[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
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (638)

Search Parameters:
Keywords = VLPs

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
56 pages, 2782 KiB  
Review
Plant-Derived Anti-Cancer Therapeutics and Biopharmaceuticals
by Ghyda Murad Hashim, Mehdi Shahgolzari, Kathleen Hefferon, Afagh Yavari and Srividhya Venkataraman
Bioengineering 2025, 12(1), 7; https://doi.org/10.3390/bioengineering12010007 - 25 Dec 2024
Viewed by 64
Abstract
In spite of significant advancements in diagnosis and treatment, cancer remains one of the major threats to human health due to its ability to cause disease with high morbidity and mortality. A multifactorial and multitargeted approach is required towards intervention of the multitude [...] Read more.
In spite of significant advancements in diagnosis and treatment, cancer remains one of the major threats to human health due to its ability to cause disease with high morbidity and mortality. A multifactorial and multitargeted approach is required towards intervention of the multitude of signaling pathways associated with carcinogenesis inclusive of angiogenesis and metastasis. In this context, plants provide an immense source of phytotherapeutics that show great promise as anticancer drugs. There is increasing epidemiological data indicating that diets rich in vegetables and fruits could decrease the risks of certain cancers. Several studies have proved that natural plant polyphenols, such as flavonoids, lignans, phenolic acids, alkaloids, phenylpropanoids, isoprenoids, terpenes, and stilbenes, could be used in anticancer prophylaxis and therapeutics by recruitment of mechanisms inclusive of antioxidant and anti-inflammatory activities and modulation of several molecular events associated with carcinogenesis. The current review discusses the anticancer activities of principal phytochemicals with focus on signaling circuits towards targeted cancer prophylaxis and therapy. Also addressed are plant-derived anti-cancer vaccines, nanoparticles, monoclonal antibodies, and immunotherapies. This review article brings to light the importance of plants and plant-based platforms as invaluable, low-cost sources of anti-cancer molecules of particular applicability in resource-poor developing countries. Full article
(This article belongs to the Section Biomedical Engineering and Biomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The chemical structures of some prominent natural alkaloids and their semisynthetic derivatives serve as effective agents in combating cancer. Reproduced from an open-access source Dhyani et al., 2022 [<a href="#B151-bioengineering-12-00007" class="html-bibr">151</a>].</p>
Full article ">Figure 2
<p>The phototherapy mechanism of action. In photodynamic therapy (PDT), photosensitizers (PS) absorb light, transitioning to an excited state. This leads to two pathways: PDT Type I, where the PS reacts with biomolecules to create reactive oxygen species (ROS), and PDT Type II, where the PS transfers energy directly to oxygen, producing ROS. ROS exhibits high oxidizing power, causing cytotoxic effects primarily near their site of generation due to their short lifespan. PS* refers to the photosensitizer’s excited state. Reproduced from an open access source Pivetta et al., 2021 [<a href="#B212-bioengineering-12-00007" class="html-bibr">212</a>].</p>
Full article ">Figure 3
<p>PVNPs as delivery therapeutic and imaging agents in cancer. (<b>A</b>) Tobacco mosaic virus (TMV) for the targeted delivery of cisplatin in Pt-resistant ovarian cancer cells [<a href="#B312-bioengineering-12-00007" class="html-bibr">312</a>] (Reprinted/Adapted with permission from [<a href="#B272-bioengineering-12-00007" class="html-bibr">272</a>] Copyright© 2018, American Chemical Society. (<b>B</b>) The preparative process for potato virus X (PVX)-HisTRAIL by coordinating the bond between a Ni-nitrilotriacetic (NTA) group on the virus; the His-tag at the N-terminus of HisTRAIL is shown with a purple triangle. Multivalent display of HisTRAIL on the elongated PVX particle permits proper binding on death receptors DR4/5 (the trimers with blue color) for activating the caspase-dependent apoptosis in cancerous cells [<a href="#B313-bioengineering-12-00007" class="html-bibr">313</a>] (Reprinted/Adapted with permission from [<a href="#B273-bioengineering-12-00007" class="html-bibr">273</a>] Copyright© 2019, American Chemical Society). (<b>C</b>) miR-181a is an important target for ovarian cancer therapy. qPCR data and cancer cell migration assays demonstrated higher knockdown efficacy when anti-miR-181a oligonucleotides were encapsulated and delivered using the VLPs resulting in reduced cancer cell invasiveness [<a href="#B314-bioengineering-12-00007" class="html-bibr">314</a>] [Adapted from open access source: 274 Citation needed]. (<b>D</b>) Schematic illustration of Gd-Cy5.5-PhMV-mPEG NPs for cancer imaging. In vivo NIR fluorescence images of PC-3 prostate tumors in athymic nude mice after the intravenous injection of Gd-Cy5.5-PhMV-DGEA [<a href="#B315-bioengineering-12-00007" class="html-bibr">315</a>] [Adapted from open access source 275: Citation needed].</p>
Full article ">Figure 4
<p>PVNPs in cancer immune and combinational therapy (<b>A</b>) Intratumoral administration of plant-derived Cowpea mosaic virus (CPMV) nanoparticles as an in situ vaccine overcomes the local immunosuppression and stimulates a potent anti-tumor response in several mouse cancer models and canine patients [<a href="#B349-bioengineering-12-00007" class="html-bibr">349</a>] (Adapted from open access source: 309, Citation needed). (<b>B</b>) The PhMV-based anti-HER2 vaccine PhMV-CH401, demonstrated efficacy as an anti-HER2 cancer vaccine. Our studies highlight that VLPs derived from PhMV are a promising platform to develop cancer vaccines [<a href="#B350-bioengineering-12-00007" class="html-bibr">350</a>] (Adapted from open access source: 310, Citation needed). (<b>C</b>) Schematic diagram of preparing CCMV VLPs containing ODN 1826 (CCMV-ODN1826) for cancer therapy [<a href="#B315-bioengineering-12-00007" class="html-bibr">315</a>] (Adapted from open access source: 275, Citation needed). (<b>D</b>) Photothermal immunotherapy of melanoma using TLR-7 agonist laden TMV with polydopamine coat [<a href="#B325-bioengineering-12-00007" class="html-bibr">325</a>]. (Adapted from open access source: 285, Citation needed). Statistical significance was measured by one-way ANOVA with Tukey’s 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. ns refers to not significant.</p>
Full article ">
30 pages, 2582 KiB  
Review
Virus-like Particles Produced in Plants: A Promising Platform for Recombinant Vaccine Development
by Eugenia S. Mardanova, Egor A. Vasyagin and Nikolai V. Ravin
Plants 2024, 13(24), 3564; https://doi.org/10.3390/plants13243564 - 20 Dec 2024
Viewed by 304
Abstract
The capsid proteins of many viruses are capable of spontaneous self-assembly into virus-like particles (VLPs), which do not contain the viral genome and are therefore not infectious. VLPs are structurally similar to their parent viruses and are therefore effectively recognized by the immune [...] Read more.
The capsid proteins of many viruses are capable of spontaneous self-assembly into virus-like particles (VLPs), which do not contain the viral genome and are therefore not infectious. VLPs are structurally similar to their parent viruses and are therefore effectively recognized by the immune system and can induce strong humoral and cellular immune responses. The structural features of VLPs make them an attractive platform for the development of potential vaccines and diagnostic tools. Chimeric VLPs can be obtained by attaching foreign peptides to capsid proteins. Chimeric VLPs present multiple copies of the antigen on their surface, thereby increasing the effectiveness of the immune response. Recombinant VLPs can be produced in different expression systems. Plants are promising biofactories for the production of recombinant proteins, including VLPs. The main advantages of plant expression systems are the overall low cost and safety of plant-produced products due to the absence of pathogens common to plants and animals. This review provides an overview of the VLP platform as an approach to developing plant-produced vaccines, focusing on the use of transient expression systems. Full article
(This article belongs to the Section Plant Genetics, Genomics and Biotechnology)
Show Figures

Figure 1

Figure 1
<p>An overview of the transient expression of recombinant proteins in plants.</p>
Full article ">Figure 2
<p>Some widely used plant transient expression systems. The T-DNA regions of plant expression vectors based on the genomes of turnip vein-clearing virus (TVCV) and crucifer-infecting TMV (cr-TMV) (magnICON), cowpea mosaic virus (CPMV) (pEAQ), bean yellow dwarf virus (BeYDV), and potato virus X (PVX) (pEff). RB and LB, the left and right T-DNA; <span class="html-italic">target</span>, gene of interest; Act2, Arabidopsis actin 2 promoter; 35S, the promoter of cauliflower mosaic virus RNA; Nos-T, the terminator of the <span class="html-italic">A. tumefaciens</span> nopaline synthase gene; Term, the terminator of transcription; <span class="html-italic">p19</span>, the gene of tomato bushy stunt virus silencing suppressor; LIR, long intergenic region; SIR, short intergenic region; Rep/RepA, replication proteins from BeYDV; <span class="html-italic">RDRP</span>, RNA-dependent RNA polymerase gene; Sgp1, the first promoter of subgenomic RNA of PVX; AMV, a translational enhancer from alfalfa mosaic virus; <span class="html-italic">p24</span>, the gene of grapevine leafroll-associated virus-2 silencing suppressor; 5′ and 3′, untranslated regions (of diverse origins).</p>
Full article ">Figure 3
<p>Scheme of transient expression in plant cells using viral expression vectors.</p>
Full article ">Figure 4
<p>General scheme of chimeric VLP formation. (<b>a</b>) Native VLPs; (<b>b</b>) chimeric VLPs obtained by genetic fusion approach; (<b>c</b>) chimeric VLPs obtained by chemical crosslinking in vitro.</p>
Full article ">Figure 5
<p>The structures of HBc (PDB 6HU4). (<b>a</b>) Monomer chains A, B, C, and D are marked in blue, green, pink, and yellow, respectively. Three-dimensional modeling was performed by SWISS-MODEL [<a href="#B130-plants-13-03564" class="html-bibr">130</a>]. (<b>b</b>) VLPs of HBc [<a href="#B131-plants-13-03564" class="html-bibr">131</a>].</p>
Full article ">
17 pages, 4221 KiB  
Article
A New Method for Indoor Visible Light Imaging and Positioning Based on Single Light Source
by Xinxin Cheng, Xizheng Ke and Huanhuan Qin
Photonics 2024, 11(12), 1199; https://doi.org/10.3390/photonics11121199 - 20 Dec 2024
Viewed by 261
Abstract
Visible light positioning (VLP) can provide indoor positioning functions under LED lighting, and it is becoming a cost-effective indoor positioning solution. However, the actual application of VLP is limited by the fact that most positioning requires at least two or more LEDs. Therefore, [...] Read more.
Visible light positioning (VLP) can provide indoor positioning functions under LED lighting, and it is becoming a cost-effective indoor positioning solution. However, the actual application of VLP is limited by the fact that most positioning requires at least two or more LEDs. Therefore, this paper introduces a positioning system based on a single LED lamp, using an image sensor as the receiver. Additionally, due to the high computational cost of image processing affecting system real-time performance, this paper proposes a virtual grid segmentation scheme combined with the Sobel operator to quickly search for the region of interest (ROI) in a lightweight image processing method. The LED position in the image is quickly determined. Finally, the position is achieved by utilizing the geometric features of the LED image. An experimental setup was established in a space of 80 cm × 80 cm × 180 cm to test the system performance and analyze the positioning accuracy of the receiver in horizontal and tilted conditions. The results show that the positioning accuracy of the method can reach the centimeter level. Furthermore, the proposed lightweight image processing algorithm reduces the average positioning time to 53.54 ms. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram of the coordinate system transformation process for the imaging model (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 2
<p>Relation diagram of coordinate system transformation.</p>
Full article ">Figure 3
<p>Visible light positioning system architecture.</p>
Full article ">Figure 4
<p>Flow chart of image recognition and detection.</p>
Full article ">Figure 5
<p>Image processing process.</p>
Full article ">Figure 6
<p>Projection of LED light source on image plane.</p>
Full article ">Figure 7
<p>System experiment platform.</p>
Full article ">Figure 8
<p>Positioning results of different heights.</p>
Full article ">Figure 8 Cont.
<p>Positioning results of different heights.</p>
Full article ">Figure 9
<p>Comparison of average positioning errors at different vertical distances.</p>
Full article ">Figure 10
<p>CDF diagram of positioning errors.</p>
Full article ">Figure 11
<p>Histogram of positioning errors.</p>
Full article ">
18 pages, 2629 KiB  
Article
Development and Evaluation of a Newcastle Disease Virus-like Particle Vaccine Expressing SARS-CoV-2 Spike Protein with Protease-Resistant and Stability-Enhanced Modifications
by Yu Chen, Fan Tian, Shunlin Hu and Xiufan Liu
Viruses 2024, 16(12), 1932; https://doi.org/10.3390/v16121932 - 18 Dec 2024
Viewed by 511
Abstract
The ongoing global health crisis caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) necessitates the continuous development of innovative vaccine strategies, especially in light of emerging viral variants that could undermine the effectiveness of existing vaccines. In this study, we developed a [...] Read more.
The ongoing global health crisis caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) necessitates the continuous development of innovative vaccine strategies, especially in light of emerging viral variants that could undermine the effectiveness of existing vaccines. In this study, we developed a recombinant virus-like particle (VLP) vaccine based on the Newcastle Disease Virus (NDV) platform, displaying a stabilized prefusion form of the SARS-CoV-2 spike (S) protein. This engineered S protein includes two proline substitutions (K986P, V987P) and a mutation at the cleavage site (RRAR to QQAQ), aimed at enhancing both its stability and immunogenicity. Using a prime-boost regimen, we administered NDV-VLP-S-3Q2P intramuscularly at different doses (2, 10, and 20 µg) to BALB/c mice. Robust humoral responses were observed, with high titers of S-protein-specific IgG and neutralizing antibodies against SARS-CoV-2 pseudovirus, reaching titers of 1:2200–1:2560 post-boost. The vaccine also induced balanced Th1/Th2 immune responses, evidenced by significant upregulation of cytokines (IFN-γ, IL-2, and IL-4) and S-protein-specific IgG1 and IgG2a. Furthermore, strong activation of CD4+ and CD8+ T cells in the spleen and lungs confirmed the vaccine’s ability to promote cellular immunity. These findings demonstrate that NDV-S3Q2P-VLP is a potent immunogen capable of eliciting robust humoral and cellular immune responses, highlighting its potential as a promising candidate for further clinical development in combating COVID-19. Full article
(This article belongs to the Section Viral Immunology, Vaccines, and Antivirals)
Show Figures

