ArticlePDF AvailableLiterature Review

Ribosomal protein S6 kinase from TOP mRNAs to cell size

Authors:

Abstract and Figures

Ribosomal protein S6 kinase (S6K) has been implicated in the phosphorylation of multiple substrates and is subject to activation by a wide variety of signals that converge at mammalian target of rapamycin (mTOR). In the course of the search for its physiological role, it was proposed that S6K activation and ribosomal protein S6 (rpS6) phosphorylation account for the translational activation of a subgroup of transcripts, the TOP mRNAs. The structural hallmark of these mRNAs is an oligopyrimidine tract at their 5'-terminus, known as the 5'-TOP motif. TOP mRNAs consists of about 90 members that encode multiple components of the translational machinery, such as ribosomal proteins and translation factors. The translation efficiency of TOP mRNAs indeed correlates with S6K activation and rpS6 phosphorylation, yet recent biochemical and genetic studies have established that, although S6K and TOP mRNAs respond to similar signals and are regulated by mTOR, they maintain no cause and effect relationship. Instead, S6K is primarily involved in regulation of cell size, and affects glucose homeostasis, but is dispensable for global protein synthesis, whereas translational efficiency of TOP mRNAs is a determinant of the cellular protein synthesis capacity. Despite extensive studies of their function and mode of regulation, the mechanism underlying the effect of S6K on the cell size, as well as the trans-acting factor that mediates the translational control of TOP mRNAs, still await their identification.
Content may be subject to copyright.
CHAPTER TWO
Ribosomal Protein S6
Phosphorylation: Four
Decades of Research
Oded Meyuhas
Department of Biochemistry and Molecular Biology, Institute for Medical Research eIsrael-Canada,
Hebrew University-Hadassah Medical School, Jerusalem, Israel
E-mail: meyuhas@cc.huji.ac.il
Contents
1. Introduction 42
2. rpS6 as an Indispensable Ribosomal Protein 43
3. Phosphorylation of rpS6 44
3.1 S6 Kinase (S6K1 and S6K2) 45
3.2 90-kDa rpS6 Kinase (RSK1eRSK4) 46
3.3 Protein Kinase A 47
3.4 Casein Kinase 1 48
3.5 rpS6 Dephosphorylation 48
4. Signals to rpS6 Phosphorylation 49
4.1 Growth Factors 49
4.1.1 PI3K/Akt/TSC/Rheb/mTORC1/S6K pathway 49
4.1.2 Ras/Raf/MEK/ERK/RSK pathway 50
4.2 Amino Acid Sufciency 51
4.3 Energy Balance 52
4.4 Oxygen Supply 54
4.5 Osmolarity 55
5. Physiological Roles of rpS6 Phosphorylation 56
5.1 Global Protein Synthesis 56
5.2 Cell Size Regulation 57
5.3 Normal Muscle Function 59
5.4 Hypertrophic Responses 59
5.5 Cell Proliferation 59
5.6 Clearance of Apoptotic Cells 60
5.7 Tumorigenicity 60
5.8 Glucose Homeostasis 62
5.9 rpS6 Phosphorylation as Diagnostic Marker 62
6. Concluding Remarks and Future Perspectives 63
References 64
International Review of Cell and Molecular Biology, Volume 320
ISSN 1937-6448
http://dx.doi.org/10.1016/bs.ircmb.2015.07.006
©2015 Elsevier Inc.
All rights reserved. 41
j
Abstract
The phosphorylation of ribosomal protein S6 (rpS6) has been described for the rst time
about four decades ago. Since then, numerous studies have shown that this modica-
tion occurs in response to a wide variety of stimuli on ve evolutionarily conserved
serine residues. However, despite a large body of information on the respective kinases
and the signal transduction pathways, the physiological role of rpS6 phosphorylation
remained obscure until genetic manipulations were applied in both yeast and mam-
mals in an attempt to block this modication. Thus, studies based on both mice and
cultured cells subjected to disruption of the genes encoding rpS6 and the respective
kinases, as well as the substitution of the phosphorylatable serine residues in rpS6,
have laid the ground for the elucidation of the multiple roles of this protein and its
posttranslational modication. This review focuses primarily on newly identied kinases
that phosphorylate rpS6, pathways that transduce various signals into rpS6 phosphor-
ylation, and the recently established physiological functions of this modication. It
should be noted, however, that despite the signicant progress made in the last
decade, the molecular mechanism(s) underlying the diverse effects of rpS6 phosphor-
ylation on cellular and organismal physiology are still poorly understood.
1. INTRODUCTION
The higher eukaryotic ribosomes are composed of two subunits desig-
nated as 40S (small) and 60S (large) subunits. The mammalian 40S subunit is
composed of a single molecule of RNA, 18S ribosomal (r) RNA, and 33
proteins, whereas the 60S subunit contains three RNA molecules, 5S,
5.8S, and 28S rRNAs, and 46 proteins (Wool et al., 1996). Of all ribosomal
proteins it is ribosomal protein S6 (rpS6) that has attracted much attention,
since it is the rst, and was for many years the only one, that has been shown
to undergo inducible phosphorylation.
The ribosome biogenesis takes place in the nucleolus starts with the syn-
thesis of 5S and 45S pre-rRNA by distinct RNA polymerases and requires
the import of ribosomal proteins from the cytoplasm. A complex pathway
that involves both endo- and exonucleolytic digestions enables the release
of mature rRNAs from the pre-rRNA. Concomitantly, the rRNAs are
extensively modied and bound by the ribosomal proteins before the assem-
bled pre-40S and pre-60S subunits are exported separately to the cytoplasm
(Fromont-Racine et al., 2003; Zemp and Kutay, 2007). High-resolution
cytological analysis has disclosed the fate of rpS6 from its biosynthesis site
in the cytoplasm to the pre-40S subunit. Thus, rpS6 enters the nucleus of
HeLa cells, reaches, via Cajal bodies, the nucleolus, where it is assembled
with other proteins and rRNA into pre-40S subunit. The latter is then
42 Oded Meyuhas
released to the nucleoplasm prior to its export through the nuclear pores to
the cytoplasm (Cisterna et al., 2006). Interestingly, the nuclear import, as
well as the nucleolar localization of human rpS6 and yeast rpS6A, relies
on motifs, whose number, nature, and position are evolutionary conserved
(Lipsius et al., 2005; Schmidt et al., 1995).
The phosphorylation of rpS6 has attracted much attention in numerous
labs since its discovery in 1974 (Gressner and Wool, 1974). However, it is
only during the last decade that the role of rpS6 and its posttranslational
modication has started being unveiled by genetic targeting of the rpS6
gene and of the respective kinases. Hence, this review includes an account
the critical role of rpS6 for mouse development, on the enzymes that
conduct its phosphorylation, the cues that affect it, the pathways that trans-
duce various signals into rpS6 phosphorylation, and the physiological role of
this modication.
2. rpS6 AS AN INDISPENSABLE RIBOSOMAL PROTEIN
rpS6 is an evolutionary conserved protein that spans 236e253 residues
in species as remote as yeast, plants, invertebrates, and vertebrates (Meyuhas,
2008), yet no homology with any ribosomal protein in Escherichia coli or
archaebacteria has been detected (Wool et al., 1996).
The role of rpS6 was rst addressed by conditional knockout of the
respective gene in adult mouse liver (Volarevic et al., 2000). Hepatocytes
that lacked rpS6 gene failed to synthesize the 40S ribosomal subunit and
consequently to proliferate following partial hepatectomy. This failure to
progress through the cell cycle correlated with a block in expression of cyclin
E gene. Nonetheless, the expression of rpS6 gene was not required for liver
growth when starved mice were refed. Moreover, the relative engagement
of liver ribosomes in translation, as exemplied by their polysomal associa-
tion, was indistinguishable between rpS6-containing and lacking hypertro-
phying livers (Volarevic et al., 2000).
The critical role of rpS6 is not conned to the regenerating liver, as
thymus-specic knockout of both rpS6 alleles, but not conditional deletion
of one allele, had devastating effect on the gland development (Sulic et al.,
2005). rpS6 heterozygosity (rpS6
wt/del
), however, had a remarkable effect on
the number of mature T cells in peripheral lymphoid organs (spleen and
lymph nodes). The deciency of one rpS6 allele led to a proportional dimi-
nution in the abundance of rpS6 and ribosome content in puried rpS6
wt/del
Ribosomal Protein S6 Phosphorylation 43
T cells, yet with no effect on their total protein content or their ability to
undergo normal stimulated cell growth (Sulic et al., 2005). Likewise, 30e
50% reduction in rpS6 content of HeLa cells by siRNA only mildly affected
global protein synthesis (Montgomery et al., 2006). Nevertheless, while
wild-type T cells progressed in vitro through several divisions upon mito-
genic stimulation, their rpS6
wt/del
counterparts failed to proliferate as a result
of a block at the G1/S checkpoint of the cell cycle, and partially due to
increased apoptosis. Interestingly, deletion of both p53 alleles almost
completely resumed the proliferative capacity of stimulated rpS6
wt/del
T cells.
These observations strongly support the notion that impaired ribosome
biogenesis, associated with rpS6 deciency, activates a p53-dependent
checkpoint to eliminate defective T cells (Sulic et al., 2005).
rpS6
wt/del
embryos died during gastrulation, at day 8.5. However,
already at day 6.5 their cells failed to show dephosphorylation and activation
of Cdk1 and to enter mitosis. Moreover, the embryonal death was preceded
by induced apoptosis. The fact that p53 gene knockout enabled rpS6
wt/del
embryo to develop past gastrulation stage, suggests that rpS6 heterozygosity
triggers a p53-mediated checkpoint during gastrulation. Interestingly, ribo-
some biogenesis is defective in rpS6
wt/del
/P53
/
embryo, as well as in the
corresponding MEFs. However, while neither cell cycle progression nor
cell growth is impaired in rpS6
wt/del
/P53
/
MEFs, compromised cell pro-
liferation was observed in the liver from rpS6
wt/del
/P53
/
embryo. This
decreased in hepatic proliferation might be explained by the relative de-
ciency of cyclins D1 and D3, observed in this organ (Panic et al., 2006).
Interestingly, unlike mammalian rpS6, lesions in Drosophila rpS6 gene
expression, due to insertion of a P element upstream of the transcription
initiation site, had a mixed response: Hyperplasia of lymphglands on the
one hand, and growth inhibition of most larval organs on the other hand
(Stewart and Denell, 1993; Watson et al., 1992). The latter response is
consistent with rpS6 having a tumor suppressor like function in Drosophila.
3. PHOSPHORYLATION OF rpS6
A pioneer study conducted by David Kabat showed that a 33-kDa
protein, termed F protein, which resided in the small ribosomal subunit un-
dergoes phosphorylation in rabbit reticulocytes (Kabat, 1970). Later, it was
identied as rpS6, and that it is the only ribosomal protein that undergoes
phosphorylation during rat liver regeneration (Gressner and Wool, 1974).
44 Oded Meyuhas
The phosphorylation sites in rpS6 in mammals and Xenopus laevis have been
mapped to ve clustered residues, S
235
,S
236
,S
240
,S
244
, and S
247
(Bandi
et al., 1993; Krieg et al., 1988; Wettenhall et al., 1992), whose location at
the carboxy terminus of higher eukaryotes is evolutionarily conserved
(Meyuhas, 2008). It has been proposed that phosphorylation progresses in
an ordered fashion, with Ser236 as the primary phosphorylation site (Flotow
and Thomas, 1992; Wettenhall et al., 1992). A similar organization of phos-
phorylation sites, relative to the carboxy terminus was described for
Drosophila melanogaster rpS6 (Radimerski et al., 2000).
The rst report on the phosphorylation of rpS6 (S10 according to an
older nomenclature) in Saccharomyces cerevisiae lagged behind that of its
mammalian counterpart (Hebert et al., 1977). Yeast rpS6 is phosphorylated
after transfer of a stationary culture to fresh nutrient medium, as well as at an
early stage of germination, and as in other eukaryotes, the protein is dephos-
phorylated during heat shock (Jakubowicz, 1985; Szyszka and Gasior, 1984).
However, yeast rpS6, unlike higher eukaryotes, bears only two phosphory-
latable serine residues (Ser232 and Ser233) that correspond to Ser235 and
Ser236 in the mammalian protein.
Numerous reports have demonstrated that rpS6 is subject to phosphor-
ylation in response to multiple physiological, pathological, and pharmaco-
logical stimuli ((Meyuhas, 2008) and references therein). Notably, this
modication can be detected in both the cytosol and the nucleus (Pende
et al., 2004). However, distinct nuclear/cytoplasmic distribution of rpS6
phosphorylated at different sits has been noticed for primary human cells
(Rosner et al., 2011), yet the physiological signicance of this compart-
mental preference is not clear.
3.1 S6 Kinase (S6K1 and S6K2)
Characterization of an S6 kinase at a molecular level was rst achieved in
Xenopus oocytes, wherein the dominant form of S6 kinase detected after
mitogenic stimulation had been puried as a 90 kDa polypeptide (Erikson
and Maller, 1985), later termed as p90 rpS6 kinase (RSK, also known as
p90
RSK
). Purication of the avian and mammalian major rpS6 kinase recov-
ered 65- to 70-kDa polypeptides (Blenis et al., 1987; Jeno et al., 1988) that
are currently referred to as S6K.
Mammalian cells contain two forms of S6K, S6K1 and S6K2 (also
known as S6Kaand S6Kbrespectively), which are encoded by two different
genes and share a very high level of overall sequence homology (Meyuhas
and Dreazen, 2009). S6K1 has cytosolic and nuclear isoforms (p70 S6K1
Ribosomal Protein S6 Phosphorylation 45
and p85 S6K1, respectively), whereas both S6K2 isoforms (p54 S6K2 and
p56 S6K2) are primarily nuclear (Martin et al., 2001) and references therein)
and partly associated with the centrosome (Rossi et al., 2007). Analysis of
rpS6 phosphorylation in mouse cells decient in either S6K1 or S6K2 sug-
gests that both are required for full S6 phosphorylation, with the predomi-
nance of S6K2 (Pende et al., 2004). Notably, the phosphorylation of the
evolutionary conserved sites of Drosophila rpS6 is carried out by dS6K that
is encoded by a single gene (Watson et al., 1996).
Two putative S6K homologs, originally named atpk1/ATPK6 and
atpk2/ATPK19, sharing 87% sequence homology were identied in Arabi-
dopsis (Mizoguchi et al., 1995; Zhang et al., 1994). They were later referred
to as atS6K1 and atS6K2, respectively (Turck et al., 1998), of which, atS6K2
was suggested to be an ortholog of the mammalian S6Ks, since this kinase,
and not atS6K2, was able to phosphorylate rpS6 (Zhang et al., 1994). Phos-
phorylation of yeast rpS6 has been known for nearly four decades (Hebert
et al., 1977), yet the identity of the respective kinase remained elusive until
recently. Initially, the Sch9 kinase was proposed to comprise the yeast S6K
(Urban et al., 2007). However, recently this role has been assigned instead,
to Ypk3, a kinase that exhibits high homology to human S6K (Gonzalez
et al., 2015). This latter notion is supported by the observation that rpS6
phosphorylation is completely abolished in cells lacking Ypk3, ypk3D,
whereas Sch9 is dispensable for rpS6 phosphorylation. Furthermore,
complementation of ypk3Dcells with human S6K restored rpS6 phosphor-
ylation in a rapamycin-sensitive manner (Gonzalez et al., 2015)(Figure 1).
3.2 90-kDa rpS6 Kinase (RSK1eRSK4)
RSKs are central mediators of extracellular signal-regulated kinase (ERK
(for further details on this pathway see Section 4.1.2.)) in regulation of
cellular division, survival, and differentiation via phosphorylation of
numerous intracellular proteins ((Romeo et al., 2012) and references
therein). Four RSK genes (RSK1eRSK4) have been identied in mammals,
and RSK orthologs have been described in D. melanogaster and Caenorhabditis
elegans, but not in yeast and plants (Hauge and Frodin, 2006; Lara et al.,
2013).
The discovery that S6K is the predominant rpS6 kinase in somatic cells
(Ballou et al., 1991; Chung et al., 1992) has led to a widely accepted belief
that RSK, despite its name, is physiologically irrelevant for rpS6 phosphor-
ylation. However, later observations have challenged this dogma. Thus,
phosphorylation of rpS6 at Ser235 and Ser236 (Ser235/236) can still be
46 Oded Meyuhas
detected, albeit at a much lower level, in cells lacking both S6K1 and S6K2.
This phosphorylation is abolished by treatment by either U0126 (a MAP
and ERK kinases (MEK) inhibitor) or PD184352 (an ERK inhibitor), indi-
cating the involvement of a MEK/ERK-dependent kinase (Pende et al.,
2004). Likewise, Ser235/236 remained partly phosphorylated in cells treated
with rapamycin, which completely inhibits S6K through inhibiting
mammalian target of rapamycin (mTOR), indicating the presence of an
mTOR-independent pathway leading to rpS6 phosphorylation at these
sites. Moreover, it has been shown that this phosphorylation is carried
out, both in vitro and in vivo, by RSK, which phosphorylates rpS6
exclusively at Ser235/236 in a response to serum, growth factors, tumor-
promoting phorbol esters, and oncogenic Ras (Roux et al., 2007).
The consensus recognition sequences of S6K and RSK are similar,
RxRxxS and R/KxRxxS respectively, where x represents any amino acid
and the carboxy-terminal S is the phosphorylated serine residue (Flotow
and Thomas, 1992; Hauge and Frodin, 2006). Notably, however, the
sequence context of serine 236 in rpS6 is the only one, among the phos-
phorylatable serine residues, that conforms to the consensus recognition
sequence of these enzymes (Figure 1).
3.3 Protein Kinase A
Protein kinase A (PKA) is a family of enzymes, whose activity is dependent
on cellular levels of cyclic AMP (cAMP) and phosphorylates a large number
of cytosolic and nuclear proteins (Kirschner et al., 2009). Agents that induce
increases in cAMP level in the pancreatic bcell line MIN6 (mouse insuli-
noma cell line 6) and in mouse islets of Langerhans lead to the phosphory-
lation of rpS6 that is conned to Ser235/236 via a pathway that is sensitive to
inhibitors of PKA. PKA was also shown to exclusively phosphorylate
recombinant rpS6 on Ser235/236 in vitro, and is likely to phosphorylate
Figure 1 Sites of rpS6 phosphorylation by different kinases and of dephosphorylation
by PP1. Arrows represent phosphorylation and the dashed lineddephosphorylation.
See text for details.
Ribosomal Protein S6 Phosphorylation 47
rpS6 on Ser235/236 in a number of other mammalian cell lines, such as
broblasts, pheochromocytoma, neuroblastoma, and kidney cells (Moore
et al., 2009).
Furthermore, treatment of mice with haloperidol, a typical antipsychotic
drug, increased rpS6 phosphorylation at Ser235/236 in medium spiny
neurons of the striatum. This phosphorylation is carried out through activa-
tion of the cAMP signaling, as was indicated by elevated Ser235/236 phos-
phorylation upon stimulation of PKA in cultured striatal neurons (Valjent
et al., 2011). Likewise, stimulation of cAMP production by forskolin, an
adenylate cyclase activator, in mouse striatal slices led to a marked increase
in the phosphorylation of S235/236, which was attenuated by pretreatment
with Rp-cAMP, a PKA inhibitor (Biever et al., 2015)(Figure 1).
3.4 Casein Kinase 1
Members of the casein kinase 1 (CK1), a family of serine/threonine kinases,
are ubiquitously expressed and regulate diverse cellular processes, through
phosphorylation of a variety of proteins, which are preferentially primed
prephosphorylated substrates (Cheong and Virshup, 2011). CK1 has recently
been shown to selectively phosphorylate rpS6 at Ser247 (Hutchinson et al.,
2011). The identication of CK1 and its specicity toward Ser247 was based
on both pharmacological inhibition of all members of the casein kinase family
and on knockdown experiments with CK1 siRNAs. However, the ability of
recombinant CK1 to phosphorylate Ser247 in in vitro kinase assay was indis-
tinguishable from that of its capacity to phosphorylate Ser235/236 or Ser240/
244 in rpS6. This apparent lack of specicity might be an artifact of the in
vitro assay (Hutchinson et al., 2011)(Figure 1).
3.5 rpS6 Dephosphorylation
The steady state level of rpS6 phosphorylation is the product of a dynamic
equilibrium between the activities of the respective kinases and the opposing
phosphatases. Nonetheless, the uctuations in rpS6 phosphorylation have
been attributed, in nearly all the relevant reports, to parallel changes in
the kinase(s) activity. In a few cases, however, the phosphorylation status
of rpS6 has been primarily ascribed to the activity of a phosphatase rather
than a kinase. Thus, Rous sarcoma virus-transformed chick embryo
broblasts show attenuated dephosphorylation of rpS6 during mitosis and
a parallel decrease in the activity of the protein phosphatase type 1 (PP1).
This observation suggests that it is the PP1 activity that might control
rpS6 phosphorylation under these circumstances (Belandia et al., 1994).
48 Oded Meyuhas
Conversely, rpS6 phosphorylation is not detectable in murine erythroleuke-
mia or other hematopoietic cells, and this constitutive dephosphorylation
state appears to be due to the action of a phosphatase that is likely to act
directly on rpS6 (Barth-Baus et al., 2002).
The demonstration that rpS6 phosphorylation is enhanced upon treat-
ment of cells with tautomycetin, a PP1 inhibitor, and that rpS6 is coimmu-
noprecipitated with anti-PP1Caantibody (Li et al., 2012), provided the rst
indication that rpS6 might indeed be a PP1 substrate. This contention has
been further supported by genetic manipulations. Thus, expression of
PP1D95N, a dominant-negative PP1 catalytic subunit, caused a marked
elevation in rpS6 phosphorylation at all ve phosphorylatable sites of
rpS6. Likewise, knockdown of PP1 catalytic subunit aresulted in similar
consequences (Hutchinson et al., 2011). Collectively, these observations
imply that PP1 is the primary phosphatase of rpS6 (Figure 1).
4. SIGNALS TO rpS6 PHOSPHORYLATION
Following the initial wave of descriptive reports on rpS6 phosphory-
lation, much effort has been invested in an attempt to establish the pathways
that transduce various signals into activation or inhibition of the respective
kinases. The pathways discussed below are only those that have been
documented to exert a parallel effect on both rpS6 phosphorylation and
the activity of the respective kinase.
4.1 Growth Factors
4.1.1 PI3K/Akt/TSC/Rheb/mTORC1/S6K pathway
Signaling to S6 phosphorylation by growth factors starts by activation of the
respective receptor tyrosine kinase. This in turn, leads to activation of class I
phosphatidylinositol 3-kinase (PI3K), either through direct binding to the
phosphorylated receptor or through tyrosine phosphorylation of scaffolding
adaptors, such as insulin receptor substrate, which then binds and activates
PI3K (Cantrell, 2001). PI3K converts the lipid phosphatidylinositol-4,5-
P2 (PIP2) into phosphatidylinositol-3,4,5-P3 (PIP3), in a reaction that can
be reversed by the PIP3 phosphatase PTEN (phosphatase and tensin homo-
log deleted from chromosome 10) (Leslie and Downes, 2002). PIP3 recruits
both 3-phosphoinositide-dependent kinase 1 (PDK1) and Akt (also known
as protein kinase B (PKB)) to the plasma membrane (Brazil and Hemmings,
2001; Lawlor and Alessi, 2001), and PDK1 phosphorylates and activates Akt
at T308 (Belham et al., 1999). Activated Akt phosphorylates tuberous
Ribosomal Protein S6 Phosphorylation 49
sclerosis complex 2 (TSC2) at multiple sites (Inoki et al., 2002; Manning
et al., 2002; Potter et al., 2002). This phosphorylation blocks the ability of
TSC2, while residing within the TSC1eTSC2 tumor suppressor dimmer
to act as a GTPase-activating protein (GAP) for Rheb (Ras-homolog
enriched in brain), thereby allowing Rheb-GTP to accumulate and operate
as an activator of the rapamycin-sensitive mammalian TOR complex 1
(mTORC1) (Avruch et al., 2006). The latter is consisting of target of rapa-
mycin (TOR), RAPTOR (regulatory associated protein of TOR) LST8
(also known as GbL), and PRAS40 (proline-rich Akt substrate 40 kDa)
(Bhaskar and Hay, 2007; Yang and Guan, 2007). Since it is mTORC1
that conveys signals to S6Ks and rpS6, it will be mentioned in the remainder
of this review, rather than mTOR, when transduction of signals is discussed.
Akt can also activate mTORC1 independently of TSC1eTSC2 by phos-
phorylating PRAS40, thereby relieving the PRAS40-mediated inhibition
of mTORC1 (Sancak et al., 2007; Vander Haar et al., 2007).
Active mTORC1 phosphorylates two translational regulators, S6Ks and
eukaryotic initiation factor 4E (eIF-4E)-binding protein (4E-BP1, 2, and 3)
(Hay and Sonenberg, 2004). Activation of S6Ks requires also phosphoryla-
tion by PDK1 in a reaction that does not need binding of PDK1 to PIP3
(Alessi et al., 1998). Finally, activated S6K phosphorylates rpS6, as well as
many other substrates (Meyuhas and Dreazen, 2009)(Figure 2).
4.1.2 Ras/Raf/MEK/ERK/RSK pathway
Activation of the second family of rpS6 kinases, the RSKs involve a distinct
signaling pathway, even though it might share the same initial event, namely
the ligand binding, with the PI3K/Akt/TORC1 pathway. Since RSK
seemed to be a minor rpS6 kinase, the pathway leading to its activation is
only very briey described here. Thus, the binding of insulin, as well as of
many other growth factors, to their receptors induces the activation of the
small GTPase Ras and consequently the recruitment of Raf to the mem-
brane for subsequent activation by phosphorylation. Raf activates
mitogen-activated protein (MAP) kinase, kinases 1 and 2 (MEK1/2), which
in turn phosphorylates and activates the extracellular-signal-regulated
kinases (ERK1 and ERK2). Activated ERKs phosphorylate and activate a
vast array of substrates localized in all cellular compartments such as the
RSK family (McCubrey et al., 2007; McKay and Morrison, 2007).
Interestingly, TSC2 is repressed by the Ras/MAPK pathway in addition
to its downregulation by the PI3K/Akt pathway, as evidenced by the obser-
vation that activated Erk1/2 directly phosphorylates TSC2 at sites that differ
50 Oded Meyuhas
from the Akt target sites, thereby causing functional inactivation of the
TSC1eTSC2 complex (Ma et al., 2005). Moreover, the MAPK-activated
RSK1 also phosphorylates TSC2 at a unique site. This RSK1-mediated
phosphorylation inhibits the TSC1eTSC2 complex and thereby increases
mTORC1 signaling toward S6K1 (Roux et al., 2004)(Figure 2).
4.2 Amino Acid Sufciency
Amino acid starvation, unlike serum starvation, fails to downregulate PI3K
or Akt (Hara et al., 1998; Wang et al., 1998), yet it results in a rapid dephos-
phorylation of S6K1 and rpS6. Furthermore, reintroduction of amino acids
restores the phosphorylation of these targets in an mTORC1-dependent
(rapamycin-sensitive) fashion (Tang et al., 2001) and references therein). It
was widely argued that the TSC1eTSC2 complex plays no role in trans-
ducing the negative signal resulting from amino acid starvation to
mTORC1 activity (Bar-Peled and Sabatini, 2014; Jewell et al., 2013).
However, recent studies have shown that inhibition of mTORC1 by amino
acid deprivation is indeed mediated by the TSC1eTSC2 complex
(Demetriades et al., 2014; Patursky-Polischuk et al., 2014). The latter is
Figure 2 Pathways transducing signals emanating from growth factors to rpS6 phos-
phorylation. Arrows represent activation and bars inhibition. See text for details.
Ribosomal Protein S6 Phosphorylation 51
required for the release of mTORC1 from its site of action, the lysosomal
membrane, and therefore, cells lacking TSC fail to efciently turn off
mTORC1 and consequently their response to amino acid starvation is
compromised (Demetriades et al., 2014).
Amino acid starvation leads to rapid dephosphorylation of S6K1, which
can be restored upon readdition of amino acids in an mTORC1-dependent
fashion ((Kim and Guan, 2011) and references therein). Members of the Rag
subfamily of Ras small GTPases (RagA, B, C, and D) and the trimeric
complex, Ragulator, are essential transducers of amino acids signals to
mTORC1 activity (Duran and Hall, 2012). Amino acid stimulation elicits
movement of mTORC1 to the lysosomal surface, where Rheb and Ragu-
lator reside. The latter recruits Rag GTPases to the lysosomes in a p62- and
vacuolar H
þ
-ATPase-dependent manner (Duran et al., 2011; Zoncu et al.,
2011), and thereby participates in mTORC1 activation (Sancak et al., 2008,
2010). In contrast, the pentameric complex, GATOR1, inhibits the
mTORC1 pathway by functioning as a GAP for RagA, whereas the
trimeric complex, GATOR2 negatively regulates GATOR1 (Bar-Peled
et al., 2013).
Another mechanism whereby amino acids could affect mTORC1 activ-
ity is via hVPS34 (vacuolar protein sorting 34). This class III PI3K (converts
phosphatidylinositol to phosphatidylinositol-3-phosphate) has been shown
to transduce the signal of amino acid sufciency to mTORC1 indepen-
dently of the TSC1eTSC2/Rheb axis (Byeld et al., 2005; Nobukuni
et al., 2005). Nevertheless, the mechanism by which hVPS34 regulates
mTOR is unknown (Figure 3).
4.3 Energy Balance
Starving mammalian cells of glucose or treating them with glycolytic (e.g.,
2-deoxyglucose (2-DG)) or mitochondrial (e.g., valinomycin, antimycin A,
oligomycin) inhibitors, depletes cellular energy and causes a concomitant
decrease in mTORC1 activity (Dennis et al., 2001; Inoki et al., 2003;
Kim et al., 2002). Activation of AMP-activated protein kinase (AMPK) is
currently the prevailing model used to explain how energy levels couple
to the regulation of mTORC1. AMPK acts as a sensor of cellular energy
status and is activated by increases in the cellular AMP: ATP ratio, caused
by metabolic stresses that either interfere with ATP production (e.g., depri-
vation for glucose or oxygen) or that accelerate ATP consumption (e.g.,
muscle contraction). Activation in response to increases in AMP levels in-
volves phosphorylation by an upstream kinase, the tumor suppressor
52 Oded Meyuhas
LKB1 (Towler and Hardie, 2007), since AMPK activation in response to
low energy conditions is blocked in LKB1 null cells (Corradetti et al.,
2004). Furthermore, LKB1 mutant cells exhibit hyperactive mTORC1
signaling (Corradetti et al., 2004; Shaw et al., 2004). Activation of AMPK
by 5-aminoimidazole-4-carboxyamide (AICAR), an AMP analog, inhibits
mTORC1-dependent phosphorylation of S6K1 (Bolster et al., 2002). Like-
wise, expression of an activated form of AMPK decreases S6K1 phosphor-
ylation, whereas a dominant-negative form of AMPK increases S6K1
phosphorylation (Kimura et al., 2003).
AMPK phosphorylates several targets to enhance catabolism and suppress
anabolism in response to low energy, and exerts this effect by directly phos-
phorylating and activating TSC2 and thereby downregulates mTORC1
(Inoki et al., 2003). Thus, the phosphorylation of S6K1 is more resistant
to glucose deprivation in TSC2
/
cells or cells, whose mutant TSC2
Figure 3 Amino acid signaling to rpS6 phosphorylation. Arrows represent activation,
bars inhibition and dotted lines putative links. See text for details.
Ribosomal Protein S6 Phosphorylation 53
cannot be phosphorylated by AMPK (Inoki et al., 2003). It appears, there-
fore, that energy depletion is sensed by AMPK and relayed to mTORC1
through the TSC1eTSC2 complex. However, it seems that cells can
convey energy stress signals to TSC2 also through upregulation of the
mRNA encoding REDD1 (Regulated in Development and Damage
Responses 1, also known as RTP801). REDD1 binds 14-3-3 and thereby
alleviates the 14-3-3-mediated inhibition of TSC1eTSC2 complex
(DeYoung et al., 2008) and consequently leads to inhibition of mTORC1
signaling to S6K1 (Sofer et al., 2005). Yet, the fact that mTORC1 is refrac-
tory to energy starvation in REDD1
/
cells, despite normal activation of
AMPK and AMPK-dependent activation of TSC2, suggests that the effect
of REDD1 on TSC2 is predominant over that of AMPK. This notion is
further supported by the observation that overexpression of REDD1 can
suppress mTORC1 activity even in the presence of dominant-negative
AMPK (Sofer et al., 2005). Taken together, these results imply that
REDD1 may act in parallel with, or downstream of AMPK toward TSC2
(Figure 4). It should be noted, however, that conicting results have shown
that acute treatment of TSC2
/
with 2-DG leads to inactivation of S6K1
(Smith et al., 2005), suggesting that signals from energy starvation might be
transduced into suppression of mTORC1 also in a TSC1eTSC2 complex-
independent fashion.
4.4 Oxygen Supply
Hypoxia has a prominent inhibitory effect on mTORC1 activity, which is
mediated in part by REDD1 (Figure 4), as REDD1
/
mouse cells are
defective in hypoxia-mediated inhibition of S6K activation (Brugarolas
et al., 2004). This effect, however, relies on an intact TSC1eTSC2 com-
plex, as S6K phosphorylation is refractory to hypoxic treatment of
TSC1
/
or TSC2
/
mouse cells (Brugarolas et al., 2004; Miloslavski
et al., 2014). Transcriptional activation of REDD1 gene under conditions
of hypoxia (Shoshani et al., 2002) is mediated by HIF-1, the master regula-
tors of oxygen homeostasis (Brahimi-Horn et al., 2007). Hypoxia can also
inhibit mTORC1 independently of REDD1 via the induction of energy
stress, possibly due to reduced oxidative phosphorylation. AMPK is upregu-
lated under these conditions, thereby activates TSC2 and inhibits mTORC1
(Liu et al., 2006)(Figure 4). It should be noted, however, that prolonged
exposure to low oxygen leads to a reduced mTORC1 activity indepen-
dently of TSC2 by an unknown mechanism (Liu et al., 2006).
54 Oded Meyuhas
4.5 Osmolarity
An increase in the concentration of solutes outside the cell relative to that
inside is termed a hyperosmotic stress. Such a stress causes water to diffuse
out of the cell, resulting in cell shrinkage, which can lead to DNA and pro-
tein damage, cell cycle arrest, and ultimately cell death (Burg et al., 2007).
Hyperosmotic stress that is induced by treating cells with either sorbitol
(Kruppa and Clemens, 1984) or high salt concentration (Naegele and
Morley, 2004) elicits reversible dephosphorylation of rpS6 in mammals. A
closer look at the sorbitol effect has suggested the involvement of a phospha-
tase, since calyculin A, a phosphatase inhibitor, was able to prevent sorbitol-
induced suppression of S6K (Parrott and Templeton, 1999).
Figure 4 Pathways transducing energy balance and hypoxic signals to rpS6 phosphor-
ylation. Arrows represent activation and bars inhibition. See text for details.
Ribosomal Protein S6 Phosphorylation 55
Hyperosmotic-dependent S6K inhibition has been shown also in
mannitol-treated tobacco leaves. Downregulation of S6K activity appears
to have a protective effect against sustained osmotic stress, as Arabidopsis
plants expressing high levels of S6K were hypersensitive to mannitol treat-
ment (Mahfouz et al., 2006; Williams et al., 2003).
5. PHYSIOLOGICAL ROLES OF rpS6
PHOSPHORYLATION
5.1 Global Protein Synthesis
An early study has shown that substitution of the two phosphorylat-
able serine residues to alanines in yeast rpS6 had no detectable effect on yeast
growth, under a wide variety of nutritional conditions (Johnson and
Warner, 1987). This observation implies that rpS6 phosphorylation has no
obvious role in protein synthesis or other cellular functions in yeast. A similar
approach has also been applied to studying the role of rpS6 phosphorylation
in regulation of protein synthesis in rpS6 phosphorylation-decient
mammalian cells (rpS6
P/
knockin MEFs). Thus, the relative rates of
global protein synthesis (incorporation of radio-labeled amino acids) and
of the accumulation of steady state levels of protein were signicantly higher
in these cells, relative to wild-type MEFs (Ruvinsky et al., 2005). It appears
therefore, that protein synthesis, at least in this cell type, is downregulated by
rpS6 phosphorylation. Though slightly faster elongation rate was deter-
mined in rpS6
P/
MEFs, the augmentation in overall protein synthesis
in these cells is mainly attributable to enhanced translation initiation by an
as yet unknown mechanism.
A lack of stimulatory effect of rpS6 phosphorylation on global protein
synthesis has also been shown in mouse liver. Thus, monitoring the relative
proportion of ribosomes engaged in translation (associated with polysomes)
has demonstrated a similar proportion in the liver of both rpS6
P/
and
wild-type mice. Furthermore, this similarity was apparent even in regener-
ating liver, in which rpS6 undergoes extensive phosphorylation only in the
wild-type (Ruvinsky et al., 2005). Consistently, mice treated for 4 weeks
with rapamycin showed a dramatic reduction in the phosphorylation of
rpS6 in both liver and muscle, yet their translational activity was indistin-
guishable from that monitored in mice treated with just a vehicle (Garelick
et al., 2013).
Notably, a lack of a role in the regulation of global protein synthesis has
been reported also for S6K-decient (S6K1
/
;S6K2
/
) mice. Thus,
56 Oded Meyuhas
myoblasts, hepatocytes, broblasts, and a whole liver from both wild-type
and S6K-decient mice displayed a similar proportion of ribosomes engaged
in polysomes (Chauvin et al., 2014; Mieulet et al., 2007). Similarly,
measuring protein synthesis by methionine incorporation, showed no differ-
ence between wild-type and S6K1
/
;S6K2
/
mice (Mieulet et al., 2007).
Clearly, these results indicate that both rpS6 phosphorylation and S6K activ-
ity are dispensable for efcient global protein synthesis.
5.2 Cell Size Regulation
Previous reports have demonstrated that the mTORC1 pathway is an inte-
gral cell growth regulator (Lee et al., 2007). Thus, treatment of mammalian
cells by rapamycin decreases their size. This mTOR-dependent regulation
of the cell size involves its downstream targets, S6K1 and 4E-BP (Fingar
et al., 2002; Ohanna et al., 2005). Indeed, S6K has been implicated as an
important positive regulator of cell and body size. Thus, most dS6K null
Drosophila exhibit embryonic lethality, with the few surviving adults having
a severely reduced body size, due to a decrease in cell size rather than a
decrease in cell number (Montagne et al., 1999). S6K1
/
mice are signif-
icantly smaller at birth, due to a proportional decrease in the size of all organs
(Shima et al., 1998). A smaller cell size in these mice was reported for
pancreatic bcells (Pende et al., 2000) and myoblasts (Ohanna et al.,
2005). In contrast, the birth weight of S6K2
/
mice, as well as the size
of their myoblasts, is similar to those of wild-type mice (Pende et al.,
2000; Ohanna et al., 2005). In accordance with the phenotypes of each of
these mutant mice, the embryonic and postnatal growth, as well as the
size of myoblasts of the double knockout mice, S6K1
/
/S6K2
/
, are
similar to those of S6K1
/
mice (Ohanna et al., 2005; Pende et al.,
2000). The fact that mammalian cell size is predominantly determined by
S6K1 and not S6K2 posed a question regarding the effector(s) of S6K1
involved in this mode of regulation.
Of the known multiple substrates of S6K1, it is rpS6 phosphorylation
that is directly involved in the control of cell size. Thus, a wide variety
of cell types derived from rpS6
P/
mice are signicantly smaller than their
wild-type counterparts. These include pancreatic bcells, interleukin-7-
dependent cells derived from fetal livers, MEFs (Granot et al., 2009;
Ruvinsky et al., 2005), muscle myotubes (Ruvinsky et al., 2009). It ap-
pears, however, that the small cell phenotype is not ubiquitous, as acinar
cells in the pancreas display a similar size regardless of the absence of
S6K1 (Pende et al., 2000) or phosphorylatable serine residues in rpS6
Ribosomal Protein S6 Phosphorylation 57
(Ruvinsky et al., 2005). Notably, even though the birth weight of
rpS6
P/
mice is similar to that of their wild-type littermates, the knock
mice start to display relative retarded weight gain at the age of 6 weeks
(Ruvinsky et al., 2005, 2009).
Several studies have demonstrated that cell cycle progression and cell
growth are separable and are therefore distinct processes, at least in some
mammalian cells (Conlon et al., 2004; Fingar and Blenis, 2004). The
apparent small size of rpS6
P/
MEFs are accompanied by accelerated divi-
sion (Ruvinsky et al., 2005), yet several lines of evidence lend support to the
notion that the small cell size phenotype reects impaired growth, rather
than being a by-product of enhanced cell division: (1) rpS6
P/
MEFs
remained smaller than their wild-type counterparts, even when progression
through the cell cycle was arrested by aphidicolin, an inhibitor of DNA po-
lymerase- ; and;(2) The size of immortalized rpS6
P/
MEFs is increased to
the extent that it equalize with that of rpS6
Pþ/þ
MEFs. Nevertheless, this
increase was not accompanied by lengthening of the doubling time, as
would be expected if the size was inversely proportional to the division
rate (Ruvinsky et al., 2005).
Interestingly, rapamycin treatment decreased the size of rpS6
Pþ/þ
MEFs, whereas the size of rpS6
P/
MEFs remained unchanged
(Ruvinsky et al., 2005). The rapamycin-resistance displayed by the latter
is reminiscent of that exhibited by S6K1
/
myoblasts (Ohanna et al.,
2005), implying that cells already displaying a small size phenotype, due
to deciency of S6K1 or of phosphorylatable serine residues in rpS6, are
not further affected by rapamycin. Furthermore, it seems that rpS6 phos-
phorylation is a critical effector of mTORC1 in regulation of cell growth
and that its absence is equivalent to inhibition of mTORC1. Notably, the
small size of S6K1
/
myoblasts is apparent, even though their rpS6 is still
phosphorylated, most probably by S6K2 (Ohanna et al., 2005). It is likely,
therefore, that this small cell size phenotype reects a reduced activity of
yet another S6K1-specic effector(s), which is involved in this mode of
regulation, such as SKAR (Richardson et al., 2004). Alternatively, if
S6K2 is inactive during muscle differentiation in early embryo, then it is
possible that S6K1 deciency is indeed equivalent to the lack of phosphor-
ylatable serine in rpS6. If the latter is the case, then it should be assumed
that once the growth of a specic cell lineage is blocked by a temporary
deciency of rpS6 phosphorylation, as a result of S6K1 deciency, the
small size is maintained thereafter, regardless of a later phosphorylation
of rpS6 by a different kinase (S6K2, for example).
58 Oded Meyuhas
5.3 Normal Muscle Function
rpS6
P/
mice suffer from muscle weakness as demonstrated by a variety of
physical performance tests (Ruvinsky et al., 2009). This physical inferiority
appears to result from two defects: (1) a decrease in total muscle mass that
reects impaired growth, rather than aberrant differentiation of myobers,
as well as a diminished abundance of contractile proteins; and (2) a reduced
content of ATP and phosphocreatine, two readily available energy sources.
However, the apparent energy deciency in this genotype does not result
from a lower mitochondrial mass or compromised activity of enzymes of
the oxidative phosphorylation, nor does it reect a decline in insulin-depen-
dent glucose uptake, or diminution in storage of glycogen or triacylglycerol
in the muscle (Ruvinsky et al., 2009). These observations have established
rpS6 phosphorylation as a determinant of muscle strength, through its role
in regulation of myober growth and energy content.
5.4 Hypertrophic Responses
The apparent role of rpS6 phosphorylation as a determinant of cell size raised
the possibility that this modication might also be critical during induced
cellular hypertrophy. One such response is the compensatory renal hyper-
trophy, which results from reduction in the number of functioning neph-
rons, as in the case in unilateral nephrectomy (Preisig, 1999). This
response is similarly blunted by either rapamycin treatment or in S6K1-
decient mice (Chen et al., 2009), indicating the role of the mTORC1/
S6K1 axis in this protective mechanism. Interestingly, among its multiple
substrates, S6K1 transduces its growth signal primarily through rpS6 phos-
phorylation, as is evident by the greatly impaired compensatory renal hyper-
trophy in rpS6
P/
mice (Xu et al., 2015). Surprisingly, induced
hypertrophy of muscle in adolescent rats following progressive resistance
exercise was shown to be associated with reduced phosphorylation of
rpS6 at Ser235/236 (Hellyer et al., 2012). Conceivably, once maximal
growth is attained following repeated bouts of exercise, signaling to rpS6
phosphorylation is silenced.
5.5 Cell Proliferation
The puzzling observations of a similar birth weight of rpS6
P/
and
rpS6
Pþ/þ
mice, despite a smaller size of rpS6
P/
embryonic cells (MEFs
and interleukin-7-dependent cells), have been reconciled by the ndings
that rpS6
P/
newborns contain a higher DNA content, which reects a
Ribosomal Protein S6 Phosphorylation 59
higher cell number (Ruvinsky et al., 2005). It is conceivable, therefore, that
a faster proliferation compensates for the smaller size of embryonic rpS6
P/
cells. Indeed, this possibility is further supported by the apparent shorter
population doubling time of rpS6
P/
MEFs, as well as the faster protein
and nucleic acids accumulation in these cells. This accelerated cell division
primarily reects a shortening of the G1 phase in rpS6
P/
MEFs (Ruvinsky
et al., 2005). Notably, the deciency of both S6K1 and S6K2, unlike the
mutation in all phosphorylatable serine residues in rpS6
P/
, had no effect
on the doubling time of MEFs or primary myoblasts (Ohanna et al., 2005;
Pende et al., 2004). This difference might reect the fact that rpS6 in
S6K1
/
/S6K2
/
is still phosphorylated at Ser235/236.
5.6 Clearance of Apoptotic Cells
Professional and amateur phagocytes rapidly clear apoptotic cells in a process
known as efferocytosis (Korns et al., 2011). It has previously been shown
that the F-box protein Pallbearer (PALL) participates in a complex func-
tioning as an E3-ubiquitin ligase. This complex promotes efcient apoptotic
cell clearance in Drosophila (Silva et al., 2007). A study with Drosophila
Schneider S2 cell line has unveiled the role of PALL in proteasomal degra-
dation of rpS6 preferentially in its phosphorylated form (Xiao et al., 2015).
Evidently, rpS6 appears to act as a negative regulator of efferocytosis, since its
knockdown enhances, whereas its overexpression decreases the engulfment
of apoptotic cells. Finally, the PALL-dependent degradation of rpS6 leads to
upregulation and activation of RAC2 GTPase that is followed by actin
remodeling to promote efferocytosis (Xiao et al., 2015). Interestingly,
rpS6 phosphorylation has been implicated in TRAIL (tumor necrosis
factor-related apoptosis-inducing ligand)-induced apoptosis in mammalian
cells. Thus, rpS6
P/
MEFs were more sensitive to TRAIL than wild-
type MEFs. Yet, they were as sensitive as wild-type cells to the topoisomer-
ase inhibitor, etoposide (Jeon et al., 2008).
5.7 Tumorigenicity
The importance of mTORC1 in cancer is well appreciated (Xu et al., 2014),
yet it is believed that the key downstream effector of this pathway in cancer
is 4E-BP (Hsieh et al., 2010). Nevertheless, S6K has also been implicated as
an important player in the development of cancer (Alliouachene et al.,
2008). Not surprisingly, therefore, that rpS6 phosphorylation has attracted
much attention as a diagnostic maker for various types of tumors (see
60 Oded Meyuhas
Section 6 below). However, it is only recently that rpS6 phosphorylation has
proven instructive for neoplastic transformation.
Pancreatic ductal adenocarcinoma (PDAC) is one of the most lethal hu-
man cancers. Oncogenic mutations in Kras are found in more than 95% of
PDACs and appear to drive the formation of precursor lesions (Vincent
et al., 2011). Several studies have shown that pharmacologic inhibition of
mTORC1 can attenuate the growth of pancreatic cancer cell lines and, at
least, a subset of PDAC in vivo (Mirzoeva et al., 2011; Morran et al.,
2014), but the identity of downstream effectors remains largely unknown.
Phosphorylation of rpS6 was increased in pancreatic acinar cells upon
implantation of the chemical carcinogen 7,12-dimethylbenz(a)anthracene
(DMBA) or transgenic expression of mutant Kras. Strikingly, the develop-
ment of pancreatic cancer precursor lesions induced by either DMBA or
mutant Kras was greatly reduced in rpS6
P/
mice. rpS6 phosphoryla-
tion-decient mice, expressing oncogenic Kras, showed increased nuclear
accumulation of the tumor suppressor, p53. This increase is accompanied
with enhanced staining of the DNA damage markers g-H2AX and 53bp1
(Trp53bp1) in areas of acinar to ductal metaplasia, suggesting that rpS6 phos-
phorylation attenuates Kras-induced DNA damage and p53-mediated tu-
mor suppression (Khalaileh et al., 2013). These results attest to the critical
role of rpS6 phosphorylation in the initiation of pancreatic cancer.
A previously published report has shown that pancreata in about 30% of
mice expressing a constitutively active myr-Akt1 in their bcells underwent
hyperplastic transformation leading to insulinoma formation. However,
deciency of S6K1, but not of S6K2, fully protected the animals from
this tumorigenesis (Alliouachene et al., 2008). The readily detectable
phosphorylated rpS6 in myr-Akt1 transgenic islets, despite S6K1 deciency,
might argue against a tumorigenic role of rpS6 phosphorylation. However,
rpS6 phosphorylation deciency in rpS6
P/
mice led to a complete protec-
tion against the development of myr-Akt1-induced insulinoma (Dreazen
et al., unpublished data). These observations imply that not only S6K1,
but also rpS6 phosphorylation can promote malignant transformation, albeit
through distinct mechanisms.
It should be noted that the role of rpS6 phosphorylation in Akt-mediated
tumorigenesis should not be referred to as a general mechanism, but rather a
tissue-specic phenomenon. Thus, Akt
T
mice, in which a constitutively
active Akt2 is expressed in immature T cells, develop spontaneous thymic
lymphomas, which cannot be prevented in rpS6
P/
;Akt
T
double mutant
mice. It appears, therefore, that rpS6 phosphorylation is dispensable for
Ribosomal Protein S6 Phosphorylation 61
transformation downstream of oncogenic Akt signaling in the thymus
(Hsieh et al., 2010).
5.8 Glucose Homeostasis
It has previously been shown that insulin secretion closely correlates with the
size of bcells (Giordano et al., 1993; Pende et al., 2000). Mice decient in
S6K1 exhibited impaired glucose homeostasis, due to insufcient insulin
secretion in response to glucose load. The reason for this defect was pro-
posed to be the small size of bcells in S6K1
/
mice (Pende et al.,
2000). This phenotype is recapitulated in rpS6
P/
mice, which show a
twofold reduction in both circulating levels and pancreatic content of insu-
lin, in addition to a higher and prolonged hyperglycemic response after
glucose challenge, compared to wild-type mice (Ruvinsky et al., 2005).
Interestingly, the apparent glucose intolerance in rpS6
P/
and
S6K1
/
mice is reminiscent of impaired glucose tolerance observed in
offsprings of rats that were undernourished during pregnancy, or in adult
human beings after prenatal exposure to famine ((Ravelli et al., 1998) and
references therein). Possibly, malnutrition during pregnancy leads to insuf-
cient signals through mTOR, an integrator of nutritional signals (Dann
et al., 2007; Proud, 2007), which in turn leads to reduced activation of
S6K1 and hypophosphorylation of rpS6 during a critical stage of pancreatic
development and consequently to impaired pancreatic function in the adult
organism. It should be pointed out, however, that the effect of perinatal
famine on the size of bcells, a hallmark of rpS6
P/
and S6K1
/
mice,
is currently unknown.
It should be noted that S6K1
/
mice, unlike rpS6
P/
mice, display an
in utero developmental defect manifested in smaller birth size (Shima et al.,
1998), and that the disruption of both S6K1 and S6K2 leads to decreased
viability due to perinatal lethality (Pende et al., 2004). Clearly, these pheno-
types attest to the involvement of S6K targets other than rpS6, in normal in
utero development.
5.9 rpS6 Phosphorylation as Diagnostic Marker
The usage of the phosphorylation state of Ser235/236 in rpS6 as a biomarker
for activation of the PI3K/mTORC1/S6K pathway in tissue samples from
tumor biopsies (Li et al., 2014; Pinto et al., 2013; Robb et al., 2006, 2007)
or transplants (Lepin et al., 2006; Li et al., 2015) has been repeatedly pro-
posed in recent years. However, these sites can also be phosphorylated by
RSK (Roux et al., 2007), and therefore, their phosphorylation cannot be
62 Oded Meyuhas
used as an indication for therapeutic strategy involving blockage of PI3K/
mTORC1/S6K signaling. Indeed, it has recently been shown that phos-
phorylation of Ser235/236 might be upregulated in tumors with activation
of the Ras/Raf/ERK pathway, rather than activation of the PI3K/
mTORC1/S6K pathway (Chow et al., 2006; Ma et al., 2007). Hence,
differential diagnosis of the activated pathway should depend on the use
of specic biomarkers, such as phospho-rpS6(Ser240/244) for the PI3K/
mTORC1/S6K pathway and phospho-ERK or phospho-RSK for the
Ras/Raf/MEK pathway. Indeed, monitoring the phosphorylation of rpS6
at Ser240/244, rather than Ser235/236, has recently been reported for
several tumors (Chaisuparat et al., 2013; Kim et al., 2013) and epidermal
hyperproliferation conditions (Ruf et al., 2014).
6. CONCLUDING REMARKS AND FUTURE
PERSPECTIVES
Many of the phenotypic manifestations of mice decient in S6K1 are
recapitulated in the rpS6 knockin mice. Thus both these mutants exhibit: (1)
small bcell size phenotype that is accompanied by hypoinsulinemia and
impaired glucose homeostasis (Pende et al., 2000; Ruvinsky et al., 2005);
(2) small myoblasts and reduced muscle mass (Ohanna et al., 2005; Ruvinsky
et al., 2009); and (3) blunted compensatory renal hypertrophy following
contralateral nephrectomy (Chen et al., 2009; Xu et al., 2015). Clearly,
based on these observations it is tempting to argue that this phenotypic sim-
ilarity simply reects the fact that rpS6 phosphorylation is a critical S6K1
effector. However, this explanation is inconsistent with the observation
that rpS6 is still fully phosphorylated in S6K1-decient mouse, due to the
compensatory activity of S6K2 (Alliouachene et al., 2008; Ohanna et al.,
2005; Shima et al., 1998). It appears, therefore, that despite the similarity
in their phenotypic manifestations, the impaired functions are caused by
distinct mechanisms operating in phosphorylation-decient and S6K1
knockout mice.
Despite a major progress in understanding the physiological roles of rpS6
phosphorylation, the mechanism underlying its highly diverse effects is
poorly understood, if at all. Several explanations can be proposed to account
for this diversity: (1) The phosphorylation of rpS6 within, or outside, the
ribosome affects the translation efciency of specic mRNAs encoding pro-
teins participating in various processes; (2) rpS6 might be one of the many
bifunctional ribosomal proteins, that can carry out extraribosomal tasks often
Ribosomal Protein S6 Phosphorylation 63
unrelated to the mechanics of protein synthesis (Warner and McIntosh,
2009); (3) Phosphorylated rpS6 might not affect protein synthesis, but
instead interacts with cellular protein(s), which consequently become active
or inactive, and thus affects the cell physiology. This notion is further
supported by reports on the coimmunoprecipitation of rpS6 with several
extraribosomal proteins, suggesting an in vivo interaction, either directly
or indirectly with these proteins (Kim et al., 2006, 2014; Schumacher
et al., 2006). Not surprisingly, therefore, rpS6
P/
mice show altered
transcription, rather than translation, of the ribosome biogenesis program
in hepatocytes (Chauvin et al., 2014). Clearly, resolving any of these mech-
anistic issues will have to wait for further studies in the coming years.
REFERENCES
Alessi, D., Kozlowski, M., Weng, Q., Morrice, N., Avruch, J., 1998. 3-Phosphoinositide-
dependent protein kinase 1 (PDK1) phosphorylates and activates the p70 S6 kinase in
vivo and in vitro. Curr. Biol. 8, 69e81.
Alliouachene, S., Tuttle, R.L., Boumard, S., Lapointe, T., Berissi, S., Germain, S.,
Jaubert, F., Tosh, D., Birnbaum, M.J., Pende, M., 2008. Constitutively active Akt1
expression in mouse pancreas requires S6 kinase 1 for insulinoma formation. J. Clin.
Invest. 118, 3629e3638.
Avruch, J., Hara, K., Lin, Y., Liu, M., Long, X., Ortiz-Vega, S., Yonezawa, K., 2006. Insulin
and amino-acid regulation of mTOR signaling and kinase activity through the Rheb
GTPase. Oncogene 25, 6361e6372.
Ballou, L., Luther, H., Thomas, G., 1991. MAP2 kinase and 70K S6 kinase lie on distinct
signalling pathways. Nature 349, 348e350.
Bandi, H.R., Ferrari, S., Kreig, J., Meyer, H.E., Thomas, G., 1993. Identication of 40 S
ribosomal protein S6 phosphorylation sites in Swiss mouse 3T3 broblasts stimulated
with serum. J. Biol. Chem. 268, 4530e4533.
Bar-Peled, L., Chantranupong, L., Cherniack, A.D., Chen, W.W., Ottina, K.A.,
Grabiner, B.C., Spear, E.D., Carter, S.L., Meyerson, M., Sabatini, D.M., 2013. A tumor
suppressor complex with GAP activity for the Rag GTPases that signal amino acid
sufciency to mTORC1. Science 340, 1100e1106.
Bar-Peled, L., Sabatini, D.M., 2014. Regulation of mTORC1 by amino acids. Trends Cell
Biol. 24, 400e406.
Barth-Baus, D., Stratton, C.A., Parrott, L., Myerson, H., Meyuhas, O., Templeton, D.J.,
Landreth, G.E., Hensold, J.O., 2002. S6 phosphorylation-independent pathways regu-
late translation of 5-terminal oligopyrimidine tract containing mRNAs in differentiating
hematopoietic cells. Nucleic Acids Res. 30, 1919e1928.
Belandia, B., Brautigan, D., Martín-Pérez, J., 1994. Attenuation of ribosomal protein S6
phosphatase activity in chicken embryo broblasts transformed by Rous sarcoma virus.
Mol. Cell. Biol. 14, 200e206.
Belham, C., Wu, S., Avruch, J., 1999. Intracellular signalling: PDK1ea kinase at the hub of
things. Curr. Biol. 9, R93eR96.
Bhaskar, P.T., Hay, N., 2007. The two TORCs and Akt. Dev. Cell 12, 487e502.
Biever, A., Puighermanal, E., Nishi, A., David, A., Panciatici, C., Longueville, S.,
Xirodimas, D., Gangarossa, G., Meyuhas, O., Herve, D., Girault, J.A., Valjent, E.,
2015. PKA-dependent phosphorylation of ribosomal protein S6 does not correlate
64 Oded Meyuhas
with translation efciency in striatonigral and striatopallidal medium-sized spiny neurons.
J. Neurosci. 35, 4113e4130.
Blenis, J., Kuo, C.J., Erikson, R.L., 1987. Identication of a ribosomal protein S6 kinase
regulated by transformation and growth-promoting stimuli. J. Biol. Chem. 262,
14373e14376.
Bolster, D.R., Crozier, S.J., Kimball, S.R., Jefferson, L.S., 2002. AMP-activated protein
kinase suppresses protein synthesis in rat skeletal muscle through down-regulated
mammalian target of rapamycin (mTOR) signaling. J. Biol. Chem. 277, 23977e23980.
Brahimi-Horn, M.C., Chiche, J., Pouyssegur, J., 2007. Hypoxia and cancer. J. Mol. Med. 85,
1301e1307.
Brazil, D.P., Hemmings, B.A., 2001. Ten years of protein kinase B signalling: a hard Akt to
follow. Trends Biochem. Sci. 26, 657e664.
Brugarolas, J., Lei, K., Hurley, R.L., Manning, B.D., Reiling, J.H., Hafen, E., Witters, L.A.,
Ellisen, L.W., Kaelin Jr., W.G., 2004. Regulation of mTOR function in response to
hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 18,
2893e2904.
Burg, M.B., Ferraris, J.D., Dmitrieva, N.I., 2007. Cellular response to hyperosmotic stresses.
Physiol. Rev. 87, 1441e1474.
Byeld, M.P., Murray, J.T., Backer, J.M., 2005. hVps34 is a nutrient-regulated lipid kinase
required for activation of p70 S6 kinase. J. Biol. Chem. 280, 33076e33082.
Cantrell, D.A., 2001. Phosphoinositide 3-kinase signalling pathways. J. Cell Sci. 114,
1439e1445.
Chaisuparat, R., Rojanawatsirivej, S., Yodsanga, S., 2013. Ribosomal protein S6 phosphor-
ylation is associated with epithelial dysplasia and squamous cell carcinoma of the oral
cavity. Pathol. Oncol. Res. 19, 189e193.
Chauvin, C., Koka, V., Nouschi, A., Mieulet, V., Hoareau-Aveilla, C., Dreazen, A.,
Cagnard, N., Carpentier, W., Kiss, T., Meyuhas, O., Pende, M., 2014. Ribosomal
protein S6 kinase activity controls the ribosome biogenesis transcriptional program.
Oncogene 33, 474e483.
Chen, J.K., Chen, J., Thomas, G., Kozma, S.C., Harris, R.C., 2009. S6 kinase 1 knockout
inhibits uninephrectomy- or diabetes-induced renal hypertrophy. Am. J. Physiol. Ren.
Physiol. 297, F585eF593.
Cheong, J.K., Virshup, D.M., 2011. Casein kinase 1: complexity in the family. Int. J. Bio-
chem. Cell Biol. 43, 465e469.
Chow, S., Minden, M.D., Hedley, D.W., 2006. Constitutive phosphorylation of the S6 ri-
bosomal protein via mTOR and ERK signaling in the peripheral blasts of acute leukemia
patients. Exp. Hematol. 34, 1183e1191.
Chung, J., Kuo, C.J., Crabtree, G.R., Blenis, J., 1992. Rapamycin-FKBP specically
blocks growth-dependent activation of and signaling by the 70 kd S6 kinases. Cell 69,
1227e1236.
Cisterna, B., Necchi, D., Prosperi, E., Biggiogera, M., 2006. Small ribosomal subunits
associate with nuclear myosin and actin in transit to the nuclear pores. FASEB J. 20,
1901e1903.
Conlon, I., Lloyd, A., Raff, M., 2004. Coordination of cell growth and cellecycle progres-
sion in proliferating mammalian cells. In: Hall, M.N., Raff, M., Thomas, G. (Eds.), Cell
Growth: Control of Cell Size. Cold Spring Harbor Laboratory Press, Cold Spring Har-
bor, NY, pp. 85e99.
Corradetti, M.N., Inoki, K., Bardeesy, N., DePinho, R.A., Guan, K.L., 2004. Regulation of
the TSC pathway by LKB1: evidence of a molecular link between tuberous sclerosis
complex and Peutz-Jeghers syndrome. Genes Dev. 18, 1533e1538.
Dann, S.G., Selvaraj, A., Thomas, G., 2007. mTOR Complex1-S6K1 signaling: at the cross-
roads of obesity, diabetes and cancer. Trends Mol. Med. 13, 252e259.
Ribosomal Protein S6 Phosphorylation 65
Demetriades, C., Doumpas, N., Teleman, A.A., 2014. Regulation of TORC1 in response to
amino acid starvation via lysosomal recruitment of TSC2. Cell 156, 786e799.
Dennis, P.B., Jaeschke, A., Saitoh, M., Fowler, B., Kozma, S.C., Thomas, G., 2001.
Mammalian TOR: a homeostatic ATP sensor. Science 294, 1102e1105.
DeYoung, M.P., Horak, P., Sofer, A., Sgroi, D., Ellisen, L.W., 2008. Hypoxia regulates
TSC1/2-mTOR signaling and tumor suppression through REDD1-mediated 14-3-3
shuttling. Genes Dev. 22, 239e251.
Duran, A., Amanchy, R., Linares, J.F., Joshi, J., Abu-Baker, S., Porollo, A., Hansen, M.,
Moscat, J., Diaz-Meco, M.T., 2011. p62 is a key regulator of nutrient sensing in the
mTORC1 pathway. Mol. Cell 44, 134e146.
Duran, R.V., Hall, M.N., 2012. Regulation of TOR by small GTPases. EMBO Rep. 13 (2),
121e128.
Erikson, E., Maller, J., 1985. A protein kinase from Xenopus eggs specic for ribosomal pro-
tein S6. Proc. Natl. Acad. Sci. U.S.A. 82, 742e746.
Fingar, D.C., Blenis, J., 2004. Target of rapamycin (TOR): an integrator of nutrient and
growth factor signals and coordinator of cell growth and cell cycle progression. Onco-
gene 23, 3151e3171.
Fingar, D.C., Salama, S., Tsou, C., Harlow, E., Blenis, J., 2002. Mammalian cell size is
controlled by mTOR and its downstream targets S6K1 and 4EBP1/eIF4E. Genes
Dev. 16, 1472e1487.
Flotow, H., Thomas, G., 1992. Substrate recognition determinants of the mitogen-activated
70K S6 kinase from rat liver. J. Biol. Chem. 267, 3074e3078.
Fromont-Racine, M., Senger, B., Saveanu, C., Fasiolo, F., 2003. Ribosome assembly in
eukaryotes. Genes Cells 313, 17e42.
Garelick, M.G., Mackay, V.L., Yanagida, A., Academia, E.C., Schreiber, K.H.,
Ladiges, W.C., Kennedy, B.K., 2013. Chronic rapamycin treatment or lack of S6K1
does not reduce ribosome activity in vivo. Cell Cycle 12, 2493e2504.
Giordano, E., Cirulli, V., Bosco, D., Rouiller, D., Halban, P., Meda, P., 1993. B-cell size
inuences glucose-stimulated insulin secretion. Am. J. Physiol. 265, C358eC364.
Gonzalez, A., Shimobayashi, M., Eisenberg, T., Merle, D.A., Pendl, T., Hall, M.N.,
Moustafa, T., 2015. TORC1 promotes phosphorylation of ribosomal protein S6 via
the AGC kinase Ypk3 in Saccharomyces cerevisiae. PLoS One 10, e0120250.
Granot, Z., Swisa, A., Magenheim, J., Stolovich-Rain, M., Fujimoto, W., Manduchi, E.,
Miki, T., Lennerz, J.K., Stoeckert Jr., C.J., Meyuhas, O., Seino, S., Permutt, M.A.,
Piwnica-Worms, H., Bardeesy, N., Dor, Y., 2009. LKB1 regulates pancreatic beta cell
size, polarity, and function. Cell Metab. 10, 296e308.
Gressner, A.M., Wool, I.G., 1974. The phosphorylation of liver ribosomal proteins in vivo.
Evidence that only a single small subunit protein (S6) is phosphorylated. J. Biol. Chem.
249, 6917e6925.
Hara, K., Yonezawa, K., Weng, Q.-P., Kozlowski, M.T., Belham, C., Avruch, J., 1998.
Amino acid sufciency and mTOR regulate p70 S6 kinase and eIF-4E BP1 through a
common effector mechanism. J. Biol. Chem. 273, 14484e14494.
Hauge, C., Frodin, M., 2006. RSK and MSK in MAP kinase signalling. J. Cell. Sci. 119,
3021e3023.
Hay, N., Sonenberg, N., 2004. Upstream and downstream of mTOR. Genes Dev. 18,
1926e1945.
Hebert, J., Pierre, M., Loeb, J.E., 1977. Phosphorylation in vitro and in vivo of ribosomal
proteins from Saccharomyces cerevisiae. Eur. J. Biochem. 72, 167e174.
Hellyer, N.J., Nokleby, J.J., Thicke, B.M., Zhan, W.Z., Sieck, G.C., Mantilla, C.B., 2012.
Reduced ribosomal protein s6 phosphorylation after progressive resistance exercise in
growing adolescent rats. J. strength Cond. Res. 26, 1657e1666.
66 Oded Meyuhas
Hsieh, A.C., Costa, M., Zollo, O., Davis, C., Feldman, M.E., Testa, J.R., Meyuhas, O.,
Shokat, K.M., Ruggero, D., 2010. Genetic dissection of the oncogenic mTOR pathway
reveals druggable addiction to translational control via 4EBP-eIF4E. Cancer Cell 17,
249e261.
Hutchinson, J.A., Shanware, N.P., Chang, H., Tibbetts, R.S., 2011. Regulation of ribosomal
protein S6 phosphorylation by casein kinase 1 and protein phosphatase 1. J. Biol. Chem.
286, 8688e8696.
Inoki, K., Li, L., Zhu, J., Wu, J., Guan, K.L., 2002. TSC2 is phosphorylated and inhibited by
Akt and suppresses mTOR signalling. Nat. Cell Biol. 4, 648e657.
Inoki, K., Zhu, T., Guan, K.L., 2003. TSC2 mediates cellular energy response to control cell
growth and survival. Cell 115, 577e590.
Jakubowicz, T., 1985. Phosphorylation-dephosphorylation changes in yeast ribosomal
proteins S2 and S6 during growth under normal and hyperthermal conditions. Acta
Biochim. Pol. 32, 7e12.
Jeno, P., Ballou, L., Novak-Hofer, I., Thomas, G., 1988. Identication and characterization
of a mitogenic-activated S6 kinase. Proc. Natl. Acad. Sci. U.S.A. 85, 406e410.
Jeon, Y.J., Kim, I.K., Hong, S.H., Nan, H., Kim, H.J., Lee, H.J., Masuda, E.S., Meyuhas, O.,
Oh, B.H., Jung, Y.K., 2008. Ribosomal protein S6 is a selective mediator of TRAIL-
apoptotic signaling. Oncogene 27, 4344e4352.
Jewell, J.L., Russell, R.C., Guan, K.L., 2013. Amino acid signalling upstream of mTOR.
Nat. Rev. Mol. Cell. Biol. 14, 133e139.
Johnson, S., Warner, J., 1987. Phosphorylation of the Saccharomyces cerevisiae equivalent of
ribosomal protein S6 has no detectable effect on growth. Mol. Cell. Biol. 7, 1338e1345.
Kabat, D., 1970. Phosphorylation of ribosomal proteins in rabbit reticulocytes. Characteriza-
tion and regulatory aspects. Biochemistry 9, 4160e4175.
Khalaileh, A., Dreazen, A., Khatib, A., Apel, R., Swisa, A., Kidess-Bassir, N., Maitra, A.,
Meyuhas, O., Dor, Y., Zamir, G., 2013. Phosphorylation of ribosomal protein S6 atten-
uates DNA damage and tumor suppression during development of pancreatic cancer.
Cancer Res. 73, 1811e1820.
Kim, D.H., Sarbassov, D.D., Ali, S.M., King, J.E., Latek, R.R., Erdjument-Bromage, H.,
Tempst, P., Sabatini, D.M., 2002. mTOR interacts with raptor to form a nutrient-
sensitive complex that signals to the cell growth machinery. Cell 110, 163e175.
Kim, J., Guan, K.L., 2011. Amino acid signaling in TOR activation. Annu. Rev. Biochem.
80, 1001e1032.
Kim, S.H., Jang, Y.H., Chau, G.C., Pyo, S., Um, S.H., 2013. Prognostic signicance and
function of phosphorylated ribosomal protein S6 in esophageal squamous cell
carcinoma. Mod. Pathol. 26, 327e335.
Kim, T.S., Jang, C.Y., Kim, H.D., Lee, J.Y., Ahn, B.Y., Kim, J., 2006. Interaction of Hsp90
with ribosomal proteins protects from ubiquitination and proteasome-dependent
degradation. Mol. Biol. Cell 17, 824e833.
Kim, Y.K., Kim, S., Shin, Y.J., Hur, Y.S., Kim, W.Y., Lee, M.S., Cheon, C.I., Verma, D.P.,
2014. Ribosomal protein S6, a target of rapamycin, is involved in the regulation of rRNA
genes by possible epigenetic changes in Arabidopsis. J. Biol. Chem. 289, 3901e3912.
Kimura, N., Tokunaga, C., Dalal, S., Richardson, C., Yoshino, K., Hara, K., Kemp, B.E.,
Witters, L.A., Mimura, O., Yonezawa, K., 2003. A possible linkage between AMP-
activated protein kinase (AMPK) and mammalian target of rapamycin (mTOR)
signalling pathway. Genes Cells 8, 65e79.
Kirschner, L.S., Yin, Z., Jones, G.N., Mahoney, E., 2009. Mouse models of altered protein
kinase A signaling. Endocr. Relat. Cancer 16, 773e793.
Korns, D., Frasch, S.C., Fernandez-Boyanapalli, R., Henson, P.M., Bratton, D.L., 2011.
Modulation of macrophage efferocytosis in inammation. Front. Immun. 2, 57.
Ribosomal Protein S6 Phosphorylation 67
Krieg, J., Hofsteenge, J., Thomas, G., 1988. Identication of the 40 S ribosomal protein S6
phosphorylation sites induced by cycloheximide. J. Biol. Chem. 263, 11473e11477.
Kruppa, J., Clemens, M.J., 1984. Differential kinetics of changes in the state of phosphory-
lation of ribosomal protein S6 and in the rate of protein synthesis in MPC 11 cells during
tonicity shifts. EMBO J. 3, 95e100.
Lara, R., Seckl, M.J., Pardo, O.E., 2013. The p90 RSK family members: common functions
and isoform specicity. Cancer Res. 73, 5301e5308.
Lawlor, M.A., Alessi, D.R., 2001. PKB/Akt: a key mediator of cell proliferation, survival and
insulin responses? J. Cell Sci. 114, 2903e2910.
Lee, C.H., Inoki, K., Guan, K.L., 2007. mTOR pathway as a target in tissue hypertrophy.
Annu. Rev. Pharmacol. Toxicol. 47, 443e467.
Lepin, E.J., Zhang, Q., Zhang, X., Jindra, P.T., Hong, L.S., Ayele, P., Peralta, M.V.,
Gjertson, D.W., Kobashigawa, J.A., Wallace, W.D., Fishbein, M.C., Reed, E.F.,
2006. Phosphorylated S6 ribosomal protein: a novel biomarker of antibody-mediated
rejection in heart allografts. Am. J. Transpl. 6, 1560e1571.
Leslie, N.R., Downes, C.P., 2002. PTEN: the down side of PI 3-kinase signalling. Cell.
Signal. 14, 285e295.
Li, F., Wei, J., Valenzuela, N.M., Lai, C., Zhang, Q., Gjertson, D., Fishbein, M.C.,
Kobashigawa, J.A., Deng, M., Reed, E.F., 2015. Phosphorylated S6 kinase and S6 ribo-
somal protein are diagnostic markers of antibody-mediated rejection in heart allografts.
J. Heart Lung Transpl. 34, 580e587.
Li, S., Kong, Y., Si, L., Chi, Z., Cui, C., Sheng, X., Guo, J., 2014. Phosphorylation of
mTOR and S6RP predicts the efcacy of everolimus in patients with metastatic renal
cell carcinoma. BMC Cancer 14, 376.
Li, Y., Mitsuhashi, S., Ikejo, M., Miura, N., Kawamura, T., Hamakubo, T., Ubukata, M.,
2012. Relationship between ATM and ribosomal protein S6 revealed by the chemical
inhibition of Ser/Thr protein phosphatase type 1. Biosci. Biotechnol. Biochem. 76,
486e494.
Lipsius, E., Walter, K., Leicher, T., Phlippen, W., Bisotti, M.A., Kruppa, J., 2005. Evolu-
tionary conservation of nuclear and nucleolar targeting sequences in yeast ribosomal pro-
tein S6A. Biochem. Biophys. Res. Commun. 333, 1353e1360.
Liu, L., Cash, T.P., Jones, R.G., Keith, B., Thompson, C.B., Simon, M.C., 2006.
Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol.
Cell 21, 521e531.
Ma, L., Chen, Z., Erdjument-Bromage, H., Tempst, P., Pandol, P.P., 2005. Phosphoryla-
tion and functional inactivation of TSC2 by Erk implications for tuberous sclerosis and
cancer pathogenesis. Cell 121, 179e193.
Ma, L., Teruya-Feldstein, J., Bonner, P., Bernardi, R., Franz, D.N., Witte, D., Cordon-
Cardo, C., Pandol, P.P., 2007. Identication of S664 TSC2 phosphorylation as a
marker for extracellular signal-regulated kinase mediated mTOR activation in tuberous
sclerosis and human cancer. Cancer Res. 67, 7106e7112.
Mahfouz, M.M., Kim, S., Delauney, A.J., Verma, D.P., 2006. Arabidopsis TARGET OF
RAPAMYCIN interacts with RAPTOR, which regulates the activity of S6 kinase in
response to osmotic stress signals. Plant Cell 18, 477e490.
Manning, B.D., Tee, A.R., Logsdon, M.N., Blenis, J., Cantley, L.C., 2002. Identication of
the tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of the
phosphoinositide -kinase/akt pathway. Mol. Cell 10, 151e162.
Martin, K.A., Schalm, S.S., Romanelli, A., Keon, K.L., Blenis, J., 2001. Ribosomal S6 kinase
2 inhibition by a potent C-terminal repressor domain is relieved by mitogen-activated
protein-extracellular signal-regulated kinase kinase-regulated phosphorylation. J. Biol.
Chem. 276, 7892e7898.
68 Oded Meyuhas
McCubrey, J.A., Steelman, L.S., Chappell, W.H., Abrams, S.L., Wong, E.W., Chang, F.,
Lehmann, B., Terrian, D.M., Milella, M., Tafuri, A., Stivala, F., Libra, M.,
Basecke, J., Evangelisti, C., Martelli, A.M., Franklin, R.A., 2007. Roles of the Raf/
MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Bio-
chim. Biophys. Acta 1773, 1263e1284.
McKay, M.M., Morrison, D.K., 2007. Integrating signals from RTKs to ERK/MAPK.
Oncogene 26, 3113e3121.
Meyuhas, O., 2008. Physiological roles of ribosomal protein S6: one of its kind. Int. Rev.
Cell Mol. Biol. 268, 1e37.
Meyuhas, O., Dreazen, A., 2009. Ribosomal protein S6 kinase: from TOP mRNAs to cell
size. Prog. Mol. Biol. Transl. Sci. 90, 109e153.
Mieulet, V., Roceri, M., Espeillac, C., Sotiropoulos, A., Ohanna, M., Oorschot, V.,
Klumperman, J., Sandri, M., Pende, M., 2007. S6 kinase inactivation impairs growth
and translational target phosphorylation in muscle cells maintaining proper regulation
of protein turnover. Am. J. Physiol. Cell Physiol. 293, C712eC722.
Miloslavski, R., Cohen, E., Avraham, A., Iluz, Y., Hayouka, Z., Kasir, J., Mudhasani, R.,
Jones, S.N., Cybulski, N., Ruegg, M.A., Larsson, O., Gandin, V., Rajakumar, A.,
Topisirovic, I., Meyuhas, O., 2014. Oxygen sufciency controls TOP mRNA transla-
tion via the TSC-Rheb-mTOR pathway in a 4E-BP-independent manner. J. Mol.
Cell Biol. 6, 255e266.
Mirzoeva, O.K., Hann, B., Hom, Y.K., Debnath, J., Aftab, D., Shokat, K., Korn, W.M.,
2011. Autophagy suppression promotes apoptotic cell death in response to inhibition
of the PI3K-mTOR pathway in pancreatic adenocarcinoma. J. Mol. Med. 89, 877e889.
Mizoguchi, T., Hayashida, N., Yamaguchi-Shinozaki, K., Kamada, H., Shinozaki, K., 1995.
Two genes that encode ribosomal-protein S6 kinase homologs are induced by cold or
salinity stress in Arabidopsis thaliana. FEBS Lett. 358, 199e204.
Montagne, J., Stewart, M.J., Stocker, H., Hafen, E., Kozma, S.C., Thomas, G., 1999.
Drosophila S6 kinase: a regulator of cell size. Science 285, 2126e2129.
Montgomery, S.A., Berglund, P., Beard, C.W., Johnston, R.E., 2006. Ribosomal protein S6
associates with alphavirus nonstructural protein 2 and mediates expression from alphavi-
rus messages. J. Virol. 80, 7729e7739.
Moore, C.E., Xie, J., Gomez, E., Herbert, T.P., 2009. Identication of cAMP-dependent
kinase as a third in vivo ribosomal protein S6 kinase in pancreatic beta-cells. J. Mol.
Biol. 389, 480e494.
Morran, D.C., Wu, J., Jamieson, N.B., Mrowinska, A., Kalna, G., Karim, S.A., Au, A.Y.,
Scarlett, C.J., Chang, D.K., Pajak, M.Z., Australian Pancreatic Cancer Genome Initia-
tive, Oien, K.A., McKay, C.J., Carter, C.R., Gillen, G., Champion, S., Pimlott, S.L.,
Anderson, K.I., Evans, T.R., Grimmond, S.M., Biankin, A.V., Sansom, O.J.,
Morton, J.P., 2014. Targeting mTOR dependency in pancreatic cancer. Gut 63,
1481e1489.
Naegele, S., Morley, S.J., 2004. Molecular cross-talk between MEK1/2 and mTOR
signaling during recovery of 293 cells from hypertonic stress. J. Biol. Chem. 279,
46023e46034.
Nobukuni, T., Joaquin, M., Roccio, M., Dann, S.G., Kim, S.Y., Gulati, P., Byeld, M.P.,
Backer, J.M., Natt, F., Bos, J.L., Zwartkruis, F.J., Thomas, G., 2005. Amino acids
mediate mTOR/raptor signaling through activation of class 3 phosphatidylinositol
3OH-kinase. Proc. Natl. Acad. Sci. U.S.A. 102, 14238e14243.
Ohanna, M., Sobering, A.K., Lapointe, T., Lorenzo, L., Praud, C., Petroulakis, E.,
Sonenberg, N., Kelly, P.A., Sotiropoulos, A., Pende, M., 2005. Atrophy of S6K1
/
skeletal muscle cells reveals distinct mTOR effectors for cell cycle and size control.
Nat. Cell Biol. 7, 286e294.
Ribosomal Protein S6 Phosphorylation 69
Panic, L., Tamarut, S., Sticker-Jantscheff, M., Barkic, M., Solter, D., Uzelac, M.,
Grabusic, K., Volarevic, S., 2006. Ribosomal protein S6 gene haploinsufciency is
associated with activation of a p53-dependent checkpoint during gastrulation. Mol.
Cell Biol. 26, 8880e8891.
Parrott, L.A., Templeton, D.J., 1999. Osmotic stress inhibits p70/85 S6 kinase through
activation of a protein phosphatase. J. Biol. Chem. 274, 24731e24736.
Patursky-Polischuk, I., Kasir, J., Miloslavski, R., Hayouka, Z., Hausner-Hanochi, M.,
Stolovich-Rain, M., Tsukerman, P., Biton, M., Mudhasani, R., Jones, S.N.,
Meyuhas, O., 2014. Reassessment of the role of TSC, mTORC1 and microRNAs
in amino acids-meditated translational control of TOP mRNAs. PLoS One 9,
e109410.
Pende, M., Kozma, S.C., Jaquet, M., Oorschot, V., Burcelin, R., Le Marchand-Brustel, Y.,
Klumperman, J., Thorens, B., Thomas, G., 2000. Hypoinsulinaemia, glucose intolerance
and diminished beta-cell size in S6K1-decient mice. Nature 408, 994e997.
Pende, M., Um, S.H., Mieulet, V., Sticker, M., Goss, V.L., Mestan, J., Mueller, M.,
Fumagalli, S., Kozma, S.C., Thomas, G., 2004. S6K1
/
/S6K2
/
mice exhibit peri-
natal lethality and rapamycin-sensitive 50-terminal oligopyrimidine mRNA translation
and reveal a mitogen-activated protein kinase-dependent S6 kinase pathway. Mol.
Cell. Biol. 24, 3112e3124.
Pinto, A.P., Degen, M., Barron, P., Crum, C.P., Rheinwald, J.G., 2013. Phosphorylated S6
as an immunohistochemical biomarker of vulvar intraepithelial neoplasia. Mod. Pathol.
26, 1498e1507.
Potter, C.J., Pedraza, L.G., Xu, T., 2002. Akt regulates growth by directly phosphorylating
Tsc2. Nat. Cell Biol. 4, 658e665.
Preisig, P., 1999. What makes cells grow larger and how do they do it? Renal hypertrophy
revisited. Exp. Nephrol. 7, 273e283.
Proud, C.G., 2007. Amino acids and mTOR signalling in anabolic function. Biochem. Soc.
Trans. 35, 1187e1190.
Radimerski, T., Mini, T., Schneider, U., Wettenhall, R.E., Thomas, G., Jeno, P., 2000.
Identication of insulin-induced sites of ribosomal protein S6 phosphorylation in
Drosophila melanogaster. Biochemistry 39, 5766e5774.
Ravelli, A.C., van der Meulen, J.H., Michels, R.P., Osmond, C., Barker, D.J., Hales, C.N.,
Bleker, O.P., 1998. Glucose tolerance in adults after prenatal exposure to famine. Lancet
351, 173e177.
Richardson, C.J., Broenstrup, M., Fingar, D.C., Julich, K., Ballif, B.A., Gygi, S., Blenis, J.,
2004. SKAR is a specic target of S6 kinase 1 in cell growth control. Curr. Biol. 14,
1540e1549.
Robb, V.A., Astrinidis, A., Henske, E.P., 2006. Frequent hyperphosphorylation of ribosomal
protein S6 in lymphangioleiomyomatosis-associated angiomyolipomas. Mod. Pathol. 19,
839e846.
Robb, V.A., Karbowniczek, M., Klein-Szanto, A.J., Henske, E.P., 2007. Activation of the
mTOR signaling pathway in renal clear cell carcinoma. J. Urol. 177, 346e352.
Romeo, Y., Zhang, X., Roux, P.P., 2012. Regulation and function of the RSK family of
protein kinases. Biochem. J. 441, 553e569.
Rosner, M., Fuchs, C., Dolznig, H., Hengstschlager, M., 2011. Different cytoplasmic/
nuclear distribution of S6 protein phosphorylated at S240/244 and S235/236. Amino
acids 40, 595e600.
Rossi, R., Pester, J.M., McDowell, M., Soza, S., Montecucco, A., Lee-Fruman, K.K., 2007.
Identication of S6K2 as a centrosome-located kinase. FEBS Lett. 581, 4058e4064.
Roux, P.P., Ballif, B.A., Anjum, R., Gygi, S.P., Blenis, J., 2004. Tumor-promoting phorbol
esters and activated Ras inactivate the tuberous sclerosis tumor suppressor complex via
p90 ribosomal S6 kinase. Proc. Natl. Acad. Sci. U.S.A. 101, 13489e13494.
70 Oded Meyuhas
Roux, P.P., Shahbazian, D., Vu, H., Holz, M.K., Cohen, M.S., Taunton, J., Sonenberg, N.,
Blenis, J., 2007. RAS/ERK signaling promotes site-specic ribosomal protein S6 phos-
phorylation via RSK and stimulates cap-dependent translation. J. Biol. Chem. 282,
14056e14064.
Ruf, M.T., Andreoli, A., Itin, P., Pluschke, G., Schmid, P., 2014. Ribosomal protein S6 is
hyperactivated and differentially phosphorylated in epidermal lesions of patients with
psoriasis and atopic dermatitis. Br. J. Dermatol. 171, 1533e1536.
Ruvinsky, I., Katz, M., Dreazen, A., Gielchinsky, Y., Saada, A., Freedman, N., Mishani, E.,
Zimmerman, G., Kasir, J., Meyuhas, O., 2009. Mice decient in ribosomal protein S6
phosphorylation suffer from muscle weakness that reects a growth defect and energy
decit. PLoS One 4, e5618.
Ruvinsky, I., Sharon, N., Lerer, T., Cohen, H., Stolovich-Rain, M., Nir, T., Dor, Y.,
Zisman, P., Meyuhas, O., 2005. Ribosomal protein S6 phosphorylation is a determinant
of cell size and glucose homeostasis. Genes Dev. 19, 2199e2211.
Sancak, Y., Bar-Peled, L., Zoncu, R., Markhard, A.L., Nada, S., Sabatini, D.M., 2010.
Ragulator-Rag complex targets mTORC1 to the lysosomal surface and is necessary
for its activation by amino acids. Cell 141, 290e303.
Sancak, Y., Peterson, T.R., Shaul, Y.D., Lindquist, R.A., Thoreen, C.C., Bar-Peled, L.,
Sabatini, D.M., 2008. The Rag GTPases bind raptor and mediate amino acid signaling
to mTORC1. Science 320, 1496e1501.
Sancak, Y., Thoreen, C.C., Peterson, T.R., Lindquist, R.A., Kang, S.A., Spooner, E.,
Carr, S.A., Sabatini, D.M., 2007. PRAS40 is an insulin-regulated inhibitor of the
mTORC1 protein kinase. Mol. Cell 25, 903e915.
Schmidt, C., Lipsius, E., Kruppa, J., 1995. Nuclear and nucleolar targeting of human ribo-
somal protein S6. Mol. Biol. Cell. 6, 1875e1885.
Schumacher, A.M., Velentza, A.V., Watterson, D.M., Dresios, J., 2006. Death-associated
protein kinase phosphorylates mammalian ribosomal protein S6 and reduces protein
synthesis. Biochemistry 45, 13614e13621.
Shaw, R.J., Bardeesy, N., Manning, B.D., Lopez, L., Kosmatka, M., DePinho, R.A.,
Cantley, L.C., 2004. The LKB1 tumor suppressor negatively regulates mTOR
signaling. Cancer Cell 6, 91e99.
Shima, H., Pende, M., Chen, Y., Fumagalli, S., Thomas, G., Kozma, S.C., 1998. Disruption
of the p70
s6k
/p85
s6k
gene reveals a small mouse phenotype and a new functional S6
kinase. EMBO J. 17, 6649e6659.
Shoshani, T., Faerman, A., Mett, I., Zelin, E., Tenne, T., Gorodin, S., Moshel, Y., Elbaz, S.,
Budanov, A., Chajut, A., Kalinski, H., Kamer, I., Rozen, A., Mor, O., Keshet, E.,
Leshkowitz, D., Einat, P., Skaliter, R., Feinstein, E., 2002. Identication of a novel hyp-
oxia-inducible factor 1-responsive gene, RTP801, involved in apoptosis. Mol. Cell.
Biol. 22, 2283e2293.
Silva, E., Au-Yeung, H.W., Van Goethem, E., Burden, J., Franc, N.C., 2007. Requirement
for a Drosophila E3-ubiquitin ligase in phagocytosis of apoptotic cells. Immunity 27,
585e596.
Smith, E.M., Finn, S.G., Tee, A.R., Browne, G.J., Proud, C.G., 2005. The tuberous sclerosis
protein TSC2 is not required for the regulation of the mammalian target of rapamycin by
amino acids and certain cellular stresses. J. Biol. Chem. 280, 18717e18727.
Sofer, A., Lei, K., Johannessen, C.M., Ellisen, L.W., 2005. Regulation of mTOR and cell
growth in response to energy stress by REDD1. Mol. Cell. Biol. 25, 5834e5845.
Stewart, M.J., Denell, R., 1993. Mutations in Drosophila gene encoding ribosomal protein S6
cause tissue overgrowth. Mol. Cell. Biol. 13, 2524e2535.
Sulic, S., Panic, L., Barkic, M., Mercep, M., Uzelac, M., Volarevic, S., 2005. Inactivation of
S6 ribosomal protein gene in T lymphocytes activates a p53-dependent checkpoint
response. Genes Dev. 19, 3070e3082.
Ribosomal Protein S6 Phosphorylation 71
Szyszka, R., Gasior, E., 1984. Phosphorylation of ribosomal proteins during differentiation of
Saccharomyces cerevisiae. Acta Biochim. Pol. 31, 375e382.
Tang, H., Hornstein, E., Stolovich, M., Levy, G., Livingstone, M., Templeton, D.J.,
Avruch, J., Meyuhas, O., 2001. Amino acid-induced translation of TOP mRNAs is fully
dependent on PI3-kinase-mediated signaling, is partially inhibited by rapamycin, and is
independent of S6K1 and rpS6 phosphorylation. Mol. Cell. Biol. 21, 8671e8683.
Towler, M.C., Hardie, D.G., 2007. AMP-activated protein kinase in metabolic control and
insulin signaling. Circ. Res. 100, 328e341.
Turck, F., Kozma, S.C., Thomas, G., Nagy, F., 1998. A heat-sensitive Arabidopsis thaliana
kinase substitutes for human p70s6k function in vivo. Mol. Cell. Biol. 18, 2038e2044.
Urban, J., Soulard, A., Huber, A., Lippman, S., Mukhopadhyay, D., Deloche, O.,
Wanke, V., Anrather, D., Ammerer, G., Riezman, H., Broach, J.R., De Virgilio, C.,
Hall, M.N., Loewith, R., 2007. Sch9 is a major target of TORC1 in Saccharomyces
cerevisiae. Mol. Cell 26, 663e674.
Valjent, E., Bertran-Gonzalez, J., Bowling, H., Lopez, S., Santini, E., Matamales, M.,
Bonito-Oliva, A., Herve, D., Hoeffer, C., Klann, E., Girault, J.A., Fisone, G., 2011.
Haloperidol regulates the state of phosphorylation of ribosomal protein S6 via
activation of PKA and phosphorylation of DARPP-32. Neuropsychopharmacology
36, 2561e2570.
Vander Haar, E., Lee, S.I., Bandhakavi, S., Grifn, T.J., Kim, D.H., 2007. Insulin signalling
to mTOR mediated by the Akt/PKB substrate PRAS40. Nat. Cell Biol. 9, 316e323.
Vincent, A., Herman, J., Schulick, R., Hruban, R.H., Goggins, M., 2011. Pancreatic cancer.
Lancet 378, 607e620.
Volarevic, S., Stewart, M.J., Ledermann, B., Zilberman, F., Terracciano, L., Montini, E.,
Grompe, M., Kozma, S.C., Thomas, G., 2000. Proliferation, but not growth, blocked
by conditional deletion of 40S ribosomal protein S6. Science 288, 2045e2047.
Wang, X., Campbell, L.E., Miller, C.M., Proud, C.G., 1998. Amino acid availability
regulates p70 S6 kinase and multiple translation factors. Biochem. J. 334, 261e267.
Warner, J.R., McIntosh, K.B., 2009. How common are extraribosomal functions of
ribosomal proteins? Mol. Cell 34, 3e11.
Watson, K.L., Chou, M.M., Blenis, J., Gelbart, W.M., Erikson, R.L., 1996. A Drosophila
gene structurally and functionally homologous to the mammalian 70-kDa s6 kinase
gene. Proc. Natl. Acad. Sci. U.S.A. 93, 13694e13698.
Watson, K.L., Konrad, K.D., Woods, D.E., Bryant, P.J., 1992. Drosophila homolog of the
human S6 ribosomal protein is required for tumor suppression in the hematopoietic
system. Proc. Natl. Acad. Sci. U.S.A. 89, 11302e11306.
Wettenhall, R.E., Erikson, E., Maller, J.L., 1992. Ordered multisite phosphorylation of Xen-
opus ribosomal protein S6 by S6 kinase II. J. Biol. Chem. 267, 9021e9027.
Williams, A.J., Werner-Fraczek, J., Chang, I.F., Bailey-Serres, J., 2003. Regulated phosphor-
ylation of 40S ribosomal protein S6 in root tips of maize. Plant Physiol. 132, 2086e2097.
Wool, I., Chan, Y.-L., Gluck, A., 1996. Mammalian ribosome: the structure and the evolu-
tion of the proteins. In: Hershey, J.W.B., Mathews, M.B., Sonenberg, N. (Eds.), Trans-
lational Control. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY,
pp. 685e718.
Xiao, H., Wang, H., Silva, E.A., Thompson, J., Guillou, A., Yates, J.R.J.I., Buchon, N.,
Franc, N.C., 2015. The pallbearer E3 ligase promotes actin remodeling via RAC in
efferocytosis by degrading the ribosomal protein S6. Dev. Cell 32, 19e30.
Xu, J., Chen, J., Dong, Z., Meyuhas, O., Chen, J.K., 2015. Phosphorylation of ribosomal
protein S6 mediates compensatory renal hypertrophy. Kidney Int. 87, 543e556.
Xu, K., Liu, P., Wei, W., 2014. mTOR signaling in tumorigenesis. Biochim. Biophys. Acta
1846, 638e654.
Yang, Q., Guan, K.L., 2007. Expanding mTOR signaling. Cell Res. 17, 666e681.
72 Oded Meyuhas
Zemp, I., Kutay, U., 2007. Nuclear export and cytoplasmic maturation of ribosomal
subunits. FEBS Lett. 581, 2783e2793.
Zhang, S.H., Lawton, M.A., Hunter, T., Lamb, C.J., 1994. atpk1, a novel ribosomal protein
kinase gene from Arabidopsis. I. Isolation, characterization, and expression. J. Biol. Chem.
269, 17586e17592.
Zoncu, R., Bar-Peled, L., Efeyan, A., Wang, S., Sancak, Y., Sabatini, D.M., 2011.
mTORC1 senses lysosomal amino acids through an inside-out mechanism that requires
the vacuolar H-ATPase. Science 334, 678e683.
Ribosomal Protein S6 Phosphorylation 73
... Subsequent reports have demonstrated that mTORC1 promotes BTB disruption via a surge of p-rpS6-S235/S236 and p-rpS6-S240/ S244 in the rat testis. Furthermore, the mTORC1 signaling pathway modulates BTB function through rpS6 by regulating actin organization (Mok et al. 2012Meyuhas 2015;Meyuhas and Dreazen 2009). Sarbassov et al. found that the mTORC2/Rictor complex regulates p-PKCα and the actin-based cytoskeleton (Sarbassov et al. 2004). ...
... mTORC1 is composed of mTOR, Raptor, PRAS40, mLST8, and Deptor. Numerous reports have focused on its roles in cell growth and proliferation manifested by regulating protein synthesis (Mok et al. 2013a;Magnuson et al. 2012;Meyuhas and Dreazen 2009;Morita et al. 2015). In contrast, mTORC2 consists of mTOR, Rictor, mLST8, deptor, Sin1 and Protor1/2. ...
Article
Full-text available
Mammalian target of rapamycin (mTOR) is a crucial signaling protein regulating a range of cellular events. Numerous studies have reported that the mTOR pathway is related to spermatogenesis in mammals. However, its functions and underlying mechanisms in crustaceans remain largely unknown. mTOR exists as two multimeric functional complexes termed mTOR complex 1 (mTORC1) and mTORC2. Herein, we first cloned ribosomal protein S6 (rpS6, a downstream molecule of mTORC1) and protein kinase C (PKC, a downstream effector of mTORC2) from the testis of Eriocheir sinensis. The dynamic localization of rpS6 and PKC suggested that both proteins may be essential for spermatogenesis. rpS6/PKC knockdown and Torin1 treatment led to defects in spermatogenesis, including germ cell loss, retention of mature sperm and empty lumen formation. In addition, the integrity of the testis barrier (similar to the blood-testis barrier in mammals) was disrupted in the rpS6/PKC knockdown and Torin1 treatment groups, accompanied by changing in expression and distribution of junction proteins. Further study demonstrated that these findings may result from the disorganization of filamentous actin (F-actin) networks, which were mediated by the expression of actin-related protein 3 (Arp3) rather than epidermal growth factor receptor pathway substrate 8 (Eps8). In summary, our study illustrated that mTORC1/rpS6 and mTORC2/PKC regulated spermatogenesis via Arp3-mediated actin microfilament organization in E. sinensis.
... Research in this field critically depends on the accurate quantification of TORC1 activities in various genetic settings and under defined physiological conditions. In mammalian cells, TORC1 activity is typically assessed via the phosphorylation levels of direct TORC1 target residues in the eukaryotic initiation factor 4E-binding protein 1 (4E-BP1; (Ma and Blenis, 2009)) and the ribosomal protein S6 (rpS6) kinase 1 (S6K1) (Fenton and Gout, 2011;Magnuson et al., 2012), or the ones in rpS6, the effector of the latter kinase (Meyuhas and Dreazen, 2009). The most commonly used proxies for TORC1 activities in the budding yeast Saccharomyces cerevisiae include, similar to S6K phosphorylation in mammalian cells, phosphorylation of the bona fide TORC1 residue Thr 737 in the protein kinase Sch9 Urban et al., 2007) and, analogous to mammalian rpS6, phosphorylation of Ser 232,233 in Rps6, which is carried out by the Sch9-related TORC1 effector kinase Ypk3 (González et al., 2015;Yerlikaya et al., 2016). ...
Article
Full-text available
The eukaryotic TORC1 kinase integrates and links nutritional, energy, and hormonal signals to cell growth and homeostasis, and its deregulation is associated with human diseases including neurodegeneration, cancer, and metabolic syndrome. Quantification of TORC1 activities in various genetic settings and defined physiological conditions generally relies on the assessment of the phosphorylation level of residues in TORC1 targets. Here we show that two commonly used TORC1 effectors in yeast, namely Sch9 and Rps6, exhibit distinct phosphorylation patterns in response to rapamycin treatment or changes in nitrogen availability, indicating that the choice of TORC1 proxies introduces a bias in decoding TORC1 activity.
... This pathway may contribute to cell size and enhanced protein synthesis. [61][62][63] In female mice, we observed enhanced P-S6 in PTEN KO L5 neurons in comparison to neighboring WT neurons ( Figures S5A and S5B). ERα heterozygosity reduced P-S6 in PTEN KO neurons. ...
Article
Full-text available
Little is known of the brain mechanisms that mediate sex-specific autism symptoms. Here, we demonstrate that deletion of the autism spectrum disorder (ASD)-risk gene, Pten, in neocortical pyramidal neurons (NSEPten knockout [KO]) results in robust cortical circuit hyperexcitability selectively in female mice observed as prolonged spontaneous persistent activity states. Circuit hyperexcitability in females is mediated by metabotropic glutamate receptor 5 (mGluR5) and estrogen receptor α (ERα) signaling to mitogen-activated protein kinases (Erk1/2) and de novo protein synthesis. Pten KO layer 5 neurons have a female-specific increase in mGluR5 and mGluR5-dependent protein synthesis. Furthermore, mGluR5-ERα complexes are generally elevated in female cortices, and genetic reduction of ERα rescues enhanced circuit excitability, protein synthesis, and neuron size selectively in NSEPten KO females. Female NSEPten KO mice display deficits in sensory processing and social behaviors as well as mGluR5-dependent seizures. These results reveal mechanisms by which sex and a high-confidence ASD-risk gene interact to affect brain function and behavior.
... Phosphorylation of RPS6 and eIF4B is proposed to stimulate mRNA translation, although it should be noted that molecular mechanisms underlying such stimulation are not well understood (discussed in [195]). What is known is that inactivation of mTORC1 negatively regulates activities of S6Ks and is believed to suppress protein synthesis (reviewed in [163]). The levels of PDCD4 are highly sensitive to mTORC1/S6K activity in mitosis [169]. ...
Article
Cellular stress is an intrinsic part of cell physiology that underlines cell survival or death. The ability of mammalian cells to regulate global protein synthesis (aka translational control) represents a critical, yet underappreciated, layer of regulation during the stress response. Various cellular stress response pathways monitor conditions of cell growth and subsequently reshape the cellular translatome to optimize translational outputs. On the molecular level, such translational reprogramming involves an intricate network of interactions between translation machinery, RNA-binding proteins, mRNAs, and non-protein coding RNAs. In this review, we will discuss molecular mechanisms, signaling pathways, and targets of translational control that contribute to cellular adaptation to stress and to cell survival or death.
... 35 As an indirect downstream effector of mTOR, rpS6 not only participates in glucose homeostasis but is also essential for regulating the size of at least some cell types. 36 In addition, it was recently shown that specific small inhibitory RNA-mediated knockdown of RPS6 disrupts 40S ribosomal biogenesis, showing that RPS6 is a critical regulator of 5′ ...
Article
Full-text available
Our previous studies illustrated that 2% H2 inhalation can protect against sepsis-associated encephalopathy (SAE) which is characterized by high mortality and has no effective treatment. To investigate the underlying role of protein phosphorylation in SAE and H2 treatment, a mouse model of sepsis was constructed by caecal ligation and puncture (CLP), then treated with H2 (CLP + H2 ). Brain tissues of the mice were collected to be analysed with tandem mass tag-based quantitative proteomics coupled with IMAC enrichment of phosphopeptides and LC-MS/MS analysis. In proteomics and phosphoproteomics analysis, 268 differentially phosphorylated proteins (DPPs) showed a change in the phosphorylated form in the CLP + H2 group (p < 0.05). Gene ontology analysis revealed that these DPPs were enriched in multiple cellular components, biological processes, and molecular functions. KEGG pathway analysis revealed that they were enriched in glutamatergic synapses, tight junctions, the PI3K-Akt signalling pathway, the HIF-1 signalling pathway, the cGMP-PKG signalling pathway, the Rap1 signalling pathway, and the vascular smooth muscle contraction. The phosphorylated forms of six DPPs, including ribosomal protein S6 (Rps6), tyrosine 3-monooxygenase/tryptophan 5-monooxygenase activation protein gamma (Ywhag/14-3-3), phosphatase and tensin homologue deleted on chromosome ten (Pten), membrane-associated guanylate kinase 1 (Magi1), mTOR, and protein kinase N2 (Pkn2), were upregulated and participated in the PI3K-Akt signalling pathway. The WB results showed that the phosphorylation levels of Rps6, Ywhag, Pten, Magi1, mTOR, and Pkn2 were increased. The DPPs and phosphorylation-mediated molecular network alterations in H2 -treated CLP mice may elucidate the biological roles of protein phosphorylation in the therapeutic mechanism of H2 treatment against SAE.
