key: cord-335075-6wo2o5pp authors: Bangaru, Sandhya; Ozorowski, Gabriel; Turner, Hannah L.; Antanasijevic, Aleksandar; Huang, Deli; Wang, Xiaoning; Torres, Jonathan L.; Diedrich, Jolene K.; Tian, Jing-Hui; Portnoff, Alyse D.; Patel, Nita; Massare, Michael J.; Yates, John R.; Nemazee, David; Paulson, James C.; Glenn, Greg; Smith, Gale; Ward, Andrew B. title: Structural analysis of full-length SARS-CoV-2 spike protein from an advanced vaccine candidate date: 2020-08-06 journal: bioRxiv DOI: 10.1101/2020.08.06.234674 sha: doc_id: 335075 cord_uid: 6wo2o5pp Vaccine efforts against the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) responsible for the current COVID-19 pandemic are focused on SARS-CoV-2 spike glycoprotein, the primary target for neutralizing antibodies. Here, we performed cryo-EM and site-specific glycan analysis of one of the leading subunit vaccine candidates from Novavax based on a full-length spike protein formulated in polysorbate 80 (PS 80) detergent. Our studies reveal a stable prefusion conformation of the spike immunogen with slight differences in the S1 subunit compared to published spike ectodomain structures. Interestingly, we also observed novel interactions between the spike trimers allowing formation of higher order spike complexes. This study confirms the structural integrity of the full-length spike protein immunogen and provides a basis for interpreting immune responses to this multivalent nanoparticle immunogen. Severe acute respiratory syndrome coronavirus (SARS-CoV) caused a global outbreak from 2002-2003 causing severe pneumonia and killing almost 900 people (1). SARS-CoV-2, belongs to the same lineage of the β-CoV genus as SARS-CoV, and recently emerged in China, spreading rapidly and infecting more than 18 million people worldwide with cases continuing to rise each day (2) . Given the global increase in population density, urbanization, and mobility, and the uncertain future behavior of the virus, vaccination is a critical tool for the response to this pandemic. The SARS-CoV-2 spike (S) trimeric glycoprotein is a focus of coronavirus vaccine development since it is a major component of the virus envelope, essential for receptor binding and virus entry, and a major target of host immune defense (3, 4) . There are several currently ongoing efforts to make spikebased vaccines using different strategies (4) (5) (6) . The CoV S protein is synthesized as an inactive precursor (S0) that gets proteolytically cleaved into S1 and S2 subunits which remain non-covalently linked to form functional prefusion trimers (7) . Like other type 1 fusion proteins, the SARS-CoV-2 S prefusion trimer is metastable and undergoes large-scale structural rearrangement from a prefusion to a thermostable post fusion conformation upon S-protein receptor binding and cleavage (8, 9) . Rearrangement exposes the hydrophobic fusion peptide (FP) allowing insertion into the host cell membrane, facilitating virus/host cell membrane alignment, fusion, and virus entry. Notably, SARS-CoV-2 S has a 4 amino acid insertion (PRRA) in the S1/S2 cleavage site compared to SARS-CoV spike resulting in a polybasic RRAR furin-like cleavage motif that enhances infection of lung cells (10, 11) . While the S2 subunit is relatively more conserved across the β-CoV genus, the S1 subunit comprising the receptor binding domain (RBD) is immunodominant and much less conserved (12) . The FP, two heptad repeats (HR1 and HR2), transmembrane (TM) domain, and cytoplasmic tail (CT) are located in the S2 subdomain that encompasses the fusion machinery. The S1 subunit of SARS-CoV-2 S folds into 4 distinct domains; the Nterminal (NTD), the C-terminal domain (CTD) containing the RBD and two subdomains, SD1 and SD2. While some human CoVs (HCoV), including OC43, exclusively use NTDsialic acid interactions as their receptor engagement, others like Middle East Respiratory Syndrome (MERS) CoV that use the CTD-RBD for primary receptor binding have also been reported to bind sialic acid receptors via their NTD to aid initial attachment to the host cells (13) (14) (15) . Although SARS-CoV-2 primarily interacts with its receptor ACE2 through the CTD-RBD, there is currently no evidence indicating possible interactions between the NTD and sialoglycans (16, 17) . The structure of the stabilized