BACH2 immunodeficiency illustrates an association between super-enhancers and haploinsufficiency

Behdad Afzali, Juha Grönholm, Jana Vandrovcova, Charlotte O'Brien, Hong-Wei Sun, Ine Vanderleyden, Fred P Davis, Ahmad Khoder, Yu Zhang, Ahmed N Hegazy, Alejandro V Villarino, Ira W Palmer, Joshua Kaufman, Norman R Watts, Majid Kazemian, Olena Kamenyeva, Julia Keith, Anwar Sayed, Dalia Kasperaviciute, Michael Mueller, Jason D Hughes, Ivan J Fuss, Mohammed F Sadiyah, Kim Montgomery-Recht, Joshua McElwee, Nicholas P Restifo, Warren Strober, Michelle A Linterman, Paul T Wingfield, Holm H Uhlig, Rahul Roychoudhuri, Timothy J Aitman, Peter Kelleher, Michael J Lenardo, John J O'Shea, Nichola Cooper, Arian D J Laurence, Behdad Afzali, Juha Grönholm, Jana Vandrovcova, Charlotte O'Brien, Hong-Wei Sun, Ine Vanderleyden, Fred P Davis, Ahmad Khoder, Yu Zhang, Ahmed N Hegazy, Alejandro V Villarino, Ira W Palmer, Joshua Kaufman, Norman R Watts, Majid Kazemian, Olena Kamenyeva, Julia Keith, Anwar Sayed, Dalia Kasperaviciute, Michael Mueller, Jason D Hughes, Ivan J Fuss, Mohammed F Sadiyah, Kim Montgomery-Recht, Joshua McElwee, Nicholas P Restifo, Warren Strober, Michelle A Linterman, Paul T Wingfield, Holm H Uhlig, Rahul Roychoudhuri, Timothy J Aitman, Peter Kelleher, Michael J Lenardo, John J O'Shea, Nichola Cooper, Arian D J Laurence

Abstract

The transcriptional programs that guide lymphocyte differentiation depend on the precise expression and timing of transcription factors (TFs). The TF BACH2 is essential for T and B lymphocytes and is associated with an archetypal super-enhancer (SE). Single-nucleotide variants in the BACH2 locus are associated with several autoimmune diseases, but BACH2 mutations that cause Mendelian monogenic primary immunodeficiency have not previously been identified. Here we describe a syndrome of BACH2-related immunodeficiency and autoimmunity (BRIDA) that results from BACH2 haploinsufficiency. Affected subjects had lymphocyte-maturation defects that caused immunoglobulin deficiency and intestinal inflammation. The mutations disrupted protein stability by interfering with homodimerization or by causing aggregation. We observed analogous lymphocyte defects in Bach2-heterozygous mice. More generally, we observed that genes that cause monogenic haploinsufficient diseases were substantially enriched for TFs and SE architecture. These findings reveal a previously unrecognized feature of SE architecture in Mendelian diseases of immunity: heterozygous mutations in SE-regulated genes identified by whole-exome/genome sequencing may have greater significance than previously recognized.

Conflict of interest statement

Competing financial interests: The authors have no competing interests to declare. Unrelated to this project, H.H.U. declares industrial project collaboration with Lilly, UCB Pharma and Vertex Pharmaceuticals. Travel support was received from Actelion, and MSD.

