Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Human RAG mutations: biochemistry and clinical implications

Key Points

  • Recombination-activating gene (RAG) mutations in humans are associated with a broad spectrum of clinical phenotypes, ranging from severe, early-onset infections to inflammation and autoimmunity.

  • There is a correlation between the severity of the clinical and immunological phenotypes and the recombination activity of the mutant RAG protein, and hypomorphic mutations that severely affect recombination activity are associated with restriction of the T cell and B cell repertoires. However, environmental factors may also contribute to determining the disease phenotype.

  • Crystal structure and cryo-electron microscopy studies have revealed the structure of the heterotetrameric RAG complex bound to DNA. Fine definition of this structure has also offered important insights into the disease-causing effects of naturally occurring RAG mutations.

  • Studies in patients and in mice have demonstrated that RAG mutations affect central and peripheral T cell and B cell tolerance, including defective expression of autoimmune regulator (AIRE), reduced number and function of regulatory T cells, impaired receptor editing and increased levels of B cell-activating factor (BAFF), allowing the rescue of self-reactive B cells.

  • A broad range of autoantibodies has been demonstrated in patients with RAG mutations presenting with inflammation and autoimmunity. Neutralizing antibodies specific for interferon-α (IFNα) and IFNω have been documented particularly in patients with a history of severe viral infections.

  • Recent data indicate that RAG expression during the early stages of lymphoid development selects cells with improved fitness. NK cells from Rag−/− mice have an activated phenotype and increased cytotoxicity. If confirmed in humans, these data may account for the high rate of graft rejection observed after unconditioned haematopoietic stem cell transplantation in patients with RAG deficiency.

Abstract

The recombination-activating gene 1 (RAG1) and RAG2 proteins initiate the V(D)J recombination process, which ultimately enables the generation of T cells and B cells with a diversified repertoire of antigen-specific receptors. Mutations of the RAG genes in humans are associated with a broad spectrum of clinical phenotypes, ranging from severe combined immunodeficiency to autoimmunity. Recently, novel insights into the phenotypic diversity of this disease have been provided by resolving the crystal structure of the RAG complex, by developing novel assays to test recombination activity of the mutant RAG proteins and by characterizing the molecular and cellular basis of immune dysregulation in patients with RAG deficiency.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Characterization and distribution of human RAG mutations.
Figure 2: Effects of mutations associated with CID–G/AI on the structure of the RAG complex.
Figure 3: RAG deficiency results in impairment of several tolerance checkpoints.
Figure 4: The interaction of genetic, immunological and environmental factors in determining the phenotype of human RAG deficiency.

Similar content being viewed by others

References

  1. Schatz, D. G., Oettinger, M. A. & Baltimore, D. The V(D)J recombination activating gene, RAG-1. Cell 59, 1035–1048 (1989).

    Article  CAS  PubMed  Google Scholar 

  2. Oettinger, M. A., Schatz, D. G., Gorka, C. & Baltimore, D. RAG-1 and RAG-2, adjacent genes that synergistically activate V(D)J recombination. Science 248, 1517–1523 (1990).

    Article  CAS  PubMed  Google Scholar 

  3. Mombaerts, P. et al. RAG-1-deficient mice have no mature B and T lymphocytes. Cell 68, 869–877 (1992).

    Article  CAS  PubMed  Google Scholar 

  4. Shinkai, Y. et al. RAG-2-deficient mice lack mature lymphocytes owing to inability to initiate V(D)J rearrangement. Cell 68, 855–867 (1992).

    Article  CAS  PubMed  Google Scholar 

  5. Schwarz, K. et al. RAG mutations in human B cell-negative SCID. Science 274, 97–99 (1996). This is the first study to report that RAG mutations in humans cause SCID.

    Article  CAS  PubMed  Google Scholar 

  6. Villa, A. et al. Partial V(D)J recombination activity leads to Omenn syndrome. Cell 93, 885–896 (1998). This is the first demonstration that hypomorphic RAG mutations are a cause of Omenn syndrome.

