Advertisement

Characterizing the Zika virus antibody response

Given the public health emergency that Zika virus poses, scientists are seeking to understand the Zika-specific immune response. Stettler et al. analyzed 119 monoclonal antibodies isolated from four donors that were infected with Zika virus during the present epidemic, including two individuals that had previously been infected with dengue virus, another member of the flavivirus family. Neutralizing antibodies primarily recognized the envelope protein domain III (EDIII) or quaternary epitopes on the intact virus, and an EDIII-targeted antibody protected mice against lethal infection. Some EDI/II-targeting antibodies cross-reacted with dengue virus in vitro and could enhance disease in dengue-infected mice. Whether dengue and Zika virus antibodies cross-react in humans remains to be tested.
Science, this issue p. 823

Abstract

Zika virus (ZIKV), a mosquito-borne flavivirus with homology to Dengue virus (DENV), has become a public health emergency. By characterizing memory lymphocytes from ZIKV-infected patients, we dissected ZIKV-specific and DENV–cross-reactive immune responses. Antibodies to nonstructural protein 1 (NS1) were largely ZIKV-specific and were used to develop a serological diagnostic tool. In contrast, antibodies against E protein domain I/II (EDI/II) were cross-reactive and, although poorly neutralizing, potently enhanced ZIKV and DENV infection in vitro and lethally enhanced DENV disease in mice. Memory T cells against NS1 or E proteins were poorly cross-reactive, even in donors preexposed to DENV. The most potent neutralizing antibodies were ZIKV-specific and targeted EDIII or quaternary epitopes on infectious virus. An EDIII-specific antibody protected mice from lethal ZIKV infection, illustrating the potential for antibody-based therapy.
After its introduction into Brazil in 2015, Zika virus (ZIKV) has spread rapidly, and in February 2016, the World Health Organization (WHO) declared it a Public Health Emergency of International Concern (13). The main route of ZIKV infection is through bites by Aedes mosquitos, but the virus may also be sexually (4) and vertically transmitted (5). Although most of the ZIKV infections are asymptomatic or cause only mild symptoms, there is evidence that ZIKV infection can lead to neurological complications, such as Guillain-Barré syndrome in adults (6) and congenital birth defects, including microcephaly in the developing fetus (5, 7, 8), likely through its ability to infect human neural progenitor cells (9).
Whereas flavivirus envelope (E) proteins mediate fusion and are the main target of neutralizing antibodies, the nonstructural protein 1 (NS1) is secreted by infected cells and is involved in immune evasion and pathogenesis (10). Two recent studies showed a high level of structural similarity between the E protein of ZIKV and that of other flaviviruses—such as dengue virus (DENV), yellow fever virus (YFV), and West Nile virus (WNV)—but also revealed distinct features that may be related to the ZIKV neurotropism (11, 12). Similarly, the structural analysis of ZIKV NS1 revealed conserved features with NS1 of other flaviviruses, although with different electrostatic characteristics (13).
A phenomenon that is characteristic of certain flaviviruses is the disease-enhancing activity of cross-reactive antibodies elicited by previous infections by heterologous viruses, termed antibody-dependent enhancement (ADE). In the case of DENV, for which four serotypes are known, there is epidemiological evidence that a primary infection protects from reinfection with the same serotype but represents a risk factor for the development of severe disease upon reinfection with a different serotype (14). The exacerbated disease is triggered by E- and prM-specific antibodies that fail to neutralize the incoming virus but instead enhance its capture by Fc receptor–expressing (FcR+) cells, leading to enhanced viral replication and activation of cross-reactive memory T cells. The resulting cytokine storm is thought to be the basis of the most severe form of disease, known as dengue hemorragic fever/dengue shock syndrome (15, 16). The role of antibodies in severe dengue is supported by studies showing that waning levels of maternal antibodies in infants represent a higher risk for development of severe dengue disease (1619). Whether individuals with antibodies induced by previous DENV infections can develop a more severe ZIKV infection or have higher risk of fetal transmission is unknown. Similarly, it is unclear whether ZIKV antibodies may affect subsequent DENV infection. Therefore, it is important to dissect the level of cross-reactive immunity at the B cell and T cell level in response to ZIKV infection.
To understand the role of antibodies in ZIKV neutralization and heterologous enhancement of flavivirus infection, we isolated a panel of 119 monoclonal antibodies (mAbs) from four ZIKV-infected donors from the current epidemic, of which two were DENV-naïve (ZA and ZD) and two had serological records of DENV infection (ZB and ZC) (table S1) (20). These mAbs were selected from Epstein-Barr virus–immortalized memory B cells on the basis of their binding to ZIKV NS1 or E proteins and for their ability to neutralize ZIKV infection, and were compared, side by side, with a panel of mAbs previously isolated from DENV-infected donors (21). The mAbs were primarily immunoglobulin G (IgG), were highly polyclonal, and carried a lower level of somatic mutations as compared with that of other acute or recurrent infections (fig. S1 and table S2).
Out of the 119 ZIKV-reactive mAbs isolated, 41 bound to NS1. ZIKV and DENV NS1 share 51 to 53% of amino acid identity (Fig. 1, A and B, and fig. S2). The mAbs isolated from ZIKV-infected, DENV-naïve donors were to a large extent ZIKV-specific and did not cross-react to DENV NS1, whereas those isolated from ZIKV-infected, DENV-immune donors showed, as expected, a higher degree of cross-reactivity (Fig. 1C, fig. S3, and table S3). The limited cross-reactivity of NS1-reactive mAbs from ZIKV-infected, DENV-naïve donors was corroborated by the lack of reactivity to DENV1-4 NS1 proteins of plasma from ZIKV-infected, DENV-naïve donors (fig. S4). Conversely, the plasma of the ZIKV-infected, DENV-immune donors reacted strongly to DENV1-4 NS1 proteins, which is consistent with a recall response to NS1 dominated by cross-reactive antibodies. Reciprocally, NS1-reactive mAbs isolated from DENV-immune, ZIKV-naïve donors showed variable patterns of cross-reactivity between DENV serotypes but, with the exception of a single mAb, did not cross-react with ZIKV NS1 (Fig. 1D, fig. S5, and table S4).
Fig. 1 Specificity and cross-reactivity of NS1-reactive mAbs and T cells derived from ZIKV- and DENV-infected donors.
(A) Sequence conservation of NS1 proteins as determined by the alignments shown in fig. S2. Conserved residues in NS1 are in purple. (B) Structure of the ZIKV NS1 dimer [Protein Data Bank (PDB) 5IY3]. The two monomers are shown with different shades of gray. The surface residues of each protomer that are conserved between ZIKV and DENV are colored as in (A). (C and D) Heat map of the reactivity of (C) 41 mAbs derived from four ZIKV-infected donors (ZA-ZD) and (D) 12 mAbs derived from two DENV-infected donors. The mAbs were tested for binding to NS1 proteins of ZIKV and DENV1-4 [median effective concentration (EC50), nanograms per milliliter]. A description of the immune status of the donors and the mAbs codes is provided in table S1. The binding curves of NS1-reactive mAbs are shown in figs. S3 and S5, and EC50 values are reported in tables S3 and S4. Data are representative of at least two independent experiments. (E) Plasma from ZIKV-infected (n = 4 donors), DENV-infected (n = 5 donors), and control donors (n = 48 donors) (1/10 dilution) were tested for their capacity to bind NS1 (blue dots) and to inhibit the binding of the biotinylated mAb ZKA35 to NS1 (red dots). (F and G) T cell libraries were generated from CXCR3+ and CXCR3 memory CD4+ T cells of the four ZIKV-infected donors and screened for reactivity against ZIKV or DENV1-4 NS1 proteins by use of 3H-thymidine incorporation. Shown is the estimated number of ZIKV NS1–specific T cells per 106 cells [mean + SEM and color-coded values for individual donors (F)]. Shown is also the proliferation [counts per minute (cpm)] of individual cultures to either ZIKV or DENV1-4 NS1 proteins (G). The cultures from different donors are color-coded, and their numbers are shown in parentheses. Dotted lines represent the cut-off value.
Given the interest in developing a diagnostic tool that would discriminate ZIKV-specific from DENV-specific antibodies, we identified antigenic sites (figs. S1 and S2) on NS1 targeted by ZIKV-specific but not by cross-reactive mAbs (fig. S6, A to D). MAbs reacting to two distinct sites on ZIKV NS1 were used in a two-determinant immunoassay so as to detect circulating NS1 protein in body fluids during acute infection (fig. S6E). In addition, mAb ZKA35 binding to the antigenic site S2 was used as a probe in a blockade-of-binding assay (22) so as to detect ZIKV-specific, but not DENV-specific, serum antibodies to NS1 (Fig. 1E) and has a potential to be developed in a serological assay in order to detect clinical and subclinical ZIKV infections at the population level.
Given the role of T cells in antibody production and in immunopathology, we also investigated the specificity and cross-reactivity of CD4+ memory T cells from the same donors using the T cell library method (Fig. 1, F and G) (23). NS1-specific memory CD4+ T cells were present in the CXCR3+ T helper 1 (TH1) compartment and, with variable frequencies, in the CXCR3 TH compartment. With a few exceptions, these TH cells were specific for either ZIKV or DENV NS1 (Fig. 1, F and G, and fig. S7A), which is consistent with a low level of T cell cross-reactivity, even in subjects who were already exposed to DENV.
The E protein is formed by three domains: EDI, which is involved in the conformational changes required for viral entry; EDII, which contains the fusion loop; and EDIII, which may be involved in binding to cellular receptors (24). DI, DII, and DIII of ZIKV and DENV share 35, 51, and 29% amino acid identity, respectively (Fig. 2A). According to the structure of the ZIKV E protein dimer (11, 12), the majority of the ZIKV- and DENV-conserved solvent-accessible residues are located in EDII, particularly in the fusion loop and the neighboring region (Fig. 2B and fig. S8). Strikingly, the majority of ZIKV EDI/II–specific mAbs isolated from the four ZIKV donors (65%; 24 of 37) cross-reacted with the E protein of all four DENV serotypes (Fig. 2C, fig. S9, and table S5). Similarly, a large proportion (67%; 31 of 46) of DENV EDI/II–reactive mAbs isolated from DENV donors were also cross-reactive with ZIKV E protein (Fig. 2D, fig. S10, and table S6). These data indicate that the human antibody response to the E protein and in particular to EDI/II is cross-reactive between ZIKV and DENV, which is consistent with previous reports on mAbs from mice immunized with flavivirus antigens (11, 25). ZIKV-immune plasma was also found to cross-react with ZIKV and DENV1-4 E proteins (fig. S11), possibly because of the abundant EDI/II cross-reactive antibody response (21, 2628).
Fig. 2 Specificity and cross-reactivity of E-reactive mAbs and T cells derived from ZIKV- and DENV-infected donors.
(A) Sequence conservation of E proteins as determined by the alignments of ZIKV isolates and DENV1-4 reference strains, as shown in fig. S8. Conserved residues in EDI, EDII, EDIII, and the hinge regions of the E protein are colored in red, yellow, blue, and green, respectively. FL, fusion loop. (B) Structure of the ZIKV E protein dimer (PDB 5JHM). The two monomers are shown with different shades of gray, and the residues conserved between ZIKV and DENV are colored as in (A). (C and D) Heat map of the reactivity of (C) 78 mAbs derived from four ZIKV donors and (D) 61 mAbs derived from nine DENV donors. NNB, neutralizing non-E-binding mAbs. The mAbs were tested for binding to E-recombinant proteins of ZIKV and DENV1-4 (EC50, nanograms per milliliter). Also shown is the virus-neutralizing activity (IC50, nanograms per milliliter). Strikethrough cells indicate not tested. Binding curves of E-reactive mAbs are shown in figs. S9 and S10. The EC50 and IC50 values of E-reactive and NNB mAbs are shown in tables S5 and S6. Data are representative of at least two independent experiments. (E and F) T cell libraries were generated from CXCR3+ and CXCR3 memory CD4+ T cells of the four ZIKV-infected donors and screened for reactivity against ZIKV or DENV1-4 E proteins by using 3H-thymidine incorporation. Shown is the estimated number of ZIKV E–specific T cells per 106 cells [mean + SEM and color-coded values for individual donors (E)]. Also shown is the proliferation (cpm) of individual cultures to either ZIKV E or DENV1-4 E proteins (F). The cultures from different donors are color-coded, and their numbers are shown in parentheses. Dotted lines represent the cut-off value. (G) CD4+ CXCR3+ T cells from peripheral blood mononuclear cells of a ZIKV-infected (ZB) or DENV-infected (DENV013) donor were labeled with carboxyfluorescein succinimidyl ester (CFSE) and stimulated with ZIKV or DENV1-4 E proteins, respectively. Proliferating T cells were isolated by means of cell sorting, relabeled with CFSE, and stimulated with the indicated proteins. Shown are the CFSE profiles on day 5.
In contrast, most of the EDIII-reactive mAbs (90%; 27 of 30) isolated from ZIKV or DENV donors were specific for either ZIKV or DENV E protein (Fig. 2, C and D; figs. S9 and S10; and tables S5 and S6). In addition, whereas the EDI/II mAbs showed only modest or even no neutralizing activity against ZIKV infection, the EDIII-reactive antibodies were overall more highly neutralizing than EDI/II–reactive antibodies (table S5). By screening for ZIKV neutralization, we also isolated several highly potent neutralizing antibodies that failed to bind to ZIKV E and EDIII proteins (Fig. 2C). These mAbs, that we define as neutralizing non-E-binding (NNB), could recognize quaternary epitopes that are displayed on the infectious virions but not on soluble proteins, as recently demonstrated for DENV (2931). E-specific CD4+ T cells were primarily detected in the CXCR3+ TH cell subset and, with only a few exceptions, were specific for ZIKV E protein (Fig. 2, E and F, and fig. S7B). Furthermore, E-reactive T cells directly isolated from ex vivo cultures of memory CD4+ T cells of ZIKV- or DENV-infected donors (fig. S12) did not cross-react upon secondary stimulation with heterologous E proteins (Fig. 2G). These findings indicate an overall low level of T cell cross-reactivity between ZIKV and DENV E proteins.
To address the biological properties of ZIKV and DENV antibodies, we used a selected mAb panel and measured in parallel their binding (Fig. 3A), neutralizing, and ADE activity (Fig. 3, B and C). This panel also contained Fc mutant versions of mAbs that do not bind to FcγR and complement [LALA mutants; leucine (L) to alanine (A) substitution at the position 234 and 235] (21, 32). The EDIII-specific mAbs ZKA64 and ZKA190 and the NNB mAb ZKA230 were highly potent in ZIKV neutralization, with median inhibitory concentration (IC50) values of 93, 9, and 10 ng/ml, respectively. Furthermore, all these mAbs enhanced infection of ZIKV in the nonpermissive K562 human erythroleukemic cell line at a broad range of concentrations, including those that fully neutralized ZIKV infection on Vero cell line. Whereas EDIII mAbs ZKA64 and ZKA190 did not enhance ZIKV infections of K562 cells above 1 μg/ml, the NNB mAb ZKA230 failed to do so, a result that might be due to the different mechanisms of neutralization of free viruses versus FcγR-internalized viruses. In contrast, the cross-reactive EDI/II–specific mAbs ZKA3 and ZKA78 only partially neutralized ZIKV infectivity, while effectively enhancing infection of K562 cells. Consistent with their cross-reactivity, these EDI/II–specific mAbs also neutralized and enhanced DENV1 infection.