Figures

Figure 1. Pedigrees and phenotype of patients…
Figure 1. Pedigrees and phenotype of patients with mutations in BACH2
(a) Pedigrees of two families with heterozygous missense coding mutations in BACH2, resulting in L24P (left) and E788K (right) amino acid substitutions. Shown are affected heterozygotes (filled symbols) and unaffected family members (open symbols). Arrows indicate probands; WT = wild-type allele; Mut = mutant allele. (b) Sanger sequencing chromatograms of the affected individuals in both families. For each individual, the two alleles of the sequenced region of BACH2 and base positions are shown above the chromatograms. Subject A.II.1 had a heterozygous T to C mutation at coding position 71 whereas patients B.II.1 and B.III.2 were heterozygous for G to A base substitutions at position 2362. (c) Computerized tomography scans showing splenomegaly (arrow in upper left) and pulmonary nodules (red circle in upper right) in patient A.II.1 and bronchiectasis (dilated airways; arrow in lower left) and fibrosis (“honeycombing” circled in lower right) in subject B.II.1. (d) Photomicrograph of a hematoxylin and eosin-stained section from a colonic biopsy from patient A.II.1 showing crypt branching and lymphocytic inflammatory infiltrate around the crypts. (e) Immunofluorescent staining of colonic biopsy from patient A.II.1, control IBD patient and healthy control for nuclear DNA (DAPI, blue), CD3 (green) and FoxP3 (orange). Shown are representative sections (left) and cumulative (mean ± sem) quantification (right) from four low power fields per patient (500–3000 CD3+ cells counted per low power field); white scale bar = 100 μm in main image and 2 μm in insets. *p<0.05, **p<0.01 by t-test.
Figure 2. Immunophenotype of patients with mutations…
Figure 2. Immunophenotype of patients with mutations in BACH2
(a–c) Treg cells (a), T cell (b) and B cell (c) immunophenotype of patient and healthy control peripheral blood cells. Shown are total FoxP3 expression (mean fluorescent intensity (MFI)) within CD4+CD25hiCD127lo cells (a), expression of the transcription factor T-bet and gut-homing receptors (CCR9 and β7-integrin) in bulk CD4+ T cells (b) and total memory (c, left) and class-switched memory B cells (c, right) in bulk B cells. (d-e) Plasmablast formation (d, left panels), IgG class switch recombination (d, right panels) and Ig secretion (e) in naïve patient and healthy control B cells activated in vitro as indicated. Shown are representative flow cytometry plots and cumulative data. N.D. = not detected; very low values are shown above the bars for clarity. In (a-d) representative flow cytometry plots are shown together with cumulative data from all patients and matched controls. Note that IgG secretion in (e) does not include patient B.III.2, who has normal IgG secretion. Bars show mean ± sem throughout. *p<0.05 **p<0.01 ***p<0.001 by t-test (a-c), one-way ANOVA (d) and Kruskal-Wallis test (e).
Figure 3. The cellular phenotype is attributable…
Figure 3. The cellular phenotype is attributable to reduced BACH2 protein expression
(a) BACH2 protein expression in primary immune cells of patients and controls. Shown are representative flow cytometry plots with MFIs indicated (left panels) and cumulative BACH2 protein expression (right panels) from patients relative to controls. (b) Cumulative BACH2 mRNA expression from naïve B cells of patients and controls. (c) Representative immunoblot for Flag and Hsp70 from lysates of HEK293T cells transfected with empty vector (EV), Flag-tagged WT or mutant murine Bach2 (L24P or E786K, the murine equivalent of E788K). Shown are a representative blot (left) and cumulative quantifications from n = 5 experiments (right). (d) PRDM1 mRNA expression in naïve B cells from patients and healthy controls: cumulative data. (e and f) PRDM1 mRNA expression in CD4+ T lymphocytes of healthy controls and patients transfected with either control or BACH2 (e) and healthy donor CD4+ T lymphocytes transfected with control or BACH2 RNAi (f). (g) Plasmablast formation, IgG class switch recombination and IgA secretion in naïve healthy control B cells transfected with control RNAi or RNAi specific for BACH2 and activated in vitro as shown. Shown are representative flow cytometry examples and cumulative data (n = 5, 5 and 4 experiments, respectively). Bars show mean ± sem; *p<0.05, **p<0.01, ***p<0.001, ****p<0.0001 by t-test (a, d), Wilcoxon test (f) and ANOVA (c, e and g).
Figure 4. BACH2 mutations produce unstable proteins
Figure 4. BACH2 mutations produce unstable proteins