Figure 1

Figure 1
<p>Construction and characterization of recombinant baculoviruses (rBVs). (<b>A</b>) Schematic diagram of the rBV constructs. rBV-S3Q2P-Ftmct includes the ectodomain of the SARS-CoV-2 S protein fused to the transmembrane and cytoplasmic tail (TM/CT) domains of the NDV F protein, with the S1/S2 cleavage site (RRAR) mutated to QQAQ and two proline substitutions (K986P, V987P) for stabilization. rBV-M-P2A-NP includes the NDV M and NP proteins, linked via the P2A sequence. (<b>B</b>,<b>C</b>) Immunofluorescence assay (IFA) of Sf9 cells infected with rBV-S3Q2P-Ftmct (<b>B</b>) and rBV-M-P2A-NP (<b>C</b>). Cells were stained with anti-S (red) or anti-NDV (green) antibodies, along with anti-GP64 (green) as a baculovirus marker. Scale bars,100 μm. (<b>D</b>,<b>E</b>) Western blot analysis of Sf9 cells infected with rBV-S3Q2P-Ftmct (<b>D</b>) and rBV-M-P2A-NP (<b>E</b>). Proteins were detected using antibodies against SARS-CoV-2 S protein, NDV NP, M, and GP64, and β-actin was used as a control. (<b>F</b>) PCR confirmation of the rBV constructs. The presence of the expected gene fragments for rBV-S3Q2P-Ftmct (lane A) and rBV-M-P2A-NP (lane B) was confirmed. (<b>G</b>) Virus titration results (TCID<sub>50</sub>/mL) of the rBVs across passages (P1–P6).</p>
Full article ">Figure 2
<p>Production and characterization of NDV-S3Q2P-VLPs. (<b>A</b>) Transmission electron microscopy (TEM) image of NDV-S3Q2P-VLPs. Sf9 cells were co-infected with rBV-S3Q2P-Ftmct and rBV-M-P2A-NP, and VLPs were purified at 72 hpi. VLPs were negatively stained and visualized under a transmission electron microscope. The zoomed-in sections (1, 2, and 3) provide magnified views of individual VLPs. Scale bars, 100 nm. (<b>B</b>) Western blot analysis of NDV-S3Q2P-VLPs. Purified VLPs were lysed and subjected to SDS-PAGE, followed by immunoblot with anti-SARS-CoV-2 S protein and anti-NDV antibodies. S protein (140 kDa), NDV NP (55 kDa), and M (45 kDa) were detected.</p>
Full article ">Figure 3
<p>Validation of 293T-hACE2 cell line and SARS-CoV-2 pseudovirus infectivity: (<b>A</b>) Western blot analysis of ACE2 expression in 293T wild-type (WT) and 293T-hACE2 cells from passages 1 to 8: Cells were lysed and subjected to SDS-PAGE, followed by immunoblotting using anti-ACE2 and anti-β-actin antibodies. (<b>B</b>) Luciferase assay to assess pseudovirus infectivity: 293T-WT and 293T-hACE2 cells were infected with SARS-CoV-2 pseudovirus or VSV-G pseudovirus, and luciferase activity was measured after 48 h to evaluate infection efficiency. (<b>C</b>) Fluorescence microscopy to visualize pseudovirus infection: 293T-WT and 293T-hACE2 cells were infected with SARS-CoV-2 pseudovirus or VSV-G pseudovirus, and GFP fluorescence was observed at 48 hpi. Scale bars, 100 μm. (<b>D</b>) Pseudovirus infection dose–response experiment: 293T-hACE2 cells were infected with increasing volumes of SARS-CoV-2 pseudovirus (10 μL, 20 μL, 30 μL, 40 μL, and 50 μL), and luciferase activity was measured at 48hpi to determine the optimal pseudovirus dose. NS means not significant, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Immunization schedule and groups: BALB/c mice (n = 6 per group) were immunized with different doses of NDV-S3Q2P-VLPs (2 μg, 10 μg, and 20 μg) or PBS as a control. Mice were immunized intramuscularly (I.M.) with a prime dose at week 0 and a boost dose at week 2. Blood samples were collected at week 2 and week 4 to assess immune responses.</p>
Full article ">Figure 5
<p>S-protein-specific IgG, IgG1, and IgG2a antibody responses induced by NDV-S3Q2P-VLPs immunization. (<b>A</b>) ELISA analysis of S-protein-specific IgG antibody titers at 2 weeks post-prime immunization. BALB/c mice were immunized with NDV-S3Q2P-VLPs at doses of 2 μg, 10 μg, or 20 μg, with PBS as the control. Serum samples were collected and serially diluted (1:10 to 1:320) to assess IgG titers. (<b>B</b>) Area under the curve (AUC) analysis of S-protein-specific IgG titers at 2 weeks post-prime immunization: AUC values were calculated to quantify the antibody responses across different doses. (<b>C</b>) ELISA analysis of S-protein-specific IgG antibody titers at 2 weeks post-boost immunization: serum samples were serially diluted (1:25 to 1:25,600) to measure the increase in IgG titers following the booster dose. (<b>D</b>) AUC analysis of S-protein-specific IgG titers at 2 weeks post-boost immunization: AUC values were calculated to evaluate the dose-dependent response after the boost. (<b>E</b>) ELISA analysis of S-protein-specific IgG1 and IgG2a antibody titers at 2 weeks post-prime immunization: serum samples were diluted at 1:40. (<b>F</b>) ELISA analysis of S-protein-specific IgG1 and IgG2a antibody titers at 2 weeks post-boost immunization: serum samples were diluted at 1:400. Statistical analyses were performed using one-way ANOVA for (<b>B</b>,<b>D</b>) and two-way ANOVA for (<b>E</b>,<b>F</b>). Data are presented as means ± standard deviation (SD). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Neutralizing-antibody titers against SARS-CoV-2 pseudovirus induced by NDV-S3Q2P-VLPs immunization: (<b>A</b>) Neutralizing-antibody titers (IC50) at 2 weeks post-prime immunization. BALB/c mice were immunized with NDV-S3Q2P-VLPs at doses of 2 μg, 10 μg, or 20 μg, with PBS as the control. Serum samples were collected and tested for neutralization of SARS-CoV-2 pseudovirus. IC50 values represent the dilution at which 50% neutralization was achieved. (<b>B</b>) Neutralizing antibody titers (IC50) at 2 weeks post-boost immunization: serum samples were collected and tested for neutralization activity against SARS-CoV-2 pseudovirus following the booster dose. Statistical analyses were performed using one-way ANOVA, and data are presented as means ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>NDV-S3Q2P-VLPs induced CD4+ and CD8+ T cell responses in lung and spleen two weeks after booster immunization: Lung and spleen single-cell suspensions were isolated from mice at 2 weeks post-boost immunization with different doses of NDV-S3Q2P-VLPs or PBS control. The cells were stained with mouse anti-CD4 and anti-CD8 antibodies to evaluate T cell populations. (<b>A</b>) Flow cytometry analysis of CD4+ and CD8+ T cell populations in the lung. (<b>B</b>) Flow cytometry analysis of CD4+ and CD8+ T cell populations in the spleen. (<b>C</b>) Proportional analysis of CD4+ and CD8+ T cells in the lung. (<b>D</b>) Proportional analysis of CD4+ and CD8+ T cells in the spleen. Data were analyzed with two-way ANOVA. * <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> &lt; 0.0001.</p>
Full article ">Figure 8
<p>Cytokine expression levels in tissues from mice immunized with NDV-S3Q2P-VLPs. Total RNA was extracted from kidney, liver, spleen, and lung tissues of NDV-S3Q2P-VLP-immunized mice or PBS control at 2 weeks post-boost immunization. qRT-PCR was used to measure the expression levels of IFN-γ, IL-2, and IL-4 in these tissues. (<b>A</b>) Fold change in IFN-γ expression in kidney, liver, spleen, and lung tissues. (<b>B</b>) Fold change in IL-2 expression in kidney, liver, spleen, and lung tissues. (<b>C</b>) Fold change in IL-4 expression in kidney, liver, spleen, and lung tissues. Cytokine expression was normalized to β-actin, and relative expression was calculated using the 2<sup>−ΔΔCt</sup> method. Statistical analyses were performed using two-way ANOVA. NS means not significant, * <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> &lt; 0.0001.</p>
Full article ">
13 pages, 5044 KiB  
Article
Development and Application of a Fully Automated Chemiluminescence Enzyme Immunoassay for the Detection of Antibodies Against Porcine Circovirus 3 Cap
by Lei Wang, Duan Li, Daoping Zeng, Xiaomin Wang, Yanlin Liu, Guoliang Peng, Zheng Xu and Changxu Song
Viruses 2024, 16(12), 1925; https://doi.org/10.3390/v16121925 - 17 Dec 2024
Viewed by 386
Abstract
Porcine circovirus 3 (PCV3) is a small non-enveloped circovirus associated with porcine dermatitis and nephropathy syndrome (PDNS). It has occurred worldwide and poses a serious threat to the pig industry. However, there is no commercially available vaccine. PCV3 capsid protein (Cap) is an [...] Read more.
Porcine circovirus 3 (PCV3) is a small non-enveloped circovirus associated with porcine dermatitis and nephropathy syndrome (PDNS). It has occurred worldwide and poses a serious threat to the pig industry. However, there is no commercially available vaccine. PCV3 capsid protein (Cap) is an ideal antigen candidate for serodiagnosis. Here, a novel fully automated chemiluminescence enzyme immunoassay (CLEIA) was developed to detect antibodies (Abs) to Cap in porcine serum. Recombinant PCV3 Cap, self-assembled into virus-like particles (VLPs), was produced using baculovirus and coupled to magnetic particles (Cap-MPs) as carriers. Combined with an alkaline phosphatase (AP)–adamantane (AMPPD) system, Cap-Abs can be rapidly measured on a fully automated chemiluminescence analyzer. Under optimal conditions, a cut-off value of 31,508 was determined, with a diagnostic sensitivity of 96.8% and specificity of 97.3%. No cross-reactivity was observed with PCV1 and PCV2 and other common porcine pathogens, and both intra-assay and inter-assay coefficients were less than 5% and 10%, respectively. Prepared Cap-MPs can be stored at 4 °C for more than 6 months. Importantly, this CLEIA had a good agreement of 95.19% with the commercially available kit, demonstrating excellent analytical sensitivity and significantly reduced operating time and labor. A serological survey was then conducted, and showed that PCV3 continues to spread widely in South China. In conclusion, our CLEIA provides time and labor-saving, and a reliable tool for PCV3 epidemiological surveillance. Full article
(This article belongs to the Special Issue Porcine Viruses 2024)
Show Figures

