ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis...

14
REVIEW Ribosomopathies: how a common root can cause a tree of pathologies Nadia Danilova 1, * and Hanna T. Gazda 2,3,4, * ABSTRACT Defects in ribosome biogenesis are associated with a group of diseases called the ribosomopathies, of which Diamond-Blackfan anemia (DBA) is the most studied. Ribosomes are composed of ribosomal proteins (RPs) and ribosomal RNA (rRNA). RPs and multiple other factors are necessary for the processing of pre-rRNA, the assembly of ribosomal subunits, their export to the cytoplasm and for the final assembly of subunits into a ribosome. Haploinsufficiency of certain RPs causes DBA, whereas mutations in other factors cause various other ribosomopathies. Despite the general nature of their underlying defects, the clinical manifestations of ribosomopathies differ. In DBA, for example, red blood cell pathology is especially evident. In addition, individuals with DBA often have malformations of limbs, the face and various organs, and also have an increased risk of cancer. Common features shared among human DBA and animal models have emerged, such as small body size, eye defects, duplication or overgrowth of ectoderm-derived structures, and hematopoietic defects. Phenotypes of ribosomopathies are mediated both by p53-dependent and -independent pathways. The current challenge is to identify differences in response to ribosomal stress that lead to specific tissue defects in various ribosomopathies. Here, we review recent findings in this field, with a particular focus on animal models, and discuss how, in some cases, the different phenotypes of ribosomopathies might arise from differences in the spatiotemporal expression of the affected genes. KEY WORDS: Ribosome biogenesis, Ribosomal protein, Ribosomopathy, Diamond-Blackfan anemia, p53, ΔNp63 Introduction Diamond-Blackfan anemia (DBA) is a congenital syndrome associated with anemia, physical malformations and cancer (Halperin and Freedman, 1989; Vlachos et al., 2012). In the majority of individuals with DBA, mutations or gene deletions of a subset of ribosomal proteins (RPs; see Box 1) are found, with RPS19 mutations accounting for about 25% of all cases (Table 1). Although some mutations are dominant negative, the major mechanism of the disease is associated with haploinsufficiency of an RP that disrupts the processing of pre-ribosomal RNA (pre- rRNA), leading to abortive ribosome biogenesis (Choesmel et al., 2007; Devlin et al., 2010; Farrar et al., 2014; Flygare et al., 2007; Gazda et al., 2004; Leger-Silvestre et al., 2005; Panic et al., 2006). DBA is a rare disease with an incidence of 5 cases per million live births, but it has attracted substantial attention as a model disease for ribosomopathies, a group of pathologies associated with defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes of ribosomopathies differ. A common feature among several ribosomopathies is p53 activation (Danilova et al., 2008b; Elghetany and Alter, 2002; Jones et al., 2008), but the mechanisms involved have not been completely elucidated. A p53-independent response to RP deficiency has also been observed (Aspesi et al., 2014; Danilova et al., 2008b; Singh et al., 2014; Torihara et al., 2011). The pathways that lead from a particular defect in ribosome biogenesis to the phenotype of a ribosomopathy are still not well understood. Several insightful reviews about the mechanisms of ribosomopathies have been published recently (Armistead and Triggs-Raine, 2014; Ellis, 2014; Golomb et al., 2014; James et al., 2014; Ruggero and Shimamura, 2014) with a focus on human data. However, cross- species analysis might also provide additional clues as to the mechanisms of these diseases. Here, we review the phenotypes that are caused by RP deficiency in humans as well as in mouse, fly and zebrafish models, with an emphasis on their common features. We also discuss the consequences of ribosomal stress at the molecular level, with a particular emphasis on p53 activation, metabolic changes, the origin of erythroid defects, and congenital malformations. We also discuss potential new directions for DBA treatment that arise from recent findings. Phenotypes caused by defects in ribosome biogenesis All ribosomopathies originate from defects in ribosome biogenesis, yet every ribosomopathy has a unique phenotype with different tissues affected. In this section, we provide an overview of the phenotypes caused by RP deficiency in human DBA and in its animal models, and compare them to phenotypes of other ribosomopathies. RP deficiency In humans, DBA is often diagnosed during the first year of life; common clinical features include anemia, low reticulocyte count, macrocytic erythrocytes (see Box 1 for a glossary of terms), increased expression of fetal hemoglobin and elevated activity of adenosine deaminase (ADA; Box 1) (Fargo et al., 2013; Halperin and Freedman, 1989). Approximately 40% of affected individuals have short stature and variable congenital malformations of craniofacial skeleton, eyes, heart, visceral organs and limbs. Notably, duplication of some structures, including triphalangeal thumb, bifid thumb and extra ribs, has been reported (Halperin and Freedman, 1989). The clinical symptoms of DBA rarely correlate with the type of causative mutation. However, individuals with mutations in RPL5 often have a cleft palate, whereas this malformation was not observed in individuals carrying RPS19 mutations (Gazda et al., 2008). The acquired haploinsufficiency of RPS14 has been shown to underlie the 1 Department of Molecular, Cell & Developmental Biology, University of California, Los Angeles, CA 90095, USA. 2 Division of Genetics and Genomics, The Manton Center for Orphan Disease Research, Boston Childrens Hospital, Boston, MA 02115, USA. 3 Department of Pediatrics, Harvard Medical School, Boston, MA 02115, USA. 4 Broad Institute, Cambridge, MA 02142, USA. *Authors for correspondence ([email protected]; hanna.gazda@childrens. harvard.edu) This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed. 1013 © 2015. Published by The Company of Biologists Ltd | Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529 Disease Models & Mechanisms

Upload: others

Post on 14-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

REVIEW

Ribosomopathies: how a common root can cause a tree ofpathologiesNadia Danilova1,* and Hanna T. Gazda2,3,4,*

ABSTRACTDefects in ribosomebiogenesisareassociatedwithagroupofdiseasescalled the ribosomopathies, of whichDiamond-Blackfan anemia (DBA)is the most studied. Ribosomes are composed of ribosomal proteins(RPs) and ribosomal RNA (rRNA). RPs and multiple other factors arenecessary for the processing of pre-rRNA, the assembly of ribosomalsubunits, their export to the cytoplasm and for the final assembly ofsubunits into a ribosome. Haploinsufficiency of certain RPs causesDBA, whereas mutations in other factors cause various otherribosomopathies. Despite the general nature of their underlyingdefects, the clinical manifestations of ribosomopathies differ. In DBA,for example, red blood cell pathology is especially evident. In addition,individuals with DBA often have malformations of limbs, the face andvarious organs, and also have an increased risk of cancer. CommonfeaturessharedamonghumanDBAandanimalmodelshaveemerged,such as small body size, eye defects, duplication or overgrowth ofectoderm-derived structures, and hematopoietic defects. Phenotypesof ribosomopathies are mediated both by p53-dependent and-independent pathways. The current challenge is to identifydifferences in response to ribosomal stress that lead to specific tissuedefects in various ribosomopathies. Here, we review recent findings inthis field, with a particular focus on animal models, and discuss how, insome cases, the different phenotypes of ribosomopathies might arisefromdifferences in thespatiotemporal expressionof theaffectedgenes.

KEY WORDS: Ribosome biogenesis, Ribosomal protein,Ribosomopathy, Diamond-Blackfan anemia, p53, ΔNp63

IntroductionDiamond-Blackfan anemia (DBA) is a congenital syndromeassociated with anemia, physical malformations and cancer(Halperin and Freedman, 1989; Vlachos et al., 2012). In themajority of individuals with DBA, mutations or gene deletions of asubset of ribosomal proteins (RPs; see Box 1) are found, withRPS19 mutations accounting for about 25% of all cases (Table 1).Although some mutations are dominant negative, the majormechanism of the disease is associated with haploinsufficiency ofan RP that disrupts the processing of pre-ribosomal RNA (pre-rRNA), leading to abortive ribosome biogenesis (Choesmel et al.,2007; Devlin et al., 2010; Farrar et al., 2014; Flygare et al., 2007;Gazda et al., 2004; Leger-Silvestre et al., 2005; Panic et al., 2006).

DBA is a rare disease with an incidence of ∼5 cases per millionlive births, but it has attracted substantial attention as a modeldisease for ribosomopathies, a group of pathologies associated withdefects in ribosome biogenesis (Armistead and Triggs-Raine, 2014;James et al., 2014). Despite this common defect, phenotypes ofribosomopathies differ. A common feature among severalribosomopathies is p53 activation (Danilova et al., 2008b;Elghetany and Alter, 2002; Jones et al., 2008), but the mechanismsinvolved have not been completely elucidated. A p53-independentresponse toRP deficiency has also been observed (Aspesi et al., 2014;Danilova et al., 2008b; Singh et al., 2014; Torihara et al., 2011). Thepathways that lead from a particular defect in ribosome biogenesis tothe phenotype of a ribosomopathy are still not well understood.Several insightful reviews about the mechanisms of ribosomopathieshave been published recently (Armistead and Triggs-Raine, 2014;Ellis, 2014; Golomb et al., 2014; James et al., 2014; Ruggero andShimamura, 2014) with a focus on human data. However, cross-species analysis might also provide additional clues as to themechanisms of these diseases.

Here, we review the phenotypes that are caused by RP deficiency inhumans as well as in mouse, fly and zebrafish models, with anemphasis on their common features.We also discuss the consequencesof ribosomal stress at the molecular level, with a particular emphasison p53 activation, metabolic changes, the origin of erythroid defects,and congenital malformations. We also discuss potential newdirections for DBA treatment that arise from recent findings.

Phenotypes caused by defects in ribosome biogenesisAll ribosomopathies originate from defects in ribosome biogenesis,yet every ribosomopathy has a unique phenotype with differenttissues affected. In this section, we provide an overview of thephenotypes caused by RP deficiency in human DBA and in itsanimal models, and compare them to phenotypes of otherribosomopathies.

RP deficiencyIn humans, DBA is often diagnosed during the first year of life;common clinical features include anemia, low reticulocyte count,macrocytic erythrocytes (see Box 1 for a glossary of terms), increasedexpression of fetal hemoglobin and elevated activity of adenosinedeaminase (ADA; Box 1) (Fargo et al., 2013; Halperin and Freedman,1989). Approximately 40% of affected individuals have short statureand variable congenital malformations of craniofacial skeleton, eyes,heart, visceral organs and limbs. Notably, duplication of somestructures, including triphalangeal thumb, bifid thumb and extra ribs,has been reported (Halperin and Freedman, 1989). The clinicalsymptoms ofDBA rarely correlatewith the type of causativemutation.However, individuals with mutations in RPL5 often have a cleftpalate, whereas this malformation was not observed in individualscarrying RPS19 mutations (Gazda et al., 2008). The acquiredhaploinsufficiency of RPS14 has been shown to underlie the

1Department of Molecular, Cell & Developmental Biology, University of California,Los Angeles, CA 90095, USA. 2Division of Genetics and Genomics, The MantonCenter for Orphan Disease Research, Boston Children’s Hospital, Boston,MA 02115, USA. 3Department of Pediatrics, Harvard Medical School, Boston, MA02115, USA. 4Broad Institute, Cambridge, MA 02142, USA.

*Authors for correspondence ([email protected]; [email protected])

This is an Open Access article distributed under the terms of the Creative Commons AttributionLicense (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use,distribution and reproduction in any medium provided that the original work is properly attributed.

1013

© 2015. Published by The Company of Biologists Ltd | Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 2: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

erythroid defect in 5q-myelodysplastic syndrome (MDS) (Ebert et al.,2008; Pellagatti et al., 2008).Individuals with DBA have an increased frequency of both solid

cancers and leukemia (Vlachos et al., 2012), and acquired mutationsin RPL5 and RPL10 are found in T-cell acute lymphoblasticleukemia (De Keersmaecker et al., 2013).A strikingly different phenotype is associated with the RPSA

protein, an RP that is involved in the maturation of the 40S subunitand also serves as a laminin receptor. Mutations in this RP lead tofamilial isolated congenital asplenia (see Box 1 for a glossary ofterms) (Bolze et al., 2013). However, this phenotype is likelyassociated with the function of RPSA as a laminin receptor becauselaminin deficiency has been shown to lead to asplenia (Miner andLi, 2000).In the fruit fly, Drosophila, the haploinsufficiency of many

RPs leads to a Minute phenotype, which is characterized bydevelopmental delay, small size, small bristles, small rough eyes,and recessive lethality at late embryonic/early larval stages(Kongsuwan et al., 1985; Marygold et al., 2005). Notably,heterozygous Rpl38 and Rpl5 Drosophila mutants have abnormallylarge wings (Marygold et al., 2005).Zebrafish embryos homozygous for RP mutations also exhibit

developmental delay and small size as well as a small head and eyes,

brain apoptosis, pericardial edema, reduced pigmentation andhematopoietic defects (Amsterdam et al., 2004a; Danilova et al.,2011; Taylor et al., 2012; Zhang et al., 2014). Similar phenotypeswere observed in zebrafish embryos in which rps19 or rpl11 wereknocked down using morpholinos; in addition, some morphants hadmissing eyes, abnormal positioning of the heart and pancreas, andincreased fin size (Chakraborty et al., 2009; Danilova et al., 2008b;Uechi et al., 2008). Adult zebrafish heterozygous for RP mutationshave a high frequency of malignant peripheral nerve sheath tumors(Amsterdam et al., 2004b).

In mice, mutations in Rps19 and Rps20 cause dark skin, reducedbody size and a reduced erythrocyte count (McGowan et al., 2008).RPS6-deficient mice also have hematopoietic defects (Keel et al.,2012). A transgenicmouse carrying anRps19R62Wmutation, whichis common in DBA, has been recently developed (Devlin et al.,2010). The constitutive expression of the transgene bearing thismutation is developmentally lethal. Its conditional expression resultsin growth retardation and anemia. Another RPS19-deficient mousecreated by transgenic RNA interference also develops symptomssimilar to those found in individuals with DBA (Jaako et al., 2011).

Mutation of the mouse Rpl24 leads to the Belly Spot and Tail(Bst) phenotype characterized by small size, eye defects, awhite ventral spot, white hind feet and various skeletalabnormalities including duplicated digits and phalanges(Fig. 1), which is similar to the anomaly noted in individualswith DBA (Oliver et al., 2004). Mutations in Rps7 also lead toskeletal malformations, ventral white spotting and eye defects(Watkins-Chow et al., 2013).

Thus, several cross-species features caused by RP deficiencyhave emerged, which include a smaller size, eye defects, congenitalmalformations often associated with the duplication or overgrowthof structures derived from the epidermal ectoderm, and varyingdegrees of anemia.

Examples of phenotypes of other ribosomopathiesTreacher Collins syndrome (TCS) is caused by mutations in genesinvolved in rRNA transcription, such as the Treacher Collins-Franceschetti syndrome 1 (TCOF1) and POLR1D and POLR1Cgenes, which encode subunits of RNA polymerase I and III(PolI/III; Box 1) (Dauwerse et al., 2011; Valdez et al., 2004). TCS isassociated with craniofacial deformities such as absent cheekbones(Dixon et al., 2006).

Shwachman Diamond syndrome (SDS) is caused by mutations inthe Shwachman-Bodian-Diamond syndrome (SBDS) gene, whichfunctions in ribosomal subunit joining (Boocock et al., 2003).SDS is characterized by decreased production of white bloodcells, exocrine pancreatic insufficiency, and problems with boneformation and growth.

Dyskeratosis congenita (DC) is caused by mutations in genes thatcode for the components of small nucleolar ribonucleoproteincomplexes, which function in rRNA processing (Ruggero et al.,2003). Individuals with DC have mucocutaneous defects and bonemarrow failure.

North American Indian childhood cirrhosis (NAIC) is caused bymutation in CIRH1A [cirrhosis, autosomal recessive 1A (cirhin)],which encodes a protein that functions in rRNA processing(Chagnon et al., 2002). The phenotype is restricted to neonataljaundice that progresses to biliary cirrhosis (Box 1) that at somepoint requires hepatic transplantation.

These examples illustrate the variability of phenotypes ofribosomopathies. A comprehensive description of these and otherribosomopathies can be found in recent reviews (Armistead and

Box 1. Glossary

5′ and 3′ external transcribed spacers (ETSs): non-functional RNAsequences of the pre-rRNA transcript that have structural roles and areexcised during pre-rRNA processing.5q-myelodysplastic syndrome (5q-MDS): a form of MDS that iscaused by loss of a part of the q arm of chromosome 5.Adenosine deaminase (ADA): an enzyme involved in the metabolismof adenosine.Anemia: a condition associated with an insufficient number of red bloodcells in blood.Asplenia: absence of spleen or very small spleen.Biliary cirrhosis: cirrhosis caused by damage to the bile ducts in the liver.Epiboly: growth of a cell layer to envelope the yolk during gastrulation.Haploinsufficiency: when a single copy of the gene is insufficient tomaintain normal function.Internal ribosomeentrysite (IRES): a sequence insidemRNA that allowsfor initiation of cap-independent translation in the middle of mRNA.Internally transcribed spacers (ITSs): spacers between 18S, 5.8S and28S rRNA in the pre-rRNA transcript that play structural roles and, likeETSs, are excised during pre-rRNA processing.Jaundice: yellow color of the skin and whites of the eyes caused byexcess bilirubin in the blood.Macrocytic erythrocytes: abnormally large red blood cells.Mechanistic target of rapamycin (mTOR): a serine/threonine kinasethat is a central regulator of cellular metabolism. It forms mTORC1 andmTORC2 complexes, which mediate cellular responses to stresses suchas DNA damage and nutrient deprivation.Myelodysplastic syndrome (MDS): a syndrome caused by mutationsin several genes, most often encoding splicing factors. It is associatedwith ineffective production of blood, which often leads to leukemia.Reticulocyte: immature erythrocyte.RNA polymerase I and III (PolI/PolIII): enzymes involved in thetranscription of non-coding RNAs.Ribosomal proteins (RPs): proteins that together with ribosomal RNA(rRNA) make up the ribosome. They are called RPS (RP from smallribosomal subunit) or RPL (RP from large ribosomal subunit) dependingonwhether theyassociatewith the small or large subunit of the ribosome.Small nucleolar RNA (snoRNA): a class of small RNAs that guidechemical modifications such as methylation and pseudouridylation ofother RNAs.

1014

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 3: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

Triggs-Raine, 2014; James et al., 2014). How a general defect canlead to the malfunction of a specific tissue remains a focus ofmany ongoing studies, which uncovered that some phenotypicconsequences of ribosomal stress are caused by p53 activation, asdiscussed in more detail in the next section.

Abortive ribosome biogenesis and p53 activationRibosome biogenesis: stages and the role of RPsThe eukaryotic ribosome is composed of a small (40S) and a large(60S) subunit (Kressler et al., 2010). The small subunit includes 18SrRNA and 33 RPs; the large subunit includes 5S rRNA, 28S rRNA,5.8S rRNA and 46 RPs. The genes encoding rRNA (rDNA) arefound in multiple copies organized into tandem repeats. 18S, 5.8Sand 28S rRNAs are transcribed as a single pre-rRNA transcript byRNA polymerase I in a substructure of the nucleus called thenucleolus. In yeast, the rDNA repeat also encodes the 5S rRNA,which is transcribed in the reverse direction (Henras et al., 2015).In human cells, the 5S rRNA precursor is transcribed from multiplegenes in the nucleoplasm by RNA polymerase III (Henras et al.,2015). Then, 5S rRNA migrates to the nucleolus for furtherprocessing and incorporation into the pre-60S subunit. The nascentpre-rRNA assembles co-transcriptionally with a subset of RPs andwith multiple other factors that facilitate the folding, modificationand cleavage of pre-rRNA and the formation of ribosomal subunits(Fig. 2). Pre-rRNA processing can take place co-transcriptionally aswell as post-transcriptionally. In yeast, the pre-40S subunit mostoften is released by cleavage within the internal transcribed spacer 1(ITS1) before the transcription of the 3′ end of the pre-rRNA isfinished. Details of pre-rRNA processing can be found in recentreviews (Henras et al., 2015; Kressler et al., 2010).

Most RPs are strictly required for the pre-rRNA processing. It isthought that their role is to assist the proper folding of the pre-rRNA(Henras et al., 2015). RPs act in a hierarchical order. Approximatelyhalf of RPs from the 40S ribosomal subunit, including RPS24 andRPS7, associate early with the 5′ end of the nascent pre-rRNA and

Fig. 1. Duplicated digits and phalanges in mice heterozygous for Rpl24.Skeletal stain of newborn forelimbs (upper) and hindlimbs (lower). Miceheterozygous for a mutation in Rpl24 (Bst/+ phenotype) show preaxialpolydactyly (0) and triphalangy of the first digit (1). Figure reproduced withpermission (Oliver et al., 2004).

Table 1. Genes mutated in DBA

Gene% of affectedindividuals Type of mutation References

RPS19 ∼25 Single-nucleotide variant (SNV) (missense,nonsense, splice site mutations),microinsertions, microdeletions, singlecopy deletions

Draptchinskaia et al., 1999; Farrar et al.,2011; Gazda et al., 2004

RPL5 ∼7 SNV (missense, nonsense, splice sitemutations), microinsertions,microdeletions

Cmejla et al., 2009; Gazda et al., 2008;Quarello et al., 2012

RPS26 ∼6.5 SNV (missense and splice site mutations),microinsertions, single copy deletions

Doherty et al., 2010; Farrar et al., 2011

RPL11 ∼5 SNV (missense, nonsense, splice sitemutations), microinsertions,microdeletions

Cmejla et al., 2009; Gazda et al., 2008;Quarello et al., 2012

RPL35A ∼3 SNV (missense, nonsense, splice sitemutations), microdeletion, single copydeletions

Farrar et al., 2008; Farrar et al., 2011

RPS10 ∼2.5 SNV (missense and nonsense),microinsertion

Doherty et al., 2010

RPS24 ∼2 SNV (nonsense), microinsertion/microdeletion, large deletion

Gazda et al., 2006b; Landowski et al., 2013

RPS17 ∼1 SNV (missense), microinsertion, single copydeletions

Cmejla et al., 2007; Farrar et al., 2011;Gazda et al., 2008

RPS7 <1 SNV (splice site mutation) Gazda et al., 2008RPL26 <1 Microdeletion Gazda et al., 2012GATA1 <1 SNV (missense, splice site mutations) Sankaran et al., 2012RPL15 <1 Large deletion Landowski et al., 2013RPS29 <1 SNV (missense mutations) Mirabello et al., 2014RPS28 <1 SNV (missense mutations) Gripp et al., 2014TSR2 <1 SNV (missense mutation) Gripp et al., 2014RPL31 <1 Single copy deletion Farrar et al., 2014RPS27 <1 Microdeletion Wang et al., 2015RPL27 <1 SNV (splice site mutation) Wang et al., 2015

Genes ordered according to the percentage of DBA-affected individuals who carry a certain mutation. RPL, RP from large ribosomal subunit; RPS, RP from smallribosomal subunit.

1015

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 4: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

are required for the initiation of cleavages at the 5′ externaltranscribed spacer (5′ETS; Box 1) and ITS1 (Fig. 2). The second setof RPs, which includes RPS19 and RPS17 along with other RPs, is

not required for 5′ETS removal but is necessary for ITS1 cleavage.Correspondingly, depletion of RPS24 or RPS7 in yeast and humancells results in the failure of maturation of the 5′ end of 18S rRNA,

Nucleolus Nucleus

Cytoplasm

Transcription

Processing

ProcessingMaturationExport

Maturation

Ribosomeformation

RPSs

RPS24or RPS7

deficiency

RPS19or RPS17deficiency

20S

Fig. 2. A simplified schematic of ribosome biogenesis in human cells. (A) 18S, 5.8S and 28S rRNAs are transcribed by Pol1 in the nucleolus as segments of along precursor pre-rRNA, which also includes two externally transcribed spacers 5′ETS and 3′ETS and two internally transcribed spacers, ITS1 and ITS2 (B; Box 1).5S rRNA is transcribed independently by PolIII in the nucleus. (B) Concomitant with transcription, the pre-rRNA assembles with accessory factors and a subset ofribosomal proteins (RPs: RPSs and RPLs). This facilitates the formation of a secondary structure necessary for the correct folding, modification and cleavage ofpre-rRNA. (C)After removal of the5′ETSandcleavage in the ITS1site, pre-40S (which contains the20Sprecursorof 18S rRNA)andpre-60Ssubunits are formedandcontinue to mature. 5S rRNA incorporates into pre-60S subunit. Subunits are then exported to the cytoplasm. (D) Once in the cytoplasm, small and large subunitsundergo final maturation, which involves the removal of remaining accessory factors and incorporation of missing RPs. (E) A functional ribosome forms aftertranscribedmRNAbinds to the40Ssubunit, which triggersassociationof the60Ssubunitwith thiscomplex.More than200accessory factors,which includehelicases,nucleases, small nucleolar RNAs (snoRNAs; Box 1), chaperones and transporters, temporally associate with the maturing ribosomal subunits at various steps. Inhuman cells, pre-rRNA processing is differentially affected by deficiency of various RPs. For example, deficiency of RPS24 or RPS7 prevents formation of the 20Sprecursor of 18S rRNA, whereas deficiency of RPS19 or RPS17 prevents conversion of the 20S precursor to a mature 18S rRNA.

1016

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 5: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

whereas, in the absence of RPS19 or RPS17, the 3′ end of the 18SrRNA cannot mature (Choesmel et al., 2007, 2008; Flygare et al.,2007; Leger-Silvestre et al., 2005; Robledo et al., 2008). In all cases,pre-40S particles that contain non-cleaved pre-rRNA cannot beexported and accumulate in the nucleus. Similarly, deficiency ofmost other RPs that are mutated in DBA, both from the small and thelarge ribosomal subunits, affects pre-rRNA processing in a uniqueway, leading to the accumulation of different rRNA precursors anddisruption of ribosome biogenesis at different steps (Robledo et al.,2008).

RP deficiency and p53 activationRibosomal biogenesis is a highly energy-consuming process thatis coupled to cell growth and requires tight regulation. It iscontrolled by the TOR pathway, which regulates the synthesis ofribosomal components in response to growth factors and nutrientavailability (see Box 1 for a glossary of terms) (Kressler et al.,2010). Ribosome biosynthesis requires coordinated activity of threeRNA polymerases, PolI, PolII and PolIII. Cellular stress such ashypoxia or DNA damage decreases ribosome biogenesis mostlyby inhibition of PolI transcription of rRNA through severalmechanisms (Boulon et al., 2010). Shutting down ribosomebiogenesis during stress preserves cellular homeostasis andensures that enough cellular resources can be relocated to aresponse to stress.One of the first indications that p53 is involved in controlling the

fidelity of ribosome biogenesis came from the finding that amutation in Bop1, which is involved in maturation of rRNAs, leadsto p53-dependent cell cycle arrest in mouse cells (Pestov et al.,2001) and that RPS6-gene haploinsufficiency activates p53 inmouse embryos and T cells (Panic et al., 2006; Sulic et al., 2005).Subsequent studies have established p53 activation as a generalresponse to RP deficiency (Barlow et al., 2010; Danilova et al.,2008b; Dutt et al., 2011; Fumagalli and Thomas, 2011; McGowanet al., 2008; Taylor et al., 2012). Acquired loss of RPS14 in5q-myelodysplastic syndrome (5q-MDS; Box 1) is also associatedwith p53 upregulation (Dutt et al., 2011; Pellagatti et al., 2008).Several hypotheses to explain how p53 is activated in DBA have

been proposed. The nucleolus has been suggested to be a universalstress sensor that is responsible for the maintenance of a low level ofp53 in the cell; the nucleolus disrupts and arrests the cell cycle inresponse to various stresses, including DNA damage, hypoxia, heatshock, nucleoside triphosphate (NTP) depletion, and others (Rubbiand Milner, 2003). However, RP deficiency does not always lead tonucleolar disruption (Fumagalli et al., 2009), suggesting that othermechanisms sense ribosomal stress as well.Oneway in which altered ribosome biogenesis might affect p53 is

through RPs that are not incorporated into ribosomes (Fig. 3).Several RPs have been shown to bind Mdm2 (mouse double minute2 homolog, which is a negative regulator of p53) and to inhibit itsbinding to p53, leading to p53 stabilization and to cell cycle arrest(Zhang and Lu, 2009). Attaining the right balance between thesynthesis of rRNA and RPs seems to be important; when rRNAsynthesis is decreased, RPs are no longer used for ribosome buildingand can stabilize p53 (Donati et al., 2011). Thus, p53 stabilizationby RPs might be a general mechanism involved in the response tovarious stresses.RPL5 and RPL11 have been proposed to be the essential players

in the regulation of p53 by RPs (Macias et al., 2010). Together with5S rRNA they form the Mdm2 regulatory complex (Sloan et al.,2013). They are protected from the proteasomal degradation thatother RPs undergo when PolI is inhibited (Bursac et al., 2012). This

raises the question of how cells respond to a deficiency of RPL5 orRPL11. In primary human lung fibroblasts and in murine embryonicstem cell lines, their depletion reportedly does not induce the p53pathway (Teng et al., 2013; Singh et al., 2014). However, Rpl11-deficient zebrafish do have an upregulated p53 pathway(Chakraborty et al., 2009; Danilova et al., 2011). Thesecontradictory results flag the need for further investigation intothe mode of action of these RPs in p53 regulation.

Besides the RPs, Mdm2 has over a hundred binding partners andsome of them might be involved in response to ribosomal stress(James et al., 2014). One of them, p14ARF, an alternate readingframe protein product of cyclin-dependent kinase inhibitor 2A(CDKN2A), is an Mdm2 inhibitor and is upregulated in response tooncogenic signaling. An increase in such signaling has beenreported in hematopoietic progenitors derived from DBA patients(Gazda et al., 2006a) and in Rpl11-deficient zebrafish (Danilovaet al., 2011) (Fig. 3).

Some data suggest that ataxia telangiectasia and Rad3 related(ATR) and ataxia telangiectasia mutated (ATM) kinases, whichare responsible for the replication stress and DNA-damagecheckpoints, might contribute to p53 upregulation in situationswhere rRNA synthesis is compromised (Fig. 3). The ATR-Chk1axis was implicated in cell cycle arrest induced by inhibition ofrRNA synthesis using actinomycin D (Ma and Pederson, 2013).Another study reported upregulation of the ATR-ATM-Chk1-p53pathway in RPS19-deficient human cells and in zebrafish modelsof DBA (Danilova et al., 2014). The exact mechanism of ATRinduction in these studies was not explored. However, someresearchers have hypothesized that the obstruction of pre-rRNAprocessing might interfere with transcription and, ultimately,with DNA replication (Bermejo et al., 2012). This hypothesis isbased on the fact that pre-rRNA processing starts before itstranscription is finished (Osheim et al., 2004; Schneider et al.,2007). Pre-mRNA splicing also happens co-transcriptionally andsplicing defects increase the formation of DNA double-strandbreaks (Li and Manley, 2005).

An additional source of p53 activation in RP-deficient cells mightarise from the altered nucleotide metabolism (Fig. 3). In RP-deficient zebrafish, the level of adenosine triphosphate (ATP),which is a source of energy, is decreased, whereas that ofdeoxythymidine triphosphate (dTTP), which is a building blockof DNA, is increased (Danilova et al., 2014). Decreased ATP leadsto activation of the energy sensor AMP-activated protein kinase(AMPK), which, in turn, activates p53 (Hardie et al., 2012). Thedisproportional increase or decrease in one of deoxyribonucleotidetriphosphates can interfere with DNA synthesis and result in p53upregulation (Austin et al., 2012; Sanchez et al., 2012).

RP deficiency in a fetal environment, in addition to p53 proteinstabilization, leads to the transcriptional upregulation of p53(Chakraborty et al., 2009; Danilova et al., 2008b) possiblybecause of the upregulation of certain growth factors such as Myc(Danilova et al., 2011).

Reactive oxygen species (ROS) can also contribute to p53stabilization in RP-deficient cells (Heijnen et al., 2014). Theinsufficient production of hemoglobin and the accumulation of theexcess of heme might contribute to oxidative stress in erythroidprogenitors (Ellis, 2014).

Various hypotheses of p53 activation are not mutually exclusive.In fact, several mechanisms might function in parallel to induce p53to guard ribosome biogenesis. The contribution of each mechanismcould differ between species, tissues, and with the particular RPinvolved.

1017

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 6: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

p53 activation in other ribosomopathiesp53 activation occurs in other ribosomopathies as well. Thecraniofacial defects in TCS are mediated by p53 upregulation(Jones et al., 2008). p53 protein overexpression has been found inbone marrow biopsies of patients with SDS (Elghetany and Alter,2002). In zebrafish and mouse models of DC, p53 is upregulated(Pereboom et al., 2011; Zhang et al., 2012; Gu et al., 2008).Knockdown of the zebrafish homolog of the gene responsible forNAIC results in p53 activation (Wilkins et al., 2013).

p53-independent responses to ribosomal stressRecent studies of transcriptional responses to RP deficiency inducedin p53-negative human cell lines revealed changes in the expressionof genes involved in metabolism, proliferation, apoptosis and cellredox homeostasis (Aspesi et al., 2014). Studies of murineembryonic stem cells haploinsufficient for Rps19 and Rpl5indicate the presence of p53-independent cell-cycle and erythroid-differentiation defects (Singh et al., 2014). Zebrafish studies alsopoint to both p53-dependent and -independent responses to RP

deficiency (Danilova et al., 2008b; Torihara et al., 2011). Yeast, anorganism that lacks the p53-Mdm2 pathway, responds to theinhibition of ribosome biogenesis by cell cycle arrest, whichinvolves a yeast functional equivalent of the human tumorsuppressor pRb (protein retinoblastoma) (Bernstein et al., 2007).Some p53-independent responses to ribosomal stress are outlinedbelow.

Reduced ribosome number or activityReduced ribosome activity might contribute to DBA phenotypes.Slowed proliferation of human RPL11- or RPL5-deficient lungfibroblasts was attributed to reduced ribosome content andtranslational capacity that suppressed the accumulation of cyclinsand thus cell cycle progression (Teng et al., 2013). Althoughreduced ribosome number and activity might still be sufficient forsurvival, they might be not adequate to translate some mRNAs thathave stringent conditions for initiation of translation (Ludwig et al.,2014). Moreover, ribosomal stress affects translation of internalribosome entry site (IRES)-containing RNAs (Armistead and

Nucleolardisruption

p53-independentpathways

Cell cycle

Lowcyclins

Decreased number of ribosomes

Altered translational profileCytoplasm

Nucleus

Nucleolus

- -

Low PIM

Low ATP

Fig. 3. Defects in ribosomal biogenesis activate p53 and other stress-response mechanisms. A schematic showing pre-rRNA transcription, andassembly of accessory factors and RPs on the nascent pre-rRNA. (A) Recent studies have suggested that problems with pre-rRNA processing can affect DNAtranscription, leading to the activation of ATR-ATM-Chk1/2 signaling (which is responsible for the replication-stress and DNA-damage checkpoints) and p53upregulation. In addition, deoxynucleoside triphosphate (dNTP) imbalance caused by RP deficiency might interfere with transcription and replication andcontribute to ATR-ATM activation. (B) Problems with pre-RNA processing compromise ribosome biogenesis and lead to nucleolar disruption. The nucleolus isinvolved in maintaining low p53 levels by exporting it for degradation. Various stressors disrupt nucleolar organization, compromising p53 export and leading top53 accumulation in the nucleus. Nucleolar disruption might also lead to the release of factors that activate p53 or cause cell cycle arrest by p53-independentmechanisms. (C) An alternative pathway of p53 activation is through free RPs that, in complex with 5S RNA, bind the p53 negative regulator MDM2, releasing p53from its control. (D) Upregulation of MYC and RAS pro-survival factors in DBA patients and in animal models suggests that they might activate p14ARF, which, incomplex with 5S RNA, also negatively regulates MDM2. Hypothetically, additional not-yet-identified nucleolar factors might also negatively interact withMDM2 and contribute to p53 upregulation. (E) p53 might also be activated by secondary changes in RP-deficient cells, such as increased levels of ROS ordecreased levels of ATP, which activates AMPK, which, in turn, activates p53. p53 then translocates to the nucleus. (F) p53 activation leads to cell cycle arrest andto the induction of downstream pathways ranging from cellular repair to apoptotic mechanisms. (G) A p53-independent response might also originate from thecytoplasm owing to a decreased number and altered activity of ribosomes, which also can lead to cell cycle arrest. For example, decreased levels of cyclins orPIM1 caused by RP deficiency might inhibit cell cycle progression. Abbreviations: AMPK, AMP-activated protein kinase; ATM, ataxia telangiectasia mutated;ATR, ataxia telangiectasia and Rad3 related; Chk1/2, checkpoint kinase 1/2; MDM2, MDM2 oncogene, E3 ubiquitin protein ligase; MYC, avianmyelocytomatosisviral oncogene homolog; p14ARF, alternate reading frame protein product of the CDKN2A, cyclin-dependent kinase inhibitor 2A; PIM, pim-1 oncogene; PolI, RNApolymerase I; RAS, rat sarcoma viral oncogene homolog; ROS, reactive oxygen species; 5S, rRNA. See Fig. 2 for a key.

1018

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 7: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

Triggs-Raine, 2014; Horos et al., 2012) (Box 1). Thus, ribosomalstress might lead to changes in the spectrum of translated mRNAs,which could differ between tissues.

Ribosomes as a platform for kinase signalingRibosomes were suggested to be a platform for signaling moleculesafter the discovery that RAC1, a receptor for protein kinase C, isa constituent of the ribosome (Nilsson et al., 2004). The PIM1proto-oncogene serine/threonine kinase is also associated withribosomes (Chiocchetti et al., 2005). In human erythroid cell lines,RP deficiency and other types of ribosomal stress such as inhibition ofrRNA production at various steps with actinomycin, camptothecin orcisplatinum cause a decrease of PIM1 levels, leading to inhibition ofthe cell cycle (Iadevaia et al., 2010). A pathway conserved from yeastto mammals involves the regulation of growth-related processes viathe interaction between ribosomes and TORC2 (see Box 1 for aglossary of terms), and ribosomal defects can inhibit TORC2(Zinzalla et al., 2011). These mechanisms might contribute to thephenotypes of RP-deficient cells.

RPs as regulatorsIn addition to their structural role in the ribosome, RPs are involvedin other cellular processes. Several RPs modulate the activity ofregulatory proteins such as NF-κB, p53, p21, Myc and nuclearreceptors, whereas others regulate translation by binding tountranslated regions of mRNAs (Warner and McIntosh, 2009).Some RPs act as a regulatory component of the ribosome to confertranscript-specific translational control. For example, translation ofvesicular stomatitis virus mRNAs depends on RPL40 (Lee et al.,2013). This RP is also required for translation of some cellularmRNAs, including those involved in the stress response (Lee et al.,2013). RPS25 facilitates interactions of the ribosome with viralIRES elements (Landry et al., 2009). In Rpl38-mutant mouseembryos, global translation is unchanged but the translation of asubset of homeoboxmRNAs is perturbed (Kondrashov et al., 2011).An imbalance in RP regulatory functions might contribute to thepathophysiology of DBA. For example, increased frequency ofphysical malformations in individuals with mutations in RPL5 andRPL11 (Gazda et al., 2008) might be related to impaired regulatoryfunctions of these RPs.

Pathological changes in RP-deficient cellsBoth p53-positive and p53-negative animal cells respond to RPdeficiency by global homeostatic changes that serve to stallproliferation and to create conditions for cellular repair, senescenceor apoptosis depending on the cell type and the intensity of the stresssignal. Here, we outline the most significant changes that have beenobserved in RP-deficient cells, both p53-dependent or -independent.A better understanding of these changes might help to find newtreatments for DBA and other ribosomopathies.

Cell cycle arrest and apoptosisAn increase in the expression of genes responsible for cell cyclearrest has been reported in various DBAmodels (Barkic et al., 2009;Danilova et al., 2008b; Miyake et al., 2008). Slower cellproliferation might promote repair and improve survival (Barkicet al., 2009). However, when repair is not possible, cells activateapoptosis. Increased apoptosis has been reported in animal andcellular models of DBA, as well as in cells from DBA patients(Avondo et al., 2009; Danilova et al., 2011, 2008b; Gazda et al.,2006a; Miyake et al., 2008; Perdahl et al., 1994). p53-negative cells

also respond to RP deficiency by the inhibition of proliferation andby the upregulation of pro-apoptotic genes (Aspesi et al., 2014).

MetabolismRpl11-deficient zebrafish and RPS19-deficient mouse fetal livercells downregulate genes that encode glycolytic enzymes andupregulate genes involved in aerobic respiration (Danilova et al.,2011). This is consistent with a known effect of p53 activation(Matoba et al., 2006). Glycolysis provides not only ATP but alsointermediates for the biosynthesis of carbohydrates, proteins, lipidsand nucleic acids, and is increased in normal cells duringproliferation and in tumors, which is known as the Warburg effect(Lunt and Vander Heiden, 2011). Suppressed glycolysis in RP-deficient cells means that lower levels of intermediates are availablefor the biosynthesis of organic macromolecules. Moreover, inRpl11-deficient zebrafish, the expression of genes encodingenzymes that are involved in the biosynthesis of lipids andproteins is downregulated, whereas expression of genes involvedin catabolism is upregulated (Danilova et al., 2011). Changes in theexpression of genes involved in metabolism have also been found inp53-negative RP-deficient cells (Aspesi et al., 2014).

