Next Article in Journal
Tracking Existing Factors Directly Affecting the Reproduction of Bumblebees: Current Knowledge
Previous Article in Journal
Virulence Adaptation by Rice Planthoppers and Leafhoppers to Resistance Genes and Loci: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advancements and Future Prospects of CRISPR-Cas-Based Population Replacement Strategies in Insect Pest Management

College of Life Science and Technology, Gansu Agricultural University, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Insects 2024, 15(9), 653; https://doi.org/10.3390/insects15090653
Submission received: 16 July 2024 / Revised: 22 August 2024 / Accepted: 28 August 2024 / Published: 30 August 2024
(This article belongs to the Section Insect Pest and Vector Management)

Abstract

:

Simple Summary

Many insects are categorized as agricultural pests due to their ability to transmit diseases and damage crops, which results in significant economic losses. Scientists have proposed two main pest control strategies: population suppression, aimed at reducing the size or distribution of pest populations, and population replacement, which involves introducing genetically modified populations to replace wild pests after an initial release. Typically, population replacement strategies use gene drive systems to spread beneficial traits throughout the target population. Current promising gene drive systems include homing endonuclease genes (HEGs), Wolbachia, maternal-effect dominant embryonic arrest (Medea), and newly adapted CRISPR/Cas genome editing systems. This review provides an overview of the recent advancements in population replacement, including insights into the development, testing, and safe implementation of CRISPR-Cas-based gene drive techniques from laboratory settings to field applications. It also discusses recent developments, identifies research gaps, and offers a comprehensive analysis of genetic control strategies for insect pests.

Abstract

Population replacement refers to the process by which a wild-type population of insect pests is replaced by a population possessing modified traits or abilities. Effective population replacement necessitates a gene drive system capable of spreading desired genes within natural populations, operating under principles akin to super-Mendelian inheritance. Consequently, releasing a small number of genetically edited insects could potentially achieve population control objectives. Currently, several gene drive approaches are under exploration, including the newly adapted CRISPR-Cas genome editing system. Multiple studies are investigating methods to engineer pests that are incapable of causing crop damage or transmitting vector-borne diseases, with several notable successful examples documented. This review summarizes the recent advancements of the CRISPR-Cas system in the realm of population replacement and provides insights into research methodologies, testing protocols, and implementation strategies for gene drive techniques. The review also discusses emerging trends and prospects for establishing genetic tools in pest management.

1. Introduction

Insects comprise the most numerous animal group, with over one million described species, representing more than half of all known living creatures on Earth [1]. The estimated total number of extant insect species equates to several million. Many insects are considered agricultural pests due to their ability to spread disease and damage crops, leading to substantial global economic losses annually [2,3]. Over recent decades, scientists have proposed various pest control methods using genetic engineering, primarily categorized into two strategies: population suppression and population replacement. Population suppression aims to decrease insect numbers or their geographical spread [4,5,6,7]. In contrast, population replacement strategies involve replacing wild-type pest populations with genetically modified strains or species, resulting in the genetically modified organisms occupying the ecological niche [8,9,10,11,12].
Suppression strategies include chemical insecticides, the sterile insect technique (SIT), the incompatible insect technique (IIT), the release of insects carrying a dominant lethal gene (RIDL), and biological control. These population suppression strategies have been extensively employed against various agricultural pest insects for decades. For instance, chemical insecticides have been widely utilized globally to control a variety of insect pests, offering advantages such as reduced labor and time requirements, low labor intensity, and high operational efficiency [13]. The sterile insect technique (SIT) is an environmentally friendly, non-resistant, non-polluting, and species-specific method used for the area-wide control of agricultural pests. It has demonstrated successful application in pests such as medflies, screwworms, fruit flies, moths, pink bollworms, codling moths, and tsetse flies [14]. The release of insects carrying a dominant lethal gene (RIDL), an advanced form of SIT, is particularly effective due to its fitness and cost advantages. RIDL has been successfully developed and tested in laboratory and open-field environments against pests like the diamondback moth, mosquito, and medfly [15]. The incompatible insect technique (IIT), akin to SIT and RIDL, relies on the sustained release of Wolbachia-induced cytoplasmic incompatibility (CI) to sterilize pests. Developed over 40 years ago, IIT has proven effective in controlling the vector-borne diseases of mosquitoes and medflies [16,17].
The population suppression methods using SIT, IIT, and RIDL techniques generally depend on long-term rearing and large-scale releases, which can be expensive to implement. These methods often encounter challenges such as inconsistent infrastructure availability, bottlenecks in efficient sex-sorting, quality control issues, extendibility problems, and suboptimal field performance. Furthermore, the repeated use of chemical insecticides can lead to the development of highly resistant insects with reduced susceptibility, contributing to the environmental pollution which is harmful to ecosystems and human health [18,19]. These factors collectively limit the effectiveness of these approaches [20]. Furthermore, most non-selective methods struggle to effectively target and control omnivorous pest insects with diverse host ranges, larvae that feed inside host fruits, and pests exhibiting opportunistic feeding and resting behaviors [19].
Therefore, there is a need to develop new pest control methods that can address these shortcomings. One promising strategy which is gaining attention is population replacement, where genetically modified populations replace wild pest populations following an initial ‘seeding’ release [21]. Population replacement strategies generally involve the deployment of gene drive systems, which combine a non-pestilent trait with a selfish genetic element that spreads rapidly within the target population once a threshold frequency is reached [22]. Compared to most conventional population suppression methods that rely on inundative releases and Mendelian inheritance patterns, gene drive strategies typically require fewer releases of individuals (either male or female) into the wild population. This makes them potentially more efficient and sustainable for long-term pest control solutions.
In this review, we consolidate the literature from the last 10 years regarding population replacement strategies for controlling agricultural insect pests. Our aim is to facilitate an understanding of the design, testing, and safe implementation of this approach from the laboratory setting to the field application. Additionally, we highlight advancements and identify areas requiring further investigation within population replacement research, thereby offering an overview of emerging insect pest management methodologies. Looking ahead, interdisciplinary collaboration including mathematical modeling, population surveys, legislative guidelines, ecology, and genetics will be essential for advancing this field of research.

2. Population Replacement Using Gene Drive Systems

2.1. Molecular Genetic Manipulation of Insects

To achieve effective population replacement for insect pest control, a gene drive system capable of biasing inheritance in its favor is essential. This requires the introduction of a selfish genetic element that autonomously spreads with super-Mendelian inheritance (greater than a 50% likelihood) within the target population, ultimately leading to the replacement of the local population with a specific genotype. This introduced construct could potentially disrupt or introduce a desired gene that lacks disease transmission capability or exhibits reduced virulence and damaging characteristics. Several gene drive approaches are currently under investigation, including HEGs (homing endonuclease genes) [23], Medea (maternal effect dominant embryonic arrest) [24,25], and CRISPR-Cas (clustered regularly interspaced short palindromic repeats-CRISPR-associated proteins) systems (Figure 1). While the concept of using gene drives to manage insect pests is not new, achieving this goal remains challenging, hampered by scientific and technical obstacles. Recent advancements in the CRISPR-Cas system, a precise site-specific gene editing technology, offer promising solutions to overcome some of these challenges. Therefore, we primarily focus on the application of CRISPR-Cas technology in gene drive systems.
The CRISPR-Cas system is a genetic engineering tool derived from an adaptive immune defense mechanism found in many archaea and bacteria. It integrates partial sequences of foreign genomes between the CRISPR repeat sequences, which are organized as arrays within specific genomic loci. These sequences are transcribed, producing RNAs that guide Cas proteins to the target and cleave the genomes of phages or plasmids. Among the various Cas proteins, several form a cascade complex which is crucial for the processes at the CRISPR loci and for DNA cleavage. There are three primary types (I, II, and III) of CRISPR-Cas systems, each characterized by distinct sets of Cas genes [26]. Type II CRISPR systems utilize the RNA-guided endonuclease SpCas9 (formerly Csn1) from Streptococcus pyogenes to cleave DNA [27,28]. In this two-component system, protospacer sequences derived from invading DNA are processed into CRISPR RNAs (crRNAs), which then hybridize with a transactivating CRISPR RNA (tracrRNA) to form a complex, guiding Cas9 to cleave the complementary target DNA sequence. Researchers have simplified this system by engineering the tracrRNA duplex into a programmable chimeric single-guide RNA (sgRNA) [29,30]. By altering the first 20 nucleotides of the sgRNA sequence, different DNA sequences can be targeted, enabling multiplexed target recognition. Additionally, the versatility of the CRISPR-Cas9 system, achieved through component modifications, facilitates gene repression, activation, epigenome editing, and RNA editing capabilities [31].
The CRISPR-Cas system offers an ideal tool for constructing gene drives. This system functions by transferring drive alleles from one chromosome to another at a specified locus that matches the sgRNA-guided genome sequence. In germ cells, this method converts heterozygotes to homozygotes, facilitating super-Mendelian inheritance. Engineered for activity in the germline or embryo of diploid organisms, gene drives ensure the production of CRISPR-Cas-bearing gametes, with heterozygous parents primarily transmitting the gene drive. This mechanism exceeds the typical 50% inheritance rate expected for a single copy of the drive allele. In principle, CRISPR-Cas-mediated gene drives could be utilized to propagate specific genes throughout populations of engineered organisms by releasing a small number of gene drive heterozygotes, potentially replacing wild populations within a few generations [32].

