17
The role of the non-homologous end-joining pathway in lymphocyte development Sean Rooney Jayanta Chaudhuri Frederick W. Alt Authors’ address Sean Rooney, Jayanta Chaudhuri, Frederick W. Alt, Howard Hughes Medical Institute, The Children’s Hospital, The Department of Genetics, Harvard Medical School and The Center for Blood Research, Boston, MA, USA. Correspondence to: Fred W. Alt Children’s Hospital 300 Longwood Avenue New Research Building, 9th Floor Boston, MA 02115, USA Tel.: þ1 617 919 2539 E-mail: [email protected] Acknowledgements This work was supported by grants from the NIH (AI35714, AI20047, and AI31541) and NCI (CA92625). F.W.A. is an investigator and J.C. an associate of the Howard Hughes Medical Institute. We apologize to those whose work we could not cite due to space constraints. Summary: One of the most toxic insults a cell can incur is a disruption of its linear DNA in the form of a double-strand break (DSB). Left unrepaired, or repaired improperly, these lesions can result in cell death or neoplastic transformation. Despite these dangers, lymphoid cells purposely introduce DSBs into their genome to maximize the diversity and effector functions of their antigen receptor genes. While the generation of breaks requires distinct lymphoid-specific factors, their resolution requires various ubiquitously expressed DNA-repair proteins, known collectively as the non-homologous end-joining pathway. In this review, we discuss the factors that constitute this pathway as well as the evidence of their involve- ment in two lymphoid-specific DNA recombination events. Introduction The adaptive arm of the immune system is predicated upon the recognition of foreign antigen by immunoglobulin (Ig) and/or T-cell receptors (TCRs) expressed on the surface of B and T lymphocytes, respectively. While the number of anti- gens is seemingly limitless, the mammalian genome contains only a finite number of antigen recognition genes. This prob- lem is dealt with in a novel fashion: the reorganization of germline DNA segments into unique antigen receptor genes, a process referred to as V(D)J recombination (1). V(D)J recombination is a site-specific event that takes place at six distinct loci: TCRb, g, a/d loci, Ig heavy chain (IgH), and k or l light chain (LC) loci. Recombination occurs between component variable (V), junctional (J), and, in some cases, diversity (D) gene segments, with fused VJ or VDJ coding sequence subsequently joined to a constant region segment through RNA splicing. Because most antigen loci have numer- ous gene segments, a significant level of antigen-receptor diversity is generated solely through multiple combinations of fused products (2). However, assortment of germline V gene segments alone is not sufficient to account for the wide range of antigen Immunological Reviews 2004 Vol. 200: 115–131 Printed in Denmark. All rights reserved Copyright ß Blackwell Munksgaard 2004 Immunological Reviews 0105-2896 115

The role of the non-homologous end-joining pathway in lymphocyte development

Embed Size (px)

Citation preview

Page 1: The role of the non-homologous end-joining pathway in lymphocyte development

The role of the non-homologous

end-joining pathway in lymphocyte

development

Sean Rooney

Jayanta Chaudhuri

Frederick W. Alt

Authors’ address

Sean Rooney, Jayanta Chaudhuri, Frederick W. Alt,

Howard Hughes Medical Institute, The

Children’s Hospital, The Department of

Genetics, Harvard Medical School and The

Center for Blood Research, Boston, MA, USA.

Correspondence to:

Fred W. Alt

Children’s Hospital

300 Longwood Avenue

New Research Building, 9th Floor

Boston, MA 02115, USA

Tel.: þ1 617 919 2539

E-mail: [email protected]

Acknowledgements

This work was supported by grants from the NIH

(AI35714, AI20047, and AI31541) and NCI (CA92625).

F.W.A. is an investigator and J.C. an associate of the

Howard Hughes Medical Institute. We apologize to those

whose work we could not cite due to space constraints.

Summary: One of the most toxic insults a cell can incur is a disruption ofits linear DNA in the form of a double-strand break (DSB). Left unrepaired,or repaired improperly, these lesions can result in cell death or neoplastictransformation. Despite these dangers, lymphoid cells purposely introduceDSBs into their genome to maximize the diversity and effector functions oftheir antigen receptor genes. While the generation of breaks requiresdistinct lymphoid-specific factors, their resolution requires variousubiquitously expressed DNA-repair proteins, known collectively as thenon-homologous end-joining pathway. In this review, we discuss thefactors that constitute this pathway as well as the evidence of their involve-ment in two lymphoid-specific DNA recombination events.

Introduction

The adaptive arm of the immune system is predicated upon

the recognition of foreign antigen by immunoglobulin (Ig)

and/or T-cell receptors (TCRs) expressed on the surface of B

and T lymphocytes, respectively. While the number of anti-

gens is seemingly limitless, the mammalian genome contains

only a finite number of antigen recognition genes. This prob-

lem is dealt with in a novel fashion: the reorganization of

germline DNA segments into unique antigen receptor genes, a

process referred to as V(D)J recombination (1).

V(D)J recombination is a site-specific event that takes place

at six distinct loci: TCRb, g, a/d loci, Ig heavy chain (IgH), and

k or l light chain (LC) loci. Recombination occurs between

component variable (V), junctional (J), and, in some cases,

diversity (D) gene segments, with fused VJ or VDJ coding

sequence subsequently joined to a constant region segment

through RNA splicing. Because most antigen loci have numer-

ous gene segments, a significant level of antigen-receptor

diversity is generated solely through multiple combinations

of fused products (2).

However, assortment of germline V gene segments alone

is not sufficient to account for the wide range of antigen

Immunological Reviews 2004

Vol. 200: 115–131

Printed in Denmark. All rights reserved

Copyright � Blackwell Munksgaard 2004

Immunological Reviews0105-2896

115

Page 2: The role of the non-homologous end-joining pathway in lymphocyte development

specificities. For example, the sequence of Va and Vb gene

segments is highly restricted for two of the three domains

responsible for antigen recognition [complementarity-deter-

mining region (CDR) 1 and CDR2] (3). By comparison,

CDR3, which spans the junction of fused gene segments, is

highly variable. Much of this junctional diversity is a result

of the imprecise repair of DNA double-strand breaks (DSBs)

introduced at specific sites within antigen-receptor gene

segments during the V(D)J reaction.

The V(D)J recombination reaction occurs in two distinct

steps. First, the lymphoid-specific endonuclease recombinase-

activating genes 1 and 2 (RAG-1 and RAG-2, collectively RAG)

introduce DSBs precisely between gene segments undergoing

recombination and their flanking recombination signal

sequences (RSSs) (4) (Fig. 1A). The RAG proteins generate

breaks by nicking a single DNA strand between the coding

flank and the adjacent RSS, leaving a 30 hydroxyl group on the

coding side and a 50 phosphoryl group on the RSS side. The

30 hydroxyl group then attacks the phosphodiester backbone of

the opposing strand, resulting in a covalently sealed hairpin

coding end and in a blunt 50 phosphorylated recombination

signal (RS) end (5).

The second half of the V(D)J reaction involves the resolution

of RAG-generated DSBs. Whereas the RS ends are fused directly

to form RS joins, coding-end hairpins must be opened and

processed prior to coding join formation (Fig. 1B). Hairpin

opening in vivo usually occurs at or within several nucleotides

of the apex (6, 7). Opening away from the apex results in

over-hanging nucleotides, which, if incorporated into the

join, generate non-germline palindromic (P)-nucleotide add-

itions (8, 9). Coding joins are further diversified through the

addition of non-templated (N)-nucleotides by the lymphoid-

specific terminal deoxynucleotidyl transferase (10–12) or

through the deletion of a small number of nucleotides from

coding sequences. Although the two distinct intermediates

generated by RAG cleavage must be dealt with in a different

manner, the repair of each is dependent upon the same ubi-

quitously expressed set of DNA-repair factors, collectively

referred to as the non-homologous end-joining (NHEJ)

pathway.

In eukaryotes, a majority of DSBs are resolved through two

repair pathways. Homologous recombination (HR) uses infor-

mation from a homologous template to accurately repair

breaks, and it is generally limited in mammalian cells to late

S-phase and G2-phase of the cell cycle, when sister chromatids

present readily available templates (13). NHEJ rejoins broken

DNA ends with little or no sequence homology, and it is the

predominant pathway for repair during G1 (14, 15), the stage

at which RAG-associated DSBs are generated (16).

The NHEJ pathway consists of at least six proteins. The

DNA-binding subunits Ku70 and Ku80, together with the

DNA-dependent protein kinase catalytic subunit (DNA-PKcs),

form the DNA-PK holoenzyme (DNA-PK), believed to be

involved early in the recognition of DSBs (17). The nuclease

Artemis forms a complex with DNA-PKcs, and it is responsible

for processing hairpin ends prior to their ligation. DNA ligase

RAG

Nicking

Nucleophilic attack

Double-strand break

HO

HO

HO

OH

OH

OH

Ku70/80

DNA-PKcsArtemis

Ligase IV/XRCC4(DNA-PKcs)

Ligase IV/XRCC4TdT

Precise RS join

Imprecise coding join

A B

Fig. 1. The V(D)J reaction. (A) Introductionof double-strand breaks by the recombinase-activating gene (RAG) endonuclease. (B) Thenon-homologous end-joining (NHEJ) factorsresponsible for recombination signal (RS)and coding joining.

Rooney et al � The NHEJ pathway in lymphocyte development

116 Immunological Reviews 200/2004

Page 3: The role of the non-homologous end-joining pathway in lymphocyte development

IV, together with its partner XRCC4, catalyzes the final step of

the reaction. In this review, we summarize the role of each

factor, with particular attention paid to the impact of its

absence on V(D)J recombination and lymphocyte develop-

ment. We also discuss the potential role of NHEJ in a second

lymphoid-specific DNA recombination event, known as IgHC

class switch recombination (CSR). Finally, we summarize the

evidence that defects in NHEJ promote aberrant recombination

events, which may ultimately lead to lymphomagenesis.

The NHEJ factors and V(D)J recombination

Ku70, Ku80, DNA-PKcs, ligase IV, and XRCC4 were each

linked to DSB repair and V(D)J recombination, either directly

or indirectly, through studies of ionizing-radiation (IR)-

sensitive Chinese hamster ovary (CHO) cell lines as well as

of the naturally occurring severe combined immunodeficiency

(SCID) mice (18, 19). Artemis was subsequently identified as

being mutated in a subset of human SCID patients (20). The

NHEJ pathway is partially conserved in yeast (21), and it may

also play a role in DSB repair in prokaryotes (22). However,

while the yeast homologs of Ku70, Ku80, XRCC4, and ligase IV

have all been identified, DNA-PKcs and, potentially, Artemis

appear restricted to higher eukaryotes (23).

The DNA-PK holoenzyme

DNA-PK was first characterized from cell extracts as a kinase

specifically activated by double-stranded DNA but not by

single-stranded DNA or RNA (24–26). Subsequent work

determined that the DNA-PK holoenzyme is a nuclear serine/

threonine kinase (27), consisting of the DNA-binding Ku

heterodimer and the DNA-PKcs (28–30).

Ku was originally identified as an autoantigen from

scleroderma–polymyositis-overlap syndrome patients (31).

Ku is a DNA-binding heterodimer, consisting of two tightly

associated subunits, Ku70 and Ku80 (sometimes referred to

as Ku86), which binds with high affinity to free double-

stranded DNA ends, DNA hairpins, single-strand nicks, or

gaps in a sequence-independent manner (32). A role for Ku

in DSB repair was demonstrated by the ability of human

Ku86 to complement the IR-sensitivity and V(D)J recom-

bination defects of xrs6 CHO cells (33–35), and subsequently

verified through gene-targeted mutation of each subunit in

mice (36, 37). The recognition of DNA ends by the Ku

heterodimer is thought to be an early event in NHEJ (38)

and may be responsible for directing repair away from HR

(39–41).

Binding of Ku to DNA is dependent upon heterodimeriza-

tion, as neither subunit can bind DNA alone (42–45). Inter-

action of the heterodimer is mediated through amino acids

449–578 of Ku70 and amino acids 439–592 of Ku80 (46),

and these two regions share significant sequence homology

(46), suggesting that the two subunits of Ku arose through a

gene-duplication event (47). Consistent with this idea, a single

DNA-binding protein with significant homology to Ku has

been identified in prokaryotes (22). Ku70 and Ku80 together

form a ring structure with a central channel large enough to

accommodate a single double-stranded DNA molecule (48).

