[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (484)

Search Parameters:
Keywords = insecticidal proteins

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 930 KiB  
Review
Molecular Mechanisms Underlying Resistance to Bacillus thuringiensis Cry Toxins in Lepidopteran Pests: An Updated Research Perspective
by Deqin Hu, Dongmei Wang, Hongsheng Pan and Xiaoning Liu
Agronomy 2025, 15(1), 155; https://doi.org/10.3390/agronomy15010155 - 10 Jan 2025
Viewed by 219
Abstract
Genetically modified crops that produce insecticidal proteins from Bacillus thuringiensis (Bt) are currently the most efficient and safest method of pest control worldwide. However, the prolonged planting period has led to a reduction in the efficacy of Bt crops due to [...] Read more.
Genetically modified crops that produce insecticidal proteins from Bacillus thuringiensis (Bt) are currently the most efficient and safest method of pest control worldwide. However, the prolonged planting period has led to a reduction in the efficacy of Bt crops due to the evolution of pest resistance in the field. This review paper examines the resistance status of lepidopteran pests to Bt crops under field conditions, elucidates the molecular mechanism underlying their resistance to Bt Cry toxins, and discusses resistance management strategies based on these mechanisms. Extensive research has demonstrated that mutations and alterations in expression patterns of midgut receptor genes are closely associated with Bt resistance. As our understanding of molecular mechanisms progresses, several innovative approaches such as DNA molecular detection techniques, engineering modified Cry toxins, and combining Bt toxin with RNAi technology have been developed for effective pest control measures. Future research will further unravel the intricate molecular mechanisms underlying this phenomenon to develop scientifically sound integrated pest management strategies. Full article
(This article belongs to the Section Pest and Disease Management)
Show Figures

Figure 1

Figure 1
<p>Resistance mechanism of midgut receptor proteins mutation (refer to Tetreau [<a href="#B20-agronomy-15-00155" class="html-bibr">20</a>]). (<b>A</b>) The mechanism of action of Cry toxins in susceptive pest, when insects ingest crystals of toxin secreted by <span class="html-italic">Bt</span> crops, the alkaline environment in the insect’s digestive tract denatures and activates these insoluble crystals. The activated toxin then binds to receptors on the apical microvillous membrane of midgut epithelial cells, resulting in a conformational change of the toxin, which allows it to enter the cell membrane. The subsequent oligomerization of the toxins leads to osmotic cell lysis, ultimately resulting in the death of the target insects. (<b>B</b>) The phenomenon of base mutations, deletions or insertions in the genes encoding APN, ALP, CAD and ABC transporters in resistant strains, which resulted in structural alterations in receptor proteins or premature stop codon generation. Following mutation or truncation, proteins were subject to the loss of significant functional regions, which in turn resulted in a reduction in the binding ability of receptor proteins to Cry toxins or an inability to be fixed to the cell membrane. This ultimately contributes to the development of resistance.</p>
Full article ">Figure 2
<p>A proposed model for hormone-mediated MAPK signaling pathway in pest resistance The main mechanisms include m<sup>6</sup>A modification by <span class="html-italic">PxMettl3</span> and <span class="html-italic">PxMettl14</span>-downregulated PxJHE in the midgut of the resistant <span class="html-italic">P. xylostella</span>, resulting in the elevation of JH titer; a midgut miRNA (miR-8545) initiated epigenetic regulatory pathway repressed <span class="html-italic">PxGLD</span> activity and elevated the 20E titer; the crosstalk between insect hormones JH and 20E enhanced FOXO dephosphorylation and <span class="html-italic">MAP4K4</span> expression. More importantly, a short interspersed nuclear element (named SE2) insertion within the resistant <span class="html-italic">MAP4K4</span> promoter harbored a potent FRE element (FRE1) which could recruit the FOXO protein more efficiently to boost the constitutive overexpression of the <span class="html-italic">MAP4K4</span> gene, resulting in the downregulation of midgut Cry1Ac receptors, thus facilitating the evolutionary adaptation of <span class="html-italic">P. xylostella</span> to the <span class="html-italic">Bt</span> Cry1Ac toxin as a resistant phenotype.</p>
Full article ">
18 pages, 2638 KiB  
Article
Glycosylation Patterns in Meccus (Triatoma) pallidipennis Gut: Implications for the Development of Vector Control Strategies
by Elia Torres-Gutiérrez, Frida Noelly Candelas-Otero, Olivia Alicia Reynoso-Ducoing, Berenice González-Rete, Mauro Omar Vences-Blanco, Margarita Cabrera-Bravo, Martha Irene Bucio-Torres and Paz María Silvia Salazar-Schettino
Microorganisms 2025, 13(1), 58; https://doi.org/10.3390/microorganisms13010058 - 1 Jan 2025
Viewed by 530
Abstract
The primary mode of transmission for Chagas disease is vector-borne transmission, spread by hematophagous insects of the Triatominae subfamily. In Mexico, the triatomine Meccus pallidipennis is particularly significant in the transmission of Trypanosoma cruzi. This study focused on analyzing protein expression and [...] Read more.
The primary mode of transmission for Chagas disease is vector-borne transmission, spread by hematophagous insects of the Triatominae subfamily. In Mexico, the triatomine Meccus pallidipennis is particularly significant in the transmission of Trypanosoma cruzi. This study focused on analyzing protein expression and modifications by glycosylation in different regions of the digestive tract of fifth-instar nymphs of M. pallidipennis. Two gut sections were dissected and extracted: the anterior midgut (AMG) and the proctodeum or rectum (RE). Proteins were extracted from each tissue sample and profiled by one- and two-dimensional electrophoresis; protein glycosylation was analyzed by lectin affinity. Our results showed significant differences in protein expression and glycosylation between both gut regions, with modifications being more frequent in the RE. The proteins HSP70, actin, and tubulin were analyzed, finding a differential expression of the latter two between AMG and RE. Understanding glycosylation patterns provides critical insights into vector–pathogen interactions that could eventually inform novel control approaches. Furthermore, the potential use of lectins as insecticidal agents highlights the broader implications of glycoprotein research in the future development of strategies on vector control to disrupt T. cruzi transmission. Full article
(This article belongs to the Special Issue Vector-Borne Zoonoses: Surveillance, Transmission and Interventions)
Show Figures

Figure 1

Figure 1
<p>Protein profiles and protein maps of the anterior midgut (AMG) and rectum (RE) of unfed, fifth-instar nymphs of <span class="html-italic">M. pallidipennis</span>. (<b>A</b>) Schematic of regions excised from <span class="html-italic">M. pallidipennis</span> gut. (<b>B</b>) Protein profiles of the AMG and RE on SDS-PAGE; MW = molecular weight, kDa = kilodaltons. Total bands identified in each gut section are shown at the bottom. (<b>C</b>) Master proteomic map of the AMG. (<b>D</b>) Master proteomic map of the RE. Red circles indicate common points in both maps. The total number of detected points is shown at the bottom of each map. (<b>E</b>) Distribution of unique points from each gut region by molecular weight (kDa); <span class="html-italic">p</span> &lt; 0.0001. (<b>F</b>). Distribution of unique points from each gut region by isoelectric point; <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Glycosylation patterns of α-mannose and α-glucose detected with concanavalin A (ConA) in the anterior midgut (AMG) and rectum (RE) of unfed, fifth-instar nymphs of <span class="html-italic">M. pallidipennis</span>. (<b>A</b>) Schematic of regions excised from <span class="html-italic">M. pallidipennis</span> gut for glycan detection with ConA. (<b>B</b>) Glycosylation profiles of the AMG and RE by 1D lectin blot. The total number of glycosylation bands identified in each gut section is shown at the bottom. (<b>C</b>) Master map of affinity to ConA in the AMG. (<b>D</b>) Master map of affinity to ConA in the RE. MW = molecular weight, kDa = kilodaltons. Red circles indicate common glycosylated spots in both samples. The total number of spots identified in each gut region is shown at the bottom of each map.</p>
Full article ">Figure 3
<p>N-acetylglucosamine and sialic acid glycosylation patterns detected by WGA in the anterior midgut (AMG) and rectum (RE) of unfed, fifth-instar nymphs of <span class="html-italic">M. pallidipennis</span>. (<b>A</b>) Schematic of regions extracted from <span class="html-italic">M. pallidipennis</span> gut for glycan detection with WGA. (<b>B</b>) Glycosylation profiles of AMG and RE by 1D lectin blot. The total number of glycosylation bands identified in each section of the intestine is shown at the bottom. (<b>C</b>) Master map of affinity to WGA in the AMG. (<b>D</b>) Master map of affinity to WGA in the RE. MW = molecular weight, kDa = kilodaltons. Red circles indicate common glycosylated spots in both samples. The total number of spots identified in each gut region is shown at the bottom of each map.</p>
Full article ">Figure 4
<p>Glycosylation patterns of N-acetylgalactosamine and β-galactose by peanut agglutinin (PNA) in the anterior midgut (AMG) and rectum (RE) of unfed, fifth-instar nymphs of <span class="html-italic">M. pallidipennis</span>. (<b>A</b>) Schematic of regions extracted from <span class="html-italic">M. pallidipennis</span> gut for glycan detection with PNA. (<b>B</b>) Glycosylation profiles of the AMG and RE by 1D lectin blot. The total number of glycosylation bands identified in each section of the gut is shown at the bottom. (<b>C</b>) Master map of PNA affinity in the AMG. (<b>D</b>) Master map of affinity for PNA in the RE. MW = molecular weight, kDa = kilodaltons. Red circles indicate common glycosylated spots in both samples. The total number of spots identified in each gut region is shown at the bottom of each map.</p>
Full article ">Figure 5
<p>Differential expression patterns of protein and glycosylation spots in the AMG and RE, and spots in common (AMG/RE). Protein spots identified in the 2D electrophoresis master maps and glycosylation spots identified in the affinity master maps for the lectins ConA, WGA, and PNA were grouped and quantified by defined isoelectric point (X-axis) and molecular weight (Y-axis) sectors and plotted on a heat map. Red and blue shades represent higher and lower levels of protein expression and glycosylation, respectively.</p>
Full article ">Figure 6
<p>Relative expression of proteins detected in the anterior midgut (AMG) and rectum (RE) of unfed, fifth-instar nymphs of <span class="html-italic">M. pallidipennis</span>. (<b>A</b>) Immunodetection of the proteins HSP70 (BRM-22), α-tubulin (DM1A), and β-actin (RM112) by Western blot. MW = molecular weight, kDa = kilodalton. (<b>B</b>) Histogram of the semiquantitative evaluation of reaction intensity detected by Western blot (Image Lab, Bio-Rad).</p>
Full article ">
15 pages, 2775 KiB  
Article
Bacillus thuringiensis Cry1A Insecticidal Toxins and Their Digests Do Not Stimulate Histamine Release from Cultured Rat Mast Cells
by Hisashi Ohto, Mayumi Ohno, Miho Suganuma-Katagiri, Takashi Hara, Yoko Egawa, Kazuya Tomimoto, Kosuke Haginoya, Hidetaka Hori, Yuzuri Iwamoto and Tohru Hayakawa
Biology 2025, 14(1), 15; https://doi.org/10.3390/biology14010015 - 27 Dec 2024
Viewed by 443
Abstract
Public acceptance of genetically modified crops engineered with Bacillus thuringiensis (Bt) insecticidal protein genes (BT-GMCs), which confer resistance to various lepidopteran insect pests, is generally lacking. As a major concern over BT-GMCs is the allergenicity of insecticidal proteins, alleviating safety concerns should help [...] Read more.
Public acceptance of genetically modified crops engineered with Bacillus thuringiensis (Bt) insecticidal protein genes (BT-GMCs), which confer resistance to various lepidopteran insect pests, is generally lacking. As a major concern over BT-GMCs is the allergenicity of insecticidal proteins, alleviating safety concerns should help increase public acceptance. In this study, three lepidopteran-specific Bt toxins, Cry1Aa, Cy1Ab, and Cry1Ac, were treated with simulated digestive fluids under various conditions. Western blotting using antiserum raised against individual segments (α-helices of domain I and β-sheets of domains II and III) of Cry1Aa showed that digestion produces a variety of polypeptides. In particular, the transmembrane α4–α5 of domain I, which may retain the ability to form pores, was the most resistant to digestion. Intact Cry1A toxins and these digests were then applied to RBL-2H3 cultured rat mast cells to determine whether the toxins directly induce histamine release. However, fluorescence microscopy revealed no specific binding of Cry1A toxins to RBL-2H3 cultured rat mast cells. In addition, neither the OPA method nor HPLC analysis detected significant histamine release from mast cells treated with Cry1A toxins and these digests. Our results provide important data supporting the safety of Cry1A toxins and potentially BT-GMCs. Full article
(This article belongs to the Section Biotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Digestion of Cry1A toxins using SGF at varying pH values. A total of 100 μg of each Cry1A toxin was digested with SGF in glycine buffer at pH 2, 3, and 4 or citrate buffer at pH 3, 4, 5, and 6. The resulting digests were separated by 14% SDS-PAGE and visualized by CBB staining. U: undigested Cry1A toxin.</p>
Full article ">Figure 2
<p>Western blotting analysis of Cry1Aa digests. Cry1Aa digests generated by SGF treatment were separated by 14% SDS-PAGE and then electroblotted onto PVDF membranes. Cry1Aa digests on the membrane were analyzed by Western blotting with six different antisera specific for α2–3, α4–5, and α6–7 helices of domain I, β1-5 and β6-11 sheets of domain II, and domain III.</p>
Full article ">Figure 3
<p>Western blotting analysis of Cry1A digests. Cry1A digests generated by SGF treatment (pH 2 and 4) and digests generated by subsequent SIF treatment (pH 8) were separated by 14% SDS-PAGE and then electroblotted onto PVDF membranes. Cry1A digests were analyzed by Western blotting using anti-α4–α5 antiserum. U: undigested Cry1A toxin. (<b>A</b>) SGF digestion at pH 2. (<b>B</b>) SGF digestion at pH 4.</p>
Full article ">Figure 4
<p>Treatment of rat mast cells with Cry1A toxins. RBL-2H3 rat mast cells were incubated with labeled Cry1A toxins at 10 and 100 nM for 60 min. BSA was used as a control. (<b>A</b>) Microscopic observation of RBL-2H3 rat mast cells. Top: Fluorescence of labeled proteins on cells after treatment with toxin at 100 nM. Bottom: Observation of cells under visual light. Bar, 20 μm. (<b>B</b>) Amount of each cell-bound Cry1A toxin, as measured using a fluorescence spectrometer. Standard deviation was calculated from triplicate experiments.</p>
Full article ">Figure 5
<p>Histamine release from rat mast cells following treatment with Cry1A toxins. (<b>A</b>) RBL-2H3 rat mast cells were incubated with undigested Cry1A toxin, and histamine released was quantified using the OPA method. The calcium ionophore A23187 and BSA were used as positive and negative controls, respectively. (<b>B</b>) Stimulation of histamine release by Cry1A polypeptides digested with SGF. (<b>C</b>) Stimulation of histamine release by Cry1A polypeptides digested sequentially with SGF and SIF.</p>
Full article ">Figure 6
<p>HPLC analysis of histamine released from rat mast cells treated with Cry1Aa digests. Histamine released from rat mast cells was analyzed by HPLC. (<b>A</b>) Authentic histamine. A single peak was detected at an RT of approximately 19.6 min. (<b>B</b>) Hanks’s buffer. No clearly discernible peak was detected. (<b>C</b>) RBL-2H3 rat mast cells. Cells were sonicated in Hanks’s buffer. (<b>D</b>) Histamine released from RBL-2H3 rat mast cells treated with ionophore A23187. (<b>E</b>) Histamine released from untreated RBL-2H3 rat mast cells. (<b>F</b>) Histamine released from RBL-2H3 rat mast cells treated with Cry1Aa digests obtained from SGF treatment.</p>
Full article ">
17 pages, 1342 KiB  
Article
Field Performance of a Genetically Modified Cowpea (Vigna unguiculata) Expressing the Cry1Ab Insecticidal Protein Against the Legume Pod Borer Maruca vitrata
by Jerry A. Nboyine, Gloria A. Adazebra, Emmanuel Y. Owusu, Philip Agrengsore, Ahmed Seidu, Salim Lamini, Mukhtaru Zakaria, James Y. Kwabena, Haruna K. Ali, Ijeoma Akaogu, Francis N. Onyekachi, Jean B. Tignegre, Prince M. Etwire, Donald J. MacKenzie, Jose M. Barrero and Thomas J. V. Higgins
Agronomy 2024, 14(12), 3055; https://doi.org/10.3390/agronomy14123055 - 21 Dec 2024
Viewed by 566
Abstract
Cowpea (Vigna unguiculata) is a vital crop in sub-Saharan Africa, but the legume pod borer (LPB), Maruca vitrata, can cause over 80% yield losses. Natural resistance to this lepidopteran pest is absent in cowpea germplasm, and insecticides are ineffective due [...] Read more.
Cowpea (Vigna unguiculata) is a vital crop in sub-Saharan Africa, but the legume pod borer (LPB), Maruca vitrata, can cause over 80% yield losses. Natural resistance to this lepidopteran pest is absent in cowpea germplasm, and insecticides are ineffective due to the pest’s cryptic behavior. To address this, a genetically modified (GM) cowpea expressing the cry1Ab protein from Bacillus thuringiensis (Bt) was developed, providing complete LPB resistance. This Bt cowpea, commercialized as Sampea 20-T in Nigeria, was recently approved in Ghana as Songotra T. To evaluate its performance and the financial returns of its cultivation, field trials were conducted across multiple locations in northern Ghana to compare it to the non-transgenic Songotra control and two commercial cultivars, Kirkhouse-Benga and Wang-Kae. Songotra T exhibited protection against LPB infestations and damage, achieving a grain yield of 2534 kg/ha compared to 1414–1757 kg/ha for the other entries. As expected, non-LPB pest infestations and damage were similar across all entries. Economic analysis revealed that Songotra T had the highest return on investment (464%), outperforming the other tested cultivars (214%). These results demonstrate the potential of GM crops to enhance yields and profitability for resource-poor farmers, underscoring the value of biotechnology for addressing critical agricultural challenges. Full article
(This article belongs to the Section Pest and Disease Management)
Show Figures

Figure 1

Figure 1
<p>Map of northern Ghana showing the spatial distribution of sites for field testing of the performance of test entries. Note: Akukayili is in the Tolon District of the Northern Region; Chinchang is in the Sissala East District of the Upper West Region; Kpasenkpe is in the West Mamprusi Municipality of the North East Region; and Settlement is in the West Gonja Municipality of the Savannah Region.</p>
Full article ">Figure 2
<p>Grain yield (kg/ha) of different cowpea entries averaged across four locations in northern Ghana. Note: bars are means ± standard error of means; bars followed by different letters are significantly different at a 5% probability threshold.</p>
Full article ">Figure 3
<p>Return on investment of cowpea varieties grown in northern Ghana. Note: return on investment for “all conventional varieties” was estimated using average yields of the crop in farmers’ fields in Ghana [<a href="#B30-agronomy-14-03055" class="html-bibr">30</a>].</p>
Full article ">
18 pages, 2411 KiB  
Article
Acephate Exposure Induces Transgenerational Ovarian Developmental Toxicity by Altering the Expression of Follicular Growth Markers in Female Rats
by Abeer Alhazmi, Saber Nahdi, Saleh Alwasel and Abdel Halim Harrath
Biology 2024, 13(12), 1075; https://doi.org/10.3390/biology13121075 - 20 Dec 2024
Viewed by 424
Abstract
Acephate is an organophosphate foliar and soil insecticide that is used worldwide. In this study, the transgenerational ovarian developmental toxicity caused by acephate, along with its in utero reprogramming mechanisms, were explored. Thirty female virgin Wistar albino rats were randomly assigned to three [...] Read more.
Acephate is an organophosphate foliar and soil insecticide that is used worldwide. In this study, the transgenerational ovarian developmental toxicity caused by acephate, along with its in utero reprogramming mechanisms, were explored. Thirty female virgin Wistar albino rats were randomly assigned to three groups: one control group and two acephate treatment groups. The treatment groups received daily low or high doses of acephate (34.2 mg/kg or 68.5 mg/kg body weight, respectively) from gestational day 6 until spontaneous labor, resulting in F1 offspring. At 28 days, a subgroup of F1 females were euthanized. The ovaries were extracted, thoroughly cleaned, and weighed before being fixed for further analysis. The remaining F1 females were mated with normal males to produce the F2 generation. The F1 female offspring presented reduced fertility and body weight, whereas the ovarian weight index and sex ratio increased in a dose-dependent manner. Structural analysis revealed altered follicular abnormalities with ovarian cells displaying pyknotic nuclei. Additionally, the gene and protein expression of Cyp19 decreased, whereas that of Gdf-9 increased in the high-dose treatment group (68.5 mg/kg). We also observed significantly increased expression levels of ovarian estrogen receptor 1 (Esr1) and insulin-like growth factor 1 (Igf1), whereas Insl3 expression was significantly decreased. The F2 female offspring presented reproductive phenotype alterations similar to those of F1 females including decreased fertility, reduced Cyp19 gene and protein expression, and structural ovarian abnormalities similar to those of polycystic ovary syndrome (PCOS). In conclusion, acephate induced ovarian developmental toxicity across two generations of rats, which may be linked to changes in the ovarian Cyp19, Gdf9, Insl3, and Igf1 levels. Full article
(This article belongs to the Section Developmental and Reproductive Biology)
Show Figures

