Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3487 2023-04-21 15:09:57 |
2 format Meta information modification 3487 2023-04-23 03:56:49 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Fantini, J.; Chahinian, H.; Yahi, N. SARS-CoV-2 and HIV Surface Envelope Glycoproteins. Encyclopedia. Available online: https://encyclopedia.pub/entry/43324 (accessed on 01 July 2024).
Fantini J, Chahinian H, Yahi N. SARS-CoV-2 and HIV Surface Envelope Glycoproteins. Encyclopedia. Available at: https://encyclopedia.pub/entry/43324. Accessed July 01, 2024.
Fantini, Jacques, Henri Chahinian, Nouara Yahi. "SARS-CoV-2 and HIV Surface Envelope Glycoproteins" Encyclopedia, https://encyclopedia.pub/entry/43324 (accessed July 01, 2024).
Fantini, J., Chahinian, H., & Yahi, N. (2023, April 21). SARS-CoV-2 and HIV Surface Envelope Glycoproteins. In Encyclopedia. https://encyclopedia.pub/entry/43324
Fantini, Jacques, et al. "SARS-CoV-2 and HIV Surface Envelope Glycoproteins." Encyclopedia. Web. 21 April, 2023.
SARS-CoV-2 and HIV Surface Envelope Glycoproteins
Edit

Although very different, in terms of their genomic organization, their enzymatic proteins, and their structural proteins, HIV and SARS-CoV-2 have an extraordinary evolutionary potential in common. Faced with various selection pressures that may be generated by treatments or immune responses, these RNA viruses demonstrate very high adaptive capacities, which result in the continuous emergence of variants and quasi-species. HIV and SARS-CoV-2 first recognize a lipid raft microdomain that acts as a landing strip for viral particles on the host cell surface. In the case of mucosal cells, which are the primary targets of both viruses, these microdomains are enriched in anionic glycolipids (gangliosides) forming a global electronegative field. Both viruses use lipid rafts to surf on the cell surface in search of a protein receptor able to trigger the fusion process. This implies that viral envelope proteins are both geometrically and electrically compatible to the biomolecules they select to invade host cells.

virus evolution HIV-1 SARS-CoV-2 electrostatic surface potential

1. Introduction

Virus-cell interactions during the early stages of the infection cycle of enveloped viruses have been the subject of numerous studies for several decades. The COVID-19 pandemic has mobilized the efforts of the scientific community around the world. Among these researchers, two (N.Y. and J.F.) extensively studied the molecular mechanisms associated with the infection of human cells by HIV-1 and other retroviruses. Nouara Yahi began her work on HIV by joining, in 1988, the retrovirus laboratory newly created in Marseille by Jean-Claude Chermann, one of the co-discoverers of the AIDS virus at the Pasteur Institute in Paris. She was in charge of HIV strain isolation [1] and antiviral drug testing [2]. Jacques Fantini joined the team in 1990. The two then started to study the infection of human epithelial intestinal cells by HIV-1 [3], a project which would eventually solve two issues: (i) the identification of galactosylceramide (GalCer) as an alternative HIV-1 receptor in these CD4-negative cells [4], shortly after the group of F. Gonzalez-Scarano in Philadelphia identified it as the portal of entry of HIV-1 in neural cells [5]; (ii) a virotoxin-induced signal transduction pathway accounting for the puzzling HIV-associated enteropathy: HIV-1 surface envelope glycoprotein gp120 binding GalCer, intracellular Ca2+ release, microtubule disruption, and impaired absorption functions [6][7][8][9][10].

