Skoči na glavni sadržaj

Prethodno priopćenje

https://doi.org/10.17113/ftb.59.01.21.6820

Potencijalno mjesto vezanja estrogena na S protein virusa SARS-CoV2

Ante Tomasović orcid id orcid.org/0000-0002-4923-1403 ; Faculty of Food Technology and Biotechnology, University of Zagreb, Pierottijeva 6, 10000 Zagreb, Croatia
Damir Stanzer orcid id orcid.org/0000-0002-8720-9063 ; Faculty of Food Technology and Biotechnology, University of Zagreb, Pierottijeva 6, 10000 Zagreb, Croatia
Ivan Krešimir Svetec orcid id orcid.org/0000-0002-0469-2652 ; Faculty of Food Technology and Biotechnology, University of Zagreb, Pierottijeva 6, 10000 Zagreb, Croatia
Marina Svetec Miklenić orcid id orcid.org/0000-0002-2438-7903 ; Faculty of Food Technology and Biotechnology, University of Zagreb, Pierottijeva 6, 10000 Zagreb, Croatia


Puni tekst: engleski pdf 3.810 Kb

str. 24-30

preuzimanja: 207

citiraj

Puni tekst: hrvatski pdf 3.810 Kb

str. 24-30

preuzimanja: 175

citiraj

Preuzmi JATS datoteku


Sažetak

Pozadina istraživanja: Tijekom trenutne SARS-CoV2 pandemije, ali i ranijih epidemija virusa SARS i MERS, primijećeno je da žene oboljele od COVID-19 imaju u prosjeku blaže simptome bolesti, te niži postotak smrtnosti od muškaraca. Poznato je da niz faktora pridonosi različitom imunološkom odgovoru i težini bolesti između spolova. Između ostalog, na sam imunološki odgovor utječe prisutnost hormona estrogena u organizmu i to posrednim djelovanjem preko signalnog puta receptora za estrogen. Međutim, postavili smo pitanje je li moguće da estrogen djeluje direktno na virus SARS-CoV2, i to možda vezanjem na glikoprotein na površini virusne čestice (S protein), čime bi se mogla umanjiti infektivnost virusa.
Eksperimentalni pristup: Radi procjene mogućnosti vezanja estrogena na S protein virusa SARS-CoV2 usporedili smo proteinske sekvence S proteina i humanog estrogen receptora β pomoću alata BLAST i HHpred, a 3D strukture potencijalnog mjesta za vezanje estrogena na S proteinu i aktivnog mjesta estrogen receptora β vizualizirali smo i usporedili pomoću programskog alata PyMOL.
Rezultati i diskusija: Usporedbom sekvence S proteina virusa SARS-CoV2 sa sekvencom humanog estrogen receptora β, identificirali smo moguće mjesto vezanja estrogena na S proteinu, a promatranjem njihovih 3D struktura utvrdili veliku sličnost tog aktivnog mjesta s aktivnim mjestom ER β. S protein virusa SARS-CoV2 uspoređen je s humanim androgen receptorom radi kontrole dobivenih rezultata, no nije utvrđena sličnost među njima. Potencijalno mjesto vezanja estrogena dio je domene S2 koja sudjeluje u fuziji membrana virusa i stanice domaćina tijekom infekcije, te je konzervirana u virusima iz porodice Coronaviridae.
Novina i znanstveni doprinos: U ovom prethodnom priopćenju prikazana je mogućnost da S protein virusa SARS-CoV2 nosi mjesto za vezanje estrogena. Ovo saznanje moglo bi potaknuti provedbu sveobuhvatnije studije stvarnih mogućnosti vezanja estrogena na S protein i eventualnog učinka koji bi to moglo imati na infektivnost virusa.

Ključne riječi

SARS-CoV2; koronavirus; S protein; estrogen; estradiol; mjesto vezanja estrogena

Hrčak ID:

256910

URI

https://hrcak.srce.hr/256910

Datum izdavanja:

18.4.2021.

Podaci na drugim jezicima: engleski

Posjeta: 905 *




INTRODUCTION

The COVID-19 pandemic that has resulted in high numbers of critically ill patients and deaths is caused by novel coronavirus SARS-CoV2. Studies on the infection and clinical management reported sex differences in severity and outcome of the COVID-19 disease (1-7). It is known that males are more susceptible to influenza and common cold than females of the same age (8-10). Two other coronaviruses that also cause severe respiratory illness as SARS-CoV-2 does, severe acute respiratory syndrome coronavirus (SARS-CoV) and the Middle East respiratory syndrome coronavirus (MERS-CoV) with a nucleotide identity to SARS-CoV-2 of 79 and 50% respectively (11), also tend to be more severe and fatal in men than in women (12-15).

Besides lifestyle factors, the sex-specific difference in coronavirus susceptibility and disease outcomes is caused by sex differences in the immune response, both innate and adaptive (16). One of the explanations for this sex difference in COVID-19 susceptibility is chromosomal: X chromosome contains more immune-related genes than the Y chromosome, and activation of these genes leads to stronger female immune response to viral infection (7,17,18). Another explanation is hormonal: hormones, such as estrogen, which females produce in larger quantities than males, help to defend against coronaviruses like MERS-CoV, SARS-CoV and SARS-CoV-2. There are various possible explanations of estrogen protective action mechanism (19-24). Among the effects of estrogen in the immune defence, there is an influence on adaptive immunity fighters, T- and B-cells, by impairing negative selection of high affinity auto-reactive B cells, modulating B cell function and leading to Th2 response (25), and induction of T cell homing by enhancing the expression of CCR5, a homing marker (26). An experiment performed with mice infected with mouse-adapted SARS-CoV MA15 (27) showed that the male mice were more susceptible to the SARS infection than females, and that the female mice that had their estrogen-producing ovaries removed or were treated with an estrogen-receptor blocker had higher fatality rates than those with normal estrogen function. The sex-specific differences were independent of T and B cell responses. Therefore, these results suggest that estrogen receptor (ER) signalling in females suppressed the accumulation and function of inflammatory monocyte macrophages in the lungs and/or directly suppressed SARS-CoV replication via effects on cellular metabolism.

Besides acting on ER signalling as a mechanism of conferring strong positive effect on disease outcomes and fatality rate (25), we considered the idea that the estrogen might act directly on the virus. We hypothesized that the estrogen molecule might bind on the virus surface and thus reduce its infectivity. Since spike (S) protein is the major surface protein the SARS-CoV2 employs to bind to the human receptor and initiate fusion with the host cell (26-29), we wondered if there was any theoretical possibility that estrogen could bind to this protein. If this were indeed the case, then estrogen and estrogen-like molecules might have a therapeutic effect. To get some insight into these questions, we first performed protein BLAST to determine if there was any similarity between SARS-CoV2 S protein and the human ER β sequences. Although these two proteins originate from different species and differ greatly in their overall size, a sequence similarity was found in regions which correspond to the estrogen binding site of the human ER β. We further explored the spatial structure of potential estrogen binding site on S protein using PyMol software and again found substantial similarities with the 3D structure of binding site on the human ER β.

MATERIALS AND METHODS

Protein sequences and 3D structures

The following protein sequences were downloaded from NCBI (30): human ER β sequence (UniProtKB/Swiss-Prot: Q92731.2), human androgen (dihydrotestosterone) receptor (UniProtKB/Swiss-Prot: P10275.3), SARS-CoV2 surface glycoprotein (NCBI reference sequence: YP_009724390.1) and SARS CoV Urbani spike protein (GenBank: AYV99817.1). The following 3D structures were downloaded from RCSB Protein Data Bank (31): crystal structure of ER β bound to estradiol (PDB ID: 5TOA (32)), crystal structure of human androgen receptor ligand-binding domain in complex with dihydrotestosterone (PDB ID: 2AMA (33)) and prefusion 2019-nCoV spike glycoprotein with a single receptor-binding domain up (PDB ID: 6VSB (29)).

Bioinformatic tools

Protein sequence alignment and analysis were performed using protein BLAST (34) and the HHpred server for remote protein homology detection and structure prediction (https://toolkit.tuebingen.mpg.de/tools/hhpred (35)). In all cases, standard algorithm parameters were used (36,37). For 3D protein structure, visualization the PyMOL software was used (38).

RESULTS AND DISCUSSION

To assess whether there could be any possibilities for the S protein of SARS-CoV2 to contain a site that might bind estrogen, we used protein BLAST to compare its sequence with human ER β. We found three regions of similarity. The two more significant were: a region of 19 amino acids starting from amino acid 319 of the ER β and amino acid 817 of S protein with 73% positives without gaps (e-value 2.5) and a region of 14 amino acids starting from amino acid 385 of the ER β and amino acid 236 of S protein with 64% positives without gaps (e-value 3.2). Although the e-values were relatively high, we found it interesting that both of these two regions are annotated in NCBI as part of the ligand-binding domain of the estrogen receptor.

Furthermore, we visualized the 3D structure of the SARS-CoV2 S protein, using PyMOL software (38), and compared it to the 3D structure of ER β to check whether the regions of S protein carrying either the 19 or 14 amino acid alignments found by BLAST bear any spatial resemblance to the estrogen binding site of ER β and thus could constitute a potential estrogen binding site on the S protein. According to the 3D structure analysis, the 19 amino acid region (319F to 337E) appears to play an important role in the ligand-binding domain on the ER β (Fig. 1a) since it protrudes into the active site with the methionine 336M as the closest amino acid to the estradiol (a major form of estrogen in women of reproductive age) when bound. The residue 336M is less than 5 Å apart from the estradiol molecule (Fig. 1b). Therefore, it is reasonable to assume that the terminal part of the 19 amino acid alignment constitutes a part of an active site on ER β. Residues involved in ligand-binding on ER β are glutamate 305E situated on the first alpha helix and the glycine 472G and histidine 475H situated on the second alpha helix, which form three polar bonds with estradiol (Fig. 1c).

Fig. 1 The 3D structure of: a) human estrogen receptor β in the vicinity of 19 amino acid sequence alignment with the S protein spanning residues from F319 do E337 (orange) with estradiol (magenta) bound to the active site, b) the alignment ends with M336 protruding into the active site and being less than 5 Å from the estradiol molecule, and c) the active site seen from different perspective offers better view of other surrounding residues likely involved in binding (yellow) and the three residues (E305, G472 and H475, marked red) which can form polar bonds with estradiol
FTB-59-24-f1

In the SARS-CoV2 S protein, the 19 amino acid alignment begins with phenylalanine 817F. In the corresponding 3D structure obtained by electron micrograph (29) residues 817F through 828L can be found. Unfortunately, the structure contains a brakeage of protein chain between the residues 811 and 815 as well as 828 and 853 resulting in separation of this part of the molecule. Moreover, the protein chain breakage visible in the 3D structure of SARS-CoV2 S protein is present at the approximately same site in the SARS-CoV Urbani 3D structure (the related coronavirus causing SARS pandemic in 2003), indicating that this particular site is sensitive to sample preparation and manipulation. However, the separated alpha helix is located near the breakage site, indicating that it has not moved considerably from its native position (Fig. 2). Instead of methionine 336M present in the analyzed 19 amino acid region on ER β, the corresponding region on S protein contains an analogue nonpolar residue isoleucine 834I scored as positive by BLAST. Although the exact position of 834I in the available model of S protein cannot be observed due to the protein chain breakage, it can be positioned in space based on the surrounding residues. We then tried to identify the amino acids that might be close in space to the 834I and could form polar bonds with estradiol molecule. Using BLAST, we identified a region of S protein (residues 1045 to 1052), which aligns with the region of ER β spanning residues 471 to 479 with 66% of them scored as positives, and which contains two key residues for estradiol binding on ER β – 472G and 475H. The main difference is that on the S protein the glycine and histidine are separated by one amino acid (Y), while two residues in between are present on the ER β (M and E). To find the potential analogue of the third amino acid responsible for estradiol binding on ER β (305E), we used the HHpred server (35). The region spanning residues 289 to 310 of ER β (containing 305E) was compared against the entire sequence of S protein. The analysis resulted in region spanning residues from 1013 to 1033 of S protein identified as similar with two gaps introduced to achieve the alignment of 305E with its potential analogue on the S protein – the glutamate at position 1031.