Figure 1

Figure 1
<p>Expression and purification of recombinant PCV3 Cap. (<b>A</b>) SDS-PAGE analysis of expression of recombinant Cap. 1,2 cell lysis forms cells infected with recombinant baculovirus, 3 cell lysis forms control, M protein marker. The blue arrow indicates expressed Cap, and the blue pentagram indicates bovine serum albumin (BSA). (<b>B</b>) SDS-PAGE analysis of Cap expression in <span class="html-italic">E. coli</span>, 1 cell lysis forms <span class="html-italic">E. coli</span> uninduced, 2 cell lysis forms <span class="html-italic">E. coli</span> induced by IPTG, 3 lysed supernatant, 4 precipitation, M protein marker. The blue arrow indicates expressed Cap. (<b>C</b>) Purification of recombinant Cap, 1. Cap from Sf9 cells, 2. Cap from <span class="html-italic">E. coli</span>., M protein marker. (<b>D</b>) IB analysis of recombinant Cap using anti-PCV3 Cap monoclonal antibody, 1. Cap from Sf9 cells, 2. Cap from <span class="html-italic">E. coli</span>., M protein marker. (<b>E</b>) TEM images of purified Cap.</p>
Full article ">Figure 2
<p>SEM images of MPs and prepared Cap-MPs. Scale bar: 5.0 μm.</p>
Full article ">Figure 3
<p>Optimization results for CLEIA. (<b>A</b>) Determination of optimal coating dosage of purified Cap. (<b>B</b>) Determination of optimal dilution of Cap-MPs. (<b>C</b>) Determination of optimal dilution of AP-conjugated goat anti-pig IgG. (<b>D</b>) Optimal reaction time (1, 3, 5, 8 min) for substrate. (<b>E</b>) Optimal antigen–antibody reaction model, one-step and two-step were assessed, respectively.</p>
Full article ">Figure 4
<p>Determination of the cut-off value. The analysis was performed on PCV3-positive serum samples (n = 29) and PCV3-negative serum samples (n = 89) using MedCalc software (Version 19.0.1). (<b>A</b>) Dot plot diagram. (<b>B</b>) ROC analysis.</p>
Full article ">Figure 5
<p>Cross-reactivity and stability test. (<b>A</b>) Specificity assay. PCV3-negatitive serum, PCV3-positive serum and PCV1, PCV2, PRV, ASFV, CSFV, PRRSV, PEDV, and FMDV positive serum were measured. Data represent the mean ±SD (standard deviation) from three independent experiments. (<b>B</b>) Prepared Cap-MPs were stored with other components at 4 °C for 0, 1, 2, 3, and 6 months. One positive sample and one negative sample were detected.</p>
Full article ">Figure 6
<p>Results of positive rates analysis. (<b>A</b>) PCV3 positive rates on farms in different stages of pigs. (<b>B</b>) PCV3 positive rates on farms in different provinces.</p>
Full article ">
13 pages, 3463 KiB  
Article
Data-Efficient Training of Gaussian Process Regression Models for Indoor Visible Light Positioning
by Jie Wu, Rui Xu, Runhui Huang and Xuezhi Hong
Sensors 2024, 24(24), 8027; https://doi.org/10.3390/s24248027 - 16 Dec 2024
Viewed by 318
Abstract
A data-efficient training method, namely Q-AL-GPR, is proposed for visible light positioning (VLP) systems with Gaussian process regression (GPR). The proposed method employs the methodology of active learning (AL) to progressively update the effective training dataset with data of low similarity to the [...] Read more.
A data-efficient training method, namely Q-AL-GPR, is proposed for visible light positioning (VLP) systems with Gaussian process regression (GPR). The proposed method employs the methodology of active learning (AL) to progressively update the effective training dataset with data of low similarity to the existing one. A detailed explanation of the principle of the proposed methods is given. The experimental study is carried out in a three-dimensional GPR-VLP system. The results show the superiority of the proposed method over both the conventional training method based on random draw and a previously proposed line-based AL training method. The impacts of the parameter of active learning on the performance of the GPR-VLP are also presented via experimental investigation, which shows that (1) the proposed training method outperforms the conventional one regardless of the number of final effective training data (E), especially for a small/moderate effective training dataset, (2) a moderate step size (k) should be chosen for updating the effective training dataset to balance the positioning accuracy and computational complexity, and (3) due to the interplay of the reliability of the initialized GPR model and the flexibility in reshaping such a model via active learning, the number of initial effective training data (m) should be optimized. In terms of data efficiency in training, the required number of training data can be reduced by ~27.8% by Q-AL-GPR for a mean positioning accuracy of 3 cm when compared with GPR. The CDF analysis shows that with the proposed training method, the 97th percentile positioning error of GPR-VLP with 300 training data is reduced from 11.8 cm to 7.5 cm, which corresponds to a ~36.4% improvement in positioning accuracy. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>): A picture of the three-dimensional VLP testbed. A total number of 1600 locations evenly distributed on four planes of different heights are used for data collection in the test. The dotted circles and solid dots on the rightmost figure show the projection of four LEDs and sampling locations on one of the four planes, respectively. The inner and outer area divided by the dashed line corresponds to the “center” and “corner” cases, respectively. (<b>b</b>): Schematic diagrams of the 3D VLP system. Note that the training dataset only needs to be constructed once for all positioning tasks at unknown locations in the future.</p>
Full article ">Figure 2
<p>Statistics of the average positioning error <math display="inline"><semantics> <mrow> <mfenced open="&#x2329;" close="&#x232A;" separators="|"> <mrow> <mi>ε</mi> </mrow> </mfenced> </mrow> </semantics></math> of 480 random test locations under different training methods after 1000 runs.</p>
Full article ">Figure 3
<p>Mean positioning error and computing time for AL under different dataset update strategies (i.e., different <span class="html-italic">k</span> values).</p>
Full article ">Figure 4
<p>Empirical (<b>a</b>) mean and (<b>b</b>) variance of the average positioning error <math display="inline"><semantics> <mrow> <mfenced open="&#x2329;" close="&#x232A;" separators="|"> <mrow> <mi>ε</mi> </mrow> </mfenced> </mrow> </semantics></math> of each run under different sizes (<math display="inline"><semantics> <mrow> <mi mathvariant="double-struck">E</mi> </mrow> </semantics></math>) of the finalized effective training dataset <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="script">D</mi> </mrow> <mrow> <mi>e</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Mean positioning accuracy of Q-AL-GPR with different numbers of initial effective training data (<span class="html-italic">m</span>). The result of GPR with the same number of effective training data (<math display="inline"><semantics> <mrow> <mi mathvariant="double-struck">E</mi> </mrow> </semantics></math> = 300) is shown by the dotted line for comparison.</p>
Full article ">Figure 6
<p>Empirical cumulative distribution function (CDF) of positioning error <math display="inline"><semantics> <mrow> <mi>ε</mi> </mrow> </semantics></math> for Q-AL-GPR and CPR under 300 effective training data after 1000 runs.</p>
Full article ">Figure 7
<p>The empirical CDF of the average positioning error <math display="inline"><semantics> <mrow> <mfenced open="&#x2329;" close="&#x232A;" separators="|"> <mrow> <mi>ε</mi> </mrow> </mfenced> </mrow> </semantics></math> for GPR and Q-AL-GPR when the training data are collected (<b>a</b>) with or (<b>b</b>) without tilt. The test data are collected with a certain angle of receiver tilt in both scenarios.</p>
Full article ">Figure 8
<p>Mean positioning error versus different numbers of final effective training data (<math display="inline"><semantics> <mrow> <mi mathvariant="double-struck">E</mi> </mrow> </semantics></math>) for the two training methods based on AL (i.e., Q-AL-GPR and line-based AL).</p>
Full article ">
23 pages, 11733 KiB  
Review
Potentiating Virus-like Particles for Mucosal Vaccination Using Material Science Approaches
by Milad Radiom
Colloids Interfaces 2024, 8(6), 68; https://doi.org/10.3390/colloids8060068 - 12 Dec 2024
Viewed by 503
Abstract
Virus-like particles (VLPs) exhibit such unique colloidal and structural properties that make them ideal candidates for various bio-nanotechnology applications, among which mucosal vaccination is particularly promising. However, since mucosal surfaces present harsh environments to VLPs, stabilization of VLP capsids or alternative delivery strategies [...] Read more.
Virus-like particles (VLPs) exhibit such unique colloidal and structural properties that make them ideal candidates for various bio-nanotechnology applications, among which mucosal vaccination is particularly promising. However, since mucosal surfaces present harsh environments to VLPs, stabilization of VLP capsids or alternative delivery strategies are necessary. Addressing these challenges requires interdisciplinary research, and the intersection of material science and immunology is presented in this review. Approaches such as crosslinking capsid coat proteins, incorporating VLPs in polymer matrices and hydrogels, or forming crystalline nano-/micro-structures show potential for developing muco-stable VLP vaccines or for delivering these vaccines in a sustainable manner. This review explores recent material science approaches that leverage VLPs as nanotools for various applications and with the potential for translation to mucosal vaccination. Full article
(This article belongs to the Special Issue Biocolloids and Biointerfaces: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Stability of virus-like particle (VLP) vaccines on mucosal surfaces is a critical factor for eliciting robust immune responses. Key determinants of antigen stability include interactions between VLPs and mucin glycoproteins leading to subsequent clearance in flowing mucus, destabilization by high concentration of proteases, aggregation due to variations in pH or secretory immunoglobulin A (sIgA), and mechanical agitation. <math display="inline"><semantics> <mrow> <mi>ξ</mi> </mrow> </semantics></math> denotes mucus porosity. (<b>B</b>) Atomic resolution structure of several VLPs including adeno-associated virus type 2 (AAV2) (PDB DOI: <a href="https://doi.org/10.2210/pdb5IPI/pdb" target="_blank">https://doi.org/10.2210/pdb5IPI/pdb</a>), the VLP constructed from the coat protein of Acinetobacter phage (AP205 VLP) (PDB DOI: <a href="https://doi.org/10.2210/pdb5LQP/pdb" target="_blank">https://doi.org/10.2210/pdb5LQP/pdb</a>), and cowpea mosaic virus (CPMV) (PDB DOI: <a href="https://doi.org/10.2210/pdb5A33/pdb" target="_blank">https://doi.org/10.2210/pdb5A33/pdb</a>). (<b>C</b>) Genetic fusion approach for VLP vaccine production: Antigens are genetically fused to coat proteins, enabling a one-step cytoplasmic assembly process (1) to generate the VLP vaccine. (<b>D</b>) Chemical conjugation approach for VLP vaccine production: Antigens are chemically attached to preassembled capsids. The process involves (1) cytoplasmic assembly of the capsid and (2) attachment of antigens to coat proteins via molecular linkers, resulting in the final VLP vaccine. (<b>E</b>) Tag/Catcher conjugation approach for VLP vaccine production: Coat proteins fused with Catcher proteins are assembled into capsids (1) during cytoplasmic assembly. In a subsequent step (2), antigens labeled with specific Tag peptides are conjugated to the Catcher proteins on the capsid surface, forming the final VLP vaccine. Alternatively, Tag peptide can be conjugated to the capsid, and Catcher protein to the antigen. Panel (<b>A</b>) is adapted from Ali et al.; Copyright © 2024 The Authors. Published by the American Chemical Society. This publication is licensed under CC-BY 4.0.</p>