    Article  CAS  PubMed  Google Scholar 

  7. de Saint-Basile, G. et al. Restricted heterogeneity of T lymphocytes in combined immunodeficiency with hypereosinophilia (Omenn's syndrome). J. Clin. Invest. 87, 1352–1359 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Rieux-Laucat, F. et al. Highly restricted human T cell repertoire in peripheral blood and tissue-infiltrating lymphocytes in Omenn's syndrome. J. Clin. Invest. 102, 312–321 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Schuetz, C. et al. An immunodeficiency disease with RAG mutations and granulomas. N. Engl. J. Med. 358, 2030–2038 (2008). By identifying patients with delayed-onset disease characterized by granulomas and autoimmunity, this manuscript has broadened the spectrum of human RAG deficiency.

    Article  CAS  PubMed  Google Scholar 

  10. Kim, M. S., Lapkouski, M., Yang, W. & Gellert, M. Crystal structure of the V(D)J recombinase RAG1–RAG2. Nature 518, 507–511 (2015). This study describes the crystal structure of the heterotetrameric complex of RAG1 and RAG2 core domains, offering important insights into RAG complex function and into mechanisms of disease in patients with RAG mutations.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Shetty, K. & Schatz, D. G. Recruitment of RAG1 and RAG2 to chromatinized DNA during V(D)J recombination. Mol. Cell. Biol. 35, 3701–3713 (2015). This study investigates the specific requirement of RSS substrates with 12–23 spacers for optimal recognition and binding by RAG1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ru, H. et al. Molecular mechanism of V(D)J recombination from synaptic RAG1-RAG2 complex structures. Cell 163, 1138–1152 (2015). This report shows by cryo-electron microscopy how the RAG complex interacts with RSSs according to the 12–23 rule.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Fischer, A., Notarangelo, L. D., Neven, B., Cavazzana, M. & Puck, J. M. Severe combined immunodeficiencies and related disorders. Nat. Rev. Disease Primers http://dx.doi.org/10.1038/nrdp.2015.61 (2015).

  14. Pai, S. Y. et al. Transplantation outcomes for severe combined immunodeficiency, 2000–2009. N. Engl. J. Med. 371, 434–446 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Nicolas, N. et al. A human severe combined immunodeficiency (SCID) condition with increased sensitivity to ionizing radiations and impaired V(D)J rearrangements defines a new DNA recombination/repair deficiency. J. Exp. Med. 188, 627–634 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Bosma, G. C., Custer, R. P. & Bosma, M. J. A severe combined immunodeficiency mutation in the mouse. Nature 301, 527–530 (1983).

    Article  CAS  PubMed  Google Scholar 

  17. Fulop, G. M. & Phillips, R. A. The scid mutation in mice causes a general defect in DNA repair. Nature 347, 479–482 (1990).

    Article  CAS  PubMed  Google Scholar 

  18. Woodbine, L., Gennery, A. R. & Jeggo, P. A. The clinical impact of deficiency in DNA non-homologous end-joining. DNA Repair (Amst.) 16, 84–96 (2014).

    Article  CAS  Google Scholar 

  19. Omenn, G. S. Familial reticuloendotheliosis with eosinophilia. N. Engl. J. Med. 273, 427–432 (1965).

    Article  CAS  PubMed  Google Scholar 

  20. Schandene, L. et al. T helper type 2-like cells and therapeutic effects of interferon-γ in combined immunodeficiency with hypereosinophilia (Omenn's syndrome). Eur. J. Immunol. 23, 56–60 (1993).

    Article  CAS  PubMed  Google Scholar 

  21. Chilosi, M. et al. CD30 cell expression and abnormal soluble CD30 serum accumulation in Omenn's syndrome: evidence for a T helper 2-mediated condition. Eur. J. Immunol. 26, 329–334 (1996).