Fig. 2 The 3D structure of S protein in the vicinity of potential estrogen binding site. The 19 amino acid sequence alignment with the region on ER β is shown in orange. Residues R815 and L828 are sites of protein chain breakage. Amino acids possibly involved in polar binding of estrogen in the potential active sites (E1031, G1046 and H1048) are marked red while other residues surrounding the potential binding site are shown in yellow
FTB-59-24-f2

Finally, using PyMol (38), we visualized the spatial position of three amino acids, which can form polar bonds with the ligand (1046G, 1048H and 1031E) on the 3D structure of S protein and compared this site to the active site for estradiol binding on ER β. Histidine residue 1048H on the S protein was rotated in space around alpha C atom to face towards the 1031E, which is likely to occur in the case of estradiol binding. The three key amino acids on the ER β responsible for polar binding of estradiol form a triangle with the length of its sides being 4.7, 12.8 and 14.8 Å (Fig. 3a). On the S protein the three analogous amino acids form a triangle with sides 4.6, 14.8 and 11.3 Å long (Fig. 3b). In other words, the dimensions of the space between the residues capable of forming polar bonds with estradiol in ER β and the analyzed site on S protein are curiously similar, with the main difference being that the longest side of the triangle in the active site of ER β is between histidine and glutamate and of the S protein between glycine and glutamate. Moreover, besides the three residues with the potential of forming polar bonds with estradiol, a number of other residues surrounding the potential binding site on S protein are identical or similar in their chemical properties to that present in the ER β active site (Figs. 3c and 3d).

Fig. 3 Comparison of estrogen binding site of ER β and potential estrogen binding site found on S protein of SARS-CoV2. Amino acids E, G and H capable of forming polar bonds with estrogen are depicted red: a) sides of the triangle between E, G and H residues are 4.7, 12.8 and 14.8 Å on ER β, and b) 4.6, 14.8 i 11.3 Å on S protein, c) identical (red and purple) and similar (yellow) residues surrounding the estrogen binding site of ER β compared with d) potential estrogen binding site on SARS-CoV2 S protein
FTB-59-24-f3

As a control, sequences and 3D structures of S protein and human androgen receptor were compared using the same methodology as described above. Although testosterone and estrogen are highly similar molecules and the estrogen and testosterone receptors share regions of notable similarity in the ligand-binding domain, we did not find sequence alignments of S protein with the androgen receptor that include all three key residues for testosterone binding (R, T and N), nor could we identify a site in the 3D structure of S protein that would share significant resemblance to the testosterone binding site (data not shown).

Taken together, our analyses suggest that SARS-CoV2 S protein features a potential estrogen binding site. Understandably, whether or not the estradiol molecule actually can bind to the potential site - and if it does bind, how it affects the S protein and the infectivity of the virus can be established only experimentally. The strong positive effect of estrogen on disease outcome and fatality rate was experimentally shown in mice using mouse-adapted SARS-CoV MA15 and the effect was explained by estrogen receptor signalling (27). The entire S protein of mouse-adapted strain differs in only one amino acid from the SARS-CoV, the human virus causing the SARS 2003 epidemic (27). Thus, we wondered if SARS-CoV, which is closely related to SARS-CoV2, also features a potential estrogen binding site and whether it is conceivable that at least to an extent the positive effect of estrogen observed in these experiments might be due to the S protein binding estrogen. The BLAST analysis revealed that in the region 800–1273, which includes the potential estrogen binding site, the two S proteins are 93% identical (with 97% amino acids scored as positives without gaps). The 3D structure of SARS-CoV also presented the same conformation in the investigated region with the potential estrogen binding site present (data not shown).