SARS-CoV-2 spike ectodomain has been solved in its prefusion conformation and exhibits a high resemblance to SARS-CoV spike (17) (18) (19) . In this report, we describe the atomic structure of a leading SARS-CoV-2 S vaccine candidate based on a full-length S gene with furin cleavage-resistant mutations in the S1/S2 cleavage site and the presence or absence of 2-proline amino acid substitutions at the apex of the central helix. Our studies reveal an overall shift in conformation of the S1 subunit compared to the previously published structures (17) (18) (19) . Interestingly, we also observed direct interactions between adjacent spike trimers; the flexible loop between residues 615-635 in the SD2 from each trimer extending and engaging a binding pocket on the NTD of the adjacent trimers resulting in higher order spike multimers. Further, sitespecific glycan analysis revealed the glycan occupancy as well as varying levels of glycan processing at the 22 N-glycosylation sequons present in the spike monomer. Thus, our studies provide in-depth structural analysis of the Novavax full-length vaccine candidate, currently being tested in humans, that appropriately recapitulates the prefusion spike. The SARS-CoV-2-3Q-2P full-length spike vaccine candidate (3Q-2P-FL) was engineered from the full-length SARS-CoV-2 spike gene (residues 1-1273) including the transmembrane domain (TM) and the cytoplasmic tail (CT) (Fig1a). The construct was modified at the S1/S2 polybasic cleavage site from RRAR to QQAQ to render it protease resistant along with 2 proline substitutions at residues K986 and V987 in the S2 fusion machinery core for enhanced stability ( Figure 1A) . A second full-length construct (3Q-FL) containing the cleavage site mutations but without the 2 proline substitutions was also made in parallel for comparative purposes. The FL spikes expressed and purified from insect cells were then formulated in 0.01% (v/v) polysorbate 80 (PS 80) detergent. To characterize the structural integrity of 3Q-2P-FL immunogen, we performed negative stain electron microscopy (nsEM) of the FL spike constituted in PS 80 in the presence of Matrix-M adjuvant, recapitulating the vaccine formulation being tested in humans. Micrographs and 2D classes of the particles revealed trimeric spike proteins present as free trimers or as multi-trimer rosettes with their transmembrane domains enclosed in cylindrical micellar cores of PS 80 detergent ( Figure 1B ). Micrographs show these 3Q-2P-FL spike nanoparticles containing as many as 14 trimers. The Matrix-M adjuvant can also be seen in the micrographs and 2D classes as spherical cages of different sizes with only very limited interaction with spike particles. To further evaluate the structural features of the 3Q-2P-FL immunogen, we performed single particle cryo-EM on the spike formulated in PS 80 detergent. The raw micrographs from cryo-EM revealed free trimers and trimer rosettes similar to those observed on the negative stain micrographs (Figure 2A ). Initial 2D classifications revealed the presence of 2 distinct classes: free spike trimers and dimers of trimers ( Figure 2A ). Each class was independently subjected to additional classification and refinement. The three-fold symmetry (C3) reconstruction of the free spike trimer resulted in a map of 3.6 Å resolution while the asymmetric reconstruction (C1) was resolved to 3.8 Å resolution ( Figure 2B and S1A, S1B). Previously published structures of soluble, stabilized SARS-CoV-2 spikes have revealed that RBDs exist in either a closed (RBDdown) or an open (RBD-up) conformation that can engage in ACE2 binding (16) (17) (18) . In contrast, we observed that all three RBDs on the 3Q-2P-FL spike trimer were present in the closed conformation in the asymmetric reconstruction; the higher resolution C3 map was consequently used for model building ( Figure 2B and S1C). Overall, the map was well resolved in both S1 and S2 subunits, particularly in the S1 NTD and CTD domains that were less resolved in previously published structures, thereby enabling us to model the full extent of these domains. Notably, the local resolution map calculated using cryoSPARC showed much of the spike trimer at substantially higher resolution than 3.6 Å ( Figure S1D ). The atomic model contains residues 14-1146 with