Entry - *179616 - RECOMBINATION-ACTIVATING GENE 2; RAG2 - OMIM

* 179616

RECOMBINATION-ACTIVATING GENE 2; RAG2


HGNC Approved Gene Symbol: RAG2

Cytogenetic location: 11p12     Genomic coordinates (GRCh38): 11:36,590,996-36,598,236 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
11p12 Combined cellular and humoral immune defects with granulomas 233650 AR 3
Omenn syndrome 603554 AR 3
Severe combined immunodeficiency, B cell-negative 601457 AR 3


TEXT

Description

RAG1 (179615) and RAG2 initiate the V(D)J recombination process, in which the variable (V), diversity (D), and joining (J) coding elements of the immunoglobulin (Ig) and T-cell receptor (TCR) genes are joined together to generate antigen-specific B- and T-cell receptors. Two molecules of RAG1 and 2 molecules of RAG2 form a heterotetramer that binds to recombination signal sequences (RSSs) flanking the V, D, and J genes and introduces double-strand breaks in the DNA, which are subsequently repaired by the nonhomologous end-joining DNA repair pathway. RAG1 contains the RNase H fold catalytic domain and regions that make direct contact with RSSs and is responsible for the enzymatic activity of the RAG complex. RAG2 promotes DNA binding by scanning the genome for enzymatic signatures characterized by histone-3 (H3; see 602810) trimethylated at lys4 (H3K4me3) and facilitates the cleavage functions of RAG1 (review by Bosticardo et al., 2021).


Cloning and Expression

Oettinger et al. (1990) demonstrated that cotransfection of the RAG1 gene (179615) with an adjacent gene, RAG2, results in at least a 1,000-fold increase in the frequency of V(D)J recombination, whereas RAG1 alone inefficiently induces recombinase activity. Oettinger et al. (1990) reported that the 2.1-kb RAG2 cDNA encodes a putative protein of 527 amino acids whose sequence is unrelated to that of RAG1. Like RAG1, RAG2 is conserved between species that carry out V(D)J recombination, and its pattern of expression correlates precisely with that of V(D)J recombinase activity. It is likely that the RAG1 and RAG2 gene products participate directly in the recombination reaction.


Gene Function

Lin and Desiderio (1994) found that in an immature B-cell line and in normal thymocytes, RAG2 protein accumulated preferentially in the G0/G1 phases of the cell cycle and declined by at least 20-fold before cells entered S phase. The amount of RAG2 protein remained low throughout the S, G2, and M phases. The amount of RAG1 protein showed considerably less fluctuation. The variation in RAG2 protein is likely to be established, at least in part, by a posttranscriptional mechanism. These observations suggested to Lin and Desiderio (1994) that V(D)J rearrangement occurs entirely or preferentially within G0/G1.

Formation of double-strand breaks at recombination signal sequences is an early step in V(D)J recombination. McBlane et al. (1995) showed that purified RAG1 and RAG2 proteins are sufficient to carry out this reaction. The cleavage reaction can be divided into 2 distinct steps, nicking and hairpin formation, each of which requires the presence of a signal sequence and both RAG proteins.

By site-directed mutagenesis of acidic amino acid residues in RAG1 and RAG2, Landree et al. (1999) and Kim et al. (1999) identified 3 RAG1 mutants that retained normal binding of recombination signal sequences but were catalytically inactive for both nicking and hairpin formation. The data suggested that 1 active site in RAG1 performs both of these steps and that at least one of these amino acids contacts and coordinates a metal ion, which is required for cleavage. The results also suggested that RAG1 contains most, if not all, of the active site of the RAG1/RAG2 V(D)J recombinase.

Hikida et al. (1996) reported that RAG1 and RAG2 are expressed in mature mouse B cells after culture with interleukin-4 (147780) in association with costimuli (lipopolysaccharide and other cytokines). Reexpression was also detected in draining lymph nodes from immunized mice. Hikida et al. (1996) noted that previously reported studies had indicated that RAG1 and RAG2 were expressed only in immature B cells.

Yu et al. (1999) investigated the regulation of RAG1 and RAG2 in vivo with bacterial artificial chromosome (BAC) transgenes containing a fluorescent indicator. Coordinate expression of RAG1 and RAG2 in B and T cells was regulated by distinct genetic elements found on the 5-prime side of the RAG2 gene. This observation suggested a mechanism by which asymmetrically disposed cis DNA elements could influence the expression of the primordial transposon and thereby capture RAGs for vertebrate evolution.

During development of B and T cells, the RAG1/RAG2 protein complex cleaves DNA at conserved recombination signal sequences (RSS) to initiate V(D)J recombination. RAG1/RAG2 also catalyzes transpositional strand transfer of RSS-containing substrates into target DNA to form branched DNA intermediates. Melek and Gellert (2000) showed that RAG1/RAG2 can resolve these intermediates by 2 pathways. RAG1/RAG2 catalyzes hairpin formation on target DNA adjacent to transposed RSS ends in a manner consistent with a model leading to chromosome translocations. Alternatively, disintegration removes transposed donor DNA from the intermediate. At high magnesium concentrations, such as those present in mammalian cells, disintegration is the favored pathway of resolution. The authors suggested that this may explain in part why RAG1/RAG2-mediated transposition does not occur at high frequency in cells.

The 'kelch' motif, named after a sequence first identified in Drosophila, is a 44- to 56-amino acid segment with low primary sequence identity. It is characterized by the presence of 4 hydrophobic residues followed by a double-glycine element and, after variable spacing, tyrosine and tryptophan residues separated by 6 amino acids. Typically, 4 to 7 of these motifs form a kelch repeat domain and a beta propeller, with each motif forming a 4-stranded beta sheet corresponding to 1 blade of the propeller (Adams et al., 2000). By hydrophobic cluster and gapped-BLAST analysis, Callebaut and Mornon (1998) determined that the N-terminal 355 residues of RAG2 are composed of a 6-fold kelch repeat forming a 6-bladed propeller in the active core. They proposed that the propeller structure of RAG2 could serve as a binding scaffold for RAG1 (179615) on one side and for DNA on the other.

Qiu et al. (2001) used site-directed mutagenesis targeting each conserved basic amino acid in RAG2, which revealed several separation-of-function mutants. Analysis of these mutants showed that RAG2 helps recognize or cleave distorted DNA intermediates and plays an essential role in the joining step of V(D)J recombination. Moreover, the discovery that some mutants blocked RAG-mediated hairpin opening in vitro provided a critical link between this biochemical activity and coding joint formation in vivo.

Corneo et al. (2007) found that removing certain portions of murine Rag proteins revealed robust alternative nonhomologous end-joining (NHEJ) activity in NHEJ-deficient cells and some alternative joining activity even in wildtype cells. Corneo et al. (2007) proposed a 2-tier model in which the Rag proteins collaborate with NHEJ factors to preserve genomic integrity during V(D)J recombination.

Matthews et al. (2007) showed that RAG2 contains a plant homeodomain (PHD) finger that specifically recognizes H3K4me3. The high-resolution crystal structure of the mouse RAG2 PHD finger bound to H3K4me3 reveals the molecular basis of H3K4me3 recognition by RAG2. Mutations that abrogate RAG2's recognition of H3K4me3 severely impaired V(D)J recombination in vivo. Reducing the level of H3K4me3 similarly led to a decrease in V(D)J recombination in vivo. Notably, a conserved tryptophan residue (W453) that constitutes a key structural component of the K4me3-binding surface and is essential for RAG2's recognition of H3K4me3 is mutated in patients with immunodeficiency syndromes (Omenn syndrome; 603554). Taken together, Matthews et al. (2007) concluded that their results identified a novel function for histone methylation in mammalian DNA recombination. Furthermore, their results provided the first evidence indicating that disrupting the read-out of histone modifications can cause an inherited human disease.

Deriano et al. (2011) showed that the RAG2 C terminus, although dispensable for recombination, is critical for maintaining genomic stability. Thymocytes from 'core' Rag2 homozygous (Rag2c/c) mice show dramatic disruption of Tcr-alpha (TCRA; see 186880)/delta (see 186810) locus integrity. Furthermore, all Rag2c/c p53 (191170)-null mice, unlike Rag1c/c p53-null and p53-null animals, rapidly develop thymic lymphomas bearing complex chromosomal translocations, amplifications, and deletions involving the Tcr-alpha/delta and Igh (147100) loci. Deriano et al. (2011) also found these features in lymphomas from Atm-null mice. Deriano et al. (2011) showed that, like ATM (607585) deficiency, core RAG2 severely destabilizes the RAG postcleavage complex. Deriano et al. (2011) concluded that their results revealed a novel genome guardian role for RAG2 and suggested that similar 'end release/end persistence' mechanisms underlie genomic instability and lymphomagenesis in Rag2c/c p53-null and Atm-null mice.