Autophagy is a mechanism used by normal cells, but especiallyby stressed cells, to replenish their energy and biosyntheticintermediates. Consistently, disrupted ribosome biogenesis leadsto the activation of autophagy both in human cells and in zebrafishembryos (Heijnen et al., 2014; Boglev et al., 2013).

Expression of genes encoding enzymes involved in ROSdetoxification, such as superoxide dismutase 2, is decreased both inRpl11 zebrafish mutants and in p53-negative RPS19-deficient humancell lines (Aspesi et al., 2014; Danilova et al., 2011). These resultsindicate that RP-depleted cells are predisposed to oxidative stress.

A common metabolic change in DBA is the upregulation of ADA(Fargo et al., 2013). Overexpression of ADA causes ATP depletion(Chen and Mitchell, 1994), which was also found in RP-deficientzebrafish (Danilova et al., 2014). ATP shortage leads to theactivation of AMPK, which inhibits biosynthesis and translation(Hardie et al., 2012). Translation is decreased in RP-deficient cells(Cmejlova et al., 2006; Gazda et al., 2006a; Signer et al., 2014)despite the overstimulation of mTOR and S6 kinase that is found inthese cells (Heijnen et al., 2014; Payne et al., 2012).

Another important metabolic change in RP-deficient cells is thedysregulation of the insulin pathway. Recent data suggest that RPdeficiency leads to the inhibition of this pathway through amechanism that is reminiscent of insulin resistance (Heijnen et al.,2014). Increased glucose levels and the upregulation of pre-proinsulin are also observed in zebrafish heterozygous for the rpl11mutation (Danilova et al., 2011). Known processes contributing tothe development of insulin resistance include p53 activation,increased ROS and increased production of proinflammatorycytokines (Minamino et al., 2009).

Structural changesMultiple changes in the expression of genes that encode structuralproteins have been found in an rpl11 zebrafish mutant, such as thedownregulation of collagens and the membrane components of redblood cells (Danilova et al., 2011). In DBA patients, the compositionof the membranes of red blood cells also changes, with theaccumulation of non-red blood cell proteins (Pesciotta et al., 2014).

Activation of the innate immune systemThe upregulation of genes involved in interferon and TNFα (tumornecrosis factor alpha) signaling has been reported in red blood cell

1019

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 8: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

progenitors and in fibroblasts obtained from DBA patients (Avondoet al., 2009; Gazda et al., 2006a), as well as in Rpl11-deficientzebrafish (Danilova et al., 2011). TNFα signaling has been found tocontribute to the hematopoietic failure observed in RPS19-deficienthuman hematopoietic progenitors and in Rps19-deficient zebrafishembryos (Bibikova et al., 2014). In RP-deficient zebrafish, theupregulation of several components of the complement system andthe downregulation of complement inhibitors have been reported(Danilova et al., 2011; Jia et al., 2013). Notably, monocytes fromthe bone marrow of DBA, DC and SDS patients have an alteredresponse to bacterial antigens (Matsui et al., 2013).

CancerThe increased incidence of cancer in an RP-deficient environmentmight be attributed to upregulation both of factors that induceapoptosis and factors that promote proliferation, as was reportedfor zebrafish rpl11 mutants (Danilova et al., 2011). Also, inhematopoietic progenitors from DBA patients, some members ofthe RAS family are upregulated, whereas the tumor suppressorsbreast cancer 2, early onset (BRCA2) and retinoblastoma 1 (RB1)are downregulated (Gazda et al., 2006a). The RP-deficientcondition therefore resembles an after-irradiation environment inwhich both pro-apoptotic and survival factors are upregulated;such an environment confers a dramatic selective advantage top53-deficient cells that avoid cell cycle arrest and apoptosis(Marusyk et al., 2010). Selection for p53-negative cells takes placeat the level of progenitor cells, which are the most sensitive tostress. Interestingly, in RP-deficient zebrafish tumors, the tp53gene is wild type and transcribed, but the p53 protein is notsynthesized (MacInnes et al., 2008). The mechanism of theselective suppression of p53 translation in these tumors is currentlyunknown.

Failure of erythropoiesis in DBAErythroid failure is the most distinctive feature of DBA, withdecreased proliferation of erythroid progenitors (Ebert et al., 2005;Flygare et al., 2005) and their increased sensitivity to apoptosis(Perdahl et al., 1994). In RP-deficient erythroid progenitors, p53 isselectively activated (Dutt et al., 2011). Hematopoietic defectsappear already at the stage of hematopoietic stem cells (HSCs).Their number in RP-deficient zebrafish is decreased (Danilovaet al., 2011) and induced pluripotent stem cells derived from DBApatients exhibit impaired differentiation (Garcon et al., 2013). In thissection, we discuss factors that might contribute to erythroid defectsin DBA.

Decreased GATA1 levelsMutations in the erythroid-specific gene GATA1 (GATA bindingprotein 1) leads to a phenotype that is currently classified as DBA(Sankaran et al., 2012). Although individuals with GATA1mutations do not have congenital malformations, their erythroidphenotype is very similar to that of DBA-affected individuals withRP mutations. Recent studies demonstrated that RP knockdown inprimary human hematopoietic progenitors or in an erythroid cellline leads to decreased GATA1 translation (Bibikova et al., 2014;Ludwig et al., 2014). A decreased level of Gata1 protein was alsofound in zebrafish mutant for rpl11 (Bibikova et al., 2014).Moreover, a global gene expression profiling of erythroid cells fromDBA patients showed that the expression of GATA1 transcriptionaltarget genes is downregulated in these cells (Ludwig et al., 2014),which is consistent with decreased GATA1 activity. These datasuggest that a decreased level of GATA1 protein underlies many

defects of erythroid cells in DBA (Ludwig et al., 2014). In supportof this, when GATA1 protein levels were increased in bone marrowmononuclear cells from DBA patients or in primary humanhematopoietic cells with reduced levels of RPL11 or RPL5, redblood cell production and cell differentiation was improved(Ludwig et al., 2014). The mechanism of GATA1 suppression inDBA might involve the highly structured 5′ end of GATA1 mRNAthat requires stringent conditions for translation initiation (Ludwiget al., 2014). These findings show that GATA1, although not theonly target of translation dysregulation in DBA, is an importantfactor in mediating the erythroid-specific defect observed in thiscondition.

Increased proliferation causes increased demand for ribosomesIt has been suggested that, because the chromatin of erythroid cellsbecomes condensed and transcriptionally inactive prior toenucleation, the rapidly proliferating immature erythroid cellsrequire very high ribosome synthesis rates in order to produceenough ribosomes to last for the duration of the mature erythrocytelife cycle (Sieff et al., 2010). This hypothesis was confirmed in fetalliver cells of RP-deficient mice, in which cell numbers increasedthree to fourfold while RNA content increased sixfold, suggestingan accumulation of an excess of ribosomes during earlyerythropoiesis (Sieff et al., 2010).

In addition, rapid proliferation itself might make cells vulnerableto stressful conditions. It is known that irradiation does the mostdamage to hematopoietic and epithelial tissues, including thethymus, spleen and small intestine, where cell turnover is the fastest;this sensitivity correlates with the increased induction of p53 inthese tissues (Gottlieb et al., 1997; Komarova et al., 1997).

Erythroid-specific alteration in transcription, splicing or translationSome data suggest that erythroid progenitors respond to stress byselective changes in transcription, splicing or in the translation of aspecific set of genes. The decreased transcription of a crucialerythropoietic factor, MYB (v-myb avian myeloblastosis viraloncogene homolog), has been reported in erythroid progenitorsfrom DBA patients (Gazda et al., 2006a) and in Rpl11-deficientzebrafish (Danilova et al., 2011). Myb protein levels are alsoreduced in RP-deficient mouse fetal liver cells (Sieff et al., 2010).

Splicing is a process that can be selectively affected under stress(Dutertre et al., 2011). Immature erythroid cells from DBA patientsexpress alternatively spliced, non-functional isoforms of FLVCR1(feline leukemia virus subgroup C cellular receptor 1), a hemeexporter required for erythropoiesis (Rey et al., 2008). These resultssuggest that FLVCR1 insufficiency contributes to the erythropoieticdefects seen in individuals with DBA.

There are indications of increased translational suppression oferythroid-specific genes in RP-deficient organisms in comparisonto other genes. In addition to the selective suppression of GATA1translation discussed above, selective suppression of globintranslation has been reported in RP-deficient zebrafish (Zhanget al., 2014). Decreased protein levels of some common factors thatare normally highly expressed in differentiating erythroid cells,such as BCL2-associated athanogene (BAG1), which encodes aHSP70 co-chaperone, and cold shock domain containing E1,RNA-binding (CSDE1), have been reported in mouse RP-deficienterythroblasts and erythroblasts cultured from DBA patients (Horoset al., 2012).

Additional suppression of GATA1 protein levels might bemediated through the upregulation of TNFα in RP-deficient cells(Bibikova et al., 2014).

1020

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 9: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

Metabolic and structural vulnerability of erythroid cellsIn addition, red blood cells have a distinct physiology that mightmake them selectively vulnerable to alterations in several commonpathways. An example is the shift from glycolysis to aerobicrespiration, which takes place in all cells after p53 upregulation(Matoba et al., 2006), including in RP-deficient cells (Danilovaet al., 2011). Although this shift occurs in all cells of RP-deficientorganisms, only erythrocytes rely almost exclusively on glycolysisand they are, therefore, selectively affected by this change.Many membrane and cytoskeleton changes seen in DBA animal

models are not specific to erythroid cells but occur in all tissues(Danilova et al., 2011). However, defects originating from thesechanges might particularly affect erythrocytes because of theespecially high requirements for a strong and deformable cellstructure to withstand circulation in narrow blood capillaries(Steiner and Gallagher, 2007).Furthermore, erythroid cells mature in association with

macrophages; alterations in membranes of RP-deficient erythroidcells might lead to their increased elimination by macrophages.

Tissue specificity and variability of defects inribosomopathiesDBA is characterized by broad phenotypic variability; moreover,some first-degree relatives of DBA-affected individuals with RPS19mutations carry an identical mutation but display no observablephenotype (Willig et al., 1999). Congenital malformations are foundin ∼40% of DBA patients and they differ between affectedindividuals. No correlation between the type of mutation and thephenotype has been found among individuals with DBA withRPS19 mutations (Willig et al., 1999). Similar to DBA, thephenotype of TCS is highly variable and some mutation carriers,parents of patients, are only mildly affected (Dixon and Dixon,2004; Teber et al., 2004). These findings point to a role for geneticmodifiers in the penetrance and severity of both DBA and TSC, and

to the importance of a patient’s overall genetic background. Belowwe discuss the origin of specific defects and factors that might causephenotypic variability in ribosomopathies.

Role of ΔNp63 in DBA congenital malformationsExtra digits or phalanges are among the notable congenital defectsobserved in individuals with DBA (Halperin and Freedman, 1989)and in the Rpl24 mouse mutant (Oliver et al., 2004) (Fig. 1). Asdiscussed above in the section devoted to the phenotypes of DBApatients and animal models, eye defects are also a commonfeature. The combination of these defects suggests that they mightoriginate during development from a misbalance between neuraland non-neural ectoderm in the early embryo. Development ofnon-neural ectoderm is controlled by a member of the p53 proteinfamily, ΔNp63. Besides its role during development, ΔNp63 is ap53 target gene and is upregulated in response to p53 activation(Bourdon, 2007). In the gastrulating zebrafish embryos, ΔNp63expression is localized to the ventral side of the embryo and marksnon-neural ectoderm; its overexpression leads to the expansion ofnon-neural ectoderm and to the suppression of neural structures(Bakkers et al., 2002). Later in development ΔNp63 controls thedevelopment of limbs and its overactivity can lead to theduplication of limb structures (Mills et al., 1999). In Rps19-deficient zebrafish embryos, the area of ΔNp63 expressionexpands to the dorsal side (Danilova et al., 2008b) (Fig. 4A).This corresponds to the expansion of non-neural ectoderm into theneural field, as confirmed by hybridization with gata2, a marker ofnon-neural ectoderm (Fig. 4A). Shrinkage of the neural fieldaffects mostly the forebrain and eyes, as illustrated by theexpression of pax2, which has a key role in the development of theCNS, eyes, urogenital tract and kidneys (Krauss et al., 1991)(Fig. 4B,C).

It would be interesting to know whether ΔNp63 upregulationcontributes to eye defects and to overgrown wings in Drosophila

Fig. 4. Origin of developmental defects in RP-deficient zebrafish embryos. (A) ΔNp63 expression (black arrow, upper panels) in early zebrafish embryosdefines the non-neural ectoderm field and overlaps with amarker of non-neural ectoderm, gata2 (black arrow, lower panels). In Rps19-deficient zebrafish [in whichrps19 expression has been knocked down with a morpholino oligonucleotide (MO)], this field is expanded (right upper and lower panels). Staining with a probe forgoosecoid (gsc), necessary for the formation of the dorsoventral axis of the embryo, marks the dorsal side (red arrow). Arrowheads point to the neural field. This isan in situ hybridization image at gastrulation, 80% epiboly (Box 1). Dorsal is to the right. wt, wild type. (B) Expression of pax2, which has a key role in thedevelopment of the CNS, eyes, urogenital tract and kidneys, is altered in Rps19-deficient zebrafish embryos. Arrows and arrowheads point, respectively, toforebrain and eye fields, which are contracted in Rps19-deficient embryos. This is an in situ hybridization image at 16 hpf. (C) Schematics showing howexpansionof non-neural ectoderm in early zebrafish embryos leads to the contraction of the neural field, especially the area of the forebrain and eye. This research wasoriginally published in Blood (Danilova et al., 2008b). © American Society of Hematology.

1021

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 10: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

Minute mutants, or to the eye defects and the duplication of fingersand phalanges in mice and humans with RP deficiency.ΔNp63 is a part of the p53 protein family network, which

involves hundreds of proteins that are differentially regulated invarious tissues during development (Danilova et al., 2008a).Because both environmental and genetic factors can modulate thisnetwork, ΔNp63 and other members of this network mightcontribute to phenotypic variation of ribosomopathies. Amultitude of other modifiers might theoretically affect cellularresponse to ribosomal stress, as was recently discussed (Farrar andDahl, 2011).

Mechanisms causing dissimilar phenotypes in DBA and TCSp53 activation contributes to the clinical features of both DBA andTCS. However, these two diseases have completely differentphenotypes. In contrast to the bone marrow failure and varioustissue defects seen in DBA, TSC defects are restricted to thecraniofacial tissues that originate from the neural crest. Most cells inthe embryo seem to survive TCOF1 or PolI/III insufficienciesdespite their crucial role in ribosome biogenesis. To explain theparadox, it was hypothesized that different progenitor cellpopulations at defined points during embryonic development havediverse ribosome requirements (James et al., 2014).Another explanation for this paradox might reside in the

spatiotemporal expression patterns of the affected genes. Inzebrafish embryos, tcof1 is expressed at a high level at the one-cell stage, which suggests the maternal origin of its mRNA (Weineret al., 2012). Moreover, tcof1 has the highest level of expression atthis stage. Therefore, enough protein can be produced in the embryofrom the maternal mRNA to pass through the early stages ofdevelopment. Neural crest cells in zebrafish develop and start tomigrate relatively late in development, at ∼15 hours post-fertilization (hpf ) through 24 hpf, and form the craniofacialskeleton, the most affected structure in TCS, even later, at∼72 hpf (Schilling and Kimmel, 1994). At this stage, all zebrafishtissues and organs, including HSCs, are already formed and,therefore, p53 activation can affect only craniofacial tissues. polr1d,the zebrafish ortholog of another gene mutated in TCS, is alsoexpressed maternally in zebrafish embryos [Thisse, B., Thisse, C.(2004) Fast Release Clones: A High Throughput ExpressionAnalysis. ZFIN Direct Data Submission (http://zfin.org)].In contrast to the genes mutated in TCS, only a trace expression of

rpl11 mRNA can be detected at the one-cell stage in zebrafishembryos; it increases from approximately 6 hpf, pointing to thebeginning of Rpl11 production (Danilova et al., 2011). p53upregulation in RP-deficient zebrafish can also be detected from∼6 hpf (Danilova et al., 2008b), pointing to the onset of problemswith ribosome biogenesis at this early time point. Early onset of p53upregulation in RP-deficient embryos would affect the formation ofall tissues and organs. The difference in timing of p53 activationmight therefore account for the multitude of defects in DBA versusthe relatively restricted phenotype of TCS.

Specialized ribosomes as drivers of tissue-specific defectsRecent studies suggest that composition of ribosomes can varybetween tissues. In yeast, many RPs are duplicated and paralogshave different functional roles (Komili et al., 2007). Based on thesedata, Komili et al. proposed the existence of many different forms offunctionally distinct ribosomes. Recent studies revealed that someanimal RPs also have paralogs with specifically tailored functions.Zebrafish Rpl22 and Rpl22L1 play different roles duringhematopoiesis (Zhang et al., 2013). Knockdown of rpl22 blocks

the development of T cells, whereas knockdown of rpl22l1 impairsthe development of HSCs. Mice also have Rpl22 and Rpl22L1paralogs; Rpl22L1 is expressed at a low level in all tissues exceptpancreas and its expression is regulated by Rpl22 (O’Leary et al.,2013). A recent study found that other RPs and their paralogs aredifferentially expressed in mouse embryos; the more recentlyevolved RP paralogs showed a much greater level of tissue-specificexpression (Wong et al., 2014).

Not only RPs but other genes involved in ribosome biogenesisalso show tissue-specific expression. Cirhin, mutations in whichcause childhood cirrhosis (NAIC), is expressed in embryonic miceat the highest levels in the liver, with much weaker expression inother tissues (Chagnon et al., 2002). In zebrafish, cirhin expressionis also high in the developing liver (Wilkins et al., 2013).

There might be other ways of creating a different ribosome.For example, RPs, rRNAs and various accessory factors involved inpre-rRNA processing are modified post-transcriptionally and,hypothetically, variations of these processes could also createaltered ribosomes. A comprehensive review of these and othermechanisms of the formation of specialized ribosomes have beenrecently presented (Xue and Barna, 2012).

If ribosomes differ among tissues then mutations in some genesinvolved in ribosome biogenesis could selectively affect only asubset of ribosomes.

Current and potential therapeutic approaches to DBAIn spite of recent progress in understanding the pathophysiologyof DBA, its treatment is still based on corticosteroids, red bloodcell transfusions and HSC transplantation (Vlachos et al., 2014).DBA is a complex disease; it is rather a tree of pathologies. Theonly way to tackle its cause is with gene therapy and thispossibility is being investigated. Replacement of the defectiveRPs in experimental systems has produced positive results(Flygare et al., 2008; Garcon et al., 2013). This approach holdspromise for the future. Until then, the individual consequences ofRP deficiency can be targeted.

Themost damaging effect of RP deficiency is the upregulation ofp53. Direct suppression of p53 in DBA is often consideredunacceptable owing to an increased risk of cancer. However,this approach might still be feasible, as illustrated by the effectsof cenersen in 5q-MDS. Cenersen is a 20-mer antisenseoligonucleotide complementary to TP53 exon 10; it suppressesp53 expression and restores erythropoiesis in 5q-MDS (Cacereset al., 2013). Zebrafish data suggest that corticosteroids might act,in part, by decreasing the expression of p53 (Danilova et al., 2011).

An indirect way of suppressing p53 activation is suggested byzebrafish studies. Treatment of RP-deficient zebrafish with amixture of nucleosides decreased the levels of p53, improvedhematopoiesis and diminished developmental defects (Danilovaet al., 2014). The efficiency of nucleoside treatment can beattributed to the fact that stress changes metabolism. Non-stressedcells use de novo nucleotide synthesis to make their DNA; therefore,very few nucleosides from food end up in the DNA of healthypeople. Because stressed cells start to use salvage pathways (Austinet al., 2012), nucleoside supplementation improves outcomes inphysiologically stressed patients (Hess and Greenberg, 2012).Nucleoside treatment in general decreases replication stress andDNA damage (Bester et al., 2011). Nucleosides in infant formulashave a positive effect on development (Singhal et al., 2010).Nucleoside mixtures are now available as over-the-countersupplements. Their application might benefit DBA patients duringthe acute phase of the disease.

1022

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 11: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

Inflammation is another damaging result of RP deficiency. Anti-inflammatory properties of lenalidomide, an immunomodulatoryagent used to treat several types of cancer, might be responsible forits positive effects in RP-deficient models (Keel et al., 2012; Narlaet al., 2011). The TNFα inhibitor etanercept also improves thecondition of RP-deficient zebrafish (Bibikova et al., 2014). Theupregulation of the complement system in zebrafish DBA models(Danilova et al., 2011; Jia et al., 2013) suggests that complementinhibitors might also merit investigation as a potential treatment forDBA. The careful consideration of the timing of treatments thattarget p53 or inflammation will still be necessary because thesepathways usually antagonize each other (Ak and Levine, 2010).An important metabolic change in RP-deficient cells that

deserves extra attention is the dysregulation of the insulinpathway. Individuals with DBA might have a condition similarto that of pre-diabetes, which could be targeted by thecorresponding drugs.Another important metabolic change, the decrease in translation,

has been recently targeted by L-leucine treatment in a DBA patient(Pospisilova et al., 2007). The treatment resulted in an increase inreticulocyte count and hemoglobin levels. Leucine treatmentstimulates mTOR and improves anemia in a mouse model ofDBA (Jaako et al., 2012). It also decreases anemia anddevelopmental defects in RP-deficient zebrafish embryos andincreases proliferation of Rps19- and Rps14-deficient erythroidprogenitors (Payne et al., 2012). It is not known, however, what theeffect of mTOR overactivation, especially for prolonged periods oftime, might be. Several ongoing clinical trials will determine thesafety of this treatment in individuals with DBA.Conflicting signaling seems to be a characteristic of RP-deficient

cells, such as the activation of p53 and inflammation. With manymodifiers involved, pathways leading from RP deficiency toerythroid defects in one patient might not be exactly the same asin another patient. Moreover, these pathways might differ in thesame patient between the acute and remission phases of the disease.Therefore, individualized therapy is likely to be necessary for thesuccessful treatment of DBA.

ConclusionA tremendous amount of knowledge about DBA has beenaccumulated since the first discovery of RP mutation in DBA.DBA studies have stimulated the field of ribosome biogenesis,resulting in a better understanding of the mechanisms of pre-rRNAprocessing and of ribosome subunit formation. The roles of the p53and p53-independent pathways in DBA have also become betterdefined.Multiple diseases have been found to originate from defectsin ribosome biogenesis. However, treatment options lag behind thisnewfound knowledge. The reason for this might lie in thecomplexity of ribosomopathies, for which a multitude ofpathways are involved. Ribosomopathy is a different disease inevery patient owing to the effects of genetic background andenvironment. The factors that modulate an individual response todeficiency of RP or other factors involved in ribosome biogenesisare still not well understood. As such, we need a betterunderstanding of the pathways downstream of p53, as well as themechanisms of the p53-independent factors that influence theetiology of these diseases.

Competing interestsThe authors declare no competing or financial interests.

Author contributionsN.D. and H.T.G. prepared and edited the manuscript.

FundingThis work was supported by National Institutes of Health grants R01 HL107558 andK02 HL111156 to H.T.G.

ReferencesAk, P. and Levine, A. J. (2010). P53 and NF-kB: different strategies for responding

to stress lead to a functional antagonism. FASEB J. 24, 3643-3652.Amsterdam, A., Nissen, R. M., Sun, Z., Swindell, E. C., Farrington, S. and

Hopkins, N. (2004a). Identification of 315 genes essential for early zebrafishdevelopment. Proc. Natl. Acad. Sci. USA 101, 12792-12797.

Amsterdam, A., Sadler, K. C., Lai, K., Farrington, S., Bronson, R. T., Lees, J. A.and Hopkins, N. (2004b). Many ribosomal protein genes are cancer genes inzebrafish. PLoS Biol. 2, e139.

Armistead, J. and Triggs-Raine, B. (2014). Diverse diseases from a ubiquitousprocess: the ribosomopathy paradox. FEBS Lett. 588, 1491-1500.

Aspesi, A., Pavesi, E., Robotti, E., Crescitelli, R., Boria, I., Avondo, F., Moniz, H.,Da Costa, L., Mohandas, N., Roncaglia, P. et al. (2014). Dissecting thetranscriptional phenotype of ribosomal protein deficiency: implications forDiamond-Blackfan anemia. Gene 545, 282-289.

Austin, W. R., Armijo, A. L., Campbell, D. O., Singh, A. S., Hsieh, T., Nathanson,D., Herschman, H. R., Phelps, M. E., Witte, O. N., Czernin, J. et al. (2012).Nucleoside salvage pathway kinases regulate hematopoiesis by linkingnucleotide metabolism with replication stress. J. Exp. Med. 209, 2215-2228.

Avondo, F., Roncaglia, P., Crescenzio, N., Krmac, H., Garelli, E., Armiraglio, M.,Castagnoli, C., Campagnoli, M. F., Ramenghi, U., Gustincich, S. et al. (2009).Fibroblasts from patients with Diamond-Blackfan anaemia show abnormalexpression of genes involved in protein synthesis, amino acid metabolism andcancer. BMC Genomics 10, 442.

Bakkers, J., Hild, M., Kramer, C., Furutani-Seiki, M. and Hammerschmidt, M.(2002). Zebrafish DeltaNp63 is a direct target of Bmp signaling and encodes atranscriptional repressor blocking neural specification in the ventral ectoderm.Dev. Cell 2, 617-627.

Barkic, M., Crnomarkovic, S., Grabusic, K., Bogetic, I., Panic, L., Tamarut, S.,Cokaric, M., Jeric, I., Vidak, S. and Volarevic, S. (2009). The p53 tumorsuppressor causes congenital malformations in Rpl24-deficient mice andpromotes their survival. Mol. Cell. Biol. 29, 2489-2504.

Barlow, J. L., Drynan, L. F., Hewett, D. R., Holmes, L. R., Lorenzo-Abalde, S.,Lane, A. L., Jolin, H. E., Pannell, R., Middleton, A. J., Wong, S. H. et al. (2010).A p53-dependent mechanism underlies macrocytic anemia in a mouse model ofhuman 5q-syndrome. Nat. Med. 16, 59-66.

Bermejo, R., Lai, M. S. and Foiani, M. (2012). Preventing replication stress tomaintain genome stability: resolving conflicts between replication andtranscription. Mol. Cell 45, 710-718.

Bernstein, K. A., Bleichert, F., Bean, J. M., Cross, F. R. and Baserga, S. J.(2007). Ribosome biogenesis is sensed at the start cell cycle checkpoint. Mol.Biol. Cell 18, 953-964.

Bester, A. C., Roniger, M., Oren, Y. S., Im, M. M., Sarni, D., Chaoat, M.,Bensimon, A., Zamir, G., Shewach, D. S. and Kerem, B. (2011). Nucleotidedeficiency promotes genomic instability in early stages of cancer development.Cell 145, 435-446.

Bibikova, E., Youn, M.-Y., Danilova, N., Ono-Uruga, Y., Konto-Ghiorghi, Y.,Ochoa, R., Narla, A., Glader, B., Lin, S. and Sakamoto, K. M. (2014). TNF-mediated inflammation represses GATA1 and activates p38 map kinase inRPS19-deficient hematopoietic progenitors. Blood 124, 3791-3798.

Boglev, Y., Badrock, A. P., Trotter, A. J., Du, Q., Richardson, E. J., Parslow,A. C., Markmiller, S. J., Hall, N. E., de Jong-Curtain, T. A., Ng, A. Y. et al.(2013). Autophagy induction is a TOR- and p53-independent cell survivalresponse in a zebrafish model of disrupted ribosome biogenesis. PLoS Genet. 9,e1003279.

Bolze, A., Mahlaoui, N., Byun, M., Turner, B., Trede, N., Ellis, S. R., Abhyankar,A., Itan, Y., Patin, E., Brebner, S. et al. (2013). Ribosomal protein SAhaploinsufficiency in humans with isolated congenital asplenia. Science 340,976-978.

Boocock, G. R. B., Morrison, J. A., Popovic, M., Richards, N., Ellis, L., Durie,P. R. and Rommens, J. M. (2003). Mutations in SBDS are associated withShwachman-Diamond syndrome. Nat. Genet. 33, 97-101.

Boulon, S., Westman, B. J., Hutten, S., Boisvert, F.-M. and Lamond, A. I. (2010).The nucleolus under stress. Mol. Cell 40, 216-227.

Bourdon, J.-C. (2007). p53 family isoforms. Curr. Pharm. Biotechnol. 8, 332-336.Bursac, S., Brdovcak, M. C., Pfannkuchen, M., Orsolic, I., Golomb, L., Zhu, Y.,

Katz, C., Daftuar, L., Grabusic, K., Vukelic, I. et al. (2012). Mutual protection ofribosomal proteins L5 and L11 from degradation is essential for p53 activationupon ribosomal biogenesis stress. Proc. Natl. Acad. Sci. USA 109, 20467-20472.

Caceres, G., McGraw, K., Yip, B. H., Pellagatti, A., Johnson, J., Zhang, L., Liu,K., Zhang, L. M., Fulp,W. J., Lee, J.-H. et al. (2013). TP53 suppression promoteserythropoiesis in del(5q) MDS, suggesting a targeted therapeutic strategy inlenalidomide-resistant patients. Proc. Natl. Acad. Sci. USA 110, 16127-16132.

Chagnon, P., Michaud, J., Mitchell, G., Mercier, J., Marion, J.-F., Drouin, E.,Rasquin-Weber, A., Hudson, T. J. and Richter, A. (2002). A missense mutation

1023

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 12: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

(R565w) in cirhin (FLJ14728) in North American Indian childhood cirrhosis.Am. J. Hum. Genet. 71, 1443-1449.

Chakraborty, A., Uechi, T., Higa, S., Torihara, H. and Kenmochi, N. (2009). Lossof ribosomal protein L11 affects zebrafish embryonic development through a p53-dependent apoptotic response. PLoS ONE 4, e4152.

Chen, E. H. and Mitchell, B. S. (1994). Hereditary overexpression of adenosinedeaminase in erythrocytes: studies in erythroid cell lines and transgenic mice.Blood 84, 2346-2353.

Chiocchetti, A., Gibello, L., Carando, A., Aspesi, A., Secco, P., Garelli, E.,Loreni, F., Angelini, M., Biava, A., Dahl, N. et al. (2005). Interactions betweenRPS19, Mutated in Diamond-Blackfan anemia, and the PIM-1 oncoprotein.Haematologica 90, 1453-1462.

Choesmel, V., Bacqueville, D., Rouquette, J., Noaillac-Depeyre, J., Fribourg, S.,Cretien, A., Leblanc, T., Tchernia, G., Da Costa, L. and Gleizes, P.-E. (2007).Impaired ribosome biogenesis in Diamond-Blackfan anemia. Blood 109,1275-1283.

Choesmel, V., Fribourg, S., Aguissa-Toure, A.-H., Pinaud, N., Legrand, P.,Gazda, H. T. and Gleizes, P.-E. (2008). Mutation of ribosomal protein RPS24 inDiamond-Blackfan Anemia results in a ribosome biogenesis disorder. Hum. Mol.Genet. 17, 1253-1263.

Cmejla, R., Cmejlova, J., Handrkova, H., Petrak, J. and Pospisilova, D. (2007).Ribosomal protein S17 gene (RPS17) is mutated in Diamond-Blackfan anemia.Hum. Mutat. 28, 1178-1182.

Cmejla, R., Cmejlova, J., Handrkova, H., Petrak, J., Petrtylova, K., Mihal, V.,Stary, J., Cerna, Z., Jabali, Y. and Pospisilova, D. (2009). Identification ofmutations in the ribosomal protein L5 (RPL5) and ribosomal protein L11 (RPL11)genes in Czech patients with Diamond-Blackfan anemia. Hum. Mutat. 30,321-327.

Cmejlova, J., Dolezalova, L., Pospisilova, D., Petrtylova, K., Petrak, J. andCmejla, R. (2006). Translational efficiency in patients with Diamond-Blackfananemia. Haematologica 91, 1456-1464.

Danilova, N., Sakamoto, K. M. and Lin, S. (2008a). P53 family in development.Mech. Dev. 125, 919-931.

Danilova, N., Sakamoto, K. M. and Lin, S. (2008b). Ribosomal protein S19deficiency in zebrafish leads to developmental abnormalities and defectiveerythropoiesis through activation of p53 protein family. Blood 112, 5228-5237.

Danilova, N., Sakamoto, K. and Lin, S. (2011). Ribosomal protein L11 mutation inzebrafish leads to haematopoietic and metabolic defects. Br. J. Haematol. 152,217-228.

Danilova, N., Bibikova, E., Covey, T. M., Nathanson, D., Dimitrova, E., Konto, Y.,Lindgren, A., Glader, B., Radu, C. G., Sakamoto, K. M. et al. (2014). The role ofthe DNA damage response in zebrafish and cellular models of Diamond Blackfananemia. Dis. Model. Mech. 7, 895-905.

Dauwerse, J. G., Dixon, J., Seland, S., Ruivenkamp, C. A. L., van Haeringen, A.,Hoefsloot, L. H., Peters, D. J. M., Boers, A. C.-D, Daumer-Haas, C., Maiwald,R. et al. (2011). Mutations in genes encoding subunits of RNA Polymerases I andIII cause Treacher Collins syndrome. Nat. Genet. 43, 20-22.

De Keersmaecker, K., Atak, Z. K., Li, N., Vicente, C., Patchett, S., Girardi, T.,Gianfelici, V., Geerdens, E., Clappier, E., Porcu, M. et al. (2013). Exomesequencing identifies mutation in CNOT3 and ribosomal genes RPL5 and RPL10in T-cell acute lymphoblastic leukemia. Nat. Genet. 45, 186-190.

Devlin, E. E., DaCosta, L., Mohandas, N., Elliott, G. and Bodine, D. M. (2010). Atransgenic mouse model demonstrates a dominant negative effect of a pointmutation in the Rps19 gene associated with Diamond-Blackfan anemia. Blood116, 2826-2835.

Dixon, J. and Dixon, M. J. (2004). Genetic background has a major effect on thepenetrance and severity of craniofacial defects in mice heterozygous for the geneencoding the nucleolar protein Treacle. Dev. Dyn. 229, 907-914.

Dixon, J., Jones, N. C., Sandell, L. L., Jayasinghe, S. M., Crane, J., Rey, J.-P.,Dixon, M. J. and Trainor, P. A. (2006). Tcof1/Treacle is required for neural crestcell formation and proliferation deficiencies that cause craniofacial abnormalities.Proc. Natl. Acad. Sci. USA 103, 13403-13408.

Doherty, L., Sheen, M. R., Vlachos, A., Choesmel, V., O’Donohue, M.-F.,Clinton, C., Schneider, H. E., Sieff, C. A., Newburger, P. E., Ball, S. E. et al.(2010). Ribosomal protein genes RPS10 and RPS26 are commonly mutated inDiamond-Blackfan anemia. Am. J. Hum. Genet. 86, 222-228.

Donati, G., Bertoni, S., Brighenti, E., Vici, M., Trere, D., Volarevic, S.,Montanaro, L. and Derenzini, M. (2011). The balance between rRNA andribosomal protein synthesis up- and downregulates the tumour suppressor p53 inmammalian cells. Oncogene 30, 3274-3288.

Draptchinskaia, N., Gustavsson, P., Andersson, B., Pettersson, M., Willig, T.-N., Dianzani, I., Ball, S., Tchernia, G., Klar, J., Matsson, H. et al. (1999). Thegene encoding ribosomal protein S19 is mutated in Diamond-Blackfan anaemia.Nat. Genet. 21, 169-175.

Dutertre, M., Sanchez, G., Barbier, J., Corcos, L. and Auboeuf, D. (2011). Theemerging role of pre-messenger RNA splicing in stress responses: sendingalternative messages and silent messengers. RNA Biol. 8, 740-747.

Dutt, S., Narla, A., Lin, K., Mullally, A., Abayasekara, N., Megerdichian, C.,Wilson, F. H., Currie, T., Khanna-Gupta, A., Berliner, N. et al. (2011).

Haploinsufficiency for ribosomal protein genes causes selective activation of p53in human erythroid progenitor cells. Blood 117, 2567-2576.

Ebert, B. L., Lee, M. M., Pretz, J. L., Subramanian, A., Mak, R., Golub, T. R. andSieff, C. A. (2005). An RNA interference model of RPS19 deficiency in Diamond-Blackfan anemia recapitulates defective hematopoiesis and rescue bydexamethasone: identification of dexamethasone-responsive genes bymicroarray. Blood 105, 4620-4626.

Ebert, B. L., Pretz, J., Bosco, J., Chang, C. Y., Tamayo, P., Galili, N., Raza, A.,Root, D. E., Attar, E., Ellis, S. R. et al. (2008). Identification of RPS14 as a 5q-syndrome gene by RNA interference screen. Nature 451, 335-339.

Elghetany, M. T. and Alter, B. P. (2002). P53 protein overexpression in bonemarrow biopsies of patients with Shwachman-Diamond syndrome has aprevalence similar to that of patients with refractory anemia. Arch. Pathol. Lab.Med. 126, 452-455.

Ellis, S. R. (2014). Nucleolar stress in Diamond Blackfan anemia pathophysiology.Biochim. Biophys. Acta 1842, 765-768.

Fargo, J. H., Kratz, C. P., Giri, N., Savage, S. A., Wong, C., Backer, K., Alter, B. P.and Glader, B. (2013). Erythrocyte adenosine deaminase: diagnostic value forDiamond-Blackfan anaemia. Br. J. Haematol. 160, 547-554.

Farrar, J. E. and Dahl, N. (2011). Untangling the phenotypic heterogeneity ofDiamond Blackfan anemia. Semin. Hematol. 48, 124-135.

Farrar, J. E., Nater, M., Caywood, E., McDevitt, M. A., Kowalski, J., Takemoto,C. M., Talbot, C. C., Jr, Meltzer, P., Esposito, D., Beggs, A. H. et al. (2008).Abnormalities of the large ribosomal subunit protein, RPL35A, in Diamond-Blackfan anemia. Blood 112, 1582-1592.

Farrar, J. E., Vlachos, A., Atsidaftos, E., Carlson-Donohoe, H., Markello, T. C.,Arceci, R. J., Ellis, S. R., Lipton, J. M. and Bodine, D. M. (2011). Ribosomalprotein gene deletions in Diamond-Blackfan anemia. Blood 118, 6943-6951.

Farrar, J. E., Quarello, P., Fisher, R., O’Brien, K. A., Aspesi, A., Parrella, S.,Henson, A. L., Seidel, N. E., Atsidaftos, E., Prakash, S. et al. (2014). Exploitingpre-rRNA processing in Diamond Blackfan anemia gene discovery and diagnosis.Am. J. Hematol. 89, 985-991.

Flygare, J., Kiefer, T., Miyake, K., Utsugisawa, T., Hamaguchi, I., Da Costa, L.,Richter, J., Davey, E. J., Matsson, H., Dahl, N. et al. (2005). Deficiency ofribosomal protein S19 in CD34+ cells generated by siRNA blocks erythroiddevelopment and mimics defects seen in Diamond-Blackfan anemia. Blood 105,4627-4634.

Flygare, J., Aspesi, A., Bailey, J. C., Miyake, K., Caffrey, J. M., Karlsson, S. andEllis, S. (2007). Human RPS19, the gene mutated in Diamond-Blackfan anemia,encodes a ribosomal protein required for the maturation of 40s ribosomalsubunits. Blood 109, 980-986.

Flygare, J., Olsson, K., Richter, J. and Karlsson, S. (2008). Gene therapy ofDiamond Blackfan anemia CD34(+) cells leads to improved erythroiddevelopment and engraftment following transplantation. Exp. Hematol. 36,1428-1435.

Fumagalli, S. and Thomas, G. (2011). The role of P53 in ribosomopathies. Semin.Hematol. 48, 97-105.

Fumagalli, S., Di Cara, A., Neb-Gulati, A., Natt, F., Schwemberger, S., Hall, J.,Babcock, G. F., Bernardi, R., Pandolfi, P. P. and Thomas, G. (2009). Absenceof nucleolar disruption after impairment of 40S ribosome biogenesis reveals anRPL11-translation-dependent mechanism of p53 induction. Nat. Cell Biol. 11,501-508.

Garcon, L., Ge, J., Manjunath, S. H., Mills, J. A., Apicella, M., Parikh, S.,Sullivan, L. M., Podsakoff, G. M., Gadue, P., French, D. L. et al. (2013).Ribosomal and hematopoietic defects in induced pluripotent stem cells derivedfrom Diamond Blackfan anemia patients. Blood 122, 912-921.

Gazda, H. T., Zhong, R., Long, L., Niewiadomska, E., Lipton, J. M., Ploszynska,A., Zaucha, J. M., Vlachos, A., Atsidaftos, E., Viskochil, D. H. et al. (2004).RNA and protein evidence for haplo-insufficiency in Diamond-Blackfan anaemiapatients with RPS19 mutations. Br. J. Haematol. 127, 105-113.

Gazda, H. T., Kho, A. T., Sanoudou, D., Zaucha, J. M., Kohane, I. S., Sieff, C. A.and Beggs, A. H. (2006a). Defective ribosomal protein gene expression alterstranscription, translation, apoptosis, and oncogenic pathways in Diamond-Blackfan anemia. Stem Cells 24, 2034-2044.

Gazda, H. T., Grabowska, A., Merida-Long, L. B., Latawiec, E., Schneider, H. E.,Lipton, J. M., Vlachos, A., Atsidaftos, E., Ball, S. E., Orfali, K. A. et al. (2006b).Ribosomal protein S24 gene is mutated in Diamond-Blackfan anemia.Am. J. Hum. Genet. 79, 1110-1118.

Gazda, H. T., Sheen, M. R., Vlachos, A., Choesmel, V., O’Donohue, M.-F.,Schneider, H., Darras, N., Hasman, C., Sieff, C. A., Newburger, P. E. et al.(2008). Ribosomal protein L5 and L11 mutations are associated with cleft palateand abnormal thumbs in Diamond-Blackfan anemia patients. Am. J. Hum. Genet.83, 769-780.

Gazda, H. T., Preti, M., Sheen, M. R., O’Donohue, M.-F., Vlachos, A., Davies,S. M., Kattamis, A., Doherty, L., Landowski, M., Buros, C. et al. (2012).Frameshift mutation in p53 regulator RPL26 is associated with multiple physicalabnormalities and a specific pre-ribosomal RNA processing defect in Diamond-Blackfan anemia. Hum. Mutat. 33, 1037-1044.

Golomb, L., Volarevic, S. and Oren, M. (2014). p53 and ribosome biogenesisstress: the essentials. FEBS Lett. 588, 2571-2579.

1024

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 13: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

Gottlieb, E., Haffner, R., King, A., Asher, G., Gruss, P., Lonai, P. and Oren, M.(1997). Transgenic mousemodel for studying the transcriptional activity of the p53protein: age- and tissue-dependent changes in radiation-induced activation duringembryogenesis. EMBO J. 16, 1381-1390.

Gripp, K. W., Curry, C., Olney, A. H., Sandoval, C., Fisher, J., Chong, J. X.-L.,Pilchman, L., Sahraoui, R., Stabley, D. L. andSol-Church, K. (2014). Diamond-Blackfan anemia with mandibulofacial dystostosis is heterogeneous, including thenovel DBA genes TSR2 and RPS28. Am. J. Med. Genet. A 164, 2240-2249.

Gu, B.-W., Bessler, M. and Mason, P. J. (2008). A pathogenic dyskerin mutationimpairs proliferation and activates a DNA damage response independent oftelomere length in mice. Proc. Natl. Acad. Sci. USA 105, 10173-10178.

Halperin, D. S. and Freedman, M. H. (1989). Diamond-Blackfan anemia: etiology,pathophysiology, and treatment. Am. J. Pediatr. Hematol. Oncol. 11, 380-394.

Hardie, D. G., Ross, F. A. and Hawley, S. A. (2012). AMPK: a nutrient and energysensor that maintains energy homeostasis. Nat. Rev. Mol. Cell. Biol. 13, 251-262.

Heijnen, H. F., van Wijk, R., Pereboom, T. C., Goos, Y. J., Seinen, C. W., vanOirschot, B. A., van Dooren, R., Gastou, M., Giles, R. H., van Solinge, W. et al.(2014). Ribosomal protein mutations induce autophagy through S6 kinaseinhibition of the insulin pathway. PLoS Genet. 10, e1004371.

Henras, A. K., Plisson-Chastang, C., O’Donohue, M.-F., Chakraborty, A. andGleizes, P.-E. (2015). An overview of pre-ribosomal RNA processing ineukaryotes. Wiley Interdiscip. Rev. RNA 6, 225-242.

Hess, J. R. andGreenberg, N. A. (2012). The role of nucleotides in the immune andgastrointestinal systems: potential clinical applications. Nutr. Clin. Pract. 27,281-294.

Horos, R., Ijspeert, H., Pospisilova, D., Sendtner, R., Andrieu-Soler, C.,Taskesen, E., Nieradka, A., Cmejla, R., Sendtner, M., Touw, I. P. et al.(2012). Ribosomal deficiencies in Diamond-Blackfan anemia impair translation oftranscripts essential for differentiation of murine and human erythroblasts. Blood119, 262-272.

Iadevaia, V., Caldarola, S., Biondini, L., Gismondi, A., Karlsson, S., Dianzani, I.and Loreni, F. (2010). PIM1 kinase is destabilized by ribosomal stress causinginhibition of cell cycle progression. Oncogene 29, 5490-5499.

Jaako, P., Flygare, J., Olsson, K., Quere, R., Ehinger, M., Henson, A., Ellis, S.,Schambach, A., Baum, C., Richter, J. et al. (2011). Mice with ribosomal proteinS19 deficiency develop bone marrow failure and symptoms like patients withDiamond-Blackfan anemia. Blood 118, 6087-6096.

Jaako, P., Debnath, S., Olsson, K., Bryder, D., Flygare, J. and Karlsson, S.(2012). Dietary L-leucine improves the anemia in a mouse model for Diamond-Blackfan anemia. Blood 120, 2225-2228.

James, A., Wang, Y., Raje, H., Rosby, R. and DiMario, P. (2014). Nucleolar stresswith and without p53. Nucleus 5, 402-426.

Jia, Q., Zhang, Q., Zhang, Z., Wang, Y., Zhang, W., Zhou, Y., Wan, Y., Cheng, T.,Zhu, X., Fang, X. et al. (2013). Transcriptome analysis of the zebrafish model ofDiamond-Blackfan anemia from RPS19 deficiency via p53-dependent and-independent pathways. PLoS ONE 8, e71782.

Jones, N. C., Lynn, M. L., Gaudenz, K., Sakai, D., Aoto, K., Rey, J.-P., Glynn,E. F., Ellington, L., Du, C., Dixon, J. et al. (2008). Prevention of theneurocristopathy Treacher Collins syndrome through Inhibition of p53 function.Nat. Med. 14, 125-133.

Keel, S. B., Phelps, S., Sabo, K. M., O’Leary, M. N., Kirn-Safran, C. B. andAbkowitz, J. L. (2012). Establishing Rps6 hemizygous mice as a model forstudying how ribosomal protein haploinsufficiency impairs erythropoiesis. Exp.Hematol. 40, 290-294.

Komarova, E. A., Chernov, M. V., Franks, R., Wang, K., Armin, G., Zelnick, C. R.,Chin, D. M., Bacus, S. S., Stark, G. R. and Gudkov, A. V. (1997). Transgenicmice with p53-responsive lacZ: p53 activity varies dramatically during normaldevelopment and determines radiation and drug sensitivity in vivo. EMBO J. 16,1391-1400.

Komili, S., Farny, N. G., Roth, F. P. and Silver, P. A. (2007). functional specificityamong ribosomal proteins regulates gene expression. Cell 131, 557-571.

Kondrashov, N., Pusic, A., Stumpf, C. R., Shimizu, K., Hsieh, A. C., Xue, S.,Ishijima, J., Shiroishi, T. and Barna, M. (2011). Ribosome-mediated specificityin Hox mRNA translation and vertebrate tissue patterning. Cell 145, 383-397.

Kongsuwan, K., Yu, Q., Vincent, A., Frisardi, M. C., Rosbash, M., Lengyel, J. A.and Merriam, J. (1985). A Drosophila minute gene encodes a ribosomal protein.Nature 317, 555-558.

Krauss, S., Johansen, T., Korzh, V. and Fjose, A. (1991). Expression of thezebrafish paired box gene pax[Zf-B] during early neurogenesis. Development113, 1193-1206.

Kressler, D., Hurt, E. and Bassler, J. (2010). Driving ribosome assembly.Biochim.Biophys. Acta 1803, 673-683.

Landowski, M., O’Donohue, M.-F., Buros, C., Ghazvinian, R., Montel-Lehry, N.,Vlachos, A., Sieff, C. A., Newburger, P. E., Niewiadomska, E., Matysiak, M.et al. (2013). Novel deletion of RPL15 identified by array-comparative genomichybridization in Diamond-Blackfan anemia. Hum. Genet. 132, 1265-1274.

Landry, D. M., Hertz, M. I. and Thompson, S. R. (2009). RPS25 is essential fortranslation initiation by the Dicistroviridae and hepatitis C viral IRESs.Genes Dev.23, 2753-2764.

Lee, A. S.-Y., Burdeinick-Kerr, R. and Whelan, S. P. J. (2013). A ribosome-specialized translation initiation pathway is required for cap-dependent translationof vesicular stomatitis virus mRNAs. Proc. Natl. Acad. Sci. USA 110, 324-329.

Leger-Silvestre, I., Caffrey, J. M., Dawaliby, R., Alvarez-Arias, D. A., Gas, N.,Bertolone, S. J., Gleizes, P.-E. and Ellis, S. R. (2005). Specific role for yeasthomologs of the Diamond Blackfan anemia-associated Rps19 protein in ribosomesynthesis. J. Biol. Chem. 280, 38177-38185.

Li, X. and Manley, J. L. (2005). Inactivation of the SR protein splicing factor ASF/SF2 results in genomic instability. Cell 122, 365-378.

Ludwig, L. S., Gazda, H. T., Eng, J. C., Eichhorn, S.W., Thiru, P., Ghazvinian, R.,George, T. I., Gotlib, J. R., Beggs, A. H., Sieff, C. A. et al. (2014). Alteredtranslation of GATA1 in Diamond-Blackfan anemia. Nat. Med. 20, 748-753.

Lunt, S. Y. and Vander Heiden, M. G. (2011). Aerobic glycolysis: meeting themetabolic requirements of cell proliferation. Annu. Rev. Cell. Dev. Biol. 27,441-464.

Ma, H. and Pederson, T. (2013). The nucleolus stress response is coupled to anATR-CHK1-mediated G2 arrest. Mol. Biol. Cell 24, 1334-1342.

Macias, E., Jin, A., Deisenroth, C., Bhat, K., Mao, H., Lindstrom, M. S. andZhang, Y. (2010). An ARF-independent c-MYC-activated tumor suppressionpathway mediated by ribosomal protein-Mdm2 interaction. Cancer Cell 18,231-243.

MacInnes, A. W., Amsterdam, A., Whittaker, C. A., Hopkins, N. and Lees, J. A.(2008). Loss of p53 synthesis in zebrafish tumors with ribosomal protein genemutations. Proc. Natl. Acad. Sci. USA 105, 10408-10413.

Marusyk, A., Porter, C. C., Zaberezhnyy, V. and DeGregori, J. (2010). Irradiationselects for p53-deficient hematopoietic progenitors. PLoS Biol. 8, e1000324.

Marygold, S. J., Coelho, C. M. A. and Leevers, S. J. (2005). Genetic analysis ofRpl38 and Rpl5, two minute genes located in the centric heterochromatin ofchromosome 2 of Drosophila melanogaster. Genetics 169, 683-695.

Matoba, S., Kang, J.-G., Patino, W. D., Wragg, A., Boehm, M., Gavrilova, O.,Hurley, P. J., Bunz, F. and Hwang, P. M. (2006). p53 regulates mitochondrialrespiration. Science 312, 1650-1653.

Matsui, K., Giri, N., Alter, B. P. and Pinto, L. A. (2013). cytokine production bybone marrow mononuclear cells in inherited bone marrow failure syndromes.Br. J. Haematol. 163, 81-92.

McGowan, K. A., Li, J. Z., Park, C. Y., Beaudry, V., Tabor, H. K., Sabnis, A. J.,Zhang, W., Fuchs, H., de Angelis, M. H., Myers, R. M. et al. (2008). Ribosomalmutations cause p53-mediated dark skin and pleiotropic effects. Nat. Genet. 40,963-970.

Mills, A. A., Zheng, B.,Wang, X.-J., Vogel, H., Roop, D. R. andBradley, A. (1999).p63 is a p53 homologue required for limb and epidermal morphogenesis. Nature398, 708-713.

Minamino, T., Orimo, M., Shimizu, I., Kunieda, T., Yokoyama, M., Ito, T., Nojima,A., Nabetani, A., Oike, Y., Matsubara, H. et al. (2009). A crucial role for adiposetissue p53 in the regulation of insulin resistance. Nat. Med. 15, 1082-1087.

Miner, J. H. and Li, C. (2000). Defective glomerulogenesis in the absence of lamininalpha5 demonstrates a developmental role for the kidney glomerular basementmembrane. Dev. Biol. 217, 278-289.

Mirabello, L., Macari, E. R., Jessop, L., Ellis, S. R., Myers, T., Giri, N., Taylor,A. M., McGrath, K. E., Humphries, J. M., Ballew, B. J. et al. (2014). Whole-exome sequencing and functional studies identify RPS29 as a novel genemutated in multicase Diamond-Blackfan anemia families. Blood 124, 24-32.

Miyake, K., Utsugisawa, T., Flygare, J., Kiefer, T., Hamaguchi, I., Richter, J. andKarlsson, S. (2008). Ribosomal Protein S19 deficiency leads to reducedproliferation and increased apoptosis but does not affect terminal erythroiddifferentiation in a cell line model of Diamond-Blackfan anemia. Stem Cells 26,323-329.

Narla, A., Dutt, S., McAuley, J. R., Al-Shahrour, F., Hurst, S., McConkey, M.,Neuberg, D. and Ebert, B. L. (2011). Dexamethasone and lenalidomide havedistinct functional effects on erythropoiesis. Blood 118, 2296-2304.

Nilsson, J., Sengupta, J., Frank, J. and Nissen, P. (2004). Regulation ofeukaryotic translation by the RACK1 protein: a platform for signallingmolecules onthe ribosome. EMBO Rep. 5, 1137-1141.

O’Leary, M. N., Schreiber, K. H., Zhang, Y., Duc, A.-C. E., Rao, S., Hale, J. S.,Academia, E. C., Shah, S. R., Morton, J. F., Holstein, C. A. et al. (2013). Theribosomal protein Rpl22 controls ribosome composition by directly repressingexpression of its own paralog, Rpl22l1. PLoS Genet. 9, e1003708.

Oliver, E. R., Saunders, T. L., Tarle, S. A. andGlaser, T. (2004). Ribosomal proteinL24 defect in belly spot and tail (Bst), a mouse Minute. Development 131,3907-3920.

Osheim, Y. N., French, S. L., Keck, K. M., Champion, E. A., Spasov, K., Dragon,F., Baserga, S. J. and Beyer, A. L. (2004). Pre-18S ribosomal RNA is structurallycompacted into the SSU processome prior to being cleaved from nascenttranscripts in Saccharomyces cerevisiae. Mol. Cell 16, 943-954.

Panic, L., Tamarut, S., Sticker-Jantscheff, M., Barkic, M., Solter, D., Uzelac, M.,Grabusic, K. and Volarevic, S. (2006). Ribosomal protein S6 genehaploinsufficiency is associated with activation of a p53-dependent checkpointduring gastrulation. Mol. Cell. Biol. 26, 8880-8891.

Payne, E. M., Virgilio, M., Narla, A., Sun, H., Levine, M., Paw, B. H., Berliner, N.,Look, A. T., Ebert, B. and Khanna-Gupta, A. (2012). L-Leucine improves the

1025

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms

Page 14: Ribosomopathies: how a common root can cause a tree of … · defects in ribosome biogenesis (Armistead and Triggs-Raine, 2014; James et al., 2014). Despite this common defect, phenotypes

anemia and developmental defects associated with Diamond-Blackfan anemiaand del(5q)MDS by activating the mTOR pathway. Blood 120, 2214-2224.

Pellagatti, A., Hellstrom-Lindberg, E., Giagounidis, A., Perry, J., Malcovati, L.,Della Porta, M. G., Jadersten, M., Killick, S., Fidler, C., Cazzola, M. et al.(2008). Haploinsufficiency of RPS14 in 5q-syndrome is associated withderegulation of ribosomal- and translation-related genes. Br. J. Haematol. 142,57-64.

Perdahl, E. B., Naprstek, B. L., Wallace, W. C. and Lipton, J. M. (1994). Erythroidfailure in Diamond-Blackfan anemia is characterized by apoptosis. Blood 83,645-650.

Pereboom, T. C., van Weele, L. J., Bondt, A. and MacInnes, A. W. (2011). Azebrafish model of dyskeratosis congenita reveals hematopoietic stem cellformation failure resulting from ribosomal protein-mediated p53 stabilization.Blood 118, 5458-5465.

Pesciotta, E. N., Sriswasdi, S., Tang, H.-Y., Speicher, D. W., Mason, P. J. andBessler, M. (2014). Dysferlin and other non-red cell proteins accumulate in the redcell membrane of Diamond-Blackfan anemia patients. PLoS ONE 9, e85504.

Pestov, D. G., Strezoska, Z. and Lau, L. F. (2001). Evidence of p53-dependentcross-talk between ribosome biogenesis and the cell cycle: effects of nucleolarprotein Bop1 on G(1)/S transition. Mol. Cell. Biol. 21, 4246-4255.

Pospisilova, D., Cmejlova, J., Hak, J., Adam, T. and Cmejla, R. (2007).Successful treatment of a Diamond-Blackfan anemia patient with amino acidleucine. Haematologica 92, e66-e67.

Quarello, P., Garelli, E., Brusco, A., Carando, A., Mancini, C., Pappi, P., Vinti, L.,Svahn, J., Dianzani, I. and Ramenghi, U. (2012). High frequency of ribosomalprotein gene deletions in italian Diamond-Blackfan anemia patients detected bymultiplex ligation-dependent probe amplification assay. Haematologica 97,1813-1817.

Rey, M. A., Duffy, S. P., Brown, J. K., Kennedy, J. A., Dick, J. E., Dror, Y. andTailor, C. S. (2008). Enhanced alternative splicing of the FLVCR1 gene inDiamond Blackfan anemia disrupts FLVCR1 expression and function that arecritical for erythropoiesis. Haematologica 93, 1617-1626.

Robledo, S., Idol, R. A., Crimmins, D. L., Ladenson, J. H., Mason, P. J. andBessler, M. (2008). The role of human ribosomal proteins in the maturation ofrRNA and ribosome production. RNA 14, 1918-1929.

Rubbi, C. P. and Milner, J. (2003). Disruption of the nucleolus mediatesstabilization of p53 in response to DNA damage and other stresses. EMBO J.22, 6068-6077.

Ruggero, D. and Shimamura, A. (2014). Marrow failure: a window into ribosomebiology. Blood 124, 2784-2792.

Ruggero, D., Grisendi, S., Piazza, F., Rego, E., Mari, F., Rao, P. H., Cordon-Cardo, C. and Pandolfi, P. P. (2003). Dyskeratosis Congenita and cancer in micedeficient in ribosomal RNA modification. Science 299, 259-262.

Sanchez, A., Sharma, S., Rozenzhak, S., Roguev, A., Krogan, N. J., Chabes, A.and Russell, P. (2012). Replication fork collapse and genome instability in adeoxycytidylate deaminase mutant. Mol. Cell. Biol. 32, 4445-4454.

Sankaran, V. G., Ghazvinian, R., Do, R., Thiru, P., Vergilio, J.-A., Beggs, A. H.,Sieff, C. A., Orkin, S. H., Nathan, D. G., Lander, E. S. et al. (2012). Exomesequencing identifies GATA1 mutations resulting in Diamond-Blackfan anemia.J. Clin. Invest. 122, 2439-2443.

Schilling, T. F. and Kimmel, C. B. (1994). Segment and cell type lineagerestrictions during pharyngeal arch development in the zebrafish embryo.Development 120, 483-494.

Schneider, D. A., Michel, A., Sikes, M. L., Vu, L., Dodd, J. A., Salgia, S., Osheim,Y. N., Beyer, A. L. and Nomura, M. (2007). Transcription elongation by RNApolymerase i is linked to efficient rRNA processing and ribosome assembly. Mol.Cell 26, 217-229.

Sieff, C. A., Yang, J., Merida-Long, L. B. and Lodish, H. F. (2010). Pathogenesisof the erythroid failure in Diamond Blackfan anaemia. Br. J. Haematol. 148,611-622.

Signer, R. A. J., Magee, J. A., Salic, A. and Morrison, S. J. (2014).Haematopoietic stem cells require a highly regulated protein synthesis Rate.Nature 509, 49-54.

Singh, S. A., Goldberg, T. A., Henson, A. L., Husain-Krautter, S., Nihrane, A.,Blanc, L., Ellis, S. R., Lipton, J. M. and Liu, J. M. (2014). p53-independent cellcycle and erythroid differentiation defects in murine embryonic stem cellshaploinsufficient for Diamond Blackfan anemia-proteins: RPS19 versus RPL5.PLoS ONE 9, e89098.

Singhal, A., Kennedy, K., Lanigan, J., Clough, H., Jenkins, W., Elias-Jones, A.,Stephenson, T., Dudek, P. and Lucas, A. (2010). Dietary nucleotides and earlygrowth in formula-fed infants: a randomized controlled trial. Pediatrics 126,e946-e953.

Sloan, K. E., Bohnsack, M. T. andWatkins, N. J. (2013). The 5S RNP couples p53homeostasis to ribosome biogenesis and nucleolar stress. Cell Rep. 5, 237-247.

Steiner, L. A. and Gallagher, P. G. (2007). Erythrocyte disorders in the perinatalperiod. Semin. Perinatol. 31, 254-261.

Sulic, S., Panic, L., Barkic, M., Mercep, M., Uzelac, M. and Volarevic, S. (2005).Inactivation of S6 ribosomal protein gene in T lymphocytes activates a p53-dependent checkpoint response. Genes Dev. 19, 3070-3082.

Taylor, A. M., Humphries, J. M., White, R. M., Murphey, R. D., Burns, C. E. andZon, L. I. (2012). Hematopoietic defects in Rps29 mutant zebrafish depend uponp53 activation. Exp. Hematol. 40, 228-237.e5.

Teber, O. A., Gillessen-Kaesbach, G., Fischer, S., Bohringer, S., Albrecht, B.,Albert, A., Arslan-Kirchner, M., Haan, E., Hagedorn-Greiwe, M., Hammans, C.et al. (2004). Genotyping in 46 patients with tentative diagnosis of TreacherCollins syndrome revealed unexpected phenotypic variation. Eur. J. Hum. Genet.12, 879-890.

Teng, T., Mercer, C. A., Hexley, P., Thomas, G. and Fumagalli, S. (2013). Loss oftumor suppressor RPL5/RPL11 does not induce cell cycle arrest but impedesproliferation due to reduced ribosome content and translation capacity.Mol. Cell.Biol. 33, 4660-4671.

Torihara, H., Uechi, T., Chakraborty, A., Shinya, M., Sakai, N. and Kenmochi, N.(2011). Erythropoiesis failure due to Rps19 deficiency is independent of anactivated tp53 response in a zebrafish model of Diamond-Blackfan anaemia.Br. J. Haematol. 152, 648-654.

Uechi, T., Nakajima, Y., Chakraborty, A., Torihara, H., Higa, S. andKenmochi, N.(2008). Deficiency of ribosomal protein S19 during early embryogenesis leads toreduction of erythrocytes in a zebrafishmodel of Diamond-Blackfan anemia.Hum.Mol. Genet. 17, 3204-3211.

Valdez, B. C., Henning, D., So, R. B., Dixon, J. and Dixon, M. J. (2004). TheTreacher Collins syndrome (TCOF1) gene product is involved in ribosomal DNAgene transcription by interacting with upstream binding factor. Proc. Natl. Acad.Sci. USA 101, 10709-10714.

Vlachos, A., Rosenberg, P. S., Atsidaftos, E., Alter, B. P. and Lipton, J. M.(2012). Incidence of neoplasia in Diamond Blackfan anemia: a report from theDiamond Blackfan anemia registry. Blood 119, 3815-3819.

Vlachos, A., Blanc, L. and Lipton, J. M. (2014). Diamond blackfan anemia: amodel for the translational approach to understanding human disease. ExpertRev. Hematol. 7, 359-372.

Wang, R., Yoshida, K., Toki, T., Sawada, T., Uechi, T., Okuno, Y., Sato-Otsubo,A., Kudo, K., Kamimaki, I., Kanezaki, R. et al. (2015). Loss of functionmutationsin RPL27 and RPS27 identified by whole-exome sequencing in Diamond-Blackfan anaemia. Br. J. Haematol. 168, 854-864.

Warner, J. R. and McIntosh, K. B. (2009). How common are extraribosomalfunctions of ribosomal proteins? Mol. Cell 34, 3-11.

Watkins-Chow, D. E., Cooke, J., Pidsley, R., Edwards, A., Slotkin, R., Leeds,K. E., Mullen, R., Baxter, L. L., Campbell, T. G., Salzer, M. C. et al. (2013).Mutation of the Diamond-Blackfan anemia gene Rps7 in mouse results inmorphological and neuroanatomical phenotypes. PLoS Genet. 9, e1003094.

Weiner, A. M. J., Scampoli, N. L. and Calcaterra, N. B. (2012). Fishing themolecular bases of Treacher Collins syndrome. PLoS ONE 7, e29574.

Wilkins, B. J., Lorent, K., Matthews, R. P. and Pack, M. (2013). p53-mediatedbiliary defects caused by knockdown of Cirh1a, the zebrafish homolog of the generesponsible for North American Indian childhood cirrhosis. PLoS ONE 8, e77670.

Willig, T. N., Draptchinskaia, N., Dianzani, I., Ball, S., Niemeyer, C., Ramenghi,U., Orfali, K., Gustavsson, P., Garelli, E., Brusco, A. et al. (1999). Mutations inribosomal protein S19 gene and Diamond Blackfan anemia: wide variations inphenotypic expression. Blood 94, 4294-4306.

Wong, Q. W.-L., Li, J., Ng, S. R., Lim, S. G., Yang, H. and Vardy, L. A. (2014).RPL39L is an example of a recently evolved ribosomal protein paralog that showshighly specific tissue expression patterns and is upregulated in ESCs and HCCTumors. RNA Biol. 11, 33-41.

Xue, S. and Barna, M. (2012). Specialized ribosomes: a new frontier in generegulation and organismal biology. Nat. Rev. Mol. Cell. Biol. 13, 355-369.

Zhang, Y. and Lu, H. (2009). Signaling to p53: ribosomal proteins find their way.Cancer Cell 16, 369-377.

Zhang, Y., Morimoto, K., Danilova, N., Zhang, B. and Lin, S. (2012). Zebrafishmodels for Dyskeratosis Congenita reveal critical roles of p53 activationcontributing to hematopoietic defects through RNA processing. PLoS ONE 7,e30188.

Zhang, Y., Duc, A.-C. E., Rao, S., Sun, X.-L., Bilbee, A. N., Rhodes, M., Li, Q.,Kappes, D. J., Rhodes, J. andWiest, D. L. (2013). Control of hematopoietic stemcell emergence by antagonistic functions of ribosomal protein paralogs. Dev. Cell24, 411-425.

Zhang, Y., Ear, J., Yang, Z., Morimoto, K., Zhang, B. and Lin, S. (2014). Defects ofprotein production in erythroid cells revealed in a zebrafish Diamond-BlackfanAnemia model for mutation in Rps19. Cell Death Dis. 5, e1352.

Zinzalla, V., Stracka, D., Oppliger, W. and Hall, M. N. (2011). Activation ofmTORC2 by association with the ribosome. Cell 144, 757-768.

1026

REVIEW Disease Models & Mechanisms (2015) 8, 1013-1026 doi:10.1242/dmm.020529

Disea

seModels&Mechan

isms