    Article  CAS  PubMed  Google Scholar 

  22. Signorini, S. et al. Intrathymic restriction and peripheral expansion of the T-cell repertoire in Omenn syndrome. Blood 94, 3468–3478 (1999).

    Article  CAS  PubMed  Google Scholar 

  23. Villa, A. et al. V(D)J recombination defects in lymphocytes due to RAG mutations: severe immunodeficiency with a spectrum of clinical presentations. Blood 97, 81–88 (2001).

    Article  CAS  PubMed  Google Scholar 

  24. Shearer, W. T. et al. Establishing diagnostic criteria for severe combined immunodeficiency disease (SCID), leaky SCID, and Omenn syndrome: the Primary Immune Deficiency Treatment Consortium experience. J. Allergy Clin. Immunol. 133, 1092–1098 (2014).

    Article  PubMed  Google Scholar 

  25. de Villartay, J. P. et al. A novel immunodeficiency associated with hypomorphic RAG1 mutations and CMV infection. J. Clin. Invest. 115, 3291–3299 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ehl, S. et al. A variant of SCID with specific immune responses and predominance of γδ T cells. J. Clin. Invest. 115, 3140–3148 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Kwan, A. et al. Newborn screening for severe combined immunodeficiency in 11 screening programs in the United States. JAMA 312, 729–738 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Al-Herz, W. & Al-Mousa, H. Combined immunodeficiency: the Middle East experience. J. Allergy Clin. Immunol. 131, 658–660 (2013).

    Article  PubMed  Google Scholar 

  29. Corneo, B. et al. Identical mutations in RAG1 or RAG2 genes leading to defective V(D)J recombinase activity can cause either T-B–severe combined immune deficiency or Omenn syndrome. Blood 97, 2772–2776 (2001).

    Article  CAS  PubMed  Google Scholar 

  30. Ijspeert, H. et al. Similar recombination-activating gene (RAG) mutations result in similar immunobiological effects but in different clinical phenotypes. J. Allergy Clin. Immunol. 133, 1124–1133 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Dalal, I. et al. Evolution of a T-B- SCID into an Omenn syndrome phenotype following parainfluenza 3 virus infection. Clin. Immunol. 115, 70–73 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Wada, T. et al. Oligoclonal expansion of T lymphocytes with multiple second-site mutations leads to Omenn syndrome in a patient with RAG1-deficient severe combined immunodeficiency. Blood 106, 2099–2101 (2005).

    Article  CAS  PubMed  Google Scholar 

  33. De Ravin, S. S. et al. Hypomorphic Rag mutations can cause destructive midline granulomatous disease. Blood 116, 1263–1271 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Avila, E. M. et al. Highly variable clinical phenotypes of hypomorphic RAG1 mutations. Pediatrics 126, e1248–e1252 (2010).

    Article  PubMed  Google Scholar 

  35. Henderson, L. A. et al. Expanding the spectrum of recombination-activating gene 1 deficiency: a family with early-onset autoimmunity. J. Allergy Clin. Immunol. 132, 969–971 (2013).

    Article  CAS  PubMed  Google Scholar 

  36. Walter, J. E. et al. Broad-spectrum antibodies against self-antigens and cytokines in RAG deficiency. J. Clin. Invest. 125, 4135–4148 (2015). This article reports on the autoantibody repertoire in a large series of patients with RAG mutations, including neutralizing IFN α - and IFN ω -specific antibodies in patients with CID–G/AI and a history of severe varicella zoster virus infection.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Sharapova, S. O. et al. Late-onset combined immune deficiency associated to skin granuloma due to heterozygous compound mutations in RAG1 gene in a 14 years old male. Hum. Immunol. 74, 18–22 (2013).

    Article  CAS  PubMed  Google Scholar 

  38. Patiroglu, T. et al. Atypical severe combined immunodeficiency caused by a novel homozygous mutation in Rag1 gene in a girl who presented with pyoderma gangrenosum: a case report and literature review. J. Clin. Immunol. 34, 792–795 (2014).