Full article ">Figure 2
<p>Examples of recent virus-like particle (VLP)-based mucosal vaccines against viral and bacterial infections. (<b>A</b>) Schematic and transmission electron microscopy (TEM) image of P22-HA<sub>head</sub> VLP vaccine. Scale bar 100 nm. (<b>B</b>) Representative example of immunity induced by P22-HA<sub>head</sub> vaccination, showing the survival of immunized mice following exposure to a lethal virus challenge. (<b>C</b>) Schematic and cryo-TEM image of T4-CoV-2 VLP vaccine and T4-HSΔ (T4 vector control lacking outer capsid proteins Hoc and Soc, or any SARS-CoV-2 antigens). The red arrowheads indicate the S-trimer displayed on capsid surface. Scale bar 100 nm. (<b>D</b>,<b>E</b>) Selected examples of mucosal immunological response, showing anti-receptor binding domain (anti-RBD) IgA (<b>D</b>) and anti-spike ectodomain trimer (anti-Secto) IgA (<b>E</b>) titers in bronchoalveolar lavage fluid. Data presented as mean ± SEM, pooled from three independent experiments (n = 12 for T4-CoV-2, n = 10 for T4-HSD, and n = 5 for PBS). Titers between the intramuscular (i.m.) and intranasal (i.n.) routes were compared using two-way ANOVA (**** <span class="html-italic">p</span> &lt; 0.0001). (<b>F</b>) Schematic and TEM image of SliC-AP205 VLP vaccine. Scale bar 200 nm. (<b>G</b>,<b>H</b>) Selected examples of immunological responses showing total IgG, IgG1, IgG2a, and IgA titers in vaginal lavage after subcutaneous immunization of mice with SpyCatcher-conjugated VLP (cVLP) alone, SliC with the N-terminal SpyTag (N-SliC) alone, or N-SliC-VLP vaccine (<b>G</b>) and IgA titers in vaginal lavages after subcutaneous and intranasal immunizations of mice with N-SliC-VLP vaccine with AddaVax adjuvant, N-SliC-VLP vaccine with CpG adjuvant, N-SliC, cVLP with AddaVax adjuvant, cVLP with CpG adjuvant, or PBS (<b>H</b>). Graphs show geometric mean titers with error bars representing 95% confidence intervals. Statistical significance among groups was determined using the Kruskal-Wallis test with Dunn’s multiple comparisons. * <span class="html-italic">p</span> &lt; 0.05. Panels (<b>A</b>) and (<b>B</b>) are adapted with permission from Sharma et al., Copyright © 2020, American Chemical Society. Panels (<b>C</b>), (<b>D</b>) and (<b>E</b>) are adapted from Zhu, Ananthaswamy et al. and Zhu, Jain et al., respectively, under the terms of Creative Commons Attribution Noncommercial License 4.0 (CC BY-NC) and Creative Com-mons Attribution 4.0 International license. Panels (<b>F</b>), (<b>G</b>) and (<b>H</b>) are adapted from Martinez et al. under the terms of Creative Commons Attribution 4.0 International license.</p>
Full article ">Figure 3
<p>(<b>A</b>) Schematic representation of one-end conjugation of polymers to virus-like particle (VLP) capsids, or conjugation from both ends (crosslinking) using bifunctional polymers. Crosslinking can be on the exterior or interior capsid surfaces or both. (<b>B</b>) A cut section of agarose gel electrophoresis of native AP205 VLP, and PEGx-crosslinked AP205 VLPs (bPEG<sub>x</sub>-, where x is the number of monomers) and simply PEGylated AP205 VLP (mPEG<sub>25</sub>−) in pig gastric fluid at pH 3.0. Incubation times are indicated. Arrows at 5 min indicate stable capsids in gastric fluid. The complete gel can be found in Ref. [<a href="#B68-colloids-08-00068" class="html-bibr">68</a>]. (<b>C</b>) Confocal microscopy images of the distribution of VLPs on human nasal epithelial tissue with motile cilia (top) and on human nasal epithelial tissue with non-motile cilia (bottom). Cyan boxes show areas of confocal microscopy along the diameter of the tissue culture. (<b>D</b>,<b>E</b>) Generation of serum IgG antibodies upon subcutaneous immunization of mice with native AP205 VLP and PEG-crosslinked AP205 VLPs, against AP205 coat protein (<b>D</b>) and against PEG (<b>E</b>). Statistical significance was determined by ordinary one-way ANOVA with Dunnett’s multiple comparisons test, with a single pooled variance on log-normalized data (* <span class="html-italic">p</span> = 0.0332; ** <span class="html-italic">p</span> = 0.0021; *** <span class="html-italic">p</span> = 0.0002). (<b>F</b>) Samples of physalis mosaic virus (PhMV) and crosslinked PhMV (EE-PhMV) at various temperatures, together with TEM images showing thermally induced morphological changes to PhMV and stability of EE-PhMV at temperature range 25–90 °C. Scale bar 100 nm. Panel (<b>B</b>) to (<b>E</b>) is adapted from Ali et al.; Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under CC-BY 4.0. Panel (<b>F</b>) is reprinted with permission from Wu et al. Copyright 2024 American Chemical Society.</p>
Full article ">Figure 4
<p>(<b>A</b>,<b>B</b>) Schematics of virus-like particle (VLP) vaccines incorporated in polymer matrices for protection against proteases and secretory immunoglobulin A (sIgA): spray-dried microparticles (<b>A</b>) and hydrogels (<b>B</b>). (<b>C</b>) Scanning electron microscopy (SEM) image of spray-dried microparticles containing M2e5x VLPs, and micropores on mouse skin created by ablative laser. (<b>D</b>) Representative immune response showing serum IgG levels generated after immunization with inactivated PR8 H1N1 and M2e5x VLP + Alhydrogel<sup>®</sup> + MPL-A<sup>®</sup> microparticles. Data are expressed as mean ± standard deviation (SD). For multiple comparisons, one-way ANOVA was performed with Tukey’s post hoc test. A <span class="html-italic">p</span>-value &lt; 0.05 was considered statistically significant (**** <span class="html-italic">p</span> &lt; 0.0001). (<b>E</b>,<b>F</b>) Experimental settings used for in vitro release analysis of CPMV from hydrogel (<b>E</b>), and data showing the release of hydrogel-incorporated Cy5-CPMV (F1, F2, and F3 samples) versus free Cy5-CPMV/PBS at 37 °C (<b>F</b>). Hydrogel formulations contained 4.5 mg/ml CPMV dispersed in low molecular weight (MW), medium MW, and high MW chitosan, and named F1, F2, and F3, respectively. (<b>G</b>) TEM images of Cy5-CPMV released from hydrogels in vitro, showing the integrity and stability of VLPs within the hydrogel matrix. (<b>H</b>) Example of immunological response in mice vaccinated with hydrogel-incorporated CPMV Covid-19 vaccine. Blank F3: negative control, F3: 200 µg of CPMV vaccine in hydrogel, 200:200 µg of CPMV vaccine in PBS, and 100 (×2): prime/boot immunization with 100 µg of CPMV vaccine in PBS. Data were statistically analyzed using one-way ANOVA with Tukey’s multiple comparison test or two-way ANOVA with pairwise comparisons followed by Holm-Šidák correction. Asterisks in the figures indicate significant differences between groups (* <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). Panels (<b>C</b>) and (<b>D</b>) are adapted from Gomes et al. under an open access Creative Common CC BY license. Panels (<b>E</b>) to (<b>H</b>) are adapted with permission from Nkanga et al., Copyright © 2022 American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>A</b>) Schematic and optical image of cucumber mosaic virus containing a tetanus toxin-derived peptide, CuMV<sub>TT</sub>, decorating microcrystalline tyrosine (MCT). Scale bar 20 µm. (<b>B</b>) Representative immunological response, showing the percentage of CD11b+ AF488-labeled CuMV<sub>TT</sub> cells in tumor after 1 or 5 days of intratumoral injection. Comparisons involving more than two groups were performed using one-way analysis of variance (ANOVA), while comparisons between two groups were conducted using the non-parametric Student’s <span class="html-italic">t</span>-test. Statistical significance is indicated as follows: **** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; ns = not significant. (<b>C</b>) Cryo-TEM image of native bacteriophage Qβ, and schematic of Qβ in a hexagonal arrangement induced by polycation pMETAC1 interactions. Scale bar 30 nm. (<b>D</b>) Infectivity of native Qβ, Qβ/pMETAC1 suprastructure, and Qβ/pMETAC1 suprastructure + centrifuge (i.e., nanostructures separated from the liquid phase as macroscopic aggregates using centrifugation (inset) and resuspended in fresh medium). Panels (<b>A</b>,<b>B</b>) are reproduced from “In situ delivery of nanoparticles formulated with micron-sized crystals protects from murine melanoma”, Mohsen et al., 10, e004643, 2022 [<a href="#B85-colloids-08-00068" class="html-bibr">85</a>] with permission from BMJ Publishing Group Ltd., London, UK; Panels (<b>C</b>,<b>D</b>) are reproduced from Tran et al. [<a href="#B86-colloids-08-00068" class="html-bibr">86</a>] under the terms of the Creative Commons CC BY license.</p>
Full article ">
15 pages, 974 KiB  
Article
Performance Improvement by FRFT-OFDM for Visible Light Communication and Positioning Systems
by Wenyang Li, Zixiong Wang and Jinlong Yu
Photonics 2024, 11(12), 1147; https://doi.org/10.3390/photonics11121147 - 5 Dec 2024
Viewed by 566
Abstract
In indoor visible light communication (VLC) and visible light positioning (VLP) systems, the performance of conventional orthogonal frequency-division multiplexing (OFDM) schemes is often compromised due to the nonlinear characteristics and limited modulation bandwidth of light-emitting diodes, the multipath effect in enclosed indoor environments, [...] Read more.
In indoor visible light communication (VLC) and visible light positioning (VLP) systems, the performance of conventional orthogonal frequency-division multiplexing (OFDM) schemes is often compromised due to the nonlinear characteristics and limited modulation bandwidth of light-emitting diodes, the multipath effect in enclosed indoor environments, and the relative positions of transmitters and receivers. This paper proposes an OFDM scheme based on the fractional Fourier transform (FRFT) to address these issues, demonstrating promising results when applied to indoor VLC and VLP systems. The FRFT, a generalization of the conventional Fourier transform (FT) in the fractional domain, captures information in both the time and frequency domains, offering greater flexibility than the FT. In this paper, we first introduce the computation method of the reality-preserving FRFT for an intensity modulation/direct detection VLC system and integrate it with OFDM to optimize system performance. By adopting FRFT-OFDM under the optimal fractional order, we enhance both the bit error ratio (BER) performance and positioning accuracy. Simulation results reveal that the FRFT-OFDM scheme with an optimized fractional order significantly improves the BER and positioning accuracy compared to the FT-OFDM scheme across most receiver positions within the indoor observation plane. For communication, the FRFT-OFDM scheme achieves over 6 dB Eb/N0 gain compared to the FT-OFDM scheme at a BER of 3×104 when the receiver is positioned at most locations in the room. For positioning, the FRFT-OFDM scheme enhances positioning accuracy by more than 1 cm relative to the FT-OFDM scheme at most locations in the room. Notably, both systems maintain the same computational complexity and spectral efficiency. Full article
(This article belongs to the Special Issue New Advances in Optical Wireless Communication)
Show Figures

