Download fulltekst

Document related concepts

Hedgehog signaling pathway wikipedia , lookup

Protein (nutrient) wikipedia , lookup

Magnesium transporter wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Protein wikipedia , lookup

Histone acetylation and deacetylation wikipedia , lookup

Cell nucleus wikipedia , lookup

Signal transduction wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

Intrinsically disordered proteins wikipedia , lookup

Protein moonlighting wikipedia , lookup

Phosphorylation wikipedia , lookup

List of types of proteins wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Silencer (genetics) wikipedia , lookup

SR protein wikipedia , lookup

Transcriptional regulation wikipedia , lookup

Transcript
Digital Comprehensive Summaries of Uppsala Dissertations
from the Faculty of Medicine 139
Functional Characterization of
the Cellular Protein p32
A Protein Regulating Adenovirus Transcription and
Splicing Through Targeting of Phosphorylation
CHRISTINA ÖHRMALM
ACTA
UNIVERSITATIS
UPSALIENSIS
UPPSALA
2006
ISSN 1651-6206
ISBN 91-554-6533-1
urn:nbn:se:uu:diva-6794
!
"# $ % ! & % % '!! ()
%
*+ ! , -&!+
.!
+ "#+ )
!
/
% ! ' "+ 0 '
1&
& 0 2& !&! && % '!!
+
0
+ 3+ #3 + + 425 36$$76#$6+
! % ! & %
% 50 % &
!!
+ & !
!!
% 8 ! ! !
+ 4 !
! 4 !
! &%
% !!
! &
% & & !
+
! !, !
! % ! 21 %
% & !
% 6
!!!
+ ! -7691)7 ,
!, 21 !!
& ! !!
''"0+ ! -7691)7:''"0 ,
!, ! 21 6
% 6% & %!
& % + ! !,
!
& 8 ! &
&
!!
+
2
! " ,
!, !!!
% ! 21
02):2)" & 150 & % 02):2)"+ ! !
& 02):2)" 150 & !
%% ! % % 02):2)" &
%%& !+ ! " ! -7691)7 ! & ! & %
02):2)"+ ;, ! &
& %% !
+ -7691)7 !!
!!
02):2)" ,!
" < 02):2)" +
)! , !
% " & % % ! = + " % 00 6 & & ! % ! %
):5)6> ! &
+ 0 %! !, !
" !!!
% ! % 150 ' 44 ,!! &%
! % ' 44 & ! &
!
% +
4 , !
!, !
-7691)7 &
! % & &
21
!
" &
! % ! 21 02):2)" & ' 44
& &
+ !
! -7691)7 " % &
& ! !!
% 8 !
+
0 " 02):2)" -7691)7 2& 150 ' 44
! " #! "$ %&'!
! ()*%+', ! ? !
.!
"#
4225 #$6#"#
425 36$$76#$6
6#@37 (!::+8+:AB
6#@37*
Till min älskade familj
Maja, Saga och Johan
Mamma och Pappa
Annika
Members of the committee
Opponent
Associate Professor Neus Visa
Department of Molecular Biology and Functional Genomics
The Wenner-Gren Institute
Stockholm University
Stockholm
Members of the committee
Professor Anders Virtanen
Department of Cell and Molecular Biology
Biomedical Center
Uppsala University
Uppsala
Professor Stefan Schwartz
Department of Medical Biochemistry and Microbiology
Biomedical Center
Uppsala University
Uppsala
Associate professor Marie Öhman
Department of Molecular Biology and Functional Genomics
The Wenner-Gren Institute
Stockholm University
Stockholm
List of manuscript
I.
Kanopka, A., Mühlemann, O., Petersen-Mahrt, S., Estmer, C.,
Öhrmalm, C., Akusjärvi, G. (1998) Regulation of adenovirus
alternative RNA splicing by dephosphorylation of SR proteins,
Nature 393:185-187
II.
Petersen-Mahrt, S. K., Estmer, C., Öhrmalm, C., Matthews,
D. A., Russell, W. C., Akusjärvi, G. (1999) The splicing factor-associated protein, p32, regulates RNA splicing by inhibiting ASF/SF2 RNA binding and phosphorylation, EMBO J
18:1014-1024
III.
Öhrmalm, C. and Akusjärvi, G (2006) Identification of a
carboxy-terminal sequence in the splicing factor-associated
protein, p32, necessary for p32-mediated inhibition of
ASF/SF2 RNA binding, Manuscript
IV.
Öhrmalm, C. and Akusjärvi, G. (2006) Cellular splicing and
transcription regulatory protein p32 represses adenovirus
major late transcription and causes hyperphosphorylation of
RNA polymerase II, In press Journal of Virology
Contents
Introduction...................................................................................................11
From gene to protein ................................................................................11
Transcription ............................................................................................12
Basal transcription of mRNA encoding genes.....................................12
Transcription cycle of RNA Pol II.......................................................12
Coupling of transcription and splicing ................................................15
The CTD kinases and phosphatases ....................................................15
Proteins involved in regulating the activity of CTD kinases and
phosphatases ........................................................................................16
Activation and Repression of Transcription ........................................17
Splicing ....................................................................................................19
Constitutive splicing ............................................................................19
Alternative splicing..............................................................................19
Spliceosome assembly .........................................................................20
SR proteins ..........................................................................................21
SR-related proteins ..............................................................................24
Phosphorylation and dephosphorylation of SR proteins......................24
Proteins regulating SR protein activity................................................24
p32............................................................................................................27
Structure of p32 ...................................................................................28
p32 protein interactions .......................................................................29
p32 and protein modifications .............................................................31
Adenovirus ...............................................................................................32
Virus life cycle.....................................................................................32
Early genes ..........................................................................................34
The Major Late Transcription Unit......................................................36
Present Investigation and discussion.............................................................39
Paper I ......................................................................................................39
Regulation of adenovirus alternative RNA splicing by
dephosphorylation of SR proteins .......................................................39
Paper II .....................................................................................................41
p32 regulates RNA splicing by inhibiting ASF/SF2 RNA binding and
phosphorylation ...................................................................................41
Paper III....................................................................................................44
Identification of a carboxy-terminal sequence in p32 necessary for
p32-mediated inhibition of ASF/SF2 RNA binding ............................44
Paper IV ...................................................................................................46
p32 represses the adenovirus major late transcription and causes
hyperphosphorylation of RNA polymerase II .....................................46
Conclusions...................................................................................................52
Paper I ......................................................................................................52
Paper II .....................................................................................................52
Paper III....................................................................................................52
Paper IV ...................................................................................................52
Acknowledgements.......................................................................................53
References.....................................................................................................55
Abbreviations
aa
AdTTflag-p32
Ad5
ASF/SF2
bp
CDK7
CDK9
CTD
C-terminal
DNA
Dox
ESE
E4-ORF4
FCP1
HeLa-NE
hpi
kDa
L1
MLP
MLTU
mRNA
NE
N-terminal
PIC
Pol II
Pol IIA
Pol IIO
PP2A
Pre-mRNA
P-TEFb
p(Y) tract
p32
RNA
amino acid
recombinant inducible adenovirus expressing
flag-tagged p32
Adenovirus 5
Alternative splicing factor/splicing factor 2
branch point
cycline-dependent kinase 7
cycline-dependent kinase 9
carboxy-terminal domain of the large subunit of
RNA Pol II
carboxy-terminal
deoxyribonucleic acid
Doxycyclin
exonic splicing enhancer
Protein of open reading frame 4 of early region 4
TFIIF- associating RNA polymerase C-terminal
domain phosphatase
Nuclear extract from HeLa cells
Hours post infection
kilo Dalton, molecular weight
Late region 1 of MLTU of adenovirus
Major late promoter
Major late transcription unit
messenger ribonucleic acid
Nuclear extract
amino-terminal
pre-initiation complex of transcription
RNA polymerase II
Hypophosphorylated RNA polymerase II
Hyperphosphorylated RNA polymerase II
Protein phosphatase 2A
Precursor mRNA
positive transcription elongation factor b
polypyrimidine rich sequence upstream of 3´ss
cellular protein with molecular weight of 32 kDa
ribonucleic acid
RRM
RS-domain
Ser 2
Ser 5
snRNA
SR-Ad
SR-HeLa
snRNP
ss
TF
3RE
RNA recognition motif
Arginine (R)- and Serine (S)-rich domain in
SR proteins
Serine at position 2
Serine at position 5
small nuclear RNA
SR proteins from adenovirus infected nuclear extract
SR proteins purified from HeLa nuclear extract
small nuclear ribonucleoprotein particle
splice site
transcription factor
IIIa repressor element
Introduction
From gene to protein
All complex organisms, like an animal or a plant, are built up of small specified cells forming different tissues that together create the organism. Within
its cell nucleus, each cell contains identical genomic material that describes
the construction of different proteins performing different functions or acting
as building blocks in the cell. For a cell to become specialized, also called
differentiated, only some genes of the genome are activated and transcribed
into pre-mRNAs, which after being processed and transported from the cell
nucleus to the cytoplasm further are translated into proteins. All these steps
are highly regulated at many different levels.
A virus consists of genetic material, DNA or RNA, and a capsid of proteins, which is enveloped or non-enveloped with membrane from the cell the
virus was produced in. The shell of proteins functions both to protect the
viral genome and to provide molecules to target cell surface receptors during
infection. The genome of a virus must be small enough to fit into the viral
capsid and consequently only contains a small number of genes. These encode the structural proteins of the virus and proteins needed to reprogram the
cell to efficiently produce new viral particles. Both eukaryotic cells and
some virus compact their genome by utilizing one gene to produce many
diverse proteins. This is achieved by a process called alternative splicing.
The pre-mRNA, which is produced during transcription of the gene, is
cleaved and different fragments are ligated in a highly regulated manner by
the spliceosome to create different mRNAs, giving the ability for one gene to
encode for many different proteins. Both transcription and splicing are regulated by cellular as well as viral proteins. The activity of these regulatory
proteins is often regulated by phosphorylation and dephosphorylation, enabling the virus or the cell to efficiently control a particular process.
In this thesis we describe how an adenoviral protein, the E4-ORF4 protein, and a cellular protein, the p32 protein, regulate the splicing event by
taking control over the activity of ASF/SF2, which is an important splicing
regulatory protein (Paper I and Paper II, respectively). Moreover we have
analyzed which parts of p32 and ASF/SF2 that interact with each other (Paper III). In paper IV we show that p32 also harbors the function as a transcription regulatory protein (Paper IV).
11
Transcription
Transcription of genes can be preformed by three different multi-subunit
complexes containing specific RNA polymerases, Pol II:
x Pol I transcribes the ribosomal genes to produce 28S, 18S and 5.8S ribosomal RNA, rRNA
x Pol II transcribes protein encoding genes to produce messenger RNA,
mRNA
x Pol III transcribes genes to produce transfere RNA, tRNA, and 5S rRNA
The polymerase is binding to the promoter, mostly situated just upstream of
the initiator site of transcription, +1, although some RNA Pol III genes have
internal promoters. After polymerase complex assembly transcription can be
further activated or repressed by different gene specific transcription coactivators and co-repressors.
Basal transcription of mRNA encoding genes
Most Pol II promoters contain a TATA-element and an initiator site, INR.
The TATA-element is situated approximately 25-30 base pairs upstream of
the transcription start site and is the binding site for the TATA-binding protein, TBP, whereas the INR is a pyrimidine rich sequence surrounding the +1
start site (-3 to +5) [181]. The INR constitutes the basal element for initiation
of transcription in promoters lacking a TATA-box [180, 219]
The general transcription factors are recruited to the promoter in a specific order (for review [147]). The first transcription factor, TF, to bind to the
promoter is TFIID, followed by TFIIB. The TATA-binding protein, TBP, is
a subunit of TFIID and binds to the minor groove of the DNA in the TATAsequence [88, 183]. The co-crystal structure of TBP binding to the TATA-box
of the adenovirus major late promoter, MLP, has been described [88]. The
additional subunits of TFIID are the TBP associated factors, TAFIIS [56]. The
Pol II and TFIIF associate and bind as a complex to the promoter/TFIID/TFIIB. Subsequently, TFIIH and TFIIE follow and the complex formed is called the pre-initiation complex, PIC. TFIIA can join the
complex at any time after TFIID has bound to DNA. The PIC causes the two
DNA strands the promoter region to melt and the first phosphodiester bond
is produced to form the protruding pre-mRNA. Thereafter, the Pol IIcomplex leaves the promoter, promoter clearance, and starts the elongation
phase of transcription.
Transcription cycle of RNA Pol II
The metazoan large subunit of Pol II has a carboxy-terminal domain, CTD,
consisting of a heptapeptide sequence (YSPTSPS) repeated 52 times. This
12
motif undergoes extensive serine phosphorylation and dephosphorylation
during the transcription cycle, especially at positions 2 and 5, and works as a
platform for the recruitment of factors involved in co-transcriptional events
like capping, splicing, and poly-adenylation of the nascent pre-mRNA [15].
The length of the CTD varies in different organisms:
x
x
x
x
Homo Sapiens the heptad motif is repeated 52 times
Drosophila melanogaster 45 times
Caenorhabditis elegans 37 times
Saccharomyces cerevisiae 27 times
It is thought that the length of the CTD increases with the complexity of
the organism, indicating that complex organisms need for a more intricate
system to recruit pre-mRNA processing factors [38]. The N-terminal part of
the CTD consists of perfect YSPTSPS repeats and has been shown to support RNA synthesis and capping, while the heptad repeats towards the C-terminal end of the CTD deviate from the consensus sequence and has been
shown to support splicing and poly-adenylation of the 3’end of the transcript
[49]. In vertebrates the C-terminus of the CTD also contains ten conserved
amino acids, aa,: ISPDDSDEEN, which are essential for high level of transcription, splicing and poly(A) site cleavage [50].
Transcription is initiated by hypophosphorylated Pol II, Pol IIA, binding
to the promoter to form the pre-initiation complex, PIC (Figure 1). After PIC
formation the CTD is phosphorylated on Ser 5 by CDK7/TFIIH. The hyperphosphorylation on Ser 5 is necessary for promoter clearance and the transition from the initiation to the elongation stage of transcription [94]. Soon
after PIC formation the DRB sensitivity-inducing factor, DSIF, is recruited,
which in turn recruits the negative elongation factor, NELF, to the Pol II
resulting in an arrest of transcription allowing time for the Ser 5 hyperphosphorylated Pol II, Pol IIO, to recruit capping enzymes to the protruding RNA
molecule [94, 168, 212]. Capping guanylyltransferase, Cgt1, and RNA guanine-7-methyltransferease bind directly to phosphorylated CTD [87]. Cgt1 is
specifically stimulated by Ser 5 phosphorylation [70, 126] and deletion studies
have shown that Pol II without a CTD produces transcripts with a lower
proportion of capped 5’ ends [126]. The capping enzymes are also suggested
to play a role in the promoter-proximal checkpoint to ensure that only correctly capped transcripts are allowed to be further transcribed [87, 120].
The CDK9 kinase of the positive elongating factor b, P-TEFb, phosphorylates the CTD on Ser 2 residues leading to a relief of the DSIF/NELF mediated arrest of transcription elongation. The increase in Ser 2 phosphorylation
renders the Pol II to be more processive during the elongation phase of transcription [35]. Phosphorylation of Ser 2 is also important for the coupling of
transcription and mRNA poly-adenylation. Thus, deletion of the yeast Ser 2
kinase, Ctk1, or inhibition of CDK9/P-TEFb causes defects in poly13
adenylation [4, 141]. Poly-adenylation factors CPSF, CstF, Ocfl 1, and Pta 1
bind preferentially to phosphorylated CTD [126].
The phosphatase FCP1 is counteracting CDK9/P-TEFb during the elongation by causing Ser 2 dephosphorylation [35]. Mutations of FCP1 cause a
decrease in the number of polymerases initiating at a promoter and an increase in the amount of Ser 2 phosphorylation during elongation, while the
level of Ser 5 phosphorylation seems to be unaffected [35]. FCP1 has been
shown to be responsible to recycle Pol II [36, 92].
FCP1
PIC formation
Elongation
Initiation
Termination
RNA pol II
CTD
CDK7
CDK9
mRNA
pre-mRNA
Capping
AA
AA
AA
A
DNA
Splicing
Poly-adenylation
Splicing factors
Exon
Ser 5 phosphorylation
Intron lariat
Cap
Ser 2 phosphorylation
Figure 1. The transcription cycle of the Pol II and the coupling of transcription and
splicing. The CTD of Pol II is phosphorylated at Ser 5 by CDK7 during initiation
and at Ser 2 by CDK9 during the elongation phase of transcription. FCP1 dephosphorylates the CTD to recycle the Pol II.
14
Coupling of transcription and splicing
During the last years many reports have provided evidence that splicing occurs co-transcriptionally (Figure 1) [reviewed in 95, 139, 148]. Truncation of the
CTD in in vivo experiments shows a failure to recruit splicing factors to transcription start sites and inefficient pre-mRNA splicing [127, 129]. Further,
in vitro splicing experiments have demonstrated that addition of purified
Pol IIO stimulates splicing, whereas Pol IIA inhibits the splicing reaction
[69]. Promoter proximal splice sites have shown to enhance transcription [52].
The strength of a splice site can also effect the Pol II processivity [144], and
stimulation of Pol II processivity affects splice site choice in alternative
splicing [143]. Different transcriptional factors and splicing proteins can interact with each other. The subunit Prp40 of the spliceosomal U1 small nuclear ribonucleoprotein particle, U1 snRNP, has also been shown to interact
directly with the CTD of Pol II [84, 138] and furthermore the human transcription elongation factor TAT-SF1 interacts with U snRNPs resulting in stimulated Pol II elongation and an increase in splicing efficiency in in vitro splicing assays [51].
Table 1. CTD kinases and phosphatases
Human
Yeast
Main target
Action
Ref.
Kin28-Ccl1
Ser 5
Facilitates promoter clearance and mRNA capping
[157,
163]
Ser 2
Promotes elongation and
recruitment of poly(A)factors
Promotes elongation
[157]
CTD kinases
CDK7/cyclinH
in TFIIH
CDK9/cyclinT
in P-TEFb
CDK8/cyclin8
in mediator NAT
CTDK-1
Ser 2
[163]
Bur1-Bur2
Ser 5
Srb10-Srb11
Ser 5, cyclin H
Inhibits PIC formation
[157]
Ser 5 and Ser 2
Recycling of Pol II
[109]
[163]
CTD Phosphatases
FCP1
Ssu72
SCP1
Ssu72
Ser 5
[99]
Ser 5
[214]
The CTD kinases and phosphatases
The kinases and phosphatases known to have a regulatory role in controlling
the phosphorylation status of the CTD are summaried in Table 1. As described above, phosphorylation of Ser 5 and Ser 2 of CTD by CDK7/TFIIH
and CDK9/P-TEFb, respectively, and dephosphorylation by FCP1 create a
15
phosphorylation/dephosphorylation cycle for the Pol II (Figure 1). Another
kinase, CDK8/cyclin C, found in the mediator complex NAT, is thought to
repress transcription by reducing the number of Pol II molecules capable of
initiating transcription by phosphorylating the CTD prior to DNA binding
and also by phosphorylating cyclinH in the CDK7/cyclinH complex of the
general transcription factor TFIIH [5, 185]. Recently, the SCP1 phosphatase
was shown to dephosphorylate Ser 5 and it is thought to play a key role in
the transition from initiation to elongation. Both FCP1 and SCP1 activities
are enhanced by the RAP74 subunit of TFIIF [214].
Proteins involved in regulating the activity of CTD kinases and
phosphatases
In all cellular processes, for example transcription and splicing, a tight control of different protein activities is of a great importance. The transcriptional
activity of different gene families varies between cell types and transcription
is strictly regulated during cell differentiation and the cell cycle. Transcription is also regulated at different stages, both before initiation and during the
elongation phase of transcription. As an example, transcriptional activators
and repressors can act by recruiting histone acetyl transferases, HATs, or
histone deacetylaces, HDACs, to enhance or inhibit transcription, respectively.
The importance of the phosphorylation status of the CTD in different
phases of transcription makes the CTD a great target for proteins that regulate transcription. The variable phosphorylation pattern of the CTD attracts
different proteins important for post- or co-transcriptional processes. By
enhancing or inhibiting phosphorylation of specific CTD residues the recruitment of these factors can be affected as well as the activity and the
processivity of Pol II. Only a few proteins have so far been reported to
stimulate or inhibit kinases and phosphatases involved in CTD phosphorylation (see Table 2 and section regulating the activities of RNA Pol II kinases
and phosphatases in Paper IV). The Pin 1, as well as its yeast homologue
Essp1, binds directly to the CTD, while others like RAP74 is part of a classical transcriptional factor, TFIIF. Paper IV demonstrates that the p32 protein
now can be added to this list of regulatory proteins controlling the activity of
CTD kinases/phosphatases (Table 2).
16
Table 2. Proteins affecting the activity of Pol II phosphatases and kinases
Protein
Effect on phophatase
Effect on kinase
Comments
Ref.
-
+ cdc2/cyclinB
Mitotic genes
[211]
+
Mitotic genes
[206]
Binds all four yeast
CTD kinases
Cellular
Pin 1
FCP1
Ess1p
RAP74
+
FCP1 and SCP1
Hce1
-
FCP1
p32
Not known
Not known
HIV-1 Tat
-
+
EBNA2
Not known
BRCA1
[32, 93,
214]
-
CDK7/TFIIH
[133]
Capping enzyme
[149]
Ser 5 and Ser 2
hyperphosphorylation
Paper IV
viral
FCP1
CDK9/P-TEFb
Not known
[1, 89]
Ser 5 hyperphosphorylation
[9]
Activation and Repression of Transcription
To enhance or inhibit the basal transcription machinery different transcriptional co-activators or co-repressors bind to enhancer elements or repressor
elements. These can be placed within the promoter region or at a very long
distance from the gene. The co-activators and co-repressors can have enzymatic activities themselves or they can recruit other proteins that introduce
different modifications like phosphorylation, acetylation, sumolylation,
ubiquitination, or methylation, on either the TFs, the co-factors or on the
DNA, resulting in changes of the transcriptional activity of the gene.
CBF/NF-Y and the CAAT-box motif
CBF/NF-Y is a heterotrimeric transcription factor consisting of three subunits; NF-YA, NF-YB and NF-YC. The CBF/NF-Y protein is evolutionary
conserved and shows more then 95% aa sequence identity between mouse,
rat and human isoforms [105]. All three subunits are needed for DNA binding
to the CAAT-box motif, one of the most common sequence elements in eukaryotic promoters. The consensus binding site for CBP/NF-Y is defined as
5´-(T/C) (A/G) (A/G) CCAAT (C/G) (A/G)-3´ [18]. Mutational studies have
17
shown that CBF/NF-Y requires all five nucleotides in the core motif for efficient binding [116]. Not all promoters that contain the CAAT-box are activated by CBF/NF-Y demonstrating that the neighboring sequences of the
motif are also important for high affinity binding of CBF/NF-Y. There are
many different transcription factors that bind to the CCAAT-box motifs, like
CTF/NF1, a 47 kDA protein that recognizes the sequence GCCAAT [166]
and the family of CCAAT/enhancer binding proteins, C/EBP, which has the
(A/G)TTGCG(C/T)AA(C/T) as recognition sequence [158]. In contrast to
CBF/NF-Y they do not require an intact sequence of the CCAAT-motif. The
CAAT-box exists in both forward and reverse direction and a comparative
investigation of 96 genes containing 178 CAAT-boxes demonstrated that
CAAT-boxes are common in both TATA-containing and TATA-less promoters [121]. In TATA-containing promoters the CAAT-box is often located
at position -100/-80 and in TATA-less promoters it is located closer to the
transcription start site, at position -60, and often in a reversed orientation
[121].
18
Splicing
The pre-mRNA, produced by Pol II during transcription, consists of both
protein encoding and non-encoding ribonucleotide sequences, called exons
and introns, respectively. In the nucleus a large protein complex, the spliceosome, removes introns and ligates selected exons of the pre-mRNA in a
two step trans-esterification reaction, called splicing, creating an mRNA.
After transport into the cytoplasm the ribosome will read the code of the
mRNA and produce the described protein, in a process called translation.
Constitutive splicing
The consensus sequence of a splice site, ss, in mammals is defined as
AG/GURAGU at the 5’ ss and YAG/N at the 3’ ss, where / indicates the
boundary between the exon and the intron. In the intron, close to the 3’ ss,
there are two important sequence elements: the branch point, YNYURAC
(placed at 18-40 nt upstream of the 3’ ss), and the polypyrimidine p(Y) tract,
the which is a stretch of uridine residues of variable length (Figure 2). A
long p(Y) tract is efficiently recruiting the splicing factor U2AF in the early
steps of spliceosome assembly, resulting in that 3’ ss usage, and is therefore
referred to as a strong splicing signal. Two different models to define splice
sites exist, the exon definition model and the intron definition model, where
of the first model is the predominant in higher eukaryotes.
Alternative splicing
Approximately 30 to 50 % of the genes in higher eukaryotes are alternatively
spliced [131, 161]. The pattern of alternative splicing varies in different cell
types and tissues, during cell differentiation and embryonic development,
and under certain physiological conditions [112]. A pre-mRNA that can be
alternative spliced contains introns that are defined by alternative 5’ ss
and/or 3’ ss and this process can also lead to inclusion or exclusion of whole
exons. Alternative splicing is highly regulated and the SR family of splicing
factors plays an important role in defining the ss selection. Cis-elements in
the pre-mRNA, like exonic splicing enhancers, ESE, exonic splicing silencers, ESS, intronic splicing enhancers, ISE, and intronic splicing silencers,
ISS, are targeted by different trans-acting factors to activate or repress splice
signals in the surrounding sequence. A typical illustration of a regulated
alternative splicing event is when an SR protein by binding to an ESE recruits the U2 auxiliary factor, U2AF, to a 3’ ss with a weak p(Y) tract and
19
thereby causing the weak 3’ ss to be activated by helping in the recruitment
of U2 snRNP to the branch point.
Spliceosome assembly
The spliceosome is assembled in a sequential manner involving almost
200 proteins associated with U small nuclear RNA, U snRNA, forming nuclear ribonucleoprotein particles, snRNPs, or acting as splicing cofactors
with different functions. In the first step, formation of the early (E) complex,
U1-snRNP binds to the 5’ ss through RNA-RNA interaction, which is stabilized by a set of co-factors (Figure 2).
5’
pre-mRNA
Exon 1
GU
AG
3’
BP p(Y)tract
Intron
Exon 2
U1
SF1
E
U2AF 35
65
BP
U1
U2
A
U2AF
BP
U4
U1
U6
U5
U2
B
U2AF
BP
Exo
n1
U2
U6
C
BP
mRNA
Exon 1
Exon 2
Exon 2
Figure 2. A simplified overview of spliceosome assembly. The different complexes
are indicated on the left side. U1,U2, and U4/U5/U6 represent the UsnRNPs. Further, the SR proteins (not present in the picture) bind to regulatory elements on the
RNA and by different protein-protein interactions they bring the two ss together.
20
The branch point binding protein, SF1/mBBP, binds to the branch point
with a binding that is stabilized by an interaction with U2AF, which in turn
binds to the p(Y) tract. The following A-complex formation is characterized
by the ATP-dependent recruitment of the U2 snRNP to the branch point. The
binding of the U4, U5 and U6 snRNPs, together called the tri-snRNP, create
the B complex and is followed by a massive rearrangement of the spliceosome in which U6 replaces U1 at the 5’ ss and interacts with U2, U5
forms a bridge between the 5’ ss and the 3’ ss, and U1 and U4 become destabilized and leaves the complex. The new complex is called the C complex
and is the catalytically active spliceosome [reviewed in 66].
SR proteins
SR proteins regulate splicing by binding to splicing repressor or splicing
enhancer elements, which exist both in exons and introns and mediate protein-protein interaction with other SR proteins, SR-related proteins, and U
snRNPs. The SR protein induced formation of RNA-protein and proteinprotein interactions helps to bring the splice sites together and is essential for
the catalytic trans-esterification event to occur. SR proteins are binding very
early during spliceosome assembly and are required for E complex formation. They help in recruitment and stabilization of U1 snRNP and U2AF to
the 5’ ss and the 3’ ss, respectively, and play an important role in recruitment
of the U4/U6.U5 tri-snRNP.
Interestingly, Valcarcel et al. has demonstrated that the RS-domain of
U2AF65, a subunit of U2AF, interacts directly with the branch point, and
recently, Shen and Green confirmed these results by showing that the RSdomain of U2AF65 binds to the branchpoint in the E complex [reviewed in
66, 173, 174, 196]. They also demonstrated that the RS-domain of an ESE
bound SR protein interacts with the branch point in the following A complex, and the RS-domain of a second SR protein interacts with the 5’ ss in
the B complex [173, 174]. The function of SR protein can also be divided into
exon dependent (binding of ESEs) and exon independent (recruitment of the
U4/U6.U5 tri-snRNP, bridging between the 5’ ss and 3’ ss) activities. The
activity of SR proteins is antagonized by members in the heterogeneous nuclear ribonucleoprotein A/B, hnRNP A/B, family of proteins (see section
Proteins regulating SR protein activity).
The classical SR proteins consists of around ten phosphoproteins which
all have one or two N-terminal RNA recognition motifs, RRMs, and a
C-terminal domain rich in arginine and serine, RS, dipeptide repeats (Figure
3). The length of the RS domain varies between 24 and 316 residues and the
molecular weight of these proteins ranges from 20 to 75 kDa [reviewed in
58]. The RRMs consist of two conserved motifs, one octamer and one
hexamer, called RNP-1 and RNP-2 respectively, which form four antiparallel E-strands packed against two D-helices (E1-D1-E2-E3-D2-E4).
21
SRp75
RRM1 GR RRM2
SRp55
RRM1 GR RRM2
SRp54
RRM1
SRp46
RRM1
SRp40
RRM1
RS
RS
RS
RS
R
RRM2
ASF/SF2 RRM1 G
RRM2
9G8
RRM1 RP Z
SC35
RRM1 PG
SCp30c
RRM1 G
SRp20
RRM1 RP
RS
RS
RS
RS
RRM2
RS
RS
Figure 3. Human classical SR proteins. RRM, RNA recognition motif; RS, arginine/serine rich domain; Z, zinc knuckle; G, glycine, P, proline, and R, arginine, rich
domains. (Modified from [6])
The criteria for an SR protein are defined as:
x Containing one or two RRMs and an RS-domain
x Ability to restore full splicing activity in an extract deficient in
SR proteins, like cytoplasmic S100 extract
x Redundant roles in constitutive splicing
x Ability to regulate alternative splicing
The SR proteins are concentrated in 20-40 distinct sub-organellar domains called speckles in the nucleus and are translocated to transcription
sites upon gene activation [128]. Phosphorylation/dephosphorylation of the
SR proteins has been suggested to play an important role in this translocationr [26, 128]. ASF/SF2, as well as SRp20 and 9G8, are shuttling between the
nucleus and cytoplasm. The presence of the RS-domain has been shown to
be required and its phosphorylation status regulates the shuttling of the
SR proteins [27]. It has been suggested that the shuttling SR proteins in some
cases are helping mRNA export since they have the ability to associate with
the RNA export factor TAP [77].
SR proteins have essential splicing factor redundant functions. Individual
targeting of six different SR proteins with dsRNA interference, RNAi, in
Caenorhabditis elegans showed no effect on the phenotype, with the exception of CeASF/SF2, which downregulation caused lethality in late embry22
onic development. However, combinations of two or several simultaneous
SR protein deletions displayed defects or lethality [111]. SR proteins have
been identified in all metazoan species that have been analyzed, but they do
not exist in all eukaryotic species, like for example Saccharomyces cerevisiae which lacks a recognizable SR protein.
ASF/SF2
The SR protein ASF/SF2 was purified to almost homogeneity from HeLa
cells and was shown to be able to complement S100 extract in an in vitro
splicing assay [98]. At the same time it was identified as an activity regulating large and small T alternative splicing in 293 cells [53]. ASF/SF2 consists
of 248 amino acids and the molecular weight is approximately 30 kDa. It has
two RNA recognition motifs RRM1 (aa 1-98) and RRM2 (aa 106-201), and
a C-terminal RS-domain (aa 201-248) containing arginine-serine dipeptides
repeat in which the 20 serine residues are potential phosphorylation sites.
Phosphorylation of ASF/SF2 is required for its specific interaction with U170K and U1 snRNP [209].
Table 3. Human SR-related proteins [modification of tables21]
Protein name
U2AF35
Number Number of
of RRMs SR domains
1
Other domains
Function
1
U2 auxiliary factor
U2AF65
3
1
U2 auxiliary factor
U1-70K
1
2
snRNP component
U5-100K
1
U4/U5.U6-27K
1
hLuc7p
1
hTra2a
1
2
hTra2b
1
2
RSF1
1
SRrp40/SRp38
SRrp86
1
DEXD/H Box
snRNP component
snRNP component
Two Zn-domains
snRNP component
Splicing regulator
Splicing regulator
GRS-domain
Splicing regulator
1
K, E/D and Q/N in RSdomain
Splicing regulator
2
EK/R-rich region
Splicing regulator
SRm160
2
SRm300
1
RS/P
Splicing coactivator
hPrp16
1
DEXD/H Box
RNA helicase
HRH1
1
DEXD/H Box
RNA helicase
Clk/Sty
1
Kinase
Protein kinase
Splicing coactivator
23
SR-related proteins
SR-related proteins, SRrp, contain an RS-domain and sometimes also an
RRM, but do not fulfill the rest of the criteria required for an SR protein.
They can be divided in groups dependent on their function in RNA processing, chromatin association and transcription, but also their enzymatic activities. About 50 SR-related proteins have been defined in humans with a bioinformatic approach, while 80 SRrps was found in C. elegans and 110 in Drosophila [20]. SRrps involved in splicing are often interacting with the classical SR proteins and are typically involved in spliceosome assembly (U1-70k,
U2AF65, and U2AF35) or splicing regulation (like SRrp86, SRp38/SRrp40,
and SRm160/300) [10].
Phosphorylation and dephosphorylation of SR proteins
Phosphorylation activates SR proteins through modulation of protein-protein
interactions but also prevents SR proteins from non-specific protein-RNA
interaction [29, 188, 209, 210]. Both hypo- and hyperphosphorylation can reduce the activity of SR proteins [154]. Phosphorylation of SR proteins are
required for the translocation from the cytoplasm into the nucleus [85, 103]
and influences the intranuclear localization [37, 60].
SR proteins are phosphorylated by a number of kinases whereof the phosphorylation preformed by Clk/Sty, SRPK (SR protein kinase) family members, as well as DNA topoisomerase I, have been shown to regulate their
activity [7, 154, 155, 164, 203]. The Clk/Sty kinase has been shown to be a nuclear kinase in several cell lines [155], whereas SRPK is predominantly cytoplasmic in interphase cells [201]. The SRPK family, consisting of SRPK1 and
SRPK2, phosphorylates serines in the RS-domain preferentially in the context of the RSR motif [201]. The Clk/Sty family, which belongs to the
LAMMER kinases, consists of Clk/Sty, Clk/Sty-2, -3 and -4, and phosphorylates both serine/threonine (S/T) and tyrosine (Y) residues [13, 74, 104]. Further, Prasad and Manley have shown that Clk/Sty is autophosphorylated and
its phosphorylation status modulates its kinase activity in vitro in a substrate
specific manner [155].
Proteins regulating SR protein activity
To regulate the activity of SR proteins, cofactors make new contact between
the SR proteins and other splicing components and thereby change the ability of the SR protein to recognize its pre-mRNA target. These cofactors
regulate the function of SR proteins by acting as antagonizing factors or by
inhibiting their RNA binding either by sequestering the SR protein into new
complexes or affecting its phosphorylation pattern. SR protein phosphorylation is regulated by different mechanisms, for example by sterical hindrance,
24
recruitment of phosphatase or inhibition of kinase activity. Some of the regulatory proteins are RNA-binding proteins while other interacts directly with
the SR protein.
Table 4. Proteins regulating SR protein activity
Protein name
RNA
SR protein ininteraction teraction
Function
Ref.
-
Antagonist of ASF/SF2
[8, 28]
Cellular proteins
hnRNP A1
+
RBM4
+
RSF1
+
+
Antagonist of ASF/SF2
hTra2-D
+
+
Complex with SR pro[189]
tein to optimize splicing
hTra2-E
+
+
Complex with SR pro[189]
tein to optimize splicing
SRrp86
+
+
Inhibit or activate specific SR proteins
[10]
p32
-
+
Block phosphorylation
of ASF/SF2
[152]
Ad E4-ORF4
+
+
Inactivates SR proteins
by bringing PP2A
[44, 82]
HSV-1 ICP27
-
-
Inhibits SRPK1
[169]
Antagonist of SR protein [102]
[101]
Viral proteins
The hnRNP A1 protein has an antagonizing effect on ASF/SF2 function
by competing for RNA binding. ASF/SF2 promotes proximal 5’ ss selection
in vitro and in vivo, while hnRNP A1 activates distal 5’ ss usage [28]. Bai et
al. later demonstrated that hnRNP A1 also antogonize ASF/SF2 in 3’ ss
choice, by enhancing the usage of distal 3’ ss instead of the ASF/SF2 induced utilisation of the proximal 3’ ss [8]. Recently, another non-SR protein
named RNA-binding motif protein 4, RBM4, was shown to use the same
nuclear import pathway as SR proteins (Transportin-SR2 import factor
pathway) and further to act antagonistically against SR proteins in ss selection [102].
The human homologues of Drosophila Transformer 2, Tra2, are called
hTra2-D and hTra2-E and function as regulators of alternative splicing both
by binding to specific RNA sequences and interacting with SR proteins [189,
190]. The Tra2 protein forms a complex with Tra onto six repeats of an enhancer element called the doublesex repeat element, dsxRE, in the Drosophila double-sex, dsx, gene and recruits an SR protein for optimal splicing.
This SR protein has been shown to be the SRp20 homologue RBP1 in Drosophila and 9G8 in mammalias [115]. The current model is that Tra2 is alter25
ing the conformation and thereby the RNA recognition specificity of certain
SR proteins.
The Drosophila splicing repressor factor, RSF1, is an RNA binding protein which has been shown to antagonize ASF/SF2 in vitro. RSF1 contains
an N-terminal RRM and a C-terminal region rich in glycine (G), arginine
(R), and serine (S) amino acids, a GRS-domain. SFR1 inhibits an early step
in spliceosome assembly by inhibiting ASF/SF2 stabilization of U1 snRNP
binding to the 5’ ss by interacting via its GRS domain with the RS-domain
of ASF/SF2. Further, expression of RSF1 has been shown to rescue developmental defects caused by p55/SRp55 overexpression in Drosophila [101].
RSF1 has also been shown to compete in RNA binding with the SR proteins
[100].
SRp86 contains both an RRM and two separated regions rich in RSresidues, but can not complement S100 extract. It inhibits splicing activated
by SC35, ASF/SF2, and SRp55. Further, SRp86 can modestly increase splicing activated by SRp20 in vitro. Transient transfections have also shown that
SRp86 affects alternative splicing in vivo [10]. Another protein which affects
the activity of SR proteins is the herpes simplex virus ICP27. It binds to
SRPK1 and inhibits its kinase activity causing SR protein hypophosphorylation [169] (see PaperI).
The proteins described above are all able to affect the activity of SR proteins (Table 4). Two of the papers (Paper I and II) in this thesis add to this
list of reports by describing that adenovirus E4-ORF4 and the human protein
p32 regulate alternative splicing by interfering with SR protein activity [82,
152]. Both E4-ORF4 and p32 affect the phosphorylation of ASF/SF2, but
they utilize different strategies (Figure 4). E4-ORF4 associates with protein
phosphatase 2A, PP2A, which interacts with the RRMs of ASF/SF2, resulting in hypophosphorylation and inactivation of ASF/SF2 as both enhancer
and repressor of splicing (Paper I and [44]). p32 blocks phosphorylation of
ASF/SF2 through sterical hinderance by binding to the RRMs of ASF/SF2,
leading to ASF/SF2 hypophosphorylation and loss of the RNA binding capacity (Paper II and paper III).
26
5’
ASF/SF2
3’
BP
Exon 1
Intron
Exon 2
3RE
E4-ORF4
p32
5’
PP2A
3’
3’
Figure 4. p32 and E4-ORF4 regulate the RNA binding capacity of ASF/SF2 by
different mechanisms. p32 block phosphorylation and sequesters ASF/SF2 into an
inactive complex, while E4-ORF4 recruits PP2A to dephosphorylate ASF/SF2.
p32
The ubiquitous cellular protein p32/HABP1/gC1q-R is a multifunctional
protein that is localized at the cell surface, in the mitochondria, cytoplasm
and nucleus of various cell types. p32 has been reported to be involved in
several activities like regulation of transcription and splicing, and receptor
binding. It was originally isolated tightly associated with ASF/SF2 during
purification from HeLa cells [98]. Deb and Datta have demonstrated that the
34-kDa hyaluronic acid-binding protein, HABP1, is identical to the ASF/SF2
binding protein p32 [39] and Ghebrehiwet et al. cloned p32 as a 33 kDa surface glycoprotein, gC1q-R, which interacts with the “globular heads” of C1q
[55]. It has also been shown that the lamin B receptor-associated protein p34
is identical to p32 [179]. The p32 gene is positioned on human chromosome
17p13.3 and on mouse chromosome 11. p32 is translated into a pre-protein
27
of 282 aa, in which the N-terminal sequence has the characteristic features of
a mitochondrial import sequence and thus the full length protein is mainly
localized in mitochondria. The mature form of p32 lacks the 73 N-terminal
aa and is found both in the cytoplasm and in the nucleus [73, 125, 202]. Amino
acids 222-282 in p32 are highly conserved in several species, ranging from
yeast to human [73, 136]. The mature protein is highly acidic with a calculated
pI of 4.15.
Structure of p32
Crystallization of p32 reveals a non-covalently tight associated trimer, forming a dough-nut shape [80]. p32 can exist as a monomer, trimer or even as a
covalently-linked hexamer, which is linked by thiol group oxidation of
Cys186, under different environmental conditions [79]. The topology of the
monomer reveals an N-terminal D-helix (A) followed by seven twisted antiparallel E–sheets, a short D-helix (B) and a c-terminal D-helix (C) (Figure 5).
Two loops connecting E1 to E2 and E3 to E4, in monomer I interact with the
connecting loop between E6 and E7 of the monomer II. The D-helices A and
C in monomer I also interact with the E-sheet and the D-helix B in monomer II [80]. When binding to hyaluronic acid p32 forms a homodimer of 68
kDa [62].
233
244
250
96
225 213 180 174
127 124
227 205 187 168 134
280
C
77
N
110
117 114
Figure 5.The secondary structure of p32. (Modified from [80]). Open thick line
marks the loop within the region of amino acid 226-239 demonstrated to be important in ASF/SF2 interaction (see Paper III), while black thick lines marks the loops
involved in trimer formation. Grey boxes represent D-helixes and white open arrows
represent E-sheets.
28
p32 protein interactions
The p32 protein binds to many different cellular, bacterial and viral proteins,
probably depending on its maturation state, quaternary structure and localization. Viral proteins like Ad pV, HSV-1 ICP27, HVS ORF73, gammaHV68 M2 and EBV EBNA-1 induce an accumulation of the mature p32
protein in the nucleus [25, 63, 106, 125, 202].
Table 5. Proteins interacting with p32
Protein name
aa of p32 interacting with
indicated protein
Reference
74-239
[98, 152], paper III
Cellular proteins
ASF/SF2
CBF/NF-Y
Not known
[33]
Cdc25
1-104
[118]
Factor XII
Not known
[81]
Fibrillanin
Not known
[213]
gC1q
76-94
[55]
high molecular weight kininogen
Not known
[67]
Hrk
221-282
[186]
Hyaluronic acid
Not known
[39]
lamin B receptor
Not known
[142]
protein kinase C
Not known
[162]
TFIIB
244-255, 260-279
[216]
vitronectin
74-96
[108]
Not known
[125]
Viral proteins
Ad pV
CMV pUL97
214-282, (243)
[124]
EBV EBNA-1
103-282, (244-282)
[202]
gamma-HV68 M2
Not known
[106]
Hepatitis C virus core protein
Not known
[90]
HIV-1 Rev
196-208
[191]
HIV-1 Tat
244-255
[17, 216, 217]
HSV-1 ICP27
Not known
[25]
HSV-1 ORF P
Not known
[24]
HVS ORF73
Not known
[63]
Rubella virus core protein
214-282
[11, 12, 132]
Not known
[23]
Not known
[140]
Bacterial proteins
Listeria monocytogenes protein
In1B
Staphylococcus aureus protein A
29
The interaction with the SR protein ASF/SF2 has been shown to play a regulatory role in splicing [152], while some p32 protein interactions, for example
with TFIIB and HIV-1 Tat, EBV EBNA-1, HSV ORF73 or gamma-HV68
M2, have indicated a role for p32 in transcription activation [63, 106, 200, 216].
Recently, Chattopadhyay et al. described that p32 can repress transcription
by interacting with the B-subunit of the transcription factor CBF/NF-Y
bound to the CAAT-box in the promoter.
It appears that the proteins interacting with p32 bind to different domains
of p32. As shown in Table 5 the N-terminal D-helix A or the C-terminal
D-helixes B and C of p32 are important for the different interactions, with an
exception of HIV-1 REV which interacts with E-sheet 6 [191]. In this thesis
we demonstrate that ASF/SF2 can interact in vitro with p32 lacking the
C-terminal D-helices (p32aa 74-239) but this interaction is lost when further
deletions are made (p32aa74-226) (see section Present Investigation and
Discussion; paper III).
Interestingly, it has been reported that many of the p32 interacting protein
are rich in arginine residues in their interacting regions like HSV-1 ICP27
[25], HVS ORF73 [63], HIV-1 REV [114], Lamin B receptor (p58) [142],
EBV EBNA-1 [202] and Rubella virus core protein [11].
Table 6. p32 homologues and their interacting proteins
Homologues PreMature origin
protein protein
Reported
interacting
protein
Interacting
domain of p32
homologues
p38
207 aa
chicken
myosin
1-28 and 187-207 [146]
p22
227 aa
p30/Mam33p 266 aa
SUAPRGA1 303 aa
YL2
181 aa
230 aa
T. Brucei
RBP16
S. cerevisiae Cytochrome b2
A. Nidulans
murine
HIV-1 REV
208 aa
Ref.
[65]
[170]
[198]
[114]
The p32 protein has homologues in all mammals and lower organisms investigated (Table 6). In S. cerevisiae the p30/Mam33p has a 53 % similarity
and 26 % overall identity with p32. This protein was identified as a mitochondrial matrix protein binding to cytochrome b2 [170]. The murine protein
YL2 shows 92 % identity with p32 and binds also to HIV-1 REV like human
p32 does [114, 191]. Further, a mitochondrial guide-RNA binding protein
named p22 in Trypanosome brucei has been identified as a homologue of
p32 and p30/Mam33p [65].
30
p32 and protein modifications
A sequence analysis of p32 reveals three potential N-glycosylation sites
(114N, 136N and 223N), one tyrosine sulfation site (188Y), one potential myristylation site (250N) and one tyrosine recognition site (268Y) [39, 54]. p32 has also
several potential phosphorylation sites for protein kinases like: protein
kinase C (205S), casein kinase II (76T, 205S, 213S, 251T and 261T), and protein
kinase ERK and cdc2 (160PELTSTP166) [39].
Several kinases can phosphorylate p32 under specific circumstances. Recently, p32 was reported to interact with the human cytomegalovirus (CMV)
protein kinase pUL97. The p32-pUL97 complex is recruited to the
lamin B receptor at the nuclear membrane where p32 and lamin B becomes
phosphorylated by pUL97 leading to a redistribution of the lamina and increased release of virus particles [124]. Furthermore, Hyaluronic acid (HA),
PMA, calyculin A and Ca2+ ionophore cause an enhanced phosphorylation
of threonine residues in p32 and MAP kinase phosphorylation of p32 causes
a translocation of p32 into the nucleus [117]. Co-immunoprecipitation studies
have demonstrated that p32 is phosphorylated by CK2 when present in a
complex consisting of p32-hnRNP K-CK2. Formation of this trimeric complex requires HSV-1 ICP27 (IE63) [25].
The only reported kinase whose activity is affected by p32 is protein
kinase CP, PKCP. p32 binds to the kinase domain of PKCP and inhibits
PKCP phosphorylation of the substrate aldolase. In contrast, p32 enhances
the autophosphorylation activity of PKCP [184].
31
Adenovirus
Adenoviruses have been isolated from every vertebrate species that have
been investigated and the classification of the family Adenoviridae consists
of the two genera Mastadenoviridae and Aviadenoviridae [14, 165]. The human adenoviruses (belonging to Mastadenoviridae) are divided into six subgroups, A-F, based on their ability to agglutinate red blood cells [68] and
their oncogenic potential in rodent cells [175]. The most studied adenoviruses
of the approximately 50 human serotypes that exist are the adenovirus type 2
and 5, Ad2 and Ad5, which both belong to subgroup C, and type 12, Ad12,
which belongs to subgroup A. Ad 12 is highly oncogenic in rodents but no
human adenovirus has so far been shown to cause any human cancers [195].
Adenovirus has a non-enveloped icosahedral capsid of 70-90 nm containing a double stranded DNA molecule of 30-38 kbp. Ad2 and Ad5 have genomes of approximately 35 kbp. The genome is divided into transcriptional
units in which the pre-mRNAs are extensively alternatively spliced to generate mRNAs that are translated into approximately 30-40 proteins dependent
on the serotype.
Virus life cycle
Dependent on the serotype human adenoviruses attach to the cell via either the cellular coxsackie and adenovirus receptor (CAR) or the MHC
class I receptor through interaction with the viral fiber extending out from
the virion [16, 72, 194]. Ad2 and Ad5 bind to the CAR protein. Two secondary
receptors, integrin DvE3 and DvE5, are helping the virus to enter the cell by
receptor-mediated endocytosis via formation of a coated pit that fuses to
endosomes [205]. In the cytosol, the virion is disassembled in the endosome
due to a decrease in pH. After transport to the nucleus, along the microtubuli, the membrane of the vacuole is disrupted and the DNA along with the
DNA-associated protein VII enter through the nuclear pore complex into the
nucleus, where the viral DNA can be transcribed and replicated.
The genome contains early (E1A, E1B, E2, E3 and E4), intermediate (IX
and IVa2) and late (MLTU) protein encoding genes and two genes for production of the highly structured VA RNA I and II (Figure 6). The first viral
gene to be transcribed is E1A which occurs approximately 1 hour after viral
attachment to the cell surface receptor. E1A activates transcription of all the
early viral and some cellular genes leading to replication of the viral genome
at approximately 6-8 hour post infection, hpi. The viral DNA replication
separates the early and the late phase of infection.
32
Figure 6. The genome of Adenovirus 2 and 5. Kindly provided by Göran Akusjärvi,
IMBIM, Uppsala University. Black arrows indicated early genes and open arrows
indicate late genes.
The major late transcription unit, MLTU, is activated in the late phase to
produce mainly the viral structural proteins. During the late phase of infection adenoviral proteins are almost exclusively produced although the transcription rate of cellular genes is not reduced. The early viral proteins
E1B-55K and E4-ORF6 are responsible for a selective transport of viral
mRNAs from nucleus to cytoplasm. During infection the cap-binding subunit eIF4E of the cellular translation initiation factor eIF4F becomes
dephosphorylated and its reduced activity causes an inhibition of cellular
translation. The viral mRNAs expressed from the MLTU can still be translated because they all contain a 5’ leader sequence, the tripartite leader,
which functions as an eIF4F-independent translational enhancer.
The VA RNA I and II are highly structured RNA molecules that are important primarily in protecting the virus against the interferone induced cellular immune system by inactivating the cellular kinase PKR. As an antiviral
defense, the cell tries to shut down the translation machinery by activating
the cellular kinase PKR, which inhibits the translational initiation factor
33
eIF2a by phosphorylation. To be activated PKR binds as dimers onto the
dsRNA produced by symmetrical transcription of the viral genome. The
short viral VA RNA I and II can keep PKR in an inactive form since PKR
binds VA RNA as monomers.
The infections cycle is complete after approximately 30 h at optimal conditions. After virus assembly the cell lyses due to disruption of the cellular
cytoskeleton and 104-105 new virus particles per cell are released.
Early genes
The proteins from the early regions are involved in different mechanisms to
drive the infection towards production of new viral particles. This is
achieved by:
x Forcing the cell into S-phase
x Production of proteins needed for replication of viral DNA
x Production of proteins protecting the infected cell from the host cell antiviral response.
The E1A transcript, produced from a constitutive promoter of the E1Aregion, can be alternative spliced into five different mRNAs; 13S and 12S,
which are produced early after infection and the 11S, 10S, and 9S, which are
produced late in infection. The shift in E1A pre-mRNA splicing is caused by
the sequestering of the SR proteins needed for E1A 13S and 12S splicing. It
is believed that the massive amount of transcripts generated from the MLTU
at late times of infection titrates out the limited amounts of SR proteins present in the cell. The E1A functions as a transcription activator of regulating
both viral and cellular genes. The E1A protein does not bind DNA directly
and is targeted to promoters through the interaction with transcription factors
bound to target promoters. By removing and inhibiting HATs, like
p300/CBP from a promoter E1A can also repress transcription. To activate
genes E1A can either phosphorylate transcription factors or bridge between
transcriptional activators and the basal transcription machinery. Furthermore,
E1A can activate transcription by displacing inhibitory corepressor complexes like binding to tumour suppressor protein RB which releases the transcription factor E2F to activate S phase genes. To summarize, the most important function of E1A is to force the cell into S-phase to allow for viral
replication.
E1A causes increased amount of p53 in the cell, which can lead to apoptosis [151]. The two proteins encoded by the E1B-region, E1B-55K and
E1B-19K, prevent from apoptosis. The E1B-55K protein can bind directly to
the p53 transcriptional activation domain and repress p53 transcriptional
activation [215]. Further, E1B-55K can form a complex with E4-ORF6 and
both proteins can inactivate p53 by targeting it for proteosomal degradation
[41, 156]. The E1B-19K protein, a homolog to the cellular anti-apoptotic
34
Bcl-2 protein, sequesters the pro-apoptotic factors Bax and Bak by forming
heterodimers and prevents them from activating caspase 3 and 9 [204].
The E2 transcription unit consists of two units: E2A, encoding E2A-72K
(also called the DNA binding protein, DBP), and E2B, encoding the precursor terminal protein pTP and the viral DNA polymerase [207]. The E2 region
has two promoters, an early promoter, which is active early in infection and
a late promoter that is activated at the intermediate phase of infection. E1A
activates the early promoter, while the late promoter is repressed by E1A
[61]. The late promoter is activated by the CAAT-box binding protein YB-1,
which is recruited to the cell nucleus by the viral E1B-55K protein [71].
The proteins encoded from the E3-region are involved in counteracting
the host immune response and are therefore dispensable in in vitro cell cultures [reviewed in 107]. Virus infections often induce a release of TNF-D and
Fas causing induction of apoptosis. The E3-14.7 kD and the E3-10.4K/14.5K
proteins can inhibit apoptosis and reduce the amount of Fas receptor by internalization [43]. The E3-19K protein can in turn reduce the amount of
MHC I receptor on the cell surface rescuing cells from death by cytotoxic Tlymphocytes [159]. The adenovirus death protein (ADP) E3-11.6K promotes
virus release by disrupting the cytoskeleton of the cell [192, 193].
The E4-region encodes for at least seven proteins which all have different
functions during the life cycle of the virus. As mentioned above, E4-ORF6
binds to E1B-55K and acts anti-apoptotic, and this complex are also responsible for the selective transport of viral mRNAs into cytoplasm. E4-ORF3
and E4-ORF6 has showed to have opposite affect on alternative splicing,
where E4-ORF3 facilitates i-leader exon inclusion, while E4-ORF6 preferentially favours i-leader exon skipping of the tripartite leader sequence of
late transcripts from MLTU [145]. E4-ORF3 and E4-ORF6 can both inhibit
cellular double-strand break repair system and the purpose seems to be to
prevent the adenoviral genome from forming concatemers [22, 45].
E4-ORF6/7 acts as a transcription co-activator by stabilizing and promoting
dimerization of the E2F transcription factor on the E2 early promoter [187].
E4-ORF4 protein
The E4-ORF4 protein is a multifunctional protein that can affect different
cellular processes like transcription and splicing. It can block E1A induced
transcription [19, 91, 135], induce p53-independent apoptosis [122, 123, 176, 177],
and cause G2/M arrest in yeast and mammalian cells [96]. In paper I we demonstrate that E4-ORF4 causes SR protein dephosphorylation during an adenovirus infection, reducing the SR protein RNA binding capacity and affecting alternative splicing [82]. It has been demonstrated that E4-ORF4 binds to
the B-subunit of serine/threonine phosphatase PP2A [44, 91] and the current
model is that E4-ORF4 performs different functions in the cell by bringing
PP2A to dephosphorylate different target proteins. Interestingly, other viral
proteins have been reported to interact with PP2A, like SV40 small T,
35
polyoma virus small T and middle T, and HIV-1 Vpr [75, 150, 197]. It is assumed that different viruses use the phosphatase to target the same cellular
key proteins controlling cellular growth [48].
The Major Late Transcription Unit
The late genes of adenovirus have a common promoter, the major late promoter, MLP, which produces one long capped pre-mRNA transcript of about
28,000 nucleotides. The late transcript can be poly-adenylated at five different poly-adenylation sites and the lates genes are therefore divided into five
3’-coterminal units, L1-L5 (Figure 6). Within each unit different mRNAs are
produced by alternative splicing. The late mRNAs encode primarily for the
late structural proteins required for capsid formation but also a few regulatory proteins [171]. All late transcripts receive through alternative splicing a
common 5’ leader sequence, the tripartite leader sequence, which functions
as a translational enhancer in late infected cells. The 201 nt long tripartite
leader is made out of three exons: 1, 2, and 3, with one exception. The L1
transcript: 52,55K+i mRNA has an extra i-leader exon inserted between
exons 2 and 3. The 52,55K and the i-leader protein are the only proteins
produced from the MLTU during the early to intermediate phase of infection.
The Major Late Promoter
The major late promoter has both a TATA-box, 5’-TATAAAAG-3’ at
position -31 to -24, and an initiator element, INR, 5’-TCACTCT-3’ at position -2 to +5 [180]. Several proteins have been reported to bind to the INR of
MLP, like the different TBP-associated factors, TAFIIS, [86], but mutational
studies have revealed that the INR of MLP is not necessary for activity in
vivo unless another MLP element is disrupted at the same time [113].
The transcription factors DEF-A and DEF-B bind to three different elements within the DNA encoding the first exon of the tripartite leader; DE1,
DE2b and DE2a [78]. There they form homo- or heterodimers with the adenovirus IVa2 protein. Further, two important upstream activating elements
have also been identified, the upstream promoter element, UPE,
5’-GCCACGTGA-3’ at position -62 to -53, [59, 167] and a reversed CAATbox, 5’-TGATTGGTTT-3’ at position -82 to -73 [116]. The transcription
factor binding to the UPE-motif was simultaneously identified by three different groups, which resulted in the names: MLTF (major late transcription
factor) [30] UEF (upstream element factor) [130], and USF (upstream stimulatory factor) [167]. The heterotrimeric transcription factor CBF/NF-Y binds to
the CAAT-box of MLP and stimulates MLP transcription in vitro [116]. Although the mechanism is unknown it is presumed that CBF/NF-Y must interact with one or more proteins in the preinitiation complex of the MLP
[182]. The TATA-box and the CAAT-box are absolutely conserved in differ36
ent serotypes in contrast to the UPE motif, the downstream elements and the
INR [182]. Analysis of the different promoter elements in the MLP has demonstrated that mutation of the UPE alone does not cause any severe effects
on transcription from the MLP [160] and the same is noticed when a single
nucleotide mutation is introduced into the CAAT-box. The two upstream
activating elements are redundant, but the combination of these mutations
causes the transcription from the MLP to decrease fifteen-fold [160]. An introduction of five nucleotide mutations either in the CAAT-box or in
neighboring sequence together with the UPE mutant has an even more severe effect on viral replication [182].
The L1-region
The L1 pre-mRNA is alternatively spliced during the viral life cycle, resulting in the accumulation of different mRNAs of the early and late phase of
infection. A common 5’ ss is joined to either the proximal 3’ ss, giving raise
to the 52,55K mRNA, or to the distal 3’ ss, creating the IIIa mRNA. The
52/55K protein is a scaffolding protein whereas the IIIa protein is a capsid
protein. The exclusive L1 mRNA produced early in infection is the 52,55K
mRNA due to a repression of the distal 3’ ss by SR proteins binding to an
intronic repressor element, called the IIIa repressor element, the 3RE [83]. At
late times of infection the IIIa mRNA becomes the predominant L1 mRNA
expressed from the L1 unit. As we describe in Paper I, SR proteins become
hypophosphorylated in late-infected cells resulting in a weakning of their
RNA binding capacities [82]. The release of SR proteins from the 3RE element is important to induce the switch towards the distal IIIa 3’ splice site
usage. Subsequently, it was shown that the IIIa branch point/pyrimidine tract
acts as an inhibitory element in uninfected extract and as an enhancer element in infected extract. This element, which is the major element controlling IIIa 3’ ss activation, was named the IIIa virus infection-dependent splicing enhancer, 3VDE [137].
The L2-, L3-, L4- and L5-regions
The L2-region encodes for five proteins where the core proteins VII and V
are histone-like proteins that bind the viral DNA to form a structural core of
the virus. pX is cleaved into two structural core proteins: X and P. Polypeptide III makes up a pentamer, called the penton base, which together with the
fiber are assembled into the structural component penton [199]. Twelve fiber
molecules are protruding out from the viral capsid and each are made up of a
homo-trimer of polypeptide IV from the L5-region.
The L3-region encodes for capsid protein VI and the homotrimeric hexon,
II, which is the major coat protein. Each penton base is surrounded by five
hexon proteins and polypeptide IX encoded from an intermediate gene function as a glue protein stabilizing the hexon proteins. The L3-region also en37
codes for the viral protease. Approximately 10-30 molecules of the L3 protease are included into each virus particle, which are needed for the uncoating of the virus. The viral protease is also active during virus maturation
when it cleaves most of the minor capsid and the core polypeptides, including the terminal protein. Binding to DNA or an 11 aa peptide cleaved from
pVI increases its activity thousand-fold [40].
One of the first proteins to be produced in the late phase of infection is the
L4-100K protein. The L4-100K protein is needed for an efficient production
of viral particles and one of its functions is to assemble the hexon protein
into trimers in the cytoplasm [31]. It is tightly bound to mRNA [3] and a temperature sensitive L4-100K mutant displays reduced translation of viral
mRNAs [64]. Studies have shown that the L4-100K protein stimulates translation of tripartite leader containing mRNAs by functionally inactivating the
cellular cap-binding translation initiation factor eIF4F [42].
Viral factors, like E4-ORF4, are needed to affect alternative splicing
within L1 and to trigger the shift towards production of proteins from the
L2-L5 units. Recently, it was shown that a transfected MLP construct only
containing L1-L3 was unable to shift from early to late protein expression.
Addition of L4-33K in trans resulted in expression of L1-IIIa, L2 and L3
proteins suggesting that L4-33K might be important for the early-late protein
expression shift from the MLTU during infection [46].
Table 7. Proteins present in the adenovirus particle
Protein
Transcription unit
Copy number
TP, terminal protein
P
V
VII
X
L3 protease
Capsid element
E2B
L2
L2
L2
L2
L3
2
~104
Not known
Not known
Not known
~30-40
IIIa
VI
VIII
IX
Capsid
L1
L3
L4
intermediate
60 monomers
60 hexamers
Not known
80 trimers
II, hexon
III, penton base
IV, fiber
L3
L2
L5
240 trimers
12 pentamers
12 trimers
Core
38
Present Investigation and discussion
Paper I
Regulation of adenovirus alternative RNA splicing by
dephosphorylation of SR proteins
In paper I we found a difference in splicing activity of SR proteins from
uninfected and adenovirus infected cells and by further investigation we
identified the adenoviral protein E4-ORF4 as a protein that can induce
SR protein dephosphorylation.
Previously, Kanopka et al. demonstrated that SR proteins like ASF/SF2
can repress splicing of the L1 IIIa pre-mRNA in an in vitro splicing assay.
ASF/SF2 binds to an intronic repressor element, named 3RE, situated just
upstream of the IIIa branch site, and blocks the formation of the A-complex.
By moving the element it was shown that SR proteins could act as repressors
or enhancers of splicing depending on where on the pre-mRNA the
SR protein binding site was situated [83]. To extend the study on how SR
proteins regulate splicing we compared the properties of SR proteins purified
from HeLa nuclear extract, SR-HeLa, with SR proteins from adenovirus
infected nuclear extract, SR-Ad. Interestingly, the SR-Ad proteins were inactive in in vitro splicing, both as splicing enhancer and splicing repressor
proteins. Using an UV cross-linking assay we could also demonstrate that
the SR-Ad proteins had a reduced RNA binding capacity.
Normally, in in vivo splicing, SR proteins are highly phosphorylated,
while SR proteins purified from adenovirus infected nuclear extract were
shown to be hypophosphorylated, with the exception of SRp20. This was
demonstrated by two-dimensional gel electrophoresis of purified SR proteins
which had been 32P-labelled in vivo.
A previous study had demonstrated that during infection with an adenovirus mutant lacking a functional E4-ORF4 protein hyperphosphorylated
forms of c-Fos and E1A proteins are accumulating, by an unknown mechanism [135]. Soon after, E4-ORF4 was shown to interact with the B subunit of
serine and threonine specific protein phosphatase 2A, PP2A, and the E4ORF4/PP2A complex possessed phosphatase activity which seemed to have
a role in transcription regulation [19, 91]. In Paper I we demonstrate that a
recombinant E4-ORF4 protein is able to induce SR-HeLa dephosphorylation
when incubated in HeLa-NE. Addition of okadaic acid, which inhibits PP2A,
39
resulted in a loss of E4-ORF4 induced SR protein dephosphorylation. Normally, addition of SR-HeLa proteins to an in vitro splicing reaction programmed with the wild type IIIa transcript causes an inhibition of splicing.
Addition of recombinant E4-ORF4 protein rescued IIIa splicing illustrating
that dephosphorylation of SR-HeLa proteins reverses the SR protein mediated repression of IIIa splicing and that incubation with E4-ORF4 gives SRHeLa the properties of SR-Ad. Transient transfection experiments were utilized to investigate the effect of E4-ORF4 in vivo. Comparing different the
IIIa constructs with or without the 3RE confirmed the result that E4-ORF4
can activate a repressed splicing also under in vivo conditions. When cotransfecting E4-ORF4 with a mini-construct containing both the 52,55K and
the IIIa 3’ ss with the 3RE element a shift towards IIIa splicing was seen.
This demonstrates that E4-ORF4 can reverse the SR protein mediated block
in IIIa 3’ ss also during alternative splicing conditions. This paper was the
first to show that alternative splicing is regulated by reversible protein phosphorylation.
In an extension, Estmer Nilsson et al. demonstrated that E4-ORF4 binds
both ASF/SF2 and SRp30c, but not the other SR proteins. E4-ORF4 binds
preferentially to the hyperphosphorylated form of the SR proteins and interacts with ASF/SF2 through its RRMs [44]. In Paper I adenovirus infection
caused a hypophosphorylation of all the classical SR proteins, with the possible exception of SRp20. In conclusion, some other viral factor, besides E4ORF4, is also needed for the SR protein hypophosphorylation.
Other viral proteins affecting the activity of SR proteins
Some viruses utilize alternative splicing to compress more information into
their genomes. It ought to be of interest for such viruses to have the ability to
regulate splicing by for example controlling the activity of the SR protein
family of splicing factors. Viral proteins known to affect the activity of
SR proteins are listed in Table 4.
The herpes simplex virus ICP27 protein has been shown to bind SRPK1
and thereby cause a translocation of SRPK1 into the nucleus. ICP27 inhibits
the activity of SRPK1 and SR proteins purified from HSV infected cells are
hypophosphorylated [169]. Why is HSV-1 inactivating the SR family of
splicing factors? It seems that the purpose for HSV-1 is somewhat different
from the adenovirus induced SR protein dephosphorylation since HSV-1
contains very few introns. Another example of a virus that inhibits
SR protein phosphorylation is the vaccinia virus [76]. The precise mechanism
is unknown, but interestingly vaccinia virus genes do not have any introns.
Both vaccinia and HSV-1 might cause hypophosphorylation of SR proteins
as a strategy to inhibit the post-transcriptional processing of cellular transcript and thereby reduce the production of cellular proteins. In the case of
adenovirus it seems that SR protein dephosphorylation gives the virus the
40
ability to regulate the shift in alternative splicing necessary to produce the
viral proteins needed late in infection.
Paper II
p32 regulates RNA splicing by inhibiting ASF/SF2 RNA binding
and phosphorylation
No splicing activity was assigned to p32 in the first study when p32 was
identified as a ASF/SF2 associated protein in HeLa cells [98]. In paper II we
therefore decided to investigate whether p32 had a regulatory role on
ASF/SF2 function rather than being a bona fide splicing factor itself. The
important finding from this paper was that p32 indeed has a function in
splicing regulation by sequestering ASF/SF2 into an inactive complex.
p32 inactivates SR proteins as splicing regulators
As expected, an in vitro splicing assay showed that splicing of a IIIa+3RE
transcript in HeLa-NE was repressed due to the binding of SR proteins to the
repressor element 3RE. Interestingly, addition of increasing amounts of recombinant p32 alleviated the repressive effect of SR protein on IIIa splicing.
As a control for the specificity of the splicing enhancer effect of p32, another
acidic protein was used, the glucose oxidase, GOD, showing no effect on the
spliced products. Complex assembly assays showed that p32 acts at an early
step of spliceosome assembly by enhancing the A complex formation. The
3RE element was then transferred to the intron of a E-globin transcript and
the same splicing activation was noticed in the presence of p32. Previous
experiments have shown that the position of an SR protein binding site in a
pre-mRNA determines if the element acts as enhancer or a repressor element
[83]. The transfer of 3RE from an intronic position to the 3’ end of a 52,55K
transcript transformed the 3RE from a repressor element to an enhancer element of splicing [218]. In such a transcript addition of p32 repressed splicing
by blocking A-complex formation. Altogether, these results demonstrated
that p32 inactivates SR proteins both as repressor and enhancer proteins of
splicing.
To test if p32 could affect splicing under in vivo conditions we utilized
transient transfection with plasmids expressing the p32 protein and the IIIa
transcripts in HeLa cells. As expected, p32 activated IIIa 3’ ss usage. The
in vitro and the in vivo data somewhat differed in the aspect that large
amounts of p32 inhibited splicing. There is a possibility of aggregation of
p32 in the cell due to overexpression of the protein, which could lead to p32
loosing its regulatory role in splicing. However, this is probably not the explanation, since p32 also inhibited IIIa-3RE when expressed in excess. More
likely, p32 is sequestering ASF/SF2 in the cytoplasm and thereby disturbing
41
the normal balance of ASF/SF2 between the cytoplasm and the nucleus.
Potentially, p32 may also affect the sub-cellular distribution of other SR
proteins as well.
p32 inactivates ASF/SF2 as a splicing regulator by blocking its RNA
binding
Further, our in vitro splicing results revealed that when the 3RE was exchanged with a consensus ASF/SF2 binding site in a IIIa transcript, p32
could still activate splicing. The addition of recombinant ASF/SF2 blocked
A complex formation on a IIIa 3’ ss transcript, and p32 was able to release
that repression. Krainer et al. have demonstrated that ASF/SF2 can influence
alternative splicing by causing a shift from distal to proximal 5’ ss usage in a
E-globin construct [97]. In agreement with their result ASF/SF2 promoted the
proximal 5’ ss in our in vitro splicing assay, and importantly p32 had the
opposite effect causing a shift from proximal towards distal 5’ ss usage. It
should be noted that p32 affects splicing not only on adenoviral transcripts,
but also on a cellular transcript indicating that it has a general role in splicing
regulation.
p32 does not bind RNA and therefore our hypothesis was that p32 directly regulated the RNA binding capacity of ASF/SF2, rather than competing for its RNA binding site like an antagonistic protein like hnRNP A1. To
test this hypothesis we used an UV cross-linking assay. As predicted, addition of p32 caused a reduction in ASF/SF2 binding to RNA. By including
hnRNP A1 in the reaction we could demonstrate that RNA was now available for the antagonist protein to bind. To demonstrate the specificity of
p32’s action the acidic protein GOD was used as a control and it did not
have any effect on the RNA binding capacity of ASF/SF2 in the UV crosslinking assay. In paper III we show additional results that p32 can reduce the
RNA binding capacity of ASF/SF2 by using two additional constructs,
IIIa+3RE and E-globin+3RE, in the UV cross-linking assay.
Furthermore, in Paper II we used a recombinant ASF/SF2 protein lacking
the RS-domain, ASF/SF2-'RS, in the UV cross-linking assay. Interestingly,
p32 could also block the RNA binding capacity of ASF/SF2-'RS. Based on
these observations we draw the conclusion that p32, at least in part, binds to
the RNA-binding domain of ASF/SF2. This result is in contrast with the
speculation of Yu et al. implying that the interaction of p32 with ASF/SF2 is
due to ionic interactions between the highly basic RS domain in ASF/SF2
and the acidic C-terminus in p32 [216, 217]. In a later study (Paper III) we
demonstrate that p32 in fact binds stronger to ASF/SF2 lacking the RS domain than full-length ASF/SF2, confirming that p32 interacts mainly with
the RRMs of ASF/SF2.
42
Notably, the homologue of p32 in T.brucei, called p22, has been shown to
stimulate the RNA binding capacity of the guide RNA (gRNA)-binding protein RBP16 [65].
Recently, Zheng et al. published a paper supporting our findings that p32
inhibits the function of ASF/SF2 [220]. Murine cells are unable to support
HIV-1 replication. The unspliced 9kb HIV-1 transcript functions both as
HIV-1 genome that should be incorporated into new viral particles and as an
mRNA encoding the structural Gag-protein and the polymerase Pol. In murine cells, the 9 kb genomic transcript was excessively spliced and the amount
of viral genomic and unspliced mRNA reduced, causing a reduction of Gagprotein and blockage of viral assembly. Introduction of human chromosome
11 in these transgenic cells had in a previous report been shown to complement these defects [178]. ASF/SF2 is known to enhance HIV-1 splicing by
binding to ESEs in the exons of Tat [153, 191]. Given that human p32 is located on chromosome 11 and that we had shown that human p32 could inhibit ASF/SF2 as a splicing factor [152], Zheng et al. tested if human p32
could rescue HIV-1 infection. Indeed, p32 inhibited ASF/SF2 and prevented
the excessive splicing of the HIV-1 genome, which relieved the posttranscriptional block and allowed HIV-1 to replicate in the mouse cells. The
mouse p32 (called mp32 or YL2) and human p32 have a sequence homology
of 92 % [114]. The difference in activity between human and mouse p32 was
mapped to a single amino acid, aa 35 in human p32. Thus, changing aa 35
from glycine to aspartic acid abrogated the capacity of human p32 to block
the excessive splicing of the HIV-1 genome.
p32 blocks phosphorylation of ASF/SF2
Hypophosphorylated SR proteins have a reduced capacity to bind to RNA.
Could p32 affect the phosphorylation status of ASF/SF2? By analyzing the
phosphorylation of ASF/SF2 in the presence of different recombinant
kinases in an in vitro kinase assay and then investigate the ability of p32 to
bind to ASF/SF2 we observed that the phosphorylation status of ASF/SF2
affected the interaction between p32 and ASF/SF2. Even though SRPK1
phosphorylated ASF/SF2 more strongly than Clk/Sty, the Clk/Sty phosphorylation of ASF/SF2 was more efficient in preventing p32 from binding
to ASF/SF2 than the SRPK1 phosphorylation indicating that Clk/Sty phosphorylates sites that are of more importance for p32 interaction with
ASF/SF2 compared to SRPK1.
Although phosphorylation of ASF/SF2 could reduce the interaction with
p32, an even more interesting finding was that p32 could actually inhibit
ASF/SF2 phosphorylation. By pre-incubating recombinant unphosphorylated
ASF/SF2 and p32 before the addition of SRPK1, Clk/Sty, or HeLa-NE the
inhibitory effect of p32 on ASF/SF2 phosphorylation was enhanced. Interestingly, p32 inhibition of ASF/SF2 phosphorylation was most effective
against Clk/Sty phosphorylation. A dot-blot far-western approach also dem43
onstrated that phosphorylation by SRPK1 probably affected the conformation of ASF/SF2 causing an indirect effect on the interaction between p32
and ASF/SF2, while phosphorylation by Clk/Sty at the specific site(s) was
directly inhibited p32 interaction. These findings also support the hypothesis
that the amino acids in ASF/SF2 targeted by Clk/Sty are of greater importance for p32-ASF/SF2 interaction.
The SRPK and Clk/Sty families of protein kinases have distinct enzymatic properties. Recently, Velazquez-Dones et al. preformed a mass spectrometric and kinetic analysis of SRPK1 and Clk/Sty phosphorylation of
ASF/SF2. They demonstrated that SRPK1 efficiently phosphorylated a short
stretch of amino acids near the N-terminal end of the RS-domain, while
Clk/Sty was able to phosphorylate all 20 serine residues in the RS-domain.
Current data indicate that the two kinases can phosphorylate ASF/SF2 sequentially, but not simultaneously and that they antagonize each other [203].
Further, Clk/Sty was shown to be a nuclear kinase in several cell lines [155]),
whereas SRPK1 is predominantly cytoplasmic in interphase cells [201]. The
current model that cytoplasmic SRPK family members are involved in the
transport of SR proteins into the nucleus, while the nuclear Clk/Sty family of
kinases are involved in regulating splicing [203] is in line with our finding
that p32 efficiently inhibited Clk/Sty phosphorylation to regulate the RNA
binding capacity of ASF/SF2.
Other cellular proteins affecting the activity of SR proteins
Not many cellular proteins have been reported to be able to affect the activity of SR proteins (see section Proteins regulating SR protein activity and
Table 4). Some proteins compete with RS-proteins in binding to a particular
RNA binding site, like hnRNP A1 and RBM4 [102]. The Tra2 proteins need
SR proteins to be able to enhance splicing [189], while others like SRr86 can
both enhance and repress SR protein induced splicing, through mechanism
that currently is unknown [10]. So far, the only cellular protein known to
affect the activity of a SR protein by controlling the phosphorylation and
thereby the RNA binding capacity of an SR protein is the cellular protein
p32 as reported in this thesis.
Paper III
Identification of a carboxy-terminal sequence in p32 necessary
for p32-mediated inhibition of ASF/SF2 RNA binding
In paper III the interaction between ASF/SF2 and p32 was studied in more
detail. We chose to study the direct binding of the two proteins by using pull
down assays with recombinant proteins and 35S-methionine labelled proteins
produced by coupled transcription/translation in wheat germ extracts.
44
Results in Paper II demonstrated that p32 could prevent ASF/SF2-'RS
from becoming phosphorylated in HeLa-NE and further that p32 could inhibit the RNA binding capacity of ASF/SF2-'RS. In paper III we demonstrated a direct interaction between ASF/SF2 lacking the RS-domain and
recombinant p32 and found that the binding of p32 to ASF/SF2-'RS was
much stronger than to full length ASF/SF2. Since the ASF/SF2 and the
ASF/SF2-'RS proteins used in this study were produced in wheat germ extract this led to the production of phosphorylated proteins. A Calf Intestine
Alkaline Phosphatase, CIAP, treatment of full length ASF/SF2 increased the
binding to p32 to the same strength as for ASF/SF2-'RS, with or without
CIAP treatment. The conclusion from this experiment was that p32 interacts
primarily with the N-terminal RRMs in ASF/SF2. Further, reduction of the
number of phosphorylated residues either by deletion or dephosphorylation
of the RS-domain strengthened the interaction between p32 and ASF/SF2.
p32 is known to bind to the lamin B receptor, p58, which is localized at
the nuclear membrane. The p58 protein contains an N-terminal RS-rich region similar to the C-terminal RS-domain found in ASF/SF2. Due to this
similarity Simos and Georgatos proposed that the RS-domain of p58 and
ASF/SF2 would provide binding sites for p32 [179]. Later, it was shown that
the N-terminal domain of p58 binds to p32 while a p58 mutant lacking the
RS-domain does not bind. Interestingly, phosphorylation of p58 by a nuclear
membrane RS kinase completely abolished binding to p32 [142]. Even
though there are similarities in the interaction properties there are also differences since we have demonstrated that p32 binds directly to the
N-terminal RRMs in ASF/SF2 and cause an inhibition of phosphorylation of
the C-terminal RS-domain in ASF/SF2. Further, phosphorylation of the
C-terminal RS-domain of ASF/SF2 prevents the p32-ASF/SF2 interaction
(Paper II and III).
We also wanted to determine which part of p32 that was important for the
p32-ASF/SF2 interaction. The RNA binding capacity of recombinant
ASF/SF2 was investigated in a UV-cross linking assay together with either
recombinant full length p32 or a C-terminal deletion mutant of p32, p32-'C
(amino acid 74-193). The recombinant p32-'C protein showed a dramatically reduced capacity to block ASF/SF2 RNA binding compared to the full
length protein. In line with this result, pull down assays demonstrated that
p32-'C protein, produced in wheat germ extract, could not bind to ASF/SF2.
Analysis of other p32 C-terminal deletion mutants in pull down assays demonstrated that the region between amino acid 226-239 in p32 is critical for
the interaction with ASF/SF2 (Figure 5). However, a GST-fusion protein
containing the residues of 226-239 of p32 was unable to bind to ASF/SF2
suggesting that aa 226-239 in p32 are necessary but not sufficient for binding to ASF/SF2.
The change of amino acid 35 from glycine to aspartic acid (G35D) in human p32 has a negative effect on HIV-1 replication in murine cells due to
45
excessive splicing of the HIV-1 transcript similarly to what have been shown
with the murine p32 protein [220] (see section Paper II). This finding indicates that amino acid 35 is important for the ability of p32 to inhibit
ASF/SF2. Zheng et al. speculated that introduction of the G35D mutation
into human p32 disrupted the asymmetrically distributed negative charges in
the doughnut shaped p32 trimer and thereby attenuated the ability of human
p32G35D to inhibit ASF/SF2 activity [220].
Our results shows that a deletion of the C-terminal amino acids 239-282,
as well as a internal deletion of amino acids 192-211, of p32 did not affect
the interaction with ASF/SF2 indicating that the N-terminus of p32 might be
involved in the direct binding. Although, a protein consisting of the Nterminal part of p32, p32-'C, consisting of aa74-193, and even p32(aa74226), could not bind to ASF/SF2 and p32-'C could not affect ASF/SF2
binding to RNA. Altogether, amino acids 226-239 in human p32 are important for the interaction with ASF/SF2, but this region is not sufficient for the
direct interaction. The conclusion from our study is that it might be important for p32 to form a trimer to be able to bind to ASF/SF2. With these findings in mind it would be interesting to see if the G35D mutation of p32 is
enough to prevent p32 from directly bind to ASF/SF2.
Paper IV
p32 represses the adenovirus major late transcription and causes
hyperphosphorylation of RNA polymerase II
In Paper IV we used recombinant adenovirus vector to overexpress the cellular protein p32 from an inducible promoter. Further, we used adenovirus as a
model organism to analyze what effect p32 has on the protein expression
from different types of viral genes. This in vivo study demonstrated that p32
has a promoter specific repressive effect on transcription which is manifested through regulation of Pol II phophorylation.
Overexpression of p32 during an adenovirus infection causes a
reduction of mRNA expression from MLP
To further investigate how p32 affects splicing in vivo we constructed a recombinant adenovirus overexpressing p32 from an inducible promoter. The
recombinant
adenovirus
Ad-TetTrip(flag-)p32,
referred
to
as
AdTT(flag-)p32 in paper IV, lacks the whole E1A and E1B-region, where
the gene encoding an N-terminal flag-epitope tagged p32 fusion protein was
inserted. The p32 gene is under the transcriptional control of an inducible
promoter which has seven binding sites for the reverse tetracycline repressor,
rtetR, and a minimal MLP. The Tet-ON system is based on a double infection strategy where the recombinant adenovirus with the gene of interest (in
46
this case the flag-p32 gene) and a recombinant helper adenovirus,
AdCMVrtTA, expressing a fusion protein consisting of the DNA binding
rtetR and the transcriptional activating domain of VP16 (Herpes simplex
virus protein) under the constitutive Cytomegalovirus (CMV) promoter, are
infecting the same cell. The rtetR protein, which is a mutant of the E.coli tetrepressor, tetR, binds to DNA in the presence of the antibiotic doxycyclin,
dox.
The purpose of the experiment was to investigate if p32 could affect the
splicing pattern of the L1 mRNAs also under a lytic infection. To be able to
analyze the late phase of infection the virus has to be able to replicate which
was why the 293 cell line was chosen for infection. 293 cells are human
embryonic neuronal cells from kindney [172] transformed with the E1A and
E1B region of adenovirus type 5 (base pair 1-4033) and therefore provide
the transcriptional activator E1A in trans [57]. Surprisingly, overexpression
of p32 resulted in reduction of the total amount of mRNA expressed from
the MLP and this led us to further analyze wether p32 had an effect on transcription.
To investigate the specificity of p32 repression of transcription the
mRNA amounts from the E2A- and the E4-region were also measured. In a
wild type infection the early E2A promoter is activated by E1A, while the
late E2A promoter is activated by E1B-55K and the CAAT-box binding
protein YB-1 (see section Early genes) [71]. Northern blot of E2A revealed
that the adenoviral E2A promoter is not affected by p32 overexpression.
This result was also confirmed by Western blot assay demonstrating a stable
expression of the E2A-72K protein. Similarly the adenoviral E4 promoter
was not repressed by p32. Interestingly, the abundance of the largest
E4 mRNA species was increased and we suggested that p32 might cause an
affect on E4 alternative splicing in p32 overexpressing cells, similarly to
what we reported in paper II. It has previously been demonstrated that inhibition of MLP by mutation analysis can affect early promoter usage and
affect the splicing pattern of early genes [47]. Thus, since p32 represses MLP
this might lead to an indirect effect on the E4 expression pattern. In conclusion, transcription of the MLP was severely inhibited by overexpressed p32,
while the E2A and E4 promoters were not repressed.
p32 repression of transcription is dependent on the presence of CAATbox
In order to investigate the MLP promoter architecture important for p32
induced repression of transcription we utilized a transient transfection strategy with different CAT reporter constructs together with a plasmid constitutively expressing p32, pCMV-p32. Transcription of the wild type MLP-CAT
construct is dramatically reduced with increasing amounts p32. The CAATbox (CCAAT) of MLP is one of two redundant upstream activating elements
47
important for MLP activity (CAAT-box and UPE element) and a mutation of
the CAAT-box causes a reduction in virus replication especially if it is combined with mutation of the UPE element (see section Major late promoter;
[160, 182]). As expected, introduction of a point mutation in the CAAT-box
(CCCAT) in the MLP-CAT construct resulted in a reduced basal expression
and destroyed the ability of p32 to further repress CAT-protein expression.
Thus, p32 mediated transcriptional inhibition seems to need an intact CAATbox in the MLP. Maity et al. has reported that the CAAT-box binding transcription factor CBF/NF-Y is able to transcriptionally activate MLP [116].
Interestingly, p32 was recently shown to interact both in vitro and in vivo
with CBF/NF-Y and further inhibit CBF/NF-Y activated transcription of the
D2(1) collagen promoter [33]. The human tripeptidyl-peptidase II promoter
contains two CAAT-boxes and has previously been demonstrated to bind
CBF/NF-Y protein [110]. By utilizing different deletion mutants of the tripeptidyl-peptidase II promoter, lacking one or two CAAT-boxes, in front of the
luciferase gene we could confirm that the inhibitory effect of p32 overexpression on transcription was dependent on the presence of intact CAATboxes.
In conclusion, transient transfections show that transcription from promoters containing CAAT-boxes known to be stimulated by the CBF/NF-Y
transcription factor, like the adenoviral MLP and the human tripeptidylpeptidase II promoter is inhibited by p32.
p32 causes hyperphosphorylation of RNA Pol II CTD and affects the
processivity of transcription elongation
In order to analyse the effect of p32 on Pol II, chromatin immunoprecipitation assays, ChIP assays, were used to measure the presence of Pol II on the
MLP and its coding region. Cells infected with the recombinant virus were
induced to express p32. The cells were cross-linked with formaldehyde at 22
hpi forming covalent bonds between interacting proteins and protein-DNA.
The DNA was fragmentized by sonication, immunoprecipitated and PCR
amplified in order to analyse the gene of interest at the moment of fixation.
Using an antibody that targets the N-terminus of the large subunit of Pol II
we analysed the polymerase distribution on the L1 region of MLTU. PCR
was used to amplify fragments from the MLP, the middle and at the end of
the L1 gene. We detected a large fraction of Pol II at the promoter region
suggesting pausing of the polymerase and also some accumulation towards
the end of the gene. Though, the pausing effect on the promoter was not at
the same extent as previously reported on eukaryotic genes [34]. As expected,
hypophosphorylated polymerase, Pol IIA, was primarily detected at the promoter region. ChIP assays have showed that phosphoserine 5 specific antibody H14 cross-links primarily to Pol II at the promoter proximal region,
while the phosphoserine 2 specific antibody H5 cross-links primarily to the
48
coding region of the gene [94]. This pattern was also found in the L1-region
of wild-type adenovirus in our ChIP assays with the H14 and H5 antibodies.
Using the Ad-TT(flag-)p32 virus and comparing the result between doxycyclin induced and uninduced samples we demonstrated that p32 caused a reduction of polymerase molecules during the elongation phase. The ChIP
assays using H14 and H5 antibodies detecting Pol II IIO demonstrated that
the polymerases that still were transcribing had an increased amount of both
Ser 5 and Ser 2 phosphorylation, respectively. Incubation of GST-CTD protein together with increasing amount of recombinant p32 in vitro confirmed
that p32 enhances the phosphorylation of the polymerase. Analysis of nuclear extract from p32 overexpressing cells revealed that a large fraction of
the endogenous Pol II was hyperphosphorylated at both Ser 5 and Ser 2 positions.
Regulating the activities of RNA Pol II phosphatases and kinases
Overexpression of p32 during transcription of the MLTU L1-region
causes both Ser 5 and Ser 2 hyperphosphorylation, which represses the processivity of the Pol II. Other proteins have been reported to affect the phosphorylation/dephosphorylation cycle of Pol II (Table 2). Interestingly, some
of these proteins are causing a hyperphosphorylation of the CTD resulting in
transcriptional inhibition similar to that by p32. The peptidyl-propyl isomerase Pin 1 causes hyper-hyperphosphorylation of CTD in vitro by both inhibiting FCP1 and stimulating the kinase cdc2/cyclin B, a mitotic kinase
which is able to phosphorylate CTD at both Ser 5 and Ser 2 [211]. By constructing an inducible stable Pin 1 cell line, Xu et al. showed that overexpression of Pin 1 inhibited transcription in vivo. Further, Pin 1 caused a
Pol II dependent inhibition of pre-mRNA splicing in vitro [211]. The yeast
homologue of Pin 1, Ess1p, binds to the CTD both in vivo and in vitro and
has been shown to interact with all four yeast CTD kinases. The present
model is that Essp1 affects transcription of mitotic genes by coordinating
multiple steps in transcription involving regulation of both isomerisation and
phosphorylation of the CTD [134, 206, 208].
Further, BRCA1, a DNA binding protein which has shown to be a component of the Pol II holoenzyme, can block the ATP binding site of
CDK7/TFIIH and thereby prevent the phosphorylation of Ser 5 [133]. The
FCP1 phosphatase is regulated by HIV-1 Tat, CK2, TFIIB and the large
subunit RAP74 of TFIIH. Phosphorylation of FCP1 by CK2 enhances FCP1
activity and strengthens its binding to RAP74, which even further stimulates
the phosphatase activity [2]. The HIV-1 Tat protein inhibits both the binding
of RAP74 to FCP1 and CK2 phosphorylation, resulting in an inhibition of
FCP1 [1]. During HIV-1 Tat transcriptional activation Tat is on the other
hand stimulating P-TEFb to phosphorylate Ser 5 and Ser 2 [89, 221].
Recently, Epsteinn-Barr virus, EBV, EBNA 2 was shown to stimulate
transcription by hyperphosphorylating Ser 5 [9]. The recruitment of Pol II to
49
the promoter region was increased in presence of EBNA 2 and showed normal level of Ser 5 phosphorylation. During the elongation phase of transcription this increased number of Pol II showed an increasing level of Ser 5
phosphorylation but no further stimulation on Ser 2 phosphorylation [9].
The regulatory proteins of CTD kinases and phosphatases are of different
origin and further the recruitment to their target enzymes differ; binding to
the DNA, to transcription activators/repressors, to classical transcription
factors, or via the CTD of Pol II. Some of the proteins in Table 2 seem to be
acting gene specific or having a cell cycle specific function, like HIV-1 Tat
and Pin1/Ess1p, while others might have a more general function since they
interact with subunits of the Pol II complex, like RAP74.
PIC formation
Initiation
Elongation
Termination
RNA pol II
CTD
CBF/NF-Y
p32
Ser 5 phosphorylation
Ser 2 phosphorylation
Figure 7. Model on p32 repression of a CBF/NF-Y activated promoter. Overexpression of p32 causes hyperphosphorylation of both Ser 5 and Ser2 in the Pol II CTD
and the processivity during the elongation phase of elongation is reduced.
50
Our data and the report of Chattopadhyay et al. support a model where
p32 is recruited to promoters containing binding sites for the transcription
factor CBF/NF-Y (paper IV and [33]). It is possible that CBF/NF-Y recruitment places p32 proximal to the initiation site leading to a direct association
to the CTD. p32 might even travel along with the elongating complex and
act on kinases/phosphatases during the elongation phase. Unfortunately, in
our study we did not succeed in demonstrating a direct interaction between
p32 and the Pol II or the MLP during elongation. This is due to low specificity of the flag epitope antibody used. Another hypothesis is that p32, bound
to CBF/NF-Y on the MLP promoter, prevents the recruitment of FCP1 to the
PIC during the transition from the initiation phase to the elongation phase of
transcription. FCP1 has been shown to have a stimulatory effect on transcription elongation and that activity is independent on its phosphatase activity [36, 119]. ChIP experiments have also showed that FCP1 remains associated with Pol II during elongation [35]. Prevention of binding or sequestering
of FPC1 would most likely lead to loss of FCP1 mediated stimulation of
elongation. This hypothesis is in line with our results since p32 causes an
increase of phosphorylation of both Ser 5 and Ser 2 and a reduced processivity of the Pol II during elongation. p32 might also recruit new CTD kinases
or stimulate the activity of protein kinases like CDK7 and CDK9. In the
hyperphosphorylated Pol II not all Ser 5 and Ser 2 residues in the 52 heptad
repeats are phosphorylated at the same time and we speculate that p32 might
stimulate phosphorylation of novel Ser 5 and/or Ser 2 residues, which might
be more difficult to dephosphorylate. Interestingly, both Pin 1 and HIV-1 Tat
inhibit the dephosphorylation activity of FCP1 and activate phosphorylation
by specific kinases. This finding opens up the possibility for a dual function
for p32 as well.
To further dissect the repressive role of p32 on transcription of CAATbox containing genes it would be interesting to analyse the position of the
p32 protein during the transcription process, as well as doing direct interaction studies between p32 and possible p32 interacting candidates like the
kinases CDK7 and CDK9, the phosphatase FCP1, and different subunits of
the Pol II would be of interest.
51
Conclusions
Paper I
x SR proteins inactivated in the late pase of adenovirus infection due to
hypophosphorylation
x E4-ORF4 and phosphatase PP2A cause the dephosphorylation of SR protein during the infection
x Alternative splicing is regulated by reversible protein phosphorylation
Paper II
x p32 inactivates ASF/SF2 both as a splicing repressor and a splicing enhancer protein
x p32 blocks the phosphorylation of ASF/SF2 and reduces the RNA binding capacity of ASF/SF2
Paper III
x P32 binds to the RBD of ASF/SF2 and dephosphorylation of ASF/SF2
strengths the binding to p32
x The region of aa 226-239 in p32 is important, but not sufficient itself for
binding to ASF/SF2
Paper IV
x p32 inhibits transcription of genes with CAAT-box containing promoters.
Further the results suggest that p32 is recruited to the promoter through
the interaction with the transcription factor CBF/NF-Y.
x p32 causes hyperphosphorylation of both Ser 5 and Ser 2 on the CTD of
the large subunit of Pol II, both in vitro and in vivo.
x p32 induced hyperphosphorylation causes a reduction in the processivity
of Pol II elongation.
52
Acknowledgements
First of all I would like to thank my supervisor Professor Göran Akusjärvi for all
patience you have had with me during all these years. I am so grateful for the all
your support and help, especially during these last interesting years. You have taught
me a lot and you have encouraged me to think how to always try to take the project
to a higher level. I also appreciate that you always have had time for me to discuss
my projects. Thank you for allowing me to proceed with the thesis, even though the
years started to fly away…
I would also like to give my warmest thanks to Docent Catharina Svensson for your
contribution to my development as a scientist. Thanks for all feed-back on my research, and all the pep-up talks you have been given me through the years! For the
nice lunch talks about everything in life out on the sunny balcony!
To my examiner Professor Göran Magnusson and to Professor Stefan Schwartz, for
taking time to share your virology knowledge.
All the virology PhD students, post-docs and scientists in the Adeno-group:
Anders Sundqvist, Anne-Christine Ström, Anette Carlsson, Anette Lindberg , Arvydas Kanopka, Baigong Yue, Bin Wang, Camilla Estmer-Nilsson, Cecilia Johansson,
Dan Edholm, Daniel Öberg, David Huang, Edyta Bajak, Ellenor Bäckström, Elisabeth Boija, Gunnar Andersson, Heidi Törminen, Josef Seibth, Kerstin Sollerbrant,
Karin Öhman-Forslund, Lamine Bouakaz, Magnus Molin, Maria Bondesson, Mattias Mannervik, Martin Lützelberger, Niel Portwood, Ning Xu, Oliver Mühlemann,
Petra Olsson, Peter Kreivi, Raj Aravelli, Saideh Berenjian, Shaoan Fan, Svend Petersen-Mahrt, Tanel Punga, Vita Dauksaite, Xiaofu Zhou and Öjar Melefors.
Without all the fun that we have had together I probably wouldn’t have made it to
the very end. I’m grateful for all the incredible memories in which you are included!
The Friday meetings-members! All of you mentioned above and Anna, Anton,
Emilio, Hongxing, Marcin, Margaret, Monika, Ulrika, Xiaomin and everyone who
has attended at theses meetings have been to so much help. Thanks for all input you
gave me on my projects and my literature seminars. It has been a tough, but great,
school to discuss experiment failures and nice data, to handle criticism and encouragement, in front of such a very competent audience.
For the staff that have always been there for me no matter if it had to do with broken
machines, material supply, administration help, computer problems or a nice talk
over a cup of coffée: Barbro Lowisin, Erika Enström, Kerstin Lidholt, Lena Möller,
Lilian Ekström, Maud Pettersson, Olav Nordli, Tony Grundin and Ylva Jansson.
53
Cecilia- There would not be a thesis without you, at least not one in this nice shape!
It has been fantastically fun to have you as my travel company the last years of this
bumpy journey! You mean so much to me and I’m not sure that I will allow you to
move abroad!
Saideh- For being my friend and for all our talks about science and more importantly
about the world “outside”. It has been invaluable for me.
Tanel- For being such a friendly person and always helping out and sharing both
your ideas and the lab stuff. Without the ChIPs, no thesis…
Svend- for your enthusiasm and for introducing me to the multifunctional p32.
Camilla- your friendship + our great cooperation during the struggling with the S1s.
Ellenor- for being my exercise-friend and still pushing me to go there…
Heidi- for sharing office and spreading a nice atmosphere.
Marika and Andrzej- For the fun we have had before the kids and after! For our very
important loooong phone calls! Marika, I know that you are always there for me and
that means a lot to me…
Anne-Christine- for your energy and never ending good mood! For cheering me up
and having pep-talks about what to really focus on.
Karin och Arne- for your friendship and support from day one, and all nice occasions at Bönan and in your home.
Micke och Anna-Karin- for all memorable vacations and weekends we have spent
together during these years, Idre, Skåne, Båven…
J2+E2+W- for your friendship and fantastic dinners. For believing in me when I have
said that “There will be a party soon…” and now: Dissertation party without you!!!
Jonas and Johanna, I promise I will report everything to the Black Book.
Ida, Lars and Wilhelm- for your support and to our future to come with CDF!
Frida och Calle- for all cosy times together.
Eva and Kurt- for all love that I get from you! For taking such good care of all of us
and our home!
My extra-parents Kerstin och Janne- for showing an endless support not only for me
but also for the rest of my family.
Annika and Tobias- for opening my eyes to another world! Annika, for being min
älskade Pyss!
Mamma och Pappa, What can I say! You did it! Your support in every level of my
life and your endless belief in me made this thesis to come true. Thanks for allowing
us to “refuel” at Tegelön during all of these summers and for all help during the
years. Jag älskar er!
Min älskade Maja och min älskade Saga! Ni är mitt allt!
Johan- Without you, no list of references! Honestly, no thesis either. Thanks for
taking care of me during all the ups and downs… You are always there for me no
matter what and I am so grateful for that. I want to thank you for making my life to
what it is. You give me so much love och jag älskar dig så…
54
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
Abbott, K.L., Archambault, J., Xiao, H., Nguyen, B.D., Roeder, R.G., Greenblatt, J., Omichinski, J.G., and Legault, P., (2005) Interactions of the HIV-1 Tat
and RAP74 proteins with the RNA polymerase II CTD phosphatase FCP1, Biochemistry. 44(8): p. 2716-2731.
Abbott, K.L., Renfrow, M.B., Chalmers, M.J., Nguyen, B.D., Marshall, A.G.,
Legault, P., and Omichinski, J.G., (2005) Enhanced binding of RNAP II CTD
phosphatase FCP1 to RAP74 following CK2 phosphorylation, Biochemistry. 44(8):
p. 2732-2745.
Adam, S.A. and Dreyfuss, G., (1987) Adenovirus proteins associated with mRNA
and hnRNA in infected HeLa cells, J Virol. 61(10): p. 3276-3283.
Ahn, S.H., Kim, M., and Buratowski, S., (2004) Phosphorylation of serine 2
within the RNA polymerase II C-terminal domain couples transcription and 3' end
processing, Mol Cell. 13(1): p. 67-76.
Akoulitchev, S., Chuikov, S., and Reinberg, D., (2000) TFIIH is negatively regulated by cdk8-containing mediator complexes, Nature. 407(6800): p. 102-106.
Akusjarvi, G. and Stevenin, J., (2003) Remodelling of the host cell RNA splicing
machinery during an adenovirus infection, Curr Top Microbiol Immunol. 272: p.
253-286.
Aubol, B.E., Chakrabarti, S., Ngo, J., Shaffer, J., Nolen, B., Fu, X.D., Ghosh,
G., and Adams, J.A., (2003) Processive phosphorylation of alternative splicing
factor/splicing factor 2, Proc Natl Acad Sci U S A. 100(22): p. 12601-12606.
Bai, Y., Lee, D., Yu, T., and Chasin, L.A., (1999) Control of 3' splice site choice
in vivo by ASF/SF2 and hnRNP A1, Nucleic Acids Res. 27(4): p. 1126-1134.
Bark-Jones, S.J., Webb, H.M., and West, M.J., (2005) EBV EBNA 2 stimulates
CDK9-dependent transcription and RNA polymerase II phosphorylation on serine
5, Oncogene.
Barnard, D.C. and Patton, J.G., (2000) Identification and characterization of a
novel serine-arginine-rich splicing regulatory protein, Mol Cell Biol. 20(9): p.
3049-3057.
Beatch, M.D., Everitt, J.C., Law, L.J., and Hobman, T.C., (2005) Interactions
between rubella virus capsid and host protein p32 are important for virus replication, J Virol. 79(16): p. 10807-10820.
Beatch, M.D. and Hobman, T.C., (2000) Rubella virus capsid associates with host
cell protein p32 and localizes to mitochondria, J Virol. 74(12): p. 5569-5576.
55
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
56
Ben-David, Y., Letwin, K., Tannock, L., Bernstein, A., and Pawson, T., (1991)
A mammalian protein kinase with potential for serine/threonine and tyrosine phosphorylation is related to cell cycle regulators, Embo J. 10(2): p. 317-325.
Benkö, M., Harrach, B., and Russel, W., (2000),Family Adenoviridae., in Virus
Taxonomy. Classification and Nomenclature of Viruses. Seventh Report of the International Commitee on Taxonomy of Viruses, M. van Regenmortel, C. Fauquet,
D. Bishop, E. Carstens, M. Estes, L. SM., J. Maniloff, M. Mayo, D. McGeoch , C.
Pringle, and R. Wickner, Editors: San Diego. p. 227-238.
Bentley, D., (2002) The mRNA assembly line: transcription and processing machines in the same factory, Curr Opin Cell Biol. 14(3): p. 336-342.
Bergelson, J.M., Cunningham, J.A., Droguett, G., Kurt-Jones, E.A., Krithivas,
A., Hong, J.S., Horwitz, M.S., Crowell, R.L., and Finberg, R.W., (1997) Isolation of a common receptor for Coxsackie B viruses and adenoviruses 2 and 5, Science. 275(5304): p. 1320-1323.
Berro, R., Kehn, K., de la Fuente, C., Pumfery, A., Adair, R., Wade, J., Colberg-Poley, A.M., Hiscott, J., and Kashanchi, F., (2006) Acetylated Tat Regulates
Human Immunodeficiency Virus Type 1 Splicing through Its Interaction with the
Splicing Regulator p32, J Virol. 80(7): p. 3189-3204.
Bi, W., Wu, L., Coustry, F., de Crombrugghe, B., and Maity, S.N., (1997) DNA
binding specificity of the CCAAT-binding factor CBF/NF-Y, J Biol Chem. 272(42):
p. 26562-26572.
Bondesson, M., Öhman, K., Mannervik, M., Fan, S., and Akusjärvi, G., (1996)
Adenovirus E4 open reading frame 4 protein autoregulates E4 transcription by inhibiting E1A transactivation of the E4 promoter, J Virol. 70(6): p. 3844-3851.
Boucher, L., Ouzounis, C.A., Enright, A.J., and Blencowe, B.J., (2001) A genome-wide survey of RS domain proteins, Rna. 7(12): p. 1693-1701.
Bourgeois, C.F., Lejeune, F., and Stevenin, J., (2004) Broad specificity of SR
(serine/arginine) proteins in the regulation of alternative splicing of pre-messenger
RNA, Prog Nucleic Acid Res Mol Biol. 78: p. 37-88.
Boyer, J., Rohleder, K., and Ketner, G., (1999) Adenovirus E4 34k and E4 11k
inhibit double strand break repair and are physically associated with the cellular
DNA-dependent protein kinase, Virology. 263(2): p. 307-312.
Braun, L., Ghebrehiwet, B., and Cossart, P., (2000) gC1q-R/p32, a C1q-binding
protein, is a receptor for the InlB invasion protein of Listeria monocytogenes, Embo
J. 19(7): p. 1458-1466.
Bruni, R. and Roizman, B., (1996) Open reading frame P--a herpes simplex virus
gene repressed during productive infection encodes a protein that binds a splicing
factor and reduces synthesis of viral proteins made from spliced mRNA, Proc Natl
Acad Sci U S A. 93(19): p. 10423-10427.
Bryant, H.E., Matthews, D.A., Wadd, S., Scott, J.E., Kean, J., Graham, S.,
Russell, W.C., and Clements, J.B., (2000) Interaction between herpes simplex virus type 1 IE63 protein and cellular protein p32, J Virol. 74(23): p. 11322-11328.
Bubulya, P.A., Prasanth, K.V., Deerinck, T.J., Gerlich, D., Beaudouin, J.,
Ellisman, M.H., Ellenberg, J., and Spector, D.L., (2004) Hypophosphorylated SR
splicing factors transiently localize around active nucleolar organizing regions in
telophase daughter nuclei, J Cell Biol. 167(1): p. 51-63.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
Caceres, J.F., Screaton, G.R., and Krainer, A.R., (1998) A specific subset of SR
proteins shuttles continuously between the nucleus and the cytoplasm, Genes Dev.
12(1): p. 55-66.
Caceres, J.F., Stamm, S., Helfman, D.M., and Krainer, A.R., (1994) Regulation
of alternative splicing in vivo by overexpression of antagonistic splicing factors,
Science. 265(5179): p. 1706-1709.
Cao, W., Jamison, S.F., and Garcia-Blanco, M.A., (1997) Both phosphorylation
and dephosphorylation of ASF/SF2 are required for pre-mRNA splicing in vitro,
Rna. 3(12): p. 1456-1467.
Carthew, R.W., Chodosh, L.A., and Sharp, P.A., (1985) An RNA polymerase II
transcription factor binds to an upstream element in the adenovirus major late
promoter, Cell. 43(2 Pt 1): p. 439-448.
Cepko, C.L. and Sharp, P.A., (1983) Analysis of Ad5 hexon and 100K ts mutants
using conformation-specific monoclonal antibodies, Virology. 129(1): p. 137-154.
Chambers, R.S., Wang, B.Q., Burton, Z.F., and Dahmus, M.E., (1995) The
activity of COOH-terminal domain phosphatase is regulated by a docking site on
RNA polymerase II and by the general transcription factors IIF and IIB, J Biol
Chem. 270(25): p. 14962-14969.
Chattopadhyay, C., Hawke, D., Kobayashi, R., and Maity, S.N., (2004) Human
p32, interacts with B subunit of the CCAAT-binding factor, CBF/NF-Y, and inhibits
CBF-mediated transcription activation in vitro, Nucleic Acids Res. 32(12): p. 36323641.
Cheng, C. and Sharp, P.A., (2003) RNA polymerase II accumulation in the promoter-proximal region of the dihydrofolate reductase and gamma-actin genes, Mol
Cell Biol. 23(6): p. 1961-1967.
Cho, E.J., Kobor, M.S., Kim, M., Greenblatt, J., and Buratowski, S., (2001)
Opposing effects of Ctk1 kinase and Fcp1 phosphatase at Ser 2 of the RNA polymerase II C-terminal domain, Genes Dev. 15(24): p. 3319-3329.
Cho, H., Kim, T.K., Mancebo, H., Lane, W.S., Flores, O., and Reinberg, D.,
(1999) A protein phosphatase functions to recycle RNA polymerase II, Genes Dev.
13(12): p. 1540-1552.
Colwill, K., Pawson, T., Andrews, B., Prasad, J., Manley, J.L., Bell, J.C., and
Duncan, P.I., (1996) The Clk/Sty protein kinase phosphorylates SR splicing factors
and regulates their intranuclear distribution, Embo J. 15(2): p. 265-275.
Dahmus, M.E., (1996) Reversible phosphorylation of the C-terminal domain of
RNA polymerase II, J Biol Chem. 271(32): p. 19009-19012.
Deb, T.B. and Datta, K., (1996) Molecular cloning of human fibroblast hyaluronic
acid-binding protein confirms its identity with P-32, a protein co-purified with
splicing factor SF2. Hyaluronic acid-binding protein as P-32 protein, co-purified
with splicing factor SF2, J Biol Chem. 271(4): p. 2206-2212.
D'Halluin, J.C., (1995) Virus assembly, Curr Top Microbiol Immunol. 199 ( Pt 1):
p. 47-66.
Dobner, T., Horikoshi, N., Rubenwolf, S., and Shenk, T., (1996) Blockage by
Adenovirus E4orf6 of transcriptional activation by the p53 tumor suppressor, science. 272: p. 1470-1473.
57
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58
Dolph, P.J., Huang, J.T., and Schneider, R.J., (1990) Translation by the adenovirus tripartite leader: elements which determine independence from cap-binding
protein complex, J Virol. 64(6): p. 2669-2677.
Elsing, A. and Burgert, H.G., (1998) The adenovirus E3/10.4K-14.5K proteins
down-modulate the apoptosis receptor Fas/Apo-1 by inducing its internalization,
Proc Natl Acad Sci U S A. 95(17): p. 10072-10077.
Estmer Nilsson, C., Petersen-Mahrt, S., Durot, C., Shtrichman, R., Krainer,
A.R., Kleinberger, T., and Akusjarvi, G., (2001) The adenovirus E4-ORF4 splicing enhancer protein interacts with a subset of phosphorylated SR proteins, Embo J.
20(4): p. 864-871.
Evans, J.D. and Hearing, P., (2003) Distinct roles of the Adenovirus E4 ORF3
protein in viral DNA replication and inhibition of genome concatenation, J Virol.
77(9): p. 5295-5304.
Farley, D.C., Brown, J.L., and Leppard, K.N., (2004) Activation of the early-late
switch in adenovirus type 5 major late transcription unit expression by L4 gene
products, J Virol. 78(4): p. 1782-1791.
Fessler, S.P. and Young, C.S., (1998) Control of adenovirus early gene expression
during the late phase of infection, J Virol. 72(5): p. 4049-4056.
Flint, S., Enquist, L., Krug, R., Ravaniello, V., and Skalka, A., The priciples of
Virology. Molecular Biology, Pathogenesis and Control. 2000, Washington DC:
ASM Press.
Fong, N. and Bentley, D.L., (2001) Capping, splicing, and 3' processing are independently stimulated by RNA polymerase II: different functions for different segments of the CTD, Genes Dev. 15(14): p. 1783-1795.
Fong, N., Bird, G., Vigneron, M., and Bentley, D.L., (2003) A 10 residue motif at
the C-terminus of the RNA pol II CTD is required for transcription, splicing and 3'
end processing, Embo J. 22(16): p. 4274-4282.
Fong, Y.W. and Zhou, Q., (2001) Stimulatory effect of splicing factors on transcriptional elongation, Nature. 414(6866): p. 929-933.
Furger, A., O'Sullivan, J.M., Binnie, A., Lee, B.A., and Proudfoot, N.J., (2002)
Promoter proximal splice sites enhance transcription, Genes Dev. 16(21): p. 27922799.
Ge, H. and Manley, J.L., (1990) A protein factor, ASF, controls cell-specific alternative splicing of SV40 early pre-mRNA in vitro, Cell. 62(1): p. 25-34.
Ghebrehiwet, B., Lim, B.L., Kumar, R., Feng, X., and Peerschke, E.I., (2001)
gC1q-R/p33, a member of a new class of multifunctional and multicompartmental
cellular proteins, is involved in inflammation and infection, Immunol Rev. 180: p.
65-77.
Ghebrehiwet, B., Lim, B.L., Peerschke, E.I., Willis, A.C., and Reid, K.B.,
(1994) Isolation, cDNA cloning, and overexpression of a 33-kD cell surface glycoprotein that binds to the globular "heads" of C1q, J Exp Med. 179(6): p. 1809-1821.
Goodrich, J.A. and Tjian, R., (1994) TBP-TAF complexes: selectivity factors for
eukaryotic transcription, Curr Opin Cell Biol. 6(3): p. 403-409.
Graham, F.L., Smiley, J., Russell, W.C., and Nairn, R., (1977) Characteristics of
a human cell line transformed by DNA from human adenovirus type 5, J Gen Virol.
36(1): p. 59-74.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
Graveley, B.R., (2000) Sorting out the complexity of SR protein functions, RNA. 6:
p. 1197-1211.
Gregor, P.D., Sawadogo, M., and Roeder, R.G., (1990) The adenovirus major
late transcription factor USF is a member of the helix-loop-helix group of regulatory proteins and binds to DNA as a dimer, Genes Dev. 4(10): p. 1730-1740.
Gui, J.F., Lane, W.S., and Fu, X.D., (1994) A serine kinase regulates intracellular
localization of splicing factors in the cell cycle [see comments], Nature. 369(6482):
p. 678-682.
Guilfoyle, R.A., Osheroff, W.P., and Rossini, M., (1985) Two functions encoded
by adenovirus early region 1A are responsible for the activation and repression of
the DNA-binding protein gene, Embo J. 4(3): p. 707-713.
Gupta, S., Batchu, R.B., and Datta, K., (1991) Purification, partial characterization of rat kidney hyaluronic acid binding protein and its localization on the cell
surface, Eur J Cell Biol. 56(1): p. 58-67.
Hall, K.T., Giles, M.S., Calderwood, M.A., Goodwin, D.J., Matthews, D.A., and
Whitehouse, A., (2002) The Herpesvirus Saimiri Open Reading Frame 73 Gene
Product Interacts with the Cellular Protein p32, J Virol. 76(22): p. 11612-11622.
Hayes, B.W., Telling, G.C., Myat, M.M., Williams, J.F., and Flint, S.J., (1990)
The adenovirus L4 100-kilodalton protein is necessary for efficient translation of
viral late mRNA species, J Virol. 64(6): p. 2732-2742.
Hayman, M.L., Miller, M.M., Chandler, D.M., Goulah, C.C., and Read, L.K.,
(2001) The trypanosome homolog of human p32 interacts with RBP16 and stimulates its gRNA binding activity, Nucleic Acids Res. 29(24): p. 5216-5225.
Hertel, K.J. and Graveley, B.R., (2005) RS domains contact the pre-mRNA
throughout spliceosome assembly, Trends Biochem Sci. 30(3): p. 115-118.
Herwald, H., Dedio, J., Kellner, R., Loos, M., and Muller-Esterl, W., (1996)
Isolation and characterization of the kininogen-binding protein p33 from endothelial cells. Identity with the gC1q receptor, J Biol Chem. 271(22): p. 13040-13047.
Hierholzer, J., (1973) Further subgrouping of the human adenoviruses by differential hemagglutination, J Infect Dis. 128: p. 541-550.
Hirose, Y., Tacke, R., and Manley, J.L., (1999) Phosphorylated RNA polymerase
II stimulates pre-mRNA splicing, Genes Dev. 13(10): p. 1234-1239.
Ho, C.K., Sriskanda, V., McCracken, S., Bentley, D., Schwer, B., and Shuman,
S., (1998) The guanylyltransferase domain of mammalian mRNA capping enzyme
binds to the phosphorylated carboxyl-terminal domain of RNA polymerase II, J Biol
Chem. 273(16): p. 9577-9585.
Holm, P.S., Bergmann, S., Jurchott, K., Lage, H., Brand, K., Ladhoff, A.,
Mantwill, K., Curiel, D.T., Dobbelstein, M., Dietel, M., Gansbacher, B., and
Royer, H.D., (2002) YB-1 relocates to the nucleus in adenovirus-infected cells and
facilitates viral replication by inducing E2 gene expression through the E2 late
promoter, J Biol Chem. 277(12): p. 10427-10434.
Hong, S.S., Karayan, L., Tournier, J., Curiel, D.T., and Boulanger, P.A., (1997)
Adenovirus type 5 fiber knob binds to MHC class I alpha2 domain at the surface of
human epithelial and B lymphoblastoid cells, Embo J. 16(9): p. 2294-2306.
Honore, B., Madsen, P., Rasmussen, H.H., Vandekerckhove, J., Celis, J.E., and
Leffers, H., (1993) Cloning and expression of a cDNA covering the complete cod-
59
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
60
ing region of the P32 subunit of human pre-mRNA splicing factor SF2, Gene.
134(2): p. 283-287.
Howell, B.W., Afar, D.E., Lew, J., Douville, E.M., Icely, P.L., Gray, D.A., and
Bell, J.C., (1991) STY, a tyrosine-phosphorylating enzyme with sequence homology
to serine/threonine kinases, Mol Cell Biol. 11(1): p. 568-572.
Hrimech, M., Yao, X., Branton, P., and Cohen, E., (2000) Human immunodeficiency virus type 1 Vpr-mediated G(2) cell cycle arrest: Vpr interferes with cell cycle signaling cascades by interaction with the B subunit of serine/threonine protein
phosphatase 2 A, Embo J. 19: p. 3956-3967.
Huang, T.S., Nilsson, C.E., Punga, T., and Akusjarvi, G., (2002) Functional
inactivation of the SR family of splicing factors during a vaccinia virus infection,
EMBO Rep. 3(11): p. 1088-1093.
Huang, Y., Gattoni, R., Stevenin, J., and Steitz, J.A., (2003) SR splicing factors
serve as adapter proteins for TAP-dependent mRNA export, Mol Cell. 11(3): p.
837-843.
Jansen-Durr, P., Mondesert, G., and Kedinger, C., (1989) Replication-dependent
activation of the adenovirus major late promoter is mediated by the increased binding of a transcription factor to sequences in the first intron, J Virol. 63(12): p.
5124-5132.
Jha, B.K., Salunke, D.M., and Datta, K., (2002) Disulfide bond formation
through Cys186 facilitates functionally relevant dimerization of trimeric hyaluronan-binding protein 1 (HABP1)/p32/gC1qR, Eur J Biochem. 269(1): p. 298-306.
Jiang, J., Zhang, Y., Krainer, A.R., and Xu, R.M., (1999) Crystal structure of
human p32, a doughnut-shaped acidic mitochondrial matrix protein, Proc Natl
Acad Sci U S A. 96(7): p. 3572-3577.
Joseph, K., Ghebrehiwet, B., Peerschke, E.I., Reid, K.B., and Kaplan, A.P.,
(1996) Identification of the zinc-dependent endothelial cell binding protein for high
molecular weight kininogen and factor XII: identity with the receptor that binds to
the globular "heads" of C1q (gC1q-R), Proc Natl Acad Sci U S A. 93(16): p. 85528557.
Kanopka, A., Mühlemann, O., Petersen-Mahrt, S., Estmer, C., Öhrmalm, C.,
and Akusjärvi, G., (1998) Regulation of adenovirus alternative RNA splicing by
dephosphorylation of SR proteins, Nature. 393(6681): p. 185-187.
Kanopka, A., Mühlemann, O., and Akusjärvi, G., (1996) Inhibition by SR proteins of splicing of a regulated adenovirus pre- mRNA, Nature. 381(6582): p. 535538.
Kao, H.Y. and Siliciano, P.G., (1996) Identification of Prp40, a novel essential
yeast splicing factor associated with the U1 small nuclear ribonucleoprotein particle, Mol Cell Biol. 16(3): p. 960-967.
Kataoka, N., Bachorik, J.L., and Dreyfuss, G., (1999) Transportin-SR, a nuclear
import receptor for SR proteins, J Cell Biol. 145(6): p. 1145-1152.
Kaufmann, J., Ahrens, K., Koop, R., Smale, S.T., and Muller, R., (1998)
CIF150, a human cofactor for transcription factor IID-dependent initiator function,
Mol Cell Biol. 18(1): p. 233-239.
Kim, H.J., Jeong, S.H., Heo, J.H., Jeong, S.J., Kim, S.T., Youn, H.D., Han,
J.W., Lee, H.W., and Cho, E.J., (2004) mRNA capping enzyme activity is coupled
to an early transcription elongation, Mol Cell Biol. 24(14): p. 6184-6193.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
Kim, J.L., Nikolov, D.B., and Burley, S.K., (1993) Co-crystal structure of TBP
recognizing the minor groove of a TATA element, Nature. 365(6446): p. 520-527.
Kim, Y.K., Bourgeois, C.F., Isel, C., Churcher, M.J., and Karn, J., (2002)
Phosphorylation of the RNA polymerase II carboxyl-terminal domain by CDK9 is
directly responsible for human immunodeficiency virus type 1 Tat-activated transcriptional elongation, Mol Cell Biol. 22(13): p. 4622-4637.
Kittlesen, D.J., Chianese-Bullock, K.A., Yao, Z.Q., Braciale, T.J., and Hahn,
Y.S., (2000) Interaction between complement receptor gC1qR and hepatitis C virus
core protein inhibits T-lymphocyte proliferation, J Clin Invest. 106(10): p. 12391249.
Kleinberger, T. and Shenk, T., (1993) Adenovirus E4orf4 protein binds to protein
phosphatase 2A, and the complex down regulates E1A-enhanced junB transcription, J Virol. 67(12): p. 7556-7560.
Kobor, M.S., Archambault, J., Lester, W., Holstege, F.C., Gileadi, O., Jansma,
D.B., Jennings, E.G., Kouyoumdjian, F., Davidson, A.R., Young, R.A., and
Greenblatt, J., (1999) An unusual eukaryotic protein phosphatase required for
transcription by RNA polymerase II and CTD dephosphorylation in S. cerevisiae,
Mol Cell. 4(1): p. 55-62.
Kobor, M.S., Simon, L.D., Omichinski, J., Zhong, G., Archambault, J., and
Greenblatt, J., (2000) A motif shared by TFIIF and TFIIB mediates their interaction with the RNA polymerase II carboxy-terminal domain phosphatase Fcp1p in
Saccharomyces cerevisiae, Mol Cell Biol. 20(20): p. 7438-7449.
Komarnitsky, P., Cho, E.J., and Buratowski, S., (2000) Different phosphorylated
forms of RNA polymerase II and associated mRNA processing factors during transcription, Genes Dev. 14(19): p. 2452-2460.
Kornblihtt, A.R., de la Mata, M., Fededa, J.P., Munoz, M.J., and Nogues, G.,
(2004) Multiple links between transcription and splicing, RNA. 10: p. 1489-1498.
Kornitzer, D., Sharf, R., and Kleinberger, T., (2001) Adenovirus E4orf4 protein
induced PP2A-dependent growth arrest in Saccharomyces cerevisiae and interacts
with the anaphase-promoting complex/cyclosome, J Cell Biol. 154: p. 331-344.
Krainer, A.R., Conway, G.C., and Kozak, D., (1990) The essential pre-mRNA
splicing factor SF2 influences 5' splice site selection by activating proximal sites,
Cell. 62(1): p. 35-42.
Krainer, A.R., Conway, G.C., and Kozak, D., (1990) Purification and characterization of pre-mRNA splicing factor SF2 from HeLa cells, Genes Dev. 4(7): p.
1158-1171.
Krishnamurthy, S., He, X., Reyes-Reyes, M., Moore, C., and Hampsey, M.,
(2004) Ssu72 Is an RNA polymerase II CTD phosphatase, Mol Cell. 14(3): p. 387394.
Labourier, E., Allemand, E., Brand, S., Fostier, M., Tazi, J., and Bourbon,
H.M., (1999) Recognition of exonic splicing enhancer sequences by the Drosophila
splicing repressor RSF1, Nucleic Acids Res. 27(11): p. 2377-2386.
Labourier, E., Bourbon, H.M., Gallouzi, I.E., Fostier, M., Allemand, E., and
Tazi, J., (1999) Antagonism between RSF1 and SR proteins for both splice-site recognition in vitro and Drosophila development, Genes Dev. 13(6): p. 740-753.
61
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
117.
62
Lai, M.C., Kuo, H.W., Chang, W.C., and Tarn, W.Y., (2003) A novel splicing
regulator shares a nuclear import pathway with SR proteins, Embo J. 22(6): p.
1359-1369.
Lai, M.C., Lin, R.I., Huang, S.Y., Tsai, C.W., and Tarn, W.Y., (2000) A human
importin-beta family protein, transportin-SR2, interacts with the phosphorylated RS
domain of SR proteins, J Biol Chem. 275(11): p. 7950-7957.
Lee, K., Du, C., Horn, M., and Rabinow, L., (1996) Activity and autophosphorylation of LAMMER protein kinases, J Biol Chem. 271(44): p. 27299-27303.
Li, X.Y., Mantovani, R., Hooft van Huijsduijnen, R., Andre, I., Benoist, C., and
Mathis, D., (1992) Evolutionary variation of the CCAAT-binding transcription factor NF-Y, Nucleic Acids Res. 20(5): p. 1087-1091.
Liang, X., Shin, Y.C., Means, R.E., and Jung, J.U., (2004) Inhibition of interferon-mediated antiviral activity by murine gammaherpesvirus 68 latencyassociated M2 protein, J Virol. 78(22): p. 12416-12427.
Lichtenstein, D.L., Toth, K., Doronin, K., Tollefson, A.E., and Wold, W.S.,
(2004) Functions and mechanisms of action of the adenovirus E3 proteins, Int Rev
Immunol. 23(1-2): p. 75-111.
Lim, B.L., Reid, K.B., Ghebrehiwet, B., Peerschke, E.I., Leigh, L.A., and Preissner, K.T., (1996) The binding protein for globular heads of complement C1q,
gC1qR. Functional expression and characterization as a novel vitronectin binding
factor, J Biol Chem. 271(43): p. 26739-26744.
Lin, P.S., Dubois, M.F., and Dahmus, M.E., (2002) TFIIF-associating carboxylterminal domain phosphatase dephosphorylates phosphoserines 2 and 5 of RNA polymerase II, J Biol Chem. 277(48): p. 45949-45956.
Lindas, A.C. and Tomkinson, B., (2005) Identification and characterization of the
promoter for the gene encoding human tripeptidyl-peptidase II, Gene. 345(2): p.
249-257.
Longman, D., Johnstone, I.L., and Caceres, J.F., (2000) Functional characterization of SR and SR-related genes in Caenorhabditis elegans, Embo J. 19(7): p. 16251637.
Lopez, A.J., (1998) Alternative splicing of pre-mRNA: developmental consequences and mechanisms of regulation, Annu Rev Genet. 32: p. 279-305.
Lu, H., Reach, M.D., Minaya, E., and Young, C.S., (1997) The initiator element
of the adenovirus major late promoter has an important role in transcription initiation in vivo, J Virol. 71(1): p. 102-109.
Luo, Y., Yu, H., and Peterlin, B.M., (1994) Cellular protein modulates effects of
human immunodeficiency virus type 1 Rev, J Virol. 68(6): p. 3850-3856.
Lynch, K.W. and Maniatis, T., (1996) Assembly of specific SR protein complexes
on distinct regulatory elements of the Drosophila doublesex splicing enhancer,
Genes Dev. 10(16): p. 2089-2101.
Maity, S.N., Golumbek, P.T., Karsenty, G., and de Crombrugghe, B., (1988)
Selective activation of transcription by a novel CCAAT binding factor, Science.
241(4865): p. 582-585.
Majumdar, M., Meenakshi, J., Goswami, S.K., and Datta, K., (2002) Hyaluronan binding protein 1 (HABP1)/C1QBP/p32 is an endogenous substrate for MAP
kinase and is translocated to the nucleus upon mitogenic stimulation, Biochem Biophys Res Commun. 291(4): p. 829-837.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
128.
129.
130.
131.
132.
Mallick, J. and Datta, K., (2005) HABP1/p32/gC1qR induces aberrant growth and
morphology in Schizosaccharomyces pombe through its N-terminal alpha helix,
Exp Cell Res. 309(2): p. 250-263.
Mandal, S.S., Cho, H., Kim, S., Cabane, K., and Reinberg, D., (2002) FCP1, a
phosphatase specific for the heptapeptide repeat of the largest subunit of RNA polymerase II, stimulates transcription elongation, Mol Cell Biol. 22(21): p. 75437552.
Mandal, S.S., Chu, C., Wada, T., Handa, H., Shatkin, A.J., and Reinberg, D.,
(2004) Functional interactions of RNA-capping enzyme with factors that positively
and negatively regulate promoter escape by RNA polymerase II, Proc Natl Acad Sci
U S A. 101(20): p. 7572-7577.
Mantovani, R., (1998) A survey of 178 NF-Y binding CCAAT boxes, Nucleic Acids
Res. 26(5): p. 1135-1143.
Marcellus, R., Chan, D., Paquette, S., Thirwell, D., Boivin, D., and Branton,
P.E., (2000) Induction of p53-independent apoptosis by the adenovirus E4orf4 protein requires binding to the Balpha subunit of protein phosphatase PP2A, J Virol.
74: p. 7869-7877.
Marcellus, R., Lavoie, D., Boivin, D., Shore, G., Ketner, G., and Branton, P.E.,
(1998) The early region 4 orf4 protein of human adenovirus type 5 induces p53independent cell death by apoptosis, J Virol. 72: p. 7144-7153.
Marschall, M., Marzi, A., aus dem Siepen, P., Jochmann, R., Kalmer, M.,
Auerochs, S., Lischka, P., Leis, M., and Stamminger, T., (2005) Cellular p32 recruits cytomegalovirus kinase pUL97 to redistribute the nuclear lamina, J Biol
Chem. 280(39): p. 33357-33367.
Matthews, D.A. and Russell, W.C., (1998) Adenovirus core protein V interacts
with p32-a protein which is associated with both the mitochondria and the nucleus,
J Gen Virol. 79(Pt 7): p. 1677-1685.
McCracken, S., Fong, N., Rosonina, E., Yankulov, K., Brothers, G., Siderovski,
D., Hessel, A., Foster, S., Shuman, S., and Bentley, D.L., (1997) 5'-Capping enzymes are targeted to pre-mRNA by binding to the phosphorylated carboxy-terminal
domain of RNA polymerase II, Genes Dev. 11(24): p. 3306-3318.
McCracken, S., Fong, N., Yankulov, K., Ballantyne, S., Pan, G., Greenblatt, J.,
Patterson, S.D., Wickens, M., and Bentley, D.L., (1997) The C-terminal domain
of RNA polymerase II couples mRNA processing to transcription, Nature.
385(6614): p. 357-361.
Misteli, T., Caceres, J.F., and Spector, D.L., (1997) The dynamics of a pre-mRNA
splicing factor in living cells, Nature. 387(6632): p. 523-527.
Misteli, T. and Spector, D.L., (1999) RNA polymerase II targets pre-mRNA splicing factors to transcription sites in vivo, Mol Cell. 3(6): p. 697-705.
Miyamoto, N.G., Moncollin, V., Egly, J.M., and Chambon, P., (1985) Specific
interaction between a transcription factor and the upstream element of the adenovirus-2 major late promoter, Embo J. 4(13A): p. 3563-3570.
Modrek, B. and Lee, C., (2002) A genomic view of alternative splicing, Nat Genet.
30(1): p. 13-19.
Mohan, K.V., Ghebrehiwet, B., and Atreya, C.D., (2002) The N-terminal conserved domain of rubella virus capsid interacts with the C-terminal region of cellu-
63
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
64
lar p32 and overexpression of p32 enhances the viral infectivity, Virus Res. 85(2):
p. 151-161.
Moisan, A., Larochelle, C., Guillemette, B., and Gaudreau, L., (2004) BRCA1
can modulate RNA polymerase II carboxy-terminal domain phosphorylation levels,
Mol Cell Biol. 24(16): p. 6947-6956.
Morris, D.P., Phatnani, H.P., and Greenleaf, A.L., (1999) Phospho-carboxylterminal domain binding and the role of a prolyl isomerase in pre-mRNA 3'-End
formation, J Biol Chem. 274(44): p. 31583-31587.
Muller, U., Kleinberger, T., and Shenk, T., (1992) Adenovirus E4orf4 protein
reduces phosphorylation of c-Fos and E1A proteins while simultaneously reducing
the level of AP-1, J Virol. 66(10): p. 5867-5878.
Muta, T., Kang, D., Kitajima, S., Fujiwara, T., and Hamasaki, N., (1997) p32
protein, a splicing factor 2-associated protein, is localized in mitochondrial matrix
and is functionally important in maintaining oxidative phosphorylation, J Biol
Chem. 272(39): p. 24363-24370.
Mühlemann, O., Yue, B.G., Petersen-Mahrt, S., and Akusjärvi, G., (2000) A
novel type of splicing enhancer regulating adenovirus pre-mRNA splicing, Mol Cell
Biol. 20(7): p. 2317-2325.
Neubauer, G., Gottschalk, A., Fabrizio, P., Seraphin, B., Luhrmann, R., and
Mann, M., (1997) Identification of the proteins of the yeast U1 small nuclear ribonucleoprotein complex by mass spectrometry, Proc Natl Acad Sci U S A. 94(2): p.
385-390.
Neugebauer, K.M., (2002) On the importance of being co-transcriptional, J Cell
Sci. 115(Pt 20): p. 3865-3871.
Nguyen, T., Ghebrehiwet, B., and Peerschke, E.I., (2000) Staphylococcus aureus
protein A recognizes platelet gC1qR/p33: a novel mechanism for staphylococcal interactions with platelets, Infect Immun. 68(4): p. 2061-2068.
Ni, Z., Schwartz, B.E., Werner, J., Suarez, J.R., and Lis, J.T., (2004) Coordination of transcription, RNA processing, and surveillance by P-TEFb kinase on heat
shock genes, Mol Cell. 13(1): p. 55-65.
Nikolakaki, E., Simos, G., Georgatos, S.D., and Giannakouros, T., (1996) A
nuclear envelope-associated kinase phosphorylates arginine-serine motifs and
modulates interactions between the lamin B receptor and other nuclear proteins, J
Biol Chem. 271(14): p. 8365-8372.
Nogues, G., Kadener, S., Cramer, P., Bentley, D., and Kornblihtt, A.R., (2002)
Transcriptional activators differ in their abilities to control alternative splicing, J
Biol Chem. 277(45): p. 43110-43114.
Nogues, G., Munoz, M.J., and Kornblihtt, A.R., (2003) Influence of pol II processivity on alternative splicing depends on splice site strength, J Biol Chem.
Nordqvist, K., Öhman, K., and Akusjärvi, G., (1994) Human adenovirus encodes
two proteins which have opposite effects on accumulation of alternatively spliced
mRNAs, Mol Cell Biol. 14(1): p. 437-445.
Okagaki, T., Nakamura, A., Suzuki, T., Ohmi, K., and Kohama, K., (2000)
Assembly of smooth muscle myosin by the 38k protein, a homologue of a subunit of
pre-mRNA splicing factor-2, J Cell Biol. 148(4): p. 653-663.
Orphanides, G., Lagrange, T., and Reinberg, D., (1996) The general transcription factors of RNA polymerase II, Genes Dev. 10(21): p. 2657-2683.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
Palancade, B. and Bensaude, O., (2003) Investigating RNA polymerase II carboxyl-terminal domain (CTD) phosphorylation, Eur J Biochem. 270(19): p. 38593870.
Palancade, B., Marshall, N.F., Tremeau-Bravard, A., Bensaude, O., Dahmus,
M.E., and Dubois, M.F., (2004) Dephosphorylation of RNA polymerase II by
CTD-phosphatase FCP1 is inhibited by phospho-CTD associating proteins, J Mol
Biol. 335(2): p. 415-424.
Pallas, D., Shahrik, L., Martin, B., Jaspers, S., Miller, T., Brautigan, D., and
Roberts, T., (1990) Polyoma small and middle T antigen and SV40 small t antigen
form stable complexes with protein phophatase 2A, Cell. 60: p. 167-176.
Perry, M. and Levine, A., (1993) Tumor-supressor p53 and the cell cycle, Curr
Opin Genet Develop. 3: p. 50-54.
Petersen-Mahrt, S.K., Estmer, C., Öhrmalm, C., Matthews, D.A., Russell,
W.C., and Akusjärvi, G., (1999) The splicing factor-associated protein, p32, regulates RNA splicing by inhibiting ASF/SF2 RNA binding and phosphorylation,
EMBO J. 18(4): p. 1014-1024.
Powell, D.M., Amaral, M.C., Wu, J.Y., Maniatis, T., and Greene, W.C., (1997)
HIV Rev-dependent binding of SF2/ASF to the Rev response element: possible role
in Rev-mediated inhibition of HIV RNA splicing, Proc Natl Acad Sci U S A. 94(3):
p. 973-978.
Prasad, J., Colwill, K., Pawson, T., and Manley, J.L., (1999) The protein kinase
Clk/Sty directly modulates SR protein activity: both hyper- and hypophosphorylation inhibit splicing, Mol Cell Biol. 19(10): p. 6991-7000.
Prasad, J. and Manley, J.L., (2003) Regulation and substrate specificity of the SR
protein kinase Clk/Sty, Mol Cell Biol. 23(12): p. 4139-4149.
Querido, E., Blanchette, P., Yan, Q., Kamura, T., Morrison, M., Boivin, D.,
Kaelin, W.G., Conaway, R.C., Conaway, J.W., and Branton, P.E., (2001) Degradation of p53 by adenovirus E4orf6 and E1B55K proteins occurs via a novel
mechanism involving a Cullin-containing complex, Genes Dev. 15(23): p. 31043117.
Ramanathan, Y., Rajpara, S.M., Reza, S.M., Lees, E., Shuman, S., Mathews,
M.B., and Pe'ery, T., (2001) Three RNA polymerase II carboxyl-terminal domain
kinases display distinct substrate preferences, J Biol Chem. 276(14): p. 1091310920.
Ramji, D.P. and Foka, P., (2002) CCAAT/enhancer-binding proteins: structure,
function and regulation, Biochem J. 365(Pt 3): p. 561-575.
Rawle, F.C., Tollefson, A.E., Wold, W.S., and Gooding, L.R., (1989) Mouse
anti-adenovirus cytotoxic T lymphocytes. Inhibition of lysis by E3 gp19K but not E3
14.7K, J Immunol. 143(6): p. 2031-2037.
Reach, M., Babiss, L.E., and Young, C.S., (1990) The upstream factor-binding
site is not essential for activation of transcription from the adenovirus major late
promoter, J Virol. 64(12): p. 5851-5860.
Roberts, G.C. and Smith, C.W., (2002) Alternative splicing: combinatorial output
from the genome, Curr Opin Chem Biol. 6(3): p. 375-383.
Robles-Flores, M., Rendon-Huerta, E., Gonzalez-Aguilar, H., MendozaHernandez, G., Islas, S., Mendoza, V., Ponce-Castaneda, M.V., GonzalezMariscal, L., and Lopez-Casillas, F., (2002) p32 (gC1qBP) is a general protein
65
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
66
kinase C (PKC)-binding protein; interaction and cellular localization of P32-PKC
complexes in ray hepatocytes, J Biol Chem. 277(7): p. 5247-5255.
Rodriguez, C.R., Cho, E.J., Keogh, M.C., Moore, C.L., Greenleaf, A.L., and
Buratowski, S., (2000) Kin28, the TFIIH-associated carboxy-terminal domain
kinase, facilitates the recruitment of mRNA processing machinery to RNA polymerase II, Mol Cell Biol. 20(1): p. 104-112.
Rossi, F., Labourier, E., Forne, T., Divita, G., Derancourt, J., Riou, J.F., Antoine, E., Cathala, G., Brunel, C., and Tazi, J., (1996) Specific phosphorylation of
SR proteins by mammalian DNA topoisomerase I, Nature. 381(6577): p. 80-82.
Russell, W.C.a.B., M, Adenoviruses (Adenoviridae): Animal viruses. Encyclopedia
of virology, ed. G.A. Webster RG. 1999, London: Academic Press. 14-21.
Santoro, C., Mermod, N., Andrews, P.C., and Tjian, R., (1988) A family of
human CCAAT-box-binding proteins active in transcription and DNA replication:
cloning and expression of multiple cDNAs, Nature. 334(6179): p. 218-224.
Sawadogo, M. and Roeder, R.G., (1985) Interaction of a gene-specific transcription factor with the adenovirus major late promoter upstream of the TATA box region, Cell. 43(1): p. 165-175.
Schroeder, S.C., Schwer, B., Shuman, S., and Bentley, D., (2000) Dynamic
association of capping enzymes with transcribing RNA polymerase II, Genes Dev.
14(19): p. 2435-2440.
Sciabica, K.S., Dai, Q.J., and Sandri-Goldin, R.M., (2003) ICP27 interacts with
SRPK1 to mediate HSV splicing inhibition by altering SR protein phosphorylation,
Embo J. 22(7): p. 1608-1619.
Seytter, T., Lottspeich, F., Neupert, W., and Schwarz, E., (1998) Mam33p, an
oligomeric, acidic protein in the mitochondrial matrix of Saccharomyces cerevisiae
is related to the human complement receptor gC1q-R, Yeast. 14(4): p. 303-310.
Shaw, A.R. and Ziff, E.B., (1980) Transcripts from the adenovirus-2 major late
promoter yield a single early family of 3' coterminal mRNAs and five late families,
Cell. 22(3): p. 905-916.
Shaw, G., Morse, S., Ararat, M., and Graham, F.L., (2002) Preferential transformation of human neuronal cells by human adenoviruses and the origin of HEK
293 cells, Faseb J. 16(8): p. 869-871.
Shen, H. and Green, M.R., (2004) A pathway of sequential arginine-serine-rich
domain-splicing signal interactions during mammalian spliceosome assembly, Mol
Cell. 16(3): p. 363-373.
Shen, H., Kan, J.L., and Green, M.R., (2004) Arginine-serine-rich domains
bound at splicing enhancers contact the branchpoint to promote prespliceosome assembly, Mol Cell. 13(3): p. 367-376.
Shenk, Adenoviridae: The viruses and their replication. 3rd ed ed. Fields Virology,
ed. B. Fields, D. Knipe, and P. Howley. Vol. 2. 1996, Philadelphia: LippincottRaven. 2111-2171.
Shtrichman, R. and Kleinberger, T., (1998) Adenovirus type 5 E4 open reading
frame protein induces apoptosis in transformed cells, J Virol. 72: p. 2975-2982.
Shtrichman, R., Sharf, R., Barr, T., Dobner, T., and Kleinberger, T., (1999)
Induction of apoptosis by E4orf4 protein is specific to transformed cells and requires an interaction with protein phosphatase 2A, Proc Natl Acad Sci U S A. 96:
p. 10080-10085.
178.
179.
180.
181.
182.
183.
184.
185.
186.
187.
188.
189.
190.
191.
192.
193.
Shukla, R.R., Marques, S.M., Kimmel, P.L., and Kumar, A., (1996) Human
chromosome 6- and 11-encoded factors support human immunodeficiency virus
type 1 Rev function in A9 cells, J Virol. 70(12): p. 9064-9068.
Simos, G. and Georgatos, S.D., (1994) The lamin B receptor-associated protein
p34 shares sequence homology and antigenic determinants with the splicing factor
2-associated protein p32, FEBS Lett. 346(2-3): p. 225-228.
Smale, S.T. and Baltimore, D., (1989) The "initiator" as a transcription control
element, Cell. 57(1): p. 103-113.
Smale, S.T., Schmidt, M.C., Berk, A.J., and Baltimore, D., (1990) Transcriptional activation by Sp1 as directed through TATA or initiator: specific requirement
for mammalian transcription factor IID, Proc Natl Acad Sci U S A. 87(12): p.
4509-4513.
Song, B. and Young, C.S., (1998) Functional analysis of the CAAT box in the
major late promoter of the subgroup C human adenoviruses, J Virol. 72(4): p.
3213-3220.
Starr, D.B. and Hawley, D.K., (1991) TFIID binds in the minor groove of the
TATA box, Cell. 67(6): p. 1231-1240.
Storz, P., Hausser, A., Link, G., Dedio, J., Ghebrehiwet, B., Pfizenmaier, K.,
and Johannes, F.J., (2000) Protein kinase C [micro] is regulated by the multifunctional chaperon protein p32, J Biol Chem. 275(32): p. 24601-24607.
Sun, Q., Mayeda, A., Hampson, R.K., Krainer, A.R., and Rottman, F.M.,
(1993) General splicing factor SF2/ASF promotes alternative splicing by binding to
an exonic splicing enhancer, Genes Dev. 7(12B): p. 2598-2608.
Sunayama, J., Ando, Y., Itoh, N., Tomiyama, A., Sakurada, K., Sugiyama, A.,
Kang, D., Tashiro, F., Gotoh, Y., Kuchino, Y., and Kitanaka, C., (2004) Physical and functional interaction between BH3-only protein Hrk and mitochondrial
pore-forming protein p32, Cell Death Differ. 11(7): p. 771-781.
Swaminathan, S. and Thimmmapaya, B., Regulation of Adenovirus E2 transcription unit. The Molecular repetoire of Adenovirus III, Biology and pathogenesia, ed.
W. Doerfler and P. Böhm. 1995, Berlin Heidelberg: Springer. 177-194.
Tacke, R., Chen, Y., and Manley, J.L., (1997) Sequence-specific RNA binding by
an SR protein requires RS domain phosphorylation: creation of an SRp40-specific
splicing enhancer, Proc Natl Acad Sci U S A. 94(4): p. 1148-1153.
Tacke, R. and Manley, J.L., (1999) Functions of SR and Tra2 proteins in premRNA splicing regulation, Proc Soc Exp Biol Med. 220(2): p. 59-63.
Tacke, R., Tohyama, M., Ogawa, S., and Manley, J.L., (1998) Human Tra2
proteins are sequence-specific activators of pre-mRNA splicing, Cell. 93(1): p. 139148.
Tange, T.O., Jensen, T.H., and Kjems, J., (1996) In vitro interaction between
human immunodeficiency virus type 1 Rev protein and splicing factor ASF/SF2associated protein, p32, J Biol Chem. 271(17): p. 10066-10072.
Tollefson, A.E., Ryerse, J.S., Scaria, A., Hermiston, T.W., and Wold, W.S.,
(1996) The E3-11.6-kDa adenovirus death protein (ADP) is required for efficient
cell death: characterization of cells infected with adp mutants, Virology. 220(1): p.
152-162.
Tollefson, A.E., Scaria, A., Hermiston, T.W., Ryerse, J.S., Wold, L.J., and
Wold, W.S., (1996) The adenovirus death protein (E3-11.6K) is required at very
67
194.
195.
196.
197.
198.
199.
200.
201.
202.
203.
204.
205.
206.
207.
208.
68
late stages of infection for efficient cell lysis and release of adenovirus from infected
cells, J Virol. 70(4): p. 2296-2306.
Tomko, R.P., Xu, R., and Philipson, L., (1997) HCAR and MCAR: the human and
mouse cellular receptors for subgroup C adenoviruses and group B coxsackieviruses, Proc Natl Acad Sci U S A. 94(7): p. 3352-3356.
Trentin, J., Yabe, Y., and Taylor, G., (1962) The quest for human cancer viruses,
Science. 137: p. 835-841.
Valcarcel, J., Gaur, R.K., Singh, R., and Green, M.R., (1996) Interaction of
U2AF65 RS region with pre-mRNA branch point and promotion of base pairing
with U2 snRNA [corrected], Science. 273(5282): p. 1706-1709.
Walter, G., Ruediger, R., Slaugther, C., and Mumby, M., (1990) Association of
protein phosphatase 2 A with polyoma virus medium tumour antigen, Proc Natl
Acad Sci U S A. 87: p. 2521-2525.
Van Den Brulle, J., Steidl, S., and Brakhage, A.A., (1999) Cloning and characterization of an Aspergillus nidulans gene involved in the regulation of penicillin
biosynthesis, Appl Environ Microbiol. 65(12): p. 5222-5228.
van Oostrum, J. and Burnett, R.M., (1985) Molecular composition of the adenovirus type 2 virion, J Virol. 56(2): p. 439-448.
Van Scoy, S., Watakabe, I., Krainer, A.R., and Hearing, J., (2000) Human p32:
a coactivator for Epstein-Barr virus nuclear antigen-1- mediated transcriptional
activation and possible role in viral latent cycle DNA replication, Virology. 275(1):
p. 145-157.
Wang, H.Y., Lin, W., Dyck, J.A., Yeakley, J.M., Songyang, Z., Cantley, L.C.,
and Fu, X.D., (1998) SRPK2: a differentially expressed SR protein-specific kinase
involved in mediating the interaction and localization of pre-mRNA splicing factors
in mammalian cells, J Cell Biol. 140(4): p. 737-750.
Wang, Y., Finan, J.E., Middeldorp, J.M., and Hayward, S.D., (1997) P32/TAP,
a cellular protein that interacts with EBNA-1 of Epstein-Barr virus, Virology.
236(1): p. 18-29.
Velazquez-Dones, A., Hagopian, J.C., Ma, C.T., Zhong, X.Y., Zhou, H., Ghosh,
G., Fu, X.D., and Adams, J.A., (2005) Mass spectrometric and kinetic analysis of
ASF/SF2 phosphorylation by SRPK1 and Clk/Sty, J Biol Chem. 280(50): p. 4176141768.
White, E., (2001) Regulation of the cell cycle and apoptosis by the oncogenes of
adenovirus, Oncogene. 20(54): p. 7836-7846.
Wickman, T., Mathias, P., Cheresh, D., and Nemerow, G., (1993) Integrins
alpha v beta 3 and alpha v beta 5 promote adenovirus internalization but not virus
attatchment, Cell. 73: p. 309-319.
Wilcox, C.B., Rossettini, A., and Hanes, S.D., (2004) Genetic interactions with Cterminal domain (CTD) kinases and the CTD of RNA Pol II suggest a role for ESS1
in transcription initiation and elongation in Saccharomyces cerevisiae, Genetics.
167(1): p. 93-105.
Vliet, V.d., (1995) Adenovirus DNA replication, Curr Top Microbiol Immunol.
199: p. 1-30.
Wu, X., Wilcox, C.B., Devasahayam, G., Hackett, R.L., Arevalo-Rodriguez, M.,
Cardenas, M.E., Heitman, J., and Hanes, S.D., (2000) The Ess1 prolyl isomerase
209.
210.
211.
212.
213.
214.
215.
216.
217.
218.
219.
220.
221.
is linked to chromatin remodeling complexes and the general transcription machinery, Embo J. 19(14): p. 3727-3738.
Xiao, S.H. and Manley, J.L., (1997) Phosphorylation of the ASF/SF2 RS domain
affects both protein-protein and protein-RNA interactions and is necessary for
splicing, Genes Dev. 11(3): p. 334-344.
Xiao, S.H. and Manley, J.L., (1998) Phosphorylation-dephosphorylation differentially affects activities of splicing factor ASF/SF2, Embo J. 17(21): p. 6359-6367.
Xu, Y.X., Hirose, Y., Zhou, X.Z., Lu, K.P., and Manley, J.L., (2003) Pin1 modulates the structure and function of human RNA polymerase II, Genes Dev. 17: p.
2765-2776.
Yamaguchi, Y., Takagi, T., Wada, T., Yano, K., Furuya, A., Sugimoto, S.,
Hasegawa, J., and Handa, H., (1999) NELF, a multisubunit complex containing
RD, cooperates with DSIF to repress RNA polymerase II elongation, Cell. 97(1): p.
41-51.
Yanagida, M., Hayano, T., Yamauchi, Y., Shinkawa, T., Natsume, T., Isobe, T.,
and Takahashi, N., (2004) Human fibrillarin forms a sub-complex with splicing
factor 2-associated p32, protein arginine methyltransferases, and tubulins alpha 3
and beta 1 that is independent of its association with preribosomal ribonucleoprotein complexes, J Biol Chem. 279(3): p. 1607-1614.
Yeo, M., Lin, P.S., Dahmus, M.E., and Gill, G.N., (2003) A novel RNA polymerase II C-terminal domain phosphatase that preferentially dephosphorylates serine 5, J Biol Chem. 278(28): p. 26078-26085.
Yew, P.R., Liu, X., and Berk, A.J., (1994) Adenovirus E1B oncoprotein tethers a
transcriptional repression domain to p53, Genes Dev. 8(2): p. 190-202.
Yu, L., Loewenstein, P.M., Zhang, Z., and Green, M., (1995) In vitro interaction
of the human immunodeficiency virus type 1 Tat transactivator and the general
transcription factor TFIIB with the cellular protein TAP, J Virol. 69(5): p. 30173023.
Yu, L., Zhang, Z., Loewenstein, P.M., Desai, K., Tang, Q., Mao, D., Symington,
J.S., and Green, M., (1995) Molecular cloning and characterization of a cellular
protein that interacts with the human immunodeficiency virus type 1 Tat transactivator and encodes a strong transcriptional activation domain, J Virol. 69(5): p.
3007-3016.
Yue, B.G. and Akusjarvi, G., (1999) A downstream splicing enhancer is essential
for in vitro pre-mRNA splicing, FEBS Lett. 451(1): p. 10-14.
Zenzie-Gregory, B., Khachi, A., Garraway, I.P., and Smale, S.T., (1993)
Mechanism of initiator-mediated transcription: evidence for a functional interaction between the TATA-binding protein and DNA in the absence of a specific recognition sequence, Mol Cell Biol. 13(7): p. 3841-3849.
Zheng, Y.H., Yu, H.F., and Peterlin, B.M., (2003) Human p32 protein relieves a
post-transcriptional block to HIV replication in murine cells, Nat Cell Biol. 5(7): p.
611-618.
Zhou, M., Halanski, M.A., Radonovich, M.F., Kashanchi, F., Peng, J., Price,
D.H., and Brady, J.N., (2000) Tat modifies the activity of CDK9 to phosphorylate
serine 5 of the RNA polymerase II carboxyl-terminal domain during human immunodeficiency virus type 1 transcription, Mol Cell Biol. 20(14): p. 5077-5086.
69
Acta Universitatis Upsaliensis
Digital Comprehensive Summaries of Uppsala Dissertations
from the Faculty of Medicine 139
Editor: The Dean of the Faculty of Medicine
A doctoral dissertation from the Faculty of Medicine, Uppsala
University, is usually a summary of a number of papers. A few
copies of the complete dissertation are kept at major Swedish
research libraries, while the summary alone is distributed
internationally through the series Digital Comprehensive
Summaries of Uppsala Dissertations from the Faculty of
Medicine. (Prior to January, 2005, the series was published
under the title “Comprehensive Summaries of Uppsala
Dissertations from the Faculty of Medicine”.)
Distribution: publications.uu.se
urn:nbn:se:uu:diva-6794
ACTA
UNIVERSITATIS
UPSALIENSIS
UPPSALA
2006