breaks only in the flexible loop (619-631) and the cleavage site (678-688) ( Figure 2C ). Interestingly, superimposition of the coordinate models of 3Q-2P-FL spike with published spike structures (PDB Id: 6VXX and 6VSB) revealed substantial domain rearrangements in the S1 subunit of 3Q-2P-FL spike compared to the other models, whereas the structure of the S2 subunit was consistent with the published data ( Figure 2D ). The S1 NTD differed the most (~14° rotation counterclockwise relative to published models when viewing down towards the viral membrane) while the CTD and subdomains showed minor local rearrangements ( Figure 2D) . Notably, we also observed shifts in the placement of residues flanking the 615-635 loop compared to the published models. This region was modeled in one of the published structures (PDB Id: 6X6P) as a helix with residues flanking the helix positioned very differently from our model as shown by the corresponding placement of residues T632 and T618 (residues flanking the gap in 3Q-2P-FL model) ( Figure 3A ). However, upon closer inspection of the cryo-EM density (EMD-22078) corresponding to residues 621-640 of the PDB model 6X6P, there is insufficient density to support the helix conformation of this region ( Figure S1E ). The resulting displacement of residues in the 3Q-2P-FL structure enables inter-protomeric interactions by creating a salt-bridge between residues Asp 614 and Lys 854 ( Figure 3B ). This observation is particularly interesting given the increased prevalence of D614G mutation in the emerging SARS-CoV-2 strains and its potential role in viral transmission and pathogenesis (20) . During refinement of an atomic model into the EM density, we observed 2 additional densities in the S1 subunit that did not correspond to any peptide or glycans within the spike ( Figure S2A ). The first density was buried within a hydrophobic pocket of the CTD created by F338, F342, Y365, Y369, F374, F377, F392, F513 ( Figure 3C and S2B). We had previously observed a non-protein density situated in the structure of porcine epidemic diarrhea virus (PEDV) that was identified to be palmitoleic acid (21) . This pocket in SARS-CoV-2 CTD corresponded with the structure of linoleic acid, a polyunsaturated fatty acid; the presence of this ligand was confirmed by mass spectrometry of 3Q-2P-FL spike ( Figure S2B and S2C). The main chain carboxyl group of linoleic acid interacts with R408 and Q409 residues of RBD from the adjacent protomer thereby making interprotomer contacts ( Figure 3C ). The second unassigned density present in NTD was relatively larger and more surface exposed than the first density, surrounded by residues N121, Y170, S172, F175, R190, H207, V227 ( Figure 3D and S2D). Analysis of the structural features of this density suggested that it may correspond to PS 80 detergent used to solubilize the membrane-bound trimers and stabilize them in solution. The aliphatic tail of PS 80 fit well into the hydrophobic pocket while the carbonyl and hydroxyl groups were well placed in proximity to residues R190 and H207 with potential for multiple hydrogen bonds between them ( Figure 3D and S2D). Overall, the density is consistent with PS 80 detergent and, given its location, provides a possible explanation for the S1 shift seen in our FL trimer density compared to the published structures. Further classification of multimeric trimer particles yielded two separate classes; a dimerof-trimers class that reconstructed to a final resolution of 4.5 Å with 2-fold symmetry and a trimer-of-trimers class that was resolved to 8.0 Å resolution ( Figure 4A , 4B and S3A). The presence of the trimer-of-trimers class revealed that each spike trimer had the ability to interact with multiple trimers simultaneously. In both reconstructions, the interaction between each pair of trimers involved the SD2 of one protomer from each trimer engaging with the NTD of the adjacent trimer ( Figure 4C ). Consequently, each trimer pair is symmetrical along a 2-fold axis with trimer axes tilted to 44.5 degrees relative to each other. The atomic model of the dimer-of-trimer EM density revealed that the interaction was mainly coordinated by the 615-635 loop. Although most of the loop residues were too flexible to resolve in the free trimer density map, the inter-trimer