Next Article in Journal
Tap Maize Yield Productivity in China: A Meta-Analysis of Agronomic Measures and Planting Density Optimization
Previous Article in Journal
The Effect of Low Temperature and Low Illumination Intensity on the Photosynthetic Characteristics and Antioxidant Enzyme Activity in the Strawberry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress and Opportunities of In Planta and Topical RNAi for the Biotechnological Control of Agricultural Pests

by
Marcos Fernando Basso
1,2,3,*,
Daniel David Noriega Vásquez
1,3,
Eduardo Romano Campos-Pinto
1,3,
Daniele Heloísa Pinheiro
1,3,
Bread Cruz
4,
Grazielle Celeste Maktura
5,6,
Giovanna Vieira Guidelli
5,6,
Henrique Marques-Souza
5,6 and
Maria Fatima Grossi-de-Sa
1,3,7,8,*
1
Embrapa Genetic Resources and Biotechnology, Brasília 70770-917, Brazil
2
Department of Biology, University of Florence, Sesto Fiorentino, 50019 Florence, Italy
3
National Institute of Science and Technology, INCT PlantStress Biotech, EMBRAPA, Brasília 70770-917, Brazil
4
Innovation Agency INOVA-UNICAMP, University of Campinas (UNICAMP), Campinas 13083-970, Brazil
5
Department of Biochemistry and Tissue Biology, Institute of Biology, University of Campinas (UNICAMP), Campinas 13083-862, Brazil
6
Graduate Program in Genetics and Molecular Biology, Institute of Biology, University of Campinas, Campinas 13083-862, Brazil
7
Graduate Program in Genomic Sciences and Biotechnology, Catholic University of Brasília, Brasília 71966-700, Brazil
8
Graduate Program in Biotechnology, Catholic University Dom Bosco, Campo Grande 79117-010, Brazil
*
Authors to whom correspondence should be addressed.
Agronomy 2025, 15(4), 859; https://doi.org/10.3390/agronomy15040859
Submission received: 5 February 2025 / Revised: 26 March 2025 / Accepted: 26 March 2025 / Published: 29 March 2025
(This article belongs to the Special Issue Plant–Microbe–Arthropod Pest Interactions in Agroecosystems)

Abstract

:
In planta RNAi or host-induced gene silencing (HIGS) has undergone significant advancements that have rendered it efficient and stable at the transgenerational level in plants for regulating host genes and targeting genes of insect pests and plant pathogens. Similarly, topical RNAi or spray-induced gene silencing (SIGS) has garnered considerable attention as an environmentally sustainable, selective, and alternative approach to chemical control of insect pests and plant pathogens. Several biotechnology companies and startups have focused their efforts on RNAi-based solutions for topical application in agriculture. Nevertheless, further technological advancements are required to enhance the efficacy of topical RNAi in agriculture, including improved dsRNA delivery systems, better target gene selection, and addressing biosafety regulatory issues. Herein, this review discusses key advances and bottlenecks in RNAi, and summarizes successful applications of these RNAi-based technologies in agriculture focusing on in planta and topical RNAi to control insect pests and plant pathogens. Furthermore, this review delves into the patenting landscape, biosafety considerations, risk evaluations, and the current regulatory status of RNAi in Latin America. Finally, it explores the contributions of RNAi to plant science, food production, and fostering a more sustainable form of agriculture.

1. Introduction

The RNA interference (RNAi) pathway has been explored in plants for proof-of-concept and biotechnological purposes (Figure 1). Initially, in transgenic plants (in planta or host-induced gene silencing, HIGS), the dsRNA expression was exploited to validate the biological function of genes and subsequently used for down-regulation of endogenous genes to achieve desired agronomic traits [1]. Virus-induced gene silencing (VIGS) has emerged as an alternative to transgenesis for proof-of-concept purposes [2]. Both in planta RNAi and VIGS have enabled the development of virus-resistant plants by expressing virus-specific dsRNAs to target viral genomes or their mRNAs [3,4]. Later, in planta RNAi was employed to confer resistance to other plant pathogens [5], insect pests [6], plant-parasitic nematodes [7,8], fungi [9], and oomycetes [10] (Figure 2). From this consolidation, transgene-free RNAi (topical or spray-induced gene silencing, SIGS) was established [11]. The first study showcasing the efficacy of topical RNAi delivery used Tobacco mosaic virus (TMV) RNA encapsulated in liposomes to deliver dsRNA molecules to tobacco protoplasts [12]. Despite these advancements, topical RNAi still encounters significant challenges; however, this approach aligns with the current pursuit of eco-sustainable and more selective agriculture technologies [13,14]. Notably, economic losses attributed to insect pests and plant pathogens in different crops have increased over the years, resulting in heightened agrochemical usage [15,16]. While considerable advancements have been achieved with in planta and topical RNAi, there are still substantial obstacles to overcome before making RNAi-based products more accessible to farmers [17]. The biosafety regulation of these RNAi-based technologies for commercial use remains a topic of discussion in Latin America, the USA, and Europe [18,19]. Herein, this review study presents advancements in (i) in planta and (ii) topical RNAi, (iii) intellectual property related to RNAi, (iv) biosafety and risk assessment of RNAi-based technologies, and (v) the major contributions of RNAi to plant science.

2. In Planta RNAi

2.1. Establishment of In Planta RNAi

In planta RNAi is based on the transgenic overexpression of a dsRNA with high sequence homology to the endogenous or exogenous mRNAs, triggering post-transcriptional gene expression (PTGS) [8,20]. This overexpression must be engineered under the control of constitutive, stress-induced, and tissue-specific promoters, making this choice fundamental for regulating the activation of the RNAi machinery depending on the type of expected outcome [21,22]. This overexpression results in a primary transcript that is further processed by Dicer-like enzymes (DCLs) to generate small interfering RNA (siRNA) [23]. Initially, the overexpression of a full-length gene sequence was utilized to induce RNAi [24]. Subsequently, it was discovered that overexpression of a shorter antisense RNA is adequate to trigger RNAi [25]. Therefore, antisense RNA expression has undergone considerable optimization to enhance PTGS and its transgenerational stability in transgenic plants [23]. The key advancement was determining the minimum length of an antisense RNA (200–400 nucleotides) capable of triggering RNAi, effectively down-regulating the target gene without causing off-target effects. Subsequently, it was shown that simultaneous overexpression of sense and antisense RNA results in increased siRNA production and more efficient PTGS [26]. However, superior RNAi responses have been achieved by overexpression of a sense and antisense RNA separated by a spacer sequence [27,28]. While these in planta RNAi proof-of-concept studies have led to high RNAi efficiency, concerns regarding the transgenerational stability of this technology still exist [6,29,30]. The transgene methylation induced by transcriptional gene silencing (TGS), the reduced transcription of RNAi-inducing RNA, and the decline in PTGS efficiency over successive generations in transgenic plants have been reported [31,32]. The overexpression of sense and antisense RNA driven by a constitutive promoter results in continuous siRNA production, which can eventually lead to TGS through transgene methylation across generations [23]. Hence, it is advisable to utilize tissue-specific promoters with targeted activity [22,33], cognate promoters [34], and stress-induced promoters [35]. The RNA structure has also been refined to ensure enhanced stability of the dsRNA, minimizing premature degradation by nucleases, promoting transgenerational stability, and augmenting PTGS efficiency [6,30,36].
Fitch et al. [37] developed transgenic papayas resistant to the Papaya ringspot virus by overexpressing the CP gene, resulting in high resistance against this virus [38]. This study served as a key reference for the biotechnological purposes of in planta RNAi. Likewise, Bonfim et al. [3] developed transgenic common beans by overexpression of sense and antisense RNAs, spaced by a PDK intron, capable of down-regulating the AC1 gene of Bean golden mosaic virus (BGMV). These transgenic beans exhibited high resistance to this virus and demonstrated transgenerational stability under field conditions [39]. This sustained transgenerational stability can be attributed to the RNAi mechanism being activated during plant infection by BGMV. Head et al. [40] further showed that transgenic maize, constitutively overexpressing a sense and antisense RNA separated by an HSP70 intron, effectively down-regulates the DvSnf7 gene of Diabrotica virgifera virgifera. This transgenic maize (event MON-87411-9) is commercially available in various countries and has proven transgenerational stability [41]. Recently, Zhang et al. [30] introduced nucleotide mismatches between the sense and antisense RNA to prevent the self-TGS. Specifically, it has been shown that replacing all cytosines with thymines maintains the hairpin RNA (hpRNA)structure and enhances the transgenerational stability of dsRNA [30]. Additionally, Hunter and Wintermantel [36], enhanced dsRNA stability and increased insect mortality by modifying the 2′-F pyrimidine nucleotides (2′-F-Uridine and 2′-F-Cytosine).
In addition to transgenerational instability, premature dsRNA processing by plant DCLs also reduces the efficiency of RNAi for pest control [42]. This premature dsRNA processing can diminish RNAi efficacy particularly when pests are unable or inefficient in uptaking siRNAs rather than dsRNA [43]. One strategy to address this challenge involves modifying the secondary structure of dsRNA to render these molecules inaccessible to plant machinery [44,45]. Abbasi et al. [46] showed that “paperclip”-structured RNAs enter cells through a clathrin-independent pathway, enhancing PTGS, while the uptake of siRNAs, short hpRNAs, and long dsRNA was suppressed. Similarly, CSIRO researchers have developed the LedRNA strategy, which facilitates high siRNA accumulation and diminishes transgene self-silencing, thereby enhancing RNAi effectiveness (WO2019051563; WO2021022325A1). Furthermore, specific structured dsRNAs may also remain inaccessible to the RNAi machinery of pests. Ribeiro et al. [6] showed that overexpression of a dsRNA resembling a viroid-like architecture in the chloroplast increased dsRNA stability in transgenic plants and enhanced its accessibility to insects. Therefore, structured dsRNA can enhance stability within the plant and make it more accessible to pests [6].

2.2. In Planta RNAi for Insect Pest Control

In addition to the stability of dsRNA in plants, the expression of sufficient amounts of dsRNA to achieve efficient uptake and PTGS in insect pests remains a major challenge [6]. Nucleases present in the insect gut degrade dsRNAs even before they enter the cells [47]. By targeting the genes encoding nucleases with the dsRNA, it was possible to enhance the RNAi efficiency in insect pests [48]. Moreover, the co-overexpression of dsRNAs and chimeric PTD-DRBD proteins containing binding domains to this dsRNA improved the efficiency of ribonucleoprotein internalization across the insect cell membrane [49]. The chloroplast-specific transgenic overexpression of dsRNAs triggered by chloroplast-specific promoters also provides greater stability to the dsRNA, as PTGS does not occur in this organelle [50,51,52]. Additionally, chloroplast-triggered overexpression of dsRNA can bypass these obstacles by accumulating high levels of protected dsRNA [53,54]. Although this strategy is not yet commonly employed due to the challenges associated with plastid transformation, transplastomic plants are being developed to combat both chewing and/or sap-sucking insect pests, such as Leptinotarsa decemlineata [42], Helicoverpa zea [55], Myzus persicae [56], Bemisia tabaci [57], Frankliniella occidentalis [51], and Manduca sexta [58]. The length and stability of the dsRNA directly impact its accumulation in chloroplasts and the efficacy of insect control [57,59]. For example, 60-nucleotide dsRNA induced a weak RNAi response in L. decemlineata, whereas 200-nucleotide dsRNAs resulted in higher insect mortality rates [50,60]. The target gene selection in insect pests also influences the effectiveness of RNAi [61], while various factors combined can contribute to the ineffectiveness of transplastomic plants in controlling insect pests [57,59]. Furthermore, the evaluation of transplastomic plants overexpressing self-replicating dsRNAs for enhanced accumulation is still pending [62].
Conversely, nuclear transgenic plants have been successfully developed to control insect pests, including D. virgifera virgifera [63], Helicoverpa armigera [64], Adelphocoris suturalis [65], Sitobion avenae [66], and Chilo suppressalis [67]. SmartStax PRO maize overexpressing a dsRNA to target the DvSnf7 gene of D. virgifera virgifera, in addition to Cry toxin, was the first commercially available RNAi-based crop to control insect pests [40]. As stated previously, the primary bottleneck impeding the efficiency of nuclear transgenic plants is the early processing of dsRNA within plant cells [42]. This issue is particularly troublesome as most species possess mechanisms for the uptake of long dsRNAs and are unable to spread siRNAs systemically [68,69]. The assertive choice of appropriate target genes in insect pests is crucial for the effectiveness of RNAi. Efficient insect control is usually achieved by targeting genes linked to a lethal phenotype or disrupting insect viability [70,71]. For instance, some potential targets for RNAi in insect pests include ATPases, proteases, kinases, ABC transporters, vitellogenin, AgraRelish, chitin synthase, laccase, Distal-less, and ecdysis-triggering hormone receptors [6,70,72,73]. Integrated approaches using multiple target genes, or strategies, such as RNAi combined with entomotoxic proteins, can reduce plant susceptibility to insect pests [74,75]. Moreover, when choosing the target gene region, it is crucial to ensure both the efficacy of RNAi in the target insect and the absence of off-target effects in non-target organisms [76].
Another challenge of in planta RNAi for controlling insect pests is the necessity to adhere to stringent regulatory frameworks [77]. In turn, in planta RNAi offers the advantage of continuous dsRNA expression, providing constant exposure to the target insect pest [78]. Consequently, this approach might be favored for its practicality in field applications. In contrast, developing transgenic plants could be more time-consuming, labor-intensive, and costly [79]. Examples of in planta RNAi applications for insect pest management and an overview of dsRNA delivery methods, RNAi strategies, targeted insect species, genes, crops, and resulting plant phenotypes are detailed in Table S1.

2.3. In Planta RNAi to Control Plant-Parasitic Nematodes

In planta RNAi studies with nematodes have provided promising results, particularly when targeting genes encoding effector proteins [8,80,81,82]. Targeting a single effector gene was sufficient to decrease plant susceptibility by 70–80% [8]. Comparatively, in planta RNAi directed at other functional genes proved effective in managing plant-parasitic nematodes but it demonstrated lower efficacy compared to effector genes [7]. Therefore, several validated genes in model plants have emerged as targets for RNAi in crops [83]. These studies suggest that pyramiding dsRNAs to target different essential genes could be an alternative for durable plant resistance under field conditions [84]. However, the main barrier to its commercial use in main crops is the fact that it is regulated as a transgenic. In planta RNAi on rootstocks could be a good strategy for fruit and vegetables propagated by grafting to control nematodes and reduce public concern regarding to transgenic products [85]. Furthermore, the additional cost of intellectual property is another factor that hampers the adoption of in planta RNAi to control nematodes in crops [86]. Examples of in planta RNAi applications for nematode control, along with an overview of dsRNA delivery methods, RNAi strategies, nematode species, target genes, crops, and plant phenotypes, are detailed in Table S2.

2.4. In Planta RNAi to Control Other Important Plant Pathogens

In planta RNAi has shown significant potential in managing viruses, subviral agents, fungi, and oomycetes (Tables S3 and S4). Notably, viral agents, which depend entirely on the plant cell, display high sensitivity to the plant’s RNAi machinery [87]. Numerous studies employing RNAi targeting different or multiple genes and genomes of RNA and DNA viruses have underscored the effectiveness of this technology in crops [88,89]. Similarly, several target genes have been identified in fungi and oomycetes for in planta RNAi targeting, resulting in pathogen suppression and reduced plant susceptibility [11,90,91]. In this context, the selection of the target gene in the specific pathogen is decisive for the effectiveness of RNAi. Therefore, in planta RNAi has proven to significantly decrease crop susceptibility to plant pathogens. However, the utilization of in planta RNAi in crops has been limited primarily due to transgenic regulations and the current practicality of implementing topical RNA. Nonetheless, both RNAi approaches have been deemed highly promising and environmentally sustainable alternatives to synthetic chemical fungicides [14,92]. The primary advantage of in planta RNAi lies is the steady and continual availability of dsRNA in crops compared to topical RNAi, which requires regular applications that may face challenges due to the instability and limited uptake of dsRNAs by these pathogens [93]. Conversely, in planta RNAi-based technologies for controlling bacteria and phytoplasmas remain limited, as these pathogens lack the RNAi machinery [94]. Nevertheless, in planta RNAi has been effectively utilized to down-regulate plant genes associated with host susceptibility, enhancing plant resistance to these pathogens [5]. Examples of in planta RNAi for control of plant viruses, subviral agents, fungi, and oomycetes, as well as an overview of the RNA delivery methods, RNAi strategies, pathogens targeted by RNAi, target genes, crops, and plant phenotypes are outlined in Tables S3 and S4.

2.5. Alternative Methods for In Planta RNAi

Symbiont microorganisms engineered to overexpress dsRNAs to cross-act on target genes of insect pests have been developed, such as Saccharomyces cerevisiae and Bacillus thuringiensis to target Drosophila suzukii, Spodoptera littoralis, and Plutella xylostella [95,96,97]. Similarly, engineered viruses have been utilized as viral vectors (virus-induced gene silencing) to suppress plant genes for proofs of concept or pest genes for biocontrol [98,99,100]. Concerns regarding transgenesis are still under regulatory consideration for the commercial deployment of these approaches. Artificial miRNAs (amiRNAs), which provide enhanced efficacy and specificity for the down-regulation of target genes offer a valuable alternative to control insect pests and pathogens but also require plant genetic engineering [20,101]. Furthermore, the stable knockout of plant genes using genome editing through CRISPR/Cas9 technology by Non-Homologous End-Joining (NHEJ or SDN1) or equivalent approaches has been reported to be relevant for commercial purposes [102,103,104]. This relevance occurs because it allows overcoming transgenic restrictions in certain countries [102,105]. In this context, both in planta RNAi and NHEJ/SDN1-mediated genome editing have been employed to down-regulate or knockout genes linked to plant susceptibility to multiple stresses, or genes encoding negative regulators of other proteins associated with increased plant resistance [106,107]. Meanwhile, genome editing of pests to knockout target genes is not yet available as an open field control measure in agriculture.

3. Topical RNAi

3.1. Establishment of Topical RNAi

Topical RNAi is a promising approach to down-regulate genes in plants, insect pests, and plant pathogens. Induction of RNAi might be achieved by ingestion of dsRNA sprayed on artificial diets or leaves and topical spray of dsRNA directly on the target organisms [17,71]. Previous studies have identified several target genes associated with lethal or deleterious effects in these organisms that can be used in topical RNAi (Tables S1–S4) [108,109,110]. The primary challenges include dsRNA instability when exposed to adverse environmental conditions, high costs for dsRNA production, low dsRNA uptake by plants or pests, and the substantial dsRNA dosages required to induce knockdown in some species [7,111]. Furthermore, factors such as leaf wettability, hydrophobic cuticle, wax layers, trichomes, cell wall barriers, and stomatal aperture and closure variations, which can all impede the efficiency of dsRNA uptake by plants, need to be considered as barriers [112,113]. Therefore, the selection of target genes and optimization of dsRNA delivery systems represent crucial steps towards achieving high control rates using minimal doses of dsRNA [114]. In contrast, topical RNAi can be applied to a wide range of organisms and can be associated with biological control and synthetic agrochemicals for crop protection [115]. It has also gained prominence in the last five years as an alternative to address the current demands for eco-sustainable technologies and the need to reduce excessive applications of non-selective agrochemicals to control pests [14]. Furthermore, it offers an important regulatory advantage by not relying on transgenic-based delivery and allowing high target specificity [17,77]. However, transforming proof-of-concept results using topical RNAi into commercial products has encountered significant challenges [116]. In this context, biotechnology companies have focused on the research and development of these topical RNAi-based products to increase efficacy, reduce costs, and streamline biosafety regulatory issues [117].

3.2. Large-Scale dsRNA Production

Topical RNAi initially involved the in vitro production of RNA in sense or antisense orientation and its delivery to target organisms through spraying, ingestion, or injection at the laboratory scale [113,118]. However, the low dose and rapid degradation of RNA highlighted the need for advancements in the stability of these molecules [119]. Consequently, dsRNA molecules have emerged as RNAi triggers because they are more effective and stable compared to sense or antisense RNA [120]. Therefore, topical RNAi depends on the large-scale production of dsRNA, either by cell-free synthesis or cell-dependent production [118]. The dsRNA production in microorganism systems such as yeast, bacteriophage, Bacillus, Pseudomonas, Corynebacterium, and Escherichia coli has been demonstrated as the most cost-effective method [118,121]. E. coli co-expressing a viral capsid protein has proven to be effective by forming complexes between the dsRNA and viral protein, protecting against degradation by nucleases [122]. However, this alternative had limited commercial application due to concerns regarding the potential viral hetero-encapsidation [123]. Given the challenges in obtaining systems that provide stability and protection for dsRNA, this approach has again been revisited [124]. Furthermore, compartmentalized dsRNA in E. coli itself has been successfully tested, facilitated by the use of an RNase III-deficient strain [119]. Even so, the cost of large-scale production, the amount of dsRNA required to be delivered to RNAi target cells, and co-contamination with other bacterial RNA/dsRNA that overload the RNAi pathway are still factors to be overcome [118]. Therefore, producing sufficient quantities of highly intact and pure dsRNA at a lower cost remains a challenge that persists to date [115].

3.3. Nanocarriers for Topical Delivery of dsRNA

The highly variable sensitivity to the RNAi and ability to access exogenous dsRNAs (such as sap-sucking and chewing insects) among target organisms pose a constraint that hampers the widespread application of topical RNAi-based products [125,126,127]. Therefore, topical RNAi often requires the formulation of dsRNAs and adjuvant molecules, ensuring their functional stability until delivered to the target cell [17]. The use of nanocarriers is considered one of the most promising approaches nowadays. Nanomaterials act not only as stabilizers, protecting dsRNA molecules against degrading agents but also as transporters, facilitating the delivery of dsRNA into target cells. Several nanocarriers improve the RNAi efficiency enhacing dsRNA stability, delivery, uptake, and endosomal escape in target organisms, [17,113,128,129]. Furthermore, nanocarriers that exhibit no harmful effects on RNA and non-target organisms, biodegradability, cellular compatibility to improve dsRNA uptake and processing into siRNAs, and low production costs have been preferred [113,130].
Most nanocarriers used to protect and deliver RNAi molecules follow the same principle of complexation, which is based on the electrostatic interactions between the negatively charged phosphate backbone of RNA molecules and cationic groups in the carrier molecule [131]. These interactions are strong enough to shield RNA molecules from most enzymes and compounds that could interact with them and lead to their degradation, while also being weak enough, as compared to covalent interactions, to release the RNA within the cell to trigger the RNAi mechanism [132]. Interestingly, protection at the extracellular level is not the sole advantage of using nanocarriers. It has been recently demonstrated to light that nanocarriers are uptaken by cells through the clathrin-mediated endocytosis mechanism and facilitate endosomal escape, ultimately preventing the dsRNA/siRNA degradation by the lysosome [128,133]. MgAl-layered double hydroxides [134], chitosan [135], E. coli-derived anucleated minicells [136], microvesicles such as exosomes [137,138,139,140,141], star polycations [128,142], clay nanostructures [143], DNA nanostructures [144], silica [145], liposomes [146], carbon dots such as polyethyleneimine [147,148], gold nanocluster [149], single-walled carbon nanotubes [150], and cell-penetrating peptides [49,151] are among the main examples of nanocarriers.
One of the most studied nanocarriers is chitosan, which has been shown to effectively enhance the RNAi effect in pests [135,152,153]. In addition, chitosan is easily degraded in the environment and has low toxicity levels when accumulated in mammal cells [154,155]. Hence, it is considered one of the most suitable nanocarriers for dsRNA delivery in agricultural systems. Dendrimers, such as star polycations, are also cheap to produce with the advantage, compared to chitosan, that the flexibility of their architecture allows them to easily modify their size, branching density, and surface functionality [156]. This is particularly interesting for topical delivery, in which interaction with different types of plant surfaces requires nanocarriers with different characteristics to penetrate a certain tissue, facilitate adsorption, and endure weather conditions. The major challenge for the development of cationic polymers is their variable stability due to the capability of most polymers to interact with molecules present in the target pests or their hosts, which can lead to the premature disassembly of nanoformulations. On the other hand, molecules that naturally occur inside cells, such as lipids, peptides, and proteins, also have the potential to deliver dsRNAs due to their biodegradable composition and high stability [157,158,159,160,161,162,163,164]. However, the main limitations of these molecules are related to production costs and eventual cytotoxicity effects [165,166]. Furthermore, chemical modifications of nanocarriers [167,168], dsRNA/nanocarrier delivery in association with surfactants to increase uptake [169], and conjugation with other chemicals to increase stability [152,170] are some alternatives to improve certain nanoformulations. Therefore, each of these nanocarriers has particular characteristics and variable efficacy for each delivery system and target pest (Table S5), from enhancing the RNAi effect to eliciting the basal defense of plants [171,172,173,174].
Despite the wide variety of nanomaterials currently available, some parameters have become standard measurements of the potential efficacy of a given nanoformulation. Ideally, nanocarriers smaller than 50 nm have been preferred as they offer a high surface-to-volume ratio, facilitating effective dsRNA binding, and are small enough to be uptaken by the plant cell wall and membrane of target organisms [143,175]. In addition, water-soluble nanocarriers with a positively charged core and surface have been preferred by facilitating dsRNA assembly [143]. Likewise, nanocarriers with a zeta potential greater than +25 mV and a low polydispersity index have shown the ability to achieve uniform distribution in the solvent [176]. However, recent studies have indicated that even nanocarriers larger than 150 nm can efficiently deliver dsRNA and enhance the RNAi effect [162,177,178]. The dsRNA fragment size is a determinant of the size of the resulting nanocarrier. While longer fragments (>60 bp) are usually required for efficient gene silencing in insects [41,179], some researchers have opted for shorter dsRNA sequences (<60 bp) for nanocarrier delivery, as the parameters such as size or charge are easier to control when using RNAi molecules with lower molecular weight [135,152]. It remains to be demonstrated whether dsRNA encapsulated into larger nanocarriers or siRNA into smaller ones is more efficient in improving gene silencing. Currently, this decision is made on a case-by-case basis; however, other factors besides efficacy could also be determinants, such are biosafety, production costs, and reproducibility.
While stability and protection provided by nanocarriers are often the emphasized characteristics in designing an RNAi nanoformulation, there is another pivotal feature that should be equally explored, that are the mechanisms of nanocomplexes disassembly and RNA release. In most cases, intracellular anionic substances act as competitors for the cationic regions in the nanocarriers, displacing and consequently releasing the RNA molecules [180,181,182]. The advancements in the pharmaceutical industry have provided a number of controlled-release methods for different nanocarriers such as chitosan [183], silica [184], peptides [133], among others. These methods include temperature- or pH-induced release and nanocarrier’s surface modifications to exploit the recognition system mediated by receptors in cell membranes. Developing nanocarriers with tunable release mechanisms will be essential in the following years for the establishment of nanocarrier-mediated delivery of RNAi as a biopesticide. Although research in medical sciences has contributed to elucidating ways to achieve controlled release, further insights specific to applications against pathogens and insect pests are still needed. More on this will be discussed in the next section “topical RNAi to control insect pests”.
Finally, as for any pesticide, conducting risk assessment studies is crucial in the development of RNAi nanocarriers. To establish these studies and protocols, it is crucial to prioritize exploring how physicochemical characteristics, tissue accumulation, concentration, and the type of nanocarriers impact the specificity of RNAi against target pests and its toxicity towards non-target species. Recent publications have shown that chitosan/dsRNA nanocarriers specific to H. armigera did not produce any effects on non-target organisms (D. melanogaster and S. litura) [135,152]. However, we consider that risk assessments on species considered biomarkers are more appropriate to evaluate the safety of nanocarriers, due to their high sensitivity to chemical compounds, as in the case of beneficial insects (e.g., pollinators, aquatic organisms, and natural predators). For instance, star polycation nanocarriers for the control of whitefly only caused significant toxicity on predatory ladybird larvae at concentrations above that required for whitefly’s management [56,185]. Risk assessment protocols for nanocarriers containing molecules other than RNA have already been explored by comparing them with traditional agrochemicals [30,168,186]. Despite RNAi being considered a highly specific strategy for targeting pests, it is essential to conduct similar toxicity evaluations using RNAi-based nanocarriers to accurately evaluate their performance in comparison to current chemical pesticides. In conclusion, the use of dsRNA complexed with nanocarriers holds the potential to impact the fields of biotechnology and agriculture in the upcoming years, and biosafety considerations are important to accelerate this process.

3.4. Topical RNAi to Control Insect Pests

The foliar spray method has been the preferred approach for topical dsRNA delivery to control insect pests, characterized by its effectiveness not only against chewing insects but also relatively effective against sap-sucking insects due to the dsRNA’s ability to spread from the leaf surface into plant tissues [187]. Research by Hunter et al. [188] revealed that dsRNAs persist for an extended period in treated insects (5–8 days) and plants (at least 57 days post-treatment) under field conditions. However, it has been noted that most insect species from the Coleoptera and Hemiptera orders exhibit a robust RNAi response compared to Lepidoptera, which has a very limited response [127,189]. Therefore, despite advancements, there is still considerable room for further improvements in this research area, especially for RNAi-recalcitrant insects in open field conditions [113,116,190]. Current strategies to address this issue include focusing on the down-regulation of lethal-associated target genes [63,108,109,110], optimizing dsRNA delivery systems [191,192], and using nanocarriers to enhance the RNAi response in target insects [193,194]. Therefore, RNAi-based insect pest management is expected to become more cost-effective with the use of nanocarriers, as the required amount of dsRNA to control the pests is likely to be reduced. Nanocarriers-mediated dsRNA delivery in insect pests has been well described in mosquitoes [195], but also on aphids and whiteflies [196,197] and lepidopteran larvae [198,199,200].
The main objective of these previous studies was the protection of dsRNAs from the hydrolytic environment of the insect gut and hemolymph to enhance the silencing effect of target genes. Furthermore, most of these studies targeted lepidopteran and hemipteran pests and explored the efficiency of dsRNA/nanocarrier complex to be delivered by feeding (62%), while the alternative approach based on transdermal/direct application was less frequent (38%) [153,194,199,201,202]. Down-regulation of target genes by dsRNAs delivered using these methods was over 50% compared to control treatments, in approximately 68% of these studies, with star polycation nanocarriers being the most used (37%), followed by chitosan (21%). In particular, Jain et al. [203] showed that foliar application of dsRNA complexed with layered double hydroxide nanocarriers on cotton exhibited significantly better performance compared to naked dsRNA, due to better dsRNA uptake and movement into plant vascular bundles and insect tissues, causing higher mortality in B. tabaci. Moreover, these authors showed that the mixture of dsRNAs targeting three genes induced higher insect mortality compared to individual dsRNAs [203]. Likewise, the mixture of dsRNAs targeting two genes conjugated with star polycation nanocarriers increased significantly the RNAi-mediated mortality of C. suppressalis [199]. The cotton boll weevil was also treated with dsRNA targeting multiple target genes and encapsulated into chitosan and PEI nanocarriers, showing high protection levels against intestinal degradation [71]. Therefore, the use of nanocarriers and multiple dsRNAs targeting different insect genes is an interesting approach to enhance the RNAi response. However, how effective these strategies might be at increasing the RNAi efficiency in other insect species and the saturation limit of the RNAi pathway remains to be explored. Finally, Ledprona is the first and only topical RNAi-based active ingredient developed for foliar application to target L. decemlineata (a chewing insect) and is commercially approved for open field use in the USA [192,204].
On the other hand, sucking insects feed on phloem or xylem sap do not have much access to cellular contents, partially depriving themselves of the RNAi present in plant cells [126,188,190]. Therefore, dsRNA delivery by topical spray may be, in some cases, insufficient to significantly down-regulate the target genes of these insect pests [205]. Furthermore, delivery of naked dsRNAs through topical spray in sufficient quantities into the vascular system of the host plant remains a challenge [205]. In turn, delivery of dsRNA in liquid solution through soaking for plant cuttings [190,191], detached leaves [206], and roots [207], or trunk injection [208] is an effective laboratory-scale method to provide systemic movement of dsRNAs through the phloem and to control sap-sucking insects. Although the economic viability of these phloem-restricted methods still needs to be evaluated for open field application, it can be enhanced and benefited by reducing the cost of dsRNA production and by optimizing nanocarrier-mediated delivery.
Optimization of nanocarriers for dsRNA/siRNA delivery in insects must consider the heterogeneous physiology among species. Particularly, internal pH (gut or hemolymph) influences the stability and efficacy of nanocarriers. Many nanomaterials used for pest management have come from studies with medical purposes; hence, their buffering capacity has been selected for the slightly acidic or physiological (pH 7.0) environment surrounding mammalian cells [131]. However, some pest species, notably from orders Lepidoptera and Orthoptera, exhibit highly alkaline midguts [209]. Structural modifications of nanocarriers are often required to overcome these conditions; for instance, the addition of functional groups such as guanidine provides an increased pKa value for polymeric nanocarriers [193]. Other nanocarriers, like branched amphiphilic peptide capsules, can be directly engineered in their primary structure to grant them stability at specific pH ranges, turning additional modifications unnecessary [162]. Further, pH not only affects nanocarriers’ stability but also RNA release. In fact, pH-responsive nanocarriers have been widely used for drug delivery [210]. By taking advantage of pH differences between insect cells and the extracellular environment, it would be possible to favor tighter encapsulation efficacy before cell uptake and promote easier release at the physiological pH of target cells. In summary, the implementation of topical RNAi for insect pest management can or not require the use of nanocarriers to be reliable for field applications. This will depend on the target species, environmental conditions, and delivery method.

3.5. Topical RNAi to Control Major Plant Pathogens

Smaller advances have been made in topical RNAi to control plant-parasitic nematodes in open fields, given the complexity of its delivery via soil [211]. However, this approach has shown promise at least through proof-of-concept tests carried out under in vitro conditions such as nematode soaking in solutions containing dsRNA [212,213]. In contrast, topical RNAi to control plant viruses has shown more significant advances, mainly using nanocarrier-mediated RNAi delivery [108,214]. Furthermore, its efficiency has been largely dependent on the selection of the target viral RNA, as well as the overtime stability of dsRNAs [89,215]. In particular, Konakalla et al. [216] showed that topical application of dsRNA molecules engineered from TMV p126 and coat protein genes enhanced tobacco resistance against TMV. Likewise, Rego-Machado et al. [217] improved tomato resistance by spraying dsRNA molecules that target coat protein genes of Tomato mosaic virus and Potato virus Y. Comparatively, the topical RNAi triggered by dsRNA engineered from begomovirus genome was not effective against phloem-limited begomoviruses [217]. Meanwhile, topical RNAi to control plant pathogenic fungi and oomycetes has advanced with promising results for multiple pathogen species, being a promising alternative to associate with chemical control [218,219]. However, the topical RNAi effectiveness in controlling these pathogens is highly dependent on the dsRNA stability, the efficiency of dsRNA uptake by the pathogen, and the function of the gene targeted by RNAi [110,220,221]. The lifetime of dsRNA on sprayed plants particularly conferred by nanocarriers and E. coli-derived anucleated minicells is another determining factor for the effectiveness of dsRNA uptake and control of these pathogens [222,223]. In particular, Koch et al. [11,224] showed that spraying barley plants with dsRNA engineered to target three Fusarium graminearum CYP450 genes required for ergosterol biosynthesis successfully inhibited fungal growth. Therefore, topical RNAi to control plant pathogens has made notable advances and is a promising alternative to overcome the issue of transgenic acceptance by consumers and reduce reliance on chemical pesticides, but it still needs major advances for use in open field crops. Examples of topical RNAi applications to control plant pathogens and an overview of dsRNA delivery methods, target genes, crops, and plant phenotypes are reported in Tables S3 and S4.

4. Intellectual Property of RNAi-Based Technologies

4.1. Patent Search Methodology

The patent search covered a 20-year period (2004–2024) and utilized both public databases (Espacenet, USPTO, and INPI) and the commercial platform Questel Orbit to ensure broad coverage. The strategy combined keyword searches (e.g., “RNA interference”, “dsRNA”, “agricultural RNAi”, and crop-specific terms like “soybean” or “sugarcane”), International Patent Classification (IPC) codes (e.g., A01H 1/00 for plant genetic modification), and Jurisdictional filters (South America-focused filings). Duplicates and irrelevant patents were manually excluded. Trends were analyzed quantitatively (annual filings) and qualitatively (technology evolution). Due to the 18-month confidentiality window for patent applications, data beyond 2022 were treated as preliminary. This approach aligns with established methodologies for patent landscaping, as described in Trippe [225].

4.2. Patent Scenario

Intellectual property is essential to protect RNAi-related technological knowledge and guarantee the right of use [226,227]. Private and public companies invest in a robust portfolio of patents in their technological area to attract and safeguard investments made in research and development, to increase the value of the company, or to overcome obstacles imposed by other competitors [86]. These companies can also expand the economic value of this portfolio by licensing their patented products to other partners, thereby increasing the impact of the technology and earning royalties [228]. Scientific research in RNAi depends on crucial inventive aspects, such as the sequence identity of dsRNA to the target gene and non-target gene, structural engineering of the dsRNA sequence, nanocarriers to protect and deliver the dsRNA, and the optimized method for large-scale RNA production [229]. Over the past 20 years, RNAi has been exploited in agriculture to control insect pests and plant pathogens, which cause significant agricultural losses worldwide each year [230]. Therefore, this study describes the patents that currently protect RNAi innovation and the current scenario of intellectual property-based RNAi businesses focused on South America.
There are currently approximately 100,000 patent families worldwide identified open access and private databases using the keyword “RNAi” through Boolean searches, based on the International Patent Classification (Figure 3A). Of this number, approximately 18,291 patent families refer to RNAi-based technologies for agriculture (Figure 3B), of which 13,644 are operationally active. Within this subset, 114 patents are for in planta RNAi, with 48% owned by the top ten players, and 65 patents are for topical RNAi, with 47% owned by the top ten players. Additionally, 1321 patents refer to dsRNA nanoformulations for agriculture, with 26% owned by the top ten players. In particular, Bayer CropScience is the leading applicant worldwide, with 4268 granted patents on RNAi-based solutions and 278 patents still pending approval. MS Technologies and Stine Seed Farm companies rank second and third, respectively, in terms of the number of RNAi patents granted or pending, followed by Bayer CropScience, Pioneer Hi-Bred International, and BASF Plant Science (Figure 3C). USA (9290), China (3646), and European Union (3117) are the top three countries with the highest number of patents related to RNAi for agriculture, followed by Canada (2088), India (1888), Japan (1690), Mexico (1581), Australia (1510), Brazil (1366) and Korea (1288). In addition to Brazil, South America is also represented by Chile (491), Argentina (425), and Colombia (370) (Figure 3D). The number of RNAi patents filed In South America for agriculture reached its highest number in 2012 (251), followed by a linear decline until 2022 (Figure 3E). In Brazil, there are approximately 10,754 active patents (granted or pending) identified by searching for the keyword ‘RNAi’. Among these, nearly 12% (1366) are destined for agriculture. Bayer CropScience holds the highest number of active patents (234), representing 8% of all active RNAi patents for agriculture in Brazil. Pioneer Hi-Bred International (187), Bayer CropScience (169), Corteva (152), and BASF (141) rank next in terms of the largest number of active RNAi patents. Embrapa is the only South American among the global companies with the highest number of RNAi patents for agriculture, with 11 patents. The DevGen company, which specializes in formulations for topical RNAi, holds eight RNAi patents for agriculture in Brazil. GreenLight Biosciences, a world leader in large-scale dsRNA production based on the cell-free synthesis method, holds seven patents in Brazil. RNAissance BioScience, the second-largest company for dsRNA production, holds four RNAi patents in Brazil.
Based on the global RNAi intellectual property scenario, South America portrays a growing interest in research and development, with an increase of approximately 8% in patents each year over the last 20 years. However, only approximately 23% of all RNAi-related patents worldwide are also protected in South America. While this percentage certainly reflects the difference in investment in research and development made in countries like the USA and China, it may also be influenced by variations in protective legislation relating to genetic material, particularly in South American countries compared to the USA. Taking Brazil as an example, the issue of patentability of genetic material, in particular DNA sequences, has been a source of controversy and legal ambiguity. Although the Brazilian Industrial Property Law [231] allows the patenting of transgenic materials and processes, the implementation of this law has been a challenge. In particular, Brazilian patent legislation requires that an invention not be obvious to a person with skills in the area, preventing patentability when any individual with average knowledge in the relevant scientific or technical field can combine different known information to achieve the same result. Thus, the question arises as to whether the isolation of a natural DNA sequence would be merely a routine task for competent biologists or geneticists, and would not qualify as an inventive activity. Furthermore, due to the history of biopiracy, there is increased scrutiny for granting patents on genetic resources. Finally, the Brazilian Patent Office (National Institute of Industrial Property) has defined unified guidelines for patent examiners on the patentability of genetic materials, reducing disparities in examination criteria and uncertainty for applicants.
In the USA, the main criteria for patentability are novelty, non-obviousness, and utility [232,233]. Therefore, a genetic sequence can be patented if it has a practical application. According to USA law applied through USPTO, it is possible to find the following explanation: “an isolated DNA molecule that has the same sequence as a naturally occurring gene is eligible for a patent because (i) an excised gene is eligible for a patent as a composition of matter or as an article of manufacture because that DNA molecule does not occur in this isolated form in nature, or (ii) synthetic DNA preparations are eligible for patents because their purified state is different from the naturally occurring compound” [234]. Many identified and isolated genetic sequences have been patented in the USA, particularly in biotechnology and diagnostics. The USA also has a large volume of jurisprudence related to the patenting of genetic materials. In particular, the Supreme Court ruling in an episode of Myriad Genetics [235] determined that naturally occurring DNA is not patentable, while complementary DNA can be patented. Therefore, this scenario of RNAi patents in South America highlights the great potential for local technological development, and the importance of agribusiness in these countries represents a promising opportunity for stakeholders who wish to participate in this market.

5. Biosafety and Risk Assessment of In Planta and Topical RNAi-Based Technologies

In planta RNAi-based products are considered transgenics worldwide and regulated by well-known biosafety laws applied to genetically modified crops [77]. In turn, topical RNAi-based products have great potential for pest and disease control with minimal environmental impact because they have the potential to be much more specific than chemical pesticides and are less residually persistent in the environment [115,236,237,238]. Therefore, in theory, these topical products are comparatively safer than chemical pesticides [239] and fit the world’s demands for eco-friendly products [14]. However, topical RNAi-based products require appropriate and specific biosafety assessment and authorization procedures for use [240,241]. The need for specific biosafety regulation led the European Food Safety Authority and the Organization for Economic Co-operation and Development (OECD) to bring together expert scientists to discuss and define guidelines for risk assessment related to topical RNAi [242,243]. These meetings stimulated scientific debate and some articles have already described the proposals for regulatory frameworks in different countries [18,77].