2.2. Performance of CRISPR-Cas-Mediated Gene Drive Populations

The CRISPR-Cas-mediated gene drive in insects was first reported in Drosophila melanogaster in March 2015 [33]. Subsequent experiments successfully applied this system to Anopheles gambiae, Anopheles stephensi, Aedes aegypti, Drosophila suzukii, Ceratitis capitata, and Plutella xylostella [34,35,36,37,38,39,40,41].
Cas9 was utilized to integrate trait or cargo genes, with Gantz pioneering the mutagenic chain reaction (MCR) in D. melanogaster. The MCR achieved transmission efficiencies averaging 97%, significantly exceeding Mendelian inheritance expectations for a recessive X-linked yellow locus [33]. Oberhofer et al. developed the ClvR (Cleave and Rescue) system in D. melanogaster, a toxin-antidote gene drive on an autosomal chromosome targeting an essential X-linked gene [37]. It imposes fitness costs on wild-type offspring, favoring ClvR-bearing chromosomes through gene conversion and maternally transmitted Cas9 activity. ClvRtko effectively deactivated over 99% of the targeted genes with minimal functional resistance mutations. Another ClvR-based strategy developed by Oberhofer et al. involves placing two distinct ClvR elements that target different essential genes at the same genomic locus but on different homologous chromosomes. The second-generation element incorporates a rescue mechanism for the existing first-generation element within the population. Upon introduction into a population already carrying the first-generation element, the second-generation element spreads to fixation, displacing the latter [42]. ClvR has shown effectiveness across multiple highly conserved genes, suggesting its adaptability and potential for implementation in target species. Furthermore, the ClvR design enables the use of multiple gRNAs as toxins, thereby reducing the likelihood of developing resistant alleles that restore target gene function.
Champer et al. also created a “same-site” toxin-antidote drive in D. melanogaster, combining both elements within a recessive lethal gene to potentially enhance efficacy using the native promoter [43]. The study demonstrates that the TARE (CRISPR Toxin-Antidote Recessive Embryo) system effectively biases inheritance of the drive allele in a large Drosophila population cage. Within six generations of a modest release, the drive allele spread to all individuals without any resistance allele formation observed. A split gene drive in D. melanogaster targeting the haplolethal gene RpL35A used two guide RNAs and a gRNA-resistant rescue allele [44]. By generation six, the drive allele reached fixation from an initial frequency of 32% when mixed with Cas9 homozygous eggs. Efficient homing and fixation were achieved with Cas9 expression from a separate locus. Offspring from crosses between heterozygous parents and the wild-type demonstrated high inheritance rates, but progeny from gene drive heterozygous females and wild-type males had reduced survival, likely due to maternal Cas9 effects. All these gene drive systems were designed to target a vital gene for inactivation, which was subsequently rescued by inserting a recoded transgene at a different locus in the genome.
Incorporating functional recoded versions of genes into drives in modification systems has been demonstrated to significantly improve their performance [45]. Nash et al. investigated the minimal genetic elements necessary for autonomous gene drives in D. melanogaster by integrating a Cas9 open reading frame with an intronic gRNA into the rcd-1r locus, demonstrating efficient propagation in both male and female germlines [46]. The study highlighted the potential applications in modular drive and effector functions, alongside considerations of the biosafety implications for minimal and potentially recoded gene drives. Sanz Juste et al. first introduced a Cas12a-driven gene drive system regulated by temperature [47]. Higher temperatures increased the super-Mendelian inheritance rates for both alleles (e1-GD and e4-GD). At 18 °C, e4-GD showed no gene drive activity, suggesting it can be temperature-controlled. These findings pave the way for advancing Cas12a-based gene drive systems in insects, aiming to improve population control strategies across different pests.
By late 2015, Gantz et al. adapted the CRISPR-Cas-based gene drive to A. stephensi, achieving transmission rates of over 99.5% with an engineered synthetic construct [34]. Hammond et al. applied a CRISPRh gene drive in An. gambiae, disrupting three female fertility genes with high transmission rates [35]. The study proposed a new strategy to suppress malaria spread, despite reduced fertility in the offspring due to Cas9-induced somatic mutations. A 25-generation cage study showed an initial rise and subsequent decline in gene drive frequency due to resistant target site mutations. By the 25th generation, transgenic heterozygotes fell below 20%, revealing challenges in sustaining the gene drive efficiency over time. In An. gambiae, a CRISPR gene drive disrupted the doublesex (dsx) gene, which is crucial for sex determination. Homozygous females with the disrupted allele were sterile with intersex traits, while males developed normally. The gene drive achieved >95% transmission efficiency in the offspring of heterozygous parents and reached 100% frequency within 12 generations in cage studies starting at 12.5%. Minimal indels (1.16%) at the gRNA target site suggest low resistance potential [36]. An efficient gene drive strain, AgNosCd-1, has been successfully demonstrated in An. gambiae. Integrated into the Agcd gene, it incorporates fluorescent and morphological markers to track its spread. Initial outcrossing with wild types resulted in >95% drive propagation in the offspring, maintaining an average efficiency of 96.7% over four generations. The gene drive imposes a low fitness penalty in both heterozygous and homozygous states, highlighting its potential as a platform for introducing desired cargo genes into populations [48]. The study documents resistance developing over generations in response to a synthetic gene drive, emphasizing the need to address resistance in gene drive design. The AgNosCd-1 gene drive system is expected to be among those prioritized for evaluation in the upcoming phase of open field trials.
Adolfi developed a recoded gene drive rescue system for modifying Anopheles stephensi populations by inserting a gene drive with recoded cDNA sequences into a critical locus, ensuring viability and fertility in a homozygous state [49]. This approach eliminates non-functional NHEJ alleles through dominant lethal or sterile effects, as demonstrated at the kh locus. Cage trials showed that a single release of gene drive males achieves an effective population modification, with ≥95% of mosquitoes harboring the drive within 5–11 generations across different initial release ratios. Hammond et al. tested new constructs targeting a female fertility locus prone to resistance, aiming to quantify reduced resistance mutations by enhancing female fertility and reducing embryonic end-joining [50]. They investigated the spread of the zpg-CRISPRh gene drive in mosquito populations, observing rapid propagation to over 97% frequency in four trials spanning 4–10 generations. The gene drive sustained frequencies above 95% for at least three generations before encountering resistance. The initial release frequency (10% or 50%) did not significantly influence its spread, demonstrating effective invasion dynamics compared to the first-generation gene drive (vas2-CRISPRh).
In other insects, Xu et al. developed a CRISPR-Cas-based split drive system in Plutella xylostella, but it failed to produce homing-based progeny [41]. Asad et al. created a CRISPR-Cas gene drive construct for the same pest, incorporating Cas9, an EGFP marker gene, and a gRNA targeting Pxyellow. Their work achieved HDR efficiencies of 6.67–12.59% and generated resistant alleles through NHEJ at rates of 80.93–86.77%. Male-derived progeny showed higher gene drive efficiency than females, indicating potential trait inheritance [51]. In contrast, Meccariello characterized the first gene drives in Ceratitis capitata and developed a gene drive targeting the transformer gene using the vasa promoter. This gene drive system demonstrated an 83.1% transmission rate, indicating that the medfly is highly suitable for homing-based gene drive strategies [40].
Table 1 summarizes homing efficiencies across various generations of CRISPR-Cas-mediated gene drive systems. These studies highlight considerable variability in gene drive efficiencies, ranging from approximately 37% to 99% in D. melanogaster and from 0 to 99% in mosquitoes. This variability highlights the influence of factors such as the timing and level of Cas expression, organism-specific traits, various gene drive modular elements and their positions, as well as the specific genomic targets.

3. Ideal Traits for Gene Drives in Pest Control

CRISPR-Cas-mediated gene drive systems offer substantial potential for pest control by replacing harmful wild-type alleles with gene drive alleles. This approach allows the engineered individuals to replace the target populations rapidly, even with a low initial frequency. Successful implementation of this strategy as a population replacement tool for pest management requires the selection of key phenotype-related genes for targeting or insertion [62,63]. Potential applications include reducing fertility, limiting development, altering food consumption, impacting mating competitiveness, skewing sex ratios, curtailing migration, decreasing longevity, reducing stress resistance, modifying behavior, affecting immune function, and lowering disease transmission—all aimed at diminishing fitness traits or species’ pest potential.
The high efficiency of the CRISPR-Cas-mediated gene drive ensures that genetically modified organisms can spread readily, even when their insertions have fitness costs [50]. Alternatively, CRISPR-Cas-based population replacement strategies can lead to population suppression by imposing significant fitness costs (e.g., sterility) or gender biases (e.g., driving Y-chromosomes, X-chromosome shredders) [64,65,66]. Moreover, these systems can propagate recessive mutations causing lethality, infertility, or single-sex offspring [67]. Theoretically, releasing a small number of insects carrying such gene drive alleles can lead to the accumulation of heterozygous individuals producing predominantly gene drive gametes. Over multiple generations, mating among these heterozygotes could potentially suppress or collapse the population, contingent upon driving efficiency and fitness costs. Laboratory studies have demonstrated this strategy using HEGs and the CRISPR-Cas system [36,68].

3.1. Blocking Pathogen Transmission

Vector insects frequently inflict severe damage on the agricultural sector and pose health risks by transmitting pathogens such as viruses, parasites, and bacteria to humans or other arthropods. Typically, these vectors acquire infectious agents from infected hosts and subsequently transmit them to new hosts through multiple feeding cycles interspersed with egg-laying phases. A functional CRISPR-Cas-mediated gene drive represents a highly efficient strategy for controlling insect vectors. This approach involves replacing disease-carrying native strains with disease-resistant variants [69]. The primary objective of such gene drives is to reduce vector competence by integrating anti-pathogen genes into the driving alleles. These anti-pathogen genes may encompass exogenous peptides targeting pathogens, single-chain antibodies (scFvs), and regulators associated with immune functions. For instance, in mosquitoes, the elevated expression of m2A10 scFvs targeting the P. falciparum circumsporozoite protein inhibits sporozoites from invading salivary glands [70]. Incorporating either m4B7 or m1C3 scFv transgenes, or employing versatile transgenic multistage effector genes as demonstrated in one recent study, leads to a nearly complete inhibition of parasite transmission [71].

3.2. Manipulation of Sex Ratios

In many pest insects, population size primarily hinges on the reproductive capacity of the females. Female insects, acting as vectors, transmit numerous insect-borne pathogens such as plasmodium, tospovirus, and Puccinia. Segregation-distorting gene drives are selfish genetic elements that manipulate the process of gametogenesis. Therefore, the implementation of gene drive systems has the potential to skew population sex ratios away from the typical Fisherian ratio of 1:1 male to female, favoring a male-dominated population by promoting ‘maleness’ or altering reproductive dynamics [72]. Persistent male bias over successive generations can precipitate rapid population decline, with complete eradication possible when populations become exclusively male [73]. For instance, the gene drive construct dsxFCRISPRh, targeting the doublesex gene to induce sterility in homozygous females, caused the complete collapse of a confined population of 600 An. gambiae mosquitoes within 12 generations, with dsxFCRISPRh heterozygotes initiated at a 12.5% allelic frequency [36]. Similarly, the SDGDdsx gene drive construct, employing zpg-Cas9, U6-gRNA, and beta2244-I-PpoI to target the doublesex gene and ribosomal DNA repeats of the X chromosome, demonstrated a pronounced male-biased sex ratio distortion and achieved population elimination within 13 generations starting from a 10% release of SDGDdsx-heterozygous males [61]. This approach of genetic sex-sorting offers more effective population suppression compared to methods like the SIT. Identification of sex determination genes like doublesex and transformer presents opportunities to manipulate population sex ratios, potentially offering a powerful strategy for controlling invasive pests, particularly vector species, through CRISPR-Cas-mediated gene drives.

3.3. Manipulation of Feeding Behaviors

Insects exhibit behavioral responses through sensory signals integrated within the central nervous system. These signals arise from physical and chemical stimuli detected by exteroceptors (sensing external stimuli), enteroceptors (sensing internal stimuli), and proprioceptors (sensing body position) [74,75,76]. Receptor cells convert these sensory inputs into electrical signals, which are distributed across various systems, including vision, hearing, smell, taste, and touch [68]. Olfactory cues, primarily processed by olfactory sensory neurons (OSNs), play a crucial role in insect behavior, guiding responses to pheromones, food, predators, shelter, and conspecifics [77,78]. Molecular studies on insects such as flies, moths, bees, and mosquitoes elucidate how olfactory systems regulate behaviors related to age, feeding, circadian rhythms, and mating [79,80,81]. Similarly, the gustatory system enables insects to detect non-volatile phytochemicals through taste sensilla strategically located on the external and internal tissues, aiding in decisions related to feeding, mate selection, and oviposition [82,83,84,85]. Gustatory receptors (GRs), analogous to olfactory receptors (ORs), are involved in chemosensation, encompassing the responses to pheromones and somatosensory stimuli [86,87,88].
CRISPR-Cas-mediated gene drive systems offer a promising approach to manipulating pest behaviors—such as feeding, food location, and mating—by targeting the sensory systems. This technique holds potential for sustainable insect pest management and disease control while concurrently safeguarding crops. For example, manipulating food-finding behavior, which is critical for the plant-feeding pests attracted to pathogen-infected plants, may reduce disease spread [89,90,91,92]. Plant viruses have the ability to influence the behavior of insect vectors both directly and indirectly by modifying plant biochemistry. Frequently, these behavioral changes result in a higher rate of interactions between the host plants and the vectors, thereby facilitating greater virus spread within the ecosystem. Additionally, gene drive systems targeting vector sensory responses to pathogen-induced chemicals could provide a sustainable strategy for insect-vector management (Figure 2). Enhancing natural enemies of insect pests through genetic strategies also presents economic benefits due to the precision of field applications.

3.4. Manipulation of Migration

Flight-mediated migration is a crucial adaptive behavior among insect pests, enabling them to exploit new environments in response to environmental variations [93]. Most insects exhibit migratory polymorphism, controlled by both endogenous genetic circannual rhythms and external factors [94,95]. The wing dimorphism of insect vectors is a crucial factor for the long-distance spread of pathogens and the development of widespread epidemics. Extreme cases demonstrate distinct morphological adaptations: migratory individuals develop long-winged forms, while flightless individuals evolve short-winged or wingless morphs with non-functional flight muscles, limiting their dispersal capabilities [96,97]. Research on wing dimorphism mechanisms in holometabolous insects (e.g., Diptera, Coleoptera, Lepidoptera) has identified a single genetic locus controlling this trait [98,99], whereas in hemimetabolous insects (e.g., Orthoptera, Hemiptera), the mechanism involves multiple gene interactions [100,101]. For instance, the rice stripe virus (RSV), transmitted by planthoppers, induces the development of long-winged males. The proportion of long-winged males correlates with RSV infection, independent of viral titers. The planthopper-specific gene Encounter, highly expressed in the testes, is crucial for this morph. Encounter is upregulated during early larval stages and influences male wing dimorphism via the insulin/insulin-like growth factor pathway, downstream of Akt, which is activated by RSV infection. The study demonstrates that a plant virus directly regulates wing dimorphism in its insect vectors, suggesting a method to disrupt viral spread [102].
Insects migrating longer distances to their overwintering sites generally exhibit longer and larger wings, suggesting that wing length and area are subject to positive selective pressure during migration [103]. Therefore, migration manipulation is not restricted to winged insects with dimorphism; insects without wing dimorphism might also be targeted. For example, migration behavior can be manipulated by targeting the genes related to long-distance migration (e.g., wing development, flight) in a pest using a gene drive. Given that thermal gliding is a key flight strategy for long-distance migration in insects, a reduction in wing area can severely impair their survival. Damaged wings lead to a progressive shift from gliding to flapping flight, resulting in increased energetic costs [104].
Strategies aimed at population replacement may offer greater resilience against pest population migrations into and out of release areas. Specifically, manipulating migration behavior through gene drive technologies which target the genes associated with long-distance flight in pests can potentially fix genetically modified alleles in the released population. This approach not only prevents reinvasion by wild pest populations capable of long-distance dispersal but also restricts the migration of pests out of the release area.

4. How Is a CRISPR-Cas-Mediated Gene Drive Made and Introduced?

The CRISPR-Cas-mediated gene drive construct typically comprises four elements: a codon-optimized Cas gene, an sgRNA gene with a specific target site, an effector gene conferring the desired phenotype, and homology arms flanking the Cas-sgRNA cassette to enable insertion at the target site [105]. A fluorescent marker gene may be included for easier identification of the transgenic progeny. Upon insertion of one copy of the gene drive cassette into a chromosome, Cas and sgRNA are independently expressed, causing a double-strand break (DSB) on the homologous chromosome lacking the gene drive [106]. This results in the gene drive cassette being copied into the wild-type homologous chromosome via homology-directed repair (HDR), producing two copies of the gene drive in the genome (Figure 3). The effectiveness of CRISPR-Cas-mediated gene drive systems for population replacement in controlling insect pests relies on several key factors: the selection of suitable promoters that specifically activate during germline cell development and gametogenesis, the phenotypic consequences of gene disruptions or insertions, the genetic stability of gene drive constructs throughout propagation, the associated fitness costs, and the efficiency of genetic transmission across generations. The application of gene drive systems to agricultural pests remains in the preliminary stages compared to their implementation in flies and mosquitoes, primarily due to these complexities.
Significantly, a germline-specific promoter employed to drive Cas expression can facilitate high rates of HDR-mediated homing exclusively within germ-line precursor cells destined to differentiate into sperm or eggs. Cas is under the control of various housekeeping and tissue-specific promoters used in loss-of-function studies across multiple species [107,108]. The majority of CRISPR-Cas gene drive constructs feature Cas expression driven by the vasa ortholog promoter, initially characterized in D. melanogaster, where it is active in both male and female germlines [109]. However, vasa’s expression extends beyond the germline, potentially leading to frequent non-homologous end-joining (NHEJ) events and the emergence of drive-resistant loci following DSBs. An improved strategy involves utilizing the nanos promoter to restrict Cas expression to male and female germ cells, which has shown reduced somatic leakage and embryonic toxicity compared to vasa [110]. Nonetheless, resistance alleles can still emerge pre- and post-fertilization with nanos-Cas gene drive constructs due to the maternal deposition of Cas-induced NHEJ alleles in genetically diverse populations [52]. Thus far, several germline-specific candidate promoters such as beta2-tubulin, zero population growth (zpg), Rcd-1r, and Sry-α have been identified in various gene drive constructs tested in confined insect populations, demonstrating minimal occurrence of NHEJ alleles with these promoters [36,54,61]. However, most of these promoters were discovered in model insects like Drosophila and mosquitoes, underscoring the urgent need for additional germline-specific promoters across different insects to optimize CRISPR/Cas-mediated gene drives in pest management. Additionally, codon optimization of the Cas nuclease is crucial. Codon usage bias varies widely between bacteria and insects, influencing Cas protein expression levels heterogeneously in the target organisms [111,112]. Codon-optimized Cas sequences have been shown to enhance expression in non-native hosts [113,114,115,116]. Moreover, the 5′-untranslated region (5’-UTR) and 3′-untranslated region (3′-UTR) play critical roles in regulating gene expression and homing efficiency [117,118]. For instance, the first intron of αTub84B could enhance Cas expression when inserted into the UTR of the Cas cassette [119,120]. Similarly, both the 5’-UTR and 3′-UTR of the β2-tubulin gene (encoded by βTub85D) have been demonstrated to improve homing efficiency in the male germline when positioned upstream and downstream of the Cas coding sequence [66,121].
Expressing sgRNA using an appropriate promoter is similarly a critical step in the CRISPR-Cas-mediated gene drive. Typically, sgRNA cassettes in gene drive constructs are driven by the RNA polymerase III-dependent U6 promoter. U6 small nuclear RNA (snRNA), a conserved component of the spliceosome found widely across species, serves this role and has been characterized in numerous organisms. The U6 promoter is well-suited for transcribing various small RNAs with precisely defined 5’ and 3’ ends, facilitating ubiquitous and high-level expression of sgRNAs [122,123]. Optimal transcription from the U6 promoter favors a guanine as the initiation site, necessitating sgRNA target sequences to start with 5’G-N19-NGG-3′ [124]. However, studies have demonstrated a tolerance for extra nucleotides mismatched at the 5′ end of sgRNA, allowing for flexibility in the target sequence selection by appending a guanine nucleotide prior to the target sequence, thereby conforming to 5’G-N20-NGG-3′ [125]. Alternatively, U3, H1, or RNA polymerase II promoters, including those identified in cell- or tissue-specific genes with variable transcription initiation sites, can be utilized for sgRNA expression to achieve precise spatio-temporal gene modification [126,127,128,129]. The variable region of the sgRNA provides a target site located on the homologous chromosome, enabling the placement of sgRNAs in tandem within gene drive constructs to simultaneously target multiple sites. To increase the likelihood of successful mutagenesis, it is essential to design and evaluate multiple sgRNAs targeting a specific gene before selecting the optimal one for use in the gene drive. The tRNA sequences could be inserted between multiple sgRNAs in the gene drive design due to their superior processing efficiency and the challenges posed by repetitive sequences during plasmid preparation [130].
The effector gene can be tightly linked with Cas and sgRNA genes that confer the desired phenotypes under the control of sex- or tissue-specific promoters, enabling expression in the appropriate pest tissues. The gene drive cassette, within the donor plasmid containing regions homologous to the target locus, is typically delivered to insects by microinjection into pre-blastoderm embryos. To enhance the germline transformation and insertion efficiency of the gene drive construct, additional dsRNA related to the NHEJ pathway (e.g., Ku70, Ku80) could be included in the injection [51,131]. The entire gene drive construct is copied into the homologous site by the cell for HDR, demonstrating the stable integration of at least ~17 kb of exogenous genes into the insect genome [34]. Importantly, transgenic progeny with non-visible phenotypes are challenging to identify and maintain; however, this can be facilitated by incorporating a fluorescent protein (e.g., EGFP, mCherry, DsRed) under the control of a strong promoter into the donor construct.
The CRISPR-based gene drive system requires further refinement before practical implementation. One critical concern is its evolutionary stability. While studies demonstrate an expected increase in the frequency of the gene drive locus over several generations, long-term effectiveness is limited by the emergence of resistance alleles in diverse natural populations, hindering the conversion of drive heterozygotes into drive homozygotes [132,133,134]. Resistance often arises from Cas-induced DSBs repaired via NHEJ or MMEJ rather than drive incorporation through HDR, particularly in mutated target sites resistant to further sgRNA recognition [133]. The rate of NHEJ-mediated repair varies widely among the species, target sites, and developmental stages during which DSBs occur [135,136]. Mathematical modeling suggests that the formation of such resistance alleles could severely impede the spread of gene drive constructs in wild populations, especially when they confer a fitness cost [137,138]. Additionally, pre-existing resistant alleles due to single-nucleotide polymorphisms (SNPs) may already exist in diverse wild populations before deployment. Enhancing the genetic stability of gene drive constructs has been proposed by using multiple sgRNAs closely spaced along the target site. In an effort to enhance stability, researchers modified the CRISPR-based dsx gene drive system by incorporating the X-chromosome shredding nuclease I-PpoI. This adaptation resulted in a male-biased sex ratio and ultimately caused the collapse of a small laboratory cage population [61].
Recent gene drive efforts have focused on creating homing split drives with a modified backup of the target gene to combat resistance. Adolfi et al. devised a recoded gene drive rescue system for Anopheles stephensi to combat the fitness costs linked to nonfunctional target sites and to prevent the emergence of resistance alleles [49]. This system exhibited promise in multi-generational laboratory populations. Another successful approach, Home and Rescue (HomeR), demonstrated strong transmission rates and limited generation of resistant alleles in fly populations, proving its effectiveness in cage studies [56]. Champer et al. introduced a homing drive in D. melanogaster targeting a haplolethal gene using two sgRNAs and a rescue allele, which significantly reduced resistance allele prevalence [53]. The drive demonstrates high efficiency, with 91% inheritance of the drive allele from heterozygotes, and successfully spreads in population cage experiments, underscoring its effectiveness for precise genetic modification. Terradas et al. offers insights into the design of the sGD (split gene drive) system to precisely manage spread and containment within native populations [45]. Strategies proposed include utilizing a linked Cas9 source for the controlled introduction and removal of the endonuclease, adjusting drive intensity through lethal or sterile mosaicism, and enhancing population-wide propagation by varying Cas9 activity levels. These adaptable systems have the potential to maintain allelic drives that affect the inheritance of advantageous traits or to distribute the desired cargo, such as supplementary anti-pathogenic molecules, in addition to the initial modifications [139]. These innovations aim to enhance gene drive effectiveness by integrating rescue alleles and specific targeting strategies [46]. Moreover, designing gene drives that disrupt the target loci with high sequence conservation suggests functional constraints that deter sequence changes likely to impair function [140]. Future improvements might focus on targeting essential genes and integrating an enhanced rescue mechanism to reduce fitness costs and boost drive efficiency.
While the CRISPR-Cas system has proven to be effective and robust, embryo microinjection for genetic modification remains severely limited due to its technical demands, time consumption, and labor intensiveness. Specifically, population replacement necessitates mass-reared transgenic insects, and this low-throughput manual injection method significantly restricts the application of gene drive technologies. There is a pressing need for the development of high-throughput genome engineering approaches to deliver gene drive components in insects, such as electroporation or paratransgenesis [141,142]. A new method called ReMOT Control has been developed for delivering CRISPR components into developing oocytes within the adult ovaries of different insect species [143,144,145]. This approach presents a viable alternative to traditional embryo microinjection methods. While effective for knock-out gene editing, its potential to deliver HDR templates for knock-in modifications is still under investigation. Overcoming these technical challenges is expected with further research efforts.

5. Benefits and Risks of Gene Drives

CRISPR-Cas-mediated gene drives for population replacement in pest control have significant biosafety implications. Modifying organisms with gene drives effectively creates new species, which can be highly invasive and spread through populations despite fitness costs. This broad dispersal may lead to unintended propagation into non-target ecosystems, causing ecological changes due to irreversible replacement of wild-type populations [146].
The practical implementation of gene drives requires robust safeguards and strategies in genetic design and confinement methods to mitigate the associated risks. Given the possibility of escape from laboratory or field trials, it is crucial to secure gene drives to prevent accidental spread to natural populations. Current strategies, mainly relying on physical confinement, may not adequately reduce the risk of escape due to human error. Since only a few escapees are needed to establish a viable drive [147,148], additional safety measures are necessary for experiments involving those drives capable of indefinite spread. One approach involves using species- or sub-population-specific sgRNA target sites in the gene drive construct to restrict the spread of the drive system to the target population [149,150]. Alternatively, with high-threshold gene drives such as underdominance, modular, or split systems, individuals containing a specifically conserved motif could be initially released to establish a pest population. Subsequently, a second gene drive population delivering the desired payload could replace the established population, contingent upon confirming that the implanted sequence does not spread to non-target populations. Once released, a gene drive cannot be recalled if Cas and sgRNAs are within the same gene cassette. Therefore, a split gene drive system, which separates the sgRNA from an unlinked Cas cassette to induce biallelic mutations, is a safer strategy. This method restricts sustained spread to the offspring of homozygous parents lacking the Cas gene, significantly reducing the risk of accidental release and non-drive inheritance patterns of the sgRNA cassette in the population [38,45,55,151]. Such split systems have demonstrated comparable efficiency to non-split constructs in D. melanogaster, Plutella xylostella, and An. gambiae [41,43,152,153,154]. Lastly, gene drive systems must be designed to avoid foreseeable risks that could harm target populations or disrupt ecosystems [154].
Ideally, a reversible mechanism should be incorporated into the gene drive construct to facilitate the removal of effector genes in cases where unexpected negative effects arise in the replaced populations. Efforts to control homing drives have been explored through various approaches. These include constructs that can overwrite existing drives in D. melanogaster [155], methods in Anopheles gambiae to mitigate suppression drives by reducing genetic load [156], and chemical inducible systems for regulating Cas expression and drive efficiency in flies [157]. D’Amato et al. first investigated the effectiveness of an AcrIIA4-based anti-drive against a gene drive in Anopheles gambiae populations. The experiments confirmed the anti-drive’s efficacy under diverse conditions, thereby supporting its potential for controlling gene drives in insects [158]. These methods aim to limit the spread of gene drives or revert organisms to their wild-type phenotype. Prior to considering field trials, extensive laboratory studies and population cage experiments are essential [159,160]. The efficacy of effector genes should be evaluated across diverse genetic backgrounds and environmental conditions, progressing from contained trials to controlled field trials, with open field releases of gene drive-carrying individuals reserved for the project’s final stages. While CRISPR-Cas-mediated gene drive research is advancing rapidly in laboratory settings, it remains in an early proof-of-concept phase. Continuous assessment and replication of gene drive-mediated replacements are crucial before transitioning from laboratory research to real-world applications that are effective, safe, and sustainable. Assessing the safety of gene drive technology involves a comprehensive approach including environmental risk assessment, laboratory and field trials, and risk mitigation strategies. Ethical considerations and stakeholder engagement are crucial, along with compliance with regulatory frameworks. Ongoing monitoring and adaptive management are essential for detecting and addressing unintended consequences. Scientific peer review and transparency further ensure responsible development and deployment. Ultimately, the successful implementation of large-scale releases hinges on scientific and technological advancements, regulatory approval, and public acceptance.

6. Conclusions

Genetic modification techniques are increasingly appealing as alternatives to nonselective methods. Genetically edited insects are deployed as control agents with heritable modifications that alter their traits. CRISPR-Cas-mediated gene drive strategies for pest control represent a more advanced approach compared to other genetic control methods, offering high species-specificity and cost-effectiveness. Self-sustaining population replacement strategies aim to reduce pest populations by leveraging rapid spread through initial releases. In contrast, self-limiting population suppression strategies focus on sustained large-scale releases to diminish pest numbers in targeted areas [161]. These genetic pest control strategies rely on successful mating between modified and wild individuals, yet mating barriers—such as postzygotic and prezygotic barriers—may hinder gene flow. Natural selection and genetic drift can further impede mating between genetically edited insects and wild populations. Highly effective gene drive systems, known for their invasive potential in wild populations, offer significant advantages in genetic insect pest control. With the rapid advancements in CRISPR-Cas technology widely applied across various organisms, this breakthrough tool holds promise for realizing effective gene drives. The potential for gene drives to contribute to sustainable and effective insect pest control appears promising.

Author Contributions

Conceptualization, Y.Z. and L.L.; methodology, Y.Z., L.L., L.W., Y.W. and Z.H.; validation, L.W., Y.W. and Z.H.; formal analysis, Y.Z., L.L., L.W., Y.W. and Z.H.; investigation, Y.Z., L.L., L.W., Y.W. and Z.H.; resources, L.W., Y.W. and Z.H.; data curation, Y.W. and Z.H.; writing—original draft preparation, Y.Z.; writing—review and editing, Y.Z., L.L., L.W., Y.W. and Z.H.; visualization, Y.Z. supervision, Y.Z.; project administration, Y.Z., L.L., L.W., Y.W. and Z.H.; funding acquisition, Y.Z. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Scientific and Technological Innovation Foundation of Gansu Agricultural University-PhD Research Startup Foundation (GAU-KYQD-2019-26); the Fuxi Youth Talent Project of Gansu Agricultural University (Gaufx-03Y03); and the National Natural Science Foundation of China (31900347).