Moreover, the entire region of SARS-Cov2 S protein ranging from residue 662 to 1270 is a conserved S2 domain belonging to the coronavirus S2 superfamily. After transcription, the S protein is cleaved to S1 and S2 subunits. The S1 is responsible for binding to the host receptor, while the S2 subunit contains machinery for fusion with the host cell which consists of several distinct regions: the fusion peptide, heptad repeats 1 and 2 (HR1 and HR2) and transmembrane domain (39). The potential estrogen binding site identified in this work lies in the region carrying a linker between the two heptad repeats (40). A BLAST analysis of SARS-CoV2 S protein region 1025 to 1055 against S proteins of 20 randomly chosen species across the Coronaviridae family showed that this site is conserved, with the three amino acids (E, G and H) likely capable of forming polar bonds with estradiol together with some of the surrounding amino acids being present across the family (Fig. 4). Given the similarities of this region to the ER β, it is possible that an ancestral viral species at some point in time integrated a section of an ER from its host (41). However, it seems that this section might have taken on an important role in the coronavirus infectivity as a part S2 domain, perhaps being involved in host-virus membrane fusion by binding of other molecules featuring four rings, such as cholesterol. Thus, binding of estrogen might compete with the natural ligand for this site or cause conformational changes leading to reduced function of S2 domain and diminished infectivity.

Fig. 4 Results of BLAST analysis of SARS-CoV2 S protein region from 1025 to 1055 against S proteins of 20 randomly chosen species across the Coronaviridae family. Dots in the alignment represent identical amino acids. Yellow squares outline the key amino acids (E, G and H) identified as having potential for estradiol binding on SARS-CoV2 S protein
FTB-59-24-f4

Given the differences in COVID19 disease outcomes and fatality rate between men and women, the usage of estradiol to reduce severity of COVID19 infection is currently being tested in a phase II clinical trial in Stony Brook University Hospital, New York, NY, USA (https://clinicaltrials.gov/ct2/show/NCT04359329). The positive and suspected to be positive patients are given estradiol patches delivering 100 μg/day of estradiol for 7 days and rates of hospitalization, transfer to intensive care unit, intubation and death are being monitored during the 30-day time frame and compared to the non-treated control. The study recruits men over the age of 18 and women over the age of 55. Hopefully, although preliminary, the results presented in this work will inspire more researchers into investigating estrogen and estrogen-like molecules for treating COVID-19 since using different drugs, doses and methods of delivery could all affect the final result.

CONCLUSIONS

In this work, we identified a conserved potential estrogen binding site on SARS-CoV2 surface glycoprotein (S protein). This result could offer additional explanation of the better disease outcomes for COVID19 positive women than men, as well as of the strong protective effect of estrogen in experimental animals. Most importantly, this result could prompt other researchers to experimentally prove whether or not the estrogen can indeed bind to the S protein and what effect would that confer on the infectivity of the virus. Hopefully, estrogen or estrogen-like molecule might be used to treat SARS-CoV2 infection or alleviate symptoms.

Notes

[1] Financial disclosure FUNDING

This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

[2] Conflicts of interest CONFLICT OF INTEREST

Authors declare no conflict of interest.

