Divergence of delta and beta variants and SARS-CoV-2 evolved in prolonged infection into distinct serological phenotypes ======================================================================================================================== * Sandile Cele * Farina Karim * Gila Lustig * San Emmanuel James * Tandile Hermanus * Eduan Wilkinson * Jumari Snyman * Mallory Bernstein * Khadija Khan * Shi-Hsia Hwa * Houriiyah Tegally * Sasha W. Tilles * Jennifer Giandhari * Ntombifuthi Mthabela * Matilda Mazibuko * Yashica Ganga * Bernadett I. Gosnell * Salim Abdool Karim * Willem Hanekom * Wesley C. Van Voorhis * Thumbi Ndung’u * COMMIT-KZN Team * Richard J. Lessells * Penny L. Moore * Mahomed-Yunus S. Moosa * Tulio de Oliveira * Alex Sigal ## Abstract SARS-CoV-2 continues to evolve variants of concern (VOC) which escape antibody neutralization and have enhanced transmission. One variant may escape immunity elicited by another, and the delta VOC has been reported to escape beta elicited immunity (1). Systematic mapping of the serological distance of current and emerging variants will likely guide the design of vaccines which can target all variants. Here we isolated and serologically characterized SARS-CoV-2 which evolved from an ancestral strain in a person with advanced HIV disease and delayed SARS-CoV-2 clearance. This virus showed evolving escape from self antibody neutralization immunity and decreased Pfizer BNT162b2 vaccine neutralization sensitivity. We mapped neutralization of evolved virus and ancestral, beta and delta variant viruses by antibodies elicited by each VOC in SARS-CoV-2 convalescent individuals. Beta virus showed moderate (7-fold) and delta slight escape from neutralizing immunity elicited by ancestral virus infection. In contrast, delta virus had stronger escape from beta elicited immunity (12-fold), and beta virus even stronger escape from delta immunity (34-fold). Evolved virus had 9-fold escape from ancestral immunity, 27-fold escape from delta immunity, but was effectively neutralized by beta immunity. We conclude that beta and delta are serologically distant, further than each is from ancestral, and that virus evolved in prolonged infection during advanced HIV disease is serologically close to beta and far from delta. These results suggest that SARS-CoV-2 is diverging into distinct serological phenotypes and that vaccines tailored to one variant may become vulnerable to infections with another. Variants of concern have evolved the ability to escape neutralization from convalescent plasma elicited by previous strains of SARS-CoV-2 (2, 3) as well as the ability to transmit more effectively (4) and may have other adaptations such as enhanced escape from the innate immune response (5). Neutralization is highly predictive of vaccine efficacy (6) and some variants show decreased neutralization by vaccine elicited immunity that may make vaccines less effective at reducing the frequency of infection (7). The alpha variant, first identified in the UK, shows relatively little escape (8, 9). However the beta variant first identified in South Africa (2, 3, 8, 10, 11), the gamma variant first identified in Brazil (12), and the lambda variant first identified in Peru (13) show neutralization escape to different degrees. The delta variant, first identified in India, has evolved a strong transmission advantage over other SARS-CoV-2 strains. It does not show a high degree of neutralization escape from plasma immunity elicited by ancestral strains but does show striking escape from beta plasma (1). Escape from neutralization involves substitutions and deletions in the spike glycoprotein of the virus which binds the ACE2 receptor on the cell surface (14). Mutations associated with neutralization escape are mostly found in the receptor binding domain (RBD) of spike, although the N-terminal domain is also important (14, 15). RBD mutations for the beta variant include the K417N, E484K and N501Y (see [https://covdb.stanford.edu/page/mutation-viewer/](https://covdb.stanford.edu/page/mutation-viewer/)). E484K and N501Y are shared with the gamma variant, which has the K417T instead of the beta K417N. Alpha shares the N501Y and in some cases has E484K. The lambda variant has the RBD L452Q and F490S substitutions. Delta has the L452R and T478K substitutions in the RBD. However, a strain of delta has evolved the K417N substitution shared with beta. HIV has the capacity to modulate immunity to other infections. The mechanism may differ between co-infecting pathogens but is thought to be at least partially because the antibody response is compromised through depletion and dysregulation of CD4 helper T cells (16, 17). ART allows PLWH to avoid the worst consequences of HIV infection, which are the result of severe depletion of CD4 T cells (18). However, lack of adherence to ART and development of drug resistance mutations leads to ongoing HIV replication, which, if left to persist for years, results in advanced HIV disease (18). We and others have recently shown that advanced HIV may lead to delayed clearance of SARS-CoV-2 and evolution of the SARS-CoV-2 virus (19, 20). Here we serologically characterized SARS-CoV-2 that had evolved in a person with advanced HIV who we previously described in a case report (20). The study participant was discharged following clinical recovery 10 days post-diagnosis according to South Africa guidelines and remained asymptomatic for most study visits. HIV viremia persisted up to day 190 due to pervasive problems