2. Structural and Functional Analysis of the SARS-CoV-2 Spike Protein

The spike protein of SARS-CoV-2 is a large glycoprotein synthesized as a precursor containing 1273 amino acid residues for the original Wuhan strain (https://www.uniprot.org/uniprotkb/P0DTC2/entry#sequences accessed on 15 November 2022). The first 18 amino acids form the signal sequence, which is cleaved to generate the mature form encompassing residues 19-1273. As shown in Figure 1A, the protein has a typical Y shape, whose upper branches correspond to the N-terminal domain (NTD) and the receptor-binding domain (RBD). The lower trunk of the Y displays two proteolytic sites: (i) S1–S2, which is cleaved by furin in the Golgi apparatus during biosynthesis and maturation in the infected cell, and (ii) S2′, an additional cleavage site that is essential for the virus to fuse at the plasma membrane of human lung cells [11][12]. The cleavage of S1–S2 by furin generates two subunits (S1 and S2) that remain non covalently attached. The S2′ site is cleaved by type II transmembrane serine protease 2 (TMPRSS2) at the cell surface or by cathepsins in the endosome [11][13]. In the absence of the furin site, or if the mutations render this site non-functional [14], the alternative for the virus is, thus, to gain entry into the cell by endocytosis, according to a classic mechanism shared by many coronaviruses [11][15]. The S1 subunit contains the NTD and the RBD, which bind to the cell surface, whereas the S2 subunit possesses the machinery necessary for the fusion between the virus envelope and the plasma membrane of the host cell [16]. The RBDs and the fusion machinery cluster in the center of the trimer, while the NTDs are pushed to the sides. There are two forms of the trimeric spike, the closed form and the open form [16]. The closed form (Figure 1B) must undergo a conformational change to make the central RBDs accessible and, thus, allow them to interact with a cellular receptor, primarily ACE2 [17]. Therefore, the attachment of the spike protein to the ACE2 receptor cannot be the first step in the process of adhesion of the virus to the surface of the host cell [18]. It is obvious that it is, indeed, the closed form of the spike that must first attach itself to the cell, which will trigger the conformational change, which will subsequently allow attachment to ACE2. Researchers compared this mechanism to docking a spacecraft on a space station [18]. First, each NTD domain seeks a favorable landing area that researchers have identified as a lipid raft, i.e., a plasma membrane microdomain enriched in gangliosides and cholesterol [19]. Indeed, researchers have shown that NTD has an excellent affinity for GM1 gangliosides, including the acetylated GM1 derivatives, which are particularly abundant on the surface of respiratory mucosal cells [20]. Since there are 3 NTDs per spike trimer, this multiplies the chances of adhesion to a lipid raft, which ensures the initial attachment of the virus to the plasma membrane of the target cell [21]. The first trimer sticks to a raft, causing a local deformation of the membrane, which invaginates. This allows other virus trimers to interact with vicinal rafts. The virus will then sail on these rafts until it encounters the ACE2 receptor, which is, itself, a raft-associated protein [22]. This fast “surfing” process [23] facilitates the formation of a trimolecular complex consisting of ACE2, raft gangliosides, and the spike protein [24]. The next step is a conformational change of the spike trimer, which triggers the unmasking of the RBD [25]. Finally, the spike protein is cleaved at the S2′ site by TMPRSS2 [26], which in turn initiates the fusion process controlled by the S2 subunit [27]. Researchers are, thus, dealing with a perfectly controlled thermodynamic mechanism, which leads inexorably from adhesion to fusion. The selection pressure that controls the evolution of SARS-CoV-2 variants at the level of the spike protein will, therefore, apply to each of these steps by independently targeting the parts of the spike protein involved. However, the parameters on which these selection pressures will act will not be the same, depending on the targeted domain. In the case of the NTD, which controls the initial step of adhesion of the virus to a lipid raft, the selection pressure will be exerted preferentially on the kinetics of adhesion. For the RBD, there will be a combination of direct effects and indirect effects. The direct effects will concern the kinetics of the RBD-ACE2 interaction, but also the affinity parameter [20]. The indirect effects will be caused by mutations facilitating conformational unmasking of the RBD, such as the D614G mutation [28]. However, the other mutations may, on the contrary, stabilize the trimer in its closed conformation, thus increasing the resistance of the virus to extreme conditions and favoring routes of contamination other than via the respiratory mucosa (for example by ingestion) [29]. The furin site may be modulated by mutations facilitating proteolytic cleavage, or instead, minimizing it, in the case of the Omicron variants. Finally, mutations in the helix domains involved in the fusion mechanism may also affect the infectivity of the virus. Yet, the analysis of SARS-CoV-2 mutations is further complicated by two parameters that must also be taken into account. First, each mutation cannot be analyzed independently of other mutations in the same functional domain [29]. Thus, the impact of a single mutation will depend not only on its own effect on the structure of the protein, but also on its contribution to a global mutational pattern. Secondly, it must be considered that, in many countries, the anti-COVID-19 vaccination rate is very high, which means that the vaccine must also be considered as a selection pressure [30]. The recent finding of a higher intra-host diversity among vaccinated individuals is also in favor of a potential vaccine-induced immune pressure [31]. Thus, many mutations are selected not on the basis of the advantage they confer, in terms of virus infectivity, but rather for their contribution to immune escape. People infected several times by the virus are also affected by this phenomenon, whether they are vaccinated or not.
Figure 1. Structural features of SARS-CoV-2 spike protein monomer (A) and pre-fusion trimer (B). NTD, N-terminal domain; RBD, receptor binding domain. In the trimer, the subunits ribbons are colored in cyan, yellow, and magenta. The two proteolytic cleavage sites (S1–S2 and S2′) are indicated in the monomer and (when visible) in the trimer. The models were modified from pdb file 7 bnm.

