14
Molecular Cell Article RBM5, 6, and 10 Differentially Regulate NUMB Alternative Splicing to Control Cancer Cell Proliferation Elias G. Bechara, 1,2 Endre Sebestye ´ n, 2 Isabella Bernardis, 2 Eduardo Eyras, 2,3 and Juan Valca ´ rcel 1,2,3, * 1 Centre de Regulacio ´ Geno ` mica 2 Universitat Pompeu Fabra 3 Institucio ´ Catalana de Recerca i Estudis Avanc ¸ ats (ICREA) Dr. Aiguader, 88, 08003 Barcelona, Spain *Correspondence: [email protected] http://dx.doi.org/10.1016/j.molcel.2013.11.010 SUMMARY RBM5, a regulator of alternative splicing of apoptotic genes, and its highly homologous RBM6 and RBM10 are RNA-binding proteins frequently deleted or mutated in lung cancer. We report that RBM5/6 and RBM10 antagonistically regulate the proliferative ca- pacity of cancer cells and display distinct positional effects in alternative splicing regulation. We identify the Notch pathway regulator NUMB as a key target of these factors in the control of cell proliferation. NUMB alternative splicing, which is frequently altered in lung cancer, can regulate colony and xeno- graft tumor formation, and its modulation recapitu- lates or antagonizes the effects of RBM5, 6, and 10 in cell colony formation. RBM10 mutations identified in lung cancer cells disrupt NUMB splicing regulation to promote cell growth. Our results reveal a key ge- netic circuit in the control of cancer cell proliferation. INTRODUCTION Lung cancer is the leading cause of cancer-related deaths worldwide, 40% of which correspond to nonsmall cell lung ade- nocarcinomas. RBM5 (also known as LUCA-15/H37), RBM6, and RBM10 are highly similar RNA-binding motif proteins sharing 30%–50% of amino acid sequence identity. RBM5 and RBM6 genes map to chromosomal region 3p21.3, which is frequently deleted in heavy smokers, lung cancer, and other tissue carcinomas (Angeloni, 2007). RBM5 protein levels are indeed downregulated in 75% of primary lung cancers (Oh et al., 2002) as well as in prostate (Zhao et al., 2012) and in some breast cancer samples (Edamatsu et al., 2000).Overex- pression of RBM5 in breast cancer samples was also reported (Oh et al., 1999; Rintala-Maki et al., 2007), suggesting that both up- and downregulation of RBM5 could play a role in tumor pro- gression. RBM6 and 10 also show altered expression in breast cancer (Rintala-Maki et al., 2007). Downregulation of RBM5 is considered one molecular signature associated with metastasis in various solid tumors (Edamatsu et al., 2000; Ramaswamy et al., 2003; Welling et al., 2002). More recently, RBM10 was found to be among the most frequently mutated genes in lung adenocarcinoma samples (Imielinski et al., 2012). The frequency of genetic lesions affecting RBM5, 6, and 10 in lung and other cancers suggests that disruption of the function of these proteins can contribute to tumor progression. Overexpression of RBM5 in various cell lines leads to growth arrest, induction of apoptosis, and retarded tumor growth when cells were injected in nude mice (Mourtada-Maarabouni et al., 2002, 2003; Oh et al., 2002, 2006). RBM5-mediated apoptosis is associated with upregulation of the proapoptotic protein BAX and downregulation of the antiapoptotic proteins BCL-2 and BCL-XL (Mourtada-Maarabouni et al., 2002; Oh et al., 2006; Sutherland et al., 2001). Downregulation of RBM5 affects the transcript levels of 35 genes involved in the control of cell proliferation and apoptosis (Mourtada-Maarabouni et al., 2006), but the key targets of RBM5, 6, and 10 and the regulatory mechanisms behind these properties remain poorly understood. RBM 5, 6, and 10 contain domains characteristic of proteins with functions in RNA metabolism and pre-mRNA splicing, including two RNA recognition motifs (RRM), two Zinc fingers, bipartite nuclear localization signals, a glycine(G)-patch, one arginine/serine-rich domain, and an OCRE domain (which con- sists of a repeated sequence of eight residues, organized around a triplet of conserved aromatic amino acids) (Figure 1A). Consis- tent with the presence of diverse RNA-binding domains, a variety of sequences have been identified as bound by these proteins. Thus, in vitro studies showed that RBM5 and 10 specifically bind poly(G) RNA tracts in vitro (Edamatsu et al., 2000) and that the RanBP-ZnF domain of these proteins binds with high affinity to ANGUAA sequences (Nguyen et al., 2011). On the other hand, the solution structure of the second RRM of RBM5 revealed spe- cific recognition of both CU-rich and GA-rich sequences (Song et al., 2012). Ex vivo and in vitro data revealed that RBM5 binds to a U/C-rich sequence in the caspase-2 pre-mRNA (Fushimi et al., 2008) and regulates alternative splicing of apoptosis- related genes (Bonnal et al., 2008; Fushimi et al., 2008), modu- lating the production of isoforms with antagonistic functions in programmed cell death. It was also shown that RBM5 targets late events in spliceosome assembly, acting as a switch for splice site pairing in the Fas receptor (CD95/APO-1) pre-mRNA (Bonnal 720 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc.

RBM5, 6, and 10 Differentially Regulate NUMB Alternative

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Molecular Cell

Article

RBM5, 6, and 10 Differentially RegulateNUMB Alternative Splicing to ControlCancer Cell ProliferationElias G. Bechara,1,2 Endre Sebestyen,2 Isabella Bernardis,2 Eduardo Eyras,2,3 and Juan Valcarcel1,2,3,*1Centre de Regulacio Genomica2Universitat Pompeu Fabra3Institucio Catalana de Recerca i Estudis Avancats (ICREA)Dr. Aiguader, 88, 08003 Barcelona, Spain

*Correspondence: [email protected]

http://dx.doi.org/10.1016/j.molcel.2013.11.010

SUMMARY

RBM5, a regulator of alternative splicing of apoptoticgenes, and its highly homologous RBM6 and RBM10are RNA-binding proteins frequently deleted ormutated in lung cancer. We report that RBM5/6 andRBM10 antagonistically regulate the proliferative ca-pacity of cancer cells and display distinct positionaleffects in alternative splicing regulation. We identifythe Notch pathway regulator NUMB as a key targetof these factors in the control of cell proliferation.NUMB alternative splicing, which is frequentlyaltered in lung cancer, can regulate colony and xeno-graft tumor formation, and its modulation recapitu-lates or antagonizes the effects of RBM5, 6, and 10in cell colony formation. RBM10 mutations identifiedin lung cancer cells disrupt NUMB splicing regulationto promote cell growth. Our results reveal a key ge-netic circuit in the control of cancer cell proliferation.

INTRODUCTION

Lung cancer is the leading cause of cancer-related deaths

worldwide, 40% of which correspond to nonsmall cell lung ade-

nocarcinomas. RBM5 (also known as LUCA-15/H37), RBM6,

and RBM10 are highly similar RNA-binding motif proteins

sharing 30%–50% of amino acid sequence identity. RBM5 and

RBM6 genes map to chromosomal region 3p21.3, which is

frequently deleted in heavy smokers, lung cancer, and other

tissue carcinomas (Angeloni, 2007). RBM5 protein levels are

indeed downregulated in �75% of primary lung cancers (Oh

et al., 2002) as well as in prostate (Zhao et al., 2012) and in

some breast cancer samples (Edamatsu et al., 2000).Overex-

pression of RBM5 in breast cancer samples was also reported

(Oh et al., 1999; Rintala-Maki et al., 2007), suggesting that both

up- and downregulation of RBM5 could play a role in tumor pro-

gression. RBM6 and 10 also show altered expression in breast

cancer (Rintala-Maki et al., 2007). Downregulation of RBM5 is

considered one molecular signature associated with metastasis

720 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier

in various solid tumors (Edamatsu et al., 2000; Ramaswamy

et al., 2003; Welling et al., 2002). More recently, RBM10 was

found to be among the most frequently mutated genes in lung

adenocarcinoma samples (Imielinski et al., 2012). The frequency

of genetic lesions affecting RBM5, 6, and 10 in lung and other

cancers suggests that disruption of the function of these proteins

can contribute to tumor progression.

Overexpression of RBM5 in various cell lines leads to growth

arrest, induction of apoptosis, and retarded tumor growth

when cells were injected in nude mice (Mourtada-Maarabouni

et al., 2002, 2003; Oh et al., 2002, 2006). RBM5-mediated

apoptosis is associated with upregulation of the proapoptotic

protein BAX and downregulation of the antiapoptotic proteins

BCL-2 and BCL-XL (Mourtada-Maarabouni et al., 2002; Oh

et al., 2006; Sutherland et al., 2001). Downregulation of RBM5

affects the transcript levels of 35 genes involved in the control

of cell proliferation and apoptosis (Mourtada-Maarabouni et al.,

2006), but the key targets of RBM5, 6, and 10 and the regulatory

mechanisms behind these properties remain poorly understood.

RBM 5, 6, and 10 contain domains characteristic of proteins

with functions in RNA metabolism and pre-mRNA splicing,

including two RNA recognition motifs (RRM), two Zinc fingers,

bipartite nuclear localization signals, a glycine(G)-patch, one

arginine/serine-rich domain, and an OCRE domain (which con-

sists of a repeated sequence of eight residues, organized around

a triplet of conserved aromatic amino acids) (Figure 1A). Consis-

tent with the presence of diverse RNA-binding domains, a variety

of sequences have been identified as bound by these proteins.

Thus, in vitro studies showed that RBM5 and 10 specifically

bind poly(G) RNA tracts in vitro (Edamatsu et al., 2000) and that

the RanBP-ZnF domain of these proteins binds with high affinity

to ANGUAA sequences (Nguyen et al., 2011). On the other hand,

the solution structure of the second RRM of RBM5 revealed spe-

cific recognition of both CU-rich and GA-rich sequences (Song

et al., 2012). Ex vivo and in vitro data revealed that RBM5 binds

to a U/C-rich sequence in the caspase-2 pre-mRNA (Fushimi

et al., 2008) and regulates alternative splicing of apoptosis-

related genes (Bonnal et al., 2008; Fushimi et al., 2008), modu-

lating the production of isoforms with antagonistic functions in

programmed cell death. It was also shown that RBM5 targets

late events in spliceosome assembly, acting as a switch for splice

site pairing in the Fas receptor (CD95/APO-1) pre-mRNA (Bonnal

Inc.

Contro

lRBM5 S

h

α-RBM10

α-RBM5

α-RBM6

α-Tubulin

Contro

lRBM6 S

h

Contro

lRBM10

Sh

Rel

ativ

e le

vels

of p

rote

ins

Contro

l

RBM5 Sh

RBM6 Sh

RBM10 S

h

Control

Num

ber

of c

olon

ies

95

110

130

A

B C

D E

F

0

400

800

600

200

RBM5 Sh RBM6 Sh RBM10 Sh

NLS

RS

RRM

RGG

ZnF

OCRE

G-Patch

RBM5

RBM6

RBM10

G

0

40

80

120

RBM5 Sh RBM6 Sh RBM10 Sh

RBM5 RBM6 RBM10

*****

***

0

1

2

3

4

RBM5 RBM6 RBM10

Exon

Intron

Promoter

Tail

IgG

KD

IgG

KD

+ UV - UV

RNASE1 RNASE1

- + ++ - + ++

95

110

130

95

110

130

RBM5

RBM6

RBM10

Fold

enr

iche

men

t

rela

tive

to s

eque

nce

leng

th

Low

er

expo

sure

Specific AB Specific AB

1 2 3 4 5 6 7 8 9 10 11

Figure 1. Regulation of Cell Proliferation by

RBM Proteins

(A) Schematic representation of the domain orga-

nization of the indicated RBM proteins and their

amino acid sequence identities.

(B) Western blot analysis of RBM protein knock-

down in stable HeLa cell lines expressing the

indicated Sh RNAs. a-tubulin was used as a

normalizing loading control.

(C) Quantification of the data in (B) using Image J

software.

(D) Results of triplicate clonogenic assays of cells

depleted of RBM5, 6, or 10, as indicated.

(E) Quantification of colony formation in the tripli-

cate experiments of (D). *p < 0.05, **p < 0.01, ***p <

0.001 (Student’s unpaired t test, n = 9).

(F) Autoradiography of 32P-labeled RNA cross-

linked to RBM5, 6, or RBM10, immunoprecipitated

from control of RBM-knocked-down (KD) HeLa

cells, either untreated or treated with different

concentrations of RNase 1, resolved in SDS-

polyacrylamide gels and transferred to a nitrocel-

lulose membrane. Rabbit IgGs were used as

controls.�UV indicates absence of ultraviolet light

irradiation.

(G) Enrichment of RBM CLIP-tags in the different

genic regions indicated, normalized to their rela-

tive sequence lengths. Quantification of results in

(C) and (E) are represented as the mean of n > 3

experiments ± SEM.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

et al., 2008). It has been reported that RBM5 and 10 are compo-

nents of prespliceosomal complexes (Behzadnia et al., 2006;

Deckert et al., 2006).

Alternative splicing of pre-mRNA precursors is a prevalent

mode of gene expression regulation in multicellular organisms

that involves the interplay between multiple sequence motifs

and cognate protein factors that favor or prevent the recognition

of splice sites by the splicing machinery (Barash et al., 2010;

Chen andManley, 2009;Wahl et al., 2009).Misregulation of alter-

native splicing contributes to cancer progression by affecting the

expression of genes or isoforms involved in cell-proliferation

control, apoptosis, DNA-damage response, energy metabolism,

Molecular Cell 52, 720–733, D

angiogenesis, and metastasis (David and

Manley, 2010; Kaida et al., 2012).

In this study, we show that depletion

of RBM5, 6, and RBM10 have distinct

effects on the ability of cancer cells

to proliferate. Using CLIP-Seq and

splicing-sensitive microarrays, we identi-

fied hundreds of exons, including many

implicated in tumor formation and pro-

gression, which are bound and/or whose

splicing is modulated by one or more of

these factors. We found that antagonistic

regulation of NUMB alternative splicing

by RBM5/6 and RBM10 can largely

account for the opposite effects of these

factors on cell proliferation, revealing a

key circuit of posttranscriptional gene

regulation that can be disrupted in cancer cells, as illustrated

by the effects of an RBM10mutant identified in lung adenocarci-

noma cells.

RESULTS

Effects of RBM5, 6, and 10 Depletion on Cell ColonyFormationHeLa cell lines stably expressing shRNAs against RBM5, 6, or 10

were generated, and the expression levels of those proteins

were reduced to 15%, 5%, and 3%, respectively, compared to

control cells (Figures 1B and 1C). Strikingly, despite the high level

ecember 12, 2013 ª2013 Elsevier Inc. 721

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

of sequence similarity between these proteins, depletion of

RBM5 and 6 caused a pronounced decrease in the clonogenic

capacity of the cells, whereas depletion of RBM10 resulted in

enhanced cell colony formation and noticeably denser colonies

(Figures 1D and 1E). Equivalent results were obtained with a sec-

ond set of shRNAs (Figures S1A and S1B). We conclude that

RBM5/6 and RBM10 have opposite effects on the capacity of

cancer cells to form colonies and proliferate.

Transcriptome-wide Identification of RBM ProteinBinding SitesTo identify direct RNA-binding sites and potential target RNAs

regulated by RBM proteins, CLIP-Seq experiments were carried

out for each of the proteins in replicates as described (Konig

et al., 2010). After irradiation of the cells with ultraviolet light,

RBM/RNAs complexes were immunoprecipitated with specific

antibodies—and complexes between 125–145 KDa, 130–150

KDa, and 140–160 KDa corresponding respectively to RBM5/

RNA, RBM6/RNA, and RBM10/RNA crosslinked species were

isolated after fractionation by SDS-polyacrylamide gel electro-

phoresis (Figure 1F). No complexes were detected using a

control antibody nor in cells in which the factors were knocked

down (Figures 1F, lanes 1 and 2). After purification, RNAs were

reverse-transcribed, PCR amplified with barcoded primers,

and sequenced using Solexa (Illumina) technology. This yielded,

respectively, 133.079, 1.614.330, and 1.307.119 clusters of

reads uniquely mapped to the genome, excluding repeat ele-

ments and excluding clusters overlapping with equivalent posi-

tions in the IgG control. Results from biological replicates

showed Spearman correlation coefficients of R2 = 0.54, 0.85,

and 0.82 for RBM5, 6, and 10, thus providing a robust set of spe-

cific RNA/RBM interactions (Figure S1C).

Cluster mapping to genic regions showed that RBM binding

sites were mainly associated with intronic regions (81%, 76%,

and 79% of total clusters for RBM5, 6, and 10, respectively).

When normalized to the general distribution of intron/exon

lengths in the genome, however, binding sites were significantly

enriched in exonic regions (Figure 1G). Clusters mapped to a to-

tal number of 20,387 protein-coding genes (Figure S1D). Gene

ontology analysis of the extensive overlap of 9,402 genes using

the software GORILLA (Eden et al., 2009) showed enrichment

in pathways involved in apoptosis, cell adhesion, actin/cytoskel-

eton reorganization, and signal transduction (Figure S1E). This

enrichment is consistent with and expands a previous study

identifying 35 genes involved in the control of cell proliferation

and apoptosis as targets of RBM5 regulation (Mourtada-Maara-

bouni et al., 2006).

To investigate the distribution of RBMs binding sites relative

to exon-intron junctions, CLIP clusters were mapped to those

boundaries after normalizing by the length of exons and introns.

The mapping revealed an enrichment of RBM5 and 6 binding

sites in exons, mostly within 50 nucleotides of the 30 or 50 splicesites, with more pronounced peaks near the 50 splice site (Fig-

ure 2A, top and middle panels). While RBM10 also showed

exonic binding sites near the 50 splice sites, two other prominent

peaks were detected within the 100 nucleotides region 50 of the30 splice sites and 30 of the 50 splice sites (Figure 2A, lower

panel).

722 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier

When reads were separately mapped to strong and weak

splice sites, a clear enrichment of exonic binding sites near

weak 50 splice sites compared to strong 50 splice sites was de-

tected (Figure 2B). These results suggest that RBM proteins

are general contributors to the splicing process but can particu-

larly contribute to the modulation of exons harboring weak 50

splice sites, without a clear enrichment in alternative over consti-

tutive exons (Figure S2B).

The distinct distribution of binding sites of RBM proteins could

be a consequence of their different RNA sequence specificities.

To determine in vivo target sequencemotifs of the RBMproteins,

enriched 5-mers present in CLIP-Seq clusters were identified.

RBM5, 6, and 10 clusters (73.7%, 61.8%, and 62.0%, respec-

tively) included the enriched 5-mers, respectively. The distribu-

tion of these motifs recapitulates the general cluster distribution

for these proteins (Figure 2C), indicating that these motifs are

indeed representative of the binding distribution of these factors.

Next, consensusmotifs were derived using the HOMER software

(Heinz et al., 2010). Two examples of the ten most significantly

enriched motifs for each RBM (Figure S2A) are shown in Fig-

ure 2D. For RBM5, UCAUC and AGUAA sequences were found

among the most significantly enriched motifs, which show simi-

larities with the previously published UC-rich RBM5 regulatory

motif (Fushimi et al., 2008) as well as with the ANGUAA motif

(Nguyen et al., 2011). RBM10 and 6 showed preference for the

CUCUGAA motif, which contains a core CUCU sequence remi-

niscent of PTB binding sites (Xue et al., 2009). Additional

U-rich motifs were also enriched in RBM10 clusters, consistent

with the in vitro observation that RBM10 interacts with poly-U

sequences (Inoue et al., 1996). This analysis therefore allowed

us to validate in a genome-wide manner motifs previously

described for individual genes and in addition expand the

sequence motif repertoires of these RBM proteins. Top

consensusmotifs were used for electrophoretic mobility retarda-

tion assays using N-terminal fragments of RBM5, 6, and 10

comprising the two RRMs and the first zinc finger domains, in

the presence of an excess of nonspecific RNA competitor.

RBM5 and 6 bound their top consensus motifs with apparent af-

finities of 50–70 Nm, and for RBM10 the apparent affinity was

150 nM (Figures 2E and S2C), confirming that these proteins

have significant intrinsic binding for the sequences identified

in vivo. Using the RNAcompete approach, Ray et al. recently re-

ported preferred sequence motifs for RBM5 and 6 (Ray et al.,

2013). RBM5 and 6 bound to these sequences with an affinity

slightly lower than to our CLIP top motifs (Figures 2E and S2C).

Single-nucleotide substitutions in the CLIP motifs led to 3- to

7-fold decreases in affinity, and binding to unrelated sequences

was undetectable (Figures 2E and S2C), indicating that RBM5, 6,

and 10 display sequence-specific RNA recognition.

Identification of Alternative Splicing Events Regulatedby RBMsTo determine the global effects of the three RBMs on the regula-

tion of alternative splicing and also to infer the functional rele-

vance of the identified RBMs/RNA interactions, RNAs isolated

from three biological replicas of each of the cell lines expressing

shRNAs against the individual RBMswere hybridized to splicing-

sensitive microarrays. Significant and consistent changes in

Inc.

cluster density at the 3’SS

RB

M5

clus

ters

den

sity

−400 −300 −200 −100 0

0.00

00.

0075

0.01

5

cluster density at the 5’SS

−50 100 200 300 400

0.00

00.

015

−400 −300 −200 −100 0

0.00

0.02

0.04

−50 100 200 300 400

0.00

0.02

0.04

−400 −300 −200 −100 0

0.00

0.03

−50 100 200 300 400

0.00

0.01

50.

03

RB

M6

clus

ters

den

sity

RB

M10

clu

ster

s de

nsity

RBM Control Random

intron intron

Strong splice site

−400 −300 −200 −100

0.00

0.05

0.1

−50 100 200 300 400

0.00

0.05

0.1

cluster density at the 3’SS cluster density at the 5’SS

intron intron

intron intron

0

intron

RB

M5

clus

ters

den

sity

intron intron

−400 −300 −200 −100 0

0.00

0.15

−50 100 200 300 400

0.00

0.15

RB

M6

clus

ters

den

sity

−50 100 200 300 400

0.00

0.10

RB

M10

clu

ster

s de

nsity

intron

−400 −300 −200 −100 0

0.00

0.05

0.10

intron

0.050.01

5

0.00

75

0.07

5

0.07

5

Weak splice site

intron

−400 −300 −200 −100 0

0.00

00.

015

−50 100 200 300 400

0.00

00.

0075

0.01

5

−400 −300 −200 −100 0

0.00

0.02

0.04

−50 100 200 300 400

0.00

0.02

0.04

−400 −300 −200 −100 0

0.00

0.03

−50 100 200 300 400

0.00

0.01

50.

03

0.01

50.

0075

intronintron

intronintron

intronintron

RBM Control Random

RB

M5

clus

ters

den

sity

RB

M6

clus

ters

den

sity

RB

M10

clu

ster

s de

nsity

clusters with motif at the 3’SS cluster with motif at the 5’SS

0E+00

2E+02

4E+02

6E+02

8E+02

-log2 P-value

RB

M5

RB

M6

RB

M10

Free RNA

RBM5/RNA

A B

C D

E

UCAUCGA AGUAACG GAAGGAA UCGUCGA AGCAACG CGCUGAA

CUCUGAA GAUCAGU AAUCCAG CGCUGAA UCGUCGA

CGAUCCC CUGUGGA CUAUCCC CCGUGGA

Free RNA

RBM6/RNA

Free RNA

RBM10/RNA

UCGUCGA

HOMER Motifs

RNACOMPETE Motifs

Mutant HOMER Motifs

- - - - - -

- - - - -

- - - - -

50 n

M Mn 001Mn 051

Negative ControlMotifs

Figure 2. Global Distribution of RBM-Bind-

ing Sites

(A) Relative enrichment of crosslinked RBM clus-

ters in exon/intron boundaries. Normalized cluster

densities for the different RBM proteins (y axis) are

represented relative to exons/introns positions (x

axis), with boundaries indicated by dashed vertical

lines. Red lines indicate RBM clusters, blue and

gray lines represent control and random clusters,

respectively.

(B) Relative enrichment of RBM clusters near

strong (dark) and weak (light) 30 and 50 splice sites

of exons.

(C) Distribution at exon/intron boundaries of clus-

ters containing significantly enriched 5-mers,

represented as in (A).

(D) Top consensus motifs identified by HOMER

among the CLIP tags for each of the RBMs. The

significance is represented as – Ln P value.

(E) Electrophoreticmobility shift assays to evaluate

binding of RBM proteins (amino-terminal regions

including the two RRMs and one zing finger, used

at concentrations of 50, 100, and 150 nM, to RNAs

that either contain one of the two top-sequence

motifs, a point mutation in the motif, a binding

sequence selected by the RNAcompete approach

(Ray et al., 2013) or an unrelated sequence as

control. The positions of protein/RNA complexes

and unbound RNA are indicated.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

alternative splicing were detected compared to control cell lines

(Figure S3A). Depletion of RBM5 affected expression of 281 tran-

scripts, 80% of which were downregulated, whereas depletion

of RBM6 and 10 affected expression of a larger set of 1202

and 1294 transcripts, respectively, more than 80% of which

were upregulated (Figure 3A). Interestingly, the overlap between

regulated targets was rather limited (Figure 3D), revealing

distinct regulatory functions for these highly similar factors. It

was previously shown that regulation of RBM5 expression

affects the transcript levels of 35 genes involved in the control

of cell proliferation and apoptosis (Mourtada-Maarabouni et al.,

2006). Enrichment analysis of KEGG pathways (Kanehisa et al.,

2012) in our data revealed, in addition to genes involved in

Molecular Cell 52, 720–733, D

apoptosis and cancer pathways, enrich-

ment in pathways linked to cardiac dis-

eases and hematopoietic cell lineages

(RBM5) or to neurodegenerative diseases

(RBM10) (Figure S3B), suggesting the

involvement of the regulatory functions

of these factors on a variety of biological

processes.

The microarray analysis also revealed

that depletion of these RBM proteins

has substantial effects on alterna-

tive splicing, most commonly affecting

cassette exons (Figure 3B). This observa-

tion is consistent with the presence of

RBM5 and 10 in spliceosomal complexes

and with known functions of RBM5 in Fas

and Caspase 2 alternative splicing (see

Introduction). Although a role for RBM6 in splicing had not

been reported, these data showed a comparable substantial

number of alternative splicing events affected by depletion of

RBM6 or 10 (Figures 3B and 3C). Similar genome-wide analyses

comparing the impact of hnRNP proteins on alternative splicing

revealed a wide repertoire of events regulated by the six mem-

bers of this family (Huelga et al., 2012), while other splicing reg-

ulators such as TDP-43 (Tollervey et al., 2011), HnRNPU (Xiao

et al., 2012), TIA1/TIAL1 (Wang et al., 2010), and MBNL (Du

et al., 2010) display a more restricted tissue/cell-type-specific

spectrum of regulatory activities, mostly on cassette exons, as

was the case for RBM proteins (Figure 3B). Using stringent

criteria, our analysis predicted that depletion of RBM5 and 10

ecember 12, 2013 ª2013 Elsevier Inc. 723

55

988 1143

226

214 151

0%

50%

100%

RBM5 RBM6 RBM10

down

up

Alternative Donor/Acceptor

Mutually Exclusive

Cassette Exon

158

837

904

29149

2450

RBM6

RBM10

RBM5 19

192

244

534

61

RBM6

RBM10

RBM5

Gene Expression events Alternative Splicing events

31 269 317

7 42 40

3 44 37

0%

50%

100%

RBM5 RBM6 RBM10 HJAY

1315

0 65

06

%A

S e

vent

s

% d

ereg

ulat

ed tr

ansc

ripts

14

27

102

8

65

71

0%

50%

100%

RBM5 RBM6 RBM10

skipping

Inclusion

A B

C D

E

BC

A3

Ex5

CC

NJL

Ex4

A

DA

D1

Ex2

HR

AS

IDX

GA

DD

45A

Ex2

TBC

1D3

Ex4

BID

Ex3

A

EIF

4E2

Ex2

A

LAS

S5

Ex1

F

NO

TCH

3 E

x16

APA

F1 E

x18

STA

T3 E

x23A

CLK

1 E

x6

MA

P4K

4 E

x21A

SR

P14

Ex3

A

EIF

4A2

Ex2

MA

DD

Ex1

6

Log2

fold

cha

nge

exon

ski

ppin

g %

Cas

sette

exo

ns

-5

-4

-3

-2

-1

0

1

2

3

4

5

RBM5

RBM6

RBM10

Figure 3. Alternative Splicing Regulation by RBMs

(A) Number and percentage of transcripts up- or downregulated more than 1.5-fold upon depletion of the indicated RBMs.

(B) Distribution of splicing changes upon depletion of the RBMs across categories of alternative splicing events. The distribution of the categories monitored by

the splicing-sensitive microarray (HJAY) is also indicated.

(C) Fraction of changes leading to skipping (red) or inclusion (blue) of high andmedium confidence cassette exon events differentially regulated upon depletion of

RBMs.

(D) Overlap between gene expression (left) or alternative splicing (right) changes observed upon depletion of RBM5, 6, and 10.

(E) Quantitative RT-PCR validation of RBM-regulated alternative splicing changes affecting cassette exons in genes involved in cell proliferation. Fold changes in

skipping, after normalization for gene expression using amplification products corresponding to flanking constitutive exons upon knockdown of RBM5 (blue), 6

(red), and 10 (green) are represented. Average values and standard deviation for three independent biological replicates are shown. The validation rates of

microarray predictions in these examples were 80% for RBM5 knockdown and 100% for 6 and 10 knockdown.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

724 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

induce exon inclusion in 63% and 59% of their target exons,

respectively, whereas RBM6 depletion promoted inclusion of

only 29% of its target cassette exons (Figure 3C). This is again

consistent with distinct mechanisms of splicing regulation by

these factors and indicates that RBM5 and 10 display mainly

repressive activities, consistent with findings for RBM5 in the

Fas and Caspase 2 systems (Bonnal et al., 2008; Fushimi

et al., 2008) and with a very recent report on RBM10 (Wang

et al., 2013), while RBM6 may act mainly as a splicing activator.

Consistent with this concept, mapping of CLIP clusters to

exons showing high or low inclusion levels in HeLa cells (quanti-

fied using ENCODE RNA-Seq data) showed an enrichment in

RBM6 cluster densities in highly included exons, while RBM5

and 10 showed no significant difference between these cate-

gories (Figure S2B).

As observed for transcriptional targets, the overlap between

alternatively spliced exons regulated by these factors was very

limited (Figure 3D). However, about 20% of the regulated events

are under the control of more than one RBM, suggesting

possible synergistic or antagonistic effects of RBM5, 6, and 10

on a subset of their targets. Validation by quantitative RT-PCR

of high- or medium-confidence microarray predictions for the

three factors was 80%–90%. Figure 3E shows the validation of

alternative splicing predictions—after normalizing by gene

expression changes—in genes previously shown to be involved

in cell proliferation, including oncogenes, apoptotic regulators,

MAP kinases, and members of the STAT and NOTCH signal

transduction pathways. A regulatory network depicting the

effects of RBM proteins on the expression and/or alternative

splicing of cancer-related and RNA processing-related genes

is shown in Figure S4A.

RNA Maps for RBM5, 6, and 10Overlap between CLIP and microarray data showed evidence of

RBM5 binding in the alternatively spliced genomic region

comprising the upstream and downstream introns as well as in

the flanking constitutive exons for 50% of the RBM5-regulated

exons. The figures reached 93% and 97% for RBM6 and 10,

respectively, a statistically highly significant enrichment

compared to the occurrence of RBM proteins binding in a

random subset of nonregulated exons. While the limited number

of RBM5-regulated events prevented us from drawing an RNA

map for this factor, RBM6 and 10 displayed distinct patterns of

binding relative to their regulated exons and the regulatory

outcome of their depletion (Figure 4A).

For RBM6, peaks of binding associated with exon skipping

were observed close to the constitutive upstream and down-

stream exons, suggesting that RBM6 promotes exon skipping

by enhancing the function of the distal splice sites. Other peaks

of binding associated with exon skipping or inclusion are

located about 200 nucleotides downstream of the 50 exon,

perhaps due to repression of regulatory sequences. Peaks

more specifically associated with alternative exon inclusion

are located within the 100 nucleotide intronic regions flanking

the regulated exon.

For RBM10, peaks of binding associated with exon skipping

were observed throughout the 50 intron and the 30 end of the

exon, while multiple peaks associated with exon inclusion were

Molec

observed in the 30 intron. The effects of RBM10 binding near

the 30 end of the exon resembles those of hnRNPA1 and could

act by similar mechanisms of 50 splice-site repression (Huelga

et al., 2012; Martinez-Contreras et al., 2006). The distribution

of repressing and activating RBM10 peaks in the upstream and

downstream introns, respectively, has resemblances with the

RNAmaps observed for other splicing regulatory factors (Licata-

losi and Darnell, 2010; Llorian et al., 2010; Yeo et al., 2009).

We conclude that, despite similarities in their binding specific-

ities (Figure 2D), RBM6 and 10 display distinct positional effects

in splicing regulation (Figure 4A). Synergistic/antagonistic effects

of RBM6 and 10 depending on the location of their binding sites

could however be detected (Figure S4B).

To evaluate the validity of these insights, we decided to focus

on an alternative event in the gene NUMB, which we found to be

regulated by RBM5, 6, and 10 (see below) and that is known to

be altered in lung cancer (Misquitta-Ali et al., 2011). A CLIP clus-

ter for RBM10 was found in the 30 splice-site region preceding

NUMB alternative exon 9. A modified version of the splicing

reporter RG6 minigene (Orengo et al., 2006) was generated re-

placing a model cassette exon by NUMB exon 9 and 100/50 nu-

cleotides of flanking intronic sequences (Figure 4B). This allowed

us to minimize the alternatively spliced region of NUMB and thus

focus on key potential regulatory elements (see below). Consis-

tent with the results of the RNA Map (Figure 4A), overexpression

of RBM10 in HeLa cells led to enhanced NUMB exon 9 exon

skipping in a cotransfection assay using the RG6-NUMB

minigene reporter (Figure 4C, lanes 1–4). Consistent with the

relevance of RBM10 binding 50 of the alternative exon, replace-

ment of the RBM10 binding site by another pyrimidine-rich re-

gion lacking consensus RBM10 motifs abolished the effects of

RBM10 overexpression (Figure 4C, lanes 5–8). These results

confirm the repressive function of RBM10 when bound 50 of analternative exon. Consistent with the effects of RBM6 depletion,

overexpression of this protein led to increased NUMB exon 9 in-

clusion (Figure 4C, lanes 9–12). In contrast, RBM5 overexpres-

sion did not change the NUMB exon 9 skipping/inclusion ratio,

presumably because of the lack of RBM5-responsive elements

in the minigene (Figure 4C, lanes 13–15).

Effects of RBM Proteins on Regulation of the NotchPathway in Breast and Lung Cancer Cell LinesTo investigate the regulation of alternative splicing by RBM5, 6,

and 10 in controlled cellular systems, cell lines derived from

normal breast and a breast tumor (MCF10A and MCF7) and

from normal fetal lung fibroblasts and a lung adenocarcinoma

(IMR9 and A549) were used to evaluate the relative expression

levels of RBM proteins and alternative splicing of their target

genes. Western blot analyses revealed significantly lower levels

of RBM5 and 6 in the cancer cell lines from the two tissues, while

the levelsofRBM10werehigher in thecancercell lines (Figure5A).

The changes in RBM protein expression were accompanied

by changes in alternative splicing of several genes involved in

cell proliferation, which were consistent in the two-cell-lines

comparisons (Figure 5B and 5C). These alternative splicing

events were found to be targets of RBM5, 6, and/or 10 in our

microarray analyses (Figure 3E) andwere also identified as direct

targets of at least one of them in our CLIP experiments.

ular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc. 725

0.00

0.05

0.100.15

0.00

0.050.10

0.15

0.00

0.05

0.100.15

0.00

0.050.10

0.15

n = 65

n = 27

n = 71

n = 102

RB

M6

Clu

ster

den

sitie

sR

BM

10 C

lust

er d

ensi

ties

A

200 b200 b200 b200 b

B

C

0

0.5

1

1.5

2

2.5

3

3.5

RBM10 WT

Rat

io S

kipp

ing/

Incl

usio

n

50bp100bpCMV p(A)

RG6-NUMB WT

RG6-NUMB minigene

RBM10 WT

RG6-NUMB WT RG6-NUMB mut

1 2 3 4 5 6 7 8

ATTGATGTTGTCTGCTCCCGTTGTAGCTAA

ATTGATTCTTTTGTCTCTCATCTTGTTGGTTCCAGCTAARG6-NUMB mut

RG6-NUMB WT

RBM6 WT RBM5 WT

9 10 11 12 13 14 15

NUMB ex9Isoforms

***

****

****

(legend on next page)

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

726 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

The differences in RBM5 and/or 6 expression between cancer

and noncancer cell lines shown in Figure 5A can contribute to

explain only three out of seven of the observed differences in

alternative splicing between the cancer cells, and the increase

in RBM10 expression cannot explain any of the observed

splicing changes (Table S1). In sharp contrast, the majority of

the changes in alternative splicing observed upon knockdown

of the RBMs that were validated in Figure 3E could help to

explain the effects of depletion of these factors on cell prolifera-

tion (Table S1). Given the recent report of frequent mutations of

RBM10 in lung adenocarcinomas (Imielinski et al., 2012), we

considered the possibility that RBM10 was mutated in the

A549 cell line. Indeed, sequencing of the RBM10 gene in A549

cells revealed a T-to-A substitution at position 1062 that leads

to replacement of valine 354 by glutamic acid, a mutation pre-

dicted to be located in alpha-helix 2 of the RRM2 domain

(Figure 5D).

Next, the effect of this mutation on the activity of RBM10 as a

regulator of NUMB exon 9 splicing was tested using the assay

described in Figure 4C. While overexpression of wild-type

RBM10 increased exon skipping (Figure 5E, lanes 1–6), titration

of the V354E mutant found in A549 cells compromised the

effects of the protein despite its higher level of expression and,

in fact, resulted in even higher exon inclusion at high levels

of RBM10 expression (Figure 5E, lanes 7–11). Similar effects

were observed in alternative splicing of the endogenous gene

and in accumulation of NUMB protein isoforms (Figure S5F).

These observations indicate that this mutation dramatically dis-

rupts RBM10 function. The reduced activity of the V354E mutant

could therefore explain (1) why the patterns of alternative splicing

observed in the A549 cell line compared to IMR9 cells are oppo-

site to what was inferred based upon the results of RBM10

knockdown (Figure 5C) and (2) why there are significantly lower

levels of the NUMB PRRS isoform in A549 cells (Figure 5C, lower

panel). Consistent with this interpretation, overexpression of

wild-type RBM10 in A549 cells was sufficient to increase skip-

ping of NUMB exon 9 from endogenous transcripts and induce

accumulation of the NUMB PRRS protein isoform (Figure 5F).

Furthermore, overexpression of wild-type RBM10 reduced the

clonogenic capacity of A549 cells (Figure 5G), consistent with

the idea that RBM10 mutation contributes to A549 cancer cell

proliferation.

Figure 4. Functional RNA Maps for RBMs and Regulation of NUMB Ex

A) RBM6 and 10 functional RNA Maps were generated adopting the method de

Experimental Procedures). The density of normalized cluster reads for each prot

(gray boxes) and flanking exonic (black boxes) and intronic (thin line) sequences

indicated by the blue and red lines, respectively. The profile of cluster densities a

The gray dashed line represents the distribution of clusters in a random subse

surrounding the alternative regulated exon (gray box).

(B) Schematic representation of the RG6-NUMBWT and mutant minigenes. Red

intronic regions, respectively. White boxes indicate constitutive exons of the mi

underlined, and two core consensus motifs are highlighted in red. This sequence

which is shown in blue italics.

(C) RBM10 overexpression induces skipping of NUMB exon 9 in transcripts of t

minigene. RNA was isolated from cells transfected with the RG6-NUMB minig

expression vector. Upper panel: semiquantitative analysis of exon inclusion/sk

between exon inclusion and skipping estimated using quantitative RT-PCR. *p <

experiments is represented as the mean of n > 3 experiments ± SEM.

Molec

NUMB Alternative Splicing and RBM-MediatedRegulation of Cell ProliferationWe decided to focus our efforts on NUMB exon 9 alternative

splicing regulation for several reasons. First, this event generates

mRNAs with opposite effects in the activation of the NOTCH

pathway: while exon 9 skipping generates a repressor of the

pathway and of cell proliferation, exon 9 inclusion correlates

with reduced NUMB protein levels and activation of the NOTCH

pathway (Misquitta-Ali et al., 2011; Westhoff et al., 2009).

Second, the NOTCH pathway is an important regulator of cell

proliferation, particularly in lung adenocarcinomas (Maraver

et al., 2012), and changes in NUMB exon 9 inclusion are among

themost frequent tumor-associated alternative splicing changes

observed in lung cancer (Misquitta-Ali et al., 2011). Third, the

event was regulated in opposite directions by RBM5/6 and

RBM10 (Figure 5C) and the direction of the changes is consistent

with the distinct effects that knockdown of these factors have in

cell proliferation assays (Figures 1D and 1E). Fourth, direct bind-

ing of RBM10 to this alternatively spliced region of NUMB was

detected in our CLIP experiments and plays a role in regulation

of NUMB exon 9 alternative splicing (Figure 4C). Fifth, differ-

ences in NUMB alternative splicing and accumulation of

NUMB alternatively spliced protein isoforms between the A549

cancer and noncancerous cell lines correlated with expression

of a RBM10 mutant that has reduced capacity to regulate

NUMB exon 9 skipping (Figures 5C and 5E). Sixth, overexpres-

sion of wild-type RBM10 in A549 cells was sufficient to increase

skipping of endogenous NUMB exon 9, induce accumulation of

the NUMB PRRS protein isoform, and decrease the clonogenic

capacity of these cells (Figures 5F and 5G). Given these prece-

dents, we decided to explore whether regulation of the NOTCH

pathway via NUMB alternative splicing can indeed explain the

effects of RBM5, 6, and 10 on cell proliferation.

We first tested whether knockdown of RBM5, 6, and 10 influ-

enced the activity of the NOTCH pathway. Figure S5A shows

that expression of the NOTCH target genes HEY1 and HEY2

(but not HES1) was affected by depletion of the RBM proteins.

To further confirm the involvement of the Notch pathway in the

activation of cell proliferation mediated by RBM10 depletion,

cells were treated with the small molecule peptidomimetic pre-

senilin inhibitor DFP-AA, which blocks Notch signaling and effec-

tively suppresses the growth of Notch1-transformed lymphoid

on 9 Alternative Splicing

scribed by Licatalosi et al. (2008) with minor modifications (see Supplemental

ein (y axis) is represented relative to the position of alternative cassette exons

(x axis). n indicates the number of exon inclusion/skipping events analyzed, as

ssociated with exon inclusion or skipping events follows the same color code.

t of 1000 nonregulated events. Black boxes represent the constitutive exons

box and lines represent NUMB exon 9 and 50/100 nucleotides of flanking 50/30

nigene, and gray boxes are vector sequences. The RBM10 binding cluster is

is replaced by a sequence from the 30 splice-site region of the following intron,

he RG6-NUMB WT minigene but not in transcripts of the RG6-NUMB mutant

enes and increasing concentrations (500 ng, 1 mg, and 2 mg) of an RBM10

ipping using primers corresponding to vector sequences. Lower panel: ratio

0.05, **p < 0.01, ***p < 0.001 (Student’s unpaired t test, n = 3). Quantification of

ular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc. 727

A B

C

E

A549

untreated

-5

-4

-3

-2

-1

0

1

2

3

4

NO

TCH

3 E

x16

BID

Ex3

a

HR

AS

IDX

MA

DD

Ex1

6

AP A

F E

x18

Breast

Lung

Log2

Fol

d ch

ange

exo

n sk

ippi

ngNUMB Ex9

MCF7

RBM10 S

h

Contro

l Sh

treated

D

F

α-RBM10

α-RBM5

α-RBM6

α-Tubulin

MCF7MCF10

A

IMR90

A549

Log2

Fol

d ch

ange

exo

n sk

ippi

ng

RBM5 Sh

RBM6 Sh

RBM10 Sh

Breast

Lung

-3

-2

-1

0

1

2

3

α-RBM10

α-Tubulin

RBM10 WT RBM10 V354E

α1 α2 β1 β2 β3 β3’ β4

RNP2 RNP1V E

RBM10-RRM2

0

0.5

1

1.5

2

2.5

3

3.5

Rat

io S

kipp

ing/

Incl

usio

n

1 2 3 4 5 6 7 8 9 10 11

(Can

cer V

s N

orm

al c

ell l

ine)

-

(DFP-AA)

0

200

400

600

800

Num

ber o

f col

onie

s

untreated

treated(DFP-AA)

*** ***

***

****

**

***

***

A549

MCF7

IMR90

MCF10A

RBM5 Sh

RBM10 S

h

Contro

l

RBM6 Sh

Numb-PRRL

Numb-PRRS

α-Tubulin

Numb-PRRL

Numb-PRRS

α-Tubulin

A549RBM10 WT - 100 500 1000

(ng)

H

-2.5

-2

-1.5

-1

-0.5

0

Log2

Rat

io N

UM

B E

x9 s

kipp

ing

A549RBM10 WT - 100 500 1000

(ng)

0

200

400

600

800

Num

ber o

f col

onie

s ***

HeLa

HeLa

A54

9

A54

9+

empt

y ve

ctor

A54

9+

RB

M10

WT

G

Figure 5. RBMs Expression and Function in

Breast and Lung Cancer Cell Lines

(A) Western blot analyses of expression of RBM5,

6, and 10 in the indicated cell lines (MCF10A,

normal breast; MCF7, breast cancer; IMR90,

normal fetal lung; A549, lung adenocarcinoma).

a-tubulin was used as a loading control.

(B) RT-qPCR analysis of differences in skipping

between cancer and normal cell lines (breast:

MCF7 versus MCF10A; lung: A549 versus IMR90)

for the exons indicated.

(C) RT-qPCR analysis of differences in NUMB

exon 9 skipping between cancer and normal cell

lines (as in B) and upon depletion of the RBM

proteins, as indicated. Lower panel: Western blot

analysis of NUMB protein isoforms of the same

samples. a-tubulin was used as loading control.

(D) Schematic representation of RBM10 RRM2

structural features and predicted location of the

V354E mutation identified in A549 cells.

(E) The V354E mutant fails to regulate NUMB

alternative splicing. RT-PCR analysis of NUMB

alternative splicing for transfection assays with

wild-type or V354E mutant RBM10 were carried

out as in Figure 4C using 100, 200, 500, and

2000 ng of expression plasmids.

(F) Rescue of NUMB splicing and NUMB PRRS

isoform expression upon expression of wild-type

RBM10 in A549 cells. The levels of endogenous

NUMB exon 9 skipping and NUMB protein iso-

forms upon transfection of the indicated amounts

of a RBM10-expression plasmid were analyzed

as in (C).

(G) Clonogenic assays under the same conditions

as in (F) were performed as in Figure 1E.

(H) Upper panel: colony formation assays using

A549, MCF7, and HeLa cells (control or RBM10-

depleted) upon treatment with the Notch pathway

inhibitor DFP-AA. Lower panel: quantification of

colony formation in the triplicate experiments. *p <

0.05, **p < 0.01, ***p < 0.001 (Student’s unpaired t

test, n = 3). Quantification of results in panels (B),

(C), and (F)–(H) are represented as the mean of

n = 3 (or >3) experiments ± SEM.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

cell lines in vitro (Weng et al., 2003). This treatment was sufficient

to suppress colony formation of MCF7 and A549 cell lines as well

as RBM10-depleted HeLa cells, whereas control knockdown

cells were only slightly affected (Figure 5G).

To further investigate the involvement of RBM proteins in

NUMB alternative splicing regulation and cell proliferation,

HeLa cell lines depleted of RBM5, 6, or 10 or A549 cells were

transfected with plasmids expressing cDNAs encoding one of

the two NUMB protein isoforms, NUMB-PRRS or NUMB-PRRL,

or empty vector as a control. Expression of NUMB PRRS (the

isoform generated upon exon 9 skipping) reduced the number

of colonies compared to control transfected cells, an effect com-

mon to A549, HeLa, andRBM10-depleted HeLa cells (Figure 6A).

These results indicate that expression of the NUMB isoform that

is reduced upon RBM10 depletion is sufficient to counteract the

effect of RBM10 depletion on cell proliferation. Conversely,

728 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier

expression of NUMB-PRRL (the isoform generated upon exon

9 inclusion) overcomes the reduction in colony formation caused

by depletion of RBM5 or 6 (Figure 6A). These results are consis-

tent with and expand previous studies showing that transient

overexpression of NUMB PRRL, but not NUMB-PRRS, promotes

cell proliferation via activation of NOTCH targets in A549 cell

lines (Misquitta-Ali et al., 2011). Furthermore, these data argue

that the effects on cell proliferation caused by depletion of

RBM proteins are, at least to a significant extent, due to altered

expression of NUMB isoforms. This conclusion is strengthened

by the observation that expression of the NUMB-PRRS isoform

did not further reduce colony formation in RBM5/6-depleted

cells (which have already a high proportion of this isoform), and

that expression of the NUMB-PRRL isoform did not further

enhance colony formation in RBM10-depleted cells (which

have already a high ratio of PRRL to PRRS) (Figure 6A).

Inc.

5’S

S-A

ON

3’S

S-A

ON

ISS

1-A

ON

ISS

2-A

ON

Exon 9 Exon 10 Exon 8

cagacgcgagggcaacaucGAUU 3’- -5’ (3’SS-AON)

(5’SS-AON)

(ISS1-AON)

(ISS2-AON)

UCGGguaagguggccacaga-5’3’-

uacuuaauugagguaacuacaa-5’3’-

uguaaguggacugacuguguc-5’3’-

3’SS-AON 5’SS-AON

ISS1-AON ISS2-AON

A549

MCF7

Control-AON

++

++

+

----

-

---

--

--

--

-

-

++---

+

+

-

--

+--

-+

++--

-+

+-

-

-

++-

--

++

++

-

----

----- no

n tu

mor

cel

ls

(M

CF1

0, IM

R90

)

A549 RBM10 sh RBM6 sh RBM5 sh 5’SS-AON ISS2-AON +

- +--

---

+-

--

-- +

---

Ctl

Numb-PRRL

Numb-PRRS

A B

C D

E

0

150

300

450

600

Control RBM5 Sh RBM6 Sh RBM10 Sh A549

Untransfected

Control AON

5'SS-AON

ISS2-AON

0

200

400

600

800

Control RBM5 Sh RBM6 Sh RBM10 Sh A549 Untransfected

pcDNA3.1 Empty

Numb-PRRS

Numb-PRRL

Num

ber o

f col

onie

s

Num

ber o

f col

onie

s

1 2 3 4 5 6 7 8 9 10 1112 13 141516 171819 20

0

0.5

1

1.5

Week 1 Week 2 Week 3 Week 4 Week 5

5’SS-AON ISS2-AON Control AON

Tum

or V

olum

e (c

m3)

1 2 3 4 5 6 7 8 9 10 1112 13 141516 171819 20

α-Tubulin

M

***

*** ***

*****

***

*** ***

***

***

**

******

Figure 6. NUMB Alternative Splicing and

RBM-Mediated Cell Proliferation

(A) NUMB isoforms can rescue the effects of RBM

depletion on cell proliferation. Quantification of

colony formation assays of HeLa cells (control or

knocked down for RBM5, 6, or 10) nontransfected,

transfected with a control plasmid, or transfected

with plasmids expressing NUMB-PRRS (exon 9

skipping) or NUMB-PRRL (exon inclusion). Similar

assays for A549 lung cancer cells are also included.

**p < 0.01, ***p < 0.001 (Student’s unpaired t test,

n = 6). Average number of colonies and standard

deviations for three independent biological repli-

cates of the experiment are represented.

(B) Schematic representation of NUMB exon 9

alternatively spliced region and sequences and

annealing sites of the modified antisense oligo-

nucleotides (AONs) complementary to the 50 or 30

splice sites or complementary to intronic splicing

silencer sequences (ISS).

(C)SemiquantitativeRT-PCR (upperpanel) analysis

of NUMB exon 9 alternative splicing upon trans-

fection of MCF7 breast cancer cells or A549 lung

cancer cells with the indicated AONs. Thepositions

of exon inclusion/skipping products are indicated.

Western blot analysis (lower panel) of NUMB pro-

tein isoforms (indicated) observed upon knock-

down of RBMs and transfection of AONs. a-tubulin

was used as a normalizing loading control.

(D) NUMB alternative splicing controls tumor cell

proliferation. Quantification of colony formation

assay of HeLa control or RBM knock down cells or

A549 cells upon transfection with the indicated

AONs. Average number of colonies and standard

deviations for three independent biological repli-

cates of the experiment are represented as in (A).

(E) NUMB alternative splicing controls xenograft

tumor growth. Xenograft tumor formation assays

were carried out by injection of A549 cell lines

transfected with control AON (blue diamonds),

ISS2-AON (red squares), or 50SS-AON (green

rectangles). X and y axis represent weeks after

injection and tumor volume, respectively. Error

bars represent standard deviation from at least

three different animals for each condition.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

Protein overexpression may lead to nonphysiological effects

due to an excess of the isoforms. To overcome these issues,

modified antisense RNA oligonucleotides (AON) were designed

to modulate the relative levels of the isoforms generated from

endogenous transcripts. We used oligonucleotides complemen-

tary to the 50 or the 30 splice sites (30SS-AON, 50SS-AON) to

promote NUMB exon 9 skipping and oligonucleotides comple-

mentary to an intronic splicing silencer (E.G.B., unpublished

data) (ISS1-AON, ISS2-AON) to promote exon 9 inclusion. The

sequences and annealing sites of those AONs are indicated in

Figure 6B. MCF7 and A549 cells were transfected with either a

single AONor a combination of two ormore AONs and the result-

ing effects on endogenous NUMBexon 9 splicing were assessed

by semiquantitative RT-PCR. The 50SS-AON oligo caused effi-

cient exon skipping, while the ISS2-AON oligo enhanced exon

9 inclusion (Figure 6C, upper panel). Intermediate ratios of inclu-

sion/skipping were achieved with mixtures of these two oligos.

Molec

The AONs did not affect the patterns of four different alternative

splicing events regulated by RBM10 (data not shown). Changes

in isoform ratios were also detectable at the protein level (Fig-

ure 6C, lower panel): consistent with NUMB exon 9 inclusion

inducing hyperubiquitination and destabilization of the NUMB

protein, which leads to lower overall levels of NUMB protein in

lung adenocarcinoma patient samples (Misquitta-Ali et al.,

2011), and with the protein analyses of Figures 5C and 5F, induc-

tion of exon inclusion by the ISS2-AON oligo led to overall

reduced levels of NUMB, while induction of exon skipping by

the 50SS-AON led to higher total levels of NUMB in A549 cells

and in HeLa cells depleted of RBM10 (Figure 6C, lower panel).

The switch in NUMB exon 9 splicing was accompanied by a

marked alteration of the expression of NOTCH targets (Figures

S5B–S5D). To evaluate whether splicing modulation of endoge-

nous transcripts by AONs results in changes in the proliferation

capacity of cells, clonogenic assays were carried out in A549

ular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc. 729

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

and HeLa cells combining the use of these reagents with knock-

down of each of the RBM proteins. Induction of exon 9 skipping

by the 50SS-AON oligo caused a strong reduction in the number

of colonies compared to a control scrambled AON, counteract-

ing the increase in colonies induced by knockdown of RBM10

(Figure 6D; lanes 3, 7, 11, 15, and 19). Therefore, the proliferative

effects of RBM10 depletion, which causes increased exon inclu-

sion, can be antagonized by increasing the ratio of inclusion/

skipping in transcripts from the endogenous NUMB gene.

Conversely, induction of exon 9 inclusion by the ISS2-AON oligo

caused a very significant increase in colony formation, counter-

acting the decrease in colonies induced by knockdown of either

RBM5 or 6 (Figure 6D; lanes 4, 8, and 12). Thus, the antiprolifer-

ative effects of RBM5/6 depletion, which causes exon skipping,

can be antagonized by increasing the ratio of inclusion/skipping

in transcripts from the endogenous NUMB gene. It is worth

pointing out that the ISS2-AON oligo did not further induce

colony formation in RBM10 knockdown cells because RBM10

depletion already led to high ratios of exon 9 inclusion (compare

lanes 13, 14, and 16 with 17, 18, and 20).

To further investigate the effects of the 50SS-AON and ISS2-

AON on cell proliferation in vivo, A549 cells were transfected

with either of the AONs and injected subcutaneously to

CB17SC-M nude mice. Simultaneous left and right injections

were performed and four nude mice/group were used. During

the first 2 weeks after injection, cells transfected with ISS2-

AON formed tumors of bigger volume compared with cells trans-

fected with the control AON (Figure 6E). This difference in size

was attenuated after 2 weeks of injection, most likely because

of titration and/or decay of the AON. In contrast, cells transfected

with 50SS-AON formed tumors only after 3 weeks of injection,

and these tumors remained smaller compared to the control,

even after 5 weeks of injection.

DISCUSSION

Our results show that three highly related RBM proteins have

overlapping, but also clearly distinguishable sets of RNA targets

and even display antagonistic properties in the regulation of cell

proliferation. Collectively, our results indicate that the ratio be-

tween isoforms of the Notch regulator NUMB is rate limiting for

the proliferation of cancer cells and that the regulatory effects

of RBM5, 6, and 10 on cell division can be recapitulated, in large

part by their effects on the regulation of NUMB alternative

splicing.

RNA Binding by RBM ProteinsThe CLIP-Seq experiments revealed that each of the RBMs

shows affinity for several sequence motifs, possibly due to the

presence of different RNA-binding domains (two RRMs, two

zinc fingers) in these proteins. For example, the prevalent

AGUAA motif identified for RBM5 resembles the ANGUAA motif

found to interact with one of its zinc finger domains in vitro

(Nguyen et al., 2011), while UC-rich motifs identified with high

confidence in our analyses are similar to a motif found important

for RBM5-mediated alternative splicing regulation of Caspase 2

(Fushimi et al., 2008) and to sequences used for structure deter-

mination of RBM5 RRM2 bound to RNA (Song et al., 2012). The

730 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier

latter study also documented interaction of RBM5 RRM2 with

GA-rich sequences, which are also present in several of the

top motifs identified by our analyses.

The identity of the enriched motifs and the degree of enrich-

ment are most similar between RBM6 and 10 (Figure S2). These

factors, however, often have different effects on alternative

splicing of common targets (e.g., Figure 3E) and distinct regula-

tory maps (Figure 4A). These observations argue that the binding

and activity of these factors is likely to be influenced by other

proteins as well as by their relative positions in the pre-mRNA

(Figure S4B).

The distribution of RBM5/6 and RBM10 binding sites relative

to exon/intron boundaries is also different (Figure 2). The enrich-

ment of RBM5/6 in exonic regions resembles that of other

splicing factors like SRSF3, SRSF4, and FOX2, opening the pos-

sibility that, like SR proteins, RBM5 and RBM6 binding in exonic

regions near splice sites helps to recruit splicing components

(Anko et al., 2012; Yeo et al., 2009). Indeed, previous work has

documented that RBM5 OCRE domain interacts with compo-

nents of spliceosomal snRNPs and other splicing factors (Bonnal

et al., 2008), although in this case the interaction was associated

with exon repression. In contrast, RBM10 binding to intronic re-

gions near splice sites (Figure 2) may regulate alternative splicing

by interfering with splice-site recognition, similar to the positional

effects that have been proposed for FET family proteins (Hoell

et al., 2011) and hnRNP C (Konig et al., 2010). Direct blockage

of the splicing factor U2AF, which recognizes the polypyrimidine

tracts associated to 30 splice sites, is a likely mechanism by

which RBM10 promotes skipping of NUMB exon 9 (Figure 4C),

which has been reported for other regulators, as well (Singh

et al., 1995; Valcarcel et al., 1993). It is interesting to consider,

however, that the mutation identified in RBM10 RRM2 in A549

cells, which abolishes RBM10-mediated NUMB regulation, is

unlikelyto affect RNA recognition and may block RBM10 activity

by failing to synergize with a corepressor. More generally, the

RNA Map of RBM10 resembles that of other splicing factors

(Licatalosi and Darnell, 2010; Llorian et al., 2010; Yeo et al.,

2009), with skipping/inclusion correlating with binding 50/30 ofthe alternative exon, respectively.

RBMs and Cell ProliferationMisregulation of alternative splicing contributes to cancer pro-

gression (David and Manley, 2010; Kaida et al., 2012; Pal et al.,

2012), and classical splicing regulators like SRSF1 display strong

oncogenic activity (Karni et al., 2007).Many of the key events and

regulatory factors involved in splicing-based decisions impact-

ing on cell growth and cancer progression, however, remain to

be discovered. A number of alternative splicing changes regu-

lated by RBM5/6 and RBM10 have the potential to contribute

to the antagonistic effects of these factors on cell proliferation

(Table S1), which is also consistent with the identification of mul-

tiple genes involved in apoptosis and cell proliferation as targets

of RBM5 (Mourtada-Maarabouni et al., 2006). This is reminiscent

of the effects of other splicing regulators like NOVA or ESRP pro-

teins, which orchestrate multiple splice site switches that can

contribute to complex cellular functions, like shaping synaptic

connections (Ule et al., 2005) or guiding epithelial-mesenchymal

transitions (Warzecha et al., 2010). Our results, however, also

Inc.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

reveal that particular alternative splicing events—in our case, in

exon 9 of the NUMB gene—can make a predominant contribu-

tion to the cell proliferation phenotype controlled by RBM5, 6,

and 10.

The gene NUMB encodes a natural inhibitor of the NOTCH

pathway (Pece et al., 2010), a pathway that is hyperactivated

in �40% of all human lung adenocarcinomas and correlates

with poor prognosis (Dang et al., 2000; Maraver et al., 2012;

Westhoff et al., 2009). NUMB promotes NOTCH ubiquitination

and proteosomal degradation of NOTCH intracellular domain

(NICD) and may also control NOTCH levels by modulating its

endocytic recycling (McGill et al., 2009; McGill and McGlade,

2003). Inclusion/skipping of exon 9 leads to protein variants

that differ in a proline-rich domain involved in protein-protein

interactions, and these isoforms display distinct activities, one

promoting cell proliferation while the other promotes cell differ-

entiation (Toriya et al., 2006; Verdi et al., 1999). Our own results

using modified oligonucleotides able to modulate the ratio

between NUMB isoforms reveal a key role for NUMB alternative

splicing regulation in cancer cell proliferation (Figure 6). Of rele-

vance, increased levels of NUMB exon 9 inclusion are frequent

in nonsmall lung cancer tumors, correlating with higher NOTCH

activity (Misquitta-Ali et al., 2011). Inhibition of the NOTCH

pathway is emerging as a promising new approach for cancer

therapy (Purow, 2012). In particular, pharmacological inhibition

of the g-secretase complex, which prevents proteolytic cleavage

of the NOTCH receptor and activation of the pathway, blocks

cancer growth in a mouse model of lung adenocarcinomas

(Maraver et al., 2012). It is therefore conceivable that modulation

of NUMB alternative splicing (e.g., using modified oligonucleo-

tides like those of Figure 6) can have therapeutic effects.

Approaches for the modulation of alternative splicing with thera-

peutic potential based upon detailed knowledge of regulatory

mechanisms are indeed emerging, including the use of modified

antisense oligonucleotides that promote inclusion of exon 7 in

the SMN2 gene, thus restoring SMN function in cells and mouse

models of spinal muscular atrophy (Hua et al., 2010, 2011; Hua

et al., 2011; Hua et al., 2008).

Our observations become particularly relevant after a recent

report of massively parallel genome sequencing of lung adeno-

carcinomas revealed that RBM10 is statistically recurrently

mutated in these tumors (12/183, 7%) (Imielinski et al., 2012).

Together with our results implicating RBM10 in lung cancer cell

proliferation and in NUMB alternative splicing regulation, these

data suggest that mutations in RBM10 contribute to lung adeno-

carcinoma progression by causing deregulation of NUMB alter-

native splicing, as we found to be the case for amutant version of

RBM10 found in the lung cancer cell line A549 and for the rescue

of NUMB splicing and decrease in cell proliferation upon expres-

sion of wild-type RBM10 in these cells (Figure 5).

Of interest, mutations in RBM10 have been also implicated in

cleft palate and TARP syndrome (Gripp et al., 2011; Johnston

et al., 2010). These developmental defects in cell proliferation/

differentiation may be influenced by alterations in the regulation

of alternative splicing targets of RBM10, as argued by very

recent results showing that an in-frame deletion in RBM10 iden-

tified in a TARP syndrome patient reduces the accumulation of

the protein in the nucleus, affecting its activity as splicing regu-

Molec

lator (Wang et al., 2013). Conceivably, NUMB and regulation of

the NOTCH pathway, which plays key roles in cell-fate determi-

nation during the development of multiple tissues, could be

among the targets affected by altered RBM10 function in these

syndromes.

EXPERIMENTAL PROCEDURES

Cell Culture Experiments

Conditions for cell culture, Sh-mediated knockdown antisense 20-O-methyl

phosphorothioate oligonucletide, and plasmid transfections are described in

the Supplemental Experimental Procedures.

Clonogenic assays were performed in triplicate using 2000 cells that were

plated in 10 cm cell culture dishes with fresh culture medium, replaced every

3 days. After 15 days, cells were washed with PBS and fixed with 5 ml of meth-

anol for 10 min at room temperature. Cells were stained with 5% Giemsa

(Sigma, Aldrich) overnight, washed, dried, and counted. Experiments were

performed in triplicates.

In Vivo and In Vitro Protein/RNA Interactions

CLIP experiments were performed following the protocol of Licatalosi et al.,

(2008) with slight modifications, which, together with the detailed computa-

tional analyses, can be found in the Supplemental Experimental Procedures.

Electrophoretic mobility shift assays were carried out using 10 fmol of radiola-

beled probes incubated on ice for 15 min with His-RBMNter in presence of

200 ng/ml yeast tRNA. Mixtures were then loaded on native 5%polyacrylamide

gels, and migration was carried out in 0.5%TBE at 150V for 2 hr, followed by

gel drying and Phosphorimager analyses.

RNA and Protein Analyses

RNA from three biological replicates of sh-RBM-infected HeLa cells was puri-

fied and analyzed using Affymetrix Human Exon Junction Arrays and validated

by RT-qPCR as described in the Supplemental Experimental Procedures.

Protein extraction and immunoblot analysis were performed as previously

described (Paronetto et al., 2011).

Xenograft Tumor Formation

A549 cells were transfected with AONs. After 48 hr, 106 cells were resus-

pended in 1 ml of Basement Membrane Matrix, Growth Factor Reduced

(GFR) (BD, Biosciences). Cells were inoculated subcutaneously in 6-week-

old CB17SC-M nude mice. Tumor formation was followed and weekly

measured until 5 weeks after injection. Animal work procedures were

approved by an institutional review board.

ACCESSION NUMBERS

Microarray data have been submitted in MIAME-compliant form to GEO data-

base (GSE47431) and can be accessed at http://regulatorygenomics.upf.edu/

Data/RBMs/.

SUPPLEMENTAL INFORMATION

Supplemental Information includes six figures, three tables, Supplemental

Experimental Procedures, and Supplemental References and can be found

with this article online at http://dx.doi.org/10.1016/j.molcel.2013.11.010.

ACKNOWLEDGMENTS

We thank Professor Benjamin Blencowe for NUMB isoform constructs and

Professor Francisco Real and all the members of Valcarcel’s lab, especially

Dr. Maria Paola Paronetto, for discussions and critical reading of the manu-

script. We are grateful to Dr. Francesca Rapino for her help with lentivirus gen-

eration, Dr. Camilla Iannone for her graphical expertise, and Anna Ribo for her

technical assistance. E.B. was supported by a Marie Curie postdoctoral

fellowship (FP7-PEOPLE-2009-IEF). Work in E.E.’s group is supported by

ular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc. 731

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

Consolider RNAREG, BIO2011-23920, and FSI2011-035 from Sandra Ibarra

Foundation. Work in J.V.’s lab is supported by Fundacion Botın, AICR, Funda-

cion Alicia Koplowitz, Consolider RNAREG, and Ministerio de Economıa y

Competitividad.

Received: June 5, 2013

Revised: September 18, 2013

Accepted: October 22, 2013

Published: December 12, 2013

REFERENCES

Angeloni, D. (2007). Molecular analysis of deletions in human chromosome

3p21 and the role of resident cancer genes in disease. Brief. Funct.

Genomics Proteomics 6, 19–39.

Anko, M.L., Muller-McNicoll, M., Brandl, H., Curk, T., Gorup, C., Henry, I., Ule,

J., and Neugebauer, K.M. (2012). The RNA-binding landscapes of two SR pro-

teins reveal unique functions and binding to diverse RNA classes. Genome

Biol. 13, R17.

Barash, Y., Calarco, J.A., Gao,W., Pan, Q.,Wang, X., Shai, O., Blencowe, B.J.,

and Frey, B.J. (2010). Deciphering the splicing code. Nature 465, 53–59.

Behzadnia, N., Hartmuth, K., Will, C.L., and Luhrmann, R. (2006). Functional

spliceosomal A complexes can be assembled in vitro in the absence of a

penta-snRNP. RNA 12, 1738–1746.

Bonnal, S., Martınez, C., Forch, P., Bachi, A., Wilm, M., and Valcarcel, J.

(2008). RBM5/Luca-15/H37 regulates Fas alternative splice site pairing after

exon definition. Mol. Cell 32, 81–95.

Chen, M., and Manley, J.L. (2009). Mechanisms of alternative splicing regula-

tion: insights from molecular and genomics approaches. Nat. Rev. Mol. Cell

Biol. 10, 741–754.

Dang, T.P., Gazdar, A.F., Virmani, A.K., Sepetavec, T., Hande, K.R., Minna,

J.D., Roberts, J.R., and Carbone, D.P. (2000). Chromosome 19 translocation,

overexpression of Notch3, and human lung cancer. J. Natl. Cancer Inst. 92,

1355–1357.

David, C.J., and Manley, J.L. (2010). Alternative pre-mRNA splicing regulation

in cancer: pathways and programs unhinged. Genes Dev. 24, 2343–2364.

Deckert, J., Hartmuth, K., Boehringer, D., Behzadnia, N.,Will, C.L., Kastner, B.,

Stark, H., Urlaub, H., and Luhrmann, R. (2006). Protein composition and elec-

tron microscopy structure of affinity-purified human spliceosomal B com-

plexes isolated under physiological conditions. Mol. Cell. Biol. 26, 5528–5543.

Du, H., Cline, M.S., Osborne, R.J., Tuttle, D.L., Clark, T.A., Donohue, J.P., Hall,

M.P., Shiue, L., Swanson, M.S., Thornton, C.A., and Ares, M., Jr. (2010).

Aberrant alternative splicing and extracellular matrix gene expression in

mouse models of myotonic dystrophy. Nat. Struct. Mol. Biol. 17, 187–193.

Edamatsu, H., Kaziro, Y., and Itoh, H. (2000). LUCA15, a putative tumour sup-

pressor gene encoding an RNA-binding nuclear protein, is down-regulated in

ras-transformed Rat-1 cells. Genes Cells 5, 849–858.

Eden, E., Navon, R., Steinfeld, I., Lipson, D., and Yakhini, Z. (2009). GOrilla: a

tool for discovery and visualization of enriched GO terms in ranked gene lists.

BMC Bioinformatics 10, 48.

Fushimi, K., Ray, P., Kar, A., Wang, L., Sutherland, L.C., and Wu, J.Y. (2008).

Up-regulation of the proapoptotic caspase 2 splicing isoform by a candidate

tumor suppressor, RBM5. Proc. Natl. Acad. Sci. USA 105, 15708–15713.

Gripp, K.W., Hopkins, E., Johnston, J.J., Krause, C., Dobyns, W.B., and

Biesecker, L.G. (2011). Long-term survival in TARP syndrome and confirma-

tion of RBM10 as the disease-causing gene. Am. J. Med. Genet. A. 155A,

2516–2520.

Heinz, S., Benner, C., Spann, N., Bertolino, E., Lin, Y.C., Laslo, P., Cheng, J.X.,

Murre, C., Singh, H., and Glass, C.K. (2010). Simple combinations of lineage-

determining transcription factors prime cis-regulatory elements required for

macrophage and B cell identities. Mol. Cell 38, 576–589.

Hoell, J.I., Larsson, E., Runge, S., Nusbaum, J.D., Duggimpudi, S., Farazi, T.A.,

Hafner, M., Borkhardt, A., Sander, C., and Tuschl, T. (2011). RNA targets of

732 Molecular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier

wild-type and mutant FET family proteins. Nat. Struct. Mol. Biol. 18, 1428–

1431.

Hua, Y., Sahashi, K., Hung, G., Rigo, F., Passini, M.A., Bennett, C.F., and

Krainer, A.R. (2010). Antisense correction of SMN2 splicing in the CNS rescues

necrosis in a type III SMA mouse model. Genes Dev. 24, 1634–1644.

Hua, Y., Vickers, T.A., Okunola, H.L., Bennett, C.F., and Krainer, A.R. (2008).

Antisense masking of an hnRNP A1/A2 intronic splicing silencer corrects

SMN2 splicing in transgenic mice. Am. J. Hum. Genet. 82, 834–848.

Hua, Y., Sahashi, K., Rigo, F., Hung, G., Horev, G., Bennett, C.F., and Krainer,

A.R. (2011). Peripheral SMN restoration is essential for long-term rescue of a

severe spinal muscular atrophy mouse model. Nature 478, 123–126.

Huelga, S.C., Vu, A.Q., Arnold, J.D., Liang, T.Y., Liu, P.P., Yan, B.Y., Donohue,

J.P., Shiue, L., Hoon, S., Brenner, S., et al. (2012). Integrative genome-wide

analysis reveals cooperative regulation of alternative splicing by hnRNP pro-

teins. Cell Rep. 1, 167–178.

Imielinski, M., Berger, A.H., Hammerman, P.S., Hernandez, B., Pugh, T.J.,

Hodis, E., Cho, J., Suh, J., Capelletti, M., Sivachenko, A., et al. (2012).

Mapping the hallmarks of lung adenocarcinoma with massively parallel

sequencing. Cell 150, 1107–1120.

Inoue, A., Takahashi, K.P., Kimura, M., Watanabe, T., andMorisawa, S. (1996).

Molecular cloning of a RNA binding protein, S1-1. Nucleic Acids Res. 24,

2990–2997.

Johnston, J.J., Teer, J.K., Cherukuri, P.F., Hansen, N.F., Loftus, S.K., Chong,

K., Mullikin, J.C., and Biesecker, L.G.; NIH Intramural Sequencing Center

(NISC) (2010). Massively parallel sequencing of exons on the X chromosome

identifies RBM10 as the gene that causes a syndromic form of cleft palate.

Am. J. Hum. Genet. 86, 743–748.

Kaida, D., Schneider-Poetsch, T., and Yoshida, M. (2012). Splicing in onco-

genesis and tumor suppression. Cancer Sci. 103, 1611–1616.

Kanehisa, M., Goto, S., Sato, Y., Furumichi, M., and Tanabe, M. (2012). KEGG

for integration and interpretation of large-scale molecular data sets. Nucleic

Acids Res. 40 (Database issue), D109–D114.

Karni, R., de Stanchina, E., Lowe, S.W., Sinha, R., Mu, D., and Krainer, A.R.

(2007). The gene encoding the splicing factor SF2/ASF is a proto-oncogene.

Nat. Struct. Mol. Biol. 14, 185–193.

Konig, J., Zarnack, K., Rot, G., Curk, T., Kayikci, M., Zupan, B., Turner, D.J.,

Luscombe, N.M., and Ule, J. (2010). iCLIP reveals the function of hnRNP par-

ticles in splicing at individual nucleotide resolution. Nat. Struct. Mol. Biol. 17,

909–915.

Licatalosi, D.D., and Darnell, R.B. (2010). RNA processing and its regulation:

global insights into biological networks. Nat. Rev. Genet. 11, 75–87.

Licatalosi, D.D., Mele, A., Fak, J.J., Ule, J., Kayikci, M., Chi, S.W., Clark, T.A.,

Schweitzer, A.C., Blume, J.E., Wang, X., et al. (2008). HITS-CLIP yields

genome-wide insights into brain alternative RNA processing. Nature 456,

464–469.

