+ All Categories
Home > Documents > Viruses as a Model System for Studies of Eukaryotic mRNA Processing

Viruses as a Model System for Studies of Eukaryotic mRNA Processing

Date post: 03-Feb-2022
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
54
Comprehensive Summaries of Uppsala Dissertations from the Faculty of Medicine 1297 Viruses as a Model System for Studies of Eukaryotic mRNA Processing BY ANETTE LINDBERG ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2003
Transcript
Page 1: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Medicine 1297

Viruses as a Model System forStudies of Eukaryotic mRNA

Processing

BY

ANETTE LINDBERG

ACTA UNIVERSITATIS UPSALIENSISUPPSALA 2003

Page 2: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

Dissertation for the degree of Doctor of Philosophy, Faculty of Medicine, in Medical Virology presented atUppsala University in 2003

ABSTRACTLindberg, A. 2003. Viruses as a model system for studies of eukaryotic mRNA processing. Acta UniversitatisUpsaliensis. Comprehensive summaries of Uppsala Dissertations from the Faculty of Medicine 1297. 52 pp.Uppsala ISBN 91-554-5775-4

Viruses depend on their hosts for the production and spread of new virus particles. For efficient virus replication,the viral genes have adapted the strategy of being recognized and processed by the cellular biosyntheticmachineries. Viruses therefore provide an important tool to study the cellular machinery regulating geneexpression. In this thesis, we have used two model DNA viruses; herpes simplex virus (HSV) and adenovirus, tostudy RNA processing at the level of pre-mRNA splicing in mammalian cells.

During a lytic infection, HSV causes an almost complete shut-off of host cell gene expression.Importantly, HSV infection causes inhibition of pre-mRNA splicing which is possibly advantageous to the virus,as only four HSV genes contain introns.

The HSV immediate early protein, ICP27, has been shown to modulate several post-transcriptionalprocesses such as polyadenylation and pre-mRNA splicing. We have studied the role of ICP27 as an inhibitor ofpre-mRNA splicing. We show that ICP27 inhibits pre-mRNA splicing in vitro in the absence of other HSVproteins. We further show that ICP27 inhibits splicing at the level of spliceosome assembly. Importantly, ICP27-induced inhibition of splicing can be reversed, either by the addition of purified SR proteins, which are essentialsplicing factors, or by the addition of an SR protein-specific kinase, SRPK1. We propose that SR proteins areprime candidates as mediators of the inhibitory effect of ICP27 on pre-mRNA splicing.

In order to learn more about how splicing is organized in the cell nucleus in vivo, we investigated howcellular splicing factors are recruited to sites of transcription and splicing in adenovirus-infected cells usingconfocal microscopy. Our results showed that the SR proteins, ASF/SF2 and SC35, are efficiently recruited tosites in the nucleus where adenovirus genes are transcribed and the resulting pre-mRNAs are processed. Ourresults demonstrate that only one of the two RNA recognition motifs (RRMs) present in the ASF/SF2 protein isrequired for its recruitment to active sites of splicing. The arginine/serine rich (RS) domain in ASF/SF2 isredundant for the translocation of the protein to active viral polymerase II genes in adenovirus-infected cells.

Keywords: ICP27; herpes simplex virus; mRNA; splicing; adenovirus; ASF/SF2

Anette Lindberg, Department of Medical Biochemistry and Microbiology,Uppsala University, Box 582, S-751 23 Uppsala, Sweden

ISSN 0282-7476ISBN 91-554-5775-4

Printed in Sweden by Eklundshofs Grafiska AB, Uppsala 2003

Page 3: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

To my family

Page 4: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

Main references

This thesis is based on the following articles, which will be referred to in the text by theirroman numerals.

I. Anette Lindberg and Jan-Peter Kreivi. 2002. Splicing Inhibition at theLevel of Spliceosome Assembly in the Presence of HerpesSimplex Virus Protein ICP27. Virology 294, 189-198

II. Anette Lindberg and Jan-Peter Kreivi. 2003. Herpes Simplex Virus proteinICP27 induced inhibition of splicing is reversed by SR-proteins. Manuscript

III. Anette Lindberg, Margarida Gama-Carvahlo, Maria Carmo-Fonseca andJan-Peter Kreivi. 2003. A single RNA recognition motif in SR proteinsdirects them to nuclear sites of adenovirus transcription. Manuscript

The reprint (Lindberg and Kreivi) was made with permission from the publisher.

Page 5: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

1

Table of contents

Page1. Abbreviations 2

2. Introduction 3

2.1 Co-transcriptional RNA processing 4

3. Introduction to pre-mRNA splicing 8

3.1 The mechanism of pre-mRNA splicing 8

3.2 The chemistry of the splicing reaction 9

3.3 Components of the splicing machinery 10

3.4 The splicing process 13

3.5 Exon definition 16

3.6 Splicing enhancers and silencers 16

3.7 Organization of splicing factors in the cell nucleus 17

4. Herpes simplex virus-1 (HSV-1) 20

4.1 Structure and genome organization 20

4.2 Herpes simplex virus infection 20

4.3 Herpes simplex virus lytic life cycle 21

5. The multifunctional HSV-1 protein ICP27 22

5.1 ICP27 interacting proteins 22

5.2 Functional domains within ICP27 23

6. The role of ICP27 at the post-transcriptional level 24

6.1 ICP27 promotes RNA 3’ end processing at weak viral poly(A) sites 24

6.2 ICP27 stabilizes and binds to 3´ ends of labile mRNA 24

6.3 ICP27 stimulates the nuclear to cytoplasmic export of intronless 24

viral mRNAs6.4 ICP27 inhibits pre-mRNA splicing 25

6.5 ICP27 homologues in other herpes viruses 25

7. Adenovirus 27

7.1 Introduction 27

7.2 Nuclear organization in adenovirus-infected cells 28

8. Present investigation 29

8.1 Project 1, Papers I & II 29

8.2 Project 2, Paper III 35

9. Acknowledgements 38

10. References 39

Page 6: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

2

1. Abbreviations

ASF/SF2 alternative splicing factor/splicing factor 2DNA deoxyribonucleic acidDBP adenovirus DNA binding proteinGFP green fluorescent proteinhnRNPs heterogenous nuclear ribonucleoprotein particlesh p i hours post infectionHSV herpes simplex virusICP27 infected cell protein 27, a herpes simplex virus proteinIGC interchromatin granule clusterskDa kilo DaltonmRNA messenger ribonucleic acidNE nuclear extracts prepared from HeLa cellsNLS nuclear localization signalRNA pol II RNA polymerase IIRRMs RNA Recognition MotifsRS-domain arginine- and serine-rich domain in the SR family of splicing factorssnRNPs small nuclear ribonucleoprotein particlesSRrp SR-related proteinSRs serine-arginine proteins, splicing factorsS100 cytoplasmic extract prepared from HeLa cellsU2AF U2 snRNP Auxiliary Factor

Page 7: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

3

2. Introduction

The genetic information of eukaryotic cells is stored as double stranded DNA

(dsDNA). The DNA is transcribed into a precursor messenger RNA (pre-mRNA), which is

processed and subsequently translated into proteins. Most newly synthesised pre-mRNA

contains stretches of non-coding RNA (introns) that must be excised by splicing before the

RNA can be translated into a protein. Although many of the splicing factors and components

of the splicing machinery are well characterized today, there are still gaps in our knowledge

about the complex process of splicing. By alternative splicing, one gene can express more

than one mRNA thus generating functionally diverse isoforms. Since the vast majority of

human genes contain introns and most pre-mRNAs undergo alternative splicing, disruption of

the normal splicing pattern is a common cause of human disease (reviewed in (Faustino and

Cooper 2003) (Blencowe 2000)).

Viruses have been used as model systems to study eukaryotic gene regulation for

decades. When a virus infects a cell, viral mRNA production is reliant, to a varying degree, on

the cellular transcription machinery. In addition, the viral mRNAs must be translated by the

cellular protein-synthesising machinery. Since viral genes have adapted to be recognized and

processed by the cellular machinery, they represent an important tool to study cellular gene

expression and regulation.

In order to study mRNA processing in mammalian cells at the level of mRNA

splicing, we have used two model DNA viruses; herpes simplex virus-1 (HSV-1) and

adenovirus. During lytic HSV-1 and adenovirus infections, almost all host cell gene

expression is turned off. Results from different studies indicate that in the case of HSV-1, a

viral protein termed ICP27 (Infected Cell Protein 27) blocks cellular gene expression by

inhibiting pre-mRNA splicing (Sandri-Goldin and Mendoza 1992; Hardy and Sandri-Goldin

1994; Bryant, Wadd et al. 2001; Sciabica, Dai et al. 2003). As only four of approximately 80

HSV genes contain introns, the virus is relatively resistant to the inhibition of splicing.

Projects I & II were initiated to investigate the role of ICP27 in the inhibition of pre-mRNA

splicing. Our aims were to establish an in vitro splicing system that would allow for the

characterization of how ICP27 inhibits the splicing process and reveal which cellular splicing

factor(s) are targeted.

In project III, we used adenovirus as a model system to learn more about how splicing

factors are organized in the cell nucleus in vivo. Relatively little is known about how splicing

Page 8: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

4

factors are recruited to sites of active RNA polymerase II transcription and pre-mRNA

splicing. Generally, splicing factors are found throughout the nucleoplasm, but many of them

are concentrated in speckles. Numerous studies have concluded that inactive splicing factors

are recycled and temporarily stored in speckles (Huang and Spector 1996; Gama-Carvalho,

Krauss et al. 1997; Misteli and Spector 1998). When transcription is initiated, the splicing

factors are suggested to be rapidly recruited from the speckles to nuclear regions where they

are needed (Misteli, Caceres et al. 1997). The recruitment of splicing factors seems to be

dependent upon transcription and protein phosphorylation, but other than that, little is known

(Kuroyanagi, Onogi et al. 1998; Misteli and Spector 1999). Previous microscopy studies of

adenovirus infected cells have shown that there is a dramatic reorganization of the nucleus

during the intermediate phase of a lytic infection (Bridge and Pettersson 1995). At this stage,

splicing factors are recruited to sites where viral genes are transcribed and processed (Bridge,

Carmo-Fonseca et al. 1993). In this project, we investigated how the SR proteins, ASF/SF2

and SC35 are recruited to sites of transcription and splicing in adenovirus-infected cells using

confocal microscopy.

2.1 Co-transcriptional RNA processing

This chapter will briefly summarize how proteins are expressed in eukaryotic cells, and

describe the link between transcription, pre-mRNA splicing, and other RNA processes (see

Fig. 1). Although capping, splicing and polyadenylation can occur as independent processes

in vitro, recent studies indicate that most pre-mRNA processing occurs co-transcriptionally

rather than post-transcriptionally in vivo (reviewed in(Daneholt 2001; Bentley 2002)). In this

thesis, the term post-transcriptional refers to processes that occur after the RNA transcript has

left the exit channel of the polymerase.

The DNA is transcribed in the nucleus by multi-subunit enzymes, the RNA

polymerases (reviewed in (Lee and Young 2000)). In eukaryotic cells, there are three different

RNA polymerases; I, II and III, which synthesise different classes of cellular RNA. RNA pol

II transcribes mRNA and some small nuclear RNAs. All genes have a promoter area from

which transcription of the gene starts. Most RNA pol II promoters contain the transcription

start site, a TATA box, and DNA sequence elements to which regulatory transcription factors

bind. The RNA pol II needs general transcription factors in order to assemble at the promoter

and initiate transcription. After initiation, phosphorylation of the C-terminal domain (CTD) of

RNA pol II switches the polymerase activity to elongation of transcription (reviewed in (Lee

and Young 2000)).

Page 9: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

RNA pol II

ds DNA

pre-mRNA

m7GhnRNPs

snRNPs

exon-junctioncomponents

splicingpolyadenylationediting

spliceosome

m7GAAAAAA

EJC

mRNA

nucleus

cytoplasm

mRNA export

nuclear pore complex

m7GAAAAAA

EJC

mRNA

translation degradation

protein

capping

Figure 1. mRNA biogenesis (adapted from (Dreyfuss, Kim et al. 2002)).

When the pre-mRNA is about 25 nucleotides long, capping enzymes bound to the

phosphorylated CTD of RNA pol II add the cap, an inverted methylated guanosine analogue,

to the 5´end. Only RNAs transcribed by pol II are capped at the 5´ end, and this is due to

direct binding of the capping enzymes to pol II (reviewed in (Schroeder, Schwer et al. 2000)

(Shatkin and Manley 2000)). The 5´ cap modification is important for protection against 5´ to

3´ exonucleases, and for recognition by the translation machinery through the cap-binding

complex (CBC), and the translation initiaton factor eIF4E (McKendrick, Thompson et al.

2001). The capping reaction can be further linked to the splicing process, as CBC also plays a

role in splicing of the first intron and promotes the nucleocytoplasmic export of snRNAs

(Izaurralde, Lewis et al. 1994; Izaurralde, Lewis et al. 1995).

Splicing of the new transcript can begin as soon as splice sites have been transcribed.

A direct connection between RNA pol II transcription and splicing have been proposed but it

5

Page 10: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

6

is still somewhat controversial (reviewed in (Hirose and Manley 2000; Neugebauer 2002)). In

vitro splicing is stimulated by phosphorylated RNA pol II or by CTD alone ((Hirose, Tacke et

al. 1999)) and different types of promoters appear to dispose a specific splicing pattern to

transcripts initiated on the promoter ((Cramer, Pesce et al. 1997; Cramer, Caceres et al.

1999)). So far, the yeast U1snRNP component, Prp40 and the SR-related proteins called

SCAFs have been shown to associate with the CTD ((Yuryev, Patturajan et al. 1996; Kim, Du

et al. 1997)).

All eukaryotic mRNAs with the exception of histone mRNAs contain poly(A) tails.

Polyadenylation occurs when the cleavage polyadenylation-specificity factor (CPSF),

together with the downstream cleavage stimulatory factor (CstF) recognizes the poly(A)

consensus sequence, AAUAAA, on the mRNA. Following endonucleolytic cleavage of the

mRNA, poly(A) synthesis takes place. The formation of the poly(A) tail and the poly(A) tail

itself are important for termination of transcription as well as for mRNA stability, splicing,

transport from the nucleus to the cytoplasm and for translation (reviewed in (Zhao, Hyman et

al. 1999)). Cleavage/polyadenylation factors have also been clearly associated with the CTD

during transcriptional elongation (Hirose and Manley 2000). Polyadenylation and splicing

appear to be connected events in the recognition of 3´ terminal exon. For example, the

poly(A) polymerase interacts with 65 kDa U2 snRNP Auxiliary Factor, U2AF65, and this

interaction probably assists in exon recognition of terminal exons and promotes the assembly

of the polyadenylation machinery within the exon (Vagner, Vagner et al. 2000).

RNA editing is a co- or post-transcriptional process that changes the nucleotide

sequence of the substrate (Schaub and Keller 2002). RNA editing occurs in many organisms,

and operates by many different molecular mechanisms; either by insertion or deletion of

nucleotides or by base modification. Adenosine to inosine conversion catalysed by adenosine

deaminases in pre-mRNA coding or non-coding regions (introns, untranslated regions) can

alter the specificity of a codon or can affect splicing by altering splice-site sequences to

produce alternative splicing (Rueter, Dawson et al. 1999).

During transcription, the pre-mRNA is bound not only by splicing factors but also by

heterogeneous ribonucleoprotein particles (hnRNPs) (reviewed in (Kim and Dreyfus 2001)).

These are thought to induce proper processing and folding of the pre-mRNAs. The hnRNPs

remain bound after splicing is completed. Nucleo-cytoplasmic export occurs through the

nuclear pore complex (NPC), which provides docking sites for transport complexes. During

the export to the cytoplasm, some hnRNPs are removed at the NPC while others follow the

mRNA to the cytoplasm and shuttle back to the nucleus.

Page 11: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

7

The composition of the proteins bound to the mRNA is altered near the exon-exon

junction following splicing. A protein complex is deposited, known as the exon-junction

complex, (EJC), which serves as a binding platform for export factors (Zhou, Luo et al. 2000).

This complex consists of at least DEK, SRm160, Y14 and RNPS1 (Kataoka, Yong et al.

2000; Le Hir, Izaurralde et al. 2000; Le Hir, Gatfield et al. 2001). Importantly, this complex

provides a link between RNA splicing and export of the RNA. Three classes of factors appear

to be involved in mRNA export; adapter proteins that bind directly to the mRNA, receptor

proteins that recognize and bind to adapter proteins, and NPC components that mediate the

export across the nuclear membrane (reviewed in (Komeili and O'Shea 2001)). So far, two

pathways for export of mRNAs have been uncovered, involving two different export

receptors, CRM1 and TAP (reviewed in (Cullen 2000)). In one of the pathways, an RNA

binding adaptor protein, REF (RNA and export factor-binding protein), interacts directly with

the export receptor, TAP (Kang and Cullen 1999) and is recruited to mRNA sites near exon-

junctions as part of the EJC (Fig. 1) (Le Hir, Gatfield et al. 2001). At the EJC, REF becomes

bound to the spliced mRNA, which is then efficiently exported to the cytoplasm. REF is most

efficiently recruited to mRNA during splicing (Luo and Reed 1999).

After reaching the cytoplasm, the mRNA is translated into a protein by the ribosome, a

multi-component complex consisting of RNA and proteins. The mRNAs in the cytoplasm are

eventually subjected to degradation (reviewed in (Guhaniyogi and Brewer 2001)). The

poly(A) tail is gradually shortened, which in turn causes decapping at the 5´ end followed by

a 5´ to 3´-exonucleolytic degradation. RNA stability can also be regulated through AU-rich

elements present in the 3´untranslated regions of short-lived mRNAs.

Even if mRNAs are exported to the cytoplasm, they may be defective, due to

incomplete splicing or splicing errors. These defective mRNAs often contain premature

translation termination stop codons, which trigger their rapid degradation by a process termed

nonsense-mediated mRNA decay (also called mRNA surveillance) (Kim and Dreyfus 2001).

Page 12: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

3. Introduction to pre-mRNA splicing

3.1 The mechanism of pre-mRNA splicing

Most newly synthesised mRNAs contain stretches of non-coding RNA (introns) that must be

excised before the RNA can be translated into a protein (see Fig. 2). The introns are removed

from the pre-mRNA by a large macromolecular complex called the spliceosome. The

consensus sequence elements at the splice sites of the transcript direct the assembly of the

spliceosome. In mammals, the 5´splice site signal is AG/GURAGU (exon/intron), (R =

purine). At the 3´ splice site, three different signals are used; the branch point sequence

(BPS), YNYURAC, (Y = pyrimidine, N = variable nucleotides), a polypyrimidine tract (10-

20 nucleotides) and the actual 3´ splice site, YAG/N (intron/exon) (Burge, Tuschl et al. 1999).

There is also a minor class of introns that have dinucleotides AU and AC at their 5´ and 3´

splice site and they are most often spliced by the minor spliceosome (Tarn and Steitz 1996).

AG GU A Y(10-20) AG

3'splice site 5'splice site

pre-mRNA

branch point sequence

polypyrimidine tract

EXON 1 EXON 2

INTRON

Figure 2. Pre-mRNA with the classical splicing signals found in the major class of introns.

The specificity of the splicing reaction is also dependent on exonic or intronic

enhancer and repressor elements (reviewed in (Blencowe 2000; Smith and Valcarcel 2000;

Hastings and Krainer 2001)). Many transcripts have multiple consensuses splice site

sequences whose activities are regulated by transacting factors binding to the enhancer or

repressor element. Alternative choices of splice sites and alternative splicing creates different

transcripts and subsequently, multiple protein products (see example in Fig. 3) (reviewed in

(Lopez 1998)). Alternative splicing is perhaps the most important mechanism for increasing

protein diversity in vertebrates (reviewed in (Graveley 2001)).

8

Page 13: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

EXON 1 EXON 2EXON 1 EXON 3

EXON 1 EXON 2 EXON 3

EXON 1 EXON 3

Figure 3. Example of alternative splicing of a pre-mRNA transcript.

3.2 The chemistry of the splicing reaction

The pre-mRNA is spliced by two trans-esterification reactions (see Fig. 4). In the first,

reaction the 2´ hydroxyl group of the conserved A residue in the BPS of the intron attacks the

3´-5´ phosphodiester bond at the 5´ splice site. This results in cleavage at the 5´ exon-intron

junction and generation of two splicing intermediates; a 5´ exon with a free hydroxyl group at

its 3´ terminus, and the lariat intermediate with the 5´ end of the intron joined with the branch

site adenosine via a 5´-2´ bond. In the second reaction, the free hydroxyl group of the 5´ exon

attacks the 3´-5´ phosphodiester bond at the 3´ splice site and displaces the intron lariat with a

3´ hydroxyl group. This results in the joining of the two exons via a 5´-3´ phosphodiester

bond generating the spliced product (reviewed in (Burge, Tuschl et al. 1999)).

GU A Y(n) AG

3' 5'

2'OH

AG

5'GpA Y(n)

5'

3'

3'OH

PA G

5'Gp

3'OH

Step 1

Step 2

Figure 4. The catalytic steps in pre-mRNA splicing (adapted from (Gesteland 1999)).

9

Page 14: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

10

3.3 Components of the splicing machinery

The spliceosome

In the major spliceosome, five small nuclear RNAs participate in the splicing reaction; U1,

U2, U4, U5 and U6 RNAs. These RNAs vary in length from 100 to 200 nucleotides and

associate with proteins to form small nuclear ribonucleoprotein particles (snRNPs) (reviewed

in (Will and Luhrmann 2001)). The RNA and protein components of snRNPs have essential

functions in recognition, selection and regulation of spliceosome assembly and catalysis. The

minor spliceosome uses the same biochemical splicing mechanism as the major spliceosome

but consists of a different set of snRNPs; U11, U12, U4atac, U6atac and U5 (reviewed in (Wu

and Krainer 1999)). The proportion of introns that are spliced by the minor (U12)

spliceosome is only about one in a thousand. This thesis will focus on splicing by the major

(U2) spliceosome.

The snRNP biogenesis

All U snRNAs, with the exception of U6 and U6atac, are transcribed by RNA pol II, capped

and exported to the cytoplasm where the U snRNP assembly is initiated (reviewed in (Will

and Luhrmann 2001)). The export of the capped U snRNAs requires an export adapter, PHAX

(phosphorylated adapter for RNA export), to mediate the interaction with the export receptor

CRM1/RanGTP (Ohno, Segref et al. 2000). In the cytoplasm, the U snRNA interacts in an

ordered, stepwise manner with seven Sm proteins. The Sm proteins interact as pre-formed

complexes with the Sm site of the U snRNA. After methylation of the cap and 3´ end

trimming, the U snRNP Sm core is imported to the nucleus. The methylated cap and the Sm

core domain form a U snRNP nuclear localization signal (NLS) required for the import of the

U snRNP. Import is mediated by the import factor, Snurportin-1, which binds to the

methylated cap on the snRNAs and to the general import factor, Importin-β.

After import, these factors dissociate and the specific proteins for each U snRNP,

which appear to be imported separately, associate with the U snRNP Sm core. Biogenesis of

the U6 snRNP, and presumably the U6atac snRNP differ from that of the other U snRNPs.

The U6 snRNA is transcribed by RNA pol III and assembly of the U6 snRNP is thought to

take place entirely in the nucleus.

Page 15: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

The SR protein family of splicing factors

The SR proteins are a family of highly conserved serine/arginine-rich RNA binding proteins.

They are essential for constitutive splicing and regulate alternative splice site selection in a

concentration-dependent manner both in vivo and in vitro (Ge and Manley 1990; Krainer,

Conway et al. 1990; Zahler, Lane et al. 1992; Zahler, Neugebauer et al. 1993; Caceres,

Stamm et al. 1994). Some SR protein functions are shared by all members, but some functions

are specific to certain family members (reviewed in (Tacke and Manley 1999; Graveley

2000). SR proteins have been identified in all metazoan species examined, and the family

currently contains 10 members (see Fig. 5). Any member of the SR protein family can

complement splicing-deficient S100 extracts (cytoplasmic extracts from HeLa cells that

contain all components necessary for in vitro splicing except a functional SR protein)

(reviwed in (Fu 1995)).

SRp55

SRp20 RRM RS

SC35 RRM RS

SRp46 RRM RS

SRp54 RRM RS

9G8 RRM RS

SRp30c RRM RSRRM

ASF/SF2 RRM RSRRM

SRp40 RRM RSRRM

RRM RSRRM

SRp75 RRM RSRRM

Figure 5. The human SR proteins (Graveley 2000).

The SR proteins have a modular structure with an N-terminal RNA binding domain

consisting of either one or two RRMs (RNA Recognition Motifs) and a C-terminal domain

containing repetitive arginine-serine dipeptides, a so called RS domain. The RNA binding

domain provides specificity for binding certain mRNAs, while the RS domain interacts with

other proteins (Wu and Maniatis 1993; Kohtz, Jamison et al. 1994; Xiao and Manley 1997)

(Caceres, Misteli et al. 1997; Mayeda, Screaton et al. 1999). The RNA binding domain can be

exchanged between SR proteins and can bind RNA in the absence of the RS domain

11

Page 16: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

12

(Chandler, Mayeda et al. 1997; Mayeda, Screaton et al. 1999). The RS domain can also be

functionally exchanged between SR proteins and it can even function when fused to a

heterologous RNA-binding domain (Chandler, Mayeda et al. 1997; Graveley and Maniatis

1998; Wang, Xiao et al. 1998).

The SR proteins are phosphorylated at serine residues within the RS domain and their

activity is thus regulated by this phosphorylation (Kanopka, Muhlemann et al. 1998; Wang,

Lin et al. 1998), (Xiao and Manley 1997; Xiao and Manley 1998). The two most extensively

studied protein kinases that have been shown to phosphorylate SR proteins are SRPK1 and

Clk/Sty (Gui, Tronchere et al. 1994; Colwill, Feng et al. 1996). Phosphorylated SR proteins

are required for the assembly of the spliceosome and dephosphorylation of the same proteins

are required for catalysis of splicing in vitro (Cao, Jamison et al. 1997). However, both hyper-

and hypophosphorylation of SR proteins can reduce their splicing activity (Kanopka,

Muhlemann et al. 1998; Prasad, Colwill et al. 1999).

SR proteins are ubiquitously expressed, but there are cell-type specific differences in

abundance. This suggests that different cell types might have a distinct pattern of relative SR

protein concentrations, which could influence regulation of alternative splicing (Zahler,

Neugebauer et al. 1993; Screaton, Caceres et al. 1995). The function of SR proteins in

alternative 5´splice site selection can be antagonized by the heterogeneous nuclear

ribonucleoprotein particle A1, hnRNPA1 (Eperon, Makarova et al. 2000). hnRNPA1 can

prevent binding of ASF/SF2 to enhancer elements in a concentration dependent manner

(Mayeda and Krainer 1992).

Only a few of the SR proteins in metazoans have been shown to be essential. SRp55 is

essential for proper development in Drosophila (Ring and Lis 1994) and ASF/SF2

(Alternative Splicing Factor/Splicing Factor 2) is essential for viability of chicken DT40 cells

(Wang, Takagaki et al. 1996). When the ASF/SF2 homologue, rsp-3 in C.elegans was

targeted by RNAi, an embryonic lethal phenotype was observed (Longman, Johnstone et al.

2000). However, individual silencing of six other SR protein-encoding genes had no effect on

embryonic development, although simultaneous silencing of all six was lethal (Longman,

Johnstone et al. 2000). The reason why only some of the SR proteins are essential is

unknown, either, the non-essential SR proteins do not participate in splicing of essential genes

or, alternatively, other SR proteins functionally substitute for the missing protein.

Page 17: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

13

SR-related proteins

There are also a number of SR proteins distinct from the classical SR proteins that are

required for splicing. These are collectively referred to as SR-related proteins, SRrps

(reviewed in (Blencowe, Bowman et al. 1999). The SRrps have an RS domain but not

necessarily an RNA binding domain. Although many of them are essential splicing factors

they cannot complement splicing-deficient S100 extracts. Examples include both subunits of

U2AF (Zhang, Zamore et al. 1992) (Zamore, Patton et al. 1992), snRNP components (like

U1-70K (Spritz, Strunk et al. 1987)), splicing regulators (like hTra2 (Dauwalder, Amaya-

Manzanares et al. 1996)), splicing co activators (like SRm160 (Blencowe B.J. 2000)), RNA

helicases (like hPrp16 (Zhou and Reed 1998)) and protein kinases (like Clk/Sty (Johnson and

Smith 1991)).

One of the SR-related proteins, SRp38 (also known as NSSR-1 (Komatsu, Kominami

et al. 1999), TASR-2 (Yang, Embree et al. 2000) or SRrp40 (Cowper, Caceres et al. 2001)),

has a typical SR protein structure but very unusual properties. Phosphorylated SRp38, cannot

activate splicing and is essentially inactive in splicing assays. However, dephosphorylation

converts SRp38 to a potent general repressor of splicing (Shin and Manley 2002) and cell

cycle-specific dephosphorylation of SRp38 has been proposed to play a role in gene silencing

during mitosis.

3.4 The splicing process

During the stepwise assembly of the spliceosome, the splicing reaction takes place in two

catalytic steps (see Fig. 6) (reviewed in (Hastings and Krainer 2001) (Reed 2000)). The first

complex formed in this pathway is the E-complex.

(i), Formation of the E-complex, also called the commitment complex, involves the

initial recognition of splice sites and commits the pre-mRNA to spliceosome assembly and

splicing. Thus, assembly of the E-complex is likely to be the key step at which alternative

splicing is regulated. It is initiated by U1 snRNP binding to the 5´ splice site by direct base

pairing of U1 snRNA, SF1 (Splicing Factor 1) binding to the adenosine nucleotide within the

BPS, 65 kDa U2AF binding to the polypyrimidine tract and 35 kDa U2AF to the AG

dinucleotide at the 3´ splice site (Gaur, Valcarcel et al. 1995; Berglund, Abovich et al. 1998)

(Eperon, Ireland et al. 1993). SF1 directs the spliceosome to the intron, and when SF1 leaves

the BPS, it assists in the base pairing of U2 snRNP to the same sequence. U2AF promotes the

binding of U2 snRNP to the BPS by direct interaction between the 65 kDa subunit and

SAP155, a component of U2 snRNP (Ruskin, Zamore et al. 1988). SR proteins also promote

Page 18: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

14

assembly of the E-complex by assisting binding of both the U1 70K protein of U1 snRNP and

U2AF35, thus contributing to the bridging of the splice sites at early stages of the reaction

(reviewed in (Graveley, Hertel et al. 2001). Selection of 3´ splice sites can be AG-dependent

or AG-independent. The 35 kDa U2AF is essential for splicing of AG-dependent introns

which require the AG dinucleotide for spliceosome assembly, but dispensible for AG-

independent introns, which have an extended pyrimidine tract (Wu, Romfo et al. 1999).

(ii), U2 snRNP is believed to bind loosely to the BPS in the E-complex through its

subunits, SF3a and SF3b. During the subsequent formation of the A-complex, an ATP-

dependent process leads to stable U2 snRNP binding to the BPS (Das, Zhou et al. 2000). The

stable U2 snRNP binding require U2AF65-associated DEAH helicase, UAP56 (U2AF65

associated protein of 56 kDa) which is believed to facilitate the RNA-RNA rearrangements

(Fleckner, Zhang et al. 1997).

(iii), Subsequent binding of U5-U4/U6 allows the transition from A-complex to B-

complex. U5 binds to sequences in both the 5´ and 3´ exons and base pairs with the 5´ end of

U1 snRNA. These interactions bring the two splice sites together.

(iiii), The last complex, the C-complex, is formed when U6 snRNP dissociates from

U4, forms base pairing with U2 snRNP and replaces U1 snRNP as the factor interacting with

the 5´ end of the intron (reviewed in (Murray and Jarrell 1999)). It is believed that RNA

rearrangements activate the spliceosome for catalysis of splicing. The binding of U2 to U6

snRNA, positions the branch point and the 5´ splice site sequence for the first catalytic step,

generating the lariat intermediate and the free 5´ exon. Before the second catalytic step, the

spliceosome undergoes additional conformational changes, creating new RNA-RNA

interactions. The U2-U6 snRNP interaction is rearranged, and it makes contact with the 5’

splice site in the intron-lariat. Finally, the 5´ splice site in the free exon is joined together with

the 3´ exon of the lariat intermediate (Burge, Tuschl et al. 1999).

The splicing factor, hSlu7 is required to hold exon 1 tightly within the spliceosome in

order to select the correct AG at the 3´ splice site (Chua and Reed 1999). All steps in the

spliceosome assembly require ATP hydrolysis, except formation of the E-complex. It is

believed that the spliceosome consumes ATP to rearrange RNA-RNA and RNA-protein

interactions during assembly and the catalytic steps of splicing (reviewed in (Staley and

Guthrie 1998)).

Page 19: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

3' splice site 5' splice sitepolypyrimidine tract

branch point

U1

U1

70K

70K

SR

SRSR

SR

U2AF35 U2

SF1

U2AF35

SF1

SR

SRSR

U2SF1 U2AF35

70KSR

SRSR

U2

U1

U1

70K

U2AF35

U4U5 U6

U6U4

U5

70KU1

U4U2

U5 U6SR

U2U5

U6

E-complex

A-complex

C-complex

B-complex

mRNA

ATP

ADP

ATP

ADP

ATP

ADP

Figure 6. The spliceosome assembly.

15

Page 20: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

3.5 Exon definition

The average vertebrate gene consists of multiple internal exons (usually 300 nucleotides or

less) separated by introns that are considerably larger (possibly mega bases in length). Berget

and co-workers observed that a 5´ splice site on the downstream side of an exon promotes

splicing of the intron immediately upstream and proposed that initial recognition requires an

interaction between 5´ and 3´ splice sites across the short exon (<350 nucleotides) (see Fig. 7)

(Berget 1995). In this exon definition model, the binding of the U1 and U2 snRNPs and

associated factors, including U2AF65

, U2AF35

and SR proteins, define the exon. In the case of

terminal exons, it has been shown that the cap structure is essential for the recognition of the

first exon (Izaurralde, Lewis et al. 1994). Terminal exons that end with a poly(A) site use

polyadenylation factors for interactions across 3´ terminal exons (Berget 1995). More

recently, Lam et al demonstrated that a single exonic splicing enhancer element, ESE

promotes the recognition of both exon/intron junctions within the same step during exon

definition. They suggested that the exonic splicing enhancer element recruits a multi-

component complex that contains the minimal complement of components of the splicing

machinery required for 5´ and 3´ splice site selection (Lam and Hertel 2002).

ESE AG GU A Y(10-20) AG GUY(10-20) AG ESEU2AF 65 35 U2AF 65 35

SR70K

U170K

U1U2SR

SRSR

Figure 7. The exon definition model.

3.6 Splicing enhancers and silencers

Pre-mRNA splicing enhancers play a critical role in the regulation of alternative splicing and

in correct splice site recognition of constitutively spliced pre-mRNAs elements (reviewed in

(Blencowe 2000; Smith and Valcarcel 2000; Hastings and Krainer 2001)). Exonic splicing

enhancers, ESEs are required for excision of introns that contain splice sites that do not

conform to the consensus sequence (reviewed in (Varani G. 1998)). Although splicing

enhancers are usually located downstream of the affected introns, they are also found within

introns and upstream of regulated 5´ splice sites (Hastings, Wilson et al. 2001). Many of these

RNA elements are recognized by SR proteins, which are thought to function by recruiting

components of the general splicing machinery to the nearby splice sites (Sun, Mayeda et al.

16

Page 21: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

17

1993; Liu, Zhang et al. 1998). The strength of splicing enhancers is determined by the relative

activities of the bound splicing factor, the number of proteins within the enhancer complex

and the distance between the enhancer and the intron (Graveley, Hertel et al. 1998a). The

enhancer elements are usually positioned less than 70 nucleotides from the 3´ splice site.

There is also increasing evidence that exons contain sequences that inhibit splicing, exonic

splicing silencers (Graveley, Hertel et al. 1998a). In at least some cases, silencing is mediated

by hnRNP proteins (heterogeneous nuclear ribonucleoproteins), such as hnRNPA/B and

hnRNP H (Del Gatto-Konczak, Olive et al. 1999). In vivo, hnRNP proteins are thought to

bind co-operatively to mRNA and form stable hnRNP complexes that in some cases can

compete with spliceosome assembly and with enhancer-bound SR proteins (Eperon,

Makarova et al. 2000; Zhu, Mayeda et al. 2001).

SR proteins may also mediate inhibition through the splicing silencers. Kanopka et al

found that SR proteins inhibit splicing of the regulated adenovirus L1 pre-mRNA by binding

an intronic repressor element, located immediately upstream of the IIIa branch site and

thereby preventing U2 snRNP recruitment to the spliceosome. They also demonstrated that

moving the IIIa repressor element to the second exon of the IIIa pre-mRNA converted the

repressor element to a classical splicing enhancer element (Kanopka, Muhlemann et al. 1996).

Thus, depending on where they bind (in the intron or exon), SR proteins may act as splicing

enhancers or silencers.

3.7 Organization of splicing factors in the cell nucleus

The splicing factors are diffusely distributed throughout the nucleus, with enrichment in

compartments commonly referred to as speckles (recently reviewed in (Misteli 2000; Lamond

and Spector 2003)). At the fluorescence-microscopic level, the speckles appear as irregular,

punctuated structures, which vary in size and shape and when examined by electron

microscopy they are seen as clusters of interchromatin granules (Thiry 1995).

Originally, speckles were characterized as sites having high concentrations of splicing

factors but more recently they have also been shown to contain transcription factors, 3´

processing factors and ribosomal proteins (Mortillaro, Blencowe et al. 1996; Schul, van Driel

et al. 1998; Mintz, Patterson et al. 1999; Dostie, Lejbkowicz et al. 2000). They are often

located close to highly transcribed genes but are not themselves sites of active transcription or

splicing (Misteli, Caceres et al. 1997; Cmarko, Verschure et al. 1999). Although several

proteins with possible structural roles in the nucleus, such as lamin A and snRNP-associated

Page 22: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

18

actin, have been found in the speckles, no underlying scaffold has so far been identified

(Nakayasu and Ueda 1984; Jagatheesan, Thanumalayan et al. 1999).

One function of the speckles is thought to be recycling and/or acting as a temporal

storage and assembly site for splicing factors, but since they contain other RNA processing

factors and accumulate polyadenylated RNA (Carter, Taneja et al. 1991; Visa, Puvion-

Dutilleul et al. 1993), there are reasons to believe that the speckles could have more functions.

The speckles consist of numerous spherical sub domains referred to as sub speckles. Each

speckle is composed of 5-50 sub speckles. Spector and co-workers have proposed that the

compartmentalization into sub speckles may represent an efficient way of organizing these

factors for their subsequent transport to transcription/RNA processing sites (Mintz and

Spector 2000).

Splicing factors are also found in other nuclear structures such as Cajal bodies,

interchromatin-associated zones and perichromatin fibrils. The Cajal bodies contain newly

assembled snRNPs and small nucleolar ribonucleoproteins (snoRNPs), the latter being active

in ribosomal RNA processing. Recent evidence suggests that Cajal bodies may be involved in

coordinating the assembly and maturation of nuclear RNPs and possibly other

macromolecular complexes (reviewed in (Ogg and Lamond 2002)). The interchromatin-

associated zones are found adjacent to speckles and contain U1 snRNA (Visa, Puvion-

Dutilleul et al. 1993). The perichromatin fibrils can be observed by electron microscopy and

are believed to represent nascent transcripts (Fakan 1994).

The majority of pre-mRNA splicing probably occurs co-transcriptionally at the site of

transcription (Beyer and Osheim 1991; Xing, Johnson et al. 1993; Bauren and Wieslander

1994). The pre-mRNA is spliced, released from the site of transcription and exported to the

cytoplasm for translation. Transcripts that are not correctly spliced or cleaved at their 3´-ends,

accumulate at the site of transcription indicating that completion of splicing and 3´-end

processing is necessary for the release of transcripts (Custodio, Carmo-Fonseca et al. 1999).

However, highly expressed pre-mRNAs or pre-mRNAs with many introns, like the RNA for

the Balbiani ring of Chironomous tentants can be spliced after its release. Wetterberg et al

analyzed the splicing in the Balbiani Ring 3, (BR3), gene in Chironomous tentants and

showed in vivo that intron excision initiates co-transcripitonally in an overall 5´ to 3´ order.

The proximal 5´ introns of the BR3 transcript are excised co-transcripitonally while the most

distal 3´ introns are excised post-transcriptionally (Wetterberg, Bauren et al. 1996). This

suggests that complete splicing is not crucial for the BR3 mRNA to leave the site of

Page 23: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

19

transcription, and that the timing of the release of the transcript depends more on the extent of

splicing that is necessary.

Changes in the overall trancriptional activity of a cell cause redistribution of splicing

factors. Inhibition of transcription results in immobilization of the splicing factors in enlarged

speckles (Misteli, Caceres et al. 1997). On the other hand, when transcription levels are high

during an adenovirus infection, the majority of the splicing factors are redistributed to viral

replication and transcription centres, where they accumulate in ring-like structures (Bridge

and Pettersson 1995). In contrast, during infection with herpes simplex virus, which only has

four genes containing introns, no obvious recruitment occurs, which suggests that the

accumulation of splicing factors at sites of active transcription is intron-dependent (Phelan,

Carmo-Fonseca et al. 1993). Similar results have been observed when transcribing artificially

intron-less genes (Huang and Spector 1996).

Several protein kinases and protein phosphatases are also localized to speckles

(Colwill, Pawson et al. 1996; Mintz, Patterson et al. 1999; Brede, Solheim et al. 2002).

Dephosphorylation is perhaps the major control mechanism for localization of at least the SR

protein family of splicing factors to speckles. Time-lapse microscopy experiments in living

cells show that release of splicing factors from speckles is inhibited in the presence of a

kinase inhibitor (Misteli, Caceres et al. 1997). Reassociation of splicing factors in speckles

requires the removal of one or more phosphate groups (Misteli and Spector 1996). This

supports the idea that speckles might be involved in regulating the pool of factors that are

accessible to the transcription/RNA processing machinery. Controlling the concentration of

splicing factors in the nucleoplasm might optimise the efficiency of splicing.

Page 24: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

20

4. Herpes Simplex virus-1

4.1 Structure and genome organisation

Herpes simplex virus-1 (HSV-1) is a member of the Herpesviridae family. Herpes viruses

infect members of all groups of vertebrates and the same host can be infected with multiple

distinct types. To date, eight human herpes viruses have been described and each causes a

characteristic disease. HSV-1, for instance, causes oral cold sores and ocular lesions whereas

infections by two other human herpes viruses, Epstein-Barr virus and human herpes virus-8,

are linked to human cancer (reviewed in (Roizman 1996) (S.J Flint 2000)).

HSV-1 is a double stranded DNA virus with an icosahedral capsid surrounded by an

envelope. The virions of the herpes viruses, contain far more proteins than any other virus

described so far. Over half of the proteins encoded by the HSV genome are present in the

virion. The virion proteins are found in the nucleocapsid, in the layer between the envelope

and the nucleocapsid (the tegument layer) and embedded in the envelope. The diameter of the

virion is approximately 200 nm. The linear genome is organised into two unique segments;

one long region of 126 kb (UL) and one short region of 26 kb (US). Both the long and the

short region have inverted repeat sequences and terminal repeats. The HSV-1 can recombine

via the inverted repeats. The viral genome contains three origins of replication, one within the

UL and two within the US (reviewed in (Roizman 1996) (S.J Flint 2000)).

4.2 Herpes simplex virus-1 infection

HSV-1 infection occurs via skin or mucosa. The virus starts a primary replication in the

infected cells immediately after infection. During this phase, the virus travels via an axonal

transport mechanism into the nervous system and infects peripheral neurones. In the neurons,

the virus establishes a latent infection and hides in dorsal root ganglia. The nervous system is

an ideal hiding site for the virus since it is not patrolled by the immune system. However, the

major factor for latency is a restricted gene expression (reviewed in (Whitley 1996) (S.J Flint

2000)).

During a latent infection, normal viral transcription is blocked by an unknown

mechanism. The terminally differentiated neurones neither replicate nor divide, so once a

viral genome is established, it does not need to replicate to persist for the life of the neuron.

When reactivated, HSV-1 travels down the sensory nerves to mucosal surfaces where it

restarts the lytic life cycle, which causes the pathologic manifestation of cold sores. It is not

Page 25: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

21

clear what causes the reactivation but there are many possible alternatives, including both

physical and psychological factors e.g. sunburn, stress and nerve damage. Some individuals

experience reactivation every two to three weeks while others experience no reactivation

(reviewed in (Whitley 1996; S.J Flint 2000)).

4.3 Herpes simplex virus-1 lytic life cycle

Herpes simplex virions attach to the cell when the viral envelope proteins bind to cellular

receptors, such as heparan sulphate, and to the extra cellular matrix. Thereafter, the virion

enters the cell by fusion of the viral envelope with the plasma membrane. The tegument

proteins are released into the cytoplasm and the uncoated virion is transported via

microtubules to the nucleus, where it docks at the nuclear pore and releases the viral DNA

into the nucleus. Some of the tegument proteins like vhs (virion host cell shut-off factor) stay

and act in the cytoplasm, while others are transported into the nucleus (reviewed in (Roizman

1996) (S.J Flint 2000).

During a productive infection, the HSV-1 genes are expressed in a temporal cascade.

After expression of the immediate-early genes (IE), the IE-gene products, in turn activates

transcription of the early genes (E) and auto regulate transcription of IE genes. The E proteins

function mainly in viral DNA replication and the production of substrates for DNA synthesis.

When sufficient levels of the viral replication proteins have accumulated in the cell, the

replication of viral genome starts and the expression of the late genes (L) starts. L proteins are

primarily structural proteins of the virion and proteins needed for virus assembly (reviewed in

(Weir 2001).

Newly replicated viral DNA is packaged into preformed capsids in the nucleus. Some

tegument proteins are added in the nucleus and others in the cytoplasm. It is not known if the

membrane with embedded glycoproteins of mature virions originates from the inner nuclear

membrane or are derived from the Golgi or the endoplasmatic reticulum. How the virion is

subsequently transported to the plasma membrane is also currently unknown. The enveloped

virion can either remain cell-associated or it can be released from the cell by exocytosis

for re-infection (reviewed in (Roizman 1996; S.J Flint 2000).

Page 26: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

5. The multifunctional HSV-1 protein, ICP27

Herpes simplex virus-1 infected cell protein, ICP27 is a 63 kDa phosphoprotein, which is

essential for lytic infection and is the only IE protein that is conserved among all known

herpes viruses. ICP27 is a multifunctional protein: it regulates expression of both viral and

cellular genes at both the transcriptional and post-transcriptional levels (reviewed in

(Clements 1998)).

5.1 ICP27-interacting proteins

ICP27 has been shown to interact with a number of proteins (Table 1).

Table 1.

Protein Activity Method Reference

hnRNP K transcription activator, regulate

translation

yeast-two hybrid screen

in vivo binding studies

(Wadd, Bryant et al.

1999)

casein kinase

2

serine/threonine protein kinase yeast-two hybrid screen

in vivo binding studies

(Wadd, Bryant et al.

1999)

common

Sm proteins

protein components of

snRNP

in vivo binding studies

with anti-Sm serum

(Sandri-Goldin and

Hibbard 1996)

SAP 145 protein component

of U2 snRNP

yeast-two hybrid screen

in vivo binding studies

(Bryant, Wadd et al.

2001)

TAP cellular export receptor yeast-two hybrid screen

in vivo binding studies

(Chen, Sciabica et

al. 2002)

REF cellular RNA export factor yeast-two hybrid screen

in vivo binding studies

(Koffa, Clements et

al. 2001)

RNA pol II

holoenzyme

catalyses synthesis of

RNA

in vivo binding studies (Zhou and Knipe

2002)

P32 Cellular

(mainly mitochondrial )

multifunctional protein

yeast-two hybrid screen

in vivo binding studies

(Bryant, Matthews et

al. 2000)

SRp20 member of the SR protein

family of splicing factors

yeast-two hybrid screen

in vivo binding studies

(Sciabica, Dai et al.

2003)

SRPK1 SR protein kinase 1, highly

specific for serine/arginine

dipeptides

in vivo binding studies (Sciabica, Dai et al.

2003)

22

Page 27: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

5.2 Functional domains within ICP27

ICP27 is an RNA binding protein that shuttles between the nucleus and the cytoplasm (see

Fig. 9). It has a nuclear export signal (Sandri-Goldin 1998) and nuclear and nucleolar

localization signals (Mears, Lam et al. 1995), that may control its sub-cellular localization.

The C-terminus of ICP27 is the most conserved part of the protein and is important for

its repressing and activating functions in viral transcription (Hardwicke, Vaughan et al. 1989;

Rice and Knipe 1990; Rice and Lam 1994). The extreme C-terminus of ICP27 contains an

Sm-homology domain, a structure that mediates protein-protein interactions between Sm

proteins associated with U snRNAs (Sandri-Goldin and Hibbard 1996). The amino-terminal

acidic region of ICP27 performs essential functions in viral DNA replication (Rice, Lam et al.

1993).

The structural domains in ICP27 involved in RNA binding are the RGG-box (Mears

and Rice 1996), and the KH-like domains, KH1, KH2 and KH3 (Soliman and Silverstein

2000). The RGG box is an arginine- and glycine-rich region similar to an RNA binding motif

found in a number of cellular proteins that are involved in nuclear RNA processing (Lengyel,

Guy et al. 2002). The KH-like domains bind single stranded RNA individually or in

combination. ICP27 has been shown to interact with a broad range of HSV-1 mRNAs through

its RGG box (Sokolowski, Scott et al. 2003).

NES 3-28

KH1288-350

KH2368-421

KH3451-512 Zink finger-like

407-512

Sm homology domain

262 327 512

repressor domain

activation domain

NLS109-138

RGG box138-152

nucleolarlocalisationsignal

CN

12 63

acidic domain

Figure 8. Schematic representation of the ICP27 protein (not to scale), (adapted from (Bryant,

Wadd et al. 2001)).

23

Page 28: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

6. The role of ICP27 at the post-transcriptional level 6.1 ICP27 promotes RNA 3´ end processing at weak viral poly(A) sites

There are several well-documented examples in both mammalian and viral systems that

employ alternative or regulated usage of poly(A) sites as a mean of controlling gene

expression (Galli, Guise et al. 1988; Falck-Pedersen and Logan 1989). Poly(A) site usage is

dependent on the inherent poly(A) site efficiency and/or on the concentrations of components

of the 3´ processing complex. Some HSV-1 late mRNAs contain poly(A) addition signals that

function poorly in the absence of ICP27 (McLauchlan, Simpson et al. 1989; McLauchlan,

Phelan et al. 1992). UV cross linking experiments have shown that ICP27 enhances binding

of protein factors, including the 64 kDa subunit of CstF, to poly(A) sites of viral mRNAs

from viral mRNAs with weak poly(A) signals (McGregor, Phelan et al. 1996). The L poly(A)

sites, that responded to ICP27 were inherently less efficient than IE and E sites. This indicates

that their poly(A) site strength and their enhanced expression in the presence of ICP27 may

influence the temporal expression of the L genes. Whether cellular poly(A) site usage is

affected during an HSV infection is still unknown.

6.2 ICP27 stabilizes and binds to 3´ ends of labile mRNA

The intrinsic lifetime of an mRNA can be a critical parameter in the regulation of gene

expression. Within the 3´ untranslated regions of many mRNAs there are AU-rich sequences

that signal rapid turnover of the mRNA. Some proteins stabilize such mRNAs by binding to

these sequences whereas other factors can induce shortening of the poly(A) tail (reviewed in

(S.J Flint 2000)). ICP27 has been shown to bind and stabilize labile 3´ ends of mRNAs

(Brown, Nakamura et al. 1995). ICP27 was, by itself, able to increase the half-life and

stimulate the steady-state accumulation of reporter genes containing AU-rich untranslated

regions. However, the mechanism behind this stimulation is still unknown.

6.3 ICP27 stimulates the nuclear to cytoplasmic export of intronless viral mRNAs

All mRNAs made in the nucleus must be transported to the cytoplasm for translation. Cellular

pre-mRNAs that contain introns and splice sites are ordinarily retained in the nucleus until

they are completely spliced or degraded. Koffa et al recently demonstrated that ICP27

interacts with REF in order to stimulate the export of intronless viral mRNAs (Koffa,

Clements et al. 2001). Shortly after, Sandri-Goldin and co-workers confirmed that ICP27

24

Page 29: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

interacts with REF for export of intronless viral mRNAs (Chen, Sciabica et al. 2002). They

also found an in vivo interaction between ICP27 and TAP but not with CRM1, demonstrating

that ICP27 interacts with REF to direct HSV-1 intronless mRNAs to the TAP export pathway.

6.4 ICP27 inhibits pre-mRNA splicing

In 1987, Martin et al demonstrated that an HSV-1 infection caused a redistribution of snRNPs

from a normal widespread speckled pattern to a highly punctuated organization in the nucleus

(Martin, Barghusen et al. 1987). It has subsequently been shown that ICP27 is necessary for

this event (Sandri-Goldin and Mendoza 1992; Phelan, Carmo-Fonseca et al. 1993; Hardy and

Sandri-Goldin 1994; Sandri-Goldin, Hibbard et al. 1995). Redistribution of splicing factors

like snRNP has also been shown during infection with adenovirus and influenza virus and

may reflect movement of the factors from active sites of splicing to inactive storage sites

(Bridge, Carmo-Fonseca et al. 1993; Fortes, Lamond et al. 1995). ICP27 was shown to

colocalize with the redistributed snRNPs at later time points of infection (Phelan, Carmo-

Fonseca et al. 1993) and Sandri-Goldin et al also demonstrated a physical interaction between

ICP27 and U1 snRNP (Sandri-Goldin and Hibbard 1996). In the same study it was shown that

pre-mRNAs with introns accumulated in the nucleus during wild type HSV infections but not

in cells infected with a mutant virus lacking ICP27. In a separate paper, it was shown how

splicing of a -globin transcript was inhibited in nuclear extracts prepared from HSV-1 wild-

type infected cells but not from cells infected with a mutant virus lacking ICP27 (Sandri-

Goldin 1994). These results suggested that ICP27 has a splicing inhibitory function, but it still

remained to be elucidated whether ICP27 is directly involved, or if another viral protein

whose production or activity is dependent on ICP27, causes the inhibition. More studies on

ICP27-induced splicing inhibition are presented and discussed in the section Present

investigation.

6.5 ICP27 homologues in other herpes viruses

ICP27 homologues have been identified in each of the herpes virus subfamilies (listed in

Table 2). The functional data that are available for ICP27 homologues indicate that although

these proteins show sequence similarity, they exhibit considerable functional diversity. For

example, the human cytomegalovirus (CMV) UL69 and varicella-zoster virus (VZV) ORF4

proteins primarily stimulate gene expression at the level of transcription (Moriuchi, Moriuchi

et al. 1994; Winkler, Rice et al. 1994; Defechereux, Debrus et al. 1997). In contrast,

functional studies of the Epstein-Barr virus (EBV) SM protein, herpes virus samiri (HVS)

25

Page 30: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

ORF57, and human herpesvirus-8 (HHV-8) ORF57, indicate that these proteins function at

post-transcriptional steps of gene expression and that they shuttle between the nucleus and the

cytoplasm (Bello, Davison et al. 1999; Boyle, Ruvolo et al. 1999; Goodwin, Hall et al. 1999;

Gupta, Ruvolo et al. 2000; Kirshner, Lukac et al. 2000). The EBV SM protein activates

expression of intronless reporter gene constructs and inhibits expression of intron-containing

constructs (Ruvolo, Wang et al. 1998). SM also stimulates viral gene expression by enhancing

levels of cytoplasmic and nuclear EBV mRNAs and by promoting EBV mRNA export

(Semmes, Chen et al. 1998). Further, Boyer et al have shown that the SM protein has the

ability to stimulate growth of an HSV-1 ICP27-null virus, suggesting that SM and ICP27 may

regulate gene expression through a common pathway that is evolutionarily conserved between

these two herpes viruses (Boyer, Swaminathan et al. 2002).

Table 2.

Herpes virus Subfamily ICP27

homologue

Reference

herpes virus varicella-zoster virus (VZV)

equine herpes virus type 1 (EHV-

1)

ORF4

UL3

(Inchauspe, Nagpal et al.

1989)

(Zhao, Holden et al. 1992)

herpes virus human cytomegalovirus (CMV) UL69 (Chee, Bankier et al.

1990)

-herpes virus Epstein-Barr virus (EBV)

Herpes virus samiri (HVS)

human herpes virus 8 (HHV8-)/

Kaposi's sarcoma associated virus

(KSHV)

SM

ORF57

ORF57

(Cook, Shanahan et al.

1994)

(Nicholas, Gompels et al.

1988)

(Bello, Davison et al.

1999)

26

Page 31: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

27

7. Adenovirus

7.1 Introduction

Over the years, many important discoveries have been accomplished by the use of adenovirus

as a model system. There are two genera of adenoviruses: Aviadenovirus (Avian) and

Mastadenovirus (mammals). The adenoviruses that infect humans include more than 50

immunologically distinct types, subdivided into six subgenera, A-F (reviewed in (Shenk

1996; S.J Flint 2000)). Adenoviruses most commonly cause respiratory tract infections, like

common cold, croup and bronchitis. Adenovirus has a linear double stranded DNA genome of

about 36 kb, which is encapsulated into a non-enveloped icosahedral capsid particle.

Adenovirus enters the cell by endocytosis, thereafter the virus particle is partially

dismantled, and the viral genome is deposited into the nucleus. The majority of the adenovirus

genes are transcribed by the host cell RNA polymerase II (reviewed in (Cann 1993; Shenk

1996; S.J Flint 2000)). In cell cultures, transcription of the viral genome starts immediately

upon entry into the nucleus and at 3-5 hours post-infection, approximately 15% of the mRNA

in the cytoplasm is virus-encoded (reviewed in (Pombo 1995)). The adenovirus genes are

classified into early and late genes. During the early phase, early region 1A (E1A) is the first

gene to be transcribed. E1A encodes for transacting factors required for the expression of

other early genes. The early phase of gene expression is followed by the onset of viral DNA

replication, which in cultured cells occurs approximately 6-8 h after infection (reviewed in

(Pombo 1995)). Replication of viral DNA is required for the transition to the late phase of

gene expression. Almost all late genes are transcribed from the major late transcription unit,

which by alternative splicing and alternative polyadenylation produces approximately 20

different mRNAs. Most of these mRNAs code for the structural components of the virion

(reviewed in (Cann 1993; Shenk 1996; S.J Flint 2000)). Each cell can produce up to 10 000

new virus particles. By 32 h after infection, the cellular content of protein and DNA is about

two fold higher than in normal cells (reviewed in (Pombo 1995)).

Page 32: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

28

7.2 Nuclear organization in adenovirus infected cells

One of the first virus-induced structural changes observed in adenovirus-infected cells is

small inclusions in the nucleus that rapidly increase in size as viral DNA synthesis continues.

These inclusions are called viral replication centres and contain viral DNA and the viral 72K

DNA binding protein (Puvion-Dutilleul and Pichard 1992). The number of replication centres

is proportional to the multiplicity of the infection. Adenovirus DNA replication starts with a

protein priming mechanism at either end of the viral genome. At each replication fork, only

one of the two daughter strands is replicated producing a new duplex and a displaced single

strand of DNA (ssDNA). In a second round of replication the displaced ssDNA serves as a

template for second strand DNA synthesis (reviewed in (Cann 1993; Shenk 1996; S.J Flint

2000)). As the replication proceeds, the replication centres become larger and

compartmentalized, with the single stranded DNA in the core of the centre surrounded by

double stranded DNA that forms a network named the peripheral replication zone that is

active in both replication and transcription (Puvion-Dutilleul and Pichard 1992; Pombo,

Ferreira et al. 1994).

Intranuclear distribution of the splicing machinery varies during the course of an

adenovirus infection (Bridge and Pettersson 1996). During the early stages of infection,

shortly after the onset of viral replication, snRNPs and SR proteins are recruited to sites of

viral transcription. They accumulate into ring-like structures surrounding the viral replication

centres (Bridge, Carmo-Fonseca et al. 1993; Pombo, Ferreira et al. 1994; Bridge and

Pettersson 1995; Bridge, Riedel et al. 1996). As the adenovirus infection proceeds and more

late viral RNA is transcribed, splicing factors again accumulate in speckle-like structures that

become progressively enlarged (Bridge, Riedel et al. 1996). These enlarged speckles

concentrate splicing factors, poly(A) RNA and spliced viral mRNA (Bridge, Riedel et al.

1996), (Bridge, Xia et al. 1995). The localization of splicing factors in enlarged speckles has

been shown to correlate with late mRNA export to the cytoplasm and production of late

proteins (Aspegren, Rabino et al. 1998). This suggests that the enlarged speckles might have a

posttranscriptional role in the production of RNA.

Page 33: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

8. Present investigation

8.1 Project 1, papers I & II

8.1.1 Introduction

This study was initiated in order to investigate how the HSV-1 ICP27 protein induces an

inhibition of pre-mRNA splicing and to increase our understanding of the basal mechanisms

behind splicing. HSV-1 appears to attack essential splicing factor(s) and thereby potentially

cause a complete shut-off of splicing. This is possible since the majority of HSV-1 genes are

intron-less. Only four of the approximately 80 HSV-1 genes contain introns, and only one is

expressed at late stages of infection. For this reason we believe that identification of the

factor(s) that are targeted by ICP27 could be of general importance since it might reveal

previously uncharacterized parts of the splicing machinery. Previously published in vivo and

in vitro studies have been based on analyses of the splicing phenotype in HSV-1 wild type or

ICP27 mutant virus infected cells. In this study, ICP27 was expressed in eukaryotic cells in

the absence of other HSV-1 proteins.

8.1.2 Results and discussion

Paper I

With the development of an ICP27-expressing adenovirus vector, we were able to produce

high amounts of the protein in eukaryotic cells in the absence of other HSV-1 proteins. The

target cells are co-infected with a viral vector that express the trans-gene ICP27 and a trans-

activator virus that constitutively produce an inducer-dependent trans-activator protein

(Edholm, Molin et al. 2001). Pre-mRNA splicing was studied in a cell-free in vitro system,

using two different 32

P-labeled exogenous pre-mRNA substrates: the cellular -globin pre-

mRNA and the standard adenovirus derived pre-mRNA splicing substrate. The labeled

exogenous pre-mRNA substrates were mixed with nuclear extracts and incubated under

optimal conditions for in vitro splicing. The splicing activity in ICP27 containing nuclear

extracts was compared with the activity in nuclear extracts prepared from cells infected with

AdICP27 and the trans-activator virus, but without addition of the inducer. Combined results

from these studies demonstrated that there is a drastic reduction of splicing products from

both the cellular and the viral pre-mRNA transcript in ICP27-containing extracts. These

results further suggest that the presence of ICP27 causes a general inhibition of splicing.

29

Page 34: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

30

To determine at which stage of the splicing process the ICP27-mediated inhibition of

splicing was manifested, we analyzed the formation of spliceosomal complexes. The results

of this experiment showed a reduced rate of formation of the A-complex in ICP27-containing

extract, indicating that the splicing inhibition is already manifested at or before the formation

of the first energy-dependent spliceosomal A-complex. Further analysis of the conversion

rates of the complexes, from A-complex to B-complex, and B-complex to C-complex showed

that all complex conversion rates were reduced in ICP27-containing extracts. However, the

largest difference between ICP27-containing extracts and control extracts appeared to be the

formation of the mature spliceosomes, the C-complex. The different RNA-protein complexes

formed in ICP27 extracts also migrated notably faster during electrophoresis than their

counterparts formed in control extracts. This altered mobility might reflect a qualitative

and/or a quantitative difference between the complexes formed in the two types of extracts.

We also found that the splicing activity in the ICP27-containing extracts could be completely

recovered by addition of small amounts of competent control extracts. This suggesed that one

or more essential splicing factors are post-transcriptionally modified in ICP27-containing

extracts. When ICP27-containing extracts were added to the control extracts, inhibition of

splicing was not transferred unless the extracts were preincubated at 30°C. Longer incubation

times increased the level of splicing inhibition. These results could indicate that an enzymatic

activity in the ICP27-containing extracts targets an essential component(s) of the splicing

machinery.

To determine whether the splicing inhibition caused by ICP27 was specific, we

compared RNA polymerase II activity in ICP27 extracts and control extracts. A DNA

fragment containing the adenovirus major late promoter was incubated with 32P-labeled CTP

in ICP27-containing extracts or control extracts under conditions optimal for RNA

polymerase II transcription. The results demonstrated that there were no major differences in

RNA polymerase II activity between the ICP27-containing extracts relative to the controls.

This confirms that the splicing inhibition in ICP27 extracts is specific rather than a general

toxic effect of an overexpressed protein.

Our observation that splicing is inhibited at an early stage of spliceosome formation,

and that ICP27 causes a shift in mobility of the spliceosomal complexes is supported by

another study performed by Bryant et al, in which they showed that ICP27 interacts with

SAP145 during an HSV-1 infection (Bryant, Wadd et al. 2001). SAP145 is a component of

the SF3b particle, which is implicated in tethering U2snRNP binding to the branch site and is

therefore needed at an early stage of spliceosome assembly. Moreover, when ICP27 was

Page 35: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

31

present, splicing in vitro was inhibited prior to the first catalytic step and B/C-complexes

formed during splicing increased in mobility and reduced in intensity.

Paper II

With the conclusions drawn from paper I, our next goal was to identify cellular splicing

factors that were targets for ICP27. We employed a biochemical complementation assay, in

which whole cell extracts were fractionated using ammonium sulfate, and magnesium

chloride precipitation steps. We found that several fractions from the ammonium sulfate

precipitation could restore the splicing activity in splicing incompetent ICP27 extracts. The

90% ammonium sulfate precipitate, however, was most efficient in restoring the splicing

activity. This fraction is highly enriched in the essential SR family of splicing factors. Since

previous studies have shown that SR proteins are prime targets for viral control (Kanopka,

Muhlemann et al. 1998; Estmer Nilsson, Petersen-Mahrt et al. 2001; Huang, Nilsson et al.

2002), we decided to continue with the 90% ammonium sulfate precipitate.

In the next step, we further purified the SR proteins present in the 90% ammonium

sulfate fraction using magnesium chloride precipitation. The highly purified SR proteins were

found to be very competent in activating splicing in ICP27-containing extracts. This suggest

that SR proteins on their own are capable of reversing ICP27-mediated splicing inhibition.

We also analyzed another essential splicing factor, U2AF65, for its capacity to restore splicing

activity in ICP27-containing extracts. In this case, we did not observe any activation of the

splicing activity, indicating that ICP27 specifically targets splicing factors of the SR protein

family.

Our next question was whether individual SR proteins are capable of restoring splicing

in ICP27 extracts. Also, it was important to exclude the possibility that the purified SR

protein fraction contained components other than SR proteins that were responsible for the

restoration of the splicing activity in ICP27-containing extracts. For this experiment we used a

bacterially produced ASF/SF2 protein. ASF/SF2 was co-expressed in E.coli together with SR

protein kinase 1, (SRPK1), which selectively phosphorylates serines in serine/arginine-rich

regions. Addition of increasing amounts of phosphorylated ASF/SF2 activated splicing in

ICP27-containing extracts. Collectively, these results suggest that ICP27 blocks splicing by

inhibiting the SR protein family of splicing factors. Thus, addition of the whole family of SR

proteins or a single SR protein is sufficient to reverse ICP27-mediated splicing inhibition.

The effects of SR proteins on splicing depend upon the phosphorylation status of the

protein. SR proteins in their active form are phosphorylated, whereas both

Page 36: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

32

hypophosphorylation and hyperphosphorylation decrease their activity as splicing factors

(Kanopka, Muhlemann et al. 1998), (Xiao and Manley 1997). To investigate whether ICP27

inhibits splicing by changing the phosphorylation status of SR proteins, we analyzed different

ICP27-containing nuclear extracts. We found that the majority of the SR proteins were

dephosphorylated in extracts containing ICP27. Finally, we analysed whether phosphorylation

of SR proteins in the ICP27-containing extracts could restore the splicing activity. For this

experiment we added purified SRPK1 to ICP27 extracts. Although the effect was less

dramatic compared to addition of ASF/SF2, or purified SR proteins, addition of SRPK1 also

resulted in an increase in the splicing activity in ICP27 extracts.

The inactivation of SR proteins by ICP27-mediated dephosphorylation is thus perhaps

reversible, since addition of a kinase to the splicing reaction probably phosphorylates SR

proteins to the extent that they can activate splicing. Another indication that SR proteins are

reversibly dephosphorylated in ICP27-containing extracts comes from the observation that

addition of SR proteins purified from ICP27-expressing cells restores splicing activity to

ICP27 extracts. Western blot and coomassie analysis of the proteins in the ICP27 SR fraction

demonstrated that they were phosphorylated. It is most likely the SR proteins we purify from

the whole cell extract become phosphorylated during the purification process and regain their

activaty as splicing factors.

Phosphorylation of SR proteins also influences their sub cellular distribution:

dephosphorylated SR proteins are localized to nuclear speckles whereas phosphorylated SR

proteins are localized to sites of active transcription (Koizumi, Okamoto et al. 1999). Previous

studies have shown that essential splicing factors are redistributed in HSV-infected cells

(Phelan, Carmo-Fonseca et al. 1993; Sandri-Goldin, Hibbard et al. 1995). Since SR proteins

in nuclear extracts prepared from ICP27-expressing cells were found to be dephosphorylated,

we analysed the distribution of ASF/SF2 in ICP27-expressing cells by immunofluoresence.

HeLa cells were transfected with a plasmid expressing GFP-ICP27 and the distribution of

endogenous ASF/SF2 was monitored by monoclonal antibody 103. In cells expressing ICP27,

ASF/SF2 accumulated in speckles, which are noticeable larger than in control cells.

Moreover, ICP27 is not present in the enlarged speckles with ASF/SF2, and instead, often

appears to be surrounding them. This result showed that expression of ICP27 causes a

redistribution of ASF/SF2 in the nucleus, and that this may be a consequence of an ICP27-

mediated dephosphorylation of ASF/SF2.

Page 37: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

33

Based on these results, we concluded that the SR protein family of splicing factors are

main targets for ICP27 and that the activity of SR proteins as regulatory splicing factors is

reduced in ICP27-expressing cells by dephosphorylation.

A major finding was recently published by Sciabica et al (Sciabica, Dai et al. 2003). In

this study they show that ICP27 interacts with SRPK1 to mediate splicing inhibition by

altering the SR protein phosphorylation. Sciabica et al also reported that during an HSV-1

infection, phosphorylation of several SR proteins was reduced and that this correlated with a

sub-nuclear redistribution of the SR proteins. Further, ICP27 was shown to interact

specifically with SRp20 both in vivo and in vitro. ICP27 interaction with SRPK1, alters

SRPK1’s activity and relocalizes it to the nucleus. In conclusion, they propose that ICP27

recruits SRPK1 to the nucleus and interacts with SR proteins at the site of spliceosome

assembly. Their finding thus supports our observation that SR proteins are dephosphorylated

in ICP27-expressing cells.

8.1.3 Perspectives

To date, two other DNA viruses, adenovirus and vaccinia virus have been reported to regulate

the cellular splicing machinery by inducing a dephosphorylation of the SR family of splicing

factors (Kanopka, Muhlemann et al. 1996; Estmer Nilsson, Petersen-Mahrt et al. 2001;

Huang, Nilsson et al. 2002). SR proteins prepared from late adenovirus-infected cells are

inactivated by a virus-induced partial de-phosphorylation. This modification enhances the

production of the alternatively spliced late mRNA IIIa, which otherwise is repressed by

phosphorylated SR proteins binding to an intronic repressor element, the 3RE (Kanopka,

Muhlemann et al. 1998). Presumably, splicing of most late stage-specific adenovirus mRNAs

is stimulated by this virus-induced dephosphorylation of SR proteins. Most adenovirus late

genes are expressed from the major late transcription unit, which by alternative splicing and

alternative polyadenylation, produce approximately 20 different mRNAs. Adenovirus is

therefore dependent on the cellular splicing machinery for processing its own pre-mRNAs.

During an adenovirus infection, the splicing machinery is not functionally blocked, and

instead, a partial dephosphorylation of SR proteins appears to promote a temporal shift into

the late stage of infection and production of late viral mRNAs.

Vaccinia virus, on the other hand, replicates in the cytoplasm and encodes for genes

lacking introns. This suggests that this virus could benefit from a total block of the splicing

machinery for maximal expression of its own genes. When SR proteins from vaccinia virus-

infected cells were analysed, they were indeed found to be hypophosphorylated and

Page 38: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

34

functionally inactivated as splicing regulatory proteins. Re-phosphorylation of these SR

proteins only partially restored the splicing enhancer or splicing repressor activity (Huang,

Nilsson et al. 2002). It is possible that vaccinia viruses use an alternative, or additional

mechanism, to alter the phosphorylation status of the SR proteins compared to adenovirus.

Adenovirus has been shown to induce dephosphorylation of SR proteins by the virus-encoded

E4ORF4, which binds to the cellular protein phosphatase 2A, (PP2A), and directs it to SR

proteins for dephosphorylation. So far, it has not been elucidated how vaccinia virus

inactivates SR proteins as splicing regulatory proteins. However, vaccinia virus encode for its

own dual protein phosphatase, VH1, which is required for virus growth and gene expression

(Liu, Lemon et al. 1995). VH1 may be a possible candidate for causing SR protein

dephosphorylation.

In this study, we present a third DNA virus, HSV-1, which uses the SR protein family

of splicing factors to regulate the splicing machinery. HSV-1 only has four intron-containing

genes, so it would benefit from blocking most of the splicing activity in the cell. In the study

presented in paper II, the SR family of splicing factors appeared to be functionally inactivated

by a reversible dephosphorylation. One question that remains is how the SR proteins are

dephosphorylated and whether ICP27 targets more components of the splicing machinery to

induce splicing inhibition. The accumulation of hypophosphorylated SR proteins in HSV-1-

infected cells could be induced by at least three different mechanisms; i, by inhibiting a

protein kinase(s) responsible for SR protein phosphorylation; ii, by inducing

dephosphorylation by a phosphatase and iii, by inducing or encoding for a protein that blocks

SR protein phosphorylation.

Sciabica et al have proposed that ICP27 prevents SRPK1 phosphorylation of SR

proteins by interacting with and changing its activity and cellular localization (Sciabica, Dai

et al. 2003). ICP27 has been shown to interact with numerous proteins (listed in Table 1, page

25). The cellular p32 is another protein with which ICP27 interacts, and which could play a

role in the ICP27 induced inhibition of splicing (Bryant, Matthews et al. 2000). p32 has been

shown to interact with and inhibit ASF/SF2 as a splicing factor by blocking phosphorylation

of its RS domain (Petersen-Mahrt, Estmer et al. 1999). It is possible that the interaction of

ICP27 with p32 is used to direct p32 to SR proteins to inhibit their phosphorylation.

Page 39: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

35

8.2 Project 2, paper III

8.2.1 Introduction

This study was undertaken in order to investigate how cellular splicing factors are recruited to

sites of splicing in the cell nucleus. We compared the recruitment of SR-proteins ASF/SF2

and SC35 to active sites of splicing with splicing factors U2AF65 and snRNPs, during an

adenovirus infection. We also analysed the contribution of the different domains of ASF/SF2

for the recruitment to active sites of viral transcription and splicing.

8.2.2 Results and discussion

In order to analyse the spatial distribution of ASF/SF2 during an adenovirus infection, HeLa

cells were infected with adenovirus and fixed and permeabilized after different times of

infection. The sub-nuclear distribution of ASF/SF2 was first determined by

immunofluorescence confocal microscopy using a monoclonal antibody directed against the

endogenous protein. The viral replication centres were visualized using a separate antibody

directed against the viral 72K DNA-binding protein. The results showed that the distribution

of ASF/SF2 during an adenovirus infection is dramatically altered. At an intermediate stage of

infection, at 12-16 hours post-infection (hpi), almost all detectable ASF/SF2 had accumulated

at sites surrounding the viral replication centres. It is in the immediate vicinity of these

structures that transcription and splicing is predominantly detected. This indicates that

ASF/SF2 is recruited to the nuclear sites where adenovirus genes are transcribed and

processed. At late stages of infection, 20 hpi, ASF/SF2 is found in so-called enlarged speckles

distant from the replication centres.

The staining pattern from endogenous ASF/SF2 detected with the monoclonal

antibody was confirmed by transient expression of a myc- or Green Fluorescent Protein-

tagged (GFP) version of the protein. The transfected cells were subsequently superinfected

with adenovirus. Both the myc- and the GFP-tagged ASF/SF2 proteins displayed similar, if

not identical, distribution pattern to endogenous ASF/SF2 during an adenovirus infection.

U2AF65 is similar to ASF/SF2 in that it contains an RNA-binding domain and an RS domain.

Previous studies have demonstrated that an adenovirus infection induces a redistribution of

U2AF65 into ring-like structures that colocalize with the sites of viral transcription (Gama-

Carvalho, Krauss et al. 1997). GFP-ASF/SF2 distribution was compared with U2AF65 in

order to examine if ASF/SF2 is recruited to the same ring-like structures as U2AF65 during an

Page 40: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

36

adenovirus infection. Both proteins were present in ring-like structures at the intermediate

stage of infection and in enlarged speckles at the late phase of infection. When superimposing

the GFP-ASF/SF2 and U2AF65 signals, it is apparent that they colocalize during the

adenovirus infection. Although U2AF65 appears to be more evenly distributed throughout the

nucleus of an infected cell compared to ASF/SF2, they show a clear colocalization in the ring-

like structures. SC35 and ASF/SF2 are similar in size and both play a role in constitutive and

alternative splicing. However, SC35 has only one RRM and it does not shuttle between the

cytoplasm and the nucleus. When ASF/SF2 distribution is compared to that of SC35 during

an adenovirus infection, they display the same pattern of redistribution. Further, a confocal

overlay of the signals shows an almost identical distribution pattern. This demonstrates that

SC35, like ASF/SF2, accumulates in the ring-like structures in which adenovirus genes are

actively transcribed and processed.

Experiments with mutant proteins showed that the two RNA-binding domains present

in ASF/SF2 are sufficient for efficient recruitment of the protein to the ring-like structures

surrounding viral replication centres. The ASF mutant, in which the two RNA binding

domains were deleted (GFP-RS), was evenly distributed throughout the nucleus and was not

notably redistributed during an adenovirus infection. The protein did not accumulate in ring-

like structures or enlarged speckles during the infection.

These results indicate that ASF/SF2 is recruited to sites of viral transcription and

splicing by a different mechanism compared to that previously shown for the essential

splicing factor U2AF65 (Gama-Carvalho, Krauss et al. 1997). Localization of U2AF65 to these

sites is totally dependent upon an intact RS domain. The RS domains of both ASF/SF2 and

U2AF65 have been shown to mediate protein-protein interactions with other splicing factors.

In the case of U2AF65, the RS domain has also been shown to be necessary for efficient

binding of U2 snRNA to the branch site in pre-mRNA, perhaps by binding to one of the two

RNA components (Valcarcel, Gaur et al. 1996). Previous studies have shown that a virus-

induced dephosporylation of the majority of the SR proteins, including ASF/SF2, late during

an adenovirus infection reduces their activity as splicing enhancer or splicing repressor

proteins (Kanopka, Muhlemann et al. 1998). Despite this, we observed that ASF/SF2

accumulates in ring-like structures where late adenoviral genes are transcribed and their pre-

mRNAs processed. This may suggest that the hypophosphorylated SR proteins are also

needed for late viral mRNA processing or that the SR proteins are required for some process

other than splicing. ASF/SF2 has been shown to interact directly both with the general mRNA

export factor TAP (Huang, Gattoni et al. 2003) and with the transcriptional coactivator, p52

Page 41: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

37

(Ge, Si et al. 1998). Thus, it is possible that ASF/SF2 may be required for efficient

transcription of adenovirus genes and/or in export of viral mRNAs.

8.2.3 Perspectives

Cellular compartments are important for functions of the cell and for the regulation of gene

expression (reviewed in (Carmo-Fonseca 2002)). Molecules can move between cellular

compartments in different ways. Nuclear compartments differ from most cytoplasmic

compartments in that they are not surrounded by membranes. Recent data from imaging of the

motion of macromolecules inside the living cells have shown that splicing factors are in

continuous flux between nuclear compartments which are not separated by a physical barriers

(Eils, Gerlich et al. 2000). As proteins diffuse through the nuclear space, they transiently

accumulate in a steady-state compartment by interactions with high-affinity binding sites.

According to this view, sequences and domains that are necessary for proper localization

within the cell, for example the nuclear retention signal within hnRNP C, would more likely

act as retention signals rather than as targeting or transport signals. Consequently, the

recruitment of splicing factors to sites of splicing would not be a directed process where the

factors are transported from storage in speckles to the active splice site. Instead, the steady-

state level of the factors would increase dynamically at the active splice site because of the

prolonged binding and interactions by the factors. The RS domain of ASF/SF2 has been

reported to function in the dissociation of the protein from speckles, and phosphorylation of

serine residues is a prerequisite for this event (Misteli, Caceres et al. 1998). It is possible that

the dephosphorylated RS domain has a higher binding affinity to the other components of the

speckles, which cause retention of the protein in speckles. The observation from this study

that the RS domain of ASF/SF2 is not needed for accumulation of the protein at sites of viral

transcription and splicing implies that interactions with the RS domain are not sufficient to

anchor the protein at the transcription/splice site. Instead it is the affinity of the RRMs of

ASF/SF2 that anchor the protein to the site of active splicing. Whether ASF/SF2 is anchored

to the activated viral gene by interaction with an RNA component and/or a protein and

whether this is unique for viral genes or is universal for metazoan genes remains to be

established.

Page 42: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

9. Acknowledgements

First, I should like to thank my enthusiastic supervisor, Peter Kreivi for all the effort put in

supervision and for incredible patience and support over the years. Next, I would like to thank

Göran Akusjärvi and Catharina Svensson for guidance and making our Friday seminars into

scientific adventures. I am also grateful for all their support during the writing of this thesis.

Special thanks to:

My family, who has contributed to this thesis in many ways, you have always encouraged me

and made me believe that anything is possible.

I would like to thank:

My father, for sharing his interest in science and biology.

My mother, for giving me some of her determination and patience, which has been very

useful during these years.

My sister and her lovely family; Niklas, Emma-Lisa, Linnea och Isak, for always being there.

A visit at your place is always a real energizer!

Sven for all love and for being my best friend, no one is more alive than you! With you by my

side even difficult things are easy.

The Kilander family; Axel, Lennart, Elisabet, Sara, Emil, Helena, Lukas och Tova for all their

encouragement.

My godmother, Eva-Lotta for all her support.

All my friends and climbing partners, you are all included, to write about all of you would fill

up another, much longer thesis. You have given me the best of times!!

I look forward to more adventures with you!

I should also like to thank:

Everyone in the lab for all the good fun, help and discussions,

Christina (I will miss our chats and good discussions in the office), Tanel (collaborations in

the future?!), Cecilia (thank you for all fun and for correcting my English), Saideh, Anette,

Josef (I liked all the “fika” you brought), Gunnar, Ellenor, Ning, Anders, Vita, Raj, Margret

(meet again in Norway?), Daniel and Shaomin……I will miss you!

Erica, for answering all my questions at any time

Maud, Kerstin, Ylva and Tony for all assistance with administration, computer service and

technical matters

Lena Möller, always there to cheer us up

Göran Magnusson and Stefan Schwartz

All colleagues at IMBIM

Former members of the group; Elisabet Boija, Camilla Estmer-Nilsson, Anders Sundkvist,

Magnus Molin, Brian Collier and Lisa Wiklund (my personal running-trainer, I will miss all

the training, talking and laughing).

38

Page 43: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

39

10. References

Aspegren, A., C. Rabino, et al. (1998). "Organization of splicing factors in adenovirus-infected cells reflectschanges in gene expression during the early to late phase transition." Exp Cell Res 245(1): 203-13.

Bauren, G. and L. Wieslander (1994). "Splicing of Balbiani ring 1 gene pre-mRNA occurs simultaneously withtranscription." Cell 76(1): 183-92.

Bello, L. J., A. J. Davison, et al. (1999). "The human herpesvirus-8 ORF 57 gene and its properties." J Gen Virol80(Pt 12): 3207-15.

Bentley, D. (2002). "The mRNA assembly line: transcription and processing machines in the same factory."Current Opinion in Cell Biology 14: 336-342.

Berget, S. M. (1995). "Exon recognition in vertebrate splicing." J Biol Chem 270(6): 2411-4.

Berglund, J. A., N. Abovich, et al. (1998). "A cooperative interaction between U2AF65 and mBBP/SF1facilitates branchpoint region recognition." Genes Dev 12(6): 858-67.

Beyer, A. L. and Y. N. Osheim (1991). "Visualization of RNA transcription and processing." Semin Cell Biol2(2): 131-40.

Blencowe B.J., B. G., Eldridge A.G., Issner R. Nickerson J., Rosonina E., P. A. Sharp (2000). "TheSRm160/300 splicing coactivator subunits." RNA 6: 111-120.

Blencowe, B. J. (2000). "Exonic splicing enhancers: mechanism of action, diversity and role in human geneticdiseases." Trends Biochem Sci 25(3): 106-10.

Blencowe, B. J., J. A. Bowman, et al. (1999). "SR-related proteins and the processing of messenger RNAprecursors." Biochem Cell Biol 77(4): 277-91.

Boyer, J. L., S. Swaminathan, et al. (2002). "The Epstein-Barr virus SM protein is functionally similar to ICP27from herpes simplex virus in viral infections." J Virol 76(18): 9420-33.

Boyle, S. M., V. Ruvolo, et al. (1999). "Association with the cellular export receptor CRM 1 mediates functionand intracellular localization of Epstein-Barr virus SM protein, a regulator of gene expression." J Virol73(8): 6872-81.

Brede, G., J. Solheim, et al. (2002). "PSKH1, a novel splice factor compartment-associated serine kinase."Nucleic Acids Res 30(23): 5301-9.

Bridge, E., M. Carmo-Fonseca, et al. (1993). "Nuclear organization of splicing small nuclear ribonucleoproteinsin adenovirus-infected cells." J Virol 67(10): 5792-802.

Bridge, E. and U. Pettersson (1995). "Nuclear organization of replication and gene expression in adenovirus-infected cells." Curr Top Microbiol Immunol 199(( Pt 1)): 99-117.

Bridge, E. and U. Pettersson (1996). "Nuclear organization of adenovirus RNA biogenesis." Exp Cell Res229(2): 233-9.

Bridge, E., K. U. Riedel, et al. (1996). "Spliced exons of adenovirus late RNAs colocalize with snRNP in aspecific nuclear domain." J Cell Biol 135(2): 303-14.

Bridge, E., D. X. Xia, et al. (1995). "Dynamic organization of splicing factors in adenovirus-infected cells." JVirol 69(1): 281-90.

Brown, C. R., M. S. Nakamura, et al. (1995). "Herpes simplex virus trans-regulatory protein ICP27 stabilizesand binds to 3' ends of labile mRNA." J Virol 69(11): 7187-95.

Page 44: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

40

Bryant, H. E., D. A. Matthews, et al. (2000). "Interaction between herpes simplex virus type 1 IE63 protein andcellular protein p32." J Virol 74(23): 11322-8.

Bryant, H. E., S. E. Wadd, et al. (2001). "Herpes simplex virus IE63 (ICP27) protein interacts with spliceosome-associated protein 145 and inhibits splicing prior to the first catalytic step." J Virol 75(9): 4376-85.

Burge, C. B., T. Tuschl, et al. (1999). Splicing of precursors to mRNAs by the spliceosome. CSH, New York,Cold Spring Harbour Laboratory Press.

Caceres, J. F., T. Misteli, et al. (1997). "Role of the modular domains of SR proteins in subnuclear localizationand alternative splicing specificity." J Cell Biol 138(2): 225-38.

Caceres, J. F., S. Stamm, et al. (1994). "Regulation of alternative splicing in vivo by overexpression ofantagonistic splicing factors." Science 265(5179): 1706-9.

Cann (1993). Principles of molecular virology.

Cao, W., S. F. Jamison, et al. (1997). "Both phosphorylation and dephosphorylation of ASF/SF2 are required forpre-mRNA splicing in vitro." Rna 3(12): 1456-67.

Carmo-Fonseca, M. (2002). "The contribution of nuclear compartmentalization to gene regulation." Cell 108(4):513-21.

Carter, K. C., K. L. Taneja, et al. (1991). "Discrete nuclear domains of poly(A) RNA and their relationship to thefunctional organization of the nucleus." J Cell Biol 115(5): 1191-202.

Chandler, S. D., A. Mayeda, et al. (1997). "RNA splicing specificity determined by the coordinated action ofRNA recognition motifs in SR proteins." Proc Natl Acad Sci U S A 94(8): 3596-601.

Chee, M. S., A. T. Bankier, et al. (1990). "Analysis of the protein-coding content of the sequence of humancytomegalovirus strain AD169." Curr Top Microbiol Immunol 154: 125-69.

Chen, I. H., K. S. Sciabica, et al. (2002). "ICP27 interacts with the RNA export factor Aly/REF to direct herpessimplex virus type 1 intronless mRNAs to the TAP export pathway." J Virol 76(24): 12877-89.

Chua, K. and R. Reed (1999). "The RNA splicing factor hSlu7 is required for correct 3' splice-site choice."Nature 402(6758): 207-10.

Clements, A. P. J. B. (1998). "Posttranscriptional regulation in herpes simplex virus." Seminars in Virology 8:309-318.

Cmarko, D., P. J. Verschure, et al. (1999). "Ultrastructural analysis of transcription and splicing in the cellnucleus after bromo-UTP microinjection." Mol Biol Cell 10(1): 211-23.

Colwill, K., L. L. Feng, et al. (1996). "SRPK1 and Clk/Sty protein kinases show distinct substrate specificitiesfor serine/arginine-rich splicing factors." J Biol Chem 271(40): 24569-75.

Colwill, K., T. Pawson, et al. (1996). "The Clk/Sty protein kinase phosphorylates SR splicing factors andregulates their intranuclear distribution." Embo J 15(2): 265-75. abs.html.

Cook, I. D., F. Shanahan, et al. (1994). "Epstein-Barr virus SM protein." Virology 205(1): 217-27.

Cowper, A. E., J. F. Caceres, et al. (2001). "Serine-arginine (SR) protein-like factors that antagonize authenticSR proteins and regulate alternative splicing." J Biol Chem 276(52): 48908-14.

Cramer, P., J. F. Caceres, et al. (1999). "Coupling of transcription with alternative splicing: RNA pol IIpromoters modulate SF2/ASF and 9G8 effects on an exonic splicing enhancer." Mol Cell 4(2): 251-8.

Page 45: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

41

Cramer, P., C. G. Pesce, et al. (1997). "Functional association between promoter structure and transcriptalternative splicing." Proc Natl Acad Sci U S A 94(21): 11456-60.

Cullen, B. R. (2000). "Nuclear RNA export pathways." Mol Cell Biol 20(12): 4181-7.

Custodio, N., M. Carmo-Fonseca, et al. (1999). "Inefficient processing impairs release of RNA from the site oftranscription." Embo J 18(10): 2855-66.

Daneholt, B. (2001). "Assembly and transport of a premessenger RNP particle." Proc Natl Acad Sci U S A98(13): 7012-7.

Das, R., Z. Zhou, et al. (2000). "Functional association of U2 snRNP with the ATP-independent spliceosomalcomplex E." Mol Cell 5(5): 779-87.

Dauwalder, B., F. Amaya-Manzanares, et al. (1996). "A human homologue of the Drosophila sex determinationfactor transformer-2 has conserved splicing regulatory functions." Proc Natl Acad Sci U S A 93(17):9004-9.

Defechereux, P., S. Debrus, et al. (1997). "Varicella-zoster virus open reading frame 4 encodes an immediate-early protein with posttranscriptional regulatory properties." J Virol 71(9): 7073-9.

Del Gatto-Konczak, F., M. Olive, et al. (1999). "hnRNP A1 recruited to an exon in vivo can function as an exonsplicing silencer." Mol Cell Biol 19(1): 251-60.

Dostie, J., F. Lejbkowicz, et al. (2000). "Nuclear eukaryotic initiation factor 4E (eIF4E) colocalizes with splicingfactors in speckles." J Cell Biol 148(2): 239-47.

Dreyfuss, G., V. N. Kim, et al. (2002). "Messenger-RNA-binding proteins and the messages they carry." NatRev Mol Cell Biol 3(3): 195-205.

Edholm, D., M. Molin, et al. (2001). "Adenovirus vector designed for expression of toxic proteins." J Virol75(20): 9579-84.

Eils, R., D. Gerlich, et al. (2000). "Quantitative imaging of pre-mRNA splicing factors in living cells." Mol BiolCell 11(2): 413-8.

Eperon, I. C., D. C. Ireland, et al. (1993). "Pathways for selection of 5' splice sites by U1 snRNPs and SF2/ASF."Embo J 12(9): 3607-17.

Eperon, I. C., O. V. Makarova, et al. (2000). "Selection of alternative 5' splice sites: role of U1 snRNP andmodels for the antagonistic effects of SF2/ASF and hnRNP A1." Mol Cell Biol 20(22): 8303-18.

Estmer Nilsson, C., S. Petersen-Mahrt, et al. (2001). "The adenovirus E4-ORF4 splicing enhancer proteininteracts with a subset of phosphorylated SR proteins." Embo J 20(4): 864-71.

Fakan, S. (1994). "Perichromatin fibrils are in situ forms of nascent transcripts." Trends Cell Biol 4: 86-90.

Falck-Pedersen, E. and J. Logan (1989). "Regulation of poly(A) site selection in adenovirus." J Virol 63(2): 532-41.

Faustino, N. A. and T. A. Cooper (2003). "Pre-mRNA splicing and human disease." Genes Dev 17(4): 419-37.

Fleckner, J., M. Zhang, et al. (1997). "U2AF65 recruits a novel human DEAD box protein required for the U2snRNP-branchpoint interaction." Genes Dev 11(14): 1864-72.

Fortes, P., A. I. Lamond, et al. (1995). "Influenza virus NS1 protein alters the subnuclear localization of cellularsplicing components." J Gen Virol 76(Pt 4): 1001-7.

Fu, X. D. (1995). "The superfamily of arginine/serine-rich splicing factors." Rna 1(7): 663-80.

Page 46: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

42

Galli, G., J. Guise, et al. (1988). "Poly(A) site choice rather than splice site choice governs the regulatedproduction of IgM heavy-chain RNAs." Proc Natl Acad Sci U S A 85(8): 2439-43.

Gama-Carvalho, M., R. D. Krauss, et al. (1997). "Targeting of U2AF65 to sites of active splicing in the nucleus."J Cell Biol 137(5): 975-87.

Gaur, R. K., J. Valcarcel, et al. (1995). "Sequential recognition of the pre-mRNA branch point by U2AF65 and anovel spliceosome-associated 28-kDa protein." Rna 1(4): 407-17.

Ge, H. and J. L. Manley (1990). "A protein factor, ASF, controls cell-specific alternative splicing of SV40 earlypre-mRNA in vitro." Cell 62(1): 25-34.

Ge, H., Y. Si, et al. (1998). "A novel transcriptional coactivator, p52, functionally interacts with the essentialsplicing factor ASF/SF2." Mol Cell 2(6): 751-9.

Gesteland, C. a. A. (1999). The RNA world.

Goodwin, D. J., K. T. Hall, et al. (1999). "The open reading frame 57 gene product of herpesvirus saimirishuttles between the nucleus and cytoplasm and is involved in viral RNA nuclear export." J Virol73(12): 10519-24.

Graveley, B. R. (2000). "Sorting out the complexity of SR protein functions." Rna 6(9): 1197-211.

Graveley, B. R. (2001). "Alternative splicing: increasing diversity in the proteomic world." Trends Genet 17(2):100-7.

Graveley, B. R., K. J. Hertel, et al. (1998a). "A systematic analysis of the factors that determine the strength ofpre- mRNA splicing enhancers." Embo J 17(22): 6747-56.

Graveley, B. R., K. J. Hertel, et al. (2001). "The role of U2AF35 and U2AF65 in enhancer-dependent splicing."Rna 7(6): 806-18.

Graveley, B. R. and T. Maniatis (1998). "Arginine/serine-rich domains of SR proteins can function as activatorsof pre-mRNA splicing." Mol Cell 1(5): 765-71.

Guhaniyogi, J. and G. Brewer (2001). "Regulation of mRNA stability in mammalian cells." Gene 265(1-2): 11-23.

Gui, J. F., H. Tronchere, et al. (1994). "Purification and characterization of a kinase specific for the serine- andarginine-rich pre-mRNA splicing factors." Proc Natl Acad Sci U S A 91(23): 10824-8.

Gupta, A. K., V. Ruvolo, et al. (2000). "The human herpesvirus 8 homolog of Epstein-Barr virus SM protein(KS- SM) is a posttranscriptional activator of gene expression." J Virol 74(2): 1038-44.

Hardwicke, M. A., P. J. Vaughan, et al. (1989). "The regions important for the activator and repressor functionsof herpes simplex virus type 1 alpha protein ICP27 map to the C-terminal half of the molecule." J Virol63(11): 4590-602.

Hardy, W. R. and R. M. Sandri-Goldin (1994). "Herpes simplex virus inhibits host cell splicing, and regulatoryprotein ICP27 is required for this effect." J Virol 68(12): 7790-9.

Hastings, M. L. and A. R. Krainer (2001). "Pre-mRNA splicing in the new millennium." Curr Opin Cell Biol13(3): 302-9.

Hastings, M. L., C. M. Wilson, et al. (2001). "A purine-rich intronic element enhances alternative splicing ofthyroid hormone receptor mRNA." Rna 7(6): 859-74.

Hirose, Y. and J. L. Manley (2000). "RNA polymerase II and the integration of nuclear events." Genes Dev14(12): 1415-29.

Page 47: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

43

Hirose, Y., R. Tacke, et al. (1999). "Phosphorylated RNA polymerase II stimulates pre-mRNA splicing." GenesDev 13(10): 1234-9.

Huang, S. and D. L. Spector (1996). "Dynamic organization of pre-mRNA splicing factors." J Cell Biochem62(2): 191-7.

Huang, S. and D. L. Spector (1996). "Intron-dependent recruitment of pre-mRNA splicing factors to sites oftranscription." J Cell Biol 133(4): 719-32.

Huang, T. S., C. E. Nilsson, et al. (2002). "Functional inactivation of the SR family of splicing factors during avaccinia virus infection." EMBO Rep 3(11): 1088-93.

Huang, Y., R. Gattoni, et al. (2003). "SR splicing factors serve as adapter proteins for TAP-dependent mRNAexport." Mol Cell 11(3): 837-43.

Inchauspe, G., S. Nagpal, et al. (1989). "Mapping of two varicella-zoster virus-encoded genes that activate theexpression of viral early and late genes." Virology 173(2): 700-9.

Izaurralde, E., J. Lewis, et al. (1995). "A cap-binding protein complex mediating U snRNA export." Nature376(6542): 709-12.

Izaurralde, E., J. Lewis, et al. (1994). "A nuclear cap binding protein complex involved in pre-mRNA splicing."Cell 78(4): 657-68.

Jagatheesan, G., S. Thanumalayan, et al. (1999). "Colocalization of intranuclear lamin foci with RNA splicingfactors." J Cell Sci 112 ( Pt 24): 4651-61.

Johnson, K. W. and K. A. Smith (1991). "Molecular cloning of a novel human cdc2/CDC28-like protein kinase."J Biol Chem 266(6): 3402-7.

Kang, Y. and B. R. Cullen (1999). "The human Tap protein is a nuclear mRNA export factor that contains novelRNA-binding and nucleocytoplasmic transport sequences." Genes Dev 13(9): 1126-39.

Kanopka, A., O. Muhlemann, et al. (1996). "Inhibition by SR proteins of splicing of a regulated adenovirus pre-mRNA." Nature 381(6582): 535-8.

Kanopka, A., O. Muhlemann, et al. (1998). "Regulation of adenovirus alternative RNA splicing bydephosphorylation of SR proteins." Nature 393(6681): 185-7.

Kataoka, N., J. Yong, et al. (2000). "Pre-mRNA splicing imprints mRNA in the nucleus with a novel RNA-binding protein that persists in the cytoplasm." Mol Cell 6(3): 673-82.

Kim, E., L. Du, et al. (1997). "Splicing factors associate with hyperphosphorylated RNA polymerase II in theabsence of pre-mRNA." J Cell Biol 136(1): 19-28.

Kim, V. N. and G. Dreyfus (2001). "Nuclear mRNA binding proteins couple pre-mRNA splicing and post-splicing events." Mol Cells 12(1): 1-10.

Kirshner, J. R., D. M. Lukac, et al. (2000). "Kaposi's sarcoma-associated herpesvirus open reading frame 57encodes a posttranscriptional regulator with multiple distinct activities." J Virol 74(8): 3586-97.

Koffa, M. D., J. B. Clements, et al. (2001). "Herpes simplex virus ICP27 protein provides viral mRNAs withaccess to the cellular mRNA export pathway." Embo J 20(20): 5769-78.

Kohtz, J. D., S. F. Jamison, et al. (1994). "Protein-protein interactions and 5'-splice-site recognition inmammalian mRNA precursors." Nature 368(6467): 119-24.

Koizumi, J., Y. Okamoto, et al. (1999). "The subcellular localization of SF2/ASF is regulated by directinteraction with SR protein kinases (SRPKs)." J Biol Chem 274(16): 11125-31.

Page 48: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

44

Komatsu, M., E. Kominami, et al. (1999). "Cloning and characterization of two neural-salient serine/arginine-rich (NSSR) proteins involved in the regulation of alternative splicing in neurones." Genes Cells 4(10):593-606.

Komeili, A. and E. K. O'Shea (2001). "New perspectives on nuclear transport." Annu Rev Genet 35: 341-64.

Krainer, A. R., G. C. Conway, et al. (1990). "Purification and characterization of pre-mRNA splicing factor SF2from HeLa cells." Genes Dev 4(7): 1158-71.

Kuroyanagi, N., H. Onogi, et al. (1998). "Novel SR-protein-specific kinase, SRPK2, disassembles nuclearspeckles." Biochem Biophys Res Commun 242(2): 357-64.

Lam, B. J. and K. J. Hertel (2002). "A general role for splicing enhancers in exon definition." Rna 8(10): 1233-41.

Lamond, A. I. and D. L. Spector (2003). "Nuclear speckles: a model for nuclear organelles." Nat Rev Mol CellBiol 4(8): 605-12.

Le Hir, H., D. Gatfield, et al. (2001). "The exon-exon junction complex provides a binding platform for factorsinvolved in mRNA export and nonsense-mediated mRNA decay." Embo J 20(17): 4987-97.

Le Hir, H., E. Izaurralde, et al. (2000). "The spliceosome deposits multiple proteins 20-24 nucleotides upstreamof mRNA exon-exon junctions." Embo J 19(24): 6860-9.

Lee, T. I. and R. A. Young (2000). "Transcription of eukaryotic protein-coding genes." Annu Rev Genet 34: 77-137.

Lengyel, J., C. Guy, et al. (2002). "Mapping of functional regions in the amino-terminal portion of the herpessimplex virus ICP27 regulatory protein: importance of the leucine-rich nuclear export signal and RGGBox RNA-binding domain." J Virol 76(23): 11866-79.

Liu, H. X., M. Zhang, et al. (1998). "Identification of functional exonic splicing enhancer motifs recognized byindividual SR proteins." Genes Dev 12(13): 1998-2012.

Liu, K., B. Lemon, et al. (1995). "The dual-specificity phosphatase encoded by vaccinia virus, VH1, is essentialfor viral transcription in vivo and in vitro." J Virol 69(12): 7823-34.

Longman, D., I. L. Johnstone, et al. (2000). "Functional characterization of SR and SR-related genes inCaenorhabditis elegans." Embo J 19(7): 1625-37.

Lopez, A. J. (1998). "Alternative splicing of pre-mRNA: developmental consequences and mechanisms ofregulation." Annu Rev Genet 32: 279-305.

Luo, M. J. and R. Reed (1999). "Splicing is required for rapid and efficient mRNA export in metazoans." ProcNatl Acad Sci U S A 96(26): 14937-42.

Martin, T. E., S. C. Barghusen, et al. (1987). "Redistribution of nuclear ribonucleoprotein antigens during herpessimplex virus infection." J Cell Biol 105(5): 2069-82.

Mayeda, A. and A. R. Krainer (1992). "Regulation of alternative pre-mRNA splicing by hnRNP A1 and splicingfactor SF2." Cell 68(2): 365-75.

Mayeda, A., G. R. Screaton, et al. (1999). "Substrate specificities of SR proteins in constitutive splicing aredetermined by their RNA recognition motifs and composite pre-mRNA exonic elements." Mol CellBiol 19(3): 1853-63.

McGregor, F., A. Phelan, et al. (1996). "Regulation of herpes simplex virus poly (A) site usage and the action ofimmediate-early protein IE63 in the early-late switch." J Virol 70(3): 1931-40.

Page 49: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

45

McKendrick, L., E. Thompson, et al. (2001). "Interaction of eukaryotic translation initiation factor 4G with thenuclear cap-binding complex provides a link between nuclear and cytoplasmic functions of the m(7)guanosine cap." Mol Cell Biol 21(11): 3632-41.

McLauchlan, J., A. Phelan, et al. (1992). "Herpes simplex virus IE63 acts at the posttranscriptional level tostimulate viral mRNA 3' processing." J Virol 66(12): 6939-45.

McLauchlan, J., S. Simpson, et al. (1989). "Herpes simplex virus induces a processing factor that stimulatespoly(A) site usage." Cell 59(6): 1093-105.

Mears, W. E., V. Lam, et al. (1995). "Identification of nuclear and nucleolar localization signals in the herpessimplex virus regulatory protein ICP27." J Virol 69(2): 935-47.

Mears, W. E. and S. A. Rice (1996). "The RGG box motif of the herpes simplex virus ICP27 protein mediates anRNA-binding activity and determines in vivo methylation." J Virol 70(11): 7445-53.

Mintz, P. J., S. D. Patterson, et al. (1999). "Purification and biochemical characterization of interchromatingranule clusters." Embo J 18(15): 4308-20.

Mintz, P. J. and D. L. Spector (2000). "Compartmentalization of RNA processing factors within nuclearspeckles." J Struct Biol 129(2-3): 241-51.

Misteli, T. (2000). "Cell biology of transcription and pre-mRNA splicing: nuclear architecture meets nuclearfunction." J Cell Sci 113(Pt 11): 1841-9.

Misteli, T., J. F. Caceres, et al. (1998). "Serine phosphorylation of SR proteins is required for their recruitment tosites of transcription in vivo." J Cell Biol 143(2): 297-307.

Misteli, T., J. F. Caceres, et al. (1997). "The dynamics of a pre-mRNA splicing factor in living cells." Nature387(6632): 523-7.

Misteli, T. and D. L. Spector (1996). "Serine/threonine phosphatase 1 modulates the subnuclear distribution ofpre-mRNA splicing factors." Mol Biol Cell 7(10): 1559-72.

Misteli, T. and D. L. Spector (1998). "The cellular organization of gene expression." Curr Opin Cell Biol 10(3):323-31.

Misteli, T. and D. L. Spector (1999). "RNA polymerase II targets pre-mRNA splicing factors to transcriptionsites in vivo." Mol Cell 3(6): 697-705.

Moriuchi, H., M. Moriuchi, et al. (1994). "Varicella-zoster virus open reading frame 4 protein is functionallydistinct from and does not complement its herpes simplex virus type 1 homolog, ICP27." J Virol 68(3):1987-92.

Mortillaro, M. J., B. J. Blencowe, et al. (1996). "A hyperphosphorylated form of the large subunit of RNApolymerase II is associated with splicing complexes and the nuclear matrix." Proc Natl Acad Sci U S A93(16): 8253-7.

Murray, H. L. and K. A. Jarrell (1999). "Flipping the switch to an active spliceosome." Cell 96(5): 599-602.

Nakayasu, H. and K. Ueda (1984). "Small nuclear RNA-protein complex anchors on the actin filaments inbovine lymphocyte nuclear matrix." Cell Struct Funct 9(4): 317-25.

Neugebauer, K. M. (2002). "On the importance of being co-transcriptional." J Cell Sci 115(Pt 20): 3865-71.

Nicholas, J., U. A. Gompels, et al. (1988). "Conservation of sequence and function between the product of the52-kilodalton immediate-early gene of herpesvirus saimiri and the BMLF1-encoded transcriptionaleffector (EB2) of Epstein-Barr virus." J Virol 62(9): 3250-7.

Page 50: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

46

Ogg, S. C. and A. I. Lamond (2002). "Cajal bodies and coilin--moving towards function." J Cell Biol 159(1): 17-21.

Ohno, M., A. Segref, et al. (2000). "PHAX, a mediator of U snRNA nuclear export whose activity is regulatedby phosphorylation." Cell 101(2): 187-98.

Petersen-Mahrt, S. K., C. Estmer, et al. (1999). "The splicing factor-associated protein, p32, regulates RNAsplicing by inhibiting ASF/SF2 RNA binding and phosphorylation." Embo J 18(4): 1014-24.

Phelan, A., M. Carmo-Fonseca, et al. (1993). "A herpes simplex virus type 1 immediate-early gene product,IE63, regulates small nuclear ribonucleoprotein distribution." Proc Natl Acad Sci U S A 90(19): 9056-60.

Pombo, A., J. Ferreira, et al. (1994). "Adenovirus replication and transcription sites are spatially separated in thenucleus of infected cells." Embo J 13(21): 5075-85.

Pombo, C.-F. (1995). "Interactions of Adenovirus with the nucleus of the Host Cell." Medical Virology 5: 213-218.

Prasad, J., K. Colwill, et al. (1999). "The protein kinase Clk/Sty directly modulates SR protein activity: bothhyper- and hypophosphorylation inhibit splicing." Mol Cell Biol 19(10): 6991-7000.

Puvion-Dutilleul, F. and E. Pichard (1992). "Segregation of viral double-stranded and single-stranded DNAmolecules in nuclei of adenovirus infected cells as revealed by electron microscope in situhybridization." Biol Cell 76(2): 139-50.

Reed, R. (2000). "Mechanisms of fidelity in pre-mRNA splicing." Curr Opin Cell Biol 12(3): 340-5.

Rice, S. A. and D. M. Knipe (1990). "Genetic evidence for two distinct transactivation functions of the herpessimplex virus alpha protein ICP27." J Virol 64(4): 1704-15.

Rice, S. A. and V. Lam (1994). "Amino acid substitution mutations in the herpes simplex virus ICP27 proteindefine an essential gene regulation function." J Virol 68(2): 823-33.

Rice, S. A., V. Lam, et al. (1993). "The acidic amino-terminal region of herpes simplex virus type 1 alphaprotein ICP27 is required for an essential lytic function." J Virol 67(4): 1778-87.

Ring, H. Z. and J. T. Lis (1994). "The SR protein B52/SRp55 is essential for Drosophila development." Mol CellBiol 14(11): 7499-506.

Roizman, B. (1996). Herpes simplex viruses and their replication. Philadelphia, PA, Lippincott-Raven.

Roizman, B. (1996). Herpesviridae. Philadelphia, PA, Lippincott-Raven.

Rueter, S. M., T. R. Dawson, et al. (1999). "Regulation of alternative splicing by RNA editing." Nature399(6731): 75-80.

Ruskin, B., P. D. Zamore, et al. (1988). "A factor, U2AF, is required for U2 snRNP binding and splicingcomplex assembly." Cell 52(2): 207-19.

Ruvolo, V., E. Wang, et al. (1998). "The Epstein-Barr virus nuclear protein SM is both a post-transcriptionalinhibitor and activator of gene expression." Proc Natl Acad Sci U S A 95(15): 8852-7.

S.J Flint, L. W. E., R.M Krug, V.R Racaniello, A.M Skalka (2000). Principles of Virology. Washington, D.C.,ASM Press.

Sandri-Goldin, R. M. (1994). "Properties of an HSV-1 regulatory protein that appears to impair host cellsplicing." Infect Agents Dis 3(2-3): 59-67.

Page 51: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

47

Sandri-Goldin, R. M. (1998). "ICP27 mediates HSV RNA export by shuttling through a leucine-rich nuclearexport signal and binding viral intronless RNAs through an RGG motif." Genes Dev 12(6): 868-79.

Sandri-Goldin, R. M. and M. K. Hibbard (1996). "The herpes simplex virus type 1 regulatory protein ICP27coimmunoprecipitates with anti-Sm antiserum, and the C terminus appears to be required for thisinteraction." J Virol 70(1): 108-18.

Sandri-Goldin, R. M., M. K. Hibbard, et al. (1995). "The C-terminal repressor region of herpes simplex virustype 1 ICP27 is required for the redistribution of small nuclear ribonucleoprotein particles and splicingfactor SC35; however, these alterations are not sufficient to inhibit host cell splicing." J Virol 69(10):6063-76.

Sandri-Goldin, R. M. and G. E. Mendoza (1992). "A herpesvirus regulatory protein appears to act post-transcriptionally by affecting mRNA processing." Genes Dev 6(5): 848-63.

Schaub, M. and W. Keller (2002). "RNA editing by adenosine deaminases generates RNA and proteindiversity." Biochimie 84(8): 791-803.

Schroeder, S. C., B. Schwer, et al. (2000). "Dynamic association of capping enzymes with transcribing RNApolymerase II." Genes Dev 14(19): 2435-40.

Schul, W., R. van Driel, et al. (1998). "A subset of poly(A) polymerase is concentrated at sites of RNA synthesisand is associated with domains enriched in splicing factors and poly(A) RNA." Exp Cell Res 238(1): 1-12.

Sciabica, K. S., Q. J. Dai, et al. (2003). "ICP27 interacts with SRPK1 to mediate HSV splicing inhibition byaltering SR protein phosphorylation." Embo J 22(7): 1608-19.

Screaton, G. R., J. F. Caceres, et al. (1995). "Identification and characterization of three members of the humanSR family of pre-mRNA splicing factors." Embo J 14(17): 4336-49.

Semmes, O. J., L. Chen, et al. (1998). "Mta has properties of an RNA export protein and increases cytoplasmicaccumulation of Epstein-Barr virus replication gene mRNA." J Virol 72(12): 9526-34.

Shatkin, A. J. and J. L. Manley (2000). "The ends of the affair: capping and polyadenylation." Nat Struct Biol7(10): 838-42.

Shenk, T. (1996). Adenoviridae: The viruses and their replication. Philadelphia, PA, Lippincott-Raven.

Shin, C. and J. L. Manley (2002). "The SR protein SRp38 represses splicing in M phase cells." Cell 111(3): 407-17.

Smith, C. W. and J. Valcarcel (2000). "Alternative pre-mRNA splicing: the logic of combinatorial control."Trends Biochem Sci 25(8): 381-8.

Sokolowski, M., J. E. Scott, et al. (2003). "Identification of herpes simplex virus RNAs that interact specificallywith regulatory protein ICP27 in vivo." J Biol Chem 278(35): 33540-9.

Soliman, T. M. and S. J. Silverstein (2000). "Identification of an export control sequence and a requirement forthe KH domains in ICP27 from herpes simplex virus type 1." J Virol 74(16): 7600-9.

Spritz, R. A., K. Strunk, et al. (1987). "The human U1-70K snRNP protein: cDNA cloning, chromosomallocalization, expression, alternative splicing and RNA-binding." Nucleic Acids Res 15(24): 10373-91.

Staley, J. P. and C. Guthrie (1998). "Mechanical devices of the spliceosome: motors, clocks, springs, andthings." Cell 92(3): 315-26.

Sun, Q., A. Mayeda, et al. (1993). "General splicing factor SF2/ASF promotes alternative splicing by binding toan exonic splicing enhancer." Genes Dev 7(12B): 2598-608.

Page 52: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

48

Tacke, R. and J. L. Manley (1999). "Determinants of SR protein specificity." Curr Opin Cell Biol 11(3): 358-62.

Tarn, W. Y. and J. A. Steitz (1996). "A novel spliceosome containing U11, U12, and U5 snRNPs excises aminor class (AT-AC) intron in vitro." Cell 84(5): 801-11.

Thiry, M. (1995). "The interchromatin granules." Histol Histopathol 10(4): 1035-45.

Vagner, S., C. Vagner, et al. (2000). "The carboxyl terminus of vertebrate poly(A) polymerase interacts withU2AF 65 to couple 3'-end processing and splicing." Genes Dev 14(4): 403-13.

Valcarcel, J., R. K. Gaur, et al. (1996). "Interaction of U2AF65 RS region with pre-mRNA branch point andpromotion of base pairing with U2 snRNA [corrected]." Science 273(5282): 1706-9.

Varani G., N. K. (1998). "RNA recognition by RNP proteins during RNA processing." Annu. Rev. Biophys.Biomol. Struct. 27: 407-45.

Visa, N., F. Puvion-Dutilleul, et al. (1993). "Intranuclear distribution of U1 and U2 snRNAs visualized by highresolution in situ hybridization: revelation of a novel compartment containing U1 but not U2 snRNA inHeLa cells." Eur J Cell Biol 60(2): 308-21.

Visa, N., F. Puvion-Dutilleul, et al. (1993). "Intranuclear distribution of poly(A) RNA determined by electronmicroscope in situ hybridization." Exp Cell Res 208(1): 19-34.

Wadd, S., H. Bryant, et al. (1999). "The multifunctional herpes simplex virus IE63 protein interacts withheterogeneous ribonucleoprotein K and with casein kinase 2." J Biol Chem 274(41): 28991-8.

Wang, H. Y., W. Lin, et al. (1998). "SRPK2: a differentially expressed SR protein-specific kinase involved inmediating the interaction and localization of pre-mRNA splicing factors in mammalian cells." J CellBiol 140(4): 737-50.

Wang, J., Y. Takagaki, et al. (1996). "Targeted disruption of an essential vertebrate gene: ASF/SF2 is requiredfor cell viability." Genes Dev 10(20): 2588-99.

Wang, J., S. H. Xiao, et al. (1998). "Genetic analysis of the SR protein ASF/SF2: interchangeability of RSdomains and negative control of splicing." Genes Dev 12(14): 2222-33.

Weir, J. P. (2001). "Regulation of herpes simplex virus gene expression." Gene 271(2): 117-30.

Wetterberg, I., G. Bauren, et al. (1996). "The intranuclear site of excision of each intron in Balbiani ring 3 pre-mRNA is influenced by the time remaining to transcription termination and different excisionefficiencies for the various introns." Rna 2(7): 641-51.

Whitley, R. (1996). Herpes simplex viruses. Philadelphia, PA, Lippincott-Raven.

Will, C. L. and R. Luhrmann (2001). "Spliceosomal UsnRNP biogenesis, structure and function." Curr Opin CellBiol 13(3): 290-301.

Winkler, M., S. A. Rice, et al. (1994). "UL69 of human cytomegalovirus, an open reading frame with homologyto ICP27 of herpes simplex virus, encodes a transactivator of gene expression." J Virol 68(6): 3943-54.

Wu, J. Y. and T. Maniatis (1993). "Specific interactions between proteins implicated in splice site selection andregulated alternative splicing." Cell 75(6): 1061-70.

Wu, Q. and A. R. Krainer (1999). "AT-AC pre-mRNA splicing mechanisms and conservation of minor intronsin voltage-gated ion channel genes." Mol Cell Biol 19(5): 3225-36.

Wu, S., C. M. Romfo, et al. (1999). "Functional recognition of the 3' splice site AG by the splicing factorU2AF35." Nature 402(6763): 832-5.

Page 53: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

49

Xiao, S. H. and J. L. Manley (1997). "Phosphorylation of the ASF/SF2 RS domain affects both protein-proteinand protein-RNA interactions and is necessary for splicing." Genes Dev 11(3): 334-44.

Xiao, S. H. and J. L. Manley (1998). "Phosphorylation-dephosphorylation differentially affects activities ofsplicing factor ASF/SF2." Embo J 17(21): 6359-67.

Xing, Y., C. V. Johnson, et al. (1993). "Higher level organization of individual gene transcription and RNAsplicing." Science 259(5099): 1326-30.

Yang, L., L. J. Embree, et al. (2000). "TLS-ERG leukemia fusion protein inhibits RNA splicing mediated byserine-arginine proteins." Mol Cell Biol 20(10): 3345-54.

Yuryev, A., M. Patturajan, et al. (1996). "The C-terminal domain of the largest subunit of RNA polymerase IIinteracts with a novel set of serine/arginine-rich proteins." Proc Natl Acad Sci U S A 93(14): 6975-80.

Zahler, A. M., W. S. Lane, et al. (1992). "SR proteins: a conserved family of pre-mRNA splicing factors." GenesDev 6(5): 837-47.

Zahler, A. M., K. M. Neugebauer, et al. (1993). "Distinct functions of SR proteins in alternative pre-mRNAsplicing." Science 260(5105): 219-22.

Zamore, P. D., J. G. Patton, et al. (1992). "Cloning and domain structure of the mammalian splicing factorU2AF." Nature 355(6361): 609-14.

Zhang, M., P. D. Zamore, et al. (1992). "Cloning and intracellular localization of the U2 small nuclearribonucleoprotein auxiliary factor small subunit." Proc Natl Acad Sci U S A 89(18): 8769-73.

Zhao, J., L. Hyman, et al. (1999). "Formation of mRNA 3' ends in eukaryotes: mechanism, regulation, andinterrelationships with other steps in mRNA synthesis." Microbiol Mol Biol Rev 63(2): 405-45.

Zhao, Y., V. R. Holden, et al. (1992). "Identification and transcriptional analyses of the UL3 and UL4 genes ofequine herpesvirus 1, homologs of the ICP27 and glycoprotein K genes of herpes simplex virus." JVirol 66(9): 5363-72.

Zhou, C. and D. M. Knipe (2002). "Association of herpes simplex virus type 1 ICP8 and ICP27 proteins withcellular RNA polymerase II holoenzyme." J Virol 76(12): 5893-904.

Zhou, Z., M. J. Luo, et al. (2000). "The protein Aly links pre-messenger-RNA splicing to nuclear export inmetazoans." Nature 407(6802): 401-5.

Zhou, Z. and R. Reed (1998). "Human homologs of yeast prp16 and prp17 reveal conservation of the mechanismfor catalytic step II of pre-mRNA splicing." Embo J 17(7): 2095-106.

Zhu, J., A. Mayeda, et al. (2001). "Exon identity established through differential antagonism between exonicsplicing silencer-bound hnRNP A1 and enhancer-bound SR proteins." Mol Cell 8(6): 1351-61.

Page 54: Viruses as a Model System for Studies of Eukaryotic mRNA Processing

Acta Universitatis UpsaliensisComprehensive Summaries of Uppsala Dissertations

from the Faculty of Medicine

Editor: The Dean of the Faculty of Medicine

Distribution:Uppsala University Library

Box 510, SE-751 20 Uppsala, Swedenwww.uu.se, [email protected]

ISSN 0282-7476ISBN 91-554-5775-4

A doctoral dissertation from the Faculty of Medicine, Uppsala University,is usually a summary of a number of papers. A few copies of the completedissertation are kept at major Swedish research libraries, while the sum-mary alone is distributed internationally through the series Comprehen-

sive Summaries of Uppsala Dissertations from the Faculty of Medicine.

(Prior to October, 1985, the series was published under the title “Abstracts ofUppsala Dissertations from the Faculty of Medicine”.)


Recommended