5.1. Risk Assessment for Human Health

The longstanding history of safe consumption of dsRNAs naturally present in all food products, even those with complementarity to human and vertebrate transcripts, strongly supports the safety of these molecules for application as agricultural biopesticides [237]. Moreover, the safe consumption of transgenic plants expressing dsRNAs in high doses has been approved over two decades (Table 1) [244]. The safe consumption is attributed to the presence of multiple biological barriers at the gastrointestinal, bloodstream, and cellular levels in mammals [237]. Indeed, small siRNAs and miRNAs are ubiquitous in plant- and animal-derived foods. Therefore, if there were no natural barriers for these molecules, the consumption of natural foods would pose a permanent risk to human health. The OECD 2019 Conference addressed the potential exposure of sensitive human populations to dsRNA-based biopesticides and no significant issues were raised regarding the safety of this technology [242]. The risk assessment for in planta RNAi-based products is typically conducted on a case-by-case basis by these authorities considering the laws applied to genetically modified crops. In turn, topical RNAi-based biopesticides can be incorporated into formulations to improve the cellular uptake of dsRNA or to protect dsRNA against nucleolytic degradation [114,193,245]. Therefore, the use of these formulations might potentially increase human exposure and elevate risks. However, considering the presence of multiple biological barriers in mammals, it is unlikely that formulated RNAi-based products developed for agricultural purposes will effectively deliver dsRNA into human cells upon ingestion [246]. To date, regulatory authorities have not adopted standardized methods for assessing the safety of topical RNAi-based products for agriculture. Nevertheless, a substantial consensus suggests that the current regulatory framework for agrochemicals could serve for risk assessment of topical RNAi-based products. Finally, transgenic rootstocks containing RNAi-based strategies to pest control or improve agronomic traits restricted to roots are a viable alternative since the aerial part is not transgenic and does not have equivalent amounts of these dsRNAs compared to the root [85]. In this sense, a particular risk assessment analysis must be carried out to authorize the transgenic planting and the commercialization of its non-transgenic fruits.

5.2. Environmental Risk Assessment

Some authors argue that the principles and guidelines used for environmental risk assessment of RNAi-based transgenic plants can be effectively adapted to topical RNAi-based products [240,247]. The primary environmental risks associated with topical RNAi-based products are linked to their potential impact on non-target organisms [243,248]. Numerous scientific articles showed that bioinformatic analyses are indispensable in the risk assessment for non-target organisms [249]. In fact, the action mode of RNAi is well understood, relying on the base pairing of siRNA with the target organism’s mRNAs [240,250]. Therefore, it is theoretically easy to design dsRNAs to specific sequences on target pests, thus mitigating off-target effects [238,251]. However, there are still knowledge gaps that justify the inability to predict the impact on non-target organisms using bioinformatics alone [241,242]. Despite the significant update of sequence databases containing information on non-target organisms, many remain unavailable. In addition, the action mode of the RNAi might vary across different organisms. For example, there may be differences in the number of mismatches to still be able to activate the RNAi machinery [249,250] and in the efficiency of systemic siRNA transport [252], which may not be accurately identified through bioinformatic analysis. For example, siRNAs resulting from DCL2 processing have demonstrated variable lengths (ranging from 20 to 22 nucleotides) across different insect species [253]. Another study revealed that differences in the number of mismatches could lead to varying effects on the mortality of non-target organisms [254]. Besides these sequence-specific mechanisms, some studies suggest that effects on the non-target organisms could occur due to sequence-unrelated mechanisms, making it challenging to predict all possible effects using only bioinformatics [252]. One theoretically possible sequence-unrelated mechanism would be immune stimulation or saturation of the enzymes involved in the RNAi machinery, which could negatively affect the fitness of non-target organisms [255]. However, considering the multiple natural barriers against exogenous dsRNAs, the likelihood of this occurring in non-target organisms is probably low [240]. Therefore, despite bioinformatics being an essential tool for risk assessment, it is not sufficient to predict all possible effects on non-target organisms from RNAi-based applications and needs to be employed in conjunction with empirical testing.

5.3. Possible Off-Target Effects of RNAi-Based Products Versus Decision Making

Surveying all potential theoretical adverse effects against non-target organisms does not contribute effectively to risk assessment to support decision making. For regulatory risk assessment of RNAi-based products, generating a comprehensive list of possible off-target effects would be challenging and inefficient in supporting decision making. Instead, problem formulation should establish criteria that define an acceptable risk framework and develop a plan to test the hypothesis that the product meets those criteria. The key to effective risk assessment of RNAi-based products is determining whether their properties indicate acceptable or unacceptable risk [247]. There is a robust body of evidence that these products, in principle, pose lower risks to non-target species compared to chemical pesticides [241,256]. However, this hypothesis needs to be tested before these products are commercially released to ensure that their use does not cause unacceptable harm to the environment [257]. Therefore, considering that agricultural activities are in practice employed in food production resulting in non-negligible environmental impacts, the main hypothesis that must be tested to support decision making is whether the use of RNAi-based products offers a significantly lower environmental impact than traditional activities. If the impact of RNAi-based biopesticides is significantly lower than that of chemical pesticides, the risk of these new technologies should be considered acceptable. Romeis and Widmer [257] suggested an empirical approach based on toxicity studies carried out in the laboratory that can be employed to test whether the effect of a specific RNAi-based product on non-target organisms constitutes an acceptable or unacceptable risk. After bioinformatics analysis, a group of closely related to phylogenetically distant species should be selected, exposed to the RNAi, and their responses analyzed [257]. This approach allows a clear comparison between the RNAi-based product and traditional chemical pesticides, making it possible to test the hypothesis that the impact of the new product is significantly lower than the current one. A combination of bioinformatics and small field trials can be performed to test the hypothesized acceptability of in planta and topical RNAi-based products for decision making. Field trials designed to assess the impact on non-target organisms are carried out in the risk assessment of transgenic plants in most countries where cultivation is permitted. It is believed that these field tests can be adapted to determine whether an RNAi-based product poses an acceptable risk. The rationale of the proposed approach is to carry out a field trial comparing two treatments: (i) RNAi-based product and (ii) conventional chemical pesticide used to control a target. Subsequently, an assessment of the microbiota and insect populations is carried out using methodologies commonly employed for transgenic plants. If these results demonstrate that the impact of the RNAi-based product is significantly lower than that of using a chemical pesticide, the environmental risk of the new technology would be considered acceptable. If the difference is large or the effect in non-target organisms is negligible, these results can in principle be extrapolated to other crop areas where the experiments were not carried out following the principle of transportability of environmental data on transgenic plants [258,259]. This principle is based on the assumption that if no biologically relevant differences are observed between a transgenic plant and its conventional counterparts in one crop region, data from these studies can be used to inform the risk assessment in another region or country, regardless of the agroclimatic zone [259]. Therefore, this principle can also be perfectly employed in the risk assessment of RNAi-based biopesticides. These studies also consider that field trials present important practical advantages over toxicity studies carried out in the laboratory to evaluate adverse effects against non-target organisms. Field trials enable the assessment of several species almost simultaneously, providing a broad view of the impact on phylogenetically distant non-target microorganisms and insects. They also allow addressing the effects of RNAi in non-target cross-species and cross-kingdoms [17,77,241]. Finally, field trials avoid the need to develop protocols for laboratory cultivation of the species to be analyzed.

5.4. Current Regulatory Status of In Planta and Topical RNAi in Latin America

In planta RNAi-based products are regulated as transgenic plants worldwide and undergo biosafety and risk assessments as determined by each country’s legislation [19]. The planned release or commercial use of transgenics needs to be addressed and approved by each country’s National Biosafety Commission. The submission and approval of a transgenic product in different countries have standard guidelines at global levels but may include some particular national legislative measures [18,19]. This global format is necessary for countries to comply with standards for importing and exporting agricultural commodities. Recently, these countries have begun to adopt additional laws to categorize and incorporate innovative technologies, such as topical RNAi and genome editing. Therefore, these new laws aim to judge on a case-by-case basis and define whether or not these new technologies are transgenic [17].
In Latin America, some countries have already adopted this new law to regulate these precision breeding technologies [19]. For both genome editing and RNAi-based products, this new law adopts as a prerogative the presence or absence of foreign DNA with the replication capacity in the final product to classify it as a transgenic or non-transgenic product. However, this same prerogative is not adopted by some other Latin American and European countries [18,260]. Once considered non-transgenic, these topical RNAi-based products must undergo prior approval from other national agricultural and environmental oversight boards to ultimately be considered commercial. Finally, the formulation of topical RNAi-based products must be addressed taking into account the encapsulation mediated by proteins, microorganisms and nanocarriers, target genes, off-target genes, non-target organisms, surfactants, and stabilizers. In particular, the regulations of these RNAi-based biopesticides are still being elaborated in Latin America countries. This regulatory subdivision, which must be drawn up on a scientific basis, will be important to allow their classification to differentiate them from non-selective synthetic chemicals.

5.5. Resistance Against Exogenous dsRNA and RNAi

The development of resistance against RNAi has already been reported in mammals [261], plants [262], plant pathogens [263,264], and insect pests [41,265]. The loss of RNAi efficiency is not completely understood in all organisms. However, the encoding of proteins with suppressive activity of the RNAi pathway, genetic variability and diversity caused by selection pressure on target genes of RNAi, hijacking of the siRNAs originating from the dsRNA interfering, disruption of amplification or systemic movement of RNA-silencing signal, and the epigenetic silencing of the transgene encoding dsRNA interfering are some examples of mechanisms that lead to resistance to RNAi [264,266,267]. Likewise, resistance against dsRNA in D. virgifera virgifera was considered polygenic, located on a single locus (CEAS 300), and associated with the reduced uptake of dsRNA in midgut cells [41]. Similarly, variation in RNAi sensitivity in L. decemlineata correlates with the expression of RNAi machinery genes and staufenC, a dsRNA-binding protein involved in dsRNA processing to small interfering RNA [268,269]. These data reveal that pyramiding or concatenating different dsRNA sequences targeting different mRNAs may not be sufficient or long-lasting to control pests. Therefore, future research is needed to understand the molecular mechanisms of resistance against dsRNA and RNAi, to measure the implications of interaction between plants versus non-target organisms, and to develop strategies to overcome these limitations. Furthermore, resistance management strategies must be implemented for the long-term sustainability of RNAi-based products in agriculture [127].

6. Major Contributions of RNAi to More Sustainable Agriculture

Through genetic engineering, several in planta RNAi-based traits were introduced in 16 crops and approved for commercial use in several countries (Table 1). These traits focused on improving plant resistance to viruses, delaying fruit softening and ripening, reducing oxidative browning of fruits, improving oil quality, and improving maize resistance to D. virgifera virgifera. The common bean resistant to BGMV (Brazil, in 2011) and potato resistant to Potato virus Y (Argentina, in 2018) were pioneers in Latin America. Bayer CropScience, Zeneca, Seminis, BASF, J.R. Simplot, DuPont/Corteva, and Okanagan Specialty Fruits are the main agencies developing these RNAi-based technologies. Despite this release, commercial use in the countries where they were approved still faces restrictions for being transgenic. However, the benefits brought by these in planta RNAi-based traits for food security and greater environmental sustainability are highlighted by several studies [14,270]. The adoption of these transgenic crops makes it possible to improve the product quality and durability until they reach the consumer, reduce the use of synthetic agrochemicals, and decrease economic losses caused by pests. In this context, topical RNAi-based products are an alternative to overcome the barriers faced by in planta RNAi and to promote greater environmental sustainability (Figure 4). Conversely, topical RNAi-based biopesticides can theoretically be applied to any pest susceptible to RNAi. Another advantage of topical RNAi-based biopesticides over the in planta RNAi is the shortest development time. Given these advantages, several agribusiness companies have added efforts in research and development in RNAi. Bayer CropScience developed the first technology portfolio based on topical RNAi for controlling insect pests, weeds, and pathogens (BioDirect™ technology). GreenLight Biosciences has launched the first commercially approved active ingredient and product, respectively Ledprona and CalanthaTM, to control the Colorado potato beetle in potatoes [192,204]. Other topical RNAi-based products are being implemented by these companies to control plant pathogens and insect pests. Likewise, new topical RNAi-based products that enhance the effects of synthetic agrochemicals on the target pest are also expected for commercial release. More recently, other companies such as Syngenta, Corteva, Sempre AgTech (WIN), AgroRNA, Nufarm, RNAissance BioScience, AgroSpheres, Agrivalle, PlantArcBio in partnership with ICL Planet, NanoSUR, Trillium Ag, and RNAgri also opened portfolios of topical RNAi-based solutions and materials to meet RNAi industry. These companies have focused on large-scale and low-cost dsRNA production using cell-free systems, while others have focused on developing nanocarriers for RNAi delivery. Therefore, these efforts and investments have made byproducts available at more affordable costs, including the production of dsRNA, nanocarriers, and stabilizers.

7. Concluding Remarks and Perspectives

Global agriculture has undergone remarkable transformations over recent years, witnessing several impactful changes across the world [271]. Notably, technological advances in various agribusiness sectors, together with the challenges arising from geopolitical conflicts, pandemics, socio-economic instability, and the profound impacts of climate change, placed sustainable food production at the forefront of discussions [14,272]. Given this scenario, the search for technologies that, in addition to providing high yield, are eco-friendly, has been intense across all agricultural sectors and countries [273]. Consequently, the agribusiness sector is currently transitioning to Agriculture 6.0, which is based on Environmental, Social, and Corporate Governance principles, that aim to prioritize technologies that minimize negative impacts on the ecosystem, such as RNAi-based products [13,14]. Despite reluctance against transgenic-based technologies, the potential of in planta RNAi to revolutionize agriculture by improving crop yield, mitigating losses caused by pathogens and insect pests, and reducing dependence on non-selective synthetic agrochemicals is becoming increasingly evident [274]. Meanwhile, topical RNAi-based products have established themselves with consistent solutions for agriculture challenges, and their continued advancements will contribute to renewable agriculture [270]. These technological advances rely on basic research to identify suitable target genes in target pests, as well as reducing the production costs of dsRNAs and nanocarriers, improving the delivery of these molecules to the target and prolonging the durability of the applied molecules [17]. Furthermore, substantial financial investments are needed to resolve existing drawbacks and facilitate the transfer of knowledge from laboratory to field [116]. These investments are crucial for transforming research findings into useful products and solutions for agriculture (Table S6) [14]. Thanks to technological advances achieved, some topical RNAi-based products are already in the regulatory approval phase for commercial use [192,204]. However, essential adjustments to the regulatory laws remain a bottleneck to be resolved in several countries [77]. By overcoming these obstacles, it will become increasingly evident that numerous RNAi-based solutions will become available, contributing significantly to agricultural and environmental sustainability [127]. Furthermore, the participation of several companies and startups focused on in planta and topical RNAi-based technologies will facilitate the offering of high-quality solutions at competitive prices (Table S7). Importantly, these new products will control insect pests, plant pathogens, and weeds, as well as enhance agronomic traits, ranging from seed treatment using non-transgenic approaches to transgenic crops [14,17]. Finally, there are also numerous opportunities with RNAi as biostimulants and priming technology for use in crops [275], which will also contribute significantly to sustainable agricultural and food production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agronomy15040859/s1. Table S1. Representative summary of successful studies using in planta and topical RNAi to control insect pests in plants. Table S2. Representative summary of successful studies using in planta and topical RNAi to control plant-parasitic nematodes in plants. Table S3. Representative summary of successful studies using in planta and topical RNAi to control viruses and subviral agents in plants. Table S4. Representative summary of successful studies using in planta and topical RNAi to control fungi and oomycetes in plants. Table S5. Summary of studies using nanoformulations for delivering dsRNA/siRNA through spray/topical application in insect pests. Table S6. Key steps from validating a RNAi strategy to product commercialization focused on drawbacks and advances to control insect pests, plant-parasitic nematodes, virus and subviral agents, fungi, and oomycetes. Table S7. Examples of companies and startups focused on in planta and topical RNAi-based technologies for agriculture.

Author Contributions

M.F.B. wrote this original manuscript. D.D.N.V., E.R.C.-P., D.H.P., B.C., G.V.G., G.C.M. and H.M.-S. contributed to the first draft of this manuscript. M.F.G.-d.-S. and D.D.N.V. provided inputs to the original manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

M.F.B. is grateful to CNPq and FAPESP for a postdoctoral research fellowship (process numbers: 108646/2024-6 and 2024/12315-0, respectively). M.F.G.-d.-S. was supported by grants from UCB, UCDB, CAPES, CNPq, FAP-DF, and INCT PlantStress Biotech.