REFERENCES

1 

Cai H. Sex difference and smoking predisposition in patients with COVID-19. Lancet Respir Med. 2020;8(4):e20. https://doi.org/10.1016/S2213-2600(20)30117-X PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32171067

2 

Zhao Y, Zhao Z, Wang Y, Zhou Y, Ma Y, Zuo W. Single-cell RNA expression profiling of ACE2, the receptor of SARS-CoV-2. Am J Respir Crit Care Med. 2020;202(5):756–9. https://doi.org/10.1164/rccm.202001-0179LE PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32663409

3 

Yang X, Yu Y, Xu J, Shu H, Xia J, Liu H, et al. Clinical course and outcomes of critically ill patients with SARS-CoV-2 pneumonia in Wuhan, China: A single-centered, retrospective, observational study. Lancet Respir Med. 2020;8(5):475–81. https://doi.org/10.1016/S2213-2600(20)30079-5 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32105632

4 

Richardson S, Hirsch JS, Narasimhan M, Crawford JM, McGinn T, Davidson KW; Northwell COVID-19 Research Consortium. Presenting characteristics, comorbidities, and outcomes among 5700 patients hospitalized with COVID-19 in the New York City area. JAMA. 2020;323(20):2052–9. https://doi.org/10.1001/jama.2020.6775 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32320003

5 

Dowd JB, Andriano L, Brazel DM, Rotondi V, Block P, Ding X, et al. Demographic science aids in understanding the spread and fatality rates of COVID-19. Proc Natl Acad Sci USA. 2020;117(18):9696–8. https://doi.org/10.1073/pnas.2004911117 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32300018

6 

La Vignera S, Cannarella R, Condorelli RA, Torre F, Aversa A, Calogero AE. Sex-specific SARS-CoV-2 mortality: Among hormone-modulated ACE2 expression, risk of venous thromboembolism and hypovitaminosis D. Int J Mol Sci. 2020;21(8):2948–53. https://doi.org/10.3390/ijms21082948 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32331343

7 

Conti P, Younes A. Coronavirus COV-19/SARS-CoV-2 affects women less than men: Clinical response to viral infection. J Biol Regul Homeost Agents. 2020;34(2):339–43. https://doi.org/10.23812/Editorial-Conti-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32253888

8 

Wang XL, Yang L, Chan KH, Chan KP, Cao PH, Lau EH, et al. Age and sex differences in rates of influenza-associated hospitalizations in Hong Kong. Am J Epidemiol. 2015;182(4):335–44. https://doi.org/10.1093/aje/kwv068 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26219977

9 

Quandelacy TM, Viboud C, Charu V, Lipsitch M, Goldstein E. Age- and sex-related risk factors for influenza-associated mortality in the United States between 1997–2007. Am J Epidemiol. 2014;179(2):156–67. https://doi.org/10.1093/aje/kwt235 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24190951

10 

Macintyre S. Gender differences in the perceptions of common cold symptoms. Soc Sci Med. 1993;36(1):15–20. https://doi.org/10.1016/0277-9536(93)90301-J PubMed: http://www.ncbi.nlm.nih.gov/pubmed/8424180

11 

Lu R, Zhao X, Li J, Niu P, Yang B, Wu H, et al. Genomic characterisation and epidemiology of 2019 novel coronavirus: Implications for virus origins and receptor binding. Lancet. 2020;395(10224):565–74. https://doi.org/10.1016/S0140-6736(20)30251-8 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32007145

12 

Karlberg J, Chong DS, Lai WY. Do men have a higher case fatality rate of severe acute respiratory syndrome than women do? Am J Epidemiol. 2004;159(3):229–31. https://doi.org/10.1093/aje/kwh056 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/14742282

13 

Alghamdi IG, Hussain II, Almalki SS, Alghamdi MS, Alghamdi MM, El-Sheemy MA. The pattern of Middle East respiratory syndrome coronavirus in Saudi Arabia: A descriptive epidemiological analysis of data from the Saudi Ministry of Health. Int J Gen Med. 2014;7:417–23. https://doi.org/10.2147/IJGM.S67061 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25187734