(a) Domain schematic of BACH2 protein and point substitutions in patients. BTB/POZ, BR-C, ttk and bab or Pox virus and Zinc finger domain; bZIP, basic leucine zipper; NES, nuclear export signal. (b), Ribbon representations of BACH2 POZ domain (crystal structure form II, PDB: 3OHV); wild-type protein (above) with expanded and rotated interface view (below); yellow, intermolecular disulfide at position 20; orange, leucine residues at position 24. (c), (top) WT POZ domain dimer interface (PDB: 3OHV); (bottom) homology model of BACH2L24P: WT POZ hetero-dimer, illustrating local changes. In each, one monomer is rendered as a partially transparent hydrophobicity surface (orange = hydrophobic, white = intermediate, blue = hydrophilic) and the other as a ribbon (green); selected side chains are shown as sticks. Cys20 (yellow) and Ile23, Leu24, and Leu27 (all orange) form a hydrophobic patch on α-helix-1; two of these patches are in close contact at the WT dimer interface. N.B. the lower diagram is not meant to represent the structure accurately but is shown merely to indicate regional changes. (d–e) Analytical ultracentrifugation of purified wild-type (WT) p.BACH2 (d) and mutant p.BACH2L24P (e) BTB/POZ domain; sedimentation direction is left to right; M = sample meniscus. WT protein is dimeric (35 kDa), as determined by sedimentation equilibrium measurements (shown in d, right), migrating with single boundary with sedimentation coefficient (S) of 2.6. The mutant exhibits several boundaries (S values from 4 to 18), indicating heterogeneous large protein aggregates (e). (f) Representative confocal microscopy of primary lymphocytes from healthy control and patient B.II.1 stained for BACH2 (green) and Hoechst (blue); arrows highlight cytoplasmic aggregates. Scale bars: 5μm in main, 2μm in inset. Bars show quantification (mean ± sem, n=3 experiments) of cells containing aggregates per high power field (HPF) and BACH2 nuclear localization. *p<0.05 by t-test.
Figure 5. Mutant forms of Bach2 do…
Figure 5. Mutant forms of Bach2 do not exert dominant negative effects
(a) Immunoblot for Flag and Hsp70 in HEK293T cells co-transfected at 1:1 ratio with Flag-tagged WT murine BACH2 and untagged WT and mutant forms of murine BACH2. Shown is a representative from n = 3 independent experiments. (b) co-immunoprecipitation of Flag- and HA-tagged WT Bach2 transfected into HEK293T cells together with untagged WT and mutant forms of murine BACH2 at 1:1:1 vector ratio. Shown is a representative example from n = 3 independent experiments (left) and quantification of the co-immunoprecipitated Flag and HA signals (right). (c) Blimp1-YFP signal in Blimp1-YFP Tg mouse CD4+ T cells co-transduced at 1:1 ratio with retrovirus supernatants encoding WT and mutant forms of murine BACH2. Shown is a representative example (left) and cumulative data (mean ± sem) from n = 4 independent experiments (right). *p<0.0001 by ANOVA.
Figure 6. Bach2 haploinsufficient mice have abnormal…
Figure 6. Bach2 haploinsufficient mice have abnormal B cell differentiation and Treg cell numbers
(a) Expression of Bach2 mRNA in B cells of Bach2+/+ and Bach2+/– mice. (b) Bach2 protein expression in splenic naïve B cells of Bach2+/+ and Bach2+/– mice. Shown is a representative example (left) and cumulative quantification (mean ± sem) (right) from n=3 independent experiments. (ce) Flow cytometry analysis of CD4+ splenocytes in Bach2+/+ and Bach2+/– mice showing percentage Foxp3+ (c), CCR9+ (d) and β7-integrin+ (e) cells. (f) IgM and IgG1 staining of B cells (upper panels) and plasma cells (lower panels) in splenocytes of Bach2+/+ and Bach2+/– mice 8 days following immunization with 4-Hydroxy-3-nitrophenylacetyl hapten-conjugated chicken gamma globulin (NP-CGG) in alum. (g) B220+Ki67+Bcl6+ germinal center B cells in splenocytes of Bach2+/+ and Bach2+/– mice 8 days after immunization with NP-CGG in alum. Shown in (c-f) are representative flow cytometry plots together with bar charts (mean ± sem). In vivo experiments were carried out twice. *p<0.05, **p<0.01, ***p<0.001 by t-test (a-b), one-way ANOVA (f) and Mann-Whitney U-test (all other panels).
Figure 7. Super-enhancer (SE)-regulated genes associate with…
Figure 7. Super-enhancer (SE)-regulated genes associate with haploinsufficiency
(a) The BACH2 locus has SE structures in multiple human immune cell types demarcated by H3K27Ac loading. Red fill denotes the presence of an SE in the BACH2 locus in a tissue. Source data are indicated. (b) Violin plots showing probability of loss of function intolerance scores in haplosufficient (HS), autosomal recessive (AR) and haploinsufficient (HI) gene sets. The white circles show median values. Source data: ExAc database. (c) Number of HS, AR or HI genes with and without associated SE architecture in humans (see also supplementary Fig. 8a and supplementary Table 3). (d) Pie charts indicating the frequency of SE (upper panels) and typical enhancer (TE; lower panels) structures in HS (left), HI (middle) and AR (right) genes. (e) Gene ontology (GO) functional annotation enrichment in HI genes. Shown are enrichment scores (blue bars) and Benjamini p-values (in orange) for the top 5 most significantly enriched terms. (f) Median probability of loss of function intolerance (black line) against SE signal size; the percentage of genes that are transcription factors (TF, red line) against SE signal size is shown in the inset. For reference, the red line asymptotes to the expected level (mean percentage of genes in the human genome that are TFs is 7.5%). Source data: ExAc and dbSuper databases. (g) Pie charts indicating the percentage of HS or HI genes that have GWAS disease associations. p-values in d and g are Fisher exact tests; NS = non-significant; GWAS = genome-wide association study.