3. Structural Dynamics of SARS-CoV-2 Spike Protein Evolution

Schematically, the NTD has two fundamental properties that explain why this domain is particularly efficient in targeting the gangliosides of the lipid rafts: (i) its contact surface with the cell is flat [23][24], and (ii) it is globally electropositive [20]. This combination of geometrical and electrostatic parameters is truly the winning formula for binding to lipid rafts, which are well-demarcated flat and electronegative landing zones [32]. Each GM1 ganglioside has an ionized sialic acid and, therefore, has a negative charge at physiological pH. The repetition of this negative charge over the raft gives these membrane domains a global negative electrostatic potential [32]. Under these conditions, the more the NTD is electropositive, the faster it will interact with the raft. Mutations that increase the electropositivity of the NTD, therefore, directly affect the kinetics of interaction of the virus with the raft [20]. In the order of appearance of SARS-CoV-2 variants, from the original Wuhan strain to the Delta variant, the electrostatic potential underwent a steady increase, becoming more and more electropositive [20]. However, the electrostatic potential cannot increase indefinitely because, beyond a certain value, the virus could stick too strongly to the membrane, which would have the effect of making it non-infectious. Therefore, it was clear that the Delta variant was the final outcome and no virus could supplant it, except to go back and compensate for the decrease in surface potential by another parameter. This is how the first Omicron variant appeared, when Delta was still predominant [33]. It remained so in several countries, especially those where vaccination was the least widespread, such as South Africa [34]. Researchers can clearly see that the analysis of variants is multiparametric and that it is sometimes necessary to look elsewhere, other than in the virus itself, to find the reasons explaining the emergence of one variant compared to another. Thus, it is in the most vaccinated countries, such as France or Denmark, that the wave of Omicron was the strongest, even though Delta was still very present [35][36][37]. It is, therefore, most likely due to an immune escape phenomenon that the first Omicron variant was able to succeed Delta, with the neutralizing antibodies induced by the vaccine being notably less effective against Omicron, compared to Delta [38][39]. A very strong argument in favor of this interpretation is given by the in vitro infection kinetics comparing the Delta and Omicron in Calu-3 cells, which have high level expressions of TMPRSS2 and in (TMPRSS2)-overexpressing VeroE6 cells [40]. In a competition assay with these cells, Delta outcompeted Omicron [40]. This means that, on a strictly virological level, Omicron has no decisive advantage over Delta. For Omicron to take over Delta, it needed outside help, the natural or vaccinal immune response. However, the selection pressures are multiple and Omicron has acquired a faculty that Delta did not have, nor the other variants that preceded it: an inoperative furin site (S1–S2) condemning the virus to enter the cells through endocytosis [40][41]. This resistance to furin cleavage is manifested at the level of the cleavage zone by a structuring that reduces the flexibility of the substrate loop.
Finally, Omicron’s mutational program is really a jigsaw puzzle, with mutations that seem to go all over the place [42]. The electropositive surface potential of the NTD is diminished, compared to Delta, but that of the RBD is increased. The contact surface of the NTD is slightly rounded, but this problem will be solved by successive Omicron variants, from BA.2 to BA.5 and BQ.1.1 today.
In order to provide a comparative analysis of SARS-CoV-2 variants, researchers have proposed a transmissibility index (T-index) taking into account the kinetic parameters (surface electrostatic potential of NTD and RBD) and affinity (NTD-gangliosides and RBD-ACE2 interactions) [20][42][43]. This index clearly accounts for the evolution of SARS-CoV-2 from the original Wuhan strain to the Delta variant, with each new variant having a higher T-index than its predecessor. The arrival of Omicron has changed the situation, since this new line of variants has a lower T-index than Delta, with its success being due both to its mechanism of entry by endocytosis (mutations of the furin site) and to the immune escape.
One of the most intriguing aspects of the structural dynamics of the evolution of SARS-CoV-2 variants is the relative rarity of insertions and deletions, which are, nevertheless, the hallmark of RNA viruses, including HIV [44][45][46]. In fact, these events also exist for SARS-CoV-2, but they are concentrated in hot spots and, in particular, at the level of the NTD [44] and, more precisely, at the level of the contact surface with the rafts [20]. One can wonder about the significance of these rearrangements in these zones of interaction with lipid rafts. If researchers consider the selection pressure induced by the neutralizing antibodies recognizing these areas, the immune escape requires mutations decreasing the affinity of these antibodies for the NTD [43]. However, ultimately, these neutralizing antibodies are kind of mirrors of the NTD’s contact surface: geometrically flat and electrostatically negatively charged, similar to rafts [18]. Under these conditions, any modification of the epitopes recognized by these antibodies could also cause a decrease of the affinity of the NTD for the rafts and, therefore, a disadvantage for the virus. The deletions in the NTD can then be interpreted as a loss of epitope, without major consequences for the binding to the rafts. The virus becomes resistant to neutralizing antibodies, while remaining infectious.
Despite the complexity of the selection pressures that determine the emergence of SARS-CoV-2 variants, the mechanisms affected by the mutations of the spike protein remain perfectly explainable. Among these mechanisms, the surface electrostatic potential plays a major role, which is, in fact, shared by many viruses, such as HIV.

4. Structural and Functional Analysis of HIV-1 gp120 Surface Envelope Glycoprotein

In the case of HIV-1, the molecular details of the mechanisms of entry are different, but the strategy is globally similar to the one used by SARS-CoV-2. The surface envelope glycoproteins that constitute the trimeric spike of HIV-1 [47] are already cleaved from a unique precursor, gp160 [48]. This precursor is cleaved by a cellular protease before the assembly of the viral particle at the plasma membrane of the infected cells [49]. Two glycoproteins are generated, the surface envelope g120 (SU), which is equivalent to S1 for SARS-CoV-2, and the transmembrane gp41 (TM), which is equivalent to S2 [50]. The amino acid residues that are critical for the binding to raft gangliosides (V3 loop for HIV-1 and part of the NTD for SARS-CoV-2) are highlighted in yellow. Although the binding to lipid rafts is typically an induced fit mechanism [51], a significant part of the amino acid chain is exposed on the surface of the viral glycoproteins, allowing for fast contact with raft gangliosides. The sequence of events leading to HIV-1 entry is, however, very similar to the fusion process of SARS-CoV-2. Researchers elucidated and published this mechanism as early as 1998 [52], before it was confirmed by many subsequent studies [53][54][55][56][57][58][59][60]. The HIV-1 virion first binds to a raft, via its gp120 V3 loop domain, which, although not totally exposed on the surface of the viral glycoprotein, is sufficiently accessible to engage functional virus-raft contacts. Since the CD4 receptor is also located in a raft [61][62], this very first step facilitates gp120-CD4 binding. A conformational change then totally unmasks the V3 loop of gp120 [63][64]. The V3 loop keeps the virus firmly attached to the raft through direct interactions with accessory glycosphingolipids (Gb3 and/or GM3) [52][65][66][67]. The virus will then sail on the cell surface (surfing step) [59][65][67] until it encounters a co-receptor, CCR5 or CXCR4, also recognized by the V3 loop [68]. The V3-coreceptor interaction disconnects CD4 from the complex and triggers the last conformational change of gp41, which eventually activates the fusion machinery [69]. As in the case of SARS-CoV-2, infection can only occur if the rafts are fully functional and free to move on the cell surface. An important specificity of HIV-1 concerns the choice of the coreceptor, CCR5 or CXCR4. Initially, HIV-1 were classified according to a phenotypic criterion and the ability to induce syncytia in cultures of infected cells [70]. The syncytium-inducing (SI) viruses appear in infected patients after the non-syncytium-inducing (NSI) viruses from which they derive by accumulation of mutations at the level of the V3 loop [71]. The more aggressive a virus is, the more syncytia it forms and the more mutations it presents in its V3 loop [72]. In general, NSI viruses recognize CCR5 [73] and SI viruses recognize CXCR4 [74]. Interestingly, this coreceptor switch is strongly correlated with an increase in the surface electrostatic potential of the V3 domain [75][76]. This increase is consistent with the fact that the surface potential of CXCR4 is much more electronegative than that of CCR5 [77].