Using chromatin immunoprecipitation analysis, Ji et al. (2010) demonstrated that mouse Rag protein binding was tightly regulated during lymphocyte development, focusing on a small region encompassing J and, where present, J-proximal D gene segments in IgH, Igk (see 147200), Tcrb (see 186930), and Tcra loci. These regions, which the authors termed recombination centers, were rich in activating histone modifications and RNA polymerase II (see 180660). Rag2 bound broadly in the genome at sites with substantial trimethylation at lys4 of H3 (see 601128). In contrast, Rag1 binding was more specific, occurring primarily with recombination signal sequences (RSS) flanking V, D, and J gene segments. Ji et al. (2010) proposed that recombination centers are specialized sites of high local RAG concentration that facilitate RSS binding and synapsis and help regulate recombination order.


Gene Structure

Oettinger et al. (1990) reported that the genomic size of RAG2 is approximately 18 kb. The convergently transcribed RAG1 and RAG2 genes are unusual in that most, if not all, of their coding and 3-prime untranslated sequences are contained in a single exon.


Biochemical Features

Crystal Structure

Matthews et al. (2007) determined the crystal structure of the RAG2 PHD-H3K4me3 complex at 1.15-angstrom resolution. The structure revealed that, instead of being closed on both sides, the back, and the top (as observed with other PHD fingers), the RAG2 PHD K4me3-binding surface is open on the top, and resembles an 'aromatic channel' rather than an 'aromatic cage.' The authors suggested that this 'channel' conformation may provide a mechanism to modulate histone binding.

Kim et al. (2015) reported the crystal structure of the mouse RAG1 (179615)-RAG2 complex at 3.2-angstrom resolution. The 230-kD RAG1-RAG2 heterotetramer is Y-shaped, with the amino-terminal domains of the 2 RAG1 chains forming an intertwined stalk. Each RAG1-RAG2 heterodimer composes 1 arm of the Y, with the active site in the middle and RAG2 at its tip. The RAG1-RAG2 structure rationalizes more than 60 mutations identified in immunodeficient patients, as well as a large body of genetic and biochemical data.


Mapping

Oettinger et al. (1990) found that RAG1 and RAG2 are only 8 kb apart.


Molecular Genetics

Schwarz et al. (1996) reported that patients with severe combined immunodeficiency can be divided into those with B lymphocytes (B-positive SCID) and those without (B-negative SCID; 601457). They searched for RAG1 and RAG2 mutations in B-negative SCID patients through the use of SSCP analysis with primer cassettes overlapping the entire RAG1 and RAG2 coding regions. Six of 14 B-negative SCID patients were found to carry mutations of the recombinase activating genes. Mutations resulted in a functional inability to form antigen receptors through genetic recombination. In 2 families, 3 patients exhibited an altered migration pattern for RAG2 amplimers. The PCR products were then sequenced. Two related patients were found to be homozygous for a missense mutation leading to cys476-to-tyr mutation in RAG2 (179616.0001). One patient was found to have inherited a RAG2 missense mutation (R229Q; 179616.0002) from the mother and a deletion involving RAG1 and RAG2 from the father. Transient transfection assays revealed that the SCID-associated RAG2 mutations exhibited either a complete loss or a marked reduction of V(D)J recombination activity.

Villa et al. (1998) reported that patients with Omenn syndrome (603554), a severe immunodeficiency characterized by the presence of activated, anergic, oligoclonal T cells, hypereosinophilia, and high IgE levels, have missense mutations in either the RAG1 (179615) or RAG2 genes that result in partial activity of the 2 proteins. Two of the amino acid substitutions map within the RAG1 homeodomain and decrease DNA binding activity, while 3 others lower the efficiency of RAG1/RAG2 interaction. These findings provided evidence indicating that the immunodeficiency manifested in patients with Omenn syndrome arises from mutations that decrease the efficiency of V(D)J recombination.

Gomez et al. (2000) identified a gly95-to-arg mutation (179616.0005) and a deletion of ile273 (179616.0006) within the predicted second beta strand of repeats 2 and 5 of the RAG2 kelch domain that led to Omenn syndrome and SCID, respectively, in 2 patients. By confocal microscopy analysis, they determined that the mutations did not impair nuclear localization but did reduce the capacity of RAG2 to interact with RAG1 and to mediate recombination signal cleavage. Furthermore, by analysis of a panel of mutants, they showed that the hydrophobic and gly-rich regions within the second strand of the beta sheet are critical for RAG1-RAG2 interaction.

Tabori et al. (2004) performed mutation analyses of PCR products of the RAG1 and RAG2 genes in 6 cases of T-negative/B-negative SCID and 8 cases of Omenn syndrome. Consanguinity was reported in 7 of the 14 families. None of the patients had a mutation in the RAG1 gene, but Tabori et al. (2004) found 4 missense mutations in the RAG2 gene in 6 of 8 Omenn syndrome patients and in 4 of 6 SCID patients (see 179616.0007).

Yu et al. (2014) performed deep sequencing on complementarity-determining region-3 (CDR3) of TCR-beta in CD4 (186940)-positive and CD8 (see 186910)-positive T cells from 2 patients with autoimmunity and/or granulomatous disease, but not severe immunodeficiency, caused by RAG1 or IL2RG (308380) mutations; 5 patients with Omenn syndrome caused by RAG1 or RAG2 mutations; 2 patients with Omenn syndrome-like phenotypes caused by a ZAP70 (176947) mutation (see 269840) or by atypical DiGeorge syndrome (188400); and 4 healthy controls. They found that patients with Omenn syndrome due to RAG1 or RAG2 mutations had poor TCR-beta diversity compared with controls and patients with Omenn syndrome not due to RAG1 or RAG2 mutations. The 2 patients with RAG1 or IL2RG mutations associated with autoimmunity and granulomatous disease did not have diminished diversity, but instead had skewed V-J pairing and CDR3 amino acid use. Yu et al. (2014) concluded that RAG enzymatic function may be necessary for normal CDR3 junctional diversity and that aberrant TCR generation, but not numeric diversity, may contribute to immune dysregulation in patients with hypomorphic forms of SCID.


Animal Model

In transgenic mice carrying a germline mutation in which a large portion in the RAG2 coding region was deleted, Shinkai et al. (1992) found that although homozygotes were viable, they failed to produce mature B or T lymphocytes. Immature lymphoid cells were present in primary lymphoid organs; however, these cells did not rearrange their immunoglobulin or T-cell receptor loci. Thus, loss of RAG2 function results in total inability to initiate V(D)J rearrangement, leading to a severe combined immunodeficiency (SCID) phenotype. Since the SCID phenotype was the only obvious abnormality detected in these mice, RAG2 function and V(D)J recombinase activity, per se, must not be required for development of cells other than lymphocytes.

Shankaran et al. (2001) found that mice lacking the lymphocyte-specific Rag2 gene, the Ifn receptor signal transcription factor Stat1 (600555), Ifngr1 (107470), or both Rag2 and Stat1, are significantly more susceptible to chemically induced tumor formation than wildtype mice, suggesting that T, NKT, and/or B cells are essential to suppress development of chemically induced tumors. Spontaneous malignant tumors did not occur in wildtype mice, occurred late in half of mice lacking either Rag2 or Stat1, but occurred early in 82% of mice lacking both genes. Transplanted chemically induced tumors from lymphocyte-deficient mice (Shankaran et al., 2001) or from Ifng-unresponsive mice (Kaplan et al., 1998), but not tumors from immunocompetent hosts, were rejected by wildtype mice, indicating that the tumors from immunodeficient mice are more immunogenic and that lymphocytes and the IFNG/STAT1 signaling pathway collaborate to shape the immunogenic phenotype of tumors that eventually form in immunocompetent hosts. Shankaran et al. (2001) proposed that tumors are imprinted by the immunologic environment in which they form and that 'cancer immunoediting' rather than 'immunosurveillance' best describes the protective and sculpting actions of the immune response on developing tumors.

Rideout et al. (2002) used immune-deficient Rag2 -/- mice as nuclear donors for transfer into enucleated oocytes and cultured the resulting blastocysts to isolate an isogenic embryonic stem (ES) cell line. One of the mutated alleles in the Rag2 -/- ES cells was repaired by homologous recombination, thereby restoring normal Rag2 gene structure. Mutant mice were treated with the repaired ES cells in 2 ways: (1) immune-competent mice were generated from the repaired ES cells by tetraploid embryo complementation and were used as bone marrow donors for transplantation, and (2) hematopoietic precursors were derived by in vitro differentiation from the repaired ES cells and engrafted into mutant mice. Mature myeloid and lymphoid cells as well as immunoglobulins became detectable 3 to 4 weeks after transplantation. These results established a paradigm for the treatment of a genetic disorder by combining therapeutic cloning with gene therapy.

Marrella et al. (2007) generated a knockin mouse model in which endogenous Rag2 was replaced with Rag2 carrying the R229Q mutation identified in patients with Omenn syndrome and SCID. These mice showed T-cell oligoclonality, a lack of circulating B cells, and peripheral eosinophilia. In addition, T-cell infiltration of gut and skin caused diarrhea, alopecia, and, in some mice, severe erythrodermia. The findings were associated with reduced thymic expression of Aire (607358) and markedly reduced regulatory T cells and NKT lymphocytes. Marrella et al. (2007) concluded that Rag2 R229Q homozygous mice mimic most symptoms of human Omenn syndrome and that the pathophysiology of Omenn syndrome involves impaired immune tolerance and defective immune regulation.


