the lmo2 oncogene regulates dna replication in ... · christophe cazauxc,2, nazar mashtalir d, el...

6
The LMO2 oncogene regulates DNA replication in hematopoietic cells Marie-Claude Sincennes a,b,1 , Magali Humbert a,1 , Benoît Grondin a , Véronique Lisi a,b , Diogo F. T. Veiga a , André Haman a , Christophe Cazaux c,2 , Nazar Mashtalir d , EL Bachir Affar d , Alain Verreault a,b,e , and Trang Hoang a,b,e,3 a Institute of Research in Immunology and Cancer, University of Montreal, Montreal, QC, Canada H3C 3J7; b Molecular Biology Program, University of Montreal, Montreal, QC, Canada H1T 2M4; c Cancer Research Center of Toulouse, Toulouse 31024, France; d Maisonneuve-Rosemont Hospital Research Center, Department of Medicine, University of Montreal, Montreal, QC, Canada H1T 2M4; and e Departments of Pharmacology and Biochemistry, University of Montreal, Montreal, QC, Canada H3T 1J2 Edited by Mark Groudine, Fred Hutchinson Cancer Research Center, Seattle, WA, and approved December 2, 2015 (received for review July 30, 2015) Oncogenic transcription factors are commonly activated in acute leukemias and subvert normal gene expression networks to reprogram hematopoietic progenitors into preleukemic stem cells, as exemplified by LIM-only 2 (LMO2) in T-cell acute lymphoblastic leukemia (T-ALL). Whether or not these oncoproteins interfere with other DNA-dependent processes is largely unexplored. Here, we show that LMO2 is recruited to DNA replication origins by interaction with three essential replication enzymes: DNA polymerase delta (POLD1), DNA primase (PRIM1), and minichromosome 6 (MCM6). Furthermore, tethering LMO2 to synthetic DNA sequences is sufficient to transform these sequences into origins of replication. We next addressed the importance of LMO2 in erythroid and thymocyte development, two lineages in which cell cycle and differen- tiation are tightly coordinated. Lowering LMO2 levels in erythroid progenitors delays G1-S progression and arrests erythropoietin-depen- dent cell growth while favoring terminal differentiation. Conversely, ectopic expression in thymocytes induces DNA replication and drives these cells into cell cycle, causing differentiation blockade. Our results define a novel role for LMO2 in directly promoting DNA synthesis and G1-S progression. LMO2 | cell cycle | DNA replication | hematopoietic cells | T-cell acute lymphoblastic leukemia M ore than 70% of recurring chromosomal translocations in T-cell acute lymphoblastic leukemia (T-ALL) involve tran- scription factors that are master regulators of cell fate. These on- cogenic transcription factors determine the gene signature and leukemic cell types (1). Whether these DNA-bound factors may have additional roles beyond modulating gene expression remains unknown. LMO2, a 17-kDa protein defined by tandem zinc finger domains, is an essential nucleation factor that assembles a multi- partite transcriptional regulatory complex on gene regulatory re- gions via direct interaction with the TAL1/SCL transcription factor, LDB1, and other DNA binding transcription factors (24, reviewed in refs. 5, 6). These complexes drive gene expression programs that determine hematopoietic cell fates at critical branchpoints both during embryonic development (7) and in adult hematopoietic stem cells (8, 9). Lmo2 function is essential in highly proliferative ery- throid progenitors (10, reviewed in refs. 5, 6). Interestingly, Lmo2 down-regulation is required for terminal erythroid differentiation (11, 12). Because commitment to terminal differentiation is coordinated with growth arrest (13), Lmo2 may have additional molecular functions that impede this critical step marked by growth cessation. In mouse models of T-ALL, LMO1 or LMO2 collaborates with SCL to inhibit the activity of two basic helixloophelix (bHLH) transcription factors that control thymocyte differentiation, E2A/ TCF3 and HEB/TCF12, causing differentiation arrest (reviewed in ref. 14). However, this inhibition is not sufficient, per se, for leu- kemogenesis, because both TAL1 and LYL1 inhibit E proteins but require interaction with LMO1/2 to activate the transcription of a self-renewal gene network in thymocytes (15, 16) and to in- duce T-ALL (17, 18). Of note, downstream target genes cannot substitute for LMO1/2 to induce T-ALL, suggesting additional functions for LMO1/2. Together, these studies underscore the dominant oncogenic properties of LMO2, as revealed by recurring retroviral integrations upstream of LMO2 in the gene therapy trial (19, 20) or by recurrent chromosomal rearrangements in T-ALL (21). As a consequence, LMO2 is misexpressed in the T lineage, where it is normally absent. In addition, LMO proteins are frequently deregulated in breast cancers (22) and neuroblastomas (23), pointing to their importance in cell transformation. In particular, in patients who eventually developed T-ALL associated with LMO2 activation after gene therapy, T-cell hyperproliferation was observed early during the preleukemic stage (19). How LMO2 affects erythroid progenitor or T-cell proliferation cannot be inferred from its downstream target genes (12, 2428). To understand LMO2 functions, we performed an unbiased screen for LMO2 interaction partners. We show that LMO2 asso- ciates with three replication proteins, minichromosome 6 (MCM6), DNA primase (PRIM1), and DNA polymerase delta (POLD1), and that LMO2 influences cell cycle progression and DNA repli- cation in hematopoietic cells, indicating an unexpected function for LMO2. Significance Understanding how cell cycle and cell differentiation are coordinated during normal hematopoiesis will reveal molecular insights in leu- kemogenesis. LIM-only 2 (LMO2) is a transcriptional regulator that controls the erythroid lineage via activation of an erythroid-specific gene expression program. Here, we uncover an unexpected function for LMO2 in controlling DNA replication via proteinprotein inter- actions with essential DNA replication enzymes. To our knowledge, this work provides the first evidence for a nontranscriptional func- tion of LMO2 that drives the cell cycle at the expense of differenti- ation in the erythroid lineage and in thymocytes when misexpressed following genetic alterations. We propose that the nontranscrip- tional control of DNA replication uncovered here for LMO2 may be a more common function of oncogenic transcription factors than previously appreciated. Author contributions: M.-C.S., M.H., B.G., and T.H. designed research; M.-C.S., M.H., B.G., V.L., D.F.T.V., A.H., and N.M. performed research; A.V. contributed new reagents/analytic tools; M.-C.S., M.H., B.G., V.L., D.F.T.V., A.H., C.C., N.M., E.B.A., and T.H. analyzed data; and M.-C.S., M.H., B.G., E.B.A., A.V., and T.H. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Freely available online through the PNAS open access option. 1 M.-C.S. and M.H. contributed equally to this work. 2 Deceased August 3, 2015. 3 To whom correspondence should be addressed. Email: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1515071113/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1515071113 PNAS | February 2, 2016 | vol. 113 | no. 5 | 13931398 MEDICAL SCIENCES Downloaded by guest on October 13, 2020

Upload: others

Post on 01-Aug-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The LMO2 oncogene regulates DNA replication in ... · Christophe Cazauxc,2, Nazar Mashtalir d, EL Bachir Affar , Alain Verreaulta,b,e, and Trang Hoanga,b,e,3 a Institute of Research

The LMO2 oncogene regulates DNA replication inhematopoietic cellsMarie-Claude Sincennesa,b,1, Magali Humberta,1, Benoît Grondina, Véronique Lisia,b, Diogo F. T. Veigaa, André Hamana,Christophe Cazauxc,2, Nazar Mashtalird, EL Bachir Affard, Alain Verreaulta,b,e, and Trang Hoanga,b,e,3

aInstitute of Research in Immunology and Cancer, University of Montreal, Montreal, QC, Canada H3C 3J7; bMolecular Biology Program, University ofMontreal, Montreal, QC, Canada H1T 2M4; cCancer Research Center of Toulouse, Toulouse 31024, France; dMaisonneuve-Rosemont Hospital ResearchCenter, Department of Medicine, University of Montreal, Montreal, QC, Canada H1T 2M4; and eDepartments of Pharmacology and Biochemistry, Universityof Montreal, Montreal, QC, Canada H3T 1J2

Edited by Mark Groudine, Fred Hutchinson Cancer Research Center, Seattle, WA, and approved December 2, 2015 (received for review July 30, 2015)

Oncogenic transcription factors are commonly activated in acuteleukemias and subvert normal gene expression networks to reprogramhematopoietic progenitors into preleukemic stem cells, as exemplifiedby LIM-only 2 (LMO2) in T-cell acute lymphoblastic leukemia (T-ALL).Whether or not these oncoproteins interferewith other DNA-dependentprocesses is largely unexplored. Here, we show that LMO2 is recruitedto DNA replication origins by interaction with three essential replicationenzymes: DNA polymerase delta (POLD1), DNA primase (PRIM1), andminichromosome 6 (MCM6). Furthermore, tethering LMO2 to syntheticDNA sequences is sufficient to transform these sequences into origins ofreplication.We next addressed the importance of LMO2 in erythroid andthymocyte development, two lineages in which cell cycle and differen-tiation are tightly coordinated. Lowering LMO2 levels in erythroidprogenitors delays G1-S progression and arrests erythropoietin-depen-dent cell growth while favoring terminal differentiation. Conversely,ectopic expression in thymocytes induces DNA replication and drivesthese cells into cell cycle, causing differentiation blockade. Our resultsdefine a novel role for LMO2 in directly promoting DNA synthesis andG1-S progression.

LMO2 | cell cycle | DNA replication | hematopoietic cells | T-cell acutelymphoblastic leukemia

More than 70% of recurring chromosomal translocations inT-cell acute lymphoblastic leukemia (T-ALL) involve tran-

scription factors that are master regulators of cell fate. These on-cogenic transcription factors determine the gene signature andleukemic cell types (1). Whether these DNA-bound factors mayhave additional roles beyond modulating gene expression remainsunknown. LMO2, a 17-kDa protein defined by tandem zinc fingerdomains, is an essential nucleation factor that assembles a multi-partite transcriptional regulatory complex on gene regulatory re-gions via direct interaction with the TAL1/SCL transcription factor,LDB1, and other DNA binding transcription factors (2–4, reviewedin refs. 5, 6). These complexes drive gene expression programs thatdetermine hematopoietic cell fates at critical branchpoints bothduring embryonic development (7) and in adult hematopoietic stemcells (8, 9). Lmo2 function is essential in highly proliferative ery-throid progenitors (10, reviewed in refs. 5, 6). Interestingly, Lmo2down-regulation is required for terminal erythroid differentiation(11, 12). Because commitment to terminal differentiation iscoordinated with growth arrest (13), Lmo2 may have additionalmolecular functions that impede this critical step marked bygrowth cessation.In mouse models of T-ALL, LMO1 or LMO2 collaborates with

SCL to inhibit the activity of two basic helix–loop–helix (bHLH)transcription factors that control thymocyte differentiation, E2A/TCF3 and HEB/TCF12, causing differentiation arrest (reviewed inref. 14). However, this inhibition is not sufficient, per se, for leu-kemogenesis, because both TAL1 and LYL1 inhibit E proteins butrequire interaction with LMO1/2 to activate the transcription ofa self-renewal gene network in thymocytes (15, 16) and to in-duce T-ALL (17, 18). Of note, downstream target genes cannot

substitute for LMO1/2 to induce T-ALL, suggesting additionalfunctions for LMO1/2.Together, these studies underscore the dominant oncogenic

properties of LMO2, as revealed by recurring retroviral integrationsupstream of LMO2 in the gene therapy trial (19, 20) or by recurrentchromosomal rearrangements in T-ALL (21). As a consequence,LMO2 is misexpressed in the T lineage, where it is normally absent.In addition, LMO proteins are frequently deregulated in breastcancers (22) and neuroblastomas (23), pointing to their importancein cell transformation. In particular, in patients who eventuallydeveloped T-ALL associated with LMO2 activation after genetherapy, T-cell hyperproliferation was observed early during thepreleukemic stage (19). How LMO2 affects erythroid progenitor orT-cell proliferation cannot be inferred from its downstream targetgenes (12, 24–28).To understand LMO2 functions, we performed an unbiased

screen for LMO2 interaction partners. We show that LMO2 asso-ciates with three replication proteins, minichromosome 6 (MCM6),DNA primase (PRIM1), and DNA polymerase delta (POLD1),and that LMO2 influences cell cycle progression and DNA repli-cation in hematopoietic cells, indicating an unexpected functionfor LMO2.

Significance

Understanding how cell cycle and cell differentiation are coordinatedduring normal hematopoiesis will reveal molecular insights in leu-kemogenesis. LIM-only 2 (LMO2) is a transcriptional regulator thatcontrols the erythroid lineage via activation of an erythroid-specificgene expression program. Here, we uncover an unexpected functionfor LMO2 in controlling DNA replication via protein–protein inter-actions with essential DNA replication enzymes. To our knowledge,this work provides the first evidence for a nontranscriptional func-tion of LMO2 that drives the cell cycle at the expense of differenti-ation in the erythroid lineage and in thymocyteswhenmisexpressedfollowing genetic alterations. We propose that the nontranscrip-tional control of DNA replication uncovered here for LMO2may be amore common function of oncogenic transcription factors thanpreviously appreciated.

Author contributions: M.-C.S., M.H., B.G., and T.H. designed research; M.-C.S., M.H., B.G.,V.L., D.F.T.V., A.H., and N.M. performed research; A.V. contributed new reagents/analytictools; M.-C.S., M.H., B.G., V.L., D.F.T.V., A.H., C.C., N.M., E.B.A., and T.H. analyzed data;and M.-C.S., M.H., B.G., E.B.A., A.V., and T.H. wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

Freely available online through the PNAS open access option.1M.-C.S. and M.H. contributed equally to this work.2Deceased August 3, 2015.3To whom correspondence should be addressed. Email: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1515071113/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1515071113 PNAS | February 2, 2016 | vol. 113 | no. 5 | 1393–1398

MED

ICALSC

IENCE

S

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 2: The LMO2 oncogene regulates DNA replication in ... · Christophe Cazauxc,2, Nazar Mashtalir d, EL Bachir Affar , Alain Verreaulta,b,e, and Trang Hoanga,b,e,3 a Institute of Research

ResultsIdentification of New LMO2 Protein–Protein Interactions in HematopoieticProgenitors. Lmo2 is expressed in c-Kit+ hematopoietic stem andprogenitor cells (HSPCs) and in immature prothymocytes, but not atlater stages of T-cell differentiation (29). To identify new LMO2binding proteins in HSPCs, we constructed a cDNA library frompurified murine Kit+Lin− hematopoietic progenitors for a yeast two-hybrid screen and used LMO2 as bait. In addition to known LMO2-interacting proteins, such as LDB1, and to proteins associated withtranscription, we unexpectedly identified interactions with three es-sential components of prereplication complexes (pre-RCs),namely, MCM6, POLD1, and PRIM1 (30) (Fig. 1A and Table S1).In comparison, a screen performed using GAL4-SCL identified onlyknown interactions (Table S1). LMO2 interaction was specific tothese three replication proteins, as confirmed by independent yeasttwo-hybrid assays with full-length cDNAs (Fig. 1 B and C). In ad-dition, we identified BAZ1A/ACF1, required for DNA replicationthrough heterochromatin (31); SetD8, for replication licensing (32);MYST2/HBO1, controlling MCM loading via ORC1 binding (33);and CCNA2-CDK1, regulating origin firing (34) (Fig. 1A). The top-ranking pathways by gene set enrichment analysis were cell cycle,DNA synthesis, and DNA replication (Fig. 1D), concurring with theview that LMO2 controls these processes via its protein partners.Finally, LMO2 coimmunoprecipitated with PRIM1, MCM6, andMYST2 in mammalian cells, and both LIM domains contributed tothis interaction (Fig. 1E), whereas LDB1 binding required mostlyLIM1, as expected (4). We conclude that in addition to its associa-tion with transcription regulators, LMO2 engages in protein–proteininteractions with DNA replication proteins.

LMO2 Associates with Replication Complexes. We next addressedthe question of whether LMO2 associates with the pre-RC inhematopoietic cells. Because pre-RC formation often requires aDNA template (35), we prepared DNA-containing chromatinextracts from CD34+Kit+ progenitors (TF-1 cells) (3, 13, 36, 37)to assess the interaction of LMO2 with endogenous replicationproteins by immunoprecipitation. We found that POLD1, PRIM1,and MCM6, but not lamin B or beta-actin (negative controls),reproducibly and specifically coimmunoprecipitate with LMO2,but not with control Ig, in TF-1 chromatin extracts (Fig. 1F). We

also detected the other members of the MCM hexameric com-plex (MCM2–MCM7), as well as ORC2, CDC6, CDT1, andPCNA, albeit with weaker efficiency for the latter three (Fig.1F). In contrast, MCM6, POLD1, and PRIM1 did not coim-munoprecipitate with SCL (Fig. 1G), whereas LMO2 and LDB1were found with SCL in this assay, as expected. Therefore, replica-tion proteins do not stably associate with the SCL transcriptioncomplex but more specifically with LMO2.

LMO2 Binds to Origins of Replication. LMO2 shares importanttranscriptional targets with SCL (15, 16, 25, 36). Nonetheless,LMO2 is rarely found at gene promoters (3%) by ChIP sequencingin multipotent progenitors, whereas SCL is more frequently ob-served within transcriptional start sites (28%) (24), suggesting thatLMO2 has SCL-independent functions. We assessed whetherLMO2 associates with DNA replication origins, using ChIP of TF-1cells with anti-LMO2, or anti-MCM5 antibodies. We first studiedseven well-characterized human DNA replication origins (38) andfound that both MCM5 and LMO2 occupied four of these repli-cation origins, c-MYC, G6PD, TOP1, and MCM4 (Fig. 2A), allmapping to early replicating G1 (ERG1) segments (39, 40) (Fig.S1A). In contrast, MCM5 did not bind the GYPA promoter, a well-defined SCL-LMO2 transcriptional target (36), whereas LMO2occupancy was confirmed, together with SCL and GATA1, twoLMO2 transcription factor partners. SCL was detected at two of theseven tested origins, although binding was 10- to 20-fold lowercompared with GYPA, whereas GATA1 binding was below thedetection limit (Fig. 2A). Therefore, LMO2 is recruited to well-characterized DNA replication origins with MCM5, in the absenceof GATA1 and frequently of SCL.These results led us to assess LMO2 occupancy of replication

initiation zones identified in two independent tiling array-basedstudies in HeLa cells, using Lambda-exonuclease digestion (38, 41)and anti-BrdU immunoprecipitation to purify origin-centered na-scent DNA strands. We focused on a subset of 45 replication ini-tiation zones that overlapped within a distance of 1 kb between thetwo studies (Fig. 2B and Fig. S1 B and C). In TF-1 cells, MCM5 andLMO2 were each detected on 20 and 19 replication initiation zones,respectively, nearly half of which were positive for both (Fig. 2B).We aligned these 45 replication initiation zones with ERG1 seg-ments identified in mammalian cells (39, 42) and found that 68% of

A

G

B

POLD1inp

ut

LMO2LaminB

MCM6MCM7

MCM2MCM3MCM5

PRIM1

ORC2CDC6

SCLLDB1

PCNA

LMO2

IgG

IP

CDT1

β-actin

F

C

EGAL4-LMO2GAL4-ΔLIM1GAL4-ΔLIM2

pcDNA3

PRIM1

MCM6

MYST2

LDB1

++

+

++

+

++

+

++

+

++

+

IP F

LAG

Inpu

t

IB: GAL4

FLAG

GAL4

02

46

CELL

CYC

LE M

ITO

TIC

SYNT

HESI

S O

F DN

ADN

A RE

PLIC

ATIO

NS

PHAS

ECE

LL C

YCLE

DNA

STRA

ND E

LONG

.G

1 S

TRAN

SITI

ON

-log(

10) F

DR

01234

PO

LD1

LDB

1M

CM

2M

CM

5M

CM

10C

DT1

RPA

32P

CN

A

MCM

6P

RIM

1

LMO

2in

tera

ctio

ns

SCLp42 *

SCLp22 *

MCM6LMO2

POLD1

inputIgG SCL

SCLIgG

IP SN

PRIM1

Transcription

Chromatinmodification

replication Cell cycle

PSME3

CCNA2

PRIM1

POLD1

MCM6

MYST2

ACF1

UHRF1

DNA

(total=52)

CDK9

PSMC1

BUB1B

HIPK2

G1-S

SETD8

MLL2PS

CCNA2222CCCCCNA2

PRIM1MMPPRIM1

POPOLDPOLD

MCM6MMMCMCM6

22PSMC11PSPSMC1

BUB1BUBUB1B

MYST2MYYS 2MYST2

ACF1FACF1SETD8ETD88ETD88

MLMMLML

A8

22T2MLLMMLL

LMO2

LDB1TAF6LFLII

GTF3C2LDBLDDBDBL

TAF6LLTTATAF6L6 LIILFLII6L6L6L6L

TF3C2F3C2GTGTFF3C2GTDBB

MCM

6PR

IM1

POLD1

LDB1

LMO2 39 16 12 26

Lamin 0 0 0 0

Empty 0 0 0 0

Tp53 0 0 0 0

D

Fig. 1. LMO2 interacts with DNA replication proteins.(A) New LMO2 protein–protein interactions in hema-topoietic progenitors revealed by yeast two-hybridscreens using LMO2 as bait (*). (B and C) LMO2 spe-cifically interacts with MCM6, PRIM1, and POLD1 byyeast two-hybrid assay. (D) Gene set enrichment anal-ysis for LMO2 interaction partners. Shown are the topseven Reactome pathways significantly enriched in theLMO2 Y2H screen [false discovery rate (FDR): q < 0.001]according to the Molecular Signatures database v5.0(PubMed identifier 16199517). (E) Both LIM domainsof LMO2 are required for interaction with PRIM1,MCM6, andMYST2. GAL4-LMO2WT or LIM1/2 mutantswere cotransfected with Flag-tagged PRIM1, MCM6, orMYST2. Protein extracts were immunoprecipitated withanti-FLAG Ab, followed by Western blotting. (F) DNAreplication proteins coimmunoprecipitate with LMO2(**). Immunoprecipitation of TF-1 chromatin extractswith Abs against LMO2. Inputs (4%) and immune pel-lets (IP) were analyzed by IB. (G) SCL does not stablyassociate with DNA replication proteins. SCL or isotypecontrol Abs were analyzed as in F (*). IB, immunoblot-ting; IP, immune pellet; SN, supernatant. Data shownare typical of at least two (*) or three (**) independentexperiments.

1394 | www.pnas.org/cgi/doi/10.1073/pnas.1515071113 Sincennes et al.

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 3: The LMO2 oncogene regulates DNA replication in ... · Christophe Cazauxc,2, Nazar Mashtalir d, EL Bachir Affar , Alain Verreaulta,b,e, and Trang Hoanga,b,e,3 a Institute of Research

LMO2-bound zones mapped to ERG1 segments, which is almosttwice the percentage of ERG1 segments found in the absence ofLMO2 (Fig. 2B). Therefore, LMO2 is recruited to a significantproportion of initiation zones in TF-1 cells, corresponding, in mostpart, to ERG1 segments.

Tethering LMO2 to DNA Recruits Replication Proteins and Induces DNAReplication. To determine whether LMO2 tethering to DNA wassufficient to stimulate DNA replication, we optimized a syntheticDNA replication assay in mammalian cells described by Takedaet al. (43). We assessed whether LMO2 fused to the DNA bindingdomain of GAL4 can drive DNA replication from GAL4 bindingsites (5XUAS) in transfected 293 cells (Fig. 2C). Newly replicatedDNA is hemimethylated or unmethylated and can be distinguishedfrom transfected, bacterially derived DNA by its resistance to Dpn1digestion (43). As expected, GAL4-ORC2 induced an approx-imately fivefold increase in newly synthesized DNA comparedwith control GAL4-VP16 (Fig. 2D). In this assay, GAL4-LMO2reproductively directed a dose-dependent increase of newly repli-cated DNA, reaching a maximum of eightfold (Fig. 2D), whereasGAL4-SCL did not produce the same results. This activity wasabrogated when GAL4-binding UAS sequences were mutated (Fig.2E) or absent (pBluescript vector). We therefore conclude thatLMO2 anchoring to DNA was sufficient to direct DNA replication.LMO2-driven DNA replication occurred most likely in the absenceof transcription because 293T cells lack hematopoietic transcrip-tion factors that recruit LMO2 to DNA, namely SCL and GATA1/2. To determine if DNA-bound GAL4-LMO2 could recruit DNAreplication proteins to the UAS template, we performed DNAcapture as previously described (44), using UAS sequencesimmobilized on beads. LMO2 recruited POLD1, MCM5, and, to alesser extent, CDT1 and PCNA to DNA, and this recruitmentrequired the integrity of the GAL4 binding site (Fig. 2F). Together,our results indicate that LMO2 tethering to DNA was sufficient tonucleate the assembly of prereplication/preinitiation complexesspecifically at the site of binding and to direct DNA replication.

LMO2 Expression Levels Control the Rate of DNA Synthesis andCellular Outcome in the Erythroid Lineage. Cell cycle is highly reg-ulated in the erythroid lineage because ∼80% of primary pro-erythroblasts (E1 stage; Fig. 3A) are in S phase, exactly at theonset of erythropoietin (Epo) dependence (45), and this pro-portion sharply drops at the E3 and E4 stages of terminal dif-ferentiation. We observed that the E1 population segregatedinto cells with high and low endogenous LMO2 protein levels(Fig. 3B), corresponding to high and low proportions of cells inS/G2/M (Fig. 3C). Strikingly, LMO2 protein levels decrease fromthe E1hi to E4 stage, directly correlating with a decrease in theproportion of cells in S phase (Fig. 3C), whereas GATA1 andSCL protein levels increase (Fig. 3D). This process is highlycoordinated because most LMO2 partners identified here, in-cluding CCNA2, are synchronously down-regulated with Lmo2during differentiation from proerythroblast to orthochromaticerythroblasts (Fig. S2).We addressed the functional importance of LMO2 by decreasing

LMO2 protein levels in erythroid progenitors via RNAi (shRNALmo2) (Fig. S3A). Lmo2 depletion in Ter119− fetal liver erythroidprogenitors decreased by twofold the proportion of cells in S phase asdetermined by DAPI staining and flow cytometry analysis, comparedwith control cells (Fig. 3E). In addition, Lmo2 depletion almost ab-rogated the proliferation of primary erythroid progenitors at a keycommitment step marked by Epo responsiveness (Fig. 3 F and G),while enhancing the generation of mature E4 cells (CD71−Ter119+)(Fig. 3H). Therefore, high LMO2 levels are required for the Epo-dependent proliferation of CD71+ erythroid precursors, whereaslowering LMO2 accelerates terminal erythroid differentiation.Decreased LMO2 levels caused a twofold decrease in the rate of

DNA synthesis monitored by the kinetics of 32P orthophosphateincorporation into synchronized mouse erythroleukemia (MEL) cells(Fig. 3I), without causing apoptosis (Fig. S3B). Consequently, thesecells failed to proliferate in culture (Fig. S3C). Decreased DNAsynthesis was unlikely due to decreased expression levels for repli-cation genes assessed by RT-quantitative PCR (Fig. S3D), consistentwith transcriptome analysis of erythroid cells and lymphoid cells

A B

E

C

NdeI DpnI DpnIUAS WT

NdeI DpnI DpnIX X X X XUAS mut

DNA replication assay (D, E)

D

Gal4-VP16Gal4-ORC2Gal4-LMO2

- -- - -- -- - -++

- -

Rep

licat

ed D

NA

(Fol

d ov

er c

ontro

l)

0

2

4

6

8

10

----

- --Gal4-SCL -- - - -

*Ear

ly re

plic

atin

g G

1 se

gmen

t

ChIP

LMO

2M

CM

5

Stu

dy 2

24 mix264 mix272 mix193 mix

5093

130155

54165194 mix

29220 K221 K

23217 K

8788

212162117126163167190 mix259 mix188 mix

53 K225 K105151218 K219 K228 K230 K237

20202206207208209210233241

Stu

dy 1

Ann

ot_O

ri (3

-4)

ERG1*

F

DNA capture assay (F) GAL4-LMO2nuc. extract

+ UAS WTOOor UAS mutO

Bound proteinsWestern blotting

MutUAS

150

100

75

50

POLD1MCM5CDT1GAL4PCNA 37

Wt

1 10 10050202 5

Immunoprecipitated chromatin (fold enrichment)

G6PD:

TOP1:

MCM4:

MYC:

HSP70:

LMNB2:

e1 e2

DNMT1:

GYPA promoter:

LMO2MCM5SCLGATA-1

n.d.

n.d.

n.d.

n.d.

n.d.

n.d.

0.5 kb

e2 G

e12e1

(TIMM13) (LMNB2)

G

e1 GGG

exon1G

e2 e1G G

e1

(PRDKC)

e1

(MCM4)

0

100

200

300

400

Ext

ra c

hrom

osom

al

plas

mid

(% c

ontro

l)

pBSWt mutUAS

VP16 LMO2

mix: early or lateK: early in K562

Gal4-

Fig. 2. LMO2 and MCM5 occupy DNA replication origins. (A) ChIP of well-characterized DNA replication origins in TF-1 cells with anti-LMO2, anti-MCM5,anti-SCL, and anti-GATA1 Abs. Shown are the ratio of enrichment over controlIg and background (c-Kit, −20 kb). GYPA promoter sequences were amplified asa control (36). (Left) Maps depicting the location of well-characterized DNAreplication origins and PCR probes. Gene exons are illustrated by black boxes.PCR primers are shown as red bars. E, E47 consensus sites; G, GATA1/2 consensussites. n.d., not detectable. (B) LMO2-bound replication initiation zones (gray,first column) are preferentially enriched within ERG1 segments. Forty-five rep-lication initiation zones found in common in two studies (38, 41) were analyzedby ChIP for LMO2 binding. The replication timing of these origins according totwo datasets is illustrated: study 1 [Gilbert and coworkers (42)] in leukemicpatients and lymphoblastoid cell lines and study 2 [Stamatoyannopoulos andcoworkers (39)] in B cells and ES cells. A higher proportion of early replicatingsegments (gray, last column) is found in LMO2-bound origins (P = 0.02). Mix:Some samples from study 1 showed early replication, whereas others showedlate replication. (C) Diagram illustrating the synthetic DNA replication assay.Plasmid DNA was transfected into HEK293 cells and detected by PCR (primerpairs as shown). (Left) Dpn1 sensitivity distinguishes unreplicated methylatedplasmid DNA (sensitive) from replicated hemi- or unmethylated DNA (resistant).(D) LMO2 fused to GAL4 stimulates the replication of a GAL4-responsive plas-mid in mammalian cells. HEK293 cells were transfected with the indicated GAL-fusion genes. After Dpn1 digestion, Dpn1-resistant extrachromosomalDNA was quantified by real-time PCR (n = 2). P < 0.05. (E) LMO2-induced DNAreplication depends on LMO2 tethering to DNA via GAL4 binding sites (UAS).The experiment was performed as in D. (F) LMO2 binding to artificial origins ofreplication is sufficient to recruit DNA replication proteins specifically. DNAcapture was performed using immobilized WT or mutant UAS sequences onbeads. Bound proteins were revealed byWestern blotting. Data are the mean ±SD of at least two independent experiments performed in duplicate.

Sincennes et al. PNAS | February 2, 2016 | vol. 113 | no. 5 | 1395

MED

ICALSC

IENCE

S

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 4: The LMO2 oncogene regulates DNA replication in ... · Christophe Cazauxc,2, Nazar Mashtalir d, EL Bachir Affar , Alain Verreaulta,b,e, and Trang Hoanga,b,e,3 a Institute of Research

in which Lmo2 was knocked down or overexpressed, respectively(12, 46). Furthermore, replication genes were not occupied by theSCL–LMO2 complex in leukemic T cells (25). Together, these ob-servations indicate a critical role for LMO2 in erythroid cell fate, viathe control of DNA replication and the cell cycle that drives Epo-dependent expansion of erythroid progenitors while impeding theircommitment to terminal maturation.

LMO2 Expression Regulates S-Phase Entry. We next addressed theimportance of LMO2 levels on the kinetics of cell cycle pro-gression (Fig. 3J). Briefly, we found that viable MEL cells withlow side and forward scatter properties are mostly in G0/G1 (Fig.S3E). These G0/G1 cells were purified and released in culturemedium to analyze their DNA content by Hoechst staining (Fig.S3E). Decreased LMO2 (shRNA Lmo2) reproducibly caused a

2-h delay in G1/S transition, occurring at 3 h after releasecompared with 1 h for control cells (Vector), as well as delayedS-phase progression (Fig. 3J and Fig. S3E), indicating thatLMO2 levels control S-phase entry in erythroblasts.To mimic retroviral integration upstream of the LMO2 locus in

the X-SCID gene therapy trial (19, 20, 46) and define the conse-quences of ectopic LMO2 expression in vivo, we delivered LMO2 inhematopoietic stem cells by retroviral infection, (Fig. 4 A–F). Upontransplantation, LMO2-transduced bone marrow cells inducedT-ALL in 60% of recipient mice, with a median of 270 days (46) (Fig.4B and Fig. S4). During the preleukemic stage (30 days), LMO2overexpression mostly affected thymocytes and led to an accumula-tion of CD44+CD25−CD4−D8− double-negative thymocyte (DN)cells and a decrease in CD4+CD8+ double-positive thymocyte(DP) cells in the thymus, reproducing the differentiation blockadeat the DN stage reported in LMO2 transgenic mice (Fig. 4C),despite the fact that the retroviral vector allowed for transgeneexpression in all cells. Elevating LMO2 in thymocytes modified thecell cycle status of thymocyte progenitors (Fig. 4D) without af-fecting bone marrow stem cells (Fig. S4B), consistent with the roleof LMO2 as a T-cell specific oncogene (46). Interestingly, elevatingLMO2 enhanced the cell cycle status of both DN1 and DP cells,suggesting that differentiation blockade (lower number of DP cells)could be due to increased cell cycle. Furthermore, when DN1thymocytes were synchronized in G0/G1 and released in culture,LMO2-expressing cells entered more rapidly into S phase com-pared with control cells (vector) (Fig. 4E), indicating that LMO2facilitates the G1-S transition and S phase progression.The above results indicate that LMO2 overexpression in hema-

topoietic cells reproduced the T-cell proliferation phenotypereported for preleukemic patients in the X-SCID gene therapy trial.Actively dividing cells are more sensitive to 5-fluorouracil (5-FU), anantimetabolite that inhibits thymidylate synthase and therefore de-pletes the pool of dTTP. Consistent with increased DNA replication,LMO2-expressing thymocytes cocultured with MS5 stromal cellsexpressing DL4 were 100-fold more sensitive thanWT thymocytes to5-FU (EC50 of 5 nM vs. 500 nM) (Fig. 4F).

DiscussionLMO2 has a well-established function in transcriptional regulationvia direct interaction with transcription factors, mostly of the bHLHfamily, SCL/TAL1, TAL2, and LYL1 or GATA proteins (2–4). Inthis study, we revealed unexpected new functions of LMO2 in he-matopoiesis and leukemogenesis through a yeast two-hybrid screenof LMO2 interaction partners in hematopoietic progenitors. Ourobservations indicate that LMO2 controls cell fate by directly pro-moting DNA replication, a hitherto unrecognized mechanism thatmight also account for its oncogenic properties.Our study unravels unexpected interactions between LMO2 and

three essential replication proteins, MCM6, PRIM1, and POLD1(30), as well as chromatin-modifying enzymes that have well-characterized roles in DNA replication, BAZ1A, SETD8, MYST2,and UHRF1 (31–33, 47). More importantly, tethering LMO2 toDNA via the GAL4-DNA binding domain was sufficient to recruitMCM5 and POLD1 to DNA and to transform UAS into origins ofreplication, indicating that LMO2 directly controls DNA replication.Unlike other transcription complexes that have a dual role in DNAreplication (48), LMO2 interaction with the RC is distinct from thewell-known recruitment of LMO2 to the SCL transcription complex.Our data are in line with the observations that there is only a partialoverlap between SCL and LMO2 chromatin occupancy in hema-topoietic progenitors (24).LMO2 binding to replication initiation zones reported here

overlaps with early G1 replication segments described in lymphoidcells (39, 42), a possibility that may be favored by its interaction withMLL2 (Fig. 1A). Indeed, the H3K4me3 histone mark found in earlyreplicating domains (40) can be controlled by MLL2. In eukaryotes,origins are licensed in excess, and not all licensed origins are active

AE1

Kit+ Lin-CD71 Ter119

B

CF

LMO2

low

high

IgG LMO2

E1

HSPC E2 E3 E4MELTF1

E

Cel

l cou

nt

DAPI

vector

NT

shRNA Lmo2

44.5%

52.2%

25.4%

CD71+Ter119-

J

CD

71

E1

E3E4

E2

Ter119

G

NT shRNALmo2

E1

(%)

Ctrl Epo

IMEL

New

ly s

ynth

esiz

ed D

NA

Vector

Time (h)

shRNALmo2

MEL

S p

hase

(% to

tal)

VectorshRNALmo2

Fold

exp

ansi

on

Time (d)

CD71+Ter119-

0 1 2 3 40

1

2

3

shRNALmo2

NT

H

time(h) E1 E2 E3 E40 3.1 0.5 0.5 3.424 0.5 0.9 3.5 2.548 0.1 1.1 5.4 96.972 0.1 2.6 4.2 94.2

Abundance ratio of E1-4 subsets in Lmo2shRNA/control

E4E2 E3

%S/

G2/

M

E1loE1hi

D E2 E3 E4LMO2

GATA1SCL

MFI4.5 3.1

7.4 18 20

E1

LMO

2 (M

FI)

Time (h)

Fig. 3. LMO2 levels determine the proliferation of erythroid progenitors.(A) Diagram of erythroid differentiation according to flow cytometry profiles.(B) LMO2 levels in bone marrow erythroid progenitors (Lin−CD71+Ter119−) cor-relate with their cell cycle status by Hoechst staining: E1low and E1high (P ≤ 0.05).The mean fluorescence intensity (MFI) for LMO2 per cell was assessed by flowcytometry. (C) Correlation between LMO2 levels and the proportion of cells inS/G2/M during erythroid differentiation (E1 to E4). (D) Expression levels of LMO2,SCL, and GATA1 during erythroid differentiation. (E) Decreased LMO2 in ery-throid progenitors reduced the fraction of cells in S phase. The shRNA Lmo2 wasdelivered in Ter119− fetal liver cells, which were then stimulated with Epo for 2 d.The cell cycle in erythroid progenitors (E1, CD71+Ter119−) was analyzed by DAPIstaining. (F) Lmo2 is required for the response of proerythroblasts to Epo. NT,nontarget (control). (G) LMO2 levels control the proliferation of erythroid progen-itors in culture. (H) Decreased LMO2 abolishes the growth of erythroid progenitors(E1) in response to Epo but favors terminally differentiating erythroid cells (E4).(I) Decreased DNA synthesis induced by shRNA-mediated Lmo2 depletion. MEL cellswere purified in G0/G1 and released in culture for different times in the presence ofα32P-dCTP. After electrophoresis, total DNAwas quantified by autoradiography (n=2). The slopes of the two curves were 0.48 ± 0.02 (Vector) and 0.29 ± 0.01 (shLmo2).*P < 0.0001. (J) Decreased proportion of cells in S phase after Lmo2 depletion. MELcells expressing shLmo2 or control (Vector) were purified in G0/G1 and analyzed forcell cycle progression at different time points in culture by Hoechst staining (n = 3).All data shown are typical of at least two independent experiments.

1396 | www.pnas.org/cgi/doi/10.1073/pnas.1515071113 Sincennes et al.

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 5: The LMO2 oncogene regulates DNA replication in ... · Christophe Cazauxc,2, Nazar Mashtalir d, EL Bachir Affar , Alain Verreaulta,b,e, and Trang Hoanga,b,e,3 a Institute of Research

during a given cell cycle, which requires recruitment of DNA poly-merases to initiate DNA synthesis (30). Accordingly, we show thatLMO2 levels influence the rate of DNA replication in MEL cells.Although we focused on the basic components of the pre-RC

in the present study, several other proteins identified in thescreen could have a regulatory function. For example, cyclinA2-CDK2 favors the onset of DNA replication at the G1-Stransition and prevents rereplication during S phase (49).These possibilities would be consistent with the effects ofLMO2 levels on the kinetics of G1-S transition that we ob-served in two cell types, delayed when Lmo2 is knocked downin murine erythroleukemic cells and, conversely, accelerated inLMO2-overexpressing DN1 thymocytes. In addition, LMO2associates with two cell cycle checkpoint proteins, CDK9 andBUB1B (Table S1). CDK9 associates with cyclin K in replica-tion stress response and prevents DNA damage in replicatingcells (50). BUB1B is an essential component of the mitoticcheckpoint. The importance of these cell cycle proteins forLMO2-induced cell proliferation remains to be addressed.Cell cycle is a highly regulated process in the erythroid lineage.

CD71+Ter119− proerythroblasts represent a critical transitionalstage marked by Epo dependency (51) and elevated DNA repli-cation (45), both shown here to require high LMO2 protein ex-pression. Conversely, terminal erythroid differentiation to theCD71−Ter119+ stage is necessarily coordinated with growth ar-rest (13, 51), which we now show to be favored by Lmo2 down-

regulation, in agreement with previous work indicating that LMO2overexpression prevents terminal erythroid differentiation (11). Weconclude that LMO2 down-regulation is required for the switch toterminal erythroid differentiation due to the implication of LMO2in DNA replication/cell cycle (Fig. 4G).It is well established that transcription factor gene networks

drive hematopoietic cell development and lineage outcome, viasynergistic or antagonistic interactions (reviewed in refs. 52 and 53).It is unclear how LMO2 inhibits cell differentiation in both theerythroid (11) and T-lymphoid lineages (reviewed in ref. 53).We now propose that LMO2-dependent DNA replication inboth lineages governs the switch between a proliferative state inprogenitors and commitment to terminal differentiation (Fig.4G). Failure to regulate this proliferative switch caused by ec-topic LMO2 expression (19) may lead to T-ALL.Oncogene-induced DNA replication stress (54) could lead to

replication errors and ultimately cause genetic lesions, such as acti-vating NOTCH1 mutations (55, 56), that convert preleukemic stemcells (15, 16) into leukemia initiating cells (15). Two other LIM-onlyproteins, LMO1 and LMO4, are also important determinants of cellcycle progression in neuroblastoma (23) and in breast cancer asso-ciated with genomic instability (22), respectively, suggesting that themechanism(s) described here may be extended to these proteins.Emerging evidence indicates that oncogenes, such as c-MYC orHOXD13, can be part of nontranscriptional complexes involvedin DNA replication (54, 57). Taken together, we propose that on-cogenic transcription factors in acute leukemias and possibly othertumor types transform cells by at least two DNA-dependent mech-anisms, the control of gene expression programs, as initially pro-posed (1), but also by deregulating DNA replication (42), with bothprocesses being important determinants of cell fate.

Materials and MethodsYeast Two-Hybrid. LMO2 full-length protein was subcloned in the pGBKT7vector and transformed in the Y187 yeast strain. The cDNA library fromc-Kit+Lin− murine bone marrow cells was constructed in the pGADT7 vectorwith the Matchmaker Library Construction and Screening Kit (BD Biosci-ences) and transformed in the AH109 yeast strain. The screen (selection ofAde+His+Leu+Trp+ colonies) was performed by yeast mating. Positive cloneswere isolated and sequenced. To confirm the interactions and check forspecificity, full-length cDNAs for Polδ1, Prim1, Mcm2-5-6-10, Cdt1, Pcna, andCaf1 were subcloned into pGADT7 and transformed in AH109 cells.

Retroviral Gene Transfer and RNAi. LMO2 retroviral gene transfer and mousebone marrow transplantation were performed as described (8). For RNAi ex-periments, vesicular stomatitis virus (VSV-G)–producing cells were transfectedwith plasmids encoding shRNAs against LMO2, with a nontarget shRNA or withthe empty pLKO.1 vector (Sigma). MEL cells or Lin− fetal liver cells were in-cubated with VSV-G supernatant for 48 h and then selected with puromycin asdescribed (8). All mice were kept under pathogen-free conditions according toinstitutional animal care and use guidelines. The protocols for gene transferand transplantation in mice were approved by the Committee of Ethics andAnimal Deontology of the University of Montreal.

Flow Cytometry, Cell Cycle Analysis, and Cell Sorting. Flow cytometry, cell cycleanalysis, and cell sorting are described in Supporting Information.

32P Orthophosphate Labeling and Quantification of de Novo DNA Synthesis.MEL cells synchronized in G0/G1 were incubated with α32P-dCTP (100 μCi/mL;PerkinElmer) in DMEM [10 mMHepes (pH 7.4), 10% (vol/vol) FCS] at 8 × 105 cellsper milliliter for the indicated times. Cells were lysed in DNA extraction buffer[80 mM Tris·HCl (pH 8.0), 8 mM EDTA, 100 mM NaCl, 0.5% (g/100 mL) SDS].After proteinase K digestion and phenol/chloroform extraction, DNA was re-solved on an alkaline (NaOH) agarose gel. The gel was dried on a Biodynemembrane (Pall Corporation). 32P incorporation was visualized by autoradi-ography and quantitated using ImageQuant software (GE Healthcare).

Protein Extraction, Immunoprecipitation, Immunoblot, Abs, RNA, and ChIP Analysis.Protein extraction, immunoprecipitation, immunoblot (IB), Abs, RNA, and ChIPanalysis methods are fully described in Supporting Information. Primer sequences

T

T

***

*

A B

C D

E F G

Fig. 4. Ectopic LMO2 expression in thymocytes triggers G1/S transition, T-cellhyperproliferation, and T-ALL development. (A) Experimental strategy to deliverLMO2 in hematopoietic cells using the murine stem cell virus (MSCV) retroviralvector. BM, bone marrow. (B) LMO2 overexpression induces T-ALL in mice.Kaplan–Meier curves showing the time of leukemia onset in mice transplantedwithMSCV-LMO2 or MSCV transduced cells. (C and D) LMO2 overexpression leadsto an increase of the DN1 thymocyte subset but a decrease in DP cells in vivo,despite higher proportions of cycling cells in both populations. (E) LMO2 over-expression induces an increase in the percentage of DN1 cells in S phase. DN1 cellswere purified in G0/G1 and analyzed for cell cycle progression by DAPI staining atdifferent time points after coculture with MS5-DL4 stromal cells. (F) LMO2-expressing thymocytes are more sensitive to 5-FU treatment in vitro. Cells werecocultured onMS5-DL4 stromal cells and treated with 5-FU at the indicated doses.Viable Thy1+ cells were scored by flow cytometry. (G) Proposed model: LMO2controls DNA replication, favors progenitor cell proliferation, and inhibits com-mitment to terminal differentiation in the erythroid and T-lymphoid lineages.Data shown are typical of at least two independent experiments.

Sincennes et al. PNAS | February 2, 2016 | vol. 113 | no. 5 | 1397

MED

ICALSC

IENCE

S

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0

Page 6: The LMO2 oncogene regulates DNA replication in ... · Christophe Cazauxc,2, Nazar Mashtalir d, EL Bachir Affar , Alain Verreaulta,b,e, and Trang Hoanga,b,e,3 a Institute of Research

used are shown in Tables S2 and S3, and Abs used for ChIP, IB, and immuno-fluorescence are listed in Table S4.

Transient Replication Assay in Mammalian Cells. The transient replicationassay is adapted from Takeda et al. (43) and described in SupportingInformation.

ACKNOWLEDGMENTS. We thank Danièle Gagné [Institute of Research in Immu-nology and Cancer (IRIC)] for her assistance with flow cytometry, VéroniqueLitalien for mouse handling, Geneviève Boucher for bioinformatic analyses,Francis Migneault and David Flaschner for assistance with the yeast two-hybridconfirmation assays, Drs. Jalila Chagraoui and Richard Martin for help with the

retroviral gene transfer and the yeast two-hybrid, and Dr. Jana Krosl for criticalcomments on the manuscript. This work was funded by the Cancer ResearchSociety, Inc. (2012–2014); the Canadian Institutes for Health Research (CIHR; GrantMOP111050, 2011–2016); Canadian Cancer Society Research Institute Grant019222 (to T.H.); the Leukemia & Lymphoma Society (2013–2015) (T.H. andE.B.A.); CIHR Grant 89928 (to A.V.); a CIHR multiuser grant to support the flowcytometry and imaging service; and a group grant from the Fonds de Recherchedu Québec-Santé to support, in part, IRIC infrastructure. M.-C.S. was supported bya Canada Graduate Scholarship Doctoral Award (CIHR) and a doctoral award fromthe Cole Foundation. M.H. was supported by a postdoctoral fellowship awardof the Swiss National Foundation (PBBEP3 144798), by Swiss Foundationfor Fellowships in Medicine and Biology, and by Novartis (P3SMP3 151720).V.L. and D.F.T.V. were supported by Cole Foundation awards.

1. Ferrando AA, et al. (2002) Gene expression signatures define novel oncogenic path-ways in T cell acute lymphoblastic leukemia. Cancer Cell 1(1):75–87.

2. Wadman I, et al. (1994) Specific in vivo association between the bHLH and LIM pro-teins implicated in human T cell leukemia. EMBO J 13(20):4831–4839.

3. Lécuyer E, et al. (2007) Protein stability and transcription factor complex assemblydetermined by the SCL-LMO2 interaction. J Biol Chem 282(46):33649–33658.

4. El Omari K, et al. (2013) Structural basis for LMO2-driven recruitment of the SCL:E47bHLHheterodimer to hematopoietic-specific transcriptional targets. Cell Reports 4(1):135–147.

5. Lécuyer E, Hoang T (2004) SCL: From the origin of hematopoiesis to stem cells andleukemia. Exp Hematol 32(1):11–24.

6. Love PE, Warzecha C, Li L (2014) Ldb1 complexes: The new master regulators oferythroid gene transcription. Trends Genet 30(1):1–9.

7. Org T, et al. (2015) Scl binds to primed enhancers in mesoderm to regulate hema-topoietic and cardiac fate divergence. EMBO J 34(6):759–777.

8. Lacombe J, et al. (2010) Scl regulates the quiescence and the long-term competence ofhematopoietic stem cells. Blood 115(4):792–803.

9. Li L, et al. (2011) Nuclear adaptor Ldb1 regulates a transcriptional program essentialfor the maintenance of hematopoietic stem cells. Nat Immunol 12(2):129–136.

10. Warren AJ, et al. (1994) The oncogenic cysteine-rich LIM domain protein rbtn2 isessential for erythroid development. Cell 78(1):45–57.

11. Visvader JE, Mao X, Fujiwara Y, Hahm K, Orkin SH (1997) The LIM-domain bindingprotein Ldb1 and its partner LMO2 act as negative regulators of erythroid differen-tiation. Proc Natl Acad Sci USA 94(25):13707–13712.

12. Fujiwara T, Lee HY, Sanalkumar R, Bresnick EH (2010) Building multifunctionality intoa complex containing master regulators of hematopoiesis. Proc Natl Acad Sci USA107(47):20429–20434.

13. Goardon N, et al. (2006) ETO2 coordinates cellular proliferation and differentiationduring erythropoiesis. EMBO J 25(2):357–366.

14. Anderson MK (2006) At the crossroads: Diverse roles of early thymocyte transcrip-tional regulators. Immunol Rev 209:191–211.

15. Gerby B, et al. (2014) SCL, LMO1 and Notch1 reprogram thymocytes into self-renewing cells. PLoS Genet 10(12):e1004768.

16. McCormack MP, et al. (2010) The Lmo2 oncogene initiates leukemia in mice by in-ducing thymocyte self-renewal. Science 327(5967):879–883.

17. Chervinsky DS, et al. (1999) Disordered T-cell development and T-cell malignancies in SCLLMO1double-transgenicmice: Parallels with E2A-deficientmice.Mol Cell Biol 19(7):5025–5035.

18. Homminga I, et al. (2012) Characterization of a pediatric T-cell acute lymphoblasticleukemia patient with simultaneous LYL1 and LMO2 rearrangements. Haematologica97(2):258–261.

19. Hacein-Bey-Abina S, et al. (2003) LMO2-associated clonal T cell proliferation in twopatients after gene therapy for SCID-X1. Science 302(5644):415–419.

20. Howe SJ, et al. (2008) Insertional mutagenesis combined with acquired somatic mu-tations causes leukemogenesis following gene therapy of SCID-X1 patients. J ClinInvest 118(9):3143–3150.

21. Boehm T, Foroni L, Kaneko Y, Perutz MF, Rabbitts TH (1991) The rhombotin family ofcysteine-rich LIM-domain oncogenes: Distinct members are involved in T-cell transloca-tions to human chromosomes 11p15 and 11p13. Proc Natl Acad Sci USA 88(10):4367–4371.

