Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Feb 1.
Published in final edited form as: Nature. 2021 Dec 23;602(7898):664–670. doi: 10.1038/s41586-021-04386-2

Broadly neutralizing antibodies overcome SARS-CoV-2 Omicron antigenic shift

Elisabetta Cameroni 1,*, John E Bowen 2,*, Laura E Rosen 3,*, Christian Saliba 1,*, Samantha K Zepeda 2, Katja Culap 1, Dora Pinto 1, Laura A VanBlargan 4, Anna De Marco 1, Julia di Iulio 3, Fabrizia Zatta 1, Hannah Kaiser 3, Julia Noack 3, Nisar Farhat 3, Nadine Czudnochowski 3, Colin Havenar-Daughton 3, Kaitlin R Sprouse 2, Josh R Dillen 3, Abigail E Powell 3, Alex Chen 3, Cyrus Maher 3, Li Yin 3, David Sun 3, Leah Soriaga 3, Jessica Bassi 1, Chiara Silacci-Fregni 1, Claes Gustafsson 5, Nicholas M Franko 6, Jenni Logue 6, Najeeha Talat Iqbal 7, Ignacio Mazzitelli 8, Jorge Geffner 8, Renata Grifantini 9, Helen Chu 6, Andrea Gori 10, Agostino Riva 11, Olivier Giannini 12,13, Alessandro Ceschi 12,14,15,16, Paolo Ferrari 12,17,18, Pietro E Cippà 13,17,19, Alessandra Franzetti-Pellanda 20, Christian Garzoni 21, Peter J Halfmann 22, Yoshihiro Kawaoka 22,23,24, Christy Hebner 3, Lisa A Purcell 3, Luca Piccoli 1, Matteo Samuele Pizzuto 1, Alexandra C Walls 2,25, Michael S Diamond 4,26,27, Amalio Telenti 3, Herbert W Virgin 3,26,28,29, Antonio Lanzavecchia 1,9,29, Gyorgy Snell 3,29, David Veesler 2,25,29, Davide Corti 1,29
PMCID: PMC9531318  NIHMSID: NIHMS1830838  PMID: 35016195

SUMMARY:

The recently emerged SARS-CoV-2 Omicron variant encodes 37 amino acid substitutions in the spike (S) protein, 15 of which are in the receptor-binding domain (RBD), thereby raising concerns about the effectiveness of available vaccines and antibody therapeutics. Here, we show that the Omicron RBD binds to human ACE2 with enhanced affinity, relative to the Wuhan-Hu-1 RBD, and binds to mouse ACE2. Marked reductions of plasma neutralizing activity were observed against Omicron compared to the ancestral pseudovirus for convalescent and vaccinated individuals, but this loss was less pronounced after a third vaccine dose. Most receptor-binding motif (RBM)-directed monoclonal antibodies (mAbs) lost in vitro neutralizing activity against Omicron, with only 3 out of 29 mAbs retaining unaltered potency, including the ACE2-mimicking S2K146 mAb1. Furthermore, a fraction of broadly neutralizing sarbecovirus mAbs neutralized Omicron through recognition of antigenic sites outside the RBM, including sotrovimab2, S2X2593 and S2H974. The magnitude of Omicron-mediated immune evasion marks a major SARS-CoV-2 antigenic shift. Broadly neutralizing mAbs recognizing RBD epitopes conserved among SARS-CoV-2 variants and other sarbecoviruses may prove key to controlling the ongoing pandemic and future zoonotic spillovers.

Keywords: SARS-CoV-2, COVID-19, antibody, vaccine, neutralizing antibodies, immune evasion

INTRODUCTION

The evolution of RNA viruses can result in immune escape and modulation of binding to host receptors through accumulation of mutations5. Previously emerged SARS-CoV-2 variants of concern (VOC) have developed resistance to neutralizing antibodies, including some clinical antibodies used as therapeutics68. The B.1.351 (Beta) VOC is endowed with the greatest magnitude of immune evasion from serum neutralizing antibodies6,7, whereas B.1.617.2 (Delta) quickly outcompeted all other circulating isolates through acquisition of mutations that enhanced transmission and pathogenicity911 and eroded the neutralizing activity of antibody responses9.

The Omicron (B.1.1.529) variant was first detected in November 2021, immediately declared by the WHO as a VOC and quickly rose in frequency worldwide. The Omicron variant is substantially mutated compared to any previously described SARS-CoV-2 isolates, including 37 S residue substitutions in the predominant haplotype (Fig. 1a and Extended Data Fig. 14). Fifteen of the Omicron mutations are clustered in the RBD, which is the main target of neutralizing antibodies after infection or vaccination12,13, suggesting that Omicron might escape infection- and vaccine-elicited Abs and therapeutic mAbs. Nine of these mutations map to the receptor-binding motif (RBM) which is the RBD subdomain directly interacting with the host receptor, ACE214.

Fig. 1. Omicron RBD shows increased binding to human ACE2 and gains binding to murine ACE2.

Fig. 1.

a, Omicron mutations are shown in a primary structure of SARS-CoV-2 S with domains and cleavage sites highlighted. b, Single-cycle kinetics SPR analysis of ACE2 binding to six RBD variants. ACE2 is injected successively at 11, 33, 100, and 300 nM (human) or 33, 100, 300, and 900 nM (mouse). Black curves show fits to a 1:1 binding model. White and gray stripes indicate association and dissociation phases, respectively. c, Quantification of human ACE2 binding data. Reporting average ± standard deviation of three replicates. Asterisks indicate that Delta was measured in a separate experiment with a different chip surface and capture tag; Delta fold-change is calculated relative to affinity of Wuhan-Hu-1 measured in parallel (91 ± 1.6 nM). d, Entry of Wu-Hu-1, Alpha, Beta, Delta, Gamma, Kappa and Omicron VSV pseudoviruses into mouse ACE2 expressing HEK293T cells. Shown are 2 biological replicates (technical triplicates). Lines, geometric mean.

Preliminary reports indicated that the neutralizing activity of plasma from Pfizer-BioNTech BNT162b2 vaccinated individuals is reduced against SARS-CoV-2 Omicron15,16, documenting a substantial, albeit not complete, escape from mRNA vaccine-elicited neutralizing antibodies. Another report also shows that vaccine effectiveness against symptomatic disease induced by the Omicron variant is significantly lower than for the Delta variant17. The potential for booster doses to ameliorate this decline in neutralization is being explored. In addition, the neutralizing activity of several therapeutic mAbs appears decreased or abolished against SARS-CoV-2 Omicron16,18.

To understand the consequences of the unprecedented number of mutations found in Omicron S, we employed a pseudovirus assay to study receptor usage and neutralization mediated by monoclonal and polyclonal antibodies as well as surface plasmon resonance to measure binding of the RBD to human and mouse ACE2 receptors.

RESULTS

The Omicron RBD binds with increased affinity to human ACE2 and gains binding to mouse ACE2

Twenty-three out of the 37 Omicron S amino acid mutations have been individually observed previously in SARS-CoV-2 variants of interest (VOI), VOC, or other sarbecoviruses, whereas the remaining 14 substitutions have not been described before (Extended Data Fig. 5a). Analysis of the GISAID database indicates that there are rarely more than 10–15 Omicron S mutations present in a given non-Omicron haplotype or Pango lineage (Extended Data Fig. 5bd). While we have not formally assessed the possibility of recombination events, persistent replication in immunocompromised individuals or inter-species ping-pong transmission5 are possible scenarios for the rapid accumulation of mutations that could have been selected based on viral fitness and immune evasion.

Several of the Omicron RBD mutations are found at positions that are key contact sites with human ACE2, such as K417N, Q493K and G496S19. Except for N501Y, which increases ACE2 binding affinity by 6-fold20,21, all other substitutions were shown by deep mutational scanning (DMS) to either reduce binding or to have no impact on human ACE2 affinity when present individually22, resulting in an overall predicted decrease of binding affinity (Supplementary Table 1). However, we found that the Omicron RBD has a 2.4-fold increased binding affinity to human ACE2 (Fig. 1b, c and Extended Data Figure 6a), suggesting epistasis of the full constellation of RBD mutations. It remains to be determined whether and how the S mutations in Omicron may influence the dynamics of RBD opening, which may also impact RBD engagement with ACE2.

The presence of the N501Y mutation has previously been described to enable some SARS-CoV-2 VOC to infect mice23. Since Omicron carries the N501Y mutation, along with 14 other RBD mutations, we investigated whether the Omicron RBD binds mouse ACE2 using surface plasmon resonance (SPR) (Fig. 1b and Extended Data Fig. 6). The Omicron RBD binds mouse ACE2 with a 1:1 binding affinity of 470 nM (Fig. 1b), whereas weak binding of the Beta RBD and very weak binding of the Alpha RBD to mouse ACE2 was observed (Fig. 1b and Extended Data Fig. 6b), consistent with previous reports23,24. Conversely, our assay did not detect any binding of the Wuhan-Hu-1, Delta, or K417N RBDs to mouse ACE2. The enhanced binding of the Omicron RBD to mouse ACE2 is likely explained by the Q493R substitution which is similar to the Q493K mutation isolated upon mouse-adaptation of SARS-CoV-219. Our binding data correlate with our observation of Omicron S-mediated but not Wuhan-Hu-1/G614 S-mediated entry of VSV pseudoviruses into mouse ACE2-expressing cells (Fig 1d), as recently reported25. Collectively, these findings highlight the plasticity of the SARS-CoV-2 RBM, which in the case of the Omicron VOC acquired enhanced binding to human and mouse ACE2 orthologues, relative to other SARS-CoV-2 isolates. The influence of these findings on viral load and replication kinetics in humans and animal models remains to be evaluated due to the interplay of additional factors besides receptor binding. Preliminary data, suggest that Omicron appears attenuated in some laboratory mouse strains (M.S.D, personal communication) and that replicates less efficiently in human lung tissue as compared to Delta26.

Extent of Omicron escape from polyclonal plasma neutralizing antibodies

To investigate the magnitude of immune evasion mediated by the 37 mutations present in Omicron S, we used Wuhan-Hu-1 S and Omicron S VSV pseudoviruses and compared plasma neutralizing activity in different cohorts of convalescent patients or individuals vaccinated with six major COVID-19 vaccines (mRNA-1273, BNT162b2, AZD1222, Ad26.COV2.S, Sputnik V and BBIBP-CorV) (Fig. 2, Supplementary Fig. 13 and Extended Data Table 1).

Fig. 2. Neutralization of Omicron SARS-CoV-2 VSV pseudovirus by plasma from COVID-19 convalescent and vaccinated individuals.

Fig. 2.

Plasma neutralizing activity in COVID-19 convalescent or vaccinated individuals (mRNA-1273, BNT162b2, AZD1222, Ad26.COV2.S (single dose), Sputnik V and BBIBP-CorV). a, Pairwise neutralizing antibody titers (ID50) against Wuhan-Hu-1 (D614G), Beta, and Omicron VOC, and SARS-CoV. Vero E6-TMPRSS2 used as target cells. Data are geometric mean of n = 3 biologically independent experiments. b, Pairwise neutralizing antibody titers of plasma (ID50) against Wuhan-Hu-1 and Omicron VOC. Data are geometric mean of n = 2 biologically independent experiments. c, Plasma neutralizing activity in dialysis patients who received 3 doses of either BNT162b2 or mRNA-1273 mRNA vaccines. Pairwise neutralizing antibody titers of plasma (ID50) against Wuhan-Hu-1 and Omicron. One representative experiment out of two is shown. Vero E6 used as target cells in b and c. Line, geometric mean of 1/ID50 titers. Shown is the percentage of samples that lost detectable neutralization against Omicron or SARS-CoV. Shown cumulative titer loss not accounting samples with 1/ID50 below the limit of detection. HCW, healthcare workers; Wu, Wuhan-Hu-1; o, Omicron VOC, b, Beta VOC. Enrolled donors’ demographics provided in Extended Data Table 1. Statistical significance is set as P<0.05 and P-values are indicated with asterisks (*=0.033; **=0.002; ***<0.001), using a paired two-sided t test (Wilcoxon rank test).

Convalescent patients and individuals vaccinated with Ad26.COV2.S (single dose), Sputnik V or BBIBP-CorV had no detectable neutralizing activity against Omicron except for one Ad26.COV2.S and three BBIBP-CorV vaccine recipients (Fig. 2a). Individuals immunized with mRNA-1273, BNT162b2, and AZD1222 had more potent neutralizing activity against Wuhan-Hu-1 and retained detectable neutralization against Omicron with a decrease of 39-, 37- and 21-fold, respectively (Fig. 2a). The dampening of neutralizing activity against Omicron was comparable to that observed against SARS-CoV, a virus that differs from Wuhan-Hu-1 by 52 residues in the RBD. Reductions of neutralization potency were less pronounced in vaccinated individuals who had been previously infected (5-fold) (Fig. 2b) and in dialysis patients (4-fold, Fig. 2c) who were boosted with a third mRNA vaccine dose. In the same cohort of dialysis patients, antibodies neutralizing the vaccine-matched Wuhan-Hu-1 strain were found to be low (less than 1/100) or undetectable in 44% of patients after the second mRNA vaccine dose27.

Collectively, these findings demonstrate a substantial and unprecedented reduction in plasma neutralizing activity against Omicron as compared to the ancestral virus, which in several cases likely falls below the protective threshold28. Our data further indicate that multiple exposures to the ancestral virus through infection or vaccination results in the production of antibodies that can neutralize divergent viruses, such as Omicron or even SARS-CoV, as a consequence of affinity maturation or epitope masking by immune-dominant RBM antibodies2830.

Broadly neutralizing sarbecovirus antibodies inhibit SARS-CoV-2 Omicron

Neutralizing mAbs with demonstrated in vivo efficacy in prevention or treatment of SARS-CoV-23137 can be divided into two groups based on whether they do or do not block S binding to ACE2. Of the eight currently authorized or approved mAbs, seven (LY-CoV555, LY-CoV016, REGN10933, REGN10933, COV2-2130, COV2-2196 and CT-P59; all synthesized based on publicly available sequences) block binding of S to ACE2 and are often used as two-mAb cocktails8. They bind to epitopes overlapping with the RBM (Fig. 3a) which is structurally and evolutionary plastic38, as illustrated by the accumulation of mutations throughout the pandemic and the genetic diversity of this subdomain among ACE2-utilizing sarbecoviruses39. Combining two such ACE2 blocking mAbs can provide greater resistance to variant viruses that carry RBM mutations31. The second class of mAbs, represented by sotrovimab, do not block ACE2 binding but neutralize SARS-CoV-2 by targeting non-RBM epitopes shared across many sarbecoviruses, including SARS-CoV4,40.

Fig. 3. Neutralization of Omicron SARS-CoV-2 VSV pseudovirus by clinical-stage mAbs.

Fig. 3.

a, RBD sequence of SARS-CoV-2 Wuhan-Hu-1 with highlighted footprints of ACE2 (light blue) and mAbs (colored according to the RBD antigenic site recognized). Omicron RBD is also shown, and amino acid substitutions are boxed. b, Neutralization of SARS-CoV-2 VSV pseudoviruses displaying Wuhan-Hu-1 (white) or Omicron (colored as in Fig. 4b) S proteins by clinical-stage mAbs. Data are representative of one independent experiment out of two. Shown is the mean of 2 technical replicates. c, Geometric mean IC50 values for Omicron (colored as in Fig. 4b) and Wuhan-Hu-1 (white) (top panel), and geometric mean fold change (bottom panel). Vero E6 used as target cells. Shown in blue (right) is neutralization of authentic virus by sotrovimab (WA1/2020 versus hCoV-19/USA/WI-WSLH-221686/2021). Non-neutralizing IC50 titers and fold change were set to 104 and 103, respectively. Orange dots for sotrovimab indicate neutralization of Omicron VSV pseudovirus carrying R346K. Data are representative of n = 2 biologically independent experiments for most mAbs, for sotrovimab against Omicron VSV n=6 and for Omicron authentic virus n=3.

We compared the in vitro neutralizing activity of these therapeutic mAbs side-by-side against Wuhan-Hu-1 S and Omicron S using VSV pseudoviruses (Fig. 3). Although sotrovimab had 3-fold reduced potency against Omicron and Omicron-R346K variant VSV pseudoviruses, all other (RBM-specific) mAbs completely lost their neutralizing activity, with the exception of the combination of COV2-2130 and COV2-2196 for which we determined a ∼100-fold reduced potency (Fig. 3bc). Moreover, sotrovimab exhibited a less than 2-fold reduction in neutralizing activity against authentic Omicron SARS-CoV-2 as compared to the WA1/2020 D614G virus (Fig. 3c and Extended Data Fig. 7), consistent with recent reports on S309, the parent of sotrovimab41,42. The 3-fold and less than 2-fold decrease in the neutralizing activity of sotrovimab against pseudoviruses and authentic virus, respectively, is within the currently defined threshold of “no change” as defined by FDA (FDA fact sheet for sotrovimab denotes no change: <5-fold reduction in susceptibility43). Overall, our findings agree with two preliminary reports16,18 and, together with serological data, support that the Omicron VOC has undergone antigenic shift.

We next tested a larger panel of 36 neutralizing NTD- or RBD-specific mAbs for which the epitopes have been characterized structurally or assigned to a given antigenic site through competition studies3,4,9,12,44,45 (Fig. 4a, Extended Data Table 2 and Extended Data Fig. 8). The four NTD-specific antibodies completely lost activity against Omicron, consistent with the presence of mutations and deletions in the NTD antigenic supersite21,46. Three out of the 22 mAbs targeting the RBD antigenic site I (RBM) retained potent neutralizing activity against Omicron, including S2K146, which binds the RBD of SARS-CoV-2, SARS-CoV and other sarbecoviruses through ACE2 molecular mimicry1. Of the nine mAbs specific for the conserved RBD site II4, only S2X2593 retained activity against Omicron, whereas neutralization was decreased by more than 10-fold or abolished for the remaining mAbs. Finally, the S2H97 mAb retained neutralizing activity against Omicron through recognition of the highly conserved cryptic site V4. The panel of 44 mAbs tested in this study includes members of each of the four classes of neutralizing mAbs, defined by their cognate RBD binding sites (site I, II, IV and V)12. Our findings show that member(s) of each of the four classes can retain Omicron neutralization: S2K146, S2X324 and S2N28 targeting site I, S2X259 targeting site II, sotrovimab targeting site IV, and S2H97 targeting site V (Fig. 4b). Several of these mAbs cross-react with and neutralize sarbecoviruses beyond the SARS-CoV-2 clade 1b1,3,4, indicating that targeting of conserved epitopes can lead to neutralization breadth and resilience to antigenic shift associated with viral evolution.

Fig. 4. Neutralization of Omicron SARS-CoV-2 VSV pseudovirus by monoclonal antibodies.

Fig. 4.

a, Mean IC50 values for Omicron (colored as in b) and Wuhan-Hu-1 (white) (top panel), and mean fold change (bottom panel) for 4 NTD mAbs and 32 RBD mAbs. Non-neutralizing IC50 titers and fold change were set to 104 and 103, respectively. Triangles for S2K146 and S2X259 indicate neutralization of Omicron carrying R346K. Vero E6 used as target cells. Data are representative of n = 2 biologically independent experiments (except for S2K146 and S2X259 where n = 6). b, The RBD sites targeted by 4 mAbs cross-neutralizing Omicron are annotated and representative antibodies (the Fv region) bound to S are shown as a composite. Colored surfaces on the RBD depict the epitopes and the RBM is shown as a black outline.

Discussion

The remarkable number of substitutions present in Omicron S marks a dramatic shift in antigenicity and is associated with immune evasion of unprecedented magnitude for SARS-CoV-2. While antigenic shift of the influenza virus is defined as genetic reassortment of the RNA genome segments, the mechanism for the abrupt appearance of a large number of mutations in SARS-CoV-2 Omicron S remains to be determined. Although recombination events are a hallmark of coronaviruses47, we and others48 propose that the Omicron shift may result from extensive viral replication in immunodeficient hosts47,49, although we cannot rule out the possibility of a contribution of inter-species ping-pong transmission5 between humans and rodents, as previously described for minks50.

Consistent with the variable decrease in plasma neutralizing antibody titers, we found that only six out of a panel of 44 neutralizing mAbs retained potent neutralizing activity against Omicron. The mAbs retaining neutralization recognize RBD antigenic sites that are conserved in Omicron and other sarbecoviruses. Notably, three of these mAbs bind to the RBM, including one which is a molecular mimic of the ACE2 receptor (S2K146)1. Collectively, these data may guide future efforts to develop SARS-CoV-2 vaccines and therapies to counteract antigenic shift and future sarbecovirus zoonotic spillovers.

MATERIALS AND METHODS

Cell lines

Cell lines used in this study were obtained from ATCC (HEK293T and Vero E6), ThermoFisher Scientific (Expi CHO cells, FreeStyle™ 293-F cells and Expi293F™ cells), Lenti-X 293T cells (Takara) or generated in-house (Vero E6/TMPRSS2)40. Vero-TMPRSS251 cells were cultured at 37°C in Dulbecco’s Modified Eagle medium (DMEM) supplemented with 10% fetal bovine serum (FBS), 10 mM HEPES pH 7.3, and 100 U/ml of penicillin–streptomycin and supplemented with 5 µg/mL of blasticidin. None of the cell lines used was authenticated. Cell lines were routinely tested for mycoplasma contamination.

Omicron prevalence analysis

The viral sequences and the corresponding metadata were obtained from GISAID EpiCoV project (https://www.gisaid.org/). Analysis was performed on sequences submitted to GISAID up to Dec 09, 2021. S protein sequences were either obtained directly from the protein dump provided by GISAID or, for the latest submitted sequences that were not incorporated yet in the protein dump at the day of data retrieval, from the genomic sequences with the exonerate52 2 2.4.0–haf93ef1_3 (https://quay.io/repository/biocontainers/exonerate?tab=tags ) using protein to DNA alignment with parameters -m protein2dna –refine full –minintron 999999 –percent 20 and using accession YP_009724390.1 as a reference. Multiple sequence alignment of all human spike proteins was performed with mafft53 7.475–h516909a_0 (https://quay.io/repository/biocontainers/mafft?tab=tags) with parameters –auto –reorder – keeplength –addfragments using the same reference as above. S sequences that contained >10% ambiguous amino acid or that were < than 80% of the canonical protein length were discarded. Figures were generated with R 4.0.2 (https://cran.r-project.org/) using ggplot2 3.3.2 and sf 0.9–7 packages. To identify each mutation prevalence, missingness (or ambiguous amino acids) was taken into account in both nominator and denominator.

Monoclonal Antibodies

Sotrovimab and VIR-7832 (VIR-783254 is derived from sotrovimab, Fc further engineered to carry GAALIE) were produced at WuXi Biologics (China). Antibody VH and VL sequences for mAbs COV2-2130 (PDB ID 7L7E), COV2-2196 (PDB ID 7L7E, 7L7D), REGN10933 (PDB ID 6XDG), REGN10987 (PDB ID 6XDG) and ADI-58125 (PCT application WO2021207597, seq. IDs 22301 and 22311) were subcloned into heavy chain (human IgG1) and the corresponding light chain (human IgKappa, IgLambda) expression vectors respectively and produced in transiently transfected Expi-CHO-S cells (Thermo Fisher, #A29133) at 37°C and 8% CO2. Cells were transfected using ExpiFectamine. Transfected cells were supplemented 1 day after transfection with ExpiCHO Feed and ExpiFectamine CHO Enhancer. Cell culture supernatant was collected eight days after transfection and filtered through a 0.2 µm filter. Recombinant antibodies were affinity purified on an ÄKTA Xpress FPLC device using 5 mL HiTrap™ MabSelect™ PrismA columns followed by buffer exchange to Histidine buffer (20 mM Histidine, 8% sucrose, pH 6) using HiPrep 26/10 desalting columns. Antibody VH and VL sequences for LY-CoV555, LY-CoV016, and CT-P59 were obtained from PDB IDs 7KMG, 7C01 and 7CM4, respectively and mAbs were produced as recombinant IgG1 by ATUM. The remaining mAbs were discovered at VIR and have been produced as recombinant IgG1 in Expi-CHO-S cells as described above. The identity of the produced mAbs was confirmed by LC-MS analysis.

IgG mass quantification by LC/MS intact protein mass analysis

Fc N-linked glycan from mAbs were removed by PNGase F after overnight non-denaturing reaction at room temperature. Deglycosylated protein (4 µg) was injected to the LC-MS system to acquire intact MS signal. Thermo MS (Q Exactive Plus Orbitrap) was used to acquire intact protein mass under denaturing condition with m/z window from 1,000 to 6,000. BioPharma Finder 3.2 software was used to deconvolute the raw m/z data to protein average mass. The theoretical mass for each mAb was calculated with GPMAW 10.10 software. Post-translational modifications such as N-terminal pyroglutamate cyclization, C-terminal lysine cleavage, and formation of 16–18 disulfide bonds were added into the calculation.

Sample donors

Samples were obtained from SARS-CoV-2 recovered and vaccinated individuals under study protocols approved by the local Institutional Review Boards (Canton Ticino Ethics Committee, Switzerland, Comitato Etico Milano Area 1). All donors provided written informed consent for the use of blood and blood derivatives (such as PBMCs, sera or plasma) for research. Samples were collected 14–28 days after symptoms onset and 14–28 days or 7–10 months after vaccination. Convalescent plasma, Ad26.COV2.S, mRNA-1273 and BNT162b2 samples were obtained from the HAARVI study approved by the University of Washington Human Subjects Division Institutional Review Board (STUDY00000959). AZD1222 samples were obtained from INGM, Ospedale Maggio Policlinico of Milan and approved by the local review board Study Polimmune. Sputnik V samples were obtained from healthcare workers at the hospital de Clínicas “José de San Martín”, Buenos Aires, Argentina. Sinopharm vaccinated individuals were enrolled from Aga Khan University under IRB of UWARN study.

Serum/plasma and mAbs pseudovirus neutralization assays

VSV pseudovirus generation used on Vero E6 cells

The plasmid encoding the Omicron SARS-CoV-2 S variant was generated by overlap PCR mutagenesis of the wild-type plasmid, pcDNA3.1(+)-spike-D1955. Replication defective VSV pseudovirus expressing SARS-CoV-2 spike proteins corresponding to the ancestral Wuhan-Hu-1 virus and the Omicron VOC were generated as previously described46 with some modifications. Lenti-X 293T cells (Takara) were seeded in 15-cm2 dishes at a density of 10e6 cells per dish and the following day transfected with 25 µg of spike expression plasmid with TransIT-Lenti (Mirus, 6600) according to the manufacturer’s instructions. One day post-transfection, cells were infected with VSV-luc (VSV-G) with an MOI 3 for 1 h, rinsed three times with PBS containing Ca2+/Mg2+, then incubated for additional 24 h in complete media at 37°C. The cell supernatant was clarified by centrifugation, aliquoted, and frozen at −80°C.

VSV pseudovirus generation used on Vero E6-TMPRSS2 cells

Comparison of Omicron SARS-CoV-2 S VSV to SARS-CoV-2 G614 S (YP 009724390.1) VSV and Beta S VSV used pseudotyped particles prepared as described previously9,56. Briefly, HEK293T cells in DMEM supplemented with 10% FBS, 1% PenStrep seeded in 10-cm dishes were transfected with the plasmid encoding for the corresponding S glycoprotein using lipofectamine 2000 (Life Technologies) following the manufacturer’s instructions. One day post-transfection, cells were infected with VSV(G*ΔG-luciferase)57 and after 2 h were washed five times with DMEM before adding medium supplemented with anti-VSV-G antibody (I1- mouse hybridoma supernatant, CRL- 2700, ATCC). Virus pseudotypes were harvested 18–24 h post-inoculation, clarified by centrifugation at 2,500 x g for 5 min, filtered through a 0.45 μm cut off membrane, concentrated 10 times with a 30 kDa cut off membrane, aliquoted and stored at −80°C.

VSV pseudovirus neutralization

Assay performed using Vero E6 cells

Vero-E6 were grown in DMEM supplemented with 10% FBS and seeded into clear bottom white 96 well plates (PerkinElmer, 6005688) at a density of 20,000 cells per well. The next day, mAbs or plasma were serially diluted in pre-warmed complete media, mixed with pseudoviruses and incubated for 1 h at 37°C in round bottom polypropylene plates. Media from cells was aspirated and 50 µl of virus-mAb/plasma complexes were added to cells and then incubated for 1 h at 37°C. An additional 100 µL of prewarmed complete media was then added on top of complexes and cells incubated for an additional 16–24 h. Conditions were tested in duplicate wells on each plate and eight wells per plate contained untreated infected cells (defining the 0% of neutralization, “MAX RLU” value) and infected cells in the presence of S309 and S2X259 at 20 µg/ml each (defining the 100% of neutralization, “MIN RLU” value). Virus-mAb/plasma-containing media was then aspirated from cells and 100 µL of a 1:2 dilution of SteadyLite Plus (Perkin Elmer, 6066759) in PBS with Ca++ and Mg++ was added to cells. Plates were incubated for 15 min at room temperature and then were analyzed on the Synergy-H1 (Biotek). Average of Relative light units (RLUs) of untreated infected wells (MAX RLUave) was subtracted by the average of MIN RLU (MIN RLUave) and used to normalize percentage of neutralization of individual RLU values of experimental data according to the following formula: (1-(RLUx – MIN RLUave) / (MAX RLUave – MIN RLUave)) x 100. Data were analyzed and visualized with Prism (Version 9.1.0). IC50 (mAbs) and ID50 (plasma) values were calculated from the interpolated value from the log(inhibitor) versus response, using variable slope (four parameters) nonlinear regression with an upper constraint of ≤100, and a lower constrain equal to 0. Each neutralization experiment was conducted on two independent experiments, i.e., biological replicates, where each biological replicate contains a technical duplicate. IC50 values across biological replicates are presented as arithmetic mean ± standard deviation. The loss or gain of neutralization potency across spike variants was calculated by dividing the variant IC50/ID50 by the parental IC50/ID50 within each biological replicate, and then visualized as arithmetic mean ± standard deviation.

Assay performed using Vero E6-TMPRSS2 cells

VeroE6-TMPRSS2 were cultured in DMEM with 10% FBS (Hyclone), 1% PenStrep and 8 µg/mL puromycin (to ensure retention of TMPRSS2) with 5% CO2 in a 37°C incubator (ThermoFisher). Cells were trypsinized using 0.05% trypsin and plated to be at 90% confluence the following day. In an empty half-area 96-well plate, a 1:3 serial dilution of sera was made in DMEM and diluted pseudovirus was then added and incubated at room temperature for 30–60 min before addition of the sera-virus mixture to the cells at 37°C. 2 hours later, 40 μL of a DMEM solution containing 20% FBS and 2% PenStrep (ThermoFisher, 10,000 units/mL of penicillin and 10,000 µg/mL of streptomycin when undiluted) was added to each well. After 17–20 hours, 40 μL/well of One-Glo-EX substrate (Promega) was added to the cells and incubated in the dark for 5–10 min prior to reading on a BioTek plate reader. Measurements were done at least in duplicate using distinct batches of pseudoviruses and one representative experiment is shown. Relative luciferase units were plotted and normalized in Prism (GraphPad). Nonlinear regression of log(inhibitor) versus normalized response was used to determine IC50 values from curve fits. Normality was tested using the D’Agostino-Pearson test and in the absence of a normal distribution, Kruskal-Wallis tests were used to compare two groups to determine whether differences reached statistical significance. Fold changes were determined by comparing individual IC50 and then averaging the individual fold changes for reporting.

Focus reduction neutralization test

The WA1/2020 strain with a D614G substitution was described previously58. The B.1.1.529 isolate (hCoV-19/USA/WI-WSLH-221686/2021) was obtained from a nasal swab and passaged on Vero-TMPRSS2 cells as described59. The B.1.1.529 isolate was sequenced (GISAID: EPI_ISL_7263803) to confirm the stability of substitutions. All virus experiments were performed in an approved biosafety level 3 (BSL-3) facility.

Serial dilutions of sotrovimab were incubated with 102 focus-forming units (FFU) of SARS-CoV-2 (WA1/2020 D614G or B.1.1.529) for 1 h at 37°C. Antibody-virus complexes were added to Vero-TMPRSS2 cell monolayers in 96-well plates and incubated at 37°C for 1 h. Subsequently, cells were overlaid with 1% (w/v) methylcellulose in MEM. Plates were harvested at 30 h (WA1/2020 D614G on Vero-TMPRSS2 cells) or 70 h (B.1.1.529 on Vero-TMPRSS2 cells) later by removal of overlays and fixation with 4% PFA in PBS for 20 min at room temperature. Plates with WA1/2020 D614G were washed and sequentially incubated with an oligoclonal pool of SARS2-2, SARS2-11, SARS2-16, SARS2-31, SARS2-38, SARS2-57, and SARS2-7160 anti-S antibodies. Plates with B.1.1.529 were additionally incubated with a pool of mAbs that cross-react with SARS-CoV-1 and bind a CR3022-competing epitope on the RBD61. All plates were subsequently stained with HRP-conjugated goat anti-mouse IgG (Sigma, A8924) in PBS supplemented with 0.1% saponin and 0.1% bovine serum albumin. SARS-CoV-2-infected cell foci were visualized using TrueBlue peroxidase substrate (KPL) and quantitated on an ImmunoSpot microanalyzer (Cellular Technologies). Antibody-dose response curves were analyzed using non-linear regression analysis with a variable slope (GraphPad Software), and the half-maximal inhibitory concentration (IC50) was calculated.

VSV pseudovirus entry assays using mouse ACE2

HEK293T (293T) cells (ATCC CRL-11268) were cultured in 10% FBS, 1% PenStrep DMEM at 37°C in a humidified 8% CO2 incubator. Transient transfection of mouse ACE2 in 293T cells was done 18–24 hours prior to infection using Lipofectamine 2000 (Life Technologies) and an HDM plasmid containing full length Mouse ACE2 (GenBank: Q8R010, synthesized by GenScript) in OPTIMEM. After 5 hr incubation at 37°C in a humidified 8% CO2 incubator, DMEM with 10% FBS was added and cells were incubated at 37°C in a humidified 8% CO2 incubator for 18–24 hr. Immediately prior to infection, 293T cells with transient expression of mouse ACE2 were washed with DMEM 1x, then plated with pseudovirus at a 1:75 dilution in DMEM. Infection in DMEM was done with cells between 60–80% confluence for 2.5 hr prior to adding FBS and PenStrep to final concentrations of 10% and 1%, respectively. Following 18–24 hr of infection, One-Glo-EX (Promega) was added to the cells and incubated in the dark for 5 min before reading on a Synergy H1 Hybrid Multi-Mode plate reader (Biotek). Cell entry levels of pseudovirus generated on different days (biological replicates) were plotted in GraphPad Prism as individual points, and average cell entry across biological replicates was calculated as the geometric mean.

Recombinant RBD protein production

SARS-CoV-2 RBD proteins for SPR binding assays (residues 328–531 of S protein from GenBank NC_045512.2 with N-terminal signal peptide and C-terminal thrombin cleavage site-TwinStrep-8xHis-tag) were expressed in Expi293F (Thermo Fisher Scientific) cells at 37°C and 8% CO2. Transfections were performed using the ExpiFectamine 293 Transfection Kit (Thermo Fisher Scientific). Cell culture supernatants were collected two to four days after transfection and supplemented with 10x PBS to a final concentration of 2.5x PBS (342.5 mM NaCl, 6.75 mM KCl and 29.75 mM phosphates). SARS-CoV-2 RBDs were purified using cobalt-based immobilized metal affinity chromatography followed by buffer exchange into PBS using a HiPrep 26/10 desalting column (Cytiva) or, for the 2nd batch of Omicron RBD used for SPR, a Superdex 200 Increase 10/300 GL column (Cytiva).

The SARS-CoV-2 Wuhan-Hu-1 and Delta (B.1.617.2) RBD-Avi constructs were synthesized by GenScript into pcDNA3.1- with an N-terminal mu-phosphatase signal peptide and a C-terminal octa-histidine tag, flexible linker, and avi tag (GHHHHHHHHGGSSGLNDIFEAQKIEWHE). The boundaries of the construct are N-328RFPN331 and 528KKST531-C9,14. Proteins were produced in Expi293F cells (ThermoFisher Scientific) grown in suspension using Expi293 Expression Medium (ThermoFisher Scientific) at 37°C in a humidified 8% CO2 incubator rotating at 130 rpm. Cells grown to a density of 3 million cells per mL were transfected using the ExpiFectamine 293 Transfection Kit (ThermoFisher Scientific) and cultivated for 3–5 days. Proteins were purified from clarified supernatants using a nickel HisTrap HP affinity column (Cytiva) and washed with ten column volumes of 20 mM imidazole, 25 mM sodium phosphate pH 8.0, and 300 mM NaCl before elution on a gradient to 500 mM imidazole. Proteins were biotinylated overnight using the BirA Biotin-Protein Ligase Kit (Avidity) and purified again using the HisTrapHP affinity column. After a wash and elution as before, proteins were buffer exchanged into 20 mM sodium phosphate pH 8 and 100 mM NaCl, and concentrated using centrifugal filters (Amicon Ultra) before being flash frozen.

Recombinant production of ACE2 orthologs

Recombinant human ACE2 (residues 19–615 from Uniprot Q9BYF1 with a C-terminal AviTag-10xHis-GGG-tag, and N-terminal signal peptide) was produced by ATUM. Protein was purified via Ni Sepharose resin followed by isolation of the monomeric hACE2 by size exclusion chromatography using a Superdex 200 Increase 10/300 GL column (Cytiva) pre-equilibrated with PBS. The mouse (Mus musculus) ACE2 ectodomain construct (GenBank: Q8R0I0) was synthesized by GenScript and placed into a pCMV plasmid. The domain boundaries for the ectodomain are residues 19–615. The native signal tag was identified using SignalP-5.0 (residues 1–18) and replaced with a N-terminal mu-phosphatase signal peptide. This construct was then fused to a sequence encoding thrombin cleavage site and a human Fc fragment or a 8x His tag at the C-terminus. ACE2-Fc and ACE2 His constructs were produced in Expi293 cells (Thermo Fisher A14527) in Gibco Expi293 Expression Medium at 37°C in a humidified 8% CO2 incubator rotating at 130 rpm. The cultures were transfected using PEI-25K (Polyscience) with cells grown to a density of 3 million cells per mL and cultivated for 4–5 days. Proteins were purified from clarified supernatants using a 1 mL HiTrap Protein A HP affinity column (Cytiva) or a 1 mL HisTrap HP affinity column (Cytiva), concentrated and flash frozen in 1x PBS, pH 7.4 (10 mM Na2HPO4, 1.8 mM KH2PO4, 2.7 mM KCl, 137 mM NaCl).

ACE2 binding measurements using surface plasmon resonance

Measurements were performed using a Biacore T200 instrument, in triplicate for monomeric human and mouse ACE2 and duplicate for dimeric mouse ACE2. A CM5 chip covalently immobilized with StrepTactin XT (IBA LifeSciences) was used for surface capture of TwinStrepTag-containing RBDs (Wuhan-Hu-1, Alpha, Beta, Omicron, K417N) and a Cytiva Biotin CAPture Kit was used for surface capture of biotinylated RBDs (Delta and Wuhan-Hu-1 used for fold-change comparison to Delta). Two different batches of Omicron RBD were used for the experiments. Running buffer was HBS-EP+ pH 7.4 (Cytiva) and measurements were performed at 25 °C. Experiments were performed with a 3-fold dilution series of human ACE2 (300, 100, 33, 11 nM) or mouse ACE2 (900, 300, 100, 33 nM) and were run as single-cycle kinetics. Monomeric ACE2 binding data were double reference-subtracted and fit to a 1:1 binding model using Biacore Evaluation software. High concentrations of dimeric mouse ACE2 exhibited significant binding to the CAP sensor chip reference flow cell.

Statistical analysis

Neutralization measurements were performed in duplicate and relative luciferase units were converted to percent neutralization and plotted with a non-linear regression model to determine IC50/ID50 values using GraphPad PRISM software (version 9.0.0). Comparisons between two groups of paired two-sided data were made with Wilcoxon rank test.

Data availability

All datasets generated and information presented in the study are available from the corresponding authors on reasonable request.

Extended Data

Extended Data Fig. 1. Schematic of mutations landscape in SARS-CoV-2 VOC, VOI and VUM (Variant Under Monitoring).

Extended Data Fig. 1.

D, deletion: ins, insertion.

Extended Data Fig. 2. Amino acid substitutions and their prevalence in the Omicron RBD.

Extended Data Fig. 2.

a, SARS-CoV-2 S in fully open conformation (PDB: 7K4N) with positions of mutated residues in Omicron highlighted on one protomer in green or red spheres in or outside the ACE2 footprint (ACE2), respectively. RBM is defined by a 6 Å cutoff in the RBD-ACE2 interface38. Not all Omicron mutations are shown. b, Substitutions and their prevalence in Omicron sequences reported in GISAID as of December 20, 2021 (ambiguous amino acid substitutions are indicated with strikethrough cells). Shown are also the substitutions found in other variants. K417N mutation in Delta is found only in a fraction of sequences.

Extended Data Fig. 3. Amino acid substitutions and their prevalence in the Omicron NTD.

Extended Data Fig. 3.

Sequences reported in GIAID as of December 20, 2021; (ambiguous amino acid substitutions are marked with strikethrough cells). Shown are also the substitutions found in other variants.

Extended Data Fig. 4. Amino acid substitutions and their prevalence in the Omicron S2.

Extended Data Fig. 4.

Sequences reported in GIAID as of December 20, 2021; (ambiguous amino acid substitutions are marked with strikethrough cells). Shown are also the substitutions found in other variants.

Extended Data Fig. 5. Characteristics of emergent mutations of Omicron.

Extended Data Fig. 5.

a, Shared mutations of micron with other sarbecovirus and with VOC. b, Since the beginning of the pandemic there is a progressive coalescence of Omicron-defining mutations into non-Omicron haplotypes that may carry as many as 10 of the Omicron-defining mutations. c, Pango lineages (dots) rarely carry more than 10–15 lineage-defining mutations. d, Exceptionally, some non-Omicron haplotypes may carry up to a maximum 19 Omicron-defining mutations. Shown are selected exceptional haplotypes. Spike G142D and Y145del may also be noted as G142del and Y145D.

Extended Data Fig. 6. SPR analysis of human and mouse ACE2.

Extended Data Fig. 6.

a, Full fit results for one representative replicate from each quantifiable SPR dataset with a monomeric analyte (1:1 binding model). b, Single-cycle kinetics SPR analysis of dimeric mouse ACE2 binding to six RBD variants. Dimeric ACE2 is injected successively at 33, 100, 300, and 900 nM. White and gray stripes indicate association and dissociation phases, respectively. The asterisk indicates where high concentrations of dimeric mouse ACE2 is non-specifically binding to the sensor chip surface (Delta experiment was performed separately from the other RBD variants, with a different capture tag and chip surface).

Extended Data Fig. 7. Neutralization of SARS-CoV-2 Omicron strain by sotrovimab in Vero-TMPRSS2 cells.

Extended Data Fig. 7.

a-f, Neutralization curves in Vero-TMPRSS2 cells comparing the sensitivity of SARS-CoV-2 strains with sotrovimab with WA1/2020 D614G and hCoV-19/USA/WI-WSLH-221686/2021 (an infectious clinical isolate of Omicron from a symptomatic individual in the United States). Shown are three independent experiments performed in technical duplicate is shown.

Extended Data Fig. 8. Neutralization of WT (D614) and Omicron SARS-CoV-2 Spike pseudotyped virus by a panel of 36 mAbs.

Extended Data Fig. 8.

a-c, Neutralization of SARS-CoV-2 VSV pseudoviruses carrying wild-type D614 (grey) or Omicron (orange) S protein by NTD-targeting (a) and RBD-targeting (b-c) mAbs (b, site I; c, sites II and V). Data are representative of one independent experiment out of two. Shown is the mean. of 2 technical replicates.

Extended Data Table 1.

Enrolled donors’ demographics.

2–4 weeks after infection/2nd vaccine dose Dating of SARS-CoV-2 infection Figure Nr. Females Males Age (average, range)

Wild type SARS-CoV-2-infected convalescent 29 10 19 56, 34–73
 Ospedale Luigi Sacco Mar-Apr 2020 2b 11 1 10 56, 34–73
 Swiss volunteers Mar 2020 2b 1 1 52, 52–52
 HAARVI (University of Washington) Mar-Apr 2020 2a 17 9 8 51, 25–78

Previously infected BNT162b2-vaccinated 29 19 10 39, 26–56
 Clinica Luganese Moncucco Mar-Nov 2020 2b 4 3 1 38, 27–54
 Ente Ospedaliero Cantonale (EOC) Mar 2020-Jan 2021 2b 18 14 4 39, 26–56
 EOC, dialysis pts Mar 2020-Jan 2021 2c 7 2 5 69, 48–87

Naïve BNT162b2-vaccinated 99 49 50 43, 24–67
 Clinica Luganese Moncucco 2b 7 4 3 42, 28–50
 Ente Ospedaliero Cantonale (EOC) 2b 18 13 5 43, 24–67
 EOC, dialysis pts 2c 55 22 33 74, 29–97
 HAARVI (University of Washington) 2a 17 10 7 45, 22–76

Naïve mRNA-1273-vaccinated 40 25 15
 Innovative Research, Novi Michigan
(1 week after 2nd dose)
2b 20 14 6 58, 34–74
 EOC, dialysis pts 2c 8 2 6 85, 81–92
 HAARVI (University of Washington) 2a 14 9 5 47, 23–79

Naïve ChAdOx1-vaccinated
 INGM, Ospedale Maggio Policlinico 2a 17 13 4 38, 29–51

Naïve Sputnik V-vaccinated
 Hospital de Clínicas José de San
San Martin, Buenos Aires
2a 11 7 4 42, 30–58

Naïve BBIBP-CorV-vaccinated
 Aga Khan University 2a 13 9 4 30, 25–39

1–19 weeks after 1st vaccine dose Nr. Females Males Age (average, range)

Naïve Ad26.COV2.S-vaccinated
 HAARVI (University of Washington) 2a 12 6 6 33, 23–60

Total 250 138 112

Extended Data Table 2.

Properties of tested mAbs.

mAb Domain (site) VH usage Source days after symptom onset IC50 Wu-Hu-1 (ng/ml) IC50 Omicron (ng/ml) PDB/EMD Ref.
sotrovimab RBD (IV) 3–23 SARS-CoV immune 90.6 260 6WPS, 7JX3 24,9,40,62,63
209 (R346K)
179
(WA1/2020)
320 (hCoV-19/USA/WI-WSLH-221686/2021)
VIR-7832* RBD (IV) 3–23 SARS-CoV immune 53.2 165 6WPS, 7JX3 24,9,40,62,63
CT-P59 RBD (I/RBM) N/A SARS-CoV-2 immune 4.3 >10’000 7CM4 64,65
COV2-2130 RBD (I/RBM) 3–15 SARS-CoV-2 immune 8.1 2772 7L7E 36,37
COV2-2196 RBD (I/RBM) 1–58 SARS-CoV-2 immune 4.3 >10’000 7L7E, 7L7D 36,37
2130+2196 3.8 418
REGN10933 RBD (I/RBM) 3–11 SARS-CoV-2 huIg mice 8.9 >10’000 6XDG 31,32,6668
REGN10987 RBD (I/RBM) 3–30 SARS-CoV-2 immune 25.1 >10’000 6XDG 31,32,6668
103933+10987 7.2 >10’000
LY-CoV555 RBD (I/RBM) 1–69 SARS-CoV-2 immune 21.3 >10’000 7KMG 34,35,69,70
LY-CoV016 RBD (I/RBM) 3–66 SARS-CoV-2 immune 59.2 >10’000 7C01 33
555+016 23 >10’000
S2D106 RBD (I/RBM) 1–69 Hosp. (98) 9.1 >10’000 7R7N 4,38
S2D8 RBD (I/RBM) 3–23 Hosp. (49) 7.3 >10’000 38
S2D97 RBD (I/RBM) 2–5 Hosp. (98) 5.3 >10’000 38
S2E12 RBD (I/RBM) 1–58 Hosp. (51) 3.7 896 7K4N, 7R6X 4,38,40,44
S2H14 RBD (I/RBM) 3–15 Sympt. (17) 625 >10’000 7JX3 4,12,38
S2H19 RBD (I/RBM) 3–15 Sympt. (45) 361 >10’000 38
S2H58 RBD (I/RBM) 1–2 Sympt. (45) 5.4 >10’000 4,38
S2H7 RBD (I/RBM) 3–66 Sympt. (17) 607 >10’000 38
S2H70 RBD (I/RBM) 1–2 Sympt. (45) 145 >10’000 38
S2H71 RBD (I/RBM) 2–5 Sympt. (45) 10.6 993 38
S2M11 RBD (I/RBM) 1–2 Hosp. (46) 1.0 >10’000 7K43 9,38,44
S2N12 RBD (I/RBM) 4–39 Hosp. (51) 11.8 10.8 38
S2N22 RBD (I/RBM) 3–23 Hosp. (51) 8.4 919 38
S2N28 RBD (I/RBM) 3–30 Hosp. (51) 5.8 17.1 38
S2X128 RBD (I/RBM) 1–69-2 Sympt. (75) 23.2 >10’000 38
S2X16 RBD (I/RBM) 1–69 Sympt. (48) 6.2 >10’000 4,38
S2X192 RBD (I/RBM) 1–69 Sympt. (75) 223 >10’000 38
S2X30 RBD (I/RBM) 1–69 Sympt. (48) 7.2 1750 38
S2X324 RBD (I/RBM) 2–5 Sympt. (125) 2.6 3.0 21
S2X58 RBD (I/RBM) 1–46 Sympt. (48) 11.1 >10’000 EMD-24607 4,38
S2K146 RBD (I/RBM) 3–43 Sympt. (35) 14.2 12.6 pending 1
S2H13 RBD (I/RBM) 3–7 Sympt. (17) 628 >10’000 7JV4 4,12
ADI-58125 RBD (II) 3–23 SARS-CoV immune 9.3 1703 71
S2H90 RBD (II) 4–61 Sympt. (81) 37 >10’000 38
S2K63v2 RBD (II) 3–30 Sympt. (118) 129 >10’000 21
S2L37 RBD (II) 3–13 Hosp. (51) 1496 >10’000 21
S2X259 RBD (II) 1–69 Sympt. (75) 81.8 193.6 7RA8, 7M7W 3
S2X35 RBD (II) 1–18 Sympt. (48) 58.6 7999 7R6W 4,12
S2X219 RBD (II) 3–53 Sympt. (75) 9.8 268.3
S304 RBD (II) 3–13 SARS-CoV immune 4603 >10’000 7JX3 4,12
S2A4 RBD (II) 3–7 Hosp. (24) 2285 >10’000 7JVC 12
S2H97 RBD (V) 5–51 Sympt. (81) 280 1368 7M7W 4
S2L50 NTD (i) 4–59 Hosp. (52) 338 >10’000 45
S2X28 NTD (i) 3–30 Sympt. (48) 423 >10’000 EMD-23584 45
S2X303 NTD (i) 2–5 Sympt. (125) 4.5 >10’000 7SOF, 7SOE 9,45
S2X333 NTD (i) 3–33 Sympt. (125) 13 >10’000 7LXW, 7LXY 9,40,45

Extended Data Table.

Binding properties and V gene usage of tested mAbs..

mAb Domain (site)* VH usage Source (DSO) IC50 WT (ng/ml) IC50 Omicron (ng/ml) Ref.
sotrovimab RBD (IV) 3–23 SARS-CoV immune donor 90.6 260.1 17
VIR-7832 RBD (IV) 3–23 SARS-CoV immune donor 53.2 164.6 17
regdanvimab RBD (I/RBM) N/A SARS- CoV-2 immune donor 4.3 und. 8,9
cilgavimab RBD (I/RBM) 3–15 SARS- CoV-2 immune donor 8.1 2772 10,11
tixagevimab RBD (I/RBM) 1–58 SARS- CoV-2 immune donor 4.3 und. 10,11
casirivimab RBD (I/RBM) 3–11 SARS-CoV-2- immunized huIg mice 8.9 und. 1216
imdevimab RBD (I/RBM) 3–30 SARS- CoV-2 immune donor 25.1 und. 1216
bamlanivimab RBD (I/RBM) 1–69 SARS-CoV-2 immune donor 21.3 und. 1720
etesevimab RBD (I/RBM) 3–66 SARS-CoV-2 immune donor 59.2 und. 21
S2D106 RBD (I/RBM) 1–69 Hosp. (98) 9.1 und. 5,22
S2D8 RBD (I/RBM) 3–23 Hosp. (49) 7.3 und. 22
S2D97 RBD (I/RBM) 2–5 Hosp. (98) 5.3 und. 22
S2E12 RBD (I/RBM) 1–58 Hosp. (51) 3.7 896 2,5,22,23
S2H14 RBD (I/RBM) 3–15 Sympt. (17) 624.8 und. 5,22,24
S2H19 RBD (I/RBM) 3–15 Sympt. (45) 361.1 und. 22
S2H58 RBD (I/RBM) 1–2 Sympt. (45) 5.4 und. 5,22
S2H7 RBD (I/RBM) 3–66 Sympt. (17) 607 und. 22
S2H70 RBD (I/RBM) 1–2 Sympt. (45) 145 und. 22
S2H71 RBD (I/RBM) 2–5 Sympt. (45) 10.6 993 22
S2M11 RBD (I/RBM) 1–2 Hosp. (46) 1.0 und. 7,22,23
S2N12 RBD (I/RBM) 4–39 Hosp. (51) 11.8 10.8 22
S2N22 RBD (I/RBM) 3–23 Hosp. (51) 8.4 919 22
S2N28 RBD (I/RBM) 3–30 Hosp. (51) 5.8 17.1 22
S2X128 RBD (I/RBM) 1–69-2 Sympt. (75) 23.2 und. 22
S2X16 RBD (I/RBM) 1–69 Sympt. (48) 6.2 und. 5,22
S2X192 RBD (I/RBM) 1–69 Sympt. (75) 223.3 und. 22
S2X30 RBD (I/RBM) 1–69 Sympt. (48) 7.2 1750 22
S2X324 RBD (I/RBM) 2–5 Sympt. (125) 2.6 3.0 25
S2X58 RBD (I/RBM) 1–46 Sympt. (48) 11.1 und. 5,22
S2K146 RBD (I/RBM) 3–43 Sympt. (35) 14.2 12.6 26
S2H13 RBD (I/RBM) 3–7 Sympt. (17) 628.3 und. 5,24
ADI-58125 RBD (I/II) 3–23 SARS-CoV immune donor 9.3 1703 Patent WO2021207597
S2H90 RBD (II) 4–61 Sympt. (81) 37.3 und. 22
S2K63v2 RBD (II) 3–30 Sympt. (118) 129.1 und. 25
S2L37 RBD (II) 3–13 Hosp. (51) 1496 und. 25
S2X259 RBD (II) 1–69 Sympt. (75) 81.8 193.6 4
S2X35 RBD (II) 1–18 Sympt. (48) 58.6 7999 5,24
S2X219 RBD (II) 3–53 Sympt. (75) 9.8 268.3 -
S304 RBD (II) 3–13 SARS-CoV immune donor 4603 und. 5,24
S2A4 RBD (II) 3–7 Hosp. (24) 2285 und. 24
S2H97 RBD (V) 5–51 Sympt. (81) 279.7 1368 5
S2L50 NTD (i) 4–59 Hosp. (52) 337.9 und. 27
S2X28 NTD (i) 3–30 Sympt. (48) 422.7 und. 27
S2X303 NTD (i) 2–5 Sympt. (125) 4.5 und. 7,27
S2X333 NTD (i) 3–33 Sympt. (125) 13 und. 2,7,27

DSO, days after symptom onset. N/A, not available.

Supplementary Material

Supp

Acknowledgements

We thank Hideki Tani (University of Toyama) for providing the reagents necessary for preparing VSV pseudotyped viruses. This study was supported by the National Institute of Allergy and Infectious Diseases (DP1AI158186 and HHSN272201700059C to D.V.), a Pew Biomedical Scholars Award (D.V.), an Investigators in the Pathogenesis of Infectious Disease Awards from the Burroughs Wellcome Fund (D.V.), Fast Grants (D.V.), the National Institute of General Medical Sciences (5T32GM008268-32 to SKZ). D.V. is an Investigator of the Howard Hughes Medical Institute. OG is funded by the Swiss Kidney Foundation. This work was supported, in part, by the National Institutes of Allergy and Infectious Diseases Center for Research on Influenza Pathogenesis (HHSN272201400008C), Center for Research on Influenza Pathogenesis and Transmission (CRIPT) (75N93021C00014), and the Japan Program for Infectious Diseases Research and Infrastructure (JP21wm0125002) from the Japan Agency for Medical Research and Development (AMED).

Footnotes

Competing interests

E.C., K.C., C.S., D.P., F.Z., A.D.M., A.L., L.P., M.S.P., D.C., H.K., J.N., N.F., J.diI., L.E.R., N.C., C.H.D., K.R.S., J.R.D., A.E.P., A.C., C.M., L.Y., D.S., L.S., L.A.P. , C.H., A.T., H.W.V. and G.S. are employees of Vir Biotechnology Inc. and may hold shares in Vir Biotechnology Inc. L.A.P. is a former employee and shareholder in Regeneron Pharmaceuticals. Regeneron provided no funding for this work. H.W.V. is a founder and hold shares in PierianDx and Casma Therapeutics. Neither company provided resources. The Veesler laboratory has received a sponsored research agreement from Vir Biotechnology Inc. HYC reported consulting with Ellume, Pfizer, The Bill and Melinda Gates Foundation, Glaxo Smith Kline, and Merck. She has received research funding from Emergent Ventures, Gates Ventures, Sanofi Pasteur, The Bill and Melinda Gates Foundation, and support and reagents from Ellume and Cepheid outside of the submitted work. M.S.D. is a consultant for Inbios, Vir Biotechnology, Senda Biosciences, and Carnival Corporation, and on the Scientific Advisory Boards of Moderna and Immunome. The Diamond laboratory has received funding support in sponsored research agreements from Moderna, Vir Biotechnology, and Emergent BioSolutions. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

REFERENCES

  • 1.Park YJ et al. Antibody-mediated broad sarbecovirus neutralization through ACE2 molecular mimicry. bioRxiv, doi: 10.1101/2021.10.13.464254 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Pinto D et al. Cross-neutralization of SARS-CoV-2 by a human monoclonal SARS-CoV antibody. Nature 583, 290–295, doi: 10.1038/s41586-020-2349-y (2020). [DOI] [PubMed] [Google Scholar]
  • 3.Tortorici MA et al. Broad sarbecovirus neutralization by a human monoclonal antibody. Nature 597, 103–108, doi: 10.1038/s41586-021-03817-4 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Starr TN et al. SARS-CoV-2 RBD antibodies that maximize breadth and resistance to escape. Nature 597, 97–102, doi: 10.1038/s41586-021-03807-6 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Telenti A et al. After the pandemic: perspectives on the future trajectory of COVID-19. Nature 596, 495–504, doi: 10.1038/s41586-021-03792-w (2021). [DOI] [PubMed] [Google Scholar]
  • 6.Wang P et al. Antibody resistance of SARS-CoV-2 variants B.1.351 and B.1.1.7. Nature 593, 130–135, doi: 10.1038/s41586-021-03398-2 (2021). [DOI] [PubMed] [Google Scholar]
  • 7.Chen RE et al. Resistance of SARS-CoV-2 variants to neutralization by monoclonal and serum-derived polyclonal antibodies. Nat Med 27, 717–726, doi: 10.1038/s41591-021-01294-w (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Corti D, Purcell LA, Snell G & Veesler D Tackling COVID-19 with neutralizing monoclonal antibodies. Cell 184, 3086–3108, doi: 10.1016/j.cell.2021.05.005 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.McCallum M & Walls AC Molecular basis of immune evasion by the Delta and Kappa SARS-CoV-2 variants. Science (2021). [DOI] [PubMed] [Google Scholar]
  • 10.Mlcochova P et al. SARS-CoV-2 B.1.617.2 Delta variant replication and immune evasion. Nature 599, 114–119, doi: 10.1038/s41586-021-03944-y (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Sheikh A et al. SARS-CoV-2 Delta VOC in Scotland: demographics, risk of hospital admission, and vaccine effectiveness. Lancet 397, 2461–2462, doi: 10.1016/S0140-6736(21)01358-1 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Piccoli L et al. Mapping Neutralizing and Immunodominant Sites on the SARS-CoV-2 Spike Receptor-Binding Domain by Structure-Guided High-Resolution Serology. Cell 183, 1024–1042 e1021, doi: 10.1016/j.cell.2020.09.037 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Greaney AJ et al. Comprehensive mapping of mutations in the SARS-CoV-2 receptor-binding domain that affect recognition by polyclonal human plasma antibodies. Cell Host Microbe 29, 463–476 e466, doi: 10.1016/j.chom.2021.02.003 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Walls AC et al. Structure, Function, and Antigenicity of the SARS-CoV-2 Spike Glycoprotein. Cell 181, 281–292 e286, doi: 10.1016/j.cell.2020.02.058 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Cele S et al. SARS-CoV-2 Omicron has extensive but incomplete escape of Pfizer BNT162b2 elicited neutralization and requires ACE2 for infection. medRxiv (2021). [Google Scholar]
  • 16.Wilhelm A et al. Reduced Neutralization of SARS-CoV-2 Omicron Variant by Vaccine Sera and monoclonal antibodies. medRxiv, doi: 10.1101/2021.12.07.21267432 (2021). [DOI] [Google Scholar]
  • 17.Andrews N et al. Effectiveness of COVID-19 vaccines against Omicron variant of concern. https://khub.net (2021).
  • 18.Cao YR et al. B.1.1.529 escapes the majority of SARS-CoV-2 neutralizing antibodies of diverse epitopes. doi: 10.1101/2021.12.07.470392 (2021). [DOI] [Google Scholar]
  • 19.Leist SR et al. A Mouse-Adapted SARS-CoV-2 Induces Acute Lung Injury and Mortality in Standard Laboratory Mice. Cell 183, 1070–1085 e1012, doi: 10.1016/j.cell.2020.09.050 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Dinnon KH 3rd et al. A mouse-adapted model of SARS-CoV-2 to test COVID-19 countermeasures. Nature 586, 560–566, doi: 10.1038/s41586-020-2708-8 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Collier DA et al. Sensitivity of SARS-CoV-2 B.1.1.7 to mRNA vaccine-elicited antibodies. Nature 593, 136–141, doi: 10.1038/s41586-021-03412-7 (2021). [DOI] [PubMed] [Google Scholar]
  • 22.Starr TN et al. Deep Mutational Scanning of SARS-CoV-2 Receptor Binding Domain Reveals Constraints on Folding and ACE2 Binding. Cell 182, 1295–1310 e1220, doi: 10.1016/j.cell.2020.08.012 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Shuai H et al. Emerging SARS-CoV-2 variants expand species tropism to murines. EBioMedicine 73, 103643, doi: 10.1016/j.ebiom.2021.103643 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Pan T et al. Infection of wild-type mice by SARS-CoV-2 B.1.351 variant indicates a possible novel cross-species transmission route. Signal Transduct Target Ther 6, 420, doi: 10.1038/s41392-021-00848-1 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Hoffmann M et al. The Omicron variant is highly resistant against antibody-mediated neutralization – implications for control of the COVID-19 pandemic. bioRxiv, doi: 10.1101/2021.12.12.472286 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Nicholis J & Chan Chi-wai M HKUMed finds Omicron SARS-CoV-2 can infect faster and better than Delta in human bronchus but with less severe infection in lung. www.med.hku.hk/en/news/press/20211215-omicron-sars-cov-2-infection. (2021).
  • 27.Bassi J et al. Poor neutralization and rapid decay of antibodies to SARS-CoV-2 variants in vaccinated dialysis patients. medRxiv, doi: 10.1101/2021.10.05.21264054v2 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Khoury DS et al. Neutralizing antibody levels are highly predictive of immune protection from symptomatic SARS-CoV-2 infection. Nat Med 27, 1205–1211, doi: 10.1038/s41591-021-01377-8 (2021). [DOI] [PubMed] [Google Scholar]
  • 29.Stamatatos L et al. mRNA vaccination boosts cross-variant neutralizing antibodies elicited by SARS-CoV-2 infection. Science, doi: 10.1126/science.abg9175 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Bergstrom JJ, Xu H & Heyman B Epitope-Specific Suppression of IgG Responses by Passively Administered Specific IgG: Evidence of Epitope Masking. Frontieres in Immunology 8, 238, doi: 10.3389/fimmu.2017.00238 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Baum A et al. Antibody cocktail to SARS-CoV-2 spike protein prevents rapid mutational escape seen with individual antibodies. Science 369, 1014–1018, doi: 10.1126/science.abd0831 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Hansen J et al. Studies in humanized mice and convalescent humans yield a SARS-CoV-2 antibody cocktail. Science 369, 1010–1014, doi: 10.1126/science.abd0827 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Shi R et al. A human neutralizing antibody targets the receptor-binding site of SARS-CoV-2. Nature 584, 120–124, doi: 10.1038/s41586-020-2381-y (2020). [DOI] [PubMed] [Google Scholar]
  • 34.Gottlieb RL et al. Effect of Bamlanivimab as Monotherapy or in Combination With Etesevimab on Viral Load in Patients With Mild to Moderate COVID-19: A Randomized Clinical Trial. JAMA 325, 632–644, doi: 10.1001/jama.2021.0202 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Jones BE et al. The neutralizing antibody, LY-CoV555, protects against SARS-CoV-2 infection in nonhuman primates. Sci Transl Med 13, doi: 10.1126/scitranslmed.abf1906 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Dong J et al. Genetic and structural basis for SARS-CoV-2 variant neutralization by a two-antibody cocktail. Nat Microbiol 6, 1233–1244, doi: 10.1038/s41564-021-00972-2 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Zost SJ et al. Potently neutralizing and protective human antibodies against SARS-CoV-2. Nature 584, 443–449, doi: 10.1038/s41586-020-2548-6 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Thomson EC et al. Circulating SARS-CoV-2 spike N439K variants maintain fitness while evading antibody-mediated immunity. Cell 184, 1171–1187 e1120, doi: 10.1016/j.cell.2021.01.037 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Starr TN et al. ACE2 binding is an ancestral and evolvable trait of sarbecoviruses. doi: 10.1101/2021.07.17.452804 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Lempp FA et al. Lectins enhance SARS-CoV-2 infection and influence neutralizing antibodies. Nature 598, 342–347, doi: 10.1038/s41586-021-03925-1 (2021). [DOI] [PubMed] [Google Scholar]
  • 41.VanBlargan LA et al. An infectious SARS-CoV-2 B.1.1.529 Omicron virus escapes neutralization by several therapeutic monoclonal antibodies. bioRxiv, doi: 10.1101/2021.12.15.472828 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Aggarwa A, Ospina Stella A & Walker G SARS-CoV-2 Omicron: evasion of potent humoral responses and resistance to clinical immunotherapeutics relative to viral variants of concern. bioRxiv (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.FDA. FACT SHEET FOR HEALTHCARE PROVIDERS EMERGENCY USE AUTHORIZATION (EUA) OF SOTROVIMAB. https://www.fda.gov/media/149534/download (2021).
  • 44.Tortorici MA et al. Ultrapotent human antibodies protect against SARS-CoV-2 challenge via multiple mechanisms. Science 370, 950–957, doi: 10.1126/science.abe3354 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.McCallum M et al. N-terminal domain antigenic mapping reveals a site of vulnerability for SARS-CoV-2. Cell 184, 2332–2347 e2316, doi: 10.1016/j.cell.2021.03.028 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.McCallum M et al. SARS-CoV-2 immune evasion by the B.1.427/B.1.429 variant of concern. Science, doi: 10.1126/science.abi7994 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Fischer W et al. HIV-1 and SARS-CoV-2: Patterns in the evolution of two pandemic pathogens. Cell Host Microbe 29, 1093–1110, doi: 10.1016/j.chom.2021.05.012 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Kupferschmidt K Where did ‘weird’ Omicron come from? Science 374, 1179, doi: 10.1126/science.acx9738 (2021). [DOI] [PubMed] [Google Scholar]
  • 49.Corey L et al. SARS-CoV-2 Variants in Patients with Immunosuppression. N Engl J Med 385, 562–566, doi: 10.1056/NEJMsb2104756 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Oude Munnink BB et al. Transmission of SARS-CoV-2 on mink farms between humans and mink and back to humans. Science 371, 172–177, doi: 10.1126/science.abe5901 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]

Additional References related to Methods

  • 51.Zang R et al. TMPRSS2 and TMPRSS4 promote SARS-CoV-2 infection of human small intestinal enterocytes. Sci Immunol 5, doi: 10.1126/sciimmunol.abc3582 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Maziarz RT et al. Control of an outbreak of human parainfluenza virus 3 in hematopoietic stem cell transplant recipients. Biol Blood Marrow Transplant 16, 192–198, doi: 10.1016/j.bbmt.2009.09.014 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Katoh K & Standley DM MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol 30, 772–780, doi: 10.1093/molbev/mst010 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Yamin R et al. Fc-engineered antibody therapeutics with improved anti-SARS-CoV-2 efficacy. Nature 599, 465–470, doi: 10.1038/s41586-021-04017-w (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Giroglou T et al. Retroviral vectors pseudotyped with severe acute respiratory syndrome coronavirus S protein. J Virol 78, 9007–9015, doi: 10.1128/JVI.78.17.9007-9015.2004 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Walls AC et al. Elicitation of broadly protective sarbecovirus immunity by receptor-binding domain nanoparticle vaccines. Cell 184, 5432–5447 e5416, doi: 10.1016/j.cell.2021.09.015 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Kaname Y et al. Acquisition of complement resistance through incorporation of CD55/decay-accelerating factor into viral particles bearing baculovirus GP64. J Virol 84, 3210–3219, doi: 10.1128/JVI.02519-09 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Plante JA et al. Spike mutation D614G alters SARS-CoV-2 fitness. Nature, doi: 10.1038/s41586-020-2895-3 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Imai M et al. Syrian hamsters as a small animal model for SARS-CoV-2 infection and countermeasure development. Proc Natl Acad Sci U S A 117, 16587–16595, doi: 10.1073/pnas.2009799117 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Liu Z et al. Identification of SARS-CoV-2 spike mutations that attenuate monoclonal and serum antibody neutralization. Cell Host Microbe 29, 477–488.e474, doi: 10.1016/j.chom.2021.01.014 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.VanBlargan LA et al. A potently neutralizing SARS-CoV-2 antibody inhibits variants of concern by utilizing unique binding residues in a highly conserved epitope. Immunity, doi: 10.1016/j.immuni.2021.08.016 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Cathcart AL et al. The dual function monoclonal antibodies VIR-7831 and VIR-7832 demonstrate potent in vitro and in vivo activity against SARS-CoV-2. bioRxiv, 2021.2003.2009.434607, doi: 10.1101/2021.03.09.434607 (2021). [DOI] [Google Scholar]
  • 63.Gupta A et al. Early Treatment for Covid-19 with SARS-CoV-2 Neutralizing Antibody Sotrovimab. N Engl J Med 385, 1941–1950, doi: 10.1056/NEJMoa2107934 (2021). [DOI] [PubMed] [Google Scholar]
  • 64.Kim C et al. A therapeutic neutralizing antibody targeting receptor binding domain of SARS-CoV-2 spike protein. Nat Commun 12, 288, doi: 10.1038/s41467-020-20602-5 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Ryu DK et al. Therapeutic effect of CT-P59 against SARS-CoV-2 South African variant. Biochem Biophys Res Commun 566, 135–140, doi: 10.1016/j.bbrc.2021.06.016 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Baum A et al. REGN-COV2 antibodies prevent and treat SARS-CoV-2 infection in rhesus macaques and hamsters. Science 370, 1110–1115, doi: 10.1126/science.abe2402 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Copin R et al. The monoclonal antibody combination REGEN-COV protects against SARS-CoV-2 mutational escape in preclinical and human studies. Cell 184, 3949–3961 e3911, doi: 10.1016/j.cell.2021.06.002 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Weinreich DM et al. REGN-COV2, a Neutralizing Antibody Cocktail, in Outpatients with Covid-19. N Engl J Med 384, 238–251, doi: 10.1056/NEJMoa2035002 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Chen P et al. SARS-CoV-2 Neutralizing Antibody LY-CoV555 in Outpatients with Covid-19. N Engl J Med 384, 229–237, doi: 10.1056/NEJMoa2029849 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Group A-TL-CS et al. A Neutralizing Monoclonal Antibody for Hospitalized Patients with Covid-19. N Engl J Med 384, 905–914, doi: 10.1056/NEJMoa2033130 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Belk J, Deveau LM, Rappazzo CG, Walker L & Wec A WO2021207597 - COMPOUNDS SPECIFIC TO CORONAVIRUS S PROTEIN AND USES THEREOF. ADAGIO THERAPEUTICS, INC. (2021). [Google Scholar]
  • 1.Cathcart AL et al. The dual function monoclonal antibodies VIR-7831 and VIR-7832 demonstrate potent in vitro and in vivo activity against SARS-CoV-2. bioRxiv, 2021.2003.2009.434607, doi: 10.1101/2021.03.09.434607 (2021). [DOI] [Google Scholar]
  • 2.Lempp FA et al. Lectins enhance SARS-CoV-2 infection and influence neutralizing antibodies. Nature 598, 342–347, doi: 10.1038/s41586-021-03925-1 (2021). [DOI] [PubMed] [Google Scholar]
  • 3.Pinto D et al. Cross-neutralization of SARS-CoV-2 by a human monoclonal SARS-CoV antibody. Nature 583, 290–295, doi: 10.1038/s41586-020-2349-y (2020). [DOI] [PubMed] [Google Scholar]
  • 4.Tortorici MA et al. Broad sarbecovirus neutralization by a human monoclonal antibody. Nature, doi: 10.1038/s41586-021-03817-4 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Starr TN et al. SARS-CoV-2 RBD antibodies that maximize breadth and resistance to escape. Nature 597, 97–102, doi: 10.1038/s41586-021-03807-6 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Gupta A et al. Early Treatment for Covid-19 with SARS-CoV-2 Neutralizing Antibody Sotrovimab. N. Engl. J. Med. 385, 1941–1950, doi: 10.1056/NEJMoa2107934 (2021). [DOI] [PubMed] [Google Scholar]
  • 7.McCallum M & Walls AC Molecular basis of immune evasion by the Delta and Kappa SARS-CoV-2 variants. Science (2021). [DOI] [PubMed] [Google Scholar]
  • 8.Kim C et al. A therapeutic neutralizing antibody targeting receptor binding domain of SARS-CoV-2 spike protein. Nat Commun 12, 288, doi: 10.1038/s41467-020-20602-5 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Ryu DK et al. Therapeutic effect of CT-P59 against SARS-CoV-2 South African variant. Biochem Biophys Res Commun 566, 135–140, doi: 10.1016/j.bbrc.2021.06.016 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Dong J et al. Genetic and structural basis for SARS-CoV-2 variant neutralization by a two-antibody cocktail. Nature Microbiology, doi: 10.1038/s41564-021-00972-2 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Zost SJ et al. Potently neutralizing and protective human antibodies against SARS-CoV-2. Nature 584, 443–449, doi: 10.1038/s41586-020-2548-6 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Baum A et al. REGN-COV2 antibodies prevent and treat SARS-CoV-2 infection in rhesus macaques and hamsters. Science 370, 1110–1115, doi: 10.1126/science.abe2402 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Baum A et al. Antibody cocktail to SARS-CoV-2 spike protein prevents rapid mutational escape seen with individual antibodies. Science 369, 1014–1018, doi: 10.1126/science.abd0831 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Copin R et al. The monoclonal antibody combination REGEN-COV protects against SARS-CoV-2 mutational escape in preclinical and human studies. Cell 184, 3949–3961 e3911, doi: 10.1016/j.cell.2021.06.002 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Hansen J et al. Studies in humanized mice and convalescent humans yield a SARS-CoV-2 antibody cocktail. Science 369, 1010–1014, doi: 10.1126/science.abd0827 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Weinreich DM et al. REGN-COV2, a Neutralizing Antibody Cocktail, in Outpatients with Covid-19. N. Engl. J. Med. 384, 238–251, doi: 10.1056/NEJMoa2035002 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Chen P et al. SARS-CoV-2 Neutralizing Antibody LY-CoV555 in Outpatients with Covid-19. N. Engl. J. Med. 384, 229–237, doi: 10.1056/NEJMoa2029849 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Gottlieb RL et al. Effect of Bamlanivimab as Monotherapy or in Combination With Etesevimab on Viral Load in Patients With Mild to Moderate COVID-19: A Randomized Clinical Trial. JAMA 325, 632–644, doi: 10.1001/jama.2021.0202 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Group A-TL-CS et al. A Neutralizing Monoclonal Antibody for Hospitalized Patients with Covid-19. N. Engl. J. Med. 384, 905–914, doi: 10.1056/NEJMoa2033130 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Jones BE et al. The neutralizing antibody, LY-CoV555, protects against SARS-CoV-2 infection in nonhuman primates. Sci Transl Med 13, doi: 10.1126/scitranslmed.abf1906 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Shi R et al. A human neutralizing antibody targets the receptor-binding site of SARS-CoV-2. Nature 584, 120–124, doi: 10.1038/s41586-020-2381-y (2020). [DOI] [PubMed] [Google Scholar]
  • 22.Thomson EC et al. Circulating SARS-CoV-2 spike N439K variants maintain fitness while evading antibody-mediated immunity. Cell 184, 1171–1187 e1120, doi: 10.1016/j.cell.2021.01.037 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Tortorici MA et al. Ultrapotent human antibodies protect against SARS-CoV-2 challenge via multiple mechanisms. Science 370, 950–957, doi: 10.1126/science.abe3354 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Piccoli L et al. Mapping Neutralizing and Immunodominant Sites on the SARS-CoV-2 Spike Receptor-Binding Domain by Structure-Guided High-Resolution Serology. Cell 183, 1024–1042 e1021, doi: 10.1016/j.cell.2020.09.037 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Collier DA et al. Sensitivity of SARS-CoV-2 B.1.1.7 to mRNA vaccine-elicited antibodies. Nature, doi: 10.1038/s41586-021-03412-7 (2021). [DOI] [PubMed] [Google Scholar]
  • 26.Park Y-J et al. Antibody-mediated broad sarbecovirus neutralization through ACE2 molecular mimicry. bioRxiv, doi: 10.1101/2021.10.13.464254 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.McCallum M et al. N-terminal domain antigenic mapping reveals a site of vulnerability for SARS-CoV-2. Cell 184, 2332–2347 e2316, doi: 10.1016/j.cell.2021.03.028 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supp

Data Availability Statement

All datasets generated and information presented in the study are available from the corresponding authors on reasonable request.

RESOURCES