with ART adherence as a result of complex social and psychological factors (20). Diagnosis of SARS-CoV-2 infection was 16 days post-symptom onset and study enrollment was 6 days post-diagnosis (20). We use time post-diagnosis as the timescale in this analysis. The CD4 count was <10 at enrollment (Fig 1A top row). It increased modestly at later timepoints, possibly due to the improved adherence to ART and a switch to dolutegravir based therapy which reduced the HIV viral load to below the level of clinical detection (20). SARS-CoV-2 was detected by qPCR until day 216 post-diagnosis (Fig 1A second row). We attempted to isolate live virus up to and including the day 216 post-diagnosis swab sample. While there was insufficient sample to isolate virus from the day 0 swab, we isolated and expanded SARS-CoV-2 from subsequent swabs until and including day 190 post-SARS-CoV-2 diagnosis (Fig 1A third row). Successful isolation indicates that live virus was shed. RBD specific IgG antibodies in the blood were at borderline detection levels (slightly above the mean negative control + 2 std) at the early timepoints (Fig 1A, fourth row) but were detected at higher levels starting day 190 peaking at day 216 (Fig S1). ![Fig S 1:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2021/09/23/2021.09.14.21263564/F3.medium.gif) [Fig S 1:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/F3) Fig S 1: Spike specific antibody levels with time post-SARS-CoV-2 diagnosis. Shown are mean (n=4 per timepoint) and standard deviation of anti-spike RBD antibody concentrations measured in the plasma of the participant with advanced HIV disease by ELISA. Red dashed line denotes the mean + 2 standard deviations of signal from a set of 6 samples of including plasma of pre-pandemic controls (n=4) and pre-pandemic commercial human serum (n=2). ![Figure 1:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2021/09/23/2021.09.14.21263564/F1.medium.gif) [Figure 1:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/F1) Figure 1: Evolution of SARS-CoV-2 with neutralization escape in a participant with advanced HIV (A) Participant characteristics over 233 days from SARS-CoV-2 diagnosis including CD4 T cell count (cells/*µ*L), presence or absence of SARS-CoV-2 by qPCR, success or failure of live virus outgrowth from swab sample, and presence or absence of anti-RBD IgG by ELISA. Because IgG levels were close to the background cutoff for some timepoints, they were marked as borderline. (B) Majority and minority SARS-CoV-2 genotypes in the swab (day 0) and outgrowth (day 6 to 190) sequences. X-axis lists substitutions and deletions in spike sequence. AF: allele frequency. (C) Live virus neutralization assay of the early day 6 (D6) SARS-CoV-2 virus isolate by self-plasma collected at days 6 to 216. Plasma dilution shown is 1:20. (D) Neutralization of D6, day 20 isolated (D20), and day 190 isolated (D190) virus by self-plasma collected days 6 to 216. (E) Neutralization of D6 virus relative to D190 virus by convalescent plasma of participants (n=8) infected by ancestral strains. Number in black above virus name is geometric mean of the PRNT50 and number in red represents fold-change of geometric mean between virus on the left and on the right of graph. (F) Neutralization of D20 isolate compared to D190 isolate by convalescent plasma from the same participants as in (E). (G) Neutralization of D190 isolate compared to ancestral virus (D614G) by Pfizer BNT162b2 elicited plasma (n=12). p-values are * <0.05; ** <0.01-0.001 as determined by the Wilcoxon rank sum test. Red horizontal line in panels D to G denotes most concentrated plasma tested. Outgrown virus was sequenced to detect majority and minority variants (Fig 1B). The mutations found in the outgrown virus were representative of the virus in the swab from the matched time point (Tables S1-S2) except for the R682W substitution at the furin cleavage site in D6. This mutation evolves *in vitro* during expansion in VeroE6 cells and likely confers moderate neutralization escape (21). E484K was first detected on day 6 (Fig 1B). This mutation persisted at day 20 and 34 but was replaced with the F490S substitution starting on day 71 when the K417T mutation was also detected. The N501Y mutation emerged on day 190 post-diagnosis. View this table: [Table S1:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/T1) Table S1: Read numbers at nucleotides which led to amino acid substitutions or deletions in virus sequenced from the swab View this table: [Table S2:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/T2) Table S2: Read numbers at nucleotides which led to amino acid substitutions or deletions in virus sequenced from outgrown stock We tested three of the isolates: viruses outgrown from day 6 and day 20 swabs (designated D6, D20) representing viruses from early infection, and virus outgrown from the day 190 swab (D190), after substantial evolution. We used a live virus neutralization assay (LVNA, materials and methods) to examine the capacity of the participant plasma up to and including day 216 to neutralize the D6, D20, and D190 isolates. LVNA reads out as the reduction in the number of infection foci at different plasma concentrations (Figure 1C). The results are fitted to a sigmoidal function to obtain the plasma dilution needed for 50% inhibition. We report PRNT50, the reciprocal of this dilution (2). Neutralization of the D6, D20, and D190 isolates by self-plasma was low at the early timepoints (Fig 1D). However, neutralization