Next Article in Journal
The Impact of Mutations on the Pathogenic and Antigenic Activity of SARS-CoV-2 during the First Wave of the COVID-19 Pandemic: A Comprehensive Immunoinformatics Analysis
Previous Article in Journal
A Combined Adjuvant TF–Al Consisting of TFPR1 and Aluminum Hydroxide Augments Strong Humoral and Cellular Immune Responses in Both C57BL/6 and BALB/c Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

VLP-Based COVID-19 Vaccines: An Adaptable Technology against the Threat of New Variants

1
Department of Immunology, Institute of Biomedical Sciences, University of São Paulo (ICB/USP), São Paulo 05508000, SP, Brazil
2
Institute of Research and Education in Child Health (PENSI), São Paulo 01228200, SP, Brazil
3
Department of Microbiology and Immunology, School of Medicine, Emory University, Claudia Nance Rollins Building, Atlanta, GA 30329, USA
4
São Carlos Institute of Physics, IFSC-USP, São Carlos 13566590, SP, Brazil
5
Institute for Biomedical Materials & Devices (IBMD), Faculty of Science, University of Technology, Sydney, NSW 2007, Australia
6
Department of Biomolecular Sciences, School of Pharmaceutical Sciences of Ribeirão Preto, University of São Paulo, Ribeirão Preto 14040903, SP, Brazil
7
Department of Chemical and Biomolecular Engeneering, North Carolina State University, Raleigh, NC 27695, USA
8
Department of Clinical and Toxicological Analyses, School of Pharmaceutical Sciences, University of São Paulo, São Paulo 05508000, SP, Brazil
9
Network of Immunity in Infection, Malignancy and Autoimmunity (NIIMA), Universal Scientific Education and Research Network (USERN), Children’s Medical Center, Tehran 1419733151, Iran
10
Laboratory of Medical Investigation in Pathogenesis and Targeted Therapy in Onco-Immuno-Hematology (LIM-31), Department of Hematology, Hospital das Clínicas HCFMUSP, Faculdade de Medicina, Universidade de São Paulo, São Paulo 0124690, SP, Brazil
*
Author to whom correspondence should be addressed.
These authors have equally contributed.
Vaccines 2021, 9(12), 1409; https://doi.org/10.3390/vaccines9121409
Submission received: 7 October 2021 / Revised: 18 November 2021 / Accepted: 20 November 2021 / Published: 30 November 2021

Abstract

:
Virus-like particles (VLPs) are a versatile, safe, and highly immunogenic vaccine platform. Recently, there are developmental vaccines targeting SARS-CoV-2, the causative agent of COVID-19. The COVID-19 pandemic affected humanity worldwide, bringing out incomputable human and financial losses. The race for better, more efficacious vaccines is happening almost simultaneously as the virus increasingly produces variants of concern (VOCs). The VOCs Alpha, Beta, Gamma, and Delta share common mutations mainly in the spike receptor-binding domain (RBD), demonstrating convergent evolution, associated with increased transmissibility and immune evasion. Thus, the identification and understanding of these mutations is crucial for the production of new, optimized vaccines. The use of a very flexible vaccine platform in COVID-19 vaccine development is an important feature that cannot be ignored. Incorporating the spike protein and its variations into VLP vaccines is a desirable strategy as the morphology and size of VLPs allows for better presentation of several different antigens. Furthermore, VLPs elicit robust humoral and cellular immune responses, which are safe, and have been studied not only against SARS-CoV-2 but against other coronaviruses as well. Here, we describe the recent advances and improvements in vaccine development using VLP technology.

1. Introduction

The SARS-CoV-2 (Severe Acute Respiratory Syndrome Coronavirus 2) is the causative agent of COVID-19 (Coronavirus Disease 2019) [1,2,3] and is responsible for the recent pandemic, which has already reported 248,467,363 cases and 5,027,183 deaths worldwide as of 5 November 2021 [4]. As new cases continue to increase worldwide, it is urgent to develop inexpensive and versatile vaccines to handle emerging variants that can affect pre-existing natural immunity and the efficacy of already approved vaccines [5]. According to the World Health Organization, 129 COVID-19 vaccines are under clinical trials, and eight are approved for emergency or definitive use worldwide, including inactivated, mRNA and viral vector vaccines (Table 1) [6,7]. Although we already have these available vaccines, there is still a need for improved versions of COVID-19 vaccines. Hence, adapting vaccines to variants of concern (VOCs) along with decreasing vaccine costs will be the goals for the next step towards eradication. A technology that has potential to address some of these issues is the virus-like particles (VLPs) vaccine platform, as it is adaptable, resembles viral structures, highly immunogenic, and can be less expensive than other platforms.
The VLPs are noninfectious nanoscale particles composed of single or multiple self-assembling viral or nonviral proteins, which mimic a native viral particle [27]. These particles, when used as immunogens, are captured and processed by antigen presenting cells (APCs) and presented by both MHC-I and MHC-II to T helper and Cytotoxic T lymphocytes (Figure 1A). The structural repetitiveness and particle size of VLPs enhance recognition and direct activation of B cells. [28,29,30,31]. Taken altogether, these properties lead to robust humoral and cellular immune responses, which are exciting for vaccination against infectious diseases [32,33,34,35,36,37,38].
VLPs are classified according to their structural composition (nonenveloped-neVLPs or enveloped-eVLPs) and to the native virus components (homologous or heterologous) [39] (Figure 1B). Homologous VLPs comprise particles that self-assemble using proteins derived from the native virus only [40]. On the other hand, heterologous VLPs contain proteins from different sources to increase immunogenicity [41]. Moreover, available bioinformatics tools can help to optimize the rational design of new and pre-existing VLPs to achieve the best immunogenic performance [32,42,43,44,45,46].
VLPs can be expressed in insect [32,47], mammalian [48,49], bacterial [50], plant [37,51] or yeast cells [33]. These particles have been used for the development of vaccines against several infectious diseases [52]. VLP-based vaccines for Hepatitis B (Engerix-B®), Hepatitis E (Hecolin®) and HPV (Cervarix®, Gardasil® and Gardasil 9®) demonstrate proof of concept for industrially scalable vaccines [33,53,54]. In addition to vaccine development, the empty shell-like structure of these nanoparticles has been studied for other biomedical purposes, such as gene therapy [53,54,55,56] and drug delivery [57,58,59,60,61,62].

2. SARS-CoV-2, VOCs, and Structural Vaccinology

The SARS-CoV-2 positive-sense single-stranded RNA genome (29 kb in length) encodes four structural and 16 non-structural proteins [3]. The structural proteins are the membrane (M), envelope (E), spike (S), and nucleocapsid (N) proteins, as seen in other coronaviruses (Figure 2A). The M protein (UNIPROT ID: P0DTC5) is a highly conserved homodimer and is the most abundant protein in mature virions [43,44]. It plays an essential role in viral assembly, membrane fusion, and particle size and shape, through interactions with other structural proteins [63,64,65,66]. The E protein (UNIPROT ID: P0DTC4) is a pentameric structure [67] found at low levels among coronaviruses [68,69]. It regulates M homodimerization [70] and, along with M, plays a central role in the vesicular transport of virions [71,72,73], ion transport [74,75,76,77,78], and pathogenesis [79,80,81]. The N protein (UNIPROT ID: P0DTC9) is a highly positive-charged, flexible, and unstable [58] phosphoprotein [82,83] that essentially performs genomic packaging [48,84,85]. The N-terminal domain of this protein [86,87] interacts with the viral RNA genome, and the C-terminal domain [84] forms a helical-shaped nucleocapsid [66,88,89]. Overall, the N protein interactions with M and viral RNA define the shape and size of the virions [85]. Unsurprisingly, N is essential in early viral replication and plays an integral role in infectivity [90].
Among these structural proteins, the main target for vaccine development is the SARS-CoV-2 S protein, which gives the characteristic crown-shaped structure of coronaviruses [88,89]. S is a highly glycosylated [93] homotrimer transmembrane protein (UNIPROT ID P0DTC2) composed of 1273 amino acids per chain. Known human sarbecoviruses (SARS-CoV-2 and SARS-CoV-1) and the alphacoronavirus NL63 invade the host cell through an interaction between the S protein and its receptor, the angiotensin-converting enzyme 2 (ACE2) [93,94,95,96,97] (Figure 2B). In general, the S protein consists of two major regions in addition to the signal peptide (SP) (1–12): the S1 subunit (13–685), and the S2 subunit (686–1273), which contains the transmembrane region (TM) (1214–1234) followed by the cytoplasmic domain (CD) (1235–1273) (Figure 3A). The S protein and the ACE2 receptor binding are mediated by the receptor-binding motif (RBM; 437–508), located in the receptor-binding domain (RBD; 319–541) [98] (Figure 3A, purple and cyan, respectively). The fusion machinery in S2 is composed of two fusion peptides (816–837 and 835–855) and two heptad regions (920–970 and 1163–1202). The first site of cleavage targeted by host proteases, such as furin and TMPRSS2, is located in the S1/S2 interface (685–686) [99,100,101] (Figure 3A, red). Removing the S1/S2 site promotes conformational changes that open the second cleavage site at S2 (815–816). The subsequent cleavage of the S2 site promotes the projection of needle-shaped fusion peptides into the host membrane [102,103], leading to cell fusion in 60–120 s in feline coronavirus [104]. The S protein presents a closed and open conformation [45,105] (Figure 3B, upper and bottom panel, respectively). With one or more RBDs projected outward, the open state constitutes the major conformation population of viable virions [105]. The increased exposure and steric freedom enable stronger interactions with the ACE2 receptor [45,106]. Therefore, mutations that stabilize this open conformation lead to positive selection, making the virus more transmissible [107,108,109].
A better understanding of the effects of these conformational changes allows us to closely monitor the emerging VOCs. The VOCs that have attracted the most attention so far are the Alpha, Beta, Gamma, and Delta variants [108], which were initially identified in the UK [109], South Africa [112], Brazil [113], and India [114], respectively (Figure 4).
The Alpha variant (B.1.1.7; main mutations in RBD: N501Y, P681H, 69/70 Δ) was first identified in the United Kingdom and was also the first VOC to receive global attention due to increased transmissibility up to 43–90% higher than the original strain [112,115]. Furthermore, the Alpha variant was associated with increased disease severity among younger patients [116]. Nevertheless, this VOC did not significantly affect the efficacy of available vaccines [117,118]. The Beta variant (B.1.351; main mutations in RBD: E484K, K417N, N501Y) was first detected in the Eastern Cape province of South Africa and was associated with decreased vaccine efficacy [18,112] and higher immune evasion than the Alpha variant [119,120]. The Gamma variant (P.1; main mutations in the RBD: K417T, E484K, N501Y, D614G) has become the primary concern in Brazil, as it is 1.7–2.4 fold more transmissible compared to the Alpha variant [113]. Indeed, the Gamma variant spread rapidly in the Amazonas, Brazil, and contributed to the collapse of the local health care system [121]. Interestingly, both the Gamma and Alpha variants have similar binding affinities between the RBD and ACE2, slightly stronger than the original strain [105]. Stronger receptor binding could explain the higher viral transmission observed in countries with the dominant circulating Gamma variant [122,123]. As mentioned above, these variants have contributed to the collapse of the public health system in many countries. However, recently, a new emerging variant (Delta) has become the highest global health concern as it was shown to affect immune responses [18,124,125] and is 55% more contagious than the Alpha variant [126]. The Delta variant (main mutations: T19R, G142D, E156G, F157Δ, R158Δ, L452R, T478K, D614G, P681R, and D950N) was first identified in India in December 2020 [114,127] and quickly became the local predominant VOC. The accumulation of convergent mutations on VOCs resulted in local variants during the COVID-19 s wave in India. As expected, due to the increased transmissibility of Delta, this variant soon became the predominant VOC in England [109], the United States [128,129], and other countries [18]. The linages of VOCs are defined by multiple convergent mutations that are hypothesized to have appeared in the condition of chronic COVID-19 infections in immunocompromised patients [130]. A recent study showed that patients who received the first dose of the BNT162b2 (Pfizer-BioNTech) and ChAdOx1 AZD1222 (Oxford-AstraZeneca) vaccines were not as protected against the Delta variant as those who received the second dose [131].
The VOCs have become a threat to pandemic control, and the vaccines need the help of biotechnology, especially structural vaccinology tools, to stay one step ahead of emerging variants. Structural vaccinology proposes the rational design and selection of antigens, including peptides, to maximize vaccine immunogenicity, safety, stability, and quality [43,132]. This critical process can accelerate the development of new vaccines through several bioinformatics, biophysical, and computational tools already available [133,134,135,136,137]. Such resources offer a rational approach to predict epitopes and measure their binding affinities and structural and dynamical properties [138,139].
Different laboratories have already reported potential B and T cell epitopes predicted based on the SARS-CoV-2 S protein that could be used as vaccine candidates, including VLPs vaccines [138,140,141]. For example, the UB-612 is a rationally designed vaccine candidate construct based on two antigenic components: an S1-RBD-sFc fusion protein and a synthetic polypeptide consisting of five SARS-CoV-2 derived peptides together with a universal peptide [132]. This alum- and CpG-adjuvanted vaccine candidate induced high titers of S1-RBD nAbs and increased Th1-oriented responses. In mice, UB-612 reduced the viral load and prevented the development of the disease in a live SARS-CoV-2 challenge. UB-612 is being developed by Vaxxinity, which is currently conducting clinical trials in Taiwan. Another successful example employing structural vaccinology is the use of a well-known structural optimization that improves the expression and stability of the S protein in MERS-CoV and was adapted to SARS-CoV-2 protein-based vaccines [45]. Two consecutive substitutions of proline residues at 986–987 on the S2 subunit (S2P) and mutations at the cleavage site S1/S2 (682–685, ‘RRAR’ to ‘GSAS’) increased protein expression and prevented host proteolytic cleavage, locking the S protein in the prefusion state [45,142]. The S2P mutations are being used in the available Pfizer-BioNTech vaccine (Comirnaty® BNT162b2) [25], which showed high efficacy (95%) [26] against symptomatic cases and a substantial reduction of hospitalizations and deaths in the United Kingdom, Israel, and other countries [143,144,145]. Other studies have shown that point mutations at residues T572I, D614N, A892P, A942P and V987P lock S in the closed state and promote a 6.4-fold increase in protein expression and thermal stability [115]. Despite the expression advantages, these mutations could make vaccines more accessible to developing countries [146].

3. Enveloped VLPs against SARS-CoV-2

As discussed above, incredible progress has been made on the understanding and improvement of SARS-CoV-2 immunogens. Significantly, the comprehensive analysis of the S protein characteristics and its antigenic regions were crucial for the rapid development of the SARS-CoV-2 vaccines. However, there is still room to test both existing and novel antigens using different platforms.
The VLP vaccine platform can be used for the production of eVLPs and neVLPs of interest [140,147]. Enveloped VLPs harbor a host-derived or synthetic membrane and generally need more complex expression systems, such as those in eukaryotic cell lines (mammalian, insect, or plant cells). For the formation of coronavirus eVLPs, structural proteins are essential for viral maturation and particle assembly [48,49,148]. Some studies have shown that the minimal requirement for the assembly of SARS-CoV-2 VLPs and other coronaviruses is the combination of M and either the E or N proteins. However, most particles include the N protein and the highly immunogenic S protein for better assembly and expression (Figure 5A) [48,49]. To date, Vero E6 cells presented the highest expression of S-containing VLPs when compared to HEK293 cells [49]. All of these initial approaches for SARS-CoV-2 VLP production show a promising use of this platform in vaccine development. However, industrial viability and large-scale production was not considered, and these are crucial features for further development and are still very challenging in the eVLP production field [100].
Although homologous VLPs are an attractive strategy for producing these particles, the combination of antigenic SARS-CoV-2 proteins with other highly expressed heterologous proteins (that could be used as alternative VLPs scaffolds) are an exciting strategy to address issues of industrial production. The NDVLP-S2P (Figure 5B) is a heterologous chimeric eVLP vaccine candidate against SARS-CoV-2 that uses the structure of a well-characterized enveloped virus, the Newcastle disease virus (NDV) [41], and is being developed by the National Institute of Allergy and Infectious Diseases (NIAID). The transmembrane domain of the NDV fusion protein was fused to SARS-CoV-2 S2P, allowing the correct display of S2P on the VLP surface [41,45,149]. The NDV-S2P VLPs were more immunogenic than the trimeric S protein alone, showing that the presentation of antigens on the surface of the VLPs is advantageous [41]. Another heterologous SARS-CoV-2 eVLPs vaccine candidate is the CoVLP from Medicago/GSK (Figure 5C) [37,51]. This vaccine is based on VLPs that display a mutated S2P protein, which comprises a plant signal peptide, GSAS substitutions in the S1/S2 site, and TM/CD regions of the Influenza H5 A/Indonesia/5/2005. The CoVLP vaccine is formulated with AS03 [141] and given in a two-dose regimen. After the second dose, immunized volunteers showed higher serum SARS-CoV-2 nAb titers than in convalescent plasma. This vaccine candidate is already in phase 3 clinical trials (NCT04636697). VBI-2902a [149] is an MLV-based eVLPs vaccine candidate containing the S protein in the prefusion state fused with the VSV-G transmembrane cytoplasmic domain (VSV-GTC) (Figure 5D). This vaccine is being developed by VBI Vaccines and is in ongoing clinical trials 1/2 (NCT04773665).
As discussed, all these approaches allow for easier incorporation of highly antigenic S proteins into the VLP scaffold, dramatically improving production and efficacy of the vaccine.

4. Non-Enveloped VLPs against SARS-CoV-2

Unlike eVLPs, neVLPs do not contain any lipid membranes. They can be produced in simpler expression systems, such as those using bacteria (i.e., Escherichia coli) and yeast (i.e., Saccharomyces cerevisiae and Pichia Pastoris) cells. The Hepatitis B virus vaccine, Engerix-B® [150], and the Human papillomavirus vaccine, Gardasil® and Gardasil 9® [33], are neVLPs-based vaccines approved by the FDA. They are produced in Saccharomyces cerevisiae, an efficient expression system that is industrially scalable and cheaper than mammalian and insect cell systems. Despite these clear advantages, bacteria and yeast cells lack complex post-translational modifications (PTM) needed to produce some proteins, such as the highly glycosylated SARS-CoV-2 S protein [75,151]. Thus, the choice of the best expression system could be a determinant for protein/VLP production, even considering neVLPs. Cervarix® is another HPV neVLP-based vaccine [152] which is highly immunogenic [153] and effective [154,155] against HPV types 16 and 18, which are the main serotypes that cause cervical cancer [148]. The Cervarix® vaccine is produced using insect cells infected with recombinant baculovirus [40,156,157].
An up-and-coming SARS-CoV-2 neVLP-based vaccine candidate is the Novavax NVAX-CoV2373 (Figure 5E) that consists of S2P protein (1–1273) locked in the prefusion state and expressed using a baculovirus/insect cell system [32]. Later, the recombinant S protein is incorporated into polysorbate 80 detergent. This vaccine is formulated in combination with Matrix-M® adjuvant [32]. The NVAX-CoV2373 vaccine was shown to be immunogenic and safe [35,36] and conferred 89.7% protection against symptomatic cases [158]. Using a different approach, the ABNCoV2 (Figure 5F) is a SARS-CoV-2 neVLPs-based vaccine candidate from the Radbound University and AdaptVac. A SpyTag-SpyCatcher was used to display the S RBD in a bacteriophage scaffold system [159]. The ABNCoV2 is another highly immunogenic candidate and induced high titers of nAbs in mice [159]. This vaccine candidate is currently in phase 1/2 clinical trials (NCT04839146).
Although these different approaches show great potential for developing a SARS-CoV-2 VLPs-based vaccine, some improvements could be made to design broad-spectrum vaccines against other coronaviruses of interest. A pancoronavirus vaccine is an exciting strategy against zoonotic coronaviruses that represent a threat to humans and may soon be responsible for the next pandemic [160,161].
It is already known that SARS-CoV-2 vaccines do not generate significant broadly neutralizing antibodies against other family members, supporting the need for a functional pancoronavirus vaccine [162]. Cohen and colleagues designed different mosaic nanoparticle vaccines based on the S RBD from several human and animal (bat and pangolin) sarbecoviruses, including the SARS-CoV-2 RBD (Figure 5G) [34]. Each vaccine induced strikingly higher cross-neutralizing antibodies than the SARS-CoV-2 S protein alone. Following the same idea, a Sortase A tagged ferritin-based VLPs conjugated to the SARS-CoV-2 S RBD (Figure 5H) was tested with alum adjuvants [163]. The protective immunity induced by this vaccine was robust against SARS-CoV-2 Alpha and Beta variants, as well as other human and bat sarbecoronaviruses. Interestingly, the cross-neutralizing antibody titers were more significant than the mRNA vaccine encoding the SARS-CoV-2 S protein. Despite this, the data suggest that mRNA and other S-based vaccines may afford some protection against other sarbecoronaviruses. Lastly, the candidate GBP510 (Figure 5I) is based on two rationally designed proteins, I53-50A and I53-50B [164,165], which self-assemble into trimers and pentamers, respectively. Combining these proteins forms a versatile 120 subunit scaffold, 28 nm wide, for the multivalent display of antigens linked to I53-50A subunits, such as SARS-CoV-2 RBD [156]. These particles were highly immunogenic in mice and resulted in 10-fold higher antibody titers at a 5-fold lower dose than the S2P protein alone [156]. In further preclinical trials with rhesus macaques, GBP510 was tested with different adjuvants, including alum, AS03, CpG, Essai O/W, and AS37. Although highly immunogenic against the original strain and the Alpha variant, a 4.5- to 16-fold reduction in neutralization was observed against the Beta variant [157]. GBP510 is already in phase I/II clinical trials (NCT04742738 and NCT04750343) sponsored by SK Bioscience Co. LTD and the Coalition for Epidemic Preparedness Innovations (CEPI).

5. Conclusions

As the cases of COVID-19 continue to grow and variants of SARS-CoV-2 emerge, the need for easily adaptable vaccine platforms remains one of the major goals in the pursuit of eradicating the disease. VLPs are an attractive platform for developing vaccines against infectious diseases such as COVID-19, due to their impressive versatility and immunological applications. The main advantage of using this vaccine platform is that VLPs can closely resemble the native virus in their size, shape and antigen display. Additionally, VLPs can be adapted to contain more than one antigen of interest, a very important feature to produce efficient and broad vaccines.
Several VLP-based vaccines have been tested in human clinical trials since the approval of the Hecolin®, Cervarix®, Gardasil® and Gardasil 9® vaccines. These vaccines consolidated this platform, and opened doors for its use to produce vaccines against several diseases, including COVID-19. Here, we discussed nine VLP-based vaccines that have been developed against SARS-CoV-2, among which five are already undergoing clinical trials. The main antigenic component of these vaccines is the S protein, specifically the RBD, which is involved in viral entry and antibody neutralization, especially in its closed conformation. By displaying multiple full-length Spikes or only its RBD on the VLPs surfaces, especially on neVLPs, these vaccines have shown high immunogenicity in clinical trials, generating neutralizing antibodies, and inducing cell-mediated responses. However, we discussed the importance of considering additional structural proteins as antigens in eVLPs-based vaccines, as they lead to more complex and native-like structures, and also generate broad-spectrum immunity against variants and other coronaviruses. As seen with other available vaccines, rational and adaptable vaccine development directly lead to a reduction of hospitalizations, deaths and spread. The recent advances in VLP production and structural vaccinology provide this platform with all the needed features to produce novel, potent, and broad-spectrum coronavirus vaccines that could help in the fight against current and future pandemics.

Author Contributions

W.A.P.-S., L.C.S.C.—Conceptualization and manuscript writing; K.P.C., L.V., V.K., A.L., N.C.—Manuscript contributions; F.E.G.G., F.L.B.d.S., O.C.-M., J.A.M.B., M.S., M.R., N.O.S.C.—Review and editing; G.C.-M.—Supervision, editing, and approval of the final version. All authors have read and agreed to the published version of the manuscript.

Funding

This research group is funded by The São Paulo Research Foundation, FAPESP-Grants 2017/10910-5, 2019/14526-0, 2020/07158-2 and 2020/05146-7. This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior-Brasil (CAPES)-Finance Code 001- and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq 305393/2020-0).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no competing interests.

References

  1. Zhu, N.; Zhang, D.; Wang, W.; Li, X.; Yang, B.; Song, J.; Zhao, X.; Huang, B.; Shi, W.; Lu, R.; et al. A Novel Coronavirus from Patients with Pneumonia in China, 2019. N. Engl. J. Med. 2020, 382, 727–733. [Google Scholar] [CrossRef]
  2. Zhou, P.; Yang, X.-L.; Wang, X.-G.; Hu, B.; Zhang, L.; Zhang, W.; Si, H.-R.; Zhu, Y.; Li, B.; Huang, C.-L.; et al. A Pneumonia Outbreak Associated with a New Coronavirus of Probable Bat Origin. Nature 2020, 579, 270–273. [Google Scholar] [CrossRef] [Green Version]
  3. Lu, R.; Zhao, X.; Li, J.; Niu, P.; Yang, B.; Wu, H.; Wang, W.; Song, H.; Huang, B.; Zhu, N.; et al. Genomic Characterisation and Epidemiology of 2019 Novel Coronavirus: Implications for Virus Origins and Receptor Binding. Lancet 2020, 395, 565–574. [Google Scholar] [CrossRef] [Green Version]
  4. Weekly Epidemiological Update on COVID-19—22 June 2021. Available online: https://www.who.int/publications/m/item/weekly-epidemiological-update-on-covid-19---22-june-2021 (accessed on 23 June 2021).
  5. Li, Y.; Tenchov, R.; Smoot, J.; Liu, C.; Watkins, S.; Zhou, Q. A Comprehensive Review of the Global Efforts on COVID-19 Vaccine Development. ACS Cent. Sci. 2021, 7, 512–533. [Google Scholar] [CrossRef] [PubMed]
  6. WHO Issues Emergency Use Listing for Eighth COVID-19 Vaccine. Available online: https://www.who.int/news/item/03-11-2021-who-issues-emergency-use-listing-for-eighth-covid-19-vaccine (accessed on 7 November 2021).
  7. COVID-19 Vaccines. Available online: https://www.who.int/emergencies/diseases/novel-coronavirus-2019/covid-19-vaccines (accessed on 7 November 2021).
  8. Gao, Q.; Bao, L.; Mao, H.; Wang, L.; Xu, K.; Yang, M.; Li, Y.; Zhu, L.; Wang, N.; Lv, Z.; et al. Development of an Inactivated Vaccine Candidate for SARS-CoV-2. Science 2020, 369, 77–81. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Zeng, G.; Pan, H.; Li, C.; Hu, Y.; Chu, K.; Han, W.; Chen, Z.; Tang, R.; Yin, W.; et al. Safety, Tolerability, and Immunogenicity of an Inactivated SARS-CoV-2 Vaccine in Healthy Adults Aged 18–59 Years: A Randomised, Double-Blind, Placebo-Controlled, Phase 1/2 Clinical Trial. Lancet Infect. Dis. 2021, 21, 181–192. [Google Scholar] [CrossRef]
  10. Bueno, S.M.; Abarca, K.; González, P.A.; Gálvez, N.M.; Soto, J.A.; Duarte, L.F.; Schultz, B.M.; Pacheco, G.A.; González, L.A.; Vázquez, Y.; et al. Interim Report: Safety and Immunogenicity of an Inactivated Vaccine against SARS-CoV-2 in Healthy Chilean Adults in a Phase 3 Clinical Trial. medRxiv 2021. [Google Scholar] [CrossRef]
  11. Tanriover, M.D.; Doğanay, H.L.; Akova, M.; Güner, H.R.; Azap, A.; Akhan, S.; Köse, Ş.; Erdinç, F.Ş.; Akalın, E.H.; Tabak, Ö.F.; et al. Efficacy and Safety of an Inactivated Whole-Virion SARS-CoV-2 Vaccine (CoronaVac): Interim Results of a Double-Blind, Randomised, Placebo-Controlled, Phase 3 Trial in Turkey. Lancet 2021, 398, 213–222. [Google Scholar] [CrossRef]
  12. Xia, S.; Duan, K.; Zhang, Y.; Zhao, D.; Zhang, H.; Xie, Z.; Li, X.; Peng, C.; Zhang, Y.; Zhang, W.; et al. Effect of an Inactivated Vaccine Against SARS-CoV-2 on Safety and Immunogenicity Outcomes: Interim Analysis of 2 Randomized Clinical Trials. JAMA 2020, 324, 951–960. [Google Scholar] [CrossRef]
  13. Al Kaabi, N.; Zhang, Y.; Xia, S.; Yang, Y.; Al Qahtani, M.M.; Abdulrazzaq, N.; Al Nusair, M.; Hassany, M.; Jawad, J.S.; Abdalla, J.; et al. Effect of 2 Inactivated SARS-CoV-2 Vaccines on Symptomatic COVID-19 Infection in Adults: A Randomized Clinical Trial. JAMA 2021, 326, 35–45. [Google Scholar] [CrossRef] [PubMed]
  14. Ella, R.; Reddy, S.; Blackwelder, W.; Potdar, V.; Yadav, P.; Sarangi, V.; Aileni, V.K.; Kanungo, S.; Rai, S.; Reddy, P.; et al. Efficacy, Safety, and Lot to Lot Immunogenicity of an Inactivated SARS-CoV-2 Vaccine (BBV152): A, Double-Blind, Randomised, Controlled Phase 3 Trial. medRxiv 2021. [Google Scholar] [CrossRef]
  15. Ella, R.; Vadrevu, K.M.; Jogdand, H.; Prasad, S.; Reddy, S.; Sarangi, V.; Ganneru, B.; Sapkal, G.; Yadav, P.; Abraham, P.; et al. Safety and Immunogenicity of an Inactivated SARS-CoV-2 Vaccine, BBV152: A Double-Blind, Randomised, Phase 1 Trial. Lancet. Infect. Dis. 2021, 21, 637–646. [Google Scholar] [CrossRef]
  16. Folegatti, P.M.; Ewer, K.J.; Aley, P.K.; Angus, B.; Becker, S.; Belij-Rammerstorfer, S.; Bellamy, D.; Bibi, S.; Bittaye, M.; Clutterbuck, E.A.; et al. Safety and Immunogenicity of the ChAdOx1 NCoV-19 Vaccine against SARS-CoV-2: A Preliminary Report of a Phase 1/2, Single-Blind, Randomised Controlled Trial. Lancet 2020, 396, 467–478. [Google Scholar] [CrossRef]
  17. Falsey, A.R.; Sobieszczyk, M.E.; Hirsch, I.; Sproule, S.; Robb, M.L.; Corey, L.; Neuzil, K.M.; Hahn, W.; Hunt, J.; Mulligan, M.J.; et al. Phase 3 Safety and Efficacy of AZD1222 (ChAdOx1 NCoV-19) COVID-19 Vaccine. N. Engl. J. Med. 2021. [Google Scholar] [CrossRef]
  18. Voysey, M.; Clemens, S.A.C.; Madhi, S.A.; Weckx, L.Y.; Folegatti, P.M.; Aley, P.K.; Angus, B.; Baillie, V.L.; Barnabas, S.L.; Bhorat, Q.E.; et al. Safety and Efficacy of the ChAdOx1 NCoV-19 Vaccine (AZD1222) against SARS-CoV-2: An Interim Analysis of Four Randomised Controlled Trials in Brazil, South Africa, and the UK. Lancet 2021, 397, 99–111. [Google Scholar] [CrossRef]
  19. Sadoff, J.; Le Gars, M.; Shukarev, G.; Heerwegh, D.; Truyers, C.; de Groot, A.M.; Stoop, J.; Tete, S.; Van Damme, W.; Leroux-Roels, I.; et al. Interim Results of a Phase 1–2a Trial of Ad26.COV2.S COVID-19 Vaccine. N. Engl. J. Med. 2021, 384, 1824–1835. [Google Scholar] [CrossRef]
  20. Bos, R.; Rutten, L.; van der Lubbe, J.E.M.; Bakkers, M.J.G.; Hardenberg, G.; Wegmann, F.; Zuijdgeest, D.; de Wilde, A.H.; Koornneef, A.; Verwilligen, A.; et al. Ad26 Vector-Based COVID-19 Vaccine Encoding a Prefusion-Stabilized SARS-CoV-2 Spike Immunogen Induces Potent Humoral and Cellular Immune Responses. NPJ Vaccines 2020, 5, 1–11. [Google Scholar] [CrossRef]
  21. Mercado, N.B.; Zahn, R.; Wegmann, F.; Loos, C.; Chandrashekar, A.; Yu, J.; Liu, J.; Peter, L.; McMahan, K.; Tostanoski, L.H.; et al. Single-Shot Ad26 Vaccine Protects against SARS-CoV-2 in Rhesus Macaques. Nature 2020, 586, 583–588. [Google Scholar] [CrossRef]
  22. Sadoff, J.; Gray, G.; Vandebosch, A.; Cárdenas, V.; Shukarev, G.; Grinsztejn, B.; Goepfert, P.A.; Truyers, C.; Fennema, H.; Spiessens, B.; et al. Safety and Efficacy of Single-Dose Ad26.COV2.S Vaccine against COVID-19. N. Engl. J. Med. 2021, 384, 2187–2201. [Google Scholar] [CrossRef]
  23. Anderson, E.J.; Rouphael, N.G.; Widge, A.T.; Jackson, L.A.; Roberts, P.C.; Makhene, M.; Chappell, J.D.; Denison, M.R.; Stevens, L.J.; Pruijssers, A.J.; et al. Safety and Immunogenicity of SARS-CoV-2 MRNA-1273 Vaccine in Older Adults. N. Engl. J. Med. 2020, 383, 2427–2438. [Google Scholar] [CrossRef]
  24. Baden, L.R.; El Sahly, H.M.; Essink, B.; Kotloff, K.; Frey, S.; Novak, R.; Diemert, D.; Spector, S.A.; Rouphael, N.; Creech, C.B.; et al. Efficacy and Safety of the MRNA-1273 SARS-CoV-2 Vaccine. N. Engl. J. Med. 2021, 384, 403–416. [Google Scholar] [CrossRef]
  25. Walsh, E.E.; Frenck, R.W.; Falsey, A.R.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Neuzil, K.; Mulligan, M.J.; Bailey, R.; et al. Safety and Immunogenicity of Two RNA-Based COVID-19 Vaccine Candidates. N. Engl. J. Med. 2020, 383, 2439–2450. [Google Scholar] [CrossRef]
  26. Polack, F.P.; Thomas, S.J.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Perez, J.L.; Pérez Marc, G.; Moreira, E.D.; Zerbini, C.; et al. Safety and Efficacy of the BNT162b2 MRNA COVID-19 Vaccine. N. Engl. J. Med. 2020, 383, 2603–2615. [Google Scholar] [CrossRef]
  27. Mohsen, M.O.; Augusto, G.; Bachmann, M.F. The 3Ds in Virus-like Particle Based-vaccines: “Design, Delivery and Dynamics”. Immunol. Rev. 2020, 296, 155–168. [Google Scholar] [CrossRef]
  28. Bachmann, M.; Rohrer, U.; Kundig, T.; Burki, K.; Hengartner, H.; Zinkernagel, R. The Influence of Antigen Organization on B Cell Responsiveness. Science 1993, 262, 1448–1451. [Google Scholar] [CrossRef]
  29. Cubas, R.; Zhang, S.; Kwon, S.; Sevick-Muraca, E.M.; Li, M.; Chen, C.; Yao, Q. Virus-like Particle (VLP) Lymphatic Trafficking and Immune Response Generation After Immunization by Different Routes. J. Immunother. 2009, 32, 118–128. [Google Scholar] [CrossRef] [Green Version]
  30. Mohsen, M.; Gomes, A.; Vogel, M.; Bachmann, M. Interaction of Viral Capsid-Derived Virus-Like Particles (VLPs) with the Innate Immune System. Vaccines 2018, 6, 37. [Google Scholar] [CrossRef] [Green Version]
  31. Win, S.J.; Ward, V.K.; Dunbar, P.R.; Young, S.L.; Baird, M.A. Cross-presentation of Epitopes on Virus-like Particles via the MHC I Receptor Recycling Pathway. Immunol. Cell Biol. 2011, 89, 681–688. [Google Scholar] [CrossRef]
  32. Bangaru, S.; Ozorowski, G.; Turner, H.L.; Antanasijevic, A.; Huang, D.; Wang, X.; Torres, J.L.; Diedrich, J.K.; Tian, J.-H.; Portnoff, A.D.; et al. Structural Analysis of Full-Length SARS-CoV-2 Spike Protein from an Advanced Vaccine Candidate. Science 2020, 370, 1089–1094. [Google Scholar] [CrossRef]
  33. Block, S.L.; Nolan, T.; Sattler, C.; Barr, E.; Giacoletti, K.E.D.; Marchant, C.D.; Castellsagué, X.; Rusche, S.A.; Lukac, S.; Bryan, J.T.; et al. Comparison of the Immunogenicity and Reactogenicity of a Prophylactic Quadrivalent Human Papillomavirus (Types 6, 11, 16, and 18) L1 Virus-Like Particle Vaccine in Male and Female Adolescents and Young Adult Women. Pediatrics 2006, 118, 2135–2145. [Google Scholar] [CrossRef] [Green Version]
  34. Cohen, A.A.; Gnanapragasam, P.N.P.; Lee, Y.E.; Hoffman, P.R.; Ou, S.; Kakutani, L.M.; Keeffe, J.R.; Wu, H.-J.; Howarth, M.; West, A.P.; et al. Mosaic Nanoparticles Elicit Cross-Reactive Immune Responses to Zoonotic Coronaviruses in Mice. Science 2021, 371, 735–741. [Google Scholar] [CrossRef] [PubMed]
  35. Keech, C.; Albert, G.; Cho, I.; Robertson, A.; Reed, P.; Neal, S.; Plested, J.S.; Zhu, M.; Cloney-Clark, S.; Zhou, H.; et al. Phase 1–2 Trial of a SARS-CoV-2 Recombinant Spike Protein Nanoparticle Vaccine. N. Engl. J. Med. 2020, 383, 2320–2332. [Google Scholar] [CrossRef]
  36. Tian, J.H.; Patel, N.; Haupt, R.; Zhou, H.; Weston, S.; Hammond, H.; Logue, J.; Portnoff, A.D.; Norton, J.; Guebre-Xabier, M.; et al. SARS-CoV-2 Spike Glycoprotein Vaccine Candidate NVX-CoV2373 Immunogenicity in Baboons and Protection in Mice. Nat. Commun. 2021, 12, 372. [Google Scholar] [CrossRef]
  37. Ward, B.J.; Makarkov, A.; Séguin, A.; Pillet, S.; Trépanier, S.; Dhaliwall, J.; Libman, M.D.; Vesikari, T.; Landry, N. Efficacy, Immunogenicity, and Safety of a Plant-Derived, Quadrivalent, Virus-like Particle Influenza Vaccine in Adults (18–64 Years) and Older Adults (≥65 Years): Two Multicentre, Randomised Phase 3 Trials. Lancet 2020, 396, 1491–1503. [Google Scholar] [CrossRef]
  38. Braun, M.; Jandus, C.; Maurer, P.; Hammann-Haenni, A.; Schwarz, K.; Bachmann, M.F.; Speiser, D.E.; Romero, P. Virus-like Particles Induce Robust Human T-Helper Cell Responses: Cellular Immune Response. Eur. J. Immunol. 2012, 42, 330–340. [Google Scholar] [CrossRef]
  39. Nooraei, S.; Bahrulolum, H.; Hoseini, Z.S.; Katalani, C.; Hajizade, A.; Easton, A.J.; Ahmadian, G. Virus-like Particles: Preparation, Immunogenicity and Their Roles as Nanovaccines and Drug Nanocarriers. J. Nanobiotechnol. 2021, 19, 59. [Google Scholar] [CrossRef]
  40. Li, M.; Cripe, T.P.; Estes, P.A.; Lyon, M.K.; Rose, R.C.; Garcea, R.L. Expression of the Human Papillomavirus Type 11 L1 Capsid Protein in Escherichia Coli: Characterization of Protein Domains Involved in DNA Binding and Capsid Assembly. J. Virol. 1997, 71, 2988–2995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Yang, Y.; Shi, W.; Abiona, O.M.; Nazzari, A.; Olia, A.S.; Ou, L.; Phung, E.; Stephens, T.; Tsybovsky, Y.; Verardi, R.; et al. Newcastle Disease Virus-Like Particles Displaying Prefusion-Stabilized SARS-CoV-2 Spikes Elicit Potent Neutralizing Responses. Vaccines 2021, 9, 73. [Google Scholar] [CrossRef] [PubMed]
  42. Brouwer, P.J.M.; Caniels, T.G.; van der Straten, K.; Snitselaar, J.L.; Aldon, Y.; Bangaru, S.; Torres, J.L.; Okba, N.M.A.; Claireaux, M.; Kerster, G.; et al. Potent Neutralizing Antibodies from COVID-19 Patients Define Multiple Targets of Vulnerability. Science 2020, 369, 643–650. [Google Scholar] [CrossRef]
  43. Dormitzer, P.R.; Ulmer, J.B.; Rappuoli, R. Structure-Based Antigen Design: A Strategy for next Generation Vaccines. Trends Biotechnol. 2008, 26, 659–667. [Google Scholar] [CrossRef]
  44. Liu, Y.; Liu, J.; Xia, H.; Zhang, X.; Fontes-Garfias, C.R.; Swanson, K.A.; Cai, H.; Sarkar, R.; Chen, W.; Cutler, M.; et al. Neutralizing Activity of BNT162b2-Elicited Serum. N. Engl. J. Med. 2021, 384, 1466–1468. [Google Scholar] [CrossRef] [PubMed]
  45. Pallesen, J.; Wang, N.; Corbett, K.S.; Wrapp, D.; Kirchdoerfer, R.N.; Turner, H.L.; Cottrell, C.A.; Becker, M.M.; Wang, L.; Shi, W.; et al. Immunogenicity and Structures of a Rationally Designed Prefusion MERS-CoV Spike Antigen. Proc. Natl. Acad. Sci. USA 2017, 114, E7348–E7357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Wec, A.Z.; Wrapp, D.; Herbert, A.S.; Maurer, D.P.; Haslwanter, D.; Sakharkar, M.; Jangra, R.K.; Dieterle, M.E.; Lilov, A.; Huang, D.; et al. Broad Neutralization of SARS-Related Viruses by Human Monoclonal Antibodies. Science 2020, 369, 731–736. [Google Scholar] [CrossRef] [PubMed]
  47. LCS, C.; BM, R.; GW, B. Production of GP64-Free Virus-like Particles from Baculovirus-Infected Insect Cells. J. Gen. Virol. 2018, 99, 265–274. [Google Scholar] [CrossRef]
  48. Plescia, C.B.; David, E.A.; Patra, D.; Sengupta, R.; Amiar, S.; Su, Y.; Stahelin, R. V SARS-CoV-2 Viral Budding and Entry Can Be Modeled Using BSL-2 Level Virus-like Particles. J. Biol. Chem. 2021, 296, 100103. [Google Scholar] [CrossRef] [PubMed]
  49. Xu, R.; Shi, M.; Li, J.; Song, P.; Li, N. Corrigendum: Construction of SARS-CoV-2 Virus-Like Particles by Mammalian Expression System. Front. Bioeng. Biotechnol. 2020, 8, 1026. [Google Scholar] [CrossRef] [PubMed]
  50. Atcheson, E.; Cabral-Miranda, G.; Salman, A.M.; Reyes-Sandoval, A. Discovery of Four New B-Cell Protective Epitopes for Malaria Using Q Beta Virus-like Particle as Platform. NPJ Vaccines 2020, 5, 92. [Google Scholar] [CrossRef] [PubMed]
  51. Ward, B.J.; Gobeil, P.; Séguin, A.; Atkins, J.; Boulay, I.; Charbonneau, P.-Y.; Couture, M.; D’Aoust, M.-A.; Dhaliwall, J.; Finkle, C.; et al. Phase 1 Randomized Trial of a Plant-Derived Virus-like Particle Vaccine for COVID-19. Nat. Med. 2021, 27, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  52. Qian, C.; Liu, X.; Xu, Q.; Wang, Z.; Chen, J.; Li, T.; Zheng, Q.; Yu, H.; Gu, Y.; Li, S.; et al. Recent Progress on the Versatility of Virus-Like Particles. Vaccines 2020, 8, 139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lee, E.B.; Kim, J.-H.; Hur, W.; Choi, J.E.; Kim, S.M.; Park, D.J.; Kang, B.-Y.; Lee, G.W.; Yoon, S.K. Liver-Specific Gene Delivery Using Engineered Virus-Like Particles of Hepatitis E Virus. Sci. Rep. 2019, 9, 1616. [Google Scholar] [CrossRef] [PubMed]
  54. Pang, H.-H.; Chen, P.-Y.; Wei, K.-C.; Huang, C.-W.; Shiue, Y.-L.; Huang, C.-Y.; Yang, H.-W. Convection-Enhanced Delivery of a Virus-Like Nanotherapeutic Agent with Dual-Modal Imaging for Besiegement and Eradication of Brain Tumors. Theranostics 2019, 9, 1752–1763. [Google Scholar] [CrossRef] [PubMed]
  55. Lam, P.; Steinmetz, N.F. Delivery of SiRNA Therapeutics Using Cowpea Chlorotic Mottle Virus-like Particles. Biomater. Sci. 2019, 7, 3138–3142. [Google Scholar] [CrossRef] [PubMed]
  56. Azizgolshani, O.; Garmann, R.F.; Cadena-Nava, R.; Knobler, C.M.; Gelbart, W.M. Reconstituted Plant Viral Capsids Can Release Genes to Mammalian Cells. Virology 2013, 441, 12–17. [Google Scholar] [CrossRef] [Green Version]
  57. Crooke, S.N.; Schimer, J.; Raji, I.; Wu, B.; Oyelere, A.K.; Finn, M.G. Lung Tissue Delivery of Virus-Like Particles Mediated by Macrolide Antibiotics. Mol. Pharm. 2019, 16, 2947–2955. [Google Scholar] [CrossRef] [PubMed]
  58. Chang, C.; Sue, S.-C.; Yu, T.; Hsieh, C.-M.; Tsai, C.-K.; Chiang, Y.-C.; Lee, S.; Hsiao, H.; Wu, W.-J.; Chang, W.-L.; et al. Modular Organization of SARS Coronavirus Nucleocapsid Protein. J. Biomed. Sci. 2006, 13, 59–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Yacoby, I.; Shamis, M.; Bar, H.; Shabat, D.; Benhar, I. Targeting Antibacterial Agents by Using Drug-Carrying Filamentous Bacteriophages. Antimicrob. Agents Chemother. 2006, 50, 2087–2097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Dashti, N.H.; Abidin, R.S.; Sainsbury, F. Programmable In Vitro Coencapsidation of Guest Proteins for Intracellular Delivery by Virus-like Particles. ACS Nano 2018, 12, 4615–4623. [Google Scholar] [CrossRef] [PubMed]
  61. Koyani, R.; Pérez-Robles, J.; Cadena-Nava, R.D.; Vazquez-Duhalt, R. Biomaterial-Based Nanoreactors, an Alternative for Enzyme Delivery. Nanotechnol. Rev. 2017, 6, 405–419. [Google Scholar] [CrossRef]
  62. Sánchez-Sánchez, L.; Cadena-Nava, R.D.; Palomares, L.A.; Ruiz-Garcia, J.; Koay, M.S.T.; Cornelissen, J.J.M.T.; Vazquez-Duhalt, R. Chemotherapy Pro-Drug Activation by Biocatalytic Virus-like Nanoparticles Containing Cytochrome P450. Enzyme Microb. Technol. 2014, 60, 24–31. [Google Scholar] [CrossRef] [PubMed]
  63. Fehr, A.R.; Perlman, S. Coronaviruses: An Overview of Their Replication and Pathogenesis. In Coronaviruses; Maier, H.J., Bickerton, E., Britton, P., Eds.; Springer: New York, NY, USA, 2015; Volume 1282, pp. 1–23. ISBN 978-1-4939-2437-0. [Google Scholar]
  64. Neuman, B.W.; Joseph, J.S.; Saikatendu, K.S.; Serrano, P.; Chatterjee, A.; Johnson, M.A.; Liao, L.; Klaus, J.P.; Yates, J.R.; Wuthrich, K.; et al. Proteomics Analysis Unravels the Functional Repertoire of Coronavirus Nonstructural Protein 3. J. Virol. 2008, 82, 5279–5294. [Google Scholar] [CrossRef] [Green Version]
  65. de Haan, C.A.; Vennema, H.; Rottier, P.J. Assembly of the Coronavirus Envelope: Homotypic Interactions between the M Proteins. J. Virol. 2000, 74, 4967–4978. [Google Scholar] [CrossRef] [PubMed]
  66. Godeke, G.-J.; de Haan, C.A.M.; Rossen, J.W.A.; Vennema, H.; Rottier, P.J.M. Assembly of Spikes into Coronavirus Particles Is Mediated by the Carboxy-Terminal Domain of the Spike Protein. J. Virol. 2000, 74, 1566–1571. [Google Scholar] [CrossRef] [Green Version]
  67. Pervushin, K.; Tan, E.; Parthasarathy, K.; Lin, X.; Jiang, F.L.; Yu, D.; Vararattanavech, A.; Soong, T.W.; Liu, D.X.; Torres, J. Structure and Inhibition of the SARS Coronavirus Envelope Protein Ion Channel. PLoS Pathog. 2009, 5, e1000511. [Google Scholar] [CrossRef] [PubMed]
  68. Godet, M.; L’Haridon, R.; Vautherot, J.-F.; Laude, H. TGEV Corona Virus ORF4 Encodes a Membrane Protein That Is Incorporated into Virions. Virology 1992, 188, 666–675. [Google Scholar] [CrossRef]
  69. Liu, D.X.; Inglis, S.C. Association of the Infectious Bronchitis Virus 3c Protein with the Virion Envelope. Virology 1991, 185, 911–917. [Google Scholar] [CrossRef]
  70. Boscarino, J.A.; Logan, H.L.; Lacny, J.J.; Gallagher, T.M. Envelope Protein Palmitoylations Are Crucial for Murine Coronavirus Assembly. J. Virol. 2008, 82, 2989–2999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Machamer, C.E.; Youn, S. The transmembrane domain of the infectious bronchitis virus E protein is required for efficient virus release. In Advances in Experimental Medicine and Biology; Springer: New York, NY, USA, 2006; Volume 581, pp. 193–198. ISBN 9780387262024. [Google Scholar]
  72. Ruch, T.R.; Machamer, C.E. The Hydrophobic Domain of Infectious Bronchitis Virus E Protein Alters the Host Secretory Pathway and Is Important for Release of Infectious Virus. J. Virol. 2011, 85, 675–685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ruch, T.R.; Machamer, C.E. A Single Polar Residue and Distinct Membrane Topologies Impact the Function of the Infectious Bronchitis Coronavirus E Protein. PLoS Pathog. 2012, 8, e1002674. [Google Scholar] [CrossRef] [Green Version]
  74. Liao, Y.; Lescar, J.; Tam, J.P.; Liu, D.X. Expression of SARS-Coronavirus Envelope Protein in Escherichia Coli Cells Alters Membrane Permeability. Biochem. Biophys. Res. Commun. 2004, 325, 374–380. [Google Scholar] [CrossRef] [PubMed]
  75. Kumar, S.; Maurya, V.K.; Prasad, A.K.; Bhatt, M.L.B.; Saxena, S.K. Structural, Glycosylation and Antigenic Variation between 2019 Novel Coronavirus (2019-NCoV) and SARS Coronavirus (SARS-CoV). Virusdisease 2020, 31, 13–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wilson, L.; Mckinlay, C.; Gage, P.; Ewart, G. SARS Coronavirus E Protein Forms Cation-Selective Ion Channels. Virology 2004, 330, 322–331. [Google Scholar] [CrossRef] [Green Version]
  77. Singh Tomar, P.P.; Arkin, I.T. SARS-CoV-2 E Protein Is a Potential Ion Channel That Can Be Inhibited by Gliclazide and Memantine. Biochem. Biophys. Res. Commun. 2020, 530, 10–14. [Google Scholar] [CrossRef]
  78. Schoeman, D.; Fielding, B.C. Coronavirus Envelope Protein: Current Knowledge. Virol. J. 2019, 16, 69. [Google Scholar] [CrossRef] [Green Version]
  79. DeDiego, M.L.; Pewe, L.; Alvarez, E.; Rejas, M.T.; Perlman, S.; Enjuanes, L. Pathogenicity of Severe Acute Respiratory Coronavirus Deletion Mutants in HACE-2 Transgenic Mice. Virology 2008, 376, 379–389. [Google Scholar] [CrossRef] [Green Version]
  80. Nieto-Torres, J.L.; Verdiá-Báguena, C.; Jimenez-Guardeño, J.M.; Regla-Nava, J.A.; Castaño-Rodriguez, C.; Fernandez-Delgado, R.; Torres, J.; Aguilella, V.M.; Enjuanes, L. Severe Acute Respiratory Syndrome Coronavirus E Protein Transports Calcium Ions and Activates the NLRP3 Inflammasome. Virology 2015, 485, 330–339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Teoh, K.-T.; Siu, Y.-L.; Chan, W.-L.; Schlüter, M.A.; Liu, C.-J.; Peiris, J.S.M.; Bruzzone, R.; Margolis, B.; Nal, B. The SARS Coronavirus E Protein Interacts with PALS1 and Alters Tight Junction Formation and Epithelial Morphogenesis. Mol. Biol. Cell 2010, 21, 3838–3852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. White, T.C.; Yi, Z.; Hogue, B.G. Identification of Mouse Hepatitis Coronavirus A59 Nucleocapsid Protein Phosphorylation Sites. Virus Res. 2007, 126, 139–148. [Google Scholar] [CrossRef]
  83. Li, F.Q.; Xiao, H.; Tam, J.P.; Liu, D.X. Sumoylation of the Nucleocapsid Protein of Severe Acute Respiratory Syndrome Coronavirus. FEBS Lett. 2005, 579, 2387–2396. [Google Scholar] [CrossRef] [Green Version]
  84. Chang, C.; Chen, C.-M.M.; Chiang, M.; Hsu, Y.; Huang, T. Transient Oligomerization of the SARS-CoV N Protein–Implication for Virus Ribonucleoprotein Packaging. PLoS ONE 2013, 8, e65045. [Google Scholar] [CrossRef]
  85. Kuo, L.; Koetzner, C.A.; Hurst, K.R.; Masters, P.S. Recognition of the Murine Coronavirus Genomic RNA Packaging Signal Depends on the Second RNA-Binding Domain of the Nucleocapsid Protein. J. Virol. 2014, 88, 4451–4465. [Google Scholar] [CrossRef] [Green Version]
  86. Fan, H.; Ooi, A.; Tan, Y.W.; Wang, S.; Fang, S.; Liu, D.X.; Lescar, J. The Nucleocapsid Protein of Coronavirus Infectious Bronchitis Virus: Crystal Structure of Its N-Terminal Domain and Multimerization Properties. Structure 2005, 13, 1859–1868. [Google Scholar] [CrossRef] [Green Version]
  87. Kang, S.; Yang, M.; Hong, Z.; Zhang, L.; Huang, Z.; Chen, X.; He, S.; Zhou, Z.; Zhou, Z.; Chen, Q.; et al. Crystal Structure of SARS-CoV-2 Nucleocapsid Protein RNA Binding Domain Reveals Potential Unique Drug Targeting Sites. Acta Pharm. Sin. B 2020, 10, 1228–1238. [Google Scholar] [CrossRef]
  88. Yao, H.; Song, Y.; Chen, Y.; Wu, N.; Xu, J.; Sun, C.; Zhang, J.; Weng, T.; Zhang, Z.; Wu, Z.; et al. Molecular Architecture of the SARS-CoV-2 Virus. Cell 2020, 183, 730–738.e13. [Google Scholar] [CrossRef]
  89. Caldas, L.A.; Carneiro, F.A.; Higa, L.M.; Monteiro, F.L.; da Silva, G.P.; da Costa, L.J.; Durigon, E.L.; Tanuri, A.; de Souza, W. Ultrastructural Analysis of SARS-CoV-2 Interactions with the Host Cell via High Resolution Scanning Electron Microscopy. Sci. Rep. 2020, 10, 16099. [Google Scholar] [CrossRef]
  90. Hurst, K.R.; Ye, R.; Goebel, S.J.; Jayaraman, P.; Masters, P.S. An Interaction between the Nucleocapsid Protein and a Component of the Replicase-Transcriptase Complex Is Crucial for the Infectivity of Coronavirus Genomic RNA. J. Virol. 2010, 84, 10276–10288. [Google Scholar] [CrossRef] [Green Version]
  91. Gardner, A.; Autin, L.; Barbaro, B.; Olson, A.J.; Goodsell, D.S. CellPAINT: Interactive Illustration of Dynamic Mesoscale Cellular Environments. IEEE Comput. Graph. Appl. 2018, 38, 51–64. [Google Scholar] [CrossRef]
  92. Tomasello, G.; Armenia, I.; Molla, G. The Protein Imager: A Full-Featured Online Molecular Viewer Interface with Server-Side HQ-Rendering Capabilities. Bioinformatics 2020, 36, 2909–2911. [Google Scholar] [CrossRef]
  93. Zhao, P.; Praissman, J.L.; Grant, O.C.; Cai, Y.; Xiao, T.; Rosenbalm, K.E.; Aoki, K.; Kellman, B.P.; Bridger, R.; Barouch, D.H.; et al. Virus-Receptor Interactions of Glycosylated SARS-CoV-2 Spike and Human ACE2 Receptor. Cell Host Microbe 2020, 28, 586–601.e6. [Google Scholar] [CrossRef]
  94. Muus, C.; Luecken, M.D.; Eraslan, G.; Sikkema, L.; Waghray, A.; Heimberg, G.; Kobayashi, Y.; Vaishnav, E.D.; Subramanian, A.; Smillie, C.; et al. Single-Cell Meta-Analysis of SARS-CoV-2 Entry Genes across Tissues and Demographics. Nat. Med. 2021, 27, 546–559. [Google Scholar] [CrossRef]
  95. Sungnak, W.; Huang, N.; Bécavin, C.; Berg, M.; Queen, R.; Litvinukova, M.; Talavera-López, C.; Maatz, H.; Reichart, D.; Sampaziotis, F.; et al. SARS-CoV-2 Entry Factors Are Highly Expressed in Nasal Epithelial Cells Together with Innate Immune Genes. Nat. Med. 2020, 26, 681–687. [Google Scholar] [CrossRef] [Green Version]
  96. Li, W.; Wicht, O.; van Kuppeveld, F.J.M.; He, Q.; Rottier, P.J.M.; Bosch, B.-J. A Single Point Mutation Creating a Furin Cleavage Site in the Spike Protein Renders Porcine Epidemic Diarrhea Coronavirus Trypsin Independent for Cell Entry and Fusion. J. Virol. 2015, 89, 8077–8081. [Google Scholar] [CrossRef] [Green Version]
  97. Yan, R.; Zhang, Y.; Li, Y.; Xia, L.; Guo, Y.; Zhou, Q. Structural Basis for the Recognition of SARS-CoV-2 by Full-Length Human ACE2. Science 2020, 367, 1444–1448. [Google Scholar] [CrossRef] [Green Version]
  98. Mercurio, I.; Tragni, V.; Busto, F.; De Grassi, A.; Pierri, C.L. Protein Structure Analysis of the Interactions between SARS-CoV-2 Spike Protein and the Human ACE2 Receptor: From Conformational Changes to Novel Neutralizing Antibodies. Cell. Mol. Life Sci. 2021, 78, 1501–1522. [Google Scholar] [CrossRef]
  99. Hoffmann, M.; Kleine-Weber, H.; Pöhlmann, S. A Multibasic Cleavage Site in the Spike Protein of SARS-CoV-2 Is Essential for Infection of Human Lung Cells. Mol. Cell 2020, 78, 779–784.e5. [Google Scholar] [CrossRef]
  100. Plotkin, S.; Robinson, J.M.; Cunningham, G.; Iqbal, R.; Larsen, S. The Complexity and Cost of Vaccine Manufacturing—An Overview. Vaccine 2017, 35, 4064–4071. [Google Scholar] [CrossRef]
  101. Shang, J.; Ye, G.; Shi, K.; Wan, Y.; Luo, C.; Aihara, H.; Geng, Q.; Auerbach, A.; Li, F. Structural Basis of Receptor Recognition by SARS-CoV-2. Nature 2020, 581, 221–224. [Google Scholar] [CrossRef] [Green Version]
  102. Gallagher, T.M. Murine Coronavirus Membrane Fusion Is Blocked by Modification of Thiols Buried within the Spike Protein. J. Virol. 1996, 70, 4683–4690. [Google Scholar] [CrossRef] [Green Version]
  103. Lavillette, D.; Barbouche, R.; Yao, Y.; Boson, B.; Cosset, F.-L.; Jones, I.M.; Fenouillet, E. Significant Redox Insensitivity of the Functions of the SARS-CoV Spike Glycoprotein. J. Biol. Chem. 2006, 281, 9200–9204. [Google Scholar] [CrossRef] [Green Version]
  104. Costello, D.A.; Millet, J.K.; Hsia, C.-Y.; Whittaker, G.R.; Daniel, S. Single Particle Assay of Coronavirus Membrane Fusion with Proteinaceous Receptor-Embedded Supported Bilayers. Biomaterials 2013, 34, 7895–7904. [Google Scholar] [CrossRef]
  105. Giron, C.C.; Laaksonen, A.; Barroso da Silva, F.L. Up State of the SARS-CoV-2 Spike Homotrimer Favors an Increased Virulence for New Variants. Front. Med. Technol. 2021, 3. [Google Scholar] [CrossRef]
  106. Letko, M.; Marzi, A.; Munster, V. Functional Assessment of Cell Entry and Receptor Usage for SARS-CoV-2 and Other Lineage B Betacoronaviruses. Nat. Microbiol. 2020, 5, 562–569. [Google Scholar] [CrossRef] [Green Version]
  107. Zhang, C.; Zheng, W.; Huang, X.; Bell, E.W.; Zhou, X.; Zhang, Y. Protein Structure and Sequence Reanalysis of 2019-NCoV Genome Refutes Snakes as Its Intermediate Host and the Unique Similarity between Its Spike Protein Insertions and HIV-1. J. Proteome Res. 2020, 19, 1351–1360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Tracking SARS-CoV-2 Variants. Available online: https://www.who.int/en/activities/tracking-SARS-CoV-2-variants/ (accessed on 5 July 2021).
  109. Davies, N.G.; Abbott, S.; Barnard, R.C.; Jarvis, C.I.; Kucharski, A.J.; Munday, J.D.; Pearson, C.A.B.; Russell, T.W.; Tully, D.C.; Washburne, A.D.; et al. Estimated Transmissibility and Impact of SARS-CoV-2 Lineage B.1.1.7 in England. Science 2020, 372. [Google Scholar] [CrossRef] [PubMed]
  110. Schrödinger The PyMOL Molecular Graphics System. Available online: https://pymol.org/2/ (accessed on 5 July 2021).
  111. Zhang Modeling of the SARS-CoV-2 Genome Using I-TASSER. Available online: https://zhanggroup.org/COVID-19/ (accessed on 24 August 2021).
  112. Tegally, H.; Wilkinson, E.; Giovanetti, M.; Iranzadeh, A.; Fonseca, V.; Giandhari, J.; Doolabh, D.; Pillay, S.; San, E.J.; Msomi, N.; et al. Detection of a SARS-CoV-2 Variant of Concern in South Africa. Nature 2021, 592, 438–443. [Google Scholar] [CrossRef]
  113. Faria, N.R.; Mellan, T.A.; Whittaker, C.; Claro, I.M.; Candido, D.d.S.; Mishra, S.; Crispim, M.A.E.; Sales, F.C.S.; Hawryluk, I.; McCrone, J.T.; et al. Genomics and Epidemiology of the P.1 SARS-CoV-2 Lineage in Manaus, Brazil. Science 2021, 372, 815–821. [Google Scholar] [CrossRef]
  114. Cherian, S.; Potdar, V.; Jadhav, S.; Yadav, P.; Gupta, N.; Das, M.; Rakshit, P.; Singh, S.; Abraham, P.; Panda, S.; et al. Convergent Evolution of SARS-CoV-2 Spike Mutations, L452R, E484Q, and P681R, in the Second Wave of COVID-19 in Maharashtra, India. BioRxiv 2021. [Google Scholar] [CrossRef]
  115. Hou, Y.J.; Chiba, S.; Halfmann, P.; Ehre, C.; Kuroda, M.; Dinnon, K.H.; Leist, S.R.; Schäfer, A.; Nakajima, N.; Takahashi, K.; et al. SARS-CoV-2 D614G Variant Exhibits Efficient Replication Ex Vivo and Transmission in Vivo. Science 2020, eabe8499. [Google Scholar] [CrossRef]
  116. Plante, J.A.; Liu, Y.; Liu, J.; Xia, H.; Johnson, B.A.; Lokugamage, K.G.; Zhang, X.; Muruato, A.E.; Zou, J.; Fontes-Garfias, C.R.; et al. Spike Mutation D614G Alters SARS-CoV-2 Fitness. Nature 2021, 592, 116–121. [Google Scholar] [CrossRef]
  117. Emary, K.R.W.; Golubchik, T.; Aley, P.K.; Ariani, C.V.; Angus, B.J.; Bibi, S.; Blane, B.; Bonsall, D.; Cicconi, P.; Charlton, S.; et al. Efficacy of ChAdOx1 NCoV-19 (AZD1222) Vaccine Against SARS-CoV-2 VOC 202012/01 (B.1.1.7). SSRN Electron. J. 2021. [Google Scholar] [CrossRef]
  118. Pritchard, E.; Matthews, P.C.; Stoesser, N.; Eyre, D.W.; Gethings, O.; Vihta, K.D.; Jones, J.; House, T.; VanSteenHouse, H.; Bell, I.; et al. Impact of Vaccination on New SARS-CoV-2 Infections in the United Kingdom. Nat. Med. 2021, 1–9. [Google Scholar] [CrossRef]
  119. Wang, P.; Casner, R.G.; Nair, M.S.; Wang, M.; Yu, J.; Cerutti, G.; Liu, L.; Kwong, P.D.; Huang, Y.; Shapiro, L.; et al. Increased Resistance of SARS-CoV-2 Variant P.1 to Antibody Neutralization. Cell Host Microbe 2021, 29, 747–751. [Google Scholar] [CrossRef] [PubMed]
  120. Garcia-Beltran, W.F.; Lam, E.C.; St. Denis, K.; Nitido, A.D.; Garcia, Z.H.; Hauser, B.M.; Feldman, J.; Pavlovic, M.N.; Gregory, D.J.; Poznansky, M.C.; et al. Multiple SARS-CoV-2 Variants Escape Neutralization by Vaccine-Induced Humoral Immunity. Cell 2021, 184, 2372–2383.e9. [Google Scholar] [CrossRef]
  121. Coutinho, R.M.; Marquitti, F.M.D.; Ferreira, L.S.; Borges, M.E.; Paixão da Silva, R.L.; Canton, O.; Portella, T.P.; Poloni, S.; Franco, C.; Plucinski, M.M.; et al. Model-Based Estimation of Transmissibility and Reinfection of SARS-CoV-2 P.1 Variant of the SARS-CoV-2. MedRxiv 2021. [Google Scholar] [CrossRef]
  122. Mullen, J.L.; Tsueng, G.; the Center for Viral Systems Biology. Outbreak.Info. Available online: outbreak.info (accessed on 7 November 2021).
  123. Elbe, S.; Buckland-Merrett, G. Data, Disease and Diplomacy: GISAID’s Innovative Contribution to Global Health. Glob. Chall. 2017, 1, 33–46. [Google Scholar] [CrossRef] [Green Version]
  124. Lustig, Y.; Zuckerman, N.; Nemet, I.; Atari, N.; Kliker, L.; Regev-Yochay, G.; Sapir, E.; Mor, O.; Alroy-Preis, S.; Mendelson, E.; et al. Neutralising Capacity against Delta (B.1.617.2) and Other Variants of Concern Following Comirnaty (BNT162b2, BioNTech/Pfizer) Vaccination in Health Care Workers, Israel. Euro Surveill. 2021, 26. [Google Scholar] [CrossRef] [PubMed]
  125. Bernal, J.L.; Andrews, N.; Gower, C.; Stowe, J.; Robertson, C.; Tessier, E.; Simmons, R.; Cottrell, S.; Roberts, R.; O’Doherty, M.; et al. Early Effectiveness of COVID-19 Vaccination with BNT162b2 MRNA Vaccine and ChAdOx1 Adenovirus Vector Vaccine on Symptomatic Disease, Hospitalisations and Mortality in Older Adults in England. medRxiv 2021. [Google Scholar] [CrossRef]
  126. Campbell, F.; Archer, B.; Laurenson-Schafer, H.; Jinnai, Y.; Konings, F.; Batra, N.; Pavlin, B.; Vandemaele, K.; Van Kerkhove, M.D.; Jombart, T.; et al. Increased Transmissibility and Global Spread of SARS-CoV-2 Variants of Concern as at June 2021. Eurosurveillance 2021, 26. [Google Scholar] [CrossRef] [PubMed]
  127. Ferreira, I.; Datir, R.; Kemp, S.; Papa, G.; Rakshit, P.; Singh, S.; Meng, B.; Pandey, R.; Ponnusamy, K.; Radhakrishnan, V.S.; et al. SARS-CoV-2 B.1.617 Emergence and Sensitivity to Vaccine-Elicited Antibodies. bioRxiv 2021. [Google Scholar] [CrossRef]
  128. Brown, C.M. Outbreak of SARS-CoV-2 Infections, Including COVID-19 Vaccine Breakthrough Infections, Associated with Large Public Gatherings—Barnstable County, Massachusetts, July 2021. MMWR. Morb. Mortal. Wkly. Rep. 2021, 70, 1059–1062. [Google Scholar] [CrossRef] [PubMed]
  129. Centers for Disease Control and Prevention CDC COVID Data Tracker. Available online: https://covid.cdc.gov/covid-data-tracker/#variant-proportions (accessed on 12 August 2021).
  130. Frost, S.D.W.; Magalis, B.R.; Kosakovsky Pond, S.L. Neutral Theory and Rapidly Evolving Viral Pathogens. Mol. Biol. Evol. 2018, 35, 1348–1354. [Google Scholar] [CrossRef]
  131. Bernal, J.L.; Andrews, N.; Gower, C.; Gallagher, E.; Simmons, R.; Thelwall, S.; Stowe, J.; Tessier, E.; Groves, N.; Dabrera, G.; et al. Effectiveness of COVID-19 Vaccines against the B.1.617.2 Variant. medRxiv 2021. [Google Scholar] [CrossRef]
  132. Guirakhoo, F.; Kuo, L.; Peng, J.; Huang, J.H.; Kuo, B.; Lin, F.; Liu, K.; Liu, Z.; Wu, G.; Ding, S.; et al. A Novel SARS-CoV-2 Multitope Protein/Peptide Vaccine Candidate Is Highly Immunogenic and Prevents Lung Infection in an Adeno Associated Virus Human Angiotensin-Converting Enzyme 2 (AAV HACE2) Mouse Model. bioRxiv 2020. [Google Scholar] [CrossRef]
  133. Fischman, S.; Ofran, Y. Computational Design of Antibodies. Curr. Opin. Struct. Biol. 2018, 51, 156–162. [Google Scholar] [CrossRef] [PubMed]
  134. Sormanni, P.; Aprile, F.A.; Vendruscolo, M. Third Generation Antibody Discovery Methods: In Silico Rational Design. Chem. Soc. Rev. 2018, 47, 9137–9157. [Google Scholar] [CrossRef]
  135. Zhao, J.; Nussinov, R.; Wu, W.-J.; Ma, B. In Silico Methods in Antibody Design. Antibodies 2018, 7, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Poveda-Cuevas, S.A.; Etchebest, C.; Barroso Da Silva, F.L. Identification of Electrostatic Epitopes in Flavivirus by Computer Simulations: The PROCEEDpKa Method. J. Chem. Inf. Model. 2020, 60, 944–963. [Google Scholar] [CrossRef] [PubMed]
  137. Robson, B. Techniques Assisting Peptide Vaccine and Peptidomimetic Design. Sidechain Exposure in the SARS-CoV-2 Spike Glycoprotein. Comput. Biol. Med. 2021, 128, 104124. [Google Scholar] [CrossRef] [PubMed]
  138. Oli, A.N.; Obialor, W.O.; Ifeanyichukwu, M.O.; Odimegwu, D.C.; Okoyeh, J.N.; Emechebe, G.O.; Adejumo, S.A.; Ibeanu, G.C. Immunoinformatics and Vaccine Development: An Overview. ImmunoTargets Ther. 2020, 9, 13–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Nezafat, N.; Sadraeian, M.; Rahbar, M.R.; Khoshnoud, M.J.; Mohkam, M.; Gholami, A.; Banihashemi, M.; Ghasemi, Y. Production of a Novel Multi-Epitope Peptide Vaccine for Cancer Immunotherapy in TC-1 Tumor-Bearing Mice. Biologicals 2015, 43, 11–17. [Google Scholar] [CrossRef] [PubMed]
  140. Vicente, T.; Roldão, A.; Peixoto, C.; Carrondo, M.J.T.; Alves, P.M. Large-Scale Production and Purification of VLP-Based Vaccines. J. Invertebr. Pathol. 2011, 107 Suppl, S42–S48. [Google Scholar] [CrossRef]
  141. Garçon, N.; Vaughn, D.W.; Didierlaurent, A.M. Development and Evaluation of AS03, an Adjuvant System Containing α-Tocopherol and Squalene in an Oil-in-Water Emulsion. Expert Rev. Vaccines 2012, 11, 349–366. [Google Scholar] [CrossRef]
  142. Wrapp, D.; Wang, N.; Corbett, K.S.; Goldsmith, J.A.; Hsieh, C.-L.; Abiona, O.; Graham, B.S.; McLellan, J.S. Cryo-EM Structure of the 2019-NCoV Spike in the Prefusion Conformation. Science 2020, 367, 1260–1263. [Google Scholar] [CrossRef] [Green Version]
  143. Abu-Raddad, L.J.; Chemaitelly, H.; Butt, A.A. Effectiveness of the BNT162b2 COVID-19 Vaccine against the B.1.1.7 and B.1.351 Variants. N. Engl. J. Med. 2021, 385, 187–189. [Google Scholar] [CrossRef] [PubMed]
  144. Dagan, N.; Barda, N.; Kepten, E.; Miron, O.; Perchik, S.; Katz, M.A.; Hernán, M.A.; Lipsitch, M.; Reis, B.; Balicer, R.D. BNT162b2 MRNA COVID-19 Vaccine in a Nationwide Mass Vaccination Setting. N. Engl. J. Med. 2021, 384, 1412–1423. [Google Scholar] [CrossRef] [PubMed]
  145. Haas, E.J.; Angulo, F.J.; McLaughlin, J.M.; Anis, E.; Singer, S.R.; Khan, F.; Brooks, N.; Smaja, M.; Mircus, G.; Pan, K.; et al. Impact and Effectiveness of MRNA BNT162b2 Vaccine against SARS-CoV-2 Infections and COVID-19 Cases, Hospitalisations, and Deaths Following a Nationwide Vaccination Campaign in Israel: An Observational Study Using National Surveillance Data. Lancet 2021, 397, 1819–1829. [Google Scholar] [CrossRef]
  146. Flood, A.; Estrada, M.; McAdams, D.; Ji, Y.; Chen, D. Development of a Freeze-Dried, Heat-Stable Influenza Subunit Vaccine Formulation. PLoS ONE 2016, 11, e0164692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Dai, L.; Zheng, T.; Xu, K.; Han, Y.; Xu, L.; Huang, E.; An, Y.; Cheng, Y.; Li, S.; Liu, M.; et al. A Universal Design of Betacoronavirus Vaccines against COVID-19, MERS, and SARS. Cell 2020, 182, 722–733.e11. [Google Scholar] [CrossRef] [PubMed]
  148. de Villiers, E.M. Heterogeneity of the Human Papillomavirus Group. J. Virol. 1989, 63, 4898–4903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Fluckiger, A.-C.; Ontsouka, B.; Bozic, J.; Diress, A.; Ahmed, T.; Berthoud, T.; Tran, A.; Duque, D.; Liao, M.; Mccluskie, L.; et al. An Enveloped Virus-like Particle Vaccine Expressing a Stabilized Prefusion Form of the SARS-CoV-2 Spike Protein Elicits Potent Immunity after a Single Dose. bioRxiv 2021. [Google Scholar] [CrossRef]
  150. Keating, G.M.; Noble, S. Recombinant Hepatitis B Vaccine (Engerix-B??): A Review of Its Immunogenicity and Protective Efficacy Against Hepatitis B. Drugs 2003, 63, 1021–1051. [Google Scholar] [CrossRef] [PubMed]
  151. Rudd, P.M.; Wormald, M.R.; Stanfield, R.L.; Huang, M.; Mattsson, N.; Speir, J.A.; DiGennaro, J.A.; Fetrow, J.S.; Dwek, R.A.; Wilson, I.A. Roles for Glycosylation of Cell Surface Receptors Involved in Cellular Immune Recognition. J. Mol. Biol. 1999, 293, 351–366. [Google Scholar] [CrossRef] [PubMed]
  152. Monie, A.; Hung, C.-F.; Roden, R.; Wu, T.-C. Cervarix: A Vaccine for the Prevention of HPV 16, 18-Associated Cervical Cancer. Biologics 2008, 2, 97–105. [Google Scholar] [PubMed]
  153. Giannini, S.L.; Hanon, E.; Moris, P.; Van Mechelen, M.; Morel, S.; Dessy, F.; Fourneau, M.A.; Colau, B.; Suzich, J.; Losonksy, G.; et al. Enhanced Humoral and Memory B Cellular Immunity Using HPV16/18 L1 VLP Vaccine Formulated with the MPL/Aluminium Salt Combination (AS04) Compared to Aluminium Salt Only. Vaccine 2006, 24, 5937–5949. [Google Scholar] [CrossRef] [PubMed]
  154. Paavonen, J.; Jenkins, D.; Bosch, F.X.; Naud, P.; Salmerón, J.; Wheeler, C.M.; Chow, S.-N.; Apter, D.L.; Kitchener, H.C.; Castellsague, X.; et al. Efficacy of a Prophylactic Adjuvanted Bivalent L1 Virus-like-Particle Vaccine against Infection with Human Papillomavirus Types 16 and 18 in Young Women: An Interim Analysis of a Phase III Double-Blind, Randomised Controlled Trial. Lancet 2007, 369, 2161–2170. [Google Scholar] [CrossRef]
  155. Harper, D.M.; Franco, E.L.; Wheeler, C.M.; Moscicki, A.-B.; Romanowski, B.; Roteli-Martins, C.M.; Jenkins, D.; Schuind, A.; Costa Clemens, S.A.; Dubin, G. Sustained Efficacy up to 4·5 Years of a Bivalent L1 Virus-like Particle Vaccine against Human Papillomavirus Types 16 and 18: Follow-up from a Randomised Control Trial. Lancet 2006, 367, 1247–1255. [Google Scholar] [CrossRef] [Green Version]
  156. Walls, A.C.; Fiala, B.; Schäfer, A.; Wrenn, S.; Pham, M.N.; Murphy, M.; Tse, L.V.; Shehata, L.; O’Connor, M.A.; Chen, C.; et al. Elicitation of Potent Neutralizing Antibody Responses by Designed Protein Nanoparticle Vaccines for SARS-CoV-2. Cell 2020, 183, 1367–1382.e17. [Google Scholar] [CrossRef]
  157. Arunachalam, P.S.; Walls, A.C.; Golden, N.; Atyeo, C.; Fischinger, S.; Li, C.; Aye, P.; Navarro, M.J.; Lai, L.; Edara, V.V.; et al. Adjuvanting a Subunit COVID-19 Vaccine to Induce Protective Immunity. Nature 2021, 594, 253–258. [Google Scholar] [CrossRef]
  158. Heath, P.T.; Galiza, E.P.; Baxter, D.N.; Boffito, M.; Browne, D.; Burns, F.; Chadwick, D.R.; Clark, R.; Cosgrove, C.; Galloway, J.; et al. Safety and Efficacy of NVX-CoV2373 Covid-19 Vaccine. N. Engl. J. Med. 2021, 385, 1172–1183. [Google Scholar] [CrossRef] [PubMed]
  159. Fougeroux, C.; Goksøyr, L.; Idorn, M.; Soroka, V.; Myeni, S.K.; Dagil, R.; Janitzek, C.M.; Søgaard, M.; Aves, K.-L.; Horsted, E.W.; et al. Capsid-like Particles Decorated with the SARS-CoV-2 Receptor-Binding Domain Elicit Strong Virus Neutralization Activity. Nat. Commun. 2021, 12, 324. [Google Scholar] [CrossRef]
  160. Zhou, H.; Ji, J.; Chen, X.; Bi, Y.; Li, J.; Wang, Q.; Hu, T.; Song, H.; Zhao, R.; Chen, Y.; et al. Identification of Novel Bat Coronaviruses Sheds Light on the Evolutionary Origins of SARS-CoV-2 and Related Viruses. Cell 2021, 184, 4380–4391.e14. [Google Scholar] [CrossRef] [PubMed]
  161. Cui, J.; Li, F.; Shi, Z.-L. Origin and Evolution of Pathogenic Coronaviruses. Nat. Rev. Microbiol. 2019, 17, 181–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Rappazzo, C.G.; Tse, L.V.; Kaku, C.I.; Wrapp, D.; Sakharkar, M.; Huang, D.; Deveau, L.M.; Yockachonis, T.J.; Herbert, A.S.; Battles, M.B.; et al. Broad and Potent Activity against SARS-like Viruses by an Engineered Human Monoclonal Antibody. Science 2021, 371, 823–829. [Google Scholar] [CrossRef] [PubMed]
  163. Powell, A.E.; Zhang, K.; Sanyal, M.; Tang, S.; Weidenbacher, P.A.; Li, S.; Pham, T.D.; Pak, J.E.; Chiu, W.; Kim, P.S. A Single Immunization with Spike-Functionalized Ferritin Vaccines Elicits Neutralizing Antibody Responses against SARS-CoV-2 in Mice. ACS Cent. Sci. 2021, 7, 183–199. [Google Scholar] [CrossRef] [PubMed]
  164. Liu, Z.; Zhou, H.; Wang, W.; Tan, W.; Fu, Y.-X.; Zhu, M. A Novel Method for Synthetic Vaccine Construction Based on Protein Assembly. Sci. Rep. 2015, 4, 7266. [Google Scholar] [CrossRef] [PubMed]
  165. Bale, J.B.; Gonen, S.; Liu, Y.; Sheffler, W.; Ellis, D.; Thomas, C.; Cascio, D.; Yeates, T.O.; Gonen, T.; King, N.P.; et al. Accurate Design of Megadalton-Scale Two-Component Icosahedral Protein Complexes. Science 2016, 353, 389–394. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The adaptive immune response generated by VLPs immunization and VLPs classification. (A) After immunization, VLPs are phagocytized by dendritic cells or macrophages. Then, they are carried out to lymphatic vessels, where the antigenic regions will be processed and presented by class II MHC molecules (CD4+ T cells) and, through cross-presentation, by class I (CD8+ T cells). Immunological pathway activation by immunization with VLPs will activate robust cellular (cytokines) and humoral (B cell-antibodies) immune responses. (B) VLPs are classified as nonenveloped (neVLPs) or enveloped VLPs (eVLPs) based on the absence or presence of a lipidic membrane, respectively. These particles can also be classified as homologous or heterologous VLPs according to their composition. Homologous VLPs are assembled using proteins from the native pathogen only (blue), and heterologous VLPs can be assembled using proteins or peptides from different sources (black and blue).
Figure 1. The adaptive immune response generated by VLPs immunization and VLPs classification. (A) After immunization, VLPs are phagocytized by dendritic cells or macrophages. Then, they are carried out to lymphatic vessels, where the antigenic regions will be processed and presented by class II MHC molecules (CD4+ T cells) and, through cross-presentation, by class I (CD8+ T cells). Immunological pathway activation by immunization with VLPs will activate robust cellular (cytokines) and humoral (B cell-antibodies) immune responses. (B) VLPs are classified as nonenveloped (neVLPs) or enveloped VLPs (eVLPs) based on the absence or presence of a lipidic membrane, respectively. These particles can also be classified as homologous or heterologous VLPs according to their composition. Homologous VLPs are assembled using proteins from the native pathogen only (blue), and heterologous VLPs can be assembled using proteins or peptides from different sources (black and blue).
Vaccines 09 01409 g001
Figure 2. SARS-CoV-2 structural proteins and the different states of the Spike protein. (A) Schematic representation of the SARS-CoV-2 viral particle. The structure of the SARS-CoV-2 viral particle is composed of four structural proteins: Membrane (M), Envelope (E), Nucleocapsid (N), and Spike (S). The S protein is found in two different states on viral particles: open state (minor population) and closed state (major population). In addition, during the membrane fusion process (host cell entry), the S protein can be found in the fusion state (fusion S). (B) Schematic representation of the binding of open-state S (PDB ID 7498) to the ACE2 receptor present in the host cell. The illustrations were made in free software (CellPaint 2.0 [91] and 3D Protein Imager [92]). The binding figure was made using the crystal structure of ACE2 bound to Spike available at the Protein Data Bank (PDB ID 7A98).
Figure 2. SARS-CoV-2 structural proteins and the different states of the Spike protein. (A) Schematic representation of the SARS-CoV-2 viral particle. The structure of the SARS-CoV-2 viral particle is composed of four structural proteins: Membrane (M), Envelope (E), Nucleocapsid (N), and Spike (S). The S protein is found in two different states on viral particles: open state (minor population) and closed state (major population). In addition, during the membrane fusion process (host cell entry), the S protein can be found in the fusion state (fusion S). (B) Schematic representation of the binding of open-state S (PDB ID 7498) to the ACE2 receptor present in the host cell. The illustrations were made in free software (CellPaint 2.0 [91] and 3D Protein Imager [92]). The binding figure was made using the crystal structure of ACE2 bound to Spike available at the Protein Data Bank (PDB ID 7A98).
Vaccines 09 01409 g002
Figure 3. Structure and domain organization of the SARS-CoV-2 Spike (S) protein. (A) The S structure comprises a cytoplasmic domain (CD, white), a transmembrane domain (TM, black), and an ectodomain, which is divided into two subunits, S1 (gray) and S2 (dark gray). The magnification shows the several disulfide bridges (DB, yellow) and the glycosylation sites (GlcNAc, green) through the S protein ectodomain. It is highlighted in red, the S1/S2 interface. The receptor-binding domain (RBD, in cyan) and the receptor-binding motif (RBM, magenta) are also shown in S1. (B) As mentioned in Figure 2, the S protein shows two conformers on viable viruses (closed and open state). The upper panel shows the S protein in the closed state (trimeric and monomeric state). The bottom panel shows the S protein in the open state (trimeric and monomeric state). Illustrations were made in PyMol [110] using the wild-type structures available from Zhang et al. [107,111].
Figure 3. Structure and domain organization of the SARS-CoV-2 Spike (S) protein. (A) The S structure comprises a cytoplasmic domain (CD, white), a transmembrane domain (TM, black), and an ectodomain, which is divided into two subunits, S1 (gray) and S2 (dark gray). The magnification shows the several disulfide bridges (DB, yellow) and the glycosylation sites (GlcNAc, green) through the S protein ectodomain. It is highlighted in red, the S1/S2 interface. The receptor-binding domain (RBD, in cyan) and the receptor-binding motif (RBM, magenta) are also shown in S1. (B) As mentioned in Figure 2, the S protein shows two conformers on viable viruses (closed and open state). The upper panel shows the S protein in the closed state (trimeric and monomeric state). The bottom panel shows the S protein in the open state (trimeric and monomeric state). Illustrations were made in PyMol [110] using the wild-type structures available from Zhang et al. [107,111].
Vaccines 09 01409 g003
Figure 4. Mapping mutations of SARS-CoV-2 variants of concern (VOCs) and phenotypes. Red: mutations; Cyan: receptor-binding domain (RBD); Magenta: receptor binding motif (RBM); Light gray: S1; Dark gray: S2; Yellow: Heptad repeat 1; Green cyan: fusion peptide 1; Slate: Fusion peptide 2; Green: Signal peptide. Illustrations were made in PyMol [110] using resources from Zhang et al. [107,111].
Figure 4. Mapping mutations of SARS-CoV-2 variants of concern (VOCs) and phenotypes. Red: mutations; Cyan: receptor-binding domain (RBD); Magenta: receptor binding motif (RBM); Light gray: S1; Dark gray: S2; Yellow: Heptad repeat 1; Green cyan: fusion peptide 1; Slate: Fusion peptide 2; Green: Signal peptide. Illustrations were made in PyMol [110] using resources from Zhang et al. [107,111].
Vaccines 09 01409 g004
Figure 5. Enveloped and nonenveloped VLPs against SARS-CoV-2.
Figure 5. Enveloped and nonenveloped VLPs against SARS-CoV-2.
Vaccines 09 01409 g005
Table 1. Summary of WHO COVID-19 approved vaccines.
Table 1. Summary of WHO COVID-19 approved vaccines.
NamePlatformAdjuvantDosageEfficacy *References
Coronavac
(Sinovac)
InactivatedAlum2 doses83.5% (95% CI, 65.4–92.1) [8,9,10,11]
BBIBP-CorV
(Sinopharm)
InactivatedAlum2 doses72.8% (95% CI, 58.1–82.4)[12,13]
BBV152-Covaxin
(Bharat Biotech)
InactivatedAlum2 doses77.8% (95% CI, 65.2–86.4)[14,15]
AZD1222–Vaxzevria
(Oxford/AstraZeneca)
Viral vectorNo2 doses74.0% (95% CI, 65.3–80.5)[16,17,18]
Covishield
(Oxford/AstraZeneca formulation)
Viral vectorNo2 doses74.0% (95% CI, 65.3–80.5)[16,17,18]
Ad26.COV2.S
(Johnson &Johnson-Janssen)
Viral vectorNo1 dose66.9% (95% CI, 59.0–73.4)[19,20,21,22]
mRNA-1273
(Moderna)
mRNANo2 doses94.1% (95% CI, 89.3–96.8)[23,24]
BNT162b-Comirnaty
(Pfizer/BioNTech)
mRNANo2 doses95% (95% CI, 90.3–97.6)[25,26]
* Against symptomatic COVID-19.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Prates-Syed, W.A.; Chaves, L.C.S.; Crema, K.P.; Vuitika, L.; Lira, A.; Côrtes, N.; Kersten, V.; Guimarães, F.E.G.; Sadraeian, M.; Barroso da Silva, F.L.; et al. VLP-Based COVID-19 Vaccines: An Adaptable Technology against the Threat of New Variants. Vaccines 2021, 9, 1409. https://doi.org/10.3390/vaccines9121409

AMA Style

Prates-Syed WA, Chaves LCS, Crema KP, Vuitika L, Lira A, Côrtes N, Kersten V, Guimarães FEG, Sadraeian M, Barroso da Silva FL, et al. VLP-Based COVID-19 Vaccines: An Adaptable Technology against the Threat of New Variants. Vaccines. 2021; 9(12):1409. https://doi.org/10.3390/vaccines9121409

Chicago/Turabian Style

Prates-Syed, Wasim A., Lorena C. S. Chaves, Karin P. Crema, Larissa Vuitika, Aline Lira, Nelson Côrtes, Victor Kersten, Francisco E. G. Guimarães, Mohammad Sadraeian, Fernando L. Barroso da Silva, and et al. 2021. "VLP-Based COVID-19 Vaccines: An Adaptable Technology against the Threat of New Variants" Vaccines 9, no. 12: 1409. https://doi.org/10.3390/vaccines9121409

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop