A ‘garbage can’ for ribosomes: how
eukaryotes degrade their ribosomes
Denis L.J. Lafontaine
1,2
1
Fonds de la Recherche Scientifique (FRS-F.N.R.S.), Institut de Biologie et de Me´ decine Mole´ culaire (IBMM), Universite´ Libre de
Bruxelles (ULB), Charleroi-Gosselies, Belgium
2
Center for Microscopy and Molecular Imaging (CMMI), Acade´ mie Wallonie Bruxelles, Charleroi-Gosselies, Belgium
Ribosome synthesis is a major metabolic activity that
involves hundreds of individual reactions, each of which
is error-prone. Ribosomal insults occur in cis (alteration
in rRNA sequences) and in trans (failure to bind to, or
loss of, an assembly factor or ribosomal protein). In
addition, specific growth conditions, such as starvation,
require that excess ribosomes are turned over efficiently.
Recent work indicates that cells evolved multiple strat-
egies to recognize specifically, and target for clearance,
ribosomes that are structurally and/or functionally
deficient, as well as in excess. This surveillance is active
at every step of the ribosome synthesis pathway and on
mature ribosomes, involves nearly entirely different
mechanisms for the small and large subunits, and
requires specialized subcellular organelles.
Ribosome synthesis is a multi-step, error-prone process
Ribosomes comprise two subunits of unequal size that
carry out specialized functions in translation: mRNA
decoding and peptidyl-transfer reaction for the small
and large subunits, respectively [1]. Each eukaryotic ribo-
some consists of 4 rRNAs and !80 ribosomal proteins. The
synthesis, maturation and transport of individual riboso-
mal components and their assembly into ribosomal sub-
units requires the intervention of !200 protein trans-
acting factors, and numerous small nucleolar RNAs (snoR-
NAs) that are involved in hundreds of individual, error-
prone, reactions [2,3] (Figure 1). Many ribosomal proteins
perform additional non ribosomal functions [4]. Likewise,
functions in processes not connected directly to ribosome
biogenesis are currently being assigned to ribosome syn-
thesis factors: connections to cell cycle progression, pre-
mRNA splicing, the DNA damage response, nuclear organ-
ization and telomere maintenance are emerging. With so
many reactions in the ribosome assembly pathway, the
possibility to introduce mistakes with potential deleterious
consequences for cell viability and human health is
immense. As such, failure to bind, or the loss of, a synthesis
factor could lead to the production of ribosomes that are
structurally defective (e.g., lacking individual or subsets of
ribosomal proteins or carrying misfolded rRNA) with func-
tional consequences in translation. To circumvent such
problems, cells have evolved multiple quality control mech-
anisms. For example, synthesis factors involved in late
cytoplasmic ribosome assembly steps can bind pre-rRNAs
at early nucleolar stages, thereby committing pre-ribo-
somes to productive synthesis pathways [5]. In other cases,
trans-acting factors with partial homology to ribosomal
proteins or translation factors might bind pre-ribosomes
to monitor and tether the structural integrity of ribosomal
protein-binding sites [6,7]. In wild-type cells, ribosome
assembly defects can result from the delayed binding of
a trans-acting factor. In these circumstances, and provid-
ing that the proper assembly reaction occurs within a
defined timeframe, faithful assembly presumably resumes;
otherwise, pre-ribosomes are identified as defective and
targeted for rapid degradation by active surveillance mech-
anisms. In addition to alterations in trans, mutations can
occur in cis either during RNA synthesis or, more fre-
quently, as a consequence of exposure to genotoxic stress.
The importance of ‘clearing the system’ of such mutations
is that they could otherwise accumulate to pathological
levels owing to the high stability (half-life of several days)
and abundance (up to 80% of total cellular RNA) of the
rRNAs. Indeed, human diseases, and in particular, neu-
rodegenerative disorders, have recently been linked to the
accumulation of such defective ribosomes [8]. Finally, there
are situations where excess ribosomes must be degraded to
recycle essential cellular building blocks; quite unexpect-
edly, this pathway also involves their specific recognition.
Recent research indicates that surveillance exerts its
action at every step of the ribosome assembly line, as well
as on the final product, that it involves dedicated subcel-
lular structures, and that distinctive pathways prevail for
monitoring the assembly and function of the small and
large ribosomal subunits. This multi-faceted ribosome
surveillance is the focus of this review (Figure 2).
Mutations in cis: non-functional ribosomal RNA decay
Several surveillance pathways have been described that
monitor the structural and functional integrity of mature
RNA molecules [9]. One such pathway that monitors
mRNAs is the ‘no-go’ decay pathway (NGD), wherein
mRNAs that induce stalled ribosomes are degraded [10].
To test whether mutations in functionally relevant and
conserved ribosomal sites affect rRNA stability, LaRivie
`
re
et al. introduced substitutions in the decoding site (DCS; in
18S rRNA) and peptidyl transferase centre (PTC; in 25S
rRNA) at positions that are essential for ribosome function
in bacteria [11] (Table S1). This led to the identification of
‘non functional rRNA decay’ or NRD [11], a pathway that
detects and eliminates functionally defective components
of mature ribosomes (Box 1). Indeed, in each case, these
Review
Corresponding author: Lafontaine, D.L.J. ([email protected]).
0968-0004/$ see front matter ! 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.tibs.2009.12.006 Available online 22 January 2010
267
changes resulted in a reduced accumulation (!5 to 10 fold)
of the mature rRNA containing the mutation.
Small ribosomal subunit NRD: a process reminiscent of
mRNA no-go decay
During mRNA NGD, a stalled ribosome triggers initiating
endonucleolytic cleavage events on the defective mRNA at
the vicinity of the pause site, and this is followed by the
exoribonucleolytic digestion of the 5’- and 3’-cleaved mRNA
products by the RNA exosome and the 5’!3’ exoRNase
Xrn1, respectively [10]. The RNA exosome is a conserved
multiprotein 3’!5’ exoRNase complex active in the syn-
thesis, degradation and surveillance of most classes of
cellular RNAs, as well as some viral RNAs [12] (Box 2).
Figure 1. Eukaryotic ribosome synthesis is a multi-step, error-prone process. Ribosome synthesis starts in the nucleolus where the core, and catalytically active, pieces of
the machinery, the rRNAs, are synthesized. Three out of the four rRNAs (18S, 5.8S and 25S rRNAs) are produced by RNA polymerase I (Pol I) from a single transcription unit.
The fourth rRNA (5S) is synthesized by RNA Pol III with a 3’-extension (inset). The 18S, 5.8S and 25S rRNAs are interspersed with non-coding sequences: the 5’- and 3’-
external transcribed spacers (5’-ETS and 3’-ETS) and internal transcribed spacers 1 and 2 (ITS1 and ITS2). An actively transcribed rDNA unit is depicted as a ‘Christmas tree’,
reminiscent of its visualization by Miller chromatin spread, with the ‘trunk’ of the ‘tree’ (brown) representing the rDNA locus and the ‘branches’ (green) corresponding to
nascent pre-rRNA transcripts. Pre-rRNA processing is initiated either post-transcriptionally (left of the ‘tree’) or co-transcriptionally (to the right). In fast-growing yeast cells,
up to 50-70% of nascent transcripts are cleaved co-transcriptionally at site A
2
within ITS1, generating pre-40S and pre-60S ribosomes [57] (other pre-rRNA cleavage sites are
described in Figure 3). In the remaining cases, a full-length transcript (35S) is generated and is assembled into a 90S pre-ribosome (not represented) that is cleaved post-
transcriptionally at site A
2
. Many ribosome synthesis factors escort pre-ribosomes from their initial site of synthesis in the nucleolus to their cytoplasmic site of function.
Different ribosome synthesis factors display distinctive patterns of association with pre-ribosomes, resulting in a progressive reduction in protein complexity. The nuclear
dwelling time for pre-40S and pre-60S subunits is distinctly different, with pre-40S subunits reaching the cytoplasm much faster than pre-60S. Translocation through the
various nucleolar domains is facilitated by specific trans-acting factors. Nuclear export involves redundant pathways. About 2000 ribosomes are exported every minute in
exponentially growing yeast cells, and control has been suggested to occur at the nuclear pore in response to nutrient availability [58,59]. There appears to be good timing
between nucleo-cytoplasmic partitioning and the acquisition of prominent ribosomal structural features. For example, final shaping of the ‘beak’, a protruding structure of
the small subunit, is thought to only occur once pre-40S has reached the cytoplasm [60]. Ribosome synthesis factors are available in limited quantities; therefore their
recycling is essential. Recent work identified a mechanochemical process powered by a AAA-type ATPase distantly related to the motor protein dynein involved in this
recycling [61]. Late sets of cytoplasmic trans-acting factors are released sequentially following further structural rearrangements fueled by a cascade of GTPase-mediated
reactions and recycled to the nucleus and nucleolus, ready to engage further rounds of assembly. The stable incorporation of each resident ribosomal protein is then
finalized. Major classes of trans-acting factors are listed. To illustrate the dynamics of ribosome synthesis and the plasticity of pre-ribosomes, several ribosome synthesis
factors (arbitrarily color-coded) are represented.
Review
Trends in Biochemical Sciences Vol.35 No.5
268
The possibility that stalling induced by defective ribosomes
might elicit a similar pathway was quite appealing and,
indeed, 18S NRD substrates are stabilized 2-fold in the
absence of the translation termination factor-like proteins
Dom34 and Hbs1 that are key components of NGD [13].
Strikingly, this is not the case for 25S NRD substrates,
thus indicating that mechanistically, NRD comprises at
least two distinct pathways.
During 18S NRD, Dom34 and Hbs1 act together in the
same pathway (their simultaneous inactivation is not
synergistic), and, consistently, the two proteins interact
functionally in vitro [14] and in vivo [15]. Small molecule
inhibitors of translation (cycloheximide and hygromycin B)
specifically stabilized 18S but not 25S NRD substrates [13],
thus providing further evidence that 18S NRD activation
requires elongating ribosomes, and that 18S and 25S
NRD are mechanistically different. The cytoplasmic
5’!3’ exoRNase Xrn1 and Ski7 (a cytoplasmic exosome
recruitment cofactor) have both been linked to NGD, and
also contribute to 18S NRD. Hbs1 and Ski7 are GTPases
homologous to the translation release factor eRF3. Delet-
ing both HBS1 and SKI7 enhanced the stabilization of 18S
Figure 2. The major pathways of eukaryotic ribosomal RNA decay. The five pathways of rRNA decay described to date. (i) Nucleolar and nuclear pre-40S and pre-60S
ribosomes are monitored actively by the ‘TRAMPexosome’ pathway during which misfolded pre-ribosomes are identified by TRAMP binding followed by polyadenylation
of the 3’-ends of defective precursor rRNAs in a step that stimulates both the recruitment and the decay activity of the exosome. It is not yet known how TRAMP detects
defective ribosomes. Polyadenylation occurs both at normal and cryptic pre-rRNA processing sites. Alterations in RNA Pol I activity, as well as in pre-rRNA processing
kinetics, can activate cryptic cleavage sites. Cytoplasmic mature subunits carrying cis mutations in functionally relevant ribosomal sites (and perhaps late cytoplasmic
precursors, Box 4) are monitored by NRD, a process that involves different mechanisms for the small and large subunits. (ii) During 18S NRD, small subunits deficient in
normal progression along the mRNA (‘stalling subunits’) are identified by the translation termination factor-related proteins Dom34 and Hbs1. They are probably cleaved
endonucleolytically by an unknown activity (by analogy to mRNA NGD), the released products are digested by Xrn1 and the RNA exosome, assisted by its cofactor Ski7 (Box
2). (iii) During 25S NRD, defective 60S subunits are targeted for proteasomal degradation by Rtt101Mms1-mediated ubiquitylation of unidentified, associated ribosomal
component(s). During conditions of nutrient deficiency and stress, excess ribosomes and pre-ribosomes are turned over by ribophagy and PMN, respectively. (iv)
Ribophagy is a specific form of macroautophagy that involves the engulfment and targeting of bulk cytosolic fractions to the vacuole by the de novo formation of an
‘isolation membrane.’ This membrane extends into an autophagosomal structure that fuses to the vacuolar membrane, and triggers the release of an autophagic body and
its subsequent degradation by resident vacuolar hydrolases. For 60S subunits, ribophagy involves the specific deubiquitylation of unidentified ribosomal components by
the Ubp3Bre5 complex. The effector(s) for 40S ribophagy remain unknown. (v) In PMN, a specific form of microautophagy, a portion of the nuclear envelope is ‘pinched
off’ by the vacuole, generating a specialized organelle termed a NVJ that comprises protein ‘velcro-like’ patches formed by the interaction between Nvj1 (small red circles)
and Vac8 (small green squares). The NVJ matures sequentially into a bulge, a tethered bleb, and finally a vesicle that is degraded by resident hydrolases. The late steps
(vesicle release) of PMN require several ATG genes also involved in macroautophagy [62]. Ribosome surveillance occurs at every step along the assembly pathway, starting
in the nucleolus, where a specialized locale coined the ‘No-body’ concentrates polyadenylated pre-rRNA species and exosomal components, continuing in the nucleoplasm
and ending in the cytoplasm. Cytoplasmic degradation involves several subcellular organelles, including P-bodies (where 18S NRD substrates, as well as mRNA NGD
substrates, are localized), the autophagosome, the proteasome and a cytosolic perinuclear compartment enriched for 25S NRD substrates. Key trans-acting factors involved
in each pathway are represented.
Review
Trends in Biochemical Sciences Vol.35 No.5
269
NRD substrates, indicating that the products of these two
genes probably operate in parallel pathways, possibly
competing for binding an empty-A site on a ribosome
stalled at a sense codon [13]. The core exosome (tested
for Rrp44, also called Dis3; Box 2) contributes, but is
apparently not rate-limiting, for 18S and 25S NRD [13].
18S NRD and mRNA NGD substrates both accumulate
in P-bodies that are conserved RNAprotein cytoplasmic
granules containing untranslated mRNAs complexed with
a set of translational repressors, the mRNA decapping
machinery and Xrn1 [16] . Strikingly, 25S NRD substrates
do not co-localize to P-bodies. Other differences between
18S and 25S NRD include the contribution of Xrn1 and
Ski7 to 18S, but not 25S, NRD [13,17].
It is quite striking that 18S rRNA NRD and mRNA NGD
operate with distinctively different kinetics (half-lives of up
!100 min and !10 min, respectively [10,13]). This differ-
ence might reflect the higher complexity and compaction of
mature ribosomal ribonucleoprotein particles (rRNP) ver-
sus mRNP substrates.
Large ribosomal subunit NRD: a role for ubiquitylation in
ribosome turnover
Fujii et al. conducted a genome-wide screen on the yeast
knock-out collection aimed at the identification of
mutations that stabilize 25S NRD substrates. This work
led to the discovery of Mms1, a component of an E3
ubiquitin ligase that had previously been characterized
as being involved in DNA repair [17]. There are two main
intracellular protein degradation systems in eukaryotes
[18]: the ubiquitin-proteasome system (UPS) that often
targets short-lived proteins and involves ubiquitylation
of the substrate and its targeting to the proteasome [18],
and autophagy, a process that leads to the degradation of
long-lived proteins and excess or aberrant organelles.
Originally identified as a protein involved in the repair
of DNA damage induced by the alkylating agent methyl
methanesulfonate (MMS), Mms1 belongs to the Mms22
module that comprises Mms1, Mms22, Rtt101 and Rtt107.
Rtt101, but not the other members of the complex, is also
required for 25S NRD [17]. Mutations in MMS1 and
RTT101 are not synergistic, indicating that their gene
products act together in the same pathway. Consistently,
Mms1 and Rtt101 interact in vitro and in vivo [19,20].
Strikingly, neither Mms1 nor Rtt101 is required for 18S
NRD, lending further credence to the idea that, mechan-
istically, NRD comprises distinct pathways.
Fujii et al. further demonstrated that the level of ubi-
quitylation observed in large ribosomal subunit fractions
increased strikingly in strains that expressed 25S NRD
rRNA, and that this modification required specifically
Mms1 and Rtt101. Although the identity of the ubiquity-
lated components remains unknown, the overexpression of
a ubiquitin variant that inhibits proteasomal function
demonstrated clearly the requirement of ribosomal com-
ponent ubiquitylation for 25S NRD.
25S NRD substrates accumulate in the cytoplasm in a
previously undescribed perinuclear compartment that, the
Box 1. When and where does NRD occur?
NRD does not depend upon rRNA synthesis, because mutations that
affect the accumulation of 18S rRNA do not impact the stability of
25S rRNA, and vice versa (both RNAs are co-expressed; Figure 1
inset). NRD is believed to occur primarily in the cytoplasm on
mature ribosomal subunits, and, perhaps, on late cytoplasmic
precursors (Box 4), for the following reasons. First, neither of the
mutations tested in the DCS or the PTC g rossly affect the
accumulation of pre-rRNA precursors. Second, NRD substrates
colocalize on velocity gradients with mature ribosomal subunit-
sized and/or polysomal fractions. Third, trans-acting factors in-
volved in NRD (Dom34, Hbs1, Xrn1 and Ski7) are all involved in
cytoplasmic processes, and neither the nuclear-specific exosomal
cofactor Rrp6 (Box 2) nor the nuclear 5’!3’ exoRNase Rat1 impact
18S or 25S NRD. Fourth, NRD substrates accumulate strikingly in
cytoplasmic structures (P-bodies for 18S NRD and cytosolic peri-
nuclear foci for 25S NRD). Fifth, NRD-decay occurs with much
slower kineti cs (with half-l ives ranging b etween !50-100 min
[11,13]) than nuclear decay (for example, in the absence of the
export factor Nmd3, nuclear 25S rRNA decays with a half-life of
!4 min [65]). Finally, 18S NRD requires ongoing translation, and
implies that the 18S rRNA is made, i.e. that the last and cytoplasmic
pre-rRNA cleavage has occurred.
Box 2. The exosome and TRAMP
The eukaryotic RNA exosome consists of a core of nine subunits, and
one or two associated RNases (Rrp44, also called Dis3, and Rrp6)
[12,26]. Six subunits of the core (homologs to bacterial phosphorolytic
RNase PH and PNPase, as well as the archaeal exosome) form a
structural inactive ‘ring’. The remaining three core subunits are S1/KH-
RNA binding proteins that provide a cap that bridges the ring subunits
together. Rrp44 is homologous to bacterial RNase II/R and is endowed
with both 3’!5’ exoRNase and endoRNase activities. Rrp44 endoRNase
activity might contribute to the progression of the exosome on its
substrates by resolving complex RNP structures. The exosome
functions in concert with many cofactors that provide substrate
specificity and modulate its functions. In the nucleus, the exosome
functions together with Rrp6 (homologous to bacterial RNase T/D),
another catalytically active 3’!5’ exoRNase subunit of the complex.
Rrp44 (Dis3) and Rrp6 are hydrolytic enzymes that share complemen-
tary biochemical properties: Rrp44 is a processive enzyme, whereas
Rrp6 is distributive. Rrp44 favors both structured and unstructured
substrates and poorly degrades poly(A) and AU-rich sequences in vitro,
whereas Rrp6 shows a marked preference for unstructured substrates
and for poly(A) and AU-rich sequences in vitro. Trf4 and Trf5 are
characterized by distinctive subcellular distribution (Trf5 is mostly
nucleolar; Trf4 is nuclear), cellular abundance (Trf4 is >3x more
abundant than Trf5), and requirement for normal growth (TRF5 deletion
has no associated phenotype whereas TRF4 loss confers a slow-growth
and cryo-sensitive phenotype). These differences probably reflect the
subtle substrate preferences of the two proteins. TRAMP-mediated
RNA polyadenylation stimulates nuclear RNA degradation; this con-
trasts with the role of canonical mRNA poly(A) tails that stimulate
mRNA export, stability and translation. The role of short poly(A) tails as
RNA ‘degradation tags’ has been compared to the role of RNA
oligoadenylation in bacteria and proposed to reflect its ancestral origin
[25,26]. The difference in function of the two types of poly(A) tails could
reflect the difference in processivity of the two RNA polymerases: Pap1
is highly processive, adding long tails of !250 adenines residues in
human cells and !60-90 in yeast, thus probably ensuring that no free 3’-
end is available until a long tail has been generated and covered by
poly(A) binding proteins, such as Pab1. By comparison, Trf4 and Trf5
display low processivity and are distributive (adding short tails of !15-
30 nucleotides) indicating that the 3’-ends of tails added by TRAMP
could be available frequently as exosome substrates.
Review
Trends in Biochemical Sciences Vol.35 No.5
270
authors suggest, could be related to the recently described
juxtanuclear quality control domain (JUNQ) that is
enriched for both misfolded proteins and proteasomes
[21]. Strikingly, localization of 25S NRD substrates to
perinuclear foci relies upon Mms1, indicating that it might
be the site of active degradation.
Mutations in trans: the TRAMP-exosome pathway and
nucleolar surveillance
It has long been appreciated that mutations in ribosome
synthesis factors that specifically inhibit nuclear pre-rRNA
processing reactions do not systematically lead to the
expected accumulation of pre-rRNA precursors. This find-
ing indicates that aberrant nucleolar and nuclear pre-
rRNAs are degraded actively by surveillance mechanisms.
This nuclear surveillance involves, at least in part, the
addition of unstructured oligoadenylate tails at the 3’-end
of flawed pre-rRNAs by a poly(A) polymerase activity
residing in TRAMP complexes, followed by their degra-
dation by the RNA exosome [22]. TRAMP complexes con-
sist of a poly(A) polymerase (either of the paralogous
proteins Trf4 or Trf5 in TRAMP4 and TRAMP5 complexes,
respectively), a zinc-knuckle-containing and putative
RNA-binding protein (either Air1 or Air2) and the DEVH
helicase Mtr4 (also called Dob1) [2325] (Box 2). In this
pathway, the addition of short poly(A) tails and/or the
actual binding of the TRAMP complex to the RNA commit
aberrant molecules to degradation by both discriminating
them from ‘normal’ RNA and by stimulating exosomal
activity [26].
The early steps of nucleolar pre-rRNA processing (clea-
vage at sites A
0
,A
1
and A
2
in yeast, Figure 3), that separate
the primary RNA polymerase I (Pol I) transcript into
precursors destined to the small and large subunits,
respectively, require a large RNP known as the SSU-
processome that consists of the box C+D snoRNA U3
and !40 associated proteins referred to as UTP (for U
Three-associated Proteins) [27,28]. The SSU-processome
comprises autonomous protein building blocks that are
loaded onto nascent pre-rRNAs and assemble into cataly-
tically active pre-rRNA processing complexes in a stepwise
and highly hierarchical process following alternative path-
ways (Figure 4). At least some of these autonomous
modules are evolutionarily conserved [29,30]. It is cur-
rently understood that failure to assemble the SSU-pro-
cessome with proper kinetics activates nucleolar
surveillance, as depletion of individual SSU-processome
components leads to early nucleolar pre-rRNA processing
inhibition (cleavage at sites A
0
-A
2
), the concomitant acti-
vation of a cryptic cleavage at site A
3
(further downstream
in internal transcribed spacer 1; ITS1, Figure 3) by the
endoRNase MRP, and the synthesis of the aberrant 23S
RNA. This 23S RNA is polyadenylated, mostly by
TRAMP5, and targeted for rapid degradation by the exo-
some (Figure 3) [31,32]. As TRAMP is a distributive
enzyme that adds short poly(A) tails, it is thought that
flawed RNAs undergo multiple rounds of TRAMP-
mediated polyadenylation, followed by exosomal digestion
to achieve greater degradation efficacy. Trf5 co-localizes
with the SSU-processome at the rDNA in living yeast cells,
indicating that this surveillance starts co-transcriptionally
on nascent pre-ribosomes [31]. Interestingly, low levels of
normal pre-rRNA intermediates are polyadenylated and
can be detected in strains defective for the exosome com-
ponent Rrp6 (Box 2) [33,34]. This indicates that every
available 3’-end is probably polyadenylated by TRAMP,
and that physiological pre-rRNA processing sites are also
used as entry points for rRNA surveillance. Moreover,
stabilized pre-rRNA fragments that terminate at multiple
positions within the coding sequence of the 18S rRNA can
be detected in surveillance-defective strains [31,32]. Thus,
several cryptic pre-rRNA processing sites appear to be
activated under conditions of defective ribosome assembly.
The accumulation of such fragments is also observed in
RNA polymerase I mutants defective for transcription
elongation, pointing to a possible regulatory role for
RNA polymerase elongation rates and pausing in ribosome
assembly [35].
The assembly of the large ribosomal subunit includes
many more nuclear steps than that of the 40S subunit, and
involves many more trans-acting factors (Figure 1). Similar
to UTP sub-complexes (Figure 4), several autonomous
structural neighborhoods have been characterized within
pre-60S ribosomes, and emerging evidence indicates that
at least some of them correspond to functional modules
(reviewed in [2]). Precursor rRNAs for large ribosomal
subunits are also monitored by the TRAMP-exosome
surveillance pathway; however, for the large ribosomal
subunit, it is mostly TRAMP4 that tags the RNAs. In
some instances, surveillance has been suggested to
occur in a specialized nucleolar domain termed the ‘No-
body’ that is enriched in both TRAMP and exosomal com-
ponents [36].
Ribosome surveillance exerted by TRAMPexosome
complexes initiates 3’ to 5’ RNA degradation at multiple
entry points; however, pre-rRNAs are long and highly
structured molecules, and it is expected that 5’!3’ nuclear
and nucleolar exoRNAse activities also contribute to pre-
rRNA decay.
Bulk ribosome and pre-ribosome decay by ribophagy
and PMN
Under limiting growth conditions, ribosome synthesis is
shut down at multiple levels. De novo synthesis is blocked
and nascent pre-ribosomes and mature ribosomes are
targeted to bulk degradative pathways. Pre-ribosomes
and mature ribosomes are turned over to recycle essential
cellular building blocks to cope with stress, and adapt to a
novel environment. It is quite striking that ubiquitin,
which plays an important role in 25S NRD, is also involved
in bulk ribosome degradation.
Ribophagy: a role for deubiquitylation in specific 60S
macroautophagy
Under starvation conditions, bulk portions of the cytosol,
including protein aggregates and entire organelles, such as
mitochondria and ribosomes, are recycled via two forms of
autophagy (micro- and macroautophagy). Macroautophagy
entails the formation of a double-membrane-bound vesicle
in the cytoplasm: the ‘autophagosome’ that sequesters
cytoplasmic material and delivers it for breakdown and
eventual recycling to the lysosome, or the vacuole in yeast
Review
Trends in Biochemical Sciences Vol.35 No.5
271
and plants (Figure 2). Autophagy can degrade nearly any
biomolecule, as the lysosome/vacuole contains non-specific
hydrolases, including proteases, nucleases, lipases, and
glycosylases. By contrast, microautophagy entails the
direct uptake of cytosolic components through invagina-
tions of the lysosome’s limiting membrane. Both macro-
and microautophagy contribute to cell survival during
starvation, and the two processes involve common and
specific sets of trans-acting factors [18,37]. Moreover,
although both micro- and macroautophagy are primarily
bulk degradative processes, selective forms of both path-
ways have been uncovered.
Figure 3. Nucleolar surveillance, the TRAMP-exosome pathway. Failure to assemble pre-ribosomes with proper kinetics leads to activation of nucleolar surveillance and
subsequent rRNA degradation. (a) In wild-type yeast cells, an early step in SSU-processome assembly is the binding of the UTP-A subcomplex (gray) to nascent pre-rRNA.
This is followed by the recruitment of other subcomplexes (UTP-B: orange; and UTP-C: green), additional individual ribosome synthesis factors (blue), the box C+D snoRNA
U3 (gold) and other snoRNAs, leading to SSU-processome catalytic activation in RNA cleavage. The nascent pre-rRNA is cleaved at sites A
0
,A
1
and A
2
(red) leading to the
production of the 20S pre-rRNA (the immediate precursor to the 18S rRNA) that is exported to the cytoplasm, dimethylated and cleaved at site D to produce mature 18S
rRNA. For simplicity, only co-transcriptional cleavage in ITS1, that occurs in !50-70% of the cases in fast-growing yeast cells, is illustrated here. Co-transcriptional cleavage
in ITS1 at site A
2
occurs when RNA polymerase I has reached the 5’-end of the 25S gene [57]. At this position, pre-40S compaction has brought the SSU-processome at its
closest position relative to the rDNA; in living yeast cells, this correlates with a strong RNA-dependent ChIP interaction of SSU-processome components [31]. In the
remaining cases, cleavage in ITS1 is delayed until transcription has reached the 3’-end of the gene, and a full-length pre-rRNA precursor (35S) has been generated (not
represented). (b) Mutations in any components of the SSU-processome (illustrated here as a red dot on UTP-B) inhibits SSU-processome assembly and/or function, thereby
resulting in inhibition of cleavage events at sites A
0
-A
2
, and activation of a downstream cleavage in ITS1 at site A
3
(purple) by endoRNase MRP. Cleavage of the nascent
transcript at site A
3
generates an aberrant 23S RNA that is rapidly polyadenylated by TRAMP5, followed by its digestion by the nuclear exosome; no mature 18S rRNA is
made. Upon surveillance inactivation (deletion of Trf5, Rrp6 or both), the 23S RNA is stabilized and not further matured (not shown). (c) The UTP-A subcomplex of the SSU-
processome is dispensable for cleavage at sites A
0
-A
2
. As described in panel B for UTP-B, a mutation in a component of UTP-A leads to alterations in pre-rRNA processing
kinetics (A
3
occurs prior to A
0
-A
2
). However, for UTP-A and by ensuring that the newly generated 23S RNA is stabilized by surveillance inactivation, this RNA is further
matured into 20S pre-rRNA and 18S rRNA. Despite this pre-rRNA processing restoration, cells are unable to survive, thus providing a strong indication that surveillance is
essential [31].
Review
Trends in Biochemical Sciences Vol.35 No.5
272
In yeast, prolonged nitrogen starvation leads to prefer-
ential targeting of both small and large ribosomal subunits
to the vacuole by selective macroautophagy [38]. This
process, specific to ribosomal subunits, is termed ‘ribo-
phagy’. For the large subunits, the increased turnover
kinetics rely upon the Ubp3 ubiquitin protease and its
activator Bre5. Mutations in UBE3 or BRE5 do not affect
detectably the increased turnover rates of small subunit
reporter constructs, indicating that, much as for 18S and
25S NRD, small and large ribosomal subunit ribophagy
relies on distinctly different pathways; notably, however,
the effector(s) for small ribosomal subunit ribophagy
remain unknown. Ubp3 deubiquitylation activity is
required for 60S ribophagy, and the ubiquitylation level
of several, still unidentified, ribosome-associated proteins
is increased in cells lacking Ubp3 [38]. It is thought that
deubiquitylation of Ubp3Bre5 target(s) assists in the
packaging of ribosomes in the autophagosome, or enables
its maturation and/or fusion to the vacuole [38]. Recent
data indicate that the ubiquitin ligase Rsp5 might help to
regulate 60S ribophagy as cells harboring mutant versions
of both Rsp5 and Ubp3 show enhanced synthetic sickness
and reduced ribosome turnover [39].
PMN: the involvement of a specialized nucleo-vacuolar
junction
Piecemeal microautophagy of the nucleus (PMN) is a se-
lective autophagic pathway in which the vacuole ‘pinches
off’ and degrades non-essential portions of the nucleus.
This process involves the formation of a specific inter-
organelle contact dubbed the nucleo-vacuole junction
(NVJ) which acts as true ‘Velcro-like’ patches formed by
interactions between the vacuolar membrane protein Vac8
and the outer nuclear membrane protein Nvj1 [40]. PMN is
a constitutive process that is induced to high levels upon
nutrient starvation. It bears relevance to ribosome degra-
dation, because there are cases where the site of vacuole
docking to the nucleus map to a position juxtaposing the
nucleolus [41] (Figure 2). In these captures, granular
material of nucleolar origin transfers directly from the
nucleolus to the vacuolar lumen; this event is indicative
of bulk destruction of pre-ribosomes.
Ribosome surveillance in human health
RNA damage occurs under normal cell growth as well as
during stress and in disease situations. Chemical modifi-
cations to nucleobases, as well as RNARNA and RNA
Figure 4. Small Subunit (SSU)-processome assembly involves concurrent parallel pathways. The SSU-processome comprises individual autonomous preassembled
building blocks that are sequentially loaded onto nascent pre-rRNA following multiple concurrent pathways, rather than via a single route [28,63]. This strategy, which
minimizes complexity, is likely to be the rule rather than the exception in ribosome assembly, and probably prevails in LSU-processome (pre-60S) formation. The currently
described SSU-processome subcomplexes include the UTP-A (gray), UTP-B (orange) and UTP-C (green), each comprising !6-7 proteins, as well as additional ribosome
synthesis factors (blue), including the Mpp10Imp3Imp4 and the Rcl1Bms1 subcomplexes (reviewed in [2]). Components of UTP-A (also referred to as t-UTPs) influence
efficient Pol I activity in yeast and humans but are not strictly required for pre-rRNA synthesis in yeast [28,3032]. UTP-C (also known as the CURI complex) integrates pre-
rRNA processing and ribosomal protein production [64]. Specific functions have not yet been assigned to UTP-B. (i) Binding of UTP-A to nascent transcripts is an early step
that initiates assembly. (ii) In one of two currently described alternative pathways, binding of UTP-A is followed by the sequential recruitment of UTP-B, U3 snoRNA and
additional ribosome synthesis factors. (iii) In an alternative route, UTP-A binding initiates the sequential recruitment of Rrp5 and UTP-C.
Review
Trends in Biochemical Sciences Vol.35 No.5
273
protein crosslinks, are introduced into RNA and RNPs as a
result of exposure to ultraviolet light, oxidation, chlori-
nation, nitration and alkylation [42]. These alterations all
constitute potential physiological triggers to ribosome
degradation pathways. Some cases of ribosomal RNA
damage induced by UV irradiation and oxidation have
been documented, notably in conjunction with human
neurodegenerative diseases such as Alzheimer’s and Par-
kinson’s, as well as atherosclerotic plaques [8] (Box 3 ). It
has been suggested that RNA oxidation could be involved
in disease progression and that RNA susceptibility to
oxidative damage is probably influenced by various factors,
including the degree of association with proteins (protec-
tion) or iron (sensitization). Mature 5.8S and 25S rRNA are
fragmented heavily in yeast cells exposed to elevated levels
of reactive oxygen species (ROS) generated by oxidative
stress (including exposure to hydrogen peroxide and mena-
dione), chronological aging, and other apoptotic cues [43
45]. Yeast cells treated with the anti-metabolite and che-
motherapeutic agent 5-fluorouracil (5-FU) accumulate
polyadenylated pre-rRNAs, and this accumulation is
exacerbated in exosome mutants, that are hypersensitive
to the drug. These findings suggest that the degradation of
5-FU containing RNAs occurs via nucleolar surveillance
[34]. Nucleolar dysfunction has been associated with
numerous human diseases, including cancer [46]. Various
‘nucleolar stresses’, such as drug-mediated interference of
rRNA synthesis, pre-rRNA processing (e.g. actinomycin D,
5-FU), or inhibition of ribosome synthesis factor function,
lead to nucleolar disruption, p53 stabilization and, ulti-
mately, cell cycle progression defects and/or apoptosis [47
50]. Of interest to the general public is the observation of
an increased abundance of polyadenylated rRNA frag-
ments in the gut of western honey bees originating from
populations infected with colony collapse disorder (CCD)
[51]. The insect gut serves as a primary interface with the
environment, as it is the principal site of pesticide detox-
ification and an integral component in the immune defense
against pathogens. CCD has been linked to picorna-like
viral infections (known to ‘hijack’ cellular ribosomes) and
the extensive use of pesticides; both stimuli could poten-
tially trigger a ribosome degradation response.
Concluding remarks
Ribosome synthesis is a major cellular metabolic activity
that can enforce a rapid ‘energy drain’ in the absence of tight
regulation [52,53]. Control is exerted at the level of the
synthesis and assembly of the pieces, as well as in the
function of the final product. Ribosome synthesis has
evolved to be fully integrated with complex nutrient-sensing
cascades, such that defective precursor ribosomes, as well as
damaged and excess mature ribosomes, are targeted specifi-
cally for rapid breakdown and recycling. Several surveil-
lance pathways that either select defective or excess pre-
ribosomes or mature particles have been identified; they
display strong specificity towards small and large ribosomal
subunits, involve specialized subcellular locales, and,
altogether, are active at each step of a ribosome’s life. The
question is now to know whether and how these pathways
interconnect (Box 4).
Box 4. Outstanding questions
1. How tightly is pre-RNA synthesis coupled to pre-rRNA proces-
sing and pre-rRNA surveillance? Do polymerase elongation
rates and/or pausing influence nucleolar surveillance?
2. How does TRAMP discriminate aberrant from normal RNAs?
Are other exosome cofactors involved in nuclear ribosome
surveillance?
3. Do any nucle(ol)ar mechanisms of pre-ribosome surveillance
initiate from the 5’-end of the pre-RNAs? There are suitable
exoRNase activities in the nucle(ol)us.
4. To what exte nt is (pre-)rRNA modification monitored by
surveillance?
5. Does PMN contribute quantitatively to nucleolar and nuclear
ribosome surveillance?
6. How does ubiquitin serve both a stabilizing and destabilizing
function in ribosome turnover? Which ribosome-associated
components are ubiquitylated in the 25S NRD and ribophagy
pathways? A ribosomal protein from each subunit (Rps31 and
Rpl40a/b) is produced as a fusion with ubiquitin [74] and Rpl28
ubiquitylation is required for optimal ribosome function [75].
What is the involvement of ubiquitin in nuclear pre-ribosome
turnover? What are the effector(s) of 40S ribophagy?
7. How are defective large ribosomal subunits functionally mon-
itored by 25S NRD?
8. Which endoRNase(s) initiate the cleavage events in mRNA NGD
and in 18S NRD? What is the fate of associated mRNAs in NRD
and, reciprocally, that of associated ribosomes in NGD? Late
cytoplasmic pre-40S ribosomes might enter translation, and
therefore be actively monitored by NRD [7678].
9. Do ribosome repair mechanisms exist? A priori, this seems
quite unlikely, owing to the compaction of mature subunits. A
demethylase is, however, known to repair alkylated RNAs in
bacteria and eukaryotes [79,80], suggesting that some rRNA
repair could occur during assembly. Evidence for ribosome
rejuvenation by ribosomal protein replacement is suggested by
in vitro translational reactivation of chemically-damaged bacter-
ial ribosomes [81].
10. How does ribosome surveillance failure contribute to human
disease etiology and progression?
Box 3. RNA damage in human health
Environmentally relevant doses of UVA and UVB are sufficient to
induce lesions in nucleic acids. UV is a source of photoproducts,
crosslinking and oxidative damage. UV crosslinking is dependent on
the geometry (distance, angle) and photoreactivity of the nucleotide
and amino acids, suggesting that certain cellular RNAs are more likely
to be protected from alterations by UV crosslinking than others (e.g.
double stranded RNA region). UV-mediated rRNA damage has been
reported in cultured mammalian cells, consistent with the formation
of photoproducts [66]. The lesions clustered around the active sites of
the large ribosomal subunit (domains V and VI near the PTC), and
were correlated with reduced translation and the activation of a
kinase-mediated stress response. Under oxidizing conditions, ribo-
somes undergo two types of damage: formation of 8-oxo-7,8-
dihydroguanosine (8-oxoG) and cross-linking of rRNA and ribosomal
proteins (shown in yeast cells treated with H
2
O
2
[67]). Most oxidative
damage results from the action of ROS that arise from metabolic
reactions. Increased levels of oxidatively damaged RNAs, as revealed
by the presence of 8-oxoG, have been reported in neurons from
patients with human neurodegenerative diseases, including Alzhei-
mer’s (AD; where it has been detected in nucleoli) and Parkinson’s, as
well as dementia, and in patients with atherosclerosis [6872]. Studies
in neurons from patients with AD revealed higher levels of oxidized
rRNA [68,69], and ribosomes purified from AD patients showed
elevated levels of associated redox-active iron [73]. Advanced human
atherosclerotic plaques have been linked to severe loss of rRNA
integrity and accumulation of elevated levels of 8-oxoG-containing
RNAs, a probable consequence of oxidative stress, ROS production
and intraplaque hemorrhage that is accompanied by iron deposition
[70,71].
Review
Trends in Biochemical Sciences Vol.35 No.5
274
How are defective (pre-)ribosomes recognized?
The exact molecular basis for recognition is not yet known;
however, the huge diversity of potential ribosome insults
makes it extremely unlikely that each one is monitored by
individual surveillance mechanisms. It is thought that each
reaction in the ribosome assembly pathway is provided with
a time frame for completion and that surveillance is trig-
gered by inefficient kinetics. Work with mature ribosomes
indicates that, in this case, it is the overall dynamics of the
translation process that is monitored. Thus far, this prop-
erty is best demonstrated for small subunits where stalling
ribosomes are identified by ‘translation-factor like’ proteins
that are also active in mRNA NGD. For 25S NRD, the
kinetics of interaction with translation factors might be
the key parameter that is monitored. It can thus be anticip-
ated that rRNA sequences that have been the most highly
conserved throughout evolution are probably prime targets
for monitoring by surveillance. Likewise, pre-rRNA modi-
fications that cluster around functional ribosomal sites and
contribute to translation efficacy, are also expected to be
monitored. The specialized functions of the two ribosomal
subunits in translation is reflected in their overall 3-D
structure and shape, with a flexible ‘Y-shape’ for the small
subunit and a monolithic block for the large subunit that
might require additional ‘effort’ (i.e., ubiquitylation, and
possible dissociation, of some of the protein components)
to expose naked RNAs to salient cellular endo- and exoR-
Nase activities. The selective cellular compartmentalization
of aberrant pre-ribosomes (No-body), and mature ribosomes
(P-bodies), in specific organelles might also contribute to the
processes of discrimination and degradation. Functional
compartmentalization could be particularly relevant to
25S NRD substrates that accumulate in cytosolic peri-
nuclear foci and primarily colocalize with 60S subunits in
velocity gradients. These findings indicate that defective
large subunits might be segregated away from the transla-
tionally active pool of ribosomes. By contrast, 18S NRD
RNAs co-migrate with monosomes, 80S particles and poly-
somes; these findings are consistent with their monitoring
during translation [11,13,17].
How are defective ribosomes tagged for degradation?
Thus far, two mechanisms have been identified that involve
either the addition of short poly(A) tails at the 3’-ends of the
RNA in the nucleus or the ubiquitylation/deubiquitylation of
unknown associated proteins in the cytoplasm. Strikingly,
ubiquitin plays both a stabilizing and destabilizing function;
clearly, its role in ribosome turnover is only starting to
unfold. That ubiquitylation also plays a role in nuclear
pre-ribosome degradation is suggested by the severe pre-
rRNA processing inhibition, and production of aberrant and
truncated rRNA fragments that are observed in yeast upon
conditional inactivation of the ubiquitin ligase Rsp5 [54].
Mammalian proteasome inhibition also leads to pre-rRNA
processing defects as well as striking nuclear accumulation
of ribosomal proteins; such findings are consistent with the
existence of further quality control systems [55,56].
Do we need ribosome surveillance?
The physiological relevance of ribosome surveillance is
illustrated by emerging connections to human diseases
(Box 3), and observations that cells harboring an inacti-
vated nucleolar pathway, 25S NRD or ribophagy, accumu-
late defective or excess ribosomes, and are either unable to
survive or show a much decreased lifespan [17,31,32,38].
Proteomic analyses of purified nucleoli and pre-ribo-
somes have left us with a myriad of trans-acting factors
to which function(s) and structure must be assigned. The
field is entering a new era in which deciphering the post-
translational modification of synthesis factors and riboso-
mal proteins will dictate future progress in understanding
the regulated assembly and turnover of ribosomes.
Acknowledgements
Makoto Kitabatake (Kyoto University), Natalia Shcherbik and Dimitri
Pestov (University of Medicine and Dentistry of New Jersey), Martin Kos
and David Tollervey (University of Edinburgh) are acknowledged for
communicating unpublished observat ions prior to publication. I
apologized to my friends and colleagues whose work could not be listed
here owing to space limitation. The laboratory is funded by the FRS-
FNRS, Re
´
gion Wallonne (Cibles) and Communaute
´
Franc¸aise de Belgique
(ARC).
Appendix A. Supplementary data
Supplementary data associated with this article can be
found, in the online version, at doi:10.1016/j.tibs.2009.
12.006.
References
1 Steitz, T.A. (2008) A structural understanding of the dynamic ribosome
machine. Nat. Rev. Mol. Cell Biol. 9, 242253
2 Henras, A.K. et al. (2008) The post-transcriptional steps of eukaryotic
ribosome biogenesis. Cell. Mol. Life Sci. 65, 23342359
3 Strunk, B.S. and Karbstein, K. (2009) Powering through ribosome
assembly. RNA 15, 20832104
4 Warner, J.R. and McIntosh, K.B. (2009) How common are
extraribosomal functions of ribosomal proteins? Mol. Cell 34, 311
5 Lafontaine, D. et al. (1995) The 18S rRNA dimethylase Dim1p is
required for pre-ribosomal RNA processing in yeast. Genes Dev. 9,
24702481
6 Senger, B. et al. (2001) The nucle(ol)ar Tif6p and Efl1p are required for
a late cytoplasmic step of ribosome synthesis. Mol. Cell 8, 1363
1373
7 Gadal, O. et al. (2002) Rlp7p is associated with 60S preribosomes,
restricted to the granular component of the nucleolus, and required for
pre-rRNA processing. J. Cell Biol. 157, 941951
8 Nunomura, A. et al. (2009) RNA oxidation in Alzheimer disease and
related neurodegenerative disorders. Acta Neuropathol. 118, 151166
9 Doma, M.K. and Parker, R. (2007) RNA quality control in eukaryotes.
Cell 131, 660668
10 Doma, M.K. and Parker, R. (2006) Endonucleolytic cleavage of
eukaryotic mRNAs with stalls in translation elongation. Nature 440,
561564
11 LaRiviere, F.J. et al. (2006) A late-acting quality control process for
mature eukaryotic rRNAs. Mol. Cell 24, 619626
12 Schmid, M. and Jensen, T.H. (2008) The exosome: a multipurpose
RNA-decay machine. Trends Biochem. Sci. 33, 501510
13 Cole, S.E. et al. (2009) A convergence of rRNA and mRNA quality
control pathways revealed by mechanistic analysis of nonfunctional
rRNA decay. Mol. Cell 34, 440450
14 Graille, M. et al. (2008) Structure of yeast Dom34: a protein related to
translation termination factor Erf1 and involved in No-Go decay. J.
Biol. Chem. 283, 7145
7154
15 Carr-Schmid, A. et al. (2002) Novel G-protein complex whose
requirement is linked to the translational status of the cell. Mol.
Cell. Biol. 22, 25642574
16 Balagopal, V. and Parker, R. (2009) Polysomes, P bodies and stress
granules: states and fates of eukaryotic mRNAs. Curr. Opin. Cell Biol.
21, 403408
17 Fujii, K. et al. (2009) A role for ubiquitin in the clearance of
nonfunctional rRNAs. Genes Dev. 23, 963974
Review
Trends in Biochemical Sciences Vol.35 No.5
275
18 Beau, I. et al. (2008) Lost to translation: when autophagy targets
mature ribosomes. Trends Cell Biol. 18, 311314
19 Suter, B. et al. (2007) Examining protein protein interactions using
endogenously tagged yeast arrays: the cross-and-capture system.
Genome Res. 17, 17741782
20 Krogan, N.J. et al. (2006) Global landscape of protein complexes in the
yeast Saccharomyces cerevisiae. Nature 440, 637643
21 Kaganovich, D. et al. (2008) Misfolded proteins partition between two
distinct quality control compartments. Nature 454, 10881095
22 Houseley, J. et al. (2006) RNA-quality control by the exosome. Nat. Rev.
Mol. Cell Biol. 7, 529 539
23 LaCava, J. et al. (2005) RNA degradation by the exosome is promoted
by a nuclear polyadenylation complex. Cell 121, 713724
24 Wyers, F. et al. (2005) Cryptic pol II transcripts are degraded by a
nuclear quality control pathway involving a new poly(A) polymerase.
Cell 121, 725737
25 Vanacova, S. et al. (2005) A new yeast poly(A) polymerase complex
involved in RNA quality control. PLoS Biology 3, e189
26 Andersen, K.R. et al. (2008) Take the ‘‘A’’ tail - quality control of
ribosomal and transfer RNA. Biochim. Biophys. Acta 1779, 532537
27 Dragon, F. et al. (2002) A large nucleolar U3 ribonucleoprotein required
for 18S ribosomal RNA biogenesis. Nature 417, 967970
28 Gallagher, J.E. et al. (2004) RNA polymerase I transcription and pre-
rRNA processing are linked by specific SSU processome components.
Genes Dev. 18, 25062517
29 Turner, A.J. et al. (2009) A novel small-subunit processome assembly
intermediate that contains the U3 snoRNP, nucleolin, RRP5, and
DBP4. Mol. Cell. Biol. 29, 30073017
30 Prieto, J.L. and McStay, B. (2007) Recruitment of factors linking
transcription and processing of pre-rRNA to NOR chromatin is
UBF-dependent and occurs independent of transcription in human
cells. Genes Dev. 21, 20412054
31 Wery, M.
et al. (2009) The nuclear poly(A) polymerase and Exosome
cofactor Trf5 is recruited cotranscriptionally to nucleolar surveillance.
