signaling to the dead box—regulation of dead-box p68 rna helicase by protein phosphorylations

10
Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations Liuqing Yang, Chunru Lin, Zhi-Ren Liu * Department of Biology, Georgia State University, University Plaza, Atlanta, GA 30303, USA Received 22 January 2005; accepted 4 March 2005 Available online 31 May 2005 Abstract P68 nuclear RNA helicase is essential for normal cell growth. The protein plays a very important role in cell development and proliferation. However, the molecular mechanism by which the p68 functions in cell developmental program is not clear. We previously observed that bacterially expressed his-p68 was phosphorylated at multiple sites including serine/threonine and tyrosine [1] [L. Yang, Z.R. Liu, Protein Expr. Purif. 35 (2004) 327]. Here we report that p68 RNA helicase is phosphorylated at tyrosine residue(s) in HeLa cells. Phosphorylation of p68 at threonine or tyrosine residues responds differently to tumor necrosis factor alpha (TNF-a) induced cell signal. Kinase inhibition and in vitro kinase assays demonstrate that p68 RNA helicase is a cellular target of p38 MAP kinase. Phosphorylation of p68 affects the ATPase and RNA unwinding activities of the protein. In addition, we demonstrate here that phosphorylation of p68 RNA helicase controls the function of the protein in the pre-mRNA splicing process. Interestingly, phosphorylation at different amino acid residues exhibits different regulatory effects. The data suggest that function(s) of p68 RNA helicase may be subjected to the regulation of multiple cell signal pathways. D 2005 Elsevier Inc. All rights reserved. Keywords: DEAD box; RNA helicase; TNF-a; Protein phosphorylation; ATPase 1. Introduction Modulation of dynamic RNA structures and complex RNA-protein interactions is essential for every biological process that is involved in RNA metabolism. It is generally believed that a super-family of so-called DEAD or DExH box RNA helicases executes this essential role in the cells [2–5]. RNA helicases use the energy derived from ATP hydrolysis to dissociate the RNA–RNA and/or RNA– protein interactions [6–8]. The DEAD or DExH box RNA helicases are characterized by a core region of 290–360 amino acids (helicase core) that consist of eight conserved sequence motifs. The proteins also have variable C- and N- terminal. RNA helicases are involved in a wide spectrum of biological processes of RNA metabolism. Thus, an individ- ual RNA helicase must be highly substrate specific both in time and space in cells to perform its own specific task. The nuclear p68 RNA helicase is a prototypical DEAD box family of RNA helicase [9,10]. As an early example of a cellular RNA helicase, the ATPase and the RNA unwinding activities of p68 RNA helicase were documented with the protein that was purified from human 293 cells [11–13]. P68 RNA helicase plays a very important role in cell proliferation and development [14,15]. The protein is expressed in all dividing cells of different vertebrates [10,15]. The protein shows clear cell cycle-related local- ization in the nucleus. Related to the role of p68 RNA helicase in cell proliferation, the protein was shown to play a critical role in the tumorigenesis process [16–18]. Never- theless, the biological functions of the protein at a molecular level are not well understood [19]. It has been suggested that p68 RNA helicase might be involved in transcription regulation [20–23] and DNA methylation pathways [24]. Recently, the experiments carried out in our laboratory 0898-6568/$ - see front matter D 2005 Elsevier Inc. All rights reserved. doi:10.1016/j.cellsig.2005.03.008 * Corresponding author. E-mail address: [email protected] (Z.-R. Liu). Cellular Signalling 17 (2005) 1495 – 1504 www.elsevier.com/locate/cellsig

Upload: liuqing-yang

Post on 04-Sep-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

www.elsevier.com/locate/cellsig

Cellular Signalling 17 (

Signaling to the DEAD box—Regulation of DEAD-box p68 RNA

helicase by protein phosphorylations

Liuqing Yang, Chunru Lin, Zhi-Ren Liu*

Department of Biology, Georgia State University, University Plaza, Atlanta, GA 30303, USA

Received 22 January 2005; accepted 4 March 2005

Available online 31 May 2005

Abstract

P68 nuclear RNA helicase is essential for normal cell growth. The protein plays a very important role in cell development and

proliferation. However, the molecular mechanism by which the p68 functions in cell developmental program is not clear. We previously

observed that bacterially expressed his-p68 was phosphorylated at multiple sites including serine/threonine and tyrosine [1] [L. Yang, Z.R.

Liu, Protein Expr. Purif. 35 (2004) 327]. Here we report that p68 RNA helicase is phosphorylated at tyrosine residue(s) in HeLa cells.

Phosphorylation of p68 at threonine or tyrosine residues responds differently to tumor necrosis factor alpha (TNF-a) induced cell signal.

Kinase inhibition and in vitro kinase assays demonstrate that p68 RNA helicase is a cellular target of p38 MAP kinase. Phosphorylation of

p68 affects the ATPase and RNA unwinding activities of the protein. In addition, we demonstrate here that phosphorylation of p68 RNA

helicase controls the function of the protein in the pre-mRNA splicing process. Interestingly, phosphorylation at different amino acid residues

exhibits different regulatory effects. The data suggest that function(s) of p68 RNA helicase may be subjected to the regulation of multiple cell

signal pathways.

D 2005 Elsevier Inc. All rights reserved.

Keywords: DEAD box; RNA helicase; TNF-a; Protein phosphorylation; ATPase

1. Introduction

Modulation of dynamic RNA structures and complex

RNA-protein interactions is essential for every biological

process that is involved in RNA metabolism. It is generally

believed that a super-family of so-called DEAD or DExH

box RNA helicases executes this essential role in the cells

[2–5]. RNA helicases use the energy derived from ATP

hydrolysis to dissociate the RNA–RNA and/or RNA–

protein interactions [6–8]. The DEAD or DExH box RNA

helicases are characterized by a core region of 290–360

amino acids (helicase core) that consist of eight conserved

sequence motifs. The proteins also have variable C- and N-

terminal. RNA helicases are involved in a wide spectrum of

biological processes of RNA metabolism. Thus, an individ-

0898-6568/$ - see front matter D 2005 Elsevier Inc. All rights reserved.

doi:10.1016/j.cellsig.2005.03.008

* Corresponding author.

E-mail address: [email protected] (Z.-R. Liu).

ual RNA helicase must be highly substrate specific both in

time and space in cells to perform its own specific task.

The nuclear p68 RNA helicase is a prototypical DEAD

box family of RNA helicase [9,10]. As an early example of

a cellular RNA helicase, the ATPase and the RNA

unwinding activities of p68 RNA helicase were documented

with the protein that was purified from human 293 cells

[11–13]. P68 RNA helicase plays a very important role in

cell proliferation and development [14,15]. The protein is

expressed in all dividing cells of different vertebrates

[10,15]. The protein shows clear cell cycle-related local-

ization in the nucleus. Related to the role of p68 RNA

helicase in cell proliferation, the protein was shown to play a

critical role in the tumorigenesis process [16–18]. Never-

theless, the biological functions of the protein at a molecular

level are not well understood [19]. It has been suggested that

p68 RNA helicase might be involved in transcription

regulation [20–23] and DNA methylation pathways [24].

Recently, the experiments carried out in our laboratory

2005) 1495 – 1504

Page 2: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–15041496

demonstrated that p68 RNA helicase is an essential human

splicing factor in vitro that plays a role in unwinding the

transient U1 :5V splice site duplex [25,26]. Consistently,

other research laboratories also suggested that p68 RNA

helicase has a functional role in the pre-mRNA splicing

process [27–30].

An interesting question is how the biological function(s)

of p68 RNA helicase is linked to cell growth signals. It was

noted early that p68 can be phosphorylated in vitro by

Protein Kinase C (PKC) [31]. Moreover, Akileswaran and

colleagues found that a kinase anchoring protein AKAP95 is

associated with p68 RNA helicase [32]. The AKAP95

functions as an anchor for cAMP-dependent kinase in the

cell nuclear matrix. These observations strongly argued that

p68 RNA helicase might be a downstream target for

multiple protein kinases. Thus, elucidation of the molecular

mechanism by which the function and activities of p68

RNA helicase are regulated by protein phosphorylation will

provide a link for the role(s) of p68 RNA helicase in cell

development programs.

We previously reported that bacterially expressed p68

RNA helicase is phosphorylated at serine/threonine and

tyrosine residues [1]. Here, we report that cellular p68

RNA helicase is phosphorylated at tyrosine residues.

Phosphorylation of p68 is coupled to TNF-induced signal

pathway. In response to TNF stimuli, p68 change its

phosphorylation status, including decrease in tyrosyl

phosphorylation and increase in threonine phosphorylation.

Phosphorylation affects the biochemical activities of the

protein. Our data also demonstrated that phosphorylations

on serine/threonine or tyrosine residues inhibit each other.

Interestingly, the different phosphorylation exhibits differ-

ent regulatory effects on its function and biochemical

activities. The data suggest that function(s) of p68 RNA

helicase is subject to the regulation of different signal

pathways.

2. Materials and methods

2.1. Recombinant p68 RNA helicase expression

Expression and purification of recombinant p68 and

other recombinant proteins were carried out by the same

experimental procedures as previously described [33,34].

2.2. Dephosphorylation and phosphorylation of p68 RNA

helicase

Dephosphorylation of recombinant p68 RNA helicase or

the immunoprecipitated cellular protein was carried out with

PP2A and/or PTP1B. Approximately 5 Ag of protein was

incubated with 4 U of the phosphatase in manufacture

suggested buffer conditions in total volume of 50 Al at 30 -Cfor 90 min. The reactions were either directly used for

ATPase, RNA unwinding and splicing assays or used for re-

phosphorylation by the procedure described in the next

paragraph.

Protein phosphatase inhibitors, okadaic acid for PP2A

and vanadium for PTP1B, were added to the above

dephosphorylation reactions. PKC or v-Abl kinase, was

added to the reaction mixture. ATP was added to a final

concentration of 100 mM. Additional 10 ACi of [g-32P]ATPwere added for 32P labeling of the proteins as indicated. The

phosphorylation reactions were further incubated at 30 -Cfor 90 min. The re-phosphorylated proteins were directly

used for ATPase, RNA unwinding and splicing assays or

Western blot assays.

2.3. Cell culture and nuclear extracts

HeLa S3 cells were grown in Ham’s F12K medium

supplement with 10% fetal bovine serum, penicillin (100 U/

ml), and streptomycin (100 Ag/ml). The nuclear extracts from

the HeLa cells were made using nuclear extract preparation

kit (Active-motif). In the cases of treatment of cells with

TNF-a, TNF-a (15 ng/ml) were added to cell culture

medium. After the treatment for indicated time, the cells

were immediately harvest for nuclear extracts preparations.

2.4. Immunoprecipitation and Western blot

Immunoprecipitation experiments were performed as

described in previous studies [26,35]. HeLa nuclear extracts

(80 Al) were diluted to 200 Al with NETS buffer (150 mM

NaCl, 50 mM Tris–HCl, pH7.5, 5 mM EDTA, 0.05%

NP40). The antibody PAbp68-rgg was then added to the

mixture. The solution was subsequently incubated at 4 -Cfor 2 h. Protein A agarose beads slurry (40 Al) was then

added and the mixture was rotated at 4 -C for 5 h. The beads

were recovered and washed with 5�600 Al NETS buffer

with 0.08% SDS. Finally, the precipitated proteins were

analyzed by SDS-PAGE followed by Western blot or were

used in a further reaction described in the appropriate text.

The Western-blot analyses were performed with the

commercially available ECL Western-blot and detection kit

(Amersham Biosciences). The supernatant from culture

medium of hybridoma cells P68-rgg were used in the blotting

experiments as 5 : 1 dilution. The polyclonal antibody

PAbp68-rgg was used in 1 :3000 dilution. The antibodies,

PY20, 16B4, and 14B3, were used with 1 :2000 dilution.

2.5. Kinase competition assays

About 5 Ag of recombinant p68 was dephosphorylated

by both PP2A and PTP1B. The dephosphorylated protein

was separated from the added protein phosphatases by Ni-

NTA beads. After extensive washes, the dephosphorylated

protein was eluted with 250 mM imidazole in protein buffer

(50 mM Tris–HCl pH=7.5, 100 mM NaCl, 0.5 mM DTT,

10% glycerol). The protein was then micro-dialysed

against the same protein buffer with 100 mM of imidazole.

Page 3: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–1504 1497

For the kinase competition, 1 /10 of the above dephos-

phorylated p68 was first incubated with one protein

kinase(s) in the presence of 3 mM of non-radioactive

ATP for 60 min at 37 -C. After the pre-incubation, the

reaction mixture underwent micro-dialysis to exchange the

kinase buffer and remove non-radioactive ATP. The second

protein kinase(s) along with 10 ACi of [g-32P]ATP were

added to the reaction mixture. The kinase reaction was

further incubated at 37 -C for 45 min. The reaction

mixture was subsequently analyzed by 10% SDS PAGE

and subjected to auto-radiography. For the kinase com-

petition with the cellular p68 RNA helicase, the cellular

p68 that was immunoprecipitated from 200 Al of HeLa

nuclear extracts was dephosphorylated on the protein A

agarose beads. After extensive washes, the dephosphory-

lated p68 was eluted at 100 mM glycine pH=2.0. The

elution fractions were collected in 500 mM Tris–HCl,

pH=8.5, 100 mM NaCl for immediate neutralization. The

protein was then micro-dialysed against protein buffer, 100

mM Tris–HCl, pH=7.5, 50 mM NaCl, 1 mM DTT.