Fig. 3 Neutralization and enhancement of ZIKV and DENV infection by mAbs and immune plasma.
(A) Binding of mAbs isolated from ZIKV-infected donors (blue) or from DENV-infected donors (red) to (left and middle) recombinant ZIKV E and (right) to DENV1 E protein. Open and solid symbols indicate mAbs specific or cross-reactive for ZIKV and DENV, respectively. (B) Neutralization of ZIKV and DENV1 as determined by the percentage of infected Vero cells. (C) ADE of ZIKV and DENV1 infection of nonpermissive K562 cells, as determined by the percentage of infected cells on day 5. (D and E) Inhibition of ADE by LALA mutant mAbs. Serial dilutions of ZIKV- and DENV3-immune plasma were incubated with ZIKV or DENV1 before addition to K562 cells in the absence or presence of a fixed amount (50 μg/ml) of ZKA64-LALA or DV82-LALA mAb. Shown is the percentage of infected K562 cells on day 4. Blue and red symbols indicate plasma derived from ZIKV- or DENV-infected donors, respectively. Data are representative of two independent experiments.
The above results suggest that cross-reactive antibodies elicited by either ZIKV or DENV infection and primarily directed to EDI/II can mediate heterologous ADE. We therefore asked whether ADE could be also induced with immune sera and whether this could be blocked by neutralizing mAbs delivered in a LALA format (Fig. 3, D and E). In a homologous setting, four ZIKV-immune plasma collected from convalescent patients showed similar capacity to enhance ZIKV infection of K562 cells, and this ADE effect was completely blocked by the EDIII-specific ZKA64-LALA mAb, but only partially inhibited by the cross-reactive EDI/II DV82-LALA mAb (Fig. 3D). In a heterologous setting, the four ZIKV-immune plasma strongly enhanced infection by DENV1 to levels comparable with those observed with DENV3 plasma. The enhancement of DENV1 infection by both ZIKV and DENV3 plasma was completely inhibited by DV82-LALA, but not by ZKA64-LALA because of its lack of cross-reactivity to DENV1 (Fig. 3E). The ADE-blocking ability of a single EDIII-specific LALA mAb could be related not only to its capacity to out-compete serum-enhancing antibodies but also to neutralize virus once internalized into endosomes, similarly to what has been reported for the WNV-neutralizing antibody E16 (33). Conversely, the lack of ADE-blocking ability of DV82-LALA might be explained by the inability of antibodies to EDI/II to effectively neutralize endocytosed virus.
The above results suggest that previous ZIKV and DENV immunity may pose a risk of a more severe disease upon exposure to heterologous virus through ADE. To address this possibility in vivo, we selected two EDI/II cross-reactive mAbs elicited by ZIKV (ZKA78) or DENV (DV82) and tested them for their capacity to enhance infection by DENV or ZIKV in animal models. Preadministration of ZKA78 or DV82, but not their LALA versions, to DENV2-infected AG129 mice led to severe symptoms and lethality by day 5 (Fig. 4A), suggesting that ZIKV immunity could predispose to enhanced DENV pathology in vivo. Given the high lethality of ZIKV infection in immunodeficient mice (3436), we investigated a possible ZIKV disease–enhancing activity of the DENV cross-reactive mAb DV82 in 129Sv/ev immunocompetent mice that are not permissive for productive ZIKV infection. However, in this mouse model we did not observe signs of enhanced disease or infection (fig. S13).
Fig. 4 In vivo enhancement of DENV2 infection by an anti-ZIKV cross-reactive mAb and ZIKV therapeutic efficacy of a potent anti-ZIKV EDIII-specific mAb.
(A) In vivo ADE. MAbs were administered intraperitoneally to AG129 mice 20 to 24 hours before intravenous inoculation with 2.5 × 105 plaque-forming units (PFU) of DENV2 D2S10. Results are representative of two independent experiments, with n = 8 mice for the ZKA78 1 and 2-μg groups; n = 6 for the DV82 and all LALA mAb groups; and n = 4 for phosphate-buffered saline (PBS) group. (Left) A Kaplan-Meier survival curve is shown; significance was determined by using the Mantel-Cox log-rank test. ZKA78 versus PBS, P = 0.0008; ZKA78 versus ZKA78-LALA, P = 0.0081; DV82 versus PBS, P = 0.0187; and DV82 versus DV82-LALA, P = 0.0255. (Right) Mice were monitored over a 10-day period and scored for morbidity and mortality by using a standardized 5-point system (38). (B) Prophylaxis and therapy of ZIKV infection. MAbs (15 mg/kg) were administered intraperitoneally to A129 mice (n = 6 mice) 24 hours before or after subcutaneous inoculation with 102 PFU of ZIKV MP1751. (Top left) A Kaplan-Meier survival curve is shown. (Top right) Mice were monitored over a 13-day period for body weight loss. (Bottom) Viral loads were measured on day 5 in blood of all animals and in blood and indicated tissues when animals were culled at the end of the study or when the humane end points were met. Lines indicate geometric means.
To develop a therapeutic approach to ZIKV infection, we tested the LALA form of the potently neutralizing EDIII-specific mAb ZKA64 in prophylactic and therapeutic settings. ZKA64-LALA completely protected A129 mice challenged with a lethal dose of ZIKV from body weight loss and lethality when given 1 day before or 1 day after ZIKV challenge (Fig. 4B).
This study reports the first characterization of the human immune response to ZIKV infection. The highly cross-reactive antibodies to EDI/II elicited by ZIKV or DENV infection that are poorly neutralizing but potently enhancing may pose a risk for heterologous ADE. We have shown in vitro and in a relevant animal model that an EDI/II cross-reactive mAb raised by ZIKV can induce lethal DENV infection. However, enhancement of ZIKV infection by DENV-elicited cross-reactive antibodies could be observed in vitro but not in vivo, a finding that may be due to the tropism of the virus or to the limitation of the mouse model. It will be important to address this issue in clinical and epidemiological studies, which may be facilitated by the development of serological diagnostics, such as the blockade-of-binding assay described in this study.
T cells in flavivirus infections play a complex role in both protection and pathogenesis (15). The low degree of CD4+ T cell cross-reactivity between DENV and ZIKV, which we observed even in individuals who are immune to both viruses, suggests that in ZIKV infection original antigenic sin may not play a pathogenic role. Our findings also suggest that the risk of cytokine storm and consequent severe disease after enhanced heterologous infection of DENV by antibodies to ZIKV may be mitigated by the poor T cell cross-reactivity.
Our study describes two classes of potent neutralizing antibodies that are specific for ZIKV and directed either against EDIII or quaternary epitopes present on infectious virus, the latter being particularly frequent, similar to what has been described in DENV (30). Given the high potency and in vivo efficacy shown in this study, these antibodies, developed in wild type or their LALA versions produced in wild-type version or in LALA version so as to avoid possible enhancement, could be used in prophylactic or therapeutic settings to prevent congenital ZIKV infection in pregnant women living in high-risk areas (37).

Acknowledgments

The data presented in this manuscript are tabulated in the main paper and in the supplementary materials. The views expressed are those of the authors and not necessarily those of the funding bodies. This work was supported by grants from the European Union (HEALTH-F3-2011-281803-Project IDAMS, “International Research Consortium on Dengue Risk Assessment, Management, and Surveillance” to F.S.), the European Research Council (grant 323183 PREDICT to F.S.), the Swiss National Science Foundation (grant 160279 to A.L.), the U.S. National Institutes of Health (grant R01 AI24493 to E.H.), and the Ministero della Salute, Fondazione IRCCS Policlinico San Matteo, Ricerca Corrente grant 8020615 (to E.P. and F.B.). A.L. is supported by the Helmut Horten Foundation. A.L. is the scientific founder of Humabs BioMed. D.C., A.L., and F.S. hold shares in Humabs BioMed. F.S. and F.B. are a scientific advisor and a consultant at Humabs BioMed, respectively. The MAbs described in this study are available under a materials transfer agreement with Humabs BioMed SA. L.V. is grateful for financial support by SNF grant 310030_166445 and Lions Club Monteceneri. A.L. is an inventor on a patent related to DENV mAbs described in this study: PCT/lB2009/007372 (Dengue virus neutralizing antibodies and uses thereof). D.C. is an inventor on a patent application held by Humabs BioMed SA related to ZIKV mAbs described in this study: PCT/EP2016/066684 (Novel antibodies specifically binding to Zika virus epitopes and uses thereof). Antibody nucleotide and amino acid sequences for the 68 mAbs described in table S2 have been deposited in GenBank (accession nos. KX496807 to KX496874).