Article
A number of neurotransmitters have been detected in tumor microenvironment and proved to modulate cancer oncogenesis and progression. We previously found that biosynthesis and secretion of neurotransmitter 5-hydroxytryptamine (5-HT) was elevated in colorectal cancer cells. In this study, we discovered that the HTR2B receptor of 5-HT was highly expressed in colorectal cancer tumor tissues, which was further identified as a strong risk factor for colorectal cancer prognostic outcomes. Both pharmacological blocking and genetic knocking down HTR2B impaired migration of colorectal cancer cell, as well as the epithelial–mesenchymal transition (EMT) process. Mechanistically, HTR2B signaling induced ribosomal protein S6 kinase B1 (S6K1) activation via the Akt/mTOR pathway, which triggered cAMP-responsive element-binding protein 1 (CREB1) phosphorylation (Ser 133) and translocation into the nucleus, then the phosphorylated CREB1 acts as an activator for ZEB1 transcription after binding to CREB1 half-site (GTCA) at ZEB1 promoter. As a key regulator of EMT, ZEB1, therefore, enhances migration and EMT process in colorectal cancer cells. We also found that HTR2B-specific antagonist (RS127445) treatment significantly ameliorated metastasis and reversed EMT process in both HCT116 cell tail-vein–injected pulmonary metastasis and CT26 cell intrasplenic-injected hepatic metastasis mouse models. Implications These findings uncover a novel regulatory role of HTR2B signaling on colorectal cancer metastasis, which provide experimental evidences for potential HTR2B-targeted anti-colorectal cancer metastasis therapy.
Article
Full-text available
Translation is critical for development as transcription in the oocyte and early embryo is silenced. To illustrate the translational changes during meiosis and consecutive two mitoses of the oocyte and early embryo, we performed a genome-wide translatome analysis. Acquired data showed significant and uniform activation of key translational initiation and elongation axes specific to M-phases. Although global protein synthesis decreases in M-phases, translation initiation and elongation activity increases in a uniformly fluctuating manner, leading to qualitative changes in translation regulation via the mTOR1/4F/eEF2 axis. Overall, we have uncovered a highly dynamic and oscillatory pattern of translational reprogramming that contributes to the translational regulation of specific mRNAs with different modes of polysomal occupancy/translation that are important for oocyte and embryo developmental competence. Our results provide new insights into the regulation of gene expression during oocyte meiosis as well as the first two embryonic mitoses and show how temporal translation can be optimized. This study is the first step towards a comprehensive analysis of the molecular mechanisms that not only control translation during early development, but also regulate translation-related networks employed in the oocyte-to-embryo transition and embryonic genome activation.
Article
Daphnia pulex is a natural food widely used in aquaculture, and S6K plays an important role in protein translation and ageing. To explore the function of S6K in the ageing process of D. pulex , we cloned and analysed the full-length cDNA of S6K. With the growth of D. pulex , the expression of S6K mRNA first increased, then peaked at 10 days, and then decreased. Western blot analysis showed that the S6K protein expression increased from 1 to 5 days of age, but did not change significantly from 5 to 25 days of age. RNAi knockdown of S6K, reduced the expression of S6K mRNA and S6K protein, and it causes changes in other genes associated with ageing in both sexes. S6K may be involved in reproductive transformation, affecting reproductive capacity, thus significantly affecting the ageing of D. pulex . miRNA with potential regulation of the S6K gene were screened by the bioinformatics method and validated using a dual luciferase reporter gene system. This study deepens our understanding of the mechanisms of ageing in D. pulex and expands our understanding of the role of S6K in the ageing process.
Article
It is well known that sublethal dose of insecticides induces life history trait changes of both target and non-target insect species, however, the underlying mechanisms remain not well understood. In this study, the effects of low concentrations of the anthranilic diamide insecticide chlorantraniliprole on the development and reproduction of the fall armyworm (FAW), Spodoptera frugiperda, were evaluated, and the underlying mechanisms were explored. The results showed that exposure of FAW to LC10 and LC30 chlorantraniliprole prolonged the larvae duration, decreased the mean weight of the larvae and pupae, and lowered the pupation rate as well as emergence rate. The fecundity of female adults was also negatively affected by treatment with low concentrations of chlorantraniliprole. Consistently, we found that exposure of FAW to LC30 chlorantraniliprole downregulated the mRNA expression of juvenile hormone (JH) esterase (SfJHE), leading to the increase of JH titer in larvae. We also found that treatment with low concentrations of chlorantraniliprole suppressed the expression of ribosomal protein S6 kinase1 (SfS6K1) in female adults, resulting in the downregulation of the gene encoding vitellogenin (SfVg). These results provided insights into the mechanisms underlying the effects of low concentrations of insecticides on insect pests, and had applied implications for the control of FAW.
Article
Full-text available
Background: Non-coding RNAs (ncRNAs), including small nucleolar RNAs (snoRNAs), are widely involved in the physiological and pathological processes of human beings. While up to date, although considerable progress has been achieved in ncRNA-related pathogenesis of non-small cell lung cancer (NSCLC), the underlying mechanisms and biological significance of snoRNAs in NSCLC still need to be further clarified. Methods: Quantitative real-time polymerase chain reaction or RNAscope was performed to verify the expression of Small Nucleolar RNA, H/ACA Box 38B (SNORA38B) in NSCLC cell lines or clinical samples. BALB/c nude mice xenograft model or C57BL/6J mice syngeneic tumor model were estimated to detect the effects of SNORA38B in tumor growth or tumor immune microenvironment in vivo. Cytometry by time of flight, enzyme-linked immunosorbent assay and flow cytometry assay were conducted to clarify the effects and mechanisms of SNORA38B-mediated tumor immunosuppressive microenvironment. The binding activity between SNORA38B and E2F transcription factor 1(E2F1) was detected by RNA immunoprecipitation and RNA pull-down assays. Then, bioinformatics analysis and chromatin immunoprecipitation were utilized to demonstrate the regulation of GRB2-associated-binding protein 2 (GAB2) by E2F1. Moreover, the combinatorial treatment of SNORA38B locked nucleic acid (LNA) and immune checkpoint blockade (ICB) was used to treat murine Lewis lung carcinoma-derived tumor burden C57BL/6J mice to clarify the effectiveness of targeting SNORA38B in NSCLC immunotherapy. Results: SNORA38B was found highly expressed in NSCLC tissues and cell lines, and associated with worse prognosis. Further results showed that SNORA38B functioned as an oncogene via facilitating cell proliferation, migration, invasion, and inhibiting cell apoptosis in vitro and promoting tumorigenesis of NSCLC cells in vivo. SNORA38B could also recruit the CD4+FOXP3+ regulatory T cells by triggering tumor cells to secrete interleukin 10, which in turn reduced the infiltration of CD3+CD8+ T cells in NSCLC tumor microenvironment (TME), favoring tumor progression and poorer immune efficacy. Mechanistically, SNORA38B mainly distributed in the nucleus, and promoted NSCLC progression by regulating GAB2 transcription to activate protein kinase B (AKT)/mammalian target of rapamycin (mTOR) pathway through directly binding with E2F1. Moreover, we found that SNORA38B LNAs were able to ameliorate CD3+CD8+ T cell infiltration in TME, which sensitized NSCLC to the treatment of ICB. Conclusions: In conclusion, our data demonstrated that SNORA38B functioned as an oncogene in NSCLC both in vitro and in vivo at least in part by regulating the GAB2/AKT/mTOR pathway via directly binding to E2F1. SNORA38B could also sensitize NSCLC to immunotherapy, which may be a critical therapeutic target for NSCLC.
Article
Full-text available
Incubation of Chinese hamster ovary cells without amino acids for up to 60 min caused a rapid marked decrease in p70 S6 kinase activity and increased binding of initiation factor eIF4E to its inhibitory regulator protein 4E-BP1. This was associated with dephosphorylation of 4E-BP1 and eIF4E and dissociation of eIF4E from eIF4G. All these effects were rapidly reversed by resupplying a mixture of amino acids and this was blocked by rapamycin and by inhibitors of phosphatidylinositol 3-kinase, implying a role for phosphatidylinositol 3-kinase in the signalling pathway linking amino acids with the control of p70 S6 kinase activity and the phosphorylation of these translation factors. Amino acid withdrawal also led to changes in the phosphorylation of other translation factors; phosphorylation of eIF4E decreased whereas elongation factor eEF2 became more heavily phosphorylated, each of these changes being associated with decreased activity of the factor in question. Earlier studies have suggested that protein kinase B (PKB) may act upstream of p70 S6 kinase. However, amino acids did not affect the activity of PKB, indicating that amino acids activate p70 S6 kinase through a pathway independent of this enzyme. Studies with individual amino acids suggested that the effects on p70 S6 kinase activity and translation-factor phosphorylation were independent of cell swelling. The data show that amino acid supply regulates multiple translation factors in mammalian cells.
Article
Full-text available
Mammalian target of rapamycin (mTOR) is a central regulator of protein synthesis whose activity is modulated by a variety of signals. Energy depletion and hypoxia result in mTOR inhibition. While energy depletion inhibits mTOR through a process involving the activation of AMP-activated protein kinase (AMPK) by LKB1 and subsequent phosphorylation of TSC2, the mechanism of mTOR inhibition by hypoxia is not known. Here we show that mTOR inhibition by hypoxia requires the TSC1/TSC2 tumor suppressor complex and the hypoxia-inducible gene REDD1/RTP801. Disruption of the TSC1/TSC2 complex through loss of TSC1 or TSC2 blocks the effects of hypoxia on mTOR, as measured by changes in the mTOR targets S6K and 4E-BP1, and results in abnormal accumulation of Hypoxia-inducible factor (HIF). In contrast to energy depletion, mTOR inhibition by hypoxia does not require AMPK or LKB1. Down-regulation of mTOR activity by hypoxia requires de novo mRNA synthesis and correlates with increased expression of the hypoxia-inducible REDD1 gene. Disruption of REDD1 abrogates the hypoxia-induced inhibition of mTOR, and REDD1 overexpression is sufficient to down-regulate S6K phosphorylation in a TSC1/TSC2-dependent manner. Inhibition of mTOR function by hypoxia is likely to be important for tumor suppression as TSC2-deficient cells maintain abnormally high levels of cell proliferation under hypoxia.
Article
Full-text available
Ribosomal protein S6 (rpS6), a component of the 40S ribosomal subunit, is phosphorylated on several residues in response to numerous stimuli. Although commonly used as a marker for neuronal activity, its upstream mechanisms of regulation are poorly studied and its role in protein synthesis remains largely debated. Here, we demonstrate that the psychostimulant d-amphetamine (d-amph) markedly increases rpS6 phosphorylation at Ser235/236 sites in both crude and synaptoneurosomal preparations of the mouse striatum. This effect occurs selectively in D1R-expressing medium-sized spiny neurons (MSNs) and requires the cAMP/PKA/DARPP-32/PP-1 cascade, whereas it is independent of mTORC1/p70S6K, PKC, and ERK signaling. By developing a novel assay to label nascent peptidic chains, we show that the rpS6 phosphorylation induced in striatonigral MSNs by d-amph, as well as in striatopallidal MSNs by the antipsychotic haloperidol or in both subtypes by papaverine, is not correlated with the translation of global or 5' terminal oligopyrimidine tract mRNAs. Together, these results provide novel mechanistic insights into the in vivo regulation of the post-translational modification of rpS6 in the striatum and point out the lack of a relationship between PKA-dependent rpS6 phosphorylation and translation efficiency. Copyright © 2015 the authors 0270-6474/15/354113-18$15.00/0.
Article
Full-text available
mTOR (the mechanistic target of rapamycin) is an atypical serine/threonine kinase involved in regulating major cellular functions including growth and proliferation. Deregulations of the mTOR signaling pathway is one of the most commonly observed pathological alterations in human cancers. To this end, oncogenic activation of the mTOR signaling pathway contributes to cancer cell growth, proliferation and survival, highlighting the potential for targeting the oncogenic mTOR pathway members as an effective anti-cancer strategy. In order to do so, a thorough understanding of the physiological roles of key mTOR signaling pathway components and upstream regulators would guide future targeted therapies. Thus, in this review, we summarize available genetic mouse models for mTORC1 and mTORC2 components, as well as characterized mTOR upstream regulators and downstream targets, and assign a potential oncogenic or tumor suppressive role for each evaluated molecule. Together, our work will not only facilitate the current understanding of mTOR biology and possible future research directions, but more importantly, provide a molecular basis for targeted therapies aiming at key oncogenic members along the mTOR signaling pathway. Copyright © 2014. Published by Elsevier B.V.
Article
Full-text available
TOP mRNAs encode components of the translational apparatus, and repression of their translation comprises one mechanism, by which cells encountering amino acid deprivation downregulate the biosynthesis of the protein synthesis machinery. This mode of regulation involves TSC as knockout of TSC1 or TSC2 rescued TOP mRNAs translation in amino acid-starved cells. The involvement of mTOR in translational control of TOP mRNAs is demonstrated by the ability of constitutively active mTOR to relieve the translational repression of TOP mRNA upon amino acid deprivation. Consistently, knockdown of this kinase as well as its inhibition by pharmacological means blocked amino acid-induced translational activation of these mRNAs. The signaling of amino acids to TOP mRNAs involves RagB, as overexpression of active RagB derepressed the translation of these mRNAs in amino acid-starved cells. Nonetheless, knockdown of raptor or rictor failed to suppress translational activation of TOP mRNAs by amino acids, suggesting that mTORC1 or mTORC2 plays a minor, if any, role in this mode of regulation. Finally, miR10a has previously been suggested to positively regulate the translation of TOP mRNAs. However, we show here that titration of this microRNA failed to downregulate the basal translation efficiency of TOP mRNAs. Moreover, Drosha knockdown or Dicer knockout, which carries out the first and second processing steps in microRNAs biosynthesis, respectively, failed to block the translational activation of TOP mRNAs by amino acid or serum stimulation. Evidently, these results are questioning the positive role of microRNAs in this mode of regulation.
Article
Clearance of apoptotic cells (efferocytosis) is achieved through phagocytosis by professional or amateur phagocytes. It is critical for tissue homeostasis and remodeling in all animals. Failure in this process can contribute to the development of inflammatory autoimmune or neurodegenerative diseases. We found previously that the PALL-SCF E3-ubiquitin ligase complex promotes apoptotic cell clearance, but it remained unclear how it did so. Here we show that the F-box protein PALL interacts with phosphorylated ribosomal protein S6 (RpS6) to promote its ubiquitylation and proteasomal degradation. This leads to RAC2 GTPase upregulation and activation and F-actin remodeling that promotes efferocytosis. We further show that the specific role of PALL in efferocytosis is driven by its apoptotic cell-induced nuclear export. Finding a role for RpS6 in the negative regulation of efferocytosis provides the opportunity to develop new strategies to regulate this process. Copyright © 2015 Elsevier Inc. All rights reserved.