    Article  PubMed  Google Scholar 

  39. Chen, K. et al. Autoimmunity due to RAG deficiency and estimated disease incidence in RAG1/2 mutations. J. Allergy Clin. Immunol. 133, 880–882 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Buchbinder, D. et al. Identification of patients with RAG mutations previously diagnosed with common variable immunodeficiency disorders. J. Clin. Immunol. 35, 119–124 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Kuijpers, T. W. et al. Idiopathic CD4+ T lymphopenia without autoimmunity or granulomatous disease in the slipstream of RAG mutations. Blood 117, 5892–5896 (2011).

    Article  CAS  PubMed  Google Scholar 

  42. Abolhassani, H. et al. A hypomorphic recombination-activating gene 1 (RAG1) mutation resulting in a phenotype resembling common variable immunodeficiency. J. Allergy Clin. Immunol. 134, 1375–1380 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Kato, T. et al. RAG1 deficiency may present clinically as selective IgA deficiency. J. Clin. Immunol. 35, 280–288 (2015).

    Article  CAS  PubMed  Google Scholar 

  44. Geier, C. B. et al. Leaky RAG deficiency in adult patients with impaired antibody production against bacterial polysaccharide antigens. PLoS ONE 10, e0133220 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Chou, J. et al. A novel homozygous mutation in recombination activating gene 2 in 2 relatives with different clinical phenotypes: Omenn syndrome and hyper-IgM syndrome. J. Allergy Clin. Immunol. 130, 1414–1416 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Reiff, A. et al. Exome sequencing reveals RAG1 mutations in a child with autoimmunity and sterile chronic multifocal osteomyelitis evolving into disseminated granulomatous disease. J. Clin. Immunol. 33, 1289–1292 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Kuo, T. C. & Schlissel, M. S. Mechanisms controlling expression of the RAG locus during lymphocyte development. Curr. Opin. Immunol. 21, 173–178 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Landree, M. A., Wibbenmeyer, J. A. & Roth, D. B. Mutational analysis of RAG1 and RAG2 identifies three catalytic amino acids in RAG1 critical for both cleavage steps of V(D)J recombination. Genes Dev. 13, 3059–3069 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Kim, D. R., Dai, Y., Mundy, C. L., Yang, W. & Oettinger, M. A. Mutations of acidic residues in RAG1 define the active site of the V(D)J recombinase. Genes Dev. 13, 3070–3080 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Bellon, S. F., Rodgers, K. K., Schatz, D. G., Coleman, J. E. & Steitz, T. A. Crystal structure of the RAG1 dimerization domain reveals multiple zinc-binding motifs including a novel zinc binuclear cluster. Nat. Struct. Biol. 4, 586–591 (1997).

    Article  CAS  PubMed  Google Scholar 

  51. Matthews, A. G. et al. RAG2 PHD finger couples histone H3 lysine 4 trimethylation with V(D)J recombination. Nature 450, 1106–1110 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Cortes, P., Ye, Z. S. & Baltimore, D. RAG-1 interacts with the repeated amino acid motif of the human homologue of the yeast protein SRP1. Proc. Natl Acad. Sci. USA 91, 7633–7637 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Deng, Z., Liu, H. & Liu, X. RAG1-mediated ubiquitylation of histone H3 is required for chromosomal V(D)J recombination. Cell Res. 25, 181–192 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Liu, Y., Subrahmanyam, R., Chakraborty, T., Sen, R. & Desiderio, S. A plant homeodomain in RAG-2 that binds hypermethylated lysine 4 of histone H3 is necessary for efficient antigen-receptor-gene rearrangement. Immunity 27, 561–571 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Shimazaki, N., Tsai, A. G. & Lieber, M. R. H3K4me3 stimulates the V(D)J RAG complex for both nicking and hairpinning in trans in addition to tethering in cis: implications for translocations. Mol. Cell 34, 535–544 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Jiang, H. et al. Ubiquitylation of RAG-2 by Skp2-SCF links destruction of the V(D)J recombinase to the cell cycle. Mol. Cell 18, 699–709 (2005).