Figure 1

Figure 1
<p>Body weight measurements of both the F1 (<b>A</b>) and F2 (<b>B</b>) treatment groups were compared with those in the control group. In the first generation, rats treated with either low-dose (34.2 mg/kg) or high-dose (68.5 mg/kg) acephate presented a significant reduction in body weight, whereas the second generation presented no significant changes. Compared with that in the control group, the sex ratio significantly increased in the first generation F1 (<b>C</b>). However, in the F2 generation (D), the sex ratio of the low-dose (34.2 mg/kg) treated females did not differ significantly from that of the control group, although the high-dose (68.5 mg/kg) treatment group presented an increased sex ratio compared with the control group. (*) <span class="html-italic">p</span> value &lt; 0.05, (**) <span class="html-italic">p</span> value &lt; 0.01, (****) <span class="html-italic">p</span> value &lt; 0.0001), ns: non signficant.</p>
Full article ">Figure 2
<p>Ovarian indices of the F1 (<b>A</b>) and F2 (<b>B</b>) treatment groups compared with those of the control group. In the first generation, rats in the low-dose (34.2 mg/kg) and high-dose (68.5 mg/kg) acephate treatment groups presented a significant increase in the ovarian index. However, no significant difference was found in the second generation. In the first generation (<b>C</b>), the number of offspring per female significantly decreased compared with that in the control group. In contrast, in the second generation (<b>D</b>), no significant difference in fertility rates was observed between the low-dose (34.2 mg/kg) group and the control group. In contrast, the high-dose acephate group (68.5 mg/kg) presented a lower fertility rate than the control group. Statistical significance was considered for all <span class="html-italic">p</span> values (* <span class="html-italic">p</span> &lt; 0.05, **<span class="html-italic">p</span>&lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> value &lt; 0.0001), ns: non significant.</p>
Full article ">Figure 3
<p>Photomicrographs of the H&amp;E-stained ovarian sections. Panels (<b>A</b>–<b>C</b>) depict the control group, showing a typical ovarian histological structure with developing follicles. In contrast, panels (<b>D</b>–<b>I</b>) illustrate the ovaries of first-generation female rats exposed to acephate prenatally. A notable characteristic of these ovaries was the increased presence of degenerating follicles compared with those in the control group. These degenerated follicles typically exhibited abnormal oocytes and a reduced number of growing follicles. (<b>G</b>) In addition to the large number of degenerating follicles, there were ovarian cysts (arrowhead in (<b>G</b>)). (<b>H</b>,<b>I</b>) There was an appearance of many vacuoles localized around the oocytes, an unusual zona pellucida, and many pyknotic granulosa cell nuclei (arrowhead) in (<b>I</b>)). DF: degenerative follicle; GC: granulosa cell; OC: oocyte; PMF: primordial follicle; PM: primary follicle; SF: secondary follicle; C: cysts; GF: Graafian follicles. Scale bar = 20 µm.</p>
Full article ">Figure 4
<p>Microscopy images of the H&amp;E-stained ovary tissue sections from second-generation female rats that were exposed (<b>A</b>–<b>I</b>). (<b>A</b>–<b>F</b>) Acephate treatment at 34.2 mg/kg resulted in many degenerating follicles, increased ovarian cysts, and many vacuoles (arrowheads in <b>E</b>). In addition, the micronucleus showed multioocyte formation (arrowhead in <b>F</b>). At a dose of 68.5 mg/kg, several corpus luteum, degenerating follicles (arrowheads in <b>H</b>)., and elevated GC pyknotic nuclei were detected (arrowheads in <b>I</b>). DF: degenerating follicle; OC: oocyte; SF: secondary follicle; TF: tertiary follicle; GF: Graafian follicles; C: cysts; CL: corpus luteum; MOF: multi oocyte formation; MN: micronucleus. Scale bar = 20 µm.</p>
Full article ">Figure 5
<p>The localization of ovarian CYP19 in first-generation females (F1) was evaluated using immunofluorescence staining and visualized with a confocal microscope. (<b>A</b>–<b>C</b>) Control ovaries, (<b>D</b>–<b>F</b>) F1 ovarian acephate-treated ovaries (34.2 mg/kg), and (<b>G</b>–<b>I</b>) F1 ovarian acephate-treated ovaries (68.5 mg/kg). The comparative fluorescence intensity (<b>J</b>) of CYP19 in the control and treated groups was analyzed via Zen 3.1 software (ZEN lite) and quantified via GraphPad Prism 10 (version 10.1.2). Additionally, the mRNA levels of <span class="html-italic">Cyp19</span> in the F1 acephate-treated groups were compared with those in the control groups via qRT–PCR (<b>K</b>). (*) <span class="html-italic">p</span> value &lt; 0.05, (**) <span class="html-italic">p</span> value &lt; 0.01, (***) <span class="html-italic">p</span> value &lt; 0.001, ns: non-significant.</p>
Full article ">Figure 6
<p>Localization and expression of CYP19 in ovarian tissue from second-generation females (F2). (<b>A</b>–<b>C</b>) Control ovaries, (<b>D</b>–<b>F</b>) F1 ovarian acephate-treated ovaries (34.2 mg/kg), and (<b>G</b>–<b>I</b>) F1 ovarian acephate-treated ovaries (68.5 mg/kg). Comparison of the fluorescence intensity of CYP19 in the control and exposure groups (<b>J</b>). The mRNA levels of <span class="html-italic">Cyp19</span> in the F2 acephate-treated groups were compared with those in the control groups via qRT–PCR (<b>K</b>) (**), <span class="html-italic">p</span> value &lt; 0.01, ns: non-significant.</p>
Full article ">Figure 7
<p>The localization of GDF9 in the ovarian tissue of first-generation females (F1) was assessed through immunofluorescence staining and visualized with a confocal microscope. (<b>A</b>–<b>C</b>) Ovaries from the control group, (<b>D</b>–<b>F</b>) F1 acephate-treated ovaries (34.2 mg/kg), (<b>G</b>–<b>I</b>) F1-treated acephate ovaries (68.5 mg/kg). Comparison of the fluorescence intensity of GDF9 in the control group and treatment groups (<b>J</b>). RT–PCR analyses of the mRNA levels of <span class="html-italic">Gdf9</span> in the first-generation acephate-treated groups compared with those in the control groups (<b>K</b>). (*) <span class="html-italic">p</span> value &lt; 0.05, (**) <span class="html-italic">p</span> value &lt; 0.01, (***) <span class="html-italic">p</span> value &lt; 0.001, (****) <span class="html-italic">p</span> value &lt; 0.0001, ns: non-significant.</p>
Full article ">Figure 8
<p>Illustration of the localization of GDF9 in ovarian tissue from females (F2), as assessed through immunofluorescence staining and visualized with a confocal microscope. (<b>A</b>–<b>C</b>) Ovaries from the control, (<b>D</b>–<b>F</b>) ovaries from 34.2 mg/kg acephate-treated F1, and (<b>G</b>–<b>I</b>) ovaries from first-generation females treated with 68.5 mg/kg acephate. Relative fluorescence intensity (<b>J</b>) of GDF9 in both the control and treatment groups. RT–PCR analyses of the <span class="html-italic">Gdf9</span> mRNA levels in the second-generation acephate-treated groups compared with those in the control groups (<b>K</b>). The scale bar represents 200 µm. Statistical significance is indicated by a (*) <span class="html-italic">p</span> value &lt; 0.05, (**) <span class="html-italic">p</span> value &lt; 0.01, (****) <span class="html-italic">p</span> value ≤ 0.0001, whereas ns denotes not significant.</p>
Full article ">Figure 9
<p>RT–PCR analyses of several genes related to folliculogenesis and steroidogenesis in the ovaries of rats from various treatment groups were performed, and the results were compared with those of the control group across multiple generations. (<b>A</b>) Changes in <span class="html-italic">Esr1</span> mRNA expression in the first generation. (<b>B</b>) Changes in <span class="html-italic">Esr1</span> mRNA expression in the second generation. (<b>C</b>) Changes in <span class="html-italic">Esr2</span> mRNA expression in the first generation. (<b>D</b>) Changes in <span class="html-italic">Esr2</span> mRNA expression in the second generation. (<b>E</b>) Changes in <span class="html-italic">Insl3</span> mRNA expression in the first generation. (<b>F</b>) Changes in <span class="html-italic">Insl3</span> mRNA expression in the second generation. (<b>G</b>) Changes in <span class="html-italic">Igf1</span> mRNA expression in the first generation. (<b>H</b>) Changes in <span class="html-italic">Igf1</span> mRNA expression in the second generation. (*) <span class="html-italic">p</span> value &lt; 0.05, (**) <span class="html-italic">p</span> value &lt; 0.01, (***) <span class="html-italic">p</span> value &lt; 0.001, (****) <span class="html-italic">p</span> value &lt; 0.0001, ns: Not significant.</p>
Full article ">
13 pages, 3992 KiB  
Article
Utilizing the Fungal Bicistronic System for Multi-Gene Expression to Generate Insect-Resistant and Herbicide-Tolerant Maize
by Yuxiao Chen, Wenjie Lv, Qun Yue, Ning Wen, Yinxiao Wang, Zhihong Lang, Wei Xu and Shengyan Li
Int. J. Mol. Sci. 2024, 25(24), 13408; https://doi.org/10.3390/ijms252413408 - 14 Dec 2024
Viewed by 432
Abstract
Developing simple and efficient multi-gene expression systems is crucial for multi-trait improvement or bioproduction in transgenic plants. In previous research, an IGG6-based bicistronic system from the nonpathogenic fungus Glarea lozoyensis efficiently expressed multiple enzyme proteins in yeast and maize, and the heterologous [...] Read more.
Developing simple and efficient multi-gene expression systems is crucial for multi-trait improvement or bioproduction in transgenic plants. In previous research, an IGG6-based bicistronic system from the nonpathogenic fungus Glarea lozoyensis efficiently expressed multiple enzyme proteins in yeast and maize, and the heterologous enzymes successfully performed their catalytic activity to reconstruct the biosynthetic pathway in the host organism. Unlike enzyme proteins, some heterologous functional proteins (such as insecticidal proteins) are dose-dependent and they need to express sufficient levels to perform their biological functions. It remains unclear whether the IGG6-based bicistronic system can achieve high expression of the functional proteins for practical applications in crops. In this study, two Bacillus thuringiensis (Bt) insecticidal genes, vip3Aa and cry1Ab, were linked via IGG6 to form a bicistron, while two glyphosate resistance genes, gr79epsps and gat, served as monocistronic selectable marker genes. Regenerated maize plants were produced through genetic transformation. RNA and immunoblot analyses revealed that the vip3Aa-IGG6-cry1Ab bicistron was transcribed as a single transcript, which was then translated into two separate proteins. Notably, the transcription and translation of cry1Ab were significantly positively correlated with those of vip3Aa. Through ELISA and leaf bioassay, we identified two transgenic maize lines, VICGG-15 and VICGG-20, that exhibited high insecticidal activity against fall armyworm (FAW; Spodoptera frugiperda) and Asian corn borer (ACB; Ostrinia furnacalis), both of which had high expression of Vip3Aa and Cry1Ab proteins. Subsequent evaluations, including silk, ear, and field bioassays, as well as glyphosate tolerance assessments, indicated that the VICGG-15 plants displayed high resistance to FAW and ACB, and could tolerate up to 3600 g acid equivalent (a.e.) glyphosate per hectare without adversely affecting phenotype or yield. Our finding established that the IGG6-based bicistronic system can achieve high expression of functional proteins in maize, and it is a potential candidate for multi-gene assembly and expression in plants. Full article
(This article belongs to the Special Issue New Insights into Plants and Insects Interactions)
Show Figures

