translation reinitiation and ribosome recycling - phd thesis

133
Reinitiation of translation and ribosome recycling are two distinct mechanisms in Saccharomyces cerevisiae A dissertation submitted to the University of Manchester Institute of Science and Technology (UMIST) for the degree of Doctor of Philosophy 2003 LUKAS RAJKOWITSCH Department of Biomolecular Sciences

Upload: others

Post on 16-Oct-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Translation reinitiation and ribosome recycling - PhD Thesis

Reinitiation of translation and ribosome recycling are two distinct mechanisms in

Saccharomyces cerevisiae

A dissertation submitted to the

University of Manchester Institute of Science and Technology

(UMIST) for the degree of Doctor of Philosophy

2003

LUKAS RAJKOWITSCH Department of Biomolecular Sciences

Page 2: Translation reinitiation and ribosome recycling - PhD Thesis

Declaration

No portion of this work referred to in this thesis has been submitted in support of an

application for another degree or qualification of this or any other university, or other

institution of learning.

Part of this work has been submitted in July 2003 to the Journal of Molecular Biology

for publication.

All work described in this thesis was carried out in accordance with Control of

Substances Hazardous to Health regulations (COSHH) in their current versions.

2

Page 3: Translation reinitiation and ribosome recycling - PhD Thesis

Dedication

for Heidi, my parents,

and all of you who supported me

on this way forth and back

1% inspiration, 99% transpiration

3

Page 4: Translation reinitiation and ribosome recycling - PhD Thesis

Acknowledgements

ACKNOWLEDGEMENTS I would like to thank Prof John McCarthy for providing me with the opportunity to

perform the work for this study in his laboratory and for the invaluable guidance

throughout its course. I am also grateful to all the other members of the

Posttranscriptional Control Group at UMIST who dramatically reduced the

theoretical and practical obstacles especially at the outset of this work and –

equally important – provided a lively surrounding I enjoyed working in. Special

thanks go to Cristina Vilela for general assistance and guidance during this

project and to Karine Berthelot as well as Carmen Ramirez Velasco for

introducing me to the wonderful worlds of in vitro transcription and mRNA

stability measurements, respectively, among many other aspects. Agoritsa

Varaklioti kindly provided access to computer resources and storage solutions,

and Mara Nardelli along with Sabina Piccinnini shared extraordinary knowledge

of sucrose gradients. Marina Ptushkina did not leave me alone with (RT-)PCR

or other nucleic acid problems, and Tobias von der Haar was always there for

information about translation in general and protein methods in particular.

Finally, John Hughes was at all times remarkably inventive in the fields of

financial administration.

Special thanks to Yoav Arava of Stanford University for providing sucrose

gradient protocols.

4

Page 5: Translation reinitiation and ribosome recycling - PhD Thesis

Abstract

ABSTRACT Historically, translation termination was viewed simply as the last step in the

process of cellular protein synthesis. Recent findings have changed this

perception dramatically, indicating that events at this stage decide about the

fate of the posttermination ribosome and about the degradation of the translated

mRNA. In particular, it remains to be elucidated how terminating ribosomes are

recycled back to start another round of translation. The circularisation model of

the polysomal mRNA suggests that this ribosome recycling might be facilitated

by 5' – 3' interactions mediated by the cap-binding complex eIF4F and the

poly(A) binding protein, Pab1p. In contrast, downstream of a short upstream

open reading frame (uORF) in the 5' untranslated region of a gene,

posttermination ribosomes can maintain the competence to (re)initiate

translation. This study shows that recycling and reinitiation are distinct

processes in Saccharomyces cerevisiae. Recycling via the 3'-UTR was

assessed by restricting ribosome movement along the mRNA using a poly(G)

stretch or the mammalian iron regulatory protein (IRP1) bound to the iron

responsive element (IRE). Although 3'-UTR structures were found to influence

translation, the main pathway of ribosome recycling does not depend on

scanning-like movement through the 3'-UTR. Changes in termination kinetics or

disruption of the Pab1 – eIF4F interaction do not affect recycling, yet the

maintenance of normal in vivo mRNP structure is important to this process.

Using bicistronic ACT1-LUC constructs, elongating yeast ribosomes were found

to maintain the competence to (re)initiate over only short distances. Thus, as

the first ORF to be translated is progressively truncated, reinitiation downstream

of an uORF of 105 nucleotides is found to be just detectable, and increases

markedly in efficiency as uORF length is reduced to 15 nucleotides.

Experiments using a strain mutated in the Cca1 nucleotidyltransferase suggest

that the uORF length-dependence of changes in reinitiation competence is

affected by peptide elongation kinetics, but that ORF length per se may also be

relevant. Thus, the loss of the reinitiation potential of elongating ribosomes

appears to render these incapable to promote intramolecular recycling via

scanning of the 3'-UTR.

5

Page 6: Translation reinitiation and ribosome recycling - PhD Thesis

I. Contents

I. TABLE OF CONTENT

ACKNOWLEDGEMENTS....................................................................................................................... 4 ABSTRACT ................................................................................................................................................ 5 I. TABLE OF CONTENT.......................................................................................................................... 6 II. ABBREVIATIONS ............................................................................................................................... 8 III. LIST OF FIGURES AND TABLES................................................................................................... 9 1. INTRODUCTION ................................................................................................................................ 10

1.1 CONTROLLING GENE EXPRESSION IN YEAST: FROM TRANSCRIPTION TERMINATION TO TRANSLATION INITIATION ................................................................................................................ 11 1.1.1. Pre-mRNA processing........................................................................................................... 11 1.1.2 Translation in the nucleus?..................................................................................................... 13 1.1.3 Transcript export from the nucleus ........................................................................................ 14

1.2 FACTORS AND MECHANISMS IN EUKARYOTIC TRANSLATION ........................................................... 15 1.2.1 Translation initiation .............................................................................................................. 15

1.2.1.1 Recruitment of the 43S preinitiation complex to the 5'-cap ........................................................... 16 1.2.1.2 Selection of the translational start site and peptide synthesis initiation ........................................ 18

1.2.2 Elongation .............................................................................................................................. 20 1.2.3 Termination............................................................................................................................ 21

1.2.3.1 eRF1: tRNA mimicry and stop codon recognition ......................................................................... 21 1.2.3.2 eRF3: from little known eRF1-binding partner to hub for virtually all translation termination

events ............................................................................................................................................ 23 1.2.4 Elements in the 5'-UTR affecting translational efficiency ..................................................... 25

1.2.4.1 AUG context .................................................................................................................................. 25 1.2.4.2 Leader length................................................................................................................................. 27 1.2.4.3 Upstream AUGs and ORFs ........................................................................................................... 27 1.2.4.4 Secondary structures and mRNA-protein interactions................................................................... 28

1.2.5 The circular-loop model of eukaryotic mRNA ...................................................................... 32 1.2.5.1 The mRNA 5' – 3' interaction......................................................................................................... 32 1.2.5.2 The 3'-UTR, a loop within the loop?.............................................................................................. 35

1.3 MRNA STABILITY IN EUKARYOTES.................................................................................................. 35 1.3.1 'Regular' pathways of mRNA turnover in yeast ..................................................................... 36

1.3.1.1 The major pathway of mRNA decay: deadenylation, decapping and 5' 3’ exonucleolytic degradation................................................................................................................................... 36

1.3.1.2 3' 5' exonucleolytic and endonucleolytic mRNA degradation ..................................................... 40 1.3.2 mRNA surveillance and nonsense-mediated decay................................................................ 41

1.3.2.1 The need for surveillance .............................................................................................................. 41 1.3.2.2 Cis- and trans-acting factors in NMD ........................................................................................... 42 1.3.2.3 mRNA surveillance by mRNP (re)organisation ? .......................................................................... 45 1.3.2.4 The evidence for mRNA surveillance in the nucleus...................................................................... 46

1.4 UORFS: REINITIATION OF TRANSLATION VS. RIBOSOME RELEASE ................................................... 48 1.4.1 Polycistronic transcripts in eukaryotes? ................................................................................. 48 1.4.2 uORF control in GCN4 translation: the stop codon context and eIF2 phosphorylation ......... 49 1.4.3 Reinitiation is affected by the intercistronic distance and the uORF length........................... 50 1.4.4 Other factors governing reinitiation ....................................................................................... 52 1.4.5 Reinitiation of translation in prokaryotes and eukaryotes ...................................................... 53 1.4.6 uORFs can regulate transcript stability .................................................................................. 54

1.5 OBJECTIVES..................................................................................................................................... 56 2. MATERIALS AND METHODS......................................................................................................... 57

2.1 CHEMICALS AND ENZYMES.............................................................................................................. 57 2.2 CELL METHODS................................................................................................................................ 58

2.2.1 Strains and cell culturing........................................................................................................ 58 2.2.2 Escherichia coli transformation ............................................................................................. 59 2.2.3 Saccharomyces cerevisiae transformation ............................................................................. 60

6

Page 7: Translation reinitiation and ribosome recycling - PhD Thesis

I. Contents

2.3 RECOMBINANT DNA METHODS....................................................................................................... 61 2.3.1 DNA manipulation ................................................................................................................. 61 2.3.2 Screening of E. coli transformants ......................................................................................... 61 2.3.3 CsCl discontinuous gradient DNA purification ..................................................................... 62 2.3.4 Plasmid constructs.................................................................................................................. 62

2.4 RNA METHODS ............................................................................................................................... 67 2.4.1 Total yeast RNA extraction.................................................................................................... 67 2.4.2 Northern blot .......................................................................................................................... 67 2.4.3 RT-PCR.................................................................................................................................. 68 2.4.4 poly(A) length determination by RNase H treatment............................................................. 69

2.5 PROTEIN METHODS .......................................................................................................................... 69 2.5.1 Luciferase assay and protein quantification ........................................................................... 69 2.5.2 Western blotting ..................................................................................................................... 70

2.6 POLYSOMAL GRADIENT ANALYSIS................................................................................................... 71 2.7 IN VITRO TRANSLATION IN CELL-FREE YEAST EXTRACTS.................................................................. 71 2.8 REPRODUCIBILITY ........................................................................................................................... 73

3. RESULTS.............................................................................................................................................. 74 3.1 RIBOSOME RECYCLING: 3'-UTR ACCESSIBILITY AND TRANSLATIONAL EFFICIENCY........................ 74

3.1.1 The potential role of 3'-UTR accessibility for ribosomal recycling on a circular mRNA...... 74 3.1.2 Structures in the 3'-UTR can reduce translational efficiency ................................................. 74 3.1.3 Recruitment of IRP1 to the 3'-UTR modulates translation and not polyadenylation ............. 79

3.1.3.1 poly(A)-length and -site verification.............................................................................................. 79 3.1.3.2 Polysomal analysis ........................................................................................................................ 81

3.1.4 Translation of a full-length ORF is insensitive to changes of the codon context ................... 83 3.1.5 The eIF4G – Pab1p interaction is not required for 3'-UTR-confered translation repression

in vivo .................................................................................................................................... 85 3.1.5.1 The absence of the cap – poly(A) interaction in vivo does not abolish the translational

inhibition by a 3'-UTR structure ................................................................................................... 85 3.1.5.2 In vitro translation in cell-free extracts is not affected by IRP1-binding to the 3'-UTR ................ 85

3.2 CORRELATION OF ORF LENGTH AND RIBOSOME REINITIATION POTENTIAL ..................................... 87 3.2.1 Reinitiation capability increases as a wild type ORF is progressively truncated ................... 87 3.2.2 Translation of a downstream ORF by reinitiation and leaky scanning can be distinguished

using start codon mutants ...................................................................................................... 91 3.2.3 Reinitiation competence is affected by elongation rate.......................................................... 94 3.2.4 Premature translation termination by either ORF truncation or stop codon insertion affects

mRNA stability differently .................................................................................................... 97 3.2.4.1 The presence of premature termination codons but not 3'-truncation of the ORF trigger

nonsense-mediated decay.............................................................................................................. 97 3.2.4.2 PTC-containing mRNAs are stabilised by a full-length downstream ORF that is not translated .. 99

4. DISCUSSION...................................................................................................................................... 101 4.1 ACCESSIBILITY OF THE 3'-UTR IS NOT A PREREQUISITE FOR POSTTERMINATION RIBOSOME

RECYCLING .................................................................................................................................... 101 4.1.1 Differences between 5'-UTR and 3'-UTR directed blockage ............................................... 101 4.1.2 Translation might be affected by a 3'UTR-conferred mRNP modification.......................... 103 4.1.3 Recycling is unlikely to be reinitiation-like since the ribosomal reinitiation potential

depends on the ORF length.................................................................................................. 105 4.2 THE RELATIONSHIP BETWEEN ORF LENGTH, NONSENSE-MEDIATED DECAY AND REINITIATION.... 106

4.2.1 PTC-insertion but not ORF 3'-truncation results in accelerated mRNA decay .................... 106 4.2.2 Insertion of an untranslated downstream ORF stabilises PTC-carrying mRNAs................. 107 4.2.3 Modulation of leaky scanning and mRNA stability by the efficiency of uORF recognition107

4.3 FUTURE PERSPECTIVES .................................................................................................................. 110 5. REFERENCES ................................................................................................................................... 112

7

Page 8: Translation reinitiation and ribosome recycling - PhD Thesis

II. Abbreviations

II. ABBREVIATIONS

Abbreviations used frequently throughout the text are listed below, for

abbreviations of chemicals and enzymes see Tables 2.1 and Table 2.2 in

Section 2.1.

Cap ..... m7Gppp-R structure at mRNA 5'-end CBC .... nuclear cap-binding complex (e)EF... (eukaryotic) translation elongation factor (e)IF .... (eukaryotic) translation initiation factor (e)RF... (eukaryotic) polypeptide chain release

factor hnRNP heterogeneous nuclear ribonucleoprotein IRE...... iron responsive element IRES ... internal ribosome entry site IRP...... IRE-binding protein mRNP . messenger ribonucleoprotein NMD.... nonsense-mediated decay NPC .... nuclear pore complex OD(x) ... optical density (absorbance) at x

nanometres Pab1p . poly(A)-binding protein Pol II ... DNA-dependent RNA polymerase II PTC..... premature termination codon RNP .... ribonucleoprotein

snRNP small nuclear ribonucleoprotein (u)ORF (upstream) open reading frame Upf ..... up-frameshift Amino acid three letter code Ade..... adenine Arg ..... argenine His ...... histidine Ile ....... Isoleucine Leu ..... leucine Lys ..... lysine phe ..... phenylalanine Ser...... Serine Thr...... threonine Trp...... tryptophan Tyr ...... tyrosine Ura ..... uracil Val ...... valine

8

Page 9: Translation reinitiation and ribosome recycling - PhD Thesis

III.List of Figures and Tables

III. LIST OF FIGURES AND TABLES Figure 1.1 The pathway of eukaryotic mRNA from the nucleus to the sites

of translation and decay in the cytoplasm ................................... 12 Figure 1.2 Steps of translation in Saccharomyces cerevisiae ...................... 17 Figure 1.3 Mechanisms of eukaryotic start codon recognition...................... 26 Figure 1.4 Deadenylation-dependent pathways of mRNA degradation in

S. cerevisiae................................................................................ 37 Figure 1.5 Nonsense-mediated decay of premature termination codon

containing mRNAs....................................................................... 43 Figure 3.1 Repression of translation via IRP1-binding to the 3'-UTR. .......... 76 Figure 3.2 Position-dependent decrease of translational efficiency by a

3'-UTR-located poly(G) tract........................................................ 79 Figure 3.3 Binding of IRP1 to an IRE located in the 3'-UTR does not

affect polyadenylation.................................................................. 80 Figure 3.4 IRP1-binding to IREs in the 5' or 3'-UTR affects the polysomal

distribution of LUC mRNA ........................................................... 82 Figure 3.5 The stop codon context of the full-length luciferase ORF does

not affect translational efficiency or mRNA stability..................... 83 Figure 3.6 Significance of 5' – 3' interactions to recycling ............................ 86 Figure 3.7 Truncation of the first ORF in a bicistronic ACT-LUC construct

leads to increased expression of the downstream ORF.............. 89 Figure 3.8 3'-truncation of an actin ORF to 63 nt or shorter triggers decay

of the monocistronic mRNA......................................................... 90 Figure 3.9 Elements that modulate reinitiation affect the low luciferase

activity displayed by the full-length actin-luciferase mRNA. ........ 91 Figure 3.10 Manipulation of the FLAG-ACT mini-ORF upstream of LUC ....... 93 Figure 3.11 Inhibition of translation elongation affects expression of a

downstream luciferase ORF in a uORF-size dependent manner ........................................................................................ 96

Figure 3.12 Impact of a premature stop codon on the translation and abundance of monocistronic ACT and bicistronic the ACT-LUC mRNA.......................................................................................... 98

Figure 4.1 Channelling is promoted by mRNP structure ............................ 104 Figure 4.2 5'-UTR sequence of ACTd105-LUC and representation of

ORFs starting at all possible upstream ATGs. .......................... 108 Figure 4.3 A ribosome flow model for the inhibition of downstream

(re)initiation by uORFs comprising uAUGs................................ 109 Table 2.1 Chemicals used........................................................................... 57 Table 2.2 Enzymes and antibodies. ............................................................ 58 Table 2.3 Correlation of rpm and RCF values in centrifuges and rotors

used. ........................................................................................... 58 Table 2.4 Sequences of oligodeoxyribonucleotides .................................... 63 Table 2.5 Sequences of FL-derived UTRs and ATG-mutated truncated

actin-luciferase constructs........................................................... 64

9

Page 10: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

1. INTRODUCTION Gene expression is the realisation of the information potential encoded in the

genome. While much of its regulation takes place before or at the stage of

transcription, the number and complexity of steps thereafter – mRNA

processing, translation and protein localisation – adds significantly to the

diversity and regulation of gene expression. The scheme in Figure 1.1 outlines

the fate of eukaryotic transcripts following transcription and illustrates that

mRNA is almost continuously processed, protected, transported, translated or

degraded, mostly within specialised cellular compartments. As a consequence,

mRNA is usually not present in the cell without RNA-binding proteins, a fact that

is reflected by the more appropriate use of the term mRNP for the protein-

associated transcript.

The complexity of gene expression in eukaryotes is also mirrored in the

variety of ways that control it; theoretically every step from transcription to

protein degradation can be used to manage the way and extent to which certain

genetic information is utilised. Therefore, in this context and throughout this

study, the characterisation of a step as 'rate limiting' should be approached

carefully since it can be misleading and meaningless. Although some steps

such as initiation of translation feature predominately in the control of

posttranscriptional gene expression, most stages contribute not excessively

unequally to maintain a possibility for balanced regulation at each step along a

pathway (compare also McCarthy, 1998).

Keeping this in mind, this study focuses on a historically somewhat

neglected step in posttranscriptional gene expression, the termination of

translation. Evidence has accumulated that this step is not only the end of

protein synthesis but also a very important regulation site for mRNA-specific

translational efficiency and transcript stability. Besides, very little is known about

the fate of the posttermination ribosome and how it is recycled to start another

round of translation. This study aims to fill this gap by utilising both a structural

approach whereby the mRNA is seen as a circular entity and a functional

assessment of the reinitiation potential of posttermination ribosomes.

10

Page 11: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

1.1 Controlling gene expression in yeast: from transcription termination to translation initiation

1.1.1. Pre-mRNA processing Processing of posttranscriptional pre-mRNA takes place in the eukaryotic

nucleus and bestows transcripts with features important for subsequent

translation, e.g. a 5'-cap structure, an intron-free open reading frame (ORF) and

a precisely defined, polyadenylated 3'-end (Figure 1.1).

Capping

Capping is the addition of GMP to the 5'-triphosphate end of the transcript in the

unusual 5'–5' direction and the subsequent methylation of position N7 of the

transferred GMP. This process is required for cell viability in Saccharomyces

cerevisiae (Shatkin, 1976; Mao et al., 1996) due to its resulting protection of

mRNAs from 5' 3' exonuclease activity (Furuichi et al., 1977; Caponigro &

Parker, 1996). In the nucleus, the m7G cap is bound by the heterodimeric cap-

binding complex CBC, which contains two proteins, CBC20 and CBC80,

respectively. Upon export through the nuclear pore complex, the nuclear

binding proteins are exchanged for the cytoplasmic translation initiation factor

eIF4E (Shatkin & Manley, 2000, for review). In the cytoplasm, cap-binding of

eIF4E is inhibiting recruitment of the decapping enzyme Dcp1p to the

transcript's 5'-end and subsequent mRNA degradation (Vilela et al., 2000;

Ramirez et al., 2002). Recruitment of eIF4E and concurrently of other

translation initiation factors to the cap also enhances translation by promoting

the engagement of the ribosomal subunits with the mRNA, linking thereby

mRNA stability with translation initiation. This is achieved partly by the

interaction of the cap-binding complex with the poly(A)-binding protein (Pab1p

in yeast, PABP in higher eukaryotes; see Section 1.2.5 for details).

A theme continuously revisited in this study is the view that processes

such as pre-mRNA modification are not a sequence of isolated steps, although

they can be separated into distinct, successive stages. As reviewed by

Proudfoot (2000) and Maniatis & Reed (2002), one of the most significant

conceptual shifts in present studies of gene expression is to focus on how

different stages interconnect. For example, RNA-processing factors appear to

substitute the original transcription initiation factors on the transcribing RNA

11

Page 12: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

polymerase II (Pol II). Thus, by 'riding along' with the Pol II elongation complex

as it makes the nascent RNA, they couple transcription with mRNA processing.

For example, capping of transcripts is catalysed by enzymes binding to a C-

terminal domain of Pol II, affecting both 5'-end methylation and transcription

elongation (reviewed by Neugebauer & Roth, 1997). This 'coupling' or 'looping

nuclear membrane

transcription initiation

Pol II

splicing

DNA

promoter

Exon

1

pre-mRNA

Exon

2

Intron

AAUAAA

AGAU

capping

p-R-ppp-m7G

transcription elongation

transcription termination

RNA cleavage

3´-end processing

polyadenylation

cap5´UTR 3´UTRORF

+ An

mRNAcircularisation (?)

localisation (?)

mRNP export

factor bindinghnRNPs, export factors

CBC exchanged for eIF4E

nuclear translation

nonsense-mediated decay (NMD) ][ ?“pioneering round“

eIFs, eEFs, eRFs etc

decay

translation

regulatory factorsRNases

NPC

nuclear membrane

transcription initiation

Pol II

splicing

DNA

promoter

Exon

1

pre-mRNA

Exon

2

Intron

AAUAAA

AGAU

capping

p-R-ppp-m7G

transcription elongation

transcription termination

RNA cleavage

3´-end processing

polyadenylation

cap5´UTR 3´UTRORF

+ Ancap5´UTR 3´UTRORF

+ An

mRNAcircularisation (?)

localisation (?)

mRNP export

factor bindinghnRNPs, export factors

CBC exchanged for eIF4E

nuclear translation

nonsense-mediated decay (NMD) ][ ?“pioneering round“

eIFs, eEFs, eRFs etc

decay

translation

regulatory factorsRNases

NPC

Figure 1.1. The pathway of eukaryotic mRNA from the nucleus to the sites of translationand decay in the cytoplasm. Transcription and mRNA processing in the nucleus are concertedmechanisms, governed by RNA polymerase II. The concept of nuclear translation of theprocessed mRNA and coupled nonsense-mediated decay of the transcript in the nucleus arecurrently under discussion. Modified from McCarthy & Kollmus (1995) and Proudfoot (2000), fordetails see text. pol II = DNA-dependent RNA polymerase II, R = purine (at this position in yeastwith 75% relative frequency an adenine, with 25% a guanine), AG/AU = dinucleotides definingthe exon boundaries, AAUAAA = positioning element for 3' RNA cleavage and polyadenylation,ORF = open reading frame, UTR = untranslated region, CBC = heterodimeric nuclear cap-binding complex, eIF4E = mostly cytoplasmic cap binding eukaryotic translation initiation factor,NPC = nuclear pore complex.

12

Page 13: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

back' of processes will be revisited when the links between translation and

mRNA decay and eventually, translation initiation and termination on a circular

mRNA are discussed.

Splicing

The next step to take place on the nascent transcript is intronic splicing.

However, there are only few intron-containing pre-mRNAs in the budding yeast,

and most of them do have only one small intron near the beginning of the

transcript (Sakurai et al., 2002). They are removed in the nucleus due to the

activity of the spliceosome, a macromolecular machine containing five small

nuclear RNAs and numerous proteins (at least 145 in the human spliceosome)

rivalling the ribosome in its complexity (Zhou et al., 2002). Whether splicing in

S. cerevisiae occurs co-transcriptionally has not yet been shown. However,

there is a tight association of these two processes in mammals (reviewed by

Neugebauer & Roth, 1997), and all posttranscriptional processing events are

influencing each other's efficiency and specificity (reviewed by Proudfoot et al.,

2002).

3'-end processing

The final alteration of nuclear pre-mRNA is 3'-cleavage and addition of a poly(A)

tail (Zhao et al., 1999). As discussed more thoroughly in Section 1.2.5, the

poly(A) tail, by providing a binding site for Pab1p, enhances the translation and

stability of mRNA (Sachs et al., 1997). Like capping and splicing, 3'-cleavage

and polyadenylation factors have also been clearly associated with the C-

terminal domain (CTD) of the transcriptionally elongating Pol II (Barilla et al.,

2001). In yeast, the termination of RNA polymerisation has been shown to

depend not only on a RNA signal but also on 3'-cleavage factors (reviewed by

Neugebauer & Roth, 1997).

1.1.2 Translation in the nucleus? Recently, evidence has accumulated to point at a translation-like scanning of

transcripts in the nucleus. Although indicated as early as 1978 by Goidl and

colleagues, more compelling evidence has been put forward by Fortes et al. in

2000 showing that the nuclear cap binding complex CBC binds to both the cap

structure and translation initiation factor eIF4G, mediating thereby translation

initiation. Moreover, Iborra and colleagues (2001) localised the sites of

13

Page 14: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

translation in permeabilised mammalian cells and found that 10-15% of the

labelled tRNA was incorporated in the nucleic fraction, challenging thereby the

long-established theory of compartmentalisation of transcription and translation

in eukaryotes. Furthermore, their findings indicate that nucleic translation is

coupled to transcription, and that the sites of translation within the nucleus

coincide with patterns of translation initiation factor eIF4E. The results of this

work have also been commented on by Hentze (2001) and, more critically, by

Dahlberg and colleagues (2003). Weaknesses pointed out include that not all

translation-related factors have been found to be present in the nucleus at

sufficient high levels, but some are even actively exported into the cytoplasm

(Bohnsack et al., 2002). Furthermore, since the nascent ribosomes in the

nucleus differ from their cytoplasmic counterparts, they would be expected to

not be translationally active or at least to translate in a different way. Thus, the

current view of translation in the nucleus is ambiguous, and needs further

clarification. However, this model could explain observations of degradation of

transcripts via the nonsense-mediated decay pathway protecting the cell from

the export of aberrant transcript to the sites of quantitative translation in the

cytoplasm. Details of the mRNA surveillance mechanism and the consequences

of these findings are discussed in Section 1.3.2.

1.1.3 Transcript export from the nucleus The next step in mRNA utilisation – translation – requires its transport from the

nucleus to the cytoplasm through the nuclear pore complexes (NPCs, Stoffler et

al., 1999). Prior, mRNPs are bound by heterogeneous nuclear

ribonucleoproteins (hnRNPs), known to shuttle between the two compartments

(Sträßer & Hurt, 1999) although they are not considered to play a direct role in

mRNA export (Zenklusen & Strutz, 2001). Much has been discovered about

nuclear export of mRNAs and its connection with upstream gene expression

steps, yet the complex and diverse routes of mRNA export are still under

investigation (Reed & Hurt, 2002). However, there is clear evidence that nuclear

export is a checkpoint for proper mRNA processing since capping, intron

removal, proper 3'-end processing and polyadenylation are all enhancers of

nuclear mRNP export (Zhao et al., 1999). Following nuclear export, mRNA can

be transported to specific sites with the cells (Farina & Singer, 2002) or even be

dormant for months until needed (Standart & Jackson, 1994).

14

Page 15: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

1.2 Factors and mechanisms in eukaryotic translation Functionally, translation can be divided in all organisms into three phases.

During initiation, the translational apparatus is assembled on the messenger

RNA and recognises the start site. Throughout elongation the actual protein

synthesis takes place, while in the last step, termination, the translation is

ended, the peptide product released and the ribosomal subunits recycled to

start another round of translation. Despite apparent homologies and analogies,

translation in eukaryotes differs fundamentally from the situation in prokaryotes,

e.g. in the way the start codon is recognised, in the size and composition of the

ribosomal subunits, and in the origin and complexity of some translation factors

(McCarthy & Gualerzi, 1990; Pain, 1996; McCarthy, 1998; Kozak, 1999; Preiss

& Hentze, 1999).

1.2.1 Translation initiation Initiation is the first step towards the translation of an mRNA, and due to its

complexity it is also the site of extensive global and transcript-specific regulation

of gene expression (Merrick & Hershey, 1996). Essentially, there are three ways

to initiate translation in eukaryotes: via the 5'-end, by recruiting the ribosome to

an internal ribosome entry site (IRES) and by reinitiation of translation by

posttermination ribosomes.

In the IRES mechanism, translation initiation is mediated by a cis-acting

element that recruits the translational machinery to an internal initiation codon

with the help of trans-acting factors (reviewed in Jackson, 2000, and Martínez-

Salas et al., 2001). While common in mammalian and plant viruses, it has only

been recently shown to work in yeast as well (Paz et al., 1999; Zhou et al.,

2001). Moreover, it should be noted that although most researchers in this field

will support the model of internal initiation, there is also an ongoing discussion

about the authenticity of some of the evidence that may result in the reduction

of proposed IRES elements (Kozak, 1992; Kozak, 2001a; see also comment in

Schneider, 2001). In addition, almost all putative IRES elements were detected

either on highly regulated mRNAs with a particularly structured leaders or when

the 5'-end dependent initiation pathway is disrupted in general, e.g. due to

starvation or virus infection (Lamphear et al., 1995; Iizuka et al., 1995; Jackson,

1996). Given that IRESs are capable of promoting internal initiation of

translation on eukaryotic polycistronic mRNAs, it is also striking that this is

15

Page 16: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

never the case (McCarthy, 1998; Kozak, 1999). Therefore, the main

physiological role of IRES-mediated ribosomal entry is apparently to render the

translation of specific ORFs independent of the almost exclusive main initiation

pathway, the recruitment via the 5'-end as in the case of a viral infection of the

cell.

1.2.1.1 Recruitment of the 43S preinitiation complex to the 5'-cap By far the most common route to initiate translation in eukaryotes is the

recruitment of the ribosome to the 5'-cap structure. This is afforded by a bridge

consisting of the cap structure, the cap-binding complex eIF4F, eIF3 and the

40S subunit. Figure 1.2 shows a consensus model since different models for

the order of assembling are currently under discussion (Sachs et al., 1997,

Gingras et al., 1999). The recent report of a reconstituted yeast translation

initiation system might aid in this effort (Algire et al., 2002). In a first step

however, the ribosomal subunits that form under physiological conditions

predominantly 80S complexes have to be separated into 40S and 60S subunits.

This happens under the influence of the factors eIF1A and eIF3, which bind to

the smaller subunit, and eIF6 that binds to the 60S subunit, thereby

sequestering the 80S complexes (Chaudhuri et al., 1999; Si & Maitra, 1999).

The smaller subunit is then charged with the initiator tRNA through binding of an

eIF2-GTP-Met-tRNAi ternary complex (Hinnebusch, 2000). The complete

complex, comprising the 40S subunit stabilised by eIF1A and eIF3 and bound to

the eIF2 ternary complex, is then termed the 43S preinitiation complex (Figure

1.2). Specificity for the mRNA 5'-end is acquired by the group 4 initiation factors,

which are also assumed to facilitate ribosomal binding through removal of

secondary structure on the RNA. Though biochemical evidence for the

described interaction has been obtained, yet again the order of 43S preinitiaion

complex assembly and mRNA recruitment is not fully resolved (Trachsel, 1996).

Certainly, a close association of the 40S subunit and mRNA before the (re-)

binding of the ternary complex can occur as implied by the current model of

reinitiation of translation (see Section 1.4).

eIF4E is the cap-binding component of the initiation factor complex eIF4F

enlisting the largest subunit, eIF4G, a multipurpose adapter capable of

recruiting a number of activities to the 5'-end. The binding of eIF4G to the

ribosome-associated initiation factor eIF3 is thought to establish physical

16

Page 17: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

contact between the mRNA and the 40S ribosomal subunit, whereas binding to

the RNA helicase eIF4A by eIF4G was proposed to be necessary for the

disruption of secondary structures in the 5' untranslated region (Figure 1.2). The

helicase activity of eIF4A is strongly stimulated by the non-essential eIF4B,

however there is no evidence that the proteins physically interact in a stable

manner (Gingras et al., 1999, for review). With the recent finding that eIF4A

associates with eIF4G in yeast (Neff & Sachs, 1999; Dominguez et al., 1999), it

is now agreed that the cap-binding complex eIF4F comprises the translation

initiation factors eIF4E, eIF4G and eIF4A in eukaryotes from mammals to plants

+ An

40S

ternary complex

eIF4F·mRNA

AUGPab1AAAn

40S ribosomal subuniteIF3eIF1A

Met-tRNAi

mRNA

eIF4E·4E-BP

Met-tRNAi·eIF2·GTP

60S

– GTPeIF3 eIF1AeIF2

eIF2·GTP

GTP GDP

eIF2·GDP eIF2B

eIF4E

4E-BP·PO4- eIF4G

eIF4F

40S (Met-tRNAi)– GTPeIF3 eIF1A

eIF2

43S preinitiation complex

48S initiationcomplex

eIF4E

eIF4G

eIF4AeIF4B

scanning

ADP ATP interaction with eIF4G

40SAUG

GTPeIF3 eIF1A

eIF2

eIF4E

eIF4G

?

eIF1A, eIF3,eIF2-GDP+P (?)

eIF540SAUG

polypeptide synthesis initiation

eEFstRNA ( )amino acids ( )

AUG + An

EAP-eEF1A-GTP

eEF2

ATP

ADP+PO4-

eEF1B

GTP GDP+PO4-

UAA

eRFs, Upf1-3p

eEFsAUG

H2Oelongation

ribosome recycling reinitiation

Upf1Upf3 Upf2

Pab1eRF3

eRF1

mRNA decay (NMD)

GTP

40S recruitment

termination

?

+ An

40S40S

ternary complex

eIF4F·mRNA

AUGPab1AAAn

40S ribosomal subuniteIF3eIF1A

Met-tRNAi

mRNA

eIF4E·4E-BP

Met-tRNAi·eIF2·GTP

60S60S

– GTPeIF3eIF3 eIF1AeIF1AeIF2eIF2

eIF2·GTP

GTP GDP

eIF2·GDP eIF2B

eIF4E

4E-BP·PO4- eIF4G

eIF4F

40S40S (Met-tRNAi)– GTPeIF3eIF3 eIF1AeIF1A

eIF2eIF2

43S preinitiation complex

48S initiationcomplex

eIF4E

eIF4GeIF4G

eIF4AeIF4B

scanning

ADP ATP interaction with eIF4G

40SAUG

GTPeIF3eIF3 eIF1AeIF1A

eIF2eIF2

eIF4E

eIF4GeIF4G

?

eIF1A, eIF3,eIF2-GDP+P (?)

eIF540SAUG

polypeptide synthesis initiation

eEFstRNA ( )amino acids ( )

AUG + An

EAP-eEF1A-GTP

eEF2

ATP

ADP+PO4-

eEF1B

GTP GDP+PO4-

UAA

eRFs, Upf1-3p

eEFsAUG

H2Oelongation

ribosome recycling reinitiation

Upf1Upf1Upf3Upf3 Upf2Upf2

Pab1Pab1eRF3eRF3

eRF1

eRF1

mRNA decay (NMD)

GTP

40S recruitment

termination

?

Figure 1.2. Steps of translation in Saccharomyces cerevisae. Recycling of eIF2 andformation of ternary complex, competition between eIF4E-BP and eIF4G for eIF4E-binding, andthe formation of the cap-binding complex are outlined. Recruitment of eIF4F via the Pab1 –eIF4G – eIF3 interaction is an alternative route for ribosomal mRNA association. Duringscanning, the helicase/annealing activity of eIF4A/4B (recruited by eIF4G) is thought to resolveRNA structures such as the depicted 5'-UTR stem-loop. The order of 43S preinitiation complexformation and the fate of the eIF4F complex after 40S binding are still under discussion. Studiesof reinitiation revealed that the release and recycling of at least initiation factor eIF2 occursgradually during translation elongation. Modified from McCarthy (1998), see text for details.eIF4E-BP = eIF4E binding protein (p20 in S. cerevisiae), PO4- = inorganic phosphate.

17

Page 18: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

and yeast (Gingras et al., 1999). Initiation factor eIF4G also contains a binding

site for Pab1p, thereby enabling the poly(A) tail to promote ribosome

recruitment to the mRNA's 5'-end independent from the cap structure.

Moreover, 5' and 3'-ends act synergistically to enhance translation rates, the

mechanism of which is thought to be mRNA circularisation due to the eIF4G –

Pab1p interaction (see Section 1.2.5 for details).

In addition to the translation initiation factors, a number of small proteins

compete with eIF4G for binding to eIF4E. Thereby, these eIF4E-binding

proteins (eIF4E-BPs) can regulate the formation of the eIF4E – eIF4G complex;

the following disruption of the cap-binding complex was observed to result in the

inhibition of cap-dependent, but not of internal initiation of translation. In turn,

phosphorylation (regulated via a number of signal transduction pathways) of

these eIF4E-BPs control their ability to bind to eIF4E (Figure 1.2; Proud, 2002,

for review). In S. cerevisiae, a small protein, p20 with an actual size of 18 kDa

co-elutes with cap-complex preparations (Zanchin & McCarthy, 1995). Its role

remains unclear since deletion of the p20 gene is not lethal and does not affect,

or only slightly increases, the growth rate in rich medium under standard

laboratory conditions (de la Cruz et al., 1997). Recent findings point to a role as

a fine regulator of protein synthesis, since p20 binds to a eIF4E site that is

partially shared by eIF4G, thereby modulating the cooperative binding of eIF4F

to the cap (Ptushkina et al., 1998).

1.2.1.2 Selection of the translational start site and peptide synthesis initiation

The prokaryotic ribosome can base pair with sequence elements such as the

Shine-Dalgarno sequence directly, locating thereby the start codon at any

position on the mRNA (Steitz & Jakes, 1975). In contrast, the eukaryotic 40S

subunit, with its associated cohort of initiation factors, is thought to traverse the

mRNA 5'-UTR from the cap in a linear and processive fashion in a 5' to 3'

direction (Figure 1.2; Kozak, 1977; Kozak, 1978; Kozak and Shatkin, 1978;

Kozak, 1989b; Kozak, 2002). This process termed 'scanning' continues until an

AUG in a favourable start codon context is recognised by the Met-tRNAi bound

to the 40S subunit in the 48S initiation complex (Figure 1.2). Initiation factor

eIF5 mediates the joining of the 60S subunit to the 48S initiation complex and

the release of most other factors. eIF5, together with eIF2, also controls the

18

Page 19: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

stringency of AUG selection, which is marked by the hydrolysis of eIF2-bound

GTP (Huang et al., 1997). The dissociation of eIF2, complexed with GDP, and

the recycling by eIF2B will be discussed together with reinitiation of translation

in Section 1.4.2. Finally, eIF5A is involved in formation of the first peptide bond,

thus linking translation initiation and elongation. These events have been

described in detail in several reviews. (McCarthy, 1998; Hershey & Merrick,

2000; Pestova et al., 2001).

Evidence for the scanning mechanism comes from many directions: The

model was first proposed when edeine-treated cells showed a lack of AUG

recognition and subsequently, almost exclusive association of mRNAs with 40S

ribosomal subunits (Kozak & Shatkin, 1978). Furthermore, a structure inserted

between the cap and the AUG codon was shown to effectively interrupt

scanning and prevent translation initiation with the mRNA co-sedimenting with

the 40S fraction (Kozak, 1986; Kozak 1989a). Findings that the cap structure

promotes translation, and that polycistronic mRNAs are almost completely

absent in eukaryotes support the scanning hypothesis (Kozak, 1989b).

Moreover, the ATP-dependence of in vitro scanning indicates that the RNA-

dependent ATPase eIF4A, recruited by the eIF4F complex, might function as

both a helicase to resolve secondary structures and a hypothetical clamp

(Kozak, 1980; Rozen et al., 1990). This was confirmed by the revelation of the

eIF4A crystal structure; the helicase core, i.e., the sequence comprising all

conserved motifs, is composed of two domains forming a cleft in which the

nucleotide will be located (Caruthers et al., 2000).

Yet, despite the overwhelming evidence for 5'-UTR scanning prior

translation initiation, the underlying mechanism still remains to be elucidated. A

main critique point is that movement of the 40S subunit alone on RNA has so

far never been experimentally demonstrated but would constitute a new

property for this macromolecule since elongation of translation is due to the

interplay of both ribosomal subunits (Londei, 1998). The only serious alternative

model for the recognition of the start codon that differs from the linear scanning

mechanism is the utilisation of an internal ribosome entry site (IRES). The

mechanisms, transcript sequences and structural features described in Section

1.2.4 can attenuate steps of translation initiation in yeast. However, they are

mainly based on predicted deviations from the scanning model but do not

19

Page 20: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

provide experimental evidence contradicting the general mechanism of 5'-UTR

scanning and AUG recognition.

1.2.2 Elongation In the process following translation initiation, the coding potential of the ORF is

realised and converted into a peptide sequence. While the ribosome moves

5' 3' in relation to the mRNA, amino acids delivered by aminoacylated tRNAs

are joined by the formation of peptidyl bonds producing a polypeptide (Figure

1.2). This energy-consuming event is promoted by eukaryotic elongation factors

(eEFs) that have been identified in their analogue (and partly homologue)

similarity to the biochemical mechanism of their prokaryotic counterparts. GTP

and aminoacyl-tRNA are recruited to the A-site of the ribosome by the highly

abundant eEF1A, homologue to the prokaryotic EF1A (formerly known as EF-

Tu). Recognition of the codon by the charged tRNA involves the hydrolysis of

GTP bound to the ternary complex; the eEF1A-GDP complex in turn is recycled

by eEF1B (analogous but dissimilar to the prokaryotic EF1B, previously called

EF-Ts). After transfer of the nascent peptide chain onto the aminoacyl-tRNA

localised in the A-site, another G-protein, eEF2 (analogous to EF2 or EF-G,

respectively), is responsible for the translocation of the charged tRNA into the

P-site and the transport of the deacylated tRNA in the E-site. The next round

starts when the A-site is loaded again with a ternary complex. Yeast and fungi

differ from other eukaryotes due to their dependence upon eukaryotic

elongation factor 3 (eEF3, Qin et al., 1990; Triana et al., 1993). eEF3 is

essential in vivo and required for each cycle of the translation elongation

process in vitro. Models predict eEF3 affects the delivery of cognate aminoacyl-

tRNA, a function performed by eEF1A, by removing deacylated tRNA from the

ribosomal exit (E) site (Anand et al., 2003).

The elongation rate is mainly determined by the amount of aminoacylated

tRNAs available as well as by the activity of various elongation factors. Because

of its processivity – the succession of equal steps – the elongation phase can

be viewed as a single entity. Any change in the rate of elongation would

therefore only alter the ribosome density on the ORF but not impose a rate

limitation on translation in general. An exception to this might be situations

whereby slowed ribosomes 'queue' to terminate or (re)initiate translation. For

example, pseudoknots are thought to delay elongating ribosomes and enable

20

Page 21: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

the ribosome to perform a -1 frameshift at a slippery site, e.g. in yeast double

stranded RNA viruses (Plant et al., 2003). However, Kozak (2001b) showed that

a pseudoknot, although inhibitory to scanning or initiating subunits, can be

readily resolved by the elongating ribosomes causing no decrease in translation

rates.

Recently, the complete three-dimensional structure of the prokaryotic 70S

ribosome and the 40 and 60S eukaryotic subunits has provided new

opportunities to study details of the elongation (see Ramakrishnan, 2002, for

review). Progress has since made to elucidate how translocation works

(Joseph, 2003) and how the nascent peptide can regulate translation (Tenson &

Ehrenberg, 2002). It can be expected that the three-dimensional reference

given by X-ray crystallography and cryo electron microscopy will help to

interpret the biochemical and genetic data as well as to design new

experiments.

1.2.3 Termination In eukaryotes, an inframe stop codon translocated into the ribosomal A-site is

identified following the association of ribosomes with the heterodimeric

polypeptide chain release factor (eRF) comprising eRF1 and eRF3 (Stansfield

et al., 1995; Zhouravleva et al., 1995). eRF1 recognises all three stop codons –

UAA, UAG and UGA – and mediates the hydrolysis of the peptide from the

tRNA in the P site (Figure 1.2, Frolova et al., 1994). eRF3, a GTPase with

homology to eEF1A, stimulates this reactions in a GTP-dependent, codon-

independent manner (Frolova et al., 1996). The mechanism envisaged currently

implies that eRF3 associates with the small ribosomal subunit and recruits

eRF1. The formation of the ternary complex ribosome-eRF3-eRF1 then triggers

GTPase activity upon stop codon recognition (Stansfield et al., 1995).

1.2.3.1 eRF1: tRNA mimicry and stop codon recognition Research on eRF1 has recently focused on the functional anatomy of the factor

as revealed by a variety of techniques (Kisselev et al., 2003, and references

therein). Alignment of structure and function has resulted in the assignment of a

ribosome-binding site, a termination codon recognition site, a peptidyl-tRNA

interaction site and an eRF3-binding site (Frolova et al., 2000). A functionally

active 'core' appears to mimic the structure of a tRNA providing thereby stop

21

Page 22: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

codon specificity in a direct protein-RNA interaction (Song et al., 2000; Bertram

et al., 2000). Contact of eRF1 with both ribosomal subunits would stabilise and

modulate complex formation and stereochemistry (Kisselev et al., 2003). eRF1

shares conserved sequence elements but no extensive homologies to the

prokaryotic counterparts RF1 and RF2. In addition, eRF1 can recognise all

three stop codons whereas RF1 terminates translation at UAG, RF2 at UGA,

and UAA is recognised by both factors (Scolnik et al., 1968). This is in

accordance with observations e.g. that yeast mutants showed an omnipotent

suppression phenotype. It was also shown that eRF1 alone can promote the

release of the elongating polypeptide from the ribosomal P-site when a stop

codon is present in the A-site in vitro (Frolova et al., 1994; Zhouravleva et al.,

1995). However, addition of eRF3 alone blocks this reaction but in turn

enhances the reactivity when it associates with GTP (Zhouravleva et al., 1995).

Stansfield and colleagues (1995) clearly showed that an over-expression of

both eRF1 and eRF3, encoded in S. cerevisiae by SUP45 and SUP35, is

necessary to enhance the efficiency of termination, and the interaction of both

factors appears to ensure to proper recognition of stop codons in vivo.

Yet, the stop codon can also be recognised as an alternative signal, i.e.,

for frame-shifting, read-through or selenocysteine incorporation, referred to as

'recoding' for reprogrammed genetic decoding (reviewed in Gesteland & Atkins,

1996). Recoding requires several regulatory elements to subvert the normal

stop codon recognition, in competition with release factors. Although the only

alternative decoding of triplets in S. cerevisiae occurs in mitochondria (Jukes &

Osawa, 1990) and the incorporation of selenocysteine has not described in this

yeast (Krol, 2002), read-through (Bonetti et al., 1995), +1 and -1 frameshifting

(Farabaugh, 1996) have been observed.

Efficiency of termination is effectively governed by the stop codon context

and variations are deliberately used in animal and plant viruses to facilitate

read-through (Gesteland et al. 1992) or even to arrest ribosomes in a coding

sequence dependent manner (Cao & Geballe, 1996). The use of the three stop

codons in S. cerevisiae is strongly biased with UAA (53.1%) > UGA (26.8%) >

UAG (20.2%). Analyses also revealed a function for the fourth base; the

frequency is then UAAA (18.2%), UAAG (13.5%), UAAU (16.6%), UGAA

(9.9%), UGAU (9.6%), and UAGA (8.4%) (Tate et al., 1996). When only highly

22

Page 23: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

expressed mRNAs were assayed, the bias was even stronger, termination of

translation therefore represents a site for posttranscriptional control. Due to its

coding potential for the nascent peptide, the influences of upstream sequences

are more difficult to assess. Indeed, Mottagui-Tabar et al. (1998) report that

both the P-site tRNA and the -2 amino acid have an influence on the efficiency

of termination. They also observed that basic -2 residues were associated with

higher read-through than acidic ones, which is the reverse to the situation in E.

coli. Bonetti et al. (1995) assumed a similar role for the C-terminal codon. It is

noteworthy that stop codon context and translation termination efficiency affect

the fate of the posttermination ribosome governing the choice between

ribosomal release and reinitiation, which is of special interest when discussing

mRNAs comprising upstream ORFs in Section 1.4.

1.2.3.2 eRF3: from little known eRF1-binding partner to hub for virtually all translation termination events

Deletion of eRF3 is lethal in eukaryotes whereas the only functionally analogue

RF3 in prokaryotes is neither essential nor does it bind to the eRF1

counterparts, RF1 and RF2 (for references see Kisselev and Buckingham,

2000). This points to a more complex role for eRF3 than just modulating eRF1

activity in a GTP-dependent manner. Indeed, eRF3 appears to comprise also

the function of a fourth prokaryotic factor, the essential ribosome recycling

factor RRF. RRF, together with elongation factor EF-2 (EF-G) disassembles the

terminating complex by promoting the translocation of the posttermination

ribosome in prokaryotes. The release of deaminoacylated tRNA is followed by

mRNA and ribosomal subunit dissociation (Hiroshima & Kaji, 1972; Hiroshima &

Kaji, 1973; Hirokowa et al., 2002; see Kaji & Hirokawa, 2000, for review). The

only eukaryotic RRF-like proteins identified so far comprise a mitochondrial

targeting signal (Zhang & Spremulli, 1998). Hence, it can be suspected that

recycling of ribosomes in eukaryotes is a property of the eRF1/eRF3 complex,

since they suffice to perform all the functions of prokaryotic RFs. Alternatively,

Buckingham et al. (1997) proposed that eRF3 alone can recycle both eRFs and

ribosomes, but the biochemical, genetic or structural evidence for this is still

missing.

Another very intriguing recent finding was the elucidation of the so-called

extrachromosomal determinant [PSI+], which was first described in S. cerevisiae

23

Page 24: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

as a modifier of nonsense suppression (Cox, 1965). It was found that the [PSI+]

factor is a product of the SUP35 gene encoding eRF3 (Chernoff et al., 1993;

Doel et al., 1994; Ter-Avanesyan et al., 1994). As well as folding into its native

structure, eRF3 is believed to be capable to adopt a second aberrant

conformation, which manifests as a prion-associated phenotype (Chernoff et al.,

1995; Tuite & Lindquist, 1996). In [PSI+] strains, eRF3 is present both as a

soluble factor and as large intracellular aggregates, resulting from the N-

terminal conferred propensity of the prion conformer to coalesce. Due to their

binding sites for eRF3, eRF1 and Upf1p (see below) could also be sequestered

in the eRF3 particles resulting in a concomitant depletion of these three factors

(Paushkin et al., 1997; Czaplinski et al., 1998). Since overexpression of eRF1

does not relieve the allo-suppressor phenotype of a [PSI+] strain but rather

enhances it, it can be assumed that eRF3 diminution itself is responsible for the

observed inefficient termination and elevated read-through (Derkatch et al.,

1998). The physiological significance of the [PSI+] phenotype appears to be in

the proposed synthesis of normally not produced proteins, i.e., translation of

mRNAs containing a nonsense codon within their reading frame or by-passing a

native stop codon to produce extended proteins (Moffat et al., 1994).

Phenotypically, [PSI+] confers upon cells resistance to heat and ethanol stress

(Eaglestone et al., 1999) and, indeed, at least one heat shock transcription

factor comprising a nonsense codon was found to be synthesised in these cells

(Lindquist et al., 1995). Moreover, the negative effects of uORFs in the 5'-UTR

of stress-related transcription factors such as GCN4, YAP1 and YAP2 might be

overcome by the elevated read-through of termination codons of these uORFs.

Some novel reports have shed some additional light onto the role of eRF3,

although the number of discovered interactions is somewhat staggering

(Mugnier & Tuite, 1999). For example, the finding that the nonsense-mediated

decay factor Upf1p (for up-frameshift) comprises a binding site for both eRF1

and eRF3 links stop codon recognition to mRNA decay (Czaplinski et al., 1998).

The consequences of this tight coupling between translation termination and the

degradation of aberrant transcripts will be discussed in Section 1.3.2. In addition

to binding sites for the ribosome, eRF1 and Upf1p, eRF3 can also associate

with poly(A)-binding protein (PABP in mammals, Hoshino et al., 1999; Pab1p in

yeast, Cosson et al., 2002). The effects on mRNA circularisation, translation

24

Page 25: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

initiation and poly(A) tail conferred mRNA stability will be examined in detail in

Section 1.2.5. Two more proteins, Mtt1p (Czaplinski et al., 2000) and Ittp1

(Urakov et al., 2001) were found to interact with both release factors and

modulate translation termination. Their relationship to transcription factors or

DNA-dependent helicases indicates a coupling of termination with other cellular

processes such as transcription and DNA replication, but these speculations

need further clarification. To add to this growing list, there are also recent

reports of an involvement of the release factors in cytoskeleton organisation and

cell cycle regulation (Valouev et al., 2002).

Hence, eRF3 obviously plays many roles in termination: GTP-dependent

interaction with the ribosome and eRF1, proofreading function, ribosome

recycling and binding-site for Pab1p, Upf1p and other proteins. This reinforces

the status of translation termination as an important control site coupling,

among others, proper stop codon recognition, efficiency of translation

termination, mRNA circularisation, mRNA decay, ribosome release and

reinitiation of translation.

1.2.4 Elements in the 5'-UTR affecting translational efficiency

1.2.4.1 AUG context Initiation of translation in eukaryotes is usually restricted to the AUG start

codon, and the surrounding sequence affects the efficiency of its recognition. In

higher eukaryotes, derivation from the ideal consensus sequence, 5'-

GCC(A/G)CCAUGG-3', leads to significantly decreased translation initiation

efficiency, especially when the two underlined bases (a purine in position -3 and

a G in position +4) are mutated (Kozak, 1987b). However, the effects of

changes are considerably smaller in S. cerevisiae than in vertebrate systems.

An analysis of 131 genes found the bias in nucleotide distribution to be 5'-

A(A/U)AAUAAUGUCU-3' (Cigan and Donahue, 1987). More recently, an index

(AUGCAI) to measure optimal AUG context has been established by Miyasaka

(1999) stating the slightly different optimal consensus sequence 5'-

(A/U)A(A/C)AA(A/C)AUGUC(U/C)-3'. A change in the most important position -3

(underlined) results in only a two-fold decrease of initiation efficiency (Cigan et

al., 1988). It is most likely due to this insensitivity that start codons other than

25

Page 26: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

AUG are hardly recognised in S. cerevisiae (Clements et al., 1988; Huang et al.,

1997; Nett et al., 2001).

AUG priority rule

The main evidence for an unidirectional 5' 3' movement of the scanning

complex derives from the observation that the first AUG is almost exclusively

utilised for initiation of translation (Kozak, 1999). Furthermore, even the

numerous cases where the first AUG does not constitute the site of translation

initiation can be explained based on the scanning model.

Leaky scanning

Non-recognition of start codons leads to 'leaky scanning' whereby the ribosomal

unstructured leader

structured leader

reinitiation

leaky scanning

shunting

IRES

An

An

An

An

An

An

IRES

unstructured leader

structured leader

reinitiation

leaky scanning

shunting

IRES

AnAn

AnAn

AnAn

AnAn

An

AnAn

IRES

Figure 1.3. Mechanisms of eukaryotic start codon recognition. Comparison of AUGidentifications mediated by scanning of the 5'-UTR or by recruitment via an internal ribosomeentry sites (IRES). An unstructured leader poses no hindrance to scanning 40S ribosomalsubunit while a sufficiently stable secondary structure can block 5'-UTR scanning. Under certaincircumstances, ribosomes can reinitiate translation having terminated at an upstream ORF(uORF), albeit with usually lower efficiency. In 'leaky scanning' an uORF can be bypassed whenits start codon is in an unfavourable sequence context, the ribosome then initiates at adownstream AUG. The translation of an uORF can also lead to negotiation of an otherwiseimpeding secondary structure by non-linear ribosome 'shunting' (bent arrow), followed byscanning to reach the downstream ORF. Internal initiation of translation does not involvescanning but recruitment of the ribosome via an IRES element to mRNAs that predominantlycomprise highly structured leaders. The 40S ribosomal subunits are depicted as spheres at sitesof translation initiation or scanning blockage. The 5' 3' scanning movement is indicated bystraight arrows, secondary structures are shown as stem-loops, ORFs as boxes.

26

Page 27: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

subunit continues to proceed along the mRNA until it initiates at a start codon

farther downstream (Figure 1.3; Kozak, 1991c and 1999). This mechanism is

widely used in mammalian cells and viruses to produce two different proteins

from the same mRNA, indicating that this mechanism is not 'sloppiness' on the

part of the translational apparatus but employed deliberately (Kozak, 2002, for a

review). The utilisation of start codons within an unfavourable codex can be

enhanced when a downstream secondary structure slows scanning and thereby

provides more time for codon / anticodon recognition. This suppression of leaky

scanning requires a critical distance of 13 –15 nt (corresponding to half the

diameter of a ribosome) between the start codon and the structure (Kozak,

1990).

1.2.4.2 Leader length Recognition of the first AUG requires a certain distance between the mRNA's 5'-

end and the start codon. In in vitro experiments, translation initiation at an AUG

located up to 12 nt from the cap was clearly impaired while placing the start

codon 20 nt downstream of the 5'-end enhanced the recognition rate greatly

(Kozak, 1991a; Kozak, 1991b; Kozak, 1991d for review). One explanation

offered for this is that 80S ribosomes poised at the start site prevent 40S

subunits from binding to the 5'-end, implying that the initiation step might be rate

limiting. Alternatively, a certain leader length might be required to facilitate

contact between the mRNA and the 40S subunit. Observations of translation

rates suggest that lengthening an unstructured leader up to 80 nt has a

proportionally stimulating effect on initiation (1991b). Further extension up to

1,800 nt does see only a moderate influence on the translational efficiency in S.

cerevisiae cells whereas in cell lysates translation was reduced to 1-2%

indicating different initiation or scanning requirements in these systems (Karine

Berthelot et al., private communications). Taken together, these results point at

a certain 5'-UTR length that is 'saturated' with scanning 40S subunits, delivering

them in vivo effectively and without great lost to the site of initiation.

1.2.4.3 Upstream AUGs and ORFs A sizeable group of natural mRNAs have upstream AUGs (uAUGs) or complete

upstream open reading frames (uORFs) in their leader sequence (Cigan and

Donahue, 1987). In accordance with the scanning model, initiation at these

27

Page 28: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

upstream sites reduces translation of downstream ORFs, albeit to varying

degrees. The two possibilities for downstream initiation in these cases are leaky

scanning, where a start codon is not recognised or reinitiation, where each start

codon is duly recognised and several ORFs are translated per mRNA (Figure

1.3). With reference to the scanning mechanism and the first-AUG rule, it

should be noted that posttermination ribosomes can scan backwards to only a

very short extent to reinitiate translation upstream of the termination site (Kozak,

2001b). Therefore, uAUGs that are not followed shortly by an in-frame stop

codon effectively inhibit (re)initiation at downstream start codons. The

phenomenon of uORFs and translation reinitiation in eukaryotes is discussed in

context with ribosome release in detail in Section 1.4.

1.2.4.4 Secondary structures and mRNA-protein interactions

5'-UTR structures can block initiation

The ability of secondary structures located 5' of start codons to be an effective

means to restrict translation initiation on a given mRNA affirms the evidence for

the 'scanning models' (Kozak, 1986; Kozak, 1989b; see Kozak, 1991d and

2002, for review). The strength of the inhibitory effects is apparently linked to

the thermodynamic stability of the structure, but there is an obvious difference

between eukaryotic systems towards the sensitivity of scanning. An estimated

folding stability of approximately -50 kcal mol-1 is required in mammalian cells to

inhibit translation by about 90% (Kozak, 1989a). In contrast, the yeast

translation apparatus shows a greater sensitivity to structured leaders (as

discussed above), stem-loops with a stability of -18 kcal mol-1 suffice to achieve

a comparable reduction (Oliveira et al., 1993b; Saggliocco et al., 1993).

Obviously to avoid longer, and therefore potentially structured leaders, in yeast

the length of the 5'-UTR sequence is usually limited to less than 100

nucleotides (Merrick, 1992). Conversely, those mRNAs that contain extended,

strongly structured leaders have been suggested to be translationally restricted

and regulated (Kozak, 1986). Moreover, the preference for adenines in the 5'-

UTR of baker's yeast (Cavener & Ray, 1991) and their reduced ability to form

stable secondary structures (Cigan & Donahue, 1987) might also reflect an

increased sensitivity of its scanning apparatus. Interestingly, in addition to the

thermodynamic stability of stem-loop structures, their effect on scanning

28

Page 29: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

inhibition increases proportionally with their G+C content (Vega Laso et al.,

1993). However, this should not be confused with impeding guanine-stretches,

whereby a sequence of about 18 Gs located in the 5'-UTR can efficiently

decrease translation and mRNA degradation (Muhlrad et al., 1995) due to

guanosine base stacking – a different mechanism (Williamson et al., 1989;

Zimmerman et al., 1975).

Mammalian systems also show a position-dependence of translational

regulation via secondary structures in the 5'-UTR. In rabbit reticulocyte lysates,

a cap-proximal stem-loop with a stability of -30 kcal mol-1 was found to be

inhibitorier than an equivalent structure placed 53 nucleotides further

downstream. This effect can be explained if the cap-proximal structure

interferes with 40S-mRNA association whereas the downstream stem-loop

would block 40S scanning driven by a thermodynamic force large enough to

unwind the structure due to the ATP-consuming activity of initiation factors

eIF4A/4B. In contrast, there is no cap-proximal potentation effect in S.

cerevisiae (Vega Laso et al., 1993; Koloteva et al., 1997, Kozak, 1991d), so that

variations in the position of a 5'-UTR stem-loop hardly alter the effect. One

suggestion put forward by Vega Laso and co-authors (1993) implies that the

focus of rate control in yeast differs from mammalian systems so that the effect

on the scanning mechanism is equivalent to the influence on ribosome binding

to the 5'-cap. These authors also found that secondary structures in in vitro

systems impose a more severe inhibition to scanning ribosomes than they do in

vivo, while simultaneously the position-dependence was lost in cell lysates.

Therefore, it is difficult to postulate a consensus model for the effects of stem-

loops in different eukaryotic systems. However, it should be noted that Kozak

(2002) also suggests that the differences monitored in the in vitro system might

be due to unconventional transcript starts producing mRNAs lacking the

structure or because of different effects on mRNA stability. These issues can be

dealt with by avoiding in vitro systems or including RNA length and stability

controls.

Ribosome shunting

The model of ribosome shunting describes a mechanism whereby major parts

of a complex leader are bypassed and not melted by scanning ribosomes

(Figure 1.3; Ryabova et al., 2002). This strategy is mostly employed by viruses,

29

Page 30: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

enabling the viral RNA to form extensive stem loops that are excluded from the

scanning process. These structures can comprise non-recognised AUGs or

even binding sites for viral proteins (Hemmings-Mieszczak & Hohn, 1999). The

negotiation of this formation usually requires a full cycle of translation on a short

upstream open reading frame (uORF) followed by non-linear ribosome

migration (ribosome shunt) across the structure. The main ORF is then

identified by subsequent linear scanning in a reinitiation-like mechanism (Figure

1.3; Hemmings-Mieszczak et al., 2000). However, these exceptions to the linear

scanning mechanism do not concern the translation of the majority of cellular

mRNAs, but are means to regulate translation initiation on viral mRNAs or on

transcripts encoding tightly controlled proteins such as growth factors or

components of the gene expression apparatus (Kozak, 2001a; Ryabova et al.,

2002). Although linear scanning represents the usual mechanism, the

eukaryotic translation apparatus obviously possesses the flexibility to combine

reinitiation with a short dissociation from the mRNA.

Secondary structures within the ORF

As mentioned above, a surprising positive effect of secondary structures on

translation was reported when these were situated immediately downstream of

an AUG. In this case, the structure seems to slow the scanning movement of

the ribosome, so that a given AUG is recognised more efficiently, even if the

context is normally unfavourable for initiation (Kozak, 1991). However, if the

structure was too stable, e.g. in the case of a pseudoknot, the scanning

ribosome could not target the start codon and translation initiation was impaired.

In contrast, the elongating ribosome has been shown to readily overcome

secondary structures located in the ORF. Kozak (2001) reported no obvious

effect when placing a pseudoknot derived from viral mRNA in the coding

domain of a CAT transcript and assaying its expression in rabbit reticulocyte

lysate. In the original viral transcript, the pseudoknot was shown to cause

ribosomal pausing during elongating, followed by frame-shifting (Somogyi et al.,

1993). Thus, the scanning ribosomal 40S subunit and the elongating 80S

ribosome differ fundamentally in their ability to resolve structured RNA during

their movement along the transcript.

30

Page 31: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

The IRE-IRP1 system

In contrast to the constitutive limitation of translation initiation imposed by

intramolecular RNA-RNA interaction such as stem-loops, the inducible

synthesis or activation of an mRNA-binding repressor protein allows negative

regulation within a defined period. For example, the human spliceosomal

protein U1A and the bacteriophage MS2 coat protein confer translational

repression on mRNAs harbouring the respective recognition site in their leader

sequence (Stripecke et al., 1994). However, the best-characterised system of

this kind in eukaryotes is based on the binding of iron-regulatory proteins (IRP1

and IRP2) to an iron responsive element (IRE) in the 5'-UTRs of mRNAs

encoding ferritin and erythroid 5-aminolevulinic acid synthetase in vertebrate

cells (Aziz & Munro, 1987; Butt et al., 1996; Rouault and Harford, 2000).

Binding is sufficient to block translation of these mRNAs, but this effect is

relieved when iron levels exceed the requirements in the cell (Rouault et al.,

1988). Oliveira and colleagues (1993a) showed that binding of IRP1 to an IRE-

containing leader results in translational repression in S. cerevisiae,

demonstrating thereby that the regulation requires no other mammalian

components other than IRP1 and IRE to function.

Due to the high affinity of IRP1 to IRE (KD = 10-10 to 10-11 M, Oliveira et al.,

1993a) this system compares easily with the stability of stem loops. However,

the upper stem-loop sequence of the IRE element that suffices for IRP1-binding

is not very stable itself (∆G -7.3 kcal mol-1, Oliveira et al., 1993b). Hence, it

imposes almost no constraints on scanning or elongating ribosomes (compare

Oliveira et al., 1993b; Vega Laso et al., 1993) and can therefore serve as an

ideal negative control in experiment investigating the accessibility of mRNA

UTRs. Similar to Kozak's observation with stable stem loops (Kozak, 1986;

Kozak, 1989a), Koloteva and colleagues (1997) observed a strong inhibition of

translation upon IRP1-binding in yeast cells to a construct with the IRE in the 5'-

UTR. In the original construct (IRE-wt) the stem-loop was 9 nt downstream of

the mRNA's 5'-end preventing efficiently binding of the 43S preinitiation

complex to the cap. However, when the IRE was placed distal to the cap, the

preinitiation complex could bind to the mRNA but not proceed due to the

imposed inhibition of 5'-UTR scanning. These features make the inducible IRE-

31

Page 32: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

IRP system highly suitable to block the movement of 40S subunits at a

predetermined site on any given mRNA in yeast cells.

1.2.5 The circular-loop model of eukaryotic mRNA

1.2.5.1 The mRNA 5' – 3' interaction The idea, that ribosomes on a closed-loop mRNA – termed then circular

polysomes – are the most efficient means to achieve expression of a given

transcript dates back a few decades (Philipps, 1965; Baglioni et al., 1969).

Later, electron microscopy revealed circular polysomes on the rough

endoplasmic reticulum, although the observation of linear and 'G' shaped

polysomes indicated that circularisation might either be not permanent or not

affect all mRNA species (Christensen et al., 1987). However, it was not until

genetic and biochemical methods were developed that the molecular basis of

this phenomenon could be explained and interpreted. The discovery that both

mRNA ends are protected by a cap structure and a poly(A)-tail and recruit

eIF4G and Pab1p, respectively, resulted in Tarun & Sachs (1996) reporting a

direct interaction of these factors that is necessary to promote translation

initiation synergistically and to inhibit mRNA degradation. Subsequently, Wells

et al. (1998) demonstrated that an in vitro reconstituted yeast eIF4E – eIF4G –

Pab1p complex could circularise a capped, polyadenylated mRNA. This

communication between the mRNA's 5' and 3'-end appears to be conserved

throughout all eukaryotes (Imataka et al., 1998; Piron et al., 1998; Wakiyama et

al., 2000), although this interaction might not be direct in mammalian cells

(Craig et al., 1998). Since transcript circularisation potentially concerns

numerous aspects of posttranscriptional control, extensive and recent reviews

on this topic are available (Jacobson, 1996; Jacobson & Peltz, 1996; Sachs et

al., 1997; Gallie, 1998; Kozak, 1999; Sachs & Varani, 2000; Sachs, 2000; Groft

& Burley, 2002).

eIF4G and Pab1p synergistically enhance translation initiation

Functionally, the interaction between initiation factor eIF4G that recruits the

ribosome to the mRNA via eIF3 (see also Figure 1.2) and Pab1p offers a

straightforward explanation, how the poly(A)-tail can promote cap-independent

translation initiation (Munroe & Jacobson, 1990; Tarun and Sachs, 1995;

Lamphear et al., 1995; Tarun et al., 1997). Moreover, Pab1p-binding to eIF4G

32

Page 33: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

increases the affinity of the cap-binding complex eIF4F (of which eIF4G is a

subunit) for the 5'-cap structure, resulting in a synergistically enhanced increase

of cap-dependent translation initiation (Sachs & Davis, 1989; Gallie, 1991;

Gerstel et al. 1992; Iizuka et al., 1994; Tarun and Sachs, 1995; Kessler &

Sachs, 1998; Otero et al., 1999). Alternatively, synergy could also base on

enhanced activity such as the ATPase within eIF4A, when all the factors have

bound. Notably, in addition to its effect on eIF4G and, consequently, 40S

recruitment, Pab1p also promotes the 60S subunit joining step (Sachs & Davis,

1989), most likely via Ski2p/Slh1p and eIF5/eIF5B (Searfoss et al., 2001).

Several groups have described that Pab1p exerts its enhancing effect on

translation initiation also in trans, on a poly(A)-deficient mRNA (e.g. Proweller &

Butler, 1996; Otero et al., 1999). Munroe & Jacobson (1990) and more recently,

Borman et al. (2002) reported that exogenous poly(A) stimulated translation via

Pab1p and eIF4G to the same extent as the presence of a poly(A) tail at the

mRNA 3'-end. Pab1p also could exert its effects in translation and stabilisation,

when it was brought to the 3'-end as a fusion protein, independent from the

poly(A) tail (Coller et al., 1998). Yet, in vivo eIF4G and eIF4B exhibit an

synergistic effect on Pab1p binding affinity for poly(A) (Le et al., 1997), and

since poly(A) deficient mRNAs are translated only poorly (Gallie, 1991) the

activity of poly(A) and thus Pab1p is likely to be almost exclusively

intermolecular and not in trans. For the sake of completeness, it should also be

noted that poly(A)-deficient mRNAs are not efficiently exported from the nucleus

(Zhao et al., 1999). Kozak (1991) suggested that in cells recruitment of Pab1p

to the poly(A) tail increases the 'local concentration' of the binding partner

eIF4G and therefore promotes (re)initiation at the polyadenylated mRNA.

mRNA circularisation and stability

The association of the cap-structure with cap-binding proteins and Pab1p-

binding to the poly(A)-tail prevents mRNA degradation directly by blocking the

access of ribonucleases and decapping enzymes to the mRNA's termini. Yet,

on a circular mRNA Pab1p also protects the 5'-cap by enhancing the affinity of

eIF4G for the cap-binding factor eIF4E (Wei et al., 1998; von der Haar et al.,

2000). This would also explain the sequence of the major mRNA degradation

pathway: deadenylation, decapping and 5' 3' exonucleolytic decay (see also

Section 1.3.1.1). Circularisation and the resulting enhancement of translation

33

Page 34: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

initiation by the 3'-end is also a means to test the integrity of RNA prior to

translation; incomplete or partly degraded mRNA would not sustain high

translation efficiency, preventing the production of truncated proteins.

Ribosome recycling via the poly(A) – Pab1p – eIF4G – cap bridge?

While the poly(A)-binding protein facilitates 40S recruitment by the initiation

factors and the subsequent 60S joining step, circularisation of the transcript

theoretically also offers the possibility of posttermination ribosomes being

directly channelled from the 3' to the 5'-end of the same mRNA, thereby

putatively enhancing the specific translational dramatically. Support for this

theory comes from measurements stating that capped transcripts lacking a

poly(A) tail do not enter translation in polysomes at a degree that was expected

if terminating ribosomes were to disassociate from an mRNA and equilibrate

with free subunits prior to re-entry into a polysome (Baglioni et al., 1969). This

suggested a preferential re-recruitment of those ribosomal subunits already

associated with an mRNA.

In a different approach, Novoa and Carrasco (1999) addressed the

question whether translation is divided between a 'pioneering' round that

recruits ribosomes from the cellular pool and the subsequent rounds that rely on

intramolecular recycling. They added poliovirus protease 2A to a pool of

translated mRNAs, cleaving the eIF4GI and II carboxy-terminally from the

eIF4E/PABP binding regions and blocking de novo translation. Under conditions

of near complete eIF4G cleavage, cellular mRNAs continued to be translated

for hours whereas, for instance, a newly transcribed reporter mRNA was very

poorly translated. This suggests that the eIF4E/PABP 'bridge' left by eIF4G

cleavage suffices to maintain a translationally competent mRNA, probably by

helping terminating ribosomes to reinitiate translation. More direct evidence is

provided by spreads of giant polysome mRNA isolated from the salivary glands

of Chironomus species. Due to the length of the growing nascent peptide the

position of the UTRs can be determined. Indeed, three to four ribosomes are

observed bound to what constitutes the 3'-UTR of these mRNAs (see Figure

1A, Kiseleva, 1989 and Figure 2, Francke et al., 1982), pointing at a

posttermination ribosomal scanning process. This could give ribosomes more

'time' to be re-recruited for another round of translation or they could be

channelled directly from the 3'-end to the cap or the site of translation initiation.

34

Page 35: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

One aim of the presented work is to assay this possibility in context with the

reinitiation (and scanning) possibilities of ribosomes having terminated at a

long, wild type ORF.

1.2.5.2 The 3'-UTR, a loop within the loop? The model discussed above assumes that the link potentially 'channelling'

posttermination ribosome to the transcript's 5'-end is formed by the poly(A)-tail,

Pab1p, eIF4G, eIF4E and the cap (see also Figure 1.4). Yet, recently a new

interaction between eukaryotic translation termination factor eRF3 and the

poly(A)-binding protein has been reported, first in mammalian cells (Hoshino et

al., 1999), then in yeast (Uchida et al., 2002). Cosson and colleagues (2002)

also showed that Pab1p acts in translation termination and that this interaction

is involved in the subsequent rounds of translation but not in de novo formation

of the initiation complex. Interestingly, they also demonstrated that eRF3-

binding disrupts the polymerisation of Pab1p, thereby carrying the termination

signal to the poly(A) tail and probably coupling translation termination and

mRNA deadenylation and decay. Thus, in addition to the possibility of

posttermination ribosomes scanning the 3'-UTR another scenario might be the

case, whereby ribosomes or subunits are directly transferred form the point of

translation termination to the 5'-cap. As well as enhancing translational

efficiency, McCarthy (1998) points out that the events at the termination site

would gain in kinetic control on translation upon coupling with the translation

initiation step that figures predominantly in terms of posttranscriptional control.

Structurally, this would result in the 3'-UTR looping out although this might only

occur at a termination event when eRF1/eRF3 are present at the stop codon.

1.3 mRNA stability in eukaryotes Transcript degradation is not simply the end of an mRNA's natural life but an

important factor in the control of gene expression. The translational efficiency of

a certain gene is dependent on its mRNA steady-state abundance, which in turn

is determined mostly by transcription and mRNA degradation rates. This is

reflected by the physical half-lives of transcripts that range in S. cerevisiae from

less than 1 minute to over 60 minutes, or from 100-fold shorter than cellular

generation times to those spanning several cell cycles (Herrick et al., 1990).

This difference in stability was found not to correlate with transcript properties

35

Page 36: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

such as ORF length, codon bias, ribosome density or abundance (Caponigro &

Parker, 1996; Wang et al., 2001b). In contrast, sequences located in the

untranslated regions or within the ORF can determine the stability of a transcript

(see below for details). Moreover, although the decay rate of 'household'

transcripts appears invariable, some mRNAs are regulated in their degradation

behaviour in answer to environmental cues (see references in McCarthy, 1998,

for example).

Most of the recent work in this field has focused on the pathways of

'regular' mRNA degradation and on the surveillance mechanism that protects

the cell from aberrant transcripts by accelerated transcript decay.

1.3.1 'Regular' pathways of mRNA turnover in yeast

1.3.1.1 The major pathway of mRNA decay: deadenylation, decapping and 5' 3’ exonucleolytic degradation

Extensive work by the Parker group on the turnover of the PGK1 mRNA in S.

cerevisiae has revealed a major pathway for the degradation of most cellular

mRNAs in yeast and higher eukaryotes (Muhlrad et al., 1995; Caponigro &

Parker, 1996; Wilusz et al., 2001).

The shortening of the poly(A) tail to oligo-length and the concomitant

release of poly(A) binding protein Pab1p marks the transition from translation-

competent to decay-susceptible mRNA (Figure 1.4). Deadenylation occurs in

two steps with an initial shortening of the poly(A) tail from 70-90 residues to 10-

15 adenosines, followed by a slower removal of the last nucleotides, termed

terminal deadenylation. Although the loss of Pab1p-association with the mRNA

reduces the translational efficiency drastically, it remains unclear whether any of

the two events are rate-limiting in mRNA decay (Caponigro & Parker, 1996). In

mammalians, a major cytoplasmic deadenylase activity has been assigned to

the poly(A)-specific deadenylating nuclease, initially termed DAN, subsequently

designated as poly(A) ribonuclease, PARN. In yeast, the deadenylation is

surprisingly less well characterised. An obvious candidate is the Pab1p-

dependent Pan2/Pan3 (poly(A) nuclease) complex that trims nascent poly(A)

tails in the nucleus. The cytoplasmic activity might be regulated by the

transcription factor complex Ccr4/Caf1 (see Wilusz et al., 2001, and references

therein). Interestingly, deletion of PAB1 in yeast led to a stabilisation of the

36

Page 37: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

poly(A) tails, implying that Pab1p simultaneously protects the poly(A) tail

and recruits the deadenylase to the mRNA's 3'-end (Mitchell & Tollervey, 2000).

Pab1p-bestowed protection is also partly due to mRNA circularisation conferred

by the Pab1p – eIF4G – eIF4E interaction (Figure 1.4), which also explains the

inhibitory effect of Pab1p on the accessibility of the 5'-end to decapping.

Moreover, in mammalian cells, the deadenylase PARN binds to the cap

structure, thereby preventing the access of the decapping enzymes until poly(A)

tail shortening is completed (Wilusz et al., 2001). However, it remains to be

elucidated exactly how the signal from the 3'-end is transferred to the 5'-end,

deadenylation

AAAAAAAAAAAAAAneIF4G

eIF4EeIFsPab1Pab1 Pab1

AREs

STEs

Xrn1

PANDcp1

+/--+ -- + -

-

poly(A) oligo(A)PAN (?), Ccr4p/Caf1p (?)

A10-15

Pab1

A10-15

eIF4E/4G

5' 3' exonuc leolytic decay

A10-15

Dcp1

Rrp4p, Rr41p/Ski6p, Ski2p, Ski3p, Ski8p, Rrp44p (exosome ?)

A10-15m7Gpp

eRF1, eRF3, Upf1-3p

A10-15

decapping

Dcp1p, Dcp2p, Lsm1-8p Vsp16, Mrt1p/Pat1p/Spb10

+ -

processivedegradation

processivedegradation

+ -

3' 5' exonuc leolytic decay

deadenylation

AAAAAAAAAAAAAAneIF4GeIF4G

eIF4EeIFsPab1Pab1Pab1Pab1 Pab1Pab1

AREsAREs

STEsSTEs

Xrn1

Xrn1

PAN

PANDcp1Dcp1

+/--+ -- + -

-

poly(A) oligo(A)PAN (?), Ccr4p/Caf1p (?)

A10-15A10-15

Pab1Pab1

A10-15

eIF4E/4G

5' 3' exonuc leolytic decay

A10-15

Dcp1

Rrp4p, Rr41p/Ski6p, Ski2p, Ski3p, Ski8p, Rrp44p (exosome ?)

A10-15m7Gpp

eRF1, eRF3, Upf1-3p

A10-15

decapping

Dcp1p, Dcp2p, Lsm1-8p Vsp16, Mrt1p/Pat1p/Spb10

+ -

processivedegradation

processivedegradation

+ -

3' 5' exonuc leolytic decay

Figure 1.4. Deadenylation-dependent pathways of mRNA degradation in S. cerevisiae. The transition from translated and protected transcript to degradation-competent mRNA is shown. Pab1p can recruit the major deadenylase (supposedly PAN or Ccr4p/Caf1p) to the poly(A) tail,which is modulated by AREs (A/U rich elements) in the 3'-UTR and by termination factors such as eRFs and Upfs. After shortening of the poly(A) tail and release of Pab1p the major decaypathway involves decapping by Dcp1 and 5' 3' exonucleolytic degradation. A minority of deadenylated mRNA is degraded by 3' 5' exonucleases, putatively involving the exosome and associated factors. Arrows and lines refer to inhibiting (-) and modulating (+/-) interactions.

37

Page 38: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

thereby ensuring the order of deadenylation and decapping.

Once deadenylated, the protective cap structure of the transcript is

removed exposing the transcript to subsequent 5' 3' degradation (Figure 1.4).

Although in vitro the S. cerevisiae enzyme Dcp1p is sufficient for efficient

decapping, the in vivo reaction is more complex. Mrt1p/Pat1p/Sbp10p appears

to be a key player, the protein co-sediments with 40S subunits and polysomal

fractions and possibly recruits the Lsm complex (Sbp8p/Lsm1p and Lsm2-8p)

that in turn interacts with Dcp1p via Dcp2p, a putative nucleoside

diphosphotase (Mitchell & Tollervey, 2000). Pab1p has also been proposed to

interact with Mrt1p, thereby protecting the cap from removal and regulating the

transition from translation to mRNA degradation. Once deadenylated,

transcripts are decapped at different rates whereas the subsequent final

degradation occurs very rapidly. Hence, it can be assumed that removal of the

cap structure is a key control point and rate-limiting step in mRNA turnover

(Caponigro & Parker, 1996). Interestingly, reductions in the rate of cellular

translation elongation also affect mRNA decapping. Treatment of S. cerevisiae

with the elongation inhibitor cycloheximide stabilises many yeast mRNAs; at

least for the PGK1 and MFA2 transcripts this is due to an inhibition of mRNA

decapping (Muhlrad et al., 1995). Since both, translated and untranslated MFA2

transcripts are stabilised, this points at a general mechanism whereby the

inhibition of decapping is a global response to the decrease in translation

elongation (Beelman & Parker, 1994). Moreover, shifting mutant yeast cells

harbouring a temperature-sensitive allele of the gene encoding tRNA nucleotidyl

transferase enzyme CCA1 to the non-permissive temperature results in a

phenotype similar to that of cycloheximide addition. Protein synthesis is

decreased rapidly, ribosome density on mRNAs is enhanced and several

mRNAs are stabilised (Peltz et al., 1992). Although the precise mode of action

is not known, an interesting possibility is that mRNA decapping is regulated in

response to reduced protein production, with increased mRNA stability

compensating partly for the decrease in translation.

Upon removal of the 5' m7GTP structure, the unprotected mRNA is rapidly

degraded by Xrn1p, a major cytoplasmic 5' 3' exonuclease (Figure 1.4).

Mutations in initiation factor eIF5A led to the accumulation of decapped mRNA,

suggesting that accessibility of the 5'-end of the mRNA is also influenced by

38

Page 39: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

translation initiation events after decapping. Moreover, Xrn1p also interacts with

the Lsm proteins, which might engage the exonuclease to the decapped 5'-end

(Mitchell & Tollervey, 2000).

Determinants of mRNA degradation

In higher eukaryotes, the mRNAs of transiently expressed genes, such as early

response genes for lymphokines, transcription factors etc have a very short half

life that correlates well with the presence of adenylate, uridylate-rich (AU-rich)

instability elements (AREs) in the 3'-UTR (Bakheet et al., 2001). AREs can

affect the efficiency of deadenylation and subsequent steps of decay via their

interaction with a plethora of ARE-binding proteins (reviewed in Wilusz et al.,

2001). For example, the overexpression of the ARE-binding protein HuR/HuA

leads to stabilisation of deadenylated, ARE-containing mRNA whereas

conversely in vivo depletion of AUF1/hnRNP D leads to a strong stabilisation of

diverse ARE-containing mRNAs (Mitchell & Tollervey, 2000). Despite intensive

investigation, it remains unclear how AREs and -binding proteins affect the rate

of deadenylation and decay rates. Possibilities are the disruptions of the PABP

– eIF4G or the 5'-cap – eIF4E interaction, thereby providing access for PARN.

In yeast, destabilising elements have been identified throughout the length

of mRNAs, e.g. in the 5'-UTR of the PPR1 and SDH Ip transcripts, or in the

coding region of several genes including MATα1, HIS3, STE3 and SPO13, yet

how these sequence function in mRNA stability still remains unexplained. The

same is true for the instability elements in the 3'-UTR e.g. of HTB1, MFA2, and

STE3 mRNAs (Caponigro & Parker, 1996). These elements are transferable, for

example by creating chimeric mRNAs composed of parts form stable and

instable mRNAs (Heaton et al., 1992). By employing this method, a stabilising

sequence STE has also been proposed to be responsible for the high stability of

PGK1 mRNA with a half-life of about 35 minutes (LaGarnadeur & Parker, 1999).

Details of the STE sequence, the STE-binding protein Pub1p and the

destabilising sequence element DSE are discussed in context with uORF-

related destabilisation in Section 1.4.6. Recently, a yeast transcript, TIF51A,

has been identified whose stability is regulated through its AU-rich 3'-UTR. Both

yeast and mammalian AREs promote deadenylation-dependent decapping in

the yeast system, and the STE/ARE binding protein Pub1p regulates the

stability mediated by the human TNFα ARE (Vasudevan & Peltz, 2001).

39

Page 40: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

In S. cerevisiae, the stability of mRNAs groups is controlled in response to

growth conditions. E.g., five different meiotic mRNAs are stabilised in response

to the shift from vegetative growth in glucose medium to sporulation conditions

in acetate medium. This effect depends on UME2p and UME5p in an unknown

way but obviously does not involve a change in the rate of deadenylation

(McCarthy, 1998). In another example, the SDH2 mRNA encoding the iron

protein of the succinate dehydrogenase complex is unstable and present at low

level when S. cerevisiae is grown with glucose. The glucose-triggered

degradation and changes in translational efficiency of this transcript and of

another example, the SUC2 mRNA, were traced to the 5'-UTR sequence, yet

the precise mechanism remains unclear (de la Cruz et al., 2002).

1.3.1.2 3' 5' exonucleolytic and endonucleolytic mRNA degradation In addition to above described 5' 3' pathway, the yeast PGK1 transcript can

also be degraded in a 3' to 5' direction, although slower than by the 5' to 3'

decay (Figure 1.4). Therefore, this pathway becomes mainly evident in the

absence of the major 5' 3' exonuclease Xrn1p. The 3' 5' exonucleolytic

activity depends on the Ski proteins and the proteins of the exosome (Jacobs et

al., 1998; van Hoof & Parker, 1999), but a mere scavenger function is unlikely

since e.g. the deadenylated MFA2 transcript poses no substrate for this kind of

decay (Caponigro & Parker, 1996). Given that for the majority of yeast mRNAs

analysed so far, decapping is the major fate of deadenylated mRNAs, 3' 5'

exonucleolytic decay could target only particular mRNAs. For example, van

Hoof et al. (2002) and Frischmeyer et al. (2002) reported recently that

ribosomes stalled at the 3'-end of transcripts lacking a stop codon are degraded

in the 3' 5' direction requiring exosome-associated Ski proteins. However, the

situation in mammals could be different, since Wang and Kiledjian (2001)

reported recently that the exosome-associated 3' 5' exonucleolytic activity

contributes significantly to in vivo mRNA decay. They also identified an

exoribonuclease-dependent scavenger decapping activity, and even in yeast at

least one of the both pathways is required for viability since ski2/xrn1 or

ski3/xrn1 double mutants induce synthetic lethality.

The translating ribosome can also arrive at the mRNA’s poly(A) tail when

the in-frame stop codon was not recognised due to a read-through event. In a

model proposed by Vasudevan et al. (2002) translation proceeds then through

40

Page 41: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

the poly(A) tail displacing Pab1p and adding poly-lysine to the C terminus of the

protein product. At the 3’-end of the non-stop RNA, the ribosome stalls and is

only released by interaction of Ski7p with the now empty A site of the ribosome.

Binding of Ski7p results in recruitment of the exosome and dissociation of the

ribosome. The non-stop mRNA is degraded 3' 5' by the exosome and the

protein product may be targeted for proteolysis by the poly-lysine tract.

However, read-through levels of most yeast mRNAs are well below 1.2%

(Bonetti et al., 1995).

mRNA decay can also be triggered by endonucleolytic cleavage, although

in contrast to an abundance of examples in higher eukaryotes (Jacobson &

Peltz, 1996) until now only one mRNA, encoding the ribosomal L2 protein, has

been found to be the target for endonucleases in yeast (Presutti et al., 1995).

1.3.2 mRNA surveillance and nonsense-mediated decay

1.3.2.1 The need for surveillance In all eukaryotes, mRNAs containing premature translation termination codons

(UAA, UGA or UAG) are highly unstable (for reviews, see e.g. Hilleren & Parker,

1999, and Czaplinski et al., 1999). This phenomenon of mRNA surveillance is

usually termed nonsense-mediated decay (NMD) since the majority of

transcripts recognised by this system comprises 'nonsense' premature

termination codons (PTCs, Peltz et al., 1993; Muhlrad & Parker, 1994; for

reviews see Jacobson & Peltz, 2000, and González et al., 1999). For example,

the presence of any of the three stop codons in the first two thirds of the PGK1

ORF reduces the half-life time from normally ~60 to 4 minutes (Hagan et al.,

1995). Although some 'normal' transcripts such as PPR1 and CTF13 are also

degraded via the NMD pathway, it is assumed that degradation of aberrant

transcripts is required to prevent the production of truncated, potentially

dangerous proteins (Czaplinski et al., 1999). Indeed, concomitant with

accelerated mRNA decay, the recognition of a PTC also decreases the

translational efficiency of the transcript (Muhlrad & Parker, 1999b). Yet, it should

be noted that in S. cerevisiae, this surveillance system is dispensable for normal

growth, indicating that stabilisation of mRNAs that are aberrantly produced does

not suffice to kill the cell (Caponigro & Parker, 1996).

41

Page 42: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

1.3.2.2 Cis- and trans-acting factors in NMD In contrast to the major degradation pathway, NMD does not require initial

deadenylation of the aberrant mRNA, but the cap of the polyadenylated mRNA

is removed directly, followed by a rapid 5' 3' degradation step. In yeast, these

activities are performed by the enzymes Dcp1p and Xrn1p that are also active

in the major deadenylation-dependent pathway, (Muhlrad & Parker, 1994;

Beelman et al., 1995). Notably, decapping of NMD substrates does not require

Mrt1p or Lsm1p, suggesting the decapping enzyme is recruited by a different

mechanism (see references in Mitchell & Tollervey, 2000). Logically, the only

situation where the distinction between normal and premature stop codon can

occur, is the translation termination event. Indeed, the activities of the

surveillance system were found to concentrate on the stop codon site involving

regular and additional termination factors (Figure 1.5). During translation

termination, the release factor eRF1 recognises the stop codon, and eRF3

modulates this activity in a GTP-dependent manner, also triggering peptide

release from the ribosome (see Section 1.2.3). In addition, the Upf1, Upf2 and

Upf3 (up-frameshift) proteins have initially been found to associate with the

termination complex and be able to modulate translation read-through of

nonsense-containing transcripts, 'saving' them from NMD (Leeds et al., 1991;

Peltz et al., 1993; Cui et al., 1995; Lee & Culbertson, 1995; Jacobson & Peltz,

2000). Upf1p demonstrates RNA binding, RNA-dependent ATPase and RNA

helicase activity (Czaplinski et al., 1995 and 1999; Weng et al., 1996a,b and

1998, Bhattacharya et al., 2000) and interacts with both, eRF1 and eRF3

(Czaplinski et al., 1998 and 1999). UPF3 encodes a protein with several

nucleus localisation signals and was shown to shuttle between the cytoplasm

and the nucleus (Lee & Culbertson, 1995; Shirley et al., 1998). Upf1p and

Upf3p do not interact directly with each other but both bind to Upf2p, which

explains both the observed complex formation and the similar effects of any

UPF deletion on mRNA decay (He et al., 1997; Atkin et al., 1997). Upf2p might

also play a crucial role in decapping, since a motif search in the

Schizosaccharomyces. pombe and human homologues of Upf2p has identified

so-called eIF4G homology (4GH) domains that span regions of the eIF4G

protein necessary to mediate the closed-loop formation of the mRNA (Mendell

et al., 2000). These data suggest a model in which Upf2p competes with eIF4G

42

Page 43: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

for 4GH domain-mediated interactions, thus disrupting or preventing the

formation of the stable, closed-loop mRNP structure and preparing the access

of Dcp1p.

Interestingly, although only Upf1p binds to eRF1 directly, all Upf proteins

CBC

An AnDSE PTC DSE

Yeast Mammals

PTC

Hrp1p

ALY/REV

AnPTC

CBC

eIF4E

nuclear membrane

eIF4G

A n

PTC

80S

Upf2

eIF4E

Upf1

Upf3

PTC

AAAAAAAAAAn

eIF4G

eIF4E

eRF3

eRF1 Upf2

Upf1Upf3

Pab1 Pab1

eRF3

eRF1 Upf2

Upf1 Upf3

eRF3

eRF1Upf2

Upf1

Upf3

eRF3

Upf2Upf1

Upf3

eRF3

Upf2Upf1

Upf3

AnPTCDcp1

eIF4EXrn1peIF4G

Dcp1

5' 3' exonuleolytic decay

binding ofmarker protein(s)

factor exchange,mRNA export

Initial translation round(nuclear, perinuclear or cytoplasmic ?):Stop codon recognition & Upf recruitment

mRNA degradation

eRFs & Upfs recycling

decapping

exon junction complex (EJC)

Identification of premature termination

Y14

CBCCBC

An AnDSE PTC DSE

Yeast Mammals

PTC

Hrp1p

ALY/REV

AnPTC AnPTC

CBCCBC

eIF4E

nuclear membrane

eIF4GeIF4G

A n

PTC

80S

Upf2Upf2

eIF4E

Upf1Upf1

Upf3Upf3

PTC

AAAAAAAAAAn

eIF4G

eIF4G

eIF4E

eRF3

eRF3

eRF1

eRF1 Upf2Upf2

Upf1Upf1Upf3Upf3

Pab1Pab1 Pab1Pab1

eRF3

eRF3

eRF1

eRF1 Upf2Upf2

Upf1Upf1 Upf3Upf3

eRF3eRF3

eRF1eRF1Upf2Upf2

Upf1Upf1

Upf3Upf3

eRF3eRF3

Upf2Upf2Upf1Upf1

Upf3Upf3

eRF3eRF3

Upf2Upf2Upf1Upf1

Upf3Upf3

AnPTCDcp1Dcp1

eIF4EXrn1peIF4GeIF4G

Dcp1Dcp1

5' 3' exonuleolytic decay

binding ofmarker protein(s)

factor exchange,mRNA export

Initial translation round(nuclear, perinuclear or cytoplasmic ?):Stop codon recognition & Upf recruitment

mRNA degradation

eRFs & Upfs recycling

decapping

exon junction complex (EJC)

Identification of premature termination

Y14

Figure 1.5. Nonsense-mediated decay of premature termination codon containing mRNAs.In both, yeast (downstream elements) and mammalian cells (exon junction complex), markerproteins bind to sites up- and especially downstream of a PTC supposedly facilitating itsrecognition when positioned within a certain distance. The location of the Upf proteins within thecell is indicated, and as a consensus of different models the recognition of premature terminationduring a first round of translation is shown to be perinuclear. Interaction of the termination factorswith the mRNA's 5'-end leads to decapping by Dcp1p although not via the pathway of normaldeadenylation-dependent decapping. Release of eRFs and recycling of Upfs are thought tooccur by mRNP remodelling.

43

Page 44: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

can interact with eRF3, but Upf2p, Upf3p and eRF1 compete with each other for

this interaction (Wang et al., 2001a). Yet, while this points to a sequential

surveillance complex assembly and rearrangement (Figure 1.5), the

intermediates remain to be characterised. Moreover, in yeast, the Upf proteins

are primarily cytoplasmic and polysomal associated (Atkin et al., 1995 and

1997; He & Jacobson, 1995), whereas the localisation and site of action of the

human homologues is currently under discussion (see Section 1.1.2 on nuclear

translation).

In both, mammalian and yeast systems, marker proteins have been

proposed that bind to specific mRNA sites and indicate via their position relative

to the terminating ribosome whether the termination codon is premature or not.

In S. cerevisiae, loosely defined sequence located between PTCs and the 3'-

UTR have been put forward to function as a signal to trigger NMD. These G/C-

rich sequences, termed downstream elements (DSE), are functional in positions

up to 150 nt downstream of the PTC and on mRNAs other from which they have

been isolated (Peltz et al., 1993; Hagan et al., 1995, Zhang et al., 1995; Ruiz-

Echevarria et al., 1996 and 1998). Recent work identified a protein,

Hrp1p/Nab4p that specifically binds to DSEs and is required for destabilisation

of PTC-containing mRNAs by interacting with Upf1p (Gonzalez et al., 2000). As

a result of this interaction, the Upf proteins could potentially identify the

existence of a close downstream DSE, which would label the termination

'premature' and result in accelerated decay. Conversely, a proper termination

complex would be located 3' of all DSEs and associated markers, thereby

preventing NMD. To explain how degradation of transcripts comprising stop

codons that can be recognised as premature, e.g. those containing upstream

ORFs, is prevented, the Peltz group identified a stabilising element (STE) in

yeast that can obviously interfere with NMD. Although a STE-binding protein,

Pub1p, has been purified that is necessary for controlling the activity of the

NMD pathway, the mode of action still remains unclear (Ruiz-Echevarria &

Peltz, 2000; see also Section 1.4.6 about uORFs).

No DSE sequences have been identified in mammalian cells, yet

premature stop codon identification also relies on a bipartite system (Maquat,

2000). During splicing, an exon-junction complex (EJC) is deposited about 20-

24 nt upstream of the splice site. Of the numerous EJC proteins, only Aly/Ref

44

Page 45: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

and Y14 bind to spliced mRNA and are exported to the cytoplasm (Figure 1.5;

Le Hir et al., 2000), but other components of the marker complex could bind to

one of these proteins and be co-transported out of the nucleus. During the initial

round of translation, the elongating ribosome strips these proteins from the

mRNA. In contrast, a ribosome terminating at a PTC could interact via Upf1p

and Upf2p with Upf3p poised at the exon-junction complex, in a manner similar

to the DSE/Hrp1 recognition by the Upf complex in yeast (Kim et al., 2001).

However, the spatial arrangements in mammalian appears to be more relaxed

since a PTC up to 645 nt upstream of the EJC can still trigger NMD (Neu-Yilik et

al., 2001).

1.3.2.3 mRNA surveillance by mRNP (re)organisation ? A current model of eukaryotic mRNA surveillance suggests that a kind of kinetic

proofreading of proper translation termination takes place and is assessed by

mRNP domain organisation (Hilleren & Parker, 1999). As mentioned above and

depicted in Figure 1.5, the (re)arrangement of the mRNP structure is

significantly shaped by a plethora of components such as the cap-binding

complex, translation initiation (and termination) factors, surveillance complex

components such as Upf1-3p, and the poly(A)-binding protein in addition to

other factors such as DSE-, STE-, ARE- and EJC-binding proteins. In case of

translation termination, the authors propose that the hydrolysis of ATP by Upf1p

acts as an 'internal clock' to verify the site of the termination codon. When

translation termination is completed due to proper mRNP domain organisation

before ATP is hydrolysed, then the mRNA is protected. Conversely, if

termination is slow, ATP hydrolysis and mRNP rearrangement changes the fate

of both, the posttermination ribosome and the mRNA. In support of this model,

Upf1p also acts in 'normal' termination, enhancing constitutive stop codon

recognition, termination complex assembly and suppressing nonsense codon

read-through (Weng et al., 1996a). This role of Upf1p depends on its ATPase

activity (Weng et al., 1998). Upf1-3p – although dispensable for normal growth –

are also thought to regulate decapping and exonucleolytic degradation of both

nonsense-containing and wild type mRNAs (He & Jacobson, 2001).

Interestingly, by virtue of its association with Pab1p and the nonsense-mediated

decay factor Upf1p, eRF3 brings together two complexes that are important for

the stability of mRNA. Thus, in this central position, Upf1p-directed ATP

45

Page 46: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

hydrolysis as well as mRNP remodelling might eventually decide between

Pab1p dissociation and successive 'regular' mRNA decay and NMD triggered

by decapping, or mRNA stabilisation and further translation.

According to the Hilleren & Parker (1999) model of NMD by mRNP

rearrangement, it is not the 'improper' end to translation but rather the lack of

proper termination in a suitable mRNP environment that triggers accelerated

decay. Hence, in addition to PTC and uORF-containing mRNAs, also other

mRNPs with a 'distorted' structure are subject to NMD. As discussed briefly,

spliced mRNAs retain an exon junction complex that signals a terminating

ribosome whether a translation termination is premature (He et al., 1993).

Unspliced transcripts are not only very likely to comprise a PTC in the intron

sequences but the lack of the EJC also explains the observation that

transfection constructs bearing at least one intron tend to yield higher

expression levels (e.g. Zhang et al., 1998). This cannot be compensated for by

using exonic 'failsafe' sequences (Neu-Yilik et al., 2001), leading to the

assumption that intron-less mammalian mRNAs may fail to form a proper

terminal mRNP domain and therefore would be subject to mRNA surveillance.

In this context, it is of interest that yeast, where only few transcripts undergo

splicing, the 3'-UTRs are strikingly homogeneous in length, averaging at about

100 nt (Graber et al., 1999). It is possible that, akin to the mammalian model,

the yeast terminating ribosome needs to interact with proteins bound to the 3'-

UTR. Indeed, in yeast, a mutation in CYC1 results in low levels of mRNAs with

extended 3'-ends, and Muhlrad & Parker (1999a) could show that these

aberrant mRNAs are subject to the NMD pathway. Hilleren and Parker suggest

that the significantly altered 3'-terminal mRNP domains of these transcripts are

marking them for degradation.

1.3.2.4 The evidence for mRNA surveillance in the nucleus Currently, a discussion is going on whether nonsense-mediated decay occurs in

the cytoplasm, during nuclear export or in the nucleus itself (for example,

Dahlberg et al., 2003; Hentze, 2001; see also Section 1.1.3). In yeast, the

nuclear-cytoplasmic shuttling of Upf3p implies a role for the nuclear

compartment in NMD (Shirley et al., 1998) but this may be restricted to mark

mRNAs for degradation outside or at the nucleus. Maderazo and co-workers

(2003) also showed recently that at least in yeast, PTC-containing mRNAs that

46

Page 47: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

were exported from the nucleus were readily recognised and degraded in the

cytoplasm upon reconstitution of the NMD pathway. In support of translation or

at least a scanning-like mechanism in the nucleus, Ishigaki et al. (2001)

reported that in mammalian immunoprecipitated PTC-comprising transcripts are

degraded in association with the nucleic cap binding protein CBP80. The

authors proposed a model, whereby the first round of translation is initiated by

the nuclear cap binding complex and serves to scan for aberrant transcripts.

Additional evidence comes from an observed nuclear splicing escape

mechanism that skip PTC-introducing mutations, and that TCR transcripts

comprising nonsense stop codons are downregulated depending on NMD and

translation factors (Wilkinson & Shyu, 2002; Wang et al., 2002). Furthermore,

Iborra and colleagues (2001) reported that mRNA translation indeed could take

place in the nucleus (see Section 1.1.2 for details). Their findings indicate that

the sites of translation within the nucleus coincides with patterns of a subunit of

the proteasome, indicating how truncated protein products from aberrant

mRNAs could be degraded immediately. Since the model of nuclear translation

– apart from other implications e.g. about the origin of the eukaryotic nucleus –

contradicts a prime orthodoxy in eukaryotic molecular biology, the evidence

obviously has to be scrutinised carefully. Dahlberg and colleagues (2003) point

out that with current methods the perinuclear cytoplasm could cofractionate with

the nucleus, making it impossible to distinguish between events in the nucleus

and those coupled with export into the cytoplasm. This possibility of the

'privileged' pioneering round to occur not in the nucleus but during nuclear

export is a main point of critique (see also Figure 1.5). It is probably at this

location the that nuclear cap-binding complex CBC is readily exchanged for

eIF4E that then can interact with Pab1p, resulting in a circularised mRNA.

Subsequent rounds of translations would then take place on a circular mRNA

with the 5' – 3' interaction offering both, protection from degradation and

enhanced translation efficiency.

47

Page 48: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

1.4 uORFs: reinitiation of translation vs. ribosome release

1.4.1 Polycistronic transcripts in eukaryotes? The majority of eukaryotic transcripts are monocistronic in a sense that they

comprise only a single major ORF. However, according to a survey in 1987 by

Kozak, in less then 10% of eukaryotic mRNAs the 5'-untranslated region is not

what it ought to be: untranslated. This mRNA part, hence better referred to as

'leader', can comprise upstream start codons (uAUGs) or upstream open

reading frames (uORFs). Following the proposed mechanism of 5'-end

ribosomal recruitment and subsequent 5' 3' scanning of the leader, these

elements exert a strong regulative role on the initiation frequency at the main

ORF (Geballe, 1996). Fittingly, the majority of uORFs is found in certain classes

of mRNAs involved in the control of cellular growth and differentiation (Kozak,

1987b; Kozak, 1991; Morris, 1995). It has been found difficult to estimate how

widespread uORFs are among the various organisms since only a minority of

databases containing precise and verified mapping of transcripts' 5'-ends. The

use of alternative transcription sites, alternative mRNA processing, and

alternative initiation codons also complicates the exact leader definition of a

given mRNA (see also Kozak, 2001a). In S. cerevisiae, however, a number of

uORF-comprising mRNAs has been positively identified (reviewed in McCarthy,

1998), and estimates suggest that up to 200 (out of approximately 6000) yeast

genes may contain this regulatory element (Vilela et al., 1998). In accordance

with published literature, throughout this work the term 5' untranslated regions

(5'-UTR) comprises the sequence of (regulative) uORFs although they are

technically translated; the expression ORF refers to a cistron encoding a full-

length gene product.

Although all uORFs can potentially undergo a full cycle of translation –

including initiation, peptide synthesis and termination, as shown by primer

extension experiments (Gaba et al., 2001) – there are differences in exactly

how they exert their restrictive influence on downstream initiation (Morris &

Geballe, 2000). The potential of a ribosome to reinitiate further downstream, as

well as the site at which it reinitiates can vary depending on both trans-acting

factors and the structure of the mRNA.

48

Page 49: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

1.4.2 uORF control in GCN4 translation: the stop codon context and eIF2 phosphorylation

In one of the most famous cases of mRNA-specific translational control, the

regulation of the GCN4 gene in yeast, amino acid starvation of cells leads to an

overall decrease in protein biosynthesis, while the expression of genes involved

in amino acid synthesis is up regulated. This is mediated via the translational

regulation of GCN4p, which itself presents a transcriptional activator

(Hinnebusch, 1984 and 1988). The initiation frequency at the GCN4 main ORF

is regulated by four uORFs that upon translation can exert a positive or negative

affect on reinitiation frequency. This control pathway is reviewed in detail by

Hinnbusch (1997).

In short, uORF1 and, to a lesser extent, uORF2 have been found to be

positive regulators of GCN4 translation since ribosomes terminating at these

ORFs are capable of downstream scanning and subsequent reinitiation. In turn,

uORF3 and especially uORF4 are promoting ribosome release at their

termination sites, thus preventing ribosomes to reach the downstream AUG of

the main ORF. The balance of initiations at the respective ORFs is regulated by

the availability of the active ternary complex Met-tRNAi-eIF2-GTP and the stop

codon context of the two upstream elements uORF1 and uORF4 that are most

relevant for regulation. The first uORF comprises an A/U-rich stop codon

context that is favouring reinitiation of translation whereas the G/C-rich

downstream element of uORF4 was shown to promote ribosome release (Grant

& Hinnebusch, 1994). These authors proposed that G/C-rich elements could

stall the terminating complex long enough for the subunits to dissociate,

probably by G/C base stacking or interaction of the mRNA site with tRNA, rRNA

or other parts of the mRNA. The GCN4 uORF4 10 nt downstream sequence

was also shown to act as a transferable element since it can reduce reinitiation

when placed downstream of the translationally non-restrictive YAP1 uORF

(Vilela et al., 1998). Sequences 5' of uORF1 have also been implied to be

required for efficient reinitiation, although the mechanism for this remains

unclear (Grant et al., 1994).

To (re)initiate translation, ribosomes need to be associated with factors

such as the active ternary complex Met-tRNAi-eIF2-GTP. In yeast, this is

regulated in response to environmental conditions via a kinase pathway. Under

starvation conditions, the kinase GCN2 is activated involving the binding of an

49

Page 50: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

uncharged tRNA and consequently phosphorylates the α subunit of eIF2

(reviewed in Hinnebusch, 1996). This enhances the affinity of eIF2 for the

nucleotide exchange factor eIF2B that is required to recycle the binary eIF2-

GDP complex (Yang & Hinnebusch, 1996). Since the exchange factor is

presumed to be considerably less abundant than eIF2, the phosphorylation of

20-30% of eIF2 is sufficient to effectively make eIF2B unavailable for nucleotide

exchange on eIF2-GDP complexes, resulting in a strong inhibition of translation

(Hershey, 1991; Rhoads, 1993).

By combining the two findings about stop codon context and eIF2

requirement, it was put forward that under normal growth conditions ribosomes

terminating at GCN4 uORF1 are able to resume scanning and acquire the

ternary complex in time to reinitiate at uORF4. Termination at this upstream

open reading frame then inhibits further scanning and translation of the GCN4

transcription factor. However, when eIF2 abundance is low, the scanning

ribosome supposedly takes longer to re-bind the ternary complex, thereby

'ignoring' the uORF4 start codon and initiating at the main ORF.

1.4.3 Reinitiation is affected by the intercistronic distance and the uORF length

In addition to intracellular eIF2 levels, the intercistronic distances between the

(u)ORFs plays a crucial role in determining whether the ribosomes can become

reinitiation-competent or not. Having terminated, ribosomes need a certain time

span to re-acquire the Met-tRNAi delivered by the ternary complex. This

minimum threshold is usually directly correlated with the length of the sequence

the ribosome can scan until it encounters a downstream AUG. Early information

came from Kozak (1987a) reporting that in human cells reinitiation becomes

more efficient when the distance between uORF and the next AUG was

lengthened without introduction of a significant secondary structure. An

intercistronic spacer of 79 nt entirely relieved the restriction on reinitiation

imposed by the uORF. In yeast cells, Grant and colleagues (1994) observed a

decrease in main ORF expression when the sequence between uORF1 and

GCN4 were shortened from 350 nt to 50 nt on a leader lacking the other three

uORFs. Similarly, under starvation conditions (and therefore lowered eIF2

levels) increasing the distance between uORF1 and uORF4 led to enhanced

ribosome release at uORF4 and therefore to a reduced frequency of reinitiation

50

Page 51: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

at the downstream GCN4 ORF (Abastado et al., 1991). In turn, shortening the

spacer from 200 nt to 32 nt resulted in increased GCN4 synthesis (Grant et al.,

1994). The requirement of a longer spacer for efficient reinitiation at the main

GCN4 ORF (>50 nt) in respect to the shorter intercistronic spacer needed for

reinitiation at uORF4 (32 nt) was explained by two types of model. The first

postulates that a slow initiation step (as at uORF4) would facilitate the binding

of the ternary complex whereas the second proposes that sequences at the

uORF4 start site can recruit additional, as yet uncharacterised, factors. Notably,

genetic experiments also implicate eIF3 in the Met-tRNAi rebinding step

(Garcia-Barrio et al., 1995). In all cases, the control would be kinetic. The

correlation between intercistronic distance and reinitiation competence has also

been observed in systems as diverse as HIV-1 (Luukkonen et al., 1995) and

fungi (Arst & Sheerins, 1996), for a review see Kozak (2002).

In analogy to the importance of the intercistronic spacer, the size of the 5'-

proximal ORF is thought to be a major constraint to reinitiation in eukaryotes.

This was initially deduced from studies in plant and animal viruses that produce

bicistronic mRNA in which the 3'-cistron is translationally silent (reviewed in

Kozak, 1978). While the scanning ribosome has to regain factors essential for

(re)initiation, especially the ternary complex, the elongating ribosome is thought

to lose them gradually. Ribosomes having translated only a short uORF

supposedly still retain these factors. Therefore, extending an upstream ORF

has a negative effect on reinitiation at a downstream ORF, but there are only

very few studies systematically exploring the uORF length – reinitiation

relationship. Luukkonnen et al. (1995) saw a gradual diminishment on viral

mRNA and predicted an almost complete absence of reinitiation capability when

uORFs reach a length of 55 codons (168 nt). Hwang and Su (1998) observed a

2-fold decrease in protein yield when uORF length was lengthened from 7 (24

nt) to 18 codons (57 nt) in a similar system. In rabbit reticulocyte lysate, Kozak

(2001) observed no change in reinitiation when the uORF was lengthened from

3 (12 nt) to 13 codons (42 nt) and then a gradual decrease. Protein yield of the

downstream ORF was decreased 3-fold when the uORF was expanded from 13

to 33 codons (102 nt), irrespective of the inserted sequence. Furthermore,

changing part of the uORF sequence to that of a pseudoknot reduced

reinitiation even more, most likely by slowing down the elongating ribosome.

51

Page 52: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

Kozak concludes that it is therefore difficult to determine a 'cut-off' length for

ORFs but that reinitiation competence is rather kinetically controlled, and that

slowing down of the ribosomal movement on the ORF therefore promotes the

release of initiation relevant factors.

1.4.4 Other factors governing reinitiation All upstream open reading frames and the main ORF of GCN4 were found to

recruit scanning ribosomes with about the same frequency (Grant et al., 1994).

However, in some cases the uORF initiation rate varies and therefore regulates

downstream reinitiation. Whether the initiation site is recognised or by-passed

by 'leaky scanning' can be affected by to the start codon context, the distance of

the start codon from the cap, the nature of the start codon (AUG or GUG), the

modulation by viral proteins and events such as ribosome shunting (see also

Section 1.2.4, Morris & Geballe, 2000).

Maybe surprisingly, also uORFs with initiation frequencies as low as 10%

have the potential to prevent ribosomes from reaching downstream ORFs by

causing ribosomes to stall either during elongation or during termination, thus

creating a blockade to subsequent scanning or translating ribosomes. In the

eukaryotic cases reported so far, the structure or sequence of the nascent

peptide chain is responsible for perturbing the normal events of uORF

translation, although the responsible peptides characterised so far do not share

a common consensus sequence (Morris & Geballe, 2000). For example, coding

sequence-dependent arrest of the ribosome at the uORF of the human

cytomegalovirus gpUL4 transcript was mapped to the termination codon by

primer extension experiments (Cao & Geballe, 1996). Similarly, the AdoMetDC

uORF failed to affect downstream initiation when the termination codon was

altered (Cao & Geballe, 1995; Mize et al., 1998). While in these two cases, the

nascent peptide inhibits peptidyl-tRNA hydrolysis during termination, the peptide

produced by the fungal uORFs of arg-2 and CPA-1 mRNAs most likely affects

the peptidyltransferase centre and thereby influences both terminating and

elongating ribosomes (Morris & Geballe, 2000).

Ribosome regulation by nascent peptides is also widespread in

prokaryotes and has been studied especially intensively in mRNAs encoding

antibiotics (see Lovett and Rogers, 1996, for review). For example, expression

of chloramphenicol acetyl transferase from the cat mRNA is subject to

52

Page 53: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

autoregulation conferred by ribosome stalling at an upstream leader. This

impediment is aided by low levels of the encoded antibiotics, which in turn

exposes the ribosome-binding site for translation of the main ORF. Similar to

the situation in eukaryotes, the peptides encoded by the inhibitory uORF are not

conserved, and they target the peptidyltransferase centre of the big ribosomal

subunit (Lovett and Rogers, 1996).

1.4.5 Reinitiation of translation in prokaryotes and eukaryotes The lack of reinitiation competence after translation of a full-length ORF also

explains the almost absolute absence of polycistronic transcripts in eukaryotes

(Kozak, 1989b). While prokaryotes can initiate translation at almost any site on

the mRNA via guidance of the Shine-Dalgarno sequence, the eukaryotic

translation apparatus is usually limited in its AUG recognition to the 5' 3'

scanning mechanism. This may explain the evolution of regulative mechanisms

in eukaryotes such as structured 5'-UTRs, leaky scanning, reinitiation and

ribosome shunt. Furthermore, prokaryotic ribosomes do not need to (re)acquire

initiation factors such as the ternary eIF2 complex. Therefore, reinitiation

competence in prokaryotes is not restricted by the length of the 5'-positioned

ORF. Furthermore, the situation for the influence of the intercistronic length is

reverse. Eukaryotic ribosomes regain the ability to initiate during

posttermination scanning, and backward scanning followed by reinitiation

upstream of the termination site is almost neglectable (Kozak, 2001b). In

contrast, 70S posttermination ribosomes seem to be able to shuffle back and

forth on the mRNA although only in a region obviously determined by the

ribosomes off-rate that lies on the order of one ribosome-equivalent length

(reviewed by McCarthy, 1998). Proximity of stop and start site enhances

reinitiation frequency as does the existence of a Shine-Dalgarno sequence. If

the intercistronic distance is too long, coupled reinitiation frequency declines but

ribosomes from the cellular pool can still enter directly at a downstream

translation initiation region (TIR). This is analogous to artificial eukaryotic

systems whereby a full-length 5'-ORF is followed by an internal ribosome entry

site (IRES) facilitating initiation at a downstream cistron (Kozak, 2001a).

In prokaryotes, the ribosome recycling factor RRF has been shown to

possess the property to release ribosomes from the mRNA, and its absence

resulted in highly elevated reinitiation of translation 3' to the termination codon

53

Page 54: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

(Janosi et al., 1998). In eukaryotes, no cytoplasmic RRF homologue has been

found (see Section 1.2.3), and the events that decide the fate of the

posttermination ribosome remain unclear. Although release factor eRF3 is a

strong candidate to possess the function of RRF in eukaryotes, the role of this

and other termination complex associated factors such as Pab1p, Upf1-3p have

to be elucidated before a model for the switch from ribosome recycling to

reinitiation can be proposed. Therefore, it also remains a distinct possibility that

posttermination ribosomes can remain on the mRNA following translation of a

full-length ORF and scan along the 3'-UTR to be channelled to the 5'-end. This

question will be addressed in part of this thesis.

1.4.6 uORFs can regulate transcript stability Beside their effect on the ability of (posttermination) ribosomes to (re)initiate,

uORFs can exert an influence on mRNA stability. Since their primary structure

is similar to mRNAs comprising premature termination codons (PTCs) they

have the potential to trigger accelerated mRNA degradation by the nonsense-

mediated pathway (NMD). However, the utilisation of this particular pathway to

reduce transcript abundance and gene expression has been reported for only a

restricted number of uORF-comprising mRNAs such as CPA1 (Ruiz-Echevarria

& Peltz, 2000) or aberrant transcripts like the his4-38 allele comprising a single

G insertion (Donahue et al., 1981; Leeds et al., 1991) and the inefficiently

spliced CYH2 precursor (He et al., 1993). Oliveira and McCarthy (1995)

reported that in yeast an uORF introduced by a single nucleotide mutation in an

artificial leader reduced the mRNA's steady-state levels to 20% accounting for

most of the simultaneously observed decrease in translational efficiency. Aided

by results from control constructs harbouring secondary structures, they

concluded that not translation inhibition per se but rather the mechanism by

which the recognition of the main ORF is rendered less efficient does determine

mRNA destabilisation. Indeed, the group could show later that the restriction of

posttermination scanning triggered accelerated mRNA decay (Linz et al., 1997).

An improved frequency of initiation and subsequent termination at the uORF

reducing downstream initiation was proposed to be responsible for the Upf1p-

dependent mRNA destabilisation. Notably, the artificial uORF tested preceded a

bacterial CAT ORF indicating that destabilisation does not require elements

unique to yeast transcripts.

54

Page 55: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

However, most of the natural uORF-containing mRNAs such as GCN4 and

(Ruiz-Echevarria et al., 1998) and YAP1 (Vilela et al., 1998) are not subject to

degradation via the NMD pathway. Therefore, Ruiz-Echevarria and Peltz (2000)

have proposed the existence of an 'escape mechanism' that relies on the

presence of an uORF-downstream stabiliser element (STE). These authors

showed that loosely defined, A/U-rich sequences confer transcript stability due

to their interaction with the RNA-binding protein Pub1p. The exact mechanism

how Pub1p-binding to STE signals the surveillance complex that the termination

is not premature and how the effects of the downstream sequence element

(DSE) responsible for triggering the NMD pathway are compensated remain

unclear. Furthermore, the not strictly defined A/U-rich sequences downstream

of GCN4 uORF1 have been shown to promote ribosome reinitiation, which in

turn could account for the observed increase in stabilisation regardless of the

Pub1p interaction (see also Section 1.4.2). The convenience of explaining NMD

by the occurrence of STEs and DSEs is also challenged by the finding that the

two uORFs of YAP2 attenuate gene expression at the level of mRNA turnover.

Described first by Vilela and colleagues in 1998 and then reinvestigated in

1999, the YAP2 5'-UTR causes accelerated mRNA decay that is largely

independent of the NMD factor Upf1p. No DSE- or STE-like element could be

identified in the YAP2 leader, but similarly to GCN4 uORF4 the destabilising two

uORFs were followed by a G/C-rich sequence promoting ribosome release and

preventing reinitiation. Thus, the authors describe a model, whereby uORF-

directed destabilisation is modulated by events at the termination site

determining whether ribosomes can resume scanning and reinitiate, or the

ribosomes are released from the mRNA triggering its decay. The ability for

ribosomes to reinitiate might also be the key to explain the absence of Upf-

dependent destabilisation by the GCN4 uORFs. Ruiz-Echevarria & Peltz (1996)

inserted a G/C-rich DSE downstream of uORF4 that triggered NMD. This effect

was relieved when the DSE was placed > 200 nt from the stop codon or an A/U-

rich STE was inserted between the termination site and the DSE. Obviously, the

decision at the translation termination site whether ribosomes are released or

proceed to reinitiate defines the fate of the mRNA and can be affected by

downstream sequences.

55

Page 56: Translation reinitiation and ribosome recycling - PhD Thesis

1. Introduction

Therefore, it remains to be tested whether the STE element exerts its

stabilising effect simply due to the sequence's A/U content and accordingly

facilitated reinitiation or because of interactions, e.g. with Pub1p (which in turn

could also promote reinitiation). The role of STEs is also somewhat complicated

by the proposal that there exists another class of these elements within ORFs

that requires translation to exert its positive effect (Peltz et al., 1993; Hagan et

al., 1995; Ruiz-Echevarria et al., 2001).

1.5 Objectives The recent progress in translational research has provided inspiring information

about the factors and the function of mRNA circularisation. The synergistic

enhancement in translation initiation was traced to the interaction of the proteins

associated with the transcripts termini; yet there has been no thorough

examination of the possibility that terminating ribosomes might be recycled

directly to the mRNA 5’-end. Since the major part of translation is performed by

ribosomes that have terminated at least once, this ‘channelling’ of subunits that

are protected from the recruitment by competing mRNAs might contribute

significantly to the overall translational efficiency of the individual transcripts.

Besides, very little is known about the fate of the posttermination ribosome

and whether the subunits dissociate immediately at the stop codon or if they

migrate into the 3’UTR, which has been shown to occur after translation of short

open reading frames. This study aims to fill this gap by blocking this putative

posttermination scanning step, and monitoring of translational efficiency and

mRNA stabilisation will reveal the contribution of this assumed intramolecular

recycling.

Another aspect of this work will focus on a methodical analysis of the

correlation of uORF length and ribosomal reinitiation competence by stepwise

truncation of a full-length ORF. This is expected to give evidence about the

differences of posttermination ribosome having translated either a short,

regulative uORF or a full-length cistron comprised by the majority of cellular

mRNAs. Additionally, some of the studies researching uORFs and reinitiation

suffer from the occurrence of leaky scanning along with reinitiation. Hence, one

of the objectives will be to rule out this mechanism obscuring the origin of

ribosomes initiating at the downstream ORF.

56

Page 57: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

2. MATERIALS AND METHODS

2.1 Chemicals and enzymes Tables 2.1 and 2.2 provide an overview of the chemicals and enzymes used.

Throughout this section, all concentrations are in weight / volume (w/v) except

where stated differently. Table 2.3 gives an overview of RCF values and the

corresponding speed in rpm in common rotors. Chemical Supplier Remark agar Merck agarose Sigma RNA grade acetate BDH amino acids Sigma ammonium chloride BDH NH4Cl ampicillin Melford amp adenosine triphosphate

Sigma ATP

bacto-tryptone Difco bacto-peptone Difco bovine serum albumin Sigma BSA 5-bromo-4-chloro-3-indolyl phosphate disodium salt

Sigma BCIP

borate BDH calcium dichloride BDH CaCl2 ·

2H2O caesium chloride Sigma CsCl, 98% cap analogue NEBL m7GpppG creatine phosphate Sigma cycloheximide Sigma D-cuciferin Sigma diethyl pyrocarbonate Sigma DEPC dimethyl formamide BDH DMF dithiothreitol Promega DTT ethylene diamine-tetraacetic acid

Sigma EDTA

ethanol BDH ethidium bromide Sigma EtBr ficoll Sigma ficoll 400 formaldehyde Fisher 37% formamide Gibco

BRL deionised

32P γ-ATP ICN 7500 mCi/ml

glass beads BDH 0.4 mm galactose BDH GAL glucose BDH GLU glycerol Merck glycine BDH glycylglycine Sigma guanosine triphosphate

Sigma GTP

Chemical Supplier Remark isopropanol BDH kanamycin Melford kan lithium acetate Sigma LiAc mannitol BDH magnesium acetate BDH MgAc magnesiumoxid acetate

BDH MgOAc

magnesium chloride BDH MgCl2· 6H2O

magnesium sulfate BDH MgSO4 methanol BDH 3-(N-morpholino) propanesulfonic acid

BDH MOPS

nitro blue tetrazolium Melford NBT phenylmethyl-sulphonyl fluoride

BDH PMSF

polyacrylamid solution

Severn Biotech

30% or 40%

polyethylene glycol BDH PEG4000, MW 4000

polyvinyl-pyrollidone BDH PVP potassium acetate BDH Kac potassium hydroxid BDH KOH potassiumoxid acetate BDH KOAc salmon sperm DNA Sigma preparation

according to Sambrock et al. (1989)

sodium acetate BDH NaAc sodium chloride BDH NaCl sodium dodecyl sulfate

BDH SDS

sodium bicarbonate BDH NaHCO3 sodium phosphate BDH NaH2PO4 sodium citrate BDH NaCitrate bis(2-hydroxyeth yl)-imino-tris(hydro xymethyl)methane

BDH Tris

urea BDH yeast extract Merck yeast nitrogen base Difco YNB (with

ammonium)

Table 2.1. Chemicals used. Abbreviations are mentioned in the remark column.

57

Page 58: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

Enymes & antibodies Supplier Remark AMV Roche avian myeloblastosis virus reverse transcriptase anti-FLAG, from rabbit Santa Cruz

Biotechnology dilution 1/1000

anti-rabbit, from goat Sigma dilution 1/10000, conjugate with alkaline phosphates (AP)

DNase Promega Klenow fragment Roche PNK NEBL polynucleotide kinase restriction enzymes NEBL, Roche various RNase A Promega Freed of DNases-according to Sambrook et al.

(1989) RNase H Promega RNasin Promega T7 RNA Polymerase NEBL

Table 2.2. Enzymes and antibodies.

Max. RCF (x g) rpm Centrifuge, rotor 860 3,000 Heraeus biofuge pico 1,240 3,000 Beckman JA-17 2,200 4,000 Beckman JA-17 2,400 5,000 Heraeus biofuge pico 8,800 8,000 Beckman JA-17 16,100 13,000 Howe Sigma 1K13 35,000 30,000 Beckman JA-25.50 230,000 50,000 Beckman 70.1 Ti 270,000 39,000 Beckman SW40 Ti

Table 2.3. Correlation of rpm and RCF values in centrifuges and rotors used. RCF, relative centrifugal force; rpm, revolutions per minute

2.2 Cell methods

2.2.1 Strains and cell culturing

Saccharomyces cerevisiae strains

W303 [MATa/MATα ade2-1/ade2-1 his3-11/his3-11 leu2-3,112/leu2-3,112

trp1∆1/ trp1∆1 ura3-52/ura3-52 can1-100/can1-100], see Thomas & Rothstein

(1989)

NT33-5 [relevant genotype: ura3-52 leu2-3,112 cca1-1] as described in Wolfe et

al. (1996)

YAS1947/pTIF4631-213TRP1CEN This strain was derived from YAS1947

[MATα ade2 his3 leu2 trp1 ura3 tif4631::LEU2 tif4632::ura3

pTIF4632URA3CEN, Tarun et al. (1997)] by plasmid shuffling. For this, the

original strain containing the URA3 plasmid was transformed with the TRP1

marker containing pTIF4631-213 vector (Tarun et al., 1997). Positive

58

Page 59: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

transformants were screened in several rounds of liquid culturing, streaking and

spreading onto YNB ura ade his or YNB trp ade his plates (see below) for loss

of the URA3 marker.

Yeast culturing

For the experiments, the yeast strains were grown at 30°C (NT33-5 at 25°C or

37°C) in YNB media (yeast nitrogen base, 6.7 g/l) and glucose or galactose (20

g/l) prepared in accordance with Sherman et al. (1986). Supplementation with

amino acids was to complement auxotrophies and to select for the respective

plasmid markers. Final concentrations [in mg/l] were: ade [40], ura [20], trp [20],

his [20], arg [20], tyr [30], leu [40], lys [30], ile [30], phe [50], val [150], thr [200],

ser [400].

By varying induction time and the carbon source, optimal conditions for

IRP1 expression were found to be as follows: 20 ml pre-culture was grown to

log-phase in YNB glucose, centrifuged for 5 minutes at 3500 rpm and washed

twice with sterile water. The cells were then resuspended and grown in YNB

galactose (YNB glucose for the control) to an OD600 of 0.8-0.9.

Escherichia coli strain and culturing

The E. coli strain Top-10F' [F'{lacIq Tn10(Tetr)} mcrA ∆(mrr-hsdRMS-mrcBC)

F80lacZ∆M15 ∆lac74 recA1 deoR araD139 ∆(ara-leu)7697 galU galK rpsL

endA1 nupG, obtained from Invitrogen] was used for DNA amplification. E. coli

media LB was prepared as described in Sambrook et al. (1989) to a final

concentration of 1% bacto-tryptone, 1% NaCl and 0.5% yeast-extract.

2.2.2 Escherichia coli transformation E. coli was transformed according to the CaCl2 method (Huff et al., 1990). For

this, bacterial cells were grown over night at 37°C in 5 ml of LB medium with

shaking (200 rpm). 2 ml of this culture were added to 100 ml LB medium, and

the cells were allowed to grow to an OD600 of 0.6. Subsequently, the cells were

cooled on ice for 10 minutes and then harvested by centrifugation for 10

minutes at 4°C and 2,200 x g. The cell pellet was resuspended in autoclaved 50

ml ice-cold 0.1 M CaCl2 and cooled on ice for 30 minutes. Afterwards, the

bacteria were pelleted again for 10 minutes at 4°C and 2,200 x g, resuspended

in 4 ml ice-cold 0.1 M CaCl2 and kept on ice over night. Sterile glycerol was

59

Page 60: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

added to a final concentration of 20%, and aliquots of the competent cells were

snap frozen in liquid nitrogen and stored for later use at -80°C.

For transformation, 100 µl of competent cells were mixed with 50 ng

plasmid DNA and incubated on ice for 30 minutes. The cells were heat-shocked

for 50 seconds at 42°C, cooled on ice for 2 minutes and then incubated with

900 µl LB medium at 37°C with shaking (200 rpm) for 45 minutes. Next, the

cells were collected at 2,400 x g and spread on LB plates (LB with 1.5% agar).

The plates contained 100 µg/ml of the appropriate antibiotic (ampicillin or

kanamycin) selecting thereby for successfully transformed cells carrying the

plasmid with the corresponding resistance gene. Colonies were formed when

the plates were incubated over night at 37°C, short-term storage was at 4°C

thereafter.

2.2.3 Saccharomyces cerevisiae transformation Yeast transformation was performed according to the lithium acetate method

(Schiestl & Gietzl, 1989). Cells were inoculated in 20 ml YEPD (10 g/l yeast

extract, 20 g/l bacto-pepton and 3 g/l glucose) and incubated with shaking (180

rpm) over night at 30°C. On the next day, about 1 ml of the overnight culture

was added to 20 ml fresh YEPD to give 5x106 cells/ml (OD600 ~ 0.25). The

culture was allowed to grow to a density of 2x107 cells/ml (OD600 ~ 0.8) and

then harvested by centrifugation at 1,240 x g.

The cell pellet was washed twice by resuspending in 20 ml sterile water

and centrifuging as before. Subsequently, the cells were transferred to a 2 ml

micro tube and washed two times in 1 ml freshly prepared TELiAc buffer (10

mM Tris pH 7.7, 1 mM EDTA, 100 mM LiAc), collection was at 860 x g. Finally,

the cells were resuspended in 1 ml TELiAC to give 2x109 cells/ml.

For transformation, 5 µg of each plasmid DNA and 20 µg salmon sperm

DNA were incubated with 50 µl (108) freshly prepared yeast cells and 300 µl

PEG/LiAc (50% PEG4000, 10 mM Tris pH 7.7, 1 mM EDTA, 100 mM LiAc) at

30°C for 30 minutes. After a 10 minutes (2 minutes for heat sensitive strains)

heat-shock at 42°C, the cells were collected by centrifugation at 860 x g for 5

minutes. Resuspended in 100 µl sterile water and plated onto appropriate

selective YNB-plates (YNB with 1.5% agar and 2% glucose added), yeast

colonies appeared 2-3 days later when incubated at 30°C (or 25°C for heat

sensitive strains).

60

Page 61: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

2.3 Recombinant DNA methods

2.3.1 DNA manipulation DNA cloning was performed using standard methods such as PCR, DNA

restriction and ligation, (Sambrook et al., 1989). To obtain the RT-PCR product

of actin and the spacer – IRE sequences (see below), the TA cloning kit

(Invitrogen) was employed following manufacturer's instructions. Mutations were

introduced by site-directed mutagenesis via PCR or insertion of double stranded

oligodeoxyribonucleotides. All constructs were confirmed by fluorescence

sequencing (Sequencing service, UMIST).

2.3.2 Screening of E. coli transformants Screening to detect transformants with plasmids carrying the desired DNA

sequence was performed by standard methods (Sambrook et al., 1989). For

mini-preparation of DNA by alkaline hydrolysis and subsequent restriction

analysis, single E. coli transformant colonies were inoculated over night at 37°C

in 2 ml LB medium with shaking (200 rpm). On the next day, 1.5 ml of this

culture was collected for 5 minutes at 2,400 x g and resuspended in 200 µl

solution I (50 mM glucose, 2 mM Tris.Cl pH 8.0, 10 mM EDTA pH 8.0, 100

µg/ml RNase A). Lysis of the cells was by addition of 400 µl solution II (1%

SDS, 0.2 M NaOH) and incubation for 2 minutes on ice. Mixing with 200 µl

solution III (3 M KAc, 5 M acetate) precipitated cell debris and genomic DNA.

After 5 minutes incubation on ice, the plasmid DNA was separated from the cell

pellet by centrifugation at 4°C and 16,100 x g for 5 minutes. The supernatant

was precipitated by adding 2 volumes of isopropanol. Having washed the pellet

with 70% ethanol, the dried DNA was resuspended in an appropriate volume to

give a concentration of about 1 µg/µl. The vector was subject to restriction

analysis to screen for recombinant DNA, positive samples were confirmed by

fluorescence sequencing.

Alternatively, up to 25 colonies were resuspended in 20 µl of a standard

PCR mix and subjected to amplification with plasmid-specific primers. After

agarose gel analysis, the plasmids of positive candidates were mini-preped

using a nucleo-spin kit (Macherey-Nagel) and sequenced.

61

Page 62: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

2.3.3 CsCl discontinuous gradient DNA purification Conventional DNA preparation implies treatment with RNases that can result in

mRNA degradation and in consecutive aberrant in vitro transcription reactions.

Hence, template plasmids were purified in CsCl gradients (adapted from

Sambrook et al., 1989). For this, DNA was prepared according to the mini-prep

protocol in Section 2.3.2 but without RNase treatment and resuspended in a

solution of 6.82 g CsCl and 3.91 g TE (10 mM Tris.Cl, 1 mM EDTA pH 8.0). To

this 0.64 g/g CsCl solution, ethidium bromide was added to a final concentration

of 0.43 mg/ml. The resulting solution was mixed and centrifuged for 5 minutes

at 8,800 x g. The clear red supernatant was loaded onto 6.5 ml of 0.75 g/l CsCl,

prelayed in a 13 ml Beckman Quickseal tube. After sealing of the tube,

centrifugation was at 15°C and 230,000 x g over night. Finally, the (lower) DNA

band was recovered with a syringe, the ethidium bromide separated by 6

extractions with butan-1-ol in TE and the remaining supernatant precipitated

with 2% polyethylene glycol. The pure DNA pellet was washed with 70%

ethanol and resuspended in water.

2.3.4 Plasmid constructs All oligodeoxyribonucleotides used in this research were obtained from Sigma-

Genosys or MWG Biotech and are listed in Table 2.4.

The IRP1-IRE system and poly(G)-containing constructs

Iron-response-element binding protein IRP1 (Butt et al., 1996; Gray et al., 1993)

was cloned as a 2.9 kb NdeI / XbaI fragment from YCpSUP-IRF (Oliveira et al.,

1993a) into YCp33Supex2 (Supex2, Oliveira et al., 1993b). Expression from this

plasmid is regulated by the galactose-inducible Gal-PGK fusion (GPF) promoter

(Vega Laso et al., 1993).

Plasmid FL comprising the constitutive promoter TEF1 and the firefly

luciferase ORF followed by a PGK1-derived 3'-UTR corresponds to YCp22FL1

in Oliveira et al. (1993a). FL-5'IRE with the IRE sequenced placed 9 nt into the

5'-UTR equals construct IRE-wt in Oliveira et al. (1993b), see also Table 2.5. To

create FL-derived constructs with the IRE in the 3'-UTR, synthetic DNA was

inserted between the XbaI and EcoRI sites of FL. PCR products derived from

reactions with the respective positive-strand primer (XbaIIRE(+), XbaI15IRE(+)

and XbaI30IRE(+)) and EcoRIIRE were used to achieve constructs FL-15-IRE,

62

Page 63: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

FL-24-IRE and FL-39-IRE. Likewise, annealed XbaI-poly(G)+/- primers were

ligated into the XbaI site of FL to produce FL-6-pG (Table 2.5). Primer Sequence) 5’ATG mut GGAATTCCATATGGATTACAAGGACGACGACGACAAGACTAAGGATTC 5’TEF CGGTCTTCAATTTCTCAAGTTTCAG 3’ATG mut GTAATCTAGAATTCTACATTAGTCACCGGCAAAACCGGCTTTACACAAACCAGAACC actDEL105- CGATAGATGGGAAGTCTAGAATTCTACATTAGTCACCGGCAAAACCGGC ActStop30+ GCTTTGGTTATTTAATGTAGAATTCGTATGTGTAAAGCC ActStop30- GGCTTTACACATACGAATTCTACATTAAATAACCAAAGC ActStop298+ GCCCCAGAATAATGTAGAATTCTTTTGACTGAAGCTCC ActStop298- GGAGCTTCAGTCAAAAGAATTCTACATTATTCTGGGGC ActStop640+ CGTGACATCAAGTAATGTAGAATTCACGTCGCCTTGGACTTCG ActStop640- CGAAGTCCAAGGCGACGTGAATTCTACATTACTTGATGTCACG ActStop931+ CCAGGTATTGCCTAATGTAGAATTCAGGAAATCACCGCTTTGGC ActStop931- GCCAAAGCGGTGATTTCCTGAATTCTACATTAGGCAATACCTGG ATGATmut+ GGAATTCCATATGATGTACAAGGACGAC ATGcorr- CAAAATATAATCGGTTCTAGAATTCTACTTTA ATGmut30- GGAATTCCATATGATGGATTACAAGGACGAC ATGmut60- GTAATCTAGAATTCTACATTAAATAACCAAAGCAGCAACCTCAG EcoRIIRE CCGGAATTCTTCTCGAGTTTTGTGTTTCTTAAGTTCAAGC EcoRIT(20) CCGGAATTCTTTTTTTTTTTTTTTTTTTT FlagATGdel+ GTTGTTTCTTAACATTTGGATTACAAG FlagATGdel- CTTGTAATCCAAATGTTAAGAAACAAC FLAGdel12- CCTCAGAATCCATATCTAGAATTCTACATTACTTGTAATCC FLAGdel30- CAATAACCAAAGCATCTAGAATTCTACATTAAGACTTGTCATCG IREXho+ CCGCTCGAGAATTATCTACTTAAGCTTC LUC-1Pro (+) TCGAGCAAACCGTAAT LUC-1Pro (-) CTAGATTACGGTTTGC LUC-2Pro (+) TCGAGCCCGTTGTAAT LUC-2Pro (-) CTAGATTACAACGGGC LUC10ds (+) TCGAGCAAATTGTAAGCGGTTACCTT LUC10ds (-) CTAGAAGGTAACCGCTTACAATTTGC LUC-1,2pro10ds (+) TCGAGCCCGCCGTAAGCGGTTACCTT LUC-1,2pro10ds (-) CTAGAAGGTAACCGCTTACGGCGGGC LucRev1546 CTCCTCCGCGCAACTTTTTCGCGGTTG LUCstopstop (+) TCGAGCAAATTGTAATAAT LUCstopstop (-) CTAGATTATTACAATTTGC LucTerm+ GAAGGGCGGAAAGTCCAAATTG LUCXho (+) GAAGGGCGGCTCGAGCAAATTG LUCXho (-) CAATTTGCTCGAGCCGCCCTTC Nde-Actin+ GGGAATTCCATATGGATTCTGAGGTTGCTGC NdeFlag+ TATGGATTACAAGGACGACGATGACAAGAC NdeFlag- TAGTCTTGTCATCGTCGTCCTTGTAATCCA P3XM+ CTATTATTTATCTCGAGATGATTATTAAG P3XM-New CTTAATAATCATCTCGAGATAAATAATAG PGK1TERM-new CCCAAGCTTTAACGTTCGCAGAATTTTCG PGK3UTRRev58 GCGTAAAGGATGGGGAAAGAGAAAAG PHS1-F GCACCGCCGCCGCAAGGAATGG RP131 GAAGTCAACTCTGATGAAGAC uORF1ds50+ TATGTCTAGAACCGATTATATTTTGTTTTTAAAGTAGATCTTCTCGAGAAAAAC uORF1ds50- TAGTTTTTCTCGAGAAGATCTACTTTAAAAACAAAATATAATCGGTTCTAGACA XbaI-Actin- CTAGTCTAGATTAGAAACACTTGTGGTGAACG XbaI-poly(G)+ CTAGAGGGGGGGGGGGGGGGGGGT XbaI-poly(G)- CTAGACCCCCCCCCCCCCCCCCCT XbaI15IRE(+) GCTCTAGAATTTATAAAAATTATCTACTTAAGCTTC XbaI30IRE(+) GCTCTAGAATTTATAAAAACAATTACCACAAAAATTATCTACTTAAGCTTC XbaIIIRE(+) GCTCTAGAAATTATCTACTTAAGCTTC XhoI-p(G)18+ TCGAGGGGGGGGGGGGGGGGGGC XhoI-p(G)18- TCGAGCCCCCCCCCCCCCCCCCC

Table 2.4. Sequences of oligodeoxyribonucleotides. Sequences of all primers used in this study are given in the 5' 3' orientation.

63

Page 64: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

For insertion of IRE close to the poly(A) site, first a XhoI-free reference

construct was created by replacing the 5'-UTR of FL BamHI / NdeI with the

longer uORF-free YAP1 leader of plasmid p∆uY1 (Vilela et al., 1998). Then,

with primers P3XM+ and P3XM-New, a new XhoI site was introduced 141

nucleotides into the 3'-UTR. The resulting construct FL' (see also Table 2.5)

displayed no significant difference to FL with respect to in vivo luciferase

expression and LUC mRNA stability (data not shown). To obtain plasmid FL-

141-IRE, the IRE sequence (derived from a XhoI restricted PCR reaction with

primers IREXho+ and EcoRIIRE) was inserted as a duplicate into the XhoI site

of FL'.

A 5'-UTR sequences FL AATTATCTACTTAAGAACACAAAACTCGAGAACATATG

FL' GGATCCTTACCGATTAAGCACAGTACCTTTACGTTATATATAGGATTGGTGTTTAGCTTTTTTTCCTGAGCCCCTGGTTGACTTGTGCAAGAACACGAGCCATTTTTAGTTTGTTTAAGGGAAGTTTTTTGCCACCCAAAACGTTTAAAGAAGGAAAAGTTGTTTCTTAACATATG

FL-5'IRE AATTATCTACTTAAGCTTCAACAGTGCTTGAACTTAAGAACACAAAACTCGAGAACATATG

B 3'-UTR sequences FL TAATCTAGAATTC – 143 nt – p(A)

FL-15-IRE TAATCTAGAAATTATCTACTTAAGCTTCAACAGTGCTTGAACTTAAGAAACACAAAACTCGAGAAGAATTC – 143 nt – p(A)

FL-24-IRE TAATCTAGAATTTATAAAAATTATCTACTTAAGCTTCAACAGTGCTTGAACTTAAGAAACACAAAACTCGAGAAGAATTC – 143 nt – p(A)

FL-39-IRE TAATCTAGAATTTATAAAAACAATTACCACAAAAATTATCTACTTAAGCTTCAACAGTGCTTGAACTTAAGAAACAAACTCGAGAAGAATTC –143 nt–CAA p(A)

FL-141-IRE TAATCTAGAATTC – 116 nt – CTCGAGAATTATCTACTTAAGCTTCAACAGTGCTTGAACTTAAGAAACACAAAACTCGAGAATTATCTACTTAAGCTTCAACAGTGCTTGAACTTAAGAAACACAAAACTCGAG – 21 nt – p(A)

FL-6-pG TAATCTAGAGGGGGGGGGGGGGGGGGGTCTAGAATTC – 143 nt – p(A)

FL-131-pG TAATCTAGAATTC – 116 nt – CTCGAGGGGGGGGGGGGGGGGGGCTCGAG – 21 nt - p(A)

C sequences of actin ORF truncations and intercistronic spacer ACTd-LUC d15 d33

ATG GAT TAC AAG GAC GAC GAT GAC AAG ACT ATG GAT TCT GAG GTT GCT GCT TTG GTTd63 d105 ______inter-ATT GAT AAC GGT TCT GGT ATG TGT AAA GCC GGT TTT GCC GGT GAC TAA TGTAGAATTCT________________cistronic_spacer________________ LUC AGAACCGATTATATTTTGTTTTTAAAGTAGATCTTCTCGAGAAAAACT ATG

ACTd ∆iAUG-LUC

ATG GAT TAC AAG GAC GAC GAC GAC AAG ACT AAG GAT TCT GAG GTT GCT GCT TTG GTT ATT GAT AAC GGT TCT GGT TTG TGT AAA GCC GGT TTT GCC GGT GAC TAA AGTAGAATTCTAGAACCGATTATATTTTGTTTTTAAAGTAGATCTTCTCGAGAAAAACT ATG

AUGAUG ACTd-LUC

ATG ATG TAC AAG GAC GAC GAC GAC AAG ACT AAG GAT TCT GAG GTT GCT GCT TTG GTT ATT GAT AAC GGT TCT GGT TTG TGT AAA GCC GGT TTT GCC GGT GAC TAA AGTAGAATTCTAGAACCGATTATATTTTGTTTTTAAAGTAGATCTTCTCGAGAAAAACT ATG

∆AUG ACTd-LUC

TTG GAT TAC AAG GAC GAC GAC GAC AAG ACT AAG GAT TCT GAG GTT GCT GCT TTG GTT ATT GAT AAC GGT TCT GGT TTG TGT AAA GCC GGT TTT GCC GGT GAC TAA AGTAGAATTCTAGAACCGATTATATTTTGTTTTTAAAGTAGATCTTCTCGAGAAAAACT ATG

Table 2.5. Sequences of FL-derived UTRs and ATG-mutated truncated actin-luciferase constructs. Sequences are given in the 5' 3' orientation. (A) 5'-UTR sequences (starting with the first transcribed nucleotide) of FL, FL' and FL-5'IRE and (B) 3'-UTR sequences of FL and 3'-UTR modified FL constructs. The bold ATG and TAA triplets refer to the luciferase start and stop-codon, respectively. The poly-guanidine stretch pG and the stems of the IRE hairpin-loop are underlined; p(A) specifies the polyadenylation site. (C) Truncated actin ORF and intercistronic sequence of actin-luciferase constructs. ATGs are indicated in bold, mutations relative to the preceding construct are underlined. d15, d33 d63 and d105 refer to the last codon of constructs with uORFs truncated to 15, 33, 63 and 105 nt, respectively. LUC specifies the luciferase start codon.

64

Page 65: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

Similarly, synthetic DNA (annealed oligodeoxyribonucleotides XhoI-p(G)18+

and XhoI-p(G)18-) was inserted into the XhoI site of FL' creating plasmid FL-

131-pG (Table 2.5).

For the experiments in strain YAS1947, plasmid BEFL was established to

place the LUC expression cassette of FL on a plasmid with the yeast ADE2

marker. For this, FL was cut at the sites PstI and HindIII, and the resulting DNA

fragment was (partly blunt-end-) ligated into BEVY-A (Miller 3rd et al., 1998) cut

at PstI and EcoRI. Construct BEFL-5'IRE was generated by replacing the

BamHI – NdeI fragment with the 5'-UTR of FL-5'IRE. BEFL-39-IRE is the result

of substituting the BEFL EcoRI – EcoRI section with the corresponding

fragment of FL-39-IRE.

LUC stop codon context modifications

FL* was derived from FL by replacing the 5'-UTR with the one of FL' and

introducing an XhoI site upstream of the LUC stop codon (Figure 3.5A) using a

two step PCR approach with primer pairs 5'TEF / LUCXho (-) and LUCXho (+) /

PGKTERM-new. This facilitated the change of the stop codon sequence by

insertion of deoxyribonucleotide pairs LUC-1Pro, LUC-2Pro, LUC-12Pro10ds,

LUC10ds and LUCstopstop as described in Figure 3.5A. FL* showed no

significant difference to FL regarding in vivo luciferase expression and LUC

mRNA stability (data not shown).

ACT1 cloning and LUC cassette appendage

The S. cerevisiae actin gene ACT1 was amplified from W303 cell lysates by RT-

PCR as described in RNA methods (Section 2.4.3) with primers Nde-Actin+ and

XbaI-Actin- introducing the conservative mutation A108C. The resulting DNA

was cloned NdeI / XbaI in vector FL' and consequently FLAG-tagged by

insertion of synthetic DNA (annealed primers NdeFlag+ / NdeFlag-) into the

NdeI site, mutating thereby the introduced second NdeI site. ACT-LUC and all

other bicistronic luciferase constructs were derived from the respective

monocistronic FLAG-tagged ACT construct by insertion of an XbaI / XbaI

spacer-luciferase cassette. This element was established by ligation of the

annealed primers uORF1ds50+/- into plasmid FL's NdeI site and comprises the

LUC ORF preceded by a 59 nucleotide intercistronic spacer sequence derived

from the reinitiation competent GCN4 uORF1 downstream sequence (Table

65

Page 66: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

2.5). Constructs ACT-pG-LUC and ACT-119-LUC were established by insertion

of a 18-guanosine stretch into the XhoI-site of ACT-LUC or by duplication of the

intercistronic spacer.

Premature termination codon constructs – series ACTptc and ACTptc-LUC

For insertion of premature stop codons in ACT1, the FLAG-actin ORF was

mutated by a two-step PCR downstream of position 960, 699, 324 or 60 to 5'-

TAATGTAGAATTC-3' using outside primers 5'Tef and PGKTERM-new and

internal primers ActStop30+/-, ActStop298+/-, ActStop640+/- and ActStop931+/-

. Thereby, the stop codon UAA and an invariable 10 nucleotides downstream

sequence was introduced giving rise to the series ACTptc-LUC.

C-terminal truncation of actin ORF – series ACTd and ACTd-LUC

In the truncated constructs, the 3'-end of the actin ORF was deleted either by

religation of the PTC-containing plasmids after restriction with EcoRI or by

deletion of sequences 3' of position 105, 33 and 15 via PCR with primers

actDEL105-, FLAGdel30- and FLAGdel12- (Table 2.5).

Mutation of actin AUGs – series ACTd∆iAUG-LUC, AUGAUGACTd-LUC and

ACTd∆iAUG-LUC

In order to eliminate internal AUGs, ACT-LUC constructs with an actin ORF

truncated to 105, 63, 33 or 15 nucleotides were mutated by PCR as shown in

Table 2.5 giving thereby rise to the ACTd∆iAUG-LUC series. Primers employed

were 5'ATG mut, 3'ATG mut, ATGmut30-, ATGmut60- and ATGcorr-.

In the AUGAUGACTd-LUC series, another PCR mutation with primer

ATGATmut+ changed the second actin codon of ACTd105∆iAUG-LUC,

ACTd63∆iAUG-LUC and ACTd33∆iAUG-LUC to AUG, introducing a tandem

start codon. Similarly, in a two-step PCR (primers 5'TEF / FlagATGdel+ and

PGKTERM-new / FlagATGdel-), the actin start codon was mutated to UUG in

the ∆AUGACT-LUC constructs, leaving the LUC start codon as the first AUG on

the mRNA (Table 2.5).

For experiments in strain NT33-5, the ACTd∆iAUG-LUC constructs were

subcloned NdeI / HindIII in the URA3 vector p∆uY1 (termed FL", Vilela et al.,

1998) carrying a luciferase expression cassette identical to the one of FL'.

66

Page 67: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

Vectors for in vitro transcription

Vectors pHST7-FL and pHST7-FL-39-IRE were derived from pHST7 (Koloteva

et al., 1997) by exchanging the NdeI – XbaI fragment with plasmids FL and FL-

39-IRE, respectively. This also removed an aberrant 3'-UTR contained in the

original plasmid pHST7-5'IRE as described in Koloteva et al. (1997).

2.4 RNA methods Glassware, plasticware and solutions used in the procedures described in this

section were treated with diethyl pyrocarbonate (DEPC) and autoclaved where

appropriate to prevent contamination with RNases.

2.4.1 Total yeast RNA extraction Total RNA was isolated using a modified version of the hot phenol procedure

(Rose et al., 1990; Köhrer & Domdey, 1991). In detail, 20 ml of each yeast

culture at OD600 ~ 0.8 were harvested for 5 minutes at 1,240 x g, washed once

with 20 ml ice-cold water and finally resuspended in 500 µl AE buffer (50 mM

sodium acetate, 10 mM EDTA pH 5.3) and 100 µl 10% SDS. RNA was

extracted in two cycles of hot phenol extraction (5 minutes shaking at 65°C,

snap-freezing in liquid nitrogen and 5 minutes centrifugation at 16,100 x g) and

consecutive two-fold phenol:chloroform:isopropanol (25:24:1, v/v/v) extraction

(5 minutes vortexing, 5 minutes centrifugation at 16,100 x g). Precipitation of the

aqueous phase was with 0.3 mM sodium acetate and ethanol in a final volume

of 2 ml. Samples were stored in ethanol short-term at -20°C or long-term at

-80C.

Before use, the samples were centrifuged at 4°C and 16,100 x g for 10

minutes, washed with 70% ice-cold ethanol and centrifuged as before. Then,

the dried pellet was resuspended in 20-50 µl DEPC-treated water by 3-5 cycles

of vigorous vortexing (1 minute) and cooling (2 minutes on ice).

2.4.2 Northern blot

Formaldehyde agarose gel electrophoresis and blotting

20 µg of total yeast RNA were denatured for 15 minutes at 65°C in a RNA

denaturing mix (6% formaldehyde, 50% formamide, 1x MOPS buffer [200 mM

MOPS pH 7.0, 50 mM sodium acetate, 10 mM EDTA pH 8.0]). Denaturation

was stopped by cooling on ice and addition of formamide loading buffer (1

67

Page 68: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

mg/ml bromophenol blue, 1 mg/ml xylene cyanol FF in 0.1 mM EDTA pH 8.0

and 100% formamide). The samples were separated on a 1.3% formaldehyde

agarose gel (1.3% agarose, 1x MOPS, 2.24 M formaldehyde in DEPC-treated

water) with continuous mixing of the MOPS buffer according to Sambrook et al.

(1989).

The RNA was then blotted onto nylon membranes (Hybond-N, Amersham

Pharmacia Biotech) according to manufacturer's instruction. The following day,

the RNA was cross-linked by UV-radiation (1,200 units in a Stratagene UV

Stratalinker 2400) and baked to the membrane at 80°C for 2 hours.

Hybridisation with γ-labelled primers 32P γ-labelling of 25 pmol deoxyribonucleotides was with polynucleotide kinase

following manufacturer's instruction. The end-labelled primer was purified from

unincorporated 32P γ-ATP using Sephadex G-25 columns (NAP-5, Amersham

Pharmacia Biotech). With respect to the sequence to be detected, the labelled

primer(s) – RP131 (PGK1 recognition), LucRev1546 (LUC), NdeFlag- (FLAG)

or PGK3UTRrev58 (FL 3'-UTR) – were denatured together with 250 µg salmon

sperm DNA. Following pre-hybridisation for at least 3 hours in 50 ml,

hybridisation was at 50°C over night in 10 ml fresh (pre-) hybridisation solution

according to Kuhn et al., 2001 (5X SSC [20x: 3 M NaCl, 300 mM NaCitrate], 20

mM NaH2PO4 pH 7.2, 7% SDS, 1x Denhardt solution [50x: 1% BSA, 1% ficoll,

1% polyvinyl-pyrollidone]). The membranes were then washed according to

Kuhn et al. (2001) in Wash Solution 1 (3x SSC, 10x Denhardt solution, 5%

SDS, 25 mM NaH2PO4 pH 7.2) and Wash Solution 2 (1x SSC, 1% SDS) at

50°C. If necessary, the membranes were stripped in 0.1x SSPE pH 7.4 (20x: 3

M NaCl, 200 mM NaH2PO4, 20 mM EDTA), 0.1% SDS at 80°C for subsequent

rehybridisation. Results were quantified on a Molecular Dynamics Typhoon

8600 PhosphorImager using the ImageQuant software, version 5.1.

2.4.3 RT-PCR 1 µg of total yeast mRNA, isolated from cell extracts as described in 2.4.1 was

subject to a first strand cDNA synthesis reaction with AMV reverse

transcriptase. RNA and 25 pmol of poly(A)-tail specific primer EcoRI-T(20) were

denatured at 65°C for 5 minutes and then cooled down to 30°C. The reaction

was started by the addition of the remaining components (according to manual

68

Page 69: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

from Roche) to a final volume of 20 µl and incubation at 42°C. The reaction was

stopped after 20 minutes by cooling the mix on ice. 4 µl of the reaction mix were

used to amplify the luciferase 3'-UTR in a subsequent PCR with primers EcoRI-

T(20) and LucTerm+. The obtained PCR product was sequenced directly with

LucTerm+.

2.4.4 poly(A) length determination by RNase H treatment For the poly(A)-tail length determination, 25 µg of total yeast RNA extract were

incubated with RNase H (Promega) and either a combination of

oligodeoxyribonucleotides LucRev1542 and EcoRI-T(20) or LucRev1542 alone

(25 pmol each) according to manufacturer's instructions. The reaction was then

denatured for 5 minutes at 95°C and separated on a 6% 8 M urea

polyacrylamide gel (40% solution with acrylamide : bis-acrylamide = 19:1) in 1x

TBE (1.78 M Tris, 1.78 M borate, 40 mM EDTA pH 8.0) at 300 V (Sambrook et

al., 1989). Next, the RNAs were electro-blotted in a Biorad Trans blot cell onto

Hybond-N membrane (Amersham Pharmacia Biotech) and hybridised with 32P

γ-labelled oligodeoxyribonucleotide PGK3UTRrev as described in 2.4.2.

2.5 Protein methods

2.5.1 Luciferase assay and protein quantification Extracts were prepared from cells grown in YNB medium to OD600 0.8-0.9

following the protocol in Oliveira et al. (1993b). 3 ml of each yeast culture were

pelleted by centrifugation (860 x g, 5 minutes) and washed twice with ice-cold

50 mM Tris.Cl pH 7.5. Resuspended in 300 µl of this buffer, the cells were

broken in three rounds of consecutive vortexing with glass beads (30 seconds)

and cooling on ice (1 minute). The cell debris was separated from the

supernatant by centrifugation at 4°C and 16,100 x g for 5 minutes. Luciferase

activity in these lysates was measured using standard procedures (de Wet et

al., 1987; Tatsumi et al., 1988). 10 µl of the supernatant was mixed with 350 µl

reaction buffer (25 mM glycylglycine pH 7.8, 5 mM ATP pH 7.5, 15 mM MgSO4)

in a test tube. Measurement of luminoscence was for 10 seconds and initiated

by the injection of a luciferin solution (0.2 mM D-luciferin in 25 mM glycylglycine

pH 7.8) in a luminometer (Berthold Lumat LB 9507).

69

Page 70: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

The total protein content per sample was estimated using a commercial kit

(Sigma) based on the biuret reaction (Smith et al., 1985). 100 µl of the sample

were mixed with 900 µl of the assay solution (50 volume parts BCA-Solution : 1

volume part 4% Copper(II)-sulphate in water) and incubated at 37°C for 30

minutes. The protein concentration was determined by comparing the OD562 of

this sample with the absorbance of standards prepared with known quantities of

BSA.

2.5.2 Western blotting For the purpose of Western blotting, 15 µg of proteins isolated by cell lysis with

glass beads (see 2.5.1) were separated by SDS polyacrylamide gel

electrophoresis (PAGE) according to Laemmli (1970). Separation was on a

12.5% polyacrylamide gel (30% solution with acrylamide : bis-acrylamide =

29:1) using the BioRad Mini Protean II system and a glycine buffer (25 mM

Tris.Cl pH 8.6, 192 mM glycine, 0.15% SDS). Prior, samples were denatured for

5 minutes at 95°C in sample buffer (final concentrations of 15.6 mM Tris.Cl pH

6.8, 5% glycerol, 0.4% SDS, 0.004% ß-mercaptoethanol, 0.001% bromophenol

blue). Runs were performed at 180 V until the bromophenol dye was 1 cm

above the bottom of the gel.

Transfer of proteins to a Hybond-P membrane (Amersham Pharmacia

Biotech) was in a Biorad semi-dry trans blot cell. Before, the membrane was

equilibrated for 10 minutes in methanol and transfer buffer (2.5 g/l glycine, 5.8

g/l Tris, 0.37 g/l SDS, 20% (v/v) methanol) each. Gel and membrane were

sandwiched between 4 sheets of Whatman paper (3M) on top and bottom; the

transfer took 30 minutes at 10 V and 400 mA.

Afterwards, the membrane was blocked by shaking in TBS (80 g/l NaCl,

2.2. g/l KCl, 18.2 g/l Tris pH 7.8) with 5% milk powder for 30 minutes. Incubation

with the rabbit anti-Flag antibody was in the cold room (4°C) over night in 2.5%

TBS. After three washes with TBS for 5 minutes at room temperature, the

second antibody (anti-rabbit conjugated with alkaline phosphatase) could bind

to the membrane in TBS for one hour at room temperature with shaking. The

membrane was washed as before and developed in the dark room with 10 ml

50 mM NaHCO3 pH 9.6, 33 µl 50 mg/ml BCIP (in 70% DMF) and 66 µl 50 mg/ml

NBT (in 100% DMF).

70

Page 71: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

2.6 Polysomal gradient analysis Plasticware, glassware and solutions (were applicable) were treated with DEPC

to guarantee RNA integrity throughout the analysis. In a procedure adopted

from Sagliocco et al. (1993), cells were grown in 150 ml YNB to an OD600 of 0.5-

0.6, at which translation was stopped by addition of cycloheximide to 100 µg/ml.

The culture was divided into 50 ml Falcon tubes and cooled on ice for 5

minutes. Then, cells were collected at 1,240 x g for 5 minutes, washed once

with 10 ml breaking buffer (10 mM Tris.Cl pH 7.4, 100 mM NaCl, 30 mM MgCl2,

100 µg/ml cycloheximide) and resuspended in 600 µl breaking buffer. The cells

were broken with an equal volume of glass beads in 6 rounds of 30 seconds

vortexing and 30 seconds cooling on ice. After separation of the cell debris by 5

minutes centrifugation at 16,100 x g and 4°C, the supernatant was layered onto

a 17%-47% sucrose gradient. This gradient was prepared in a Beckman

ultracentrifugation tube (No 331374) using a standard continuous gradient mixer

and 6 ml of each DEPC-treated 17% and 47% sucrose solution (in 50 mM

Tris.Cl pH 7.4, 50 mM NH4Cl, 12 mM MgCl2, 1 mM DTT). Centrifugation was for

3.5 hours at 270,000 x g in a SW 40Ti swing-out rotor.

The gradient was then immediately fractioned in a density gradient

fractionator (ISCO model 185) with a UV flow-through absorbance detector

(ISCO UA-5) using 50% sucrose (in DEPC-treated water). The RNA of 0.6 ml

fractions was instantly extracted in two rounds of

phenol:chloroform:isoamylacohol (25:24:1, v/v) extractions and precipitated with

0.3 M sodium acetate in ethanol. For mRNA detection in the respective

fractions, the complete RNA sample was subjected to the Northern blotting

procedure described in 2.4.1.

2.7 In vitro translation in cell-free yeast extracts Template preparation by plasmid digestion or PCR amplification

For transcription, the CsCl-purified pHST7-derived templates were digested with

NsiI according to manufacturer's instruction. After control for complete

restriction on a 1% agarose gel (Sambrook et al., 1989), the ends were filled up

with the Klenow polymerase fragment at 37°C for 30 minutes following the

producer's protocol. Purification from proteins was by phenol (pH 8.0),

phenol:chloroform:isoamylacohol (P:C:I = 25:24:1, v/v) and C:I (25:24, v/v)

71

Page 72: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

extraction. The DNA was precipitated with 0.3 M sodium acetate in ethanol,

salts removed by washing with 70% ethanol, and the dried DNA pellet was

resuspended to give a final concentration of approximately 1 µg/µl, which was

controlled by agarose electrophoresis and UV visualisation together with a DNA

sample of known concentration.

To obtain transcripts comprising the T7 promoter but lacking the 30

adenosine stretch of pHST7, templates were raised from plasmids pHST7-FL

and pHST7-FL-39-IRE by PCR amplification with primers pHS1-F and P3XM-R.

These products were purified from the original plasmids using the PCR

purification kit from Qiagen.

In vitro transcription

Templates were transcribed with T7 polymerase (NEBL) according to

manufacturer's specifications in the presence of cap-analogue m7GpppG for 2

hours at 37°C. After incubation with DNase and RNasin in compliance with

manufacturer's specifications, DNA degradation was controlled by agarose

electrophoresis.

The transcripts were separated from proteins and free nucleotides by

phenol/chloroform/isoamylacohol extractions and precipitation with ammonium

acetate followed by gel filtration (NAP-5 column, Pharmacia) and another round

of precipitations with ammonium acetate and sodium acetate.

Preparation of cell-free extracts

Cell-free translation extracts were prepared from diploid yeast strain W303

transformed with either the mock plasmid YCpSupex (Oliveira et al., 1993b) or

YCpSUP-IRP1. Preparation from cells grown in 2x 400 ml YNB galactose to

OD600 0.8 followed the general procedures described (Gasior et al., 1979;

Gerstel et al., 1992; Iizuka and Sarnow, 1997). The cells were cooled for 20

minutes on ice, collected at 1,240 x g for 5 minutes, and the cell pellet was

washed twice with ice-cold sterile water. In 50 ml tubes, the cells were then

washed three times in Buffer A (30 mM HEPES/KOH pH 7.4, 100 mM KAc, 2

mM MgAc) with 85 g/l mannitol by centrifugation as before. The cell pellet was

dried and resuspended in the cold room in 1.5 ml buffer A with mannitol for

each gramm cell pellet. Phenylmethyl-sulphonyl fluoride (PMSF) was added to

1 mM and glass beads (treated with HCl and extensively washed) to 3 g per g

72

Page 73: Translation reinitiation and ribosome recycling - PhD Thesis

2. Materials and Methods

pellet. The cells were broken manually in 5-6 cycles of 1 minute vigorous

shaking and 1 minute cooling on ice. The debris was separated by two

centrifugations at 4°C and 35,000 x g for 5 and 10 minutes, respectively. 4 ml of

the supernatant was loaded onto a GF25 column equilibrated with buffer A and

1 mM PMSF. The samples were collected in 1.5 - 2 ml fractions, assayed for

OD260/OD280, and translational activity was measured as described below. 80 µl

aliquots of active fractions were snap-frozen in liquid nitrogen and stored at -

80°C.

In vitro translation

In vitro translation was for 60 minutes at 25°C with 15 µl of cell-free extract and

the described amount of capped RNA in a final volume of 30 µl. The reaction

buffer contained final concentrations of 22 mM HEPES pH 7.4, 120 mM KOAc,

2 mM MgOAc, 0.75 mM ATP, 0.1 mM GTP, 25 mM creatine phosphate, 1.7 mM

DTT, 0.4 mg/ml creatine kinase, 1 U/µl Rnasin and 2 µl / 30 µl amino acid

mixture complete (Promega). The reaction was stopped by incubation on ice,

luciferase activity was measured in duplicate as described (see 2.5.1).

2.8 Reproducibility All experiments shown have been done at least three times with similar results,

and results of representative experiments are shown together with mean values

± standard deviation (STD). To correct the protein-content specific luciferase

activity (V1) for luc mRNA abundance (V2), the overall mRNA-specific translation

efficiency (V12) was calculated as V12 (± STD12) = V1 (± STD1) / V2 (± STD2) with

STD12 = ((STD1/V1)2+(STD2/V2)2))0.5.

73

Page 74: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

3. RESULTS

3.1 Ribosome recycling: 3'-UTR accessibility and translational efficiency

3.1.1 The potential role of 3'-UTR accessibility for ribosomal recycling on a circular mRNA

In the closed loop model of the polysome, the 5' and the 3'-ends of the mRNA

work together to enhance translational efficiency (see also Section 1.2.5). Due

to the vicinity of termination and initiation site, only interrupted by the 5'-UTR

and 3'-UTR, respectively, circularisation of the mRNA could facilitate the

intramolecular recycling of ribosomes. In one model, ribosomes are thought to

be guided back to the 5'-end via the bridging afforded by the proposed

interaction chain 3'-UTR – poly(A) tail – PABP – eIF4G. Blocking the access of

the ribosome to part or all of the 3'-UTR should interfere with posttermination

scanning and 'channelling' to the cap, and thus inhibit reinitiation within the

circular mRNP structure. We therefore devised experiments that would allow us

to modulate the behaviour of ribosomes downstream of termination.

3.1.2 Structures in the 3'-UTR can reduce translational efficiency As outlined in Section 1.2.4.4, the binding of IRP1 to IRE sequences is an ideal

and inducible system to establish a stable secondary structure on any position

within a transcript. Whereas the IRE itself only forms a very weak stem-loop and

can be resolved easily by scanning or translating ribosomes, IRP1-binding to

IRE is very efficient and forms a stable complex (Oliveira et al., 1993a).

Koloteva et al. (1997) showed this complex to impede efficiently ribosomal 40S

scanning along the mRNA when the IRE is located cap-distal. Moreover, stable

secondary structures (other than IRE-IRP1) have also been demonstrated to

block posttermination scanning 3' of an uORF (Vilela et al., 1999).

Previous studies (Koloteva et al., 1997; Oliveira et al., 1993a; Linz et al.,

1997), reported a 90% inhibition of translation upon IRP1-binding to an IRE

located 9 nt into the 5'-UTR of a luciferase reporter mRNA, most likely by

preventing the 43S preinitiation complex from binding to the cap structure. This

construct, FL-5'IRE (originally termed IRE-wt in Oliveira et al., 1993a; see also

table 2.5), was used as a control construct to ensure reproducibility of

74

Page 75: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

repression by IRP1 in the S. cerevisiae strain (diploid W303) and under the

growth conditions used (see Material and Methods). We employed the bipartite

system established by Oliveira et al., 1993a, whereby in a yeast double

transformant arrangement, the IRP1 ORF was placed under the control of an

galactose-inducible promoter (GAL1-PGK1 fusion promoter, GPF, Oliveira et

al., 1993b), and the IRE on a luciferase reporter mRNA that is expressed

constitutively by the TEF1 promoter (Figure 3.1A; Vega Laso et al., 1993).

To determine whether restriction of the 3'-UTR accessibility by IRE-IRP1

structures affects translational activity of a luciferase reporter mRNA in vivo,

constructs with an IRE sequence located downstream of the LUC termination

codon were assessed for their luciferase activity. In the onset of our work we

considered that inserting the IRE at several positions relative to the stop codon

might affect termination and posttermination behaviour of the ribosomes

differently (see Figure 3.1B, Table 2.5). In construct FL-15-IRE, the IRE stem-

loop is sufficiently close (15 nt) for the IRE-IRP1 complex to disturb a

terminating ribosome covering 16 nt downstream of the termination site as

shown by primer extension experiments in Wang et al. (1999). In contrast, the

mRNA of FL-24-IRE leaves enough space between termination site and IRE-

IRP1 for ribosomes to terminate properly, but migration into the 3'-UTR is

restricted. In FL-39-IRE, ribosomes could scan shortly along the 3'-UTR but

would soon be blocked 5' of the induced complex. In FL-141-IRE, the IRP1

target sequence was inserted in duplicate 141 nt downstream of the LUC stop

codon and 37 nt upstream of the poly(A) tail (Table 2.5 and Figure 3.1B).

Initially derived as a cloning by-product, the tandem IRE sequences are

expected to recruit two IRP molecules to the 3'-UTR. They are therefore

predicted to impose a very efficient blockage to potential 3'-UTR scanning when

ribosomes would almost have reached the poly(A) tail, potentially leading to

'queuing' of ribosomes (or subunits).

Taking luciferase activity of the IRE-deficient plasmid FL as a reference

value, co-expression of the human IRP1 gene led to differing degrees of

inhibition, depending on the position of the IRE. For easier comparison, the

luciferase values from cells harboured in either glucose or galactose (Figure

3.1C) are presented as GAL/GLU quotient (Figure 3.1D). Translational activity

is depicted as LUC values corrected for the protein content of the sample only

75

Page 76: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

Figure 3.1. Repression of translationvia IRP1-binding to the 3'-UTR. (A) TheYCpSUP-IRP1 plasmid contains thehuman IRP1 ORF under the control of thegalactose-inducible GAL-PGK fusionpromoter (PGPF). The YCp22FL1 (FL)plasmid carries the firefly luciferasereporter gene LUC downstream of theconstitutive TEF1 promoter PTEF1. A =AflII, B = BamHI, C = ClaI, E = EcoRI, H= HindIII, N = NdeI, Xb = XbaI, Xh = XhoI.(B) mRNA maps of FL-derived luciferaseconstructs containing an IRE hairpin-loopin either the 5' or 3'-UTR. Numbersindicate lengths in nucleotides, theindicated IRE stem-loop covers 29 nt. (C)Relative luciferase activities arepresented for both non-induced (glucose,GLU) and induced (galactose, GAL)conditions. The values are normalised toFL and corrected for the protein contentof the sample.

52%

69%

63%

100%

100% 10

%

46%

74%

72%

107%

100% 5%

0%

25%

50%

75%

100%

125%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

E

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for

108%

112%

102%

106%

97%

100%

97%

98%

93%

110%

99%

100%

0%

25%

50%

75%

100%

125%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

E

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for

co-transformed with mock plasmid Supex2

YCpSup-IRP1 CEN4 ARS1 URA3PGFP

BH C XbA H

IRP1 PGK1 terminator

Xh

YCp22FL1 CEN4 ARS1 TRP1LUCPTEF1

BC Xb/EA N H

PGK1 terminator

Xh

A

LUC An167

LUC An167

LUC37

LUC An

LUC An

LUC An167

FL-24-IRE

FL-39-IRE

FL-141-IRE

FL

FL-5’IRE

FL-15-IRE

B

E

D

F

FL FL-5

’IRE

FL-1

5-IR

E

FL-2

4-IR

E

FL-3

9-IR

E

FL-1

41-IR

E

glucoseLUC

PGK1

100±11

100±17

93±19

100±6

106±8

86±24

LUC/PGK1[%] ± STD

C

100% 87%

91%

81%

116%

92%

60%

101%

65%

114%

126% 11

%

0%

25%

50%

75%

100%

125%

150%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

ERel

ativ

e Lu

cife

rase

Act

ivity

co

rrec

ted

for p

rote

in c

onte

nt o

nly GLU

GAL

galactoseLUC

PGK1

100±8

94±20

208±23

86±16

84±21

133±12

LUC/PGK1[%] ± STD

An15

153

153

24

39

141 27

52%

69%

63%

100%

100% 10

%

46%

74%

72%

107%

100% 5%

0%

25%

50%

75%

100%

125%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

E

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for

108%

112%

102%

106%

97%

100%

97%

98%

93%

110%

99%

100%

0%

25%

50%

75%

100%

125%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

E

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for

108%

112%

102%

106%

97%

100%

97%

98%

93%

110%

99%

100%

0%

25%

50%

75%

100%

125%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

E

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for

co-transformed with mock plasmid Supex2

YCpSup-IRP1 CEN4 ARS1 URA3PGFP

BH C XbA H

IRP1 PGK1 terminator

Xh

CEN4 ARS1 URA3PGFP

BH C XbA H

IRP1 PGK1 terminator

Xh

YCp22FL1 CEN4 ARS1 TRP1LUCPTEF1

BC Xb/EA N H

PGK1 terminator

Xh

CEN4 ARS1 TRP1LUCPTEF1

BC Xb/EA N H

PGK1 terminator

Xh

A

LUCLUC An167

LUCLUC An167

LUCLUC37

LUC AnLUCLUC An

LUC AnLUCLUC An

LUCLUC An167

FL-24-IRE

FL-39-IRE

FL-141-IRE

FL

FL-5’IRE

FL-15-IRE

B

E

D

F

FL FL-5

’IRE

FL-1

5-IR

E

FL-2

4-IR

E

FL-3

9-IR

E

FL-1

41-IR

E

glucoseLUC

PGK1

100±11

100±17

93±19

100±6

106±8

86±24

100±11

100±17

93±19

100±6

106±8

86±24

LUC/PGK1[%] ± STD

C

100% 87%

91%

81%

116%

92%

60%

101%

65%

114%

126% 11

%

0%

25%

50%

75%

100%

125%

150%

FL

FL-5'IR

E

FL-15-IR

E

FL-24-IR

E

FL-39-IR

E

FL-141-IR

ERel

ativ

e Lu

cife

rase

Act

ivity

co

rrec

ted

for p

rote

in c

onte

nt o

nly GLU

GAL

galactoseLUC

PGK1

100±8

94±20

208±23

86±16

84±21

133±12

LUC/PGK1[%] ± STD

An15

153

153

24

39

141 27

(D) Ratios of luciferase activities with and without induction, corrected for either protein content of thesample only or for both protein content and the relative LUC mRNA abundance in the sample. (E) Therelative mRNA abundance data was derived from steady-state Northern blots probed for LUC and PGK1mRNAs. LUC abundance is expressed as LUC/PGK1 ratio, the averaged values are given in percent ±standard deviation and are normalised to FL. (F) Control experiment assaying luciferase activities of cellsco-transformed with the mock plasmid Supex2 expressing no ORF upon culturing in galactose. The valuesare normalised to FL and corrected for either protein content of the sample only or for both protein contentand the relative LUC mRNA abundance in the sample (Northern blot not shown).

76

Page 77: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

or for both the protein content and the LUC mRNA abundance. In accordance

with the respective publications (Koloteva et al., 1997: Linz et al., 1997; Oliveira

et al., 1993b) this illustrates the effect of transcript stabilisation upon mRNA-

specific translational efficiency. The difference between these two values

therefore allows locating the reason for the change in LUC activity, e.g. whether

a reduction of the transcript abundance causes an observed decrease in gene

expression.

The control construct FL-5'IRE showed a repression of about 95% upon

IRP1 expression after correction for protein and RNA levels (Figure 3.1D). This

degree of inhibition is in agreement with previous measurements using this

construct (Oliveira et al., 1993b) and therefore confirms the effectiveness of the

experimental set-up. Where the IRE was sufficiently close to the stop codon to

allow terminating ribosomes to destabilise this element's structure (FL-15-IRE),

there was no significant effect on translation (Figure 3.1D). In earlier work,

Niepel and colleagues (1999) found that positioning a very stable (43-base pair)

stem-loop 4 nucleotides downstream of the stop codon of a LUC mRNA

resulted in the loss of half of the measurable luciferase activity. However, these

results were generated from experiments with mRNA electroporated into yeast

spheroblasts, and are therefore not directly comparable with the ones presented

here. It is also important to note that the stem-loops used by these authors were

far more stable than the IRE structure, which were shown previously to be

readily destabilised by even a scanning yeast ribosome (Oliveira et al., 1993a).

In the case of FL-15-IRE, therefore, a terminating ribosome would be expected

to be readily capable of disrupting or occluding an IRE element, thus preventing

inhibitory effects that might be associated with IRP1-binding.

After correction for protein and mRNA abundance in the sample, 28% and

26% inhibition was measured when the IRE was moved further 3' (FL-24-IRE

and FL-39-IRE, respectively). The IRP1 binding sites of both constructs allow

proper termination, and the positioning of IRE 15 nt further downstream in FL-

39-IRE seems to be without additional effect on translation. Locating two IRE

elements close to the site of polyadenylation (FL-141-IRE) led to approximately

50% inhibition of translation upon IRP1-binding (Figure 3.1D). The latter

construct recruits two IRP proteins to the 3'-UTR and potentially establishes a

more stable structural barrier. However, the observed enhanced repression of

77

Page 78: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

translation is more likely to be position-dependent since a poly(G) tract at this

very position resulted in a comparable decrease in LUC activity (see below).

Growth in carbon sources different from glucose is known to change the

translational pattern in S. cerevisiae (Kuhn et al., 2001); moreover, IRP1-

binding to IRE could disturb protein interactions with the 3'-UTR. The first issue

was addressed by co-transforming cells either with a 'mock' plasmid (Supex2)

containing no inducible ORF. By comparing LUC expression after growth in

glucose and galactose, all constructs were found to be expressed to the same

extent and without change in mRNA stability (Figure 3.1F). In a different control

experiment, the IRE-IRP1 system and thus the necessity of harbouring the cells

in galactose was circumvented. For this, constructs with an intrinsic secondary

structure consisting of a 18 guanosine stretch close to the stop codon (FL-6-pG)

or further downstream (FL-131-pG) were created (Figure 3.2A). Muhlrad et al.

(1995) showed that this poly(G) tract inserted into the 5'-UTR can inhibit

translation and stabilise PGK1 mRNA. Indeed, a position-dependent inhibitory

effect similar to the one with IRE-bearing constructs was observed with these 3'-

UTR located poly(G) tracts, although repression was not as effective as with

binding of IRP1 to the mRNA (Figure 3.2B).

The luciferase assay results indicate that the insertion of a stable

obstruction in the 3'-UTR imposes a position-dependent effect on translation.

Whereas the terminating ribosome can obviously destabilise the IRE-IRP1

interaction, placing the complex further downstream reduces the translation

activity of the mRNA although the effect is not as pronounced as with a 5'-UTR

located structure. The effect is more evident when two IRP1 molecules bind to a

site close to the poly(A) tail. With reference to the lower extent of translational

repression, it can be concluded that 3'-UTR scanning is not an absolute

requirement for the posttermination ribosome to be able to recycle and

contribute to the overall translational efficiency of a specific mRNA. mRNA

abundance was calculated from Northern blots showing the steady-state levels

of the LUC transcript and the reference mRNA PGK1 (Figure 3.1E, Figure

3.2.C). Two-fold stabilisation of FL-5'IRE upon IRP1-binding was observed

previously by Koloteva et al. (1997) and Linz et al., (1997) and was interpreted

to be an effect of translational inhibition of the luciferase transcripts. The same

result has been seen with intrinsic structural elements e.g. stem-loops of the

78

Page 79: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

mRNA (Oliveira et al., 1993b). In agreement with this finding is the observation

that translational repression of the 3'-UTR modified constructs FL-131-pG and

FL-141-IRE was coupled with mRNA stabilisation. The smaller increase in

transcript stability – when compared to FL-5'IRE – also correlates to the lesser

extent with translation repression of the 3'-UTR modified constructs.

3.1.3 Recruitment of IRP1 to the 3'-UTR modulates translation and not polyadenylation

3.1.3.1 poly(A)-length and -site verification The insertion of the IRE sequences and poly(G) tracts was supplementary to

the 3'-UTR sequence, i.e. they did not replace any parts of the trailer that might

function as binding sites for 3'-UTR specific factors. Moreover, the insertion

sites were carefully planned to avoid disturbance of the poly(A) signals

(compare sequences in Table 2.5). However, changes in the structure of the 3'-

UTR such as by IRP1-binding could theoretically influence the state of

adenylation of the mRNA, which in turn could modulate translation initiation and

mRNA stability. Therefore, I examined the average length of the poly(A) tails on

the IRE-bearing mRNAs from cell extracts with IRP1 being co-expressed.

LUC An

LUC An131 ntG18

LUC An6 ntG18

FL

FL-6-pG

FL-131-pG

C

LUCPGK1

FL FL-6

-pG

FL-1

31-p

G

92±8

100±7

LUC/PGK1[%] ± STD

117±9

A72

%

98%

100%

62%

107%

100%

0%

25%

50%

75%

100%

125%

FL

FL-6-pG

FL-131-pG

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content onlyprotein and RNA content

corrected forBLUC An

LUC An131 ntG18

LUC An6 ntG18

FL

FL-6-pG

FL-131-pG

LUC An

LUC An131 ntG18

LUC An6 ntG18

LUC AnLUC An

LUC An131 ntG18LUC An131 ntG18G18

LUC An6 ntG18LUC An6 ntG18G18

FL

FL-6-pG

FL-131-pG

C

LUCPGK1

FL FL-6

-pG

FL-1

31-p

G

92±8

100±7

LUC/PGK1[%] ± STD

117±9

LUCPGK1

FL FL-6

-pG

FL-1

31-p

G

92±8

100±7

LUC/PGK1[%] ± STD

117±9

A72

%

98%

100%

62%

107%

100%

0%

25%

50%

75%

100%

125%

FL

FL-6-pG

FL-131-pG

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content onlyprotein and RNA content

corrected for

72%

98%

100%

62%

107%

100%

0%

25%

50%

75%

100%

125%

FL

FL-6-pG

FL-131-pG

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content onlyprotein and RNA content

corrected forB

Figure 3.2. Position-dependent decrease of translational efficiency by a 3'-UTR-locatedpoly(G) tract. (A) mRNA maps of FL and derived constructs carrying a poly(G) stretch in the 3'-UTR. (B) Relative luciferase activities of poly(G)-bearing constructs, normalised to FL andcorrected for either protein content of the sample only, or for both protein content and the relativeLUC mRNA abundance in the sample. (C) Typical steady-state Northern blot of extracts fromstrains carrying the respective constructs, probed for LUC and PGK1 mRNAs. The averagedvalues for relative LUC mRNA abundance are given under the respective lanes of the Northernblot.

79

Page 80: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

RNase H treatment of mRNAs extracted from cells expressing IRP1 was used

to cause cleavage either at the end of the LUC ORF alone, or at the end of the

LUC ORF and within the poly(A) tail simultaneously. The size of the two 3'

fragments was discernible after separation on a 6% polyacrylamide gel and

corresponds to the average in vivo poly(A) length of each mRNA. In a Northern

blot hybridised with a 3'-UTR targeted 32P-labelled oligodeoxyribonucleotide

A

LUC AAAAAn

poly(T)

LucRev

200

500

400

300

600

+ +

FL+

+ +

FL-1

5-IR

E+

+ +

FL-2

4-IR

E+

+ +

FL-3

9-IR

E+

+ +

FL-1

41-IR

E+LucRev

poly(T)M

in nt

poly(A) tailpoly(A) site3'UTRB

AAAAAn

3'UTR only

3'UTR + poly(A)

A

LUC AAAAAnLUC AAAAAn

poly(T)

LucRev

200

500

400

300

600

+ +

FL+

+ +

FL-1

5-IR

E+

+ +

FL-2

4-IR

E+

+ +

FL-3

9-IR

E+

+ +

FL-1

41-IR

E+LucRev

poly(T)M

in nt

200

500

400

300

600

200

500

400

300

600

+ +

FL+

+ +

FL-1

5-IR

E+

+ +

FL-2

4-IR

E+

+ +

FL-3

9-IR

E+

+ +

FL-1

41-IR

E+LucRev

poly(T)M

in nt

poly(A) tailpoly(A) site3'UTRB

AAAAAn

3'UTR only

3'UTR + poly(A)

Figure 3.3. Binding of IRP1 to an IRE located in the 3'-UTR does not affectpolyadenylation. (A) Total RNA extracted from each yeast strain co-expressing the IRP1 geneand one of the LUC constructs was subjected to RNase H treatment in the presence ofspecifically targeted oligodeoxyribonucleotides. Incubation with the oligodeoxyribonucleotideLucRev led to digestion upstream of the LUC stop-codon, while the further addition of theoligodeoxyribonucleotide poly(T) caused degradation of the poly(A) tail as well. The lengthdifference of the single (white arrows) and double digestion products (black arrows) detected inNorthern blots by probing for a 3'-UTR sequence corresponds to the poly(A) tail length. Thelengths (in nucleotides) of RNA markers are indicated on the right hand side. (B) Dideoxyfluorescent derivative sequencing of RT-PCR products confirms that the wild typepolyadenylation site is utilised in all of the IRE-bearing constructs. The representative sequenceshown corresponds to the polyadenylated 3'-UTR of construct FL-39-IRE, co-expressed in thecell with IRP1.

80

Page 81: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

comparison with IRE-free construct FL revealed that neither IRE-insertion nor

IRP1-binding to its target sequence affected the distribution of poly(A) tail

lengths at any of the positions examined (Figure 3.3A). The average poly(A)

length of about 70 nt observed correlates to those published for yeast mRNAs

(Brown & Sachs, 1998).

Modifications of the 3'-UTR sequence could not only affect poly(A) tail

length but also change the choice of the poly(A) site (Düvel et al., 1999). Thus,

LUC mRNAs extracted from cells expressing IRP1 were subjected to first strand

synthesis and subsequent PCR amplification of their 3'-UTR. Evidence was

then obtained via sequencing of the RT-PCR products that utilisation of the wild

type polyadenylation site was unaffected (Figure 3.3B).

3.1.3.2 Polysomal analysis Efficiently translated mRNAs are bound to more than one ribosome in a

structure called the polysome. For further investigation of the change in the

overall translational state conferred by IRP1-binding to the 3'-UTR, reference

construct FL, control construct FL-5'IRE and two mRNAs with the IRE in the 3'-

UTR – FL-39-IRE and FL-141-IRE – were subjected to sucrose gradient

analysis. As described in Material and Methods, extracts of cells expressing the

respective construct in the presence of IRP1 were separated by centrifugation

in a sucrose gradient, which was analysed thereafter for absorbance at 260 nm.

From the resulting graph (Figure 3.4A), the peaks corresponding to 40S (43S),

80S, di-, tri- and further polysomes can be derived.

As indicated by the representative polysomal profile of construct FL-141-

IRE, overall translation in strain W303 is not impaired by the growth in

galactose. Most ribosomes are not dissociated in 40S and 60S subunits but

present in one or more copies per mRNA. The respective sucrose fractions

were then assessed in Northern blots for their LUC and PGK1 mRNA

abundance (Figure 3.4B). For better visualisation, the LUC/PGK1 mRNA

abundance ratios of these blots have been normalised to the values of the

respective 80S peak and are depicted as bars in Figure 3.4C. In the Northern

blot of construct FL, most of the LUC mRNA is detectable in the polysomal

fractions, although PGK1 is translated even more efficiently (Figures 3.4B,C).

The substantial shift of FL-5'IRE mRNA towards the 40S / 43S and monosomal

fractions can be interpreted as a result of IRP1-binding to an IRE close to the

81

Page 82: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

cap. The strong inhibition of 80S formation and therefore, translation initiation is

in agreement with observations by Koloteva et al. (1997) and Linz et al. (1997).

With IRP1 being co-expressed, the major part of both, FL-39-IRE and FL-

141-IRE mRNAs, was localised in the lower polysomal or 80S fraction (Figure

3.4B,C). Moreover, the abundance of transcripts in the polysomal fractions

matches the degree of IRP-conferred repression of luciferase activity observed

for the individual constructs (FL-39-IRE: 74%, FL-141-IRE: 47% as in Figure

fraction

5‘IRE

FL-141-IRE

FL-39-IRE

FL

B101 2 3 4 5 6 7 8 9

PGK1

LUC

PGK1

LUC

PGK1

LUC

PGK1

LUC

A

–40

S

–60

S–

80S

OD

260

––––

–43

S

polysomes

1 2 3 4 5 6 7 8 9 10

FL-141-IRE

FL-39-IRE

FL-5'IREFL

0%

100%

200%

300%

400%

Rel

ativ

e LU

C/P

GK

1 m

RN

AA

bund

ance

polysomal fraction

C

fraction

5‘IRE

FL-141-IRE

FL-39-IRE

FL

B101 2 3 4 5 6 7 8 9 101 2 3 4 5 6 7 8 9

PGK1

LUC

PGK1

LUC

PGK1

LUC

PGK1

LUC

A

–40

S

–60

S–

80S

OD

260

––––

–43

S

polysomes

–40

S

–60

S–

80S

OD

260

––––

–43

S

polysomes

1 2 3 4 5 6 7 8 9 10

FL-141-IRE

FL-39-IRE

FL-5'IREFL

0%

100%

200%

300%

400%

Rel

ativ

e LU

C/P

GK

1 m

RN

AA

bund

ance

polysomal fraction

C

Figure 3.4. IRP1-binding toIREs in the 5' or 3'-UTR affectsthe polysomal distribution ofLUC mRNA. Sucrose gradientanalysis was performed onextracts from cells containing aLUC reporter construct afterinduction of IRP1 synthesis for12 hours in galactose medium.(A) The diagram shows thefractional distribution of opticaldensity at 260 nm in a sucrosegradient. The representativeprofile was derived from cell-extracts containing FL-141-IRE.Fractions corresponding tomRNA bound to 40S/43S, 60S or80S ribosomes are indicated. (B)Northern blots of the numberedfractions as indicated, theluciferase construct probed for isspecified alongside each panel.Blots were co-hybridised with aPGK1-specific probe to revealthe polysomal distribution of achromosomally encoded controlmRNA. (C) Diagram illustratingthe relative abundance of LUC toPGK1 mRNA among thepolysomal fractions. For bettercomparison, the values arenormalised to the magnitude ofthe respective 80S peak (fraction/ lane 4).

82

Page 83: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

3.1D). The reference mRNA PGK1 is present in the high-molecular fractions for

all the luciferase constructs tested (Figure 3.4B).

These alterations of the polysomal profile can be taken as additional

evidence that IRP1 recruitment downstream of the stop codon changes the

function of the mRNA's translational status in a position-dependent manner.

3.1.4 Translation of a full-length ORF is insensitive to changes of the codon context

As discussed in Section 1.4, translation termination at a short uORF is sensitive

to the stop codon context (Grant & Hinnebusch, 1994a). By modulation of

ribosome release from the termination site, the necessity of downstream

scanning for efficient LUC translation can be investigated. Accordingly, the

FL

XhoI -2 -1 stop XbaIC TCG AGC AAA TTG TAA TCTAGAC TCG AGC AAA CCG TAA TCTAGAC TCG AGC CCG TTG TAA TCTAGAC TCG AGC CCG CCG TAA GCG GTT ACC TCTAGAC TCG AGC AAA TTG TAA GCG GTT ACC TCTAGAC TCG AGC AAA TTG TAA TAA TCTAGA

LUC An

C

A

FL*

-1Pr

o

-2Pr

o

-1,2

Pro1

0ds

10ds

stop

stop

LUC

PGK1

112±9

100±5

LUC/PGK1[%] ± STD

86±12

98±10

89±9

96±6

FL*-1Pro-2Pro

-1,2Pro10ds10ds

stopstop

101%

95%

103%

94%

100%

98%

100%

109%

88%

105%

105%

106%

0%

25%

50%

75%

100%

125%

FL*

-1Pro

-2Pro

-1,2P

ro10

ds10

ds

stops

top

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content only protein and RNA contentcorrected forB

FL

XhoI -2 -1 stop XbaIC TCG AGC AAA TTG TAA TCTAGAC TCG AGC AAA CCG TAA TCTAGAC TCG AGC CCG TTG TAA TCTAGAC TCG AGC CCG CCG TAA GCG GTT ACC TCTAGAC TCG AGC AAA TTG TAA GCG GTT ACC TCTAGAC TCG AGC AAA TTG TAA TAA TCTAGA

LUC AnLUC An

C

A

FL*

-1Pr

o

-2Pr

o

-1,2

Pro1

0ds

10ds

stop

stop

LUC

PGK1

112±9

100±5

LUC/PGK1[%] ± STD

86±12

98±10

89±9

96±6

FL*

-1Pr

o

-2Pr

o

-1,2

Pro1

0ds

10ds

stop

stop

LUC

PGK1

FL*

-1Pr

o

-2Pr

o

-1,2

Pro1

0ds

10ds

stop

stop

LUC

PGK1

112±9

100±5

LUC/PGK1[%] ± STD

86±12

98±10

89±9

96±6

FL*-1Pro-2Pro

-1,2Pro10ds10ds

stopstop

101%

95%

103%

94%

100%

98%

100%

109%

88%

105%

105%

106%

0%

25%

50%

75%

100%

125%

FL*

-1Pro

-2Pro

-1,2P

ro10

ds10

ds

stops

top

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content only protein and RNA contentcorrected for

101%

95%

103%

94%

100%

98%

100%

109%

88%

105%

105%

106%

0%

25%

50%

75%

100%

125%

FL*

-1Pro

-2Pro

-1,2P

ro10

ds10

ds

stops

top

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content only protein and RNA contentcorrected forB

Figure 3.5. The stop codon context of the full-length luciferase ORF does not affecttranslational efficiency or mRNA stability. (A) mRNA map of FL-derived construct FL*containing an XhoI restriction site upstream of the stop codon facilitating the change of the stopcodon context. Modifications took the form of mutations of codons at positions -1 and -2upstream of the LUC stop codon to CCG (proline), the introduction of the 10 nucleotidedownstream sequence from GCN4 uORF4, or the introduction of an additional stop codon. (B)Relative luciferase activities of stop codon context mutants, normalised to FL* and corrected foreither protein content of the sample only or for both protein content and the relative LUC mRNAabundance in the sample. (C) Typical steady-state Northern blot of FL* and derivatives, probedfor LUC and PGK1 mRNAs. The averaged values for relative LUC mRNA abundance are givenunder the respective lanes of the Northern blot.

83

Page 84: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

effects of changing the stop codon context so as to maximise termination

efficiency were explored.

One approach was to introduce CCG codons encoding proline into the last

two positions of the coding sequence, since these have been shown to promote

efficient translation termination (Figure 3.5A). Introduction of these codons

suppresses read-through (Mottagui-Tabar et al., 1998), most likely by affecting

peptidyl-tRNA hydrolysis rates. In a further approach, an additional stop codon

immediately after the normal luciferase gene stop codon provided the ribosome

with a second opportunity to recognise a site of peptide chain termination. The

combination of two stop codons can efficiently suppress read-through in E. coli

(Major et al., 2002). Finally, in a further construct the ten nucleotides

downstream of the stop codon were changed to those of GCN4 uORF4. The

latter downstream context is known to suppress the reinitiation competence of

terminating ribosomes (Grant and Hinnebusch, 1994a), even in the context of a

different mRNA (Vilela et al., 1999).

LUC measurements and Northern blots of mRNA steady-state levels

revealed that none of these approaches affected the specific translation rates or

the stabilities of the LUC mRNA significantly (Figures 3.5B,C). This is in

contrast to termination at a short uORF such as the ones in the GCN4 leader.

Termination there is sensitive to the stop codon context, and the G/C ratio has

been shown to modulate the balance of reinitiation and ribosomal release

(Grant and Hinnebusch, 1994a). One way to explain the observed results is to

assume that the ribosome terminating after translation of a long ORF differs

from those at the stop codon of a short uORF in its sensitivity towards the stop

codon context. Although e.g. the ribosomal reinitiation capability depends on the

ORF length (see below), there is no direct evidence supporting this assumption.

Notably, as described above, also recruitment of IRP1 to an IRE

immediately 3' of the termination site did not affect translational efficiency or

mRNA stability (see Figure 3.1D, construct FL-15-IRE). If, as it seems likely,

this and the above measures both promote release of ribosomal subunits from

the LUC mRNA, the results suggest that progression of ribosomes beyond the

stop codon into the 3'-UTR is not required for normal recycling to occur.

Alternatively, enhanced recruitment of ribosomal subunits from the free cellular

pool can compensate a (moderate) loss of intramolecular recycled ribosomes.

84

Page 85: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

3.1.5 The eIF4G – Pab1p interaction is not required for 3'-UTR-confered translation repression in vivo

3.1.5.1 The absence of the cap – poly(A) interaction in vivo does not abolish the translational inhibition by a 3'-UTR structure

Circularisation of mRNA by the eIF4G – Pab1p interaction requires the

presence of a poly(A) tail (Tarun and Sachs, 1996). If translational efficiency is

controlled by ribosomal 'channelling' via 3'-UTR scanning and subsequent

transfer via mentioned 5' – 3' end interaction, the disruption of the circularisation

would be expected to abolish the negative effect of 3'-UTR inaccessibility. I

therefore constructed strains, in which the impact of blocking the 3'-UTR could

be assessed against a cellular background that is defective in the eIF4G –

Pab1p interaction. For this, a mutated version of eIF4G1 (tif4631-213) in which

the binding site for Pab1p is no longer functional was obtained (Tarun et al.,

1997). This mutant was expressed from a plasmid that had been introduced into

a strain (YAS1947) in which the genes encoding eIF4G1 and eIF4G2 have both

been disrupted. Luciferase measurements revealed that despite the absence of

the eIF4G – Pab1p interaction, IRP1-binding to an IRE in the 3'-UTR still

partially repressed translation (BEFL-39-IRE; Figure 3.6A,B). The degree of

inhibition (18%) was somewhat less than that observed in the wild type

background (26%, Figure 3.1D). However, it was also observed that the degree

of inhibition caused by IRP1-binding to an IRE in the 5'-UTR was reduced in the

YAS1947 background (BEFL-5'IRE; 73% versus 95% for FL-5'IRE), which

means that the change in inhibitory potential of the IRE/IRP1 complex was not

specific to the 3'-UTR. This demonstrates that the eIF4G – Pab1p interaction is

not required for IRE-mediated modulation of translation via the 3'-UTR in vivo.

3.1.5.2 In vitro translation in cell-free extracts is not affected by IRP1-binding to the 3'-UTR

In an alternative approach to the above in vivo system, a cell-free translation

assay offered the possibility to determine whether the influence of IRE-mediated

blocking of the 3'-UTR was reproducible in vitro and could be changed by the

absence of the poly(A) tail. For this, translation competent cell extracts were

prepared from yeast cells either expressing the mock plasmid Supex2 or

YCpSup-IRP1. Aliquots were then programmed with 0-1 µg mRNA

corresponding to constructs FL, FL-5'IRE and FL-39-IRE.

85

Page 86: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

The analysis revealed, as reported previously (Gerstel et al., 1992; Niepel

et al., 1999; Iizuka et al., 1994), both capping (data not shown) and

polyadenylation (Figure 3.6C,D) enhanced translation of the reporter mRNA in

the cell-free system. Moreover, there was synergy between the effects of these

two modifications, so that the combined effects resulted in a sum effect that was

greater than the arithmetic sum of the individual contributions (compare Sachs,

2000). The enhanced translational activity of polyadenylated 5'-IRE mRNA in

cell extracts without IRP1 presence can be explained by the elongation of the

leader due to IRE insertion (Figure 3.6C).

97%

64%

100%

82%

27%

100%

0%

25%

50%

75%

100%

125%

BEFL

BEFL-5'IRE

BEFL-39-IRE

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for C

B

LUC

BEF

L

BEF

L-5’

IRE

BEF

L-39

IRE

BEF

L

BEF

L-5’

IRE

BEF

L-39

IRE

PGK1

galactoseglucose

110±10

100±12

LUC/PGK1[%] ± STD

123±17

133±12

91±8

315±27

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.00

FL FLpA 39IRE 39IREpA 5'IREpA

Rel

ativ

e Lu

cife

rase

Act

ivity

104 R

LU u

nits

6

5

4

3

2

1

0

D

FLFLpA 39IRE 39IREpA 5'IREpA

Rel

ativ

e Lu

cife

rase

Act

ivity

106 R

LU u

nits

5

4

3

2

1

0

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.00

D

A

97%

64%

100%

82%

27%

100%

0%

25%

50%

75%

100%

125%

BEFL

BEFL-5'IRE

BEFL-39-IRE

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for

97%

64%

100%

82%

27%

100%

0%

25%

50%

75%

100%

125%

BEFL

BEFL-5'IRE

BEFL-39-IRE

Rel

ativ

e Lu

cife

rase

Act

ivity

GAL

/GLU

protein content onlyprotein and RNA content

corrected for C

B

LUC

BEF

L

BEF

L-5’

IRE

BEF

L-39

IRE

BEF

L

BEF

L-5’

IRE

BEF

L-39

IRE

PGK1

galactoseglucose

110±10

100±12

LUC/PGK1[%] ± STD

123±17

133±12

91±8

315±27

LUC

BEF

L

BEF

L-5’

IRE

BEF

L-39

IRE

BEF

L

BEF

L-5’

IRE

BEF

L-39

IRE

PGK1

galactoseglucose

110±10

100±12

LUC/PGK1[%] ± STD

123±17

133±12

91±8

315±27

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.00

FL FLpA 39IRE 39IREpA 5'IREpA

Rel

ativ

e Lu

cife

rase

Act

ivity

104 R

LU u

nits

6

5

4

3

2

1

0

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.00

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.000.00 0.25 0.50 0.75 1.00

FL FLpA 39IRE 39IREpA 5'IREpA

FL FLpA 39IRE 39IREpA 5'IREpA

Rel

ativ

e Lu

cife

rase

Act

ivity

104 R

LU u

nits

6

5

4

3

2

1

0

Rel

ativ

e Lu

cife

rase

Act

ivity

104 R

LU u

nits

6

5

4

3

2

1

0

D

FLFLpA 39IRE 39IREpA 5'IREpA

Rel

ativ

e Lu

cife

rase

Act

ivity

106 R

LU u

nits

5

4

3

2

1

0

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.00

D

FLFLpA 39IRE 39IREpA 5'IREpA

D

FLFLpA 39IRE 39IREpA 5'IREpA

Rel

ativ

e Lu

cife

rase

Act

ivity

106 R

LU u

nits

5

4

3

2

1

0

Rel

ativ

e Lu

cife

rase

Act

ivity

106 R

LU u

nits

5

4

3

2

1

0

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.00

capped luciferase RNA (µg)0.00 0.25 0.50 0.75 1.000.00 0.25 0.50 0.75 1.00

D

A

Figure 3.6. Significance of 5' – 3' interactions to recycling. (A) Ratios of luciferase activitieswith and without induction, corrected for either protein content of the sample only or for bothprotein content and the relative LUC mRNA abundance in the sample. The samples wereextracted from YAS1947 cells (expressing eIF4GI-213 mutated in its Pab1p-binding) and co-transformed with YCpSUP-IRP1 and the target IRE constructs. Growth was under either non-induced (glucose, GLU) or induced (galactose, GAL) conditions. The BEFL plasmids, whichcarry an ADE2 marker, correspond to their counterparts in the FL series that have a TRP1marker. (B) Northern blots probed for LUC and PGK1 mRNAs. (C,D) In vitro transcribed andcapped mRNAs were translated in extracts derived from cells containing either (C) the controlplasmid YCpSupex without an inducible ORF or (D) YCpSUP-IRP1. The luciferase activities ofthe in vitro translation reactions are presented as titrations of added mRNA over the range 0-1µg. Open symbols represent translation data from non-polyadenylated mRNAs, while the fullsymbols correspond to mRNAs with a 30 nucleotide poly(A) tail. Bars indicate standarddeviations.

86

Page 87: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

However, one particular observation strongly influenced the interpretation

of the effects of downstream elements on translation in vivo. The IRP1/IRE

interaction in the 5'-UTR clearly inhibits translation as effectively as in vivo, yet

has no effect on in vitro translation in the 3'-UTR position, irrespective of the

presence or absence of a poly(A) tail (compare Figures 3.6C,D). It was also

observed that primarily 80S monosomes were formed on the reporter mRNA in

the cell-free system (private correspondence, Karine Berthelot). It follows that

the inhibitory effect of the IRP1/IRE interaction occurring on the 3'-UTR is

therefore restricted to polysomal structures in vivo.

3.2 Correlation of ORF length and ribosome reinitiation potential

3.2.1 Reinitiation capability increases as a wild type ORF is progressively truncated

Tightly connected with our initial question – whether ribosomes entering the 3'-

UTR contribute to the overall translational mRNA efficiency – is the definition of

the circumstances affecting posttermination reinitiation competence. Initially,

analysis of translational control on the GCN4 mRNA of S. cerevisiae

demonstrated that termination on a short ORF can allow ribosomes to retain

their ability to initiate translation downstream of the stop codon was (reviewed in

Hinnebusch, 1997), and this holds true for other uORFs (upstream ORFs;

McCarthy, 1998; Geballe & Sachs, 2000). It therefore can be derived that

termination on a short ORF leaves at least the 40S ribosomal subunit in a state

that must be distinct from the state that normally results from termination on the

main ORFs of genes. Since uORFs and wild type ORFs differ primarily in their

length, we needed to consider the effects of this attribute on the ribosomal

reinitiation capability.

However, only few studies have explored the effect of varying the uORF

length (Luukkonen et al., 1995; Hwang et al., 1998; Kozak, 2001), and this

research was done in mammalian and virus systems. A systematic investigation

into the relationship of reinitiation capacity and length of the upstream ORF in

yeast could not be found in the published literature. In a first step towards filling

this gap and thereby elucidating the basis of the transition from (re)initiation-

competent to (re)initiation-incompetent ribosomes that seems to occur during

87

Page 88: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

elongation, we studied the dependence of reinitiation capacity on the length of

the upstream ORF in a bicistronic mRNA.

For this, a FLAG-tagged, full-length actin ORF was inserted out-of-frame

upstream of luciferase in plasmid FL'. The 59 nucleotide intercistronic region

comprises the A/T-rich sequence 5'-TGTAGAATTCTAGA-3' followed by the

GCN4 uORF1 downstream sequence (see Table 2.5). The latter has been

shown to promote reinitiation on the GCN4 mRNA, probably due to its A/T

content (Grant & Hinnebusch, 1994b). Bicistronic ACT-LUC mRNAs with 3'-

truncated versions of the actin ORF were created in order to shorten the ORF to

lengths of 963 to 15 nucleotides (giving the ACTd-LUC constructs, Figure 3.7A).

The truncation procedure (see Materials and Methods section) left the

sequence context at the start codon identical for all constructs as well as the

sequence 3' of the ACT stop codon (compare Table 2.5).

As expected, downstream of a wild type full-length ORF, posttermination

initiation at the LUC ORF, and therefore LUC activity, was at a minimal level

(Figure 3.7B). Reinitiation only became readily detectably with a truncated actin

ORF of 105 nucleotides or less, whereby efficiency increased markedly with

first-ORF lengths of 63 nucleotides and shorter (Figure 3.7B). Luciferase

activities were corrected for mRNA abundance and protein, revealing a

maximum reinitiation-dependent level of luciferase (encoded by ACTd15-LUC)

equivalent to 5% of the monocistronic luciferase construct FL' (Figure 3.7B;

Table 2.5). This means that even the shortest ACT-derived ORF encoding only

four amino acids is relatively poor at promoting reinitiation on the LUC ORF.

Translation of the first ORF was confirmed by detection of the full-length FLAG-

tagged actin protein in a Western blot (Figure 3.7D). Truncated FLAG-actin

products were not readily quantifiable because of partial degradation.Constructs

that showed luciferase activity generated by reinitiation were also destabilised;

the mRNA abundance of constructs with an actin ORF of 105 nt or less was

reduced by up to 73% relative to the full-length control (Figure 3.7C). Thus,

translation termination at the ends of these shortened first ORFs partially

destabilises the mRNA, although the degree of destabilisation is smaller than

that normally associated with the nonsense-mediated decay pathway (Zhang et

al., 1995). Similar effects of uORFs on the stability of mRNAs have been

reported for other uORF-containing 5'-UTRs (Ruiz- Echevarria & Peltz, 2000;

88

Page 89: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

89

0.00

1%

0.02

2%

0.02

6%

0.02

0%

0.10

7%

0.00

1%

0.02

4%

0.02

6%

0.02

2%

0.28

4% 0.75

1%

0.60

4% 1.34

7%5.

006%

1.35

7% 2.00

9%

0%

1%

2%

3%

4%

5%

6%

7%

ACT-LUC

ACTd963

-LUC

ACTd672

-LUC

ACTd327

-LUC

ACTd105

-LUC

ACTd63-L

UC

ACTd33-L

UC

ACTd15-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content onlyprotein and RNA content

corrected forB

A

C D

FL'

AC

T-LU

C

AC

Td96

3-LU

C

AC

Td67

2-LU

C

AC

Td32

7-LU

C

AC

Td10

5-LU

C

AC

Td63

-LU

C

AC

Td33

-LU

C

AC

Td15

-LU

C

ACT-LUC

PGK1 LUC

84±10

100±5

LUC/PGK1[%] ± STD

103±5

38±3

95±5

88±7

45±5

37±6

27±4

ACT-LUC

ACTd963-LUC

ACTd672-LUC

ACTd327-LUC

ACTd105-LUC

ACTd63-LUC

ACTd33-LUC

ACTd15-LUC

FL'

An

63

963

An

AnFLAG-actin luciferase170 1157 165359 153

672An

An

105An

33An

327An

15An

32.5

16.5

47.5

25.0

Act1p

kDa

AC

T-LU

CA

CTd

963-

LUC

AC

Td67

2-LU

CA

CTd

327-

LUC

AC

Td10

5-LU

CA

CTd

63-L

UC

AC

Td33

-LU

CA

CTd

15-L

UC

0.00

1%

0.02

2%

0.02

6%

0.02

0%

0.10

7%

0.00

1%

0.02

4%

0.02

6%

0.02

2%

0.28

4% 0.75

1%

0.60

4% 1.34

7%5.

006%

1.35

7% 2.00

9%

0%

1%

2%

3%

4%

5%

6%

7%

ACT-LUC

ACTd963

-LUC

ACTd672

-LUC

ACTd327

-LUC

ACTd105

-LUC

ACTd63-L

UC

ACTd33-L

UC

ACTd15-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content onlyprotein and RNA content

corrected for

0.00

1%

0.02

2%

0.02

6%

0.02

0%

0.10

7%

0.00

1%

0.02

4%

0.02

6%

0.02

2%

0.28

4% 0.75

1%

0.60

4% 1.34

7%5.

006%

1.35

7% 2.00

9%

0%

1%

2%

3%

4%

5%

6%

7%

ACT-LUC

ACTd963

-LUC

ACTd672

-LUC

ACTd327

-LUC

ACTd105

-LUC

ACTd63-L

UC

ACTd33-L

UC

ACTd15-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

protein content onlyprotein and RNA content

corrected forB

A

C D

FL'

AC

T-LU

C

AC

Td96

3-LU

C

AC

Td67

2-LU

C

AC

Td32

7-LU

C

AC

Td10

5-LU

C

AC

Td63

-LU

C

AC

Td33

-LU

C

AC

Td15

-LU

C

ACT-LUC

PGK1 LUC

84±10

100±5

LUC/PGK1[%] ± STD

103±5

38±3

95±5

88±7

45±5

37±6

27±4

FL'

AC

T-LU

C

AC

Td96

3-LU

C

AC

Td67

2-LU

C

AC

Td32

7-LU

C

AC

Td10

5-LU

C

AC

Td63

-LU

C

AC

Td33

-LU

C

AC

Td15

-LU

C

ACT-LUC

PGK1 LUC

84±10

100±5

LUC/PGK1[%] ± STD

103±5

38±3

95±5

88±7

45±5

37±6

27±4

ACT-LUC

ACTd963-LUC

ACTd672-LUC

ACTd327-LUC

ACTd105-LUC

ACTd63-LUC

ACTd33-LUC

ACTd15-LUC

FL'

An

63

963

An

AnFLAG-actin luciferase170 1157 165359 153

672An

An

105An

33An

327An

15An

ACT-LUC

ACTd963-LUC

ACTd672-LUC

ACTd327-LUC

ACTd105-LUC

ACTd63-LUC

ACTd33-LUC

ACTd15-LUC

FL'

An

63

963

An

AnFLAG-actin luciferase170 1157 165359 153

672An

An

105An

33An

327An

15An

32.5

16.5

47.5

25.0

Act1p

kDa

AC

T-LU

CA

CTd

963-

LUC

AC

Td67

2-LU

CA

CTd

327-

LUC

AC

Td10

5-LU

CA

CTd

63-L

UC

AC

Td33

-LU

CA

CTd

15-L

UC

32.5

16.5

47.5

25.0

Act1p

kDa

32.5

16.5

47.5

25.0

Act1p

kDa

AC

T-LU

CA

CTd

963-

LUC

AC

Td67

2-LU

CA

CTd

327-

LUC

AC

Td10

5-LU

CA

CTd

63-L

UC

AC

Td33

-LU

CA

CTd

15-L

UC

Figure 3.7. Truncation of the first ORF in a bicistronic ACT-LUC construct leads toincreased expression of the downstream ORF. (A) Physical maps of FLAG-tagged actin –luciferase (ACT-LUC) mRNAs. The actin ORF was progressively truncated at its 3'-end, whilethe length and sequence of the intercistronic spacer was maintained. Numbers indicate lengthsin nucleotides. (B) Relative luciferase activities of constructs, normalised to FL' (not shown) andcorrected for either protein content of the sample only or for both protein content and the relativeLUC mRNA abundance in the sample. (C) Typical steady-state Northern blot of ACT-LUC andtruncated versions, probed for LUC and PGK1 mRNAs. The averaged values for relative LUCmRNA abundance are given under the respective lanes of the Northern blot. (D) Western blotshowing expression of the actin ORF using a FLAG-antibody. The full-length FLAG-tagged actinprotein is indicated.

Page 90: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

Vilela et al., 1998 and 1999; McCarthy, 1998; Oliveira & McCarthy, 1995; see

Section 1.4.6 for details). A set of monocistronic control plasmids bearing the

same series of actin truncations but lacking the LUC gene was also

constructed. In the resulting mRNAs, the termination codon in each of the actin

truncation constructs was always 153 nucleotides away from the poly(A) tail

(Figure 3.8A). The abundance of these mRNAs was reduced by more than 50%

when the uORF was truncated to 63 nt or shorter (Figure 3.8B). The degrees of

destabilisation and the correlations to actin ORF length are generally similar to

the ones observed with the bicistronic constructs, although some stabilisation in

the presence of the LUC ORF is evident.

While displaying only 0.01% of the luciferase activity of the monocistronic

reference construct FL', the full-length bicistronic ACT-LUC mRNA clearly and

reproducible showed expression of the downstream ORF. Since the ORFs of all

bicistronic constructs are out of frame (see also Table 2.5), read-through of the

stop codon can be excluded. To test whether the translation of the downstream

luciferase ORF is due to reinitiation, additional constructs were assayed. The

insertion of an impeding poly(G) stretch decreased this value to 0.0003% (ACT-

pG -LUC in Figure 3.9) corresponding to the luminoscence background activity

of mock-transformed cells (data not shown). In contrast, extending the

intercistronic spacer by duplication of the sequence in construct ACT-119-LUC

resulted in a 31% increase of LUC activity. Thus, the few ribosomes arriving at

AACT

ACTd963

ACTd672

ACTd327

ACTd105

ACTd63

ACTd33

ACTd15

An

63

963

An

AnFLAG-actin170 1157 153

672An

15An

105An

33An

327An

B

AC

T

AC

Td96

3

AC

Td67

2

AC

Td32

7

AC

Td10

5

AC

Td63

AC

Td33

AC

Td15

ACT

PGK1

116±16

ACT/PGK1[%] ± STD

131±20

83±14

42±9

100±9

27±2

19±3

137±11

AACT

ACTd963

ACTd672

ACTd327

ACTd105

ACTd63

ACTd33

ACTd15

An

63

963

An

AnFLAG-actin170 1157 153

672An

15An

105An

33An

327An

ACT

ACTd963

ACTd672

ACTd327

ACTd105

ACTd63

ACTd33

ACTd15

An

63

963

An

AnFLAG-actin170 1157 153

672An

15An

105An

33An

327An

B

AC

T

AC

Td96

3

AC

Td67

2

AC

Td32

7

AC

Td10

5

AC

Td63

AC

Td33

AC

Td15

ACT

PGK1

116±16

ACT/PGK1[%] ± STD

131±20

83±14

42±9

100±9

27±2

19±3

137±11

AC

T

AC

Td96

3

AC

Td67

2

AC

Td32

7

AC

Td10

5

AC

Td63

AC

Td33

AC

Td15

ACT

PGK1

116±16

ACT/PGK1[%] ± STD

131±20

83±14

42±9

100±9

27±2

19±3

137±11

Figure 3.8. 3'-truncation of an actin ORF to 63 nt or shorter triggers decay of themonocistronic mRNA. (A) Map of FLAG-tagged ACT mRNA and 3'-truncation derivatives;numbers indicate lengths in nucleotides. (B) Typical Northern blot of full-length and truncatedFLAG-actin mRNAs, probed for PGK1 and FLAG-tagged ACT mRNAs. The averaged values forrelative LUC mRNA abundance are given under the respective lanes of the Northern blot.

90

Page 91: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

the downstream ORF in ACT-LUC indeed appear to be reinitiating since both

blocking scanning posttermination ribosomes and facilitating factor recruitment

have been shown to modulate the frequency reinitiation at uORFs (see Section

1.4). However, these experiments are operating at the lower limit of the

sensitivity of luciferase measurement, and due to the enzymatic nature of the

assay values cannot be representative in this range.

3.2.2 Translation of a downstream ORF by reinitiation and leaky scanning can be distinguished using start codon mutants

Having established the first-ORF length range for which the reinitiation

competence of the terminating ribosomes becomes readily measurable, we

investigated the reinitiation-uORF-length relationship in more detail. We started

by generating control constructs in which all the AUGs upstream of LUC were

eliminated by nucleotide substitution (series ∆AUGACTd-LUC, Figure 3.10A,

0.00

100%

0.00

004%

0.00

135%

0.00

119%

0.00

154%

0.00

003%

0.0000%

0.0005%

0.0010%

0.0015%

0.0020%

ACT-LUC

ACT-pG-LUC

ACT-119-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

for protein onlyfor protein and RNA content

corrected forB

AACT-LUC

ACT-pG-LUC

ACT-119-LUC An

119

170 1157 1653AnFLAG-actin luciferase

59 153

An

G18

ACT-LUC

PGK1

AC

T-LU

C

FL'

AC

T-pG

-LU

C

AC

T-11

9-LU

C

LUC

100±5

84±10

LUC/PGK1[%] ± STD

88±7

104±11

C

0.00

100%

0.00

004%

0.00

135%

0.00

119%

0.00

154%

0.00

003%

0.0000%

0.0005%

0.0010%

0.0015%

0.0020%

ACT-LUC

ACT-pG-LUC

ACT-119-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

for protein onlyfor protein and RNA content

corrected for

0.00

100%

0.00

004%

0.00

135%

0.00

119%

0.00

154%

0.00

003%

0.0000%

0.0005%

0.0010%

0.0015%

0.0020%

ACT-LUC

ACT-pG-LUC

ACT-119-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

for protein onlyfor protein and RNA content

corrected forB

AACT-LUC

ACT-pG-LUC

ACT-119-LUC An

119

170 1157 1653AnFLAG-actin luciferase

59 153

An

G18

ACT-LUC

ACT-pG-LUC

ACT-119-LUC An

119

170 1157 1653AnFLAG-actin luciferase

59 153

An

G18

An

119An

119

170 1157 1653AnFLAG-actin luciferase

59 153170 1157 1653AnFLAG-actin luciferase

59 153

An

G18

An

G18

ACT-LUC

PGK1

AC

T-LU

C

FL'

AC

T-pG

-LU

C

AC

T-11

9-LU

C

LUC

100±5

84±10

LUC/PGK1[%] ± STD

88±7

104±11

ACT-LUC

PGK1

AC

T-LU

C

FL'

AC

T-pG

-LU

C

AC

T-11

9-LU

C

LUC

100±5

84±10

LUC/PGK1[%] ± STD

88±7

104±11

C

Figure 3.9. Elements that modulate reinitiation affect the low luciferase activity displayedby the full-length actin-luciferase mRNA. (A) Physical maps of FLAG-tagged actin–luciferase(ACT-LUC) mRNA and derivates with a poly(G) stretch inserted between the ORFs (ACT-pG-LUC) or with the intercistronic spacer extended (ACT-119-LUC). Numbers indicate lengths innucleotides. (B) Relative luciferase activities of constructs, normalised to FL' (not shown) andcorrected for either protein content of the sample only or for both protein content and the relativeLUC mRNA abundance in the sample. (C) Typical steady-state Northern blot of ACT-LUC andmodified transcripts, probed for LUC and PGK1 mRNAs. The averaged values for relative LUCmRNA abundance are given under the respective lanes of the Northern blot.

91

Page 92: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

Table 2.5). mRNAs with a 5'-UTR derived from the 33 or 63 nucleotide long

ACT ORFs showed 47 and 44% luciferase activity relative to FL', respectively,

after correction for protein and mRNA abundance (Figure 3.10B). However, the

translational activity of construct ∆AUGACTd105-LUC with a longer 5'-UTR was

reduced to 14%. This unexpectedly low translation efficiency was probably due

to the presence of particularly stable secondary structure in this long 5'-UTR

(predicted to have a maximum stability of -75 kcal mol-1, compared to -33 kcal

mol-1 for FL', -45 kcal mol-1 for ∆AUGACTd33-LUC, and -57 kcal mol-1 for

∆AUGACTd63-LUC; according to the mfold algorithm; Zuker et al., 1999). The

reduced translation efficiency of this transcript also correlates with a 2-fold

stabilisation of the ∆AUGACTd105-LUC mRNA. A similar effect was seen upon

IRP1-binding to the leader of construct FL-5'IRE (compare Figure 3.1 and

Figure 3.10). Since we were primarily interested in defining the cut-off point for

reinitiation competence, not all of the construct series included the 15

nucleotide uORF.

As part of a strategy to characterise the contribution of leaky scanning to

initiation on the LUC ORF, the series ACTd∆iAUG-LUC was created, whereby

all AUGs upstream of luciferase with the exception of the actin start codon were

eliminated by nucleotide substitution (Figure 3.10A, Table 2.5). This prevents

initiation on internal codons within the first ORF, for example at any of the three

actin-internal AUGs that are present in the ACTd105-LUC construct (Figure

3.10A, Table 2.5). Moreover, the AUGs that were removed included one that

overlaps in the -1 frame with the UAA stop codon, and which would therefore be

capable of diverting some of the terminating ribosomes into reinitiating at this

site. By normalising to the respective monocistronic construct of the

∆AUGACTd-LUC series we eliminated influences of the different 5'-UTR

sequences upstream of LUC and derived a revised plot of uORF length against

reinitiation (Figure 3.10C).

Consequently, introduction of a 33 nucleotide long (internal AUG-free)

uORF reduced LUC activity to only 54% (after correction for protein and mRNA

abundance), a 63 nt long ACT ORF to 28%, and even ACTd105∆iAUG-LUC,

with an uORF length of 105 nucleotides, still displayed 12% LUC activity. This

compares with 4.3%, 3.1% and 2.0% for the wild type like ACTd-LUC series

comprising actin-internal start codons (Figure 3.10C).

92

Page 93: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

Some of this initiation at the LUC start codon was expected to be due to

failed recognition of the ACT open reading frame. For increased identification of

Normalised to FL'

0%

10%

20%

30%

40%

50%

60%

15 33 63 105uORF length (in nt)

Rel

ativ

e Lu

cife

rase

Act

ivity

AA

CT-

LUC

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

AU

GA

UG

AC

Td10

5-LU

C

AU

GA

UG

AC

Td63

-LU

C

AU

GA

UG

AC

Td33

-LU

C

∆A

UG

AC

Td10

5-LU

C

∆A

UG

AC

Td63

-LU

C

∆A

UG

AC

Td33

-LU

C

ACT-LUC

PGK1

45±4

84±10

LUC/PGK1[%] ± STD

83±6

42±6

99±4

72±7

53±6

46±7

213±6

140±6

105±11

B

D

170 105 165359 153

ACTd105-LUC

ACTd105∆iAUG-LUC

AUGAUGACTd105-LUC

∆AUGACTd105-LUC * An

* ** An

** ** **ACT AnLUC

* * An

C Normalised to AUGACT d-LUC

0%

10%

20%

30%

40%

50%

60%

33 63 105uORF length (in nt)

Rel

ativ

e Lu

cife

rase

A

ctiv

ity

*

∆AUGACTd-LUC

ACTd∆iAUG-LUCAUGAUGACTd-LUC

ACTd-LUC#

Normalised to FL'

0%

10%

20%

30%

40%

50%

60%

15 33 63 105uORF length (in nt)

Rel

ativ

e Lu

cife

rase

Act

ivity

AA

CT-

LUC

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

AU

GA

UG

AC

Td10

5-LU

C

AU

GA

UG

AC

Td63

-LU

C

AU

GA

UG

AC

Td33

-LU

C

∆A

UG

AC

Td10

5-LU

C

∆A

UG

AC

Td63

-LU

C

∆A

UG

AC

Td33

-LU

C

ACT-LUC

PGK1

45±4

84±10

LUC/PGK1[%] ± STD

83±6

42±6

99±4

72±7

53±6

46±7

213±6

140±6

105±11

AC

T-LU

C

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

AU

GA

UG

AC

Td10

5-LU

C

AU

GA

UG

AC

Td63

-LU

C

AU

GA

UG

AC

Td33

-LU

C

∆A

UG

AC

Td10

5-LU

C

∆A

UG

AC

Td63

-LU

C

∆A

UG

AC

Td33

-LU

C

ACT-LUC

PGK1

AC

T-LU

C

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

AU

GA

UG

AC

Td10

5-LU

C

AU

GA

UG

AC

Td63

-LU

C

AU

GA

UG

AC

Td33

-LU

C

∆A

UG

AC

Td10

5-LU

C

∆A

UG

AC

Td63

-LU

C

∆A

UG

AC

Td33

-LU

C

ACT-LUC

PGK1

45±4

84±10

LUC/PGK1[%] ± STD

83±6

42±6

99±4

72±7

53±6

46±7

213±6

140±6

105±11

B

D

170 105 165359 153

ACTd105-LUC

ACTd105∆iAUG-LUC

AUGAUGACTd105-LUC

∆AUGACTd105-LUC * An

* ** An

** ** **ACT AnLUC

* * An

170 105 165359 153170 105 165359 153

ACTd105-LUC

ACTd105∆iAUG-LUC

AUGAUGACTd105-LUC

∆AUGACTd105-LUC * An* AnAn

* ** An* ** AnAn

** ** **ACT AnLUC** ** **ACT AnAnLUC

* * An* * AnAn

C Normalised to AUGACT d-LUC

0%

10%

20%

30%

40%

50%

60%

33 63 105uORF length (in nt)

Rel

ativ

e Lu

cife

rase

A

ctiv

ity

*

∆AUGACTd-LUC

ACTd∆iAUG-LUCAUGAUGACTd-LUC

ACTd-LUC

∆AUGACTd-LUC

ACTd∆iAUG-LUCAUGAUGACTd-LUC

ACTd-LUC#

Figure 3.10. Manipulation of the FLAG-ACT mini-ORF upstream of LUC. (A) Maps ofACTd105-LUC and further truncated derivatives exemplify four series of constructs. Asterisksindicate the positions of AUGs upstream of the LUC ORF (see Table 2.5 for sequences). uORF-internal start codons are eliminated in the ACTd∆iAUG-LUC series, while a tandem start codonis introduced at the uORF start of this internal AUG-free construct in the AUGAUGACTd-LUCseries. All AUGs upstream of the LUC start codon are mutated in ∆AUGACTd-LUC. (B,C)Relative luciferase activities of AUG-mutated constructs, corrected for both protein content andthe relative LUC mRNA abundance in the sample and normalised to (B) FL' or (C) the respectivemonocistronic ∆AUGACTd-LUC construct. (#) For the ∆AUGACTd-LUC series comprising noupstream AUG, 'uORF length' indicates the size of 5'UTR inserted sequences. (D) Northern blotswere probed for LUC and PGK1 mRNAs; a typical example is shown. The averaged values forrelative ACT-LUC mRNA abundance normalized to FL’ are given under the respective lanes ofthe Northern blots.

93

Page 94: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

the uORF by scanning ribosomes, a tandem actin start codon was introduced,

yielding the AUGAUGACTd-LUC series to provide the ribosome with a second

opportunity to initiate translation at the uORF in the correct reading frame.

Indeed, luciferase activity of these constructs is reduced by 33% to 48% when

compared to the series ACTd∆iAUG-LUC, which means that the contribution to

LUC initiation from leaky scanning rather than reinitiation has been minimised in

these constructs. The uORF-length / luciferase activity plot obtained with the

AUGAUGACTd-LUC constructs (Figure 3.10C) therefore represents the

relationship that is least distorted by leaky scanning.

However, a striking feature of the plots of uORF length versus relative

luciferase activity is that, although there are variations in quantitative terms,

there is a consistent trend of reduction in reinitiation as length increases (Figure

6C). The construct AUGAUGACTd33-LUC showed 33% luciferase activity,

displaying thereby a high reinitiation rate, whereby an efficiently recognised

uORF of 105 nt almost completely abolishes downstream initiation. This data

provide a strong indication that 35 codons are close to representing the

maximum length of uORF that can support significant reinitiation.

The effectiveness of uORF recognition by ribosomes is also reflected in

mRNA abundance (Figures 3.7C, Figure 3.10D). In both series where initiation

at the ACT start codon or within the uORF was efficient (AUGAUGACTd105-

LUC, ACTd-LUC), the mRNA steady-state levels were generally somewhat

lower than in the constructs with a higher level of 'leaky scanning' or of primary

initiation at the LUC start codon (ACTd∆iAUG-LUC, ∆AUGACTd-LUC).

Recognition and translation of the short uORFs apparently leads to enhanced

mRNA degradation.

3.2.3 Reinitiation competence is affected by elongation rate One potential explanation for the loss of reinitiation competence during the

course of elongation is that one or more of the eIFs remain(s) associated with

the elongating ribosome subsequent to formation of the initial peptide bond, but

then dissociates from the ribosome at a finite rate (McCarthy, 1998). According

to this model, the proportion of ribosomes lacking the eIF(s) would increase with

time, resulting in an apparent ORF length-dependent loss of reinitiation

competence. In order to test this model we carried out experiments in a

temperature-sensitive mutant strain (S. cerevisiae NT33-5) that is defective in

94

Page 95: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

the Cca1 tRNA nucleotidyltransferase. At the non-permissive temperature for

this mutant, the availability of functional tRNAs in the cell declines, and the rate

of peptide elongation is reduced (Wolfe et al., 1996). At least some mRNA

species are stabilised by this mutation (Peltz et al., 1992).

Measurements with the ACTd∆iAUG-LUC constructs in the mutant

background revealed a general decrease in LUC activity encoded by uORF-

containing mRNAs at the non-permissive temperature (37°C) compared to the

permissive temperature (25°C; Figure 3.11A,B). The data suggest that reducing

the rate of elongation affects the reinitiation rate downstream of an uORF. At

the same time, the relationship between uORF-length and LUC activity is very

different in this strain, whereby the reduction in reinitiation competence as the

uORF increases from 63 to 105 nucleotides is particularly marked. Stalling of

translation elongation also resulted in a more than 10-fold general stabilisation

of LUC mRNA at 37°C, whereas the PGK1 transcript is unstable in heat-

shocked cells (Figure 3.11C). LUC and PGK1 mRNAs are known to respond

differently to inhibition of translation as caused by the shift in temperature in the

Cca1p mutant. Whereas a translationally inhibitory secondary structure in the 5'-

UTR elevates LUC mRNA levels 2-fold (Figure 3.1E), it reduces the half-life of a

PGK1 transcript from ~35 to 3.5 minutes (Muhlrad et al., 1995). Consistently,

LUC mRNA was about 10 to 20-times more stable than PGK1 transcripts when

the cells were shifted from 25°C to 37°C (Figure 3.11C). The increased

sensitivity of reinitiation towards uORF length when aminoacylated tRNAs are

less abundant (compare construct ACTd105∆iAUG-LUC with the dicistronic

constructs harbouring a shorter ORF in Figure 3.11B) is in agreement with a

model whereby the elongating ribosome loses factors necessary for reinitation

(see Section 1.4.3). However, for a reliable comparison, expression of the

tested dicistronic constructs in a reference strain (W303) at the temperatures

assayed (25°C and 37°C) would have to be monitored. In addition, reduced

levels of charged tRNAs result in the activation of the GCN2 / GCN4 pathway

with initiation factor eIF2 becoming less available for translation initiation

(Hinnebusch, 2000). Thus, the observed increased dependence of reinitiation

on uORF length could also stem partly from this indirect effect of Cca1 tRNA

nucleotidyltransferase inactivation. Experiments e.g. in a GCN2 mutant strain

95

Page 96: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

could help to separate effects of altered eIF2 abundance and uORF length

dependence.

In a study using rabbit reticulocyte lysates, Kozak (2001b) observed a

decrease in reinitiation competence when a pseudoknot that is capable of

slowing down elongating ribosomes was inserted into a 54 nucleotide uORF.

This parallels the effect we see in the mutant strain where elongation is

restricted due to a decrease in the level of aminoacylated tRNAs. Both our

study and that of Kozak are consistent with the idea that the kinetics of

corrected for

87%

106%

90%

57%

100%

63%

84%

77%

19%100%

0%

25%

50%

75%

100%

125%

FL"

ACTd105

iAUG-L

UC

ACTd63

iAUG-L

UC

ACTd33

iAUG-L

UC

ACTd15

iAUG-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

37

ºC /

25ºC

protein content onlyprotein and RNA content

∆∆ ∆ ∆

81%

86%

64%2%100%

51%

73%

49%

0.4%

100%

0%

25%

50%

75%

100%

125%

FL"

ACTd105

iAUG-L

UC

ACTd63

iAUG-L

UC

ACTd33

iAUG-L

UC

ACTd15

iAUG-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

corr

ecte

d fo

r pro

tein

and

RN

A co

nten

t

25ºC37ºC

∆∆ ∆ ∆

A

B C

FL"

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

FL”

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

LUCACT-LUC

PGK1

25ºC 37ºC

51±7

100±8

LUC/PGK1[%] ± STD

90±14

1224±174

129±17

81±5

1869±228

1850±280

1391±199

1376±213

corrected for

87%

106%

90%

57%

100%

63%

84%

77%

19%100%

0%

25%

50%

75%

100%

125%

FL"

ACTd105

iAUG-L

UC

ACTd63

iAUG-L

UC

ACTd33

iAUG-L

UC

ACTd15

iAUG-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

37

ºC /

25ºC

protein content onlyprotein and RNA content

∆∆ ∆ ∆

87%

106%

90%

57%

100%

63%

84%

77%

19%100%

0%

25%

50%

75%

100%

125%

FL"

ACTd105

iAUG-L

UC

ACTd63

iAUG-L

UC

ACTd33

iAUG-L

UC

ACTd15

iAUG-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

37

ºC /

25ºC

protein content onlyprotein and RNA content

∆∆ ∆ ∆

81%

86%

64%2%100%

51%

73%

49%

0.4%

100%

0%

25%

50%

75%

100%

125%

FL"

ACTd105

iAUG-L

UC

ACTd63

iAUG-L

UC

ACTd33

iAUG-L

UC

ACTd15

iAUG-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

corr

ecte

d fo

r pro

tein

and

RN

A co

nten

t

25ºC37ºC

∆∆ ∆ ∆

81%

86%

64%2%100%

51%

73%

49%

0.4%

100%

0%

25%

50%

75%

100%

125%

FL"

ACTd105

iAUG-L

UC

ACTd63

iAUG-L

UC

ACTd33

iAUG-L

UC

ACTd15

iAUG-L

UC

Rel

ativ

e Lu

cife

rase

Act

ivity

corr

ecte

d fo

r pro

tein

and

RN

A co

nten

t

25ºC37ºC

∆∆ ∆ ∆

A

B C

FL"

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

FL”

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

LUCACT-LUC

PGK1

25ºC 37ºC

51±7

100±8

LUC/PGK1[%] ± STD

90±14

1224±174

129±17

81±5

1869±228

1850±280

1391±199

1376±213

FL"

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

FL”

AC

Td10

5∆iA

UG

-LU

C

AC

Td63

∆iA

UG

-LU

C

AC

Td33

∆iA

UG

-LU

C

AC

Td15

∆iA

UG

-LU

C

LUCACT-LUC

PGK1

25ºC 37ºC

51±7

100±8

LUC/PGK1[%] ± STD

90±14

1224±174

129±17

81±5

1869±228

1850±280

1391±199

1376±213

Figure 3.11. Inhibition of translation elongation affects expression of a downstreamluciferase ORF in a uORF-size dependent manner. (A) Luciferase activities of cell lysatesextracted from strain NT33-5 carrying a heat-sensitive allele of the ATP(CTP):tRNAnucleotidyltransferase-encoding gene CCA1 grown for 4 hours at either 25°C or at the non-permissive temperature of 37°C. The values are normalised to FL" (which corresponds to FL butcarries a URA3 marker) for each temperature and corrected for both protein and RNAconcentrations in the sample. (B) Ratios of luciferase activities of cells grown at 37°C or 25°C,corrected for either protein content of the sample only or for both protein content and the relativeLUC mRNA abundance in the sample. (C) Northern blots were probed for LUC and PGK1mRNAs, the averaged values for relative LUC mRNA abundance are given under the respectivelanes of the Northern blot.

96

Page 97: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

elongation influence reinitiation competence. This can be explained by a model

in which one or more of the eIFs does not immediately leave the ribosome

subsequent to initiation of peptide synthesis, thus maintaining the ribosome in a

reinitiation-competent state. Assuming a finite (relatively low) rate of release of

the eIF(s), reinitiation competence will be progressively lost as a function of time

and / or progression through the uORF. However, the data we have obtained

with the cca1 mutant indicate that the rate of loss of the eIF(s) in such a model

would not simply be a straightforward function of time, but rather also influenced

by other factors. Consequently, the elongation rate influences the maximum

length of uORF that can allow ribosomes to retain their reinitiation competence.

3.2.4 Premature translation termination by either ORF truncation or stop codon insertion affects mRNA stability differently

3.2.4.1 The presence of premature termination codons but not 3'-truncation of the ORF trigger nonsense-mediated decay

It is known that translation termination can trigger accelerated decay of mRNA

(McCarthy, 1998; Jacobson & Peltz, 2000). The strongest effect is seen when

premature termination codons (PTCs) in a reading frame cause nonsense-

mediated decay (Jacobson & Peltz, 2000; Maquat, 2000, Section 1.3.2).

Previous studies of nonsense-mediated control in yeast have focused on

premature termination within an otherwise intact gene (Gonzalez et al., 2001).

Yet, the 3' truncation of the actin ORF in the ACTd series results in a special

kind of premature termination, whereby the length and the composition of the 3'-

UTR remains unaffected (Figure 3.8). For comparison, therefore, we generated

a set of premature termination codon (PTC) constructs by introducing a UAA

stop codon at several positions within the actin ORF. Translation on these

mRNA is terminated 963, 672, 327 or 63 nt downstream of the start codon, and

the remaining actin ORF – together with the FL' inherent trailer – acts as an 3'-

UTR (Figure 3.12). In contrast to the ACTd-LUC 3'-UTRs, this additional

sequence extends the distance between the stop codon and the poly(A) tail by

about 200 – 1000 nt, depending on the site of the PTC. Moreover, the

remainder of the actin ORF also comprises several start and in-frame stop

codons, creating thereby the possibility for downstream ribosomes to undergo

cycles of translation at these ORFs of varying lengths.

97

Page 98: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

G

ACT-LUC

ACTptc963-LUC

ACTptc672-LUC

ACTptc327-LUC

ACTptc63-LUC

AnFLAG-actin luciferase170 1157 165359 153

327 PTCAn

An

963 PTC

672An

PTC

63AnFLAG-actin

PTC

ACT

ACTptc963

ACTptc672

ACTptc327

ACTptc63

C

E

B

ACT

AC

Tptc

963

AC

Tptc

672

AC

Tptc

327

AC

Tptc

63

AC

T

PGK1

100±5

14±7

11±9

12±10

110±12

ACT/PGK1[%] ± STD

ACT-LUC

PGK1

FL'

AC

T-LU

C

AC

Tptc

963-

LUC

AC

Tptc

672-

LUC

AC

Tptc

327-

LUC

AC

Tptc

63-L

UC

LUC

100±12

119±6

LUC/PGK1[%] ± STD

56±4

65±6

76±6

80±2

F

32.5

16.5

47.5

25.0

Act1p

kDa

AC

T-LU

C

AC

Tptc

963-

LUC

AC

Tptc

672-

LUC

AC

Tptc

327-

LUC

AC

Tptc

63-L

UC

A

0.00

005%

0.00

011%

0.00

006%

0.00

013%

0.00

100%

0.00

119%

0.00

008%

0.00

023%

0.00

009%

0.00

024%

0.0000%

0.0005%

0.0010%

0.0015%

ACT-LUC

ACTptc9

63-LUC

ACTptc67

2-LUC

ACTptc32

7-LUC

ACTptc63

-LUC

Rel

ativ

e Lu

cife

rase

Act

ivity protein content only

protein and RNA content

corrected forD

100%

110%

14%

11%

12%

100%

65%

80%

56%

76%

0%

25%

50%

75%

100%

125%

ACT

ACTptc96

3

ACTptc67

2

ACTptc32

7

ACTptc63

Rel

ativ

e m

RN

A a

bund

ance

LUC

/PG

K1

monocistronicdicistronic (+LUC)

An

63

963

AnFLAG-actin

AnFLAG-actin170 1157

672An

327

PTC

PTC

PTC

PTCAn

153

G

ACT-LUC

ACTptc963-LUC

ACTptc672-LUC

ACTptc327-LUC

ACTptc63-LUC

AnFLAG-actin luciferase170 1157 165359 153

327 PTCAn

An

963 PTC

672An

PTC

63AnFLAG-actin

PTC

ACT-LUC

ACTptc963-LUC

ACTptc672-LUC

ACTptc327-LUC

ACTptc63-LUC

AnFLAG-actin luciferase170 1157 165359 153

327 PTCAn

An

963 PTC

672An

PTC

63AnFLAG-actin

PTC

AnFLAG-actin luciferase170 1157 165359 153

327 PTCAn

An

963 PTC

672An

PTC

63AnFLAG-actin

PTC

ACT

ACTptc963

ACTptc672

ACTptc327

ACTptc63

C

E

B

ACT

AC

Tptc

963

AC

Tptc

672

AC

Tptc

327

AC

Tptc

63

AC

T

PGK1

100±5

14±7

11±9

12±10

110±12

ACT/PGK1[%] ± STD

ACT

AC

Tptc

963

AC

Tptc

672

AC

Tptc

327

AC

Tptc

63

AC

T

PGK1

100±5

14±7

11±9

12±10

110±12

ACT/PGK1[%] ± STD

ACT-LUC

PGK1

FL'

AC

T-LU

C

AC

Tptc

963-

LUC

AC

Tptc

672-

LUC

AC

Tptc

327-

LUC

AC

Tptc

63-L

UC

LUC

100±12

119±6

LUC/PGK1[%] ± STD

56±4

65±6

76±6

80±2

ACT-LUC

PGK1

FL'

AC

T-LU

C

AC

Tptc

963-

LUC

AC

Tptc

672-

LUC

AC

Tptc

327-

LUC

AC

Tptc

63-L

UC

LUC

100±12

119±6

LUC/PGK1[%] ± STD

56±4

65±6

76±6

80±2

F

32.5

16.5

47.5

25.0

Act1p

kDa

AC

T-LU

C

AC

Tptc

963-

LUC

AC

Tptc

672-

LUC

AC

Tptc

327-

LUC

AC

Tptc

63-L

UC

32.5

16.5

47.5

25.0

Act1p

kDa

32.5

16.5

47.5

25.0

Act1p

kDa

AC

T-LU

C

AC

Tptc

963-

LUC

AC

Tptc

672-

LUC

AC

Tptc

327-

LUC

AC

Tptc

63-L

UC

A

0.00

005%

0.00

011%

0.00

006%

0.00

013%

0.00

100%

0.00

119%

0.00

008%

0.00

023%

0.00

009%

0.00

024%

0.0000%

0.0005%

0.0010%

0.0015%

ACT-LUC

ACTptc9

63-LUC

ACTptc67

2-LUC

ACTptc32

7-LUC

ACTptc63

-LUC

Rel

ativ

e Lu

cife

rase

Act

ivity protein content only

protein and RNA content

corrected for

0.00

005%

0.00

011%

0.00

006%

0.00

013%

0.00

100%

0.00

119%

0.00

008%

0.00

023%

0.00

009%

0.00

024%

0.0000%

0.0005%

0.0010%

0.0015%

ACT-LUC

ACTptc9

63-LUC

ACTptc67

2-LUC

ACTptc32

7-LUC

ACTptc63

-LUC

Rel

ativ

e Lu

cife

rase

Act

ivity protein content only

protein and RNA content

corrected forD

100%

110%

14%

11%

12%

100%

65%

80%

56%

76%

0%

25%

50%

75%

100%

125%

ACT

ACTptc96

3

ACTptc67

2

ACTptc32

7

ACTptc63

Rel

ativ

e m

RN

A a

bund

ance

LUC

/PG

K1

monocistronicdicistronic (+LUC)

An

63

963

AnFLAG-actin

AnFLAG-actin170 1157

672An

327

PTC

PTC

PTC

PTCAn

153

Figure 3.12. Impact of a premature stop codon on the translation and abundance ofmonocistronic ACT and bicistronic the ACT-LUC mRNA. (A) Maps of FLAG-tagged ACT mRNAand mutants with a premature termination codon (PTC). Numbers indicate lengths in nucleotides. (B) Typicalsteady-state Northern blot of FLAG-actin and PTC-mutants, probed for FLAG-ACT and PGK1 mRNAs. Theaveraged values for relative FLAG-ACT mRNA abundance are given under the respective lanes of the Northernblot. (C) Maps of FLAG-tagged ACT-LUC mRNAs with a premature termination codon (PTC) was inserted atseveral positions. The remaining (white-boxed) ACT ORF forms part of the intercistronic spacer. Numbers indicatelengths in nucleotides. (D) Relative luciferase activities of constructs, normalised to FL' (not shown) and correctedfor either protein content of the sample only or for both protein and ACT-LUC RNA concentrations. (E) Western blotshowing expression of the actin ORF using a FLAG-antibody. The full-length actin protein is indicated. (F) TypicalNorthern blot of FL' and bicistronic constructs, probed for LUC and PGK1 mRNAs. The averaged values for relativemRNA abundance are given under each lane. (G) Comparison, for each construct, of the relative ACT mRNAabundance with and without the downstream LUC ORF.

98

Page 99: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

These monocistronic reference mRNAs showed significant mRNA

destabilisation when a PTC was located in the first two thirds of the actin ORF.

Whereas PTC insertion in construct ACTptc963 with 83% of the full-length

FLAG-actin ORF being translated had no obvious effect on mRNA stability,

reducing the translated region to 58% (ACTptc672) or less (ACTptc327,

ACTptc63) resulted in NMD-like accelerated decay (Figure 3.12B). However, to

verify degradation via the NMD pathway, the stability of these constructs would

have to be assessed in a genetic background deficient for components of this

pathway (e.g. the Upf proteins).

3.2.4.2 PTC-containing mRNAs are stabilised by a full-length downstream ORF that is not translated

To test for reinitiation downstream of PTCs, a luciferase cassette was inserted

downstream of the wild type actin stop codon of ACTptc constructs, yielding the

ACTptc-LUC constructs. The resulting bicistronic constructs differed from

ACTd-LUC in that the spacer also comprises the actin sequence downstream of

the PTC (compare Figures 3.7A; 3.11C). In this case, reinitiation at the LUC

ORF remained highly inefficient, even with the very shortest actin-derived first

ORF. Intercistronic AUG codons from the original actin ORF lie downstream of

the newly created PTCs, and these are all capable of 'capturing' reinitiation-

competent ribosomal subunits. Initiation and translation of the actin ORF was

efficient as shown by detection of the full-length actin protein in an anti-FLAG

Western blot (Figure 3.12E).

However, introduction of the downstream LUC ORF markedly stabilised

the decay-susceptible PTC-containing mRNAs (as shown in the comparison

provided by Figure 3.12G). Since the dicistronic constructs display no significant

luciferase activity, mRNA stabilisation appears not to be linked to translation of

the downstream ORF. Moreover, this effect seems to be restricted to mRNAs

susceptible to NMD, since mRNAs with truncated actin ORFs were not

stabilised to the same degree by insertion of a LUC ORF in the 3'-UTR

(compare Figures 3.7C, 3.8B and Figures 3.12B, 3.12F, 3.12G). In summary,

the relative levels of mRNA abundances and LUC translation rates for the

truncated bicistronic and monocistronic transcripts showed no evident

relationship to the mRNA degradation and reinitiation situation at PTC-carrying

transcripts. Therefore, the situation of the terminating ribosome must be

99

Page 100: Translation reinitiation and ribosome recycling - PhD Thesis

3. Results

different at these two kinds of shortened ORFs, affecting its subsequent

posttermination fate (reinitiation) and the effect on mRNA stability.

100

Page 101: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

4. DISCUSSION

4.1 Accessibility of the 3'-UTR is not a prerequisite for posttermination ribosome recycling

4.1.1 Differences between 5'-UTR and 3'-UTR directed blockage In this study, we show that imposing a block to ribosomal movement along the

3'-UTR of an in vivo expressed reporter mRNA did have a limited and position-

dependent effect on translational efficiency. When compared to the 95%

reduction obtained by restricting access to the 5'-UTR, IRP1-binding 24 nt or 39

nt downstream of the luciferase stop codon decreased translation only by about

27%. When the IRP1 binding site was close enough to the stop codon for the

terminating ribosomes to interfere with the complex, no effect on LUC activity

was apparent, suggesting that the terminating ribosome has sufficient energy to

destabilise IRP1-binding to IRE (construct FL-15-IRE in Figure 3.1).

Alternatively, the stop codon 3'-proximal structure might even aid in recognition

of the termination site as it was observed for the identification of start codons in

a suboptimal context by 40S subunits (Kozak, 1991). IRP1-association with the

3'-UTR at a position close to the poly(A) tail resulted in a slight (~25%) increase

in mRNA stability and a ~50% decrease in translational efficiency (FL-141-IRE).

This more pronounced effect could be due to the close vicinity of the poly(A) tail

to the IRP1-binding site and because of the presence of tandem copies of IRE,

presumably resulting in a more stable IRP1-blockage.

Niepel and colleagues (1999) observed a ~50% decrease in translational

efficiency when they placed a strong stem loop close adjacent to the stop

codon, but their experiments differed from ours since they programmed yeast

spheroblasts and used an inherent stable structure that is not readily

destabilised unlike IRE that is not bound by IRP1 (see also Section 3.1.2). They

also observed a gradual loss of this inhibitory effect when inserting a 20 nt or

100 nt long spacers between the translation termination and the polyadenylation

site. The monitored reduction by 33% and 22%, respectively, was only

detectable in yeast spheroblasts, whereas in mammalian or plant cell-free and

protoblast systems movement of the structure downstream of the stop codon

relieved the repression completely. This indicates not only an enhanced

101

Page 102: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

sensitivity of yeast cells towards secondary structures in the 3'-UTR (as already

known for the 5'-UTR), but also confirms the position-dependence of the

translational repression by the IRE – IRP1 system.

The observed IRE – IRP1 induced reduction in reporter gene activity was

exclusively caused by a decrease in translational efficiency of the LUC mRNA,

since (1) the mRNA was not destabilised, (2) the poly(A) tail length and

polyadenylation sites were not affected, and (3) the LUC mRNA distribution

among the polysomal profiles correlated with the decrease in translation. In the

system presented, the IRP1-IRE complex almost completely abolished gene

expression when located in the 5'-UTR, most likely due to an inhibition of 43S-

cap association and scanning of the leader (Koloteva et al., 1997). However,

the 3'-UTR located blockage imposed a comparably low effect on LUC

translation, indicating thereby that the accessibility of the sequences

downstream of the termination site to scanning ribosomes is only of minor

importance for the translational efficiency. Assuming that Pab1p-contributed

increase in translation initiation frequency is partly due to 'channelling' and

recycling of ribosomes or subunits, it can be deduced that this does not occur

via 3'-UTR scanning, contrasting with the reinitiation mechanism that relies on

scanning downstream of termination. This perception is supported by the finding

that mutating the LUC stop codon context as to facilitate ribosomal release at

the termination site and to prevent reinitiation did not affect translational

efficiency of the mRNA. Moreover, repression by IRP-binding to the 3'-UTR did

not rely on the existence of an efficient 5' – 3' interactions as revealed by

assaying construct FL-39-IRE in a yeast strain expressing an eIF4GI mutant

defective for Pab1p binding. Similarly, Niepel and colleagues (1999) observed

no interruption of the poly(A) – cap synergy by insertion of 3'-UTR secondary

structures in yeast spheroblasts.

However, in the in vivo system assayed the yeast cells were allowed to

grow in rich medium at a favourable temperature (30°C). In this optimal

environment it cannot be excluded that a reduction in translational activity

caused by the 3'-UTR blockage was compensated by the enhanced utilisation

of an alternative pathway. For example, blocking the intramolecular recycling

pathway could cause an enhanced de novo recruitment of ribosomal subunits

from the cellular pool. In further experiments, it would be interesting to monitor

102

Page 103: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

the effect of 3'-UTR secondary structures in cells grown under stress conditions

or in the stat phase with enhanced competition between mRNAs for recruitment

of ribosomes.

4.1.2 Translation might be affected by a 3'UTR-conferred mRNP modification

The data suggest that modification of the 3'-UTR affects the translatability of a

cellular mRNA by virtue of the fact that it changes mRNP structure, rather than

because it blocks the passage of ribosomes through the 3'-UTR. This would

also explain the 2-fold reduction in translation when placing the IRE-IRP1

complex further downstream, close to the poly(A) tail. This could not be

explained by blocking a processive event such as the scanning-like association

of ribosomes with the 3'-UTR but indicates a more indirect, position-dependent

effect like changes of the mRNP structure. In the yeast cell-free system, the

normal mRNP structure is already partially disrupted, and this explains why

changes in the 3'-UTR of this kind have no detectable effect. This does not

exclude that mRNAs are circularised in vitro (as shown by Wells et al., 1998).

The presence of a poly(A) tail stimulates translation of mRNAs regardless of its

conformation caused by IRP1-binding (Figure 3.6C,D) or by secondary

structures (Niepel et al., 1999). However, at least the quantitative principles of

translational control on a template that is located within a non-physiological

mRNP structure will be different to those applying to normal cellular polysomes.

In this context, the results of Borman and colleagues (58) are worthy of note.

They observed that, in rabbit reticulocyte lysate, exogenous poly(A) stimulates

translation of capped, nonadenylated mRNA to the same degree as the

presence of a poly(A) tail at the 3'-end of the mRNA. This underlines the fact

that mRNP functions differently in a cell-free system, at least to the extent that

internal 5' – 3' interactions are not required for translation efficiency to be

maximised.

Similar to the proposed reliance on mRNP domain organisation for mRNA

surveillance (Hilleren & Parker, 1999), translation depends on mRNP domains

such as those defined by the cap-binding complex, the factors associated with

translation initiation and termination sites, and the 3'-UTR / poly(A) attached

proteins. Although not directly interfering with ribosomal scanning, downstream

structures could therefore disturb this proper organisation and e.g. disrupt

103

Page 104: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

termination site – poly(A) interactions granted by the eRF3 – Pab1p association.

Likewise, Curatola et al. (1995) showed that a secondary structure at any

position between the stop codon and a 3'-UTR located ARE sequence can

prevent ARE-induced degradation in HeLA cell extracts. They suggest this

might due blocked propagation of conformational or compositional changes to

the 3'-UTR RNP that might result from translation of the 3'-terminus of the

coding region rather than from entrance of translation-linked factors into the 3'-

UTR (i.e. posttermination scanning). Interestingly, IRE-insertion and IRP1-

binding to the 3'-UTR did not result in destabilisation of LUC mRNA. This

indicates that the induced mRNP distortion only affected translation, but not any

of the mRNA stability determinants such as the length of the poly(A) tail, its

association with Pab1p or the eIF4G – Pab1p interaction.

A modified 'circular loop' model of mRNA

In context of the mRNP model, the interaction between the termination site and

nAAAA

ORF

60S4E

eIF4G

Pab1p eR

F3eR

F1

ribosomal release

initiation

termination

elongation

scanning

40S recruit-

ment

'chan

nellin

g'

40S40S

nAAAA

ORF

60S4E

eIF4G

Pab1p eR

F3eR

F1

ribosomal release

initiation

termination

elongation

scanning

40S recruit-

ment

'chan

nellin

g'

40S40S40S40S

Figure 4.1. Channelling is promoted by mRNP structure. The protein bridge Pab1-eIF4G-eIF4E canmediate interactions between the 5’ and 3’-ends of mRNA. The interaction between eRF3 and Pab1 mayalso link the termination site with the poly(A) tail. According to our data, while such structures within thecellular mRNP promote recycling of ribosomes, this does not require continuous scanning along the 3'-UTR. Instead, juxtaposition of the sites of termination and initiation facilitates the process. This contrastswith the process of reinitiation, which is only found to occur at a significant efficiency downstream of uORFsof up to 35 codons in length. Reinitiation is strictly dependent upon scanning between the site of terminationand the downstream start site. The dotted hairpin alludes to the positioning of structural elements in the 3'-UTR that were used in this work.

104

Page 105: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

the poly(A) tail would cause the 3'-UTR to loop out of an already circular mRNA,

as proposed by Uchida and colleagues (2002). Channelling of ribosomes or

subunits would occur not via the 3'-UTR – Pab1p – eIF4F – cap interaction but

due to the proximity of termination site and 5'-UTR afforded by the eRF3 –

Pab1p – eIF4F – cap bridge (Figure 4.1). As a consequence, this would make

properties such as secondary structures in the 3'-UTR ineffective since they

would not block a 'bottleneck' step such as 3'-UTR scanning. Interestingly,

eRF3 remains associated with 40S subunits after termination (Didichenko et al.,

1991; Eustice et al., 1986), enabling an uninterrupted guidance of the subunit

via Pab1p to the 5'-end of a circular mRNA without requiring involvement of the

3'-UTR sequence.

4.1.3 Recycling is unlikely to be reinitiation-like since the ribosomal reinitiation potential depends on the ORF length

In this study, we also consequently investigated the dependence of ribosomal

reinitiation capability on ORF-length. We found that the reinitiation of translation

at a downstream ORF is practically zero when the ribosome has translated a

full-length wild type like open reading frame of about 1.2 kb. Reinitiation was

only readily detectable with the ORF 3'-truncted to 35 codons (105 nt) or less.

This is in agreement with a predicted cut-off length of 55 codons (168 nt) on

viral mRNA (Luukkonnen et al., 1995) and a gradual decrease in reinitiation by

extending the uORF up to 33 codons (102 nt) in a rabbit reticulocyte cell-free

lysate (Kozak, 2001). Moreover, experiments in a cca1 mutant with reduced

aminoacylated tRNA levels at a non-permissive temperature suggest a model

whereby the length of the ORF in relationship with the elongation cycle kinetics

determines the stage in elongation at which the ribosome is rendered unable to

promote posttermination reinitiation. Introduction of short uORFs of varying

length resulted in a more pronounced inhibition of downstream initiation when

translation elongation was slowed down by above measure (Figure 3.11).

Hence, the properties of ribosomes terminating at a short uORF and at a long

ORF differ fundamentally, most likely in their association with factors required

for reinitiation such as eIF3 and GTP-charged eIF2. In agreement with the

limited effect of 3'-UTR structures on translational efficiency, this finding

provides additional evidence that intramolecular ribosome recycling does not

rely on a posttermination scanning and reinitiation-like step. In this context, it

105

Page 106: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

will be of great interest to determine at which step in recycling ribosomes

reacquire the ability to initiate translation. Our results indicate that in contrast to

reinitiation after a short uORF, this recruitment appears not to be facilitated by a

posttermination scanning step.

4.2 The relationship between ORF length, nonsense-mediated decay and reinitiation

4.2.1 PTC-insertion but not ORF 3'-truncation results in accelerated mRNA decay

To further investigate the role of ORF length and the 3'-UTR on translation and

mRNA degradation, we assessed the stability of two groups of mRNAs: The first

comprised a single, progressively 3'-truncated ORF with a 3'-UTR of constant

length (Figure 3.7, Figure 3.8). The second had the ORF interrupted by a

premature termination codon (PTC), thus extending the 3'-UTR accordingly.

Interestingly, only insertion of the PTC, but not 3'-truncation resulted in

accelerated decay. Rapid mRNA degradation due to the presence of a

premature stop codon has been shown for other mRNAs to occur via the NMD

pathway at comparable rates (see Section 1.3.2.2 for details). Abolishing this

pathway, e.g. in a mutant background deficient for one of the Upf proteins could

verify the mechanism of mRNA degradation for both the PTC-containing and

the truncated ORFs. According to the model by Peltz and colleagues (1993),

the observed difference in the degradation rates would be due to the lack of a

downstream destabilising sequence element (DSE) in the truncated constructs.

In agreement, the only construct that is not destabilised has the PTC located

about 200 nt upstream of the regular termination site, probably deficient for a

downstream DSE. However, this mRNA also comprises only a single

downstream ORF whereas the other PTC-carrying transcripts contain several

downstream AUGs and inframe stop codons in the extended sequence between

stop codon and poly(A) tail. In an alternative model, it might therefore be

possible for ribosomes terminating at the prematurely encountered stop codons

to resume scanning and to reinitiate. Consequently, the numerous downstream

ORFs all terminate in a distance from the 3'-UTR and the resulting unfavourable

mRNP environment might be responsible for triggering the accelerated mRNA

decay. Conversely, the truncated constructs all have a 3'-UTR with a constant

106

Page 107: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

length which might help to stabilise these transcripts. This might be a feature

characteristic for S. cerevisiae since in this organism 3'-UTRs are strikingly

homogenous in length (see also Section 1.3.2.3).

4.2.2 Insertion of an untranslated downstream ORF stabilises PTC-carrying mRNAs

The introduction of a PTC into the first two thirds of the actin ORF increases the

effective length of the 3’UTR, resulting in marked destabilisation of the mRNA.

This is entirely consistent with previous work on the effects of PTCs on the

stability of other yeast messages (including PGK1 and URA1; reviewed in

McCarthy, 1998, and Jacobson & Peltz, 1996). However, this effect was

relieved to a certain degree when a full-length ORF was inserted downstream of

the PTC-containing ORF. We found that the downstream LUC ORF is not

translated in these constructs but observed an unexpected 4 to 7-fold

stabilisation of the mRNAs upon downstream ORF insertion (Figure 3.12G). No

such stabilisation was found when the luciferase ORF was appended to

truncated actin ORFs, but there was a general tendency to destabilisation when

the actin ORF was truncated to 105 nt or less (compare Figure 3.7 and Figure

3.8). This corresponds to a progressive replacement of the stable actin mRNA

sequence by the less stable LUC ORF.

Another earlier study has shown that extension of the 3'-UTR of the CUP1

ORF resulted in NMD-like destabilisation (Muhlrad & Parker, 1999). Although

addition of the (untranslated) LUC ORF practically extends the 3'-UTR by 1,700

nt, our findings differ from this and point to a general stabilising function of the

added sequence that is also not correlated to its relative position to the

(premature) termination codon. This indicates that several parts of the LUC

reading frame might contribute to determine mRNA stability, probably by

defining an overall mRNP structure different from the PTC-mRNA with a single

ORF.

4.2.3 Modulation of leaky scanning and mRNA stability by the efficiency of uORF recognition

The presence of an upstream ORF derived from the wild type, FLAG-tagged

yeast actin sequence in the leader of a luciferase reporter mRNA resulted in a

high repression of downstream translation, even when the uORF was truncated

to 15 residues (4 amino acids). Since this restriction was upheld in constructs

107

Page 108: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

ranging from 15 to 105 nt, a 'ribosome stalling' effect of the sequence 5' of the

uORF stop codon (that varied in these constructs) can be ruled out.

To illustrate the uORF initiation and termination events at construct

ACTd105-LUC, Figure 4.2 depicts all 5’-UTR located uORFs arising by the

utilisation of the ATGs upstream of the LUC ORF. At this wild type sequence,

leaky scanning across the first ATG will result in ribosomes translating the

second uORF. Translation of this uORF2 is shifted in frame (by +1) with respect

to uORF1 and terminates 46 nt upstream of the main LUC ORF. Furthermore,

all other uORFs terminate in this region and all of them are shorter in length

than uORF1. Therefore, ribosomes having translated any part of the leader can

be expected to have comparable reinitiation capabilities. Moreover, due to

overlapping ORFs it can also be excluded that a particular ribosomes translates

more than one uORF. Therefore, the observed low reinitiation rate for this

construct cannot be the result of leaky scanning or multiple translation events

upstream of the main LUC ORF. This is also true for the constructs with a

truncated uORF. However, the highly restrictive control on reinitiation was vastly

relieved (10-fold with a uORF of 33 nt) when the native uORF-internal start

codons were mutated (Figure 3.10).

To investigate the contribution of ribosomes arriving at the luciferase start

codon via 'leaky scanning', we inserted a double uORF start codon in these

internal AUG-free mutants. Indeed, downstream translation was reduced

proportionally indicating to be more reliant on reinitiation in these constructs. To

explain the even more stringent effect of wild type uORF sequences with

inherent uAUGs and uORFs, a model might be useful that entails the negative

effects of initiating and terminating ribosomes (and subunits) on scanning and

elongation (Figure 4.3). Whereas in the (artificial) constructs with only a single

or double (inframe) uORF AUG, scanning, elongation and (re)scanning occurs

________________________________________________main uORF________________________________________________----spacer--->>ATGGATTACAAGGACGACGATGACAAGACTATGGATTCTGAGGTTGCTGCTTTGGTTATTGATAACGGTTCTGGTATGTGTAAAGCCGGTTTTGCCGGTGACTAATGTAGAATTCTAGAA________________________________________________main uORF________________________________________________----spacer--->>ATGGATTACAAGGACGACGATGACAAGACTATGGATTCTGAGGTTGCTGCTTTGGTTATTGATAACGGTTCTGGTATGTGTAAAGCCGGTTTTGCCGGTGACTAATGTAGAATTCTAGAA

Figure 4.2. 5'-UTR sequence of ACTd105-LUC and representation of ORFs starting at allpossible upstream ATGs. The 5' to 3' sequence of the 105 nt long uORF is representedtogether with the first nucleotides of the 59 nt long intercistronic spacer. Boxes indicate thesequences covered by ORFs starting at all possible uATGs with gray boxes indicating ORFs ina different reading frame. Note that all uORFs extend to a maxiumum of 13 nt into the spacer,terminating thereby at least 46 nt upstream of the (not depicted) downstream LUC ORF. ATGsare in bold, termination codons are italic.

108

Page 109: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

in an ordered manner, the situation at a uORF with internal initiation and

termination sites might be more complex. Ribosomes that have arrived at these

sites by leaky scanning or translation of (internal) uORFs could potentially

hinder the movement of scanning or elongating ribosomes, thus slowing them

down or causing them to abort their association with the mRNA, resulting in

ribosomal release. The effects of these internal ORF would therefore be greater

than their contribution to the reinitiation cycles on these transcripts. Ribosomes

stalled at termination sites of uORFs only translated by 10% of the subunits

have been found to efficiently block movement of subsequent ribosomes and

subsequent (re)initiation at the main ORF start codon (see also Section 1.4.4).

Similarly, the described system will only work within a leader where the (main)

uORF is not recognised by all scanning subunits. This is true for the uORF start

codon context used in this study, as revealed by the decrease in downstream

initiation upon creation insertion of a second uORF start codon Interestingly,

GCN4, YAP1, ermC, cm1A and potentially other natural uORFs do not

*

* *

*

* **

** *

∆AUGACTd33-LUC

ACTd33∆iAUG-LUC

AUGAUGACTd33-LUC

ACTd33-LUC (wild type)

LUCA

B

C

D

100%

54%

33%

4%

LUC activity

scanning

* = AUG

ribosome release

uORF

translation

posttermination scanning

*

* *

*

* **

** *

∆AUGACTd33-LUC

ACTd33∆iAUG-LUC

AUGAUGACTd33-LUC

ACTd33-LUC (wild type)

LUCA

B

C

D

100%

54%

33%

4%

LUC activity

scanning

* = AUG

ribosome release

uORF

translation

posttermination scanning

Figure 4.3. A ribosome flow model for the inhibition of downstream (re)initiation byuORFs comprising uAUGs. The fates of (40S) ribosomal subunits entering the 5’-ends of AUG-mutated bicistronic ACTd33-LUC mRNAs are depicted; the thickness of the respectivearrows corresponds to a semi-quantitative representation. The LUC activity values normalised to ∆AUGACTd33-LUC are taken from Figure 3.10B. (A) Virtually all scanning 40S subunits start translation at the main ORF on an uORF-free mRNA. (B) A single uAUG with an inframe stop codon diverts most of the scanning subunits to translation of the uORF, a small part will reach the downstream AUG via ‘leaky scanning’. Having translated the uORF, part of the ribosomeswill be released from the mRNA, some subunits will resume scanning and reinitiate translation.(C) A tandem start codon enhances recognition of the uORF; initiation at the downstream ORF will drop because of significantly reduced ‘leaky scanning’ and an elevated release of ribosomesat the uORF stop codon. (D) On the ‘wild-type’ uORF of construct ACTd105-LUC 40S subunits cannot only initiate at 3 uAUGs, but scanning, elongating and terminating ribosomes arerepeatedly blocked by ribosomes at translation initiation and termination sites. This blockagemultiplies the negative effects of all uORFs, resulting in only a small fraction of ribosomesreaching the LUC main ORF.

109

Page 110: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

comprise internal AUGs, and in cases of overlapping uORF such as in YAP2

and connexin-41 these appear to be used deliberately as a means of regulation.

The efficiency of uORF recognition and initiation upstream of the LUC

ORF is also related closely to the stability of the uORF-containing mRNA

(Figure 3.10). Effective translation of the upstream ORF and lack of proper

termination results in accelerated mRNA decay, whereas translation of

luciferase due to leaky scanning or absence of uORFs stabilises the mRNA.

4.3 Future perspectives The results of this study indicate that posttermination scanning of the 3'-UTR

contributes only minimally to ribosome recycling within the mRNA. Recent work

suggests that the sites of termination and initiation are held in close proximity

via the interaction chain eRF3 – Pab1p – eIF4F, thus providing a different

model for ribosomal 'channelling'. Preliminary experiments by Uchida et al.

(2002) indicate that eRF3 – Pab1p binding is involved only in subsequent

rounds of translation, and not in the de novo formation of the 80S complex. In

order to test this type of model further, it will be a major challenge to

demonstrate the existence of 'channelling' ribosomes of this kind and to

determine whether the whole 80S ribosome or its subunits are guided back to

the 5'-end. Genetically, it might be possible to dissect the contribution of Pab1p

and eRF3 to ribosome recycling by fusing the eIF4G-binding site of Pab1p to

eRF3 in a mutant background with the eIF4G – Pab1p interaction disrupted

(e.g. strain YAS1947).

The second major fate for the posttermination ribosome (in addition to

recycling via release or 'channelling') is reinitiation. Although it appears now to

be established that ribosomes lose the ability to reinitiate translation during

translation, a comprehensive investigation into the dependence on the various

initiation factors is still missing. For this, the truncated actin – luciferase series

of constructs can be used to isolate posttermination ribosomes at different

stages. Crosslinked at an intercistronic secondary structure, the associated

(initiation) factors could be analysed, e.g. by co-immunoprecipitation. Likewise,

by varying the length of the intercistronic spacer, reassociation of the factors

could be observed.

110

Page 111: Translation reinitiation and ribosome recycling - PhD Thesis

4. Discussion

In another aspect of this work, it was shown that uORFs can reduce

downstream initiation to an unexpected high extent when ribosomes are also

able to initiate and terminate within the uORF via 'leaky scanning'. These

ribosomes appear to block scanning or elongating ribosomes in a manner

similar to ribosomes stalled at a uORF termination site. It would be worthwhile

to investigate the occurrence of this phenomenon on mRNAs with 'wild type'

uORFs and whether it is used to regulate gene expression on these transcripts.

111

Page 112: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

5. REFERENCES Abastado, J.-P., Miller, P. F., Jackson, B. M., and Hinnebusch, A. G. 1991.

Suppression of ribosomal reinitiation at upstream open reading frames in amino acid-starved cells forms the basis for GCN4 translational control. Mol. Cell. Biol. 11:486-496.

Algire, M. A., Maag, D., Savio, P., Acker, M. G., Tarun, S. Z. Jr., Sachs, A. B., Asano, K., Nielsen, K. H., Olsen, D. S., Phan, L., Hinnebusch, A. G., and Lorsch, J. R. 2002. Development and characterization of a reconstituted yeast translation initiation system. RNA 8:382-397.

Anand, M., Chakraburtty, K., Marton, M. J., Hinnebusch, A. G., and Kinzy, T. G. 2003. Functional interactions between yeast translation eukaryotic elongation factor (eEF) 1A and eEF3. J. Biol. Chem. 278:6985-6991.

Arst, H. N. Jr., and Sheerins, A. 1996. Translational initiation competence, ‘leaky scanning’ and translational reinitiation in areA mRNA of Aspergillus nidulans. Mol. Microbiol. 19:1019-1024.

Atkin, A. L., Altamura, N., Leeds, P., and Culbertson, M. R. 1995. The majority of yeast UPF1 co-localizes with polyribosomes in the cytoplasm. Mol. Biol. Cell. 6:611-625.

Atkin, A. L., Schenkman, L. R., Eastham, M., Dahlseil, J. F., Lelivelt, M. J., and Culbertson, M. R. 1997. Relationship between yeast polyribosomes and Upf proteins required for nonsense mRNA decay. J. Biol. Chem. 272:22163-22172.

Aziz, N., and Munro, H. N. 1987. Iron regulates ferritin mRNA translation through a segment of its 5’ untranslated region. Proc. Natl. Acad. Sci. USA 84:8478-8482.

Baglioni, C., Vesco, C., Jacobs-Lorena, M. 1969. The role of ribosomal subunits in mammalian cells. Cold Spring Harbor Symp. Quant. Biol. 34:555–566.

Bakheet, T., Frevel, M., Williams, B. R., Greer, W., and Khabar, K. S. 2001. ARED: human AU-rich element-containing mRNA database reveals an unexpectedly diverse functional reportoire of encoded proteins. Nucleic Acids Res. 29:246-254.

Barilla, D., Lee, B. A., and Proudfoot, N. 2001. Cleavage/polyadenylation factor IA associates with the carboxy-terminal domain of RNA polymerase II in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 98:445-450.

Beelman, C. A., and Parker, R. 1994. Differential effects of translational inhibition in cis and in trans on the decay of the unstable yeast MFA2 mRNA. J. Biol. Chem. 269:9687-9692.

Belgrader, P., Cheng, J., and Maquat, L. E. 1993. Evidence to implicate translation by ribosomes in the mechanism by which nonsense codons reduce the nuclear level of human triosephosphate isomerase mRNA. Proc. Natl. Acad. Sci. USA 90:482-486.

112

Page 113: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Bertram, G., Bell, H. A., Ritchie, D. W., Fullerton, G., and Stansfield, I. 2000. Terminating eukaryotic translation: domain 1 of release factor eRF1 functions in stop codon recognition. RNA 6:1236-1247.

Bhattacharya, A., Czaplinski, K., Trifillis, P., He, F., Jacobson, A., and Peltz, S. W. 2000. Characterization of the biochemical properties of the human Upf1 gene product that is involved in nonsense-mediated mRNA decay. RNA 6:1226-1235.

Bohnsack, M. T., Regener, K., Schwappach, B., Saffrich, R., Paraskeva, E., Hartmann, E., and Gorlich, D. Exp5 exports eEF1A via tRNA from nuclei and synergizes with other transport pathways to confine translation to the cytoplasm. EMBO J. 21:6205-6215.

Bonetti, R., Fu, L. Moon, J., and Bedwell, D. M. 1995. The efficiency of translation termination is determined by a synergistic interplay between upstream and downstream sequences in Saccharomyces cerevisiae. J. Mol. Biol. 251:334-345.

Borman, A. M., Michel, Y. M., Malnou, C. E., and Kean, K. M. 2002. Free poly(A) stimulates capped mRNA translation in vitro through the eIF4G-poly(A)-binding protein interaction. J. Biol. Chem. 277:36818-36824.

Brown, C. E., and Sachs, A. 1998. Poly(A) tail length control in Saccharomyces cerevisiae occurs by message-specific deadenylation. Mol. Cell. Biol. 18:6548-6559.

Buckingham, R. H., Gentzmann, G., and Kisselev, L. 1997. Polypeptide chain release factors. Mol. Microbiol. 147:255-269.

Butt, J., Kim, H. Y., Basilion, J. P., Cohen, S., Iwai, K., Philpott, C. C., Altschul, S., Klausner, R. D., and Rouault, T. A. 1996. Differences in the RNA binding sites of the iron regulatory proteins and potential target diversity. Proc. Natl. Acad. Sci. USA. 93:4345-4349.

Cao, J. H., and Geballe, A. P. 1995. Translational inhibition by a human cytomegalovirus upstream open reading frame despite inefficient utilization of its AUG codon. J. Virol. 69:1030-1036.

Cao, J., and Geballe, A. P. 1996. Coding sequence-dependent ribosomal arrest at termination of translation. Mol. Cell. Biol. 16:603-608.

Caponigro, G., and Parker, R. 1996. Mechanisms and control of mRNA turnover in Saccharomyces cerevisiae. Microbiol. Rev. 60:233-249.

Caruthers, J. M., Johnson, E. R., McKay, D. B. 2000. Crystal structure of yeast initiation factor 4A, a DEAD-box RNA helicase. Proc. Natl. Acad. Sci. USA 97: 3080–3085.

Cavener, D. R., and Ray, S. C. 1991. Eukaryotic start and stop translation sites. Nucleic Acids Res. 19:3185-3192.

Chaudhuri, J., Chowdhury, D., and Maitra, U. 1999. Distinct functions of eukaryotic translation initiation factors eIF1A and eIF3 in the formation of the 40 S ribosomal preinitiation complex. J. Biol. Chem. 274:17975-17980.

113

Page 114: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Chernoff, Y. O., Derkach, I. L., and Inge-Vechtomov, S. G. 1993. Mulcticopy SUP35 gene induces de-novo appearance of psi-like factors in the yeast Saccharomyces cerevisiae. Curr. Genet. 24:268-270.

Chernoff, Y. O., Lindquist, S. L., Ono, B., Inge-Vechtomov, S. G., and Liebman, S. W. 1995. Role of chaperon protein Hsp104 in propagation and maintenance of the yeast prion-like factor [PSI+]. Science 268:880-883.

Christensen, A. K., Kahn, L. E., and Bourne, C. M. 1987. Circular polysomes predominate on the rough endoplasmic reticulum of somatotropes and mammotropes in the rat anterior pituitary. Am. J. Anat. 178:1-10.

Cigan, A. M., and Donahue, T. F. 1987. Sequence and structural features associated with translation initiator regions in yeast - a review. Gene 59:1-18.

Cigan, A. M., Pabich, E. K., and Donahue, T. F. 1988. Mutational analysis of the HIS4 translational initiator region in Saccharomyces cerevisiae. Mol. Cell. Biol. 8:2964-2975.

Clements, J. M., Laz, T. M., and Sherman, F. 1988. Efficiency of translation initiation by non-AUG codons in Saccharomyces cerevisiae. Mol. Cell. Biol. 8:4533-4536.

Coller, J. M., Gray, N. K., and Wickens, M. P. 1998. mRNA stabilzation by poly(A) binding protein is independent of poly(A) and requires translation. Genes Dev. 12:3226-3235.

Cosson, B., Couturier, A., Chabelskaya, S., Kiktev, D., Inge-Vechtomov, S., Philippe, M., and Zhouravleva, G. 2002. Poly(A)-binding protein acts in translation termination via eukaryotic release factor 3 interaction and does not influence [PSI+] propagation. Mol. Cell. Biol. 22:3301-3315.

Cox, B. S. 1965. Ψ, a cytoplasmic suppressor of super-suppressor in yeast. Heredity 20:505-521.

Craig, A. W. B., Haghighat, A., Yu, A. T. K., and Sonenberg, N. 1998. Interaction of polyadenylate-binding protein with the eIF4G homologue PAIP enhances translation. Nature 392:520–523.

Cui, Y., Hagan, K. W., Zhang, S., and Peltz, S. W. 1995. Identification and characterization of genes that are required for the accelerated degradation of mRNAs containing a premature translational termination codon. Genes Dev. 9:423-436.

Curatola, A. M., Nadal, M. S., and Schneider, R. 1995. Rapid degradation of AU-rich elements (ARE) mRNAs is activated by ribosome transit and blocked by secondary structures at any position 5' to the ARE. Mol. Cell. Biol. 15:6331-6340.

Czaplinski, K., Weng, Y., Hagan, K. W., Peltz, S. W. 1995. Purification and characterization of the Upf1 protein: a factor involved in translation and mRNA degradation. RNA 1:610–623.

Czaplinski, K., Ruiz-Echevarria, M. J., Paushkin, S. V., Han, X., Weng, Y., Perlick, H. A., Dietz, H. C., Ter-Avanesyan, M. D., and Peltz, S. W.

114

Page 115: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

1998. The surveillance complex interacts with the translation release factors to enhance termination and degrade aberrant mRNAs. Genes Dev. 12:1665-1677.

Czaplinski, K., Ruiz-Echevarria, M. J., González, C. I., and Peltz, S. W. 1999. Should we kill the messenger? The role of the surveillance complex in translation termination and mRNA turnover. Bioessays 21:685-696.

Czaplinski, K., Majlesi, N., Banerjee, T., and Peltz, S. W. 2000. MttI is a Upf1-like helicase that interacts with the translation termination factors and whose overexpression can modulate termination efficiency. RNA 6:730-743.

Dahlberg, J. E., Lund, E., and Goodwin, E. B. 2003. Nuclear translation: what is the evidence? RNA 9:1-8.

Day, D. A., and Tuite, M. F. 1998. Post-transcriptional gene regulatory mechanism in eukaryotes: an overview. J. Endocrinology 157:361-371.

de la Cruz B., J., Prieto S., and Scheffler, I. E. 2002. The role of the 5’ untranslated region (UTR) in glucose-dependent mRNA decay. Yeast 19:887-902.

de la Cruz, J., Lost, I., Kressler, D., and Linder, P. 1997. The p20 and DED1 proteins have antagonistic roles in eIF4E-dependent translation in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 94:5201-5206.

de Wet, J. R., Wood, K. V., de Luca, M., Helinski, D. R., and Subramani, S. 1987. Firefly luciferase gene: structure and expression in mammalian cells. Mol. Cell. Biol. 7:725-737.

Derkatch, I. L., Bradley, M. E., and Liebman, S. W. 1998. Overexpression of the SUP45 gene encoding a Sup35p-binding protein inhibits the induction of the de novo appearance of the [PSI+] prion. Proc. Natl. Acad. Sci. USA 95:2400-2405.

Didichenko, S. A., Ter-Avanesyan, M. D., and Smirnov, V. N. 1991 Ribosome-bound EF-1 alpha-like protein of yeast Saccharomyces cerevisiae. Eur. J. Biochem. 198:705-711.

Doel, S. M., McCready, S. J., Nierras, C. R., and Cox, B. S. 1994. The dominant PNM2 mutation which eliminates the Ψ factor of Saccharomyces cerevisiae is a result of missense mutation in the SUP35 gene. Genetics 137:1-12.

Dominguez, D., Altmann, M., Benz, J., Baumann, U., and Trachsel, H. 1999. Interaction of translation initiation factor eIF4G with eIF4A in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 274:26720-26726.

Donahue, T. F., Farabaugh, P. J., and Fink, G. R. 1981. Suppressible glycine and proline four base codons. Science 212:455-457.

Düvel, K., Egli, C. M., and Braus, G. H. 1999. A single point mutation in the yeast TRP4 affects efficiency of mRNA 3’ end processing and alters selection of the poly(A) site. Nucleic Acids Res. 27:1289-1295.

115

Page 116: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Eaglestone, S. S., Cox, B. S., and Tuite, M. F. 1999. Translation termination efficiency can be regulated in Saccharomyces cerevisiae by environmental stress through a prion-mediated mechanism. EMBO J. 18:1974-1981.

Eustice, D. C., Wakem, L. P., Wilhelm, J. M., and Sherman, F. 1986. Altered 40 S ribosomal subunits in omnipotent suppressors of yeast. J. Mol. Biol. 188:207-214.

Farabaugh, P. J. 1996. Programmed translational frameshifting. Microbiol. Rev. 60:103-134.

Farina, K. L., and Singer, R. H. 2002. The nuclear connection in RNA transport and localization. Trends Cell. Biol. 12:466-472.

Fortes, P., Inada, T., Preiss, T., Hentze, M. W., Mattaj, I. W., and Sachs, A. B. 2000. The yeast nuclear cap binding complex can interact with translation factor eIF4G and mediate translation initiation. Mol. Cell 6:191-196.

Francke, C., Edstrom, J. -E., McDowall, A. W., and Miller Jr, O. L. 1982. Electron microscopic visualization of a discrete class of giant translation units in salivary gland cells of Chironomus tentans. EMBO J. 1:59–62.

Frischmeyer, P. A., van Hoof, A., O’Donnell, K., Guerrerio, A. L., Parker, R., and Dietz, H. C. 2002. An mRNA surveillance mechanism that eliminates transcripts lacking termination codons. Science 295:2258-2261.

Frolova, L., Legoff, X., Rasmussen, H. H., Chepergin, S., Drugeon, G., Kress, M., Arman, I., Celis, J. E., Philippe, M., Justesen, J., and Kisselev, L. 1994. A highly conserved eukaryotic protein family possessing properties of polypeptide-chain release factor. Nature 372:701-703.

Frolova, L., Le Goff, X., Zhouravleva, G., Davydova, E., Philippe, M. and Kisselev, L. 1996. Eukaryotic polypeptide chain release factor eRF3 is an eRF1- and ribosome-dependent guanosine triphosphatase. RNA 2:334-341.

Frolova, L., Merkulova, T. I., and Kisselev, L. L. 2000. Translation termination in eukaryotes: polypeptide release factor eRF1 is composed of functionally and structurally distinct sites. RNA 6:381-390.

Furuichi, Y., LaFiandra, Y., and Shatkin, A. J. 1997. 5’-terminal structure and mRNA stability. Nature 266:235-239.

Gallie, D. R. 1991. The cap and poly(A) tail function synergistically to regulate mRNA translational efficiency. Genes Dev. 5:2108-2116.

Gallie, D. R. 1998. A tale of two termini: A functional interaction between the termini of an mRNA is a prerequisite for efficient translation initiation. Gene 216:1-11.

Gaba, A., Wang, Z., Krishnamoorthy, T., Hinnebusch, A. G., and Sachs, M. S. 2001. Physical evidence for distinct mechanism of translational control by upstream open reading frames. EMBO J. 20:6453-6463.

116

Page 117: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Garcia-Barrio, M. T., Naranda, T., Vazquez de Aldana, C. R., Cuesta, R., Hinnebusch, A. G., Hershey, J. W. B., and Tamame, M. 1995. GCD10, a translational repressor of GCN4, is the RNA-binding subunit of eukaryotic translation initiation factor-3. Genes Dev. 9:1781–1796.

Gasior, E., Herrera, F., McLaughlin, C. S., and Moldave, K. 1979. The analysis of intermediary reactions involved in the protein synthesis, in a cell-free extract of Saccharomyces cerevisiae that translates natural messenger ribonucleic acid. J. Biol. Chem. 254:3970-3976.

Geballe, A. P. 1996. Translational control mediated by upstream AUG codons, p. 173-197. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Geballe, A. P., and Sachs, M. S. 2000. Translational control by upstream open reading frames, p. 595-614. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Gerstel, B., Tuite, M. F., and McCarthy, J. E. G. 1992. The effects of 5’ capping, 3’ polyadenylation and leader composition upon translation and stability of mRNA in a cell-free extract derived from the yeast Saccharomyces cerevisiae. Mol. Microbiol. 6:2339-2348.

Gesteland, R. F., and Atkins, J. F. 1996. Recoding: dynamic reprogramming of translation. Annu. Rev. Biochem. 65:741-768.

Gesteland, R. F., Weiss, R. B., and Atkins, J. F. 1992. Recoding: reprogramming genetic decoding. Science 257:1640-1641.

Gingras, A.-C., Raught, B., and Sonenberg, N. 1999. eIF4 initiation factors: Effectors of mRNA recruitment to ribosomes and regulators of translation. Annu. Rev. Biochem. 68:913-963.

González C. I., Ruiz-Echevarria, M. J., Vasudevan, S., Henry, M. F., and Peltz, S. W. 2000. The yeast hnRNP-like protein Hrp1/Nab4 marks a transcript for nonsense-mediated mRNA decay. Mol. Cell 5:489-499.

González C. I., Bhattacharya, A., Wang, W., and Peltz, S. W. 2001. Nonsense-mediated mRNA decay in Saccharomyces cerevisiae. Gene 274:15-25.

Goidl, J. A., and Allen, W. R. 1978. Does protein synthesis occur within the nucleus? Trends Biochem. Sci. 3:225-

Graber, J. H., Cantor, C. R., Mohr, S. C., and Smith, T. F. 1999. Genomic detection of new yeast pre-mRNA 3’-end-processing signals. Nucleic Acids Res. 27:888-894.

Grant, C. M., and Hinnebusch, A. G. 1994. Effect of sequence context at stop codons on efficiency of reinitiation in GCN4 translational control. Mol. Cell. Biol. 14:606-618.

Grant, C. M., Miller, P. F., and Hinnebusch, A. G. 1994. Requirements for intercistronic distance and level of eukaryotic initiation factor 2 activity in

117

Page 118: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

reinitiation on GCN4 mRNA vary with the downstream cistron. Mol. Cell. Biol. 14:2616-2628.

Grant, C. M., Miller, P. F., and Hinnebusch, A. G. 1995. Sequences 5’ of the first upstream open reading frame in GCN4 mRNA are required for efficient translational reinitiation. Nucleic Acids Res. 23:3980-3988.

Gray, N. K., Quick, S., Goossen, B., Constable, A., Hirling, H., Kühn, L. C., and Hentze, M. W. 1993. Recombinant iron-regulatory factor functions as an iron-responsive-element-binding protein, a translational repressor and an aconitase. A functional assay for translational repression and direct demonstration of the iron switch. Eur. J. Biochem. 218:657-667.

Groft, C. M., and Burley, S. K. 2002. Recognition of eIF4G by rotavirus NSP3 reveals a basis for mRNA circularization. Mol. Cell 9:1273-1283.

Hagan, K. W., Ruiz-Echevarria, M. J., Quan, Y., and Peltz, S. W. 1995. Characterization of cis-acting sequences and decay intermediates involved in nonsense-mediated mRNA turnover. Mol. Cell. Biol. 15:809-823.

Hamm, J., and Mattaj, I. W. 1990. Monomethylated cap structures facilitate RNA export from the nucleus. Cell 63:109-118.

He, F., and Jacobson, A. 2001. Upf1p, Nmd2p, and Upf3p regulate the decapping and exonucleolytic degradation of both nonsense-containing mRNAs and wild type mRNAs. Mol. Cell. Biol. 21:1515-1530.

He, F., Peltz, S. W., Donahue, J. L., Rosbash, M., and Jacobson, A. 1993. Stabilization and ribosome association of unspliced pre-mRNA in a yeast upf1- mutant. Proc. Natl. Acad. Sci. USA 90:7034-7038.

He, F., Brown, A. H., and Jacobson, A. 1997. Upf1p, Nmd2p, and Upf3p are interacting components of the yeast nonsense-mediated mRNA decay pathway. Mol. Cell. Biol. 17:1580–1594.

Heatin, B., Decker, C., Muhlrad, D., Donahue, J., Jacobson, A., and Parker, R. 1992. Analysis of chimeric mRNAs derived from the STE3 mRNA identifies multiple regions within yeast mRNAs that modulate mRBA decay. Nucleic Acids Res. 20:5365-5373.

Hemmings-Mieszczak, M., and Hohn, T. 1999. A stable hairpin preceded by a short open reading frame promotes nonlinear ribosome migration on a synthetic mRNA leader. RNA 5:1149-1157.

Hemmings-Mieszczak, M., Hohn, T., and Preiss, T. 2000. Termination and peptide release at the upstream open reading frame are required for downstream translation on synthetic shunt-competent mRNA leaders. Mol. Cell. Biol. 20:6212-6223.

Hentze, M. W. 2001. Believe it or not – translation in the nucleus. Science 293:1058-1059.

Herrick, D., Parker, R., and Jacobson, A. 1990. Identification and comparison of stable and unstable mRNAs in Saccharomyces cerevisiae. Mol. Cell. Biol. 10:2269-2284.

118

Page 119: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Hershey, J. W. B. 1991. Translational control in mammalian cells. Annu. Rev. Biochem. 60:717-755.

Hershey, J. W. B., and Merrick, W. C. 2000. Pathway and mechanism of initiation of protein synthesis, p. 33-88. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Hilleren, P., and Parker, R. 1999. mRNA surveillance in eukaryotes: kinetic proofreading of proper translation termination as assessed by mRNP domain organization? RNA 5:711-719.

Hinnebusch, A. G. 1984. Evidence for translational regulation of the activator of general amino acid control in yeast. Proc. Natl. Acad. Sci. USA 81:6442-6446.

Hinnebusch, A. G. 1988. Mechanisms of gene regulation in the general control of amino acid biosynthesis in Saccharomyces cerevisiae. Microbiol. Rev. 52:248-273.

Hinnebusch, A. G. 1996. Translational control of GCN4: gene-specific regulation by phosphorylation of eIF-2, p. 199-204. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Hinnebusch, A. G. 1997. Translational regulation of yeast GCN4. A window on factors that control initiator-tRNA binding to the ribosome. J. Biol. Chem. 272:21661-21664.

Hinnebusch, A. G. 2000. Mechanism and regulation of initiator methionyl-tRNA binding to ribosomes, p. 185-243. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Hirokowa, G., Kiel, M. C., Muto, A., Selmer, M., Raj, V. S., Lijas, A., Igarashi, K., Kaji, H., and Kaji, A. 2002. Posttermination complex disassembly by ribosome recycling factor, a functional tRNA mimic. EMBO J. 21:2272-2281.

Hiroshima, A., and Kaji, A. 1972. Factor-dependent release of ribosomes from messenger RNA. Requirement for two heat-stable factors. J. Mol. Biol. 65:43-58.

Hiroshima, A., and Kaji, A. 1973. Role of elongation factor G and a protein factor on the release of ribosomes from messenger ribonucleic acid. J. Biol. Chem. 248:7580-7587.

Hoshino, S., Imai, M., Kobayashi, T., Uchida, N., and Katada, T. 1999. The eukaryotic polypeptide chain releasing factor (eRF3/GSPT) carrying the translation termination signal to the 3’-Poly(A) tail of mRNA. Direct association of eRF3/GSPT with polyadenylate-binding protein. J. Biol. Chem. 274:16677-16680.

Huang, H.-K., Yoon, H., Hannig, E. M., and Donahue, T. F. 1997. GTP hydrolysis controls stringent selection of the AUG start codon during

119

Page 120: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

translation initiation in Saccharomyces cerevisiae. Genes Dev. 11:2396-2413.

Huff, J. P., Grant, B. J., Penning, C. A., and Sullivan, K. F. 1990. Optimization of routine transformation of Escherichia coli with plasmid DNA. Biotechniques 9:570-572, 574, 576-577.

Hwang, W.-L., and Su, T.-S. 1998. Translational regulation of hepatitis B virus polymerase gene by termination-reinitiation of an upstream minicistron in a length-dependent manner. J. Gen. Virol. 79:2181-2189.

Iborra, F.J., Jackson, D. A., and Cook, P. A. 2001. Coupled translation and transcription within nuclei of mammalian cells. Science 293:1139-1142.

Iizuka, N., Najita, L., Franzusoff, A., and Sarnow, P. 1994. Cap-dependent and cap-independent translation by internal initiation of mRNAs in cell extracts prepared from Saccharomyces cerevisiae. Mol. Cell. Biol. 14:7322-7330.

Iizuka, N., and Sarnow, P. 1997. Translation-competent extracts from Saccharomyces cerevisiae: effects of L-A RNA, 5’ cap, and 3’ poly(A) tail on translational efficiency of mRNAs. Methods 11:353-360.

Imataka, H., Gradi, A., and Sonenberg, N. 1998. A newly identified N-terminal amino acid sequence of human eIF4G binds poly(A) binding protein and functions in poly(A) dependent translation. EMBO J. 17:7480-7489.

Ishigaki, Y., Li, X., Serin, G., and Maquat, L. E. 2001. Evidence for a pioneer round of mRNA translation: mRNAs subject to nonsense-mediated decay in mammalian cells are bound by CBP80 and CBP20. Cell 106:607-617.

Jackson, R. J. 2000. Comparative view of initiation site selection mechanisms, p. 127-184. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Jackson, R. J. 1996. A comparative view of initiation site selection mechanisms, p. 71-112 In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Jacobs, J. S., Anderson, A. R., and Parker, R. P. 1998. The 3’ to 5’ degradation of yeast mRNAs is a general mechanism for mRNA turnover that requires the SKI2 DEVH box protein and the 3’ to 5’ exonucleases of the exosome complex. EMBO J. 17:1497-1506.

Jacobson, A. 1996. Poly(A) metabolism and translation: The close-loop model, p 451-480. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Jacobson, A., and Peltz, S. W. 1996. Interrelationships of the pathways of mRNA decay and translation in eukaryotic cells. Annu. Rev. Biochem. 65:693-739.

Jacobson, A., and Peltz, S. W. 2000. Destabilization of nonsense-containing transcripts in Saccharomyces cerevisiae, p. 827-847. In Hershey, J. W.

120

Page 121: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Janosi, L., Mottagui-Tabar, S., Isaksson, L. A., Sekine, Y., Ohtsubo, E., Zhang, S., Goon, S., Nelken, S., Shuda, M., and Kaji, A. 1998. Evidence for in vivo ribosome recycling, the fourth step in protein biosynthesis. EMBO J. 17:2252-1151.

Joseph, S. 2003. After the ribosome structure: how does translocation work? RNA 9:160-164.

Jukes, T. H, and Osawa, S. 1990. The genetic code in mitochondria and chloroplasts. Experientia 46:1117-11126.

Kaji, A., and Hirokowa, G. 2000. Ribosome recycling factor: an essential factor for protein synthesis, p. 527-539. In Garret, R.A., Douthwaite, S. R., Liljas, A., Matheson, A.T., Moore, P.B., and Noller, H. F. (eds), The ribosome: structure, function antibiotics and cellular interactions. ASM Press, Washington, D.C.

Kessler, S. H., and Sachs, A. B. 1998. RNA recognition motif 2 of yeast Pab1p is required for its functional interaction with eukaryotic translation initiation factor 4G. Moll. Cell. Biol. 18:51-57.

Kim, V. N., Kataoka, N., and Dreyfuss, G. 2001. Role of the nonsense-mediated decay factor hUpf3 in the splicing-dependent exon-exon junction complex. Science 293:1832-1836.

Kiseleva, E. V. 1989. Secretory protein synthesis in Chironomus salivary gland cells is not coupled with protein translocation across endoplasmic reticulum membranes. FEBS Lett. 257:251–253.

Kisselev, L. L., and Buckingham, R. H. 2000. Translational termination comes of age. Trends Biochem. Sci. 25:561-566.

Köhrer, K., and Domdey, H. 1991. Preparation of high molecular weight RNA. Methods Enzymol. 194:398-401.

Koloteva, N., Müller, P. P., and McCarthy, J. E. G. 1997. The position dependence of translational regulation via RNA-RNA and RNA-Protein interactions in the 5’-untranslated region of eukaryotic mRNA is a function of the thermodynamic competence of 40 S ribosomes in translation initiation. J. Biol. Chem. 272:16531-16539.

Kozak, M. 1978. How do eucaryotic ribosomes select initiation regions in messenger mRNA? Cell 15:1109-1123.

Kozak, M. 1980. Role of ATP in binding and migration of 40S ribosomal subunits. Cell 22:459-467.

Kozak, M. 1986. Influences of mRNA secondary structures on initiation by eukaryotic ribosomes. Proc. Natl. Acad. Sci. 83:2850-2854.

Kozak, M. 1987a. Effects of intercistronic length on the efficiency of reinitiation by eukaryotic ribosomes. Mol. Cell. Biol. 7:3438-3445.

121

Page 122: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Kozak, M. 1987b. An analysis of 5’-noncoding sequences from 699 vertebrate messenger RNAs. Nuc. Acids Res. 15:8125-8148.

Kozak, M. 1989a. Circumstances and mechanisms of inhibition of translation by secondary structure in eukaryotic mRNA. Mol. Cell. Biol. 9:5134-5142.

Kozak, M. 1989b. The scanning model for translation: An update. J. Cell Biol. 108:229-241.

Kozak, M. 1990. Downstream secondary structure facilitates recognition of initiator codons by eukaryotic ribosomes. Proc. Natl. Acad. Sci. USA. 87:8301-8305.

Kozak, M. 1991a. A short leader sequence impairs the fidelity of initiation by eukaryotic ribosomes. Gene Expr. 1:111-115.

Kozak, M. 1991b. Effects of long 5’ leader sequences on initiation by eukaryotic ribosomes in vitro. Gene Expr. 1:117-25.

Kozak, M. 1991c. An analysis of vertebrate mRNA sequences: intimidations of translational control. J. Cell. Biol. 115:887-903.

Kozak, M. 1991d. Structural features in eukaryotic mRNAs that modulate the initiation of translation. J. Cell. Biol. 266:19867-19870.

Kozak, M. 1992. A consideration of alternative models for the initiation of translation in eukaryotes. Crit. Rev. Biochem. Mol. Biol. 27:385-402.

Kozak, M. 1999. Initiation of translation in prokaryotes and eukaryotes. Gene 234:187-208.

Kozak, M. 2001a. New ways of initiating translation in eukaryotes? Mol. Cell. Biol. 21:1899-1907.

Kozak, M. 2001b. Constraints on reinitiation of translation in mammals. Nucleic Acids Res. 29:5226-5232.

Kozak, M. 2002. Pushing the limit of the scanning mechanism for initiation of translation. Gene 299:1-34.

Kozak, M., and Shatkin, A. J. 1978. Migration of 40S ribosomal subunits in the presence of edeine. J. Biol. Chem. 253:6568-6577.

Krol, A. 2002. Evolutionarily different RNA motifs and RNA–protein complexes to achieve selenoprotein synthesis. Biochimie 84:765-774.

Kuhn, K. M., DeRisi, J. L., Brown, P. O., and Sarnow, P. 2001. Global and specific translational regulation in the genomic response of Saccharomyces cerevisiae to a rapid transfer form a fermentable to a nonfermentable carbon source. Mol. Cell. Biol. 21:916-927.

LaGrandeur, T., and Parker, R. 1999. The cis acting sequences for the differential decay of the unstable MFA2 and stable PGK1 transcripts in yeast include the context of the translational start codon. RNA 5:420-433.

Laemmli, U. K. 1970. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 227:680-685.

122

Page 123: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Lamphear B. J., Kirchweger, R., Skern, T., and Rhoads, R. E. 1995. Mapping of functional domains in eukaryotic protein synthesis initiation factor 4G (eIF4G) with picornaviral proteases. Implications for cap-dependent and cap-independent translational initiation. J. Biol. Chem. 270:21975-21983.

Le, H., Tanguay, R. L., Balasta, M. L., Wei, C.-C., Browning, K. S., Metz, A. M., Goss, D. J., and Gallie, D. R. 1997. The translation initiation factors eIFiso4G and eIF-4B interact with the poly(A)-binding protein to increase its RNA binding affinity. J. Biol. Chem. 272:16247–16255.

Le Hir, H., Izaurralde, E., Maquat, L. E., and Moore, M. J. 2000. The spliceosome deposits multiple proteins 20-24 nucleotides upstream of mRNA exon-exon junctions. EMBO J. 19:6860-6869.

Lee, B. S., and Culbertson, M. R. Identification of an additional gene required for eukaryotic nonsense mRNA turnover. Proc. Natl. Acad. Sci. USA 92:10354-10358.

Leeds, P., Peltz, S. W., Jacobson, A., and Culbertson, M. R. The product of the yeast UPF1 gene is required for rapid turnover of mRNAs containing a premature translation termination codon. Genes Dev. 5:2303-2314.

Lindquist, S., and Kim, G. 1996. Heat-shock protein 104 expression is sufficient for thermotolerance in yeast. Proc. Natl. Acad. Sci. USA. 93:5301-5306.

Linz, B., Koloteva, N., Vasilescu, S., and McCarthy, J. E. G. 1997. Disruption of ribosomal scanning on the 5’-untranslated region, and not restriction of initiation per se, modulates the stability of nonaberrant mRNAs in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 272:9131-9140.

Londei, P. 1998. A hypothesis on the mechanism of translational initiation. Biochim. Biophys. Acta 1396:169-178.

Lovett, P. S., and Rogers, E. J. 1996. Ribsome regulation by the nascent peptide. Microbiol. Rev. 60:366-385.

Luukkonen, B. G. M., Tan, W., and Schwartz, S. 1995. Efficiency of reinitiation of translation on human immunodeficiency virus type 1 mRNAs is determined by the length of the upstream open reading frame and the intercistronic distance. J. Virol. 69:4086-4049.

Maderazo, A. B., Belk, J. P., He, F., and Jacobson A. 2003. Nonsense-containing mRNAs that accumulate in the absence of a functional nonsense-mediated mRNA decay pathway are destabilized rapidly upon its restitution. Mol. Cell. Biol. 23:842-851.

Maniatis, T., and Reed, R. 2002. An extensive network of coupling gene expression machines. Nature 416:499-506.

Mao, X., Schwer, B., and Shuman, S. 1996. Mutational analysis of the Saccharomyces cerevisiae ABD1 gene: cap methyltransferase activity is essential for cell growth. Mol. Cell. Biol. 16:457-480.

Major, L. L., Edgar, T. D., Yee Yip, P., Isaksson, L. A., and Tate, W. P. 2002. Tandem termination signals: myth or reality? FEBS Lett. 514:84-89.

123

Page 124: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Maquat, L. E. 2000. Nonsense-mediated RNA decay in mammalian cells: A splicing-dependent means to down-regulate the levels of mRNAs that prematurely terminate translation, p. 849-868. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Martínez-Salas, E., Ramos, R., Lafuente, E., and López de Quinto, S. 2001. Functional interactions in internal translation initiation directed by viral and cellular IRES elements. J. Gen. Vir. 82:973-984.

McCarthy, J. E. G. 1998. Posttranscriptional control of gene expression. Microbiol. Mol. Biol. Rev. 62:1492-1553.

McCarthy, J. E., and Gualerzi, C. 1990. Translational control of prokaryotic gene expression. Trends Genet. 6:78-85.

McCarthy, J. E. G., and Kollmus, H. 1995. Cytoplasmic mRNA-protein interactions in eukaryotic gene expression. Trends Biochem. Sci. 20:191-197.

Mendell, J. T., Medghalchi, S. M., Lake, R. G., Noensie, E. N., and Dietz, H. C. 2001. Novel Upf2p orthologues suggest a functional link between translation initiation and nonsense surveillance complexes. Mol. Cell. Biol. 20:8944-8957.

Merrick, W. C. 1992. Mechanism and regulation of eukaryotic protein synthesis. Microbiol. Rev. 56:291-315.

Merrick, W. C., and Hershey, J. W. B. 1996. The pathway and mechanism of eukaryotic protein synthesis, p. 31-70. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Miller 3rd, C. A., Martinat, M. A., and Hyman, L. E. 1998. Assessment of aryl hydrocarbon receptor complex interactions using pBEVY plasmids: expression vectors with bi-directional promoters for use in Saccharomyces cerevisiae. Nucleic Acids Res. 26:3577-3583.

Mitchell, P., and Tollervey, D. 2000. mRNA stability in eukaryotes. Curr. Opinion Gen. Dev. 10:193-198.

Miyasaka, H. 1999. The positive relationship between codon usage bias and translation initiation AUG context in Saccharomyces cerevisiae. Yeast 15:633-637.

Mize, G. J., Ruan, H. J., Low, J. J., and Morris, D. R. 1998. The inhibitory upstream open reading frame from mammalian S-adenosylmethionine decarboxylase mRNA has a strict sequence specificity in critical positions. J. Biol. Chem. 273:32500-32505.

Moffat, J. G., and Tate, W. P. 1994. A single proteolytic cleavage in release factor 2 stabilizes ribosome binding and abolishes peptidyl-tRNA hydrolysis activity. J. Biol. Chem. 269:18899-18903.

Monroe, D., and Jacobson, A. 1990. mRNA poly(A) tail, a 3’ enhancer of translational initiation. Mol. Cell. Biol. 10:3441-3445.

124

Page 125: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Morley, S. J., Curtis, P. S., and Pain, V. M. 1997. eIF4G: translation’s mystery factor begins to yield its secrets. RNA 3:1085-1104.

Morris, D. R. 1995. Growth control in mammalian cells. Prog. Nucleic Acids Res. Mol. Biol. 51:339-363.

Morris, D. R., and Geballe, A. P. 2000. Upstream open reading frames as regulator of mRNA translation. Mol. Cell. Biol. 20:8635-8642.

Mottagui-Tabar, S., Tuite, M. F., and Isaksson, L. A. 1998. The influence of 5’ codon context on translation termination in Saccharomyces cerevisiae. Eur. J. Biochem. 257:249-254.

Mugnier, P., and Tuite, M. F. 1999. Translation termination and its regulation in eukaryotes: Recent insights provided by studies in yeast. Biochemistry 64:1360-1366.

Muhlrad, D., and Parker, R. 1994. Premature translation termination triggers mRNA decapping. Nature 370:578-581.

Muhlrad, D., Decker, C. J., and Parker, R. 1995. Turnover mechanism of the stable yeast PGK1 mRNA. Mol. Cell. Biol. 15:2145-2156.

Muhlrad, D., and Parker, R. 1999a. Aberrant mRNAs with extended 3’ UTRs are substrates for rapid degradation by mRNA surveillance. RNA 5:1299-1307.

Muhlrad, D., and Parker, R. 1999b. Recognition of yeast mRNAs as 'nonsense containing' leads to both inhibition of mRNA translation and mRNA degradation: implications for the control of mRNA decapping. Mol. Cell. Biol. 11:3971-3978.

Munroe, D., and Jacobson, A. 1990. mRNA poly(A) tail, a 3’ enhancer of translation initiation. Mol. Cell. Biol. 10:3441-3455.

Namy, O., Hatin, I., and Rousset, J. P. 2001. Impact of the six nucleotides downstream of the stop codon on translation termination. EMBO Rep. 2:787-793.

Neff, C. L., and Sachs, A. B. 1999. Eukaryotic translation initiation factors 4G and 4A from Saccharomyces cerevisiae interact physically and functionally. Mol. Cell. Biol. 19:5557-5564.

Nett, J. H., Kessl, J., Wenz, T., and Trumpower, B. L. 2001. The AUG start codon in Saccharomyces cerevisiae NSF1 gene can be substituted for by UUG without increased initiation of translation at downstream codons. Eur. J. Biochem. 268:5209-5214.

Neu-Yilik, G., Gehring, N. H., Thermann, R., Frede, U., Hentze, M. W., and Kulozik, A. E. 2001. Splicing and 3’ end formation in the definition of nonsense-mediated decay-competent human β-globin mRNPs. EMBO J. 20:532-540.

Neugebauer, K. M., and Roth, M. B. 1997. Transcription units as RNA processing units. Genes Dev. 11:3279-3285.

Niepel, M., Ling, J., and Gallie, D. R. 1999. Secondary structure in the 5’-leader or 3’-untranslated region reduces protein yield but does not affect

125

Page 126: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

the functional interaction between the 5’-cap and the poly(A) tail. FEBS Lett. 462:79-84.

Novoa, I., and Carrasco, L. 1999. Cleavage of eukaryotic translation initiation factor 4G by exogenous added hybrid proteins containing poliovirus 2Apro in HeLa cells: effects on gene expression. Mol. Cell. Biol. 19:2445-2454.

Oliveira, C. C., Goossen, B., Zanchin, N. I. T., McCarthy, J. E. G., Hentze, M. W., and Stripecke, R. 1993a. Translational repression by the human iron-regulatory factor (IRF) in Saccharomyces cerevisiae. Nucleic Acids Res. 21:5316-5322.

Oliveira, C. C., van den Heuvel, J. J., and McCarthy, J. E. G. 1993b. Inhibition of translational initiation in Saccharomyces cerevisiae by secondary structure: the role of the stability and position of stem-loops in the mRNA leader. Mol. Microbiol. 9:521-532.

Oliveira, C. C., and McCarthy, J. E. G. 1995. The relationship between eukaryotic translation and mRNA stability. J. Biol. Chem. 270:8936-8943.

Otero, L. J., Ashe, M. P., and Sachs, A. B. 1999. The yeast poly(A)-binding protein Pab1p stimulates in vitro poly(A)-dependent and cap-dependent translation by distinct mechanisms. EMBO J. 18:3153-3163.

Pain, V. M. 1996. Initiation of protein synthesis in eukaryotic cells. Eur. J. Biochem. 236:747-771.

Paushkin, S. V., Kushnirov, V. V., Smirnov, V. N., and Ter-Avanesyan, M. D. 1997. Interaction between yeast Sup45p (eRF1) and Sup35p (eRF3) polypeptide chain release factors: implications for prion-dependent regulation. Mol. Cell. Biol. 17:2798-2805.

Paz, I., Abramovitz. L., and Choder, M. 1999. Starved Saccharomyces cerevisiae cells have the capacity to support internal initiation of translation. J. Biol. Chem. 274:21741-21745.

Peltz, S. W., Donahue, J. L., and Jacobson, A. 1992. A mutation in the tRNA nucleotidyltransferase gene promotes stabilization of mRNAs in Saccharomyces cerevisiae. Mol. Cell. Biol. 12:5778-5784.

Peltz, S. W., Brown, A. H., and Jacobson, A. 1993. mRNA destabilization triggered by premature translation termination depends on three mRNA sequence elements and at least on one trans-acting factor. Genes Dev. 7:1737-1754.

Pestova, T. V., Kolupaeva, V. G., Lomakin, I. B., Pilipenko, E. V., Shatsky, I. N., Agol, V. I., and Hellen, C. T. 2001. Molecular mechanism of translation initiation in eukaryotes. Proc. Natl. Acad. Sci. USA 98:7029-7036.

Philipps, G. R. 1965. Haemoglobin synthesis and polysomes in intact factors eIFiso4G and eIF-4B interact with the poly(A)-binding reticulocytes. Nature 205:567–570.

126

Page 127: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Piron, M., Vende, P., Cohne, J., and Poncet, D. 1998. Rotavirus RNA-binding protein NSP3 interacts with eIF4GI and evicts the poly(A)-binding protein from eIF4F. EMBO J. 17:5811-5821.

Plant, E. P., Jacobs, K. L., Harger, J. W., Meskauskas, A., Jacobs, J. L., Baxter, J. L., Petrov, A. N., and Dinman, J. D. 2003. The 9-A solution: how mRNA pseudoknots promote efficient programmed -1 ribosomal frameshifting. RNA 9:168-174.

Preiss, T., and Hentze, M. W. 1999. From factors to mechanisms: translation and translational control in eukaryotes. Curr. Opin. Genet. Dev. 9:515-521.

Presutti, C., Villa, T., Hall, D., Pertica, C., and Bozzoni, I. 1995. Identification of the cis-elements mediating the autogenous control of ribosomal protein L2 mRNA stability in yeast. EMBO J. 14:4022-4030.

Proud, C. G. 2002. Regulation of mammalian translation factors by nutrients. Eur. J. Biochem. 269:5338-5349.

Proudfoot, N. 2000. Connecting transcription to messenger RNA processing. Trends Biocem. Sci. 25:290-293.

Proudfoot, N., Furger, A., and Dye, J. M. 2002. Integrating mRNA processing with transcription. Cell 108:501-512.

Proweller, A., and Butler, J. S. 1996. Ribosomal association of poly(A)-binding protein in poly(A)-deficient Saccharomyces cerevisiae. J. Biol. Chem. 271:10859-10865.

Ptushkina, M., von der Haar, T., Vasilescu, S., Frank, R., Birkenhäger, R., and McCarthy, J. E. G. 1998. Cooperative modulation by eIF4G of eIF4E-binding to the mRNA 5’cap in yeast involves a site partially shared by p20. EMBO J. 17:4798-4808.

Qin, S. L., Xie, A. G., Bonato, M. C., and McLaughlin, C. S. 1990. Sequence analysis of the translation elongation factor 3 from Saccharomyces cerevisiae. J. Biol. Chem. 265:1903-1912.

Ramakrishnan, V. 2002. Ribosome structure and the mechanism of translation. Cell 108:557-572.

Ramirez, C. V., Vilela, C., Berthelot, K., and McCarthy, J. E. G. 2002. Modulation of eukaryotic mRNA stability via the cap-binding translation complex eIF4F. J. Mol. Biol. 318:951-962.

Reed, R., and Hurt, E. 2002. A conserved mRNA export machinery coupled to pre-mRNA splicing. Cell 108:523-531.

Rhoads, R. E. 1993. Regulation of eukaryotic protein synthesis by initiation factors. J. Biol. Chem. 268:3017-3020.

Rose, M. D., Winston, F., and Hieter, P. 1990. Methods in yeast genetics: A laboratory manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Rouault, T. A., and Harford, J. B. 2000. Translational control of ferritin synthesis, p.655-670. In Hershey, J. W. B., Matthews, M., and

127

Page 128: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Rouault, T. A., Hentze, M. W., Caughman, S. W., Harford, J. B., and Klausner, R. D. 1988. Binding of a cytosolic protein to the iron-responsive element of human ferritin messenger mRNA. Science 241:1207-1210.

Rozen, F., Edery, I., Meerovitch, K., Dever, T. E., Merrick, W. C., and Sonenberg, N. 1990. Bidirectional RNA helicase activity of eucaryotic translation initiation factors eIF4A and eIF4F. Mol. Cell. Biol. 10:1134-1144.

Ruiz-Echevarria, M. J., Czaplinski, K., and Peltz, S. W. 1996. Utilizing the GCN4 leader region to investigate the role of the sequence determinants in nonsense-mediated mRNA decay. EMBO J. 15:2810-2819.

Ruiz-Echevarria, M., González, C. I., and Peltz, S. W. 1998. Identifying the right stop: determining how the surveillance complex recognises and degrades an aberrant mRNA. EMBO J. 17:575-589.

Ruiz-Echevarria, M., and Peltz, S. W. 2000. The RNA binding protein Pub1 modulates the stability of transcripts containing upstream open reading frames. Cell 101:741-751.

Ruiz-Echevarria, M., Munshi, R., Tomback, J., Kinzy, T. G., and Peltz, S. W. 2001. Characterization of a general stabilizer element that blocks deadenylation dependent mRNA decay. J. Biol. Chem. 276:30995-31003.

Ryabova, L. A., Pooggin, M. M., and Hohn, T. 2002. Viral strategies of translation initiation: ribosomal shunt and reinitiation. Prog. Nucleic Acid Res. Mol. Biol. 72:1-39.

Sachs, A. 2000. Physical and functional interactions between the mRNA cap structure and the poly(A) tail, p. 447-465. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Sachs, A., Sarnow, P., and Hentze, M. 1997. Starting at the beginning, middle, and end: translation initiation in eukaryotes. Cell 89:831-838.

Sachs, A. B, and Davis, R. W. 1989. The poly(A) binding protein is required for poly(A) shortening and 60S ribosomal subunit-dependent translation initiation. Cell 58:857-867.

Sachs, A. B., and Varani, G. 2000. Eukaryotic translation initiation: there are (at least) two sides to every story. Nat. Struct. Biol. 7:356-361.

Sagliocco, F. A., Vega Laso, M. R., Zhu, D., Tuite, M. F., McCarthy, J. E. G., and Brown, A. J. 1993. The influence of 5’-secondary structures upon ribosome binding to mRNA during translation in yeast. J. Biol. Chem. 268:26522-26530.

Sakurai, A., Fujimori, S., Kochiwa, H., Katimura-Abe, S., Washio, T., and Saito, R. 2002. On biased distribution of introns in various eukaryotes. Gene 300:89-95.

128

Page 129: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Sambrook, J., Fritsch, E. F., and Maniatis, T. 1989. Molecular cloning: A laboratory manual. 2nd edn. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY

Schiestl, R. H., and Gietz, R. D. 1989. High efficiency transformation of intact yeast cells using single stranded nucleic acids as carrier. Curr. Genet. 16:339-346.

Schneider, R. 2001. New ways of initiating translation in eukaryotes? Mol. Cell. Biol. 21:8238-8246.

Scolnik, E. Tompkins, R., Caskey, T., and Nirenberg, M. 1968. Release factors differ in specificity for terminator codons. Proc. Natl. Acad. Sci. USA 61:768-774.

Searfoss, A., Dever, T. E., and Wickner, R. 2001. Linking the 3’ poly(A) tail to the subunit joining step of translation initiation: relations of Pab1p, eukaryotic translation initiation factor 5B (Fun12p), and Ski2P-Slh1p. Mol. Cell. Biol. 21:4900-4908.

Shatkin, A. J. 1976. Capping of eukaryotic mRNAs. Cell 9:645-653.

Shatkin, A. L., and Manley, J. L. 2000. The ends of the affair: capping and polyadenylation. Nat. Struct. Biol. 7:838-842.

Sherman, F., Fink, G. R., and Hicks, J. B. 1986. Laboratory course manual for methods in yeast genetics. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY

Shirley, R. L., Lelivelt, M. J., Schenkman, L. R., Dahlseid, J. N., and Culbertson, M. R. 1998. A factor required for nonsense-mediated mRNA decay in yeast is exported from the nucleus to the cytoplasm by a nuclear export signal sequence. J. Cell Sci. 111:3129–3143.

Si, K., and Maitra, U. 1999. The Saccharomyces cerevisiae homologue of mammalian translation initiation factor 6 does not function as a translation initiation factor. Mol. Cell. Biol. 19:1416-1426.

Smith, P. K., Krohn, R. I., Hermanson, G. T., Mallia, A. K., Gartner, F. H., Provenzano, M. D., Fujimoto, E. K., Goeke, N. M., Olson, B. J., and Klenk, D. C. 1985. Measurement of protein using bicinchoninic acid. Anal. Biochem. 150:76-85.

Somogyi, P., Jenner, A. J., Brierley, I., and Inglis, S. C. 1993. Ribosomal pausing during translation of an RNA pseudoknot. Mol. Cell. Biol. 9:5134-5142.

Song, H., Mugnier, P., Das, A. K., Webb, H. M., Evans, D. R., Tuite, M. F., Hemmings, B. A., and Barford, D. 2000. The crystal structure of human eukaryotic release factor eRF1–mechanism of stop codon recognition and peptidyl-tRNA hydrolysis. Cell 100:311-321.

Standart, N., and Jackson, R. 1994. Translational regulation. Y the message is masked? Curr. Biol. 4:939-941.

Stansfield, I., Jones, K. M., Kushnirov, V. V., Dagkesamanskaya, A. R., Poznyakovski, A. I., Paushkin, S. V., Nierras, C. R., Cox, B. S., Ter-

129

Page 130: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Avanesyan, M. D., and Tuite M. F. 1995. The products of the SUP45 (eRF1) and SUP35 genes interact to mediate translation termination in Saccharomyces cerevisiae. EMBO J. 14:4365-4373.

Steitz, J. A., and Jakes, K. 1975. How ribosomes select initiator regions in mRNA: base pair formation between the 3’terminus of 18S rRNA and the mRNA during initiation of protein synthesis in Escherichia coli. Proc. Natl. Acad. Sci. USA 72:4734-4738.

Stoffler, D., Fahrenkrog, B., and Aebi, U. 1999. The nuclear pore complex: from molecular architecture to functional dynamics. Curr. Opin. Cell. Biol. 11:391-401.

Sträßer, K., and Hurt, E. 1999. Nuclear RNA export in yeast. FEBS Let. 452:77-81.

Stripecke, R., Oliveira, C. C., McCarthy, J. E. G., and Hentze, M. W. 1994. Proteins binding to 5’UTR sites: a general mechanism for translational regulation of mRNAs in human and yeast cells. Mol. Cell. Biol. 14:5898-5909.

Tarun, S. Z. Jr., and Sachs, A. B. 1995. A common function for mRNA 5’ and 3’ ends in translation initiation in yeast. Genes Dev. 9:2997-3007.

Tarun, S. Z. Jr., and Sachs, A. B. 1996. Association of the yeast poly(A) tail binding protein with translation initiation factor eIF-4G. EMBO J. 15:7168-7177.

Tarun, S. Z. Jr., Wells, S. E., Deardoff, J. A., and Sachs, A. B. 1997. Translation factor eIF4G mediates in vitro poly(A) tail-dependent translation. Proc. Natl. Acad. Sci. USA 94:9046-9051.

Tate, W. P., Poole, E. S., Dalphin, M. E., Major, L. L., Crawford, D. J., and Mannering, S. A. 1996. The translational stop signal: codon with a context, or extended factor recognition element? Biochimie 78:945-952.

Tatsumi, H., Masuda, T., and Nakano, E. 1988. Synthesis of enzymatically active firefly luciferase in yeast. Agric. Biol. Chem. 52:1123-1127.

Tenson, T., and Ehrenberg, M. 2002. Regulatory nascent peptides in the ribosomal tunnel. Cell 108:591-594.

Ter-Avanesyan, M. D., Dagkesamanskaya, A. R., Krushnirov, V. V., and Smirnov, V. N. 1994. The SUP35 omnipotent suppressor gene is involved in the maintenance of the non-Mendelian determinant [PSI+] in the yeast Saccharomyces cerevisiae. Genetics 137:671-676.

Thomas, B. J., and Rothstein, R. 1989. Elevated recombination rates in transcriptionally active DNA. Cell 56:619-630.

Trachsel, H. 1996. Binding of initiator methionyl-tRNA to ribosomes, p. 113-138. In Hershey, J. W. B., Matthews, M., and Sonenberg, N. (eds), Translational control of gene expression. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y.

Triana, F. J. K., Nierhaus, H., Ziehler, J., and Chakratburtty, K. 1993. Defining the function of EF-3, a unique elongation factor in low fungi, p.

130

Page 131: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

327-338. In Nierhaus, K. H. et al. (eds), The translational apparatus: Structure, function, regulation, evolution. Plenum Press, N.Y.

Tuite, M. F., and Lindquist, S. L. 1996. Maintenance and inheritance of yeast prion. Trends. Genet. 12:467-471.

Uchida, N., Hoshino, S., Imataka, H., Sonenberg, N., and Katada, T. 2002. A novel role of the mammalian GSPT/eRF3 associating with poly(A)-binding protein in cap/poly(A)-dependent translation. J. Biol. Chem. 277:50286-50292.

Urakov, V. N., Valouev, I. A., Lewitin, E. I., Paushkin, S. V., Kosorukov, V. S., Kushnirov, V. V, Smirnov, V. N., and Ter-Avanesyan, M. D. 2001. Itt1p, a novel protein inhibiting translation termination in Saccharomyces cerevisiae. BMC Mol. Biol. 2:9.

van Hoof, A., and Parker, R. 1999. The exosome: a proteasome for RNA? Cell 99:347-350.

van Hoof, A., Frischmeyer, P. A., Dietz, H. C., and Parker, R. 2002. Exosome-mediated recognition and degradation of mRNAs lacking a termination codon. Science 295:2262-2264.

Valouev, I. A., Kushnirov, V. V., and Ter-Avanesyan, M. D. 2002. Yeast polypeptide chain release factors eRF1 and eRF3 are involved in cytoskeleton organization and cell cycle regulation. Cell. Motil. Cytoskeleton. 52:161-173.

Vasudevan, S., and Peltz, S. W. 2001. Regulated ARE-mediated mRNA decay in Saccharomyces cerevisiae. Mol. Cell 7:1191-1200.

Vasudevan, S., Peltz, S. W., and Wilusz, C. J. 2002. Non-stop decay--a new mRNA surveillance pathway. Bioessays 24:785-788.

Vega Laso, M. R., Zhu, D., Sagliocco, F., Brown, A. J., Tuite, M. F., and McCarthy, J. E. G. 1993. Inhibition of translation initiation in the yeast Saccharomyces cerevisiae as a function of the stability and the position of hairpin structures in the mRNA leader. J. Biol. Chem. 268:6453-6462.

Vilela, C., Linz, B., Rodrigues-Pousada, C., and McCarthy, J. E. G. 1998. The yeast transcription factor genes YAP1 and YAP2 are subject to differential control at the levels of both translation and mRNA stability. Nucleic Acids Res. 26:1150-1159.

Vilela, C., Ramirez, C. V., Linz, B., Rodrigues-Pousada, C., and McCarthy. J. E. 1999. Posttermination ribosome interactions with the 5’UTR modulate yeast mRNA stability. EMBO J. 18:3139-3152.

Vilela, C., Velasco, R., Ptushkina, M., and McCarthy, J. E. G. 2000. The eukaryotic mRNA decapping protein Dcp1 interacts physically and functionally with the eIF4F translation initiation complex. EMBO J. 19:4372-4382.

von der Haar T., Ball, P. D., and McCarthy J. E. 2000. Stabilization of eukaryotic initiation factor 4E binding to the mRNA 5’-cap by domains of eIF4G. J. Biol. Chem. 275:30551-30555.

131

Page 132: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Wakiyama, M., Imataka, H., and Sonenberg, N. 2000. Interaction of eIF4G with poly(A)-binding protein stimulates translation and is critical for Xenopus oocyte maturation. Curr. Biol. 10:1147-1150.

Wang, Z., Gaba, A., and Sachs, M. S. 1999. A highly conserved mechanism of regulated ribosome stalling mediated by fungal arginine attenuator peptides that appears independent of the charging status of arginyl-tRNAs. J. Biol. Chem. 274:37565-37574.

Wang, W., Czaplinksi, K., Rao, Y., and Peltz, S. W. 2001a. The role of Upf proteins in modulating the translational read-through of nonsense-containing transcripts. EMBO J. 20:880-890.

Wang, Y., Long Liu, C., Storey, J. D., Tibshirani, R. J., Herschlag, D., and Brown, P. O. 2001b. Precision and functional specificity in mRNA decay. Proc. Natl. Acad. Sci. USA 99:8560-5865.

Wang, J., Vock, V. M., Li, S., Olivas, O. R., and Wilkinson, M. F. 2002. A quality-control pathway that downregulates aberrant TCR transcripts by a mechanism requiring UPF2 and translation. J. Biol. Chem. 277:18489-18493.

Wei, C.-C., Balasta, M. L., Ren, J., and Goss, D. J. 1998. Wheat germ poly(A) binding protein enhances the binding affinity of eukaryotic initiation factor 4F and (iso)4F for cap analogues. Biochemistry 37:1910-1916.

Weng, Y., Czaplinski, K., and Peltz, S. W. 1996a. Genetic and biochemical characterization of mutations in the ATPase and helicase regions of the Upf1 protein. Mol. Cell. Biol. 16:5477-5490.

Weng, Y., Czaplinski, K., and Peltz, S. W. 1996b. Identification and characterization of mutations in the UPF1 gene that affect nonsense suppression and the formation of the Upf protein complex but not mRNA turnover. Mol. Cell. Biol. 16:5491-5506.

Weng, Y., Czaplinski, K., and Peltz, S. W. 1998. ATP is a cofactor of the Upf1 protein that modulates its translation termination and RNA binding activities. RNA 4:205-214.

Wells, S. E., Hillner, P. E, Vale, R. D., and Sachs, A. B. 1998. Circularization of mRNA by eukaryotic translation initiation factors. Mol. Cell 2:135-140.

Wiliamson, J. R., Raghurman, M. K., and Cech, T. R. 1989. Monovalent cation-induced structure of telomeric DNA; the G-quartet model. Cell 59:817-880.

Wilkinson, M. F., and Shyu, A.-B. 2002. RNA surveillance by nuclear scanning? Nature Cell Biol. 4:144-147.

Wilusz, C. J., Wormington, M., and Peltz, S. W. 2001. The cap-to-tail guide to mRNA turnover. Nature 2:237-246.

Wolfe, C. L., Hopper, A. K., and Martin, N. C. 1996. Mechanisms leading to and the consequences of altering the normal distribution of ATP(CTP):tRNA nucleotidyltransferase in yeast. J. Biol. Chem. 271:4679-4686.

132

Page 133: Translation reinitiation and ribosome recycling - PhD Thesis

5. References

Yang, W., and Hinnebusch, A. G. 1996. Identification of a regulatory subcomplex in the guanine nucleotide exchange factor eIF2B that mediates inhibition by phosphorylated eIF2. Mol. Cell. Biol. 16:6603-6616.

Zanchin, N. I. T., and McCarthy, J. E. G. 1995. Characterization of the in vivo phosphorylation sites of the mRNA-cap-binding complex proteins eukaryotic initiation factor-4E and p20 in Saccharomyces cerevisiae. J. Biol. Chem. 270:26505-26510.

Zenklusen, D., and Strutz, F. 2001. Nuclear export of mRNA. FEBS Let. 498:150-156.

Zhang S, Ruiz-Echevarria, M. J., Quan, Y., and Peltz, S. W. 1995. Identification and characterization of a sequence motif involved in nonsense-mediated mRNA decay. Mol. Cell. Biol. 15:2231-2244.

Zhang, Y., and Spremulli, L. L. 1998. Identification and cloning of human mitochondrial translational release factor 1 and the ribosome recycling factor. Biochim. Biophys. Acta 1443:245-250.

Zhang, J., Sun, X., Qian, Y., LaDuca, J. P., and Maquat, L. E. 1998. At least one intron is required for the nonsense-mediated decay of triosephosphate isomerase mRNA: A possible link between nuclear splicing and cytoplasmic translation. Mol. Cell. Biol. 18:5272–5283.

Zhao, J., Hyman, L., and Moore, C. 1999. Formation of mRNA 3’ ends in eukaryotes: mechanism, regulation and interrelationship with other steps in mRNA synthesis. Microbiol. Mol. Biol. Rev. 63:405-455.

Zhou, W., Edelman, G. M., and Mauro, V. P. 2001. Transcript leader regions of two Saccharomyces cerevisiae mRNAs contain internal ribosome entry sites that function in living cells. Proc. Natl. Acad. Sci. USA 98:1531-1536.

Zhou, Z., Licklider, L. J., Gygi, S. P., and Reed, R. 2002. Comprehensive proteomic analysis of the human spliceosome. Nature 419:182-185.

Zhouravleva, G., Frolova, L., Le Goff, X., Le Guellex, R., Inge-Vechtomov, S. G., Kisselev, L., and Philippe, M. 1995. Termination of translation in eukaryotes is governed by two interacting polypeptide chain release factors, eRF1 and eRF3. EMBO J. 14:4065-4072.

Zimmerman, S. B., Cohen, S. H., and Davies, D. R. 1975. X-ray fiber diffractiom and model-building study of polyguanylic acid and polyinosininc acid. J. Mol. Biol. 98:181-192.

Zuker, M., Mathews, D. H., and Turner, D. H. 1999. Algorithms and thermodynamics for RNA secondary structure prediction: a practical guide, p. 11-43. In Barciszewski, J., and Clark, B. F. C. (eds) RNA biochemistry and biotechnology, NATO ASI Series, Kluwer Academic Publishers.

133