The heterodimer does not make contact with a single DNA

base, and it has only limited contact with the sugar-phosphate

backbone, in agreement with the lack of sequence-specific

binding for Ku. However, the ring structure fits well into

both the major and minor groove of duplex DNA, orientating

the DNA in a particular plane, a feature that may be relevant

for aligning DNA ends prior to their ligation (49, 50).

Upon binding to DNA, a 12-amino-acid sequence at the

C-terminus of Ku80 is exposed and recruits DNA-PKcs (46,

51). This, in turn, stimulates the inward translocation of Ku,

allowing DNA-PKcs to bind directly to DNA ends (52, 53). At

470 kDa, DNA-PKcs is one of the largest known cellular pro-

teins (47). Despite its large size, the only clearly defined motif

within the subunit is a C-terminal phosphatidylinositol

3-kinase (PI3K) domain (54, 55), characteristic of kinases

with lipid substrates (56). However, DNA-PKcs is specific

for protein substrates, preferentially phosphorylating S/T-Q

motifs (serine or threonine followed by a glutamine) (27, 57,

58). As a result, DNA-PKcs is categorized as a member of the

PI3K-related protein kinase (PIKK) family, which includes the

cell-cycle checkpoint proteins antaxia telangiectasia mutated

(ATM) and ATM- and RAD3-related (ATR) (59).

The role of DNA-PKcs in V(D)J recombination was first

appreciated through studies of the IR-sensitive SCID mouse

and transient transfecion of V3 CHO cell line. Both V3 cells

and SCID mice have a specific defect in coding with RAG and

V(D)J recombination subtrates join formation, while RS join

formation is relatively unaffected (60–63). Furthermore, the

rare coding joins recovered from both SCID and V3 cells have

elevated levels of P-nucleotide additions, indicating that RAG-

generated hairpins are opened aberrantly (8). While the V3

defect results from the deletion of prkdc (the gene encoding

DNA-PKcs) (64), SCID mice harbor a single point mutation at

amino acid 4046 of DNA-PKcs, resulting in a truncated protein

(65–67). The truncation leaves the kinase domain intact;

however, the deleted region is highly conserved among PIKK

family members (66), and the loss of this domain greatly

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 117

Page 4: The role of the non-homologous end-joining pathway in lymphocyte development

reduces the kinase activity of DNA-PKcs in vitro (68). The

requirement for DNA-PKcs kinase activity in V(D)J recombin-

ation was further demonstrated by the inability of a kinase-

dead DNA-PKcs to complement the coding join defect of V3

cells (69). However, it was only with the identification of the

nuclease Artemis (20) that the ability of DNA-PKcs to regulate

coding join formation was fully understood (see below).

Moreover, autophosphorylation of DNA-PKcs appears to be

important for proper end processing. DNA-PKcs has seven

different sites that undergo phosphorylation in vivo, six of

which are clustered in a 38-amino-acid region in the middle

of the protein (70). This modification appears relevant to

NHEJ in vivo, as mutation of the sites within this cluster

abrogates the ability of DNA-PKcs to rescue the IR sensitivity

of V3 cells (71, 72), without affecting the kinase activity of

DNA-PKcs in vitro (72). A recent report suggests that auto-

phosphorylation occurs following the synapse of two DNA-

PK-bound DNA ends, possibly resulting in a conformational

change of the holoenzyme that facilitates further processing

and/or ligation of DNA ends (73).

Several studies have demonstrated that DNA-PKcs is also

necessary for the proper repair of blunt-ended breaks (74–77),

indicating that the holoenzyme has a role in NHEJ beyond

processing incompatible DNA ends. DNA-PK may serve as a

scaffold for recruiting other factors necessary for DSB repair,

such as the XRCC4–ligase IV complex (78–82). DNA-PK has

also been shown to interact with multiple DNA molecules

simultaneously (83, 84), leading to the speculation that the

holoenzyme may mediate the synapse of DNA ends. Neither

the recruitment of the XRCC4–ligase IV complex nor the

synapse formation requires the kinase activity of DNA-PKcs

(81, 84). Together, these data suggest that kinase-dead ver-

sions of DNA-PKcs, such as the murine SCID mutation, retain

partial function. In support of this notion, it has been recently

reported that the kinase activity of DNA-PKcs, but not DNA-

PKcs itself, is expendable for transposition of the sleeping

beauty element (85).

Ku70, Ku80, and DNA-PKcs have all been inactivated in

mice by gene targeting, and deletion of each results in defect-

ive T- and B-cell development, due to the inability to repair

RAG-induced DSBs (37, 86–92). Covalently sealed coding

ends accumulate in thymi of these mice (86, 88, 92), demon-

strating that the holoenzyme is employed for hairpin opening

in vivo. Furthermore, coding join formation is greatly reduced

in fibroblasts isolated from each of these mice, and the rare

joins that are recovered display large deletions of the coding

flanks, indicating that the coding join formation that does

occur is grossly aberrant (86, 92, 93). Moreover, it is worth

emphasizing that coding joins isolated from these cells,

including DNA-PKcs-null fibroblasts, do not display the ele-

vated levels of P-nucleotide additions that characterize joins

from kinase-dead SCID mice. Therefore, coding join formation

in the absence of the kinase activity may be distinct from what

occurs in the absence of the entire protein, supporting the idea

that DNA-PKcs has multiple functions in this process.

The frequency and fidelity of RS joining is also defective in

the absence of each subunit of the DNA-PK holoenzyme, albeit

more severely so for Ku than for DNA-PKcs. The level of RS

joining is greatly reduced in Ku-deficient fibroblasts assayed

by transient transfection of a V(D)J recombination substrate.

Furthermore, the rare RS joins recovered from these cells

exhibit deletions and regions of microhomology (86). It has

been suggested that the Ku heterodimer facilitates alignment

of DNA ends prior to ligation (49, 50) and thus may obviate

the need for such microhomology in normal RS joining. SCID

cells have little or no impairment in the level of RS joining and

most are precise (60, 61). However, DNA-PKcs-null fibro-

blasts have a more variable defect in both the level and the

precision of RS joining, although a number of joins remain

precise (93, 94). While more studies need to be performed,

the potential differences in the frequency and the fidelity of RS

joining between SCID and DNA-PKcs-deficient cells further

support the notion that the SCID mutation is not equivalent

to the complete loss of DNA-PKcs.

Consistent with Ku having functions beyond the context of

the DNA-PK holoenzyme, the overall phenotypic effects of

DNA-PKcs deficiency in mice are clearly not equivalent to

the loss of the Ku heterodimer. Aside from the defects in

lymphocyte development, the only obvious phenotype of

DNA-PKcs-null mice is an increased level of organismal and

cellular sensitivity to IR, consistent with a general defect in

DSB repair (89, 90, 92, 95). Ku-deficient mice have increased

levels of neuronal apoptosis (96), are small (approximately

50% by body weight) compared to their littermates (37, 86,

87), and incur age-specific changes significantly earlier than

controls (97). Likewise, fibroblasts isolated from Ku-deficient

mice display growth defects, premature cellular senescence,

and hypersensitivity to IR, generally greater than that of DNA-

PKcs-deficient cells (37, 86). Inactivation of the cell-cycle

checkpoint protein p53 suppresses the growth defects of

Ku-deficient cells (98, 99); however, it does not rescue the

dwarfism observed in these mice.

Finally, inactivation of Ku70, Ku80, or DNA-PKcs in the

context of ATM-deficiency is synergistically lethal at a very

early embryonic stage (100, 101), suggesting that these two

PIKK kinases may play overlapping roles in development. Such

Rooney et al � The NHEJ pathway in lymphocyte development

118 Immunological Reviews 200/2004

Page 5: The role of the non-homologous end-joining pathway in lymphocyte development

potential overlapping roles likely do not involve NHEJ, as

ATM deficiency actually suppresses the embryonic lethality

associated with XRCC4- or ligase IV-deficient mice (101). In

this regard, both Ku and DNA-PKcs appear to have other

functions not directly related to NHEJ, including telomere main-

tenance (102), phosphorylation of the histone variant H2AX

(103–106), regulation of the WRN nuclease (107–109), DNA

replication (110, 111), cell signaling (112, 113), or regulating

apoptosis (114, 115).

Artemis

Approximately 20% of human SCID patients are characterized

by a complete absence of mature T and B cells, despite normal

natural killer (NK)-cell development (T-B-SCID) (116). Muta-

tions in rag1 or in rag2 have been observed in a majority of

these patients, and fibroblasts from these patients display

normal radiosensitivity. However, fibroblasts from a subset

of these patients are highly radiosensitive (RS-SCID) (117).

V(D)J recombination is defective in these patients, as evi-

denced by grossly abnormal D-to-J rearrangements of the

IgH locus (118). Furthermore, RS-SCID fibroblasts have a

specific defect in coding join formation, while RS join forma-

tion appears unaffected (119). While this profile is similar to

that of SCID mice, RS-SCID fibroblasts have normal DNA-PKcs

kinase activity and normal Ku DNA-binding activity (119,

120). Genetic analysis localized the rs-scid mutation to a region

on chromosome 10 (121), associated with the high incidence

of T-B-SCID observed in Athabascan-speaking Native Ameri-

cans (SCIDA) (122). The rs-scid gene was cloned and was

found to encode a novel 77.6 kDa protein named Artemis

(20). Similar defects in Artemis have subsequently been

shown to underlie related human SCID cohorts (123–125).

Artemis is a member of the metallo-b-lactamase superfamily

of proteins (20), enzymes that use co-ordinated zinc(II) ions

to activate water molecules for the hydrolysis of covalent

bonds (126). First identified in prokaryotes based on the

ability to inactivate bactericidal compounds, b-lactamases

have evolved to recognize a wide variety of substrates (127).

While having little primary homology, metallo-b-lactamases

share a characteristic secondary feature, known as a metallo-b-lactamase fold (126). Artemis has been further classified into a

subfamily of metallo-b-lactamases that act on nucleic acid

substrates and share a second highly conserved motif, referred

to as the b-CASP domain (128).

The homology of Artemis with enzymes that catalyze the

hydrolysis of covalent bonds, as well as the parallel phenotypes

of RS-SCID patients and SCID mice, led to the speculation that

coding-end processing might be mediated by DNA-PKcs

through the regulation of Artemis (20). Biochemical analysis

strongly supported such a model, as DNA-PKcs forms a com-

plex with Artemis in vivo and phosphorylates Artemis in vitro

(129). Furthermore, a DNA-PKcs–Artemis complex preferen-

tially opens RAG-generated hairpins two nucleotides 30 of the

apex in vitro (129), similar to what has been observed in vivo

(6, 7). DNA-PKcs also alters the nuclease activity of Artemis

from an intrinsic single-strand-specific 50-to-30 exonuclease to

an endonuclease that works on both 50 and 30 overhangs

(129), an activity that may be relevant to the role of Artemis

in general DSB repair.

Like DNA-PKcs-null mice, Artemis-deficient (ArtN/N) mice

appear relatively normal, with a defective lymphocyte devel-

opment as the only obvious phenotype (93). The block in

lymphocyte development observed in ArtN/N mice is due to a

specific defect in coding join formation, and covalently sealed

hairpin intermediates can be detected in thymi of ArtN/N mice,

validating Artemis as a nuclease responsible for processing

coding ends in vivo. However, Artemis-deficient mice display

a variable level of leaky CD4þ T-cell development, and coding

joins can be recovered from Artemis-deficient cell lines, albeit

at greatly reduced levels (93, 130). Similar to in SCID mice,

the joins recovered in the absence of Artemis are characterized

by an increase in the size and frequency of P-nucleotide add-

ition, indicating that the little coding join formation that does

occur is aberrant.

Unlike what has been observed for a single RS-SCID infant

(131), it does not appear that maternal CD4þ T cells contri-

bute to the leaky T-cell development in ArtN/N mice (Rooney

and Alt, unpublished data). Instead, it seems more likely that

an alternative factor(s) is capable of opening hairpins in the

absence of Artemis. Several other proteins have been shown to

possess hairpin-opening activity in vitro, including the Mre11–

Rad50–Nbs1 complex (132) and the RAG proteins (133,

134). While the hairpin-opening activities of both have char-

acteristics that are inconsistent with normal coding-end pro-

cessing in vivo (9), these or other factors still may be able to

compensate for the loss of Artemis at a low level.