14 

Matsuyama R, Nishiura H, Kutsuna S, Hayakawa K, Ohmagari N. Clinical determinants of the severity of Middle East respiratory syndrome (MERS): A systematic review and meta-analysis. BMC Public Health. 2016;16(1):1203–12. https://doi.org/10.1186/s12889-016-3881-4 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27899100

15 

Chen J, Subbarao K. The Immunobiology of SARS. Annu Rev Immunol. 2007;25:443–72. https://doi.org/10.1146/annurev.immunol.25.022106.141706 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/17243893

16 

Klein SL, Marriott I, Fish EN. Sex-based differences in immune function and responses to vaccination. Trans R Soc Trop Med Hyg. 2015;109(1):9–15. https://doi.org/10.1093/trstmh/tru167 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25573105

17 

Klein SL, Flanagan KL. Sex differences in immune responses. Nat Rev Immunol. 2016;16(10):626–38. https://doi.org/10.1038/nri.2016.90 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27546235

18 

Robinson DP, Huber SA, Moussawi M, Roberts B, Teuscher C, Watkins R, et al. Sex chromosome complement contributes to sex differences in coxsackievirus B3 but not influenza A virus pathogenesis. Biol Sex Differ. 2011;2:8. https://doi.org/10.1186/2042-6410-2-8 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21806829

19 

Jones BG, Sealy RE, Penkert RR, Surman SL, Birshtein BK, Xu B, et al. From influenza virus infections to lupus: Synchronous estrogen receptor α and RNA polymerase II binding within the immunoglobulin heavy chain locus. Viral Immunol. 2020;33(4):307–15. https://doi.org/10.1089/vim.2019.0144 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32105583

20 

Heath WR, Kato Y, Steiner TM, Caminschi I. Antigen presentation by dendritic cells for B cell activation. Curr Opin Immunol. 2019;58:44–52. https://doi.org/10.1016/j.coi.2019.04.003 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31071588

21 

Laffont S, Seillet C, Guery JC. Estrogen receptor-dependent regulation of dendritic cell development and function. Front Immunol. 2017;8:108. https://doi.org/10.3389/fimmu.2017.00108 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28239379

22 

Mai T, Zan H, Zhang J, Hawkins JS, Xu Z, Casali P. Estrogen receptors bind to and activate the HOXC4/HoxC4 promoter to potentiate HoxC4-mediated activation-induced cytosine deaminase induction, immunoglobulin class switch DNA recombination, and somatic hypermutation. J Biol Chem. 2010;285(48):37797–810. https://doi.org/10.1074/jbc.M110.169086 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/20855884

23 

Pauklin S, Sernández IV, Bachmann G, Ramiro AR, Petersen-Mahrt SK. Estrogen directly activates AID transcription and function. J Exp Med. 2009;206(1):99–111. https://doi.org/10.1084/jem.20080521 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19139166

24 

Jones BG, Penkert RR, Surman SL, Sealy RE, Pelletier S, Xu B, et al. Matters of life and death: How estrogen and estrogen receptor binding to the immunoglobulin heavy chain locus may influence outcomes of infection, allergy, and autoimmune disease. Cell Immunol. 2019;346:103996. https://doi.org/10.1016/j.cellimm.2019.103996 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31703914

25 

Grimaldi CM, Jeganathan V, Diamond B. Hormonal regulation of B cell development: 17 b-estradiol impairs negative selection of high-affinity DNA-reactive B cells at more than one developmental checkpoint. J Immunol. 2006;176(5):2703–10. https://doi.org/10.4049/jimmunol.176.5.2703 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/16493025

26 

Mo R, Chen J, Grolleau-Julius A, Murphy HS, Richardson BC, Yung RL. Estrogen regulates CCR gene expression and function in T lymphocytes. J Immunol. 2005;174(10):6023–9. https://doi.org/10.4049/jimmunol.174.10.6023 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15879095