5. Structural Dynamics of HIV-1 gp120 Surface Envelope Glycoprotein Evolution

It, therefore, appears that the evolution of SARS-CoV-2 and HIV-1 variants and quasi-species is driven by an adaptation of the contact surfaces of the virus with the areas and membrane receptors controlling the adhesion of viral particles to target cells. In fact, the analysis of the evolution of the V3 loop of HIV-1 gp120 shows an accumulation of basic amino acid residues (lysine and arginine), which increase the electropositivity of this domain [78]. It is this evolution of the surface potential that allows the virus to sequentially use the CCCR5 then CXCR4 co-receptors, while maintaining a good attractiveness for lipid rafts. From this point of view, the virus behaves like an evolutionary probe, capable of estimating the electrostatic potential of its targets and of making its own potential evolve, according to this parameter. Interestingly, this important feature was established several years before HIV-1 coreceptors were identified [79]. Indeed, the evolution of the V3 loop sequences shows a progressive enrichment in basic amino acids. As a result, a strong correlation can be established between the net charge of the V3 loop and the type of coreceptor used. Below a net charge of +3, CCR5 will be preferred. Above the value +4, the virus will then be able to use CXCR4 [75]. Viruses able to use both CCR5 and CXCR4 (referred to as R5X 4 strains) have an intermediary net charge in the 3–4 range [75].
At the phenotypic level, quasi-species using CXCR4 are more aggressive, and they can infect cellular targets, other than T4 lymphocytes and macrophages. However, since secreted gp120 has virotoxin properties [10], infection is not requested to induce deleterious effects in the intestinal epithelium [80] and nerve cells [81], through direct binding of the viral glycoprotein to cell surface glycosphingolipids, such as galactosylceramide [4][5].

6. Considerations on the Electrostatic Surface Potential

As discussed above, the surface electrostatic potential plays a major role in the selection mechanisms that are directly responsible for the emergence of new viral populations with increased tropism and/or infectivity. At this point, researchers must make an important distinction between kinetic effects and affinity enhancement [20]. Regarding the lipid raft recognition domains, it is clear that the increase in surface electrostatic potential does not translate into an increase in the affinity of viruses for gangliosides. There are several reasons for this. First, the interaction of viral glycoproteins with raft gangliosides involves several distinct gangliosides [32]. It is, therefore, possible to measure an overall avidity for this cluster of gangliosides, which is the sum of the affinity of each individual ganglioside for the viral glycoprotein. The variation in free energy (ΔG) associated with this multi-partner reaction is already very large, and it can no longer increase significantly [20]. Indeed, in the case of SARS-CoV-2, this ΔG shows very little variation from one variant to another, despite the accumulation of mutations in the spike protein. In contrast, the increase in surface potential gives a clear kinetic advantage, allowing the most electropositive viruses to adhere to target cells faster than their competitors.
This advantage is conferred by the particular associations of mutations concentrated in the V3 loop of HIV-1 gp120 [82] and at the surface of the NTD and of the RBD in the case of the spike protein of SARS-CoV-2 [20]. In the latter case, the NTD and the RBD can evolve in concert by increasing their surface electrostatic potential, which can be visualized when observing the upper face of the spike protein trimer. This is the case for all SARS-CoV-2 variants, from the original Wuhan strain to the Delta variant, but not for the Omicron series. This series of variants differs significantly from their predecessors because, for them, the increase in surface potential essentially concerns the RBD. This unexpected asymmetry explains why the T-index of the first Omicron variant is lower than that of the last Delta variant [42].