While both DNA-PKcs and Artemis-deficient fibroblasts are

clearly more sensitive to IR than wildtype, they are not as

sensitive as Ku70-, Ku80-, XRCC4-, or ligase IV-deficient

fibroblasts (92, 93). Similarly, mice deficient in DNA-PKcs

or Artemis lack the more severe phenotypes observed in other

NHEJ-deficient mice. These findings support the idea that the

DNA-PKcs–Artemis complex might have arisen later in evolu-

tion to deal with a subset of DSBs that required endonucleo-

lytic processing, such as those with staggered or damaged ends

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 119

Page 6: The role of the non-homologous end-joining pathway in lymphocyte development

(23, 93, 130). However, in the context of general DSB repair,

it remains to be determined exactly what types of ends require

processing by Artemis and how this function overlaps with

other nucleases implicated in this process, such as WRN, FEN-1,

or the Mre11 complex (135–137).

The XRCC4–ligase IV complex

Following the processing and alignment of DNA ends, V(D)J

recombination appears to be completed via the ligation of two

coding ends or two RS ends by the XRCC4–ligase IV complex.

Xrcc4 was identified through complementation of the IR sensi-

tivity of XR-1 CHO cells (138), and encodes a 326-amino-acid

product with no obvious homology with any known proteins.

XRCC4 was subsequently found to co-purify with DNA ligase

IV (139, 140), and the role of this complex in both NHEJ and

V(D)J recombination was validated through gene-targeted

mutation in mice (141, 142).

The interaction of ligase IV with XRCC4 is thought to

stimulate ligase IV activity in several ways. Ligase IV is severely

reduced in XR-1 cells (143), suggesting that XRCC4 stabilizes

ligase IV protein expression. Like Ku, XRCC4 may have a role

in aligning DSBs prior to their ligation (50, 80, 144, 145).

Moreover, XRCC4 can stimulate the adenylation of lysine

residues within the catalytic core of ligase IV, the first step in

the formation of a new phosphodiester bond (145). In this

context, human patients have been identified with reduced

levels of ligase IV activity (146, 147), and the mutations

underlying this defect prevent either the interaction with

XRCC4 (146) or the adenylation or the ligase IV active site

(147).

Although XRCC4 itself has been reported to have DNA end-

binding activity (145), the XRCC4–ligase IV complex most

likely is recruited to DNA ends through interactions with the

DNA-PK holoenzyme (78–81). In this regard, the C-terminus

of ligase IV has two BRCT (BRCA-1 C-terminal) domains, a

motif found in several DNA-repair proteins and recently

shown to be phosphoserine or phosphothreonine-specific-

binding module (148–150). It is interesting to speculate that

these repeats may be involved in the interaction between

DNA-PK and the XRCC4–ligase IV complex, possibly through

the binding of autophosphorylation sites on DNA-PKcs.

As mentioned, gene-targeted mutations of LIGASE IV or

XRCC4 have verified a role for both in V(D)J recombination.

While homozygous mutations of either gene result in embryo-

nic lethality, XRCC4- and ligase IV-deficient embryos exhibit

a complete block in lymphoid development (141, 142), and

IgH rearrangements cloned from ligase IV-deficient progenitor

B cells grown in culture are grossly aberrant (141). Further-

more, fibroblasts from these mice fail to support either coding

join or RS join formation by transient transfection (141, 142).

In addition to V(D)J recombination defects, XRCC4- and

ligase IV-deficient mice share a number of other phenotypes,

indicative of the absolute requirement for this complex in

NHEJ. XRCC4- and ligase IV-deficient mice generally die in

utero by day E16.5. Although the exact cause of death has not

been established, these embryos display massive apoptosis of

post-mitotic neurons in the central nervous system (142, 151,

152). It remains unclear why these cells, in particular, are

sensitive to DSB repair defects (142), given that the embryonic

lethality and neuronal apoptosis can be suppressed by inactiva-

tion of p53 or ATM (101, 153–155), it appears that both are

the result of cell-cycle checkpoint responses to the persistence

of unresolved DSBs, rather than the inability to repair DSBs per

se. However, inactivation of p53 does not rescue NHEJ per se,

therefore, lymphoid development remains defective in these

mice (153, 154), as progenitor cells lack the functional anti-

gen receptors required to promote further differentiation

(156, 157). Similarly, fibroblasts isolated from XRCC4- or

ligase IV-deficient mice exhibit substantial growth defects

and premature cellular senescence, as well as hypersensitivity

to IR (141, 142).

Finally, the embryonic lethality of ligase IV-deficient mice

can also be suppressed by inactivation of Ku80 (158). Similar

observations have been reported in a chicken pre-B-cell line

(39, 40), and together these data support a model where the

binding of the Ku heterodimer to DSBs prevents repair by

another pathway, such as HR, even in the absence of func-

tional NHEJ (39–41). Alternatively, it has been proposed that

the Ku facilitates a critical loss of DNA through the serial

recruitment of the DNA-PKcs–Artemis nucleolytic complex to

ends left unrepaired in the absence of ligase IV (158). In this

context, it remains to be determined whether inactivation of

DNA-PKcs or Artemis also rescues the embryonic lethality

associated with ligase IV deficiencies.

The role of NHEJ in IgH CSR

Upon migration to peripheral lymphoid organs, mature B

lymphocytes can be stimulated to undergo a second DNA

recombination event. In the appropriate context, B cells switch

the effector function of their IgH through the replacement of

the m constant-gene exons (Cm) with one of several sets of

downstream constant-region gene exons (referred to as CHgenes) (Fig. 2), while retaining the specificity of their variable

region. This process is referred to as IgH CSR (159).

Rooney et al � The NHEJ pathway in lymphocyte development

120 Immunological Reviews 200/2004

Page 7: The role of the non-homologous end-joining pathway in lymphocyte development

Unlike V(D)J recombination, CSR does not require the RAG

proteins or short, canonical RSSs (160). Instead, CSR as well as

the related process of somatic hypermutation (SHM) is depen-

dent upon activation-induced deaminase (AID) (161), which

converts cytidine to uridine in the context of single-stranded

DNA substrates (162–165). Germline transcription through

CH genes stimulated to undergo recombination is also

required, apparently to generate single-stranded DNA sub-

strates for AID (163), and disruption of cis-acting elements

promoting transcription of a particular CH gene blocks switch-

ing to that isotype (159).

With the exception of Cd (which switches through normal

RNA splicing), each CH gene is organized into a discrete unit

comprised of a 50 promoter region, an I exon, a highly

repetitive, G/C-rich element referred to as the switch region

(S region), and finally the CH exons. Transcription of S

regions, both in vivo and in vitro, has been shown to result in

the formation of a stable RNA : DNA hybrid between the

switch transcript and the C-rich template strand, displacing

the non-template G-strand in an R-loop structure (166, 167).

Inversion of S regions changes the template strand from C-rich

to G-rich and severely inhibits both R-loop formation in vitro

(166, 167) and CSR in vivo (168). Furthermore, a synthetic

sequence, unrelated to S regions but capable of forming R

loops in vitro, supports CSR in vivo (168). Additionally, in vitro

studies suggest that transcribed S regions may also assume

secondary structures, such as stem-loops resulting from the

high density of palindromic sequence within the S region or

four-stranded G-quartets (G4 DNA) stabilized by base pairing

between G residues (169, 170).

These observations provide an important link between the

requirement for germline transcription and the activity of AID

in CSR, and they have led to the following model: transcrip-

tion through S regions induces the formation of single-

stranded DNA that serves as a substrate for AID (163). The

generation of stable R loops may be one means to generate

such single-stranded DNA structures, although other mechan-

isms must exist (168, 171). AID-mediated deamination of

cytidine to uridine results in a dU/dG mismatch, which then

serves as a substrate for either the base excision repair (BER) or

the mismatch repair (MMR) pathways, both of which generate

a single-strand nick or gap at the site of the mismatch (172).

Deamination of the proximal cytidine bases on opposing

strands would result in a staggered DSB within S regions

(Fig. 3). Consistent with such a model, mutations of either

BER or MMR factors alter CSR in mice (173–179).

There is significant evidence in support of CSR proceeding

through a DSB intermediate. First, CSR between two S regions

results in the deletion of the intervening sequence (180), and the

excised products can be detected as extrachromosomal circles

following CSR (181–183). Second, DSBs in S regions of B cells

stimulated to undergo CSR can be detected by ligation-mediated

polymerase chain reaction (LM-PCR) (184), and these breaks

appear to be AID dependent (185). Finally, g-H2AX foci, which

form at sites of DNA damage, associate with the IgH locus of

activated B cells in an AID-dependent manner (186). However, a

word of caution is needed. LM-PCR also detects DSBs in V genes

in an AID-independent manner (187–189), and g-H2AX foci also

can be found in V regions following induction of AID (190).

Therefore, while it appears quite likely that DSBs are an integral

part of the CSR mechanism, it remains to be formally proven.

Allowing for CSR-associated DSBs, several studies suggest

that the repair of these breaks may involve a subset of the NHEJ

factors. Switch junctions cloned from wildtype B cells display

little, if any, homology (191), demonstrating that HR is not

involved in this process. Moreover, in studies of mature B cells

Class switch product Excised circle

IgG2a

DNA-PKcsNHEJ?H2AX

Double-strand break

UNG/APE1

AID

R-loop

Cγ2a

γ3 γ1

γ2bµδ

u u u u u

c c

Cµ Cγ2a

Cγ2a

Cγ2a

Cγ2a

µ δ γ 3 γ 1 γ 2b γ 2a ε α

c c c

IgM

A

B

c

u

Sγ2a

Fig. 2. Immunoglobulin (Ig) heavy chain (HC) class switchrecombination (CSR). (A) Germline organization of the IgH constantregions. (B) Model for the generation of activation-induced deaminase(AID)-dependent double-strand breaks during CSR.

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 121

Page 8: The role of the non-homologous end-joining pathway in lymphocyte development

reconstituted by IgH and LC knock-in alleles, inactivation of

certain NHEJ factors dramatically reduces CSR. Both Ku70- and

Ku80-deficient B cells are unable to undergo CSR either in vivo

or in vitro (192, 193). However, as deletion of either Ku70 or

Ku80 results in a general cellular proliferation defect (37, 86),

it is unclear whether Ku-deficient B cells are defective in CSR

per se or simply die due to other causes during the proliferative

stage of B-cell activation prior to CSR. While the finding that a

subset of Ku80-deficient B cells that undergo multiple of the

former rounds of cell division still fail to switch is strongly

suggestive (194), it is possible that this population represents

cells that were not fully activated for CSR.

Like Ku70/80, DNA-PKcs appears to be involved, either

directly or indirectly, in CSR, as deletion of DNA-PKcs blocks

CSR to all isotypes except for IgG1 (193). Unlike Ku-deficient

cells, this effect appears to be specific as DNA-PKcs-null B cells

proliferate normally in vitro and switching is not rescued by

crossing p53 deficiency (193). By comparison, B cells from

SCID mice appear to undergo CSR to all IgH isotypes, albeit at

reduced levels (195, 196). While differences in strain back-

ground and specificity of the knock-in Ig receptor must be

accounted for, the difference in CSR observed in DNA-PKcs-

null and SCID mice may reflect the complete loss of function in

the former and partial loss of function in the latter. In this case,

activities of DNA-PKcs not directly related to its kinase activity

may be those relevant to CSR.

The role of DNA-PK in CSR is clearly different from its role

in V(D)J recombination. Whereas coding end hairpins must

Chromosome 12

D V

RAG

DSB replication

Invasion

Chromosome 15

Random breakage

C12;15 translocation

Replication

Replication

Breakage

Fusion

Telomere donation fromthird chromosome

Complicon with co-amplified IgH and c-myc

Breakage-fusion-bridge cycles

c-myc

J

J

J

Fig. 3. Breakage–fusion–bridge cycle

promotes lymphogenesis in the absenceof proper double-strand break repair and

functional cell-cycle checkpoints. Modeldetailing the molecular events leading toco-amplification of IgH/c-myc in NHEJ/p53pro-B cell tumors is mediated throughbreakage–fusion–bridge cycle.