Figure 1

Figure 1
<p>PCR screening of regenerated maize plants. (<b>A</b>) Schematic diagram of the maize genetic transformation constructs VICGG. The glyphosate resistance genes <span class="html-italic">gat</span> and <span class="html-italic">gr79-epsps</span> were used as selected marker genes for maize transformation. The black arrow represents the primer position. (<b>B</b>) PCR analysis of the <span class="html-italic">vip3Aa</span>, <span class="html-italic">cry1Ab</span>, <span class="html-italic">gr79,</span> and <span class="html-italic">gat</span> genes in corresponding transgenic maize plants.</p>
Full article ">Figure 2
<p>Correlation between <span class="html-italic">cry1Ab</span> and <span class="html-italic">vip3Aa</span>, as well as <span class="html-italic">gat</span> and <span class="html-italic">gr79</span>, at the transcription and translation levels in the 18 VICGG transgenic maize lines. (<b>A</b>,<b>B</b>) Correlation between <span class="html-italic">cry1Ab</span> and <span class="html-italic">vip3Aa</span> at the transcription and translation levels. (<b>C</b>,<b>D</b>) Correlation between <span class="html-italic">gat</span> and <span class="html-italic">gr79</span> at the transcription and translation levels. All data were presented as the mean of three biological replicates. Pearson’s correlation coefficients and their statistical significance were determined using GraphPad prism.</p>
Full article ">Figure 3
<p>RT-PCR and immunoblot analysis of the transcript and protein forms of <span class="html-italic">IGG6</span>-mediated bicistron in VICGG plants. (<b>A</b>) RT-PCR analysis of the transcript forms of <span class="html-italic">IGG6</span>-mediated bicistron. Primers P1/P2, P3/P4 and P1/P4 were used to detect the <span class="html-italic">vip3Aa</span>, <span class="html-italic">cry1Ab</span> and the bicistronic mRNA, respectively. The maize <span class="html-italic">actin1</span> gene was used as control. M: Trans5K DNA Marker, P: positive control, N: wild-type maize plants, B: blank. (<b>B</b>) The BLAST result of the bicistronic mRNA’s PCR amplicon. (<b>C</b>) Immunoblot analysis of the sizes of Vip3Aa and Cry1Ab protein. The β-Actin protein was used as control. M: PageRuler<sup>TM</sup> prestained protein ladder,10 to 180 kDa. N: wild-type maize plants.</p>
Full article ">Figure 4
<p>Laboratory bioassays of VICGG transgenic maize leaves with fall armyworm (FAW) and Asian corn borer (ACB). (<b>A</b>,<b>B</b>) The mortality rates of FAW and ACB larvae feeding on the leaves of wild-type and VICGG transgenic maize plants. Data represent means ± SD (n = 3 biological replicates). (<b>C</b>) The appearance of wild-type and transgenic maize leaves after insect bioassays with FAW and ACB. Photographs were taken after 5 days of infestation. Scale bar = 0.5 cm.</p>
Full article ">Figure 5
<p>Laboratory bioassays of VICGG-15 transgenic maize silk and ear with fall armyworm (FAW) and Asian corn borer (ACB). (<b>A</b>,<b>B</b>) The appearance of wild-type and VICGG-15 transgenic maize silks and ears after insect bioassays with FAW and ACB. Photographs were taken after 5 days of infestation. Scale bar = 0.5 cm. (<b>C</b>,<b>D</b>) The mortality rates of FAW larvae feeding on the silks and ears of wild-type and VICGG-15 plants. (<b>E</b>,<b>F</b>) The mortality rates of ACB larvae feeding on the silks and ears of wild-type and VICGG-15 plants. Data represent means ± SD (n = 3 biological replicates).</p>
Full article ">Figure 6
<p>Field bioassays of VICGG-15 transgenic maize plants with Asian corn borer (ACB). (<b>A</b>) The appearance of wild-type and VICGG-15 transgenic maize plants after insect bioassays with ACB. Photographs were taken after 3 days of infestation. (<b>B</b>) The resistance rating level of wild-type and VICGG-15 transgenic maize plants to ACB. Data represent means ± SD (n = 30).</p>
Full article ">Figure 7
<p>Glyphosate tolerance analysis and agronomic traits investigation of VICGG-15. (<b>A</b>) Pictures of VICGG-15 and B104 (WT) plants were recorded 10 days after treatment with glyphosate (900, 1800, and 3600 g a.e.ha<sup>−1</sup>), water as negative control. Aerial view taken by Unmanned Aerial Vehicle. (<b>B</b>) The plant height of VICGG-15 at 1, 2, and 4 weeks after glyphosate treatment. Data are means ± SD (n = 30), n.s: no significance (<span class="html-italic">p</span> &gt; 0.05, one-way ANOVA). (<b>C</b>) Ear phenotype of B104 and VICGG-15. Scale bar = 2 cm.</p>
Full article ">
19 pages, 33776 KiB  
Article
Effect of Insecticides Imidacloprid and Alpha-Cypermethrin on the Development of Pea (Pisum sativum L.) Nodules
by Artemii P. Gorshkov, Pyotr G. Kusakin, Maxim G. Vorobiev, Anna V. Tsyganova and Viktor E. Tsyganov
Plants 2024, 13(23), 3439; https://doi.org/10.3390/plants13233439 - 7 Dec 2024
Viewed by 666
Abstract
Insecticides are used commonly in agricultural production to defend plants, including legumes, from insect pests. It is a known fact that insecticides can have a harmful effect on the legume–rhizobial symbiosis. In this study, the effects of systemic seed treatment insecticide Imidor Pro [...] Read more.
Insecticides are used commonly in agricultural production to defend plants, including legumes, from insect pests. It is a known fact that insecticides can have a harmful effect on the legume–rhizobial symbiosis. In this study, the effects of systemic seed treatment insecticide Imidor Pro (imidacloprid) and foliar insecticide Faskord (alpha-cypermethrin) on the structural organization of pea (Pisum sativum L.) nodules and their transcriptomic activity were investigated. The plants were treated as recommended by the manufacturer (10 mg/mL for Imidor Pro and 50 µg/mL for Faskord) and twofold concentrations were used for both insecticides. Insecticides had no visible effect on the growth of pea plants. The nodules also showed no visible changes, except for the variant treated with twofold concentration of Imidor Pro. However, the dry weight of shoots and roots differed significantly in insecticide-treated plants compared to untreated plants in almost all treatments. The number of nodules decreased in variants with Imidor Pro treatment. At the ultrastructural level, both insecticides caused cell wall deformation, poly-β-hydroxybutyrate accumulation in bacteroids, expansion of the peribacteroid space in symbiosomes, and inclusions in vacuoles. Treatment with Faskord caused chromatin condensation in nucleus. Imidor Pro treatment caused hypertrophy of infection droplets by increasing the amount of matrix, as confirmed by immunofluorescence analysis of extensins. Transcriptome analysis revealed upregulation of expression of a number of extensin-like protein-coding genes in nodules after the Imidor Pro treatment. Overall, both insecticides caused some minor changes in the legume–rhizobial system when used at recommended doses, but Faskord, an enteric contact insecticide, has fewer negative effects on symbiotic nodules and legume plants; of these two insecticides, it is preferred in pea agricultural production. Full article
(This article belongs to the Special Issue Application of Agrochemical Technologies in Crop Protection)
Show Figures