ALLELIC VARIANTS ( 10 Selected Examples):

.0001 SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

RAG2, CYS476TYR
  
RCV000014009...

Schwarz et al. (1996) found 2 related patients with B-negative SCID (601457) who were homozygous for a cys476-to-tyr substitution mutation in RAG2 caused by a 2634G-A transition.


.0002 SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

OMENN SYNDROME, INCLUDED
RAG2, ARG229GLN
  
RCV000014010...

In a patient with T-, B- SCID (601457), Schwarz et al. (1996) identified compound heterozygosity for 2 mutations in the RAG2 gene: a 1887G-A transition, resulting in an arg229-to-gln (R229Q) substitution, and a deletion encompassing RAG1 and RAG2. A polymorphism in the RAG1 gene (A156V; 179615.0004) was also identified.

In a patient with T-, B- SCID, Corneo et al. (2001) identified compound heterozygosity for the R229Q mutation and R39G (179616.0008). A sib with Omenn syndrome (603554) had the same genotype.


.0003 OMENN SYNDROME

RAG2, CYS41TRP
  
RCV000014012...

Villa et al. (1998) found a cys41-to-trp mutation in heterozygous state in a patient with Omenn syndrome (603554). This mutation occurs in a domain with HimA homology.


.0004 OMENN SYNDROME

RAG2, MET285ARG
  
RCV000014013...

Villa et al. (1998) found a met285-to-arg mutation in heterozygous state in a patient with Omenn syndrome (603554). This mutation occurs in a domain with topoII homology.


.0005 OMENN SYNDROME

RAG2, GLY95ARG
  
RCV000014014...

Gomez et al. (2000) identified a heterozygous G-to-A transition at nucleotide 1484 of the RAG2 gene, leading to a gly95-to-arg substitution, in a male patient with Omenn syndrome (603554). The patient died at age 5 months.


.0006 SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

RAG2, 3-BP DEL, NT2018
  
RCV000014015

In a 2-month-old male patient with B cell-negative (CD19 (107265) less than 1%) SCID (601457) with maternal T-cell engraftment, Gomez et al. (2000) identified a homozygous 3-bp (nucleotides 2018 to 2020) in-frame deletion in the RAG2 gene. The mutation resulted in the removal of ile273 with the remainder of the protein intact. The patient received a bone marrow transplant and was alive and well 2 years later.


.0007 SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

RAG2, TRP215ILE
  
RCV000014016...

In a patient with T cell-negative/B cell-negative SCID (601457), Tabori et al. (2004) identified an 1845C-T transition in the RAG2 gene, resulting in a trp215-to-ile (W215I) change.


.0008 SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

OMENN SYNDROME, INCLUDED
RAG2, ARG39GLY
  
RCV000014017...

In a patient with T-, B- SCID (601457), Corneo et al. (2001) identified compound heterozygosity for 2 mutations in the RAG2 gene: a 1316A-G transition, resulting in an arg39-to-gly (R39G) substitution, and R229Q (179616.0002). A sib with Omenn syndrome (603554) had the same genotype.


.0009 COMBINED CELLULAR AND HUMORAL IMMUNE DEFECTS WITH GRANULOMAS

RAG2, THR77ASN
  
RCV000014019...

In a patient with combined cellular and humoral immune defects associated with granulomas (233650), Schuetz et al. (2008) identified compound heterozygous mutations in the RAG2 gene: a thr77-to-asn (T77N) substitution in the catalytic core of the protein and a gly451-to-ala (G451A; 179616.0010) substitution in the PHD-like region. The patient presented at age 10 years with a history of severe infections and massive splenomegaly. She was found to have hypogammaglobulinemia and defective T-cell function. Noninfectious granulomas were present in the spleen and lungs. In vitro functional expression studies showed that the mutant proteins had significantly impaired function. Schuetz et al. (2008) concluded that the relatively late onset and low incidence of repeated infections observed in this patient reflected a low level of residual RAG2 activity.


.0010 COMBINED CELLULAR AND HUMORAL IMMUNE DEFECTS WITH GRANULOMAS

RAG2, GLY451ALA
  
RCV000014020...

For discussion of the gly451-to-ala (G451A) mutation in the RAG2 gene that was found in compound heterozygous state in a patient with combined cellular and humoral immune defects associated with granulomas (233650) by Schuetz et al. (2008), see 179616.0009.


REFERENCES

  1. Adams, J., Kelso, R., Cooley, L. The kelch repeat superfamily of proteins: propellers of cell function. Trends Cell Biol. 10: 17-24, 2000. [PubMed: 10603472, related citations] [Full Text]

  2. Bosticardo, M., Pala, F., Notarangelo, L. D. RAG deficiencies: recent advances in disease pathogenesis and novel therapeutic approaches. Europ. J. Immun. 51: 1028-1038, 2021. [PubMed: 33682138, images, related citations] [Full Text]

  3. Callebaut, I., Mornon, J.-P. The V(D)J recombination activating protein RAG2 consists of a six-bladed propeller and a PHD fingerlike domain, as revealed by sequence analysis. Cell. Molec. Life Sci. 54: 880-891, 1998. [PubMed: 9760994, related citations] [Full Text]

  4. Corneo, B., Moshous, D., Gungor, T., Wulffraat, N., Philippet, P., Le Deist, F., Fischer, A., de Villartay, J.-P. 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. [PubMed: 11313270, related citations] [Full Text]

  5. Corneo, B., Wendland, R. L., Deriano, L., Cui, X., Klein, I. A., Wong, S.-Y., Arnal, S., Holub, A. J., Weller, G. R., Pancake, B. A., Shah, S., Brandt, V. L., Meek, K., Roth, D. B. Rag mutations reveal robust alternative end joining. Nature 449: 483-486, 2007. [PubMed: 17898768, related citations] [Full Text]

  6. Deriano, L., Chaumeil, J., Coussens, M., Multani, A., Chou, Y., Alekseyenko, A. V., Chang, S., Skok, J. A., Roth, D. B. The RAG2 C terminus suppresses genomic instability and lymphomagenesis. Nature 471: 119-123, 2011. [PubMed: 21368836, images, related citations] [Full Text]

  7. Gomez, C. A., Ptaszek, L. M., Villa, A., Bozzi, F., Sobacchi, C., Brooks, E. G., Notarangelo, L. D., Spanopoulou, E., Pan, Z. Q., Vezzoni, P., Cortes, P., Santagata, S. Mutations in conserved regions of the predicted RAG2 kelch repeats block initiation of V(D)J recombination and result in primary immunodeficiencies. Molec. Cell. Biol. 20: 5653-5664, 2000. [PubMed: 10891502, images, related citations] [Full Text]

  8. Hikida, M., Mori, M., Takai, T., Tomochika, K., Hamatani, K., Ohmori, H. Reexpression of RAG-1 and RAG-2 genes in activated mature mouse B cells. Science 274: 2092-2094, 1996. [PubMed: 8953042, related citations] [Full Text]

  9. Ji, Y., Resch, W., Corbett, E., Yamane, A., Casellas, R., Schatz, D. G. The in vivo pattern of binding of RAG1 and RAG2 to antigen receptor loci. Cell 141: 419-431, 2010. Note: Erratum: Cell 143: 170 only, 2010. [PubMed: 20398922, images, related citations] [Full Text]

  10. Kaplan, D. H., Shankaran, V., Dighe, A. S., Stockert, E., Aguet, M., Old, L. J., Schreiber, R. D. Demonstration of an interferon gamma-dependent tumor surveillance system in immunocompetent mice. Proc. Nat. Acad. Sci. 95: 7556-7561, 1998. [PubMed: 9636188, images, related citations] [Full Text]

  11. 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. [PubMed: 10601033, images, related citations] [Full Text]

  12. Kim, M.-S., Lapkouski, M., Yang, W., Gellert, M. Crystal structure of the V(D)J recombinase RAG1-RAG2. Nature 518: 507-511, 2015. [PubMed: 25707801, images, related citations] [Full Text]

  13. 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. [PubMed: 10601032, images, related citations] [Full Text]

  14. Lin, W.-C., Desiderio, S. Cell cycle regulation of V(D)J recombination-activating protein RAG-2. Proc. Nat. Acad. Sci. 91: 2733-2737, 1994. [PubMed: 8146183, related citations] [Full Text]

  15. Marrella, V., Poliani, P. L., Casati, A., Rucci, F., Frascoli, L., Gougeon, M.-L., Lemercier, B., Bosticardo, M., Ravanini, M., Battaglia, M., Roncarolo, M. G., Cavazzana-Calvo, M., Facchetti, F., Notarangelo, L. D., Vezzoni, P., Grassi, F., Villa, A. A hypomorphic R229Q Rag2 mouse mutant recapitulates human Omenn syndrome. J. Clin. Invest. 117: 1260-1269, 2007. [PubMed: 17476358, images, related citations] [Full Text]

  16. Matthews, A. G. W., Kuo, A. J., Ramon-Maiques, S., Han, S., Champagne, K. S., Ivanov, D., Gallardo, M., Carney, D., Cheung, P., Ciccone, D. N., Walter, K. L., Utz, P. J., Shi, Y., Kutateladze, T. G., Yang, W., Gozani, O., Oettinger, M. A. RAG2 PHD finger couples histone H3 lysine 4 trimethylation with V(D)J recombination. Nature 450: 1106-1110, 2007. [PubMed: 18033247, images, related citations] [Full Text]

  17. McBlane, J. F., van Gent, D. C., Ramsden, D. A., Romeo, C., Cuomo, C. A., Gellert, M., Oettinger, M. A. Cleavage at a V(D)J recombination signal requires only RAG1 and RAG2 proteins and occurs in two steps. Cell 83: 387-395, 1995. [PubMed: 8521468, related citations] [Full Text]

  18. Melek, M., Gellert, M. RAG1/2-mediated resolution of transposition intermediates: two pathways and possible consequences. Cell 101: 625-633, 2000. [PubMed: 10892649, related citations] [Full Text]

  19. 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. [PubMed: 2360047, related citations] [Full Text]

  20. Qiu, J., Kale, S. B., Schultz, H. Y., Roth, D. B. Separation-of-function mutants reveal critical roles for RAG2 in both the cleavage and joining steps of V(D)J recombination. Molec. Cell 7: 77-87, 2001. [PubMed: 11172713, related citations] [Full Text]

  21. Rideout, W. M., III, Hochedlinger, K., Kyba, M., Daley, G. Q., Jaenisch, R. Correction of a genetic defect by nuclear transplantation and combined cell and gene therapy. Cell 109: 17-27, 2002. [PubMed: 11955443, related citations] [Full Text]

  22. Schuetz, C., Huck, K., Gudowius, S., Megahed, M., Feyen, O., Hubner, B., Schneider, D. T., Manfras, B., Pannicke, U., Willemze, R., Knuchel, R., Gobel, U., Schulz, A., Borkhardt, A., Friedrich, W., Schwarz, K., Niehues, T. An immunodeficiency disease with RAG mutations and granulomas. New Eng. J. Med. 358: 2030-2038, 2008. [PubMed: 18463379, related citations] [Full Text]

  23. Schwarz, K., Gauss, G. H., Ludwig, L., Pannicke, U., Li, Z., Linder, D., Friedrich, W., Seger, R. A., Hansen-Hagge, T. E., Desiderio, S., Lieber, M. R., Bartram, C. R. RAG mutations in human B cell-negative SCID. Science 274: 97-99, 1996. [PubMed: 8810255, related citations] [Full Text]

  24. Shankaran, V., Ikeda, H., Bruce, A. T., White, J. M., Swanson, P. E., Old, L. J., Schreiber, R. D. IFN-gamma and lymphocytes prevent primary tumour development and shape tumour immunogenicity. Nature 410: 1107-1111, 2001. [PubMed: 11323675, related citations] [Full Text]

  25. Shinkai, Y., Rathbun, G., Lam, K.-P., Oltz, E. M., Stewart, V., Mendelsohn, M., Charron, J., Datta, M., Young, F., Stall, A. M., Alt, F. W. RAG-2-deficient mice lack mature lymphocytes owing to inability to initiate V(D)J rearrangement. Cell 68: 855-867, 1992. [PubMed: 1547487, related citations] [Full Text]

  26. Tabori, U., Mark, Z., Amariglio, N., Etzioni, A., Golan, H., Biloray, B., Toren, A., Rechavi, G., Dalal, I. Detection of RAG mutations and prenatal diagnosis in families presenting with either T-B- severe combined immunodeficiency or Omenn's syndrome. Clin. Genet. 65: 322-326, 2004. [PubMed: 15025726, related citations] [Full Text]

  27. Villa, A., Santagata, S., Bozzi, F., Giliani, S., Frattini, A., Imberti, L., Gatta, L. B., Ochs, H. D., Schwarz, K., Notarangelo, L. D., Vezzoni, P., Spanopoulou, E. Partial V(D)J recombination activity leads to Omenn syndrome. Cell 93: 885-896, 1998. [PubMed: 9630231, related citations] [Full Text]

  28. Yu, W., Misulovin, Z., Suh, H., Hardy, R. R., Jankovic, M., Yannoutsos, N., Nussenzweig, M. C. Coordinate regulation of RAG1 and RAG2 by cell type-specific DNA elements 5-prime of RAG2. Science 285: 1080-1084, 1999. [PubMed: 10446057, related citations] [Full Text]

  29. Yu, X., Almeida, J., Darko, S., van der Burg, M., DeRavin, S. S., Malech, H., Gennery, A., Chinn, I., Markert, M. L., Douek, D. C., Milner, J. D. Human syndromes of immunodeficiency and dysregulation are characterized by distinct defects in T-cell receptor repertoire development. J. Allergy Clin. Immun. 133: 1109-1115, 2014. [PubMed: 24406074, images, related citations] [Full Text]


Matthew B. Gross - updated : 05/01/2024
Ada Hamosh - updated : 02/03/2016
Paul J. Converse - updated : 1/23/2014
Paul J. Converse - updated : 8/19/2011
Ada Hamosh - updated : 6/10/2011
Cassandra L. Kniffin - updated : 5/19/2008
Ada Hamosh - updated : 4/22/2008
Paul J. Converse - updated : 10/24/2007
Ada Hamosh - updated : 10/11/2007
Cassandra L. Kniffin - updated : 10/20/2004
Victor A. McKusick - updated : 4/22/2004
Stylianos E. Antonarakis - updated : 5/3/2002
Paul J. Converse - updated : 4/25/2001
Stylianos E. Antonarakis - updated : 2/1/2001
Paul J. Converse - updated : 10/31/2000
Stylianos E. Antonarakis - updated : 6/21/2000
Paul J. Converse - updated : 4/13/2000
Ada Hamosh - updated : 8/12/1999
Stylianos E. Antonarakis - updated : 6/29/1998
Moyra Smith - updated : 12/19/1996
Moyra Smith - updated : 10/3/1996
Creation Date:
Victor A. McKusick : 7/9/1990
mgross : 05/01/2024
alopez : 07/31/2023
alopez : 02/03/2016
mcolton : 7/21/2015
mcolton : 1/30/2015
mgross : 2/11/2014
mgross : 2/11/2014
mgross : 2/11/2014
mcolton : 1/23/2014
mgross : 10/7/2013
mgross : 10/4/2013
mgross : 2/5/2013
mgross : 9/19/2011
mgross : 9/19/2011
terry : 8/19/2011
alopez : 6/20/2011
terry : 6/10/2011
wwang : 5/20/2008
ckniffin : 5/19/2008
alopez : 5/15/2008
terry : 4/22/2008
mgross : 11/1/2007
terry : 10/24/2007
alopez : 10/16/2007
terry : 10/11/2007
carol : 10/28/2004
terry : 10/28/2004
ckniffin : 10/20/2004
tkritzer : 4/23/2004
terry : 4/22/2004
mgross : 5/3/2002
alopez : 4/25/2001
mgross : 2/1/2001
mgross : 10/31/2000
mgross : 6/21/2000
carol : 4/13/2000
alopez : 8/12/1999
alopez : 8/12/1999
carol : 2/18/1999
terry : 2/18/1999
carol : 10/5/1998
carol : 6/29/1998
alopez : 7/10/1997
mark : 12/19/1996
mark : 10/5/1996
mark : 10/4/1996
mark : 10/4/1996
mark : 10/3/1996
mark : 12/4/1995
jason : 7/15/1994
carol : 8/11/1992
carol : 6/2/1992
carol : 6/1/1992
carol : 5/29/1992
carol : 5/26/1992

* 179616

RECOMBINATION-ACTIVATING GENE 2; RAG2


HGNC Approved Gene Symbol: RAG2

SNOMEDCT: 307650006, 722067005;  


Cytogenetic location: 11p12     Genomic coordinates (GRCh38): 11:36,590,996-36,598,236 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
11p12 Combined cellular and humoral immune defects with granulomas 233650 Autosomal recessive 3
Omenn syndrome 603554 Autosomal recessive 3
Severe combined immunodeficiency, B cell-negative 601457 Autosomal recessive 3

TEXT

Description

RAG1 (179615) and RAG2 initiate the V(D)J recombination process, in which the variable (V), diversity (D), and joining (J) coding elements of the immunoglobulin (Ig) and T-cell receptor (TCR) genes are joined together to generate antigen-specific B- and T-cell receptors. Two molecules of RAG1 and 2 molecules of RAG2 form a heterotetramer that binds to recombination signal sequences (RSSs) flanking the V, D, and J genes and introduces double-strand breaks in the DNA, which are subsequently repaired by the nonhomologous end-joining DNA repair pathway. RAG1 contains the RNase H fold catalytic domain and regions that make direct contact with RSSs and is responsible for the enzymatic activity of the RAG complex. RAG2 promotes DNA binding by scanning the genome for enzymatic signatures characterized by histone-3 (H3; see 602810) trimethylated at lys4 (H3K4me3) and facilitates the cleavage functions of RAG1 (review by Bosticardo et al., 2021).


Cloning and Expression

Oettinger et al. (1990) demonstrated that cotransfection of the RAG1 gene (179615) with an adjacent gene, RAG2, results in at least a 1,000-fold increase in the frequency of V(D)J recombination, whereas RAG1 alone inefficiently induces recombinase activity. Oettinger et al. (1990) reported that the 2.1-kb RAG2 cDNA encodes a putative protein of 527 amino acids whose sequence is unrelated to that of RAG1. Like RAG1, RAG2 is conserved between species that carry out V(D)J recombination, and its pattern of expression correlates precisely with that of V(D)J recombinase activity. It is likely that the RAG1 and RAG2 gene products participate directly in the recombination reaction.


Gene Function

Lin and Desiderio (1994) found that in an immature B-cell line and in normal thymocytes, RAG2 protein accumulated preferentially in the G0/G1 phases of the cell cycle and declined by at least 20-fold before cells entered S phase. The amount of RAG2 protein remained low throughout the S, G2, and M phases. The amount of RAG1 protein showed considerably less fluctuation. The variation in RAG2 protein is likely to be established, at least in part, by a posttranscriptional mechanism. These observations suggested to Lin and Desiderio (1994) that V(D)J rearrangement occurs entirely or preferentially within G0/G1.

Formation of double-strand breaks at recombination signal sequences is an early step in V(D)J recombination. McBlane et al. (1995) showed that purified RAG1 and RAG2 proteins are sufficient to carry out this reaction. The cleavage reaction can be divided into 2 distinct steps, nicking and hairpin formation, each of which requires the presence of a signal sequence and both RAG proteins.

By site-directed mutagenesis of acidic amino acid residues in RAG1 and RAG2, Landree et al. (1999) and Kim et al. (1999) identified 3 RAG1 mutants that retained normal binding of recombination signal sequences but were catalytically inactive for both nicking and hairpin formation. The data suggested that 1 active site in RAG1 performs both of these steps and that at least one of these amino acids contacts and coordinates a metal ion, which is required for cleavage. The results also suggested that RAG1 contains most, if not all, of the active site of the RAG1/RAG2 V(D)J recombinase.

Hikida et al. (1996) reported that RAG1 and RAG2 are expressed in mature mouse B cells after culture with interleukin-4 (147780) in association with costimuli (lipopolysaccharide and other cytokines). Reexpression was also detected in draining lymph nodes from immunized mice. Hikida et al. (1996) noted that previously reported studies had indicated that RAG1 and RAG2 were expressed only in immature B cells.

Yu et al. (1999) investigated the regulation of RAG1 and RAG2 in vivo with bacterial artificial chromosome (BAC) transgenes containing a fluorescent indicator. Coordinate expression of RAG1 and RAG2 in B and T cells was regulated by distinct genetic elements found on the 5-prime side of the RAG2 gene. This observation suggested a mechanism by which asymmetrically disposed cis DNA elements could influence the expression of the primordial transposon and thereby capture RAGs for vertebrate evolution.

During development of B and T cells, the RAG1/RAG2 protein complex cleaves DNA at conserved recombination signal sequences (RSS) to initiate V(D)J recombination. RAG1/RAG2 also catalyzes transpositional strand transfer of RSS-containing substrates into target DNA to form branched DNA intermediates. Melek and Gellert (2000) showed that RAG1/RAG2 can resolve these intermediates by 2 pathways. RAG1/RAG2 catalyzes hairpin formation on target DNA adjacent to transposed RSS ends in a manner consistent with a model leading to chromosome translocations. Alternatively, disintegration removes transposed donor DNA from the intermediate. At high magnesium concentrations, such as those present in mammalian cells, disintegration is the favored pathway of resolution. The authors suggested that this may explain in part why RAG1/RAG2-mediated transposition does not occur at high frequency in cells.

The 'kelch' motif, named after a sequence first identified in Drosophila, is a 44- to 56-amino acid segment with low primary sequence identity. It is characterized by the presence of 4 hydrophobic residues followed by a double-glycine element and, after variable spacing, tyrosine and tryptophan residues separated by 6 amino acids. Typically, 4 to 7 of these motifs form a kelch repeat domain and a beta propeller, with each motif forming a 4-stranded beta sheet corresponding to 1 blade of the propeller (Adams et al., 2000). By hydrophobic cluster and gapped-BLAST analysis, Callebaut and Mornon (1998) determined that the N-terminal 355 residues of RAG2 are composed of a 6-fold kelch repeat forming a 6-bladed propeller in the active core. They proposed that the propeller structure of RAG2 could serve as a binding scaffold for RAG1 (179615) on one side and for DNA on the other.

Qiu et al. (2001) used site-directed mutagenesis targeting each conserved basic amino acid in RAG2, which revealed several separation-of-function mutants. Analysis of these mutants showed that RAG2 helps recognize or cleave distorted DNA intermediates and plays an essential role in the joining step of V(D)J recombination. Moreover, the discovery that some mutants blocked RAG-mediated hairpin opening in vitro provided a critical link between this biochemical activity and coding joint formation in vivo.

Corneo et al. (2007) found that removing certain portions of murine Rag proteins revealed robust alternative nonhomologous end-joining (NHEJ) activity in NHEJ-deficient cells and some alternative joining activity even in wildtype cells. Corneo et al. (2007) proposed a 2-tier model in which the Rag proteins collaborate with NHEJ factors to preserve genomic integrity during V(D)J recombination.

Matthews et al. (2007) showed that RAG2 contains a plant homeodomain (PHD) finger that specifically recognizes H3K4me3. The high-resolution crystal structure of the mouse RAG2 PHD finger bound to H3K4me3 reveals the molecular basis of H3K4me3 recognition by RAG2. Mutations that abrogate RAG2's recognition of H3K4me3 severely impaired V(D)J recombination in vivo. Reducing the level of H3K4me3 similarly led to a decrease in V(D)J recombination in vivo. Notably, a conserved tryptophan residue (W453) that constitutes a key structural component of the K4me3-binding surface and is essential for RAG2's recognition of H3K4me3 is mutated in patients with immunodeficiency syndromes (Omenn syndrome; 603554). Taken together, Matthews et al. (2007) concluded that their results identified a novel function for histone methylation in mammalian DNA recombination. Furthermore, their results provided the first evidence indicating that disrupting the read-out of histone modifications can cause an inherited human disease.

Deriano et al. (2011) showed that the RAG2 C terminus, although dispensable for recombination, is critical for maintaining genomic stability. Thymocytes from 'core' Rag2 homozygous (Rag2c/c) mice show dramatic disruption of Tcr-alpha (TCRA; see 186880)/delta (see 186810) locus integrity. Furthermore, all Rag2c/c p53 (191170)-null mice, unlike Rag1c/c p53-null and p53-null animals, rapidly develop thymic lymphomas bearing complex chromosomal translocations, amplifications, and deletions involving the Tcr-alpha/delta and Igh (147100) loci. Deriano et al. (2011) also found these features in lymphomas from Atm-null mice. Deriano et al. (2011) showed that, like ATM (607585) deficiency, core RAG2 severely destabilizes the RAG postcleavage complex. Deriano et al. (2011) concluded that their results revealed a novel genome guardian role for RAG2 and suggested that similar 'end release/end persistence' mechanisms underlie genomic instability and lymphomagenesis in Rag2c/c p53-null and Atm-null mice.

Using chromatin immunoprecipitation analysis, Ji et al. (2010) demonstrated that mouse Rag protein binding was tightly regulated during lymphocyte development, focusing on a small region encompassing J and, where present, J-proximal D gene segments in IgH, Igk (see 147200), Tcrb (see 186930), and Tcra loci. These regions, which the authors termed recombination centers, were rich in activating histone modifications and RNA polymerase II (see 180660). Rag2 bound broadly in the genome at sites with substantial trimethylation at lys4 of H3 (see 601128). In contrast, Rag1 binding was more specific, occurring primarily with recombination signal sequences (RSS) flanking V, D, and J gene segments. Ji et al. (2010) proposed that recombination centers are specialized sites of high local RAG concentration that facilitate RSS binding and synapsis and help regulate recombination order.


Gene Structure

Oettinger et al. (1990) reported that the genomic size of RAG2 is approximately 18 kb. The convergently transcribed RAG1 and RAG2 genes are unusual in that most, if not all, of their coding and 3-prime untranslated sequences are contained in a single exon.


Biochemical Features

Crystal Structure

Matthews et al. (2007) determined the crystal structure of the RAG2 PHD-H3K4me3 complex at 1.15-angstrom resolution. The structure revealed that, instead of being closed on both sides, the back, and the top (as observed with other PHD fingers), the RAG2 PHD K4me3-binding surface is open on the top, and resembles an 'aromatic channel' rather than an 'aromatic cage.' The authors suggested that this 'channel' conformation may provide a mechanism to modulate histone binding.

Kim et al. (2015) reported the crystal structure of the mouse RAG1 (179615)-RAG2 complex at 3.2-angstrom resolution. The 230-kD RAG1-RAG2 heterotetramer is Y-shaped, with the amino-terminal domains of the 2 RAG1 chains forming an intertwined stalk. Each RAG1-RAG2 heterodimer composes 1 arm of the Y, with the active site in the middle and RAG2 at its tip. The RAG1-RAG2 structure rationalizes more than 60 mutations identified in immunodeficient patients, as well as a large body of genetic and biochemical data.


Mapping

Oettinger et al. (1990) found that RAG1 and RAG2 are only 8 kb apart.


Molecular Genetics

Schwarz et al. (1996) reported that patients with severe combined immunodeficiency can be divided into those with B lymphocytes (B-positive SCID) and those without (B-negative SCID; 601457). They searched for RAG1 and RAG2 mutations in B-negative SCID patients through the use of SSCP analysis with primer cassettes overlapping the entire RAG1 and RAG2 coding regions. Six of 14 B-negative SCID patients were found to carry mutations of the recombinase activating genes. Mutations resulted in a functional inability to form antigen receptors through genetic recombination. In 2 families, 3 patients exhibited an altered migration pattern for RAG2 amplimers. The PCR products were then sequenced. Two related patients were found to be homozygous for a missense mutation leading to cys476-to-tyr mutation in RAG2 (179616.0001). One patient was found to have inherited a RAG2 missense mutation (R229Q; 179616.0002) from the mother and a deletion involving RAG1 and RAG2 from the father. Transient transfection assays revealed that the SCID-associated RAG2 mutations exhibited either a complete loss or a marked reduction of V(D)J recombination activity.

Villa et al. (1998) reported that patients with Omenn syndrome (603554), a severe immunodeficiency characterized by the presence of activated, anergic, oligoclonal T cells, hypereosinophilia, and high IgE levels, have missense mutations in either the RAG1 (179615) or RAG2 genes that result in partial activity of the 2 proteins. Two of the amino acid substitutions map within the RAG1 homeodomain and decrease DNA binding activity, while 3 others lower the efficiency of RAG1/RAG2 interaction. These findings provided evidence indicating that the immunodeficiency manifested in patients with Omenn syndrome arises from mutations that decrease the efficiency of V(D)J recombination.

Gomez et al. (2000) identified a gly95-to-arg mutation (179616.0005) and a deletion of ile273 (179616.0006) within the predicted second beta strand of repeats 2 and 5 of the RAG2 kelch domain that led to Omenn syndrome and SCID, respectively, in 2 patients. By confocal microscopy analysis, they determined that the mutations did not impair nuclear localization but did reduce the capacity of RAG2 to interact with RAG1 and to mediate recombination signal cleavage. Furthermore, by analysis of a panel of mutants, they showed that the hydrophobic and gly-rich regions within the second strand of the beta sheet are critical for RAG1-RAG2 interaction.

Tabori et al. (2004) performed mutation analyses of PCR products of the RAG1 and RAG2 genes in 6 cases of T-negative/B-negative SCID and 8 cases of Omenn syndrome. Consanguinity was reported in 7 of the 14 families. None of the patients had a mutation in the RAG1 gene, but Tabori et al. (2004) found 4 missense mutations in the RAG2 gene in 6 of 8 Omenn syndrome patients and in 4 of 6 SCID patients (see 179616.0007).

Yu et al. (2014) performed deep sequencing on complementarity-determining region-3 (CDR3) of TCR-beta in CD4 (186940)-positive and CD8 (see 186910)-positive T cells from 2 patients with autoimmunity and/or granulomatous disease, but not severe immunodeficiency, caused by RAG1 or IL2RG (308380) mutations; 5 patients with Omenn syndrome caused by RAG1 or RAG2 mutations; 2 patients with Omenn syndrome-like phenotypes caused by a ZAP70 (176947) mutation (see 269840) or by atypical DiGeorge syndrome (188400); and 4 healthy controls. They found that patients with Omenn syndrome due to RAG1 or RAG2 mutations had poor TCR-beta diversity compared with controls and patients with Omenn syndrome not due to RAG1 or RAG2 mutations. The 2 patients with RAG1 or IL2RG mutations associated with autoimmunity and granulomatous disease did not have diminished diversity, but instead had skewed V-J pairing and CDR3 amino acid use. Yu et al. (2014) concluded that RAG enzymatic function may be necessary for normal CDR3 junctional diversity and that aberrant TCR generation, but not numeric diversity, may contribute to immune dysregulation in patients with hypomorphic forms of SCID.


Animal Model

In transgenic mice carrying a germline mutation in which a large portion in the RAG2 coding region was deleted, Shinkai et al. (1992) found that although homozygotes were viable, they failed to produce mature B or T lymphocytes. Immature lymphoid cells were present in primary lymphoid organs; however, these cells did not rearrange their immunoglobulin or T-cell receptor loci. Thus, loss of RAG2 function results in total inability to initiate V(D)J rearrangement, leading to a severe combined immunodeficiency (SCID) phenotype. Since the SCID phenotype was the only obvious abnormality detected in these mice, RAG2 function and V(D)J recombinase activity, per se, must not be required for development of cells other than lymphocytes.

Shankaran et al. (2001) found that mice lacking the lymphocyte-specific Rag2 gene, the Ifn receptor signal transcription factor Stat1 (600555), Ifngr1 (107470), or both Rag2 and Stat1, are significantly more susceptible to chemically induced tumor formation than wildtype mice, suggesting that T, NKT, and/or B cells are essential to suppress development of chemically induced tumors. Spontaneous malignant tumors did not occur in wildtype mice, occurred late in half of mice lacking either Rag2 or Stat1, but occurred early in 82% of mice lacking both genes. Transplanted chemically induced tumors from lymphocyte-deficient mice (Shankaran et al., 2001) or from Ifng-unresponsive mice (Kaplan et al., 1998), but not tumors from immunocompetent hosts, were rejected by wildtype mice, indicating that the tumors from immunodeficient mice are more immunogenic and that lymphocytes and the IFNG/STAT1 signaling pathway collaborate to shape the immunogenic phenotype of tumors that eventually form in immunocompetent hosts. Shankaran et al. (2001) proposed that tumors are imprinted by the immunologic environment in which they form and that 'cancer immunoediting' rather than 'immunosurveillance' best describes the protective and sculpting actions of the immune response on developing tumors.

Rideout et al. (2002) used immune-deficient Rag2 -/- mice as nuclear donors for transfer into enucleated oocytes and cultured the resulting blastocysts to isolate an isogenic embryonic stem (ES) cell line. One of the mutated alleles in the Rag2 -/- ES cells was repaired by homologous recombination, thereby restoring normal Rag2 gene structure. Mutant mice were treated with the repaired ES cells in 2 ways: (1) immune-competent mice were generated from the repaired ES cells by tetraploid embryo complementation and were used as bone marrow donors for transplantation, and (2) hematopoietic precursors were derived by in vitro differentiation from the repaired ES cells and engrafted into mutant mice. Mature myeloid and lymphoid cells as well as immunoglobulins became detectable 3 to 4 weeks after transplantation. These results established a paradigm for the treatment of a genetic disorder by combining therapeutic cloning with gene therapy.

Marrella et al. (2007) generated a knockin mouse model in which endogenous Rag2 was replaced with Rag2 carrying the R229Q mutation identified in patients with Omenn syndrome and SCID. These mice showed T-cell oligoclonality, a lack of circulating B cells, and peripheral eosinophilia. In addition, T-cell infiltration of gut and skin caused diarrhea, alopecia, and, in some mice, severe erythrodermia. The findings were associated with reduced thymic expression of Aire (607358) and markedly reduced regulatory T cells and NKT lymphocytes. Marrella et al. (2007) concluded that Rag2 R229Q homozygous mice mimic most symptoms of human Omenn syndrome and that the pathophysiology of Omenn syndrome involves impaired immune tolerance and defective immune regulation.


ALLELIC VARIANTS 10 Selected Examples):

.0001   SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

RAG2, CYS476TYR
SNP: rs121918573, gnomAD: rs121918573, ClinVar: RCV000014009, RCV000681602, RCV003460470, RCV003764566

Schwarz et al. (1996) found 2 related patients with B-negative SCID (601457) who were homozygous for a cys476-to-tyr substitution mutation in RAG2 caused by a 2634G-A transition.


.0002   SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

OMENN SYNDROME, INCLUDED
RAG2, ARG229GLN
SNP: rs121917894, gnomAD: rs121917894, ClinVar: RCV000014010, RCV000014011, RCV000681586, RCV001049147, RCV001813982, RCV003398499, RCV003473086

In a patient with T-, B- SCID (601457), Schwarz et al. (1996) identified compound heterozygosity for 2 mutations in the RAG2 gene: a 1887G-A transition, resulting in an arg229-to-gln (R229Q) substitution, and a deletion encompassing RAG1 and RAG2. A polymorphism in the RAG1 gene (A156V; 179615.0004) was also identified.

In a patient with T-, B- SCID, Corneo et al. (2001) identified compound heterozygosity for the R229Q mutation and R39G (179616.0008). A sib with Omenn syndrome (603554) had the same genotype.


.0003   OMENN SYNDROME

RAG2, CYS41TRP
SNP: rs121917895, gnomAD: rs121917895, ClinVar: RCV000014012, RCV000681574, RCV001851843

Villa et al. (1998) found a cys41-to-trp mutation in heterozygous state in a patient with Omenn syndrome (603554). This mutation occurs in a domain with HimA homology.


.0004   OMENN SYNDROME

RAG2, MET285ARG
SNP: rs121917896, ClinVar: RCV000014013, RCV000681589

Villa et al. (1998) found a met285-to-arg mutation in heterozygous state in a patient with Omenn syndrome (603554). This mutation occurs in a domain with topoII homology.


.0005   OMENN SYNDROME

RAG2, GLY95ARG
SNP: rs36001797, gnomAD: rs36001797, ClinVar: RCV000014014, RCV000671154, RCV001043887, RCV003398500, RCV003473087

Gomez et al. (2000) identified a heterozygous G-to-A transition at nucleotide 1484 of the RAG2 gene, leading to a gly95-to-arg substitution, in a male patient with Omenn syndrome (603554). The patient died at age 5 months.


.0006   SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

RAG2, 3-BP DEL, NT2018
SNP: rs2133313409, ClinVar: RCV000014015

In a 2-month-old male patient with B cell-negative (CD19 (107265) less than 1%) SCID (601457) with maternal T-cell engraftment, Gomez et al. (2000) identified a homozygous 3-bp (nucleotides 2018 to 2020) in-frame deletion in the RAG2 gene. The mutation resulted in the removal of ile273 with the remainder of the protein intact. The patient received a bone marrow transplant and was alive and well 2 years later.


.0007   SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

RAG2, TRP215ILE
SNP: rs35691292, gnomAD: rs35691292, ClinVar: RCV000014016, RCV000433357, RCV000645697, RCV000681585, RCV001081978, RCV001104047, RCV001104048, RCV003492295

In a patient with T cell-negative/B cell-negative SCID (601457), Tabori et al. (2004) identified an 1845C-T transition in the RAG2 gene, resulting in a trp215-to-ile (W215I) change.


.0008   SEVERE COMBINED IMMUNODEFICIENCY, B CELL-NEGATIVE

OMENN SYNDROME, INCLUDED
RAG2, ARG39GLY
SNP: rs121917897, ClinVar: RCV000014017, RCV000014018, RCV000681573, RCV003991458

In a patient with T-, B- SCID (601457), Corneo et al. (2001) identified compound heterozygosity for 2 mutations in the RAG2 gene: a 1316A-G transition, resulting in an arg39-to-gly (R39G) substitution, and R229Q (179616.0002). A sib with Omenn syndrome (603554) had the same genotype.


.0009   COMBINED CELLULAR AND HUMORAL IMMUNE DEFECTS WITH GRANULOMAS

RAG2, THR77ASN
SNP: rs121918574, ClinVar: RCV000014019, RCV000681578

In a patient with combined cellular and humoral immune defects associated with granulomas (233650), Schuetz et al. (2008) identified compound heterozygous mutations in the RAG2 gene: a thr77-to-asn (T77N) substitution in the catalytic core of the protein and a gly451-to-ala (G451A; 179616.0010) substitution in the PHD-like region. The patient presented at age 10 years with a history of severe infections and massive splenomegaly. She was found to have hypogammaglobulinemia and defective T-cell function. Noninfectious granulomas were present in the spleen and lungs. In vitro functional expression studies showed that the mutant proteins had significantly impaired function. Schuetz et al. (2008) concluded that the relatively late onset and low incidence of repeated infections observed in this patient reflected a low level of residual RAG2 activity.


.0010   COMBINED CELLULAR AND HUMORAL IMMUNE DEFECTS WITH GRANULOMAS

RAG2, GLY451ALA
SNP: rs121918575, gnomAD: rs121918575, ClinVar: RCV000014020, RCV000521152, RCV000681597, RCV000821298, RCV001027616, RCV001731284, RCV003398501

For discussion of the gly451-to-ala (G451A) mutation in the RAG2 gene that was found in compound heterozygous state in a patient with combined cellular and humoral immune defects associated with granulomas (233650) by Schuetz et al. (2008), see 179616.0009.


REFERENCES

  1. Adams, J., Kelso, R., Cooley, L. The kelch repeat superfamily of proteins: propellers of cell function. Trends Cell Biol. 10: 17-24, 2000. [PubMed: 10603472] [Full Text: https://doi.org/10.1016/s0962-8924(99)01673-6]

  2. Bosticardo, M., Pala, F., Notarangelo, L. D. RAG deficiencies: recent advances in disease pathogenesis and novel therapeutic approaches. Europ. J. Immun. 51: 1028-1038, 2021. [PubMed: 33682138] [Full Text: https://doi.org/10.1002/eji.202048880]

  3. Callebaut, I., Mornon, J.-P. The V(D)J recombination activating protein RAG2 consists of a six-bladed propeller and a PHD fingerlike domain, as revealed by sequence analysis. Cell. Molec. Life Sci. 54: 880-891, 1998. [PubMed: 9760994] [Full Text: https://doi.org/10.1007/s000180050216]

  4. Corneo, B., Moshous, D., Gungor, T., Wulffraat, N., Philippet, P., Le Deist, F., Fischer, A., de Villartay, J.-P. 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. [PubMed: 11313270] [Full Text: https://doi.org/10.1182/blood.v97.9.2772]

  5. Corneo, B., Wendland, R. L., Deriano, L., Cui, X., Klein, I. A., Wong, S.-Y., Arnal, S., Holub, A. J., Weller, G. R., Pancake, B. A., Shah, S., Brandt, V. L., Meek, K., Roth, D. B. Rag mutations reveal robust alternative end joining. Nature 449: 483-486, 2007. [PubMed: 17898768] [Full Text: https://doi.org/10.1038/nature06168]

  6. Deriano, L., Chaumeil, J., Coussens, M., Multani, A., Chou, Y., Alekseyenko, A. V., Chang, S., Skok, J. A., Roth, D. B. The RAG2 C terminus suppresses genomic instability and lymphomagenesis. Nature 471: 119-123, 2011. [PubMed: 21368836] [Full Text: https://doi.org/10.1038/nature09755]

  7. Gomez, C. A., Ptaszek, L. M., Villa, A., Bozzi, F., Sobacchi, C., Brooks, E. G., Notarangelo, L. D., Spanopoulou, E., Pan, Z. Q., Vezzoni, P., Cortes, P., Santagata, S. Mutations in conserved regions of the predicted RAG2 kelch repeats block initiation of V(D)J recombination and result in primary immunodeficiencies. Molec. Cell. Biol. 20: 5653-5664, 2000. [PubMed: 10891502] [Full Text: https://doi.org/10.1128/MCB.20.15.5653-5664.2000]

  8. Hikida, M., Mori, M., Takai, T., Tomochika, K., Hamatani, K., Ohmori, H. Reexpression of RAG-1 and RAG-2 genes in activated mature mouse B cells. Science 274: 2092-2094, 1996. [PubMed: 8953042] [Full Text: https://doi.org/10.1126/science.274.5295.2092]

  9. Ji, Y., Resch, W., Corbett, E., Yamane, A., Casellas, R., Schatz, D. G. The in vivo pattern of binding of RAG1 and RAG2 to antigen receptor loci. Cell 141: 419-431, 2010. Note: Erratum: Cell 143: 170 only, 2010. [PubMed: 20398922] [Full Text: https://doi.org/10.1016/j.cell.2010.03.010]

  10. Kaplan, D. H., Shankaran, V., Dighe, A. S., Stockert, E., Aguet, M., Old, L. J., Schreiber, R. D. Demonstration of an interferon gamma-dependent tumor surveillance system in immunocompetent mice. Proc. Nat. Acad. Sci. 95: 7556-7561, 1998. [PubMed: 9636188] [Full Text: https://doi.org/10.1073/pnas.95.13.7556]

  11. 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. [PubMed: 10601033] [Full Text: https://doi.org/10.1101/gad.13.23.3070]

  12. Kim, M.-S., Lapkouski, M., Yang, W., Gellert, M. Crystal structure of the V(D)J recombinase RAG1-RAG2. Nature 518: 507-511, 2015. [PubMed: 25707801] [Full Text: https://doi.org/10.1038/nature14174]

  13. 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. [PubMed: 10601032] [Full Text: https://doi.org/10.1101/gad.13.23.3059]

  14. Lin, W.-C., Desiderio, S. Cell cycle regulation of V(D)J recombination-activating protein RAG-2. Proc. Nat. Acad. Sci. 91: 2733-2737, 1994. [PubMed: 8146183] [Full Text: https://doi.org/10.1073/pnas.91.7.2733]

  15. Marrella, V., Poliani, P. L., Casati, A., Rucci, F., Frascoli, L., Gougeon, M.-L., Lemercier, B., Bosticardo, M., Ravanini, M., Battaglia, M., Roncarolo, M. G., Cavazzana-Calvo, M., Facchetti, F., Notarangelo, L. D., Vezzoni, P., Grassi, F., Villa, A. A hypomorphic R229Q Rag2 mouse mutant recapitulates human Omenn syndrome. J. Clin. Invest. 117: 1260-1269, 2007. [PubMed: 17476358] [Full Text: https://doi.org/10.1172/JCI30928]

  16. Matthews, A. G. W., Kuo, A. J., Ramon-Maiques, S., Han, S., Champagne, K. S., Ivanov, D., Gallardo, M., Carney, D., Cheung, P., Ciccone, D. N., Walter, K. L., Utz, P. J., Shi, Y., Kutateladze, T. G., Yang, W., Gozani, O., Oettinger, M. A. RAG2 PHD finger couples histone H3 lysine 4 trimethylation with V(D)J recombination. Nature 450: 1106-1110, 2007. [PubMed: 18033247] [Full Text: https://doi.org/10.1038/nature06431]

  17. McBlane, J. F., van Gent, D. C., Ramsden, D. A., Romeo, C., Cuomo, C. A., Gellert, M., Oettinger, M. A. Cleavage at a V(D)J recombination signal requires only RAG1 and RAG2 proteins and occurs in two steps. Cell 83: 387-395, 1995. [PubMed: 8521468] [Full Text: https://doi.org/10.1016/0092-8674(95)90116-7]

  18. Melek, M., Gellert, M. RAG1/2-mediated resolution of transposition intermediates: two pathways and possible consequences. Cell 101: 625-633, 2000. [PubMed: 10892649] [Full Text: https://doi.org/10.1016/s0092-8674(00)80874-0]

  19. 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. [PubMed: 2360047] [Full Text: https://doi.org/10.1126/science.2360047]

  20. Qiu, J., Kale, S. B., Schultz, H. Y., Roth, D. B. Separation-of-function mutants reveal critical roles for RAG2 in both the cleavage and joining steps of V(D)J recombination. Molec. Cell 7: 77-87, 2001. [PubMed: 11172713] [Full Text: https://doi.org/10.1016/s1097-2765(01)00156-3]

  21. Rideout, W. M., III, Hochedlinger, K., Kyba, M., Daley, G. Q., Jaenisch, R. Correction of a genetic defect by nuclear transplantation and combined cell and gene therapy. Cell 109: 17-27, 2002. [PubMed: 11955443] [Full Text: https://doi.org/10.1016/s0092-8674(02)00681-5]

  22. Schuetz, C., Huck, K., Gudowius, S., Megahed, M., Feyen, O., Hubner, B., Schneider, D. T., Manfras, B., Pannicke, U., Willemze, R., Knuchel, R., Gobel, U., Schulz, A., Borkhardt, A., Friedrich, W., Schwarz, K., Niehues, T. An immunodeficiency disease with RAG mutations and granulomas. New Eng. J. Med. 358: 2030-2038, 2008. [PubMed: 18463379] [Full Text: https://doi.org/10.1056/NEJMoa073966]

  23. Schwarz, K., Gauss, G. H., Ludwig, L., Pannicke, U., Li, Z., Linder, D., Friedrich, W., Seger, R. A., Hansen-Hagge, T. E., Desiderio, S., Lieber, M. R., Bartram, C. R. RAG mutations in human B cell-negative SCID. Science 274: 97-99, 1996. [PubMed: 8810255] [Full Text: https://doi.org/10.1126/science.274.5284.97]

  24. Shankaran, V., Ikeda, H., Bruce, A. T., White, J. M., Swanson, P. E., Old, L. J., Schreiber, R. D. IFN-gamma and lymphocytes prevent primary tumour development and shape tumour immunogenicity. Nature 410: 1107-1111, 2001. [PubMed: 11323675] [Full Text: https://doi.org/10.1038/35074122]

  25. Shinkai, Y., Rathbun, G., Lam, K.-P., Oltz, E. M., Stewart, V., Mendelsohn, M., Charron, J., Datta, M., Young, F., Stall, A. M., Alt, F. W. RAG-2-deficient mice lack mature lymphocytes owing to inability to initiate V(D)J rearrangement. Cell 68: 855-867, 1992. [PubMed: 1547487] [Full Text: https://doi.org/10.1016/0092-8674(92)90029-c]

  26. Tabori, U., Mark, Z., Amariglio, N., Etzioni, A., Golan, H., Biloray, B., Toren, A., Rechavi, G., Dalal, I. Detection of RAG mutations and prenatal diagnosis in families presenting with either T-B- severe combined immunodeficiency or Omenn's syndrome. Clin. Genet. 65: 322-326, 2004. [PubMed: 15025726] [Full Text: https://doi.org/10.1111/j.1399-0004.2004.00227.x]

  27. Villa, A., Santagata, S., Bozzi, F., Giliani, S., Frattini, A., Imberti, L., Gatta, L. B., Ochs, H. D., Schwarz, K., Notarangelo, L. D., Vezzoni, P., Spanopoulou, E. Partial V(D)J recombination activity leads to Omenn syndrome. Cell 93: 885-896, 1998. [PubMed: 9630231] [Full Text: https://doi.org/10.1016/s0092-8674(00)81448-8]

  28. Yu, W., Misulovin, Z., Suh, H., Hardy, R. R., Jankovic, M., Yannoutsos, N., Nussenzweig, M. C. Coordinate regulation of RAG1 and RAG2 by cell type-specific DNA elements 5-prime of RAG2. Science 285: 1080-1084, 1999. [PubMed: 10446057] [Full Text: https://doi.org/10.1126/science.285.5430.1080]

  29. Yu, X., Almeida, J., Darko, S., van der Burg, M., DeRavin, S. S., Malech, H., Gennery, A., Chinn, I., Markert, M. L., Douek, D. C., Milner, J. D. Human syndromes of immunodeficiency and dysregulation are characterized by distinct defects in T-cell receptor repertoire development. J. Allergy Clin. Immun. 133: 1109-1115, 2014. [PubMed: 24406074] [Full Text: https://doi.org/10.1016/j.jaci.2013.11.018]


Contributors:
Matthew B. Gross - updated : 05/01/2024
Ada Hamosh - updated : 02/03/2016
Paul J. Converse - updated : 1/23/2014
Paul J. Converse - updated : 8/19/2011
Ada Hamosh - updated : 6/10/2011
Cassandra L. Kniffin - updated : 5/19/2008
Ada Hamosh - updated : 4/22/2008
Paul J. Converse - updated : 10/24/2007
Ada Hamosh - updated : 10/11/2007
Cassandra L. Kniffin - updated : 10/20/2004
Victor A. McKusick - updated : 4/22/2004
Stylianos E. Antonarakis - updated : 5/3/2002
Paul J. Converse - updated : 4/25/2001
Stylianos E. Antonarakis - updated : 2/1/2001
Paul J. Converse - updated : 10/31/2000
Stylianos E. Antonarakis - updated : 6/21/2000
Paul J. Converse - updated : 4/13/2000
Ada Hamosh - updated : 8/12/1999
Stylianos E. Antonarakis - updated : 6/29/1998
Moyra Smith - updated : 12/19/1996
Moyra Smith - updated : 10/3/1996

Creation Date:
Victor A. McKusick : 7/9/1990

Edit History:
mgross : 05/01/2024
alopez : 07/31/2023
alopez : 02/03/2016
mcolton : 7/21/2015
mcolton : 1/30/2015
mgross : 2/11/2014
mgross : 2/11/2014
mgross : 2/11/2014
mcolton : 1/23/2014
mgross : 10/7/2013
mgross : 10/4/2013
mgross : 2/5/2013
mgross : 9/19/2011
mgross : 9/19/2011
terry : 8/19/2011
alopez : 6/20/2011
terry : 6/10/2011
wwang : 5/20/2008
ckniffin : 5/19/2008
alopez : 5/15/2008
terry : 4/22/2008
mgross : 11/1/2007
terry : 10/24/2007
alopez : 10/16/2007
terry : 10/11/2007
carol : 10/28/2004
terry : 10/28/2004
ckniffin : 10/20/2004
tkritzer : 4/23/2004
terry : 4/22/2004
mgross : 5/3/2002
alopez : 4/25/2001
mgross : 2/1/2001
mgross : 10/31/2000
mgross : 6/21/2000
carol : 4/13/2000
alopez : 8/12/1999
alopez : 8/12/1999
carol : 2/18/1999
terry : 2/18/1999
carol : 10/5/1998
carol : 6/29/1998
alopez : 7/10/1997
mark : 12/19/1996
mark : 10/5/1996
mark : 10/4/1996
mark : 10/4/1996
mark : 10/3/1996
mark : 12/4/1995
jason : 7/15/1994
carol : 8/11/1992
carol : 6/2/1992
carol : 6/1/1992
carol : 5/29/1992
carol : 5/26/1992