interaction stabilized the loop so that it could be fully resolved ( Figure 4D ). The loop reaches into a pocket on the adjacent NTD, interacting with residues 621-PVAIHADQ-628 in the loop with NTD residues Q183, H146, Y248, L249, V70 and S71 ( Figure 4D ). We observed subtle changes in the NTD binding pocket in the loop-bound state compared to the free trimer model that allow better accommodation of the loop in the pocket. The residues Y145 and H146 in the binding pocket appear to switch positions in the loop-bound state resulting in a salt bridge interaction between H146 and D627 and potential stacking between W152 and H146 ( Figure 4E ). We also observed minor displacement of residues 68-75 and 248-250 surrounding the pocket. In addition to the main loop interaction resulting in higherorder oligomers, we also observed N282 glycans extending out towards the symmetry related chain in the adjacent trimer ( Figure S3B were replaced by a GS linker ( Figure 4F ). To investigate if the absence of 2 stabilizing proline mutations impacted the spike stability and formation of higher order multimers, we performed cryo-EM studies of the SARS-CoV-2-3Q-FL (without 2P) protein formulated in PS 80 detergent. The raw micrographs and 2D classes revealed the presence of free trimers as well as trimer-trimer complexes as observed with 3Q-2P-FL, indicating that the proline stabilization is not necessary for the formation of these higher order complexes ( Figure S4A ). The 3D refinement of free trimers was refined to 4 Å resolution imposing C3 symmetry as we observed that the RBDs were present in closed conformation similar to 3Q-2P-FL ( Figure S4B ). Fitting the 3Q-2P-FL model into the 3Q-FL map revealed identical conformation of the spike protein further supporting that the presence of 2P in the full-length immunogen does not lead to any structural changes in the spike protein ( Figure S4C ). Glycans on viral glycoproteins play a wide role in protein folding, stability, immune recognition and potentially in immune evasion. Site-specific glycosylation of the SARS-CoV-2 prefusion spike protein produced in SF9 insect cells was analyzed using our recently described mass spectrometry proteomics-based method, involving treatment with proteases followed by sequential treatment with the endoglycosidases (Endo H and PNGase F) to introduce mass signatures in peptides with N-linked sequons (Asn-X-Thr/Ser) to assess the extent of glycosylation and the degree of glycan processing from high mannose/hybrid type to complex type (24) . Although the method was developed to assess the degree of processing of N-linked glycans in mammalian cells, it is also applicable for analyzing glycosylation of SF9 insect cells. The primary differences in glycan processing of N-linked glycans in SF9 insect cells are: 1) the production of truncated paucimannose glycans, and 2) the potential to introduce either one (α1,6) or two (α1,6/α1,3) fucose substitutions into the core GlcNAc attached to Asn. Although α1,3 fucose substitution is known to prevent cleavage by PNGase F (25), this is not a factor when analyzing glycosylation from SF9 cells since they contain α1,6-fucosylatransferase, which is found in mammalian cells, but only contain trace amounts α-1,3fucosyltransferase activity, if any (26) . The paucimannose glycans are highly processed like complex type glycans and not cleaved by EndoH, but are cleaved by PNGase F. Thus, for SF9 insect cell-produced glycoproteins, the use of endoglycosidases to introduce mass signatures is analogous to analysis of glycoproteins produced in mammalian cells, with EndoH removing high mannose/hybrid glycans leaving a GlcNAc-Asn (+203), followed by treatment with PNGase F in O18 water which removes the remaining paucimannose and complex type glycans and while converting Asn to Asp (+3), and the Asn of unoccupied sites remains unaltered (+0). Our analysis detected glycosylation at all 22 potential N-linked glycan sequons present on SARS-CoV-2 spike ( Figure 4G ). Overall, there was high glycan occupancy of over >98%, with only two sites, 603 and 657, more than 5% unoccupied. Interestingly, we did not see clear glycan density at either 603 or 657 in the cryo-EM reconstruction of the 3Q-2P-FL spike. Most sites showed extensive glycan processing to complex/paucimannose type glycans, with only four sites that exhibit ≥40% oligomannose. The glycan analysis also confirmed the presence of glycans at sites 1158, 1173 and 1194 present in the membrane-proximal region of the spike not resolved by cryo-EM. The extensive site-specific glycan processing of the SARS-CoV-2 prefusion spike protein in SF9 insect cells seen here is similar to that recently reported for the spike protein produced in mammalian HEK293F cells (27) . The coronavirus disease (COVID-19) caused by SARS-CoV-2 poses a serious health threat and was declared a pandemic by the World Health Organization (WHO). In quick response to this rapidly evolving situation, several SARS-CoV-2 spike-based vaccine candidates are being developed and tested at various stages of clinical trials (4) (5) (6) . In this study, we performed structural analysis of the Novavax SARS-CoV We also observed two non-spike densities within the spike trimer that corresponded with linoleic acid and polysorbate 80 detergent. Linoleic acid, an essential free fatty acid, was buried within a hydrophobic pocket in the CTD with its main chain carboxyl group making contacts with the adjacent RBD in closed conformation. A recent report by Toelzer et al. also identified this density and attributed it to the presence of linoleic acid (34) . The second large density occupied by PS 80 is situated in the NTD and is relatively more surface exposed. Since PS 80 is unique to the formulation of the Novavax 3Q-2P-FL immunogen, this observation is specific to this structure. However, there is a possibility of other ligands occupying this pocket in the place of PS 80. The presence of these binding pockets for different ligands in the spike structure provide potential targets for drug design against SARS-CoV-2. The widely used SARS-CoV-2 spike ectodomain construct with mutated cleavage site and 2P substitution has been shown to partially exist in all RBD 'down' conformation or in one RBD 'up' conformation (17, 18) . Surprisingly, we observed that all the RBDs in the 3Q-2P-FL spike immunogen were present in a down confirmation, which could be a cause for concern for eliciting neutralizing antibodies that compete with ACE2 binding. However, binding analysis of the 3Q-2P-FL immunogen to ACE2 by both bio-layer interferometry and ELISA clearly show binding to ACE2, indicating that the RBD is dynamic and the receptor binding site accessible (31) . Another study on the prefusion structure of a full-length spike protein reported similar findings with RBDs clamped down as a consequence of potential clashes between S2 residues 828-853 and SD1 when RBD is in open conformation (35) . It is also possible that the interprotomeric contacts made by linoleic acid observed in our structure preferentially lead to the observed RBD positioning. Our structural work is consistent with the burgeoning body of structures available of the spike protein, albeit with the important differences described above. Hence, this advanced protein subunit vaccine candidate currently being tested in humans appears stable, homogeneous, and locked in the antigenically preferred prefusion conformation. Further, the tight clustering of the spikes in the nanoparticle formulation may lead to stronger immune responses over soluble trimers alone, consistent with other viral glycoprotein immunogens (HA and RSV F) (38, 39) . It appears that the remarkable speed at which this vaccine was designed did not compromise the quality of the immunogen, and that building off the previous success in formulating RSV F and influenza HA nanoparticle immunogens could readily be extended to SARS-CoV-2 spike, particularly in the background of the 2P mutations previously shown to stabilize many other β-CoV spikes (40) . With structural, biophysical, and antigenic characterization now complete, evaluation in humans will provide the true proof-of-principle for this vaccine concept. We thank Bill Anderson, Hannah L. Turner and Charles A. Bowman for their help with electron microscopy, data acquisition and data processing. We thank Bill Webb and Linh Truc Hoang for their assistance with mass spectrometry and data processing. We thank Lauren Holden for her assistance with the manuscript. Authors would also like to thank Figures S1 to S4 Table S1 References ( Figs. S1 to S4 Table S1 SUPPLEMENTARY MATERIALS 2-3Q-2P full-length spike formulated in PS80 and Matrix adjuvant were diluted to approximately 20 µg/mL with TBS. The sample was directly deposited onto carbon-coated 400-mesh copper grids and stained immediately with 2% (w/v) uranyl formate for 90 seconds. Grids were imaged at 120 KeV on Tecnai T12 Spirit with a 4k x 4k Eagle CCD camera at 52,000x magnification and -1.5 μm nominal defocus. Micrographs were collected using Leginon and the images were transferred to Appion for processing (41, 42) . Particle stacks were generated in Appion with particles picked using a difference-of-Gaussians picker (DoG-picker) and 2D classes generated by MSA/MRA (43, 44) . (45) . MotionCor2 was used for alignment and dose weighting of the frames (46) . Micrographs were transferred to CryoSPARC 2.9 for further processing (47) . CTF estimations were performed using GCTF and micrographs were selected using the Curate Exposures tool in CryoSPARC based on their CTF resolution estimates (cutoff 5 Å) for downstream particle picking, extraction and iterative rounds of 2D classification and selection (48) . Particles selected from 2D classes were used for 3D refinement of free trimers for 3Q-2P-FL and 3Q-FL datasets in CryoSPARC. Final subsets of clean trimer particles were refined with C3 symmetry and local resolution for the free trimer was calculated using the local resolution function in CryoSPARC. Particles corresponding to dimers-of-trimers classes in CryoSPARC were transferred to Relion 3.0 for iterative rounds of 3D classification to separate dimers-of-trimers and trimers-of-trimers (49) . Final subsets of clean particles from dimers-of-trimers class were refined with C2 symmetry and the trimers-of-trimers class with C1 symmetry. Model building and refinement. The 3.6 Å C3-symmetric free trimer map and the 4.5 Å C2-symmetric dimers-of-trimers maps were used for model building and refinement. Initial model building was performed manually in Coot using PDB 6VXX as a template followed by iterative rounds of Rosetta relaxed refinement and Coot manual refinement to generate the final models (50, 51) . EMRinger and MolProbity were run following each round of Rosetta refinement to evaluate and choose the best refined models (52, 53) . The coordinates were manually placed and refined into the respective map densities using Coot. For Rosetta refinement, each ligand was saved in MOL2 format and Rosetta parameter files were generated using the molfile_to_params.py function (51) . Final map and model statistics are summarized in Table S1 . Figures were generated using UCSF Chimera and UCSF Chimera X (55, 56) . Mass spectrometry to identify fatty acids in the SARS-3Q-2P-FL protein was performed as described previously (21) . We obtained several candidates in this screen that were narrowed down to 6 candidates based on their intensity and the m/z range of 250-300. A sample of the SARS-CoV-2 prefusion spike protein expressed in the SF9 insect cell line was prepared for MS analysis as previously described with minor modifications (24) . In brief, the protein (50 µg) was denatured and aliquots (10 µg The MS data were processed essentially as described previously (24) . The data were searched against the proteome database and quantified using peak area in Integrated Proteomics Pipeline-IP2. Since the processing pathway in SF9 cell line (insect cell line) is similar to mammalian cells for oligomannose and hybrid structures cleaved by Endo-H, and then diverges to produce a combination of paucimannose and complex type glycans, peptides with N+203 were identified as having oligomannose type glycans, and peptides with N+3 are assigned as peptides with complex and paucimannose type glycans. SARS: the first pandemic of the 21st century A Novel Coronavirus from Patients with Pneumonia in China Mechanisms of coronavirus cell entry mediated by the viral spike protein SARS-CoV-2 SPIKE PROTEIN: an optimal immunological target for vaccines SARS-CoV-2 Vaccines: Status Report Developing Covid-19 Vaccines at Pandemic Speed Structure, Function, and Evolution of Coronavirus Spike Proteins Tectonic conformational changes of a coronavirus spike glycoprotein promote membrane fusion The coronavirus spike protein is a class I virus fusion protein: structural and functional characterization of the fusion core complex The proximal origin of SARS-CoV-2 A Multibasic Cleavage Site in the Spike Protein of SARS-CoV-2 Is Essential for Infection of Human Lung Cells The receptor binding domain of the viral spike protein is an immunodominant and highly specific target of antibodies in SARS-CoV-2 patients Human coronaviruses OC43 and HKU1 bind to 9-O-acetylated sialic acids via a conserved receptor-binding site in spike protein domain A Identification of sialic acid-binding function for the Middle East respiratory syndrome coronavirus spike glycoprotein Structures of MERS-CoV spike glycoprotein in complex with sialoside attachment receptors Structural basis of receptor recognition by SARS-CoV-2 Structure, Function, and Antigenicity of the SARS-CoV-2 Spike Glycoprotein Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation Characterization of the SARS-CoV-2 S Protein: Biophysical, Biochemical, Structural, and Antigenic Analysis. bioRxiv The D614G mutation in the SARS-CoV-2 spike protein reduces S1 shedding and increases infectivity. bioRxiv Structure and immune recognition of the porcine epidemic diarrhea virus spike protein. bioRxiv In-Silico evidence for two receptors based strategy of SARS-CoV-2. bioRxiv N-terminal domain (NTD) of SARS-CoV-2 spike-protein structurally resembles MERS-CoV NTD sialoside-binding pocket Differential processing of HIV envelope glycans on the virus and soluble recombinant trimer Peptide-N4-(N-acetyl-beta-glucosaminyl)asparagine amidase F cannot release glycans with fucose attached alpha 1----3 to the asparagine-linked N-acetylglucosamine residue Distinct N-glycan fucosylation potentials of three lepidopteran cell lines Site-specific glycan analysis of the SARS-CoV-2 spike Antigen Receptor Nanoclusters: Small Units with Big Functions Molecular determinants of immunogenicity: the immunon model of immune response Polyvalent antigens stabilize B cell antigen receptor surface signaling microdomains SARS-CoV-2 spike glycoprotein vaccine candidate NVX-CoV2373 elicits immunogenicity in baboons and protection in mice A pH-dependent switch mediates conformational masking of SARS-CoV-2 spike. bioRxiv Structures, conformations and distributions of SARS-CoV-2 spike protein trimers on intact virions. bioRxiv Unexpected free fatty acid binding pocket in the cryo-EM structure of SARSCoV-2 spike protein. bioRxiv Distinct conformational states of SARS-CoV-2 spike protein. bioRxiv Controlling the SARS-CoV-2 Spike Glycoprotein Conformation. bioRxiv Glycans on the SARS-CoV-2 Spike Control the Receptor Binding Domain Conformation. bioRxiv Respiratory syncytial virus fusion glycoprotein expressed in insect cells form protein nanoparticles that induce protective immunity in cotton rats Improved Titers against Influenza Drift Variants with a Nanoparticle Vaccine Immunogenicity and structures of a rationally designed prefusion MERS-CoV spike antigen Leginon: a system for fully automated acquisition of 1000 electron micrographs a day Appion: an integrated, database-driven pipeline to facilitate EM image processing DoG Picker and TiltPicker: software tools to facilitate particle selection in single particle electron microscopy Topology representing network enables highly accurate classification of protein images taken by cryo electron-microscope without masking Automated molecular microscopy: the new Leginon system MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination Gctf: Real-time CTF determination and correction New tools for automated high-resolution cryo-EM structure determination in RELION-3 Structural analysis of glycoproteins: building N-linked glycans with Coot Automated structure refinement of macromolecular assemblies from cryo-EM maps using Rosetta EMRinger: side chain-directed model and map validation for 3D cryoelectron microscopy MolProbity: More and better reference data for improved all-atom structure validation electronic Ligand Builder and Optimization Workbench (eLBOW): a tool for ligand coordinate and restraint generation Meeting modern challenges in visualization and analysis UCSF Chimera--a visualization system for exploratory research and analysis Analysis Tool Web Services from the EMBL-EBI Isolation of potent SARS-CoV-2 neutralizing antibodies and protection from disease in a small animal model