Figure 1

Figure 1
<p>Block diagram of RPFRFT-based OFDM system.</p>
Full article ">Figure 2
<p>Hammerstein model.</p>
Full article ">Figure 3
<p>VLC and VLP link with LOS and NLOS paths.</p>
Full article ">Figure 4
<p>Subcarrier allocation in the fractional Fourier domain for 4-LED VLC and VLP systems.</p>
Full article ">Figure 5
<p>BER performance comparison at various FRFT orders <span class="html-italic">p</span>.</p>
Full article ">Figure 6
<p>BER performance against <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>b</mi> </msub> <mo>/</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math> for FRFT-OFDM and FT-OFDM schemes using 4-QAM and 16-QAM.</p>
Full article ">Figure 7
<p>Localization errors against <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>b</mi> </msub> <mo>/</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math> for FRFT-OFDM and FT-OFDM schemes with 3- or 4-LED positioning.</p>
Full article ">Figure 8
<p>Localization errors at various positions throughout the room under the optimal order <span class="html-italic">p</span> of the FRFT.</p>
Full article ">Figure 9
<p>Improvement in positioning accuracy at various positions throughout the room under the optimal order <span class="html-italic">p</span> of the FRFT compared to the FT.</p>
Full article ">
17 pages, 7850 KiB  
Article
A Preclinical Immunogenicity Study of the Recombinant Human Papillomavirus Nine-Valent Virus-like Particle Vaccine
by Dan Xu, Jia-Dai Li, Jiao An, Xin-Xing Ma, Xiao-Liang Wang, Zheng Zhou, Hai-Ping Liu, Mei-Jun Diao, Yuan-Xiang Jiang, Ling-Yun Zhou, Xin Tong and Chen-Liang Zhou
Vaccines 2024, 12(12), 1356; https://doi.org/10.3390/vaccines12121356 - 30 Nov 2024
Viewed by 678
Abstract
Background: Cervical cancer is associated with persistent infection of high-risk human papillomaviruses (HPVs). Prophylactic HPV vaccines have been recommended and have significant efficacy in preventing cervical cancer. Multivalent HPV vaccines have a better preventative effect on HPV-related diseases. However, there is currently only [...] Read more.
Background: Cervical cancer is associated with persistent infection of high-risk human papillomaviruses (HPVs). Prophylactic HPV vaccines have been recommended and have significant efficacy in preventing cervical cancer. Multivalent HPV vaccines have a better preventative effect on HPV-related diseases. However, there is currently only one nine-valent HPV vaccine on the market: Gardasil® 9. The development of new HPV vaccines is still urgent in order to achieve the goal of eliminating cervical cancer as proposed by the WHO. Methods: In this study, we developed a nine-valent recombinant HPV virus-like particle (VLP) vaccine (HPV-9 vaccine) containing HPV type 6, 11, 16, 18, 31, 33, 45, 52, and 58 antigens, with an adjuvant of aluminum phosphate (AlPO4). The type-specific L1 proteins were recombinantly expressed using Pichia pastoris, followed by self-assembly into VLPs. Immunogenicity studies of the HPV-9 vaccine were performed using rodents (mice and rats) and non-human primates (macaques) as animal models. Results: Immunogenicity studies showed that the HPV-9 vaccine is able to elicit a robust and long-lasting neutralizing antibody response in rodents (mice and rats) and non-human primates (cynomolgus macaque) models. The HPV-9 vaccine shows immunogenicity comparable to that of Walrinvax® and Gardasil® 9. Conclusions: In summary, this study provides a comprehensive investigation of the immunogenicity of the HPV-9 vaccine, including its immune persistence. These findings, derived from using models of diverse animal species, contribute valuable insights into the potential efficacy of the vaccine candidate in clinical settings. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of recombinant expressed HPV L1 proteins. (<b>A</b>) Recombinant expressed HPV L1 proteins of each type were purified and subjected to reduced SDS-PAGE, arrowheads indicated the major band of HPV L1s. (<b>B</b>) Purified HPV L1 proteins were self-assembled into VLPs, and the nano-particle size was detected through using the Zetasizer Nano instrument. (<b>C</b>) Self-assembled VLPs of all nine types were characterized through transmission electron microscopy (TEM).</p>
Full article ">Figure 1 Cont.
<p>Characterization of recombinant expressed HPV L1 proteins. (<b>A</b>) Recombinant expressed HPV L1 proteins of each type were purified and subjected to reduced SDS-PAGE, arrowheads indicated the major band of HPV L1s. (<b>B</b>) Purified HPV L1 proteins were self-assembled into VLPs, and the nano-particle size was detected through using the Zetasizer Nano instrument. (<b>C</b>) Self-assembled VLPs of all nine types were characterized through transmission electron microscopy (TEM).</p>
Full article ">Figure 2
<p>Immunogenicity study of the nine monovalent AlPO<sub>4</sub>-adsorbed HPV VLPs. (<b>A</b>) Scheme of mice-based immunization routine. Briefly, female BALB/c mice were randomly divided into groups (with 10 mice/group), followed by intraperitoneal injection of 0.002 μg, 0.02 μg, or 0.2 μg monovalent AlPO<sub>4</sub>-adsorbed VLPs twice at a 2-week interval, with injection of the adjuvant AlPO<sub>4</sub> alone as the placebo. Serum samples were collected for detection of binding antibody titers and neutralizing antibody titers. (<b>B</b>) Binding antibody titers were detected through ELISAs using HPV VLPs of each type as antigens, serum from immunized mice as the 1st antibody with series dilutions, and diluted HRP-conjugated Goat anti-Mouse (H+L) IgG as the 2nd antibody. <span class="html-italic">p</span>-value less than 0.05 was considered statistically significant (<span class="html-italic">p</span> ≤ 0.05 (*), ≤0.01 (**), ≤0.001 (***), and &lt;0.0001 (****)). (<b>C</b>) Detection of neutralizing antibody titers was conducted based on the HPV pseudovirus of each type we prepared. <span class="html-italic">p</span>-value less than 0.05 was considered statistically significant (* <span class="html-italic">p</span> ≤ 0.05 (*), ≤0.01 (**), ≤0.001 (***), and &lt;0.0001 (****)).</p>
Full article ">Figure 2 Cont.
<p>Immunogenicity study of the nine monovalent AlPO<sub>4</sub>-adsorbed HPV VLPs. (<b>A</b>) Scheme of mice-based immunization routine. Briefly, female BALB/c mice were randomly divided into groups (with 10 mice/group), followed by intraperitoneal injection of 0.002 μg, 0.02 μg, or 0.2 μg monovalent AlPO<sub>4</sub>-adsorbed VLPs twice at a 2-week interval, with injection of the adjuvant AlPO<sub>4</sub> alone as the placebo. Serum samples were collected for detection of binding antibody titers and neutralizing antibody titers. (<b>B</b>) Binding antibody titers were detected through ELISAs using HPV VLPs of each type as antigens, serum from immunized mice as the 1st antibody with series dilutions, and diluted HRP-conjugated Goat anti-Mouse (H+L) IgG as the 2nd antibody. <span class="html-italic">p</span>-value less than 0.05 was considered statistically significant (<span class="html-italic">p</span> ≤ 0.05 (*), ≤0.01 (**), ≤0.001 (***), and &lt;0.0001 (****)). (<b>C</b>) Detection of neutralizing antibody titers was conducted based on the HPV pseudovirus of each type we prepared. <span class="html-italic">p</span>-value less than 0.05 was considered statistically significant (* <span class="html-italic">p</span> ≤ 0.05 (*), ≤0.01 (**), ≤0.001 (***), and &lt;0.0001 (****)).</p>
Full article ">Figure 3
<p>The immunogenicity and immune persistence of the HPV-9 vaccine in BALB/c mice. (<b>A</b>) As the immunization routine scheme shows, female BALB/c mice were randomly divided into groups, with 10 mice per group. The mice were intraperitoneally injected with a 1/100 clinical dose of Gardasil<sup>®</sup> 9 and HPV-9 vaccine or Walrinvax<sup>®</sup> or the adjuvant as the placebo at weeks 0, 2, and 6; neutralizing assays were conducted at weeks 2, 4, 8, 12, 20, 24 and 28. (<b>B</b>) Two weeks after the 3rd vaccination, the neutralizing antibody titers tested through neutralizing experiments demonstrated that the HPV-9 vaccine’s effectiveness is comparable to that of Gardasil<sup>®</sup> 9 and Walrinvax<sup>®</sup> (<span class="html-italic">p</span> ≤ 0.01 (**), ≤0.001 (***), and &lt;0.0001 (****)). (<b>C</b>) To evaluate the long-term protection induced by the HPV-9 vaccine in comparison with Gardasil<sup>®</sup> 9 or Walrinvax<sup>®</sup>, neutralizing assays were conducted to test the antibody-specific neutralizing antibodies in serum from vaccinated mice. The arrowheads represent the immunization timepoints.</p>
Full article ">Figure 4
<p>The immunogenicity and immune persistence of the HPV-9 vaccine in SD rats. (<b>A</b>) As the immunization routine scheme shows, female SD rats were randomly divided into groups (5 rats per group) and then intramuscularly injected with vaccines, namely the HPV-9 vaccine, Walrinvax<sup>®</sup>, and Gardasil<sup>®</sup> 9, at 1/10, 1/1000, or 1/10,000 of the clinical dose, respectively, at weeks 0, 2, and 6. Rats injected with the AlPO<sub>4</sub> adjuvant alone were used as the placebo control. Neutralizing assays of serum samples were conducted at weeks 2, 4, 8, 12, 20, 24, and 28. (<b>B</b>) Two weeks after the 3rd vaccination, the neutralizing antibody levels tested through neutralizing experiments showed that all vaccines induced dose-dependent immunogenicity, which was comparable among the vaccines (<span class="html-italic">p</span> ≤ 0.05 (*), ≤0.01 (**), and ≤0.001 (***)). (<b>C</b>) To evaluate the long-term protective effect induced by the HPV-9 vaccine compared to Gardasil<sup>®</sup> 9 or Walrinvax<sup>®</sup>, neutralizing assays were conducted to test the antibody-specific neutralizing antibodies in serum from vaccinated rats. Durable immune protection was induced with 1/10 of the clinical dose. The arrowheads represent the vaccination timepoints.</p>
Full article ">Figure 4 Cont.
<p>The immunogenicity and immune persistence of the HPV-9 vaccine in SD rats. (<b>A</b>) As the immunization routine scheme shows, female SD rats were randomly divided into groups (5 rats per group) and then intramuscularly injected with vaccines, namely the HPV-9 vaccine, Walrinvax<sup>®</sup>, and Gardasil<sup>®</sup> 9, at 1/10, 1/1000, or 1/10,000 of the clinical dose, respectively, at weeks 0, 2, and 6. Rats injected with the AlPO<sub>4</sub> adjuvant alone were used as the placebo control. Neutralizing assays of serum samples were conducted at weeks 2, 4, 8, 12, 20, 24, and 28. (<b>B</b>) Two weeks after the 3rd vaccination, the neutralizing antibody levels tested through neutralizing experiments showed that all vaccines induced dose-dependent immunogenicity, which was comparable among the vaccines (<span class="html-italic">p</span> ≤ 0.05 (*), ≤0.01 (**), and ≤0.001 (***)). (<b>C</b>) To evaluate the long-term protective effect induced by the HPV-9 vaccine compared to Gardasil<sup>®</sup> 9 or Walrinvax<sup>®</sup>, neutralizing assays were conducted to test the antibody-specific neutralizing antibodies in serum from vaccinated rats. Durable immune protection was induced with 1/10 of the clinical dose. The arrowheads represent the vaccination timepoints.</p>
Full article ">Figure 5
<p>Comparing the immunogenicity of the HPV-9 vaccine with that of Gardasil<sup>®</sup> 9 in macaques. (<b>A</b>) Macaques were randomly divided into groups (3 females and 3 males in each group), as the immunization routine scheme shows. Serum samples were collected as the control, followed by intramuscular injection of 1 clinical dose of the HPV-9 vaccine or Gardasil at weeks 0, 8, and 24 and the collection of blood at weeks 4, 12, and 26. (<b>B</b>) After the 2nd or 3rd vaccination, neutralizing antibody titers were detected at weeks 12 and 26, which indicated that the immunogenicity of the HPV-9 vaccine is comparative to that of Gardasil<sup>®</sup> 9 (<span class="html-italic">p</span> &gt; 0.05 (ns), and ≤0.001 (***)). (<b>C</b>) Trend regarding neutralizing antibodies at each timepoint compared with Gardasil<sup>®</sup> 9. The arrowheads represent the vaccination timepoints.</p>
Full article ">Figure 5 Cont.
<p>Comparing the immunogenicity of the HPV-9 vaccine with that of Gardasil<sup>®</sup> 9 in macaques. (<b>A</b>) Macaques were randomly divided into groups (3 females and 3 males in each group), as the immunization routine scheme shows. Serum samples were collected as the control, followed by intramuscular injection of 1 clinical dose of the HPV-9 vaccine or Gardasil at weeks 0, 8, and 24 and the collection of blood at weeks 4, 12, and 26. (<b>B</b>) After the 2nd or 3rd vaccination, neutralizing antibody titers were detected at weeks 12 and 26, which indicated that the immunogenicity of the HPV-9 vaccine is comparative to that of Gardasil<sup>®</sup> 9 (<span class="html-italic">p</span> &gt; 0.05 (ns), and ≤0.001 (***)). (<b>C</b>) Trend regarding neutralizing antibodies at each timepoint compared with Gardasil<sup>®</sup> 9. The arrowheads represent the vaccination timepoints.</p>
Full article ">
30 pages, 3897 KiB  
Article
Efficient Genome Editing Using ‘NanoMEDIC’ AsCas12a-VLPs Produced with Pol II-Transcribed crRNA
by Sofiia E. Borovikova, Mikhail V. Shepelev, Dmitriy V. Mazurov and Natalia A. Kruglova
Int. J. Mol. Sci. 2024, 25(23), 12768; https://doi.org/10.3390/ijms252312768 - 27 Nov 2024
Viewed by 734
Abstract
Virus-like particles (VLPs) are an attractive vehicle for the delivery of Cas nuclease and guide RNA ribonucleoprotein complexes (RNPs). Most VLPs are produced by packaging SpCas9 and its sgRNA, which is expressed from the RNA polymerase III (Pol III)-transcribed U6 promoter. VLPs assemble [...] Read more.
Virus-like particles (VLPs) are an attractive vehicle for the delivery of Cas nuclease and guide RNA ribonucleoprotein complexes (RNPs). Most VLPs are produced by packaging SpCas9 and its sgRNA, which is expressed from the RNA polymerase III (Pol III)-transcribed U6 promoter. VLPs assemble in the cytoplasm, but U6-driven sgRNA is localized in the nucleus, which hinders the efficient formation and packaging of RNPs into VLPs. In this study, using the nuclease packaging mechanism of ‘NanoMEDIC’ VLPs, we produced VLPs with AsCas12a and exploited its ability to process pre-crRNA. This allowed us to direct crRNA in the cytoplasm as part of a Pol II-driven transcript where AsCas12a excised mature crRNA, thus boosting RNP incorporation into VLPs. CMV-driven crRNA increased Venus and CCR5 transgene knockout levels in 293 cells from 30% to 50–90% and raised the level of endogenous CXCR4 knockout in Jurkat T cells from 1% to 20%. Changing a single crRNA to an array of three or six identical crRNAs improved CXCR4 knockout rates by up to 60–70%. Compared to SpCas9-VLPs, the editing efficiencies of AsCas12a-VLPs were higher, regardless of promoter usage. Thus, we showed that AsCas12a and CMV-driven crRNA could be efficiently packaged into VLPs and mediate high levels of gene editing. AsCas12a-VLPs are a new and promising tool for the delivery of RNPs into mammalian cells that will allow efficient target genome editing and may be useful for gene therapy applications. Full article