    Article  CAS  PubMed  Google Scholar 

  57. Branzei, D. & Foiani, M. Regulation of DNA repair throughout the cell cycle. Nat. Rev. Mol. Cell Biol. 9, 297–308 (2008).

    Article  CAS  PubMed  Google Scholar 

  58. Lee, Y. N. et al. A systematic analysis of recombination activity and genotype-phenotype correlation in human recombination-activating gene 1 deficiency. J. Allergy Clin. Immunol. 133, 1099–1108 (2014). This comprehensive analysis of expression and recombination activity of naturally occurring human RAG1 mutant proteins reveals a genotype–phenotype correlation in this disease.

    Article  CAS  PubMed  Google Scholar 

  59. Mo, X., Bailin, T. & Sadofsky, M. J. A. C-terminal region of RAG1 contacts the coding DNA during V(D)J recombination. Mol. Cell. Biol. 21, 2038–2047 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Wong, S. Y., Lu, C. P. & Roth, D. B. A. RAG1 mutation found in Omenn syndrome causes coding flank hypersensitivity: a novel mechanism for antigen receptor repertoire restriction. J. Immunol. 181, 4124–4130 (2008).

    Article  CAS  PubMed  Google Scholar 

  61. Santagata, S. et al. N-terminal RAG1 frameshift mutations in Omenn's syndrome: internal methionine usage leads to partial V(D)J recombination activity and reveals a fundamental role in vivo for the N-terminal domains. Proc. Natl Acad. Sci. USA 97, 14572–14577 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Simkus, C., Anand, P., Bhattacharyya, A. & Jones, J. M. Biochemical and folding defects in a RAG1 variant associated with Omenn syndrome. J. Immunol. 179, 8332–8340 (2007).

    Article  CAS  PubMed  Google Scholar 

  63. Couedel, C. et al. Analysis of mutations from SCID and Omenn syndrome patients reveals the central role of the Rag2 PHD domain in regulating V(D)J recombination. J. Clin. Invest. 120, 1337–1344 (2010). This study reveals the crucial regulatory role of the non-core PHD region of RAG2 by showing that mutations in the PHD compromise recombination activity by affecting protein stability, nuclear localization and/or interaction with histones.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Itan, Y. et al. The human gene damage index as a gene-level approach to prioritizing exome variants. Proc. Natl Acad. Sci. USA 112, 13615–13620 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Noordzij, J. G. et al. The immunophenotypic and immunogenotypic B-cell differentiation arrest in bone marrow of RAG-deficient SCID patients corresponds to residual recombination activities of mutated RAG proteins. Blood 100, 2145–2152 (2002).

    CAS  PubMed  Google Scholar 

  66. Yu, X. et al. Human syndromes of immunodeficiency and dysregulation are characterized by distinct defects in T-cell receptor repertoire development. J. Allergy Clin. Immunol. 133, 1109–1115 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Wesemann, D. R. et al. Immature B cells preferentially switch to IgE with increased direct Sμ to Sɛ recombination. J. Exp. Med. 208, 2733–2746 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Cassani, B. et al. Homeostatic expansion of autoreactive immunoglobulin-secreting cells in the Rag2 mouse model of Omenn syndrome. J. Exp. Med. 207, 1525–1540 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Walter, J. E. et al. Expansion of immunoglobulin-secreting cells and defects in B cell tolerance in Rag-dependent immunodeficiency. J. Exp. Med. 207, 1541–1554 (2010). References 68 and 69 demonstrate in animal models and in patients that hypomorphic RAG mutations are associated with abnormalities of central and peripheral B cell tolerance, resulting in the expansion of autoantibody-secreting cell populations.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Anderson, M. S. et al. Projection of an immunological self shadow within the thymus by the Aire protein. Science 298, 1395–1401 (2002).