Rooney et al � The NHEJ pathway in lymphocyte development

122 Immunological Reviews 200/2004

Page 9: The role of the non-homologous end-joining pathway in lymphocyte development

be processed through the co-ordinate activity of DNA-PKcs

and Artemis, CSR occurs at wildtype levels in the absence of

Artemis (Rooney, Manis, and Alt, in preparation). Therefore,

processing of DSBs by Artemis is not required for CSR, despite

evidence that these breaks have staggered ends (197). It is

possible that staggered breaks are not the only relevant DSB in

CSR, that a CSR-specific nuclease is required to process these

ends (198, 199), or that CSR-associated breaks are repaired by

a different pathway. Additionally, given that both Ku and

DNA-PKcs have functions that lie beyond NHEJ (101), the

impact of mutations on these factors, while specific for CSR,

cannot be taken as absolute evidence for a role of the NHEJ

pathway in this process. Therefore, examination of CSR in the

absence of XRCC4 or ligase IV, the two factors thought to be

exclusively involved in all NHEJ repair, is needed to directly

assess the relevance of this pathway to CSR.

Several factors outside of NHEJ have been implicated in the

repair of CSR-associated breaks. In particular, H2AX-deficient

mice are defective for CSR (186); however, intra-switch-

region deletions (ISRs), a related AID-dependent process

(200), remain intact (194). H2AX is phosphorylated

(g-H2AX) in response to DNA damage, and subsequently, it

recruits a number of different DNA-repair factors to the site

of DSBs. g-H2AX has also been proposed to facilitate repair

by anchoring broken DNA ends in proximity through its

interaction with several other factors including 53BP1, the

Mre11–Rad50–Nbs1 complex, and MDC1 (201). Similarly,

H2AX may be required to ensure the long-range interaction

of two S regions necessary for proper CSR (194, 201). Con-

sistent with a role for H2AX in CSR, H2AX haploinsuffi-

ciency, in the context of p53 deficiency, results in the

appearance of B-cell lymphomas with translocations invol-

ving IgH S regions (202). Moreover, loss of the DNA damage

response factor 53BP1 also severely impairs CSR, further

supporting the notion that H2AX and 53BP1 may work

together, as well as with other damage response factors, in

the joining phase of the CSR (Manis et al., submitted for

publication).

The exact timing of AID-induced DSBs remains unresolved.

Single-strand mutations accumulate during G1 in a tumor cell

line stimulated to undergo SHM (189). Similarly, AID-

dependent g-H2AX foci are recruited to switch regions in

activated B cells during G1/early S phase (186). However, it

remains possible that in the case of CSR, AID introduces

mutations on a single DNA strand in G1, and following the

activity of the BER/MMR pathway(s), a single-strand nick is

converted to a DSB by DNA replication of the IgH locus in S

phase. Such a model would be in agreement with the observa-

tion that CSR requires cell division (203–205) and is blocked

by DNA replication inhibitors. As the choice of DSB repair

pathways can be greatly influenced by the stage of the cell

cycle that breaks occur (14, 15), determining the exact timing

of AID-induced breaks remains quite important.

The consequences of defective DSB repair in

lymphocytes

The importance of proper DNA maintenance is underscored

by the observation that inactivation of DNA repair pathways,

from yeast to mammalian cell lines, results in an increased

level of genomic instability (206, 207). While NHEJ is

required for V(D)J recombination, this process is a specific

example of the more general function of this pathway: to

appropriately resolve DSBs, regardless of the causative agent.

As a result, fibroblasts from NHEJ-deficient mice have an

increased level of IR-sensitivity, as well as elevated levels of

spontaneously occurring chromosomal aberrations (93, 99,

153, 208–211).

The types of karyotypic abnormalities observed in these cells

include broken chromosomes, broken chromatids, chromo-

some gaps, chromosome fusions, aneuploidy, and as well as

translocations. In particular, recurrent translocations are diag-

nostic of a number of human tumors (212), and may result in

the deletion of tumor suppressor genes, the deregulation or

amplification of protooncogenes, or the production of novel

fusion proteins, all of which can promote oncogenesis (213).

Therefore, by limiting such genome instability, NHEJ has a

role in suppressing tumor formation.

When coupled with the inactivation of cell-cycle check-

points, the absence of NHEJ can lead to tumor formation

(213). While mice doubly deficient in any of the six known

NHEJ factors and p53 (NHEJ/p53) appear relatively normal at

birth, these mice all have an increased mortality rate and an

elevated level of tumor incidence (98, 99, 153, 154, 214–

216). Nearly all tumors arising in NHEJ/p53-deficient mice

(with the notable exception of Artemis/p53 mice, see below)

are B220þIgM–CD43þ progenitor (pro)-B-cell lymphomas,

characterized by a marker der(12)t(12;15) translocation

(referred to as C12;15) (99, 153, 217–220) and co-

amplification of IgH and c-myc on a complex translocation

product, referred to as a complicon (220).

Based on the identification of several key structural inter-

mediates by fluorescence in situ hybridization and chromo-

some painting techniques, the molecular mechanisms

underlying the transformation of NHEJ/p53 pro-B cells have

been elucidated (217, 220) (Fig. 3). In the absence of p53,

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 123

Page 10: The role of the non-homologous end-joining pathway in lymphocyte development

unrepaired RAG-induced DSBs at the IgH locus on chromosome

12 are allowed to persist into S phase, where the broken

chromosome 12 is replicated and recombines with chromosome

15 to produce a clonal C12;15, as well as a dicentric 12;15

chromosome. This dicentric intermediate is then subjected to

multiple rounds of a breakage–fusion–bridge (BFB) cycle,

terminated following recombination with the distal portion of

a third unrelated chromosome, thereby resulting in the forma-

tion of a complicon harboring co-amplified IgH and c-myc.

Despite the progress made in understanding the role of

NHEJ in suppressing lymphomas in mice, it remains unclear

what role defects in NHEJ play in human tumors. While there

is some evidence that polymorphisms of NHEJ factors, par-

ticularly ligase IV, can influence the susceptibility to various

cancers (147, 221–224), such analysis is complicated by the

fact that Artemis is the only NHEJ factor known to be inacti-

vated in humans, potentially due to a more severe impact of

NHEJ mutations in human versus mouse cells (225). In this

context, hypomorphic alleles of Artemis have been associated

with human lymphoid malignancies harboring chromosomal

translocations (226), although it is unclear whether this is

secondary to infection with Epstein–Barr virus.

Artemis/p53-deficient mice also develop pro-B cell tumors

at an age comparable to other NHEJ/p53 mice, and c-myc

amplification and C12;15 translocation were observed in a

subset of Artemis/p53 tumors (216). However, rather than

c-myc, a majority of Artemis/p53 pro-B cells exhibit amplifica-

tion of the related oncogene N-myc. As amplification of both

N-myc and c-myc appears to occur through similar BFB cycles

(216), it is of interest to speculate why only c-myc amplifica-

tion is observed in other NHEJ/p53 mice. While mutation of

either c-myc or N-myc is lethal embryonically (227, 228),

replacement of c-myc with N-myc has little effect on develop-

ment (229), suggesting that the functional difference between

the two lies in their expression patterns. Given the appearance

of both N-myc and c-myc amplification in Artemis/p53 tumors,

it is likely that the oncogenic potential of N-myc and c-myc is

also largely equivalent. Therefore, whether due to strain dif-

ferences or the distinct impact of the loss of different NHEJ

factors, the conserved amplification of c-myc in most NHEJ/

p53 pro-B-cell tumors appears to be the result of preferential

translocation of IgH into the regions surrounding c-myc, rather

than selection for c-myc.

Switch translocations

Aberrant repair of V(D)J-associated DSBs is not alone in its

ability to promote chromosomal translocations. Translocation

breakpoints involving IgH S regions and c-myc have been cloned

from a number of lymphomas from both mice and humans

(213), indicating that inappropriate resolution of CSR inter-

mediates can promote lymphomagenesis as well. Unlike the

complex translocations observed in NHEJ/p53 pro-B cells,

these translocations tend to be simple and balanced, with the

switch region translocated directly into either the first exon or

the first intron of c-myc. Whether NHEJ deficiencies are capable

of promoting switch-region translocation in mature B cells

remains to be determined. However, H2AXþ/Dp53–/– mice

develop a wide spectrum of lymphoid tumors, including

B-lineage tumors with translocation breakpoints involving

Sm and c-myc (202), suggesting that haploinsufficiency for

H2AX can promote CSR-associated translocations in a

p53-deficient background.

Concluding remarks

While much is understood about the NHEJ pathway, there is

clearly much left to be determined. Both genetic and biochem-

ical data suggest that V(D)J recombination requires additional

factors (75, 230, 231); these may include polynucleotide

kinase, which restores 50 phosphate ends (232), or one of

several gap-filling polymerases (233–236). Similarly, general

DSB repair by NHEJ may require other nucleases to deal with

ends that cannot be processed by Artemis, such as WRN (107,

136), FEN-1 (135), or the Mre11 complex (137).

One additional factor that modulates the resolution of V(D)J

intermediates is RAG itself. Paired cleavage of gene segments

ensures that RAG does not introduce a single DSB within the

genome (237). Furthermore, following the generation of two

DSBs, RAG remains bound in a post-synaptic cleavage com-

plex, holding DNA ends in proximity to one another (238). In

this manner, RAG appears to facilitate the productive inter-

action between cleaved gene segments. Similarly, it remains to

be determined how the RAG complex influences the choice of

DNA repair pathways and how the post-synaptic cleavage

complex interacts with various members of the NHEJ pathway.

Further examination of the potential role of DNA damage

response pathways in V(D)J recombination is warranted.

While H2AX or 53BP1 is not required for V(D)J recombin-

ation per se (239–242), H2AX may prevent aberrant processing

of V(D)J ends (201). Furthermore, in cells with multiple DSBs,

NHEJ tends to join ends from a given DSB back to each other,

rather than to those of another DSB (210). Therefore, these

ends must be juxtaposed with one another within the context

of chromosomal DNA. While this may be achieved in V(D)J

recombination by the RAG proteins, the anchoring of general

Rooney et al � The NHEJ pathway in lymphocyte development

124 Immunological Reviews 200/2004

Page 11: The role of the non-homologous end-joining pathway in lymphocyte development

DSBs may be promoted by g-H2AX and by its associated DNA

damage factors (201). Further elucidation of such potential

functions of these proteins may yield significant insights into

both lymphocyte-specific and more general translocations.

Unlike V(D)J recombination, little is understood about the

resolution of CSR. Clearly, AID and germline transcription are

required, but the intermediates generated by this process

remain ill-defined. Particularly, while most studies point to

the introduction of DSBs during CSR, direct evidence has

remained elusive. In the same context, the role of NHEJ in

CSR needs to be clarified. Examination of CSR in the absence of

XRCC4 or ligase IV should provide the most tractable system to

evaluate this question, as both factors are required for all NHEJ

reactions, and they do not appear to have functions outside of

this pathway. Similarly, it remains to be determined how the

productive interaction of two switch regions is ensured. H2AX

has been proposed to anchor DNA ends through multiple

protein–protein and protein–DNA interactions (201). There-

fore, systematic evaluation of the role of H2AX-associated

factors in CSR may give further insight into the joining phase

of this reaction.

A complete index of the function(s) of DNA-PK is also

required. This complex appears to have a multitude of roles,

not all of which are involved in NHEJ per se (101, 243), and

some of which, such as the protection of telomeres from

fusions (102), appear to be in direct contrast with its role in

NHEJ. Therefore, the function of the holoenzyme most likely

is modulated in some manner, possibly by interacting with

various partners in different contexts or by post-translational

modification.

References

1. Jung D, Alt FW. Unraveling V(D)J

recombination: insights into gene regulation.

Cell 2004;116:299–311.

2. Nossal GJ. The double helix and immunology.

Nature 2003;421:440–444.

3. Arden B. Conserved motifs in T-cell receptor

CDR1 and CDR2: implications for ligand and

CD8 co-receptor binding. Curr Opin

Immunol 1998;10:74–81.

4. Roth DB, Nakajima PB, Menetski JP,

Bosma MJ, Gellert M. Double-strand breaks

associated with V(D)J recombination at the

TCR delta locus in murine thymocytes. Curr

Top Microbiol Immunol 1992;182:115–124.

5. McBlane JF, van Gent DC, Ramsden DA,

Romeo C, Cuomo CA, Gellert M, Oettinger MA.

Cleavage at a V(D)J recombination signal

requires only RAG1 and RAG2 proteins and

occurs in two steps. Cell 1995;83:387–395.

6. Schlissel MS. Structure of nonhairpin coding-

end DNA breaks in cells undergoing V(D)J

recombination. Mol Cell Biol 1998;18:

2029–2037.

7. Livak F, Schatz DG. Identification of V(D)J

recombination coding end intermediates in

normal thymocytes. J Mol Biol 1997;267:1–9.

8. Lewis SM. P nucleotides, hairpin DNA and

V(D)J joining: making the connection. Semin

Immunol 1994;6:131–141.

9. Schlissel MS. Does Artemis end the hunt for

the hairpin-opening activity in V(D)J

recombination? Cell 2002;109:1–4.

10. Alt FW, Baltimore D. Joining of

immunoglobulin heavy chain gene segments:

implications from a chromosome with

evidence of three D-JH fusions. Proc Natl

Acad Sci USA 1982;79:4118–4122.

11. Gilfillan S, Dierich A, Lemeur M, Benoist C,

Mathis D. Mice lacking TdT: mature animals

with an immature lymphocyte repertoire.

Science 1993;261:1175–1178.

12. Komori T, Okada A, Stewart V, Alt FW. Lack

of N regions in antigen receptor variable

region genes of TdT-deficient lymphocytes

[published erratum appears in Science 1993;

262: 1957]. Science 1993;261:1171–1175.

13. Helleday T. Pathways for mitotic homologous

recombination in mammalian cells. Mutat Res

2003;532:103–115.

14. Takata M, et al. Homologous recombination

and non-homologous end-joining pathways

of DNA double-strand break repair have

overlapping roles in the maintenance of

chromosomal integrity in vertebrate cells.

EMBO J 1998;17:5497–5508.

15. Rothkamm K, Kruger I, Thompson LH,

Lobrich M. Pathways of DNA double-strand

break repair during the mammalian cell cycle.

Mol Cell Biol 2003;23:5706–5715.

16. Lin WC, Desiderio S. V(D)J recombination

and the cell cycle. Immunol Today

1995;16:279–289.

17. Smith GC, Jackson SP. The DNA-dependent

protein kinase. Genes Dev 1999;13:916–934.

18. Bassing CH, Swat W, Alt FW. The mechanism

and regulation of chromosomal V(D)J

recombination. Cell 2002;109:S45–S55.

19. Taccioli GE, Alt FW. Potential targets for

autosomal SCID mutations. Curr Opin

Immunol 1995;7:436–440.

20. Moshous D, et al. Artemis, a novel DNA

double-strand break repair/V(D)J

recombination protein, is mutated in human

severe combined immune deficiency. Cell

2001;105:177–186.

21. Lewis LK, Resnick MA. Tying up loose ends:

nonhomologous end-joining in

Saccharomyces cerevisiae. Mutat Res

2000;451:71–89.

22. Weller GR, et al. Identification of a DNA

nonhomologous end-joining complex in

bacteria. Science 2002;297:1686–1689.

23. Jeggo P, O’Neill P. The Greek Goddess,

Artemis, reveals the secrets of her cleavage.

DNA Repair (Amst) 2002;1:771–777.

24. Jackson SP, MacDonald JJ, Lees-Miller S,

Tjian R. GC box binding induces

phosphorylation of Sp1 by a DNA-dependent

protein kinase. Cell 1990;63:155–165.

25. Carter T, Vancurova I, Sun I, Lou W,

DeLeon S. A DNA-activated protein kinase

from HeLa cell nuclei. Mol Cell Biol 1990;10:

6460–6471.

26. Walker AI, Hunt T, Jackson RJ, Anderson CW.

Double-stranded DNA induces the

phosphorylation of several proteins including

the 90 000 mol. wt. heat-shock protein in

animal cell extracts. EMBO J 1985;4:139–

145.

27. Anderson CW, Lees-Miller SP. The nuclear

serine/threonine protein kinase DNA-PK. Crit

Rev Eukaryot Gene Expr 1992;2:283–314.

28. Suwa A, Hirakata M, Takeda Y, Jesch SA,

Mimori T, Hardin JA. DNA-dependent protein

kinase (Ku protein-p350 complex) assembles

on double-stranded DNA. Proc Natl Acad Sci

USA 1994;91:6904–6908.

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 125

Page 12: The role of the non-homologous end-joining pathway in lymphocyte development

29. Gottlieb TM, Jackson SP. The DNA-dependent

protein kinase: requirement for DNA ends

and association with Ku antigen. Cell

1993;72:131–142.

30. Dvir A, Peterson SR, Knuth MW, Lu H,

Dynan WS. Ku autoantigen is the regulatory

component of a template-associated protein

kinase that phosphorylates RNA polymerase II.

Proc Natl Acad Sci USA 1992;89:11920–11924.

31. Mimori T, Akizuki M, Yamagata H, Inada S,

Yoshida S, Homma M. Characterization of a

high molecular weight acidic nuclear protein

recognized by autoantibodies in sera from

patients with polymyositis-scleroderma

overlap. J Clin Invest 1981;68:611–620.

32. Featherstone C, Jackson SP. Ku, a DNA repair

protein with multiple cellular functions?

Mutat Res 1999;434:3–15.

33. Taccioli GE, et al. Ku80: product of the

XRCC5 gene and its role in DNA repair and

V(D)J recombination. Science

1994;265:1442–1445.

34. Errami A, et al. Ku86 defines the genetic

defect and restores X-ray resistance and V(D)J

recombination to complementation group 5

hamster cell mutants. Mol Cell Biol

1996;16:1519–1526.

35. Smider V, Rathmell WK, Lieber MR, Chu G.

Restoration of X-ray resistance and V(D)J

recombination in mutant cells by Ku cDNA.

Science 1994;266:288–291.

36. Gu Y, Jin S, Gao Y, Weaver DT, Alt FW.

Ku70-deficient embryonic stem cells have

increased ionizing radiosensitivity, defective

DNA end-binding activity, and inability to

support V(D)J recombination. Proc Natl Acad

Sci USA 1997;94:8076–8081.

37. Nussenzweig A, et al. Requirement for Ku80

in growth and immunoglobulin V(D)J

recombination. Nature 1996;382:551–555.

38. Doherty AJ, Jackson SP. DNA repair: how

Ku makes ends meet. Curr Biol

2001;11:R920–R924.

39. Fukushima T, et al. Genetic analysis of the

DNA-dependent protein kinase reveals an

inhibitory role of Ku in late S-G2 phase DNA

double-strand break repair. J Biol Chem

2001;276:44413–44418.

40. Adachi N, Ishino T, Ishii Y, Takeda S,

Koyama H. DNA ligase IV-deficient cells are

more resistant to ionizing radiation in the

absence of Ku70: implications for DNA

double-strand break repair. Proc Natl Acad Sci

USA 2001;98:12109–12113.

41. Pierce AJ, Hu P, Han M, Ellis N, Jasin M.

Ku DNA end-binding protein modulates

homologous repair of double-strand breaks

in mammalian cells. Genes Dev

2001;15:3237–3242.

42. Wu X, Lieber M. Protein–protein and

protein–DNA interaction regions within the

DNA end-binding protein Ku70–Ku86. Mol

Cell Biol 1996;16:5186–5193.

43. Ochem AE, et al. Functional properties of the

separate subunits of human DNA helicase II/

Ku autoantigen. J Biol Chem

1997;272:29919–29926.

44. Ono M, Tucker PW, Capra JD. Production and

characterization of recombinant human Ku

antigen. Nucleic Acids Res 1994;22:

3918–3924.

45. Griffith AJ, Blier PR, Mimori T, Hardin JA. Ku

polypeptides synthesized in vitro assemble into

complexes which recognize ends of double-

stranded DNA. J Biol Chem 1992;267:

331–338.

46. Gell D, Jackson SP. Mapping of protein–

protein interactions within the DNA-

dependent protein kinase complex. Nucleic

Acids Res 1999;27:3494–3502.

47. Dynan WS, Yoo S. Interaction of Ku protein

and DNA-dependent protein kinase catalytic

subunit with nucleic acids. Nucleic Acids Res

1998;26:1551–1559.

48. Walker JR, Corpina RA, Goldberg J. Structure

of the Ku heterodimer bound to DNA and its

implications for double-strand break repair.

Nature 2001;412:607–614.

49. Feldmann E, Schmiemann V, Goedecke W,

Reichenberger S, Pfeiffer P. DNA double-

strand break repair in cell-free extracts from

Ku80-deficient cells: implications for Ku

serving as an alignment factor in non-

homologous DNA end joining. Nucleic Acids

Res 2000;28:2585–2596.

50. Kabotyanski EB, Gomelsky L, Han JO,

Stamato TD, Roth DB. Double-strand break

repair in Ku86- and XRCC4-deficient cells.

Nucleic Acids Res 1998;26:5333–5342.

51. Singleton BK, Torres-Arzayus MI,

Rottinghaus ST, Taccioli GE, Jeggo PA. The C

terminus of Ku80 activates the DNA-

dependent protein kinase catalytic subunit.

Mol Cell Biol 1999;19:3267–3277.

52. Calsou P, Frit P, Humbert O, Muller C,

Chen DJ, Salles B. The DNA-dependent

protein kinase catalytic activity regulates DNA

end processing by means of Ku entry into

DNA. J Biol Chem 1999;274:7848–7856.

53. Yoo S, Kimzey A, Dynan WS. Photocross-

linking of an oriented DNA repair complex.

Ku bound at a single DNA end. J Biol Chem

1999;274:20034–20039.

54. Hartley KO, et al. DNA-dependent protein

kinase catalytic subunit: a relative of

phosphatidylinositol 3-kinase and the ataxia

telangiectasia gene product. Cell

1995;82:849–856.

55. Poltoratsky VP, Shi X, York JD, Lieber MR,

Carter TH. Human DNA-activated protein

kinase (DNA-PK) is homologous to

phosphatidylinositol kinases. J Immunol

1995;155:4529–4533.

56. Djordjevic S, Driscoll PC. Structural insight

into substrate specificity and regulatory

mechanisms of phosphoinositide 3-kinases.

Trends Biochem Sci 2002;27:426–432.

57. Lees-Miller SP, Sakaguchi K, Ullrich SJ,

Appella E, Anderson CW. Human DNA-

activated protein kinase phosphorylates

serines 15 and 37 in the amino-terminal

transactivation domain of human p53. Mol

Cell Biol 1992;12:5041–5049.

58. O’Neill T, et al. Utilization of oriented

peptide libraries to identify substrate motifs

selected by ATM. J Biol Chem

2000;275:22719–22727.

59. Durocher D, Jackson SP. DNA-PK, ATM and

ATR as sensors of DNA damage: variations on a

theme? Curr Opin Cell Biol 2001;13: 225–231.

60. Blackwell TK, et al. Isolation of scid pre-B

cells that rearrange kappa light chain genes:

formation of normal signal and abnormal

coding joins. EMBO J 1989;8:735–742.

61. Lieber MR, et al. The defect in murine severe

combined immune deficiency: joining of

signal sequences but not coding segments in

V(D)J recombination. Cell 1988;55:7–16.

62. Malynn BA, et al. The scid defect affects the

final step of the immunoglobulin VDJ

recombinase mechanism. Cell 1988;54:

453–460.

63. Taccioli GE, Cheng HL, Varghese AJ,

Whitmore G, Alt FW. A DNA repair defect in

Chinese hamster ovary cells affects V(D)J

recombination similarly to the murine scid

mutation. J Biol Chem 1994;269:7439–7442.

64. Peterson SR, Kurimasa A, Oshimura M,

Dynan WS, Bradbury EM, Chen DJ. Loss of

the catalytic subunit of the DNA-dependent

protein kinase in DNA double-strand-break-

repair mutant mammalian cells. Proc Natl

Acad Sci USA 1995;92:3171–3174.

65. Danska JS, Holland DP, Mariathasan S,

Williams KM, Guidos CJ. Biochemical and

genetic defects in the DNA-dependent

protein kinase in murine scid lymphocytes.

Mol Cell Biol 1996;16:5507–5517.