RNA 15, 406419
32 Dez, C. et al. (2007) Roles of the HEAT repeat proteins Utp10 and
Utp20 in 40S ribosome maturation. RNA 13, 15161527
33 Kuai, L. et al. (2004) Polyadenylation of rRNA in Saccharomyces
cerevisiae. Proc. Natl. Acad. Sci. U. S. A. 101, 85818586
34 Fang, F. et al. (2004) 5-fluorouracil enhances exosome-dependent
accumulation of polyadenylated rRNAs. Mol. Cell. Biol. 24, 10766
10776
35 Schneider, D.A. et al. (2007) Transcription elongation by RNA
polymerase I is linked to efficient rRNA processing and ribosome
assembly. Mol. Cell 26, 217229
36 Dez, C. et al. (2006) Surveillance of nuclear-restricted pre-ribosomes
within a subnucleolar region of Saccharomyces cerevisiae. EMBO J. 25,
15341546
37 Nakatogawa, H. et al. (2009) Dynamics and diversity in autophagy
mechanisms: lessons from yeast. Nat. Rev. Mol. Cell Biol. 10, 458
467
38 Kraft, C. et al. (2008) Mature ribosomes are selectively degraded upon
starvation by an autophagy pathway requiring the Ubp3p/Bre5p
ubiquitin protease. Nat. Cell Biol. 10, 602610
39 Kraft, C. and Peter, M. (2008) Is the Rsp5 ubiquitin ligase involved in
the regulation of ribophagy? Autophagy 4, 838840
40 Pan, X. et al. (2000) Nucleus-vacuole junctions in Saccharomyces
cerevisiae are formed through the direct interaction of Vac8p with
Nvj1p. Mol. Biol. Cell 11, 24452457
41 Roberts, P. et al. (2003) Piecemeal microautophagy of nucleus in
Saccharomyces cerevisiae. Mol. Biol. Cell 14, 129141
42 Wurtmann, E.J. and Wolin, S.L. (2009) RNA under attack: cellular
handling of RNA damage. Crit. Rev. Biochem. Mol. Biol. 44, 3449
43 Mroczek, S. and Kufel, J. (2008) Apoptotic signals induce specific
degradation of ribosomal RNA in yeast. Nucleic Acids Res. 36, 2874
2888
44 Thompson, D.M.
et al. (2008) tRNA cleavage is a conserved response to
oxidative stress in eukaryotes. RNA 14, 20952103
45 Thompson, D.M. and Parker, R. (2009) The RNase Rny1p cleaves
tRNAs and promotes cell death during oxidative stress in
Saccharomyces cerevisiae. J. Cell Biol. 185, 4350
46 Derenzini, M. et al. (2009) What the nucleolus says to a tumour
pathologist. Histopathology 54, 753762
47 Mayer, C. and Grummt, I. (2005) Cellular stress and nucleolar
function. Cell Cycle 4, 10361038
48 Rubbi, C.P. and Milner, J. (2003) Disruption of the nucleolus mediates
stabilization of p53 in response to DNA damage and other stresses.
EMBO J. 22, 60686077
49 Azuma, M. et al. (2006) Perturbation of rRNA synthesis in the bap28
mutation leads to apoptosis mediated by p53 in the zebrafish central
nervous system. J. Biol. Chem. 281, 1330913316
50 Pestov, D.G. et al. (2001) Evidence of p53-dependent cross-talk between
ribosome biogenesis and the cell cycle: effects of nucleolar protein Bop1
on G(1)/S transition. Mol. Cell. Biol. 21, 42464255
51 Johnson, R.M. et al. (2009) Changes in transcript abundance relating to
colony collapse disorder in honey bees (Apis mellifera). Proc. Natl.
Acad. Sci. U. S. A. 106, 1479014795
52 Lempiainen, H. and Shore, D. (2009) Growth control and ribosome
biogenesis. Curr. Opin. Cell Biol. 21, 19
53 Jorgensen, P. et al. (2004) Forging the factory: ribosome synthesis and
growth control in budding yeast. In Cell Growth: Control of Cell Size
(Hall, M.N. et al., eds), pp. 329370, CSHL Press
54 Neumann, S. et al. (2003) Formation and nuclear export of tRNA, rRNA
and mRNA is regulated by the ubiquitin ligase Rsp5p. EMBO Rep. 4,
11561162
55 Lam, Y.W. et al. (2007) Analysis of nucleolar protein dynamics reveals
the nuclear degradation of ribosomal proteins. Curr. Biol. 17, 749760
56 Stavreva, D.A. et al. (2006) Potential roles for ubiquitin and the
proteasome during ribosome biogenesis. Mol. Cell Biol. 26, 51315145
57 Osheim, Y.N.
et al. (2004) Pre-18S ribosomal RNA is structurally
compacted into the SSU processome prior to being cleaved from
nascent transcripts in Saccharomyces cerevisiae. Mol. Cell 16, 943954
58 Honma, Y. et al. (2006) TOR regulates late steps of ribosome
maturation in the nucleoplasm via Nog1 in response to nutrients.
EMBO J. 25, 38323842
59 Vanrobays, E. et al. (2008) TOR regulates the subcellular distribution
of DIM2, a KH domain protein required for cotranscriptional ribosome
assembly and pre-40S ribosome export. RNA 14, 20612073
60 Schafer, T. et al. (2006) Hrr25-dependent phosphorylation state
regulates organization of the pre-40S subunit. Nature 441, 651655
61 Ulbrich, C. et al. (2009) Mechanochemical removal of ribosome
biogenesis factors from nascent 60S ribosomal subunits. Cell 138,
911922
62 Krick, R. et al. (2008) Piecemeal microautophagy of the nucleus
requires the core macroautophagy genes. Mol. Biol. Cell 19, 4492
4505
63 Perez-Fernandez, J. et al. (2007) The 90S preribosome is a
multimodular structure that is assembled through a hierarchical
mechanism. Mol. Cell. Biol. 27, 54145429
64 Rudra, D. et al. (2007) Potential interface between ribosomal protein
production and pre-rRNA processing. Mol. Cell Biol. 27, 48154824
65 Ho, J.H. and Johnson, A.W. (1999) NMD3 encodes an essential
cytoplasmic protein required for stable 60S ribosomal subunits in
Saccharomyces cerevisiae. Mol. Cell Biol. 19, 23892399
66 Iordanov, M.S. et al. (1998) Ultraviolet radiation triggers the ribotoxic
stress response in mammalian cells. J. Biol. Chem. 273, 1579415803
67 Mirzaei, H. and Regnier, F. (2006) Protein-RNA cross-linking in the
ribosomes of yeast under oxidative stress. J. Proteome Res. 5, 3249
3259
68 Ding, Q. et al. (2006) Decreased RNA, and increased RNA oxidation, in
ribosomes from early Alzheimer’s disease. Neurochem. Res. 31, 705
710
69 Ding, Q. et al. (2005) Ribosome dysfunction is an early event in
Alzheimer’s disease. J. Neurosci.
25, 91719175
70 Martinet, W. et al. (2005) RNA damage in human atherosclerosis:
pathophysiological significance and implications for gene expression
studies. RNA Biol. 2, 47
71 Martinet, W. et al. (2004) React ive oxygen species induce RNA damage
in human atherosclerosis. Eur. J. Clin. Invest. 34, 323327
72 Nunomura, A. et al. (1999) RNA oxidation is a prominent feature of
vulnerable neurons in Alzheimer’s disease. J. Neurosci. 19, 19591964
73 Honda, K. et al. (2005) Ribosomal RNA in Alzheimer disease is oxidized
by bound redox-active iron. J. Biol. Chem. 280, 2097820986
74 Finley, D. et al. (1989) The tails of ubiquitin precursors are ribosomal
proteins whose fusion to ubiquitin facilitates ribosome biogenesis.
Nature 338, 394401
Review
Trends in Biochemical Sciences Vol.35 No.5
276
75 Spence, J. et al. (2000) Cell cycle-regulated modification of the ribosome
by a variant multiubiquitin chain. Cell 102, 6776
76 Lacombe, T. et al. (2009) Linear ubiquitin fusion to Rps31 and its
subsequent cleavage are required for the efficient production and
functional integrity of 40S ribosomal subunits. Mol. Microbiol. 72,
6984
77 Pertschy, B. et al. (2009) RNA helicase Prp43 and its co-factor Pfa1
promote 20S to 18S rRNA processing catalyzed by the endonuclease
Nob1. J. Biol. Chem. 284, 3507935091
78 Soudet, J. et al. (2010) Immature small ribosomal subunits can engage
in translation initiation in Saccharomyces cerevisiae. EMBO J 29, 80
92
79 Aas, P.A. et al. (2003) Human and bacterial oxidative demethylases
repair alkylation damage in both RNA and DNA. Nature 421, 859863
80 Ougland, R. et al. (2004) AlkB restores the biological function of mRNA
and tRNA inactivated by chemical methylation. Mol. Cell 16, 107116
81 Pulk, A. et al. (2009) Ribosome reactivation by replacement of damaged
proteins. Mol. Microbiol Epub ahead of print PMID 19968789
Review
Trends in Biochemical Sciences Vol.35 No.5
277