Llorian, M., Schwartz, S., Clark, T.A., Hollander, D., Tan, L.Y., Spellman, R.,

Gordon, A., Schweitzer, A.C., de la Grange, P., Ast, G., and Smith, C.W.

(2010). Position-dependent alternative splicing activity revealed by global

profiling of alternative splicing events regulated by PTB. Nat. Struct. Mol.

Biol. 17, 1114–1123.

Maraver, A., Fernandez-Marcos, P.J., Herranz, D., Canamero, M., Munoz-

Martin, M., Gomez-Lopez, G., Mulero, F., Megıas, D., Sanchez-Carbayo, M.,

Shen, J., et al. (2012). Therapeutic effect of g-secretase inhibition in

KrasG12V-driven non-small cell lung carcinoma by derepression of DUSP1

and inhibition of ERK. Cancer Cell 22, 222–234.

Martinez-Contreras, R., Fisette, J.F., Nasim, F.U., Madden, R., Cordeau, M.,

and Chabot, B. (2006). Intronic binding sites for hnRNP A/B and hnRNP F/H

proteins stimulate pre-mRNA splicing. PLoS Biol. 4, e21.

McGill, M.A., Dho, S.E., Weinmaster, G., and McGlade, C.J. (2009). Numb

regulates post-endocytic trafficking and degradation of Notch1. J. Biol.

Chem. 284, 26427–26438.

Inc.

Molecular Cell

NUMB Splicing Regulation by RBM Proteins

McGill, M.A., and McGlade, C.J. (2003). Mammalian numb proteins promote

Notch1 receptor ubiquitination and degradation of the Notch1 intracellular

domain. J. Biol. Chem. 278, 23196–23203.

Misquitta-Ali, C.M., Cheng, E., O’Hanlon, D., Liu, N., McGlade, C.J., Tsao,

M.S., and Blencowe, B.J. (2011). Global profiling and molecular characteriza-

tion of alternative splicing events misregulated in lung cancer. Mol. Cell. Biol.

31, 138–150.

Mourtada-Maarabouni, M., Keen, J., Clark, J., Cooper, C.S., and Williams,

G.T. (2006). Candidate tumor suppressor LUCA-15/RBM5/H37 modulates

expression of apoptosis and cell cycle genes. Exp. Cell Res. 312, 1745–1752.

Mourtada-Maarabouni, M., Sutherland, L.C., Meredith, J.M., and Williams,

G.T. (2003). Simultaneous acceleration of the cell cycle and suppression of

apoptosis by splice variant delta-6 of the candidate tumour suppressor

LUCA-15/RBM5. Genes Cells 8, 109–119.

Mourtada-Maarabouni, M., Sutherland, L.C., and Williams, G.T. (2002).

Candidate tumour suppressor LUCA-15 can regulate multiple apoptotic path-

ways. Apoptosis 7, 421–432.

Nguyen, C.D., Mansfield, R.E., Leung, W., Vaz, P.M., Loughlin, F.E., Grant,

R.P., and Mackay, J.P. (2011). Characterization of a family of RanBP2-type

zinc fingers that can recognize single-stranded RNA. J. Mol. Biol. 407,

273–283.

Oh, J.J., Grosshans, D.R., Wong, S.G., and Slamon, D.J. (1999). Identification

of differentially expressed genes associated with HER-2/neu overexpression

in human breast cancer cells. Nucleic Acids Res. 27, 4008–4017.

Oh, J.J., Razfar, A., Delgado, I., Reed, R.A., Malkina, A., Boctor, B., and

Slamon, D.J. (2006). 3p21.3 tumor suppressor gene H37/Luca15/RBM5

inhibits growth of human lung cancer cells through cell cycle arrest and

apoptosis. Cancer Res. 66, 3419–3427.

Oh, J.J., West, A.R., Fishbein, M.C., and Slamon, D.J. (2002). A candidate

tumor suppressor gene, H37, from the human lung cancer tumor suppressor

locus 3p21.3. Cancer Res. 62, 3207–3213.

Orengo, J.P., Bundman, D., and Cooper, T.A. (2006). A bichromatic fluores-

cent reporter for cell-based screens of alternative splicing. Nucleic Acids

Res. 34, e148.

Pal, S., Gupta, R., and Davuluri, R.V. (2012). Alternative transcription and

splicing in cancer. Pharmacol. Ther. 136, 283–294.

Paronetto, M.P., Minana, B., and Valcarcel, J. (2011). The Ewing sarcoma

protein regulates DNA damage-induced alternative splicing. Mol. Cell 43,

353–368.

Pece, S., Confalonieri, S., Romano, P.R., and Di Fiore, P.P. (2010). NUMB-ing

down cancer by more than just a NOTCH. Biochim. Biophys. Acta. 1815,

26–43.

Purow, B. (2012). Notch inhibition as a promising new approach to cancer

therapy. Adv. Exp. Med. Biol. 727, 305–319.

Ramaswamy, S., Ross, K.N., Lander, E.S., and Golub, T.R. (2003). Amolecular

signature of metastasis in primary solid tumors. Nat. Genet. 33, 49–54.

Ray, D., Kazan, H., Cook, K.B., Weirauch, M.T., Najafabadi, H.S., Li, X.,

Gueroussov, S., Albu, M., Zheng, H., Yang, A., et al. (2013). A compendium

of RNA-binding motifs for decoding gene regulation. Nature 499, 172–177.

Rintala-Maki, N.D., Goard, C.A., Langdon, C.E., Wall, V.E., Traulsen, K.E.,

Morin, C.D., Bonin, M., and Sutherland, L.C. (2007). Expression of RBM5-

related factors in primary breast tissue. J. Cell. Biochem. 100, 1440–1458.

Singh, R., Valcarcel, J., and Green, M.R. (1995). Distinct binding specificities

and functions of higher eukaryotic polypyrimidine tract-binding proteins.

Science 268, 1173–1176.

Song, Z., Wu, P., Ji, P., Zhang, J., Gong, Q., Wu, J., and Shi, Y. (2012). Solution

structure of the second RRM domain of RBM5 and its unusual binding charac-

ters for different RNA targets. Biochemistry 51, 6667–6678.

Molec

Sutherland, L.C., Lerman, M., Williams, G.T., and Miller, B.A. (2001). LUCA-15

suppresses CD95-mediated apoptosis in Jurkat T cells. Oncogene 20, 2713–

2719.

Tollervey, J.R., Curk, T., Rogelj, B., Briese, M., Cereda, M., Kayikci, M., Konig,

J., Hortobagyi, T., Nishimura, A.L., Zupunski, V., et al. (2011). Characterizing

the RNA targets and position-dependent splicing regulation by TDP-43. Nat.

Neurosci. 14, 452–458.

Toriya, M., Tokunaga, A., Sawamoto, K., Nakao, K., and Okano, H. (2006).

Distinct functions of human numb isoforms revealed by misexpression in the

neural stem cell lineage in the Drosophila larval brain. Dev. Neurosci. 28,

142–155.

Ule, J., Ule, A., Spencer, J., Williams, A., Hu, J.S., Cline, M., Wang, H., Clark,

T., Fraser, C., Ruggiu, M., et al. (2005). Nova regulates brain-specific splicing

to shape the synapse. Nat. Genet. 37, 844–852.

Valcarcel, J., Singh, R., Zamore, P.D., and Green, M.R. (1993). The protein

Sex-lethal antagonizes the splicing factor U2AF to regulate alternative splicing

of transformer pre-mRNA. Nature 362, 171–175.

Verdi, J.M., Bashirullah, A., Goldhawk, D.E., Kubu, C.J., Jamali, M., Meakin,

S.O., and Lipshitz, H.D. (1999). Distinct human NUMB isoforms regulate differ-

entiation vs. proliferation in the neuronal lineage. Proc. Natl. Acad. Sci. USA 96,

10472–10476.

Wahl, M.C., Will, C.L., and Luhrmann, R. (2009). The spliceosome: design

principles of a dynamic RNP machine. Cell 136, 701–718.

Wang, Y., Gogol-Doring, A., Hu, H., Frohler, S., Ma, Y., Jens, M., Maaskola, J.,

Murakawa, Y., Quedenau, C., Landthaler, M., et al. (2013). Integrative analysis

revealed the molecular mechanism underlying RBM10-mediated splicing

regulation. EMBO Mol. Med. 5, 1431–1442.

Wang, Z., Kayikci, M., Briese, M., Zarnack, K., Luscombe, N.M., Rot, G.,

Zupan, B., Curk, T., and Ule, J. (2010). iCLIP predicts the dual splicing effects

of TIA-RNA interactions. PLoS Biol. 8, e1000530.

Warzecha, C.C., Jiang, P., Amirikian, K., Dittmar, K.A., Lu, H., Shen, S., Guo,

W., Xing, Y., and Carstens, R.P. (2010). An ESRP-regulated splicing pro-

gramme is abrogated during the epithelial-mesenchymal transition. EMBO J.

29, 3286–3300.

Welling, D.B., Lasak, J.M., Akhmametyeva, E., Ghaheri, B., and Chang, L.S.

(2002). cDNA microarray analysis of vestibular schwannomas. Otol.

Neurotol. 23, 736–748.

Weng, A.P., Nam, Y., Wolfe, M.S., Pear, W.S., Griffin, J.D., Blacklow, S.C., and

Aster, J.C. (2003). Growth suppression of pre-T acute lymphoblastic leukemia

cells by inhibition of notch signaling. Mol. Cell. Biol. 23, 655–664.

Westhoff, B., Colaluca, I.N., D’Ario, G., Donzelli, M., Tosoni, D., Volorio, S.,

Pelosi, G., Spaggiari, L., Mazzarol, G., Viale, G., et al. (2009). Alterations of

the Notch pathway in lung cancer. Proc. Natl. Acad. Sci. USA 106, 22293–

22298.

Xiao, R., Tang, P., Yang, B., Huang, J., Zhou, Y., Shao, C., Li, H., Sun, H.,

Zhang, Y., and Fu, X.D. (2012). Nuclear matrix factor hnRNP U/SAF-A exerts

a global control of alternative splicing by regulating U2 snRNP maturation.

Mol. Cell 45, 656–668.

Xue, Y., Zhou, Y., Wu, T., Zhu, T., Ji, X., Kwon, Y.S., Zhang, C., Yeo, G., Black,

D.L., Sun, H., et al. (2009). Genome-wide analysis of PTB-RNA interactions

reveals a strategy used by the general splicing repressor to modulate exon

inclusion or skipping. Mol. Cell 36, 996–1006.

Yeo, G.W., Coufal, N.G., Liang, T.Y., Peng, G.E., Fu, X.D., and Gage, F.H.

(2009). An RNA code for the FOX2 splicing regulator revealed by mapping

RNA-protein interactions in stem cells. Nat. Struct. Mol. Biol. 16, 130–137.

Zhao, L., Li, R., Shao, C., Li, P., Liu, J., andWang, K. (2012). 3p21.3 tumor sup-

pressor gene RBM5 inhibits growth of human prostate cancer PC-3 cells

through apoptosis. World J. Surg. Oncol. 10, 247.

ular Cell 52, 720–733, December 12, 2013 ª2013 Elsevier Inc. 733