Supplementary Material

Summary

Figs. S1 to S13
Tables S1 to S6
References (3952)

Resources

File (stettler-sm.pdf)

References and Notes

1
Petersen L. R., Jamieson D. J., Powers A. M., and Honein M. A., Zika virus. N. Engl. J. Med. 374, 1552–1563 (2016).
2
Fauci A. S. and Morens D. M., Zika virus in the Americas—Yet another arbovirus threat. N. Engl. J. Med. 374, 601–604(2016).
3
Heymann D. L., Hodgson A., Sall A. A., Freedman D. O., Staples J. E., Althabe F., Baruah K., Mahmud G., Kandun N., Vasconcelos P. F., Bino S., and Menon K. U., Zika virus and microcephaly: Why is this situation a PHEIC? Lancet 387, 719–721 (2016).
4
Musso D., Roche C., Robin E., Nhan T., Teissier A., and Cao-Lormeau V. M., Potential sexual transmission of Zika virus. Emerg. Infect. Dis. 21, 359–361 (2015).
5
Mlakar J., Korva M., Tul N., Popović M., Poljšak-Prijatelj M., Mraz J., Kolenc M., Resman Rus K., Vesnaver Vipotnik T., Fabjan Vodušek V., Vizjak A., Pižem J., Petrovec M., and Avšič Županc T., Zika virus associated with microcephaly. N. Engl. J. Med. 374, 951–958 (2016).
6
Cao-Lormeau V.-M., Blake A., Mons S., Lastère S., Roche C., Vanhomwegen J., Dub T., Baudouin L., Teissier A., Larre P., Vial A. L., Decam C., Choumet V., Halstead S. K., Willison H. J., Musset L., Manuguerra J. C., Despres P., Fournier E., Mallet H. P., Musso D., Fontanet A., Neil J., and Ghawché F., Guillain-Barré Syndrome outbreak associated with Zika virus infection in French Polynesia: A case-control study. Lancet 387, 1531–1539 (2016).
7
Rubin E. J., Greene M. F., and Baden L. R., Zika virus and microcephaly. N. Engl. J. Med. 374, 984–985 (2016).
8
Calvet G., Aguiar R. S., Melo A. S., Sampaio S. A., de Filippis I., Fabri A., Araujo E. S., de Sequeira P. C., de Mendonça M. C., de Oliveira L., Tschoeke D. A., Schrago C. G., Thompson F. L., Brasil P., Dos Santos F. B., Nogueira R. M., Tanuri A., and de Filippis A. M., Detection and sequencing of Zika virus from amniotic fluid of fetuses with microcephaly in Brazil: A case study. Lancet Infect. Dis. 16, 653–660 (2016).
9
Tang H., Hammack C., Ogden S. C., Wen Z., Qian X., Li Y., Yao B., Shin J., Zhang F., Lee E. M., Christian K. M., Didier R. A., Jin P., Song H., and Ming G. L., Zika virus infects human cortical neural progenitors and attenuates their growth. Cell Stem Cell 18, 587–590 (2016).
10
Muller D. A. and Young P. R., The flavivirus NS1 protein: Molecular and structural biology, immunology, role in pathogenesis and application as a diagnostic biomarker. Antiviral Res. 98, 192–208 (2013).
11
Dai L., Song J., Lu X., Deng Y. Q., Musyoki A. M., Cheng H., Zhang Y., Yuan Y., Song H., Haywood J., Xiao H., Yan J., Shi Y., Qin C. F., Qi J., and Gao G. F., Structures of the Zika Virus Envelope Protein and Its Complex with a Flavivirus Broadly Protective Antibody. Cell Host Microbe 19, 696–704 (2016).
12
Sirohi D., Chen Z., Sun L., Klose T., Pierson T. C., Rossmann M. G., and Kuhn R. J., The 3.8 Å resolution cryo-EM structure of Zika virus. Science 352, 467–470 (2016).
13
Song H., Qi J., Haywood J., Shi Y., and Gao G. F., Zika virus NS1 structure reveals diversity of electrostatic surfaces among flaviviruses. Nat. Struct. Mol. Biol. 23, 456–458 (2016).
14
Halstead S. B., Dengue antibody-dependent enhancement: Knowns and unknowns. Microbiol. Spectr. 2, 249–271 (2014).
15
Screaton G., Mongkolsapaya J., Yacoub S., and Roberts C., New insights into the immunopathology and control of dengue virus infection. Nat. Rev. Immunol. 15, 745–759 (2015).
16
Halstead S. B., Neutralization and antibody-dependent enhancement of dengue viruses. Adv. Virus Res. 60, 421–467 (2003).
17
Halstead S. B., Lan N. T., Myint T. T., Shwe T. N., Nisalak A., Kalyanarooj S., Nimmannitya S., Soegijanto S., Vaughn D. W., and Endy T. P., Dengue hemorrhagic fever in infants: Research opportunities ignored. Emerg. Infect. Dis. 8, 1474–1479 (2002).
18
Nguyen T. H., Lei H. Y., Nguyen T. L., Lin Y. S., Huang K. J., Le B. L., Lin C. F., Yeh T. M., Do Q. H., Vu T. Q., Chen L. C., Huang J. H., Lam T. M., Liu C. C., and Halstead S. B., Dengue hemorrhagic fever in infants: A study of clinical and cytokine profiles. J. Infect. Dis. 189, 221–232 (2004).
19
Rothman A. L., Dengue: Defining protective versus pathologic immunity. J. Clin. Invest. 113, 946–951 (2004).
20
Ginier M., Neumayr A., Günther S., Schmidt-Chanasit J., and Blum J., Zika without symptoms in returning travellers: What are the implications? Travel Med. Infect. Dis. 14, 16–20 (2016).
21
Beltramello M., Williams K. L., Simmons C. P., Macagno A., Simonelli L., Quyen N. T., Sukupolvi-Petty S., Navarro-Sanchez E., Young P. R., de Silva A. M., Rey F. A., Varani L., Whitehead S. S., Diamond M. S., Harris E., Lanzavecchia A., and Sallusto F., The human immune response to Dengue virus is dominated by highly cross-reactive antibodies endowed with neutralizing and enhancing activity. Cell Host Microbe 8, 271–283 (2010).
22
Ohana B., Lipson M., Vered N., Srugo I., Ahdut M., and Morag A., Novel approach for specific detection of herpes simplex virus type 1 and 2 antibodies and immunoglobulin G and M antibodies. Clin. Diagn. Lab. Immunol. 7, 904–908 (2000).
23
Geiger R., Duhen T., Lanzavecchia A., and Sallusto F., Human naive and memory CD4+ T cell repertoires specific for naturally processed antigens analyzed using libraries of amplified T cells. J. Exp. Med. 206, 1525–1534 (2009).
24
Modis Y., Ogata S., Clements D., and Harrison S. C., A ligand-binding pocket in the dengue virus envelope glycoprotein. Proc. Natl. Acad. Sci. U.S.A. 100, 6986–6991 (2003).
25
Monath T. P., Schlesinger J. J., Brandriss M. W., Cropp C. B., and Prange W. C., Yellow fever monoclonal antibodies: Type-specific and cross-reactive determinants identified by immunofluorescence. Am. J. Trop. Med. Hyg. 33, 695–698 (1984).
26
Smith S. A., Zhou Y., Olivarez N. P., Broadwater A. H., de Silva A. M., and Crowe J. E. Jr., Persistence of circulating memory B cell clones with potential for dengue virus disease enhancement for decades following infection. J. Virol. 86, 2665–2675 (2012).
27
Lai C.-Y., Tsai W. Y., Lin S. R., Kao C. L., Hu H. P., King C. C., Wu H. C., Chang G. J., and Wang W. K., Antibodies to envelope glycoprotein of dengue virus during the natural course of infection are predominantly cross-reactive and recognize epitopes containing highly conserved residues at the fusion loop of domain II. J. Virol. 82, 6631–6643 (2008).
28
Lin H.-E., Tsai W. Y., Liu I. J., Li P. C., Liao M. Y., Tsai J. J., Wu Y. C., Lai C. Y., Lu C. H., Huang J. H., Chang G. J., Wu H. C., and Wang W. K., Analysis of epitopes on dengue virus envelope protein recognized by monoclonal antibodies and polyclonal human sera by a high throughput assay. PLOS Negl. Trop. Dis. 6, e1447 (2012).
29
Teoh E. P., Kukkaro P., Teo E. W., Lim A. P., Tan T. T., Yip A., Schul W., Aung M., Kostyuchenko V. A., Leo Y. S., Chan S. H., Smith K. G., Chan A. H., Zou G., Ooi E. E., Kemeny D. M., Tan G. K., Ng J. K., Ng M. L., Alonso S., Fisher D., Shi P. Y., Hanson B. J., Lok S. M., and MacAry P. A., The structural basis for serotype-specific neutralization of dengue virus by a human antibody. Sci. Transl. Med. 4, 139ra83 (2012).
30
Dejnirattisai W., Wongwiwat W., Supasa S., Zhang X., Dai X., Rouvinski A., Jumnainsong A., Edwards C., Quyen N. T. H., Duangchinda T., Grimes J. M., Tsai W.-Y., Lai C.-Y., Wang W.-K., Malasit P., Farrar J., Simmons C. P., Zhou Z. H., Rey F. A., Mongkolsapaya J., and Screaton G. R., A new class of highly potent, broadly neutralizing antibodies isolated from viremic patients infected with dengue virus. Nat. Immunol. 16, 170–177 (2015).
31
Rouvinski A., Guardado-Calvo P., Barba-Spaeth G., Duquerroy S., Vaney M. C., Kikuti C. M., Navarro Sanchez M. E., Dejnirattisai W., Wongwiwat W., Haouz A., Girard-Blanc C., Petres S., Shepard W. E., Desprès P., Arenzana-Seisdedos F., Dussart P., Mongkolsapaya J., Screaton G. R., and Rey F. A., Recognition determinants of broadly neutralizing human antibodies against dengue viruses. Nature 520, 109–113 (2015).
32
Hessell A. J., Hangartner L., Hunter M., Havenith C. E., Beurskens F. J., Bakker J. M., Lanigan C. M., Landucci G., Forthal D. N., Parren P. W., Marx P. A., and Burton D. R., Fc receptor but not complement binding is important in antibody protection against HIV. Nature 449, 101–104 (2007).
33
Thompson B. S., Moesker B., Smit J. M., Wilschut J., Diamond M. S., and Fremont D. H., A therapeutic antibody against west nile virus neutralizes infection by blocking fusion within endosomes. PLOS Pathog. 5, e1000453 (2009).
34
Dowall S. D., Graham V. A., Rayner E., Atkinson B., Hall G., Watson R. J., Bosworth A., Bonney L. C., Kitchen S., and Hewson R., A susceptible mouse model for Zika virus infection. PLOS Negl. Trop. Dis. 10, e0004658–e13 (2016).
35
Lazear H. M., Govero J., Smith A. M., Platt D. J., Fernandez E., Miner J. J., and Diamond M. S., A mouse model of Zika virus pathogenesis. Cell Host Microbe 19, 720–730 (2016).
36
Aliota M. T., Caine E. A., Walker E. C., Larkin K. E., Camacho E., and Osorio J. E., Characterization of lethal Zika virus infection in AG129 mice. PLOS Negl. Trop. Dis. 10, e0004682–e11 (2016).
37
Johansson M. A., Mier-Y-Teran-Romero L., Reefhuis J., Gilboa S. M., and Hills S. L., Zika and the risk of microcephaly. N. Engl. J. Med. 375, 1–4 (2016).
38
Orozco S., Schmid M. A., Parameswaran P., Lachica R., Henn M. R., Beatty R., and Harris E., Characterization of a model of lethal dengue virus 2 infection in C57BL/6 mice deficient in the alpha/beta interferon receptor. J. Gen. Virol. 93, 2152–2157 (2012).
39
Whitehead S. S., Falgout B., Hanley K. A., Blaney J. E. Jr., Markoff L., and Murphy B. R., A live, attenuated dengue virus type 1 vaccine candidate with a 30-nucleotide deletion in the 3′ untranslated region is highly attenuated and immunogenic in monkeys. J. Virol. 77, 1653–1657 (2003).
40
Whitehead S. S., Hanley K. A., Blaney J. E. Jr., Gilmore L. E., Elkins W. R., and Murphy B. R., Substitution of the structural genes of dengue virus type 4 with those of type 2 results in chimeric vaccine candidates which are attenuated for mosquitoes, mice, and rhesus monkeys. Vaccine 21, 4307–4316 (2003).
41
Blaney J. E. Jr., Hanson C. T., Firestone C. Y., Hanley K. A., Murphy B. R., and Whitehead S. S., Genetically modified, live attenuated dengue virus type 3 vaccine candidates. Am. J. Trop. Med. Hyg. 71, 811–821 (2004).
42
Durbin A. P., Whitehead S. S., McArthur J., Perreault J. R., Blaney J. E. Jr., Thumar B., Murphy B. R., and Karron R. A., rDEN4delta30, a live attenuated dengue virus type 4 vaccine candidate, is safe, immunogenic, and highly infectious in healthy adult volunteers. J. Infect. Dis. 191, 710–718 (2005).
43
Traggiai E., Becker S., Subbarao K., Kolesnikova L., Uematsu Y., Gismondo M. R., Murphy B. R., Rappuoli R., and Lanzavecchia A., An efficient method to make human monoclonal antibodies from memory B cells: Potent neutralization of SARS coronavirus. Nat. Med. 10, 871–875 (2004).
44
Pappas L., Foglierini M., Piccoli L., Kallewaard N. L., Turrini F., Silacci C., Fernandez-Rodriguez B., Agatic G., Giacchetto-Sasselli I., Pellicciotta G., Sallusto F., Zhu Q., Vicenzi E., Corti D., and Lanzavecchia A., Rapid development of broadly influenza neutralizing antibodies through redundant mutations. Nature 516, 418–422 (2014).
45
Brochet X., Lefranc M.-P., and Giudicelli V., IMGT/V-QUEST: The highly customized and integrated system for IG and TR standardized V-J and V-D-J sequence analysis. Nucleic Acids Res. 36 (web server), W503 (2008).
46
Arnaout R., Lee W., Cahill P., Honan T., Sparrow T., Weiand M., Nusbaum C., Rajewsky K., and Koralov S. B., High-resolution description of antibody heavy-chain repertoires in humans. PLOS ONE 6, e22365 (2011).
47
Tiller T., Meffre E., Yurasov S., Tsuiji M., Nussenzweig M. C., and Wardemann H., Efficient generation of monoclonal antibodies from single human B cells by single cell RT-PCR and expression vector cloning. J. Immunol. Methods 329, 112–124 (2008).
48
Sievers F., Wilm A., Dineen D., Gibson T. J., Karplus K., Li W., Lopez R., McWilliam H., Remmert M., Söding J., Thompson J. D., and Higgins D. G., Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539–539 (2011).
49
Kumar S., Stecher G., and Tamura K., MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).
50
van den Broek M. F., Müller U., Huang S., Aguet M., and Zinkernagel R. M., Antiviral defense in mice lacking both alpha/beta and gamma interferon receptors. J. Virol. 69, 4792–4796 (1995).
51
Shresta S., Sharar K. L., Prigozhin D. M., Snider H. M., Beatty P. R., and Harris E., Critical roles for both STAT1-dependent and STAT1-independent pathways in the control of primary dengue virus infection in mice. J. Immunol. 175, 3946–3954 (2005).
52
Lanciotti R. S., Kosoy O. L., Laven J. J., Velez J. O., Lambert A. J., Johnson A. J., Stanfield S. M., and Duffy M. R., Genetic and serologic properties of Zika virus associated with an epidemic, Yap State, Micronesia, 2007. Emerg. Infect. Dis. 14, 1232–1239 (2008).