Subsequent kinase competition assays were essential the

same as that with the recombinant protein.

2.6. ATPase and RNA unwinding assays

ATPase activities were determined by measuring the

released inorganic phosphate during ATP hydrolysis using a

direct colorimetric assay as previously reported [33,36,37].

RNA unwinding activities were determined by the method

similar to that which was described in our previous report

[33].

2.7. In vitro pre-mRNA splicing

Splicing reactions were carried out with pPIP10A in 40%

HeLa nuclear extracts or p68 depleted HeLa nuclear

extracts. p68 RNA helicase was depleted from the nuclear

extracts by the experimental procedure that was described in

our previous report [25]. To reconstitute the splicing

activity, the phosphorylated/dephosphorylated his-tag pro-

tein was added to the p68-depleted nuclear extracts to a final

concentration of ¨20 ng/Al. The mixture was incubated at

30 -C for 15 min under normal in vitro splicing conditions.

About 25 fmol of pre-mRNA pPIP10A was then added to

the 10 Al of pre-incubated extracts and the splicing reaction

was incubated at 30 -C for an additional 150 min. The

splicing products were analyzed by 12% urea-PAGE.

3. Results

3.1. P68 RNA helicase is phosphorylated at tyrosine

residues in HeLa cells

We previously reported that bacterially expressed p68

RNA helicase was serine/threonine and tyrosine phosphory-

lated [1]. We questioned whether the endogenous p68 in

the human cells was also serine/threonine and tyrosine

phosphorylated. To this end, we used human HeLa cells as

an example. HeLa nuclear extracts were freshly made from

cultured HeLa cells. To prevent the dephosphorylation of

the target proteins by endogenous protein phosphatases,

Okadaic acid, inhibitors to PP2A, or phosphotyrosine

phosphatase inhibitor set (Vanadium, Calbiochem) were

added to the nuclear extracts. The cellular p68 RNA

helicase was immuno-precipitated from the HeLa nuclear

extracts by a polyclonal antibody PAbp68-rgg, raised

against bacterially expressed recombinant C-terminal

domain (aa 437–614) of p68. Our Western blot experi-

ments verified that the cellular p68 is the only antigen in

the HeLa nuclear extracts that is recognized by this

antibody (data not shown). The immuno-precipitates were

subsequently subjected to Western blot analyses using

specific antibodies against the phospho amino acids. The

immuno-precipitates from the HeLa nuclear extracts that

were supplemented with Okadaic acid were probed by the

antibodies 16B4 (anti-phosphoserine) or 14B3 (anti-phos-

phothreonine). The immuno-precipitates from the nuclear

extracts that were supplemented with vanadium were

probed by the antibody PY20 (anti-phosphotyrosine). It

is evident that p68 RNA helicase precipitated from the

HeLa nuclear extracts was not recognized by 14B3 and

16B4 (Fig. 1A, C, lane 2). The precipitated protein,

however, was recognized by the antibody PY20 (Fig. 1B).

To further verify that the protein recognized by the

antibody PY20 was identical to p68 RNA helicase, we

carried out another Western blot experiment using a

monoclonal antibody (p68-rgg) raised against bacterially

expressed C-terminal p68 (Fig. 1D). Overlay of these two

Western blots indicated that the protein recognized by

PY20 was identical to p68 RNA helicase. Our previous

experiments demonstrated that the antibody PY20 specif-

ically recognizes tyrosine phosphorylated protein [1].

Thus, our immunoprecipitation and subsequent Western

blot experiments indicated that the cellular p68 RNA

helicase in HeLa cells was tyrosyl phosphorylated. The

precipitated p68 could not be recognized by 14B3 and

16B4 may suggest that p68 is not phosphorylated at the

serine/threonine residues. However, it is also possible that

the antibodies do not recognize all serine/threonine

phosphorylation sites. Phosphorylation of p68 in HeLa

cells was further confirmed by metabolism 32P labeling

(data will be reported elsewhere).

3.2. Phosphorylation of p68 RNA helicase is controlled by

TNF-induced cell signals

In an effort to identify the cell signal pathway(s) that is

responsible for stimulation of the phosphorylations of p68

RNA helicase, we tested the effects of a few growth

hormones and cytokines on p68 phosphorylations. Tumor

necrosis factor (TNF-a) induces a cell signaling pathway

Page 4: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

Fig. 2. TNF-a induced cell signals regulate phosphorylation of p68 RNA

helicase. Western blot of cellular p68 RNA helicase from HeLa nuclear

extracts by antibodies, (A) p68-rgg, (B) PY20, and (C) 14B3. P68 RNA

helicase was immunoprecipitated from HeLa nuclear extracts by the

polyclonal antibody PAbp68-rgg. The HeLa nuclear extracts were made

after the cells were treated with 15 ng/ml of TNF-a for indicated time

points.

Fig. 1. Cellular p68 RNA helicase from HeLa cells is phosphorylated at tyrosine residue. Western-blot analyses of cellular p68 RNA helicase by monoclonal

antibodies 16B4 (A), PY20 (B), and 14B3 (C), and p68-rgg (D). The cellular p68 RNA helicase was immunoprecipitated via the polyclonal antibody PAbp68-

rgg. The Western blots were carried out with; 600 ng of BSA treated with PP2A then PKC and v-Abl kinases (lane 1 in A and C), 600 ng of BSA treated with

PTP1B then PKC and v-Abl kinases (lane 1 in D), and the immunoprecipitated p68 was treated with; no treatment (lane 2 in A, C), PP2A then PKC (lane 3 in

A, C), PP2A then v-Abl (lane 4 in A, C), PTP1B (lane 2 in D), PTP1B then PKC (lane 3 in D), and PTP1B then v-Abl (lane 4 in D). The precipitated cellular

p68 RNA helicase was dephosphorylated and/or phosphorylated via appropriate protein kinases or protein phosphatases as indicated prior to the Western blot

analyses. In (B), the p68 was immunoprecipitated via PAbp68-rgg raised from two separate rabbits (8207 and 8209).

L. Yang et al. / Cellular Signalling 17 (2005) 1495–15041498

that simultaneously activates a number of downstream

targets, including a number of protein kinases, e.g. JNK,

p38 MAP kinase [38–40]. Thus, we treated HeLa cells

with TNF-a (Sigma-Aldrich). Nuclear extracts were made

from the TNF-a treated cells. The cellular p68 RNA

helicase was immunoprecipitated from the nuclear extracts

via the polyclonal antibody PAbp68-rgg and subsequent

probed by Western blot using antibodies against specific

phospho amino acids. It was evident that the tyrosine

phosphorylation(s) of p68 responded very quickly to the

TNF-a treatment. The phosphorylation at tyrosine became

completely undetectable if the cells were treated with TNF-

a for 30 min (Fig. 2B, lane 3–5). We next asked if the

serine/threonine phosphorylations of p68 were affected by

TNF-a treatment in HeLa cells. Very similar experiments

were performed. P68 RNA helicase was immunoprecipi-

tated from the HeLa nuclear extracts that were made from

TNF-a treated cells. The precipitated cellular p68 was

subsequently probed with monoclonal antibodies against

phospho-serine or phospho-threonine. Unlike in our

observations made with untreated cells, the Western blots

analyses clearly indicated that cellular p68 RNA helicase

was phosphorylated at threonine residue(s) after the cells

were treated with TNF-a (Fig. 2C, lane 2 and 3), but not

at serine residue(s) (data not shown). Interestingly,

phosphorylation of p68 at threonine residue(s) only

occurred in a very narrow time window of TNF-

a treatments. The phosphorylation(s) was detected at 5

min of TNF-a treatment, reached the maxim at about 30

min, and rapidly decreased thereafter (Fig. 2C, lane 2 and

3, and data not shown). Our data indicated that phosphor-

ylations of p68 RNA helicase were controlled by TNF-a

induced cell signal pathway.

3.3. P68 is phosphorylated at threonine residue(s) by p38

MAP kinase

One important downstream target in the TNF-a induced

signal pathway is the activation of p38 MAP kinase [38,41].

Since we observed the threonine phosphorylation of p68 in

response to the TNF-a treatment, we reasoned that p68

Page 5: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–1504 1499

RNA helicase might be targeted by p38 MAP kinase in

cells. To test this possibility, we utilized a commercially

available p38 MAP kinase specific inhibitor, SB203580

(Sigma-Aldrich). HeLa cells were pre-treated with various

concentrations of SB203580. Following the treatments, the

cells were treated by TNF-a for 30 min. Nuclear extracts

were made from the cells. The cellular p68 RNA helicase

was immunoprecipitated from the nuclear extracts and

examined by Western blots using anti-phosphothreonine

antibody. It was very clear that the threonine phosphor-

ylation of p68 is inhibited by this kinase inhibitor (Fig. 3B,

lane 3, 4, and 5). To confirm phosphorylation of p68 RNA

helicase at threonine by p38 MAP kinase, we carried out

phosphorylation reaction with recombinant p68 using the

same HeLa nuclear extracts. Bacterially expressed recombi-

Fig. 3. P68 RNA helicase is a target of p38 MAP kinase. Western blot of cellular p6

RNA helicase was immunoprecipitated from HeLa nuclear extracts by PAbp68-r

different concentrations (indicated) of p38 MAP kinase inhibitor SB205380 follow

RNA helicase (lane 2–4), HCV-NS3 (lane 5), and yeast Saccharomyces cerevisia

untreated (lane 2), treated with TNF-a (15 ng/ml) for 30 min (lane 3–6), and the c

(lane 4). Top panel is Western blot with antibody 14B3 and bottom panel is Cooma

(600 ng) were incubated with corresponding HeLa nuclear extracts (30%) in 30 Aextracts by Ni-NTA beads. The precipitates were separated in 10% SDS-PAGE. T

were indicated on right side. (D) In vitro phosphorylation of; p38 MAP kinase

recombinant p68 (lane 4), recombinant HCV-NS3 (lane 5), and recombinant De

incubated with 0.5 U of p38 MAP kinase in 10 Al reactions in the presence of 4 m

14B3. The middle panel is autoradiography of the SDS-PAGE and bottom panel is

his-p68, his-Ded1p, his-HCV-NS3, and p38 MAP kinase were indicated on right

nant p68 was completely dephosphorylated by PP2A and

PTP1B prior to the addition to the HeLa extracts. It is

evident that the recombinant p68 was phosphorylated at

threonine in the nuclear extracts made from TNF-a treated

cells (Fig. 3C, lane 3). However, under the same conditions,

the recombinant p68 was not phosphorylated if the cells

were pre-treated with p38 MAP kinase inhibitor (Fig. 3C

lane 4). As controls, other recombinant RNA helicases,

HCV-NS3 and yeast Ded1p were not phosphorylated by the

HeLa nuclear extracts made from TNF-a treated HeLa cells

under the same conditions (Fig. 3C, lane 5 and 6). To further

verify phosphorylation of p68 at threonine residue(s) by p38

MAP kinase, we carried out in vitro phosphorylation assay

using a commercially available p38 MAP kinase (Calbio-

chem). Our assays clearly showed that the dephosphorylated

8 RNA helicase from HeLa nuclear extracts by, (A) p68-rgg, (B) 14B3. P68

gg. The HeLa nuclear extracts were made after the cells were treated with

ed by TNF-a treatment for 30 min. (C) Phosphorylation of recombinant; p68

e Ded1p (lane 6) in HeLa nuclear extracts made from HeLa cells that were;

ells were pre-treated with SB203580 (30 AM) prior to the TNF-a treatment

ssie staining of the SDS-PAGE. The dephosphorylated recombinant proteins

l for 60 min. The recombinant proteins were precipitated from the nuclear

he protein bands corresponding to his-p68, his-Ded1p, and his-HCV-NS3

(lane 2), recombinant p68 without addition of p38 MAP kinase (lane 3),

d1p (lane 6) by p38 MAP kinase. Recombinant proteins (¨200 ng) were

M ATP and 3 ACi of [g-32P]ATP. Top panel is Western blot with antibody

Coomassie staining of the SDS-PAGE. The protein bands corresponding to

side.

Page 6: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–15041500

recombinant p68 RNA helicase became phosphorylated at

threonine residue(s) by p38 MAP kinase (Fig. 3D, lane 4).

Under the same conditions, the recombinant HCV-NS3 and

yeast Ded1p were not phosphorylated by p38 MAP kinase

(Fig. 3D, lane 5 and 6). The experiments provided evidence

that p68 RNA helicase is phosphorylated at threonine

residue(s) by p38 MAP kinase in response to TNF-a

induced signal.

3.4. The serine/threonine phosphorylation(s) and tyrosine

phosphorylation(s) inhibit each other

The preceding data demonstrated the phosphorylations of

p68 RNA helicase at serine/threonine and tyrosine residues.

The phosphorylations of the protein were coupled to TNF-a

induced cell signal pathway and genotoxic agent mediated

DNA damage signal. Our next question was whether the

serine/threonine phosphorylations and/or tyrosyl phosphor-

ylation affected each other. To this end, we employed

competition kinase assays. The dephosphorylated recombi-

nant p68 RNA helicase was examined in two kinase

reactions. We examined two kinase reactions with

[g-32P]ATP by PKC or v-Abl kinases under the conditions

that the dephosphorylated protein was first phosphorylated

with non-radioactive ATP by PKC or/and v-Abl kinases

prior to [g-32P]ATP phosphorylations by another protein

Fig. 4. (A) PKC and v-Abl kinase competition assays with recombinant p68 RNA

immunoprecipitated from HeLa nuclear extracts. (C) P38 MAP kinase and v-Abl k

is auto-radiography of the SDS-PAGE. The bottom panel is Coomassie staining of

and p38 MAP kinase were indicated on right side. The dephosphorylated p68 was

the presence of 3 mM of non-radioactive ATP for 60 min at 37 -C (competition). A

along with 10 ACi of [g-32P]ATP were added to the competition reaction mixture. T

reaction). The reaction mixture was subsequently analyzed by 10% SDS-PAGE a

kinase. If dephosphorylated p68 was pre-treated with PKC

and non-radioactive ATP, phosphorylation by v-Abl kinase

was then very weak (Fig. 4A, lane 3). Similarly, If

dephosphorylated p68 was pre-treated with v-Abl kinase

and non-radioactive ATP, phosphorylation by PKC was very

weak (Fig. 4A, lane 5). Similar kinase competition experi-

ments were also carried out with p38 MAP kinase and v-Abl

kinase. Our experiments showed that phosphorylation of

p68 by p38 MAP kinase inhibited the phosphorylation by

v-Abl kinase (Fig. 4C, lane 7). The v-Abl phosphorylation

also inhibited p38 MAPK phosphorylation (Fig. 4C, lane 5).

We also tested the kinase competitions with cellular p68 that

was purified from HeLa nuclear extracts. The cellular p68

from HeLa nuclear extracts was first immobilized on protein

A agarose beads by the polyclonal antibody PAbp68-rgg

and dephosphorylated by PTP1B. The proteins were eluted

and further separated by a gel-filtration. The dephosphor-

ylation reaction went to completion as indicated by Western

blot (data not shown). The kinase competitions similar to

that were used to the recombinant protein were applied here

with the dephosphorylated cellular p68. The results obtained

with the cellular p68 were essentially the same as that

obtained with the recombinant protein (Fig. 4B). The kinase

competition results demonstrated that phosphorylations on

serine/threonine residues inhibited the phosphorylations on

tyrosine residues and vice-versa.

helicase. (B) PKC and v-Abl kinase competition assays with cellular p68

inase competition assays with recombinant p68 RNA helicase. The top panel

the SDS-PAGE. The protein bands corresponding to his-p68, v-Abl kinase,

first incubated with appropriate protein kinase(s) indicated in each panel in

fter the pre-incubation, the second protein kinase(s), indicated in each panel

he kinase reaction was further incubated at 37 -C for another 45 min (kinase

nd subjected to auto-radiography.

Page 7: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–1504 1501

3.5. Phosphorylation affects the ATPase and RNA unwind-

ing activities of p68 RNA helicase

In our previous study, we showed that the recombinant

p68 is an RNA-dependent ATPase and ATP-dependent

RNA helicase. Therefore, we asked whether or not p68

phosphorylation affects the ATPase and RNA unwinding

activities of the protein. We have observed that phosphor-

ylations on tyrosine or serine/threonine residues have

different effects on the ATPase activity of the protein [1].

We next tested the effects of the phosphorylation of p68 on

the RNA unwinding activity of the protein. Since the

bacterially expressed recombinant p68 was phosphorylated

at serine/threonine and tyrosine residues. Thus, the effects of

dephosphorylation at serine/threonine or tyrosine residue(s)

on the RNA unwinding activity of the protein were

monitored. The unwinding substrate RNA used in our study

contains a short RNA duplex (¨22 bp in length) and long

186 nt and 88 nt 3V overhangs on both sides [33]. It is

evident that the recombinant p68 RNA helicase treated with

PTP1B unwound the substrate RNA (Fig. 5, lane 4). The

proteins that were dephosphorylated by PP2A failed to

unwind the substrate (Fig. 5, lane 5). Thus, our experiments

of ATPase and RNA unwinding assays demonstrated that

phosphorylation on p68 RNA helicase control the ATPase

and RNA unwinding activities of the protein.

3.6. Phosphorylations on tyrosine abolish the function of

p68 RNA helicase in the pre-mRNA splicing process

We have previously showed that p68 RNA helicase is an

essential human splicing factor in vitro [25]. Thus, we asked

whether or not the phosphorylation of p68 RNA helicase

affect the function(s) of the protein in the in vitro pre-

mRNA splicing process. The effects of serine/threonine or

tyrosine phosphorylation on the function of p68 RNA

Fig. 5. RNA unwinding activity of p68 RNA helicase is differently affected by p

RNA helicase. dsRNA, 2.5 fmol, unwinding substrate were incubated with 150 ng

60 min. The his-tag p68 was pretreated with; no treatment (lane 3), 1 Al of PTP1B p

1 Al PP2A (lane 6), 1.5 Al of v-Abl kinase was added to PTP1B and PP2A treated h

PKC was added to PTP1B and PP2A treated his-p68 after addition of protein phosp

for 8 min. The sample was loaded on the gel immediately after heat denature. La

helicase in the spliceosome were tested in previously

established in vitro reconstitution system [25]. To obtain

the recombinant p68 with serine/threonine or tyrosine

phosphorylation(s), the recombinant protein was completely

dephosphorylated by PP2A and PTP1B. The dephosphory-

lated protein was then re-phosphorylated by PKC or v-Abl

kinase. The PKC or v-Abl re-phosphorylated p68 RNA

helicase was added back to the p68 depleted HeLa extracts.

The splicing activity of the reconstituted HeLa extracts was

monitored. The splicing substrate pPIP10A (derivative of

major late transcript of adenovirus) was used in the splicing

assays. Our experiments showed that the splicing activity of

the p68 depleted HeLa nuclear extracts was recovered by

addition of dephosphorylated (Fig. 6, lane 7) or serine/

threonine re-phosphorylated p68 (Fig. 6, lane 8). The

splicing activity, however, could not be restored by addition

of the tyrosyl phosphorylated p68 (Fig. 6, lane 9). The

results suggested that tyrosyl phosphorylation of p68

inhibited pre-mRNA splicing process in HeLa nuclear

extracts. To further confirm the effects of p68 on the

splicing activity were indeed due to tyrosyl phosphorylation,

we re-phosphorylated the dephosphorylated his-p68 via

ATP-g-S by Abl kinase. The tyrosyl thio-phosphorylated

his-p68 could not restore the splicing activities of p68

depleted HeLa extracts (Fig. 6, lane 11).

4. Discussion

In this report, we demonstrated that p68 RNA helicase is

phosphorylated at multiple amino acid residues, including

serine/threonine and tyrosine. We presented evidence to

show that protein phosphorylation play a very important

role in regulating the biochemical activities. Protein

phosphorylation is a major mechanism that links the protein

function with various important cell signal pathways, such

hosphorylations. RNA unwinding activities of the recombinant his-tag p68

of the recombinant p68 RNA helicase at 37 -C in the unwinding buffer for

rotein tyrosine phosphatase (lane 4), 1 Al of PP2A (lane 5), 1 Al PTP1B and

is-p68 after addition of protein phosphatase inhibitors (lane 7), and 1.5 Al ofhatase inhibitors (lane 8). Lane 1 is the dsRNA substrate denatured at 95 -C

ne 2 is the dsRNA alone.

Page 8: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

Fig. 6. Tyrosyl phosphorylation inhibits the function of p68 in the in vitro

pre-mRNA splicing. Splicing reactions were carried out with pPIP10A in

40% intact HeLa nuclear extracts (lane 2–4) or p68 depleted HeLa nuclear

extracts (lane 5–11) at 30 -C for 150 min. The HeLa nuclear extracts were

supplemented with; no supplements (lane 2), v-Abl kinase (lane 3), and

protein tyrosine phosphatase inhibitor vanadium (lane 4), and the

recombinant his-tag p68 RNA helicase (¨20 ng/Al). The recombinant

p68 RNA helicase was pre-treated with appropriate protein kinases and/or

protein phosphatases as indicated in each lane (lane 5–11) prior to the

addition of p68 to the p68-depleted extracts. The splicing products were

analyzed by 12% urea-denature PAGE and subjected to the autoradiog-

raphy. Lane 1 is the splicing substrate pPIP10A alone.

L. Yang et al. / Cellular Signalling 17 (2005) 1495–15041502

as many growth hormones and cytokines. The function of

p68 RNA helicase has been shown to be critical for cell

normal growth and regulation. Dysregulating the cellular

function of the p68 leads to tumor progression and cell

proliferation [15,16,18,42]. Thus, regulation of the cellular

function(s) of p68 RNA helicase by protein phosphorylation

provides an excellent explanation for the role of p68 in the

cell development program and proliferation. The p68 RNA

helicase is phosphorylated at multiple sites, including

serine/threonine and tyrosine amino acid residues, which

reflect a complex regulation mechanism. P68 seems to be an

essential multifunctional RNA helicase that is required to

unwind RNA structures or dissociate RNA protein inter-

actions in multiple cellular processes. It is conceivable that

different phosphorylation may be employed to regulate the

function(s) of p68 in response to different external stimuli.

Obviously, the physiological importance of phosphory-

lation on p68 RNA helicase has to be established by

identifying the cell signal pathway(s) and the cellular

protein kinase(s) that are responsible for phosphorylation

of p68. Regulation of phosphorylation/dephosphorylation of

p68 by TNF-a signal pathway certainly established one

connection between cell signal pathways and p68 phosphor-

ylation. TNF-a has been shown in regulating cell apoptosis,

proliferation, and survival [38,43]. Given the role played by

p68 RNA helicase in cell developmental programs, it is not

difficult to imagine that p68 is one cellular target that

execute TNF-a signal downstream. Interestingly, phosphor-

ylation at tyrosine residue(s) or phosphorylation on threo-

nine residue(s) was reciprocally regulated by TNF-a

treatment. This opposite regulatory effects correlated very

well with the different effects of the tyrosine or threonine

phosphorylation on the biochemical activities of the protein,

as well as the function of the protein in the spliceosome.

TNF-a stimulated threonine phosphorylation(s) of p68 was

quickly dephosphorylated upon a longer time of TNF-a

treatment, which suggested that TNF-a also activated a

protein phosphatase to dephosphorylate p68 at threonine

residue(s). Decreases and disappearance of tyrosine phos-

phorylation(s) of p68 upon the TNF-a treatment is intrigu-

ing. It is likely that TNF-a treatment activated a protein

tyrosine phosphatase (PTP) that dephosphorylated the

tyrosyl phosphorylation(s). In supporting this conjecture,

we observed that tyrosyl phosphorylation(s) of recombinant

p68 was dephosphorylated in HeLa nuclear extracts that

were made from TNF-a treated HeLa cells, while the

tyrosyl phosphorylation(s) was not dephosphorylated in

HeLa nuclear extracts that were made from untreated HeLa

cells (data not shown). In addition, it is known that TNF-a

induced cell signal inhibits cell proliferation through

activation of a protein tyrosine phosphatase (SHP-1) [44].

Given that HeLa cell is a human cancer cell line, it is

tempted to speculate that tyrosine dephosphorylation and/or

threonine phosphorylation(s) of p68 may be an apoptotic

response to TNF-a treatment.

Protein tyrosine kinase(s) (PTK) that is responsible for

phosphorylation of p68 on the tyrosine residues is an open

question. Since p68 RNA helicase was shown to predom-

inately localize in the cell nuclear [4], it would be expected

that a nuclear PTK phosphorylates p68 RNA helicase in

cells. The c-Abl kinase is a protein tyrosine kinase that

localizes to both cell nuclear and cytoplasm [45]. The kinase

is a regulator of the DNA damage response system [46,47].

We showed that recombinant p68 could be phosphorylated

by bacterially expressed truncated form of MuLV v-Abl

kinase and c-Abl. Cellular p68 purified by immunoprecipi-

tation from HeLa nuclear extracts could also be phosphory-

lated by the Abl kinases. Moreover, we observed that c-Abl

kinase co-immunoprecipitated with p68 in an immunopre-

cipitation experiment with nuclear extracts made from A549

and Caco-2 cells (data will be reported elsewhere). Thus, c-

Abl kinase is likely a candidate PTK that phosphorylates

p68 RNA helicase on tyrosine residues. It is also possible

that p68 RNA helicase may be a target of protein tyrosine

kinases that are Fmis_-localized to the cell nuclear. It has

Page 9: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–1504 1503

been demonstrated that members of the src family of PTK,

the intracellular tyrosine kinase BRK/Sik, and the receptor

tyrosine kinases (RTK) can change their localization

patterns in response to external stimuli [48–50]. Thus, all

these protein tyrosine kinases are all potential candidates

that phosphorylate p68 RNA helicase.

Which cellular serine/threonine protein kinase that

directly targets p68 under the treatment of TNF-a is also

an intrigue question. We identified p68 RNA helicase as a

target of p38 MAP kinase in response to the TNF-a

treatment, which further confirm the role of p68 in TNF-a

induced signal pathway. P38 MAP kinase usually target

transcription factors or other protein kinases [51–54]. Thus,

phosphorylation of p68 by p38 MAP kinase seems

consistent with the potential role of p68 RNA helicase in

transcription activation [20–22]. The p68 RNA helicase has

an IQ motif, a potential PKC phosphorylation site, in its C-

terminal domain. Both the protein purified from rat PC12

cells and the recombinant protein expressed in E. coli can be

phosphorylated by PKC in vitro [31,34]. In addition, p68

RNA helicase was shown to associate with PKC in the

nuclear matrix of the neuroblastoma cells [55]. All these

observations seem to argue that p68 is a substrate of PKC.

Nevertheless, solid evidence to demonstrate the phosphor-

ylation of p68 RNA helicase by PKC in cells is not

presented. P68 RNA helicase was also shown to associate

with a cAMP-dependent kinase (PKA) anchoring protein in

HEK293 and COS cells [32]. Thus, it is also possible that

p68 may be phosphorylated by cAMP-dependent protein

kinase. We reason that phosphorylation(s) of p68 by PKC,

PKA, may be cell/tissue type specific. Recently, we indeed

observed that p68 RNA helicase that was immunoprecipi-

tated from nuclear extracts made from normal human lung

tissue and MCF-7 cells was threonine phosphorylated (data

will be reported elsewhere). We noted that, in contrast to our

observation, p68 RNA helicase was observed as an

unphosphorylated form in HeLa cells by metabolic inor-

ganic 32P labeling the cells [56]. We do not know the

experimental conditions for labeling and detection of protein

phosphorylation used by Uhlmann-Schiffler and co-work-

ers. It is, however, possible that not all phospho proteins in

the cells were detected by the metabolic labeling.

P68 RNA helicase was shown to function in the pre-

mRNA splicing process [25]. In this report, we further

demonstrated that the function of p68 in the spliceosome is

regulated by protein tyrosine phosphorylation. The pre-

mRNA splicing process is an essential step in the eukaryotic

gene expression pathway [57–59]. It is likely that this

process is subjected to regulation by a number of cell

growth signals [57]. Tyrosine phosphorylation is involved in

major signal pathways that regulate the cell growth and

differentiation [60,61]. It has been shown that both pre-

mRNA splicing and mRNA transport are regulated by the

tyrosine kinase activity of src [62]. Therefore, controlling

the pre-mRNA splicing by tyrosine phosphorylation of p68

RNA helicase may represent a critical regulatory point for

this important process in the eukaryotic gene expression

pathway. In our previous report, we showed that p68 RNA

helicase functioned at the transient U1:5Vss duplex [25,26].

We suggested that p68 may unwind the U1:5Vss duplex

during the spliceosome assembly process. In this report, we

demonstrated that the tyrosyl phosphorylated p68 RNA

helicase lost its dsRNA-dependent ATPase and RNA

unwinding activities and the tyrosyl phosphorylated p68

do not support splicing. Although, the bacterially expressed

his-p68 without the treatment of protein phosphatases could

not restore the splicing activity of p68 depleted extracts

most of the time (data not shown), curiously, we repeatedly

observed that the his-p68 without dephosphorylation some-

times could partially restore the splicing activity of the p68

depleted extracts (Fig. 5, lane 6). Interestingly, when the

untreated recombinant his-p68 could recover the splicing

activity, the same recombinant protein often also demon-

strated RNA unwinding activity (data not shown). Our

explanation is that a fraction of his-p68 occasionally was not

completely phosphorylated at tyrosine residue(s) by E. coli.

Thus, the tyrosyl unphosphorylated his-p68 unwound RNA

and supported splicing.

Interestingly, the cellular p68 RNA helicase is tyrosine

phosphorylated in HeLa cells. Nevertheless, the pre-mRNA

splicing process is not inhibited in the intact HeLa nuclear

extracts. The tyrosine phosphorylated recombinant protein,

however, did not restore the splicing activity of p68 depleted

extracts. It is possible that only a portion of cellular p68 is

tyrosyl phosphorylated. Another possibility is that the

tyrosine phosphorylation on the endogenous p68 is dephos-

phorylated by a protein tyrosine phosphatase that is

associated with the spliceosome. Alternatively, the tyrosine

phosphorylations on the cellular p68 are dephosphorylated

before the protein is assembled into the spliceosome

complexes. The tyrosine phosphorylation on the recombi-

nant p68 is somehow not dephosphorylated by the protein

tyrosine phosphatase. In this regard, it will be interesting to

probe the phosphorylation status of endogenous p68 in the

spliceosome complexes. In addition, we cannot rule out

another possibility that the tyrosine phosphorylation sites by

v-Abl kinase on the recombinant protein differ from that of

endogenous protein in the HeLa cells. Currently, we are

mapping the tyrosine phosphorylation site(s) of the cellular

p68 from the HeLa cells and the bacterially expressed

recombinant p68 RNA helicase.

Acknowledgments

We thank Roger Bridgeman for antibody p68-rgg

production. We are grateful to Jenny Yang, Phang C. Tai,

April Ellis, Heena Dey, Amit Khanna, and Shubhalaxmi

Kayarthodi for detailed critical comments on the manu-

script. This work is supported in part by research grants

from National Institute of Health (GM063874) and Georgia

Cancer Coalition to ZRL.

Page 10: Signaling to the DEAD box—Regulation of DEAD-box p68 RNA helicase by protein phosphorylations

L. Yang et al. / Cellular Signalling 17 (2005) 1495–15041504

References

[1] L. Yang, Z.R. Liu, Protein Expr. Purif. 35 (2004) 327.

[2] P. Linder, M.C. Daugeron, Nat. Struct. Biol. 7 (2000) 97.

[3] P. Linder, N.K. Tanner, J. Banroques, Trends Biochem. Sci. 26 (2001)

339.

[4] A. Luking, U. Stahl, U. Schmidt, Crit. Rev. Biochem. Mol. Biol. 33

(1998) 259.

[5] A. Pause, N. Sonenberg, Curr. Opin. Struck. Biol. 3 (1993) 953.

[6] E. Jankowsky, C.H. Gross, S. Shuman, A.M. Pyle, Science 291 (2001)

121.

[7] N.K. Tanner, P. Linder, Mol. Cell 8 (2001) 251.

[8] M. Jaramillo, T.E. Dever, W.C. Merrick, N. Sonenberg, Mol. Cell.

Biol. 11 (1991) 5992.

[9] L. Crawford, K. Leppard, D. Lane, E. Harlow, J. Virol. 42 (1982) 612.

[10] D.P. Lane, W.K. Hoeffler, Nature 288 (1980) 167.

[11] R.D. Iggo, D.P. Lane, EMBO J. 8 (1989) 1827.

[12] M.J. Ford, I.A. Anton, D.P. Lane, Nature 332 (1988) 736.

[13] H. Hirling, M. Scheffner, T. Restle, H. Stahl, Nature 339 (1989) 562.

[14] U.A. Heinlein, J. Pathol. 184 (1998) 345.

[15] R.J. Stevenson, S.J. Hamilton, D.E. MacCallum, P.A. Hall, F.V. Fuller-

Pace, J. Pathol. 184 (1998) 351.

[16] M. Causevic, R.G. Hislop, N.M. Kernohan, F.A. Carey, R.A. Kay, R.J.

Steele, F.V. Fuller-Pace, Oncogene 20 (2001) 7734.

[17] P. Dubey, R.C. Hendrickson, S.C. Meredith, C.T. Siegel, J. Shabano-

witz, J.C. Skipper, V.H. Engelhard, D.F. Hunt, H. Schreiber, J. Exp.

Med. 185 (1997) 695.

[18] Y. Wei, M.H. Hu, YiChuan XueBao 28 (2001) 991.

[19] S.R. Schmid, P. Linder, Mol. Microbiol. 6 (1992) 283.

[20] H. Endoh, K. Maruyama, Y. Masuhiro, Y. Kobayashi, M. Goto, H. Tai,

J. Yanagisawa, D. Metzger, S. Hashimoto, S. Kato, Mol. Cell. Biol. 19

(1999) 5363.

[21] T. Fujita, Y. Kobayashi, O. Wada, Y. Tateishi, L. Kitada, Y.

Yamamoto, H. Takashima, A. Murayama, T. Yano, T. Baba, S. Kato,

Y. Kawabe, J. Yanagisawa, J. Biol. Chem. 278 (2003) 26704.

[22] M. Watanabe, J. Yanagisawa, H. Kitagawa, K. Takeyama, S. Ogawa,

Y. Arao, M. Suzawa, Y. Kobayashi, T. Yano, H. Yoshikawa, Y.

Masuhiro, S. Kato, EMBO J. 20 (2001) 1341.

[23] K.L. Rossow, R. Janknecht, Oncogene 22 (2003) 151.

[24] J.P. Jost, S. Schwarz, D. Hess, H. Angliker, F.V. Fuller-Pace, H. Stahl,

S. Thiry, M. Siegmann, Nucleic Acids Res. 27 (1999) 3245.

[25] Z.R. Liu, Mol. Cell. Biol. 22 (2002) 5443.

[26] Z.R. Liu, B. Sargueil, C.W. Smith, Mol. Cell. Biol. 18 (1998) 6910.

[27] K. Hartmuth, H. Urlaub, H.P. Vornlocher, C.L. Will, M. Gentzel, M.

Wilm, R. Luhrmann, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 16719.

[28] A. Honig, D. Auboeuf, M.M. Parker, B.W. O’Malley, S.M. Berget,

Mol. Cell. Biol. 22 (2002) 5698.

[29] S. Guil, R. Gattoni, M. Carrascal, J. Abian, J. Stevenin, M. Bach-Elias,

Mol. Cell. Biol. 23 (2003) 2927.

[30] M.S. Jurica, L.J. Licklider, S.R. Gygi, N. Grigorieff, M.J. Moore,

RNA 8 (2002) 426.

[31] M.K. Buelt, B.J. Glidden, D.R. Storm, J. Biol. Chem. 269 (1994)

29367.

[32] L. Akileswaran, J.W. Taraska, J.A. Sayer, J.M. Gettemy, V.M.

Coghlan, J. Biol. Chem. 276 (2001) 17448.

[33] Y. Huang, Z.R. Liu, J. Biol. Chem. 277 (2002) 12810.

[34] L. Yang, J. Yang, Y. Huang, Z.R. Liu, Biochem. Biophys. Res.

Commun. 314 (2004) 622.

[35] Z.R. Liu, B. Laggerbauer, R. Luhrmann, C.W. Smith, RNA 3 (1997)

1207.

[36] K.-M. Chan, D. Delfert, K.D. Junger, Anal. Biochem. 157 (1986) 375.

[37] G.E. Pugh, S.M. Nicol, F.V. Fuller-Pace, J. Mol. Biol. 292 (1999) 771.

[38] U. Gaur, B.B. Aggarwal, Biochem. Pharmacol. 66 (2003) 1403.

[39] Y. Deng, X. Ren, L. Yang, Y. Lin, X. Wu, Cell 115 (2003) 61.

[40] K. Heyninck, A. Wullaert, R. Beyaert, Biochem. Pharmacol. 66 (2003)

1409.

[41] M.E. Guicciardi, G.J. Gores, J. Clin. Invest. 111 (2003) 1813.

[42] V.C. Ogilvie, B.J. Wilson, S.M. Nicol, N.A. Morrice, L.R. Saunders,

G.N. Barber, F.V. Fuller-Pace, Nucleic Acids Res. 31 (2003) 1470.

[43] B.B. Aggarwal, Nat. Rev., Immunol. 3 (2003) 745.

[44] H. Nakagami, T.X. Cui, M. Iwai, T. Shiuchi, Y. Takeda-Matsubara, L.

Wu, M. Horiuchi, Arterioscler. Thromb. Vasc. Biol. 22 (2002) 238.

[45] A.M. Pendergast, Adv. Cancer Res. 85 (2002) 51.

[46] J.Y. Wang, Oncogene 19 (2000) 5643.

[47] T. Rich, R.L. Allen, A.H. Wyllie, Nature 407 (2000) 777.

[48] M.S. Serfas, A.L. Tyner, Oncol. Res. 13 (2003) 409.

[49] J.J. Derry, G.S. Prins, V. Ray, A.L. Tyner, Oncogene 22 (2003) 4212.

[50] G. Carpenter, Curr. Opin. Cell Biol. 15 (2003) 143.

[51] X.Z. Wang, D. Ron, Science 272 (1996) 1347.

[52] G.L. Johnson, R. Lapadat, Science 298 (2002) 1911.

[53] W.T. Gerthoffer, C.A. Singer, Respir. Physiol. Neurobiol. 137 (2003)

237.

[54] M. Xiu, J. Kim, E. Sampson, C.Y. Huang, R.J. Davis, K.E. Paulson,

A.S. Yee, Mol. Cell. Biol. 23 (2003) 8890.

[55] U. Rosenberger, I. Lehmann, C. Weise, P. Franke, F. Hucho, K.

Buchner, J. Cell. Biochem. 86 (2002) 394.

[56] H. Uhlmann-Schiffler, O.G. Rossler, H. Stahl, J. Biol. Chem. 277

(2002) 1066.

[57] J. Soret, J. Tazi, Prog. Mol. Subcell. Biol. 31 (2003) 89.

[58] A. Kramer, Annu. Rev. Biochem. 65 (1996) 367.

[59] A.I. Lamond, BioEssays 15 (1993) 595.

[60] A.C. Porter, R.R. Vaillancourt, Oncogene 17 (1998) 1343.

[61] S.V. Russello, S.K. Shore, Front. Biosci. 8 (2003) s1068.

[62] P. Gondran, F. Dautry, Oncogene 18 (1999) 2547.