References

  1. Chermann, J.C.; Donker, G.; Yahi, N.; Salaun, D.; Guettari, N.; Gayet, O.; Hirsh, I. Discrepancies in AIDS virus data. Nature 1991, 351, 277–278.
  2. Venaud, S.; Yahi, N.; Fehrentz, J.L.; Guettari, N.; Nisato, D.; Hirsch, I.; Chermann, J.C. Inhibition of HIV by an anti-HIV protease synthetic peptide blocks an early step of viral replication. Res. Virol. 1992, 143, 311–319.
  3. Fantini, J.; Yahi, N.; Chermann, J.C. Human immunodeficiency virus can infect the apical and basolateral surfaces of human colonic epithelial cells. Proc. Natl. Acad. Sci. USA 1991, 88, 9297–9301.
  4. Yahi, N.; Baghdiguian, S.; Moreau, H.; Fantini, J. Galactosyl ceramide (or a closely related molecule) is the receptor for human immunodeficiency virus type 1 on human colon epithelial HT29 cells. J. Virol. 1992, 66, 4848–4854.
  5. Harouse, J.M.; Bhat, S.; Spitalnik, S.L.; Laughlin, M.; Stefano, K.; Silberberg, D.H.; Gonzalez-Scarano, F. Inhibition of entry of HIV-1 in neural cell lines by antibodies against galactosyl ceramide. Science 1991, 253, 320–323.
  6. Fantini, J.; Yahi, N.; Baghdiguian, S.; Chermann, J.C. Human colon epithelial cells productively infected with human immunodeficiency virus show impaired differentiation and altered secretion. J. Virol. 1992, 66, 580–585.
  7. Dayanithi, G.; Yahi, N.; Baghdiguian, S.; Fantini, J. Intracellular calcium release induced by human immunodeficiency virus type 1 (HIV-1) surface envelope glycoprotein in human intestinal epithelial cells: A putative mechanism for HIV-1 enteropathy. Cell Calcium 1995, 18, 9–18.
  8. Delézay, O.; Yahi, N.; Tamalet, C.; Baghdiguian, S.; Boudier, J.A.; Fantini, J. Direct effect of type 1 human immunodeficiency virus (HIV-1) on intestinal epithelial cell differentiation: Relationship to HIV-1 enteropathy. Virology 1997, 238, 231–242.
  9. Fantini, J.; Maresca, M.; Hammache, D.; Yahi, N.; Delézay, O. Glycosphingolipid (GSL) microdomains as attachment platforms for host pathogens and their toxins on intestinal epithelial cells: Activation of signal transduction pathways and perturbations of intestinal absorption and secretion. Glycoconj. J. 2000, 17, 173–179.
  10. Maresca, M.; Mahfoud, R.; Garmy, N.; Kotler, D.P.; Fantini, J.; Clayton, F. The virotoxin model of HIV-1 enteropathy: Involvement of GPR15/Bob and galactosylceramide in the cytopathic effects induced by HIV-1 gp120 in the HT-29-D4 intestinal cell line. J. Biomed. Sci. 2003, 10, 156–166.
  11. Jackson, C.B.; Farzan, M.; Chen, B.; Choe, H. Mechanisms of SARS-CoV-2 entry into cells. Nat. Rev. Mol. Cell Biol. 2022, 23, 3–20.
  12. 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.e775.
  13. Negi, G.; Sharma, A.; Dey, M.; Dhanawat, G.; Parveen, N. Membrane attachment and fusion of HIV-1, influenza A, and SARS-CoV-2: Resolving the mechanisms with biophysical methods. Biophys. Rev. 2022, 14, 1109–1140.
  14. Sasaki, M.; Uemura, K.; Sato, A.; Toba, S.; Sanaki, T.; Maenaka, K.; Hall, W.W.; Orba, Y.; Sawa, H. SARS-CoV-2 variants with mutations at the S1/S2 cleavage site are generated in vitro during propagation in TMPRSS2-deficient cells. PLoS Pathog. 2021, 17, e1009233.
  15. Kawase, M.; Shirato, K.; Matsuyama, S.; Taguchi, F. Protease-Mediated Entry via the Endosome of Human Coronavirus 229E. J. Virol. 2009, 83, 712–721.
  16. Cai, Y.; Zhang, J.; Xiao, T.; Peng, H.; Sterling, S.M.; Walsh, R.M., Jr.; Rawson, S.; Rits-Volloch, S.; Chen, B. Distinct conformational states of SARS-CoV-2 spike protein. Science 2020, 369, 1586–1592.
  17. 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.
  18. Fantini, J.; Chahinian, H.; Yahi, N. Leveraging coronavirus binding to gangliosides for innovative vaccine and therapeutic strategies against COVID-19. Biochem. Biophys. Res. Commun. 2021, 538, 132–136.
  19. Fantini, J.; Garmy, N.; Mahfoud, R.; Yahi, N. Lipid rafts: Structure, function and role in HIV, Alzheimer’s and prion diseases. Expert Rev. Mol. Med. 2002, 4, 1–22.
  20. Fantini, J.; Yahi, N.; Azzaz, F.; Chahinian, H. Structural dynamics of SARS-CoV-2 variants: A health monitoring strategy for anticipating Covid-19 outbreaks. J. Infect. 2021, 83, 197–206.
  21. Sun, X.L. The role of cell surface sialic acids for SARS-CoV-2 infection. Glycobiology 2021, 31, 1245–1253.
  22. Palacios-Rápalo, S.N.; De Jesús-González, L.A.; Cordero-Rivera, C.D.; Farfan-Morales, C.N.; Osuna-Ramos, J.F.; Martínez-Mier, G.; Quistián-Galván, J.; Muñoz-Pérez, A.; Bernal-Dolores, V.; Del Ángel, R.M.; et al. Cholesterol-Rich Lipid Rafts as Platforms for SARS-CoV-2 Entry. Front. Immunol. 2021, 12, 796855.