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Embedded figures cannot be submitted, and we discourage the use of figures within eLetters in general. If a figure is essential, please include a link to the figure within the text of the eLetter. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 353 | Issue 6301
19 August 2016

Article versions

You are viewing the most recent version of this article.

Submission history

Received: 10 April 2016
Accepted: 5 July 2016
Published in print: 19 August 2016

Permissions

Request permissions for this article.

Acknowledgments

The data presented in this manuscript are tabulated in the main paper and in the supplementary materials. The views expressed are those of the authors and not necessarily those of the funding bodies. This work was supported by grants from the European Union (HEALTH-F3-2011-281803-Project IDAMS, “International Research Consortium on Dengue Risk Assessment, Management, and Surveillance” to F.S.), the European Research Council (grant 323183 PREDICT to F.S.), the Swiss National Science Foundation (grant 160279 to A.L.), the U.S. National Institutes of Health (grant R01 AI24493 to E.H.), and the Ministero della Salute, Fondazione IRCCS Policlinico San Matteo, Ricerca Corrente grant 8020615 (to E.P. and F.B.). A.L. is supported by the Helmut Horten Foundation. A.L. is the scientific founder of Humabs BioMed. D.C., A.L., and F.S. hold shares in Humabs BioMed. F.S. and F.B. are a scientific advisor and a consultant at Humabs BioMed, respectively. The MAbs described in this study are available under a materials transfer agreement with Humabs BioMed SA. L.V. is grateful for financial support by SNF grant 310030_166445 and Lions Club Monteceneri. A.L. is an inventor on a patent related to DENV mAbs described in this study: PCT/lB2009/007372 (Dengue virus neutralizing antibodies and uses thereof). D.C. is an inventor on a patent application held by Humabs BioMed SA related to ZIKV mAbs described in this study: PCT/EP2016/066684 (Novel antibodies specifically binding to Zika virus epitopes and uses thereof). Antibody nucleotide and amino acid sequences for the 68 mAbs described in table S2 have been deposited in GenBank (accession nos. KX496807 to KX496874).