References

    1. Bousfiha A, et al. The 2015 IUIS Phenotypic Classification for Primary Immunodeficiencies. J Clin Immunol. 2015;35:727–738.
    1. Picard C, et al. Primary Immunodeficiency Diseases: an Update on the Classification from the International Union of Immunological Societies Expert Committee for Primary Immunodeficiency 2015. J Clin Immunol. 2015;35:696–726.
    1. Arason GJ, Jorgensen GH, Ludviksson BR. Primary immunodeficiency and autoimmunity: lessons from human diseases. Scand J Immunol. 2010;71:317–328.
    1. Notarangelo LD. Primary immunodeficiencies. J Allergy Clin Immunol. 2010;125:S182–94.
    1. Conley ME, Casanova J-L. Discovery of single-gene inborn errors of immunity by next generation sequencing. Curr Opin Immunol. 2014;30:17–23.
    1. Deau M-C, et al. A human immunodeficiency caused by mutations in the PIK3R1 gene. J Clin Invest. 2015;125:1764–1765.
    1. Lo B, et al. Patients with LRBA deficiency show CTLA4 loss and immune dysregulation responsive to abatacept therapy. Science. 2015;349:436–440.
    1. Cunningham-Rundles C. The many faces of common variable immunodeficiency. Hematology Am Soc Hematol Educ Program. 2012;2012:301–305.
    1. Rieux-Laucat F, Casanova J-L. Immunology. Autoimmunity by haploinsufficiency. Science. 2014;345:1560–1561.
    1. Lo B, et al. CHAI and LATAIE: new genetic diseases of CTLA-4 checkpoint insufficiency. Blood. 2016;128:1037–1042.
    1. Vahedi G, et al. STATs shape the active enhancer landscape of T cell populations. Cell. 2012;151:981–993.
    1. Whyte WA, et al. Master Transcription Factors and Mediator Establish Super-Enhancers at Key Cell Identity Genes. Cell. 2013;153:307–319.
    1. Lovén J, et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell. 2013;153:320–334.
    1. Vahedi G, et al. Super-enhancers delineate disease-associated regulatory nodes in T cells. Nature. 2015;520:558–562.
    1. Roychoudhuri R, et al. BACH2 represses effector programs to stabilize T(reg)-mediated immune homeostasis. Nature. 2013;498:506–510.
    1. Igarashi K, Ochiai K, Itoh-Nakadai A, Muto A. Orchestration of plasma cell differentiation by Bach2 and its gene regulatory network. Immunol Rev. 2014;261:116–125.
    1. Ferreira MAR, et al. Identification of IL6R and chromosome 11q13.5 as risk loci for asthma. Lancet. 2011;378:1006–1014.
    1. Cooper JD, et al. Meta-analysis of genome-wide association study data identifies additional type 1 diabetes risk loci. Nat Genet. 2008;40:1399–1401.
    1. Franke A, et al. Genome-wide meta-analysis increases to 71 the number of confirmed Crohn's disease susceptibility loci. Nat Genet. 2010;42:1118–1125.
    1. Dubois PCA, et al. Multiple common variants for celiac disease influencing immune gene expression. Nat Genet. 2010;42:295–302.
    1. Jin Y, et al. Genome-wide association analyses identify 13 new susceptibility loci for generalized vitiligo. Nat Genet. 2012;44:676–680.
    1. International Multiple Sclerosis Genetics Consortium et al. Genetic risk and a primary role for cell-mediated immune mechanisms in multiple sclerosis. Nature. 2011;476:214–219.
    1. Nakayama Y, et al. A limited number of genes are involved in the differentiation of germinal center B cells. J Cell Biochem. 2006;99:1308–1325.
    1. Ochiai K, et al. Plasmacytic transcription factor Blimp-1 is repressed by Bach2 in B cells. J Biol Chem. 2006;281:38226–38234.