27 

Channappanavar R, Fett C. Matthias Mack, Ten Eyck PP, Meyerholz DK, Perlman S. Sex-based differences in susceptibility to SARS-CoV infection. J Immunol. 2017;198(10):4046–53. https://doi.org/10.4049/jimmunol.1601896 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28373583

28 

Wan Y, Shang J, Graham R, Baric RS, Li F. Receptor recognition by novel coronavirus from Wuhan: An analysis based on decade-long structural studies of SARS. J Virol. 2020;94(7):e00127–20. https://doi.org/10.1128/JVI.00127-20 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/31996437

29 

Wrapp D, Wang N, Corbett KS, Goldsmith JA, Hsieh C, Abiona O, et al. Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation. Science. 2020;367(6483):1260–3. https://doi.org/10.1126/science.abb2507 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32075877

30 

National Center for Biotechnology Information (NCBI). Bethesda, MD, USA: National Library of Medicine (US), National Center for Biotechnology Information; 1988. Available from:https://www.ncbi.nlm.nih.gov/protein.

31 

Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, et al. The Protein Data Bank. Nucleic Acids Res. 2000;28(1):235–42. https://doi.org/10.1093/nar/28.1.235 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10592235

32 

Souza PCT, Textor LC, Melo DC, Nascimento AS, Skaf MS, Polikarpov I. An alternative conformation of ERβ bound to estradiol reveals H12 in a stable antagonist position. Sci Rep. 2017;7(1):3509. https://doi.org/10.1038/s41598-017-03774-x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28615710

33 

Pereira de Jésus-Tran K, Côté PL, Cantin L, Blanchet J, Labrie F, Breton R. Comparison of crystal structures of human androgen receptor ligand-binding domain complexed with various agonists reveals molecular determinants responsible for binding affinity. Protein Sci. 2006;15(5):987–99. https://doi.org/10.1110/ps.051905906 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/16641486

34 

BLAST [Internet]. Bethesda, MD, USA: National Library of Medicine (US), National Center for Biotechnology Information; 2004. [cited 15.05.2020.]. Available from:https://blast.ncbi.nlm.nih.gov/Blast.cgi.

35 

Söding J, Biegert A, Lupas AN. The HHpred interactive server for protein homology detection and structure prediction. Nucleic Acids Res. 2005;33 Suppl. 2:W244-8. https://doi.org/10.1093/nar/gki408 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15980461

36 

Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997;25(17):3389–402. https://doi.org/10.1093/nar/25.17.3389 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9254694

37 

Altschul SF, Wootton JC, Gertz EM, Agarwala R, Morgulis A, Schäffer AA. Yi- Yu YK. Protein database searches using compositionally adjusted substitution matrices. FEBS J. 2005;272(20):5101–9. https://doi.org/10.1111/j.1742-4658.2005.04945.x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/16218944

38 

PyMOL™, v. 2.3.4, Schrödinger, LLC., New York, NY, USA; 2020. Available from:https://pymol.org/.

39 

Li F. Structure, function, and evolution of coronavirus spike proteins. Annu Rev Virol. 2016;3:237–61. https://doi.org/10.1146/annurev-virology-110615-042301 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27578435

40 

Xia S, Liu M, Wang C, Xu W, Lan Q, Feng S, et al. Inhibition of SARS-CoV-2 (previously 2019-nCoV) infection by a highly potent pan-coronavirus fusion inhibitor targeting its spike protein that harbors a high capacity to mediate membrane fusion. Cell Res. 2020;30(4):343–55. https://doi.org/10.1038/s41422-020-0305-x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32231345

41 

Weiss RA. Exchange of genetic sequences between viruses and hosts. In: Hunter E, Bister K, editors. Viruses, genes, and cancer. Current Topics in Microbiology and Immunology, vol. 407. Cham, Switzerland: Springer; 2017. pp. 1-29. https://doi.org/ https://doi.org/10.1007/82_2017_21


This display is generated from NISO JATS XML with jats-html.xsl. The XSLT engine is libxslt.