22. Montañez-Wiscovich ME, et al. (2009) LMO4 is an essential mediator of ErbB2/HER2/Neu-induced breast cancer cell cycle progression. Oncogene 28(41):3608–3618.

23. Wang K, et al. (2011) Integrative genomics identifies LMO1 as a neuroblastoma on-cogene. Nature 469(7329):216–220.

24. Wilson NK, et al. (2010) Combinatorial transcriptional control in blood stem/progenitor cells: Genome-wide analysis of ten major transcriptional regulators.Cell Stem Cell 7(4):532–544.

25. Sanda T, et al. (2012) Core transcriptional regulatory circuit controlled by the TAL1complex in human T cell acute lymphoblastic leukemia. Cancer Cell 22(2):209–221.

26. Palii CG, et al. (2011) Differential genomic targeting of the transcription factor TAL1in alternate haematopoietic lineages. EMBO J 30(3):494–509.

27. KassoufMT, et al. (2010) Genome-wide identification of TAL1’s functional targets: Insightsinto its mechanisms of action in primary erythroid cells. Genome Res 20(8):1064–1083.

28. Palomero T, et al. (2006) Transcriptional regulatory networks downstream of TAL1/SCL in T-cell acute lymphoblastic leukemia. Blood 108(3):986–992.

29. Herblot S, Steff AM, Hugo P, Aplan PD, Hoang T (2000) SCL and LMO1 alter thymocytedifferentiation: Inhibition of E2A-HEB function and pre-T alpha chain expression.Nat Immunol 1(2):138–144.

30. Remus D, Diffley JF (2009) Eukaryotic DNA replication control: Lock and load, thenfire. Curr Opin Cell Biol 21(6):771–777.

31. Collins N, et al. (2002) An ACF1-ISWI chromatin-remodeling complex is required forDNA replication through heterochromatin. Nat Genet 32(4):627–632.

32. Beck DB, et al. (2012) The role of PR-Set7 in replication licensing depends on Suv4-20h. Genes Dev 26(23):2580–2589.

33. Wu ZQ, Liu X (2008) Role for Plk1 phosphorylation of Hbo1 in regulation of repli-cation licensing. Proc Natl Acad Sci USA 105(6):1919–1924.

34. Katsuno Y, et al. (2009) Cyclin A-Cdk1 regulates the origin firing program in mam-malian cells. Proc Natl Acad Sci USA 106(9):3184–3189.

35. Méndez J, Stillman B (2000) Chromatin association of human origin recognitioncomplex, cdc6, and minichromosome maintenance proteins during the cell cycle:Assembly of prereplication complexes in late mitosis. Mol Cell Biol 20(22):8602–8612.

36. Lahlil R, Lécuyer E, Herblot S, Hoang T (2004) SCL assembles a multifactorial complexthat determines glycophorin A expression. Mol Cell Biol 24(4):1439–1452.

37. Kitamura T, et al. (1995) Efficient screening of retroviral cDNA expression libraries.Proc Natl Acad Sci USA 92(20):9146–9150.

38. Cadoret JC, et al. (2008) Genome-wide studies highlight indirect links between humanreplication origins and gene regulation. Proc Natl Acad Sci USA 105(41):15837–15842.

39. Hansen RS, et al. (2010) Sequencing newly replicated DNA reveals widespread plas-ticity in human replication timing. Proc Natl Acad Sci USA 107(1):139–144.

40. Ryba T, et al. (2010) Evolutionarily conserved replication timing profiles predict long-rangechromatin interactions and distinguish closely related cell types.Genome Res 20(6):761–770.

41. Karnani N, Taylor CM, Malhotra A, Dutta A (2010) Genomic study of replication ini-tiation in human chromosomes reveals the influence of transcription regulation andchromatin structure on origin selection. Mol Biol Cell 21(3):393–404.

42. Ryba T, et al. (2012) Abnormal developmental control of replication-timing domainsin pediatric acute lymphoblastic leukemia. Genome Res 22(10):1833–1844.

43. Takeda DY, Shibata Y, Parvin JD, Dutta A (2005) Recruitment of ORC or CDC6 to DNAis sufficient to create an artificial origin of replication in mammalian cells. Genes Dev19(23):2827–2836.

44. Grondin B, et al. (2007) c-Jun homodimers can function as a context-specific co-activator. Mol Cell Biol 27(8):2919–2933.

45. Pop R, et al. (2010) A key commitment step in erythropoiesis is synchronized with thecell cycle clock through mutual inhibition between PU.1 and S-phase progression.PLoS Biol 8(9):e1000484.

46. Treanor LM, et al. (2011) Functional interactions between Lmo2, the Arf tumor sup-pressor, and Notch1 in murine T-cell malignancies. Blood 117(20):5453–5462.

47. Nishiyama A, et al. (2013) Uhrf1-dependent H3K23 ubiquitylation couples mainte-nance DNA methylation and replication. Nature 502(7470):249–253.

48. Karmakar S, Mahajan MC, Schulz V, Boyapaty G, Weissman SM (2010) A multiproteincomplex necessary for both transcription and DNA replication at the β-globin locus.EMBO J 29(19):3260–3271.

49. Girard F, Strausfeld U, Fernandez A, Lamb NJ (1991) Cyclin A is required for the onsetof DNA replication in mammalian fibroblasts. Cell 67(6):1169–1179.

50. Yu DS, et al. (2010) Cyclin-dependent kinase 9-cyclin K functions in the replicationstress response. EMBO Rep 11(11):876–882.

51. Zhang J, Socolovsky M, Gross AW, Lodish HF (2003) Role of Ras signaling in erythroiddifferentiation of mouse fetal liver cells: Functional analysis by a flow cytometry-based novel culture system. Blood 102(12):3938–3946.

52. Graf T, Enver T (2009) Forcing cells to change lineages. Nature 462(7273):587–594.53. Yui MA, Rothenberg EV (2014) Developmental gene networks: A triathlon on the

course to T cell identity. Nat Rev Immunol 14(8):529–545.54. Srinivasan SV, Dominguez-Sola D, Wang LC, Hyrien O, Gautier J (2013) Cdc45 is a critical

effector of myc-dependent DNA replication stress. Cell Reports 3(5):1629–1639.55. Tremblay M, et al. (2010) Modeling T-cell acute lymphoblastic leukemia induced by

the SCL and LMO1 oncogenes. Genes Dev 24(11):1093–1105.56. Weng AP, et al. (2004) Activating mutations of NOTCH1 in human T cell acute lym-

phoblastic leukemia. Science 306(5694):269–271.57. Salsi V, et al. (2009) HOXD13 binds DNA replication origins to promote origin li-

censing and is inhibited by geminin. Mol Cell Biol 29(21):5775–5788.58. Mashtalir N, et al. (2014) Autodeubiquitination protects the tumor suppressor BAP1

from cytoplasmic sequestration mediated by the atypical ubiquitin ligase UBE2O.Mol Cell 54(3):392–406.

59. An X, et al. (2014) Global transcriptome analyses of human and murine terminalerythroid differentiation. Blood 123(22):3466–3477.

60. Wong P, et al. (2011) Gene induction and repression during terminal erythropoiesisare mediated by distinct epigenetic changes. Blood 118(16):e128–e138.

61. Demers C, et al. (2007) Activator-mediated recruitment of the MLL2 methyltransferasecomplex to the beta-globin locus. Mol Cell 27(4):573–584.

62. DeVilbiss AW, et al. (2015) Epigenetic Determinants of Erythropoiesis: Role of theHistone Methyltransferase SetD8 in Promoting Erythroid Cell Maturation and Sur-vival. Mol Cell Biol 35(12):2073–2087.

1398 | www.pnas.org/cgi/doi/10.1073/pnas.1515071113 Sincennes et al.

Dow

nloa

ded

by g

uest

on

Oct

ober

13,

202

0