Figure 1

Figure 1
<p>Phenotypes of pea plants (<span class="html-italic">Pisum sativum</span> L.) of the cv. ‘Frisson’. Untreated plants and plants treated with recommended by the manufacturer (1×), double-concentrated (2×) solutions of Imidor Pro and Faskord. Scale bars = 1 cm.</p>
Full article ">Figure 2
<p>Nodule phenotypes of pea plants (<span class="html-italic">Pisum sativum</span> L.) of the cv. ‘Frisson’. Untreated plants (<b>A</b>) and plants treated with recommended by the manufacturer (1×; <b>B</b>,<b>D</b>) and twofold-concentrated (2×; <b>C</b>,<b>E</b>) solutions of Imidor Pro (<b>B</b>,<b>C</b>) and Faskord (<b>D</b>,<b>E</b>). Bars = 2 mm.</p>
Full article ">Figure 3
<p>Growth parameters of pea plants (<span class="html-italic">Pisum sativum</span> L.) of the cv. ‘Frisson’. (<b>A</b>) Mean number of nodules per plant. (<b>B</b>) Mean shoot dry weight. (<b>C</b>) Mean root dry weight. Different letters indicate groups with a significant difference according to Tukey’s HSD test (<span class="html-italic">p</span>-value &lt; 0.05; n = 20). Vertical bars represent standard deviation.</p>
Full article ">Figure 4
<p>Histological organization of the nodules of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’ treated with insecticides. (<b>A</b>,<b>B</b>,<b>E</b>,<b>F</b>,<b>I</b>,<b>J</b>,<b>M</b>,<b>N</b>) Treatment with Imidor Pro. (<b>C</b>,<b>D</b>,<b>G</b>,<b>H</b>,<b>K</b>,<b>L</b>,<b>O</b>,<b>P</b>) Treatment with Faskord. (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>,<b>I</b>,<b>K</b>,<b>M</b>,<b>O</b>) Treatment with insecticides at the concentration recommended by the manufacturer. (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>,<b>J</b>,<b>L</b>,<b>N</b>,<b>P</b>) Treatment with a twofold-concentrated solution of insecticides. (<b>A</b>–<b>D</b>) Longitudinal section of a nodule. (<b>E</b>–<b>H</b>) Nodule meristematic cells. (<b>I</b>–<b>L</b>) Cells in the infection zone. (<b>M</b>–<b>P</b>) Infected cells in the nitrogen fixation zone. I, meristem; II, infection zone; III, nitrogen fixation zone; IV, senescence zone; n, nucleus; v, vacuole; arrows indicate infection droplets; *, enlarged infection droplets; #, cells with enlightened cytoplasm; s, starch; ic, infected cell; uic, uninfected cell. Triangles indicate vacuole fusion. Bars (<b>A</b>–<b>D</b>) = 100 µm, (<b>E</b>–<b>P</b>) = 10 µm.</p>
Full article ">Figure 5
<p>Ultrastructural organization of cell walls in nodules of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’. (<b>A</b>) Section of a nodule of an untreated 20-day-old plant. (<b>B</b>,<b>C</b>) Treatment with insecticide Imidor Pro. (<b>D</b>,<b>E</b>) Treatment with insecticide Faskord. Treatment with concentration recommended by the manufacturer (<b>B</b>,<b>D</b>) and with a twofold-concentrated solution of insecticides (<b>C</b>,<b>E</b>). cw, cell wall; ba, bacteroid; s; starch; a, amyloplast; id, infection droplet. Arrows indicate different cell wall abnormalities (curvature, swelling, uneven density). Bars (<b>A</b>–<b>E</b>) = 2 µm.</p>
Full article ">Figure 6
<p>Ultrastructural organization of bacteroids in infected cells in nodules of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’. (<b>A</b>) Section of a nodule of an untreated 20-day-old plant. (<b>B</b>,<b>C</b>) Treatment with insecticide Imidor Pro. (<b>D</b>,<b>E</b>) Treatment with insecticide Faskord. Treatment with concentration recommended by the manufacturer (<b>B</b>,<b>D</b>) and with a twofold-concentrated solution of insecticides (<b>C</b>,<b>E</b>). cw, cell wall; m, mitochondrion; ba, bacteroid; s, starch. Arrows indicate poly-β-hydroxybutyrate drops in bacteroids; arrowheads indicate the expansion of the peribacteroid space in symbiosomes. Bars (<b>B</b>–<b>E</b>) = 2 µm, (<b>A</b>) = 1 µm.</p>
Full article ">Figure 7
<p>Ultrastructural organization of vacuoles in nodules of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’. (<b>A</b>) Nodule section of an untreated 20-day-old plant. (<b>B</b>,<b>C</b>) Treatment with insecticide Imidor Pro. (<b>D</b>,<b>E</b>) Treatment with insecticide Faskord. Treatment with concentration recommended by the manufacturer (<b>B</b>,<b>D</b>) and with a twofold-concentrated solution of insecticides (<b>C</b>,<b>E</b>). n, nucleus; cw, cell wall; ba, bacteroid; v, vacuole; s, starch. Arrows indicate protein complexes in vacuoles; triangles indicate inclusions in uninfected cells, presumably of a phenolic nature. Bars (<b>A</b>–<b>E</b>) = 2 nm.</p>
Full article ">Figure 8
<p>Ultrastructural organization of infection droplets in infected cells in nodules of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’. (<b>A</b>) Nodule section of an untreated 20-day-old plant. (<b>B</b>,<b>C</b>) Treatment with insecticide Imidor Pro with concentration recommended by the manufacturer (<b>B</b>) and with a twofold-concentrated solution of insecticides (<b>C</b>). n, nucleus; it, infection thread; id, infection droplet; v, vacuole; cw, cell wall; ba, bacteroid; b, bacterium; a, amyloplast. Bars (<b>B</b>) = 5 µm, (<b>A</b>,<b>C</b>) = 2 µm.</p>
Full article ">Figure 9
<p>Ultrastructural organization of cell nuclei in nodules of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’. (<b>A</b>) Nodule section of an untreated 20-day-old plant. (<b>B</b>,<b>C</b>) Treatment with insecticide Faskord with concentration recommended by the manufacturer (<b>B</b>) and with a twofold-concentrated solution of insecticides (<b>C</b>). n, nucleus; cw, cell wall; ba, bacteroid; v, vacuole. Arrows indicate the formation of coarse chromatin clumps in the nuclei. Bars (<b>A</b>–<b>C</b>) = 2 µm.</p>
Full article ">Figure 10
<p>Effect of insecticide treatment of pea (<span class="html-italic">Pisum sativum</span> L.) cv. ‘Frisson’ on the matrix composition of infection droplets in nodule cells. (<b>A</b>–<b>E</b>) Immunolocalization of extensin labeled with JIM11 monoclonal antibody (MAb). (<b>A</b>) Section of nodules of untreated plants. Treatment with Imidor Pro (<b>B</b>,<b>C</b>) and Faskord (<b>D</b>,<b>E</b>) with concentration recommended by the manufacturer (<b>B</b>,<b>D</b>) and with a twofold-concentrated solution of insecticides (<b>C</b>,<b>E</b>). (<b>F</b>) Mean fluorescence intensity. The secondary antibodies used were goat anti-rat (<b>A</b>–<b>E</b>) IgG MAb conjugated with Alexa Fluor 488. ic, infected cell; n, nucleus; v, vacuole. Arrows indicate infection droplets. Different letters indicate groups with a significant difference according to Tukey’s HSD test (<span class="html-italic">p</span>-value &lt; 0.05; n = 30–45). Vertical bars represent standard deviation. Bars = 10 µm.</p>
Full article ">
8 pages, 3070 KiB  
Communication
A Novel Polymerase Chain Reaction (PCR)-Based Method for the Rapid Identification of Chrysodeixis includens and Rachiplusia nu
by Guilherme A. Gotardi, Natália R. F. Batista, Tamylin Kaori Ishizuka, Luiz H. Marques, Mário H. Dal Pogetto, Amit Sethi, Mark L. Dahmer and Timothy Nowatzki
Insects 2024, 15(12), 969; https://doi.org/10.3390/insects15120969 - 4 Dec 2024
Viewed by 806
Abstract
Chrysodeixis includens and Rachiplusia nu are two species belonging to the Plusiinae subfamily within the Noctuidae family. Due to their morphological similarity, the identification of their larvae is difficult and time-consuming. A rapid and accurate identification of these two species is essential for [...] Read more.
Chrysodeixis includens and Rachiplusia nu are two species belonging to the Plusiinae subfamily within the Noctuidae family. Due to their morphological similarity, the identification of their larvae is difficult and time-consuming. A rapid and accurate identification of these two species is essential for their management as these species exhibit differential susceptibilities to insecticides and crops engineered to express Bacillus thuringiensis (Bt) proteins, and a molecular tool can easily provide this differentiation. Currently, molecular analysis can identify these species through genetic sequencing, an expensive and time-consuming process. In our study, after sequencing part of the mtDNA cytochrome c oxidase I (COI) gene and based on the differences found in the gene of each species, a set of species-specific primers was developed: one reverse primer common to both species and two forward primers, specific to each species, amplifying fragments of 199 base pairs (bp) for C. includens and 299 bp for R. nu. Our results indicate that the primers were specific for these species, enabling the identification of individuals directly through agarose gel. The new methodology proved to be accurate, rapid, and reliable for the correct identification of these two species of loopers. Full article
(This article belongs to the Section Insect Molecular Biology and Genomics)
Show Figures

Figure 1

Figure 1
<p>Alignment of partial mtCOI gene sequences for <span class="html-italic">C. includens</span>, <span class="html-italic">R. nu</span>, <span class="html-italic">Trichoplusia ni</span>, <span class="html-italic">A. gamma</span>, and <span class="html-italic">A. egena</span>. The black dots indicate the nucleotide positions conserved across all the species. The nucleotides in white text over black backgrounds highlight differences between <span class="html-italic">C. includens</span> and <span class="html-italic">R. nu</span>. The sequences underlined in red indicate the position of the primers.</p>
Full article ">Figure 2
<p>Multiple sequence alignment of the mitochondrial cytochrome c oxidase subunit I (mtCOI) gene region for <span class="html-italic">R. nu</span> and <span class="html-italic">C. includens</span>, highlighting nucleotide variability between the species. The reference sequences used were KC354734.1 for <span class="html-italic">R. nu</span> and MT180766.1 for <span class="html-italic">C. includens</span>. Identical nucleotides are displayed in white text over a black background, while mismatched nucleotides are shown in plain text.</p>
Full article ">Figure 3
<p>SYBR Safe-stained 2% agarose gel electrophoresis of mtCOI PCR products. Lanes 1–5 show a 199 bp amplicon confirming <span class="html-italic">C. includens</span>. Lanes 6–10 show a 299 bp amplicon indicating <span class="html-italic">R. nu</span>. Lane M contains a 1 kb DNA ladder for size estimation, and lane N is a negative control.</p>
Full article ">Figure 4
<p>Agarose gel electrophoresis of mtCOI gene amplicons for validating method with field-collected insect samples. M: 1 kb DNA ladder; Lanes 1–6, 9, 10, and 12–19: 299 bp amplicon for <span class="html-italic">R. nu</span>; Lanes 7, 8, and 11: 199 bp amplicon for <span class="html-italic">C. includens</span>.</p>
Full article ">
26 pages, 4599 KiB  
Article
Enhancing Antimicrobial Efficacy of Sandalwood Essential Oil Against Salmonella enterica for Food Preservation
by Andrea Verešová, Margarita Terentjeva, Zhaojun Ban, Li Li, Milena Vukic, Nenad Vukovic, Maciej Ireneusz Kluz, Rania Ben Sad, Anis Ben Hsouna, Alessandro Bianchi, Ján Kollár, Joel Horacio Elizondo-Luévano, Natália Čmiková, Stefania Garzoli and Miroslava Kačániová
Foods 2024, 13(23), 3919; https://doi.org/10.3390/foods13233919 - 4 Dec 2024
Viewed by 1064
Abstract
The growing emphasis on food safety and healthier lifestyles, driven by industrial expansion and scientific priorities, has highlighted the necessity of managing harmful microorganisms to guarantee food quality. A significant challenge in this domain is the control of pathogens that are capable of [...] Read more.
The growing emphasis on food safety and healthier lifestyles, driven by industrial expansion and scientific priorities, has highlighted the necessity of managing harmful microorganisms to guarantee food quality. A significant challenge in this domain is the control of pathogens that are capable of forming biofilms, entering a sessile state that enhances their resistance to broad-spectrum antibiotics. Essential oils, renowned for their antibacterial properties, present a promising natural alternative for food preservation. In this study, we analyzed the chemical composition of Santalum album essential oil (SAEO) using GC-MS, identifying (Z)-α-santalol (57.1%) as the primary constituent. Antimicrobial activity was confirmed through disc diffusion and minimum inhibitory concentration (MIC) assays against Gram-positive and Gram-negative bacteria and yeast from the genus Candida. Additionally, in situ experiments demonstrated that vapor-phase SAEO effectively inhibited Serratia marcescens on the food model, supporting its potential as a natural preservative. MBIC assays, crystal violet staining, and MALDI-TOF MS analysis on S. enterica biofilms were used to further evaluate the antibiofilm effects of SAEO. The crystal violet assay revealed a strong antibiofilm effect, while the MALDI-TOF MS analysis showed changes in the bacterial protein profiles on both glass and plastic surfaces. SAEO also showed significant anti-Salmonella activity on vacuum-packed carrot slices. SAEO outperformed the control samples. The insecticidal activity against Megabruchidius dorsalis was also studied in this work, and the best insecticidal activity was found at the highest concentrations. These findings indicate that SAEO could serve as a valuable component in food preservation, with notable antibacterial and antibiofilm benefits. Full article
(This article belongs to the Section Food Microbiology)
Show Figures

