against Xanthomonas euvesicatoria infection. Chem. Biol. Technol. Agric. 2022, 9, 84. [Google Scholar] [CrossRef]
  288. Yu, H.; Wang, L.; Qu, J.; Wang, X.; Huang, F.; Jiao, Y.; Zhang, Y. Bi2O3/TiO2@ reduced graphene oxide with enzyme-like properties efficiently inactivates Pseudomonas syringae pv. tomato DC3000 and enhances abiotic stress tolerance in tomato. Environ. Sci. Nano 2022, 9, 118–132. [Google Scholar] [CrossRef]
  289. Bityutskii, N.P.; Yakkonen, K.L.; Lukina, K.A.; Semenov, K.N. Fullerenol increases effectiveness of foliar iron fertilization in iron-deficient cucumber. PLoS ONE 2020, 15, e0232765. [Google Scholar] [CrossRef]
  290. Gyawali, B.; Rahimi, R.; Alizadeh, H.; Mohammadi, M. Graphene Quantum Dots (GQD)-Mediated dsRNA Delivery for the Control of Fusarium Head Blight Disease in Wheat. ACS Appl. Bio Mater. 2024, 7, 1526–1535. [Google Scholar] [CrossRef]
  291. Wang, X.; Cai, A.; Wen, X.; Jing, D.; Qi, H.; Yuan, H. Graphene oxide-Fe3O4 nanocomposites as high-performance antifungal agents against Plasmopara viticola. Sci. China Mater. 2017, 60, 258–268. [Google Scholar] [CrossRef]
  292. Cheng, J.; Sun, Z.; Li, X.; Yu, Y. Effects of modified nanoscale carbon black on plant growth, root cellular morphogenesis, and microbial community in cadmium-contaminated soil. Environ. Sci. Pollut. Res. 2020, 27, 18423–18433. [Google Scholar] [CrossRef] [PubMed]
  293. Gahoi, P.; Omar, R.A.; Verma, N.; Gupta, G.S. Rhizobacteria and Acylated homoserine lactone-based nanobiofertilizer to improve growth and pathogen defense in Cicer arietinum and Triticum aestivum Plants. ACS Agric. Sci. Technol. 2021, 1, 240–252. [Google Scholar] [CrossRef]
  294. Joshi, A.; Sharma, A.; Nayyar, H.; Verma, G.; Dharamvir, K. Carbon nanofibers suppress fungal inhibition of seed germination of maize (Zea mays) and barley (Hordeum vulgare L.) crop. AIP Conf. Proc. 2015, 1675, 030034. [Google Scholar]
  295. Shakiba, S.; Astete, C.E.; Paudel, S.; Sabliov, C.M.; Rodrigues, D.F.; Louie, S.M. Emerging investigator series: Polymeric nanocarriers for agricultural applications: Synthesis, characterization, and environmental and biological interactions. Environ. Sci. Nano 2020, 7, 37–67. [Google Scholar] [CrossRef]
  296. Astete, C.E.; Sabliov, C.M.; Watanabe, F.; Biris, A. Ca2+ cross-linked alginic acid nanoparticles for solubilization of lipophilic natural colorants. J. Agric. Food Chem. 2009, 57, 7505–7512. [Google Scholar] [CrossRef]
  297. Goh, K.L.; Manikam, J.; Qua, C.S. High-dose rabeprazole–amoxicillin dual therapy and rabeprazole triple therapy with amoxicillin and levofloxacin for 2 weeks as first and second line rescue therapies for H elicobacter pylori treatment failures. Aliment. Pharmacol. Ther. 2012, 35, 1097–1102. [Google Scholar] [CrossRef] [PubMed]
  298. Goh, C.H.; Heng, P.W.; Chan, L.W. Alginates as a useful natural polymer for microencapsulation and therapeutic applications. Carbohydr. Polym. 2012, 88, 1–12. [Google Scholar] [CrossRef]
  299. Ye, F.; Astete, C.E.; Sabliov, C.M. Entrapment and delivery of α-tocopherol by a self-assembled, alginate-conjugated prodrug nanostructure. Food Hydrocoll. 2017, 72, 62–72. [Google Scholar] [CrossRef]
  300. Chauhan, N.; Dilbaghi, N.; Gopal, M.; Kumar, R.; Kim, K.H.; Kumar, S. Development of chitosan nanocapsules for the controlled release of hexaconazole. Int. J. Biol. Macromol. 2017, 97, 616–624. [Google Scholar] [CrossRef]
  301. Murugeshu, A.; Astete, C.; Leonardi, C.; Morgan, T.; Sabliov, C.M. Chitosan/PLGA particles for controlled release of α-tocopherol in the GI tract via oral administration. Nanomedicine 2011, 6, 1513–1528. [Google Scholar] [CrossRef]
  302. Chuacharoen, T.; Sabliov, C.M. Stability and controlled release of lutein loaded in zein nanoparticles with and without lecithin and pluronic F127 surfactants. Colloids Surf. A Physicochem. Eng. Asp. 2016, 503, 11–18. [Google Scholar] [CrossRef]
  303. Prasad, A.; Astete, C.E.; Bodoki, A.E.; Windham, M.; Bodoki, E.; Sabliov, C.M. Zein nanoparticles uptake and translocation in hydroponically grown sugar cane plants. J. Agric. Food Chem. 2017, 66, 6544–6551. [Google Scholar] [CrossRef]
  304. Ristroph, K.D.; Astete, C.E.; Bodoki, E.; Sabliov, C.M. Zein nanoparticles uptake by hydroponically grown soybean plants. Environ. Sci. Technol. 2017, 51, 14065–14071. [Google Scholar] [CrossRef]
  305. Pascoli, M.; Lopes-Oliveira, P.J.; Fraceto, L.F.; Seabra, A.B.; Oliveira, H.C. State of the art of polymeric nanoparticles as carrier systems with agricultural applications: A minireview. Energy Ecol. Environ. 2018, 3, 137–148. [Google Scholar] [CrossRef]
  306. Wang, Y.; Li, M.; Ying, J.; Shen, J.; Dou, D.; Yin, M.; Whisson, S.C.; Birch, P.R.; Yan, S.; Wang, X. High-efficiency green management of potato late blight by a self-assembled multicomponent nano-bioprotectant. Nat. Commun. 2023, 14, 5622. [Google Scholar] [CrossRef] [PubMed]
  307. Tian, Y.; Huang, Y.; Zhang, X.; Tang, G.; Gao, Y.; Zhou, Z.; Li, Y.; Wang, H.; Yu, X.; Li, X.; et al. Self-assembled nanoparticles of a prodrug conjugate based on pyrimethanil for efficient plant disease management. J. Agric. Food Chem. 2022, 70, 11901–11910. [Google Scholar] [CrossRef] [PubMed]
  308. Xiang, H.; Meng, J.; Shao, W.; Zeng, D.; Ji, J.; Wang, P.; Zhou, X.; Qi, P.; Liu, L.; Yang, S. Plant protein-based self-assembling core–shell nanocarrier for effectively controlling plant viruses: Evidence for nanoparticle delivery behavior, plant growth promotion, and plant resistance induction. Chem. Eng. J. 2023, 464, 142432. [Google Scholar] [CrossRef]
  309. Dong, B.R.; Jiang, R.; Chen, J.F.; Xiao, Y.; Lv, Z.Y.; Chen, W.S. Strategic nanoparticle-mediated plant disease resistance. Crit. Rev. Biotechnol. 2023, 43, 22–37. [Google Scholar] [CrossRef]
  310. Alghuthaymi, M.A.; Ahmad, A.; Khan, Z.; Khan, S.H.; Ahmed, F.K.; Faiz, S.; Nepovimova, E.; Kuča, K.; Abd-Elsalam, K.A. Exosome/liposome-like nanoparticles: New carriers for CRISPR genome editing in plants. Int. J. Mol. Sci. 2021, 22, 7456. [Google Scholar] [CrossRef]
  311. Hafeez, R.; Guo, J.; Ahmed, T.; Jiang, H.; Raza, M.; Shahid, M.; Ibrahim, E.; Wang, Y.; Wang, J.; Yan, C.; et al. Bio-formulated chitosan nanoparticles enhance disease resistance against rice blast by physiomorphic, transcriptional, and microbiome modulation of rice (Oryza sativa L.). Carbohydr. Polym. 2024, 334, 122023. [Google Scholar] [CrossRef]
  312. Mejdoub-Trabelsi, B.; Touihri, S.; Ammar, N.; Riahi, A.; Daami-Remadi, M. Effect of chitosan for the control of potato diseases caused by Fusarium species. J. Phytopathol. 2020, 168, 18–27. [Google Scholar] [CrossRef]
  313. Lin, M.; Fang, S.; Zhao, X.; Liang, X.; Wu, D. Natamycin-loaded zein nanoparticles stabilized by carboxymethyl chitosan: Evaluation of colloidal/chemical performance and application in postharvest treatments. Food Hydrocoll. 2020, 106, 105871. [Google Scholar] [CrossRef]
  314. Oliveira-Pinto, P.R.; Mariz-Ponte, N.; Sousa, R.M.; Torres, A.; Tavares, F.; Ribeiro, A.; Cavaco-Paulo, A.; Fernandes-Ferreira, M.; Santos, C. Satureja montana essential oil, zein nanoparticles and their combination as a biocontrol strategy to reduce bacterial spot disease on tomato plants. Horticulturae 2021, 7, 584. [Google Scholar] [CrossRef]
  315. Khairy, A.M.; Tohamy, M.R.; Zayed, M.A.; Mahmoud, S.F.; El-Tahan, A.M.; El-Saadony, M.T.; Mesiha, P.K. Eco-friendly application of nano-chitosan for controlling potato and tomato bacterial wilt. Saudi J. Biol. Sci. 2022, 29, 2199–2209. [Google Scholar] [CrossRef]
  316. El Gamal, A.Y.; Atia, M.M.; Sayed, T.E.; Abou-Zaid, M.I.; Tohamy, M.R. Antiviral activity of chitosan nanoparticles for controlling plant-infecting viruses. S. Afr. J. Sci. 2022, 118, 1–9. [Google Scholar] [CrossRef] [PubMed]
  317. Zhang, C.; Long, Y.; Li, J.; Li, M.; Xing, D.; An, H.; Wu, X.; Wu, Y. A chitosan composite film sprayed before pathogen infection effectively controls postharvest soft rot in kiwifruit. Agronomy 2020, 10, 265. [Google Scholar] [CrossRef]
  318. Kumar, N.V.; Basavegowda, V.R.; Murthy, A.N. Synthesis and characterization of copper-chitosan based nanofungicide and its induced defense responses in Fusarium wilt of banana. Inorg. Nano Met. Chem. 2022, 1–9. [Google Scholar] [CrossRef]
  319. Ghule, M.R.; Ramteke, P.K.; Ramteke, S.D.; Kodre, P.S.; Langote, A.; Gaikwad, A.V.; Holkar, S.K.; Jambhekar, H. Impact of chitosan seed treatment of fenugreek for management of root rot disease caused by Fusarium solani under in vitro and in vivo conditions. 3 Biotech 2021, 11, 290. [Google Scholar] [CrossRef]
  320. Zhang, C.; Li, Q.; Li, J.; Su, Y.; Wu, X. Chitosan as an adjuvant to enhance the control efficacy of low-dosage pyraclostrobin against powdery mildew of Rosa roxburghii and improve its photosynthesis, yield, and quality. Biomolecules 2022, 12, 1304. [Google Scholar] [CrossRef]
  321. Abdel-Rahman, F.A.; Monir, G.A.; Hassan, M.S.; Ahmed, Y.; Refaat, M.H.; Ismail, I.A.; El-Garhy, H.A. Exogenously applied chitosan and chitosan nanoparticles improved apple fruit resistance to blue mold, upregulated defense-related genes expression, and maintained fruit quality. Horticulturae 2021, 7, 224. [Google Scholar] [CrossRef]
  322. Abdelkhalek, A.; Qari, S.H.; Abu-Saied, M.A.; Khalil, A.M.; Younes, H.A.; Nehela, Y.; Behiry, S.I. Chitosan nanoparticles inactivate alfalfa mosaic virus replication and boost innate immunity in Nicotiana glutinosa plants. Plants 2021, 10, 2701. [Google Scholar] [CrossRef]
  323. Chouhan, D.; Dutta, A.; Kumar, A.; Mandal, P.; Choudhuri, C. Application of nickel chitosan nanoconjugate as an antifungal agent for combating Fusarium rot of wheat. Sci. Rep. 2022, 12, 14518. [Google Scholar] [CrossRef]
  324. Fan, Z.; Qin, Y.; Liu, S.; Xing, R.; Yu, H.; Li, K.; Li, P. Fluoroalkenyl-grafted chitosan oligosaccharide derivative: An exploration for control nematode Meloidogyne incognita. Int. J. Mol. Sci. 2022, 23, 2080. [Google Scholar] [CrossRef]
  325. Khan, A.; Tariq, M.; Ahmad, F.; Mennan, S.; Khan, F.; Asif, M.; Nadeem, H.; Ansari, T.; Shariq, M.; Siddiqui, M.A. Assessment of nematicidal efficacy of chitosan in combination with botanicals against Meloidogyne incognita on carrot. Acta Agric. Scand. Sect. B Soil. Plant Sci. 2021, 71, 225–236. [Google Scholar] [CrossRef]
  326. Xu, X.; Peng, X.; Huan, C.; Chen, J.; Meng, Y.; Fang, S. Development of natamycin-loaded zein-casein composite nanoparticles by a pH-driven method and application to postharvest fungal control on peach against Monilinia fructicola. Food Chem. 2023, 404, 134659. [Google Scholar] [CrossRef] [PubMed]
  327. Bidyarani, N.; Srivastav, A.K.; Gupta, S.K.; Kumar, U. Synthesis and physicochemical characterization of rhamnolipid-stabilized carvacrol-loaded zein nanoparticles for antimicrobial application supported by molecular docking. J. Nanoparticle Res. 2020, 22, 307. [Google Scholar] [CrossRef]
  328. Bidyarani, N.; Kumar, U. Synthesis of rotenone loaded zein nano-formulation for plant protection against pathogenic microbes. RSC Adv. 2019, 9, 40819–40826. [Google Scholar] [CrossRef] [PubMed]
  329. Khatua, A.; Prasad, A.; Priyadarshini, E.; Virmani, I.; Ghosh, L.; Paul, B.; Meena, R.; Barabadi, H.; Patel, A.K.; Saravanan, M. CTAB-PLGA Curcumin Nanoparticles: Preparation, Biophysical Characterization and Their Enhanced Antifungal Activity against Phytopathogenic Fungus Pythium ultimum. ChemistrySelect 2020, 5, 10574–10580. [Google Scholar] [CrossRef]
  330. De Angelis, G.; Simonetti, G.; Chronopoulou, L.; Orekhova, A.; Badiali, C.; Petruccelli, V.; Portoghesi, F.; D’Angeli, S.; Brasili, E.; Pasqua, G.; et al. A novel approach to control Botrytis cinerea fungal infections: Uptake and biological activity of antifungals encapsulated in nanoparticle based vectors. Sci. Rep. 2022, 12, 7989. [Google Scholar] [CrossRef] [PubMed]
  331. Han, J.; Zhao, L.; Zhu, H.; Dhanasekaran, S.; Zhang, X.; Zhang, H. Study on the effect of alginate oligosaccharide combined with Meyerozyma guilliermondii against Penicillium expansum in pears and the possible mechanisms involved. Physiol. Mol. Plant Pathol. 2021, 115, 101654. [Google Scholar] [CrossRef]
  332. Zhuo, R.; Li, B.; Tian, S. Alginate oligosaccharide improves resistance to postharvest decay and quality in kiwifruit (Actinidia deliciosa cv. Bruno). Hortic. Plant J. 2022, 8, 44–52. [Google Scholar] [CrossRef]
  333. Bouissil, S.; Guérin, C.; Roche, J.; Dubessay, P.; El Alaoui-Talibi, Z.; Pierre, G.; Michaud, P.; Mouzeyar, S.; Delattre, C.; El Modafar, C. Induction Induction of defense gene expression and the resistance of date palm to Fusarium oxysporum f. sp. albedinis in response to alginate extracted from Bifurcaria bifurcata. Mar. Drugs 2022, 20, 88. [Google Scholar] [CrossRef]
  334. Ben Salah, I.; Aghrouss, S.; Douira, A.; Aissam, S.; El Alaoui-Talibi, Z.; Filali-Maltouf, A.; El Modafar, C. Seaweed polysaccharides as bio-elicitors of natural defenses in olive trees against verticillium wilt of olive. J. Plant Interact. 2018, 13, 248–255. [Google Scholar] [CrossRef]
  335. Ngoc, D.T.; Du, B.D.; Tuan, L.N.; Thach, B.D.; Kien, C.T.; Phu, D.V.; Hien, N.Q. Study on antifungal activity and ability against rice leaf blast disease of nano Cu-Cu2O/alginate. Indian J. Agric. Res. 2020, 54, 802–806. [Google Scholar] [CrossRef]
  336. Yadav, A.; Yadav, K.; Ahmad, R.; Abd-Elsalam, K.A. Emerging frontiers in nanotechnology for precision agriculture: Advancements, hurdles and prospects. Agrochemicals 2023, 2, 220–256. [Google Scholar] [CrossRef]
  337. Hsueh, Y.H.; Ke, W.J.; Hsieh, C.T.; Lin, K.S.; Tzou, D.Y.; Chiang, C.L. ZnO nanoparticles affect Bacillus subtilis cell growth and biofilm formation. PLoS ONE 2015, 10, e0128457. [Google Scholar] [CrossRef]
  338. Rajput, V.D.; Minkina, T.; Sushkova, S.; Tsitsuashvili, V.; Mandzhieva, S.; Gorovtsov, A.; Nevidomskyaya, D.; Gromakova, N. Effect of nanoparticles on crops and soil microbial communities. J. Soils Sediments 2018, 18, 2179–2187. [Google Scholar] [CrossRef]
  339. Tang, R.; Zhu, D.; Luo, Y.; He, D.; Zhang, H.; El-Naggar, A.; Palansooriya, K.N.; Chen, K.; Yan, Y.; Lu, X.; et al. Nanoplastics induce molecular toxicity in earthworm: Integrated multi-omics, morphological, and intestinal microorganism analyses. J. Hazard. Mater. 2023, 442, 130034. [Google Scholar] [CrossRef]
  340. Tiede, K.; Hassellöv, M.; Breitbarth, E.; Chaudhry, Q.; Boxall, A.B. Considerations for environmental fate and ecotoxicity testing to support environmental risk assessments for engineered nanoparticles. J. Chromatogr. A 2009, 1216, 503–509. [Google Scholar] [CrossRef]
  341. Pietroiusti, A.; Stockmann-Juvala, H.; Lucaroni, F.; Savolainen, K. Nanomaterial exposure, toxicity, and impact on human health. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2018, 10, e1513. [Google Scholar] [CrossRef] [PubMed]
  342. Larese, F.F.; D’Agostin, F.; Crosera, M.; Adami, G.; Renzi, N.; Bovenzi, M.; Maina, G. Human skin penetration of silver nanoparticles through intact and damaged skin. Toxicology 2009, 255, 33–37. [Google Scholar] [CrossRef]
  343. Singh, N.; Manshian, B.; Jenkins, G.J.; Griffiths, S.M.; Williams, P.M.; Maffeis, T.G.; Wright, C.J.; Doak, S.H. NanoGenotoxicology: The DNA damaging potential of engineered nanomaterials. Biomaterials 2009, 30, 3891–3914. [Google Scholar] [CrossRef]
  344. Fu, P.P.; Xia, Q.; Hwang, H.M.; Ray, P.C.; Yu, H. Mechanisms of nanotoxicity: Generation of reactive oxygen species. J. Food Drug Anal. 2014, 22, 64–75. [Google Scholar] [CrossRef]
  345. Jafari, S.M.; Katouzian, I.; Akhavan, S. Safety and regulatory issues of nanocapsules. In Nanoencapsulation Technologies for the Food and Nutraceutical Industries; Elsevier: Amsterdam, The Netherlands, 2017; pp. 545–590. [Google Scholar]
Figure 1. Plant disease control methods from conventional methods to the modern era.
Figure 1. Plant disease control methods from conventional methods to the modern era.
Agronomy 14 02175 g001
Figure 2. Limitations of conventional control strategies for plant disease management.
Figure 2. Limitations of conventional control strategies for plant disease management.
Agronomy 14 02175 g002
Figure 3. Merits and demerits of chemical pesticides.
Figure 3. Merits and demerits of chemical pesticides.
Agronomy 14 02175 g003
Figure 4. Summary of the benefits and limitations of biopesticides.
Figure 4. Summary of the benefits and limitations of biopesticides.
Agronomy 14 02175 g004
Figure 5. Capabilities and challenges associated with meganucleases.
Figure 5. Capabilities and challenges associated with meganucleases.
Agronomy 14 02175 g005
Figure 6. Pros and Cons of Zinc Finger Nucleases.
Figure 6. Pros and Cons of Zinc Finger Nucleases.
Agronomy 14 02175 g006
Figure 7. Capabilities and constraints associated with TALENs.
Figure 7. Capabilities and constraints associated with TALENs.
Agronomy 14 02175 g007
Figure 8. Pros and Cons of the CRISPR Cas system.
Figure 8. Pros and Cons of the CRISPR Cas system.
Agronomy 14 02175 g008
Figure 9. Challenges and constraints associated with nanotechnology.
Figure 9. Challenges and constraints associated with nanotechnology.
Agronomy 14 02175 g009
Table 1. Applications of metal NPs in plant disease management.
Table 1. Applications of metal NPs in plant disease management.
MNPsConcentrationSource of SynthesisSize of NPs (nm)Disease & Pathogen ManagementReference
Au200 ppmBiosynthesis N/ASheath blight of rice caused by Rhizoctonia solani[159]
2 ppmTrichoderma atroviride50–75Phomopsis canker in tea pant[160]
0, 20, 40, 60, and 80 ppmBacillus sonorensis10–35 Dematophora necatrix, Fusarium oxysporum, Alternaria aternata, Alternaria mali, Sclerotium rolfsii and Colletotrichum capsici[161]
150 ppmMetarhizium anisopliae9–54Rice blast disease[162]
Zn50, 100, 250, and 500 ppmN/ADLS: 30–40
TEM: 15–20
Aspergillus niger[163]
0–65 mg/LN/A>50Peronospora tabacina[164]
Ag25 ppm, 37.5 ppm and 50 ppmN/A3 to 10 Fusarium oxysporum f. sp. radicis-lycopersici [165]
2, 4 and 10 µg/mLSerratia sp10 to 20 Spot blotch of wheat (Bipolaris sorokiniana)[166]
17.24 μg/mLAzadirachta indica15 Bakanae of rice (F. fujikuroi and
F. proliferatum)
[167]
50 ppmAzadirachta indicaN/AEarly blight of tomato (Alternaria solani)[168]
40 mg/L Allium sativum bulbN/ASpot blotch of wheat (B. sorokiniana)[169]
40 ppmAvena fatua5 to 25 Fusarium oxysporum f. sp. lycopersici [170]
40 mg/LAllium sativum bulb extractN/ABlack leg and soft rot of potato [169]
7.8 μg/mLPseudomonas canadensis21 and 52 Brown blotch of mushroom
(Pseudomonas. tolaasii.)
[171]
100 ppmPleurotus ostreatusN/AFusarium oxysporum[172]
0.35 mg/100 uLHyppophae rhamnoidesN/ARalstonia solanacearum and Pseudomonas syringae[173]
N/AMalva parviflora L.50.6Helminthosporium rostratum, Fusarium solani, Fusarium oxysporum, and Alternaria alternata.[174]
150 ppmPenicillium verrucosum10–12 F. chlamydosporum and Aspergillus flavus[175]
15 μg/mLleaf extract of rice16.5Xanthomonas oryzae pv. oryzae[176]
20 μg mLBacillus sp.22–41Red rot of sugarcane (Colletotrichum falcatum)[177]
N/APseudomonas poae19.8–44.9,Head blight of wheat (Fusarium graminearum)[178]
Cu20 ppmCuSO4
precursor
35–70 Macrophomina phaseolina, Bipolaris maydis, and Fusarium verticillioides In maize[179]
50 ppmCuSO4
precursor
35–70 Rhizoctonia solani in maize[179]
30 ppmCuSO4 precursor35–70 Erwinia carotovora and Ralstonia solanacearum in maize[179]
≥80 ppmChemical reduction of Cu2+ with reductive agent of NaHB426.5 Fusarium oxysporum[180]
500 ppm and 1000 ppmEucalyptus and
Mint leaves
10–130
23–39
Colletotrichum capsici of chilli [181]
N/APseudomonas fluorescens, Trichoderma atroviride and Streptomyces griseusN/ARed root-rot disease in tea (Poria hypolateritia), collar canker (Phomopsis theae)[182]
20 ppm
50 ppm
30 ppm
CuSO4 precursor 35–70 Macrophomina phaseolina, Bipolaris maydis, and Fusarium verticillioides.
Rhizoctonia solani.
Erwinia carotovora and Ralstonia solanacearum.
[179]
100 µg/mLBacillus altitudinis strain WM-2/229.11–78.56Bacterial fruit blotch (BLB) of watermelon (Acidovorax citrulli)[183]
N/AChemical synthesis28 Bacterial leaf blight (BLB) of rice (Xanthomonas oryzae pv. Oryzae)[184]
100, 150 and 200 ppmChemical synthesis and green synthesis126 and 85Root knot nematode (Meloidogyne Incongita)[185]
Table 2. Applications of metal oxide NPs in plant disease management.
Table 2. Applications of metal oxide NPs in plant disease management.
MONPsConcentrationSource of SynthesisSize of NPs (nm)Disease & Pathogen ManagementReference
ZnO130.1 and 104.9 µg/mLCarica papaya leaf extractN/AS. sclerotiorum[192]
10% of 3 mL/LPVP/ZnSO4 irradiated to 30 kGy.38 Black mould of pomegranate (Aspergillus niger)[193]
5% of 9 mL/L.PVP/ZnSO4 irradiated to 30 kGy.38 Green mould of orange (Penicillium digitatum)[193]
20 mg/LCinnamomum camphora13.92, 15.19 and 21.13 Early blight of tomato (Alternaria solani) [194]
18.0 µg/mLMatricaria chamomilla flower extract 8.9 to 32.6 Bacterial wilt of tomato (Ralstonia solanacearum)[195]
1.0 mg/mLTrachyspermum ammi48.52 Fruit rot (Rhizoctonia solani)[196]
100 ppm Eucalyptus globules52–70Alteraria blotch (Alternaria mali), Botryosphaeria canker of apple (Botryosphaeria dothidea) [197]
250 ppmN/AN/APurple Blotch disease in onion (Alternaria porri)[198]
100 µg/mLPicea smithiana extract25Bacterial leaf spot of tomato
Bacterial wilt of tomato
[199]
100 µg/mLTrichoderma harzianum25–60 Fusarium wilt of tomato (F. oxysporum)[200]
200 μg/mLCannabis sativa L.13.51Fusarium virguliforme in soybean[201]
100 mg/mL−1lemon peels16.8 Citrus black rot (Alternaria citri)[202]
Ag2O0.10 and 0.20 g/Lsolid homogeneous solution of silver oxide material38.23 Pseudomonas syringae pv. tomato, Xanthomonas campestris pv. vesicatoria, Pectobacterium carotovorum subsp. carotovorum, Ralstonia solanacearum, Fusarium oxysporum f. sp. lycopersici and Alternaria solani in tomato [203]
CuON/AHibiscus rosa-sinensis L. flower extract28.1 Xanthomonas oryzae pv. oryzae[204]
10 ppmcoffee powder85–100 Fusarium wilt in chickpea[205]
200 μg/mLCannabis sativa L.7.36 Fusarium virguliforme in soybean[201]
N/AChemical synthesis25.54 and 25.83 Root rot disease in cucumber (Fusarium solani)[206]
10, 15, 30, 50, 70, 100, and 150 mg/LPseudomonas fluorescens and Trichoderma viride40–100 and 20–80 Gummosis of citrus (Phytophthora parasitica)[207]
5–350 μg/mLCassia fistula12–38Fusarium wilt of tomato (Fusarium oxysporum f. sp. lycopersici)[208]
200 ppmTrichoderma asperellum22Alternaria brassicae[209]
200.0 μg/mLHibiscus rosa-sinensis L.28.1Bacterial leaf blight of rice (Xanthomonas oryzae pv. Oryzae)[204]
200 ppmJatropha curcas5 to 15 Root-knot nematode in chickpea (Meloidogyne incognita)[210]
100 mg/mL lemon peels18 Citrus black rot (Alternaria citri) [202]
SiO2100 mg/LN/A54–76P. syringae[211]
50, 100, 150, 200 and 250 ppmbioleaching of sand 22.5 Meloidogyne javanica[36]
2, 20, 200 and 2000 ppm.Green synthesis58.6 Vigna radiata L.[212]
200 µg/mLCrocus sativus L.N/ABacterial leaf blight of rice (Xanthomonas oryzae pv. Oryzae)[213]
250 to 1000 mg/kgagro-wasteN/AFusarium oxysporum (Fusarium wilt in Eruca sativa)[214]
150 ppmGreen synthesisN/APepper bacterial leaf spot (Xanthomonas vesicatoria)[215]
25, 50, and 100 µg/mLsaffron extract9.92 and 19.8 Rhizoctonia solani[216]
50 mg/LMilled/acid leaching rice husk15 Bakanae of rice (F. fujikuroi)[217]
TiO2100 and 200 mg/LN/A21Pectobacterium betavasculorum, Rhizoctonia solani, and Meloidogyne incognita in beetroot[218]
40 mg/LMoringa oleifera Lam10–100 Spot blotch of wheat (Bipolaris sorokiniana)[219]
0.20 mg/mLN/A 5–15 Phomopsis vexans, Ralstonia solanacearum, and root-knot nematode (Meloidogyne incognita) in egg plant[220]
50 µg/mLmixing of TiO2 solution with a lemon fruit extract at room temperature41.5 Soft rot of sweet potato (Dickeya dadantii)[221]
100 ppmAspergillus versicolor47.15 Leaf blight of tomato (Alternaria alternata)[222]
25, 50 and 75 μL Trianthema portulacastrum, Chenopodium quinoa and by chemical conventional (sol-gel) method<15Wheat rust (Ustillago tritici)[223]
40 mg/LMoringa oleifera<100 Stripe rust (Puccinia striiformis f. sp. tritici)[224]
150 µg/mLChemical (sol gel) synthesis20 Tobacco mosaic virus in chili pepper[225]
400 mg/LGreen synthesis by African palm oil and Chemical synthesis by sol gel method14.60 ± 0.44 and 12.30 ± 0.54Fusarium solani[226]
CeO250 and 250 mg/L N/A8 ± 1 Fusarium wilt (Fusarium oxysporum f. sp. lycopersici)[227]