    Article  CAS  PubMed  Google Scholar 

  71. Watanabe, N. et al. Hassall's corpuscles instruct dendritic cells to induce CD4+CD25+ regulatory T cells in human thymus. Nature 436, 1181–1185 (2005).

    Article  CAS  PubMed  Google Scholar 

  72. Akiyama, T., Tateishi, R., Akiyama, N., Yoshinaga, R. & Kobayashi, T. J. Positive and negative regulatory mechanisms for fine-tuning cellularity and functions of medullary thymic epithelial cells. Front. Immunol. 6, 461 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Cavadini, P. et al. AIRE deficiency in thymus of 2 patients with Omenn syndrome. J. Clin. Invest. 115, 728–732 (2005). This is the first demonstration that RAG mutations in humans compromise thymic architecture and expression of AIRE and AIRE-dependent TRAs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Poliani, P. L. et al. Early defects in human T-cell development severely affect distribution and maturation of thymic stromal cells: possible implications for the pathophysiology of Omenn syndrome. Blood 114, 105–108 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Rucci, F. et al. Abnormalities of thymic stroma may contribute to immune dysregulation in murine models of leaky severe combined immunodeficiency. Front. Immunol. 2, 15 (2011).

    Article  PubMed Central  Google Scholar 

  76. Marrella, V. et al. Anti-CD3ɛ mAb improves thymic architecture and prevents autoimmune manifestations in a mouse model of Omenn syndrome: therapeutic implications. Blood 120, 1005–1014 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Marrella, V. et al. A hypomorphic R229Q Rag2 mouse mutant recapitulates human Omenn syndrome. J. Clin. Invest. 117, 1260–1269 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Meager, A. et al. Anti-interferon autoantibodies in autoimmune polyendocrinopathy syndrome type 1. PLoS Med. 3, e289 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Cassani, B. et al. Defect of regulatory T cells in patients with Omenn syndrome. J. Allergy Clin. Immunol. 125, 209–216 (2010).

    Article  CAS  PubMed  Google Scholar 

  80. Matangkasombut, P. et al. Lack of iNKT cells in patients with combined immune deficiency due to hypomorphic RAG mutations. Blood 111, 271–274 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Jankovic, M., Casellas, R., Yannoutsos, N., Wardemann, H. & Nussenzweig, M. C. RAGs and regulation of autoantibodies. Annu. Rev. Immunol. 22, 485–501 (2004).

    Article  CAS  PubMed  Google Scholar 

  82. Lesley, R. et al. Reduced competitiveness of autoantigen-engaged B cells due to increased dependence on BAFF. Immunity 20, 441–453 (2004).

    Article  CAS  PubMed  Google Scholar 

  83. Thien, M. et al. Excess BAFF rescues self-reactive B cells from peripheral deletion and allows them to enter forbidden follicular and marginal zone niches. Immunity 20, 785–798 (2004).

    Article  CAS  PubMed  Google Scholar 

  84. Borghesi, L. et al. B lineage-specific regulation of V(D)J recombinase activity is established in common lymphoid progenitors. J. Exp. Med. 199, 491–502 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Pilbeam, K. et al. The ontogeny and fate of NK cells marked by permanent DNA rearrangements. J. Immunol. 180, 1432–1441 (2008).

    Article  CAS  PubMed  Google Scholar 

  86. Karo, J. M., Schatz, D. G. & Sun, J. C. The RAG recombinase dictates functional heterogeneity and cellular fitness in natural killer cells. Cell 159, 94–107 (2014). This study demonstrates that RAG expression during the early stages of lymphoid development allows selection of cells with higher fitness and improved ability to respond to infections and cellular stress.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Schuetz, C. et al. SCID patients with ARTEMIS versus RAG deficiencies following HCT: increased risk of late toxicity in ARTEMIS-deficient SCID. Blood 123, 281–289 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Hacein- Bey-Abina, S. et al. Efficacy of gene therapy for X-linked severe combined immunodeficiency. N. Engl. J. Med. 363, 355–364 (2010).