of D6 and D20 was evident by plasma sampled from day 190 and was clear in the plasma sampled from day 216. The D6 isolate was the most sensitive to neutralization by day 216 plasma. Neutralization declined for D20 and further for D190, suggesting sequential evolution of escape (Fig 1D). We also tested the D6, D20, and D190 isolates against plasma from other convalescent participants also infected with ancestral variants in the same infection wave (see Table S3 for infected study participant details from all infection waves and associated infecting sequence accession numbers where available). Neutralization of D190 by the matched-wave plasma was decreased dramatically relative to D6, with PRNT50 for D190 being 9.3-fold lower despite the presence of the E484K mutation in D6 (Figure 1E). The difference was smaller between D190 and D20 (5.1-fold, Figure 1F), consistent with evolution of some neutralization escape in D20 relative to D6. We also tested neutralization of D190 virus using Pfizer BNT162b2 vaccinated participants (Table S4). BNT162b2-elicited plasma neutralization capacity was decreased against D190 by 5-fold relative to D614G (Figure 1G). These results support evolution of escape from neutralization of D190 virus. View this table: [Table S3:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/T3) Table S3: Characteristics of SARS-CoV-2 convalescent study participants View this table: [Table S4:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/T4) Table S4: Characteristics of Pfizer BNT162b2 vaccinated participants We next assessed whether the highly evolved D190 virus serologically resembles a variant of concern. We therefore tested the D190 isolate against convalescent plasma samples obtained from infections by three different variants. South Africa so far has had three infection waves (Fig 2A). The first infection wave consisted of ancestral strains with the D614G substitution. The second infection wave was dominated by the beta variant. The third wave, currently in progress, was dominated by the delta variant. We have obtained both viral isolates and plasma from convalescent individuals for each infection wave, and we sequenced viruses eliciting the plasma immunity (Table S3). ![Figure 2:](http://medrxiv.org/https://www.medrxiv.org/content/medrxiv/early/2021/09/23/2021.09.14.21263564/F2.medium.gif) [Figure 2:](http://medrxiv.org/content/early/2021/09/23/2021.09.14.21263564/F2) Figure 2: Mapping neutralization of variants and the *in vivo* evolved D190 virus. (A) Infection waves in South Africa and variant frequencies in each. First infection wave consisted of ancestral strains with the D614G mutation, second wave was dominated by the beta variant, and third wave was dominated by the delta variant, with a low frequency of alpha variant infections. (B) Neutralization of the beta variant compared to ancestral virus with the D614G mutation by plasma from convalescent participants infected by ancestral strains (n=8). (C) Neutralization of the delta variant compared to ancestral virus by same plasma as (B). (D) Neutralization of the D190 isolate compared to ancestral virus by same plasma as (B). (E) Neutralization of the delta variant compared to beta variant by plasma from participants infected by the beta variant (n=9). (F) Neutralization of D190 compared to beta variant by same plasma as (E). (G) Neutralization of the beta variant compared to delta variant by plasma of participants infected by the delta variant (n=10). (H) Neutralization of D190 compared to delta variant by the same plasma as (G). (I) Neutralization of the beta variant compared to delta variant by the same plasma as (G) using a pseudo-virus neutralization assay.Neutralization was measured by a reduction in luciferase gene expression after single-round infection with beta and delta spike-pseudotyped viruses. Titers were calculated as the reciprocal plasma dilution (ID50) causing 50% reduction of relative light units (J) Summary map (not to scale) of serological distances as measured by fold-decrease in neutralization. p-values are * <0.05; ** <0.01; \***| < 0.001, \**\*|\* < 0.0001 as determined by the Wilcoxon rank sum test. Red horizontal line in panels B to I denotes most concentrated plasma tested. We first mapped the neutralization capacity of plasma antibodies triggered by each variant against the homologous virus (from the same wave) versus heterologous variants from the different waves. We found that ancestral plasma PRNT50 declined 7.2-fold against the beta variant (Fig 2B), but only two-fold against delta, a drop that was not statistically significant (Fig 2C). Neutralization by ancestral plasma declined 8.8-fold relative to ancestral virus for the D190 isolate (Fig 2D), similar to beta escape. Using beta-elicited plasma, we observed a 12.4-fold decline of neutralization of delta, compared beta virus neutralization (Fig 2E). In contrast, the D190 virus was neutralized relatively well by beta plasma with only a 2.6-fold reduction (Fig 2F). A much more dramatic decline was observed when delta plasma was used to neutralize beta virus or the D190 isolate. This showed 33.6-fold and 27.1-fold drops in neutralization capacity, respectively, compared to neutralization of delta virus (Fig 2G and 2H). Because we have never previously observed the degree of escape seen with beta virus in delta plasma, we repeated the experiments using a pseudovirus neutralization assay where the beta and delta spike was engineered and expressed on a lentivirus vector (3). We observed increased sensitivity to neutralization with this system relative to LVNA. However, the drop in neutralization of beta virus with delta plasma, at 26.1-fold, was strikingly similar (Fig 2I). Mapping these results (Fig 2J) shows that beta and delta are serologically far apart, with ancestral virus forming a bridge between the variants, with the beta further removed from ancestral than delta. The distance from ancestral is consistent with beta being an escape variant and so evolving escape mutations such as E484K and K417N which make it serologically distinct from ancestral strains. The D190 evolved virus was serologically similar to beta, an intermediate distance from the ancestral, and far from delta. The D190 was the most evolved and the most resistant to neutralization of the isolates tested from the case of advanced HIV with prolonged SARS-CoV-2 infection, even though the early D6 isolate had the E484K mutation (and the in vitro evolved R682W which is reported to confer a moderate decrease in neutralization). Among the RBD mutations in the D190 isolate, K417T is found in the gamma, and F490S is found in the lambda variant. Despite these key sequence differences from beta, the evolved virus shows serologically similar behavior to beta, which is escape from ancestral plasma and a much more pronounced escape from delta plasma neutralization. The driving force behind the evolution may have been the presence of very low levels of antibody targeting SARS-CoV-2 RBD. In another study, we have observed that the fraction of PLWH co-infected with SARS-CoV-2 in whom ART was not detected was increased 2-fold between the first and second infection waves in South Africa. Consistent with this, the fraction of people with detectable HIV viremia doubled (22). If immunosuppression by advanced HIV is indeed a driver of SARS-CoV-2 evolution, then we should expect additional SARS-CoV-2 evolution, unless ART coverage is increased. Divergence of one virus into serologically distinct strains is observed in other viruses, examples being polio (23) and dengue (24). Upon infection, immunity is developed to the infecting but not the other serotypes. Whether this will become the case for antibody immunity to SARS-CoV-2 variants is still unclear. However, in this study we do observe a widening serological divide between variants, with the ancestral virus at the hub. The *in vivo* evolved SARS-CoV-2 we tested also fits this pattern, which may give some indication of the behavior of future variants. This may imply that a vaccine which is broadly effective across variants should either be designed around the ancestral strain, or with all the diverging variants. The use of only one variant may generate immunity which is vulnerable to infection by another variant. ## Methods ### Ethical statement Nasopharyngeal and oropharyngeal swab samples and plasma samples were obtained from hospitalized adults with PCR-confirmed SARS-CoV-2 infection who were enrolled in a prospective cohort study approved by the Biomedical Research Ethics Committee at the University of KwaZulu–Natal (reference BREC/00001275/2020). ### Whole-genome sequencing, genome assembly and phylogenetic analysis cDNA synthesis was performed on the extracted RNA using random primers followed by gene-specific multiplex PCR using the ARTIC V.3 protocol ([https://www.protocols.io/view/covid-19-artic-v3-illumina-library-construction-an-bibtkann](https://www.protocols.io/view/covid-19-artic-v3-illumina-library-construction-an-bibtkann)). In brief, extracted RNA was converted to cDNA using the Superscript IV First Strand synthesis system (Life Technologies) and random hexamer primers. SARS-CoV-2 whole-genome amplification was performed by multiplex PCR using primers designed using Primal Scheme ([http://primal.zibraproject.org/](http://primal.zibraproject.org/)) to generate 400-bp amplicons with an overlap of 70 bp that covers the 30 kb SARS-CoV-2 genome. PCR products were cleaned up using AmpureXP purification beads (Beckman Coulter) and quantified using the Qubit dsDNA High Sensitivity assay on the Qubit 4.0 instrument (Life Technologies). We then used the Illumina Nextera Flex DNA Library Prep kit according to the manufacturer’s protocol to prepare indexed paired-end libraries of genomic DNA. Sequencing libraries were normalized to 4 nM, pooled and denatured with 0.2 N sodium acetate. Then, a 12-pM sample library was spiked with 1% PhiX (a PhiX Control v.3 adaptor-ligated library was used as a control). We sequenced libraries on a 500-cycle v.2 MiSeq Reagent Kit on the Illumina MiSeq instrument (Illumina). We assembled paired-end fastq reads using Genome Detective 1.126 ([https://www.genomedetective.com](https://www.genomedetective.com)) and the Coronavirus Typing Tool. We polished the initial assembly obtained from Genome Detective by aligning mapped reads to the reference sequences and filtering out low-quality mutations using the bcftools 1.7-2 mpileup method. Mutations were confirmed visually with BAM files using Geneious software (Biomatters). All of the sequences were deposited in GISAID ([https://www.gisaid.org/](https://www.gisaid.org/)). We analysed sequences from the seven different time points (D0, D6, D20, D34, D71, D106 and D190) as well as a random sample of genotypes from South Africa (n=480) against the global reference dataset (n = 3,170) using a custom pipeline based on a local version of NextStrain ([https://github.com/nextstrain/ncov](https://github.com/nextstrain/ncov)).The pipeline contains several Python scripts that manage the analysis workflow. It performs alignment of genotypes in MAFFT, phylogenetic tree inference in IQ-Tree20, tree dating and ancestral state construction and annotation. All seven sequences from the patient clustered in a monophyletic clade ([https://nextstrain.org/groups/ngs-sa/COVID19-AHRI-2021.05.27?label=clade:HIV%20patient](https://nextstrain.org/groups/ngs-sa/COVID19-AHRI-2021.05.27?label=clade:HIV%20patient)) that are well separated from the rest of the phylogeny. ### Cells Vero E6 cells (ATCC CRL-1586, obtained from Cellonex in South Africa) were propagated in complete DMEM with 10% fetal bovine serum (Hylone) containing 1% each of HEPES, sodium pyruvate, L-glutamine and nonessential amino acids (Sigma-Aldrich). Vero E6 cells were passaged every 3–4 days. The H1299-E3 cell line for first-passage SARS-CoV-2 expansion, derived as described in (2), was propagated in complete RPMI with 10% fetal bovine serum containing 1% each of HEPES, sodium pyruvate, L-glutamine and nonessential amino acids. H1299 cells were passaged every second day. Cell lines have not been authenticated. The cell lines have been tested for mycoplasma contamination and are mycoplasma negative. ### Virus expansion All work with live virus was performed in Biosafety Level 3 containment using protocols for SARS-CoV-2 approved by the Africa Health Research Institute Biosafety Committee. We used ACE2-expressing H1299-E3 cells for the initial isolation (P1 stock) followed by passaging in Vero E6 cells (P2 and P3 stocks, where P3 stock was used in experiments). ACE2-expressing H1299-E3 cells were seeded at 1.5 × 105 cells per mL and incubated for 18–20 h. After one DPBS wash, the sub-confluent cell monolayer was inoculated with 500 μL universal transport medium diluted 1:1 with growth medium filtered through a 0.45-μm filter. Cells were incubated for 1 h. Wells were then filled with 3 mL complete growth medium. After 8 days of infection, cells were trypsinized, centrifuged at 300 rcf for 3 min and resuspended in 4 mL growth medium. Then 1 mL was added to Vero E6 cells that had been seeded at 2 × 105 cells per mL 18–20 h earlier in a T25 flask (approximately 1:8 donor-to-target cell dilution ratio) for cell-to-cell infection. The coculture of ACE2-expressing H1299-E3 and Vero E6 cells was incubated for 1 h and the flask was then filled with 7 mL of complete growth medium and incubated for 6 days. The viral supernatant (P2 stock) was aliquoted and stored at −80 °C and further passaged in Vero E6 cells to obtain the P3 stock used in experiments as follows: a T25 flask (Corning) was seeded with Vero E6 cells at 2 × 105 cells per mL and incubated for 18–20 h. After one DPBS wash, the sub-confluent cell monolayer was inoculated with 500 μL universal transport medium diluted 1:1 with growth medium and filtered through a 0.45-μm filter. Cells were incubated for 1 h. The flask was then filled with 7 mL of complete growth medium. After infection for 4 days, supernatants of the infected culture were collected, centrifuged at 300 rcf for 3 min to remove cell debris and filtered using a 0.45-μm filter. Viral supernatant was aliquoted and stored at −80 °C. ### Live virus neutralization assay Vero E6 cells were plated in a 96-well plate (Corning) at 30,000 cells per well 1 day before infection. Approximately 5 mL sterile water was added between wells to prevent wells at the edge drying more rapidly, which we have observed to cause edge effects resulting in lower number of foci. Plasma was separated from EDTA-anticoagulated blood by centrifugation at 500 rcf for 10 min and stored at −80 °C. Aliquots of plasma samples were heat-inactivated at 56 °C for 30 min and clarified by centrifugation at 10,000 rcf for 5 min, after which the clear middle layer was used for experiments. Inactivated plasma was stored in single-use aliquots to prevent freeze–thaw cycles. For experiments, plasma was serially diluted twofold from 1:100 to 1:1,600; this is the concentration that was used during the virus–plasma incubation step before addition to cells and during the adsorption step. As a positive control, the GenScript A02051 anti-spike monoclonal antibody was added. Virus stocks were used at approximately 50-100 focus-forming units per microwell and added to diluted plasma; antibody–virus mixtures were incubated for 1 h at 37 °C, 5% CO2. Cells were infected with 100 μL of the virus–antibody mixtures for 1 h, to allow adsorption of virus. Subsequently, 100 μL of a 1X RPMI 1640 (Sigma-Aldrich, R6504), 1.5% carboxymethylcellulose (Sigma-Aldrich, C4888) overlay was added to the wells without removing the inoculum. Cells were fixed at 18 h after infection using 4% paraformaldehyde (Sigma-Aldrich) for 20 min. For staining of foci, a rabbit anti-spike monoclonal antibody (BS-R2B12, GenScript A02058) was used at 0.5 μg/mL as the primary detection antibody. Antibody was resuspended in a permeabilization buffer containing 0.1% saponin (Sigma-Aldrich), 0.1% BSA (Sigma-Aldrich) and 0.05% Tween-20 (Sigma-Aldrich) in PBS. Plates were incubated with primary antibody overnight at 4 °C, then washed with wash buffer containing 0.05% Tween-20 in PBS. Secondary goat anti-rabbit horseradish peroxidase (Abcam ab205718) antibody was added at 1 μg/mL and incubated for 2 h at room temperature with shaking. The TrueBlue peroxidase substrate (SeraCare 5510-0030) was then added at 50 μL per well and incubated for 20 min at room temperature. Plates were then dried for 2 h and imaged using a Metamorph-controlled Nikon TiE motorized microscope with a 2X objective or ELISPOT instrument with built-in image analysis (C.T.L). For microscopy images, automated image analysis was performed using a custom script in MATLAB v.2019b (Mathworks), in which focus detection was automated and did not involve user curation. Plasma dilutions used were 1:10, 1:20, 1:40, 1:80, 1:160, 1:320, 1:640, 1:1280 for self-plasma and 1:25, 1:50, 1:100, 1:200, 1:400, 1:800, 1:1600 for all other plasma samples tested. ### Pseudovirus neutralization assay SARS-CoV-2 pseudotyped lentiviruses were prepared by co-transfecting the HEK 293T cell line with either the SARS-CoV-2 Beta spike (L18F, D80A, D215G, K417N, E484K, N501Y, D614G, A701V, 242-244 del) or the Delta spike (T19R, R158G L452R, T478K, D614G, P681R, D950N, 156-157 del) plasmids in conjunction with a firefly luciferase encoding lentivirus backbone plasmid. For the neutralization assay, heat-inactivated plasma samples from vaccine recipients were incubated with the SARS-CoV-2 pseudotyped virus for 1 hour at 37°C, 5% CO2. Subsequently, 1×104 HEK 293T cells engineered to over-express ACE-2 were added and incubated at 37°C, 5% CO2 for 72 hours upon which the luminescence of the luciferase gene was measured. CB6 was used as a positive control. ### Receptor binding domain ELISA Plasma samples were tested for anti-SARS-CoV-2 IgG. Flat bottom microplates (ThermoFisher Scientific) were coated with 500 ng/mL of the receptor binding domain (RBD) protein (provided by Dr Galit Alter from the Ragon Institute) and incubated overnight at 4°C. Plates were blocked with a 200 µL/well tris-buffered saline containing 1% BSA (TBSA) and incubated at room temperature (RT) for 1h. Samples were diluted in TBSA with 0.05% Tween-20 (TBSAT) to 1:100. Subsequently, goat anti-human IgG (1:5000)-horseradish peroxidase conjugated secondary antibodies (Jackson ImmunoResearch) were added a 100 µL/well and incubated at RT for 1h. Bound secondary antibodies were detected using 100 μL/well 1-step Ultra TMB substrate (ThermoFisher Scientific). Plates were incubated at RT for 3 min in the dark before addition of 1 N sulphuric acid stop solution at 100 μL/well. Plates were washed with 1X high salt TBS containing 0.05% Tween-20, three times each after coating and blocking, and five times each after the sample and secondary antibody. The concentration of anti-RBD expressed as ng/mL equivalent of anti-SARS-CoV-2 monoclonal, CR3022 (Genscript). We used pre-pandemic plasma samples as negative controls to define seroconversion cut-offs calculated as mean + 2 std of the negative samples. ### Statistics and fitting All statistics and fitting were performed using MATLAB v.2019b. Neutralization data were fit to Tx=1/1+(D/ID50). Here Tx is the number of foci normalized to the number of foci in the absence of plasma on the same plate at dilution D and ID50 is the plasma dilution giving 50% neutralization. PRNT50 = 1/ID50. Values of PRNT50 <1 are set to 1 (undiluted), the lowest measurable value. ## Data Availability All data available in the manuscript, GISAID SARS-CoV-2 sequence repository, or upon reasonable request from the authors. [https://www.gisaid.org/](https://www.gisaid.org/) ## COMMIT-KZN Team Moherndran Archary, Department of Paediatrics and Child Health, University of KwaZulu-Natal, Durban, South Africa. Philip Goulder, Africa Health Research Institute and Department of Paediatrics, Oxford, UK Nokwanda Gumede, Africa Health Research Institute, Durban, South Africa. Ravindra K. Gupta, Africa Health Research Institute and Cambridge Institute of Therapeutic Immunology & Infectious Disease, Cambridge, UK. Guy Harling, Africa Health Research Institute and the Institute for Global Health, University College London, London, UK. Rohen Harrichandparsad, Department of Neurosurgery, University of KwaZulu-Natal, Durban, South Africa. Kobus Herbst, Africa Health Research Institute and the South African Population Research Infrastructure Network, Durban, South Africa. Prakash Jeena, Department of Paediatrics and Child Health, University of KwaZulu-Natal, Durban, South Africa. Zesuliwe Jule, Africa Health Research Institute, Durban, South Africa. Thandeka Khoza, Africa Health Research Institute, Durban, South Africa. Nigel Klein, Africa Health Research Institute and the Institute of Child Health, University College London, London, UK. Henrik Kloverpris, Africa Health Research Institute, Durban, South Africa. Alasdair Leslie, Africa Health Research Institute, Durban, South Africa. Rajhmun Madansein, Department of Cardiothoracic Surgery, University of KwaZulu-Natal, Durban, South Africa. Mohlopheni Marakalala, Africa Health Research Institute, Durban, South Africa. Yoliswa Miya, Africa Health Research Institute, Durban, South Africa. Mosa Moshabela, College of Health Sciences, University of KwaZulu-Natal, Durban, South Africa. Nokukhanya Msomi, Department of Virology, University of KwaZulu-Natal, Durban, South Africa. Kogie Naidoo, Centre for the AIDS Programme of Research in South Africa, Durban, South Africa. Zaza Ndhlovu, Africa Health Research Institute, Durban, South Africa. Kennedy Nyamande, Department of Pulmonology and Critical Care, University of KwaZulu-Natal, Durban, South Africa. Vinod Patel, Department of Neurology, University of KwaZulu-Natal, Durban, South Africa. Dirhona Ramjit, Africa Health Research Institute, Durban, South Africa. Kajal Reedoy, Africa Health Research Institute, Durban, South Africa. Theresa Smit, Africa Health Research Institute, Durban, South Africa. Adrie Steyn, Africa Health Research Institute, Durban, South Africa. Emily Wong, Africa Health Research Institute, Durban, South Africa. ## Acknowledgements This study was supported by the Bill and Melinda Gates award INV-018944 (AS), National Institutes of Health award R01 AI138546 (AS), South African Medical Research Council awards (AS, TdO, PLM) and National Institutes of Health U01 AI151698 (WVV). PLM is supported by the South African Research Chairs Initiative of the Department of Science and Innovation and the NRF (Grant No 9834). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. We thank Hylton Rodel for advice on figure preparation. ## Footnotes * § A list of authors and affiliations appears at the end of the paper. * Received September 14, 2021. * Revision received September 14, 2021. * Accepted September 23, 2021. * © 2021, Posted by Cold Spring Harbor Laboratory This pre-print is available under a Creative Commons License (Attribution-NoDerivs 4.0 International), CC BY-ND 4.0, as described at [http://creativecommons.org/licenses/by-nd/4.0/](http://creativecommons.org/licenses/by-nd/4.0/) ## References 1. 1.Liu C, Ginn HM, Dejnirattisai W, Supasa P, Wang B, Tuekprakhon A, Nutalai R, Zhou D, Mentzer AJ, Zhao Y. Reduced neutralization of SARS-CoV-2 B. 1.617 by vaccine and convalescent serum. Cell. 2021;184(16):4220-36. e13. 2. 2.Cele S, Gazy I, Jackson L, Hwa SH, Tegally H, Lustig G, Giandhari J, Pillay S, Wilkinson E, Naidoo Y, Karim F, Ganga Y, Khan K, Bernstein M, Balazs AB, Gosnell BI, Hanekom W, Moosa MS, Network for Genomic Surveillance in South A, Team C-K, Lessells RJ, de Oliveira T, Sigal A. Escape of SARS-CoV-2 501Y.V2 from neutralization by convalescent plasma. Nature. 2021;593(7857):142–6. 3. 3.Wibmer CK, Ayres F, Hermanus T, Madzivhandila M, Kgagudi P, Oosthuysen B, Lambson BE, de Oliveira T, Vermeulen M, van der Berg K, Rossouw T, Boswell M, Ueckermann V, Meiring S, von Gottberg A, Cohen C, Morris L, Bhiman JN, Moore PL. SARS-CoV-2 501Y.V2 escapes neutralization by South African COVID-19 donor plasma. Nat Med. 2021;27(4):622–5. 4. 4.Campbell F, Archer B, Laurenson-Schafer H, Jinnai Y, Konings F, Batra N, Pavlin B, Vandemaele K, Van Kerkhove MD, Jombart T. Increased transmissibility and global spread of SARS-CoV-2 variants of concern as at June 2021. Eurosurveillance. 2021;26(24):2100509. 5. 5.Guo K, Barrett BS, Mickens KL, Hasenkrug KJ, Santiago ML. Interferon Resistance of Emerging SARS-CoV-2 Variants. bioRxiv. 2021. 6. 6.Khoury DS, Cromer D, Reynaldi A, Schlub TE, Wheatley AK, Juno JA, Subbarao K, Kent SJ, Triccas JA, Davenport MP. Neutralizing antibody levels are highly predictive of immune protection from symptomatic SARS-CoV-2 infection. Nature medicine. 2021:1–7. 7. 7.Madhi SA, Baillie V, Cutland CL, Voysey M, Koen AL, Fairlie L, Padayachee SD, Dheda K, Barnabas SL, Bhorat QE, Briner C, Kwatra G, Ahmed K, Aley P, Bhikha S, Bhiman JN, Bhorat AE, du Plessis J, Esmail A, Groenewald M, Horne E, Hwa SH, Jose A, Lambe T, Laubscher M, Malahleha M, Masenya M, Masilela M, McKenzie S, Molapo K, Moultrie A, Oelofse S, Patel F, Pillay S, Rhead S, Rodel H, Rossouw L, Taoushanis C, Tegally H, Thombrayil A, van Eck S, Wibmer CK, Durham NM, Kelly EJ, Villafana TL, Gilbert S, Pollard AJ, de Oliveira T, Moore PL, Sigal A, Izu A, Group N-S, Wits VCG. Efficacy of the ChAdOx1 nCoV-19 Covid-19 Vaccine against the B.1.351 Variant. N Engl J Med. 2021;384(20):1885–98. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1056/NEJMoa2102214&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=33725432&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) 8. 8.Garcia-Beltran WF, Lam EC, Denis KS, Nitido AD, Garcia ZH, Hauser BM, Feldman J, Pavlovic MN, Gregory DJ, Poznansky MC. Multiple SARS-CoV-2 variants escape neutralization by vaccine-induced humoral immunity. Cell. 2021;184(9):2372-83. e9. 9. 9.Supasa P, Zhou D, Dejnirattisai W, Liu C, Mentzer AJ, Ginn HM, Zhao Y, Duyvesteyn HME, Nutalai R, Tuekprakhon A, Wang B, Paesen GC, Slon-Campos J, Lopez-Camacho C, Hallis B, Coombes N, Bewley KR, Charlton S, Walter TS, Barnes E, Dunachie SJ, Skelly D, Lumley SF, Baker N, Shaik I, Humphries HE, Godwin K, Gent N, Sienkiewicz A, Dold C, Levin R, Dong T, Pollard AJ, Knight JC, Klenerman P, Crook D, Lambe T, Clutterbuck E, Bibi S, Flaxman A, Bittaye M, Belij-Rammerstorfer S, Gilbert S, Hall DR, Williams MA, Paterson NG, James W, Carroll MW, Fry EE, Mongkolsapaya J, Ren J, Stuart DI, Screaton GR. Reduced neutralization of SARS-CoV-2 B.1.1.7 variant by convalescent and vaccine sera. Cell. 2021;184(8):2201–11 e7. 10. 10.Tegally H, Wilkinson E, Giovanetti M, Iranzadeh A, Fonseca V, Giandhari J, Doolabh D, Pillay S, San EJ, Msomi N, Mlisana K, von Gottberg A, Walaza S, Allam M, Ismail A, Mohale T, Glass AJ, Engelbrecht S, Van Zyl G, Preiser W, Petruccione F, Sigal A, Hardie D, Marais G, Hsiao NY, Korsman S, Davies MA, Tyers L, Mudau I, York D, Maslo C, Goedhals D, Abrahams S, Laguda-Akingba O, Alisoltani-Dehkordi A, Godzik A, Wibmer CK, Sewell BT, Lourenco J, Alcantara LCJ, Kosakovsky Pond SL, Weaver S, Martin D, Lessells RJ, Bhiman JN, Williamson C, de Oliveira T. Detection of a SARS-CoV-2 variant of concern in South Africa. Nature. 2021;592(7854):438–43. [PubMed](http://medrxiv.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) 11. 11.Zhou D, Dejnirattisai W, Supasa P, Liu C, Mentzer AJ, Ginn HM, Zhao Y, Duyvesteyn HME, Tuekprakhon A, Nutalai R, Wang B, Paesen GC, Lopez-Camacho C, Slon-Campos J, Hallis B, Coombes N, Bewley K, Charlton S, Walter TS, Skelly D, Lumley SF, Dold C, Levin R, Dong T, Pollard AJ, Knight JC, Crook D, Lambe T, Clutterbuck E, Bibi S, Flaxman A, Bittaye M, Belij-Rammerstorfer S, Gilbert S, James W, Carroll MW, Klenerman P, Barnes E, Dunachie SJ, Fry EE, Mongkolsapaya J, Ren J, Stuart DI, Screaton GR. Evidence of escape of SARS-CoV-2 variant B.1.351 from natural and vaccine-induced sera. Cell. 2021;184(9):2348–61 e6. 12. 12.Dejnirattisai W, Zhou D, Supasa P, Liu C, Mentzer AJ, Ginn HM, Zhao Y, Duyvesteyn HME, Tuekprakhon A, Nutalai R, Wang B, Lopez-Camacho C, Slon-Campos J, Walter TS, Skelly D, Costa Clemens SA, Naveca FG, Nascimento V, Nascimento F, Fernandes da Costa C, Resende PC, Pauvolid-Correa A, Siqueira MM, Dold C, Levin R, Dong T, Pollard AJ, Knight JC, Crook D, Lambe T, Clutterbuck E, Bibi S, Flaxman A, Bittaye M, Belij-Rammerstorfer S, Gilbert SC, Carroll MW, Klenerman P, Barnes E, Dunachie SJ, Paterson NG, Williams MA, Hall DR, Hulswit RJG, Bowden TA, Fry EE, Mongkolsapaya J, Ren J, Stuart DI, Screaton GR. Antibody evasion by the P.1 strain of SARS-CoV-2. Cell. 2021;184(11):2939–54 e9. 13. 13.Acevedo ML, Alonso-Palomares L, Bustamante A, Gaggero A, Paredes F, Cortés CP, Valiente-Echeverría F, Soto-Rifo R. Infectivity and immune escape of the new SARS-CoV-2 variant of interest Lambda. medRxiv. 2021:2021.06.28.21259673. 14. 14.Barnes CO, Jette CA, Abernathy ME, Dam K-MA, Esswein SR, Gristick HB, Malyutin AG, Sharaf NG, Huey-Tubman KE, Lee YE. SARS-CoV-2 neutralizing antibody structures inform therapeutic strategies. Nature. 2020;588(7839):682–7. [PubMed](http://medrxiv.org/lookup/external-ref?access_num=33045718&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) 15. 15.McCallum M, De Marco A, Lempp FA, Tortorici MA, Pinto D, Walls AC, Beltramello M, Chen A, Liu Z, Zatta F. N-terminal domain antigenic mapping reveals a site of vulnerability for SARS-CoV-2. Cell. 2021;184(9):2332-47. e16. 16. 16.Avelino-Silva VI, Miyaji KT, Mathias A, Costa DA, de Carvalho Dias JZ, Lima SB, Simoes M, Freire MS, Caiaffa-Filho HH, Hong MA. CD4/CD8 ratio predicts yellow fever vaccine-induced antibody titers in virologically suppressed HIV-infected patients. JAIDS Journal of Acquired Immune Deficiency Syndromes. 2016;71(2):189–95. 17. 17.Ho DD, Neumann AU, Perelson AS, Chen W, Leonard JM, Markowitz M. Rapid turnover of plasma virions and CD4 lymphocytes in HIV-1 infection. Nature. 1995;373(6510):123–6. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1038/373123a0&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=7816094&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=A1995QB06300049&link_type=ISI) 18. 18.Murphy EL, Collier AC, Kalish LA, Assmann SF, Para MF, Flanigan TP, Kumar PN, Mintz L, Wallach FR, Nemo GJ. Highly active antiretroviral therapy decreases mortality and morbidity in patients with advanced HIV disease. Annals of internal medicine. 2001;135(1):17–26. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.7326/0003-4819-135-1-200107030-00005&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=11434728&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000169680500003&link_type=ISI) 19. 19.Hoffman SA, Costales C, Sahoo MK, Palanisamy S, Yamamoto F, Huang C, Verghese M, Solis DC, Sibai M, Subramanian A. SARS-CoV-2 Neutralization Resistance Mutations in Patient with HIV/AIDS, California, USA. Emerging Infectious Diseases. 2021;27(10). 20. 20.Karim F, Moosa MY, Gosnell B, Sandile C, Giandhari J, Pillay S, Tegally H, Wilkinson E, San EJ, Msomi N. Persistent SARS-CoV-2 infection and intra-host evolution in association with advanced HIV infection. medRxiv. 2021. 21. 21.Johnson BA, Xie X, Bailey AL, Kalveram B, Lokugamage KG, Muruato A, Zou J, Zhang X, Juelich T, Smith JK, Zhang L, Bopp N, Schindewolf C, Vu M, Vanderheiden A, Winkler ES, Swetnam D, Plante JA, Aguilar P, Plante KS, Popov V, Lee B, Weaver SC, Suthar MS, Routh AL, Ren P, Ku Z, An Z, Debbink K, Diamond MS, Shi P-Y, Freiberg AN, Menachery VD. Loss of furin cleavage site attenuates SARS-CoV-2 pathogenesis. Nature. 2021;591(7849):293–9. 22. 22.Karim F, Gazy I, Cele S, Zungu Y, Krause R, Bernstein M, Ganga Y, Rodel H, Mthabela N, Mazibuko M, Khan K, Muema D, Ramjit D, Ndung’u T, Hanekom W, Gosnell BI, Team C-K, Lessells R, Wong E, de Oliveira T, Moosa M-YS, Lustig G, Leslie A, Kløverpris H, Sigal A. HIV status alters disease severity and immune cell responses in *β* variant SARS-CoV-2 infection wave. medRxiv. 2021:2020.11.23.20236828. 23. 23.Minor PD, Ferguson M, Evans DM, Almond JW, Icenogle JP. Antigenic structure of polioviruses of serotypes 1, 2 and 3. Journal of General Virology. 1986;67(7):1283–91. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.1099/0022-1317-67-7-1283&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=2425046&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=A1986D294800006&link_type=ISI) 24. 24.Guzman MG, Alvarez M, Rodriguez-Roche R, Bernardo L, Montes T, Vazquez S, Morier L, Alvarez A, Gould EA, Kourí G. Neutralizing antibodies after infection with dengue 1 virus. Emerging infectious diseases. 2007;13(2):282. [CrossRef](http://medrxiv.org/lookup/external-ref?access_num=10.3201/eid1302.060539&link_type=DOI) [PubMed](http://medrxiv.org/lookup/external-ref?access_num=17479892&link_type=MED&atom=%2Fmedrxiv%2Fearly%2F2021%2F09%2F23%2F2021.09.14.21263564.atom) [Web of Science](http://medrxiv.org/lookup/external-ref?access_num=000244111000014&link_type=ISI)