30 mg/LAcorus calamusas rhizomes42 Wheat stripe rust (Puccinia striformis)[228]
100 mg/L Chenopodium quinoa L.7–10 Ustilago tritici in wheat[229]
Fe2O320 µg/mLGreen synthesis by hydrothermal processHRTEM: 5 ± 1.0 DLS: 7.5
XRD: 5.95
Fusarium wilt (F. oxysporum)[230]
1.0 mg/mL Green synthesis49 Fusarium fruit rot (Fusarium oxysporum)[231]
1.0 mg/mLTrichoderma harzianum17.78 brown rot of apple (Fusarium oxysporum)[232]
0, 10, 50, 250 and 500
6.2 μg/mL
Mentha spicata21–26 E. coli and B. cereus[233]
6 mg/mLSkimmia laureola leaf extract56–350 Bacterial wilt (Ralstonia solanacearum)[234]
1.0 mg/mLAzadirachta indica24 brown rot of sweet oranges (Fusarium oxysporum)[235]
100 and 200 ppmThyme plantN/AGray mold of strawberry (Botrytis cinerea)[236]
1.0 mg/mLCalotropis procera49 Fusarium fruit rot of loquat [231]
15, 10 and 5 mg/LDried Guava29–41 Alternaria solani[237]
1.0 mg/mLCalotropis procera32 Fruit rot of cherry (Aspergillus flavus)[238]
Al2O31, 6 and 50 mg/100 mLPepsi cans4–10 A. flavus, Fusarium sp. and Alternaria sp.[239]
150 mg/mLColletotrichum sp.39 ± 35 F. oxysporum[240]
CoO200 µg/mLHibiscus rosa sinensis flower34.9 Bacterial leaf blight of rice (Xanthomonas oryzae pv. oryzae)[241]
Fe3O40.01–15 μg/mLSpinach20 Fusarium Wilt of tomato (Fusarium oxysporum)[242]
10 µg/mLSpinacia oleracea4 Fusarium wilt of tomato[208]
1000–1400 ppm N/A50 ± 5 Onion white rot (Sclerotium cepivorum)[243]
10, 40 and 80 mg/LChemical co-precipitation synthesis60 to 72 Acremonium Wilt of sorghum (Acremonium striticum)[244]
MgO100 μg/mLGreen synthesis 29–60Rhizoctonia solani, Acidovorax oryzae[245]
4 μg/mLAqueous Rosemary extract<20Xanthomonas oryzae pv. oryzae[246]
16.0 μg/mLPaenibacillus polymyxa10.9Xanthomonas oryzae oryzae[247]
75, 150, 300, and 500 μg/mLStrawberry 100 Root-knot nematode (Meloidogynidae)[233]
3 mg/mLChemical synthesisN/ABrown rot of potato (R. solanacearum)[248]
50 and 100 mg/LLemon fruit extractsN/AAlternaria leaf blight of carrot[249]
20 mg/mLAcinetobacter johnsonii strain RTN118 to 45 Acidovorax oryzae[245]
15.36 μg/mLBurkholderia rinojensis26.70 Fusarium oxysporum f. sp. lycopersici[250]
200 ppmS. cerevisiae27Callosobruchus maculatus[251]
50 μg/mLGreen synthesisN/APhytophthora infestans[252]
75, 150 and 200 mg/LN/A52.5 to 57.3 Black scurf of potato (Rhizoctonia solani)[253]
74.81, 82.94, and 91.19 mg/g Magnesium nitrate hexahydrates precursor52.97 ± 1.43 Powdery mildew of peppers (Oidiopsis sicula)[254]
500, 1500, 2500 mg/LMagnesium nitrate hexahydrate precursor21.8 Clubroot caused by Plasmodiophora brassica[255]
79.43 ppmAlcoholic extract of the bark of the walnut tree28.55 Thielaviopsis paradoxa and Thielaviopsis punctulata[256]
100, 200, and/or 1000 μg/mLN/A20Bacterial spot of tomato (Xanthomonas perforans)[257]
Table 3. Applications of organic NPs in plant disease management.
Table 3. Applications of organic NPs in plant disease management.
Organic NPsConcentrationNanocompositeDisease & Pathogen ManagementRerefernces
Micelles0.013 to 0.042 mg/mLCarboxymethylchitosan (CMCS) micellesSmart delivery of agrochemicals[259]
N/AHexaconazole/dazomet-micelle Bio-fungicidal activity against Ganoderma boninense[260]
30 mg/LLinear Supramolecular Block Copolymer MicellesRhizoctonia solani[261]
10, 20, 40, and 80 μL/LHumidity-Responsive Cinnamon Essential Oil Nano micellesHigh antifungal activity against Botrytis cinerea and nano-vesicles for preservation of fruit or vegetable[262]
Liposomes10 µg/mLLiposomes bounded amphotericinF. oxysporum f. sp. ciceris in chickpea[263]
0, 1 and 3 g/LTea tree oil solid liposomes (TTO-SLPs)Brown rot of peach fruit caused by Monilinia fructicola[264]
0.046 mg/LNano-Insecticide through Encapsulation of insecticides in Polymeric LiposomesFall armyworm Spodoptera frugiperda[265]
136.59 and 83.99 mg/L,
315.78 and 154.34 mg/L
Eleocharis dulcis peel extract (EDPE) nanoliposomes Megoura crassicauda and Acyrthosiphon pisum[266]
Table 4. Characteristics of Organic Nanoparticles in Disease Management.
Table 4. Characteristics of Organic Nanoparticles in Disease Management.