Data Availability Statement

All data were sourced from the scientific literature referenced in the References section. No additional data were generated.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chapman, A.D. Numbers of Living Species in Australia and the World, 2nd ed.; Australian Government, Department of the Environment and Energy: Canberra, Australia, 2009.
  2. Van Der Goes Van Naters, W.; Carlson, J.R. Insects as chemosensors of humans and crops. Nature 2006, 444, 302–307. [Google Scholar] [CrossRef] [PubMed]
  3. Ferry, N.; Gatehouse, A. Transgenic crop plants for resistance to biotic stress. In Transgenic Crop Plants; Springer: Berlin/Heidelberg, Germany, 2010; pp. 1–65. [Google Scholar]
  4. Knipling, E. Possibilities of insect control or eradication through the use of sexually sterile males. J. Econ. Entomol. 1955, 48, 459–462. [Google Scholar] [CrossRef]
  5. Casida, J.E.; Quistad, G.B. Golden age of insecticide research: Past, present, or future? Annu. Rev. Entomol. 1998, 43, 1–16. [Google Scholar] [CrossRef] [PubMed]
  6. Thomas, D.D.; Donnelly, C.A.; Wood, R.J.; Alphey, L.S. Insect population control using a dominant, repressible, lethal genetic system. Science 2000, 287, 2474–2476. [Google Scholar] [CrossRef]
  7. Xi, Z.; Dean, J.L.; Khoo, C.; Dobson, S.L. Generation of a novel Wolbachia infection in Aedes albopictus (Asian tiger mosquito) via embryonic microinjection. Insect Biochem. Mol. Biol. 2005, 35, 903–910. [Google Scholar] [CrossRef]
  8. Fraley, R.T.; Rogers, S.G.; Horsch, R.B.; Sanders, P.R.; Flick, J.S.; Adams, S.P.; Bittner, M.L.; Brand, L.A.; Fink, C.L.; Fry, J.S. Expression of bacterial genes in plant cells. Proc. Natl. Acad. Sci. USA 1983, 80, 4803–4807. [Google Scholar] [CrossRef]
  9. Charlesworth, B.; Langley, C.H. The population genetics of Drosophila transposable elements. Annu. Rev. Genet. 1989, 23, 251–287. [Google Scholar] [CrossRef] [PubMed]
  10. Ranner, H. Investigations on the biology and control of the cherry fruit fly, Rhagoletis cerasi L. (Diptera, Trypetidae)—V. Experiments on the control of the cherry fruit fly by means of the incompatible insect technique (IIT). Pflanzenschutzberichte 1990, 51, 1–16. [Google Scholar]
  11. Beeman, R.; Friesen, K.; Denell, R. Maternal-effect selfish genes in flour beetles. Science 1992, 256, 89–92. [Google Scholar] [CrossRef]
  12. Burt, A. Site-specific selfish genes as tools for the control and genetic engineering of natural populations. Proc. R. Soc. Lond. B Biol. Sci. 2003, 270, 921–928. [Google Scholar] [CrossRef]
  13. Knipling, E.F. The Basic Principles of Insect Population Suppression and Management; U.S. Department of Agriculture: Washington, DC, USA, 1979; pp. 1–659.
  14. Marec, F.; Vreysen, M.J. Advances and challenges of using the sterile insect technique for the management of pest Lepidoptera. Insects 2019, 10, 371. [Google Scholar] [CrossRef]
  15. Harris, A.F.; Nimmo, D.; McKemey, A.R.; Kelly, N.; Scaife, S.; Donnelly, C.A.; Beech, C.; Petrie, W.D.; Alphey, L. Field performance of engineered male mosquitoes. Nat. Biotechnol. 2011, 29, 1034–1037. [Google Scholar] [CrossRef] [PubMed]
  16. Bourtzis, K. Wolbachia-based technologies for insect pest population control. Adv. Exp. Med. Biol. 2008, 627, 104–113. [Google Scholar] [PubMed]
  17. Laven, H. Eradication of Culex pipiens fatigans through cytoplasmic incompatibility. Nature 1967, 216, 383–384. [Google Scholar] [CrossRef] [PubMed]
  18. Walker, J.T.; Suckling, D.M.; Wearing, C.H. Past, present, and future of integrated control of apple pests: The New Zealand experience. Annu. Rev. Entomol. 2017, 62, 231–248. [Google Scholar] [CrossRef]
  19. Guedes, R.; Smagghe, G.; Stark, J.; Desneux, N. Pesticide-induced stress in arthropod pests for optimized integrated pest management programs. Annu. Rev. Entomol. 2016, 61, 43–62. [Google Scholar] [CrossRef] [PubMed]
  20. Cayol, J.P. Changes in sexual behavior and life history traits of tephritid species caused by mass-rearing processes. In Fruit Flies (Tephritidae); Aluja, M., Norrbom, A.L., Eds.; CRC Press: Boca Raton, FL, USA, 2001; pp. 861–878. ISBN 978-0-429-12467-9. [Google Scholar]
  21. Curtis, C. Possible use of translocations to fix desirable genes in insect pest populations. Nature 1968, 218, 368–369. [Google Scholar] [CrossRef]
  22. Sinkins, S.P.; Gould, F. Gene drive systems for insect disease vectors. Nat. Rev. Genet. 2006, 7, 427–435. [Google Scholar] [CrossRef]
  23. Windbichler, N.; Menichelli, M.; Papathanos, P.A.; Thyme, S.B.; Li, H.; Ulge, U.Y.; Hovde, B.T.; Baker, D.; Monnat, R.J.; Burt, A. A synthetic homing endonuclease-based gene drive system in the human malaria mosquito. Nature 2011, 473, 212–215. [Google Scholar] [CrossRef]
  24. Beeman, R.W.; Friesen, K.S. Properties and natural occurrence of maternal-effect selfish genes (‘Medea’ factors) in the Red Flour Beetle, Tribolium castaneum. Heredity 1999, 82, 529–534. [Google Scholar] [CrossRef]
  25. Buchman, A.; Marshall, J.M.; Ostrovski, D.; Yang, T.; Akbari, O.S. Synthetically engineered Medea gene drive system in the worldwide crop pest Drosophila suzukii. Proc. Natl. Acad. Sci. USA 2018, 115, 4725–4730. [Google Scholar] [CrossRef]
  26. Barrangou, R.; van der Oost, J. RNA-mediated Adaptive Immunity in Bacteria and Archaea. In CRISPR-Cas Systems; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  27. Jinek, M.; Jiang, F.; Taylor, D.W.; Sternberg, S.H.; Kaya, E.; Ma, E.; Anders, C.; Hauer, M.; Zhou, K.; Lin, S. Structures of Cas9 endonucleases reveal RNA-mediated conformational activation. Science 2014, 343, 1247997. [Google Scholar] [CrossRef]
  28. Garneau, J.E.; Dupuis, M.-È.; Villion, M.; Romero, D.A.; Barrangou, R.; Boyaval, P.; Fremaux, C.; Horvath, P.; Magadán, A.H.; Moineau, S. The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA. Nature 2010, 468, 67–71. [Google Scholar] [CrossRef] [PubMed]
  29. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A. Multiplex genome engineering using CRISPR/Cas systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef]
  30. Sternberg, S.H.; Redding, S.; Jinek, M.; Greene, E.C.; Doudna, J.A. DNA interrogation by the CRISPR RNA-guided endonuclease Cas9. Nature 2014, 507, 62–67. [Google Scholar] [CrossRef]
  31. Wang, H.; Russa, M.L.; Qi, L.S. CRISPR/Cas9 in genome editing and beyond. Annu. Rev. Biochem. 2016, 85, 227–264. [Google Scholar] [CrossRef]
  32. Raban, R.; Marshall, J.M.; Hay, B.A.; Akbari, O.S. Manipulating the destiny of wild populations using CRISPR. Annu. Rev. Genet. 2023, 57, 361–390. [Google Scholar] [CrossRef] [PubMed]
  33. Gantz, V.M.; Bier, E. The mutagenic chain reaction: A method for converting heterozygous to homozygous mutations. Science 2015, 348, 442–444. [Google Scholar] [CrossRef] [PubMed]
  34. Gantz, V.M.; Jasinskiene, N.; Tatarenkova, O.; Fazekas, A.; Macias, V.M.; Bier, E.; James, A.A. Highly efficient Cas9-mediated gene drive for population modification of the malaria vector mosquito Anopheles stephensi. Proc. Natl. Acad. Sci. USA 2015, 112, E6736–E6743. [Google Scholar] [CrossRef]
  35. Hammond, A.; Galizi, R.; Kyrou, K.; Simoni, A.; Siniscalchi, C.; Katsanos, D.; Gribble, M.; Baker, D.; Marois, E.; Russell, S. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nat. Biotechnol. 2016, 34, 78–83. [Google Scholar] [CrossRef]
  36. Kyrou, K.; Hammond, A.M.; Galizi, R.; Kranjc, N.; Burt, A.; Beaghton, A.K.; Nolan, T.; Crisanti, A. A CRISPR-Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nat. Biotechnol. 2018, 36, 1062–1066. [Google Scholar] [CrossRef] [PubMed]
  37. Oberhofer, G.; Ivy, T.; Hay, B.A. Cleave and Rescue, a novel selfish genetic element and general strategy for gene drive. Proc. Natl. Acad. Sci. USA 2019, 116, 6250–6259. [Google Scholar] [CrossRef] [PubMed]
  38. Li, M.; Yang, T.; Kandul, N.P.; Bui, M.; Gamez, S.; Raban, R.; Bennett, J.; Sánchez C, H.M.; Lanzaro, G.C.; Schmidt, H.; et al. Development of a confinable gene drive system in the human disease vector Aedes aegypti. eLife 2020, 9, e51701. [Google Scholar] [CrossRef] [PubMed]
  39. Yadav, A.K.; Butler, C.; Yamamoto, A.; Patil, A.A.; Lloyd, A.L.; Scott, M.J. CRISPR/Cas9-based split homing gene drive targeting doublesex for population suppression of the global fruit pest Drosophila suzukii. Proc. Natl. Acad. Sci. USA 2023, 120, e2301525120. [Google Scholar] [CrossRef]
  40. Meccariello, A.; Hou, S.; Davydova, S.; Fawcett, J.D.; Siddall, A.; Leftwich, P.T.; Krsticevic, F.; Papathanos, P.A.; Windbichler, N. Gene drive and genetic sex conversion in the global agricultural pest Ceratitis capitata. Nat. Commun. 2024, 15, 372. [Google Scholar] [CrossRef]
  41. Xu, X.; Harvey-Samuel, T.; Siddiqui, H.A.; Ang, J.X.D.; Anderson, M.E.; Reitmayer, C.M.; Lovett, E.; Leftwich, P.T.; You, M.; Alphey, L. Toward a CRISPR-Cas9-based gene drive in the diamondback moth Plutella xylostella. CRISPR J. 2022, 5, 224–236. [Google Scholar] [CrossRef]
  42. Oberhofer, G.; Ivy, T.; Hay, B.A. Gene drive and resilience through renewal with next generation Cleave and Rescue selfish genetic elements. Proc. Natl. Acad. Sci. USA 2020, 117, 9013–9021. [Google Scholar] [CrossRef]
  43. Champer, J.; Lee, E.; Yang, E.; Liu, C.; Clark, A.G.; Messer, P.W. A toxin-antidote CRISPR gene drive system for regional population modification. Nat. Commun. 2020, 11, 1082. [Google Scholar] [CrossRef]
  44. Champer, J.; Yang, E.; Lee, E.; Liu, J.; Clark, A.G.; Messer, P.W. A CRISPR homing gene drive targeting a haplolethal gene removes resistance alleles and successfully spreads through a cage population. Proc. Natl. Acad. Sci. USA 2020, 117, 24377–24383. [Google Scholar] [CrossRef]
  45. Terradas, G.; Buchman, A.B.; Bennett, J.B.; Shriner, I.; Marshall, J.M.; Akbari, O.S.; Bier, E. Inherently confinable split-drive systems in Drosophila. Nat. Commun. 2021, 12, 1480. [Google Scholar] [CrossRef]
  46. Nash, A.; Capriotti, P.; Hoermann, A.; Papathanos, P.A.; Windbichler, N. Intronic gRNAs for the construction of minimal gene drive systems. Front. Bioeng. Biotechnol. 2022, 10, 857460. [Google Scholar] [CrossRef]
  47. Sanz Juste, S.; Okamoto, E.M.; Nguyen, C.; Feng, X.; López Del Amo, V. Next-generation CRISPR gene-drive systems using Cas12a nuclease. Nat. Commun. 2023, 14, 6388. [Google Scholar] [CrossRef]
  48. Carballar-Lejarazú, R.; Ogaugwu, C.; Tushar, T.; Kelsey, A.; Pham, T.B.; Murphy, J.; Schmidt, H.; Lee, Y.; Lanzaro, G.C.; James, A.A. Next-generation gene drive for population modification of the malaria vector mosquito, Anopheles gambiae. Proc. Natl. Acad. Sci. USA 2020, 117, 22805–22814. [Google Scholar] [CrossRef] [PubMed]
  49. Adolfi, A.; Gantz, V.M.; Jasinskiene, N.; Lee, H.-F.; Hwang, K.; Terradas, G.; Bulger, E.A.; Ramaiah, A.; Bennett, J.B.; Emerson, J. Efficient population modification gene-drive rescue system in the malaria mosquito Anopheles stephensi. Nat. Commun. 2020, 11, 5553. [Google Scholar] [CrossRef]
  50. Hammond, A.; Karlsson, X.; Morianou, I.; Kyrou, K.; Beaghton, A.; Gribble, M.; Kranjc, N.; Galizi, R.; Burt, A.; Crisanti, A. Regulating the expression of gene drives is key to increasing their invasive potential and the mitigation of resistance. PLoS Genet. 2021, 17, e1009321. [Google Scholar] [CrossRef] [PubMed]
  51. Asad, M.; Liu, D.; Li, J.; Chen, J.; Yang, G. Development of CRISPR/Cas9-mediated gene-drive construct targeting the phenotypic gene in Plutella xylostella. Front. Physiol. 2022, 13, 938621. [Google Scholar] [CrossRef]
  52. Champer, J.; Reeves, R.; Oh, S.Y.; Liu, C.; Liu, J.; Clark, A.G.; Messer, P.W. Novel CRISPR/Cas9 gene drive constructs reveal insights into mechanisms of resistance allele formation and drive efficiency in genetically diverse populations. PLoS Genet. 2017, 13, e1006796. [Google Scholar] [CrossRef] [PubMed]
  53. Champer, J.; Liu, J.; Oh, S.Y.; Reeves, R.; Luthra, A.; Oakes, N.; Clark, A.G.; Messer, P.W. Reducing resistance allele formation in CRISPR gene drive. Proc. Natl. Acad. Sci. USA 2018, 115, 5522–5527. [Google Scholar] [CrossRef]