    1. Muto A, et al. The transcriptional programme of antibody class switching involves the repressor Bach2. Nature. 2004;429:566–571.
    1. Kuwahara M, et al. The Menin-Bach2 axis is critical for regulating CD4 T-cell senescence and cytokine homeostasis. Nat Commun. 2014;5:3555.
    1. Povoleri GAM, et al. Thymic versus induced regulatory T cells - who regulates the regulators? Front Immunol. 2013;4:169.
    1. Rosbrook GO, Stead MA, Carr SB, Wright SC. The structure of the Bach2 POZ-domain dimer reveals an intersubunit disulfide bond. Acta Crystallogr. D Biol Crystallogr. 2012;68:26–34.
    1. Uhlig HH, et al. The diagnostic approach to monogenic very early onset inflammatory bowel disease. Gastroenterology. 2014;147:990–1007.e3.
    1. Deane S, Selmi C, Naguwa SM, Teuber SS, Gershwin ME. Common variable immunodeficiency: etiological and treatment issues. Int Arch Allergy Immunol. 2009;150:311–324.
    1. Salzer U, Grimbacher B. Monogenetic defects in common variable immunodeficiency: what can we learn about terminal B cell differentiation? Curr Opin Rheumatol. 2006;18:377–382.
    1. Iwata M, et al. Retinoic acid imprints gut-homing specificity on T cells. Immunity. 2004;21:527–538.
    1. Cassani B, et al. Gut-Tropic T Cells That Express Integrin α4β7 and CCR9 Are Required for Induction of Oral Immune Tolerance in Mice. Gastroenterology. 2011;141:2109–2118.
    1. Igarashi K, Ochiai K, Muto A. Architecture and dynamics of the transcription factor network that regulates B-to-plasma cell differentiation. J Biochem. 2007;141:783–789.
    1. Seidman JG, Seidman C. Transcription factor haploinsufficiency: when half a loaf is not enough. J Clin Invest. 2002;109:451–455.
    1. Hnisz D, et al. Super-enhancers in the control of cell identity and disease. Cell. 2013;155:934–947.
    1. Qian J, et al. B Cell Super-Enhancers and Regulatory Clusters Recruit AID Tumorigenic Activity. Cell. 2014;159:1524–1537.
    1. Huang N, Lee I, Marcotte EM, Hurles ME. Characterising and predicting haploinsufficiency in the human genome. PLoS Genet. 2010;6:e1001154.
    1. Lek M, et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature. 2016;536:285–291.
    1. Creyghton MP, et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc Natl Acad Sci. 2010;107:21931–21936.
    1. Khan A, Zhang X. dbSUPER: a database of super-enhancers in mouse and human genome. Nucleic Acids Res. 2016;44:D164–71.
    1. Parker SCJ, et al. Chromatin stretch enhancer states drive cell-specific gene regulation and harbor human disease risk variants. Proc Natl Acad Sci. 2013;110:17921–17926.
    1. Roychoudhuri R, et al. BACH2 regulates CD8(+) T cell differentiation by controlling access of AP-1 factors to enhancers. Nat Immunol. 2016;17:851–860.
    1. Shinnakasu R, et al. Regulated selection of germinal-center cells into the memory B cell compartment. Nat Immunol. 2016;17:861–869.
    1. 1000 Genomes Project Consortium et al. A map of human genome variation from population-scale sequencing. Nature. 2010;467:1061–1073.
    1. Abolhassani H, Aghamohammadi A, Hammarstrom L. Monogenic mutations associated with IgA deficiency. Expert Rev Clin Immunol. 2016;12:1–15.
    1. Johnson ML, et al. Age-related changes in serum immunoglobulins in patients with familial IgA deficiency and common variable immunodeficiency (CVID) Clin Exp Immunol. 1997;108:477–483.