(This article belongs to the Special Issue CRISPR-Cas Systems and Genome Editing—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Selection of guide RNAs (gRNA). (<b>A</b>) Scheme of gRNA target sites for SpCas9 and AsCas12a in the selected loci: <span class="html-italic">Venus</span>, <span class="html-italic">CXCR4</span>, and <span class="html-italic">CCR5</span>; (<b>B</b>) <span class="html-italic">Venus</span> knockout level was measured by flow cytometry in 293-Venus clone #8 on day 6 after transfection; (<b>C</b>) <span class="html-italic">CXCR4</span> and (<b>D</b>) <span class="html-italic">CCR5</span> knockout levels were measured in CEM/CCR5 and CEM/CCR5 clone #8, respectively, stained with the corresponding antibodies on day 5 after electroporation [<a href="#B24-ijms-25-12768" class="html-bibr">24</a>,<a href="#B26-ijms-25-12768" class="html-bibr">26</a>,<a href="#B27-ijms-25-12768" class="html-bibr">27</a>,<a href="#B28-ijms-25-12768" class="html-bibr">28</a>,<a href="#B29-ijms-25-12768" class="html-bibr">29</a>,<a href="#B30-ijms-25-12768" class="html-bibr">30</a>,<a href="#B31-ijms-25-12768" class="html-bibr">31</a>,<a href="#B32-ijms-25-12768" class="html-bibr">32</a>].</p>
Full article ">Figure 2
<p>Generation of AsCas12acontaining virus-like particles (VLPs). (<b>A</b>) Scheme of plasmids encoding AsCas12a and its FRB fusion variants. (<b>B</b>) Representative Western blot evaluating the level of AsCas12a and its FRB fusion variants in transfected 293T cells. (<b>C</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout level in CEM/CCR5 induced by AsCas12a and its FRB fusion variants. (<b>D</b>) Workflow of VLP production. Original plasmids used by Gee et al. to produce ‘NanoMEDIC’ particles are highlighted in green, plasmids generated and used in this study are highlighted in red. (<b>E</b>) Representative Western blot evaluating the nuclease content in lysates of 293T producer cells and VLPs targeting <span class="html-italic">Venus</span>. (<b>F</b>) Flow cytometry analysis of the <span class="html-italic">Venus</span> knockout in 293-Venus clone #8 cells mediated by VLPs with AsCas12a or SpCas9. Shaded bars correspond to target cells preliminary transfected with the plasmid encoding the corresponding gRNA, dashed bars depict target cells transfected with the plasmid coding for control gRNA. Results from three independent experiments are shown as individual data points and as mean ± standard deviation; different symbols correspond to independent experiments. Mean values were compared by (<b>C</b>) one-sample <span class="html-italic">t</span>-test with Bonferroni correction (*) <span class="html-italic">p</span> &lt; 0.025 or (<b>F</b>) three-way ANOVA (with VLP dose, presence of crRNA in target cells, and nuclease type as factors) with subsequent Sidak’s multiple comparison test (**) <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Expression of AsCas12a crRNA under the control of the RNA polymerase II (Pol II) promoter allows efficient genome editing. (<b>A</b>) Scheme of plasmids encoding crRNA under the control of the U6 or CMV promoter. (<b>B</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout level in CEM/CCR5 cells mediated by AsCas12a or AsCas12a H800A in combination with one of the crRNA plasmids shown in (<b>A</b>). (<b>C</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout level in CEM/CCR5 cells mediated by AsCas12a and increasing amounts of the pCMV-mClover-trpl-DR-cr1X4-DR plasmid. (<b>D</b>,<b>E</b>) Flow cytometry analysis of the <span class="html-italic">CCR5</span> (<b>D</b>) or <span class="html-italic">Venus</span> (<b>E</b>) knockout levels in 293T/CD4/CCR5 clone #19 or 293-Venus clone #8 cells, respectively, induced by AsCas12a and crRNA expressed under the control of the U6 or CMV promoter. (<b>F</b>) Scheme of plasmids encoding crRNA (AsCas12a) or sgRNA (SpCas9) under the control of the U6 or CMV promoter. (<b>G</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout level in CEM/CCR5 cells mediated by AsCas12a in combination with one of the crRNA/sgRNA plasmids shown in (<b>F</b>). Results from 3–5 independent experiments are shown as individual data points and as mean ± standard deviation; different symbols correspond to independent experiments. Mean values were compared by one-way ANOVA for independent samples with subsequent Tukey’s test for multiple comparisons. (*) <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> &lt; 0.0001, (ns)—not significant.</p>
Full article ">Figure 4
<p>AsCas12a and crRNA can be expressed from a single Pol II-driven transcript that is compatible with the ‘NanoMEDIC’ system. (<b>A</b>) Scheme of single plasmids encoding AsCas12a and crRNA. (<b>B</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout level in CEM/CCR5 cells electroporated with one of the plasmid variants shown in (<b>A</b>). (<b>C</b>) Schematic of separate and single plasmids encoding AsCas12a with or without the FRB domain. (<b>D</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout level in CEM/CCR5 cells electroporated with one of the plasmid variants shown in (<b>C</b>) by AsCas12a and increasing amounts of the pCMV-mClover-trpl-DR-crRNA-DR plasmid. Results from three independent experiments are shown as individual data points and as mean ± standard deviation; different symbols correspond to independent experiments. Mean values were compared by one-way ANOVA for independent samples with subsequent Tukey’s test for multiple comparisons. (*) <span class="html-italic">p</span> &lt; 0.05, (**) <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Pol II-driven crRNA is compatible with multiplex genome editing. (<b>A</b>) Scheme of plasmids encoding crRNAs. (<b>B</b>,<b>C</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> (<b>B</b>) and <span class="html-italic">CCR5</span> (<b>C</b>) knockout levels in CEM/CCR5 clone #8 cells electroporated with the AsCas12a plasmid together with one of the plasmids shown in (<b>A</b>). The following amounts of crRNA plasmids were used: 0.48 pmol for pKS-U6-crRNA and 0.96 pmol for pCMV-based plasmids. Results from three independent experiments are shown as individual data points as mean ± standard deviation. Mean values were compared by one-way ANOVA for independent samples with subsequent Tukey’s test for multiple comparisons. (*) <span class="html-italic">p</span> &lt; 0.05, (**) <span class="html-italic">p</span> &lt; 0.01 (<b>B</b>).</p>
Full article ">Figure 6
<p>Generation of AsCas12a-VLPs with CMV promoter-driven crRNA. (<b>A</b>) Scheme of plasmids encoding AsCas12a and crRNA used for VLP production. (<b>B</b>,<b>D</b>) Flow cytometry analysis of <span class="html-italic">Venus</span> (<b>B</b>) and <span class="html-italic">CCR5</span> (<b>D</b>) knockout levels in 293-Venus clone #8 and 293T/CD4/CCR5 clone #19 cells, respectively, transduced with VLP preparations #1–4. Results from three independent experiments are shown as individual data points and as mean ± standard deviation; different symbols correspond to independent experiments. Mean values were compared by two-way ANOVA (with VLP dose and VLP type as factors) with subsequent Tukey’s multiple comparison 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; 0.0001, (ns)—not significant (shown only for a 50 µL dose). (<b>C</b>,<b>E</b>) Representative Western blot evaluating the nuclease content in lysates of 293T producer cells and VLPs targeting <span class="html-italic">Venus</span> (<b>C</b>) or <span class="html-italic">CCR5</span> (<b>E</b>).</p>
Full article ">Figure 7
<p>AsCas12a-VLPs produced with CMV-driven crRNA allow efficient genome editing in Jurkat T lymphocytes. (<b>A</b>) Scheme of plasmids encoding AsCas12a and crRNA used for VLP production. (<b>B</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout levels in Jurkat T cells transduced with VLPs #1–4. (<b>C</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout levels in Jurkat T cells electroporated with SpCas9 or AsCas12a RNPs. (<b>D</b>–<b>F</b>) Flow cytometry analysis of the <span class="html-italic">CXCR4</span> knockout levels in Jurkat T cells that were transduced with VLPs #3 produced with 1.66 µg or 4.98 µg of the crRNA plasmid (<b>D</b>), VLPs #3 produced with 1.66 µg of the crRNA plasmid or the plasmid coding for 3 or 6 identical spacers (<b>E</b>), and VLPs #3 coated with VSVG or VSVG+BaEVRless (data points related to independent experiments are shown by different shapes) (<b>F</b>).</p>
Full article ">
14 pages, 2348 KiB  
Article
Chimeric Virus-like Particles of Physalis Mottle Virus as Carriers of M2e Peptides of Influenza a Virus
by Elena A. Blokhina, Eugenia S. Mardanova, Anna A. Zykova, Marina A. Shuklina, Liudmila A. Stepanova, Liudmila M. Tsybalova and Nikolai V. Ravin
Viruses 2024, 16(11), 1802; https://doi.org/10.3390/v16111802 - 20 Nov 2024
Viewed by 792
Abstract
Plant viruses and virus-like particles (VLPs) are safe for mammals and can be used as a carrier/platform for the presentation of foreign antigens in vaccine development. The aim of this study was to use the coat protein (CP) of Physalis mottle virus (PhMV) [...] Read more.
Plant viruses and virus-like particles (VLPs) are safe for mammals and can be used as a carrier/platform for the presentation of foreign antigens in vaccine development. The aim of this study was to use the coat protein (CP) of Physalis mottle virus (PhMV) as a carrier to display the extracellular domain of the transmembrane protein M2 of influenza A virus (M2e). M2e is a highly conserved antigen, but to induce an effective immune response it must be linked to an adjuvant or carrier VLP. Four tandem copies of M2e were either fused to the N-terminus of the full-length PhMV CP or replaced the 43 N-terminal amino acids of the PhMV CP. Only the first fusion protein was successfully expressed in Escherichia coli, where it self-assembled into spherical VLPs of about 30 nm in size. The particles were efficiently recognized by anti-M2e antibodies, indicating that the M2e peptides were exposed on the surface. Subcutaneous immunization of mice with VLPs carrying four copies of M2e induced high levels of M2e-specific IgG antibodies in serum and protected animals from a lethal influenza A virus challenge. Therefore, PhMV particles carrying M2e peptides may become useful research tools for the development of recombinant influenza vaccines. Full article
(This article belongs to the Special Issue Nanovaccines against Viral Infection)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic representation of the recombinant protein structures. 4M2eh, four tandem copies of the M2e peptide; PhMV, the PhMV coat protein; PhMV_del, a truncated coat protein of PhMV with the deletion of the N-terminal region (shown in dark blue in PhMV). (<b>b</b>) 3D modeling of the structures of PhMV CP and 19s-4M2eh-19s-PhMV proteins using Alphafold v.2.3.1 [<a href="#B43-viruses-16-01802" class="html-bibr">43</a>]. Visualization was performed using the SWISS MODEL server [<a href="#B44-viruses-16-01802" class="html-bibr">44</a>]. The N- and C- termini are shown.</p>
Full article ">Figure 2
<p>Expression of recombinant proteins in <span class="html-italic">E.coli</span>. Proteins isolated from <span class="html-italic">E.coli</span> were analyzed by SDS-PAGE (<b>a</b>) or Western blotting with antibodies against the hexahistidine tag (<b>b</b>). M, molecular weight marker (sizes are shown in kD). Total proteins isolated from the <span class="html-italic">E. coli</span> strain without expression vector (lane 1); total proteins isolated after induction from <span class="html-italic">E. coli</span> strains carrying plasmids pQE30_PhMV (lane 2), pQE30_PhMV_del (lane 3), pQE30_4M2eh-PhMV (lane 4), pQE30_4M2eh-PhMV_del (lane 5), pQE30_19s-4M2eh-19s-PhMV (lane 6), and pQE30_19s-4M2eh-19s-PhMV_del (lane 7). The positions of recombinant proteins in SDS-PAGE are marked with red asterisks.</p>
Full article ">Figure 3
<p>Purification of VLPs formed by PhMV (<b>a</b>) and 19s-4M2eh-19s-PhMV (<b>b</b>) proteins. The proteins were analyzed by SDS–PAGE. M, molecular weight marker (sizes are shown in kD). Lanes: 1,—clarified cell lysate; 2,—VLPs after the first step of purification (precipitation with ammonium sulfate); 3,—VLPs after the second step of purification using ultracentrifugation (<b>a</b>) or nickel affinity chromatography (<b>b</b>).</p>
Full article ">Figure 4
<p>Analysis of the structure of VLPs formed by PhMV (<b>a</b>) and 19s-4M2eh-19s-PhMV (<b>b</b>) proteins using electron microscopy. Scale bar, 100 nm.</p>
Full article ">Figure 5
<p>Antigenic properties of VLPs. Two-fold dilutions of VLPs formed by PhMV and 19s-4M2eh-19s-PhMV proteins were loaded onto ELISA plates and then probed with anti-M2e antibodies.</p>
Full article ">Figure 6
<p>Immunogenicity and protective efficiency of VLPs. (<b>a</b>) Titers of anti-M2e IgG in sera of immunized mice. Data are presented as geometric mean titers (GMTs) and values observed in individual mice. (<b>b</b>) Protective efficacy of recombinant proteins. Statistically significant differences between groups are indicated (**, <span class="html-italic">p</span> &lt; 0.01; *, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
20 pages, 8040 KiB  
Article
A Modified Novel Validated High-Throughput Hemagglutinin Inhibition Assay Using Recombinant Virus-like Particles and Human Red Blood Cells for the Objective Evaluation of Recombinant Hemagglutinin Nanoparticle Seasonal Influenza Vaccine
by Timothy S. Vincent, Mingzhu Zhu, Anand Parekh, Urvashi Patel, Shane Cloney-Clark, Andrew Klindworth, David Silva, Andrew Gorinson, Karlee Miranda, Mi Wang, Zachary Longacre, Bin Zhou, Iksung Cho, Rongman Cai, Raj Kalkeri, Louis Fries, Vivek Shinde and Joyce S. Plested
Microorganisms 2024, 12(11), 2358; https://doi.org/10.3390/microorganisms12112358 - 19 Nov 2024
Viewed by 863
Abstract
Currently available seasonal influenza vaccines confer variable protection due to antigenic changes resulting from the accumulation of diverse mutations. The analysis of new seasonal influenza vaccines is challenging in part due to the limitations of the traditional hemagglutination inhibition (HAI) assay with A/H3N2 [...] Read more.
Currently available seasonal influenza vaccines confer variable protection due to antigenic changes resulting from the accumulation of diverse mutations. The analysis of new seasonal influenza vaccines is challenging in part due to the limitations of the traditional hemagglutination inhibition (HAI) assay with A/H3N2 strains. An improved and objective novel HAI assay was developed with recombinant virus-like particles (VLPs) and an egg-derived virus as agglutinins, the oseltamivir treatment of VLPs, human red blood cells, and using an automated image reader-based analysis of hemagglutination. HAI validation was demonstrated using four VLPs and egg-derived strains, with 46–56 serum samples tested 12 times in duplicate per strain. The validated HAI assay was precise as indicated by the percent geometric coefficient of variation for intra-, inter-, and total assay precision, as well as accurate as evidenced by percent bias measurements. The assay exhibited linearity, specificity for homologous type/subtype strains, and sensitivity with a starting dilution of 1:10. Assay robustness and sample stability were demonstrated as a percentage difference compared to reference condition. Validated HAI results were equivalent for the single and duplicate sample testing and correlated well with a qualified live wild-type influenza microneutralization assay. These findings demonstrate the suitability of this high-throughput novel modified validated HAI assay for evaluating vaccine immunogenicity and efficacy. Full article
(This article belongs to the Section Medical Microbiology)
Show Figures

Figure 1

Figure 1
<p>Linearity results of egg-derived virus HAI assay for influenza virus strains. (<b>a</b>) A/Kansas/14/2017, (<b>b</b>) A/Brisbane/02/2018, (<b>c</b>) B/Maryland/15/2016, and (<b>d</b>) B/Phuket/3073/2013. GMT, geometric mean titer; HAI, hemagglutination inhibition; <span class="html-italic">R</span><sup>2</sup>, coefficient of determination.</p>
Full article ">Figure 1 Cont.
<p>Linearity results of egg-derived virus HAI assay for influenza virus strains. (<b>a</b>) A/Kansas/14/2017, (<b>b</b>) A/Brisbane/02/2018, (<b>c</b>) B/Maryland/15/2016, and (<b>d</b>) B/Phuket/3073/2013. GMT, geometric mean titer; HAI, hemagglutination inhibition; <span class="html-italic">R</span><sup>2</sup>, coefficient of determination.</p>
Full article ">Figure 2
<p>Linearity results of VLP HAI assay for strains. (<b>a</b>) A/Kansas/14/2017, (<b>b</b>) A/Brisbane/02/2018, (<b>c</b>) B/Maryland/15/2016, (<b>d</b>) B/Phuket/3073/2013, (<b>e</b>) A/California/94/2019, (<b>f</b>) A/Cardiff/0508/2019, (<b>g</b>) A/Netherlands/1268/2019, and (<b>h</b>) A/Tokyo/EH1801/2018. GMT, geometric mean titer; HAI, hemagglutination inhibition; <span class="html-italic">R</span><sup>2</sup>, coefficient of determination; VLP, virus-like particle.</p>
Full article ">Figure 2 Cont.
<p>Linearity results of VLP HAI assay for strains. (<b>a</b>) A/Kansas/14/2017, (<b>b</b>) A/Brisbane/02/2018, (<b>c</b>) B/Maryland/15/2016, (<b>d</b>) B/Phuket/3073/2013, (<b>e</b>) A/California/94/2019, (<b>f</b>) A/Cardiff/0508/2019, (<b>g</b>) A/Netherlands/1268/2019, and (<b>h</b>) A/Tokyo/EH1801/2018. GMT, geometric mean titer; HAI, hemagglutination inhibition; <span class="html-italic">R</span><sup>2</sup>, coefficient of determination; VLP, virus-like particle.</p>
Full article ">Figure 3
<p>Comparison of clinical samples from qNIV-E-301 using duplicate results compared to random titer (singleton) for key study parameters. (<b>a</b>) Geometric mean ratio, (<b>b</b>) Seroprotection rate, and (<b>c</b>) Seroconversion rate for homologous and drifted seasonal influenza strains. Geometric mean ratio (GMR) was defined as the ratio of post-vaccination and pre-vaccination HAI GMTs within the same treatment group [<a href="#B23-microorganisms-12-02358" class="html-bibr">23</a>]. Seroprotection was defined as a titer of ≥1:40 (a titer that gives a 50% reduction of disease) [<a href="#B7-microorganisms-12-02358" class="html-bibr">7</a>]. Seroconversion was defined as HAI titer post-vaccination meeting one of the following criteria: either pre-vaccination titer &lt; 1:10 and post-vaccination titer ≥ 1:40, or pre-vaccination titer ≥ 1:10 and at least a 4-fold increase in post-vaccination titer [<a href="#B24-microorganisms-12-02358" class="html-bibr">24</a>,<a href="#B25-microorganisms-12-02358" class="html-bibr">25</a>]. CI, confidence interval; GMR, geometric mean ratio; GMT, geometric mean titer; HAI, hemagglutination inhibition; SCR, seroconversion rate; SPR, seroprotection rate; VLP, virus-like particle.</p>
Full article ">Figure 3 Cont.
<p>Comparison of clinical samples from qNIV-E-301 using duplicate results compared to random titer (singleton) for key study parameters. (<b>a</b>) Geometric mean ratio, (<b>b</b>) Seroprotection rate, and (<b>c</b>) Seroconversion rate for homologous and drifted seasonal influenza strains. Geometric mean ratio (GMR) was defined as the ratio of post-vaccination and pre-vaccination HAI GMTs within the same treatment group [<a href="#B23-microorganisms-12-02358" class="html-bibr">23</a>]. Seroprotection was defined as a titer of ≥1:40 (a titer that gives a 50% reduction of disease) [<a href="#B7-microorganisms-12-02358" class="html-bibr">7</a>]. Seroconversion was defined as HAI titer post-vaccination meeting one of the following criteria: either pre-vaccination titer &lt; 1:10 and post-vaccination titer ≥ 1:40, or pre-vaccination titer ≥ 1:10 and at least a 4-fold increase in post-vaccination titer [<a href="#B24-microorganisms-12-02358" class="html-bibr">24</a>,<a href="#B25-microorganisms-12-02358" class="html-bibr">25</a>]. CI, confidence interval; GMR, geometric mean ratio; GMT, geometric mean titer; HAI, hemagglutination inhibition; SCR, seroconversion rate; SPR, seroprotection rate; VLP, virus-like particle.</p>
Full article ">Figure 4
<p>Correlation analysis of the validated HAI assay with qualified MN assay for strains. (<b>a</b>) A/Brisbane/02/2018, (<b>b</b>) A/Kansas/14/2017, (<b>c</b>) B/Maryland/15/2016, and (<b>d</b>) B/Phuket/3073/2013. The dotted line shows 95% CI. CI, confidence interval; GMT, geometric mean titer; HAI, hemagglutination inhibition; MN, microneutralization; <span class="html-italic">R</span><sup>2</sup>, coefficient of determination.</p>
Full article ">Figure 4 Cont.
<p>Correlation analysis of the validated HAI assay with qualified MN assay for strains. (<b>a</b>) A/Brisbane/02/2018, (<b>b</b>) A/Kansas/14/2017, (<b>c</b>) B/Maryland/15/2016, and (<b>d</b>) B/Phuket/3073/2013. The dotted line shows 95% CI. CI, confidence interval; GMT, geometric mean titer; HAI, hemagglutination inhibition; MN, microneutralization; <span class="html-italic">R</span><sup>2</sup>, coefficient of determination.</p>
Full article ">
20 pages, 2059 KiB  
Review
Engineering Escherichia coli-Derived Nanoparticles for Vaccine Development
by Shubing Tang, Chen Zhao and Xianchao Zhu
Vaccines 2024, 12(11), 1287; https://doi.org/10.3390/vaccines12111287 - 18 Nov 2024
Viewed by 1279
Abstract
The development of effective vaccines necessitates a delicate balance between maximizing immunogenicity and minimizing safety concerns. Subunit vaccines, while generally considered safe, often fail to elicit robust and durable immune responses. Nanotechnology presents a promising approach to address this dilemma, enabling subunit antigens [...] Read more.
The development of effective vaccines necessitates a delicate balance between maximizing immunogenicity and minimizing safety concerns. Subunit vaccines, while generally considered safe, often fail to elicit robust and durable immune responses. Nanotechnology presents a promising approach to address this dilemma, enabling subunit antigens to mimic critical aspects of native pathogens, such as nanoscale dimensions, geometry, and highly repetitive antigen display. Various expression systems, including Escherichia coli (E. coli), yeast, baculovirus/insect cells, and Chinese hamster ovary (CHO) cells, have been explored for the production of nanoparticle vaccines. Among these, E. coli stands out due to its cost-effectiveness, scalability, rapid production cycle, and high yields. However, the E. coli manufacturing platform faces challenges related to its unfavorable redox environment for disulfide bond formation, lack of post-translational modifications, and difficulties in achieving proper protein folding. This review focuses on molecular and protein engineering strategies to enhance protein solubility in E. coli and facilitate the in vitro reassembly of virus-like particles (VLPs). We also discuss approaches for antigen display on nanocarrier surfaces and methods to stabilize these carriers. These bioengineering approaches, in combination with advanced nanocarrier design, hold significant potential for developing highly effective and affordable E. coli-derived nanovaccines, paving the way for improved protection against a wide range of infectious diseases. Full article
Show Figures

Figure 1

Figure 1
<p>Commonly used nanocarriers to display antigens. (<b>A</b>) 3D structure of VLPs. (<b>B</b>) 3D structure of protein-based nanoparticles. T represents the triangulation of nanoparticles, and mer is the abbreviation of protomers.</p>
Full article ">Figure 2
<p>Nanoparticles offer several key advantages over conventional vaccines, leading to more potent and targeted immune responses. Unlike smaller proteins or particles (&lt;10 nm) that are quickly cleared from the bloodstream, nanoparticles (20–100 nm) efficiently target draining lymph nodes, the primary site of immune response initiation. Furthermore, nanoparticles are preferentially engulfed by antigen-presenting cells (APCs), particularly dendritic cells (DCs), maximizing antigen presentation and immune activation.</p>
Full article ">Figure 3
<p>Display of antigens using the genetic fusion approach. (<b>A</b>) Antigens can be inserted at the N-terminus, C-terminus or MIR site of HBc. (<b>B</b>) Two consecutive HBc are designed as nanocarriers with the insertion of a single antigen at the MIR site to reduce steric hindrance.</p>
Full article ">Figure 4
<p>Three chemoenzymatic methods to loading antigens onto HBc nanocarrier. (<b>A</b>) Soratase-mediated site-specific tagging is introduced into split HBc nanoscaffold. (<b>B</b>) Split intein-mediated <span class="html-italic">trans</span>-splicing is utilized to conjugate antigens onto split HBc nanoscaffold. (<b>C</b>) SpyCatcher/SpyTag mediated covalent conjugation is applied to couple antigens onto HBc nanoscaffold.</p>
Full article ">Figure 5
<p>Advanced nanotechnologies used in <span class="html-italic">E. coli</span>-derived vaccine development.</p>
Full article ">
14 pages, 11775 KiB  
Article
Development of a Novel Chimeric ND-GP cVLPs Vaccine for the Prevention of Goose-Derived Newcastle Disease and Gosling Plague
by Jindou Li, Jiaxin Ding, Chunhong Guo, Xiaohong Xu, Chunhui Shan, Jing Qian and Zhuang Ding
Microorganisms 2024, 12(11), 2266; https://doi.org/10.3390/microorganisms12112266 - 8 Nov 2024
Viewed by 727
Abstract
Goose-derived Newcastle disease (ND) and gosling plague (GP) are serious threats to the goose industry. Conventional vaccines have made significant contributions to preventing GP and ND. Nevertheless, the renewal of conventional vaccines and the application of novel vaccines are urgently needed to align [...] Read more.
Goose-derived Newcastle disease (ND) and gosling plague (GP) are serious threats to the goose industry. Conventional vaccines have made significant contributions to preventing GP and ND. Nevertheless, the renewal of conventional vaccines and the application of novel vaccines are urgently needed to align with eco-friendly and efficient breeding concepts and achieve the final goal of epidemic purification. Therefore, based on the Newcastle disease virus-like particles (ND VLPs) vector platform, we developed novel chimeric ND-GP bivalent cVLPs (ND-GP cVLPs) displaying the NDV HN protein and the GPV VP3 protein. In vivo, immunization experiments revealed that geese immunized with 30 µg, 50 µg, or 70 µg of the ND-GP cVLPs and commercial vaccines produced highly effective hemagglutination inhibitory antibodies against NDV and neutralizing antibodies against GPV, respectively. Furthermore, 70 µg of the ND-GP cVLPs effectively protected against virulent NDV and GPV, reducing tissue damage from viral infection and virus shedding in the oropharynx and cloaca. In conclusion, we provide eco-friendly and efficient novel ND-GP cVLPs for preventing goose-derived ND and GP. Our findings provide the basis for using ND VLPs as foreign protein carriers for the developing of multi-conjugate vaccines. Full article
(This article belongs to the Topic Advances in Vaccines and Antimicrobial Therapy)
Show Figures

Figure 1

Figure 1
<p>The rescue and identification of the rBV-rVP3. (<b>A</b>) The amplification of the cHN and cVP3 fragments. Lane 1 contained a 177 bp cHN fragment, while lane 2 contained a 1644 bp cVP3 fragment. (<b>B</b>) The rVP3 fragment was obtained by overlap extension PCR with the cHN and cVP3 as templates. Lane 1 contained a 1806 bp fragment of the rVP3. (<b>C</b>) A restriction analysis of the rpFastbac1-rVP3 was performed using the restriction enzymes <span class="html-italic">Sal</span> I and <span class="html-italic">Hind</span> III. The short fragment was approximately 1800 bp long for the rVP3 gene, while the long fragment was 4708 bp long for the pFastbac1 linear vector. (<b>D</b>) The rBV-rVP3 genome was identified using rVP3-specific and M13 universal primers. Lane 1 was an approximately 1800 bp rVP3 fragment, and lane 2 was an approximately 4100 bp fragment containing the Tn7 transposon gene and the rVP3 gene. (<b>E</b>) The rVP3 protein was detected via Western blotting. (<b>F</b>) The morphological examination of the sf9 cells infected with rBV-rVP3 compared to the uninfected control sf9 cells. (<b>G</b>) The immunofluorescence detection of the rBV-rVP3-infected sf9 cells compared to the control sf9 cells.</p>
Full article ">Figure 2
<p>Assembly and identification of ND-GP cVLPs. (<b>A</b>–<b>C</b>) Identification of M, HN, and cVP3 proteins comprising ND-GP cVLPs through Western blotting. Analysis of morphological composition of NDV virions (<b>D</b>) and ND-GP cVLPs (<b>E</b>) by TEM. (<b>F</b>) Schematic representation of structure of ND-GP cVLPs.</p>
Full article ">Figure 3
<p>Monitoring of antibody titers induced by different immunization doses of ND-GP cVLPs. Detection of HI antibody titers against NDV (<b>A</b>) and VNA titers against GPV (<b>B</b>). A <span class="html-italic">p</span>-value of &lt; 0.05 was considered to indicate statistical significance (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Survival rate and weight increase rate of geese immunized with 70 µg of ND-GP cVLPs after challenge. Survival rate of geese post-virulent NDV (<b>A</b>) and post-virulent GPV challenge (<b>B</b>). Weight growth rate of geese after virulent NDV (<b>C</b>) or virulent GPV challenge (<b>D</b>) was monitored.</p>
Full article ">Figure 5
<p>Histopathological observations of the lung and intestine after the NDV challenge and the brain and intestine after GPV challenge. (<b>A</b>) Histological observation of the lung and intestine in the PBS group, ND-GP cVLPs group, and commercial vaccine group post-NDV challenge. (<b>B</b>) Histological observation of the brain and intestine in the PBS group, ND-GP cVLPs group, and commercial vaccine group post-NDV challenge.</p>
Full article ">
19 pages, 4273 KiB  
Article
Immunogenicity Assessment of a 14-Valent Human Papillomavirus Vaccine Candidate in Mice
by Lei Bei, Shuman Gao, Dandan Zhao, Yajuan Kou, Siyu Liang, Yurong Wu, Xiao Zhang, Dan Meng, Jianbo Lu, Chunxia Luo, Xuefeng Li, Yang Wang, Hongbin Qiu and Liangzhi Xie
Vaccines 2024, 12(11), 1262; https://doi.org/10.3390/vaccines12111262 - 8 Nov 2024
Viewed by 946
Abstract
Background: Cervical cancer ranks as the fourth most common cancer affecting women globally, with HPV as the primary etiology agent. Prophylactic HPV vaccines have substantially reduced the incidence of cervical cancer. Methods: This study assessed the immunogenicity of SCT1000, a 14-valent recombinant virus-like [...] Read more.
Background: Cervical cancer ranks as the fourth most common cancer affecting women globally, with HPV as the primary etiology agent. Prophylactic HPV vaccines have substantially reduced the incidence of cervical cancer. Methods: This study assessed the immunogenicity of SCT1000, a 14-valent recombinant virus-like particle (VLP) vaccine developed by Sinocelltech, Ltd. using pseudovirion-based neutralization assays (PBNAs) and total IgG Luminex immunoassays (LIAs). Currently in phase III clinical trials in China, SCT1000 targets the same HPV types as Gardasil 9®, plus five additional high-risk types, thereby covering twelve high-risk HPV types implicated in 96.4% of cervical cancer cases. Results: In murine models, a dose of 1.85 μg per mouse was identified as optimal for evaluating SCT1000’s immunogenicity in a three-dose regimen, as measured by PBNA and total IgG LIA across all 14 HPV types. SCT1000 induced high levels of protective antibodies, which were sustained for at least four months following the third dose. The vaccine also demonstrated stable and consistent immunogenicity in mouse potency assays under both long-term and accelerated conditions. Additionally, our studies revealed a strong correlation between the two serological tests used. Conclusions: SCT1000 elicited robust, durable, and consistent humoral immune responses across all 14 HPV types, indicating its potential as a broad-spectrum vaccine candidate against HPV types 6/11/16/18/31/33/35/39/45/51/52/56/58/59. The significant correlations observed between PBNA and total IgG LIA support the use of the Luminex-based total IgG method as a reliable and effective alternative for immunogenicity assessment in preclinical and future clinical vaccine development. Full article
(This article belongs to the Special Issue Vaccine Strategies for HPV-Related Cancers)
Show Figures

Figure 1

Figure 1
<p>Principles of PBNA and total IgG LIA. (<b>A</b>). Illustration of the cell-based PBNA. Neutralizing antibodies block GFP expression in target cells by binding to pseudoviruses and inhibiting pseudovirus infection. Non-neutralizing antibodies do not block infection and GFP expression. The increased presence of type-specific neutralizing antibodies results in a decrease in the number of cells expressing GFP, which can be quantified using the ImmunoSpot Analyzer (CTL). (<b>B</b>). Luminex-based total IgG LIA principle. The assay uses HPV VLPs coupled to fluorescent Luminex magnetic beads to quantify type-specific antibodies in mouse sera. Antibodies bound to type-specific epitopes on the VLPs are detected using a PE-labeled anti-mouse secondary antibody. The fluorescence signal value (MFI) positively correlates with the level of total type-specific IgG in the serum.</p>
Full article ">Figure 2
<p>Immunogenicity analysis of a 14-valent HPV candidate vaccine administered to BALB/c mice at different dosages in Study #1. In a dose escalation study (Study #1), geometric mean titers (GMTs) of 14 HPV types were measured using PBNA and total IgG LIA. The 0.5-, 1-, 1.25-, and 2-fold dilutions represent total HPV doses of 0.93 μg, 1.85 μg, 2.31 μg, and 3.70 μg per mouse for the 14-valent vaccine candidate, respectively. The solid black line and left vertical axis represent the PBNA titers, and the solid red triangles and right vertical axis represent total IgG LIA results (MFI). Error bars represent 95% confidence intervals (CIs) for GMTs.</p>
Full article ">Figure 3
<p>Immunogenicity analysis of a 14-valent HPV candidate vaccine administered to BALB/c mice at a single dosage in Study #2. A total of 60 mice were vaccinated with a 1-fold dilution (1.85 μg) of the candidate vaccine. Sera were tested with PBNA (<b>A</b>) and total IgG LIA (<b>B</b>). Geometric mean titers (GMTs) for 14 HPV types were measured. Error bars represent 95% confidence intervals (CIs) of GMTs.</p>
Full article ">Figure 4
<p>Immunogenicity analysis of a 14-valent HPV candidate vaccine administered to BALB/c mice in Study #3. In Study #3, geometric mean titers (GMTs) for 14 HPV types were measured by PBNA and total IgG LIA on days 28, 35, 56, 63, and 105. The solid blue lines and left vertical axis represent the PBNA results. The solid red squares and right vertical axis represent the total IgG LIA results. Error bars represent the 95% confidence intervals (CIs) of the GMT.</p>
Full article ">Figure 5
<p>The consistency and stability of the 14-valent HPV vaccine candidate in a mouse potency assay. (<b>A</b>). Mouse efficacy of three batches of 14-valent HPV vaccine candidates stored at 2~8 °C. The vertical axis represents ED<sub>50</sub> and standard error (SE) of the mean. (<b>B</b>). Mouse efficacy of three batches of 14-valent HPV vaccine candidates stored at 2~8 °C for 0, 3, 6, 9, 12, 18, 24, and 38 months. (<b>C</b>). Mouse efficacy of three batches of 14-valent HPV vaccine candidates stored at 25 °C for 0, 12, and 24 weeks. (<b>D</b>). Mouse efficacy of three batches of 14-valent HPV vaccine candidates stored at 25 °C for 0, 4, and 12 weeks.</p>
Full article ">Figure 6
<p>Correlation coefficients in individual studies and combined Study #1, Study #2, and Study #3. The figure shows the correlation between neutralizing antibody titers measured by PBNA and MFI readings measured by total IgG LIA when used alone and in combination in studies #1, #2, and #3. The horizontal axis represents the neutralizing antibody titers, and the vertical axis represents MFI readings. The horizontal and vertical dashed lines represent the cutoff values of total IgG LIA and PBNA for 14 HPV types, respectively. The slanted line represents linear regression. Pearson correlation coefficients (r values) are shown in the figure. Different symbols and colors represent data from different studies. Purple squares with purple regression lines represent data from Study #1, blue triangles with blue regression lines represent data from Study #2, green inverted triangles with green regression lines represent data from Study #3, and the red regression lines represent the combined or total data from all three studies (n = 150).</p>
Full article ">
Back to TopTop