  54. Carrami, E.M.; Eckermann, K.N.; Ahmed, H.M.; Sánchez, C.H.M.; Dippel, S.; Marshall, J.M.; Wimmer, E.A. Consequences of resistance evolution in a Cas9-based sex conversion-suppression gene drive for insect pest management. Proc. Natl. Acad. Sci. USA 2018, 115, 6189–6194. [Google Scholar] [CrossRef]
  55. López Del Amo, V.; Bishop, A.L.; Sánchez C, H.M.; Bennett, J.B.; Feng, X.; Marshall, J.M.; Bier, E.; Gantz, V.M. A transcomplementing gene drive provides a flexible platform for laboratory investigation and potential field deployment. Nat. Commun. 2020, 11, 352. [Google Scholar] [CrossRef]
  56. Kandul, N.P.; Liu, J.; Bennett, J.B.; Marshall, J.M.; Akbari, O.S. A confinable home-and-rescue gene drive for population modification. eLife 2021, 10, e65939. [Google Scholar] [CrossRef] [PubMed]
  57. Yang, E.; Metzloff, M.; Langmüller, A.M.; Xu, X.; Clark, A.G.; Messer, P.W.; Champer, J. A homing suppression gene drive with multiplexed gRNAs maintains high drive conversion efficiency and avoids functional resistance alleles. G3 2022, 12, jkac081. [Google Scholar] [CrossRef]
  58. Chen, J.; Xu, X.; Champer, J. Assessment of distant-site rescue elements for CRISPR toxin-antidote gene drives. Front. Bioeng. Biotechnol. 2023, 11, 1138702. [Google Scholar] [CrossRef]
  59. Chen, W.; Guo, J.; Liu, Y.; Champer, J. Population suppression by release of insects carrying a dominant sterile homing gene drive targeting doublesex in Drosophila. bioRxiv 2023. [Google Scholar] [CrossRef]
  60. Ma, S.; Ni, X.; Chen, S.; Qiao, X.; Xu, X.; Champer, J.; Huang, J. A small-molecule approach to restore female sterility phenotype targeted by a homing suppression gene drive in the fruit pest Drosophila suzukii. PLoS Genet. 2024, 20, e1011226. [Google Scholar] [CrossRef] [PubMed]
  61. Simoni, A.; Hammond, A.M.; Beaghton, A.K.; Galizi, R.; Taxiarchi, C.; Kyrou, K.; Meacci, D.; Gribble, M.; Morselli, G.; Burt, A. A male-biased sex-distorter gene drive for the human malaria vector Anopheles gambiae. Nat. Biotechnol. 2020, 38, 1054–1060. [Google Scholar] [CrossRef] [PubMed]
  62. Ying, Y.; Aumann, R.A.; Häcker, I.; Schetelig, M.F. CRISPR-based genetic control strategies for insect pests. J. Integr. Agric. 2023, 22, 651–668. [Google Scholar]
  63. Burt, A. Heritable strategies for controlling insect vectors of disease. Phil. Trans. R. Soc. B 2014, 369, 20130432. [Google Scholar] [CrossRef]
  64. Akbari, O.S.; Matzen, K.D.; Marshall, J.M.; Huang, H.; Ward, C.M.; Hay, B.A. A synthetic gene drive system for local, reversible modification and suppression of insect populations. Curr. Biol. 2013, 23, 671–677. [Google Scholar] [CrossRef]
  65. Galizi, R.; Doyle, L.A.; Menichelli, M.; Bernardini, F.; Deredec, A.; Burt, A.; Stoddard, B.L.; Windbichler, N.; Crisanti, A. A synthetic sex ratio distortion system for the control of the human malaria mosquito. Nat. Commun. 2014, 5, 3977. [Google Scholar] [CrossRef]
  66. Fasulo, B.; Meccariello, A.; Morgan, M.; Borufka, C.; Papathanos, P.A.; Windbichler, N. A fly model establishes distinct mechanisms for synthetic CRISPR/Cas9 sex distorters. PLoS Genet. 2020, 16, e1008647. [Google Scholar] [CrossRef] [PubMed]
  67. Deredec, A.; Burt, A.; Godfray, H.C. The population genetics of using homing endonuclease genes in vector and pest management. Genetics 2008, 179, 2013–2026. [Google Scholar] [CrossRef] [PubMed]
  68. Osta, M.A.; Christophides, G.K.; Kafatos, F.C. Effects of mosquito genes on Plasmodium development. Science 2004, 303, 2030–2032. [Google Scholar] [CrossRef]
  69. Franz, A.W.; Sanchez-Vargas, I.; Adelman, Z.N.; Blair, C.D.; Beaty, B.J.; James, A.A.; Olson, K.E. Engineering RNA interference–based resistance to dengue virus type 2 in genetically modified Aedes aegypti. Proc. Natl. Acad. Sci. USA 2006, 103, 4198–4203. [Google Scholar] [CrossRef]
  70. Isaacs, A.T.; Jasinskiene, N.; Tretiakov, M.; Thiery, I.; Zettor, A.; Bourgouin, C.; James, A.A. Transgenic Anopheles stephensi coexpressing single-chain antibodies resist Plasmodium falciparum development. Proc. Natl. Acad. Sci. USA 2012, 109, E1922–E1930. [Google Scholar] [CrossRef]
  71. Dong, Y.; Simões, M.L.; Dimopoulos, G. Versatile transgenic multistage effector-gene combinations for Plasmodium falciparum suppression in Anopheles. Sci. Adv. 2020, 6, eaay5898. [Google Scholar] [CrossRef]
  72. Hamilton, W.D. Extraordinary sex ratios. Science 1967, 156, 477–488. [Google Scholar] [CrossRef]
  73. Courtier-Orgogozo, V.; Morizot, B.; Boëte, C. Agricultural pest control with CRISPR-based gene drive: Time for public debate: Should we use gene drive for pest control? EMBO Rep. 2017, 18, 878–880. [Google Scholar] [CrossRef]
  74. Renou, M.; Anton, S. Insect olfactory communication in a complex and changing world. Curr. Opin. Insect Sci. 2020, 42, 1–7. [Google Scholar] [CrossRef]
  75. Jury, S.; Chabot, C.C.; Goldstein, J.; Harzsch, S. Sensory biology and behaviour. In Ecophysiology of the European Green Crab (Carcinus Maenas) and Related Species; Elsevier: Amsterdam, The Netherlands, 2024; pp. 123–157. [Google Scholar]
  76. Jiang, C.; Xu, H.; Yang, L.; Liu, J.; Li, Y.; Takei, K.; Xu, W. Neuromorphic antennal sensory system. Nat. Commun. 2024, 15, 2109. [Google Scholar] [CrossRef]
  77. Gadenne, C.; Barrozo, R.B.; Anton, S. Plasticity in insect olfaction: To smell or not to smell? Annu. Rev. Entomol. 2016, 61, 317–333. [Google Scholar] [CrossRef] [PubMed]
  78. Pírez, N.; Klappenbach, M.; Locatelli, F.F. Experience-dependent tuning of the olfactory system. Curr. Opin. Insect Sci. 2023, 60, 101117. [Google Scholar] [CrossRef] [PubMed]
  79. Zhao, Z.; Zung, J.L.; Hinze, A.; Kriete, A.L.; Iqbal, A.; Younger, M.A.; Matthews, B.J.; Merhof, D.; Thiberge, S.; Ignell, R. Mosquito brains encode unique features of human odour to drive host seeking. Nature 2022, 605, 706–712. [Google Scholar] [CrossRef] [PubMed]
  80. McBride, C.S.; Baier, F.; Omondi, A.B.; Spitzer, S.A.; Lutomiah, J.; Sang, R.; Ignell, R.; Vosshall, L.B. Evolution of mosquito preference for humans linked to an odorant receptor. Nature 2014, 515, 222. [Google Scholar] [CrossRef]
  81. Zhao, J.; Chen, A.Q.; Ryu, J.; Del Mármol, J. Structural basis of odor sensing by insect heteromeric odorant receptors. Science 2024, 384, eadn6384. [Google Scholar] [CrossRef]
  82. Gomes, J.V.; Singh-Bhagania, S.; Cenci, M.; Chacon Cordon, C.; Singh, M.; Butterwick, J.A. The molecular basis of sugar detection by an insect taste receptor. Nature 2024, 629, 228–234. [Google Scholar] [CrossRef]
  83. Ma, D.; Hu, M.; Yang, X.; Liu, Q.; Ye, F.; Cai, W.; Wang, Y.; Xu, X.; Chang, S.; Wang, R. Structural basis for sugar perception by Drosophila gustatory receptors. Science 2024, 383, eadj2609. [Google Scholar] [CrossRef]
  84. Zhang, S.S.; Wang, P.C.; Ning, C.; Yang, K.; Li, G.C.; Cao, L.L.; Huang, L.Q.; Wang, C.Z. The larva and adult of Helicoverpa armigera use differential gustatory receptors to sense sucrose. eLife 2024, 12, RP91711. [Google Scholar] [CrossRef]
  85. Arntsen, C.; Guillemin, J.; Audette, K.; Stanley, M. Tastant-receptor interactions: Insights from the fruit fly. Front. Nutr. 2024, 11, 1394697. [Google Scholar] [CrossRef]
  86. Pentzold, S.; Burse, A.; Boland, W. Contact chemosensation of phytochemicals by insect herbivores. Nat. Prod. Rep. 2017, 34, 478–483. [Google Scholar] [CrossRef]
  87. Zhang, S.; Tang, J.; Li, Y.; Li, D.; Chen, G.; Chen, L.; Yang, Z.; He, N. The silkworm gustatory receptor BmGr63 is dedicated to the detection of isoquercetin in mulberry. Proc. Biol. Sci. 2022, 289, 20221427. [Google Scholar] [CrossRef]
  88. Frank, H.M.; Walujkar, S.; Walsh, R.M.; Laursen, W.J.; Theobald, D.L.; Garrity, P.A.; Gaudet, R. Structural basis of ligand specificity and channel activation in an insect gustatory receptor. Cell Rep. 2024, 43, 4. [Google Scholar] [CrossRef] [PubMed]
  89. Kos, M.; Houshyani, B.; Overeem, A.J.; Bouwmeester, H.J.; Weldegergis, B.T.; van Loon, J.J.; Dicke, M.; Vet, L.E. Genetic engineering of plant volatile terpenoids: Effects on a herbivore, a predator and a parasitoid. Pest Manag. Sci. 2013, 69, 302–311. [Google Scholar] [CrossRef] [PubMed]
  90. Qian, J.; Zhu, C.; Jian, G.; Zeng, L.; Yang, Y. Release patterns and potential utility of herbivore-induced plant volatiles in crops: A review. Environ. Exp. Bot. 2024, 219, 105659. [Google Scholar] [CrossRef]
  91. Mann, R.S.; Ali, J.G.; Hermann, S.L.; Tiwari, S.; Pelz-Stelinski, K.S.; Alborn, H.T.; Stelinski, L.L. Induced release of a plant-defense volatile ‘deceptively’ attracts insect vectors to plants infected with a bacterial pathogen. PLoS Pathog. 2012, 8, e1002610. [Google Scholar] [CrossRef]
  92. Volpe, H.X.; Carmo-Sousa, M.; Luvizotto, R.A.; de Freitas, R.; Esperança, V.; Darolt, J.C.; Pegoraro, A.A.; Magalhães, D.M.; Favaris, A.P.; Wulff, N.A. The greening-causing agent alters the behavioral and electrophysiological responses of the Asian citrus psyllid to a putative sex pheromone. Sci. Rep. 2024, 14, 455. [Google Scholar] [CrossRef]
  93. Grimaldi, D.; Engel, M.S. Evolution of the Insects; Cambridge University Press: New York, NY, USA, 2005; ISBN 13 978-0-521-82149-0. [Google Scholar]
  94. Xu, H.J.; Zhang, C.X. Insulin receptors and wing dimorphism in rice planthoppers. Philos. Trans. R. Soc. Lond. B. Biol Sci. 2017, 372, 20150489. [Google Scholar] [CrossRef]
  95. Wagner, D.L.; Liebherr, J.K. Flightlessness in insects. Trends Ecol. Evol. 1992, 7, 216–220. [Google Scholar] [CrossRef]
  96. Zera, A.J.; Denno, R.F. Physiology and ecology of dispersal polymorphism in insects. Annu. Rev. Entomol. 1997, 42, 207–230. [Google Scholar] [CrossRef]
  97. Zhao, Z.; Zera, A.J. Differential lipid biosynthesis underlies a tradeoff between reproduction and flight capability in a wing-polymorphic cricket. Proc. Natl. Acad. Sci. USA 2002, 99, 16829–16834. [Google Scholar] [CrossRef]
  98. Simpson, S.J.; Sword, G.A.; Lo, N. Polyphenism in insects. Curr. Biol. 2011, 21, R738–R749. [Google Scholar] [CrossRef] [PubMed]
  99. Nijhout, H.F. Control mechanisms of polyphenic development in insects: In polyphenic development, environmental factors alter some aspects of development in an orderly and predictable way. Bioscience 1999, 49, 181–192. [Google Scholar] [CrossRef]
  100. Roff, D.A. The genetic basis of wing dimorphism in the sand cricket, Gryllus firmus and its relevance to the evolution of wing dimorphisms in insects. Heredity 1986, 57, 221. [Google Scholar] [CrossRef]
  101. Xu, H.J.; Xue, J.; Lu, B.; Zhang, X.C.; Zhuo, J.C.; He, S.F.; Ma, X.F.; Jiang, Y.Q.; Fan, H.W.; Xu, J.Y. Two insulin receptors determine alternative wing morphs in planthoppers. Nature 2015, 519, 464. [Google Scholar] [CrossRef] [PubMed]
  102. Yu, J.; Zhao, W.; Chen, X.; Lu, H.; Xiao, Y.; Li, Q.; Luo, L.; Kang, L.; Cui, F. A plant virus manipulates the long-winged morph of insect vectors. Proc. Natl. Acad. Sci. USA 2024, 121, e2315341121. [Google Scholar] [CrossRef]
  103. Roff, D.A.; Fairbairn, D.J. The evolution and genetics of migration in insects. Bioscience 2007, 57, 155–164. [Google Scholar] [CrossRef]
  104. Korkmaz, R.; Rajabi, H.; Eshghi, S.; Gorb, S.N.; Büscher, T.H. The frequency of wing damage in a migrating butterfly. Insect Sci. 2023, 30, 1507–1517. [Google Scholar] [CrossRef] [PubMed]
  105. Bier, E. Gene drives gaining speed. Nat. Rev. Genet. 2022, 23, 5–22. [Google Scholar] [CrossRef] [PubMed]
  106. Nambiar, T.S.; Baudrier, L.; Billon, P.; Ciccia, A. CRISPR-based genome editing through the lens of DNA repair. Mol. Cell 2022, 82, 348–388. [Google Scholar] [CrossRef]
  107. Li, M.; Bui, M.; Yang, T.; Bowman, C.S.; White, B.J.; Akbari, O.S. Germline Cas9 expression yields highly efficient genome engineering in a major worldwide disease vector, Aedes aegypti. Proc. Natl. Acad. Sci. USA 2017, 114, E10540–E10549. [Google Scholar] [CrossRef]
  108. Port, F.; Boutros, M. Tissue-specific CRISPR-Cas9 screening in Drosophila. Methods Mol. Biol. 2022, 2540, 157–176. [Google Scholar] [PubMed]
  109. Yajima, M.; Wessel, G.M. Essential elements for translation: The germline factor Vasa functions broadly in somatic cells. Development 2015, 142, 1960–1970. [Google Scholar] [CrossRef] [PubMed]
  110. Port, F.; Chen, H.M.; Lee, T.; Bullock, S.L. Optimized CRISPR/Cas tools for efficient germline and somatic genome engineering in Drosophila. Proc. Natl. Acad. Sci. USA 2014, 111, 2967–2976. [Google Scholar] [CrossRef] [PubMed]
  111. Liu, Y.; Yang, Q.; Zhao, F. Synonymous but not silent: The codon usage code for gene expression and protein folding. Annu. Rev. Biochem. 2021, 90, 375–401. [Google Scholar] [CrossRef]
  112. Zhou, M.; Guo, J.; Cha, J.; Chae, M.; Chen, S.; Barral, J.M.; Sachs, M.S.; Liu, Y. Non-optimal codon usage affects expression, structure and function of clock protein FRQ. Nature 2013, 495, 111. [Google Scholar] [CrossRef]
  113. Gao, J.; Wang, G.; Ma, S.; Xie, X.; Wu, X.; Zhang, X.; Wu, Y.; Zhao, P.; Xia, Q. CRISPR/Cas9-mediated targeted mutagenesis in Nicotiana tabacum. Plant Mol. Biol. 2015, 87, 99–110. [Google Scholar] [CrossRef]
  114. Gratz, S.J.; Cummings, A.M.; Nguyen, J.N.; Hamm, D.C.; Donohue, L.K.; Harrison, M.M.; Wildonger, J.; O’Connor-Giles, K.M. Genome engineering of Drosophila with the CRISPR RNA-guided Cas9 nuclease. Genetics 2013, 194, 1029–1035. [Google Scholar] [CrossRef] [PubMed]
  115. Zhou, H.; Liu, B.; Weeks, D.P.; Spalding, M.H.; Yang, B. Large chromosomal deletions and heritable small genetic changes induced by CRISPR/Cas9 in rice. Nucleic Acids Res. 2014, 42, 10903–10914. [Google Scholar] [CrossRef] [PubMed]
  116. Kulkarni, A.; Yu, W.; Moon, A.S.; Pandey, A.; Hanley, K.A.; Xu, J. Programmable CRISPR interference for gene silencing using Cas13a in mosquitoes. J. Genom. 2020, 8, 30. [Google Scholar] [CrossRef]
  117. Anderson, M.A.; Gonzalez, E.; Edgington, M.P.; Ang, J.X.; Purusothaman, D.K.; Shackleford, L.; Nevard, K.; Verkuijl, S.A.; Harvey-Samuel, T.; Leftwich, P.T. A multiplexed, confinable CRISPR/Cas9 gene drive can propagate in caged Aedes aegypti populations. Nat. Commun. 2024, 15, 729. [Google Scholar] [CrossRef]
  118. Rylee, J.C.; Nin-Velez, A.; Mahato, S.; Helms, K.J.; Wade, M.J.; Zentner, G.E.; Zelhof, A.C. Generating and testing the efficacy of transgenic Cas9 in Tribolium castaneum. Insect. Mol. Biol. 2022, 31, 543–550. [Google Scholar] [CrossRef] [PubMed]
  119. Tracey Jr, W.D.; Ning, X.; Klingler, M.; Kramer, S.G.; Gergen, J.P. Quantitative analysis of gene function in the Drosophila embryo. Genetics 2000, 154, 273–284. [Google Scholar] [CrossRef] [PubMed]
  120. Duncker, B.; Davies, P.; Walker, V. Introns boost transgene expression in Drosophila melanogaster. Mol. Gen. Genet. 1997, 254, 291–296. [Google Scholar] [CrossRef] [PubMed]
  121. Haber, D.A.; Arien, Y.; Lamdan, L.B.; Alcalay, Y.; Zecharia, C.; Krsticevic, F.; Yonah, E.S.; Avraham, R.D.; Krzywinska, E.; Krzywinski, J. Targeting mosquito X-chromosomes reveals complex transmission dynamics of sex ratio distorting gene drives. Nat. Commun. 2024, 15, 4983. [Google Scholar] [CrossRef] [PubMed]
  122. Zeng, B.; Zhan, S.; Wang, Y.; Huang, Y.; Xu, J.; Liu, Q.; Li, Z.; Huang, Y.; Tan, A. Expansion of CRISPR targeting sites in Bombyx mori. Insect Biochem. Mol. Biol. 2016, 72, 31–40. [Google Scholar] [CrossRef] [PubMed]
  123. Huang, Y.; Wang, Y.; Zeng, B.; Liu, Z.; Xu, X.; Meng, Q.; Huang, Y.; Yang, G.; Vasseur, L.; Gurr, G.M. Functional characterization of Pol III U6 promoters for gene knockdown and knockout in Plutella xylostella. Insect Biochem. Mol. Biol. 2017, 89, 71–78. [Google Scholar] [CrossRef]
  124. Jinek, M.; East, A.; Cheng, A.; Lin, S.; Ma, E.; Doudna, J. RNA-programmed genome editing in human cells. eLife 2013, 2, e00471. [Google Scholar] [CrossRef]
  125. Fu, Y.; Sander, J.D.; Reyon, D.; Cascio, V.M.; Joung, J.K. Improving CRISPR-Cas nuclease specificity using truncated guide RNAs. Nat. Biotechnol. 2014, 32, 279. [Google Scholar] [CrossRef]
  126. Ma, H.; Wu, Y.; Dang, Y.; Choi, J.G.; Zhang, J.; Wu, H. Pol III promoters to express small RNAs: Delineation of transcription initiation. Mol. Ther. Nucleic Acids 2014, 3, E161. [Google Scholar] [CrossRef]
  127. Yoshioka, S.; Fujii, W.; Ogawa, T.; Sugiura, K.; Naito, K. Development of a mono-promoter-driven CRISPR/Cas9 system in mammalian cells. Sci. Rep. 2015, 5, 18341. [Google Scholar] [CrossRef]
  128. Nissim, L.; Perli, S.D.; Fridkin, A.; Perez-Pinera, P.; Lu, T.K. Multiplexed and programmable regulation of gene networks with an integrated RNA and CRISPR/Cas toolkit in human cells. Mol. Cell 2014, 54, 698–710. [Google Scholar] [CrossRef] [PubMed]
  129. Mali, P.; Yang, L.; Esvelt, K.M.; Aach, J.; Guell, M.; Dicarlo, J.E.; Norville, J.E.; Church, G.M. RNA-Guided Human Genome Engineering via Cas9. Science 2013, 339, 823. [Google Scholar] [CrossRef] [PubMed]
  130. Xu, L.; Zhao, L.; Gao, Y.; Xu, J.; Han, R. Empower multiplex cell and tissue-specific CRISPR-mediated gene manipulation with self-cleaving ribozymes and tRNA. Nucleic Acids Res. 2017, 45, e28. [Google Scholar] [CrossRef] [PubMed]
  131. Zhu, L.; Mon, H.; Xu, J.; Lee, J.M.; Kusakabe, T. CRISPR/Cas9-mediated knockout of factors in non-homologous end joining pathway enhances gene targeting in silkworm cells. Sci. Rep. 2015, 5, 18103. [Google Scholar] [CrossRef] [PubMed]
  132. Reed, F.A. CRISPR/Cas9 gene drive: Growing pains for a new technology. Genetics 2017, 205, 1037–1039. [Google Scholar] [CrossRef] [PubMed]
  133. Unckless, R.L.; Clark, A.G.; Messer, P.W. Evolution of resistance against CRISPR/Cas9 gene drive. Genetics 2017, 205, 827–841. [Google Scholar] [CrossRef] [PubMed]
  134. Pham, T.B.; Phong, C.H.; Bennett, J.B.; Hwang, K.; Jasinskiene, N.; Parker, K.; Stillinger, D.; Marshall, J.M.; Carballar-Lejarazú, R.; James, A.A. Experimental population modification of the malaria vector mosquito, Anopheles stephensi. PLoS Genet. 2019, 15, e1008440. [Google Scholar] [CrossRef] [PubMed]
  135. Fu, Y.W.; Dai, X.Y.; Wang, W.T.; Yang, Z.X.; Zhao, J.J.; Zhang, J.P.; Wen, W.; Zhang, F.; Oberg, K.C.; Zhang, L. Dynamics and competition of CRISPR-Cas9 ribonucleoproteins and AAV donor-mediated NHEJ, MMEJ and HDR editing. Nucleic Acids Res. 2021, 49, 969–985. [Google Scholar] [CrossRef]
  136. Roy, S.; Juste, S.S.; Sneider, M.; Auradkar, A.; Klanseck, C.; Li, Z.; Julio, A.H.F.; Del Amo, V.L.; Bier, E.; Guichard, A. Cas9/Nickase-induced allelic conversion by homologous chromosome-templated repair in Drosophila somatic cells. Sci. Adv. 2022, 8, eabo0721. [Google Scholar] [CrossRef]
  137. Eckhoff, P.A.; Wenger, E.A.; Godfray, H.C.J.; Burt, A. Impact of mosquito gene drive on malaria elimination in a computational model with explicit spatial and temporal dynamics. Proc. Natl. Acad. Sci. USA 2017, 114, E255–E264. [Google Scholar] [CrossRef]
  138. Kim, J.; Harris, K.D.; Kim, I.K.; Shemesh, S.; Messer, P.W.; Greenbaum, G. Incorporating ecology into gene drive modelling. Ecol. Lett. 2023, 26, S62–S80. [Google Scholar] [CrossRef] [PubMed]
  139. Guichard, A.; Haque, T.; Bobik, M.; Xu, X.-R.S.; Klanseck, C.; Kushwah, R.B.S.; Berni, M.; Kaduskar, B.; Gantz, V.M.; Bier, E. Efficient allelic-drive in Drosophila. Nat. Commun. 2019, 10, 1640. [Google Scholar] [CrossRef] [PubMed]
  140. Noble, C.; Olejarz, J.; Esvelt, K.M.; Church, G.M.; Nowak, M.A. Evolutionary dynamics of CRISPR gene drives. Sci. Adv. 2017, 3, e1601964. [Google Scholar] [CrossRef] [PubMed]
  141. Ando, T.; Fujiwara, H. Electroporation-mediated somatic transgenesis for rapid functional analysis in insects. Development 2013, 140, 454–458. [Google Scholar] [CrossRef] [PubMed]
  142. Qin, W.; Dion, S.L.; Kutny, P.M.; Zhang, Y.; Cheng, A.W.; Jillette, N.L.; Malhotra, A.; Geurts, A.M.; Chen, Y.G.; Wang, H. Efficient CRISPR/Cas9-mediated genome editing in mice by zygote electroporation of nuclease. Genetics 2015, 200, 423–430. [Google Scholar] [CrossRef] [PubMed]
  143. Chaverra-Rodriguez, D.; Macias, V.M.; Hughes, G.L.; Pujhari, S.; Suzuki, Y.; Peterson, D.R.; Kim, D.; McKeand, S.; Rasgon, J.L. Targeted delivery of CRISPR-Cas9 ribonucleoprotein into arthropod ovaries for heritable germline gene editing. Nat. Commun. 2018, 9, 3008. [Google Scholar] [CrossRef] [PubMed]
  144. Chaverra-Rodriguez, D.; Dalla Benetta, E.; Heu, C.C.; Rasgon, J.L.; Ferree, P.M.; Akbari, O.S. Germline mutagenesis of Nasonia vitripennis through ovarian delivery of CRISPR-Cas9 ribonucleoprotein. Insect. Mol. Biol. 2020, 29, 569–577. [Google Scholar] [CrossRef] [PubMed]
  145. Macias, V.M.; McKeand, S.; Chaverra-Rodriguez, D.; Hughes, G.L.; Fazekas, A.; Pujhari, S.; Jasinskiene, N.; James, A.A.; Rasgon, J.L. Cas9-mediated gene-editing in the malaria mosquito Anopheles stephensi by ReMOT Control. G3 2020, 10, 1353–1360. [Google Scholar] [CrossRef]
  146. Marshall, J.M.; Akbari, O.S. Can CRISPR-based gene drive be confined in the wild? A question for molecular and population biology. ACS Chem. Biol. 2018, 13, 424–430. [Google Scholar] [CrossRef]
  147. Unckless, R.L.; Messer, P.W.; Connallon, T.; Clark, A.G. Modeling the manipulation of natural populations by the mutagenic chain reaction. Genetics 2015, 201, 425–431. [Google Scholar] [CrossRef]
  148. Noble, C.; Adlam, B.; Church, G.M.; Esvelt, K.M.; Nowak, M.A. Current CRISPR gene drive systems are likely to be highly invasive in wild populations. eLife 2018, 7, e33423. [Google Scholar] [CrossRef] [PubMed]
  149. Combs, M.A.; Golnar, A.J.; Overcash, J.M.; Lloyd, A.L.; Hayes, K.R.; O’Brochta, D.A.; Pepin, K.M. Leveraging eco-evolutionary models for gene drive risk assessment. Trends Genet. 2023, 39, 609–623. [Google Scholar] [CrossRef] [PubMed]
  150. Esvelt, K.M.; Smidler, A.L.; Catteruccia, F.; Church, G.M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife 2014, 3, 3277–3288. [Google Scholar] [CrossRef] [PubMed]
  151. Akbari, O.S.; Bellen, H.J.; Bier, E.; Bullock, S.L.; Burt, A.; Church, G.M.; Cook, K.R.; Duchek, P.; Edwards, O.R.; Esvelt, K.M. Safeguarding gene drive experiments in the laboratory. Science 2015, 349, aac7932. [Google Scholar] [CrossRef] [PubMed]
  152. Ellis, D.A.; Avraam, G.; Hoermann, A.; Wyer, C.A.; Ong, Y.X.; Christophides, G.K.; Windbichler, N. Testing non-autonomous antimalarial gene drive effectors using self-eliminating drivers in the African mosquito vector Anopheles gambiae. PLoS Genet. 2022, 18, e1010244. [Google Scholar] [CrossRef] [PubMed]
  153. Oberhofer, G.; Ivy, T.; Hay, B.A. Split versions of Cleave and Rescue selfish genetic elements for measured self limiting gene drive. PLoS Genet. 2021, 17, e1009385. [Google Scholar] [CrossRef]
  154. Champer, J.; Chung, J.; Lee, Y.L.; Liu, C.; Yang, E.; Wen, Z.; Clark, A.G.; Messer, P.W. Molecular safeguarding of CRISPR gene drive experiments. eLife 2019, 8, e41439. [Google Scholar] [CrossRef]
  155. Xu, X.-R.S.; Bulger, E.A.; Gantz, V.M.; Klanseck, C.; Heimler, S.R.; Auradkar, A.; Bennett, J.B.; Miller, L.A.; Leahy, S.; Juste, S.S. Active genetic neutralizing elements for halting or deleting gene drives. Mol. Cell 2020, 80, 246–262.e4. [Google Scholar] [CrossRef]
  156. Taxiarchi, C.; Beaghton, A.; Don, N.I.; Kyrou, K.; Gribble, M.; Shittu, D.; Collins, S.P.; Beisel, C.L.; Galizi, R.; Crisanti, A. A genetically encoded anti-CRISPR protein constrains gene drive spread and prevents population suppression. Nat. Commun. 2021, 12, 3977. [Google Scholar] [CrossRef]
  157. Chae, D.; Lee, J.; Lee, N.; Park, K.; Moon, S.J.; Kim, H.H. Chemical controllable gene drive in Drosophila. ACS Synth. Biol. 2020, 9, 2362–2377. [Google Scholar] [CrossRef]
  158. D’Amato, R.; Taxiarchi, C.; Galardini, M.; Trusso, A.; Minuz, R.L.; Grilli, S.; Somerville, A.G.; Shittu, D.; Khalil, A.S.; Galizi, R. Anti-CRISPR Anopheles mosquitoes inhibit gene drive spread under challenging behavioural conditions in large cages. Nat. Commun. 2024, 15, 952. [Google Scholar] [CrossRef] [PubMed]
  159. Trump, B.D.; Cummings, C.L.; Kuzma, J.; Linkov, I. Synthetic Biology 2020: Frontiers in Risk Analysis and Governance; Springer Nature: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  160. Macias, V.; James, A.A. Impact of genetic modification of vector populations on the malaria eradication agenda. In Genetic Control of Malaria and Dengue; Elsevier Academic Press: London, UK, 2015; pp. 423–444. [Google Scholar]
  161. Alphey, L. Genetic control of mosquitoes. Annu. Rev. Entomol. 2014, 33, 59. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Gene drives manipulate inheritance to rapidly spread altered genes through populations. In Mendelian inheritance (A), offspring have a 50% chance of inheriting an altered gene (shown in blue), when an individual carrying the altered gene on only one of two homologous chromosomes is crossed with a wild-type (WT) individual (shown in red). Under gene drive inheritance (B), crossing a trans-heterozygous organism with a WT organism will result in the gene drive system converting WT chromosomes to the altered gene in the germline cells, if the process is confined to the germline. As a result, offspring will inherit the altered gene 100% of the time. This leads to the rapid replacement of WT organisms with a specific genotype over fewer generations. The spread of a selfish gene through a population depends on the gene drive system, the organisms, and the ecological conditions. In a diploid organism (C) with a site-specific HEG (homing endonuclease gene) or CRISPR/Cas (shown in blue) integrated into one homologous chromosome, the endonuclease causes a DNA break in the wild-type chromosome. Homology-dependent repair then copies the modified allele into the wild-type chromosome. This process, called homing, results in HEG- or CRISPR-Cas-bearing gametes. The Medea (maternal effect dominant embryonic arrest) system (D) uses a maternal toxin and a zygotic antidote to drive gene spread. Heterozygous Medea females crossed with WT or heterozygous Medea males produce only viable embryos that inherit the Medea element. When Medea-bearing males are crossed with WT females, all embryos survive because the maternal toxin is not expressed. This system creates a frequency-dependent fitness advantage by killing the offspring of Medea heterozygous mothers that did not inherit Medea, allowing the spread of a cargo gene. The CRISPR-Cas-based toxin-antidote system uses a maternal toxin and zygotic antidote to promote the survival of offspring with the toxin-antidote element, thereby spreading a cargo gene throughout the population. Lethality is indicated by a gray background, +: WT gamete; M: Medea-bearing gamete; M♀: heterozygous Medea female; M♂: heterozygous Medea male; +♀: WT female; +♂: WT male.
Figure 1. Gene drives manipulate inheritance to rapidly spread altered genes through populations. In Mendelian inheritance (A), offspring have a 50% chance of inheriting an altered gene (shown in blue), when an individual carrying the altered gene on only one of two homologous chromosomes is crossed with a wild-type (WT) individual (shown in red). Under gene drive inheritance (B), crossing a trans-heterozygous organism with a WT organism will result in the gene drive system converting WT chromosomes to the altered gene in the germline cells, if the process is confined to the germline. As a result, offspring will inherit the altered gene 100% of the time. This leads to the rapid replacement of WT organisms with a specific genotype over fewer generations. The spread of a selfish gene through a population depends on the gene drive system, the organisms, and the ecological conditions. In a diploid organism (C) with a site-specific HEG (homing endonuclease gene) or CRISPR/Cas (shown in blue) integrated into one homologous chromosome, the endonuclease causes a DNA break in the wild-type chromosome. Homology-dependent repair then copies the modified allele into the wild-type chromosome. This process, called homing, results in HEG- or CRISPR-Cas-bearing gametes. The Medea (maternal effect dominant embryonic arrest) system (D) uses a maternal toxin and a zygotic antidote to drive gene spread. Heterozygous Medea females crossed with WT or heterozygous Medea males produce only viable embryos that inherit the Medea element. When Medea-bearing males are crossed with WT females, all embryos survive because the maternal toxin is not expressed. This system creates a frequency-dependent fitness advantage by killing the offspring of Medea heterozygous mothers that did not inherit Medea, allowing the spread of a cargo gene. The CRISPR-Cas-based toxin-antidote system uses a maternal toxin and zygotic antidote to promote the survival of offspring with the toxin-antidote element, thereby spreading a cargo gene throughout the population. Lethality is indicated by a gray background, +: WT gamete; M: Medea-bearing gamete; M♀: heterozygous Medea female; M♂: heterozygous Medea male; +♀: WT female; +♂: WT male.
Insects 15 00653 g001
Figure 2. The concept involves utilizing a gene drive to manipulate insect pest behavior by targeting their sensory systems. Insects utilize their olfactory system to detect volatile pheromones and environmental odors, influencing behaviors such as food source localization, host selection, mating, oviposition, and predator avoidance. Their gustatory system plays a role in detecting nonvolatile phytochemicals, which impacts feeding, mating, and oviposition decisions. The application of a gene drive to manipulate these sensory mechanisms represents a promising avenue for sustainable insect pest management strategies.
Figure 2. The concept involves utilizing a gene drive to manipulate insect pest behavior by targeting their sensory systems. Insects utilize their olfactory system to detect volatile pheromones and environmental odors, influencing behaviors such as food source localization, host selection, mating, oviposition, and predator avoidance. Their gustatory system plays a role in detecting nonvolatile phytochemicals, which impacts feeding, mating, and oviposition decisions. The application of a gene drive to manipulate these sensory mechanisms represents a promising avenue for sustainable insect pest management strategies.
Insects 15 00653 g002
Figure 3. Schematic depiction of a non-split CRISPR-Cas-mediated gene drive mechanism. The construct comprises essential components: a Cas gene, sgRNA, effector genes, and flanking homology arms (illustrated in green). A fluorescent marker gene facilitates the identification of transgenic progeny. Upon insertion into a chromosome, independent expression of Cas and sgRNAs initiates a DSB on the homologous chromosome within germ cells. The inserted gene drive cassette remains intact due to disruption at the sgRNA recognition site. Repair via HDR results in duplication of the entire cassette into the wild-type homologous chromosome, leading to the presence of two gene drive cassettes in the genome, predominantly in sperm or eggs. Split gene drives, separating sgRNA from an unlinked Cas cassette to induce biallelic mutations, exhibit efficiency comparable to non-split constructs, maintaining the foundational gene drive principle.
Figure 3. Schematic depiction of a non-split CRISPR-Cas-mediated gene drive mechanism. The construct comprises essential components: a Cas gene, sgRNA, effector genes, and flanking homology arms (illustrated in green). A fluorescent marker gene facilitates the identification of transgenic progeny. Upon insertion into a chromosome, independent expression of Cas and sgRNAs initiates a DSB on the homologous chromosome within germ cells. The inserted gene drive cassette remains intact due to disruption at the sgRNA recognition site. Repair via HDR results in duplication of the entire cassette into the wild-type homologous chromosome, leading to the presence of two gene drive cassettes in the genome, predominantly in sperm or eggs. Split gene drives, separating sgRNA from an unlinked Cas cassette to induce biallelic mutations, exhibit efficiency comparable to non-split constructs, maintaining the foundational gene drive principle.
Insects 15 00653 g003
Table 1. Homing efficiencies of genetically edited populations with CRISPR-Cas-mediated gene drive systems.
Table 1. Homing efficiencies of genetically edited populations with CRISPR-Cas-mediated gene drive systems.
OrganismGene Drive ConstructCas9 CassettegRNA
Cassette
MarkerTarget GenePhenotypeHoming Rate (G1)Reference
D. melanogasterMCRvasa-Cas9U6:3-gRNAyellowyellow body color97%[33]
nanos-Cas9
vasa-Cas9
U6:3-gRNA3xP3-DsRedyellowyellow body color62%
52%
[52]
vasa-Cas9U6:3-gRNA3xP3-DsRedyellow
white
yellow body color
white-eye
37–53%
56%
[53]
nanos-Cas9U6:3-gRNA3xP3-DsRedwhite
white-2gRNA
cinnabar
cinnabar
yellow
white-eye
white-eye
brilliant orange eye
brilliant orange eye
yellow body color
59%
76%
54%
38%
40–62%
CHERcd-1r-Cas9U6:3-gRNA3xP3-DsRedtransformersex conversion from females to males56%[54]
ClvRnanos-Cas9U6:3-gRNA3xP3-GFPtkolethality99%[37]
tGDvasa-Cas9U6:3-gRNA3xP3-DsRed
3xP3-EGFP
yellow
ebony
white
yellow body color
dark body color
white-eye
67–98%[55]
ClvRnanos-Cas9U6:3-gRNA3xP3-GFP
opie-td-tomato
dbe
TfIIA-S
recessive lethal>99% in ♀;
>94.7 to >99% in ♂
[42]
TAREnanos-Cas9U6:3-gRNA3xP3-DsRed
3xP3-GFP
hairyrecessive lethal88–95% in ♀[43]
nanos-Cas9U6:3-gRNA3xP3-DsRed
3xP3-GFP
RpL35Ahaplolethal91%[44]
sGDnanos-Cas9
vasa-Cas9
U6:3-gRNA3xP3-tdTomato
3XP3-eGFP
Opie2-DsRed
rab5
rab11
spo11
prosalpha2
recessive lethal
recessive lethal
sterility
recessive lethal
64.8–99.9%[45]
HomeRnanos-Cas9U6:3-gRNA3xP3-EGFPPolG2lethality99.6% in ♀;
75.0% in ♂
[56]
nanos-Cas9U6:3-gRNA3xP3-DsRedyellow-grecessive female sterility86.4% in ♀;
90.4% in ♂
[57]
rcd-1r-Cas9U6:3-gRNA3xP3-CFPrcd-1rmale fertility77.1% in ♂;
80.5% in ♀
[46]
TARE
TADE
nanos-Cas9U6:3-gRNA3xP3-DsRedRpL35A
hairy
haplolethal
recessive lethal
51–54%[58]
vasa-Cas12aU6:3-gRNAOpie2-GFPebonydark body color52–89%[47]
nanos-Cas9
CG4415-Cas9
rcd-1r-Cas9
U6:3-gRNA3xP3-DsRed
3xP3-GFP
doublesexdominant female sterility70.0–85.5%[59]
Drosophila suzukiinanos-Cas9U6:3-gRNApUb-DsRed
hsp83-ZsGreen
doublesexdominant female sterility94–99%[39]
DsTdc2CRISPRvasa-Cas9U6:3-gRNAPUb-DsRedTyrosine decarboxylase 2recessive female fertility53.6–58.2%[60]
Anopheles stephensiAsMCRkh2vasa-Cas9U6-gRNA3xP3-GFPkynurenine hydroxylasewhitewhite-eyeG1: 99.5%
G3: 98.8%
G4: 97.2%
[34]
Anopheles gambiaeCRISPRhvasa2-Cas9U6-gRNA3xP3-RFPAGAP007280
AGAP005958
AGAP011377
sterilityG2: 99%, G3: 97.6%
G2: 95.8%, G3: 92.8%
G2: 82.8%, G3: 75.5%,
[35]
Anopheles gambiaedsxFCRISPRhzpg-Cas9U6-gRNA3xP3-RFPdoublesexsterility in homozygous females95.9% in ♂;
99.4% in ♀
[36]
Anopheles gambiaeSDGDdsxzpg-Cas9U6-gRNA3xP3-DsReddoublesex
X chromosome
distort sex ratios (male only)92% in ♂;
99% in ♀
[61]
Anopheles gambiaeAgNosCd-1nanos-Cas9U6-gRNA3xP3-CFPAgcdred-eye96.7% (G1–G4)[48]
Anopheles stephensiReckhvasa-Cas9U6A-gRNA3xP3-GFPkynurenine hydroxylasewhitewhite-eye99.8% in ♂;
57% in ♀
[49]
Anopheles gambiaenos-CRISPRh
zpg-CRISPRh
exu-CRISPRh
zpg-Cas9
nanos-Cas9
exu-Cas9
U6-gRNA3xP3-DsRedAGAP007280recessive female sterility93.5% in ♂, 97.8% in ♀;
99.6% in ♂, 99.1% in ♀;
65.0% in ♂; 0 in ♀
[50]
Ceratitis capitatavasa-Cas9U6-gRNApUb-DsRedtransformersex conversion from females to males83.1%[40]
Plutella xylostellavasa-Cas9
meiw68-Cas9
nanos-Cas9
U6-gRNAHr5ie1-DsRedyellow
kmo
yellow-pigmentation yellow-eyeno significant deviation from 50% inheritance[41]
Plutella xylostellananos-Cas9U6-gRNAHr5ie1-EGFPyellowyellow-pigmentation6.67–12.59%[51]
–: not available; ♀: female; ♂: male.
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

Zhao, Y.; Li, L.; Wei, L.; Wang, Y.; Han, Z. Advancements and Future Prospects of CRISPR-Cas-Based Population Replacement Strategies in Insect Pest Management. Insects 2024, 15, 653. https://doi.org/10.3390/insects15090653

AMA Style

Zhao Y, Li L, Wei L, Wang Y, Han Z. Advancements and Future Prospects of CRISPR-Cas-Based Population Replacement Strategies in Insect Pest Management. Insects. 2024; 15(9):653. https://doi.org/10.3390/insects15090653

Chicago/Turabian Style

Zhao, Yu, Longfeng Li, Liangzi Wei, Yifan Wang, and Zhilin Han. 2024. "Advancements and Future Prospects of CRISPR-Cas-Based Population Replacement Strategies in Insect Pest Management" Insects 15, no. 9: 653. https://doi.org/10.3390/insects15090653

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