  23. Seyran, M.; Takayama, K.; Uversky, V.N.; Lundstrom, K.; Palù, G.; Sherchan, S.P.; Attrish, D.; Rezaei, N.; Aljabali, A.A.A.; Ghosh, S.; et al. The structural basis of accelerated host cell entry by SARS-CoV-2. FEBS J. 2021, 288, 5010–5020.
  24. Fantini, J.; Di Scala, C.; Chahinian, H.; Yahi, N. Structural and molecular modelling studies reveal a new mechanism of action of chloroquine and hydroxychloroquine against SARS-CoV-2 infection. Int. J. Antimicrob. Agents 2020, 55, 105960.
  25. Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L.; et al. Structure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptor. Nature 2020, 581, 215–220.
  26. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Krüger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.-H.; Nitsche, A. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 2020, 181, 271–280.e278.
  27. Pak, A.J.; Yu, A.; Ke, Z.; Briggs, J.A.G.; Voth, G.A. Cooperative multivalent receptor binding promotes exposure of the SARS-CoV-2 fusion machinery core. Nat. Commun. 2022, 13, 1002.
  28. Guérin, P.; Yahi, N.; Azzaz, F.; Chahinian, H.; Sabatier, J.M.; Fantini, J. Structural Dynamics of the SARS-CoV-2 Spike Protein: A 2-Year Retrospective Analysis of SARS-CoV-2 Variants (from Alpha to Omicron) Reveals an Early Divergence between Conserved and Variable Epitopes. Molecules 2022, 27, 3851.
  29. Colson, P.; Lavagna, C.; Delerce, J.; Groshenry, G.; Yahi, N.; Fantini, J.; La Scola, B.; Althaus, T. First Detection of the SARS-CoV-2 Omicron BA.5/22B in Monaco. Microorganisms 2022, 10, 1952.
  30. Yahi, N.; Chahinian, H.; Fantini, J. Infection-enhancing anti-SARS-CoV-2 antibodies recognize both the original Wuhan/D614G strain and Delta variants. A potential risk for mass vaccination? J. Infect. 2021, 83, 607–635.
  31. Al-Khatib, H.A.; Smatti, M.K.; Ali, F.H.; Zedan, H.T.; Thomas, S.; Ahmed, M.N.; El-Kahlout, R.A.; Al Bader, M.A.; Elgakhlab, D.; Coyle, P.V.; et al. Comparative analysis of within-host diversity among vaccinated COVID-19 patients infected with different SARS-CoV-2 variants. iScience 2022, 25, 105438.
  32. Azzaz, F.; Yahi, N.; Di Scala, C.; Chahinian, H.; Fantini, J. Ganglioside binding domains in proteins: Physiological and pathological mechanisms. Adv. Protein Chem. Struct. Biol. 2022, 128, 289–324.
  33. Gowrisankar, A.; Priyanka, T.M.C.; Banerjee, S. Omicron: A mysterious variant of concern. Eur. Phys. J. Plus 2022, 137, 100.
  34. Mahase, E. Omicron: South Africa says fourth wave peak has passed as it lifts curfew. BMJ Br. Med. J. 2022, 376, o7.
  35. Haschka, T.; Vergu, E.; Roche, B.; Poletto, C.; Opatowski, L. Retrospective analysis of SARS-CoV-2 omicron invasion over delta in French regions in 2021-22: A status-based multi-variant model. BMC Infect. Dis. 2022, 22, 815.
  36. Bal, A.; Simon, B.; Destras, G.; Chalvignac, R.; Semanas, Q.; Oblette, A.; Quéromès, G.; Fanget, R.; Regue, H.; Morfin, F.; et al. Detection and prevalence of SARS-CoV-2 co-infections during the Omicron variant circulation in France. Nat. Commun. 2022, 13, 6316.
  37. Grabowski, F.; Kochańczyk, M.; Lipniacki, T. The Spread of SARS-CoV-2 Variant Omicron with a Doubling Time of 2.0-3.3 Days Can Be Explained by Immune Evasion. Viruses 2022, 14, 294.
  38. Zhang, L.; Li, Q.; Liang, Z.; Li, T.; Liu, S.; Cui, Q.; Nie, J.; Wu, Q.; Qu, X.; Huang, W. The significant immune escape of pseudotyped SARS-CoV-2 variant Omicron. Emerg. Microbes Infect. 2022, 11, 1–5.
  39. Willett, B.J.; Grove, J.; MacLean, O.A.; Wilkie, C.; De Lorenzo, G.; Furnon, W.; Cantoni, D.; Scott, S.; Logan, N.; Ashraf, S.; et al. SARS-CoV-2 Omicron is an immune escape variant with an altered cell entry pathway. Nat. Microbiol. 2022, 7, 1161–1179.
  40. Zhao, H.; Lu, L.; Peng, Z.; Chen, L.-L.; Meng, X.; Zhang, C.; Ip, J.D.; Chan, W.-M.; Chu, A.W.-H.; Chan, K.-H.; et al. SARS-CoV-2 Omicron variant shows less efficient replication and fusion activity when compared with Delta variant in TMPRSS2-expressed cells. Emerg. Microbes Infect. 2022, 11, 277–283.
  41. Willett, B.J.; Grove, J.; MacLean, O.; Wilkie, C.; Logan, N.; De Lorenzo, G.; Furnon, W.; Scott, S.; Manali, M.; Szemiel, A. The hyper-transmissible SARS-CoV-2 Omicron variant exhibits significant antigenic change, vaccine escape and a switch in cell entry mechanism. MedRxiv 2022.
  42. Fantini, J.; Yahi, N.; Colson, P.; Chahinian, H.; La Scola, B.; Raoult, D. The puzzling mutational landscape of the SARS-2-variant Omicron. J. Med. Virol. 2022, 94, 2019–2025.
  43. Jaafar, R.; Boschi, C.; Aherfi, S.; Bancod, A.; Le Bideau, M.; Edouard, S.; Colson, P.; Chahinian, H.; Raoult, D.; Yahi, N.; et al. High Individual Heterogeneity of Neutralizing Activities against the Original Strain and Nine Different Variants of SARS-CoV-2. Viruses 2021, 13, 2177.