Organic NPsConcentrationSize (nm)Disease & Pathogen ManagementReference
Dendrimers24 μg 1.1, 1.8, and 3.2 Cotton bollworm cells and larvae (H. armigera)[267]
500, 1000, 2000 and 5000 ppm 20 to 30 Phytophthora infestans[268]
FerritinN/AN/AChanges in the regulation of iron homeostasis are involved in increasing resistance to Common scab caused by Streptomyces scabies[269]
N/AN/AEnhanced resistance against fire blight of pear caused by Erwinia amylovora[270]
Table 5. Properties of Nanoparticles in Disease and Pathogen Management.
Table 5. Properties of Nanoparticles in Disease and Pathogen Management.
ConcentrationSourceSizeDisease & Pathogen ManagementReference
Carbon nanotubes100 mg/LN/A30–50 nmStem and fruit rot and leaf blight of tomato caused by Alternaria solani[279]
200 mg/LN/A20–30 nmPowdery mildew of roses caused by Podosphaera pannosa[280]
19 and 23 mg/mL.Pulsed laser ablation in liquid (PLAL)23 nmFusarium oxysporum[281]
100, 200 and 500 mg/LN/A20–30 nmTobacco mosaic virus in Nicotiana benthamiana[282]
100 and 500 mg/LChemical synthesis52 ± 1.2 nmStalk rot caused by Fusarium verticillioides in Maize[283]
Fullerenes100, 200 and 500 mg/LN/A50 nmTobacco mosaic virus in Nicotiana benthamiana[282,284]
100 mg/LN/A50 ± 5 nmCucurbit Chlorotic Yellows Virus (CCYV) Infecting Nicotiana benthamiana[284]
Table 6. Properties of Nanocomposites in Disease and Pathogen Management.
Table 6. Properties of Nanocomposites in Disease and Pathogen Management.
ConcentrationNanocompositeSizeDisease & Pathogen ManagementReference
Graphene150 μg/mLReduced graphene oxide/silver nanocomposite (rGO-Ag)7–26 nmChocolate spot disease of broad bean caused by Botrytis fabae [285]
200 mg/LReduced graphene oxide copper oxide (rGO-CuO)0.55 to 3.74 nmPowdery mildew of roses caused by Podosphaera pannosa[280]
1 mg/LReduced Graphene Oxide Nanosheet-Decorated Copper Oxide (rGO-CuO)5, 20 and 50 nmFusarium wilt and root rot [286]
50 and 500 µg/mLReduced graphene oxide based Cu and Ag NPs (rGO-Cu/Ag)SEM: 2.4 nm
EDS: 40 nm
Bacterial spot of tomato and pepper caused by Xanthomonas euvesicatoria[287]
1280 μg/mLBi2O3/TiO2@reduced graphene oxide (rGO)N/APseudomonas syringae tomato[288]
N/AGraphene quantum dots (GQD)2–5 nmFusariusm head blight of wheat caused by Fusarium graminearum[289,290]
50 and 250 μg/mLGraphene oxide-Fe3O4 nanocomposites (GO- Fe3O4)30–36nmDowny mildew of grapevine (Plasmopara viticola)[291]
Table 7. Impact of Nanoparticle Types on Crop Performance.
Table 7. Impact of Nanoparticle Types on Crop Performance.
NP TypeCrop EffectReference
Carbon BlackModified nanoscale carbon black (MCB)Ryegrass and chardReduction of heavy metals, increased plant growth and enhanced microbial communities[292]
Carbon nanofibersAcylated homoserine-coated iron-carbon nanofibersChickpeaSuppression of Fusarium oxyssporum f. sp. ciceris[292]
Carbon nanofiber
CNF-Cu
ChicoryImproved water absorption, germination rate, shoot and root ratio and protein content[289]
Acylated homoserine lactone coated-iron carbon nanofiber (AHL/Fe-CNF)Cicer arietinum and Triticum aestivumFusarium wilt of chickpea and root rot of wheat caused by Fusarium oxysporum f. sp. ciceris and Cochliobolus sativus[293]
Carbon nanofibers (CNFs)Maize and barleyResistance against fungal diseases and enhanced seed germination[294]
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

Akhtar, H.; Usman, M.; Binyamin, R.; Hameed, A.; Arshad, S.F.; Aslam, H.M.U.; Khan, I.A.; Abbas, M.; Zaki, H.E.M.; Ondrasek, G.; et al. Traditional Strategies and Cutting-Edge Technologies Used for Plant Disease Management: A Comprehensive Overview. Agronomy 2024, 14, 2175. https://doi.org/10.3390/agronomy14092175

AMA Style

Akhtar H, Usman M, Binyamin R, Hameed A, Arshad SF, Aslam HMU, Khan IA, Abbas M, Zaki HEM, Ondrasek G, et al. Traditional Strategies and Cutting-Edge Technologies Used for Plant Disease Management: A Comprehensive Overview. Agronomy. 2024; 14(9):2175. https://doi.org/10.3390/agronomy14092175

Chicago/Turabian Style

Akhtar, Hira, Muhammad Usman, Rana Binyamin, Akhtar Hameed, Sarmad Frogh Arshad, Hafiz Muhammad Usman Aslam, Imran Ahmad Khan, Manzar Abbas, Haitham E. M. Zaki, Gabrijel Ondrasek, and et al. 2024. "Traditional Strategies and Cutting-Edge Technologies Used for Plant Disease Management: A Comprehensive Overview" Agronomy 14, no. 9: 2175. https://doi.org/10.3390/agronomy14092175

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