Figure 1

Figure 1
<p>Isometric representation of the data from <a href="#foods-13-03919-t004" class="html-table">Table 4</a>: (<b>a</b>) Apple; (<b>b</b>) Pear. Color scale: Red ≤ 0%; Green: 0–50%; Blue: ≥ 50%.</p>
Full article ">Figure 2
<p>Isometric representation of the data from <a href="#foods-13-03919-t005" class="html-table">Table 5</a>: (<b>a</b>) Carrot; (<b>b</b>) Potato. Color scale: Red ≤ 0%; Green: 0–50%; Blue: ≥50%.</p>
Full article ">Figure 3
<p>Representative MALDI-TOF mass spectra of <span class="html-italic">S. enterica</span>: (<b>A</b>) 3rd day, (<b>B</b>) 5th day, (<b>C</b>) 7th day, (<b>D</b>) 9th day, (<b>E</b>) 12th day, and (<b>F</b>) 14th day. SE = <span class="html-italic">S. enterica</span>; G = glass; P = plastic; PC = planktonic cells.</p>
Full article ">Figure 4
<p>Dendrogram of <span class="html-italic">S. enterica</span> generated using MSPs of the planktonic cells and the control. SE = <span class="html-italic">S. enterica</span>; G = glass; P = plastic; PC = planktonic cells.</p>
Full article ">Figure 5
<p>Optical density (OD850) of <span class="html-italic">S. enterica</span> over time with and without SAEO treatment at varying temperatures.</p>
Full article ">Figure 6
<p>Growth rate (μ) of <span class="html-italic">S. enterica</span> over time with and without SAEO treatment at increasing temperatures.</p>
Full article ">Figure 7
<p>Total bacterial count (log CFU/g) in vacuum-packed carrots after heat treatment (50–65 °C, 5–20 min) on days 1 and 7. Data represent means ± SD (<span class="html-italic">n</span> = 3). Groups: Control (no treatment), Control Vacuum (vacuum-packed only), Essential Oil (1% SAEO), <span class="html-italic">S. enterica</span> (inoculated), and <span class="html-italic">S. enterica</span> + SAEO.</p>
Full article ">Figure 8
<p>Coliform bacteria count (log CFU/g) in vacuum-packed carrots after heat treatment (50–65 °C, 5–20 min) on days 1 and 7. Data represent means ± SD (<span class="html-italic">n</span> = 3). Groups: Control (no treatment), Control Vacuum (vacuum-packed only), Essential Oil (1% SAEO), <span class="html-italic">S. enterica</span> (inoculated), and <span class="html-italic">S. enterica</span> + SAEO.</p>
Full article ">Figure 9
<p>Krona diagram: species, genera, and families that were isolated from the sous vide carrot on the first day of storage.</p>
Full article ">Figure 10
<p>Krona diagram: species, genera, and families that were isolated from the sous vide carrot on the seventh day of storage.</p>
Full article ">
11 pages, 1513 KiB  
Article
Objective Assessment of the Damage Caused by Oulema melanopus in Winter Wheat with Intensive Cultivation Technology Under Field Conditions
by Sándor Keszthelyi, Richárd Hoffmann and Helga Lukács
AgriEngineering 2024, 6(4), 4538-4548; https://doi.org/10.3390/agriengineering6040259 - 28 Nov 2024
Viewed by 589
Abstract
Oulema melanopus L., 1758 (Coleoptera: Chrysomelidae) is one of the significant pests affecting cereal crops in Europe. Its damage is evident in the destruction of leaves during the spring growing season, leading to substantial impacts on both the quantity and quality of the [...] Read more.
Oulema melanopus L., 1758 (Coleoptera: Chrysomelidae) is one of the significant pests affecting cereal crops in Europe. Its damage is evident in the destruction of leaves during the spring growing season, leading to substantial impacts on both the quantity and quality of the harvested yields. The study aimed to evaluate the extent of leaf surface damage, changes in chlorophyll content caused by this pest, and the subsequent effects on yield quality. To achieve this, two experimental parcels were established, each subjected to different pesticide treatments during the spring vegetation cycle, but notably, with the difference that one parcel did not receive insecticide applications. The phytosanitary status, yield quantity, and quality parameters of thes parcels were compared. Chlorophyll content in damaged and undamaged plants was monitored in vivo using SPAD measurements, while the extent of leaf surface damage was assessed through image analysis using GIMP software 2.10.32. Harvested grain underwent milling and baking analysis, with milling and baking-quality indicators measured using a NIR grain analyzer. The results revealed that omitting springtime insecticide treatments during the emergence of O. melanopus led to significant reductions in leaf area and yield quality. In untreated parcels, leaf decession followed linear progression, reaching 35–40% within 20 days. This damage correlated with the decline in SPAD index values, indicating a 40–50% reduction in chlorophyll content dependent photosynthetic activity. Consequently, there were substantial decreases in milling and baking qualities, including nearly 30% reductional protein-content indicators and 10% in the Hagberg falling number. In summary, our large-scale field experiments demonstrated that persistent O. melanopus damage in wheat fields significantly reduced both the quantity and quality of yields, particularly protein content. These facts underscore the economic importance of timely pest-control measures to mitigate damage and preserve crop value. Full article
(This article belongs to the Section Pre and Post-Harvest Engineering in Agriculture)
Show Figures

Figure 1

Figure 1
<p>Developmental stages of <span class="html-italic">Oulema melanopus</span> and its host, winter wheat, and the applied chemical treatment and scheduled samplings in the experimental field.</p>
Full article ">Figure 2
<p>Estimation of damaged wheat leaf area using GIMP image analysis software. (<b>a</b>) scanned leaf area, (<b>b</b>) determination of the number of pixels of the whole leaf area, (<b>c</b>) designated leaf parts, (<b>c1</b>,<b>c2</b>) determination of the number of pixels of leaf spots damaged by <span class="html-italic">Oulema melanopus</span>, (<b>d</b>) image analysis parameters of the damaged leaf area.</p>
Full article ">Figure 3
<p>Trends of leaf surface destroyed by <span class="html-italic">Oulema melanopus</span>, depending on insecticide treatment (<b>a</b>) and representative samples (originating from untreated parcels) from the different sampling times (<b>b</b>). (a–e) aggravation of leaf damage as a function of time.</p>
Full article ">
13 pages, 2761 KiB  
Article
Characterization and Expression Patterns of Heat Shock Protein 70 Genes from Paracoccus marginatus in Response to Temperature and Insecticide Stress
by Yanting Chen, Jianwei Zhao, Mengzhu Shi, Fei Ruan, Jianwei Fu, Wanxue Liu and Jianyu Li
Agriculture 2024, 14(12), 2164; https://doi.org/10.3390/agriculture14122164 - 28 Nov 2024
Viewed by 649
Abstract
The objective of this study was to identify the Hsp70s in Paracoccus marginatus and explore their roles in P. marginatus’s resistance to temperature and insecticide stress. The full-length cDNA sequences of PmHsp70s were obtained by PCR cloning and sequencing. The physicochemical and [...] Read more.
The objective of this study was to identify the Hsp70s in Paracoccus marginatus and explore their roles in P. marginatus’s resistance to temperature and insecticide stress. The full-length cDNA sequences of PmHsp70s were obtained by PCR cloning and sequencing. The physicochemical and structural characteristics of PmHsp70s were analyzed, and a phylogenetic tree was constructed. The gene expressions of PmHsp70s were detected using qRT-PCR to explore the impacts of temperature and insecticide stress on P. marginatus. A total of 12 PmHsp70s were identified and cloned. The amino acids encoded by PmHsp70s were found to contain highly conserved regions characteristic of the Hsp70 family. The subcellular localization results showed that the majority of PmHsp70s were located in the cytoplasm. A total of 13 unique conserved motifs were identified for the PmHsp70s, of which 9 were shared motifs. The phylogenetic tree showed that the 12 PmHsp70s could be clustered into five branches, with the closest evolutionary relationship observed with the Phenacoccus solenopsis. The expression of the majority of PmHsp70s was up-regulated in P. marginatus when subjected to heat stress, with the higher expression fold change observed for PmHsp70-9, PmHsp70-11, and PmHsp70-12. The expression of specific PmHsp70s was notably suppressed under cold stress, whereas the expression of others was markedly enhanced. Upon exposure to chlorfenapyr and lambda-cyhalothrin, the expressions of PmHsp70-11 and PmHsp70-12 were significantly up-regulated with the highest expression fold change, respectively. The results revealed the significance of specific PmHsp70s in the resistance of P. marginatus to temperature and insecticide stress. This study improved our understanding of the mechanisms underlying P. marginatus’s adaptive responses to unfavorable environmental conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Conserved motifs of PmHsp70s in <span class="html-italic">P. marginatus</span>. The PmHsp70 family exhibited a variety of motifs, which are marked by the use of a color box.</p>
Full article ">Figure 2
<p>Phylogenetic analysis of Hsp70 from <span class="html-italic">P. marginatus</span> and other insect species based on neighbor-joining method. The classification of the PmHsp70s was based on the number of sub-branches into which they were grouped. The resulting classification was as follows: branch A (PmHsp70-2, PmHsp70-4, PmHsp70-5, PmHsp70-6, and PmHsp70-7), branch B (PmHsp70-9, PmHsp70-11, and PmHsp70-12), branch C (PmHsp70-3 and PmHsp70-10), branch D (PmHsp70-1), and branch E (PmHsp70-8). The different colors represent the different species.</p>
Full article ">Figure 3
<p>Expression patterns of <span class="html-italic">PmHsp70s</span> in <span class="html-italic">P. marginatus</span> under temperature stress. The expression of the target gene in <span class="html-italic">P. marginatus</span> at 26 °C was set as the control with a relative expression value = 1. Data are represented as mean ± standard error (SE). Asterisks (*) indicate a statistically significant difference between the control and the treatment (<span class="html-italic">p</span> ≤ 0.05), while double asterisks (**) indicate a highly statistically significant difference between the control and the treatment (<span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">Figure 4
<p>Expression patterns of <span class="html-italic">PmHsp70s</span> in <span class="html-italic">P. marginatus</span> under insecticide stress. CK: control; CH: chlorfenapyr; CY: lambda-cyhalothrin. The expression of the target gene in <span class="html-italic">P. marginatus</span> without insecticide treatment was set as the control with a relative expression value = 1. Data are represented as mean ± SE. Asterisks (*) indicate a statistically significant difference between the control and the treatment (<span class="html-italic">p</span> ≤ 0.05), while double asterisks (**) indicate a statistically significant difference between the control and the treatment (<span class="html-italic">p</span> ≤ 0.01).</p>
Full article ">
16 pages, 3173 KiB  
Article
Functional Analysis of Two Carboxylesterase Genes Involved in Beta-Cypermethrin and Phoxim Resistance in Plutella xylostella (L.)
by Ran Li, Liang Liang, Yujia Zhao, Junyi Zhang, Zhiyuan Hao, Haibo Zhao and Pei Liang
Agronomy 2024, 14(12), 2781; https://doi.org/10.3390/agronomy14122781 - 23 Nov 2024
Viewed by 631
Abstract
Enhanced expression of carboxylesterase (CarE) genes is an important mechanism of insecticide resistance in pests. However, their roles in multi-insecticide resistance have rarely been reported. Herein, two CarE genes (PxαE6 and PxαE9) were identified; their relative expression levels in three multi-insecticide-resistant [...] Read more.
Enhanced expression of carboxylesterase (CarE) genes is an important mechanism of insecticide resistance in pests. However, their roles in multi-insecticide resistance have rarely been reported. Herein, two CarE genes (PxαE6 and PxαE9) were identified; their relative expression levels in three multi-insecticide-resistant populations of P. xylostella (HN, GD-2017 and GD-2019) were 2.69- to 15.32-fold higher than those in the sensitive population, and they were considerably overexpressed at the larval stage and in the midgut of the 4th instar. PxαE6 and PxαE9 knockdown increased the susceptibility of GD-2019 larvae to phoxim or/and beta-cypermethrin. The recombinant PxαE6 and PxαE9 expressed in Escherichia coli exhibited high hydrolysis activity towards α-NA. GC–MS and LC–MS/MS assays revealed that PxαE9 could metabolize beta-cypermethrin and phoxim with efficiency determinations of 51.6% and 21.1%, respectively, while PxαE6 could metabolize phoxim with an efficiency of 12.0%. Homology modelling, molecular docking and molecular-dynamics simulation analyses demonstrated that beta-cypermethrin or/and phoxim could fit well into the active pocket and stably bind to PxαE6 or PxαE9. These results show that PxαE6 and PxαE9 overexpression were involved in resistance to beta-cypermethrin or/and phoxim in multi-insecticide-resistant P. xylostella populations, a finding which sheds light on the molecular mechanisms of multi-insecticide resistance in insect pests. Full article
Show Figures

Figure 1

Figure 1
<p>CarE-specific activity of third-instars of <span class="html-italic">Plutella xylostella</span> from SS, HN, GD-2017 and GD-2019 populations. Means with distinct lowercase letters are significantly different.</p>
Full article ">Figure 2
<p>Relative expression of <span class="html-italic">PxaE6</span> and <span class="html-italic">PxαE9</span> in four populations (<b>A</b>,<b>B</b>) and various developmental stages (<b>C</b>,<b>D</b>) and body parts/tissues (<b>E</b>,<b>F</b>) of <span class="html-italic">Plutella xylostella</span>. SS, susceptible population. Three field populations: HN, GD-2017 and GD-2019. L1–L4, first to fourth instars. MT, malpighian tubule. The results are depicted as the mean ± standard deviation (<span class="html-italic">n</span> = 3), with bars labeled by distinct lowercase letters indicating a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05) according to a one-way ANOVA followed by Tukey’s post hoc test.</p>
Full article ">Figure 3
<p>Relative expression of <span class="html-italic">PxaE6</span> (<b>A</b>) and <span class="html-italic">PxαE9</span> (<b>C</b>) in the 3rd larvae with dsRNA of selected <span class="html-italic">PxαE9</span> or ds<span class="html-italic">EGFP</span>. Mortalities of ds<span class="html-italic">PxαE6</span>- (<b>B</b>) and ds<span class="html-italic">PxαE9</span>-injected (<b>D</b>) third instars of <span class="html-italic">P. xylostella</span> 48 h post treatment with LC<sub>50</sub>. CYP, beta-cypermethrin; CHL, chlorantraniliprole; PHO, phoxim; TEB, tebufenozide; MET: metaflumizone. The asterisk * denotes a significant distinction between the treatment and control groups (Student’s <span class="html-italic">t</span>-test; <span class="html-italic">p</span> &lt; 0.05). Lowercase letters indicating a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05) according to a one-way ANOVA followed by Tukey’s post hoc test.</p>
Full article ">Figure 4
<p>Analysis of recombinant PxαE6 and PxαE9 by SDS-PAGE (<b>A</b>) and WB assay (<b>B</b>). The recombinant proteins were fractionated on 10% gels, and WB analysis was conducted using an anti-His tag antibody. Lanes 1, pColdII vector with 0.5 mM IPTG; M, protein ladder; Lanes 2, recombinant vector of PxαE6 with 0.5 mM IPTG; Lanes 3, purified proteins of PxαE6; Lanes 4, recombinant vector of PxαE9 with 0.5 mM IPTG; Lanes 5, purified proteins of PxαE9. WB: Western blot.</p>
Full article ">Figure 5
<p>The catalytic kinetic parameters of recombinant PxαE6 (<b>A</b>) and PxαE9 (<b>B</b>). <span class="html-italic">V</span><sub>max</sub>, maximum velocity; <span class="html-italic">K</span><sub>m</sub>, Michaelis–Menten constant. Values are presented as mean ± standard error (SE).</p>
Full article ">Figure 6
<p>Residues of phoxim after incubation with purified recombinant PxαE6 (<b>A</b>). Residues of beta-cypermethrin (<b>B</b>) and phoxim (<b>C</b>) after incubation with purified PxαE9. Heat-inactivated recombinant PxαE6 or PxαE9 (boiled PxαE6 or PxαE9) and blank controls (no proteins were added) were used as double controls. The data were presented as mean ± standard deviation (SD) with n = 3 replicates. Bars labeled with distinct lowercase letters indicated significant differences (<span class="html-italic">p</span> &lt; 0.05) based on one-way ANOVA followed by Tukey’s multiple comparison test.</p>
Full article ">Figure 7
<p>The overall structure (left) and binding model (right) of PxαE6 and PxaE9 relative to insecticides. (<b>A</b>) PxαE6 and phoxim. (<b>B</b>) PxαE9 and beta-cypermethrin. (<b>C</b>) PxαE9 and phoxim. Note: Beta-cypermethrin and phoxim are represented in green and cyan, respectively. Gray, H; blue, N; red, O; green, Cl; firebrick, Br; yellow, S; orange, P. Residues of catalytic triad (Ser-Glu-His) are presented with deep-blue slate lines. Oxyanion holes are presented with green lines, anion sites are blue-colored lines and acyl binding pockets are shown with orange lines. Other amino acids which can form hydrogen bonds with pesticides are shown with yellow lines. The hydrogen bonds are shown with red dotted lines.</p>
Full article ">Figure 8
<p>(<b>A</b>–<b>D</b>) The binding models of PxαE9 and (<span class="html-italic">S</span>)-(1<span class="html-italic">R</span>, 3<span class="html-italic">R</span>)-beta-cypermethrin, (<span class="html-italic">R</span>)-(1<span class="html-italic">S</span>, 3<span class="html-italic">S</span>) -beta-cypermethrin, (<span class="html-italic">S</span>)-(1<span class="html-italic">R</span>, 3<span class="html-italic">S</span>) -beta-cypermethrin and (<span class="html-italic">R</span>)-(1<span class="html-italic">S</span>, 3<span class="html-italic">R</span>)-beta-cypermethrin. Note: The beta-cypermethrin molecule is depicted by green lines. Residues of catalytic triad (Ser-Glu-His) are represented with deep-blue slate lines. Anion sites are depicted with blue lines, oxyanion holes are represented with green lines, and acyl binding pockets are represented by orange lines. The hydrogen bonds are illustrated with red dotted lines.</p>
Full article ">
15 pages, 5982 KiB  
Article
V-ATPase C Acts as a Receptor for Bacillus thuringiensis Cry2Ab and Enhances Cry2Ab Toxicity to Helicoverpa armigera
by Pin Li, Yuge Zhao, Ningbo Zhang, Xue Yao, Xianchun Li, Mengfang Du, Jizhen Wei and Shiheng An
Insects 2024, 15(11), 895; https://doi.org/10.3390/insects15110895 - 15 Nov 2024
Viewed by 874
Abstract
Cry2Ab is a significant alternative Bacillus thuringiensis (Bt) protein utilized for managing insect resistance to Cry1 toxins and broadening the insecticidal spectrum of crops containing two or more Bt genes. Unfortunately, the identified receptors fail to fully elucidate the mechanism of [...] Read more.
Cry2Ab is a significant alternative Bacillus thuringiensis (Bt) protein utilized for managing insect resistance to Cry1 toxins and broadening the insecticidal spectrum of crops containing two or more Bt genes. Unfortunately, the identified receptors fail to fully elucidate the mechanism of action underlying Cry2Ab. Previous studies have demonstrated the involvement of vacuolar H+-ATPase subunits A, B, and E (V-ATPase A, B, and E) in Bt insecticidal activities. The present study aims to investigate the contribution of V-ATPase C to the toxicities of Cry2Ab against Helicoverpa armigera. The feeding of Cry2Ab in H. armigera larvae resulted in a significant decrease in the expression of V-ATPase C. Further investigations confirmed the interaction between V-ATPase C and activated Cry2Ab protein according to Ligand blot and homologous and heterologous competition assays. Expressing endogenous HaV-ATPase C in Sf9 cells resulted in an increase in Cry2Ab cytotoxicity, while the knockdown of V-ATPase C by double-stranded RNAs (dsRNA) in midgut cells decreased Cry2Ab cytotoxicity. Importantly, a higher toxicity of the mixture containing Cry2Ab and V-ATPase C against insects was also observed. These findings demonstrate that V-ATPase C acts as a binding receptor for Cry2Ab and is involved in its toxicity to H. armigera. Furthermore, the synergy between V-ATPase C protein and Cry2Ab protoxins provides a potential strategy for enhancing Cry2Ab toxicity or managing insect resistance. Full article
(This article belongs to the Section Insect Molecular Biology and Genomics)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">HaV-ATPase C</span> expression in newly hatched (<b>A</b>) and fifth-instar (<b>B</b>) larvae following sublethal Cry2Ab exposure. The standard error of the average from three biological replicates is represented by each error bar. Bars represent the mean (±SD) of three biological replicates. Significant differences at each time point are denoted with asterisks based on Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, DPSSOFT: DPS9.01).</p>
Full article ">Figure 2
<p>SDS-PAGE analysis of the HaV-ATPase <span class="html-italic">C</span> protein and Ligand blotting analysis of the interaction between HaV-ATPase C and <span class="html-italic">Bt</span> toxin. (<b>A</b>) The prokaryotic expression of HaV-ATPase <span class="html-italic">C</span> protein was confirmed. (<b>B</b>) SDS-PAGE analysis of the purified HaV-ATPase <span class="html-italic">C</span> protein. (<b>C</b>) The interaction between HaV-ATPase C protein and activated Cry2Ab toxin. (<b>D</b>) The interaction between HaV-ATPase <span class="html-italic">C</span> protein and Vip3Aa protoxin.</p>
Full article ">Figure 3
<p>Special binding of HaV-ATPase <span class="html-italic">C</span> to activated Cry2Ab. (<b>A</b>) Saturation binding assays between HaV-ATPase <span class="html-italic">C</span> and activated Cry2Ab. The quantitative determination of the relative intensity of three replicates was conducted using Image J 1.8.0, followed by normalization with the result under the treatment of 30 μg biotinylated Cry2Ab. Bars labeled with different capital letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.01) based the LSD test conducted using DPSSOFT: DPS9.01. (<b>B</b>) Homologous competition between labeled Cry2Ab with unlabeled Cry2Ab to HaV-ATPase C protein. (<b>C</b>) Heterologous competition between labeled Cry2Ab with unlabeled Vip3Aa protoxin to HaV-ATPase C protein. Bars represent the mean (±SD) of three biological replicates. The bars with *** (<span class="html-italic">p</span> &lt; 0.001) exhibit significant differences and ns shows no significant differences, based on Student’s <span class="html-italic">t</span>-test (DPSSOFT: DPS9.01).</p>
Full article ">Figure 4
<p>Impact of expression of <span class="html-italic">HaV-ATPase C</span> in Sf9 cells on the cytotoxicity exerted by Cry2Ab and Vip3Aa. (<b>A</b>) Sf9 cells were transfected with RFP-pIEx plasmid (empty vector, control) 24 h. (<b>B</b>) Sf9 cells were transfected with <span class="html-italic">HaV-ATPase C</span> -pIEx-RFP plasmid 24 h. (<b>C</b>) Expression of HaV-ATPase <span class="html-italic">C</span> protein in Sf9 was analyzed by Western blot. (<b>D</b>) The cell mortalities of transfected cells in toxin-free treatment. (<b>E</b>) Cell mortalities were observed upon exposure to 200 μg/mL Vip3Aa protoxin. (<b>F</b>) Cell mortalities were observed upon exposure to 200 μg/mL activated Cry2Ab. Bars represent the mean (±SD) of three biological replicates. The bars with different capital letters (<span class="html-italic">p</span> &lt; 0.01) or lowercase letters (<span class="html-italic">p</span> &lt; 0.05) exhibit significant differences, as determined by the least significant difference (LSD) test using DPSSOFT: DPS9.01.</p>
Full article ">Figure 5
<p>Effects of <span class="html-italic">HzV-ATPase C</span> silencing on the toxicity of activated Cry2Ab and Vip3Aa towards MG cells. (<b>A</b>) The positioning of the dsRNA fragment within the <span class="html-italic">HaV-ATPase C</span> gene. (<b>B</b>) The expression levels of <span class="html-italic">HzV-ATPase C</span> under different treatments. (<b>C</b>) The cell mortalities of different treatments without toxins treated. (<b>D</b>) Cell mortalities were observed upon exposure to 150 μg/mL Vip3Aa protoxin. (<b>E</b>) Cell mortalities were observed upon exposure to 150 μg/mL activated Cry2Ab. Bars represent the mean (±SD) of three biological replicates. The bars with different capital letters (<span class="html-italic">p</span> &lt; 0.01) or lowercase letters (<span class="html-italic">p</span> &lt; 0.05) exhibit significant differences, as determined by the least significant difference (LSD) test using DPSSOFT: DPS9.01.</p>
Full article ">Figure 6
<p>Effect of HaV-ATPase <span class="html-italic">C</span> protein on the larvae weight (<b>A</b>): the toxicity of Cry2Ab (<b>B</b>) and Vip3Aa (<b>C</b>). (<b>A</b>) The effects of HaV-ATPase C protein on larval weight after 7 days of exposure. (<b>B</b>) The impact of incorporating HaV-ATPase <span class="html-italic">C</span> on the toxicity of Cry2Ab towards larvae. (<b>C</b>) The impact of incorporating HaV-ATPase <span class="html-italic">C</span> on the toxicity of Vip3Aa towards larvae. HaV-ATPase <span class="html-italic">C</span> dissolved in PBS, and <span class="html-italic">Bt</span> toxins dissolved in Na<sub>2</sub>CO<sub>3</sub>. The values 1×, 5×, 15×, and 50× represent the concentrations of V-ATPase C at 1 time, 5 times, 15 times, and 50 times that of <span class="html-italic">Bt</span> toxins. The mortality values were expressed as mean ± SD. The ns indicated no significant difference among the effects of HaV-ATPase C protein on larval weight after 7 days exposure. Bars represent the mean (±SD) of three biological replicates. Bars labeled with different capital letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.01) based the LSD test conducted using DPSSOFT: DPS9.01.</p>
Full article ">
17 pages, 5448 KiB  
Article
Biophysical Analysis of Vip3Aa Toxin Mutants Before and After Activation
by Pongsatorn Khunrach, Wahyu Surya, Boonhiang Promdonkoy, Jaume Torres and Panadda Boonserm
Int. J. Mol. Sci. 2024, 25(22), 11970; https://doi.org/10.3390/ijms252211970 - 7 Nov 2024
Viewed by 753
Abstract
Cry toxins from Bacillus thuringiensis are effective biopesticides that kill lepidopteran pests, replacing chemical pesticides that indiscriminately attack both target and non-target organisms. However, resistance in susceptible pests is an emerging problem. B. thuringiensis also produces vegetative insecticidal protein (Vip3A), which can kill [...] Read more.
Cry toxins from Bacillus thuringiensis are effective biopesticides that kill lepidopteran pests, replacing chemical pesticides that indiscriminately attack both target and non-target organisms. However, resistance in susceptible pests is an emerging problem. B. thuringiensis also produces vegetative insecticidal protein (Vip3A), which can kill insect targets in the same group as Cry toxins but using different host receptors, making the combined application of Cry and Vip3A an exciting possibility. Vip3A toxicity requires the formation of a homotetramer. Hence, screening of Vip3A mutants for increased stability requires orthogonal biophysical assays that can test both tetrameric integrity and monomeric robustness. For this purpose, we have used herein for the first time a combination of analytical ultracentrifugation (AUC), mass photometry (MP), differential static light scattering (DSLS) and differential scanning fluorimetry (DSF) to test five mutants at domains I and II. Although all mutants appeared more stable than the wild type (WT) in DSLS, mutants that showed more dissociation into dimers in MP and AUC experiments also showed earlier thermal unfolding by DSF at domains IV–V. All of the mutants were less toxic than the WT, but toxicity was highest for domain II mutations N242C and F229Y. Activation of the protoxin was complete and resulted in a form with a lower sedimentation coefficient. Future high-resolution structural data may lead to a deeper understanding of the increased stability that will help with rational design while retaining native toxicity. Full article
(This article belongs to the Special Issue Molecular Insights into Protein Structure and Folding)
Show Figures