  44. Akaishi, T.; Fujiwara, K.; Ishii, T. Insertion/deletion hotspots in the Nsp2, Nsp3, S1, and ORF8 genes of SARS-related coronaviruses. BMC Ecol. Evol. 2022, 22, 123.
  45. Tamalet, C.; Izopet, J.; Koch, N.; Fantini, J.; Yahi, N. Stable rearrangements of the beta3-beta4 hairpin loop of HIV-1 reverse transcriptase in plasma viruses from patients receiving combination therapy. AIDS 1998, 12, F161–F166.
  46. Tamalet, C.; Yahi, N.; Tourrès, C.; Colson, P.; Quinson, A.M.; Poizot-Martin, I.; Dhiver, C.; Fantini, J. Multidrug resistance genotypes (insertions in the beta3-beta4 finger subdomain and MDR mutations) of HIV-1 reverse transcriptase from extensively treated patients: Incidence and association with other resistance mutations. Virology 2000, 270, 310–316.
  47. Li, Z.; Li, W.; Lu, M.; Bess, J., Jr.; Chao, C.W.; Gorman, J.; Terry, D.S.; Zhang, B.; Zhou, T.; Blanchard, S.C.; et al. Subnanometer structures of HIV-1 envelope trimers on aldrithiol-2-inactivated virus particles. Nat. Struct. Mol. Biol. 2020, 27, 726–734.
  48. Willey, R.L.; Bonifacino, J.S.; Potts, B.J.; Martin, M.A.; Klausner, R.D. Biosynthesis, cleavage, and degradation of the human immunodeficiency virus 1 envelope glycoprotein gp160. Proc. Natl. Acad. Sci. USA 1988, 85, 9580–9584.
  49. Chatterjee, S.; Basak, S.; Khan, N.C. Morphogenesis of human immunodeficiency virus type 1. Pathobiol. J. Immunopathol. Mol. Cell. Biol. 1992, 60, 181–186.
  50. Turner, B.G.; Summers, M.F. Structural biology of HIV. J. Mol. Biol. 1999, 285, 1–32.
  51. Fantini, J.; Chahinian, H.; Yahi, N. A Vaccine Strategy Based on the Identification of an Annular Ganglioside Binding Motif in Monkeypox Virus Protein E8L. Viruses 2022, 14, 2531.
  52. Hammache, D.; Yahi, N.; Piéroni, G.; Ariasi, F.; Tamalet, C.; Fantini, J. Sequential interaction of CD4 and HIV-1 gp120 with a reconstituted membrane patch of ganglioside GM3: Implications for the role of glycolipids as potential HIV-1 fusion cofactors. Biochem. Biophys. Res. Commun. 1998, 246, 117–122.
  53. Hug, P.; Lin, H.M.; Korte, T.; Xiao, X.; Dimitrov, D.S.; Wang, J.M.; Puri, A.; Blumenthal, R. Glycosphingolipids promote entry of a broad range of human immunodeficiency virus type 1 isolates into cell lines expressing CD4, CXCR4, and/or CCR5. J. Virol. 2000, 74, 6377–6385.
  54. Popik, W.; Alce, T.M.; Au, W.C. Human immunodeficiency virus type 1 uses lipid raft-colocalized CD4 and chemokine receptors for productive entry into CD4(+) T cells. J. Virol. 2002, 76, 4709–4722.
  55. Sorice, M.; Garofalo, T.; Misasi, R.; Longo, A.; Mattei, V.; Sale, P.; Dolo, V.; Gradini, R.; Pavan, A. Evidence for cell surface association between CXCR4 and ganglioside GM3 after gp120 binding in SupT1 lymphoblastoid cells. FEBS Lett. 2001, 506, 55–60.
  56. Nguyen, D.H.; Giri, B.; Collins, G.; Taub, D.D. Dynamic reorganization of chemokine receptors, cholesterol, lipid rafts, and adhesion molecules to sites of CD4 engagement. Exp. Cell Res. 2005, 304, 559–569.
  57. Carter, G.C.; Bernstone, L.; Sangani, D.; Bee, J.W.; Harder, T.; James, W. HIV entry in macrophages is dependent on intact lipid rafts. Virology 2009, 386, 192–202.
  58. Campbell, S.M.; Crowe, S.M.; Mak, J. Lipid rafts and HIV-1: From viral entry to assembly of progeny virions. J. Clin. Virol. 2001, 22, 217–227.
  59. Mañes, S.; del Real, G.; Lacalle, R.A.; Lucas, P.; Gómez-Moutón, C.; Sánchez-Palomino, S.; Delgado, R.; Alcamí, J.; Mira, E.; Martínez, A.C. Membrane raft microdomains mediate lateral assemblies required for HIV-1 infection. EMBO Rep. 2000, 1, 190–196.
  60. Kozak, S.L.; Heard, J.M.; Kabat, D. Segregation of CD4 and CXCR4 into distinct lipid microdomains in T lymphocytes suggests a mechanism for membrane destabilization by human immunodeficiency virus. J. Virol. 2002, 76, 1802–1815.
  61. Fragoso, R.; Ren, D.; Zhang, X.; Su, M.W.; Burakoff, S.J.; Jin, Y.J. Lipid raft distribution of CD4 depends on its palmitoylation and association with Lck, and evidence for CD4-induced lipid raft aggregation as an additional mechanism to enhance CD3 signaling. J. Immunol. 2003, 170, 913–921.
  62. Sorice, M.; Garofalo, T.; Misasi, R.; Longo, A.; Mikulak, J.; Dolo, V.; Pontieri, G.M.; Pavan, A. Association between GM3 and CD4-Ick complex in human peripheral blood lymphocytes. Glycoconj. J. 2000, 17, 247–252.
  63. Sattentau, Q.J.; Moore, J.P. Conformational changes induced in the human immunodeficiency virus envelope glycoprotein by soluble CD4 binding. J. Exp. Med. 1991, 174, 407–415.
  64. Pinter, A.; Honnen, W.J.; Tilley, S.A. Conformational changes affecting the V3 and CD4-binding domains of human immunodeficiency virus type 1 gp120 associated with env processing and with binding of ligands to these sites. J. Virol. 1993, 67, 5692–5697.
  65. Hammache, D.; Yahi, N.; Maresca, M.; Piéroni, G.; Fantini, J. Human erythrocyte glycosphingolipids as alternative cofactors for human immunodeficiency virus type 1 (HIV-1) entry: Evidence for CD4-induced interactions between HIV-1 gp120 and reconstituted membrane microdomains of glycosphingolipids (Gb3 and GM3). J. Virol. 1999, 73, 5244–5248.
  66. Hammache, D.; Piéroni, G.; Yahi, N.; Delézay, O.; Koch, N.; Lafont, H.; Tamalet, C.; Fantini, J. Specific interaction of HIV-1 and HIV-2 surface envelope glycoproteins with monolayers of galactosylceramide and ganglioside GM3. J. Biol. Chem. 1998, 273, 7967–7971.
  67. Rawat, S.S.; Johnson, B.T.; Puri, A. Sphingolipids: Modulators of HIV-1 infection and pathogenesis. Biosci. Rep. 2005, 25, 329–343.
  68. Cormier, E.G.; Dragic, T. The crown and stem of the V3 loop play distinct roles in human immunodeficiency virus type 1 envelope glycoprotein interactions with the CCR5 coreceptor. J. Virol. 2002, 76, 8953–8957.
  69. Gallo, S.A.; Finnegan, C.M.; Viard, M.; Raviv, Y.; Dimitrov, A.; Rawat, S.S.; Puri, A.; Durell, S.; Blumenthal, R. The HIV Env-mediated fusion reaction. Biochim. Et Biophys. Acta (BBA)—Biomembr. 2003, 1614, 36–50.
  70. Richman, D.D.; Bozzette, S.A. The Impact of the Syncytium-Inducing Phenotype of Human Immunodeficiency Virus on Disease Progression. J. Infect. Dis. 1994, 169, 968–974.
  71. Rosen, O.; Sharon, M.; Quadt-Akabayov, S.R.; Anglister, J. Molecular switch for alternative conformations of the HIV-1 V3 region: Implications for phenotype conversion. Proc. Natl. Acad. Sci. USA 2006, 103, 13950–13955.
  72. Kuiken, C.L.; De Jong, J.; Baan, E.; Keulen, W.; Tersmette, M.; Goudsmit, J. Evolution of the V3 envelope domain in proviral sequences and isolates of human immunodeficiency virus type 1 during transition of the viral biological phenotype. J. Virol. 1992, 66, 4622–4627.
  73. Xiao, L.; Owen, S.M.; Goldman, I.; Lal, A.A.; deJong, J.J.; Goudsmit, J.; Lal, R.B. CCR5 Coreceptor Usage of Non-Syncytium-Inducing Primary HIV-1 Is Independent of Phylogenetically Distinct Global HIV-1 Isolates: Delineation of Consensus Motif in the V3 Domain That Predicts CCR-5 Usage. Virology 1998, 240, 83–92.
  74. Björndal, A.; Deng, H.; Jansson, M.; Fiore, J.R.; Colognesi, C.; Karlsson, A.; Albert, J.; Scarlatti, G.; Littman, D.R.; Fenyö, E.M. Coreceptor usage of primary human immunodeficiency virus type 1 isolates varies according to biological phenotype. J. Virol. 1997, 71, 7478–7487.
  75. Balasubramanian, C.; Chillemi, G.; Abbate, I.; Capobianchi, M.R.; Rozera, G.; Desideri, A. Importance of V3 loop flexibility and net charge in the context of co-receptor recognition. A molecular dynamics study on HIV gp120. J. Biomol. Struct. Dyn. 2012, 29, 879–891.
  76. Repits, J.; Sterjovski, J.; Badia-Martinez, D.; Mild, M.; Gray, L.; Churchill, M.J.; Purcell, D.F.; Karlsson, A.; Albert, J.; Fenyö, E.M. Primary HIV-1 R5 isolates from end-stage disease display enhanced viral fitness in parallel with increased gp120 net charge. Virology 2008, 379, 125–134.
  77. Kalinina, O.V.; Pfeifer, N.; Lengauer, T. Modelling binding between CCR5 and CXCR4 receptors and their ligands suggests the surface electrostatic potential of the co-receptor to be a key player in the HIV-1 tropism. Retrovirology 2013, 10, 130.
  78. López de Victoria, A.; Kieslich, C.A.; Rizos, A.K.; Krambovitis, E.; Morikis, D. Clustering of HIV-1 subtypes based on gp120 V3 loop electrostatic properties. BMC Biophys. 2012, 5, 3.
  79. Shioda, T.; Levy, J.A.; Cheng-Mayer, C. Small amino acid changes in the V3 hypervariable region of gp120 can affect the T-cell-line and macrophage tropism of human immunodeficiency virus type 1. Proc. Natl. Acad. Sci. USA 1992, 89, 9434–9438.
  80. Clayton, F.; Kotler, D.P.; Kuwada, S.K.; Morgan, T.; Stepan, C.; Kuang, J.; Le, J.; Fantini, J. Gp120-induced Bob/GPR15 activation: A possible cause of human immunodeficiency virus enteropathy. Am. J. Pathol. 2001, 159, 1933–1939.
  81. Brenneman, D.E.; Westbrook, G.L.; Fitzgerald, S.P.; Ennist, D.L.; Elkins, K.L.; Ruff, M.R.; Pert, C.B. Neuronal cell killing by the envelope protein of HIV and its prevention by vasoactive intestinal peptide. Nature 1988, 335, 639–642.
  82. Yokoyama, M.; Naganawa, S.; Yoshimura, K.; Matsushita, S.; Sato, H. Structural dynamics of HIV-1 envelope Gp120 outer domain with V3 loop. PloS ONE 2012, 7, e37530.
More
Information
Subjects: Virology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 560
Entry Collection: COVID-19
Revisions: 2 times (View History)
Update Date: 23 Apr 2023
1000/1000
Video Production Service