    Article  Google Scholar 

  89. Hacein- Bey-Abina, S. et al. A modified γ-retrovirus vector for X-linked severe combined immunodeficiency. N. Engl. J. Med. 371, 1407–1417 (2014).

    Article  CAS  Google Scholar 

  90. Aiuti, A. et al. Gene therapy for immunodeficiency due to adenosine deaminase deficiency. N. Engl. J. Med. 360, 447–458 (2009).

    Article  CAS  PubMed  Google Scholar 

  91. Candotti, F. et al. Gene therapy for adenosine deaminase-deficient severe combined immune deficiency: clinical comparison of retroviral vectors and treatment plans. Blood 120, 3635–3646 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Pike-Overzet, K. et al. Correction of murine Rag1 deficiency by self-inactivating lentiviral vector-mediated gene transfer. Leukemia 25, 1471–1483 (2011).

    Article  CAS  PubMed  Google Scholar 

  93. Lagresle-Peyrou, C. et al. Long-term immune reconstitution in RAG-1-deficient mice treated by retroviral gene therapy: a balance between efficiency and toxicity. Blood 107, 63–72 (2006).

    Article  CAS  PubMed  Google Scholar 

  94. van Til, N. P. et al. Recombination-activating gene 1 (Rag1)-deficient mice with severe combined immunodeficiency treated with lentiviral gene therapy demonstrate autoimmune Omenn-like syndrome. J. Allergy Clin. Immunol. 133, 1116–1123 (2014).

    Article  CAS  PubMed  Google Scholar 

  95. Yates, F. et al. Gene therapy of RAG-2−/− mice: sustained correction of the immunodeficiency. Blood 100, 3942–3949 (2002).

    Article  CAS  PubMed  Google Scholar 

  96. van Til, N. P. et al. Correction of murine Rag2 severe combined immunodeficiency by lentiviral gene therapy using a codon-optimized RAG2 therapeutic transgene. Mol. Ther. 20, 1968–1980 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Feeney, A. J., Goebel, P. & Espinoza, C. R. Many levels of control of V gene rearrangement frequency. Immunol. Rev. 200, 44–56 (2004).

    Article  CAS  PubMed  Google Scholar 

  98. Teng, G. et al. RAG represents a widespread threat to the lymphocyte genome. Cell 162, 751–765 (2015). This study demonstrates that RAG1 associates with chromatin at a very large number of promoters and enhancers in developing lymphocytes. During evolution, protection from illegitimate DNA cleavage and genomic instability has been accomplished by reducing the number of cryptic RSS heptamer sequences near RAG1-binding sites that are not associated with TCR and immunoglobulin genes.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Hesse, J. E., Lieber, M. R., Gellert, M. & Mizuuchi, K. Extrachromosomal DNA substrates in pre-B cells undergo inversion or deletion at immunoglobulin V-(D)-J joining signals. Cell 19, 775–783 (1987).

    Article  Google Scholar 

  100. Gauss, G. H. & Lieber, M. R. Unequal signal and coding joint formation in human V(D)J recombination. Mol. Cell. Biol. 13, 3900–3906 (1987).

    Article  Google Scholar 

  101. Liang, H. E. et al. The “dispensable” portion of RAG2 is necessary for efficient V-to-DJ rearrangement during B and T cell development. Immunity 17, 639–651 (2002).

    Article  CAS  PubMed  Google Scholar 

  102. Bredemeyer, A. L. et al. ATM stabilizes DNA double-strand-break complexes during V(D)J recombination. Nature 442, 466–470 (2006).