Authors

Affiliations

Karin Stettler*
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Martina Beltramello*
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Diego A. Espinosa
Division of Infectious Diseases and Vaccinology, School of Public Health, University of California, Berkeley, Berkeley, CA, USA.
Victoria Graham
National Infection Service, Public Health England, Porton Down, Salisbury, Wiltshire, UK.
Antonino Cassotta
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Institute for Microbiology, ETH Zurich, Wolfgang-Pauli-Strasse 10, 8093 Zurich, Switzerland.
Siro Bianchi
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Fabrizia Vanzetta
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Andrea Minola
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Stefano Jaconi
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Federico Mele
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Mathilde Foglierini
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Mattia Pedotti
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Luca Simonelli
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Stuart Dowall
National Infection Service, Public Health England, Porton Down, Salisbury, Wiltshire, UK.
Barry Atkinson
National Infection Service, Public Health England, Porton Down, Salisbury, Wiltshire, UK.
Elena Percivalle
Molecular Virology Unit, Microbiology and Virology Department, Fondazione Istituto di Ricovero e Cura a Carattere Scientifico (IRCCS) Policlinico San Matteo, Pavia, Italy.
Cameron P. Simmons
Centre for Tropical Medicine, Nuffield Department of Medicine, University of Oxford, UK.
Oxford University Clinical Research Unit, Center for Tropical Medicine, Ho Chi Minh City, Vietnam.
Department of Microbiology and Immunology, University of Melbourne, Peter Doherty Institute, 792 Elizabeth Street, Melbourne VIC 3000, Australia.
Luca Varani
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Johannes Blum
Swiss Tropical and Public Health Institute, Socinstrasse 57, 4002 Basel, Switzerland.
University of Basel, Petersgraben 4, 4031 Basel, Switzerland.
Fausto Baldanti
Molecular Virology Unit, Microbiology and Virology Department, Fondazione Istituto di Ricovero e Cura a Carattere Scientifico (IRCCS) Policlinico San Matteo, Pavia, Italy.
Elisabetta Cameroni
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.
Roger Hewson
National Infection Service, Public Health England, Porton Down, Salisbury, Wiltshire, UK.
Eva Harris
Division of Infectious Diseases and Vaccinology, School of Public Health, University of California, Berkeley, Berkeley, CA, USA.
Antonio Lanzavecchia
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Institute for Microbiology, ETH Zurich, Wolfgang-Pauli-Strasse 10, 8093 Zurich, Switzerland.
Federica Sallusto,§ [email protected]
Institute for Research in Biomedicine, Università della Svizzera italiana, Via Vincenzo Vela 6, 6500 Bellinzona, Switzerland.
Davide Corti,§ [email protected]
Humabs BioMed SA, Via Mirasole 1, 6500 Bellinzona, Switzerland.

Notes

*
These authors contributed equally to this work.
These authors contributed equally to this work.
These authors contributed equally to this work.
§
Corresponding author. Email: [email protected] (D.C.); [email protected] (F.S.)

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. Subcapsular sinus macrophages maximize germinal center development in non-draining lymph nodes during blood-borne viral infection, Science Immunology, 9, 93, (2024)./doi/10.1126/sciimmunol.adi4926
    Abstract
  2. Antibody-Dependent Enhancement Activity of a Plant-Made Vaccine against West Nile Virus, Vaccines, 11, 2, (197), (2023).https://doi.org/10.3390/vaccines11020197
    Crossref
  3. Identification of Zika Virus NS1-Derived Peptides with Potential Applications in Serological Tests, Viruses, 15, 3, (654), (2023).https://doi.org/10.3390/v15030654
    Crossref
  4. Evaluation of the immunogenicity and efficacy of an rVSV vaccine against Zika virus infection in macaca nemestrina, Frontiers in Virology, 3, (2023).https://doi.org/10.3389/fviro.2023.1108420
    Crossref
  5. Structural and immunological basis of cross-reactivity between dengue and Zika infections: Implications in serosurveillance in endemic regions, Frontiers in Microbiology, 14, (2023).https://doi.org/10.3389/fmicb.2023.1107496
    Crossref
  6. Diagnostic and vaccine potential of Zika virus envelope protein (E) derivates produced in bacterial and insect cells, Frontiers in Immunology, 14, (2023).https://doi.org/10.3389/fimmu.2023.1071041
    Crossref
  7. Targeting ‘immunogenic hotspots’ in Dengue and Zika virus: an in silico approach to a common vaccine candidate , Future Virology, 18, 2, (87-106), (2023).https://doi.org/10.2217/fvl-2022-0104
    Crossref
  8. Pathogenesis of Zika Virus Infection, Annual Review of Pathology: Mechanisms of Disease, 18, 1, (181-203), (2023).https://doi.org/10.1146/annurev-pathmechdis-031521-034739
    Crossref
  9. Influence of Dosing Regimen and Adjuvant Type on the Immunogenicity of Novel Recombinant Zika Virus-Like Particles, Microbiology Spectrum, 11, 1, (2023).https://doi.org/10.1128/spectrum.02885-22
    Crossref
  10. Re-mining serum proteomics data reveals extensive post-translational modifications upon Zika and dengue infection, Molecular Omics, (2023).https://doi.org/10.1039/D2MO00258B
    Crossref
  11. See more
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Check Access

Log in to view the full text

AAAS ID LOGIN

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

Log in via OpenAthens.
Log in via Shibboleth.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Media

Figures

Multimedia

Tables

Share

Share

Share article link

Share on social media