Figure 1

Figure 1
<p>Location of residues mutated in the Vip3Aa protoxin (PDB: 6TFJ). (<b>A</b>) Atetrameric model of Vip3Aa, where one of the monomers is shown in color and the other three only as a transparent surface. The four residues mutated are indicated (red orange) and highlighted with a dotted rectangle and arrow. The five domains (I–V) of the toxin monomer are color-coded in orange (1–198), grey (199–325), blue (326–536), yellow (537–675), and cyan (676–789), respectively. (<b>B</b>) Same as (<b>A</b>) but showing a single monomer. (<b>C</b>) A close-up of the four residues mutated, seen from the tetrameric interface, where N242, T167, and E168 were mutated to Cys and F229 to Tyr. (<b>D</b>) The top view of the protoxin tetrameric interface, where two T167 residues are in close proximity (bold) whereas the other two are too far away to interact. Nearby residues E168 and N242, which may form interactions, are also shown.</p>
Full article ">Figure 2
<p>Molecular mass of Vip3Aa protoxin samples measured using mass photometry. (<b>A</b>) SDS−PAGE gel corresponding to freshly purified WT Vip3Aa and mutants. The black arrow indicates the band with expected molecular mass for the monomer. (<b>B</b>) Mass photometry histograms (particle count) at the indicated mass were fitted to Gaussian distributions (solid lines). Symmetrical peaks centered at 0 kDa are typical noise peaks and were not fitted. Vertical grey lines are shown to guide the eye and indicate the expected mass of monomer, dimer, and tetramer species. For T167C, the peak at 48 kDa could not be fitted into two peaks like for the other mutants. Percentages shown are based on total counts.</p>
Full article ">Figure 3
<p>SV analysis of Vip3A protoxin and activated toxin. (<b>A</b>) A comparison of the tetrameric shapes of the toxin before (PDB: 6TFJ) and after activation (PDB: 6TFK). In the latter, the N-terminal fragment is missing and only reached residue 95, and therefore it was extended up to residue 25 using the predicted AF2-structure of the domains I–III (overlayed with the experimental structure). (<b>B</b>,<b>C</b>) A c(s) plot of Vip3A protoxin at 0.5 mg/mL in Tris buffer at 20 °C normalized by band height (<b>B</b>) and proportion of different oligomers (<b>C</b>) calculated from the relative area under the bands in (<b>B</b>). (<b>D</b>,<b>E</b>) Same as (<b>B</b>,<b>C</b>), but for the Vip3A activated toxin. (<b>F</b>) The c(s) plots of the protoxin, with the y-axis not normalized. (<b>G</b>) Same as (<b>F</b>) for the activated toxin. The dotted line is shown to guide the eye. Dots in panels (<b>B</b>,<b>D</b>) represent number of monomers in the oligomer.</p>
Full article ">Figure 4
<p>Thermal aggregation curves of protoxin WT Vip3Aa and mutants. (<b>A</b>) Light scattering intensity was normalized, plotted as a function of temperature, and fitted to the Boltzmann equation by non-linear regression to obtain the temperature of aggregation, T<sub>agg</sub>. Each curve is a representative of three independent experiments conducted. The arrow indicates the increase in aggregation temperature between the WT and mutant N242C (~11.5 °C). (<b>B</b>) Average T<sub>agg</sub> values, with error bars representing one SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Thermal denaturation shift assay. (<b>A</b>) Melting curves of wild-type Vip3Aa and mutants. The melting temperatures (indicated) are marked by vertical dashed lines. Tm1 and Tm2 values are summarized in (<b>B</b>) with colors to guide the eye: green: more stable than the WT; blue: similar to the WT; red: both Tm1 and Tm2 lower than in the WT.</p>
Full article ">Figure 6
<p>Toxicity against <span class="html-italic">S. exigua</span> results. Orange bars: graphical representation in a logarithmic scale of LC<sub>50</sub>, with fiducial limits represented by a vertical bar; cyan bars: LC<sub>50</sub> normalized for the proportion of protoxin tetramer according to AUC and mass photometry.</p>
Full article ">Figure 7
<p>Comparison of the possible interactions of protoxin residues 242 and 168. In the WT AF2-predicted structure, N242 forms hydrogen bonds with E168 of the same monomer (cyan), and the latter forms another via main chain atoms with Q174 of a different monomer (magenta). In the experimental structure (6TFJ), these three residues are close, although too far to form hydrogen bonds. In the E168C mutant, the AF2-predicted model shows that the introduced cysteine can form hydrogen bonds with T170 of a neighboring monomer in addition to the aforementioned interaction with Q174, also present in the AF2-predicted models of mutants N242C and the double mutant. N242C may form three possible interactions with the same monomer, one of which is with E168.</p>
Full article ">
15 pages, 1463 KiB  
Review
Using Egg Parasitoids to Manage Caterpillars in Soybean and Maize: Benefits, Challenges, and Major Recommendations
by Adeney de F. Bueno, Weidson P. Sutil, M. Fernanda Cingolani and Yelitza C. Colmenarez
Insects 2024, 15(11), 869; https://doi.org/10.3390/insects15110869 - 5 Nov 2024
Viewed by 1042
Abstract
The use of egg parasitoids in Augmentative Biological Control (ABC) is a highly effective strategy within the integrated pest management (IPM) of lepidopteran defoliators. Safer than chemical insecticides, these natural antagonists have demonstrated significant efficacy. Trichogramma pretiosum and Telenomus remus, known for their [...] Read more.
The use of egg parasitoids in Augmentative Biological Control (ABC) is a highly effective strategy within the integrated pest management (IPM) of lepidopteran defoliators. Safer than chemical insecticides, these natural antagonists have demonstrated significant efficacy. Trichogramma pretiosum and Telenomus remus, known for their high parasitism rates, are the most extensively used and studied parasitoids for controlling economically important lepidopterous in crops such as soybean and maize. Brazil, a leading adopter of crops expressing Bacillus thuringiensis (Bt) proteins, faces growing field-evolved resistance to Cry proteins in soybean and maize. This resistance, particularly of Rachiplusia nu in soybean and Spodoptera frugiperda in maize, has become more prominent in recent years, increasing insecticide use. Therefore, this article reviews the current status of egg parasitoids adoption in ABC against lepidopteran pests, emphasizing the role of Tr. pretiosum and the potential of Te. remus as sustainable alternatives to chemical insecticides to manage pests in both non-Bt and Bt crops. Additionally, we provide recommendations for using these parasitoids in ABC programs and discuss the challenges that must be addressed to optimize the adoption of biocontrol agents in ABC programs for maximum benefit. Full article
Show Figures

Figure 1

Figure 1
<p>Adoption of <span class="html-italic">Bt</span> soybeans in Brazil (%) over the years (<b>A</b>) and in different states of Brazil in the 2019/20 season (<b>B</b>). Adapted from Bueno &amp; Silva [<a href="#B31-insects-15-00869" class="html-bibr">31</a>]. Brazilian States: Bahia (BA), Maranhão (MA), Mato Grosso do Sul (MS), Mato Grosso (MT), Paraná (PR), Santa Catarina (SC), and São Paulo (SP).</p>
Full article ">Figure 2
<p>Life cycle of <span class="html-italic">Trichogramma pretiosum</span> parasitizing <span class="html-italic">Anticarsia gemmatalis</span> eggs. Adapted from Bueno et al. [<a href="#B47-insects-15-00869" class="html-bibr">47</a>].</p>
Full article ">Figure 3
<p>Life cycle of <span class="html-italic">Telenomus remus</span> parasitizing <span class="html-italic">Spodoptera frugiperda</span> eggs. Adapted from Colmenarez et al. [<a href="#B40-insects-15-00869" class="html-bibr">40</a>].</p>
Full article ">
Back to TopTop