14
2014 http://informahealthcare.com/bmg ISSN: 1040-9238 (print), 1549-7798 (electronic) Editor: Michael M. Cox Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177 ! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/10409238.2014.887051 REVIEW ARTICLE Transcriptome-wide studies uncover the diversity of modes of mRNA recruitment to eukaryotic ribosomes Ivan N. Shatsky 1 , Sergey E. Dmitriev 1,2 , Dmitri E. Andreev 1 , and Ilya M. Terenin 1,2 1 Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, Moscow, Russia and 2 Engelhardt Institute of Molecular Biology, Russian Academy of Sciences, Moscow, Russia Abstract The conventional paradigm of translation initiation in eukaryotes states that the cap-binding protein complex eIF4F (consisting of eIF4E, eIF4G and eIF4A) plays a central role in the recruitment of capped mRNAs to ribosomes. However, a growing body of evidence indicates that this paradigm should be revised. This review summarizes the data which have been mostly accumulated in a post-genomic era owing to revolutionary techniques of transcriptome-wide analysis. Unexpectedly, these techniques have uncovered remarkable diversity in the recruitment of cellular mRNAs to eukaryotic ribosomes. These data enable a preliminary classification of mRNAs into several groups based on their requirement for particular components of eIF4F. They challenge the widely accepted concept which relates eIF4E-dependence to the extent of secondary structure in the 5 0 untranslated regions of mRNAs. Moreover, some mRNA species presumably recruit ribosomes to their 5 0 ends without the involvement of either the 5 0 m 7 G-cap or eIF4F but instead utilize eIF4G or eIF4G-like auxiliary factors. The long-standing concept of internal ribosome entry site (IRES)-elements in cellular mRNAs is also discussed. Keywords 5 0 UTR secondary structure, cap-independent translation enhancer, cellular IRESs, DAP5, eIF4F, eIF4G, ribosome profiling, translation initiation factor eIF4E History Received 4 November 2013 Revised 17 January 2014 Accepted 21 January 2014 Published online 13 February 2014 Introduction Translational control mainly operates at the initiation step of polypeptide synthesis. The basis of translation initiation in eukaryotes was elucidated in the 1970 and 80s, starting with the discovery of the mRNA scanning mechanism by Kozak (1989). This important event occurred in a ‘‘pre-genomic era’’ when the set of available model mRNAs was extremely limited and was not representative of the entire diversity of mRNA structures. These ‘‘single mRNA’’ studies resulted in a widely recognized model of cap-dependent translation initiation which is believed to be used by the overwhelming majority of cellular mRNAs in all eukaryotes. Somewhat later (in the 1990s), an alternative mechanism – internal initiation of translation – was discovered (Jang et al., 1988; Pelletier & Sonenberg, 1988) and dissected (Pestova et al., 1996a,b, 1998) for uncapped genomic RNAs from several viruses replicating in the cytoplasm. According to this mechanism, the ribosome recognizes the mRNA at a specific internal ribosome entry site (IRES) within its 5 0 untranslated region (5 0 UTR). Whether this alternative mechanism of ribosome recruitment to mRNA is also used by cellular mRNAs, in particular by mammalian mRNAs, is still a matter of debate (see below). With the beginning of the transcriptome-wide era (a discussion of respective techniques can be found in Kapeli & Yeo, 2012), we now have the ability to analyze large numbers of structurally diverse mRNAs while manip- ulating specific key components of the cell or signaling pathways. Some of the emerging data cannot be explained on the basis of either the standard model of mRNA recruitment to ribosomes or with the help of IRES-elements. These data suggest that something very important has been missed in the past, presumably since a large variety of mRNA structures was not taken into account. Without filling these gaps in our ideas on the molecular mechanisms of recruitment of various mRNAs to ribosomes, we cannot advance our studies of the mechanisms of translational control. The aim of this review is to share our observations while mining the literature of the last decade. We decided to focus on the very first events of mRNA recruitment onto ribosomes (40S subunits) and will not discuss the subsequent steps of translation initiation (such as ribosomal scanning, start codon selection or subunit joining) since these issues have been covered in a number of comprehensive recent reviews (Alekhina & Vassilenko, 2012; Hinnebusch & Lorsch, 2012; Sonenberg & Hinnebusch, 2009;). We will not describe all known cases of translational control, either. Instead we wish to highlight those studies which do not easily align to the Address for correspondence: Dr. Ivan N. Shatsky, PhD, Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, Leninskie gory 1, Khokhlov str., bldg. 40, Moscow 119234, Russia. Tel: +7495 939 4857. E-mail: [email protected] Critical Reviews in Biochemistry and Molecular Biology Downloaded from informahealthcare.com by University of Windsor on 06/05/14 For personal use only.

Transcriptome-wide studies uncover the diversity of modes of mRNA recruitment to eukaryotic ribosomes

  • Upload
    ilya-m

  • View
    214

  • Download
    1

Embed Size (px)

Citation preview

2014

http://informahealthcare.com/bmgISSN: 1040-9238 (print), 1549-7798 (electronic)

Editor: Michael M. CoxCrit Rev Biochem Mol Biol, 2014; 49(2): 164–177

! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/10409238.2014.887051

REVIEW ARTICLE

Transcriptome-wide studies uncover the diversity of modes of mRNArecruitment to eukaryotic ribosomes

Ivan N. Shatsky1, Sergey E. Dmitriev1,2, Dmitri E. Andreev1, and Ilya M. Terenin1,2

1Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, Moscow, Russia and 2Engelhardt Institute of Molecular

Biology, Russian Academy of Sciences, Moscow, Russia

Abstract

The conventional paradigm of translation initiation in eukaryotes states that the cap-bindingprotein complex eIF4F (consisting of eIF4E, eIF4G and eIF4A) plays a central role in therecruitment of capped mRNAs to ribosomes. However, a growing body of evidence indicatesthat this paradigm should be revised. This review summarizes the data which have been mostlyaccumulated in a post-genomic era owing to revolutionary techniques of transcriptome-wideanalysis. Unexpectedly, these techniques have uncovered remarkable diversity in therecruitment of cellular mRNAs to eukaryotic ribosomes. These data enable a preliminaryclassification of mRNAs into several groups based on their requirement for particularcomponents of eIF4F. They challenge the widely accepted concept which relateseIF4E-dependence to the extent of secondary structure in the 50 untranslated regions ofmRNAs. Moreover, some mRNA species presumably recruit ribosomes to their 50 ends withoutthe involvement of either the 50 m7G-cap or eIF4F but instead utilize eIF4G or eIF4G-likeauxiliary factors. The long-standing concept of internal ribosome entry site (IRES)-elements incellular mRNAs is also discussed.

Keywords

50 UTR secondary structure, cap-independenttranslation enhancer, cellular IRESs, DAP5,eIF4F, eIF4G, ribosome profiling,translation initiation factor eIF4E

History

Received 4 November 2013Revised 17 January 2014Accepted 21 January 2014Published online 13 February 2014

Introduction

Translational control mainly operates at the initiation step of

polypeptide synthesis. The basis of translation initiation in

eukaryotes was elucidated in the 1970 and 80s, starting with

the discovery of the mRNA scanning mechanism by Kozak

(1989). This important event occurred in a ‘‘pre-genomic

era’’ when the set of available model mRNAs was extremely

limited and was not representative of the entire diversity of

mRNA structures. These ‘‘single mRNA’’ studies resulted in

a widely recognized model of cap-dependent translation

initiation which is believed to be used by the overwhelming

majority of cellular mRNAs in all eukaryotes. Somewhat later

(in the 1990s), an alternative mechanism – internal initiation

of translation – was discovered (Jang et al., 1988; Pelletier &

Sonenberg, 1988) and dissected (Pestova et al., 1996a,b,

1998) for uncapped genomic RNAs from several viruses

replicating in the cytoplasm. According to this mechanism,

the ribosome recognizes the mRNA at a specific internal

ribosome entry site (IRES) within its 50 untranslated region

(50 UTR). Whether this alternative mechanism of ribosome

recruitment to mRNA is also used by cellular mRNAs, in

particular by mammalian mRNAs, is still a matter of debate

(see below).

With the beginning of the transcriptome-wide era

(a discussion of respective techniques can be found in

Kapeli & Yeo, 2012), we now have the ability to analyze

large numbers of structurally diverse mRNAs while manip-

ulating specific key components of the cell or signaling

pathways. Some of the emerging data cannot be explained on

the basis of either the standard model of mRNA recruitment

to ribosomes or with the help of IRES-elements. These data

suggest that something very important has been missed in the

past, presumably since a large variety of mRNA structures

was not taken into account. Without filling these gaps in our

ideas on the molecular mechanisms of recruitment of various

mRNAs to ribosomes, we cannot advance our studies of the

mechanisms of translational control.

The aim of this review is to share our observations while

mining the literature of the last decade. We decided to focus

on the very first events of mRNA recruitment onto ribosomes

(40S subunits) and will not discuss the subsequent steps of

translation initiation (such as ribosomal scanning, start codon

selection or subunit joining) since these issues have been

covered in a number of comprehensive recent reviews

(Alekhina & Vassilenko, 2012; Hinnebusch & Lorsch, 2012;

Sonenberg & Hinnebusch, 2009;). We will not describe all

known cases of translational control, either. Instead we wish

to highlight those studies which do not easily align to the

Address for correspondence: Dr. Ivan N. Shatsky, PhD, BelozerskyInstitute of Physico-Chemical Biology, Lomonosov Moscow StateUniversity, Leninskie gory 1, Khokhlov str., bldg. 40, Moscow119234, Russia. Tel: +7495 939 4857. E-mail: [email protected]

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

current paradigm of primary mRNA recognition by

ribosomes.

The conventional mechanism of mRNA recruitmentto ribosomes in eukaryotes and factors involved inthis process

The classical mechanism (Sonenberg & Hinnebusch, 2009)

states that translation initiation starts with the binding of

factor eIF4E to the cap of mRNA. In the mammalian cell,

eIF4E1 (a prevalent version of eIF4E) may be: (1) in a free

state, (2) in a complex with its repressor, eIF4E-Binding

Protein (4E-BP), or (3) part of an active heterotrimer, termed

as eIF4F (eIF4E�eIF4G�eIF4A). eIF4G plays the role of a

scaffold in the trimer eIF4F. This large multifunctional

protein (M.W. �175 kDa) interacts with several ligands

involved in the accommodation of mRNA on the 40S

ribosomal subunit (Figure 1A). For mammalian eIF4G,

these include interactions with: eIF4E (which holds the

mRNA by the cap), eIF3 (which bridges the complex to the

40S subunit), eIF4A (an ATP-dependent RNA helicase), and

PABP (Poly(A) Binding Protein) bound to the mRNA

poly(A)-tail. In addition, eIF4G can also bind the mRNA

itself, contacting not only its 50 UTR but in some cases the 30

UTR (Park et al., 2011a), by means of distinct RNA-binding

sites. Yeast eIF4G has at least three sites that interact with

mRNA (Berset et al., 2003; Park et al., 2011a), while

mammalian eIF4G has two known RNA-binding sites (Prevot

et al., 2003). No information is available about the nucleotide

sequence preferences of eIF4G for RNA binding.

All these interactions result in the mRNA pseudo-

circularization which is thought to reinforce interactions that

help keeping eIF4F on the mRNA. Indeed, the interaction of

eIF4G with PABP increases the affinity of eIF4E for the cap

(Borman et al., 2000). However, our knowledge about its

Figure 1. eIF4G and eIF4G-like protein families. (A) Some characterized representatives of eIF4G family. The identified structural domains aredesignated inside the boxes whereas the binding sites for initiation factors, PABP and Mnk1 and 2 kinases are shown above the diagrams. Thehorizontal arrows denote alternative translation starts in eIF4G1 of Homo sapiens. The horizontal brackets delimit either sequences missing in someisoforms or C-terminal truncated ones. The vertical long arrows under the boxes denote cleavage sites for picornavirus proteases 2A (2A-pro) and L(L-pro), caspase 3 (Casp3) and HIV-1/2 proteases. The sites of interaction with RNA are indicated by arrows above the diagrams. (B) MIF4G domaincontaining proteins that have been shown to participate in the mRNA recruitment to ribosomes. The binding sites for particular initiation factors areshown with the same color as in (A). The designations for conserved domains are: MIF4 for ‘‘Middle domain of eIF4G’’ (pfam02854); MA3 for‘‘Domain in DAP-5, eIF4G, MA-3 and other proteins’’ (pfam02847); W2 for ‘‘C-terminal domain of eIF4-gamma/eIF5/eIF2b-epsilon’’ (pfam02020);KH for ‘‘hnRNP K homology RNA-binding domain, type I’’ (pfam00013).

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 165

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

contribution to the translation of individual natural mRNAs,

i.e. possessing distinct 50 and 30 UTRs, is still very limited

(Kopeina et al., 2008; Park et al., 2011b).

The activity of helicase eIF4A is important for both

accommodation of the 43S preinitiation complex (43S PIC) at

the 50 end of mRNA and subsequent mRNA scanning assisted

by eIF4B (Parsyan et al., 2011). eIF4B is indispensable for

eIF4A functioning during scanning of the 50 UTRs that have

even minor base-pairings within their sequences (Dmitriev

et al., 2003). It is possible that the primary accommodation of

the scanning apparatus on some 50 UTRs that have highly

stable stems at their 50 termini is also promoted by additional

DEAD-box helicases, e.g. DDX3 (Ded1p in yeast) or

DDX9/RHA (for review, see Marintchev, 2013; Soto-Rifo

et al., 2012). Unfolding of especially stable stem-loops during

the scanning may also require DHX29 (Pisareva et al., 2008).

For some mRNAs, eIF4B may be substituted by its homolog

eIF4H (for review, see Parsyan et al., 2011).

The chain of interactions eIF4E–eIF4G–eIF3–40S is one

of the principal targets for translational regulation. Although a

direct interaction of eIF4G with eIF3 has not been found in

yeast, it may be mediated by some other translation compo-

nents, e.g. by eIF5 or eIF4B. Yeast also demonstrate a less

pronounced dependence on eIF4G for mRNA recruitment

in vitro, with eIF3 playing the major role (Jivotovskaya et al.,

2006; Mitchell et al., 2010). However, the importance

of eIF4G for increasing the rate and extent of mRNA

recruitment in yeast is nevertheless in no doubt (Mitchell

et al., 2010).

Regulation of eIF4F integrity by eIF4E-bindingproteins and mTOR signaling

It should be mentioned that though eIF4E and eIF4G form a

complex, their individual concentrations in proliferating cells

are quite different. eIF4E1 is a very abundant protein, and

even when its concentration in mammalian cells was

decreased 10-fold by RNA interference, no significant

reduction in global protein synthesis was observed

(Yanagiya et al., 2012). This high molar excess of eIF4E is

partially neutralized in cells by multiple specific repressors

4E-BPs (see below). Therefore, it is the ratio of eIF4E1 to

active (hypophosphorylated) 4E-BPs what should be con-

sidered (‘‘active’’ eIF4E1 concentration) rather than the

absolute concentration of the factor (Alain et al., 2012).

In higher eukaryotes, the global translational regulation of

eIF4E is thought to be based on 4E-BPs 1, 2 and 3. These

small proteins harbor the eIF4E-binding motif YXXXXL�(where X is any amino acid and � is a hydrophobic residue)

that interacts with the same dorsal surface of eIF4E respon-

sible for its interaction with eIF4G (Mader et al., 1995).

In a hypophosphorylated (active) state, 4E-BPs compete with

eIF4G for eIF4E binding to sequester eIF4E from the eIF4F

complex (for review, see Roux & Topisirovic, 2012).

On receiving a proliferative signal, these eIF4E repressors

are multiply phosphorylated by the complex C1 of mam-

malian Target Of Rapamycin kinase (mTORC1) and revers-

ibly inactivated. As a consequence, eIF4E is liberated,

associates with eIF4G-eIF4A (thereby forming eIF4F),

and directs the mRNA to the 40S ribosomal subunit.

mTORC1 also activates the ribosomal protein S6 kinases

p70S6K1/2, which subsequently phosphorylate eIF4B and

Programmed Cell Death 4 (PDCD4), an inhibitor of eIF4A

that is targeted to proteolysis after the phosphorylation

(for review, see Roux & Topisirovic, 2012). Although

direct effects of eIF4B modifications are open to question

(reviewed by Shagam et al., 2012), both eIF4B and PDCD4

phosphorylation are necessary in vivo for the manifestation of

mTOR activation effects (Dennis et al., 2012). mTOR can

directly bind eIF3 and positively affect the eIF4G–eIF3

interaction (Harris et al., 2006). Its indirect targets also

include eIF4G, translation elongation factor eEF2 and ribo-

somal protein S6. mTOR also controls the transcription of

rRNAs and tRNAs (Laplante & Sabatini, 2012). Such a

multiplicity of complex events caused by activation of this

huge protein kinase may diminish to some extent a direct role

of 4E-BPs in response to proliferative signals. It is important

to note that overexpression of a completely non-phosphor-

ylatable mutant of 4E-BP1 in cells inhibited translation of

a reporter mRNA by only �60% and overexpression of

wt 4E-BP1 inhibited it by only 20% (Mothe-Satney et al.,

2000). Similarly, modest effects of recombinant 4E-BP1

addition were also observed in vitro (Andreev et al., 2009;

Haghighat et al., 1995).

mTOR is a downstream effector in the PI3K/AKT pathway

that has been linked to cancer, obesity, diabetes, neuro-

degeneration and aging (Laplante & Sabatini, 2012).

A number of other signaling cascades that translate external

signals into the cell converge on this pathway, thus affecting

mTOR activity: these include special GTPases that sense

amino acid availability, receptors for growth factors

(e.g. insulin, EGF and interferons) or various molecules

responsible for energy supply. A detailed description of

mTOR complexes Q1 and Q2 and their functions, as well as

implication of other signaling pathways in translation regu-

lation can be found in (Laplante & Sabatini, 2012; Roux &

Topisirovic, 2012; Wang & Proud, 2011).

All eukaryotes express proteins that can repress the

eIF4E–eIF4G interaction thus regulating large number of

mRNAs. Examples in yeast are Caf20p and Eap1p with MWs

of 18 and 70 kDa, respectively (Cridge et al., 2010). Except

for the eIF4E interaction motif, they are quite distinct from

4E-BPs of higher eukaryotes. They interact with eIF4E while

being associated with 30 UTRs of the mRNAs they regulate.

It has been shown that Caf20p and Eap1p together control

translation of �1000 mRNAs and each repressor regulates

its own class of mRNAs. Apart from general regulators of

eIF4E activity, eukaryotic cells also contain eIF4E-binding

proteins that regulate translation of subsets of mRNAs.

Among them, it is worth mentioning Cup in Drosophila,

Maskin in oocytes of Xenopus laevis and Neuroguidin

in neurons (for references see reviews by Darnell &

Richter, 2012; Lasko, 2012; Richter, 2007). However, this

mode of regulation requires the formation of complexes

of eIF4E-binding proteins with additional protein factors

which in turn recognize specific RNA-motives within 30

UTRs of respective mRNAs. This leads to pseudo-

circularization of mRNA, and therefore resembles the afore-

mentioned mechanism of action proposed for yeast Caf20p

and Eap1p.

166 I. N. Shatsky et al. Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

Large variations in eIF4E-dependence for variousindividual mRNAs

The current literature, as a rule, proposes only two modes of

mRNA recruitment in eukaryotes – the classical mechanism

just described and the cap-independent IRES-driven mode of

initiation (Figure 2A and B). Until recently, no other

mechanisms of cellular mRNAs binding with 40S ribosomal

subunits were known. For the classical mechanism, the

interaction of the cap with eIF4E (mostly its eIF4E1 isoform

in mammals) is regarded as an obligatory first step in

recruitment of mRNAs to 40S ribosomes for the overwhelm-

ing majority of cellular mRNAs. Nevertheless, there is a

general agreement that the eIF4E dependence varies among

individual mRNAs, as some mRNAs still continue to bind

ribosomes under stress conditions when eIF4E1 is inactivated

by 4E-BPs and the scaffold factor eIF4G is cleaved by virus

proteases or caspases (Figure 1A). Such stress-resistant

cellular mRNAs are believed to harbor IRES-elements

within their 50 UTRs (for review, see Ruggero, 2013).

Conversely, another subset of mRNAs is believed to have

an increased requirement for eIF4E. Until recently this

property has been ascribed to those mRNAs that have long

and highly structured 50 UTRs (Gingras et al., 1999) and this

opinion still predominates (for references, see Jia et al., 2012).

Originally it came from two well-known reports where

mRNAs containing artificial perfect stem-loops of different

length in their 50 UTR were shown to require more eIF4E and

eIF4A for efficient translation (Koromilas et al., 1992; Svitkin

et al., 2001). Indeed, it is logical to assume that the structured

50 UTRs are more dependent on eIF4E since they may need

more helicase eIF4A for unwinding their stem-loop structures

and eIF4A is delivered to the 50 terminus through the

eIF4E�eIF4G complex. The recent unexpected finding that

eIF4E binding to eIF4G not only brings the complex to the 50

cap, but also stimulates in vitro the eIF4A-helicase activity in

a cap-independent manner, which has been presented as a fact

in favor of this hypothesis (Feoktistova et al., 2013). This is an

important point since highly structured 50 UTRs are charac-

teristic of many mRNAs encoding oncogenic growth and

transcription factors or regulatory protein kinases and these

considerations have been employed to explain why many

cancer cells (�30%) have elevated levels of eIF4E. The same

arguments have been used to explain why some specific

mRNAs with highly structured 50 UTRs are activated after the

knock-out of 4E-BPs in mice (Colina et al., 2008; Gkogkas

et al., 2013). Nevertheless, this concept should be revised: as

we shall see below, recent transcriptome-wide data do not

support it.

The initial testing of the correlation between the eIF4E

dependence and the length and degree of secondary structure

in natural 50 leaders was performed in 2009 (Andreev et al.,

2009). The analysis was carried out both in a cell-free system

and in RNA-transfected cells. It should be emphasized that for

in vitro assays a cytoplasmic extract from cultured mamma-

lian cells was used since the conventional rabbit reticulocyte

lysate was shown to be inadequate for translation initiation

studies of mRNAs possessing complex 50 leaders (Dmitriev

et al., 2009). Similarly, the RNA transfection technique was

chosen to avoid artifacts associated with DNA-encoded

reporters (Dmitriev et al., 2007; see also below). This study

showed that the 50 cap stimulated translation of mRNAs with

short and less structured 50 UTRs much more than some

mRNAs with long and highly structured 50 leaders. These

results were confirmed further by in vitro experiments

with 4E-BP1. Among the tested mRNAs, the lowest cap-

dependence, and therefore eIF4E-dependence, was the 50

UTR of Apaf-1 mRNA which possesses four highly structured

domains (though no IRES activity was detected within its

sequence) (Andreev et al., 2009).

Transcriptome-wide identification of mRNAs withincreased requirement for eIF4E

What classes of natural mRNAs are particularly sensitive to

the ‘‘active’’ eIF4E concentration? Or do 4E-BPs regulate

global protein synthesis? Of earlier reports on this point one

should mention the paper by (Mamane et al., 2007) where the

authors carried out a microarray analysis of polysomal mRNA

from an eIF4E-inducible NIH 3T3 cell line. Among 294

transcripts that shifted upon eIF4E induction to the heavier

polysome fractions, were mRNAs coding for components of

the translational machinery (ribosomal proteins, translation

factors, etc.) These mRNAs are known to possess relatively

short and low structured 50 UTRs.

The decisive contribution to clarifying this issue has been

done by recent transcriptome-wide studies by Thoreen et al.

(2012) and Hsieh et al. (2012). These authors treated mouse

p53 �/� embryonic fibroblasts (MEF) or human p53 �/�prostate cancer cells PC3, respectively, with inhibitors of the

ATP binding site of mTOR, thorin 1 and PP242. Both drugs

inhibit the catalytic subunit of mTOR and activate 4E-BP’s

(see above) much more efficiently than the well-known drug

rapamycin that acts allosterically. Changes in the translational

activity of individual mRNAs were assessed by ribosome

profiling, a novel powerful method of genome wide analysis

of mRNA activities in living cells (Ingolia et al., 2012).

A positive aspect in the design of both studies was a

short-time interval between the drug addition and the

moment when the cells were harvested, lysed and used in

ribosome profiling. In other words, the authors analyzed an

immediate response. This is an obvious advantage over the

experiments with 4E-BP knockout murine cells (see above)

where essential adaptive changes during development of

knockout murine stem cells (MSC) into embryos cannot be

excluded.

Surprisingly, the inhibition of eIF4E activity resulted in a

strong suppression of translation (from 4-to-8-fold) of only

120–150 mRNAs. The authors confirmed that this inhibition

was largely due to dephosphorylated 4E-BPs. The affected

mRNAs mostly coded for ribosomal proteins, translation

factors and proteins indispensable for growth rather than

oncogenic kinases, receptors or transcription factors. In fact,

this result is not as surprising as it seems at first glance: one of

the principal hallmarks of cancer cells is their enhanced

ability to grow and proliferate and mTOR is the one of the

main downstream regulators that governs these features. One

should bear in mind that growth and proliferation first

necessitate the synthesis of more ribosomes and translation

factors since the translational capacity of any cell is limited

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 167

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

Fig

ure

2.

Div

erse

mo

del

so

fre

cru

itm

ent

of

euk

aryo

tic

40

Sri

bo

som

esto

mR

NA

s.(A

)T

he

can

on

ical

eIF

4E

-dep

end

ent

mec

han

ism

of

mR

NA

bin

din

gby

40

Sri

boso

mal

sub

un

its.

(B)

Th

ein

tern

alen

try

of

40

Sri

bo

som

alsu

bu

nit

so

nto

50 U

TR

so

fm

RN

As.

Her

e,th

eIR

ES

-dri

ven

mo

de

isex

emp

lifi

edby

the

mec

han

ism

use

dby

pic

orn

avir

us

IRE

S-e

lem

ents

.It

requ

ires

ast

able

bin

din

go

fa

key

mR

NA

recr

uit

ing

fact

or

(e.g

.eI

F4

G)

toa

spec

ific

stru

ctu

ral

elem

ent

wit

hin

the

IRE

S.

An

oth

erp

uta

tive

elem

ent(

s)o

fth

eIR

ES

may

pro

mo

teth

eac

com

mo

dat

ion

of

mR

NA

inte

rnal

seg

men

tin

the

mR

NA

bin

din

gch

ann

elo

f4

0S

rib

oso

me(

som

eIR

ES

sfr

om

oth

erv

iru

sfa

mil

ies

do

no

tre

qu

ire

scan

nin

g).

(C)

Th

e50 C

ITE

-dri

ven

mR

NA

recr

uit

men

t.T

his

mec

han

ism

also

requ

ires

ast

ruct

ura

lel

emen

tw

ith

anin

crea

sed

affi

nit

yto

akey

mR

NA

recr

uit

men

tfa

cto

r.H

ow

ever

,th

isel

emen

tis

no

tca

pab

leo

fp

laci

ng

the

adja

cen

tse

qu

ence

of

mR

NA

into

the

mR

NA

bin

din

gch

ann

el.

Th

eref

ore

,th

ere

cruit

men

tca

nb

ein

itia

ted

excl

usi

vel

yat

the

mR

NA

50

end

.(D

)T

he

30

CIT

E-d

irec

ted

mR

NA

recr

uit

men

tem

plo

yed

by

som

ep

lan

tv

iru

sR

NA

s(s

eeth

ete

xt)

.

168 I. N. Shatsky et al. Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

and determined to a large extent by translation apparatus

availability. Conversely, mTOR inhibition immediately stops

the formation and function of protein synthesizing factories as

the most resource and energy consuming process. That is why

mTOR activity is linked to all or almost all other signal

pathways sensing environmental changes. In the case of

prostate cancer PC3, a strong decrease of translation was also

noted for four mRNAs that determined mesenchymal charac-

teristics of PC3 cells and their capability of metastasizing

(vimentin, YB-1, MTA1, CD44) (Hsieh et al., 2012).

Remarkably, the 50 UTRs of the majority of these highly

eIF4E-dependent mRNAs did not demonstrate an increased

length and/or GC-content. Just the opposite, they revealed

some slight bias to a shorter length and a lower base-pairing.

Their common distinctive feature was the presence of a 50

terminal oligopyrimidine tract (50 TOP). This feature is

known to determine the ease with which mRNAs are recruited

to polysomes after transferring starved cells to rich medium.

Strikingly, the authors found no more than 30% of the total

number (about 5000) of examined individual mRNAs which

were sensitive to some extent to mTOR inhibition. The

translation of other mRNAs either did not change at all or

were even stimulated, presumably due to less competition for

ribosomes from the very abundant mRNAs encoding compo-

nents of the translational apparatus. Among the stimulated

mRNAs there were some that mediate cell survival after drug

treatment (XIAP, Bcl2), proliferation (c-Myc) or apoptosis

(Apaf-1). All these transcripts have long and highly structured

50 UTRs but do not pass through stringent criteria for IRES-

containing mRNAs (reviewed by Shatsky et al., 2010, see also

below). Thus, translation of the majority of mRNAs in the

investigated cells depended very little on the concentration of

‘‘free’’ eIF4E1, and is unlikely to be directly controlled by

mTOR kinase. Moreover, the activity of some of them,

e.g. the stress-induced translation of Hsp70 mRNA, has been

reported to be negatively regulated by PI3K–mTORC1

signaling (Sun et al., 2011).

As the function of eIF4E in the mRNA recruitment into

48S complex may be realized exclusively via its interaction

with eIF4G, it is logical to assume that the translation of

mRNAs coding for components of translational apparatus

should be highly dependent on eIF4G, as well. Indeed,

polysome analysis of mRNAs after eIF4G knockdown showed

(Thoreen et al., 2012) that the translation of mRNAs highly

sensitive to mTOR inhibition (mainly 50 TOP and 50 TOP-like

mRNAs) is also sensitive to eIF4GI depletion. Thus, the

translation of mRNAs directly controlled by 4E-BPs depends

both on eIF4E (mostly eIF4E1) and eIF4G1, and, presumably,

on eIF4G3, the minor eIF4G variant with a domain organ-

ization identical to that of eIF4G1.

The mRNAs with a relaxed requirement for eIF4E butwith a strong dependence on eIF4G

It is well known that the binding of eIF4G to mRNA

dramatically enhances the binding of eIF4E to the 50 cap

(Haghighat & Sonenberg, 1997). In fact, some natural

mRNAs even possess special RNA structures within their 50

UTRs that facilitate recruitment of eIF4E�eIF4G complexes

(Wallace et al., 2010). It is easy to assume, therefore, that

different RNA sequences at the 50 termini of mRNAs would

have different affinity to eIF4G and thus could modulate the

requirement of the mRNA for eIF4E. Indeed, there are many

mRNAs whose translation is only marginally affected by

eIF4E depletion but strongly requires eIF4G. Many breast

cancer cells manifest an enhanced expression of eIF4GI and

concentration of this factor is critical for their survival after

irradiation therapy (Badura et al., 2012). eIF4G1 depletion

results in a significant drop in the survival of these tumor cells

(unlike knockdown of the less abundant eIF4G3) and in

growth arrest. Using microarray profiling of polysomal

mRNAs, a set of mRNAs was identified that were particularly

dependent on eIF4GI levels. They coded for survivin, XIAP,

HIF1a, GADD45a, p53, ATM, Chk1 and many other protein

factors involved in the DNA damage response (Badura et al.,

2012). eIF4E knockdown does not cause a strong effect on the

expression of these proteins (a maximum of 2-fold).

Conversely, eIF4G1 overexpression strongly improved sur-

vival, promoted assembly of the DNA Damage Response

complex (DDR-complex), suppressed apoptosis and overcame

autophagy. As shown in the same paper (Badura et al., 2012),

these properties were not accounted for by the presence of

IRES-elements within the corresponding 50 UTRs.

These data demonstrate that some mRNAs can do without

eIF4E, that is, they retain substantial translational activity

when it is deficient. A cellular mRNA that is independent of

the 50 cap and eIF4E, at least under stress conditions, but

nevertheless uses a 50 end-dependent mode of translation

initiation has been identified (Andreev et al., 2012). The

occurrence of mRNAs with such properties was additionally

supported by creating an artificial 50 UTR possessing such

properties (Terenin et al., 2013).

Other members of eIF4E and eIF4G families and theirpossible functions

The existence of homologues of both eIF4E and eIF4G

additionally expands the number of essential modifications of

the conventional mechanism of mRNA recruitment to 40S

ribosomes. Eukaryotic organisms have multiple eIF4E homo-

logues but the function of most of them is unknown (Joshi

et al., 2004, 2005; Rhoads, 2009). As a rule, one of them is

the most abundant and ubiquitous in every organism.

In mammalian cells this form is represented by eIF4E1.

As described above, it is mostly responsible for growth and

proliferation. It may be used by many, if not all mRNAs but

with different impacts on translation activity. Other forms are

presumably implicated in some specialized functions and

expressed at specific steps of development and differentiation.

For instance, Drosophila has eight eIF4E isoforms. One of

them, eIF4E2 (d4EHP), specifically interacts with bicoid

(Bcd) to suppress the caudal (cad) translation during devel-

opment. This translational suppression occurs via the Bcd

binding region (BBR) located in the cad 30 UTR.

Simultaneous interactions of eIF4E2 with the cap structure

and of Bcd with BBR renders cad mRNA translationally

inactive (Cho et al., 2005). Drosophila eIF4E3 (not to be

confused with mammalian eIF4E3) has been recently shown

to be essential for spermatogenesis (Hernandez & Vazquez-

Pianzola, 2005; Hernandez et al., 2012; Lasko, 2012).

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 169

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

Mammalian cells have three major forms of eIF4E:

eIF4E1, eIF4E2 (4EHP) and eIF4E3 (Joshi et al., 2004,

2005; Rhoads, 2009). Like in Drosophila, mammalian eIF4E2

does not interact with eIF4G and therefore is regarded as a

competitor for eIF4E1. Alone, it possesses a lower affinity for

the cap than eIF4E1 but we may not rule out that its

complexes with specific protein partners may modify this

affinity: a good example of this kind is the enhancement of

the cap-eIF4E interaction by eIF4G (Haghighat & Sonenberg,

1997; von Der Haar et al., 2000). Mammalian eIF4E2

interacts with the GIGYF2 protein and the zinc finger protein

598. This complex directs the inhibitory action of eIF4E2 to

specific mRNA targets to repress translation of a subset of

mRNAs during embryonic development (Morita et al., 2012).

The mechanism of repression is presumably similar to that

just described for eIF4E2 from Drosophila.

eIF4E3 forms a complex with eIF4G but does not bind

4E-BPs (Joshi et al., 2004). Its overexpression inhibits the

eIF4E1-cap interaction and therefore suppresses cellular

growth. The structural organization of its cap-binding surface

is strikingly distinct from that for eIF4E1 and 2 (Joshi et al.,

2004; Osborne et al., 2013). eIF4E3 is expressed only in some

tissues suggesting a role in translation of some specific

individual mRNAs rather than regulation of eIF4E1 – cap

recognition. It was found to be frequently deleted in oral

squamous cell carcinoma (Cha et al., 2011). We should note

again that the positive or inhibitory action of a particular

eIF4E homologue may be determined by protein partners with

which it interacts. For instance, eIF4E2 has recently been

shown to promote, rather than to inhibit, the translation

initiation of VEGF and other proteins under hypoxia (Uniacke

et al., 2012). To do this, eIF4E2 binds within the 30 UTRs of

corresponding mRNAs with assistance of auxiliary proteins,

one of which recognizing a specific nucleotide sequence. To

the best of our knowledge, this is thus far the only report on a

positive function of eIF4E2 in translation.

eIF4G also has several homologues (Figure 1A).

In different organisms, from yeast to mammals, there are

two homologous forms of this factor encoded by separate

genes, e.g. eIF4G1 (TIF4631) and eIF4G2 (TIF4632) in yeast

and eIF4G1 (eIF4GI) and eIF4G3 (eIF4GII) in vertebrates.

In addition, each has several variants arising from alternative

splicing, use of alternative promoters or variations in the

position of transcription starts (Byrd et al., 2005; Coldwell &

Morley, 2006; Coldwell et al., 2012). Interestingly,

Caenorhabditis elegans has a single gene for IFG-1, an

eIF4G homologue: the diversity of isoforms is maintained by

an alternative splicing and up to five isoforms have been

suggested to be synthesized. Importantly, the second major

isoform lacks a putative eIF4E-binding site and thus cannot be

purified via cap-binding chromatography (Contreras et al.,

2008). This raises the possibility that such truncated IFG-1

represents an ancient form of unusual eIF4G variant DAP5 in

mammalian cells (or eIF4G2, see below).

In general, various variants of eIF4G are able to function-

ally replace each other, at least partially. They are apparently

completely interchangeable in yeast (Clarkson et al., 2010).

In mammalian cells, the longest form of eIF4G3 is also

capable of substituting for eIF4G1 (Coldwell et al., 2012).

However, it seems very likely that specific combinations of

eIF4G and eIF4E forms do exist in higher eukaryotes.

Whether properties of eIF4F depend on which variants of

eIF4E and eIF4G form a complex is an intriguing question.

A Drosophila eIF4G variant, eIF4G-2, plays a critical role in

spermatocyte meiosis (Baker & Fuller, 2007; Franklin-

Dumont et al., 2007). It has been hypothesized that eIF4G-2

forms with eIF4E-3, a specific variant of eIF4F operating in

spermatocytes during this process (Hernandez et al., 2012).

Finally, the specialization of eIF4G isoforms can be achieved

through their specific phosphorylation (see Dobrikov et al.,

2014; Srivastava et al., 2012; and references therein).

Phosphorylation of eIF4E (Ser 209 in human eIF4E) by

MNK1/2 kinases in metazoans is one more way to diversify

the function of eIF4F. MNKs are recruited to eIF4E through

an interaction with the C-terminal part of eIF4G (Figure 1A).

It is poorly understood how this phosphorylation modifies

properties of eIF4E. On the one hand, the oncogenic potential

of unphosphorylatable mutant of eIF4E dramatically drops.

On the other hand, the phosphorylation was shown to reduce

its affinity to the cap. In addition to variable effects on general

translation, the phosphorylation of eIF4E was shown to

selectively affect translation of a subset of mRNAs, i.e. it may

serve as a translational switch between distinct classes of

mRNAs (for references, see Roux & Topisirovic, 2012).

eIF4F-independent ways of mRNA-binding toeukaryotic ribosomes

The existence of eIF4F-independent mode(s) of mRNA

recruitment to ribosomes has been strongly supported by an

identification of mRNAs that continue to be translated in

poliovirus infected cells (Macejak & Sarnow, 1991). The

poliovirus infection is known to produce the viral protease 2A

that rapidly cleaves eIF4G1 between amino acid 681 and 682

(and with a much slower kinetics also eIF4G3) (Figure 1A).

The cleavage disconnects the eIF4E binding site of eIF4G

from those for eIF4A and eIF3, thereby abolishing the

recruitment of mRNA to the 40S subunit by a cap-dependent

way. It has been thought for a long time that the continued

translation of some mRNAs in poliovirus-infected cells is

indicative of an IRES-driven mode (Figure 2B) of translation

initiation. More recent data strongly suggest that 50 UTRs of

some mammalian mRNAs can do without the cap (and

therefore without eIF4E) even if they have no IRES (Andreev

et al., 2009, 2012; Terenin et al., 2013).

Mammalian eIF4G2 (other names: Death Associated

Protein 5 (DAP5), Novel APOBEC Target1 (NAT1) or p97;

not to be confused with eIF4GII, or eIF4G3) deserves special

attention (Figure 1A). This mysterious version of eIF4G was

discovered simultaneously in 1997 by four separate labora-

tories even though each group worked in quite different fields

(Imataka et al., 1997; Levy-Strumpf et al., 1997; Shaughnessy

et al., 1997; Yamanaka et al., 1997). eIF4G2 (MW 97 kDa)

appears to be an N-terminal truncated form of eIF4G which

lacks binding sites for eIF4E and PABP. At least in higher

eukaryotes its coding sequence invariably starts with a non-

AUG codon (always GUG in vertebrates) (Takahashi et al.,

2005). eIF4G2 demonstrates a non-continuous distribution

among Metazoans, being present in all Chordata and some

invertebrates (insects, sea urchins, mollusks and some types

170 I. N. Shatsky et al. Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

of worms – but unexpectedly absent from C. elegans where it

may be replaced by a truncated variant of eIF4G, see above).

Such a complex distribution suggests a role for eIF4G2 in

some higher functions associated with multicellularity. Since

it cannot associate with eIF4E, eIF4G2 has long been

regarded as a negative regulator of cap-dependent translation

which is transformed into a positive regulator after cleavage

with Caspase-3 into p86 under apoptotic conditions (Marash

& Kimchi, 2005). It is believed to serve some cellular mRNAs

that harbor IRES-elements (Figure 2B) operating under stress

conditions. However, this hypothesis still requires experi-

mental support (see next section about cellular IRESs).

Surprisingly, apart from apoptosis (Marash & Kimchi, 2005;

Ozpolat et al., 2008), eIF4G2 has been implicated in the

opposite responses, e.g. in cell survival during mitosis and

differentiation (Liberman et al., 2009; Marash et al., 2008).

Null mutation of eIF4G2 in mice leads to an embryonic

lethality since the factor is essential for gastrulation. The

expression of retinoic acid-responsive genes, such as the cell-

cycle inhibitor p21 (WAF1), is also selectively impaired in

eIF4G2 (�/�) cells (Yamanaka et al., 2000). Knockdown of

eIF4G2 expression by siRNAs suppresses all-trans-retinoic

acid-induced granulocytic differentiation (Ozpolat et al.,

2008).

Overall, the data obtained in the first 10 years after the

discovery of eIF4G2 gave the impression that this factor is

implicated in translation of a rather small subset of mRNAs

involved in highly specialized functions. However, later

reports, including transcriptome-wide studies, suggest that

eIF4G2 governs the translation of a larger set of mRNAs and

may be required by cells growing under normal conditions

(Lee & McCormick, 2006; Nousch et al., 2007; Ramirez-

Valle et al., 2008). eIF4G2 was found to be associated with

the polysome fraction in actively proliferative cells (Nousch

et al., 2007). Remarkably, depletion on eIF4G1 alone by

siRNAs resulted in only 20–30% inhibition of total protein

synthesis, whereas simultaneous depletion of both eIF4G1

and eIF4G2 leads to 60% inhibition and cessation of cell

division. Moreover, overexpression of eIF4G2 does not

compensate the eIF4G1 depletion indicating that these two

factors direct translation of distinct sets of mRNAs (Ramirez-

Valle et al., 2008). Like all other members of the eIF4G

family, eIF4G2 contains the MIF4G domain that binds

helicase eIF4A albeit with a lower affinity than the respective

domain from eIF4G1 (Virgili et al., 2013). The mechanism of

functioning and partners of eIF4G2 are not known. It is

ubiquitous and relatively abundant in both cultured cells and

normal tissues (Geiger et al., 2013; Schwanhausser et al.,

2011) supporting the view that this factor is a translational

activator of many mRNAs even under normal conditions.

The eIF4G2 is not the only unusual homologue of eIF4G1/

3 with the MIF4G domain. In recent years, a number of other

MIF4G domain – containing proteins have been further

investigated (Figure 1B). They seem to escort eukaryotic

PolII-transcripts through all the steps of RNA metabolism:

starting from transcription and splicing (CBP80, NOM1/

SGD1, CWC22), through nuclear export and mRNA quality

control (all the above plus GLE1 and UPF2) to translation in

the cytoplasm (CBP80, GLE1, PAIP1, CTIF and SLIP1) and

to mRNA decay (NOT1). However, not all of them may be

regarded as factors involved in mRNA recruitment to

ribosomes, i.e. the process discussed in this review. Other

MIF4G proteins may be implicated in later steps of translation

initiation or its regulation (see, for example, Bolger & Wente,

2011). Complex of CBP80 and nuclear cap-binding protein

CBC20 is thought to be involved in the pioneer round of

translation of mRNAs (Maquat et al., 2010) exported out of

the nucleus. The mRNA recruitment to ribosomes with this

factor is assisted by CTIF (CBP80/20-dependent Translation

Initiation Factor) and eIF3g (Choe et al., 2012; Kim et al.,

2009). However, it is quite possible that CBP80/20 complex is

also involved in translation of a specific set of mRNAs, e.g.

the mRNAs activated under stress conditions (Sharma et al.,

2012). Its yeast functional analog cbc1 plays an active role in

translation: it is necessary for rapid reprograming of trans-

lation after hyper-osmotic shock (Garre et al., 2012). Paip1

stabilizes the interaction between eIF4G and PABP by

binding to eIF3, which in turn binds to eIF4G and the

40S ribosome (Martineau et al., 2008). The factor SLIP1

is needed for translation initiation on some replication-

dependent histon mRNAs which are not polyadenylated and

instead have a specific stem-loop structure at the 30 end

(Cakmakci et al., 2008; Neusiedler et al., 2012; von Moeller

et al., 2013).

Hernandez et al. have recently reported the discovery of

Mextli (Mxt), a novel eIF4E-binding protein in Drosophila

that promotes translation (Hernandez et al., 2013). It contains

(Figure 1B) a MIF4G domain, an hnRNP K homology

RNA-binding domain and a consensus eIF4E-binding site.

Strikingly, its eIF4E binding site is located at the C-terminus

of the factor. Mxt is expressed at high levels in ovarian

germline stem cells and early-stage cystocytes and forms a

complex with both eIF4E1 and eIF4E3. Thus, Mxt is a novel

type of scaffolding protein that binds RNA, eIF3 and eIF4Es.

Interestingly, Mxt interacts with a DEAH-box helicase

CG3225 (ortholog of mammalian DHX35) instead of

‘‘canonical’’ eIF4A. Although Mxt was not found in mam-

malian cells, its discovery is a very important event since it

hints at the existence of other non-canonical eIF4G analogs in

eukaryotic cells.

An even more intriguing question concerns the possible

participation of ‘‘alternative’’ cap-binding complexes in

translation initiation under certain conditions. For example,

a proteomic study of m7G-associated proteins from quiescent

plant cells revealed an unexpected set of polypeptides where

eIF4A was completely absent but several other DEAD-box

proteins (plant-specific or homologous to mammalian

DDX3X, DDX3Y or DDX16) appeared, together with nuclear

cap-binding protein CBP20 (Bush et al., 2009). In another

study, translation induction of a specific mRNA set encoding

enzymes of the pentose phosphate pathway was detected after

glucose depletion conditions that induced eIF4A loss from

eIF4F (Castelli et al., 2011).

The possibility of eIF4F- and eIF4G-independent modes of

translation initiation in eukaryotic cells is indirectly supported

by experiments with eIF4G depletion. As mentioned above,

its knock-down in mammalian cells resulted in a surprisingly

small effect. It strongly affected only 20% of cellular mRNAs

(Ramirez-Valle et al., 2008). Although the eIF4G depletion in

yeast cells was eventually lethal, it had a small effect on the

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 171

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

translational potential of most mRNAs (Park et al., 2011b).

Depleting eIF4G just narrowed the range of translational

efficiencies transcriptome-wide: the yeast mRNAs with better

than average efficiency were translated relatively worse,

whereas the mRNAs with lower than average efficiency were

translated relatively better. Importantly, the mRNAs most

dependent on eIF4G showed an average 50 UTR length at or

below the mean for all yeast genes. The authors even

concluded that eIF4G (eIF4GI + eIF4GII) is not essential for

translation of most mRNAs but enhances the differentiation of

translational efficiencies transcriptome-wide (Park et al.,

2011b).

The concept of cellular IRES-elements twodecades later

Many researchers, in particular those who are not specialized

in translation, believe that the cap-independent mode of

translation initiation is synonymous to IRES-driven initiation.

The data discussed in the previous section of this review

strongly contradict this view. Nevertheless, the concept of

cellular IRESs as a principal cap-independent mechanism of

translation initiation in eukaryotes is still dominant since it

explains the deviations from the standard mechanism in a

plausible, simple and hence a very attractive way.

Virus IRES-elements are no longer a scientific hypothesis.

Their existence is a firmly established fact. Some IRESs from

viruses with a positive RNA genome are characterized in

detail (for recent review see, Jackson, 2013) though several

important gaps still remain. Based on their structures, factor

requirements and modes of function, the virus IRESs are even

amenable to some classification. In contrast, in spite of the

22-years long history (Macejak & Sarnow, 1991) of cellular

IRESs, we do not know yet what they look like, how they

work and even whether they exist. The existence of none of

them has been proven with the use of a set of rigorous control

experiments. Most of them have been proposed on the basis of

cell transfection experiments with dicistronic DNA construc-

tions, i.e. by an approach extremely prone to unavoidable

artifacts (reviewed by Shatsky et al., 2010). Recent work by

Lemp et al. (2012) who detected cryptic transcripts from an

ubiquitous plasmid origin of replication used in an over-

whelming majority of reporter plasmids has finally high-

lighted the severe short-comings of using DNA-based

constructs for IRES analysis. A set of strict criteria that

should be met and a set of control experiments that should be

performed before researchers confirm the existence of a

cellular IRES are presented by Andreev et al. (2009) and

Jackson (2013). One of the most important criterion, the ratio

of absolute translation efficiencies of a 50 UTR in the capped

monocystronic context versus dicistronic position, is routinely

absent from published reports. As a result, readers cannot

judge whether the authors deal with a physiologically relevant

IRES or with a background activity.

Another widely held opinion, which is used to support the

concept of cellular IRESs, is that the scanning 50 end-

dependent mechanism cannot operate in cells with an

inactivated eIF4E or cleaved eIF4G1 can also be challenged

(Andreev et al., 2012; and see the text above). Finally, an

mRNA with a low cap-dependence that still requires the 50

end has been recently demonstrated by our lab to be highly

translationally active. To this end, we created a model mRNA

that had such properties: its translation depended very little on

the cap both in vitro and in transfected cells and occurred via

a 50 end-dependent scanning mechanism (Terenin et al.,

2013). Additionally, this report indicated that the presence

within the 50 UTR of a high affinity binding site for a key

translation initiation component (factor) was necessary but

not sufficient for internal initiation. Indeed, the well-

characterized virus IRESs appear to harbor, along with such

an affinity site, additional specific domains that promote

placing the initiation region of the mRNA into the

mRNA-binding channel of the ribosome. The best studied

example is domain II of HCV IRES (Filbin & Kieft, 2011;

Malygin et al., 2013). Otherwise, if such an accommodation

‘‘device’’ is lacking, the only point at which the process of

mRNA ‘‘threading’’ into the 40S subunit channel can be

initiated is the very 50 end of the mRNA (Figure 2). The

requirement for a free 50 terminus is not determined

exclusively by the cap and eIF4E. Even more it is dictated

by a great preference that the scanning machinery has for an

unpaired mRNA end to start unwinding stem-loops within 50

UTRs of mRNAs (Rajagopal et al., 2012 and references cited

there in). The only exclusion from this general rule may be

A- or U-rich nucleotide tracts: their lower base-pairing

potential may enable them to spontaneously occupy the

mRNA binding groove of the 40S ribosome without assist-

ance of the RNA-unwinding machinery (Gilbert et al., 2007;

Terenin et al., 2005).

Finally, compelling evidence that the initiation from a free

50 end is only partially determined by recognition of the

cap by eIF4E is offered by RNAs from some plant viruses.

The mRNAs of viruses from the Potyviridae, Comoviridae,

Tombusviridae and Luteoviridae families and the

Sobemovirus and Umbravirus genera lack the m7GpppN-

cap structure. These viruses employ special elements termed

Cap Independent Translation Enhancers (CITE) (Figures 2C

and D) (reviewed by Miller et al., 2007). The majority of

CITEs are located in the 30 UTRs (30 CITEs). They are

thought to recruit components of the translational apparatus

(e.g. eIF4E or eIF4G) and then deliver them to the 50 end of

mRNA through long distance base pairing between 50 and 30

UTRs (Nicholson et al., 2010; Stupina et al., 2008; Treder

et al., 2008). The CITE’s can also function when placed

within 50 UTRs of mRNAs (Guo et al., 2000) (Figure 2C).

The initiation on these plant virus RNAs starts at the 50 end

and employs a scanning mechanism. We have suggested that a

similar mechanism may also be used by mRNAs from animal

cells (Shatsky et al., 2010) and demonstrated its feasibility

(Terenin et al., 2013).

Thus, there may be two quite different cap-independent

ways of mRNA recruitment onto the 40S ribosomal subunit in

mammalian cells, an IRES-driven mode and a 50 CITE-based

mechanism (Figure 2B and C). They are easy to distinguish:

the latter mode should manifest a very low translation activity

when the corresponding nucleotide sequence is placed

into the intercistronic position of a dicistronic mRNA in

comparison with that in a natural capped monocistronic

context. For a true IRES-element, these activities should be

quite similar. This test (capped monocistronic versus

172 I. N. Shatsky et al. Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

Figure 3. Speculative modes of mRNA recruitment to ribosomes used by distinct classes of cellular mRNAs. (A) The mRNAs that are highlydependent on eIF4E availability: blocking of eIF4E by dephosphorylated 4E-BPs results in dissociation of eIF4G + eIF4A from such mRNAs and theirtranslational inactivation. (B) The mRNAs which are less dependent on the eIF4E – m7G cap interaction and therefore less sensitive to mTORinhibition. These mRNAs may have a higher affinity of eIF4G + eIF4A to their 50 UTRs (may have CITEs) and remain active in the translationinitiation even under complete inhibition of eIF4E. (C) Specialized combinations of eIF4E and eIF4G. Specific variants of these factors are denoted aseIF4Ex and eIF4Gx. The corresponding forms of eIF4F may bind to particular mRNAs and activate them for translation during development anddifferentiation. Some of these combinations may be refractory to standard 4E-BPs. (D) The mRNAs that use eIF4G2 or eIF4G-like factors rather thaneIF4F. These factors may recognize the corresponding 50 UTRs with the help of auxiliary mRNA binding proteins (denoted with ‘‘Y’’). Such mRNAsare not necessarily cap-independent: they may use non-canonical cap-binding molecules (denoted as ‘‘X’’) which communicate with eIF4G-like factorsthrough specific mRNA-binding proteins. Alternatively, such mRNAs may employ IRES- or CITE-based mechanisms. All the modes of mRNArecruitment which are illustrated in this figure require a free 50 terminus and involve scanning from the 50 end. The IRES-driven translation initiation isdepicted in Figure 2(B).

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 173

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

intercistronic) should be performed both under normal and

stress conditions, to check whether or not the putative IRES is

induced by stress.

Concluding remarks and hypotheses

A growing body of evidence hints at the existence of

significant modifications in the standard mechanism of

mRNA recruitment to eukaryotic ribosomes. They may

differ in the set of recruitment factors or their concentrations.

In addition, subsets of mRNAs may employ, at least under

some specific conditions, mechanisms that do not involve

eIF4E or even eIF4F. We do not exclude that IRES-elements

may also be implicated in the translation initiation of cellular

mRNAs. We only stress that all putative IRESs should be

verified and carefully characterized using stringent criteria to

show that internal ribosome binding activities do have

reasonable and physiologically justified values. These

values should be at least comparable with those of the 50

end-dependent initiation for the same 50 UTRs. On the

basis of information available to date, we can propose

several speculative modes of mRNA recruitment onto 40S

ribosomes that are used by distinct classes of eukaryotic

mRNAs:

Class 1. Highly eIF4E- and eIF4G-dependent mRNAs

(Figure 3A): The number of such mRNAs is relatively small

but they make up a substantial portion of the total amount of

mRNA within polysomes of proliferating cells. Their trans-

lation rapidly responds to changes in nutrient availability and

growth factors. They are strongly and directly controlled by

mTORC1, mostly via phosphorylation of 4E-BPs and are

represented by mRNAs responsible for growth and prolifer-

ation. Interestingly, restriction of their translation extends life

span (Harrison et al., 2009; Syntichaki et al., 2007; Zid et al.,

2009). They are overwhelmingly represented by mRNAs that

encode components of translational apparatus (ribosomal

proteins, some initiation factors, elongation factors eEF1 and

eEF2, PABP, etc.) as well as some factors participating in cell

cycle and epithelial-mesenchymal transition. All 50 TOP

mRNAs belong to this class. Their translation initiation is

strongly dependent on the concentration of free eIF4E, but

also dependent on eIF4G1/3. Importantly, eIF4G alone is not

able to support their efficient recruitment to 40S ribosomes,

presumably, because of a low affinity to the corresponding

UTRs. That is why disruption of the interaction of eIF4E with

eIF4G upon 4E-BPs activation results in a rapid dissociation

of eIF4G and eIF4A from such mRNAs. This process may be

additionally promoted by 50 TOP-binding proteins, TIA-1 and

TIAR (Damgaard & Lykke-Andersen, 2011).

Class 2. eIF4G-controlled mRNAs (Figure 3B): mRNAs of

this class may contain within their 50 UTRs sites with a high

affinity to eIF4G. Such mRNAs require less eIF4E or can

even do without this factor (e.g. when they harbor CITEs),

especially in non-dividing cells where the competition for

ribosomes with mRNAs of class 1 is low or even absent. It is

not excluded that a more stable and specific binding of eIF4G

to a 50 UTR may produce conformational changes in this

factor which in turn enable a more stable binding of eIF4E.

In this case, these mRNAs might successfully compete with

4E-BPs for eIF4E binding. It is worth mentioning that the

translation of the mRNAs which code for all eIF4G variants is

resistant to mTOR inhibition (4E-BPs activation) (Thoreen

et al., 2012).

Class 3. mRNAs controlled by specific variants of eIF4E,

eIF4G and their combinations (Figure 3C): As noted above,

the use of alternative forms of eIF4E and eIF4G during

mitosis and development is well documented, especially in

Drosophila. It is logical to predict the existence of their

specific combinations operating with particular mRNAs.

However, this issue has not been explored and we do not

know how these alternative forms and combinations are

selected by specific mRNAs. It may well be that this selection

is assisted by other auxiliary factors or mRNA-binding

proteins.

Class 4. mRNAs that do not use eIF4F (Figure 3D): They

may utilize a truncated form of eIF4G1, eIF4G2 (DAP5) or

other functional analogs of eIF4G1 (eIF4G-like proteins). The

mRNAs controlled by such factors are not necessary cap-

independent. They may use a cap-binding protein that

communicates with an eIF4G-like protein via a specific

mRNA-binding protein. Alternatively, DAP5 or similar

proteins may be a critical factor that governs internal

initiation on some cellular IRESs.

Finally, we have no idea about the factor requirement for

ribosome recycling through preformed polysomes. It may

well be that polysomes assembled on some mRNAs no longer

need eIF4E, eIF4G or even eIF4A for new rounds of

translation. This could be an alternative explanation why

eIF4G depletion produced a relatively minor effect on

eukaryotic translation. Some fragmentary data on insensitiv-

ity of preformed polysomes to a cap-analog (Amrani et al.,

2008) or non-hydrolysable derivative of ATP (Kopeina et al.,

2008) should encourage further experiments in this direction.

Certainly, there may be many more specific mechanisms

that determine translation initiation on eukaryotic mRNA.

This may be an exciting direction for further studies

of eukaryotic translation and its regulation. In addition,

knowledge of these mechanisms would be useful if we want

to design drugs targeting the translation of specific classes

of mRNAs in particular cells. To succeed in this tremendous

and very ambitious task, we should intensify not only

transcriptome-wide studies of translation but also focus

special attention on developing advanced biochemical tech-

niques and sophisticated in vitro approaches to study the

underlying mechanisms in detail. Otherwise, our knowledge

will remain just a myriad of effects, observations and their

arbitrary interpretations. Development of such techniques

certainly lags behind genome- and transcriptome-wide

studies.

Acknowledgements

We apologize to our colleagues for not being able to discuss

other important contributions due to space constraints. The

authors are very grateful to Gary Loughran (Cork Institute,

Ireland) for critical reading of the manuscript and valuable

suggestions.

Declaration of interest

The authors report no declaration of interest.

174 I. N. Shatsky et al. Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

We acknowledge financial support from the Russian

Foundation of Basic Researches to I.N.S. (grant N 11-04-

01010) and S.E.D. (grant N 12-04-33196).

References

Alain T, Morita M, Fonseca BD, et al. (2012). eIF4E/4E-BP ratiopredicts the efficacy of mTOR targeted therapies. Cancer Res 72:6468–76.

Alekhina OM, Vassilenko KS. (2012). Translation initiation in eukary-otes: versatility of the scanning model. Biochemistry (Mosc) 77:1465–77.

Amrani N, Ghosh S, Mangus DA, Jacobson A. (2008). Translationfactors promote the formation of two states of the closed-loop mRNP.Nature 453:1276–80.

Andreev DE, Dmitriev SE, Terenin IM, et al. (2009). Differentialcontribution of the m7G-cap to the 50 end-dependent translationinitiation of mammalian mRNAs. Nucleic Acids Res 37:6135–47.

Andreev DE, Dmitriev SE, Zinovkin R, et al. (2012). The 50 untranslatedregion of Apaf-1 mRNA directs translation under apoptosis conditionsvia a 50 end-dependent scanning mechanism. FEBS Lett 586:4139–43.

Badura M, Braunstein S, Zavadil J, Schneider RJ. (2012). DNA damageand eIF4G1 in breast cancer cells reprogram translation for survivaland DNA repair mRNAs. Proc Natl Acad Sci USA 109:18767–72.

Baker CC, Fuller MT. (2007). Translational control of meiotic cell cycleprogression and spermatid differentiation in male germ cells by anovel eIF4G homolog. Development 134:2863–9.

Berset C, Zurbriggen A, Djafarzadeh S, et al. (2003). RNA-bindingactivity of translation initiation factor eIF4G1 from Saccharomycescerevisiae. RNA 9:871–80.

Bolger TA, Wente SR. (2011). Gle1 is a multifunctional DEAD-boxprotein regulator that modulates Ded1 in translation initiation. J BiolChem 286:39750–9.

Borman AM, Michel YM, Kean KM. (2000). Biochemical characterisa-tion of cap-poly(A) synergy in rabbit reticulocyte lysates: theeIF4G-PABP interaction increases the functional affinity of eIF4Efor the capped mRNA 50-end. Nucleic Acids Res 28:4068–75.

Bush MS, Hutchins AP, Jones AM, et al. (2009). Selective recruitment ofproteins to 50 cap complexes during the growth cycle in Arabidopsis.Plant J 59:400–12.

Byrd MP, Zamora M, Lloyd RE. (2005). Translation of eukaryotictranslation initiation factor 4GI (eIF4GI) proceeds from multiplemRNAs containing a novel cap-dependent internal ribosome entry site(IRES) that is active during poliovirus infection. J Biol Chem 280:18610–22.

Cakmakci NG, Lerner RS, Wagner EJ, et al. (2008). SLIP1, a factorrequired for activation of histone mRNA translation by the stem-loopbinding protein. Mol Cell Biol 28:1182–94.

Castelli LM, Lui J, Campbell SG, et al. (2011). Glucose depletioninhibits translation initiation via eIF4A loss and subsequent 48Spreinitiation complex accumulation, while the pentose phosphatepathway is coordinately up-regulated. Mol Biol Cell 22:3379–93.

Cha JD, Kim HJ, Cha IH. (2011). Genetic alterations in oral squamouscell carcinoma progression detected by combining array-basedcomparative genomic hybridization and multiplex ligation-dependentprobe amplification. Oral Surg Oral Med Oral Pathol Oral RadiolEndod 111:594–607.

Cho PF, Poulin F, Cho-Park YA, et al. (2005). A new paradigm fortranslational control: inhibition via 50-30 mRNA tethering by Bicoidand the eIF4E cognate 4EHP. Cell 121:411–23.

Choe J, Oh N, Park S, et al. (2012). Translation initiation on mRNAsbound by nuclear cap-binding protein complex CBP80/20 requiresinteraction between CBP80/20-dependent translation initiation factorand eukaryotic translation initiation factor 3g. J Biol Chem 287:18500–9.

Clarkson BK, Gilbert WV, Doudna JA. (2010). Functional overlapbetween eIF4G isoforms in Saccharomyces cerevisiae. PLoS One 5:e9114.

Coldwell MJ, Morley SJ. (2006). Specific isoforms of translationinitiation factor 4GI show differences in translational activity. MolCell Biol 26:8448–60.

Coldwell MJ, Sack U, Cowan JL, et al. (2012). Multiple isoforms of thetranslation initiation factor eIF4GII are generated via use of

alternative promoters, splice sites and a non-canonical initiationcodon. Biochem J 448:1–11.

Colina R, Costa-Mattioli M, Dowling RJ, et al. (2008). Translationalcontrol of the innate immune response through IRF-7. Nature 452:323–8.

Contreras V, Richardson MA, Hao E, Keiper BD. (2008). Depletionof the cap-associated isoform of translation factor eIF4Ginduces germline apoptosis in C. elegans. Cell Death Differ 15:1232–42.

Cridge AG, Castelli LM, Smirnova JB, et al. (2010). Identifying eIF4E-binding protein translationally-controlled transcripts reveals links tomRNAs bound by specific PUF proteins. Nucleic Acids Res 38:8039–50.

Damgaard CK, Lykke-Andersen J. (2011). Translational coregulation of50TOP mRNAs by TIA-1 and TIAR. Genes Dev 25:2057–68.

Darnell JC, Richter JD. (2012). Cytoplasmic RNA-binding proteins andthe control of complex brain function. Cold Spring Harb Perspect Biol4:a012344.

Dennis MD, Jefferson LS, Kimball SR. (2012). Role ofp70S6K1-mediated phosphorylation of eIF4B and PDCD4proteins in the regulation of protein synthesis. J Biol Chem 287:42890–9.

Dmitriev SE, Andreev DE, Adyanova ZV, et al. (2009). Efficient cap-dependent translation of mammalian mRNAs with long and highlystructured 50-untranslated regions in vitro and in vivo. Mol Biol(Mosk) 43:108–13.

Dmitriev SE, Andreev DE, Terenin IM, et al. (2007). Efficienttranslation initiation directed by the 900-nucleotide-long and GC-rich 50 untranslated region of the human retrotransposon LINE-1mRNA is strictly cap dependent rather than internal ribosome entrysite mediated. Mol Cell Biol 27:4685–97.

Dmitriev SE, Terenin IM, Dunaevsky YE, et al. (2003). Assembly of 48Stranslation initiation complexes from purified components withmRNAs that have some base pairing within their 50 untranslatedregions. Mol Cell Biol 23:8925–33.

Dobrikov MI, Shveygert M, Brown MC, Gromeier M. (2014). MitoticPhosphorylation of eukaryotic initiation factor 4G1 (eIF4G1) atSer1232 by Cdk1: cyclin B inhibits eIF4A helicase complex bindingwith RNA. Mol Cell Biol 34:439–51.

Feoktistova K, Tuvshintogs E, Do A, Fraser CS. (2013). Human eIF4Epromotes mRNA restructuring by stimulating eIF4A helicase activity.Proc Natl Acad Sci USA 110:13339–44.

Filbin ME, Kieft JS. (2011). HCV IRES domain IIb affects theconfiguration of coding RNA in the 40S subunit’s decoding groove.RNA 17:1258–73.

Franklin-Dumont TM, Chatterjee C, Wasserman SA, Dinardo S. (2007).A novel eIF4G homolog, Off-schedule, couples translational controlto meiosis and differentiation in Drosophila spermatocytes.Development 134:2851–61.

Garre E, Romero-Santacreu L, De Clercq N, et al. (2012). Yeast mRNAcap-binding protein Cbc1/Sto1 is necessary for the rapid reprogram-ming of translation after hyperosmotic shock. Mol Biol Cell 23:137–50.

Geiger T, Velic A, Macek B, et al. (2013). Initial quantitative proteomicmap of 28 mouse tissues using the SILAC mouse. Mol CellProteomics 12:1709–22.

Gilbert WV, Zhou K, Butler TK, Doudna JA. (2007). Cap-independenttranslation is required for starvation-induced differentiation in yeast.Science 317:1224–7.

Gingras AC, Raught B, Sonenberg N. (1999). eIF4 initiation factors:effectors of mRNA recruitment to ribosomes and regulators oftranslation. Annu Rev Biochem 68:913–63.

Gkogkas CG, Khoutorsky A, Ran I, et al. (2013). Autism-related deficitsvia dysregulated eIF4E-dependent translational control. Nature 493:371–7.

Guo L, Allen E, Miller WA. (2000). Structure and function of a cap-independent translation element that functions in either the 30 or the 50

untranslated region. RNA 6:1808–20.Haghighat A, Mader S, Pause A, Sonenberg N. (1995). Repression of

cap-dependent translation by 4E-binding protein 1: competition withp220 for binding to eukaryotic initiation factor-4E. EMBO J 14:5701–9.

Haghighat A, Sonenberg N. (1997). eIF4G dramatically enhances thebinding of eIF4E to the mRNA 50-cap structure. J Biol Chem 272:21677–80.

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 175

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

Harris TE, Chi A, Shabanowitz J, et al. (2006). mTOR-dependentstimulation of the association of eIF4G and eIF3 by insulin. EMBO J25:1659–68.

Harrison DE, Strong R, Sharp ZD, et al. (2009). Rapamycin fed late inlife extends lifespan in genetically heterogeneous mice. Nature 460:392–5.

Hernandez G, Han H, Gandin V, et al. (2012). Eukaryotic initiationfactor 4E-3 is essential for meiotic chromosome segregation,cytokinesis and male fertility in Drosophila. Development 139:3211–20.

Hernandez G, Miron M, Han H, et al. (2013). Mextli is a noveleukaryotic translation initiation factor 4E-binding protein that pro-motes translation in Drosophila melanogaster. Mol Cell Biol 33:2854–64.

Hernandez G, Vazquez-Pianzola P. (2005). Functional diversity of theeukaryotic translation initiation factors belonging to eIF4 families.Mech Dev 122:865–76.

Hinnebusch AG, Lorsch JR. (2012). The mechanism of eukaryotictranslation initiation: new insights and challenges. Cold Spring HarbPerspect Biol 4:a011544.

Hsieh AC, Liu Y, Edlind MP, et al. (2012). The translational landscapeof mTOR signalling steers cancer initiation and metastasis. Nature485:55–61.

Imataka H, Olsen HS, Sonenberg N. (1997). A new translationalregulator with homology to eukaryotic translation initiation factor 4G.EMBO J 16:817–25.

Ingolia NT, Brar GA, Rouskin S, et al. (2012). The ribosome profilingstrategy for monitoring translation in vivo by deep sequencing ofribosome-protected mRNA fragments. Nat Protoc 7:1534–50.

Jackson RJ. (2013). The current status of vertebrate cellular mRNAIRESs. Cold Spring Harb Perspect Biol 5:a011569.

Jang SK, Krausslich HG, Nicklin MJ, et al. (1988). A segment of the 50

nontranslated region of encephalomyocarditis virus RNA directsinternal entry of ribosomes during in vitro translation. J Virol 62:2636–43.

Jia Y, Polunovsky V, Bitterman PB, Wagner CR. (2012). Cap-dependenttranslation initiation factor eIF4E: an emerging anticancer drug target.Med Res Rev 32:786–814.

Jivotovskaya AV, Valasek L, Hinnebusch AG, Nielsen KH. (2006).Eukaryotic translation initiation factor 3 (eIF3) and eIF2 can promotemRNA binding to 40S subunits independently of eIF4G in yeast. MolCell Biol 26:1355–72.

Joshi B, Cameron A, Jagus R. (2004). Characterization of mammalianeIF4E-family members. Eur J Biochem 271:2189–203.

Joshi B, Lee K, Maeder DL, Jagus R. (2005). Phylogenetic analysis ofeIF4E-family members. BMC Evol Biol 5:48.

Kapeli K, Yeo GW. (2012). Genome-wide approaches to dissect the rolesof RNA binding proteins in translational control: implications forneurological diseases. Front Neurosci 6:144.

Kim KM, Cho H, Choi K, et al. (2009). A new MIF4G domain-containing protein, CTIF, directs nuclear cap-binding protein CBP80/20-dependent translation. Genes Dev 23:2033–45.

Kopeina GS, Afonina ZA, Gromova KV, et al. (2008). Step-wiseformation of eukaryotic double-row polyribosomes and circulartranslation of polysomal mRNA. Nucleic Acids Res 36:2476–88.

Koromilas AE, Lazaris-Karatzas A, Sonenberg N. (1992). mRNAscontaining extensive secondary structure in their 50 non-coding regiontranslate efficiently in cells overexpressing initiation factor eIF-4E.EMBO J 11:4153–8.

Kozak M. (1989). The scanning model for translation: an update. J CellBiol 108:229–41.

Laplante M, Sabatini DM. (2012). mTOR signaling in growth controland disease. Cell 149:274–93.

Lasko P. (2012). mRNA localization and translational control inDrosophila oogenesis. Cold Spring Harb Perspect Biol 4:a012294.

Lee SH, McCormick F. (2006). p97/DAP5 is a ribosome-associatedfactor that facilitates protein synthesis and cell proliferation bymodulating the synthesis of cell cycle proteins. EMBO J 25:4008–19.

Lemp NA, Hiraoka K, Kasahara N, Logg CR. (2012). Cryptic transcriptsfrom a ubiquitous plasmid origin of replication confound tests for cis-regulatory function. Nucleic Acids Res 40:7280–90.

Levy-Strumpf N, Deiss LP, Berissi H, Kimchi A. (1997). DAP-5, a novelhomolog of eukaryotic translation initiation factor 4G isolated as aputative modulator of gamma interferon-induced programmed celldeath. Mol Cell Biol 17:1615–25.

Liberman N, Marash L, Kimchi A. (2009). The translation initiationfactor DAP5 is a regulator of cell survival during mitosis. Cell Cycle8:204–9.

Macejak DG, Sarnow P. (1991). Internal initiation of translationmediated by the 50 leader of a cellular mRNA. Nature 353:90–4.

Mader S, Lee H, Pause A, Sonenberg N. (1995). The translationinitiation factor eIF-4E binds to a common motif shared by thetranslation factor eIF-4 gamma and the translational repressors4E-binding proteins. Mol Cell Biol 15:4990–7.

Malygin AA, Kossinova OA, Shatsky IN, Karpova GG. (2013). HCVIRES interacts with the 18S rRNA to activate the 40S ribosomefor subsequent steps of translation initiation. Nucleic Acids Res41:8706–14.

Mamane Y, Petroulakis E, Martineau Y, et al. (2007). Epigeneticactivation of a subset of mRNAs by eIF4E explains its effects on cellproliferation. PLoS One 2:e242.

Maquat LE, Hwang J, Sato H, Tang Y. (2010). CBP80-promoted mRNPrearrangements during the pioneer round of translation, nonsense-mediated mRNA decay, and thereafter. Cold Spring Harb Symp QuantBiol 75:127–34.

Marash L, Kimchi A. (2005). DAP5 and IRES-mediated translationduring programmed cell death. Cell Death Differ 12:554–62.

Marash L, Liberman N, Henis-Korenblit S, et al. (2008). DAP5 promotescap-independent translation of Bcl-2 and CDK1 to facilitate cellsurvival during mitosis. Mol Cell 30:447–59.

Marintchev A. (2013). Roles of helicases in translation initiation: amechanistic view. Biochim Biophys Acta 1829:799–809.

Martineau Y, Derry MC, Wang X, et al. (2008). Poly(A)-binding protein-interacting protein 1 binds to eukaryotic translation initiation factor3 to stimulate translation. Mol Cell Biol 28:6658–67.

Miller WA, Wang Z, Treder K. (2007). The amazing diversity of cap-independent translation elements in the 30-untranslated regions ofplant viral RNAs. Biochem Soc Trans 35:1629–33.

Mitchell SF, Walker SE, Algire MA, et al. (2010). The50-7-methylguanosine cap on eukaryotic mRNAs serves both tostimulate canonical translation initiation and to block an alternativepathway. Mol Cell 39:950–62.

Morita M, Ler LW, Fabian MR, et al. (2012). A novel 4EHP-GIGYF2translational repressor complex is essential for mammalian develop-ment. Mol Cell Biol 32:3585–93.

Mothe-Satney I, Yang D, Fadden P, et al. (2000). Multiple mechanismscontrol phosphorylation of PHAS-I in five (S/T)P sites that governtranslational repression. Mol Cell Biol 20:3558–67.

Neusiedler J, Mocquet V, Limousin T, et al. (2012). INT6 interacts withMIF4GD/SLIP1 and is necessary for efficient histone mRNAtranslation. RNA 18:1163–77.

Nicholson BL, Wu B, Chevtchenko I, White KA. (2010). Tombusvirusrecruitment of host translational machinery via the 30 UTR. RNA 16:1402–19.

Nousch M, Reed V, Bryson-Richardson RJ, et al. (2007). The eIF4G-homolog p97 can activate translation independent of caspase cleavage.RNA 13:374–84.

Osborne MJ, Volpon L, Kornblatt JA, et al. (2013). eIF4E3 actsas a tumor suppressor by utilizing an atypical mode of methyl-7-guanosine cap recognition. Proc Natl Acad Sci U S A 110:3877–82.

Ozpolat B, Akar U, Zorrilla-Calancha I, et al. (2008). Death-associatedprotein 5 (DAP5/p97/NAT1) contributes to retinoic acid-inducedgranulocytic differentiation and arsenic trioxide-induced apoptosis inacute promyelocytic leukemia. Apoptosis 13:915–28.

Park EH, Walker SE, Lee JM, et al. (2011a). Multiple elements in theeIF4G1 N-terminus promote assembly of eIF4G1*PABP mRNPsin vivo. EMBO J 30:302–16.

Park EH, Zhang F, Warringer J, et al. (2011b). Depletion of eIF4G fromyeast cells narrows the range of translational efficiencies genome-wide. BMC Genomics 12:68.

Parsyan A, Svitkin Y, Shahbazian D, et al. (2011). mRNA helicases:the tacticians of translational control. Nat Rev Mol Cell Biol 12:235–45.

Pelletier J, Sonenberg N. (1988). Internal initiation of translation ofeukaryotic mRNA directed by a sequence derived from poliovirusRNA. Nature 334:320–5.

Pestova TV, Hellen CU, Shatsky IN. (1996a). Canonical eukaryoticinitiation factors determine initiation of translation by internalribosomal entry. Mol Cell Biol 16:6859–69.

176 I. N. Shatsky et al. Crit Rev Biochem Mol Biol, 2014; 49(2): 164–177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.

Pestova TV, Shatsky IN, Fletcher SP, et al. (1998). A prokaryotic-likemode of cytoplasmic eukaryotic ribosome binding to the initiationcodon during internal translation initiation of hepatitis C and classicalswine fever virus RNAs. Genes Dev 12:67–83.

Pestova TV, Shatsky IN, Hellen CU. (1996b). Functional dissection ofeukaryotic initiation factor 4F: the 4A subunit and the central domainof the 4G subunit are sufficient to mediate internal entry of 43Spreinitiation complexes. Mol Cell Biol 16:6870–8.

Pisareva VP, Pisarev AV, Komar AA, et al. (2008). Translation initiationon mammalian mRNAs with structured 50UTRs requires DExH-boxprotein DHX29. Cell 135:1237–50.

Prevot D, Decimo D, Herbreteau CH, et al. (2003). Characterization of anovel RNA-binding region of eIF4GI critical for ribosomal scanning.EMBO J 22:1909–21.

Rajagopal V, Park EH, Hinnebusch AG, Lorsch JR. (2012). Specificdomains in yeast translation initiation factor eIF4G strongly bias RNAunwinding activity of the eIF4F complex toward duplexes with 50-overhangs. J Biol Chem 287:20301–12.

Ramirez-Valle F, Braunstein S, Zavadil J, et al. (2008). eIF4GI linksnutrient sensing by mTOR to cell proliferation and inhibition ofautophagy. J Cell Biol 181:293–307.

Rhoads RE. (2009). eIF4E: new family members, new binding partners,new roles. J Biol Chem 284:16711–15.

Richter JD. (2007). CPEB: a life in translation. Trends Biochem Sci 32:279–85.

Roux PP, Topisirovic I. (2012). Regulation of mRNA translation bysignaling pathways. Cold Spring Harb Perspect Biol 4:a012252.

Ruggero D. (2013). Translational control in cancer etiology. Cold SpringHarb Perspect Biol 5:a012336.

Schwanhausser B, Busse D, Li N, et al. (2011). Global quantification ofmammalian gene expression control. Nature 473:337–42.

Shagam LI, Terenin IM, Andreev DE, et al. (2012). In vitro activity ofhuman translation initiation factor eIF4B is not affected byphosphomimetic amino acid substitutions S422D and S422E.Biochimie 94:2484–90.

Sharma A, Yilmaz A, Marsh K, et al. (2012). Thriving under stress:selective translation of HIV-1 structural protein mRNA during Vpr-mediated impairment of eIF4E translation activity. PLoS Pathog 8:e1002612.

Shatsky IN, Dmitriev SE, Terenin IM, Andreev DE. (2010). Cap- andIRES-independent scanning mechanism of translation initiation as analternative to the concept of cellular IRESs. Mol Cells 30:285–93.

Shaughnessy Jr JD, Jenkins NA, Copeland NG. (1997). cDNA cloning,expression analysis, and chromosomal localization of a gene with highhomology to wheat eIF-(iso)4F and mammalian eIF-4G. Genomics39:192–7.

Sonenberg N, Hinnebusch AG. (2009). Regulation of translationinitiation in eukaryotes: mechanisms and biological targets. Cell136:731–45.

Soto-Rifo R, Rubilar PS, Limousin T, et al. (2012). DEAD-box proteinDDX3 associates with eIF4F to promote translation of selectedmRNAs. EMBO J 31:3745–56.

Srivastava T, Fortin DA, Nygaard S, et al. (2012). Regulation ofneuronal mRNA translation by CaM-kinase I phosphorylation ofeIF4GII. J Neurosci 32:5620–30.

Stupina VA, Meskauskas A, Mccormack JC, et al. (2008). The 30

proximal translational enhancer of Turnip crinkle virus binds to 60Sribosomal subunits. RNA 14:2379–93.

Sun J, Conn CS, Han Y, et al. (2011). PI3K-mTORC1 attenuates stressresponse by inhibiting cap-independent Hsp70 translation. J BiolChem 286:6791–800.

Svitkin YV, Pause A, Haghighat A, et al. (2001). The requirement foreukaryotic initiation factor 4A (elF4A) in translation is in directproportion to the degree of mRNA 50 secondary structure. RNA 7:382–94.

Syntichaki P, Troulinaki K, Tavernarakis N. (2007). eIF4E function insomatic cells modulates ageing in Caenorhabditis elegans. Nature445:922–6.

Takahashi K, Maruyama M, Tokuzawa Y, et al. (2005). Evolutionarilyconserved non-AUG translation initiation in NAT1/p97/DAP5(EIF4G2). Genomics 85:360–71.

Terenin IM, Andreev DE, Dmitriev SE, Shatsky IN. (2013). Anovel mechanism of eukaryotic translation initiation that isneither m7G-cap-, nor IRES-dependent. Nucleic Acids Res 41:1807–16.

Terenin IM, Dmitriev SE, Andreev DE, et al. (2005). A cross-kingdominternal ribosome entry site reveals a simplified mode of internalribosome entry. Mol Cell Biol 25:7879–88.

Thoreen CC, Chantranupong L, Keys HR, et al. (2012). A unifyingmodel for mTORC1-mediated regulation of mRNA translation. Nature485:109–13.

Treder K, Kneller EL, Allen EM, et al. (2008). The 30 cap-independenttranslation element of Barley yellow dwarf virus binds eIF4F via theeIF4G subunit to initiate translation. RNA 14:134–47.

Uniacke J, Holterman CE, Lachance G, et al. (2012). An oxygen-regulated switch in the protein synthesis machinery. Nature 486:126–9.

Virgili G, Frank F, Feoktistova K, et al. (2013). Structural analysis of theDAP5 MIF4G domain and its interaction with eIF4A. Structure 21:517–27.

Von Der Haar T, Ball PD, McCarthy JE. (2000). Stabilization ofeukaryotic initiation factor 4E binding to the mRNA 50-Cap bydomains of eIF4G. J Biol Chem 275:30551–5.

Von Moeller H, Lerner R, Ricciardi A, et al. (2013). Structural andbiochemical studies of SLIP1-SLBP identify DBP5 and eIF3g asSLIP1-binding proteins. Nucleic Acids Res 41:7960–71.

Wallace A, Filbin ME, Veo B, et al. (2010). The nematode eukaryotictranslation initiation factor 4E/G complex works with a trans-splicedleader stem-loop to enable efficient translation of trimethylguanosine-capped RNAs. Mol Cell Biol 30:1958–70.

Wang X, Proud CG. (2011). mTORC1 signaling: what we still don’tknow. J Mol Cell Biol 3:206–20.

Yamanaka S, Poksay KS, Arnold KS, Innerarity TL. (1997). A noveltranslational repressor mRNA is edited extensively in livers contain-ing tumors caused by the transgene expression of the apoB mRNA-editing enzyme. Genes Dev 11:321–33.

Yamanaka S, Zhang XY, Maeda M, et al. (2000). Essential role ofNAT1/p97/DAP5 in embryonic differentiation and the retinoic acidpathway. EMBO J 19:5533–41.

Yanagiya A, Suyama E, Adachi H, et al. (2012). Translationalhomeostasis via the mRNA cap-binding protein, eIF4E. Mol Cell46:847–58.

Zid BM, Rogers AN, Katewa SD, et al. (2009). 4E-BP extends lifespanupon dietary restriction by enhancing mitochondrial activity inDrosophila. Cell 139:149–60.

DOI: 10.3109/10409238.2014.887051 Transcriptome-wide studies of mRNA binding to ribosomes 177

Cri

tical

Rev

iew

s in

Bio

chem

istr

y an

d M

olec

ular

Bio

logy

Dow

nloa

ded

from

info

rmah

ealth

care

.com

by

Uni

vers

ity o

f W

inds

or o

n 06

/05/

14Fo

r pe

rson

al u

se o

nly.