66. Blunt T, Gell D, Fox M, Taccioli GE,

Lehmann AR, Jackson SP, Jeggo PA.

Identification of a nonsense mutation in the

carboxyl-terminal region of DNA-dependent

protein kinase catalytic subunit in the scid

mouse. Proc Natl Acad Sci USA

1996;93:10285–10290.

67. Araki R, et al. Nonsense mutation at Tyr-

4046 in the DNA-dependent protein kinase

catalytic subunit of severe combined immune

deficiency mice. Proc Natl Acad Sci USA

1997;94:2438–2443.

68. Beamish HJ, et al. The C-terminal conserved

domain of DNA-PKcs, missing in the SCID

mouse, is required for kinase activity. Nucleic

Acids Res 2000;28:1506–1513.

Rooney et al � The NHEJ pathway in lymphocyte development

126 Immunological Reviews 200/2004

Page 13: The role of the non-homologous end-joining pathway in lymphocyte development

69. Kurimasa A, et al. Requirement for the kinase

activity of human DNA-dependent protein

kinase catalytic subunit in DNA strand break

rejoining. Mol Cell Biol 1999;19:3877–3884.

70. Douglas P, et al. Identification of in vitro and

in vivo phosphorylation sites in the catalytic

subunit of the DNA-dependent protein

kinase. Biochem J 2002;368:243–251.

71. Chan DW, Chen BP, Prithivirajsingh S,

Kurimasa A, Story MD, Qin J, Chen DJ.

Autophosphorylation of the DNA-dependent

protein kinase catalytic subunit is required for

rejoining of DNA double-strand breaks.

Genes Dev 2002;16:2333–2338.

72. Ding Q, et al. Autophosphorylation of the

catalytic subunit of the DNA-dependent

protein kinase is required for efficient end

processing during DNA double-strand break

repair. Mol Cell Biol 2003;23:5836–5848.

73. Weterings E, Verkaik NS, Bruggenwirth HT,

Hoeijmakers JH, van Gent DC. The role of

DNA dependent protein kinase in synapsis

of DNA ends. Nucleic Acids Res 2003;31:

7238–7246.

74. Kuhne C, Tjornhammar ML, Pongor S,

Banks L, Simoncsits A. Repair of a minimal

DNA double-strand break by NHEJ requires

DNA-PKcs and is controlled by the ATM/ATR

checkpoint. Nucleic Acids Res

2003;31:7227–7237.

75. Baumann P, West SC. DNA end-joining

catalyzed by human cell-free extracts. Proc

Natl Acad Sci USA 1998;95:14066–14070.

76. Verkaik NS, et al. Different types of V(D)J

recombination and end-joining defects in

DNA double-strand break repair mutant

mammalian cells. Eur J Immunol

2002;32:701–709.

77. van Heemst D, Brugmans L, Verkaik NS, van

Gent DC. End-joining of blunt DNA double-

strand breaks in mammalian fibroblasts is

precise and requires DNA-PK and XRCC4.

DNA Repair (Amst) 2004;3:43–50.

78. Nick McElhinny SA, Snowden CM,

McCarville J, Ramsden DA. Ku recruits the

XRCC4–ligase IV complex to DNA ends. Mol

Cell Biol 2000;20:2996–3003.

79. Kysela B, et al. Ku stimulation of DNA ligase

IV-dependent ligation requires inward

movement along the DNA molecule. J Biol

Chem 2003;278:22466–22474.

80. Chen L, Trujillo K, Sung P, Tomkinson AE.

Interactions of the DNA ligase IV–XRCC4

complex with DNA ends and the DNA-

dependent protein kinase. J Biol Chem

2000;275:26196–26205.

81. Calsou P, Delteil C, Frit P, Drouet J, Salles B.

Coordinated assembly of Ku and p460

subunits of the DNA-dependent protein

kinase on DNA ends is necessary for XRCC4–

ligase IV recruitment. J Mol Biol

2003;326:93–103.

82. Hsu HL, Yannone SM, Chen DJ. Defining

interactions between DNA-PK and ligase IV/

XRCC4. DNA Repair (Amst) 2002;1:225–235.

83. Merkle D, et al. The DNA-dependent protein

kinase interacts with DNA to form a protein-

DNA complex that is disrupted by

phosphorylation. Biochemistry

2002;41:12706–12714.

84. DeFazio LG, Stansel RM, Griffith JD, Chu G.

Synapsis of DNA ends by DNA-dependent

protein kinase. EMBO J 2002;21:3192–3200.

85. Izsvak Z, Stuwe EE, Fiedler D, Katzer A,

Jeggo PA, Ivics Z. Healing the wounds

inflicted by sleeping beauty transposition by

double-strand break repair in Mammalian

somatic cells. Mol Cell 2004;13:279–290.

86. Gu Y, et al. Growth retardation and leaky

SCID phenotype of Ku70-deficient mice.

Immunity 1997;7:653–665.

87. Ouyang H, et al. Ku70 is required for DNA

repair but not for T cell antigen receptor gene

recombination in vivo. J Exp Med

1997;186:921–929.

88. Zhu C, Bogue MA, Lim DS, Hasty P, Roth DB.

Ku86-deficient mice exhibit severe combined

immunodeficiency and defective processing

of V(D)J recombination intermediates. Cell

1996;86:379–389.

89. Kurimasa A, et al. Catalytic subunit of

DNA-dependent protein kinase: impact on

lymphocyte development and tumorigenesis.

Proc Natl Acad Sci USA 1999;96:1403–1408.

90. Taccioli GE, et al. Targeted disruption of the

catalytic subunit of the DNA-PK gene in mice

confers severe combined immunodeficiency and

radiosensitivity. Immunity 1998;9:355–366.

91. Jhappan C, Morse HC III, Fleischmann RD,

Gottesman MM, Merlino G. DNA-PKcs: a

T-cell tumour suppressor encoded at the mouse

scid locus. Nat Genet 1997;17:483–486.

92. Gao Y, Chaudhuri J, Zhu C, Davidson L,

Weaver DT, Alt FW. A targeted DNA-PKcs-

null mutation reveals DNA-PK-independent

functions for KU in V(D)J recombination.

Immunity 1998;9:367–376.

93. Rooney S, et al. Leaky scid phenotype

associated with defective V(D)J coding end

processing in Artemis-deficient mice. Mol

Cell 2002;10:1379–1390.

94. Fukumura R, et al. Signal joint formation is

also impaired in DNA-dependent protein

kinase catalytic subunit knockout cells.

J Immunol 2000;165:3883–3889.

95. Jhappan C, Yusufzai TM, Anderson S,

Anver MR, Merlino G. The p53 response to

DNA damage in vivo is independent of

DNA-dependent protein kinase. Mol Cell Biol

2000;20:4075–4083.

96. Gu Y, et al. Defective embryonic

neurogenesis in Ku-deficient but not DNA-

dependent protein kinase catalytic subunit-

deficient mice. Proc Natl Acad Sci USA

2000;97:2668–2673.

97. Vogel H, Lim DS, Karsenty G, Finegold M,

Hasty P. Deletion of Ku86 causes early onset

of senescence in mice. Proc Natl Acad Sci

USA 1999;96:10770–10775.

98. Lim DS, Vogel H, Willerford DM, Sands AT,

Platt KA, Hasty P. Analysis of ku80-mutant

mice and cells with deficient levels of p53.

Mol Cell Biol 2000;20:3772–3780.

99. Difilippantonio MJ, et al. DNA repair

protein Ku80 suppresses chromosomal

aberrations and malignant transformation.

Nature 2000;404:510–514.

100. Gurley KE, Vo K, Kemp CJ. DNA double-

strand breaks, p53, and apoptosis during

lymphomagenesis in scid/scid mice. Cancer

Res 1998;58:3111–3115.

101. Sekiguchi J, et al. Genetic interactions

between ATM and the nonhomologous

end-joining factors in genomic stability and

development. Proc Natl Acad Sci USA

2001;98:3243–3248.

102. de Lange T. Protection of mammalian

telomeres. Oncogene 2002;21:532–540.

103. Paull TT, Rogakou EP, Yamazaki V,

Kirchgessner CU, Gellert M, Bonner WM. A

critical role for histone H2AX in recruitment

of repair factors to nuclear foci after DNA

damage. Curr Biol 2000;10:886–895.

104. Park EJ, Chan DW, Park JH, Oettinger MA,

Kwon J. DNA-PK is activated by

nucleosomes and phosphorylates H2AX

within the nucleosomes in an acetylation-

dependent manner. Nucleic Acids Res

2003;31:6819–6827.

105. Furuta T, et al. Phosphorylation of histone

H2AX and activation of Mre11, Rad50, and

Nbs1 in response to replication-dependent

DNA double-strand breaks induced by

mammalian DNA topoisomerase I cleavage

complexes. J Biol Chem 2003;278:

20303–20312.

106. Burma S, Chen BP, Murphy M, Kurimasa A,

Chen DJ. ATM phosphorylates histone

H2AX in response to DNA double-strand

breaks. J Biol Chem 2001;276:42462–42467.

107. Karmakar P, Snowden CM, Ramsden DA,

Bohr VA. Ku heterodimer binds to both ends

of the Werner protein and functional

interaction occurs at the Werner N-terminus.

Nucleic Acids Res 2002;30:3583–3591.

108. Yannone SM, et al. Werner syndrome

protein is regulated and phosphorylated by

DNA-dependent protein kinase. J Biol Chem

2001;276:38242–38248.

109. Cooper MP, Machwe A, Orren DK, Brosh RM,

Ramsden D, Bohr VA. Ku complex interacts

with and stimulates the Werner protein.

Genes Dev 2000;14:907–912.

110. Novac O, Matheos D, Araujo FD, Price GB,

Zannis-Hadjopoulos M. In vivo association

of Ku with mammalian origins of

DNA replication. Mol Biol Cell 2001;12:

3386–3401.

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 127

Page 14: The role of the non-homologous end-joining pathway in lymphocyte development

111. Shao RG, Cao CX, Zhang H, Kohn KW,

Wold MS, Pommier Y. Replication-

mediated DNA damage by camptothecin

induces phosphorylation of RPA by DNA-

dependent protein kinase and dissociates

RPA: DNA-PK Complexes. EMBO J

1999;18:1397–1406.

112. Lucero H, Gae D, Taccioli GE. Novel

localization of the DNA-PK complex in lipid

rafts: a putative role in the signal

transduction pathway of the ionizing

radiation response. J Biol Chem

2003;278:22136–22143.

113. Chu W, et al. DNA-PKcs is required for

activation of innate immunity by

immunostimulatory DNA. Cell

2000;103:909–918.

114. Wang S, et al. The catalytic subunit of

DNA-dependent protein kinase selectively

regulates p53-dependent apoptosis but not

cell-cycle arrest. Proc Natl Acad Sci USA

2000;97:1584–1588.

115. Sawada M, Sun W, Hayes P, Leskov K,

Boothman DA, Matsuyama S. Ku70

suppresses the apoptotic translocation of

Bax to mitochondria. Nat Cell Biol

2003;5:320–329.

116. Fischer A, et al. Naturally occurring primary

deficiencies of the immune system. Annu

Rev Immunol 1997;15:93–124.

117. Cavazzana-Calvo M, Le Deist F, De Saint

Basile G, Papadopoulo D, De Villartay JP,

Fischer A. Increased radiosensitivity of

granulocyte macrophage colony-forming

units and skin fibroblasts in human

autosomal recessive severe combined

immunodeficiency. J Clin Invest

1993;91:1214–1218.

118. Schwarz K, Hansen-Hagge TE, Knobloch C,

Friedrich W, Kleihauer E, Bartram CR.

Severe combined immunodeficiency (SCID)

in man: B cell-negative (B-) SCID patients

exhibit an irregular recombination pattern

at the JH locus. J Exp Med 1991;174:

1039–1048.

119. 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 1998;188:627–634.

120. Nicolas N, et al. Lack of detectable defect

in DNA double-strand break repair and

DNA-dependent protein kinase activity in

radiosensitive human severe combined

immunodeficiency fibroblasts. Eur J

Immunol 1996;26:1118–1122.

121. Moshous D, et al. A new gene involved in

DNA double-strand break repair and V(D)J

recombination is located on human

chromosome 10p. Hum Mol Genet

2000;9:583–588.

122. Li L, Drayna D, Hu D, Hayward A,

Gahagan S, Pabst H, Cowan MJ. The gene

for severe combined immunodeficiency

disease in Athabascan-speaking Native

Americans is located on chromosome 10p.

Am J Hum Genet 1998;62:136–144.

123. Kobayashi N, et al. Novel Artemis gene

mutations of radiosensitive severe combined

immunodeficiency in Japanese families.

Hum Genet 2003;112:348–352.

124. Li L, et al. A founder mutation in Artemis,

an SNM1-like protein, causes SCID in

Athabascan-speaking Native Americans.

J Immunol 2002;168:6323–6329.

125. Noordzij JG, et al. Radiosensitive SCID

patients with Artemis gene mutations show

a complete B-cell differentiation arrest at the

pre-B-cell receptor checkpoint in bone

marrow. Blood 2003;101:1446–1452.

126. Wang Z, Fast W, Valentine AM, Benkovic SJ.

Metallo-beta-lactamase: structure and

mechanism. Curr Opin Chem Biol

1999;3:614–622.

127. Aravind L. An evolutionary classification of

the metallo-beta-lactamase fold proteins. In

Silico Biol 1999;1:69–91.

128. Callebaut I, Moshous D, Mornon JP,

de Villartay JP. Metallo-beta-lactamase fold

within nucleic acids processing enzymes:

the beta-CASP family. Nucleic Acids Res

2002;30:3592–3601.

129. Ma Y, Pannicke U, Schwarz K, Lieber MR.

Hairpin opening and overhang processing

by an Artemis/DNA-dependent protein

kinase complex in nonhomologous end

joining and V(D)J recombination. Cell

2002;108:781–794.

130. Rooney S, et al. Defective DNA repair and

increased genomic instability in Artemis-

deficient murine cells. J Exp Med

2003;197:553–565.

131. Kobayashi N, et al. Expansion of clonotype-

restricted HLA-identical maternal CD4(þ)

T cells in a patient with severe combined

immunodeficiency and a homozygous

mutation in the Artemis gene. Clin

Immunol 2003;108:159–166.

132. Paull TT, Gellert M. Nbs1 potentiates ATP-

driven DNA unwinding and endonuclease

cleavage by the Mre11/Rad50 complex.

Genes Dev 1999;13:1276–1288.

133. Besmer E, et al. Hairpin coding end opening

is mediated by RAG1 and RAG2 proteins.

Mol Cell 1998;2:817–828.

134. Shockett PE, Schatz DG. DNA hairpin

opening mediated by the RAG1 and RAG2

proteins. Mol Cell Biol 1999;19:

4159–4166.

135. Wu X, Wilson TE, Lieber MR. A role for

FEN-1 in nonhomologous DNA end

joining: the order of strand annealing and

nucleolytic processing events. Proc Natl

Acad Sci USA 1999;96:1303–1308.

136. Li B, Comai L. Requirements for the

nucleolytic processing of DNA ends by the

Werner syndrome protein-Ku70/80

complex. J Biol Chem 2001;276:

9896–9902.

137. Huang J, Dynan WS. Reconstitution of the

mammalian DNA double-strand break end-

joining reaction reveals a requirement for an

Mre11/Rad50/NBS1-containing fraction.

Nucleic Acids Res 2002;30:667–674.

138. Li Z, et al. The XRCC4 gene encodes a novel

protein involved in DNA double-strand

break repair and V(D)J recombination. Cell

1995;83:1079–1089.

139. Grawunder U, et al. Activity of DNA ligase

IV stimulated by complex formation with

XRCC4 protein in mammalian cells. Nature

1997;388:492–495.

140. Critchlow SE, Bowater RP, Jackson SP.

Mammalian DNA double-strand break

repair protein XRCC4 interacts with DNA

ligase IV. Curr Biol 1997;7:588–598.

141. Frank KM, et al. Late embryonic lethality

and impaired V(D)J recombination in mice

lacking DNA ligase IV. Nature

1998;396:173–177.

142. Gao Y, et al. A critical role for DNA end-

joining proteins in both lymphogenesis and

neurogenesis. Cell 1998;95:891–902.

143. Bryans M, Valenzano MC, Stamato TD.

Absence of DNA ligase IV protein in XR-1

cells: evidence for stabilization by XRCC4.

Mutat Res 1999;433:53–58.

144. Leber R, Wise TW, Mizuta R, Meek K. The

XRCC4 gene product is a target for and

interacts with the DNA-dependent protein

kinase. J Biol Chem 1998;273:1794–1801.

145. Modesti M, Hesse JE, Gellert M. DNA

binding of Xrcc4 protein is associated with

V(D)J recombination but not with

stimulation of DNA ligase IV activity. EMBO

J 1999;18:2008–2018.

146. O’Driscoll M, et al. DNA ligase IV mutations

identified in patients exhibiting

developmental delay and immuno-

deficiency. Mol Cell 2001;8:1175–1185.

147. Riballo E, et al. Identification of a defect in

DNA ligase IV in a radiosensitive leukaemia

patient. Curr Biol 1999;9:699–702.

148. Manke IA, Lowery DM, Nguyen A, Yaffe MB.

BRCT repeats as phosphopeptide-binding

modules involved in protein targeting.

Science 2003;302:636–639.

149. Rodriguez M, Yu X, Chen J, Songyang Z.

Phosphopeptide binding specificities of

BRCT domains. J Biol Chem

2003;278:52914–52918.

150. Yu X, Chini CC, He M, Mer G, Chen J. The

BRCT domain is a phospho-protein binding

domain. Science 2003;302:639–642.

Rooney et al � The NHEJ pathway in lymphocyte development

128 Immunological Reviews 200/2004

Page 15: The role of the non-homologous end-joining pathway in lymphocyte development

151. Barnes DE, Stamp G, Rosewell I, Denzel A,

Lindahl T. Targeted disruption of the gene

encoding DNA ligase IV leads to lethality

in embryonic mice. Curr Biol 1998;8:

1395–1398.

152. Sekiguchi JM, et al. Non-homologous end-

joining proteins are required for V(D)J

recombination, normal growth and

neurogenesis. Cold Spring Harb Symp

Quant Biol 1999;64:169–181.

153. Gao Y, et al. Interplay of p53 and DNA-

repair protein XRCC4 in tumorigenesis,

genomic stability and development. Nature

2000;404:897–900.

154. Frank KM, et al. DNA ligase IV deficiency in

mice leads to defective neurogenesis and

embryonic lethality via the p53 pathway.

Mol Cell 2000;5:993–1002.

155. Lee Y, Barnes DE, Lindahl T, McKinnon PJ.

Defective neurogenesis resulting from DNA

ligase IV deficiency requires Atm. Genes Dev

2000;14:2576–2580.

156. Love PE, Chan AC. Regulation of thymocyte

development: only the meek survive. Curr

Opin Immunol 2003;15:199–203.

157. Niiro H, Clark EA. Regulation of B-cell fate

by antigen-receptor signals. Nat Rev

Immunol 2002;2:945–956.

158. Karanjawala ZE, et al. The embryonic

lethality in DNA ligase IV-deficient mice is

rescued by deletion of Ku: implications for

unifying the heterogeneous phenotypes of

NHEJ mutants. DNA Repair (Amst)

2002;1:1017–1026.

159. Manis JP, Tian M, Alt FW. Mechanism and

control of class-switch recombination.

Trends Immunol 2002;23:31–39.

160. Lansford R, Manis JP, Sonoda E, Rajewsky K,

Alt FW. Ig heavy chain class switching in

Rag-deficient mice. Int Immunol

1998;10:325–332.

161. Muramatsu M, Kinoshita K, Fagarasan S,

Yamada S, Shinkai Y, Honjo T. Class switch

recombination and hypermutation require

activation-induced cytidine deaminase

(AID), a potential RNA editing enzyme. Cell

2000;102:553–563.

162. Bransteitter R, Pham P, Scharff MD,

Goodman MF. Activation-induced cytidine

deaminase deaminates deoxycytidine on

single-stranded DNA but requires the action

of RNase. Proc Natl Acad Sci USA

2003;100:4102–4107.

163. Chaudhuri J, Tian M, Khuong C, Chua K,

Pinaud E, Alt FW. Transcription-targeted

DNA deamination by the AID antibody

diversification enzyme. Nature

2003;422:726–730.

164. Dickerson SK, Market E, Besmer E,

Papavasiliou FN. AID mediates

hypermutation by deaminating single

stranded DNA. J Exp Med 2003;197:

1291–1296.

165. Sohail A, Klapacz J, Samaranayake M, Ullah

A, Bhagwat AS. Human activation-induced

cytidine deaminase causes transcription-

dependent, strand-biased C to U

deaminations. Nucleic Acids Res

2003;31:2990–2994.

166. Yu K, Chedin F, Hsieh CL, Wilson TE,

Lieber MR. R-loops at immunoglobulin class

switch regions in the chromosomes of

stimulated B cells. Nat Immunol

2003;4:442–451.

167. Tian M, Alt FW. Transcription-induced

cleavage of immunoglobulin switch regions

by nucleotide excision repair nucleases in

vitro. J Biol Chem 2000;275:24163–24172.

168. Shinkura R, Tian M, Smith M, Chua K,

Fujiwara Y, Alt FW. The influence of

transcriptional orientation on endogenous

switch region function. Nat Immunol

2003;4:435–441.

169. Tashiro J, Kinoshita K, Honjo T.

Palindromic but not G-rich sequences are

targets of class switch recombination. Int

Immunol 2001;13:495–505.

170. Sen D, Gilbert W. Formation of parallel

four-stranded complexes by guanine-rich

motifs in DNA and its implications for

meiosis. Nature 1988;334:364–366.

171. Mussmann R, Courtet M, Schwager J, Du

Pasquier L. Microsites for immunoglobulin

switch recombination breakpoints from

xenopus to mammals. Eur J Immunol

1997;27:2610–2619.

172. Petersen-Mahrt SK, Harris RS, NeubergerMS.

AID mutates E. coli suggesting a DNA

deamination mechanism for antibody

diversification. Nature 2002;418:99–103.

173. Rada C, Williams GT, Nilsen H, Barnes DE,

Lindahl T, Neuberger MS. Immunoglobulin

isotype switching is inhibited and somatic

hypermutation perturbed in UNG-deficient

mice. Curr Biol 2002;12:1748–1755.

174. Schrader CE, Edelmann W, Kucherlapati R,

Stavnezer J. Reduced isotype switching in

splenic B cells from mice deficient in

mismatch repair enzymes. J Exp Med

1999;190:323–330.

175. Schrader CE, Vardo J, Stavnezer J. Role for

mismatch repair proteins Msh2, Mlh1, and

Pms2 in immunoglobulin class switching

shown by sequence analysis of

recombination junctions. J Exp Med

2002;195:367–373.

176. Schrader CE, Vardo J, Stavnezer J. Mlh1 can

function in antibody class switch

recombination independently of Msh2. J Exp

Med 2003;197:1377–1383.

177. Ehrenstein MR, Neuberger MS. Deficiency

in Msh2 affects the efficiency and local

sequence specificity of immunoglobulin

class-switch recombination: parallels with

somatic hypermutation. EMBO J

1999;18:3484–3490.

178. Ehrenstein MR, Rada C, Jones AM,

Milstein C, Neuberger MS. Switch junction

sequences in PMS2-deficient mice reveal a

microhomology-mediated mechanism of Ig

class switch recombination. Proc Natl Acad

Sci USA 2001;98:14553–14558.

179. Imai K, et al. Human uracil-DNA glycosylase

deficiency associated with profoundly

impaired immunoglobulin class-switch

recombination. Nat Immunol

2003;4:1023–1028.

180. Honjo T, Kataoka T. Organization of

immunoglobulin heavy chain genes and

allelic deletion model. Proc Natl Acad Sci

USA 1978;75:2140–2144.

181. Iwasato T, Shimizu A, Honjo T,

Yamagishi H. Circular DNA is excised by

immunoglobulin class switch

recombination. Cell 1990;62:143–149.

182. Matsuoka M, Yoshida K, Maeda T, Usuda S,

Sakano H. Switch circular DNA formed in

cytokine-treated mouse splenocytes:

evidence for intramolecular DNA deletion in

immunoglobulin class switching. Cell

1990;62:135–142.

183. von Schwedler U, Jack HM, Wabl M.

Circular DNA is a product of the

immunoglobulin class switch

rearrangement. Nature 1990;345:452–456.

184. Wuerffel RA, Du J, Thompson RJ, Kenter AL.

IgS gamma 3 DNA-specific double strand

breaks are induced in mitogen-activated B

cells and are implicated in switch

recombination. J Immunol

1997;159:4139–4144.

185. Catalan N, Selz F, Imai K, Revy P, Fischer A,

Durandy A. The block in immunoglobulin

class switch recombination caused by

activation-induced cytidine deaminase

deficiency occurs prior to the generation of

DNA double strand breaks in switch mu

region. J Immunol 2003;171:2504–2509.

186. Petersen S, et al. AID is required to initiate

Nbs1/gamma-H2AX focus formation and

mutations at sites of class switching. Nature

2001;414:660–665.

187. Papavasiliou FN, Schatz DG. The activation-

induced deaminase functions in a

postcleavage step of the somatic

hypermutation process. J Exp Med

2002;195:1193–1198.

188. Bross L, Muramatsu M, Kinoshita K,

Honjo T, Jacobs H. DNA double strand

breaks: prior to but not sufficient in

targeting hypermutation. J Exp Med

2002;195:1187–1192.

189. Faili A, et al. AID-dependent somatic

hypermutation occurs as a DNA single-

strand event in the BL2 cell line. Nat

Immunol 2002;3:815–821.

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 129

Page 16: The role of the non-homologous end-joining pathway in lymphocyte development

190. Woo CJ, Martin A, Scharff MD. Induction of

somatic hypermutation is associated with

modifications in immunoglobulin variable

region chromatin. Immunity 2003;19:

479–489.

191. Dunnick W, Hertz GZ, Scappino L,

Gritzmacher C. DNA sequences at

immunoglobulin switch region

recombination sites [published erratum

appears in Nucleic Acids Res

1993;21:2285]. Nucleic Acids Res

1993;21:365–372.

192. Casellas R, et al. Ku80 is required for

immunoglobulin isotype switching. EMBO J

1998;17:2404–2411.

193. Manis JP, Dudley D, Kaylor L, Alt FW. IgH

class switch recombination to IgG1 in DNA-

PKcs-deficient B cells. Immunity

2002;16:607–617.

194. Reina-San-Martin B, Difilippantonio S,

Hanitsch L, Masilamani RF, Nussenzweig A,

Nussenzweig MC. H2AX is required for

recombination between immunoglobulin

switch regions but not for intra-switch

region recombination or somatic

hypermutation. J Exp Med 2003;197:

1767–1778.

195. Cook AJ, Oganesian L, Harumal P, Basten A,

Brink R, Jolly CJ. Reduced switching in SCID

B cells is associated with altered somatic

mutation of recombined S regions.

J Immunol 2003;171:6556–6564.

196. Bosma GC, et al. DNA-dependent protein

kinase activity is not required for

immunoglobulin class switching. J Exp Med

2002;196:1483–1495.

197. Chen X, Kinoshita K, Honjo T. Variable

deletion and duplication at recombination

junction ends: implication for staggered

double-strand cleavage in class-switch

recombination. Proc Natl Acad Sci USA

2001;98:13860–13865.

198. Muller JR, Marcu KB. Stimulation of murine

B lymphocytes induces a DNA exonuclease

whose activity on switch-mu DNA is

specifically inhibited by other germ-line

switch region RNAs. J Immunol

1998;160:3337–3341.

199. Kenter AL, Tredup J. High expression of a

30�50 exonuclease activity is specific to B

lymphocytes. Mol Cell Biol 1991;11:

4398–4404.

200. Dudley DD, Manis JP, Zarrin AA, Kaylor L,

Tian M, Alt FW. Internal IgH class switch

region deletions are position-independent

and enhanced by AID expression. Proc Natl

Acad Sci USA 2002;99:9984–9989.

201. Bassing CH, Alt FW. H2AX may function as

an anchor to hold broken chromosomal

DNA ends in close proximity. Cell Cycle

2004;3:149–153.

202. Bassing CH, et al. Histone H2AX. A dosage-

dependent suppressor oncogenic

translocations and tumors. Cell

2003;114:359–370.

203. Lundgren M, et al. Cell cycle regulation of

immunoglobulin class switch

recombination and germ-line transcription:

potential role of Ets family members. Eur J

Immunol 1995;25:2042–2051.

204. McCall MN, Hodgkin PD. Switch

recombination and germ-line transcription

are division-regulated events in B

lymphocytes. Biochim Biophys Acta

1999;1447:43–50.

205. Hodgkin PD, Lee JH, Lyons AB. B cell

differentiation and isotype switching is

related to division cycle number. J Exp Med

1996;184:277–281.

206. Kolodner RD, Putnam CD, Myung K.

Maintenance of genome stability in

Saccharomyces cerevisiae. Science

2002;297:552–557.

207. Hoeijmakers JH. Genome maintenance

mechanisms for preventing cancer. Nature

2001;411:366–374.

208. Bailey SM, et al. DNA double-strand break

repair proteins are required to cap the ends

of mammalian chromosomes. Proc Natl

Acad Sci USA 1999;96:14899–14904.

209. d’Adda di Fagagna F, et al. Effects of DNA

nonhomologous end-joining factors on

telomere length and chromosomal stability

in mammalian cells. Curr Biol

2001;11:1192–1196.

210. Ferguson DO, et al. The nonhomologous

end-joining pathway of DNA repair is

required for genomic stability and the

suppression of translocations. Proc Natl

Acad Sci USA 2000;97:6630–6633.

211. Karanjawala ZE, Grawunder U, Hsieh CL,

Lieber MR. The nonhomologous DNA end

joining pathway is important for

chromosome stability in primary

fibroblasts. Curr Biol 1999;9:1501–1504.

212. Chen Z, Sandberg AA. Molecular cytogenetic

aspects of hematological malignancies:

clinical implications. Am J Med Genet

2002;115:130–141.

213. Mills KD, Ferguson DO, Alt FW. The role

of DNA breaks in genomic instability and

tumorigenesis. Immunol Rev 2003;194:

77–95.

214. Guidos CJ, Williams CJ, Grandal I,

Knowles G, Huang MT, Danska JS. V(D)J

recombination activates a p53-dependent

DNA damage checkpoint in scid lymphocyte

precursors. Genes Dev 1996;10:

2038–2054.

215. Nacht M, et al. Mutations in the p53 and

SCID genes cooperate in tumorigenesis.

Genes Dev 1996;10:2055–2066.

216. Rooney S, et al. Artemis and p53 cooperate

to suppress oncogenic N-myc amplification

in progenitor B cells. Proc Natl Acad Sci USA

2004;101:2410–2415.

217. Difilippantonio MJ, et al. Evidence for

replicative repair of DNA double-strand

breaks leading to oncogenic translocation

and gene amplification. J Exp Med

2002;196:469–480.

218. Gladdy RA, et al. The RAG-1/2

endonuclease causes genomic instability and

controls CNS complications of

lymphoblastic leukemia in p53/Prkdc-

deficient mice. Cancer Cell 2003;3:37–50.

219. Vanasse GJ, et al. Genetic pathway to

recurrent chromosome translocations in

murine lymphoma involves V(D)J

recombinase. J Clin Invest 1999;103:

1669–1675.

220. Zhu C, et al. Unrepaired DNA breaks in P53-

deficient cells lead to oncogenic gene

amplification subsequent to translocations.

Cell 2002;109:811–821.

221. Kuschel B, et al. Variants in DNA double-

strand break repair genes and breast cancer

susceptibility. Hum Mol Genet

2002;11:1399–1407.

222. Ruttan CC, Glickman BW. Coding variants

in human double-strand break DNA repair

genes. Mutat Res 2002;509:175–200.

223. Han J, Hankinson SE, Ranu H, De Vivo I,

Hunter DJ. Polymorphisms in DNA double-

strand break repair genes and breast cancer

risk in the Nurses’ Health Study.

Carcinogenesis 2004;25:189–195.

224. Roddam PL, Rollinson S, O’Driscoll M,

Jeggo PA, Jack A, Morgan GJ. Genetic

variants of NHEJ DNA ligase IV can affect

the risk of developing multiple myeloma,

a tumour characterised by aberrant class

switch recombination. J Med Genet

2002;39:900–905.

225. Li G, Nelsen C, Hendrickson EA. Ku86 is

essential in human somatic cells. Proc Natl

Acad Sci USA 2002;99:832–837.

226. Moshous D, et al. Partial T and B

lymphocyte immunodeficiency and

predisposition to lymphoma in patients

with hypomorphic mutations in Artemis.

J Clin Invest 2003;111:381–387.

227. Charron J, et al. Embryonic lethality in mice

homozygous for a targeted disruption of

the N-myc gene. Genes Dev 1992;6:

2248–2257.

228. Davis AC, Wims M, Spotts GD, Hann SR,

Bradley A. A null c-myc mutation causes

lethality before 10.5 days of gestation in

homozygotes and reduced fertility in

heterozygous female mice. Genes Dev

1993;7:671–682.

Rooney et al � The NHEJ pathway in lymphocyte development

130 Immunological Reviews 200/2004

Page 17: The role of the non-homologous end-joining pathway in lymphocyte development

229. Malynn BA, et al. N-myc can functionally

replace c-myc in murine development,

cellular growth, and differentiation. Genes

Dev 2000;14:1390–1399.

230. Udayakumar D, Bladen CL, Hudson FZ,

Dynan WS. Distinct pathways of

nonhomologous end joining that are

differentially regulated by DNA-dependent

protein kinase-mediated phosphorylation.

J Biol Chem 2003;278:41631–41635.

231. Dai Y, et al. Nonhomologous end joining

and V(D)J recombination require an

additional factor. Proc Natl Acad Sci USA

2003;100:2462–2467.

232. Chappell C, Hanakahi LA, Karimi-Busheri F,

Weinfeld M, West SC. Involvement of

human polynucleotide kinase in double-

strand break repair by non-homologous end

joining. EMBO J 2002;21:2827–2832.

233. Mahajan KN, McElhinny SAN, Mitchell BS,

Ramsden DA. Association of DNA

polymerase mu (pol mu) with Ku and ligase

IV: role for pol mu in end-joining double-

strand break repair. Mol Cell Biol

2002;22:5194–5202.

234. Lee JW, et al. Implication of DNA

polymerase lambda in alignment-based gap

filling for nonhomologous DNA end joining

in human nuclear extracts. J Biol Chem

2004;279:805–811.

235. Bebenek K, Garcia-Diaz M, Blanco L,

Kunkel TA. The frameshift infidelity of

human DNA polymerase lambda.

Implications for function. J Biol Chem

2003;278:34685–34690.

236. Zhang Y, Wu X, Yuan F, Xie Z, Wang Z.

Highly frequent frameshift DNA synthesis

by human DNA polymerase mu. Mol Cell

Biol 2001;21:7995–8006.

237. Hiom K, Gellert M. Assembly of a 12/23

paired signal complex: a critical control

point in V(D)J recombination. Mol Cell

1998;1:1011–1019.

238. Agrawal A, Schatz DG. RAG1 and RAG2

form a stable postcleavage synaptic complex

with DNA containing signal ends in V(D)J

recombination. Cell 1997;89:43–53.

239. Ward IM, Minn K, van Deursen J, Chen J.

p53 binding protein 53BP1 is required for

DNA damage responses and tumor

suppression in mice. Mol Cell Biol

2003;23:2556–2563.

240. Morales JC, et al. Role for the BRCA1

C-terminal repeats (BRCT) protein 53BP1 in

maintaining genomic stability. J Biol Chem

2003;278:14971–14977.

241. Celeste A, et al. Genomic instability in mice

lacking histone H2AX. Science

2002;296:922–927.

242. Bassing CH, et al. Increased ionizing

radiation sensitivity and genomic instability

in the absence of histone H2AX. Proc Natl

Acad Sci USA 2002;99:8173–8178.

243. Gurley KE, Kemp CJ. Synthetic lethality

between mutation in Atm and DNA-PK(cs)

during murine embryogenesis. Curr Biol

2001;11:191–194.

Rooney et al � The NHEJ pathway in lymphocyte development

Immunological Reviews 200/2004 131