Conflicts of Interest

Authors Marcos Fernando Basso, Daniel David Noriega Vásquez, Eduardo Romano Campos-Pinto, Daniele Heloísa Pinheiro, and Maria Fatima Grossi-de-Sa were employed by the public company Embrapa Genetic Resources and Biotechnology. Author Bread Cruz was employed by the public Innovation Agency INOVA-UNICAMP. Authors Grazielle Celeste Maktura, Giovanna Vieira Guidelli, and Henrique Marques-Souza were employed by the University of Campinas (UNICAMP, public university). All 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. Ali, N.; Datta, S.K.; Datta, K. RNA interference in designing transgenic crops. GM Crops 2010, 1, 207–213. [Google Scholar] [CrossRef] [PubMed]
  2. Rössner, C.; Lotz, D.; Becker, A. VIGS goes viral: How VIGS transforms our understanding of plant science. Annu. Rev. Plant Biol. 2022, 73, 703–728. [Google Scholar] [CrossRef] [PubMed]
  3. Bonfim, K.; Faria, J.C.; Nogueira, E.O.; Mendes, E.A.; Aragão, F.J. RNAi-mediated resistance to Bean golden mosaic virus in genetically engineered common bean (Phaseolus vulgaris). Mol. Plant-Microbe Interact. 2007, 20, 717–726. [Google Scholar] [CrossRef]
  4. Aregger, M.; Borah, B.K.; Seguin, J.; Rajeswaran, R.; Gubaeva, E.G.; Zvereva, A.S.; Windels, D.; Vazquez, F.; Blevins, T.; Farinelli, L.; et al. Primary and secondary siRNAs in geminivirus-induced gene silencing. PLoS Pathog. 2012, 8, e1002941. [Google Scholar] [CrossRef]
  5. Yang, B.; Sugio, A.; White, F.F. Os8N3 is a host disease-susceptibility gene for bacterial blight of rice. Proc. Natl. Acad. Sci. USA 2006, 103, 10503–10508. [Google Scholar] [CrossRef]
  6. Ribeiro, T.P.; Vasquez, D.D.N.; Macedo, L.L.P.; Lourenço-Tessutti, I.T.; Valença, D.C.; Oliveira-Neto, O.B.; Paes-de-Melo, B.; Rodrigues-Silva, P.L.; Firmino, A.A.P.; Basso, M.F.; et al. Stabilized double-stranded RNA strategy improves cotton resistance to CBW (Anthonomus grandis). Int. J. Mol. Sci. 2022, 23, 13713. [Google Scholar] [CrossRef]
  7. Basso, M.F.; Lourenço-Tessutti, I.T.; Mendes, R.A.G.; Pinto, C.E.M.; Bournaud, C.; Gillet, F.-X.; Togawa, R.C.; de Macedo, L.L.P.; de Almeida Engler, J.; Grossi-de-Sa, M.F. MiDaf16-like and MiSkn1-like gene families are reliable targets to develop biotechnological tools for the control and management of Meloidogyne incognita. Sci. Rep. 2020, 10, 6991. [Google Scholar] [CrossRef]
  8. Mendes, R.A.G.; Basso, M.F.; Amora, D.X.; Silva, A.P.; Paes-de-Melo, B.; Coiti Togawa, R.; Saliba Albuquerque, E.V.; Lisei-de-Sa, M.E.; Lima Pepino Macedo, L.; Lourenço-Tessutti, I.T.; et al. In planta RNAi approach targeting three M. incognita effector genes disturbed the process of infection and reduced plant susceptibility. Exp. Parasitol. 2022, 238, 108246. [Google Scholar] [CrossRef]
  9. Zhang, T.; Jin, Y.; Zhao, J.H.; Gao, F.; Zhou, B.J.; Fang, Y.Y.; Guo, H.S. Host-induced gene silencing of the target gene in fungal cells confers effective resistance to the cotton wilt disease pathogen Verticillium dahliae. Mol. Plant 2016, 9, 939–942. [Google Scholar] [CrossRef]
  10. Cheng, W.; Lin, M.; Chu, M.; Xiang, G.; Guo, J.; Jiang, Y.; Guan, D.; He, S. RNAi-based gene silencing of RXLR effectors protects plants against the oomycete pathogen Phytophthora capsici. Mol. Plant-Microbe Interact. 2022, 35, 440–449. [Google Scholar] [CrossRef]
  11. Koch, A.; Biedenkopf, D.; Furch, A.; Weber, L.; Rossbach, O.; Abdellatef, E.; Linicus, L.; Johannsmeier, J.; Jelonek, L.; Goesmann, A.; et al. An RNAi-based control of Fusarium graminearum infections through spraying of long dsRNAs involves a plant passage and is controlled by the fungal silencing machinery. PLoS Pathog. 2016, 12, e1005901. [Google Scholar] [CrossRef] [PubMed]
  12. Nagata, T.; Okada, K.; Takebe, I.; Matsui, C. Delivery of Tobacco mosaic virus RNA into plant protoplasts mediated by reverse-phase evaporation vesicles (Liposomes). Mol. Gen. Genet. 1981, 184, 161–165. [Google Scholar] [CrossRef]
  13. Grossi-de-Sa, M.F.; Basso, M.F. Ciências agrárias e as revoluções na produção de alimentos: Do passado ao futuro. In Segurança alimentar e nutricional: O papel da ciência brasileira no combate à fome; Hungria, M.H., Ed.; Academia Brasileira de Ciências: Rio de Janeiro, Brazil, 2024; pp. 48–55. [Google Scholar]
  14. Basso, M.; Neves, M.; Grossi-de-Sá, M.F. Agriculture evolution, sustainability and trends, focusing on Brazilian agribusiness: A review. Front. Sustain. Food Syst. 2024, 7, 1296337. [Google Scholar] [CrossRef]
  15. Savary, S.; Willocquet, L.; Pethybridge, S.J.; Esker, P.; McRoberts, N.; Nelson, A. The global burden of pathogens and pests on major food crops. Nat. Ecol. Evol. 2019, 3, 430–439. [Google Scholar] [CrossRef]
  16. Torres, J.B.; Rolim, G.G.; Arruda, L.S.; dos Santos, M.P.; Leite, S.A.; Neves, R.C.d.S. Insecticides in Use and Risk of Control Failure of Boll Weevil (Coleoptera: Curculionidae) in the Brazilian Cerrado. Neotrop. Entomol. 2022, 51, 613–627. [Google Scholar] [CrossRef] [PubMed]
  17. Távora, F.T.P.K.; de Assis Dos Santos Diniz, F.; de Moraes Rêgo-Machado, C.; Chagas Freitas, N.; Barbosa Monteiro Arraes, F.; Chumbinho de Andrade, E.; Furtado, L.L.; Osiro, K.O.; Lima de Sousa, N.; Cardoso, T.B.; et al. CRISPR/Cas- and topical RNAi-based technologies for crop management and improvement: Reviewing the risk assessment and challenges towards a more sustainable agriculture. Front. Bioeng. Biotechnol. 2022, 10, 913728. [Google Scholar] [CrossRef]
  18. Dietz-Pfeilstetter, A.; Mendelsohn, M.; Gathmann, A.; Klinkenbuß, D. Considerations and regulatory approaches in the USA and in the EU for dsRNA-based externally applied pesticides for plant protection. Front. Plant Sci. 2021, 12, 682387. [Google Scholar] [CrossRef]
  19. Rosado, A.; Eriksson, D. Biosafety legislation and the regulatory status of the products of precision breeding in the Latin America and the Caribbean region. Plants People Planet 2022, 4, 214–231. [Google Scholar] [CrossRef]
  20. Basso, M.F.; Ferreira, P.C.G.; Kobayashi, A.K.; Harmon, F.G.; Nepomuceno, A.L.; Molinari, H.B.C.; Grossi-de-Sa, M.F. MicroRNAs and new biotechnological tools for its modulation and improving stress tolerance in plants. Plant Biotechnol. J. 2019, 17, 1482–1500. [Google Scholar] [CrossRef]
  21. Basso, M.F.; Lourenço-Tessutti, I.T.; Busanello, C.; Pinto, C.E.M.; de Oliveira Freitas, E.; Ribeiro, T.P.; de Almeida Engler, J.; de Oliveira, A.C.; Morgante, C.V.; Alves-Ferreira, M.; et al. Insights obtained using different modules of the cotton uceA1.7 promoter. Planta 2020, 251, 56. [Google Scholar] [CrossRef]
  22. Fragoso, R.R.; Arraes, F.B.M.; Lourenço-Tessutti, I.T.; Miranda, V.J.; Basso, M.F.; Ferreira, A.V.J.; Viana, A.A.B.; Lins, C.B.J.; Lins, P.C.; Moura, S.M.; et al. Functional characterization of the pUceS8.3 promoter and its potential use for ectopic gene overexpression. Planta 2022, 256, 69. [Google Scholar] [CrossRef] [PubMed]
  23. Basso, M.F.; Arraes, F.B.M.; Grossi-de-Sa, M.; Moreira, V.J.V.; Alves-Ferreira, M.; Grossi-de-Sa, M.F. Insights into genetic and molecular elements for transgenic crop development. Front. Plant Sci. 2020, 11, 509. [Google Scholar] [CrossRef]
  24. Lindbo, J.A.; Silva-Rosales, L.; Proebsting, W.M.; Dougherty, W.G. Induction of a highly specific antiviral state in transgenic plants: Implications for regulation of gene expression and virus resistance. Plant Cell 1993, 5, 1749–1759. [Google Scholar] [CrossRef]
  25. Baulcombe, D.C.; Chapman, S.; Santa Cruz, S. Jellyfish green fluorescent protein as a reporter for virus infections. Plant J. 1995, 7, 1045–1053. [Google Scholar] [CrossRef] [PubMed]
  26. Hamilton, A.; Voinnet, O.; Chappell, L.; Baulcombe, D. Two classes of short interfering RNA in RNA silencing. EMBO J. 2002, 21, 4671–4679. [Google Scholar] [CrossRef]
  27. Mao, Y.-B.; Tao, X.-Y.; Xue, X.-Y.; Wang, L.-J.; Chen, X.-Y. Cotton plants expressing CYP6AE14 double-stranded RNA show enhanced resistance to bollworms. Transgenic Res. 2011, 20, 665–673. [Google Scholar] [CrossRef] [PubMed]
  28. Yan, P.; Shen, W.; Gao, X.; Li, X.; Zhou, P.; Duan, J. High-throughput construction of intron-containing hairpin RNA vectors for RNAi in plants. PLoS ONE 2012, 7, e38186. [Google Scholar] [CrossRef]
  29. Hu, D.; Chen, Z.-Y.; Zhang, C.; Ganiger, M. Reduction of Phakopsora pachyrhizi infection on soybean through host- and spray-induced gene silencing. Mol. Plant Pathol. 2020, 21, 794–807. [Google Scholar] [CrossRef]
  30. Zhang, D.; Zhong, C.; Smith, N.A.; de Feyter, R.; Greaves, I.K.; Swain, S.M.; Zhang, R.; Wang, M.B. Nucleotide mismatches prevent intrinsic self-silencing of hpRNA transgenes to enhance RNAi stability in plants. Nat. Commun. 2022, 13, 3926. [Google Scholar] [CrossRef]
  31. Erdmann, R.M.; Picard, C.L. RNA-directed DNA methylation. PLoS Genet. 2020, 16, e1009034. [Google Scholar] [CrossRef]
  32. Butel, N.; Yu, A.; Le Masson, I.; Borges, F.; Elmayan, T.; Taochy, C.; Gursanscky, N.R.; Cao, J.; Bi, S.; Sawyer, A.; et al. Contrasting epigenetic control of transgenes and endogenous genes promotes post-transcriptional transgene silencing in Arabidopsis. Nat. Commun. 2021, 12, 2787. [Google Scholar] [CrossRef]
  33. Joshi, I.; Kumar, A.; Kohli, D.; Bhattacharya, R.; Sirohi, A.; Chaudhury, A.; Jain, P.K. Gall-specific promoter, an alternative to the constitutive CaMV35S promoter, drives host-derived RNA interference targeting Mi-msp2 gene to confer effective nematode resistance. Front. Plant Sci. 2022, 13, 1007322. [Google Scholar] [CrossRef]
  34. Li, J.; Jiang, D.; Zhou, H.; Li, F.; Yang, J.; Hong, L.; Fu, X.; Li, Z.; Liu, Z.; Li, J.; et al. Expression of RNA-interference/antisense transgenes by the cognate promoters of target genes is a better gene-silencing strategy to study gene functions in rice. PLoS ONE 2011, 6, e17444. [Google Scholar] [CrossRef]
  35. Moura, S.M.; Freitas, E.O.; Ribeiro, T.P.; Paes-de-Melo, B.; Arraes, F.B.M.; Macedo, L.L.P.; Paixão, J.F.R.; Lourenço-Tessutti, I.T.; Artico, S.; da Cunha Valença, D.; et al. Discovery and functional characterization of novel cotton promoters with potential application to pest control. Plant Cell Rep. 2022, 41, 1589–1601. [Google Scholar] [CrossRef] [PubMed]
  36. Hunter, W.B.; Wintermantel, W.M. Optimizing efficient RNAi-mediated control of hemipteran pests (psyllids, leafhoppers, whitefly): Modified pyrimidines in dsRNA triggers. Plants 2021, 10, 1782. [Google Scholar] [CrossRef] [PubMed]
  37. Fitch, M.M.M.; Manshardt, R.M.; Gonsalves, D.; Slightom, J.L.; Sanford, J.C. Virus resistant papaya plants derived from tissues bombarded with the coat protein gene of Papaya ringspot virus. Nat. Biotechnol. 1992, 10, 1466–1472. [Google Scholar] [CrossRef]
  38. Ferreira, S.A.; Pitz, K.Y.; Manshardt, R.; Zee, F.; Fitch, M.; Gonsalves, D. Virus coat protein transgenic papaya provides practical control of Papaya ringspot virus in Hawaii. Plant Dis. 2002, 86, 101–105. [Google Scholar] [CrossRef]
  39. Aragão, F.J.; Nogueira, E.O.; Tinoco, M.L.; Faria, J.C. Molecular characterization of the first commercial transgenic common bean immune to the Bean golden mosaic virus. J. Biotechnol. 2013, 166, 42–50. [Google Scholar] [CrossRef]
  40. Head, G.P.; Carroll, M.W.; Evans, S.P.; Rule, D.M.; Willse, A.R.; Clark, T.L.; Storer, N.P.; Flannagan, R.D.; Samuel, L.W.; Meinke, L.J. Evaluation of SmartStax and SmartStax PRO maize against western corn rootworm and northern corn rootworm: Efficacy and resistance management. Pest Manag. Sci. 2017, 73, 1883–1899. [Google Scholar] [CrossRef]
  41. Khajuria, C.; Ivashuta, S.; Wiggins, E.; Flagel, L.; Moar, W.; Pleau, M.; Miller, K.; Zhang, Y.; Ramaseshadri, P.; Jiang, C.; et al. Development and characterization of the first dsRNA-resistant insect population from western corn rootworm, Diabrotica virgifera virgifera LeConte. PLoS ONE 2018, 13, e0197059. [Google Scholar] [CrossRef]
  42. Zhang, J.; Khan, S.A.; Hasse, C.; Ruf, S.; Heckel, D.G.; Bock, R. Pest control. Full crop protection from an insect pest by expression of long double-stranded RNAs in plastids. Science 2015, 347, 991–994. [Google Scholar] [CrossRef] [PubMed]
  43. Bolognesi, R.; Ramaseshadri, P.; Anderson, J.; Bachman, P.; Clinton, W.; Flannagan, R.; Ilagan, O.; Lawrence, C.; Levine, S.; Moar, W.; et al. Characterizing the mechanism of action of double-stranded RNA activity against western corn rootworm (Diabrotica virgifera virgifera LeConte). PLoS ONE 2012, 7, e47534. [Google Scholar] [CrossRef]
  44. Svoboda, P.; Di Cara, A. Hairpin RNA: A secondary structure of primary importance. Cell. Mol. Life Sci. 2006, 63, 901–908. [Google Scholar] [CrossRef] [PubMed]
  45. Kutter, C.; Svoboda, P. miRNA, siRNA, piRNA: Knowns of the unknown. RNA Biol. 2008, 5, 181–188. [Google Scholar] [CrossRef]
  46. Abbasi, R.; Heschuk, D.; Kim, B.; Whyard, S. A novel paperclip double-stranded RNA structure demonstrates clathrin-independent uptake in the mosquito Aedes aegypti. Insect Biochem. Mol. Biol. 2020, 127, 103492. [Google Scholar] [CrossRef]
  47. Guan, R.B.; Li, H.C.; Fan, Y.J.; Hu, S.R.; Christiaens, O.; Smagghe, G.; Miao, X.X. A nuclease specific to lepidopteran insects suppresses RNAi. J. Biol. Chem. 2018, 293, 6011–6021. [Google Scholar] [CrossRef]
  48. Garcia, R.A.; Lima Pepino Macedo, L.; Cabral do Nascimento, D.; Gillet, F.X.; Moreira-Pinto, C.E.; Faheem, M.; Moreschi Basso, A.M.; Mattar Silva, M.C.; Grossi-de-Sa, M.F. Nucleases as a barrier to gene silencing in the cotton boll weevil, Anthonomus grandis. PLoS ONE 2017, 12, e0189600. [Google Scholar] [CrossRef]
  49. Gillet, F.X.; Garcia, R.A.; Macedo, L.L.P.; Albuquerque, E.V.S.; Silva, M.C.M.; Grossi-de-Sa, M.F. Investigating engineered ribonucleoprotein particles to improve oral RNAi delivery in crop insect pests. Front. Physiol. 2017, 8, 256. [Google Scholar] [CrossRef]
  50. He, G. Engineering chloroplasts for insect pest control. Proc. Natl. Acad. Sci. USA 2022, 119, e2205125119. [Google Scholar] [CrossRef]
  51. Wu, M.; Dong, Y.; Zhang, Q.; Li, S.; Chang, L.; Loiacono, F.V.; Ruf, S.; Zhang, J.; Bock, R. Efficient control of western flower thrips by plastid-mediated RNA interference. Proc. Natl. Acad. Sci. USA 2022, 119, e2120081119. [Google Scholar] [CrossRef]
  52. Liao, J.; Rong, H.; You, L.; Xia, K.; Wang, M.; Han, P.; Li, C.; Zhang, J. Identification of leaf chloroplast-specific promoter to efficiently control of Colorado potato beetle with reduced dsRNA accumulation in potato tubers. Pest Manag. Sci. 2023, 79, 3326–3333. [Google Scholar] [CrossRef]
  53. Ren, B.; Cao, J.; He, Y.; Yang, S.; Zhang, J. Assessment on effects of transplastomic potato plants expressing Colorado potato beetle β-Actin double-stranded RNAs for three non-target pests. Pestic. Biochem. Physiol. 2021, 178, 104909. [Google Scholar] [CrossRef]
  54. Xu, W.; Zhang, M.; Li, Y.; He, W.; Li, S.; Zhang, J. Complete protection from Henosepilachna vigintioctopunctata by expressing long double-stranded RNAs in potato plastids. J. Integr. Plant Biol. 2023, 65, 1003–1011. [Google Scholar] [CrossRef] [PubMed]
  55. Jin, S.; Singh, N.D.; Li, L.; Zhang, X.; Daniell, H. Engineered chloroplast dsRNA silences cytochrome p450 monooxygenase, V-ATPase and chitin synthase genes in the insect gut and disrupts Helicoverpa zea larval development and pupation. Plant Biotechnol. J. 2015, 13, 435–446. [Google Scholar] [CrossRef] [PubMed]
  56. Dong, Y.; Wu, M.; Zhang, Q.; Fu, J.; Loiacono, F.V.; Yang, Y.; Wang, Z.; Li, S.; Chang, L.; Bock, R.; et al. Control of a sap-sucking insect pest by plastid-mediated RNA interference. Mol. Plant 2022, 15, 1176–1191. [Google Scholar] [CrossRef]
  57. Dong, Y.; Yang, Y.; Wang, Z.; Wu, M.; Fu, J.; Guo, J.; Chang, L.; Zhang, J. Inaccessibility to double-stranded RNAs in plastids restricts RNA interference in Bemisia tabaci (whitefly). Pest Manag. Sci. 2020, 76, 3168–3176. [Google Scholar] [CrossRef]
  58. Burke, W.G.; Kaplanoglu, E.; Kolotilin, I.; Menassa, R.; Donly, C. RNA interference in the tobacco hornworm, Manduca sexta, using plastid-encoded long double-stranded RNA. Front. Plant Sci. 2019, 10, 313. [Google Scholar] [CrossRef]
  59. Fu, J.; Xu, S.; Lu, H.; Li, F.; Li, S.; Chang, L.; Heckel, D.G.; Bock, R.; Zhang, J. Resistance to RNA interference by plant-derived double-stranded RNAs but not plant-derived short interfering RNAs in Helicoverpa armigera. Plant Cell Environ. 2022, 45, 1930–1941. [Google Scholar] [CrossRef]
  60. He, W.; Xu, W.; Xu, L.; Fu, K.; Guo, W.; Bock, R.; Zhang, J. Length-dependent accumulation of double-stranded RNAs in plastids affects RNA interference efficiency in the Colorado potato beetle. J. Exp. Bot. 2020, 71, 2670–2677. [Google Scholar] [CrossRef]
  61. Bally, J.; McIntyre, G.J.; Doran, R.L.; Lee, K.; Perez, A.; Jung, H.; Naim, F.; Larrinua, I.M.; Narva, K.E.; Waterhouse, P.M. In-plant protection against Helicoverpa armigera by production of long hpRNA in chloroplasts. Front. Plant Sci. 2016, 7, 1453. [Google Scholar] [CrossRef]
  62. Sheridan, C. First self-amplifying mRNA vaccine approved. Nat. Biotechnol. 2024, 42, 4. [Google Scholar] [CrossRef]
  63. Baum, J.A.; Bogaert, T.; Clinton, W.; Heck, G.R.; Feldmann, P.; Ilagan, O.; Johnson, S.; Plaetinck, G.; Munyikwa, T.; Pleau, M.; et al. Control of coleopteran insect pests through RNA interference. Nat. Biotechnol. 2007, 25, 1322–1326. [Google Scholar] [CrossRef] [PubMed]
  64. Mao, Y.-B.; Cai, W.-J.; Wang, J.-W.; Hong, G.-J.; Tao, X.-Y.; Wang, L.-J.; Huang, Y.-P.; Chen, X.-Y. Silencing a cotton bollworm P450 monooxygenase gene by plant-mediated RNAi impairs larval tolerance of gossypol. Nat. Biotechnol. 2007, 25, 1307–1313. [Google Scholar] [CrossRef]
  65. Luo, J.; Liang, S.; Li, J.; Xu, Z.; Li, L.; Zhu, B.; Li, Z.; Lei, C.; Lindsey, K.; Chen, L.; et al. A transgenic strategy for controlling plant bugs (Adelphocoris suturalis) through expression of double-stranded RNA homologous to fatty acyl-coenzyme A reductase in cotton. New Phytol. 2017, 215, 1173–1185. [Google Scholar] [CrossRef] [PubMed]
  66. Sun, Y.; Sparks, C.; Jones, H.; Riley, M.; Francis, F.; Du, W.; Xia, L. Silencing an essential gene involved in infestation and digestion in grain aphid through plant-mediated RNA interference generates aphid-resistant wheat plants. Plant Biotechnol. J. 2019, 17, 852–854. [Google Scholar] [CrossRef]
  67. Mao, C.; Zhu, X.; Wang, P.; Sun, Y.; Huang, R.; Zhao, M.; Hull, J.J.; Lin, Y.; Zhou, F.; Chen, H.; et al. Transgenic double-stranded RNA rice, a potential strategy for controlling striped stem borer (Chilo suppressalis). Pest Manag. Sci. 2022, 78, 785–792. [Google Scholar] [CrossRef]
  68. Shukla, J.N.; Kalsi, M.; Sethi, A.; Narva, K.E.; Fishilevich, E.; Singh, S.; Mogilicherla, K.; Palli, S.R. Reduced stability and intracellular transport of dsRNA contribute to poor RNAi response in lepidopteran insects. RNA Biol. 2016, 13, 656–669. [Google Scholar] [CrossRef]
  69. Li, H.; Bowling, A.J.; Gandra, P.; Rangasamy, M.; Pence, H.E.; McEwan, R.E.; Khajuria, C.; Siegfried, B.D.; Narva, K.E. Systemic RNAi in western corn rootworm, Diabrotica virgifera virgifera, does not involve transitive pathways. Insect Sci. 2018, 25, 45–56. [Google Scholar] [CrossRef]
  70. Firmino, A.A.P.; Pinheiro, D.H.; Moreira-Pinto, C.E.; Antonino, J.D.; Macedo, L.L.P.; Martins-de-Sa, D.; Arraes, F.B.M.; Coelho, R.R.; Fonseca, F.C.A.; Silva, M.C.M.; et al. RNAi-mediated suppression of laccase2 impairs cuticle tanning and molting in the cotton boll weevil (Anthonomus grandis). Front. Physiol. 2020, 11, 591569. [Google Scholar] [CrossRef]
  71. Vasquez, D.D.N.; Pinheiro, D.H.; Teixeira, L.A.; Moreira-Pinto, C.E.; Macedo, L.L.P.; Salles-Filho, A.L.O.; Silva, M.C.M.; Lourenço-Tessutti, I.T.; Morgante, C.V.; Silva, L.P.; et al. Simultaneous silencing of juvenile hormone metabolism genes through RNAi interrupts metamorphosis in the cotton boll weevil. Front. Mol. Biosci. 2023, 10, 1073721. [Google Scholar] [CrossRef]
  72. Moreira-Pinto, C.E.; Coelho, R.R.; Leite, A.G.B.; Silveira, D.A.; de Souza, D.A.; Lopes, R.B.; Macedo, L.L.P.; Silva, M.C.M.; Ribeiro, T.P.; Morgante, C.V.; et al. Increasing Anthonomus grandis susceptibility to Metarhizium anisopliae through RNAi-induced AgraRelish knockdown: A perspective to combine biocontrol and biotechnology. Pest Manag. Sci. 2021, 77, 4054–4063. [Google Scholar] [CrossRef] [PubMed]
  73. Ramkumar, G.; Asokan, R.; Prasannakumar, N.R.; Kariyanna, B.; Karthi, S.; Alwahibi, M.S.; Elshikh, M.S.; Abdel-Megeed, A.; Ghaith, A.; Senthil-Nathan, S.; et al. RNA interference suppression of v-ATPase B and juvenile hormone binding protein genes through topically applied dsRNA on tomato leaves: Developing biopesticides to control the South American pinworm, Tuta absoluta (Lepidoptera: Gelechiidae). Front Physiol. 2021, 12, 742871. [Google Scholar] [CrossRef] [PubMed]
  74. Ni, M.; Ma, W.; Wang, X.; Gao, M.; Dai, Y.; Wei, X.; Zhang, L.; Peng, Y.; Chen, S.; Ding, L.; et al. Next-generation transgenic cotton: Pyramiding RNAi and Bt counters insect resistance. Plant Biotechnol. J. 2017, 15, 1204–1213. [Google Scholar] [CrossRef] [PubMed]
  75. Wei, J.-Z.; Lum, A.; Schepers, E.; Liu, L.; Weston, R.T.; McGinness, B.S.; Heckert, M.J.; Xie, W.; Kassa, A.; Bruck, D.; et al. Novel insecticidal proteins from ferns resemble insecticidal proteins from Bacillus thuringiensis. Proc. Natl. Acad. Sci. USA 2023, 120, e2306177120. [Google Scholar] [CrossRef]
  76. Mehlhorn, S.; Hunnekuhl, V.S.; Geibel, S.; Nauen, R.; Bucher, G. Establishing RNAi for basic research and pest control and identification of the most efficient target genes for pest control: A brief guide. Front. Zool. 2021, 18, 60. [Google Scholar] [CrossRef]
  77. Schutter, K.; Taning, C.N.T.; Van Daele, L.; Van Damme, E.J.M.; Dubruel, P.; Smagghe, G. RNAi-based biocontrol products: Market status, regulatory aspects, and risk assessment. Front. Insect Sci. 2022, 1, 818037. [Google Scholar] [CrossRef]
  78. Christiaens, O.; Whyard, S.; Vélez, A.M.; Smagghe, G. Double-stranded RNA technology to control insect pests: Current status and challenges. Front. Plant Sci. 2020, 11, 451. [Google Scholar] [CrossRef]
  79. Goodfellow, S.; Zhang, D.; Wang, M.B.; Zhang, R. Bacterium-mediated RNA interference: Potential application in plant protection. Plants 2019, 8, 572. [Google Scholar] [CrossRef]
  80. Chaudhary, S.; Dutta, T.K.; Tyagi, N.; Shivakumara, T.N.; Papolu, P.K.; Chobhe, K.A.; Rao, U. Host-induced silencing of Mi-msp-1 confers resistance to root-knot nematode Meloidogyne incognita in eggplant. Transgenic Res. 2019, 28, 327–340. [Google Scholar] [CrossRef]
  81. Moreira, V.J.V.; Pinheiro, D.H.; Lourenço-Tessutti, I.T.; Basso, M.F.; Lisei-de-Sa, M.E.; Silva, M.C.M.; Danchin, E.G.J.; Guimarães, P.M.; Grynberg, P.; Brasileiro, A.C.M.; et al. In planta RNAi targeting Meloidogyne incognita Minc16803 gene perturbs nematode parasitism and reduces plant susceptibility. J. Pest Sci. 2023, 97, 411–427. [Google Scholar] [CrossRef]
  82. Moreira, V.J.V.; Lourenço-Tessutti, I.T.; Basso, M.F.; Lisei-de-Sa, M.E.; Morgante, C.V.; Paes-de-Melo, B.; Arraes, F.B.M.; Martins-de-Sa, D.; Silva, M.C.M.; de Almeida Engler, J.; et al. Minc03328 effector gene downregulation severely affects Meloidogyne incognita parasitism in transgenic Arabidopsis thaliana. Planta 2022, 255, 44. [Google Scholar] [CrossRef]
  83. Mejias, J.; Chen, Y.; Bazin, J.; Truong, N.M.; Mulet, K.; Noureddine, Y.; Jaubert-Possamai, S.; Ranty-Roby, S.; Soulé, S.; Abad, P.; et al. Silencing the conserved small nuclear ribonucleoprotein SmD1 target gene alters susceptibility to root-knot nematodes in plants. Plant Physiol. 2022, 189, 1741–1756. [Google Scholar] [CrossRef]
  84. Lisei-de-Sá, M.E.; Rodrigues-Silva, P.L.; Morgante, C.V.; de Melo, B.P.; Lourenço-Tessutti, I.T.; Arraes, F.B.M.; Sousa, J.P.A.; Galbieri, R.; Amorim, R.M.S.; de Lins, C.B.J.; et al. Pyramiding dsRNAs increases phytonematode tolerance in cotton plants. Planta 2021, 254, 121. [Google Scholar] [CrossRef] [PubMed]
  85. Sidorova, T.; Miroshnichenko, D.; Kirov, I.; Pushin, A.; Dolgov, S. Effect of Grafting on Viral Resistance of Non-transgenic Plum Scion Combined With Transgenic PPV-Resistant Rootstock. Front. Plant Sci. 2021, 12, 621954. [Google Scholar] [CrossRef]
  86. Rodrigues-Silva, P.; Moura, S.; Dantas, L.; Domiciano, G.; Silva, M.; Campos, M.; Grossi-de-Sa, M. RNAi and Bt approaches to insect-pest control: Analyses and perspectives on trends in global patent publications. Biotechnol. Res. Innov. 2023, 6, e2022204. [Google Scholar] [CrossRef]
  87. Voloudakis, A.E.; Kaldis, A.; Patil, B.L. RNA-based vaccination of plants for control of viruses. Annu. Rev. Virol. 2022, 9, 521–548. [Google Scholar] [CrossRef]
  88. Beyene, G.; Chauhan, R.D.; Ilyas, M.; Wagaba, H.; Fauquet, C.M.; Miano, D.; Alicai, T.; Taylor, N.J. A virus-derived stacked RNAi construct confers robust resistance to cassava brown streak disease. Front. Plant Sci. 2017, 7, 2052. [Google Scholar] [CrossRef]
  89. Tabein, S.; Jansen, M.; Noris, E.; Vaira, A.M.; Marian, D.; Behjatnia, S.A.A.; Accotto, G.P.; Miozzi, L. The induction of an effective dsRNA-mediated resistance against Tomato spotted wilt virus by exogenous application of double-stranded RNA largely depends on the selection of the viral RNA target region. Front. Plant Sci. 2020, 11, 533338. [Google Scholar] [CrossRef]
  90. Walker, P.L.; Ziegler, D.J.; Giesbrecht, S.; McLoughlin, A.; Wan, J.; Khan, D.; Hoi, V.; Whyard, S.; Belmonte, M.F. Control of white mold (Sclerotinia sclerotiorum) through plant-mediated RNA interference. Sci. Rep. 2023, 13, 6477. [Google Scholar] [CrossRef]
  91. Jin, B.J.; Chun, H.J.; Choi, C.W.; Lee, S.H.; Cho, H.M.; Park, M.S.; Baek, D.; Park, S.-Y.; Lee, Y.-H.; Kim, M.C. Host-induced gene silencing is a promising biological tool to characterize the pathogenicity of Magnaporthe oryzae and control fungal disease in rice. Plant Cell Environ. 2023, 47, 319–336. [Google Scholar] [CrossRef]
  92. Pérez, C.E.B.; Cabral, G.B.; Aragão, F.J.L. Host-induced gene silencing for engineering resistance to Fusarium in soybean. Plant Pathol. 2021, 70, 417–425. [Google Scholar] [CrossRef]
  93. Wytinck, N.; Manchur, C.L.; Li, V.H.; Whyard, S.; Belmonte, M.F. dsRNA uptake in plant pests and pathogens: Insights into RNAi-based insect and fungal control technology. Plants 2020, 9, 1780. [Google Scholar] [CrossRef] [PubMed]
  94. Rusk, N. Prokaryotic RNAi. Nat. Methods 2012, 9, 220–221. [Google Scholar] [CrossRef] [PubMed]
  95. Murphy, K.A.; Tabuloc, C.A.; Cervantes, K.R.; Chiu, J.C. Ingestion of genetically modified yeast symbiont reduces fitness of an insect pest via RNA interference. Sci. Rep. 2016, 6, 22587. [Google Scholar] [CrossRef]
  96. Caccia, S.; Astarita, F.; Barra, E.; Di Lelio, I.; Varricchio, P.; Pennacchio, F. Enhancement of Bacillus thuringiensis toxicity by feeding Spodoptera littoralis larvae with bacteria expressing immune suppressive dsRNA. J. Pest Sci. 2020, 93, 303–314. [Google Scholar] [CrossRef]
  97. Jiang, Y.X.; Chen, J.Z.; Li, M.W.; Zha, B.H.; Huang, P.R.; Chu, X.M.; Chen, J.; Yang, G. The combination of Bacillus thuringiensis and its engineered strain expressing dsRNA increases the toxicity against Plutella xylostella. Int. J. Mol. Sci. 2021, 23, 444. [Google Scholar] [CrossRef]
  98. Kolliopoulou, A.; Taning, C.N.T.; Smagghe, G.; Swevers, L. Viral Delivery of dsRNA for Control of Insect Agricultural Pests and Vectors of Human Disease: Prospects and Challenges. Front. Physiol. 2017, 8, 399. [Google Scholar] [CrossRef] [PubMed]
  99. Xu, J.; Wang, X.; Li, Y.; Zeng, J.; Wang, G.; Deng, C.; Guo, W. Host-induced gene silencing of a regulator of G protein signalling gene (VdRGS1) confers resistance to Verticillium wilt in cotton. Plant Biotechnol. J. 2018, 16, 1629–1643. [Google Scholar] [CrossRef]
  100. Yang, Y.; Wang, S.; Zhao, C.; Jiang, X.; Gao, D. Responses of non-structural carbohydrates and biomass in plant to heavy metal treatment. Sci. Total Environ. 2024, 909, 168559. [Google Scholar] [CrossRef]
  101. Paciorek, T.; Chiapelli, B.J.; Wang, J.Y.; Paciorek, M.; Yang, H.; Sant, A.; Val, D.L.; Boddu, J.; Liu, K.; Gu, C.; et al. Targeted suppression of gibberellin biosynthetic genes ZmGA20ox3 and ZmGA20ox5 produces a short stature maize ideotype. Plant Biotechnol. J. 2022, 20, 1140–1153. [Google Scholar] [CrossRef]
  102. Basso, M.F.; Duarte, K.E.; Santiago, T.R.; de Souza, W.R.; Garcia, B.O.; da Cunha, B.D.B.; Kobayashi, A.K.; Molinari, H.B.C. Efficient genome editing and gene knockout in Setaria viridis with CRISPR/Cas9 directed gene editing by the non-homologous end-joining pathway. Plant Biotechnol. 2021, 38, 227–238. [Google Scholar] [CrossRef]
  103. Tuncel, A.; Pan, C.; Sprink, T.; Wilhelm, R.; Barrangou, R.; Li, L.; Shih, P.M.; Varshney, R.K.; Tripathi, L.; Van Eck, J.; et al. Genome-edited foods. Nat. Rev. Bioeng. 2023, 1, 799–816. [Google Scholar] [CrossRef]
  104. Siddiquee, R.; Pong, C.H.; Hall, R.M.; Ataide, S.F. A programmable seekRNA guides target selection by IS1111 and IS110 type insertion sequences. Nat. Commun. 2024, 15, 5235. [Google Scholar] [CrossRef]
  105. Gao, H.; Gadlage, M.J.; Lafitte, H.R.; Lenderts, B.; Yang, M.; Schroder, M.; Farrell, J.; Snopek, K.; Peterson, D.; Feigenbutz, L.; et al. Superior field performance of waxy corn engineered using CRISPR-Cas9. Nat. Biotechnol. 2020, 38, 579–581. [Google Scholar] [CrossRef]
  106. Shelake, R.M.; Kadam, U.S.; Kumar, R.; Pramanik, D.; Singh, A.K.; Kim, J.Y. Engineering drought and salinity tolerance traits in crops through CRISPR-mediated genome editing: Targets, tools, challenges, and perspectives. Plant Commun. 2022, 3, 100417. [Google Scholar] [CrossRef] [PubMed]
  107. Usovsky, M.; Gamage, V.A.; Meinhardt, C.G.; Dietz, N.; Triller, M.; Basnet, P.; Gillman, J.D.; Bilyeu, K.D.; Song, Q.; Dhital, B.; et al. Loss-of-function of an α-SNAP gene confers resistance to soybean cyst nematode. Nat. Commun. 2023, 14, 7629. [Google Scholar] [CrossRef]
  108. Worrall, E.A.; Bravo-Cazar, A.; Nilon, A.T.; Fletcher, S.J.; Robinson, K.E.; Carr, J.P.; Mitter, N. Exogenous application of RNAi-inducing double-stranded RNA inhibits aphid-mediated transmission of a plant virus. Front. Plant Sci. 2019, 10, 265. [Google Scholar] [CrossRef]
  109. Mai, J.; Liao, L.; Ling, R.; Guo, X.; Lin, J.; Mo, B.; Chen, W.; Yu, Y. Study on RNAi-based herbicide for Mikania micrantha. Synth. and Syst. Biotechnol. 2021, 6, 437–445. [Google Scholar] [CrossRef]
  110. Qiao, L.; Lan, C.; Capriotti, L.; Ah-Fong, A.; Nino Sanchez, J.; Hamby, R.; Heller, J.; Zhao, H.; Glass, N.L.; Judelson, H.S.; et al. Spray-induced gene silencing for disease control is dependent on the efficiency of pathogen RNA uptake. Plant Biotechnol. J. 2021, 19, 1756–1768. [Google Scholar] [CrossRef]
  111. Bachman, P.; Fischer, J.; Song, Z.; Urbanczyk-Wochniak, E.; Watson, G. Environmental fate and dissipation of applied dsRNA in soil, aquatic systems, and plants. Front. Plant Sci. 2020, 11, 21. [Google Scholar] [CrossRef]
  112. Bennett, M.; Deikman, J.; Hendrix, B.; Iandolino, A. Barriers to efficient foliar uptake of dsRNA and molecular barriers to dsRNA activity in plant cells. Front. Plant Sci. 2020, 11, 816. [Google Scholar] [CrossRef] [PubMed]
  113. Hoang, B.T.L.; Fletcher, S.J.; Brosnan, C.A.; Ghodke, A.B.; Manzie, N.; Mitter, N. RNAi as a foliar spray: Efficiency and challenges to field applications. Int. J. Mol. Sci. 2022, 23, 6639. [Google Scholar] [CrossRef]
  114. Schutter, K.; Christiaens, O.; Taning, C.N.T.; Smagghe, G. Boosting dsRNA Delivery in Plant and Insect Cells with Peptide- and Polymer-Based Carriers: Case-Based Current Status and Future Perspectives; CABI: Oxfordshire, UK, 2021; pp. 102–116. [Google Scholar]
  115. Fletcher, S.J.; Reeves, P.T.; Hoang, B.T.; Mitter, N. A Perspective on RNAi-based biopesticides. Front. Plant Sci. 2020, 11, 51. [Google Scholar] [CrossRef] [PubMed]
  116. Rank, A.P.; Koch, A. Lab-to-field transition of RNA spray applications—How far are we? Front. Plant Sci. 2021, 12, 755203. [Google Scholar] [CrossRef] [PubMed]
  117. Jalaluddin, N.S.M.; Othman, R.Y.; Harikrishna, J.A. Global trends in research and commercialization of exogenous and endogenous RNAi technologies for crops. Crit. Rev. Biotechnol. 2019, 39, 67–78. [Google Scholar] [CrossRef]
  118. Hough, J.; Howard, J.D.; Brown, S.; Portwood, D.E.; Kilby, P.M.; Dickman, M.J. Strategies for the production of dsRNA biocontrols as alternatives to chemical pesticides. Front. Bioeng. Biotechnol. 2022, 10, 980592. [Google Scholar] [CrossRef]
  119. Necira, K.; Makki, M.; Sanz-García, E.; Canto, T.; Djilani-Khouadja, F.; Tenllado, F. Topical application of Escherichia coli-encapsulated dsRNA induces resistance in Nicotiana benthamiana to potato viruses and involves RDR6 and combined activities of DCL2 and DCL4. Plants 2021, 10, 644. [Google Scholar] [CrossRef]
  120. Killiny, N.; Hajeri, S.; Tiwari, S.; Gowda, S.; Stelinski, L.L. Double-stranded RNA uptake through topical application, mediates silencing of five CYP4 genes and suppresses insecticide resistance in Diaphorina citri. PLoS ONE 2014, 9, e110536. [Google Scholar] [CrossRef]
  121. Aalto, A.P.; Sarin, L.P.; van Dijk, A.A.; Saarma, M.; Poranen, M.M.; Arumäe, U.; Bamford, D.H. Large-scale production of dsRNA and siRNA pools for RNA interference utilizing bacteriophage phi6 RNA-dependent RNA polymerase. RNA 2007, 13, 422–429. [Google Scholar] [CrossRef]
  122. Huang, L.; Jin, J.; Deighan, P.; Kiner, E.; McReynolds, L.; Lieberman, J. Efficient and specific gene knockdown by small interfering RNAs produced in bacteria. Nat. Biotechnol. 2013, 31, 350–356. [Google Scholar] [CrossRef]
  123. Ghoshal, K.; Theilmann, J.; Reade, R.; Maghodia, A.; Rochon, D. Encapsidation of Host RNAs by Cucumber Necrosis Virus Coat Protein during both Agroinfiltration and Infection. J. Virol. 2015, 89, 10748–10761. [Google Scholar] [CrossRef]
  124. Wuthisathid, K.; Chaijarasphong, T.; Chotwiwatthanakun, C.; Somrit, M.; Sritunyalucksana, K.; Itsathitphaisarn, O. Co-expression of double-stranded RNA and viral capsid protein in the novel engineered Escherichia coli DualX-B15(DE3) strain. BMC Microbiol. 2021, 21, 88. [Google Scholar] [CrossRef]
  125. Singh, I.K.; Singh, S.; Mogilicherla, K.; Shukla, J.N.; Palli, S.R. Comparative analysis of double-stranded RNA degradation and processing in insects. Sci. Rep. 2017, 7, 17059. [Google Scholar] [CrossRef]
  126. Dong, Y.; Zhang, Q.; Mao, Y.; Wu, M.; Wang, Z.; Chang, L.; Zhang, J. Control of two insect pests by expression of a mismatch corrected double-stranded RNA in plants. Plant Biotechnol. J. 2024, 22, 2010–2019. [Google Scholar] [CrossRef] [PubMed]
  127. Hanamasagar, Y.; Ramesha, N.M.; Mahapatra, S.; Panigrahi, C.K.; Vidhya, C.S.; Agnihotri, N.; Satapathy, S.N. Advancing RNAi for Sustainable Insect Management: Targeted Solutions for Eco-friendly Pest Control. J. Exp. Agric. Int. 2024, 46, 740–775. [Google Scholar] [CrossRef]
  128. Ma, Z.; Zheng, Y.; Chao, Z.; Chen, H.; Zhang, Y.; Yin, M.; Shen, J.; Yan, S. Visualization of the process of a nanocarrier-mediated gene delivery: Stabilization, endocytosis and endosomal escape of genes for intracellular spreading. J. Nanobiotechnol. 2022, 20, 124. [Google Scholar] [CrossRef]
  129. Jalaluddin, N.S.M.; Asem, M.; Harikrishna, J.A.; Ahmad Fuaad, A.A.H. Recent progress on nanocarriers for topical-mediated RNAi strategies for crop protection—A review. Molecules 2023, 28, 2700. [Google Scholar] [CrossRef]
  130. Wang, X.; Ji, S.; Bi, S.; Tang, Y.; Zhang, G.; Yan, S.; Wan, F.; Lü, Z.; Liu, W. A promising approach to an environmentally friendly pest management solution: Nanocarrier-delivered dsRNA towards controlling the destructive invasive pest Tuta absoluta. Environ. Sci. Nano 2023, 10, 1003–1015. [Google Scholar] [CrossRef]
  131. Arjunan, N.; Thiruvengadam, V.; Sushil, S.N. Nanoparticle-mediated dsRNA delivery for precision insect pest control: A comprehensive review. Mol. Biol. Rep. 2024, 51, 355. [Google Scholar] [CrossRef]
  132. Afonin, K.A.; Viard, M.; Kagiampakis, I.; Case, C.L.; Dobrovolskaia, M.A.; Hofmann, J.; Vrzak, A.; Kireeva, M.; Kasprzak, W.K.; KewalRamani, V.N.; et al. Triggering of RNA interference with RNA-RNA, RNA-DNA, and DNA-RNA nanoparticles. ACS Nano 2015, 9, 251–259. [Google Scholar] [CrossRef]
  133. Zhou, H.; Wan, F.; Jian, Y.; Guo, F.; Zhang, M.; Shi, S.; Yang, L.; Li, S.; Liu, Y.; Ding, W. Chitosan/dsRNA polyplex nanoparticles advance environmental RNA interference efficiency through activating clathrin-dependent endocytosis. Int. J. Biol. Macromol. 2023, 253, 127021. [Google Scholar] [CrossRef]
  134. Almeida, A.A.; Santos, R.M.M.; Alves Rosa, M.A.; Pulcinelli, S.H.; John, V.M.; Santilli, C.V. MgAl-Layered Double Hydroxide Nanoparticles as Smart Nanofillers To Control the Rheological Properties and the Residual Porosity of Cement-Based Materials. ACS Appl. Nano Mater. 2022, 5, 7896–7907. [Google Scholar] [CrossRef]
  135. Kolge, H.; Kadam, K.; Galande, S.; Lanjekar, V.; Ghormade, V. New frontiers in pest control: Chitosan nanoparticles-shielded dsRNA as an effective topical RNAi spray for gram podborer biocontrol. ACS Appl. Bio Mater. 2021, 4, 5145–5157. [Google Scholar] [CrossRef]
  136. Islam, M.T.; Davis, Z.; Chen, L.; Englaender, J.; Zomorodi, S.; Frank, J.; Bartlett, K.; Somers, E.; Carballo, S.M.; Kester, M.; et al. Minicell-based fungal RNAi delivery for sustainable crop protection. Microb. Biotechnol. 2021, 14, 1847–1856. [Google Scholar] [CrossRef]
  137. Cai, Q.; He, B.; Wang, S.; Fletcher, S.; Niu, D.; Mitter, N.; Birch, P.R.J.; Jin, H. Message in a bubble: Shuttling small RNAs and proteins between cells and interacting organisms using extracellular vesicles. Annu. Rev. Plant Biol. 2021, 72, 497–524. [Google Scholar] [CrossRef]
  138. Amiri, A.; Bagherifar, R.; Ansari Dezfouli, E.; Kiaie, S.H.; Jafari, R.; Ramezani, R. Exosomes as bio-inspired nanocarriers for RNA delivery: Preparation and applications. J. Transl. Med. 2022, 20, 125. [Google Scholar] [CrossRef] [PubMed]
  139. Lu, Y.; Huang, W.; Li, M.; Zheng, A. Exosome-based carrier for RNA delivery: Progress and challenges. Pharmaceutics 2023, 15, 598. [Google Scholar] [CrossRef] [PubMed]
  140. Cai, Q.; Halilovic, L.; Shi, T.; Chen, A.; He, B.; Wu, H.; Jin, H. Extracellular vesicles: Cross-organismal RNA trafficking in plants, microbes, and mammalian cells. Extracell. Vesicles Circ. Nucl. Acids 2023, 4, 262–282. [Google Scholar] [CrossRef]
  141. Chen, Y.-X.; Cai, Q. Plant exosome-like nanovesicles and their role in the innovative delivery of RNA therapeutics. Biomedicines 2023, 11, 1806. [Google Scholar] [CrossRef]
  142. Yan, S.; Qian, J.; Cai, C.; Ma, Z.; Li, J.; Yin, M.; Ren, B.; Shen, J. Spray method application of transdermal dsRNA delivery system for efficient gene silencing and pest control on soybean aphid Aphis glycines. J. Pest Sci. 2020, 93, 449–459. [Google Scholar] [CrossRef]
  143. Mitter, N.; Worrall, E.A.; Robinson, K.E.; Li, P.; Jain, R.G.; Taochy, C.; Fletcher, S.J.; Carroll, B.J.; Lu, G.Q.; Xu, Z.P. Clay nanosheets for topical delivery of RNAi for sustained protection against plant viruses. Nat. Plants 2017, 3, 16207. [Google Scholar] [CrossRef] [PubMed]
  144. Zhang, H.; Demirer, G.S.; Zhang, H.; Ye, T.; Goh, N.S.; Aditham, A.J.; Cunningham, F.J.; Fan, C.; Landry, M.P. DNA nanostructures coordinate gene silencing in mature plants. Proc. Natl. Acad. Sci. USA 2019, 116, 7543–7548. [Google Scholar] [CrossRef] [PubMed]
  145. Torney, F.; Trewyn, B.G.; Lin, V.S.; Wang, K. Mesoporous silica nanoparticles deliver DNA and chemicals into plants. Nat. Nanotechnol. 2007, 2, 295–300. [Google Scholar] [CrossRef] [PubMed]
  146. Tenchov, R.; Bird, R.; Curtze, A.E.; Zhou, Q. Lipid nanoparticles-from liposomes to mRNA vaccine delivery, a landscape of research diversity and advancement. ACS Nano 2021, 15, 16982–17015. [Google Scholar] [CrossRef]
  147. Buchman, Y.K.; Lellouche, E.; Zigdon, S.; Bechor, M.; Michaeli, S.; Lellouche, J.-P. Silica Nanoparticles and Polyethyleneimine (PEI)-Mediated Functionalization: A New Method of PEI Covalent Attachment for siRNA Delivery Applications. Bioconjugate Chem. 2013, 24, 2076–2087. [Google Scholar] [CrossRef]
  148. Schwartz, S.H.; Hendrix, B.; Hoffer, P.; Sanders, R.A.; Zheng, W. Carbon dots for efficient small interfering RNA delivery and gene silencing in plants. Plant Physiol. 2020, 184, 647–657. [Google Scholar] [CrossRef]
  149. Zhang, H.; Cao, Y.; Xu, D.; Goh, N.S.; Demirer, G.S.; Cestellos-Blanco, S.; Chen, Y.; Landry, M.P.; Yang, P. Gold-nanocluster-mediated delivery of siRNA to intact plant cells for efficient gene knockdown. Nano Lett. 2021, 21, 5859–5866. [Google Scholar] [CrossRef]
  150. Demirer, G.S.; Zhang, H.; Goh, N.S.; Pinals, R.L.; Chang, R.; Landry, M.P. Carbon nanocarriers deliver siRNA to intact plant cells for efficient gene knockdown. Sci. Adv. 2020, 6, eaaz0495. [Google Scholar] [CrossRef]
  151. McGraw, E.; Roberts, J.D.; Kunte, N.; Westerfield, M.; Streety, X.; Held, D.; Avila, L.A. Insight into cellular uptake and transcytosis of peptide nanoparticles in Spodoptera frugiperda cells and isolated midgut. ACS Omega 2022, 7, 10933–10943. [Google Scholar] [CrossRef]
  152. Kolge, H.; Kadam, K.; Ghormade, V. Chitosan nanocarriers mediated dsRNA delivery in gene silencing for Helicoverpa armigera biocontrol. Pestic. Biochem. Physiol. 2023, 189, 105292. [Google Scholar] [CrossRef]
  153. Sandal, S.; Singh, S.; Bansal, G.; Kaur, R.; Mogilicherla, K.; Pandher, S.; Roy, A.; Kaur, G.; Rathore, P.; Kalia, A. Nanoparticle-shielded dsRNA delivery for enhancing RNAi efficiency in cotton spotted bollworm Earias vittella (Lepidoptera: Nolidae). Int. J. Mol. Sci. 2023, 24, 9161. [Google Scholar] [CrossRef]
  154. Ma, Z.; Garrido-Maestu, A.; Lee, C.; Chon, J.; Jeong, D.; Yue, Y.; Sung, K.; Park, Y.; Jeong, K.C. Comprehensive in vitro and in vivo risk assessments of chitosan microparticles using human epithelial cells and Caenorhabditis elegans. J. Hazard. Mater. 2018, 341, 248–256. [Google Scholar] [CrossRef]
  155. Zoe, L.H.; David, S.R.; Rajabalaya, R. Chitosan nanoparticle toxicity: A comprehensive literature review of in vivo and in vitro assessments for medical applications. Toxicol. Rep. 2023, 11, 83–106. [Google Scholar] [CrossRef] [PubMed]
  156. Yan, S.; Hu, Q.; Li, J.; Chao, Z.; Cai, C.; Yin, M.; Du, X.; Shen, J. A Star Polycation Acts as a Drug Nanocarrier to Improve the Toxicity and Persistence of Botanical Pesticides. ACS Sustain. Chem. Eng. 2019, 7, 17406–17413. [Google Scholar] [CrossRef]
  157. Avila, L.A.; Chandrasekar, R.; Wilkinson, K.E.; Balthazor, J.; Heerman, M.; Bechard, J.; Brown, S.; Park, Y.; Dhar, S.; Reeck, G.R.; et al. Delivery of lethal dsRNAs in insect diets by branched amphiphilic peptide capsules. J. Control. Release 2018, 273, 139–146. [Google Scholar] [CrossRef]
  158. Santos, V.S.; Badan Ribeiro, A.P.; Andrade Santana, M.H. Solid lipid nanoparticles as carriers for lipophilic compounds for applications in foods. Food Res. Int. 2019, 122, 610–626. [Google Scholar] [CrossRef] [PubMed]
  159. Wessel, E.M.; Tomich, J.M.; Todd, R.B. Biodegradable Drug-Delivery Peptide Nanocapsules. ACS Omega 2019, 4, 20059–20063. [Google Scholar] [CrossRef]
  160. Dhandapani, R.K.; Gurusamy, D.; Palli, S.R. Development of catechin, poly-l-lysine, and double-stranded RNA nanoparticles. ACS Appl. Bio Mater. 2021, 4, 4310–4318. [Google Scholar] [CrossRef]
  161. Khan, F.; Kim, M.; Kim, Y. Greenhouse test of spraying dsRNA to control the western flower thrips, Frankliniella occidentalis, infesting hot peppers. BMC Biotechnol. 2023, 23, 10. [Google Scholar] [CrossRef]
  162. Carroll, E.; Kunte, N.; McGraw, E.; Gautam, S.; Range, R.; Noveron-Nunez, J.A.; Held, D.W.; Avila, L.A. Gene silencing in adult Popillia japonica through feeding of double-stranded RNA (dsRNA) complexed with branched amphiphilic peptide capsules (BAPCs). Front. Insect Sci. 2023, 3, 1151789. [Google Scholar] [CrossRef]
  163. Su, C.; Liu, S.; Sun, M.; Yu, Q.; Li, C.; Graham, R.I.; Wang, X.; Wang, X.; Xu, P.; Ren, G. Delivery of methoprene-tolerant dsRNA to improve RNAi efficiency by modified liposomes for pest control. ACS Appl. Mater. Interfaces 2023, 15, 13576–13588. [Google Scholar] [CrossRef]
  164. Evangelista, A.R.; Amoroso, C.G.; Nitride, C.; Andolfo, G. Seed storage allergens tackled via next-generation research assistant. Front. Food Sci. Technol. 2024, 4, 1372770. [Google Scholar] [CrossRef]
  165. Strachan, J.B.; Dyett, B.P.; Nasa, Z.; Valery, C.; Conn, C.E. Toxicity and cellular uptake of lipid nanoparticles of different structure and composition. J. Colloid Interface Sci. 2020, 576, 241–251. [Google Scholar] [CrossRef] [PubMed]
  166. Sharma, S.; Parveen, R.; Chatterji, B.P. Toxicology of Nanoparticles in Drug Delivery. Curr. Pathobiol. Rep. 2021, 9, 133–144. [Google Scholar] [CrossRef]
  167. Jain, R.G.; Fletcher, S.J.; Manzie, N.; Robinson, K.E.; Li, P.; Lu, E.; Brosnan, C.A.; Xu, Z.P.; Mitter, N. Foliar application of clay-delivered RNA interference for whitefly control. Nat. Plants 2022, 8, 535–548. [Google Scholar] [CrossRef] [PubMed]
  168. Li, M.; Sun, X.; Yin, M.; Shen, J.; Yan, S. Recent advances in nanoparticle-mediated co-delivery system: A promising strategy in medical and agricultural field. Int. J. Mol. Sci. 2023, 24, 5121. [Google Scholar] [CrossRef]
  169. Zheng, Y.; Hu, Y.; Yan, S.; Zhou, H.; Song, D.; Yin, M.; Shen, J. A polymer/detergent formulation improves dsRNA penetration through the body wall and RNAi-induced mortality in the soybean aphid Aphis glycines. Pest Manag. Sci. 2019, 75, 1993–1999. [Google Scholar] [CrossRef]
  170. Lyu, Z.; Xiong, M.; Mao, J.; Li, W.; Jiang, G.; Zhang, W. A dsRNA delivery system based on the rosin-modified polyethylene glycol and chitosan induces gene silencing and mortality in Nilaparvata lugens. Pest Manag. Sci. 2023, 79, 1518–1527. [Google Scholar] [CrossRef]
  171. De Vega, D.; Holden, N.; Hedley, P.E.; Morris, J.; Luna, E.; Newton, A. Chitosan primes plant defence mechanisms against Botrytis cinerea, including expression of Avr9/Cf-9 rapidly elicited genes. Plant Cell Environ. 2021, 44, 290–303. [Google Scholar] [CrossRef]
  172. Noman, M.; Ahmed, T.; Wang, J.; Ijaz, M.; Shahid, M.; Islam, M.S.; Azizullah; Manzoor, I.; Li, D.; Song, F. Nano-enabled crop resilience against pathogens: Potential, mechanisms and strategies. Crop Health 2023, 1, 15. [Google Scholar] [CrossRef]
  173. Wang, M.; Dean, R.A. Host induced gene silencing of Magnaporthe oryzae by targeting pathogenicity and development genes to control rice blast disease. Front. Plant Sci. 2022, 13, 959641. [Google Scholar] [CrossRef] [PubMed]
  174. Timmermann, T.; Poupin, M.J.; Vega, A.; Urrutia, C.; Ruz, G.A.; González, B. Gene networks underlying the early regulation of Paraburkholderia phytofirmans PsJN induced systemic resistance in Arabidopsis. PLoS ONE 2019, 14, e0221358. [Google Scholar] [CrossRef] [PubMed]
  175. Bao, W.; Wang, J.; Wang, Q.; O’Hare, D.; Wan, Y. Layered Double Hydroxide Nanotransporter for Molecule Delivery to Intact Plant Cells. Sci. Rep. 2016, 6, 26738. [Google Scholar] [CrossRef]
  176. Petrônio, M.S.; Barros-Alexandrino, T.T.; Lima, A.M.F.; Assis, O.B.G.; Inoue-Nagata, A.K.; Nakasu, E.Y.T.; Tiera, M.J.; Pilon, L. Physicochemical and toxicity investigation of chitosan-based dsRNA nanocarrier formation. Biointerface Res. Appl. Chem. 2022, 12, 5266–5279. [Google Scholar]
  177. Lv, H.; Li, X.; Li, J.; Yu, C.; Zeng, Q.; Ning, G.; Wan, H.; Li, J.; Ma, K.; He, S. Overcoming resistance in insect pest with a nanoparticle-mediated dsRNA and insecticide co-delivery system. Chem. Eng. J. 2023, 475, 146239. [Google Scholar] [CrossRef]
  178. Li, X.; Xiao, J.; Cheng, X.; Zhang, H.; Zheng, W. Nanomaterial-encapsulated dsRNA of ecdysone-induced early gene E75, a potential RNAi-based SIT strategy for pest control against Bactrocera dorsalis. Int. J. Biol. Macromol. 2024, 263, 130607. [Google Scholar] [CrossRef]
  179. Ivashuta, S.; Zhang, Y.; Wiggins, B.E.; Ramaseshadri, P.; Segers, G.C.; Johnson, S.; Meyer, S.E.; Kerstetter, R.A.; McNulty, B.C.; Bolognesi, R.; et al. Environmental RNAi in herbivorous insects. RNA 2015, 21, 840–850. [Google Scholar] [CrossRef]
  180. Okuda, T.; Niidome, T.; Aoyagi, H. Cytosolic soluble proteins induce DNA release from DNA—Gene carrier complexes. J. Control. Release 2004, 98, 325–332. [Google Scholar] [CrossRef]
  181. Bertschinger, M.; Backliwal, G.; Schertenleib, A.; Jordan, M.; Hacker, D.L.; Wurm, F.M. Disassembly of polyethylenimine-DNA particles in vitro: Implications for polyethylenimine-mediated DNA delivery. J. Control. Release 2006, 116, 96–104. [Google Scholar] [CrossRef]
  182. Zuckerman, J.E.; Choi, C.H.; Han, H.; Davis, M.E. Polycation-siRNA nanoparticles can disassemble at the kidney glomerular basement membrane. Proc. Natl. Acad. Sci. USA 2012, 109, 3137–3142. [Google Scholar] [CrossRef]
  183. Herdiana, Y.; Wathoni, N.; Shamsuddin, S.; Muchtaridi, M. Drug release study of the chitosan-based nanoparticles. Heliyon 2022, 8, e08674. [Google Scholar] [CrossRef] [PubMed]
  184. Bakhshian Nik, A.; Zare, H.; Razavi, S.; Mohammadi, H.; Torab Ahmadi, P.; Yazdani, N.; Bayandori, M.; Rabiee, N.; Izadi Mobarakeh, J. Smart drug delivery: Capping strategies for mesoporous silica nanoparticles. Microporous Mesoporous Mater. 2020, 299, 110115. [Google Scholar] [CrossRef]
  185. Chu, D.; Xu, S.; Li, X.; Li, H.; Miao, X.; An, S.; Guan, R. Combination of diflubenzuron and RNAi technology to improve the control effect of Helicoverpa armigera. Entomol. Gen. 2024, 44, 223–232. [Google Scholar] [CrossRef]
  186. Hou, R.; Li, C.; Tan, Y.; Wang, Y.; Huang, S.; Zhao, C.; Zhang, Z. Eco-friendly O-carboxymethyl chitosan base chlorfenapyr nanopesticide for effective pest control and reduced toxicity to honey bees. Int. J. Biol. Macromol. 2023, 224, 972–983. [Google Scholar] [CrossRef]
  187. Biedenkopf, D.; Will, T.; Knauer, T.; Jelonek, L.; Furch, A.C.U.; Busche, T.; Koch, A. Systemic spreading of exogenous applied RNA biopesticides in the crop plant Hordeum vulgare. ExRNA 2020, 2, 12. [Google Scholar] [CrossRef]
  188. Hunter, W.B.; Glick, E.; Paldi, N.; Bextine, B.R. Advances in RNA interference: dsRNA treatment in trees and grapevines for insect pest suppression. Southwest. Entomol. 2012, 37, 85–87. [Google Scholar]
  189. Baum, J.A.; Roberts, J.K. Chapter Five—Progress Towards RNAi-Mediated Insect Pest Management. In Advances in Insect Physiology; Dhadialla, T.S., Gill, S.S., Eds.; Academic Press: Cambridge, MA, USA, 2014; Volume 47, pp. 249–295. [Google Scholar]
  190. Ghosh, S.K.B.; Hunter, W.B.; Park, A.L.; Gundersen-Rindal, D.E. Double strand RNA delivery system for plant-sap-feeding insects. PLoS ONE 2017, 12, e0171861. [Google Scholar] [CrossRef]
  191. Andrade, E.; Hunter, W. RNA interference—Natural gene-based technology for highly specific pest control (HiSPeC). In RNA Interference; Ibrokhim, Y.A., Ed.; IntechOpen: Rijeka, Croatia, 2016; Chapter 19. [Google Scholar]
  192. Rodrigues, T.B.; Mishra, S.K.; Sridharan, K.; Barnes, E.R.; Alyokhin, A.; Tuttle, R.; Kokulapalan, W.; Garby, D.; Skizim, N.J.; Tang, Y.-W.; et al. First sprayable double-stranded RNA-based biopesticide product targets proteasome subunit beta type-5 in colorado potato beetle (Leptinotarsa decemlineata). Front. Plant Sci. 2021, 12, 728652. [Google Scholar] [CrossRef]
  193. Christiaens, O.; Tardajos, M.G.; Martinez Reyna, Z.L.; Dash, M.; Dubruel, P.; Smagghe, G. Increased RNAi efficacy in Spodoptera exigua via the formulation of dsRNA with guanylated polymers. Front. Physiol. 2018, 9, 316. [Google Scholar] [CrossRef]
  194. Gurusamy, D.; Mogilicherla, K.; Palli, S.R. Chitosan nanoparticles help double-stranded RNA escape from endosomes and improve RNA interference in the fall armyworm, Spodoptera frugiperda. Arch. Insect Biochem. Physiol. 2020, 104, e21677. [Google Scholar] [CrossRef]
  195. Pei, Y.; Hao, H.; Zuo, Y.; Xue, Y.; Aioub, A.A.A.; Hu, Z. Functional validation of CYP304A1 associated with haedoxan A detoxification in Aedes albopictus by RNAi and transgenic drosophila. Pest Manag. Sci. 2023, 79, 447–453. [Google Scholar] [CrossRef] [PubMed]
  196. Yan, S.; Shen, J. Application of nanoparticle-mediated RNAi for efficient gene silencing and pest control on soybean aphids. Methods Mol. Biol. 2022, 2360, 307–315. [Google Scholar] [CrossRef]
  197. Zhang, J.; Li, H.; Zhong, X.; Tian, J.; Segers, A.; Xia, L.; Francis, F. RNA-interference-mediated aphid control in crop plants: A review. Agriculture 2022, 12, 2108. [Google Scholar] [CrossRef]
  198. Dhandapani, R.K.; Gurusamy, D.; Palli, S.R. Protamine-lipid-dsRNA nanoparticles improve RNAi efficiency in the Fall Armyworm, Spodoptera frugiperda. J. Agric. Food Chem. 2022, 70, 6634–6643. [Google Scholar] [CrossRef]
  199. Niu, L.; Yan, H.; Sun, Y.; Zhang, D.; Ma, W.; Lin, Y. Nanoparticle facilitated stacked-dsRNA improves suppression of the Lepidoperan pest Chilo suppresallis. Pesticide Biochem. Physiol. 2022, 187, 105183. [Google Scholar] [CrossRef]
  200. Sharma, Y.; Padha, S.; Dhar, A.; Baweja, V.; Singh, I.K. RNAi-Based Biopesticides: A Review of Recent Studies in Lepidopteran Insects. Proc. Zool. Soc. 2023, 76, 373–381. [Google Scholar] [CrossRef]
  201. Geng, K.; Zhang, Y.; Zhao, X.; Zhang, W.; Guo, X.; He, L.; Liu, K.; Yang, H.; Hong, H.; Peng, J.; et al. Fluorescent nanoparticle-RNAi-mediated silencing of sterol carrier protein-2 gene expression suppresses the growth, development, and reproduction of Helicoverpa armigera. Nanomaterials 2023, 13, 245. [Google Scholar] [CrossRef] [PubMed]
  202. Long, G.-J.; Liu, X.-Z.; Guo, H.; Zhang, M.-Q.; Gong, L.-L.; Ma, Y.-F.; Dewer, Y.; Mo, W.-J.; Ding, L.-W.; Wang, Q.; et al. Oral-based nanoparticle-wrapped dsRNA delivery system: A promising approach for controlling an urban pest, Blattella germanica. J. Pest Sci. 2023, 97, 739–775. [Google Scholar] [CrossRef]
  203. Jain, M.; Bansal, J.; Rajkumar, M.S.; Garg, R. An integrated transcriptome mapping the regulatory network of coding and long non-coding RNAs provides a genomics resource in chickpea. Commun. Biol. 2022, 5, 1106. [Google Scholar] [CrossRef]
  204. Pallis, S.; Alyokhin, A.; Manley, B.; Rodrigues, T.; Barnes, E.; Narva, K. Effects of low doses of a novel dsRNA-based biopesticide (Calantha) on the Colorado potato beetle. J. Econ. Entomol. 2023, 116, 456–461. [Google Scholar] [CrossRef]
  205. Christiaens, O.; Smagghe, G. The challenge of RNAi-mediated control of hemipterans. Curr. Opin. Insect Sci. 2014, 6, 15–21. [Google Scholar] [CrossRef] [PubMed]
  206. Rossi, M.; Ottati, S.; Bucci, L.; Fusco, A.; Abbà, S.; Bosco, D.; Marzachì, C.; Galetto, L. Lab-scale method for plant-mediated delivery of dsRNAs to phloem-feeding leafhoppers. J. Pest Sci. 2024, 97, 455–467. [Google Scholar] [CrossRef]
  207. Li, H.; Guan, R.; Guo, H.; Miao, X. New insights into an RNAi approach for plant defence against piercing-sucking and stem-borer insect pests. Plant Cell Environ. 2015, 38, 2277–2285. [Google Scholar] [CrossRef] [PubMed]
  208. Wise, J.C.; Wise, A.G.; Rakotondravelo, M.; Vandervoort, C.; Seeve, C.; Fabbri, B. Trunk injection delivery of dsRNA for RNAi-based pest control in apple trees. Pest Manag. Sci. 2022, 78, 3528–3533. [Google Scholar] [CrossRef]
  209. Harrison, J.F. Insect acid-base physiology. Annu. Rev. Entomol. 2001, 46, 221–250. [Google Scholar] [CrossRef]
  210. Gao, W.; Chan, J.M.; Farokhzad, O.C. pH-Responsive nanoparticles for drug delivery. Mol. Pharm. 2010, 7, 1913–1920. [Google Scholar] [CrossRef] [PubMed]
  211. Zotti, M.; Dos Santos, E.A.; Cagliari, D.; Christiaens, O.; Taning, C.N.T.; Smagghe, G. RNA interference technology in crop protection against arthropod pests, pathogens and nematodes. Pest Manag. Sci. 2018, 74, 1239–1250. [Google Scholar] [CrossRef]
  212. Banerjee, S.; Banerjee, A.; Gill, S.S.; Gupta, O.P.; Dahuja, A.; Jain, P.K.; Sirohi, A. RNA interference: A novel source of resistance to combat plant parasitic nematodes. Front. Plant Sci. 2017, 8, 834. [Google Scholar] [CrossRef]
  213. Iqbal, S.; Fosu-Nyarko, J.; Jones, M.G.K. Attempt to silence genes of the RNAi pathways of the root-knot nematode, Meloidogyne incognita results in diverse responses including increase and no change in expression of some genes. Front. Plant Sci. 2020, 11, 328. [Google Scholar] [CrossRef]
  214. Mitter, N.; Worrall, E.A.; Robinson, K.E.; Xu, Z.P.; Carroll, B.J. Induction of virus resistance by exogenous application of double-stranded RNA. Curr. Opin. Virol. 2017, 26, 49–55. [Google Scholar] [CrossRef]
  215. Delgado-Martín, J.; Ruiz, L.; Janssen, D.; Velasco, L. Exogenous application of dsRNA for the control of viruses in cucurbits. Front. Plant Sci. 2022, 13, 895953. [Google Scholar] [CrossRef] [PubMed]
  216. Konakalla, N.C.; Kaldis, A.; Berbati, M.; Masarapu, H.; Voloudakis, A.E. Exogenous application of double-stranded RNA molecules from TMV p126 and CP genes confers resistance against TMV in tobacco. Planta 2016, 244, 961–969. [Google Scholar] [CrossRef]
  217. Rego-Machado, C.M.; Nakasu, E.Y.T.; Silva, J.M.F.; Lucinda, N.; Nagata, T.; Inoue-Nagata, A.K. siRNA biogenesis and advances in topically applied dsRNA for controlling virus infections in tomato plants. Sci. Rep. 2020, 10, 22277. [Google Scholar] [CrossRef]
  218. Kalyandurg, P.B.; Sundararajan, P.; Dubey, M.; Ghadamgahi, F.; Zahid, M.A.; Whisson, S.C.; Vetukuri, R.R. Spray-induced gene silencing as a potential tool to control potato late blight disease. Phytopathol. Res. 2021, 111, 2168–2175. [Google Scholar] [CrossRef] [PubMed]
  219. Degnan, R.M.; McTaggart, A.R.; Shuey, L.S.; Pame, L.J.S.; Smith, G.R.; Gardiner, D.M.; Nock, V.; Soffe, R.; Sale, S.; Garrill, A.; et al. Exogenous double-stranded RNA inhibits the infection physiology of rust fungi to reduce symptoms in planta. Mol. Plant Pathol. 2023, 24, 191–207. [Google Scholar] [CrossRef] [PubMed]
  220. Song, X.S.; Gu, K.X.; Duan, X.X.; Xiao, X.M.; Hou, Y.P.; Duan, Y.B.; Wang, J.X.; Yu, N.; Zhou, M.G. Secondary amplification of siRNA machinery limits the application of spray-induced gene silencing. Mol. Plant Pathol. 2018, 19, 2543–2560. [Google Scholar] [CrossRef]
  221. Ackah, M.; Jin, X.; Zhang, Q.; Amoako, F.K.; Wang, L.; Attaribo, T.; Zhao, M.; Yuan, F.; Herman, R.A.; Qiu, C.; et al. Long noncoding RNA transcriptome analysis reveals novel lncRNAs in Morus alba ‘Yu-711’ response to drought stress. Plant Genome 2024, 17, e20273. [Google Scholar] [CrossRef]
  222. Mosa, M.A.; Youssef, K. Topical delivery of host induced RNAi silencing by layered double hydroxide nanosheets: An efficient tool to decipher pathogenicity gene function of Fusarium crown and root rot in tomato. Physiol. Mol. Plant Pathol. 2021, 115, 101684. [Google Scholar] [CrossRef]
  223. Niño-Sánchez, J.; Sambasivam, P.T.; Sawyer, A.; Hamby, R.; Chen, A.; Czislowski, E.; Li, P.; Manzie, N.; Gardiner, D.M.; Ford, R.; et al. BioClay™ prolongs RNA interference-mediated crop protection against Botrytis cinerea. J. Integr. Plant Biol. 2022, 64, 2187–2198. [Google Scholar] [CrossRef]
  224. Koch, A.; Kumar, N.; Weber, L.; Keller, H.; Imani, J.; Kogel, K.H. Host-induced gene silencing of cytochrome P450 lanosterol C14α-demethylase-encoding genes confers strong resistance to Fusarium species. Proc. Natl. Acad. Sci. USA 2013, 110, 19324–19329. [Google Scholar] [CrossRef]
  225. Trippe, A.J. Guidelines for Preparing Patent Landscape Reports; WIPO: Geneva, Switzerland, 2015. [Google Scholar]
  226. Banerjee, G. Assessing visibility of research organization: A fuzzy analytic network process approach. J. Sci. Ind. Res. 2014, 73, 283–289. [Google Scholar]
  227. Park, S.; Lee, S.-J.; Jun, S. A network analysis model for selecting sustainable technology. Sustainability 2015, 7, 13126–13141. [Google Scholar] [CrossRef]
  228. Kay, L.; Youtie, J.; Shapira, P. Signs of things to come? What patent submissions by small and medium-sized enterprises say about corporate strategies in emerging technologies. Technol. Forecast. Soc. Change 2014, 85, 17–25. [Google Scholar] [CrossRef]
  229. Das, P.R.; Sherif, S.M. Application of exogenous dsRNAs-induced RNAi in agriculture: Challenges and triumphs. Front. Plant Sci. 2020, 11, 946. [Google Scholar] [CrossRef] [PubMed]
  230. Hernández-Soto, A.; Chacón-Cerdas, R. RNAi crop protection advances. Int. J. Mol. Sci. 2021, 22, 12148. [Google Scholar] [CrossRef]
  231. Government of Brazil. Lei nº 9.279, de 14 de maio de 1996. Regula Direitos e Obrigações Relativos à Propriedade Industrial; Diário Oficial da União: Brasília, DF, Brazil, 1996; Volume 134, p. 8353.
  232. US Patent Law. 35 USC Sec. 103. United States Code, Supplement 5, Title 35—Conditions for Patentability; Non-Obvious Subject Matter. 2001.
  233. US Patent Law. 35 USC Sec. 102. United States Code, Supplement 5, Title 35—Patents, Part II—Patentability of Inventions and Grant of Patents, Chapter 10—Patentability of Inventions. 2001.
  234. Federal Register, Environmental Protection Agency (EPA). 2001. Available online: https://www.govinfo.gov/content/pkg/FR-2001-01-05/pdf/01-84.pdf (accessed on 15 January 2025).
  235. Justia US Supreme Court. Certiorari to the United States Court of Appeals for the Federal Circuit. Assoc. for Molecular Pathology v. Myriad Genetics, Inc., 569 U.S. 576. Available online: https://supreme.justia.com/cases/federal/us/569/576/ (accessed on 15 January 2025).
  236. Bramlett, M.; Plaetinck, G.; Maienfisch, P. RNA-based biocontrols-A new paradigm in crop protection. Engineering 2020, 6, 522–527. [Google Scholar] [CrossRef]
  237. Rodrigues, T.B.; Petrick, J.S. Safety considerations for humans and other vertebrates regarding agricultural uses of externally applied RNA molecules. Front. Plant Sci. 2020, 11, 407. [Google Scholar] [CrossRef]
  238. Liu, S.; Jaouannet, M.; Dempsey, D.A.; Imani, J.; Coustau, C.; Kogel, K.H. RNA-based technologies for insect control in plant production. Biotechnol. Adv. 2020, 39, 107463. [Google Scholar] [CrossRef]
  239. Albright, V.C., 3rd; Wong, C.R.; Hellmich, R.L.; Coats, J.R. Dissipation of double-stranded RNA in aquatic microcosms. Environ. Toxicol. Chem. 2017, 36, 1249–1253. [Google Scholar] [CrossRef]
  240. Taning, C.N.; Arpaia, S.; Christiaens, O.; Dietz-Pfeilstetter, A.; Jones, H.; Mezzetti, B.; Sabbadini, S.; Sorteberg, H.G.; Sweet, J.; Ventura, V.; et al. RNA-based biocontrol compounds: Current status and perspectives to reach the market. Pest Manag. Sci. 2020, 76, 841–845. [Google Scholar] [CrossRef]
  241. Christiaens, O.; Sweet, J.; Dzhambazova, T.; Urru, I.; Smagghe, G.; Kostov, K.; Arpaia, S. Implementation of RNAi-based arthropod pest control: Environmental risks, potential for resistance and regulatory considerations. J. Pest Sci. 2022, 95, 1–15. [Google Scholar] [CrossRef]
  242. Mendelsohn, M.L.; Gathmann, A.; Kardassi, D.; Sachana, M.; Hopwood, E.M.; Dietz-Pfeilstetter, A.; Michelsen-Correa, S.; Fletcher, S.J.; Székács, A. Summary of discussions from the 2019 OECD conference on RNAi based pesticides. Front. Plant Sci. 2020, 11, 740. [Google Scholar] [CrossRef]
  243. OECD. Considerations for the Environmental Risk Assessment of the Application of Sprayed or Externally Applied dsRNA-based Pesticides. ENV/JM/MONO 2020 Paris 2020, 26. Available online: https://www.oecd.org/en/publications/considerations-for-the-environmental-risk-assessment-of-the-application-of-sprayed-or-externally-applied-ds-rna-based-pesticides_576d9ebb-en.html (accessed on 15 January 2025).
  244. Sherman, J.H.; Munyikwa, T.; Chan, S.Y.; Petrick, J.S.; Witwer, K.W.; Choudhuri, S. RNAi technologies in agricultural biotechnology: The Toxicology Forum 40th Annual Summer Meeting. Regul. Toxicol. Pharmacol. 2015, 73, 671–680. [Google Scholar] [CrossRef] [PubMed]
  245. Christiaens, O.; Dzhambazova, T.; Kostov, K.; Arpaia, S.; Joga, M.R.; Urru, I.; Sweet, J.; Smagghe, G. Literature review of baseline information on RNAi to support the environmental risk assessment of RNAi-based GM plants. EFSA Support. Publ. 2018, 15, 1424E. [Google Scholar] [CrossRef]
  246. Petrick, J.S.; Brower-Toland, B.; Jackson, A.L.; Kier, L.D. Safety assessment of food and feed from biotechnology-derived crops employing RNA-mediated gene regulation to achieve desired traits: A scientific review. Regul. Toxicol. Pharmacol. 2013, 66, 167–176. [Google Scholar] [CrossRef]
  247. Raybould, A.; Burns, A. Problem formulation for off-target effects of externally applied double-stranded RNA-based products for pest control. Front. Plant Sci. 2020, 11, 424. [Google Scholar] [CrossRef]
  248. Mamta, B.; Rajam, M.V. RNAi technology: A new platform for crop pest control. Physiol. Mol. Biol. Plants 2017, 23, 487–501. [Google Scholar] [CrossRef] [PubMed]
  249. Castellanos, N.L.; Smagghe, G.; Taning, C.N.T.; Oliveira, E.E.; Christiaens, O. Risk assessment of RNAi-based pesticides to non-target organisms: Evaluating the effects of sequence similarity in the parasitoid wasp Telenomus podisi. Sci. Total Environ. 2022, 832, 154746. [Google Scholar] [CrossRef]
  250. Taning, C.N.T.; Gui, S.; De Schutter, K.; Jahani, M.; Castellanos, N.L.; Christiaens, O.; Smagghe, G. A sequence complementarity-based approach for evaluating off-target transcript knockdown in Bombus terrestris, following ingestion of pest-specific dsRNA. J. Pest Sci. 2021, 94, 487–503. [Google Scholar] [CrossRef]
  251. Zhang, J.; Khan, S.A.; Heckel, D.G.; Bock, R. Next-generation insect-resistant plants: RNAi-mediated crop protection. Trends Biotechnol. 2017, 35, 871–882. [Google Scholar] [CrossRef]
  252. Chen, X.; Koo, J.; Gurusamy, D.; Mogilicherla, K.; Reddy Palli, S. Caenorhabditis elegans systemic RNA interference defective protein 1 enhances RNAi efficiency in a lepidopteran insect, the fall armyworm, in a tissue-specific manner. RNA Biol. 2021, 18, 1291–1299. [Google Scholar] [CrossRef] [PubMed]
  253. Santos, D.; Mingels, L.; Vogel, E.; Wang, L.; Christiaens, O.; Cappelle, K.; Wynant, N.; Gansemans, Y.; Van Nieuwerburgh, F.; Smagghe, G.; et al. Generation of virus- and dsRNA-derived siRNAs with species-dependent length in insects. Viruses 2019, 11, 738. [Google Scholar] [CrossRef]
  254. Haller, S.; Widmer, F.; Siegfried, B.D.; Zhuo, X.; Romeis, J. Responses of two ladybird beetle species (Coleoptera: Coccinellidae) to dietary RNAi. Pest Manag. Sci. 2019, 75, 2652–2662. [Google Scholar] [CrossRef]
  255. Niu, J.; Smagghe, G.; De Coninck, D.I.; Van Nieuwerburgh, F.; Deforce, D.; Meeus, I. In vivo study of Dicer-2-mediated immune response of the small interfering RNA pathway upon systemic infections of virulent and avirulent viruses in Bombus terrestris. Insect Biochem. Mol. Biol. 2016, 70, 127–137. [Google Scholar] [CrossRef]
  256. Joga, M.R.; Zotti, M.J.; Smagghe, G.; Christiaens, O. RNAi Efficiency, Systemic Properties, and Novel Delivery Methods for Pest Insect Control: What We Know So Far. Front. Physiol. 2016, 7, 553. [Google Scholar] [CrossRef]
  257. Romeis, J.; Widmer, F. Assessing the risks of topically applied dsRNA-based products to non-target arthropods. Front. Plant Sci. 2020, 11, 679. [Google Scholar] [CrossRef]
  258. Vesprini, F.; Maggi, A.I.; López Olaciregui, M.; Módena, N.A. Transportability of conclusions from confined field trials: A case study using the virus resistant transgenic bean developed in Brazil. Front. Bioeng. Biotechnol. 2020, 8. [Google Scholar] [CrossRef]
  259. Bachman, P.M.; Anderson, J.A.; Burns, A.; Chakravarthy, S.; Goodwin, L.; Privalle, L.; Song, S.; Storer, N. Data transportability for studies performed to support an environmental risk assessment for genetically modified (GM) crops. J. Regul. Sci. 2021, 9, 38–44. [Google Scholar]
  260. Tachikawa, M.; Matsuo, M. Divergence and convergence in international regulatory policies regarding genome-edited food: How to find a middle ground. Front. Plant Sci. 2023, 14, 1105426. [Google Scholar] [CrossRef]
  261. Zheng, Z.M.; Tang, S.; Tao, M. Development of resistance to RNAi in mammalian cells. Ann. N. Y. Acad. Sci. 2005, 1058, 105–118. [Google Scholar] [CrossRef]
  262. Dadami, E.; Dalakouras, A.; Zwiebel, M.; Krczal, G.; Wassenegger, M. An endogene-resembling transgene is resistant to DNA methylation and systemic silencing. RNA Biol. 2014, 11, 934–941. [Google Scholar] [CrossRef] [PubMed]
  263. Laurie, J.D.; Ali, S.; Linning, R.; Mannhaupt, G.; Wong, P.; Güldener, U.; Münsterkötter, M.; Moore, R.; Kahmann, R.; Bakkeren, G.; et al. Genome comparison of barley and maize smut fungi reveals targeted loss of RNA silencing components and species-specific presence of transposable elements. Plant Cell 2012, 24, 1733–1745. [Google Scholar] [CrossRef] [PubMed]
  264. Pollari, M.; De, S.; Wang, A.; Mäkinen, K. The potyviral silencing suppressor HCPro recruits and employs host ARGONAUTE1 in pro-viral functions. PLoS Pathog. 2020, 16, e1008965. [Google Scholar] [CrossRef]
  265. Mishra, G.P.; Aski, M.S.; Bosamia, T.; Chaurasia, S.; Mishra, D.C.; Bhati, J.; Kumar, A.; Javeria, S.; Tripathi, K.; Kohli, M.; et al. Insights into the host-pathogen interaction pathways through RNA-seq analysis of Lens culinaris Medik. in response to Rhizoctonia bataticola infection. Genes 2021, 13, 90. [Google Scholar] [CrossRef]
  266. Mishra, S.; Dee, J.; Moar, W.; Dufner-Beattie, J.; Baum, J.; Dias, N.P.; Alyokhin, A.; Buzza, A.; Rondon, S.I.; Clough, M.; et al. Selection for high levels of resistance to double-stranded RNA (dsRNA) in Colorado potato beetle (Leptinotarsa decemlineata Say) using non-transgenic foliar delivery. Sci. Rep. 2021, 11, 6523. [Google Scholar] [CrossRef] [PubMed]
  267. Kyre, B.R.; Dupuis, J.R.; Zúñiga, G.; Rieske, L.K. Variation in RNA interference sensitivity in the southern pine beetle, Dendroctonus frontalis (Coleoptera: Curculionidae). Biol. J. Linn. Soc. 2024, 142, 147–154. [Google Scholar] [CrossRef]
  268. Yoon, J.-S.; Sahoo, D.K.; Maiti, I.B.; Palli, S.R. Identification of target genes for RNAi-mediated control of the Twospotted Spider Mite. Sci. Rep. 2018, 8, 14687. [Google Scholar] [CrossRef]
  269. Mehlhorn, S.G.; Geibel, S.; Bucher, G.; Nauen, R. Profiling of RNAi sensitivity after foliar dsRNA exposure in different European populations of Colorado potato beetle reveals a robust response with minor variability. Pestic Biochem. Physiol. 2020, 166, 104569. [Google Scholar] [CrossRef]
  270. Neves, M.F.; Casagrande, B.P.; Cambaúva, V.; Teixeira, G.d.O.; Toledo, P.J.F. Agriculture 6.0: A New Proposal for the Future of Agribusiness. Revista de Gestão Social e Ambiental 2023, 17, e04004. [Google Scholar] [CrossRef]
  271. Rezaei, E.E.; Webber, H.; Asseng, S.; Boote, K.; Durand, J.L.; Ewert, F.; Martre, P.; MacCarthy, D.S. Climate change impacts on crop yields. Nat. Rev. Earth Environ. 2023, 4, 831–846. [Google Scholar] [CrossRef]
  272. Martha Júnior, G.B.; Lopes, M.A. Charting new sustainable agricultural innovation pathways in Brazil. Sci. Agric. 2023, 80. [Google Scholar] [CrossRef]
  273. Silva, F.T.d.; Baierle, I.C.; Correa, R.G.d.F.; Sellitto, M.A.; Peres, F.A.P.; Kipper, L.M. Open innovation in agribusiness: Barriers and challenges in the transition to Agriculture 4.0. Sustainability 2023, 15, 8562. [Google Scholar] [CrossRef]
  274. Munaweera, T.I.K.; Jayawardana, N.U.; Rajaratnam, R.; Dissanayake, N. Modern plant biotechnology as a strategy in addressing climate change and attaining food security. Agric. Food Secur. 2022, 11, 26. [Google Scholar] [CrossRef]
  275. Dalakouras, A.; Papadopoulou, K.K. Epigenetic modifications: An unexplored facet of exogenous RNA application in plants. Plants 2020, 9, 673. [Google Scholar] [CrossRef]
Figure 1. A brief timeline of RNAi in plants highlighting advancements from proofs of concept to commercial products. Abbreviations: RNAi: RNA interference; PRSV: Papaya ringspot virus; VIGS: virus-induced gene silencing; HIGS: host-induced gene silencing; BGMV: Bean golden mosaic virus; SIGS: spray-induced gene silencing; ESG: Environmental, Social, and Corporate Governance; CRISPR/Cas9: clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9).
Figure 1. A brief timeline of RNAi in plants highlighting advancements from proofs of concept to commercial products. Abbreviations: RNAi: RNA interference; PRSV: Papaya ringspot virus; VIGS: virus-induced gene silencing; HIGS: host-induced gene silencing; BGMV: Bean golden mosaic virus; SIGS: spray-induced gene silencing; ESG: Environmental, Social, and Corporate Governance; CRISPR/Cas9: clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein 9 (Cas9).
Agronomy 15 00859 g001
Figure 2. Cellular-level schematic depicting different RNAi-based strategies in agriculture. The gene silencing mechanism is represented separately for plants (green box), insects (blue box), and other pests and diseases (orange box). In the host-induced gene silencing (HIGS) strategy, a double-strand RNA (dsRNA) molecule is produced from the transcription of a transgene inserted in the plant genome. In the spray-induced gene silencing (SIGS) strategy, a dsRNA molecule or formulated RNAi is topically delivered onto the target organism (plant or pest). For both DNA and RNA viruses, as well as in virus-induced gene silencing (VIGS)-mediated delivery system, the engineered RNA viral is primarily targeted by the plant’s RNAi machinery, originating small-interfering RNA (siRNA) of 21–24 nucleotides in length. The dsRNA molecules originated from SIGS and HIGS strategies are cleaved by RNase III-type endoribonucleases known as Dicer or Dicer-like (DCL) enzymes to produce siRNAs, which are then processed in the host cell (black arrows) or cross-transferred to the cell of the plant’s pest or pathogen (dash lines). Most organisms can efficiently uptake dsRNA or siRNA from the environment (black dash lines); however, insect cells exhibit deficient uptake or siRNA and dsRNA is preferred (purplish purple lines). To cross-transfer dsRNA from plants to other organisms, a protected form of dsRNA is required to avoid processing by DLC proteins. In the cytoplasm, most siRNAs are typically incorporated into the RNA-induced silencing complex (RISC). The active part of RISC, argonaute enzymes act along siRNAs by scanning and cleaving the sequence-homologous messenger RNAs (mRNAs) or exogenous RNAs (viral or subviral RNAs). Therefore, mRNA degradation leads to reduced translation of the encoded protein in a process known as post-transcriptional gene silencing (PTGS). As a consequence, target proteins from both plants and pests are not produced in regular amounts, and this intrinsic molecular factor can be orchestrated to improve agronomic traits or pest control. Furthermore, most cell types are capable of amplifying the RNAi signal by the de novo production of secondary siRNAs. An enzyme denominated of RNA-dependent RNA polymerase (RdRP) uses the RNA template obtained from the dsRNA processing to execute this mechanism. Insects lack RdRP genes and RNAi signal amplification in these organisms is still poorly understood.
Figure 2. Cellular-level schematic depicting different RNAi-based strategies in agriculture. The gene silencing mechanism is represented separately for plants (green box), insects (blue box), and other pests and diseases (orange box). In the host-induced gene silencing (HIGS) strategy, a double-strand RNA (dsRNA) molecule is produced from the transcription of a transgene inserted in the plant genome. In the spray-induced gene silencing (SIGS) strategy, a dsRNA molecule or formulated RNAi is topically delivered onto the target organism (plant or pest). For both DNA and RNA viruses, as well as in virus-induced gene silencing (VIGS)-mediated delivery system, the engineered RNA viral is primarily targeted by the plant’s RNAi machinery, originating small-interfering RNA (siRNA) of 21–24 nucleotides in length. The dsRNA molecules originated from SIGS and HIGS strategies are cleaved by RNase III-type endoribonucleases known as Dicer or Dicer-like (DCL) enzymes to produce siRNAs, which are then processed in the host cell (black arrows) or cross-transferred to the cell of the plant’s pest or pathogen (dash lines). Most organisms can efficiently uptake dsRNA or siRNA from the environment (black dash lines); however, insect cells exhibit deficient uptake or siRNA and dsRNA is preferred (purplish purple lines). To cross-transfer dsRNA from plants to other organisms, a protected form of dsRNA is required to avoid processing by DLC proteins. In the cytoplasm, most siRNAs are typically incorporated into the RNA-induced silencing complex (RISC). The active part of RISC, argonaute enzymes act along siRNAs by scanning and cleaving the sequence-homologous messenger RNAs (mRNAs) or exogenous RNAs (viral or subviral RNAs). Therefore, mRNA degradation leads to reduced translation of the encoded protein in a process known as post-transcriptional gene silencing (PTGS). As a consequence, target proteins from both plants and pests are not produced in regular amounts, and this intrinsic molecular factor can be orchestrated to improve agronomic traits or pest control. Furthermore, most cell types are capable of amplifying the RNAi signal by the de novo production of secondary siRNAs. An enzyme denominated of RNA-dependent RNA polymerase (RdRP) uses the RNA template obtained from the dsRNA processing to execute this mechanism. Insects lack RdRP genes and RNAi signal amplification in these organisms is still poorly understood.
Agronomy 15 00859 g002
Figure 3. The patent scenario of RNAi-based technologies. Patentability scenarios may have defined the number of patents filed in South America. (A) Technology investment trend over 20 years worldwide measured by the number of patent families filed exclusively for RNAi. (B) Technology investment trend over 20 years worldwide, measured by the number of patent families filed for RNAi in agriculture. (C) Number of patent families granted by assignees for RNAi in agriculture. (D) Number of patent families filed for RNAi in agriculture, categorized by country of protection. PCT: Patent Cooperation Treaty. (E) Technology investment trend over 20 years in South America, measured by the numbers of patent families filed for RNAi in agriculture.
Figure 3. The patent scenario of RNAi-based technologies. Patentability scenarios may have defined the number of patents filed in South America. (A) Technology investment trend over 20 years worldwide measured by the number of patent families filed exclusively for RNAi. (B) Technology investment trend over 20 years worldwide, measured by the number of patent families filed for RNAi in agriculture. (C) Number of patent families granted by assignees for RNAi in agriculture. (D) Number of patent families filed for RNAi in agriculture, categorized by country of protection. PCT: Patent Cooperation Treaty. (E) Technology investment trend over 20 years in South America, measured by the numbers of patent families filed for RNAi in agriculture.
Agronomy 15 00859 g003
Figure 4. Comparison of RNAi-based technologies from lab-to-field. Main processes and experiments for (1) bioprospecting suitable target genes, (2) target validation, stabilization of dsRNA structure, RNAi strategy optimization (in planta or topical RNAi) and dsRNA delivery system, (3) proof of concept for evaluating potential product in a controlled environment, (4) broad evaluation of the commercial product under field conditions, biosafety, intellectual property, commercial regulation, and technology transfer to the productivity sector, and (5) commercial use by farmers of innovative RNAi-based technology for practical applications to control pests and improve specific agronomic traits. Abbreviations: VIGS: virus-induced gene silencing, HIGS: host-induced gene silencing, and SIGS: spray-induced gene silencing.
Figure 4. Comparison of RNAi-based technologies from lab-to-field. Main processes and experiments for (1) bioprospecting suitable target genes, (2) target validation, stabilization of dsRNA structure, RNAi strategy optimization (in planta or topical RNAi) and dsRNA delivery system, (3) proof of concept for evaluating potential product in a controlled environment, (4) broad evaluation of the commercial product under field conditions, biosafety, intellectual property, commercial regulation, and technology transfer to the productivity sector, and (5) commercial use by farmers of innovative RNAi-based technology for practical applications to control pests and improve specific agronomic traits. Abbreviations: VIGS: virus-induced gene silencing, HIGS: host-induced gene silencing, and SIGS: spray-induced gene silencing.
Agronomy 15 00859 g004
Table 1. Current crops with improved agronomic traits developed by host-induced gene silencing (HIGS) and successfully approved for commercial use in different countries. Information from the International Service for the Acquisition of Agri-biotech Applications (ISAAA, 2024).
Table 1. Current crops with improved agronomic traits developed by host-induced gene silencing (HIGS) and successfully approved for commercial use in different countries. Information from the International Service for the Acquisition of Agri-biotech Applications (ISAAA, 2024).
CropTrade NameEvent Code (Name)DeveloperTarget GeneModified Plant TraitRegulatory ApprovalsFirst Approval
Carica papayaRainbow, SunUpCUH-CP551-8 (55-1)Cornell University and the University of HawaiiCPImproved plant resistance to Papaya ringspot virusCanada, Japan, and USA1996
not availableCUH-CP631-7 (63-1)Cornell University and the University of HawaiiCPUSA1996
Huanong No. 1Huanong No. 1South China Agricultural UniversityRepChina2006
not availableUFL-X17CP-6 (X17-2)University of FloridaCPUSA2008
Solanum lycopersicumFLAVR SAVR™CGN-89564-2 (FLAVR SAVR™)MonsantoPGsDelayed fruit softeningMexico, Canada, and USA1992
not availableSYN-ØØØØB-6 (B)Zeneca Plant Science and PetoseedPGsDelayed fruit softeningMexico and USA1994
not availableSYN-ØØØDA-9 (Da)Zeneca Plant Science and PetoseedPGsDelayed fruit softeningMexico and USA1995
not availableSYN-ØØØØF-1 (F)Zeneca Plant Science and PetoseedPGsDelayed fruit softeningMexico, Canada, and USA1995
not available1345-4DNA Plant Technology CorporationACCReduced synthesis of endogenous ethyleneMexico, Canada, and USA1995
not availableHuafan No. 1Huazhong Agricultural UniversityACODelayed ripening by suppressing the production of ethylene by silencing the ACO gene China1997
not availablePK-TM8805R (8805R)Beijing UniversityCPImproved plant resistance to Cucumber mosaic virusChina1999
Cucurbita peponot availableSEM-ØZW2Ø-7 (ZW20)Seminis Vegetable Seeds (Canada) and Monsanto (Asgrow)CPImproved plant resistance to Zucchini yellow mosaic virus, and Watermelon mosaic virus 2USA1994
not availableSEM-ØCZW3-2 (CZW3)Seminis Vegetable Seeds (Canada) and Monsanto (Asgrow)CPImproved plant resistance to Cucumber mosaic virus, Zucchini yellow mosaic virus, and Watermelon mosaic virus 2Canada and USA1996
Dianthus caryophyllusnot availableFLO-ØØØ66-8 (66)Florigene Pty Ltd.ACCReduced synthesis of ethyleneAustralia and Norway1995
Capsicum annuumnot availablePK-SP01Beijing UniversityCPImproved plant resistance to Cucumber mosaic virusChina1998
Solanum tuberosumHi-Lite NewLeaf™ Y potatoHLMT15-15
HLMT15-3
HLMT15-46
MonsantoCPImproved plant resistance to Potato virus YUSA1998
Shepody NewLeaf™ Y potatoNMK-89935-9 (SEMT15-02)
SEMT15-07
NMK-8993Ø-4 (SEMT15-15)
MonsantoCPImproved plant resistance to Potato virus YAustralia, Canada, Japan, Mexico, New Zealand, Philippines, South Korea, and USA1998
New Leaf™ Y Russet Burbank potatoNMK-89653-6 (RBMT15-101)
NMK-89684-1 (RBMT21-129)
RBMT21-152
NMK-89896-6 (RBMT22-082)
RBMT22-186
RBMT22-238
RBMT22-262
MonsantoORF1 and ORF2Improved plant resistance to Potato leaf roll virusAustralia, Canada, Japan, Mexico, New Zealand, Philippines, South Korea, and USA1998
Amflora™BPS-25271-9 (EH92-527-1)BASFgbssReduced levels of amylose and increased levels of amylopectin in starch granulesEuropean Union2010
Starch PotatoBPS-A1Ø2Ø-5 (AM04-1020)BASFgbssUSA2014
Simplot InnateSPS-ØØØZ6-5 (Gen2-Z6)J.R. Simplot Co.asn1, ppo5, PhL, and VlnvImproved black spot bruise tolerance, reduced levels of cold-induced sweetening, and reduced levels of acrylamideCanada and USA2020
not availableSPS-ØØØW8-4 (W8)J.R. Simplot Co.asn1, ppo5, R1, PhL, and VlnvAustralia, Canada, New Zealand, and USA2015
Innate® AcclimateSPS-ØØX17-5 (X17)J.R. Simplot Co.asn1, ppo5, R1, PhL, and VlnvAustralia, Canada, New Zealand, Philippines, and USA2016
Innate® HibernateSPS-ØØØY9-7 (Y9)J.R. Simplot Co.asn1, ppo5, R1, PhL, and VlnvAustralia, Canada, New Zealand, Philippines, and USA2016
Innate® CultivateSPS-ØØE12-8 (E12)J.R. Simplot Co.asn1, ppo5, PhL, and R1Australia, Canada, Japan, Malaysia, Mexico, New Zealand, Philippines, Singapore, and USA2014
not availableSPS-ØØE24-2 (E24)J.R. Simplot Co.asn1, ppo5, PhL, and R1USA2014
not availableSPS-ØØE56-7 (E56)J.R. Simplot Co.asn1, ppo5, PhL, and R1Australia and New Zealand2017
Innate® Generate
Innate® Accelerate
SPS-ØØF10-7 (F10)
SPS-ØØØJ3-4 (J3)
J.R. Simplot Co.asn1, ppo5, PhL, and R1Australia, Canada, Mexico, New Zealand, and USA2014
not availableSPS-ØØF37-7 (F37)
SPS-ØØG11-9 (G11)
SPS-ØØH37-9 (H37)
SPS-ØØH50-4 (H50)
SPS-ØØJ78-7 (J78)
J.R. Simplot Co.asn1, ppo5, PhL, and R1USA2014
not availableSPS-ØØJ55-2 (J55)J.R. Simplot Co.asn1, ppo5, PhL, and R1Canada and USA2014
Innate® InvigorateSPS–ØØV11–6 (V11)J.R. Simplot Co.asn1, ppo5, PhL, and R1Australia, New Zealand, and USA2016
not availableTIC-AR233-5 Technoplant ArgentinaCPImproved plant resistance to Potato virus YArgentina2018
Nicotiana tabacumnot availableVector 21-41Vector Tobacco Inc.NtQPT1Reduced production of nicotinic acidUSA2002
Medicago sativaHarvXtra™MON-ØØ179-5 (KK179)Monsanto and Forage Genetics InternationalCCOMTReduces content of guaiacyl (G) ligninAustralia, Canada, Japan, Mexico, New Zealand, Philippines, Singapore, South Korea, and USA2013
Zea maysnot availableMON-87411-9 (MON87411)MonsantoDvSnf7Improved plant resistance against Diabrotica virgifera virgiferaArgentina, Australia, Brazil, Canada, Colombia, Japan, Mexico, New Zealand, Philippines, South Korea, Taiwan, and USA2014
Malus domesticaArctic™ “Golden Delicious” AppleOKA-NBØØ1-8 (GD743)Okanagan Specialty Fruits IncorporatedPPOsApples with a non-browning phenotypeCanada and USA2015
Arctic™OKA-NBØØ2-9 (GS784)Okanagan Specialty Fruits IncorporatedPPOsApples with a non-browning phenotypeCanada and USA2015
Arctic™ Fuji AppleOKA-NBØØ3-1 (NF872)Okanagan Specialty Fruits IncorporatedPPOsApples with a non-browning phenotypeCanada and USA2018
Ananas comosusRoséFDP-ØØ114-5 (EF2-114)Del Monte Fresh Produceb-Lyc and e-LycIncreased lycopene accumulationCanada and USA2016
Gossypium hirsutumnot availableTAM-66274-5 (TAM66274)Texas A&M AgriLife Research UniversitydCSReduced gossypol biosynthesisUSA2018
Carthamus tinctoriusnot availableGOR-73226-6 (Event 26)
GOR-7324Ø-2 (Event 40)
Go Resources Pty Ltd.fatB and fad2.2Modified oil/fatty acidAustralia2018
Glycine maxnot availableDD-Ø26ØØ5-3 (260-05)DuPont (Pioneer Hi-Bred International Inc.)gm-fad2-1Blocked conversion of oleic acid to linoleic acidAustralia, Canada, Japan, New Zealand, and USA1997
Treus™, Plenish™DP-3Ø5423-1 (DP305423)DuPont (Pioneer Hi-Bred International Inc.)gm-fad2-1Blocked conversion of oleic acid to linoleic acidAustralia, Brazil, Canada, China, Colombia, European Union, Indonesia, Iran, Japan, Malaysia, Mexico, New Zealand, Philippines, Singapore, South Africa, South Korea, Taiwan, Turkey, and USA2008
Vistive Gold™MON-877Ø5-6 (MON87705)Monsantofad2.1A and fatb1-ABlocked conversion of oleic acid to linoleic acid and decreased transport of saturated fatty acids out of the plastidAustralia, Canada, China, Colombia, European Union, Indonesia, Japan, Mexico, New Zealand, Nigeria, Philippines, Singapore, South Korea, Taiwan, Turkey, Vietnam, and USA2011
Phaseolus vulgarisBRS FC401 RMDEMB-PVØ51-1 (EMBRAPA 5.1)EmbrapaAC1Improved plant resistance to Bean golden mosaic virusBrazil2011
Prunus domesticanot availableARS-PLMC5-6 (C-5)USDA—Agricultural Research ServiceCPImproved plant resistance to Plum pox virusUSA2007
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

Basso, M.F.; Vásquez, D.D.N.; Campos-Pinto, E.R.; Pinheiro, D.H.; Cruz, B.; Maktura, G.C.; Guidelli, G.V.; Marques-Souza, H.; Grossi-de-Sa, M.F. Progress and Opportunities of In Planta and Topical RNAi for the Biotechnological Control of Agricultural Pests. Agronomy 2025, 15, 859. https://doi.org/10.3390/agronomy15040859

AMA Style

Basso MF, Vásquez DDN, Campos-Pinto ER, Pinheiro DH, Cruz B, Maktura GC, Guidelli GV, Marques-Souza H, Grossi-de-Sa MF. Progress and Opportunities of In Planta and Topical RNAi for the Biotechnological Control of Agricultural Pests. Agronomy. 2025; 15(4):859. https://doi.org/10.3390/agronomy15040859

Chicago/Turabian Style

Basso, Marcos Fernando, Daniel David Noriega Vásquez, Eduardo Romano Campos-Pinto, Daniele Heloísa Pinheiro, Bread Cruz, Grazielle Celeste Maktura, Giovanna Vieira Guidelli, Henrique Marques-Souza, and Maria Fatima Grossi-de-Sa. 2025. "Progress and Opportunities of In Planta and Topical RNAi for the Biotechnological Control of Agricultural Pests" Agronomy 15, no. 4: 859. https://doi.org/10.3390/agronomy15040859

APA Style

Basso, M. F., Vásquez, D. D. N., Campos-Pinto, E. R., Pinheiro, D. H., Cruz, B., Maktura, G. C., Guidelli, G. V., Marques-Souza, H., & Grossi-de-Sa, M. F. (2025). Progress and Opportunities of In Planta and Topical RNAi for the Biotechnological Control of Agricultural Pests. Agronomy, 15(4), 859. https://doi.org/10.3390/agronomy15040859

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