    1. Aghamohammadi A, et al. Progression of selective IgA deficiency to common variable immunodeficiency. Int Arch Allergy Immunol. 2008;147:87–92.
    1. Rutishauser RL, et al. Transcriptional repressor Blimp-1 promotes CD8(+) T cell terminal differentiation and represses the acquisition of central memory T cell properties. Immunity. 2009;31:296–308.
    1. Khoder A, et al. Regulatory B cells are enriched within the IgM memory and transitional subsets in healthy donors but are deficient in chronic GVHD. Blood. 2014;124:2034–2045.
    1. Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 2009;25:1754–1760.
    1. McKenna A, et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 2010;20:1297–1303.
    1. Wang K, Li M, Hakonarson H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 2010;38:e164–e164.
    1. Naviaux RK, Costanzi E, Haas M, Verma IM. The pCL vector system: rapid production of helper-free, high-titer, recombinant retroviruses. J Virol. 1996;70:5701–5705.
    1. Wingfield PT, et al. Biophysical and functional characterization of full-length, recombinant human tissue inhibitor of metalloproteinases-2 (TIMP-2) produced in Escherichia coli. Comparison of wild type and amino-terminal alanine appended variant with implications for the mechanism of TIMP functions. J Biol Chem. 1999;274:21362–21368.
    1. Liu X, Jian X, Boerwinkle E. dbNSFP: a lightweight database of human nonsynonymous SNPs and their functional predictions. Hum Mutat. 2011;32:894–899.
    1. Liu X, Jian X, Boerwinkle E. dbNSFP v2.0: A Database of Human Non‐synonymous SNVs and Their Functional Predictions and Annotations. Hum Mutat. 2013;34:E2393–E2402.
    1. Itan Y, et al. The mutation significance cutoff: gene-level thresholds for variant predictions. Nature Methods. 2016;13:109–110.
    1. Dang VT, Kassahn KS, Marcos AE, Ragan MA. Identification of human haploinsufficient genes and their genomic proximity to segmental duplications. Eur J Hum Genet. 2008;16:1350–1357.
    1. Huang DW, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nature Protocols. 2009;4:44–57.
    1. Huang DW, Sherman BT, Lempicki RA. Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res. 2009;37:1–13.
    1. Langmead B, Trapnell C, Pop M, Salzberg SL. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 2009;10:R25.
    1. Heinz S, et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol Cell. 2010;38:576–589.
    1. Whyte WA, et al. Master Transcription Factors and Mediator Establish Super-Enhancers at Key Cell Identity Genes. Cell. 2013;153:307–319.
    1. Bible PW, et al. PAPST, a User Friendly and Powerful Java Platform for ChIP-Seq Peak Co-Localization Analysis and Beyond. PLoS ONE. 2015;10:e0127285.
    1. Aken BL, et al. The Ensembl gene annotation system. Database (Oxford) 2016;2016:baw093.
    1. Quinlan AR, Hall IM. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010;26:841–842.
    1. Zhang H-M, et al. AnimalTFDB: a comprehensive animal transcription factor database. Nucleic Acids Res. 2012;40:D144–9.
    1. Hart M, et al. Loss of discrete memory B cell subsets is associated with impaired immunization responses in HIV-1 infection and may be a risk factor for invasive pneumococcal disease. J Immunol. 2007;178:8212–8220.
    1. Kircher M, et al. A general framework for estimating the relative pathogenicity of human genetic variants. Nat Genet. 2014;46:310–315.

Source: PubMed

3
Subscribe