    Article  CAS  PubMed  Google Scholar 

  103. Rigoni, R. et al. Intestinal microbiota sustains inflammation and autoimmunity induced by hypomorphic RAG defects. J. Exp. Med. 213, 355–375 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

L.D.N. was supported by grants from the National Institute of Allergy and Infectious Diseases, US National Institutes of Health (NIH; 5R01AI100887) and the March of Dimes (1-FY13-500). J.E.W. was supported by the NIH grant 5K08AI103035.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Luigi D. Notarangelo.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

PowerPoint slides

Supplementary information

Supplementary information S1 (table)

Missense RAG1 and RAG2 mutations: protein expression, recombination activity, structural effects and in vivo clinical phenotype (PDF 364 kb)

Glossary

Non-homologous end joining pathway

(NHEJ pathway). An error-prone pathway that mediates joining of DNA double-strand breaks without requiring a homologous template. In mammals, the NHEJ pathway involves several proteins — Ku70, Ku80, DNA-dependent protein kinase catalytic subunit (DNA-PKcs), Artemis, Cernunnos (also known as XLF and NHEJ1), X-ray repair cross-complementing protein 4 (XRCC4) and DNA ligase IV. Genetic defects of Artemis are the most common cause of radiosensitive severe combined immunodeficiency (SCID) in humans.

Haematopoietic stem cell transplantation

(HSCT). A therapeutic procedure that involves transfusion of donor HSCs into a recipient. HLA matching between the recipient and the donor determines compatibility. HSCT from a haploidentical donor (such as a parent) is associated with a high risk of graft-versus-host disease, unless T cells are depleted. Chemotherapy is often used before HSCT to eliminate the recipient's blood cells and favour engraftment of donor-derived HSCs but is not strictly necessary in infants with severe combined immunodeficiency (SCID).

Cellular radiosensitivity

Susceptibility of cells to the damaging effects of ionizing radiation, resulting in genomic instability, tumour development or cell death. Genetic defects that affect mechanisms of repair of DNA double-strand breaks are associated with increased cellular radiosensitivity.

SCID mouse

An animal model of severe combined immunodeficiency (SCID) that is characterized by lack of T cells and B cells, and increased cellular radiosensitivity. The SCID mouse carries mutations of the Prkdc gene, encoding DNA-dependent protein kinase catalytic subunit (DNA-PKcs).

Purifying selection

In population genetics, purifying selection refers to the selective removal of deleterious alleles from a given population. Purging of these genetic variants occurs when they cause early death or affect the reproductive fitness of affected individuals.

CDR3 spectratyping

(Complementarity-determining region 3 spectratyping). A PCR-based method that measures the diversity of T cell and B cell repertoires, based on the length of the CDR3 of immunoglobulin and T cell receptor transcripts. A Gaussian distribution of CDR3 length is detected in polyclonal T cells and B cells, whereas a single peak is observed in patients with leukaemia or lymphoma, and an altered distribution may be detected in patients with infections, autoimmune diseases or severe impairment of T cell and/or B cell development.

Autoimmune polyendocrinopathy candidiasis and ectodermal dystrophy

(APECED). A monogenic autoimmune disease caused by mutations of the autoimmune regulator (AIRE) gene, affecting central T cell tolerance. Common manifestations of this disease include autoimmune hypoparathyroidism, Addison disease, type 1 diabetes, candidiasis, alopecia and nail dystrophy.

Codon-optimized

Transgenic products that are generated through a process replacing the original codons with synonymous codons for which a more abundant tRNA is available. This facilitates the rate of translation and ultimately results in the production of higher amounts of the protein.

Whole-exome sequencing

A process by which all exons contained in the genome (collectively comprising the exome) are amplified and subjected to high-throughput sequencing. DNA genetic variants that are present in a given individual are identified by comparing the exome of that subject to the normal reference sequence.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Notarangelo, L., Kim, MS., Walter, J. et al. Human RAG mutations: biochemistry and clinical implications. Nat Rev Immunol 16, 234–246 (2016). https://doi.org/10.1038/nri.2016.28

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nri.2016.28

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing