ArticlePDF Available

14-3-3σ controls mitotic translation to facilitate cytokinesis

Authors:

Abstract

14-3-3 proteins are crucial in a wide variety of cellular responses including cell cycle progression, DNA damage checkpoints and apoptosis. One particular 14-3-3 isoform, sigma, is a p53-responsive gene, the function of which is frequently lost in human tumours, including breast and prostate cancers as a result of either hypermethylation of the 14-3-3sigma promoter or induction of an oestrogen-responsive ubiquitin ligase that specifically targets 14-3-3sigma for proteasomal degradation. Loss of 14-3-3sigma protein occurs not only within the tumours themselves but also in the surrounding pre-dysplastic tissue (so-called field cancerization), indicating that 14-3-3sigma might have an important tumour suppressor function that becomes lost early in the process of tumour evolution. The molecular basis for the tumour suppressor function of 14-3-3sigma is unknown. Here we report a previously unknown function for 14-3-3sigma as a regulator of mitotic translation through its direct mitosis-specific binding to a variety of translation/initiation factors, including eukaryotic initiation factor 4B in a stoichiometric manner. Cells lacking 14-3-3sigma, in marked contrast to normal cells, cannot suppress cap-dependent translation and do not stimulate cap-independent translation during and immediately after mitosis. This defective switch in the mechanism of translation results in reduced mitotic-specific expression of the endogenous internal ribosomal entry site (IRES)-dependent form of the cyclin-dependent kinase Cdk11 (p58 PITSLRE), leading to impaired cytokinesis, loss of Polo-like kinase-1 at the midbody, and the accumulation of binucleate cells. The aberrant mitotic phenotype of 14-3-3sigma-depleted cells can be rescued by forced expression of p58 PITSLRE or by extinguishing cap-dependent translation and increasing cap-independent translation during mitosis by using rapamycin. Our findings show how aberrant mitotic translation in the absence of 14-3-3sigma impairs mitotic exit to generate binucleate cells and provides a potential explanation of how 14-3-3sigma-deficient cells may progress on the path to aneuploidy and tumorigenesis.
LETTERS
14-3-3scontrols mitotic translation to facilitate
cytokinesis
Erik W. Wilker
1
, Marcel A. T. M. van Vugt
1
, Stephen C. Artim
1
, Paul H. Huang
2
, Christian P. Petersen
1
,
H. Christian Reinhardt
1
, Yun Feng
1
, Phillip A. Sharp
1
, Nahum Sonenberg
3
, Forest M. White
2
& Michael B. Yaffe
1,2
14-3-3 proteins are crucial in a wide variety of cellular responses
including cell cycle progression, DNA damage checkpoints and
apoptosis. One particular 14-3-3 isoform, s, is a p53-responsive
gene, the function of which is frequently lost in human tumours,
including breast and prostate cancers as a result of either hyper-
methylation of the 14-3-3spromoter or induction of an oestrogen-
responsive ubiquitin ligase that specifically targets 14-3-3sfor
proteasomal degradation
1–9
. Loss of 14-3-3sprotein occurs not
only within the tumours themselves but also in the surrounding
pre-dysplastic tissue (so-called field cancerization), indicating that
14-3-3smight have an important tumour suppressor function
that becomes lost early in the process of tumour evolution
3,9
. The
molecular basis for the tumour suppressor function of 14-3-3s
is unknown. Here we report a previously unknown function for
14-3-3sas a regulator of mitotic translation through its direct
mitosis-specific binding to a variety of translation/initiation fac-
tors, including eukaryotic initiation factor 4B in a stoichiometric
manner. Cells lacking 14-3-3s, in marked contrast to normal
cells, cannot suppress cap-dependent translation and do not
stimulate cap-independent translation during and immediately
after mitosis. This defective switch in the mechanism of trans-
lation results in reduced mitotic-specific expression of the endo-
genous internal ribosomal entry site (IRES)-dependent form of
the cyclin-dependent kinase Cdk11 (p58 PITSLRE), leading to
impaired cytokinesis, loss of Polo-like kinase-1 at the midbody,
and the accumulation of binucleate cells. The aberrant mitotic
phenotype of 14-3-3s-depleted cells can be rescued by forced
expression of p58 PITSLRE or by extinguishing cap-dependent
translation and increasing cap-independent translation during
mitosis by using rapamycin. Our findings show how aberrant
mitotic translation in the absence of 14-3-3simpairs mitotic exit
to generate binucleate cells and provides a potential explanation of
how 14-3-3s-deficient cells may progress on the path to aneu-
ploidy and tumorigenesis.
Because many tumour suppressor proteins function during spe-
cific parts of the cell cycle, we examined whether the binding of
ligands to 14-3-3sshowed a cell cycle dependence. U2OS cells were
synchronized by a double thymidine block, and lysates were prepared
at various times after release (Supplementary Fig. 1a). Immunopre-
cipitation of endogenous 14-3-3sfrom these lysates showed a
marked increase in 14-3-3s-bound proteins in the mitotic and
immediate post-mitotic periods (Fig. 1a, time points at 12 and 16 h
after release). This same mitotic enrichment in ligand binding to 14-
3-3swas also observed in other cell types, including HCT116 cells
and HeLa cells (Supplementary Fig.1b–d), but was not observed when
ligands of another endogenous 14-3-3 isoform, 14-3-3b, were ana-
lysed (Supplementary Fig. 1e) or in 14-3-3simmunoprecipitations
from asynchronous cells after DNA damage (Supplementary Fig. 2).
The mitosis-dependent binding of ligands to 14-3-3sseemed to be
both dependent on phosphorylation, because binding was lost when
lysates were treated with protein phosphatase 1 (Supplementary Fig.
1d), and direct, as shown with a two-dimensional ‘far-western assay
(Fig. 1d).
The mitotic targets of 14-3-3swere identified by mass spectro-
metry. This analysis revealed that a large proportion of 14-3-3s
ligands were proteins involved in the process of translation, including
several initiation factors involved in mediating cap-dependent trans-
lation (Fig. 1b and Supplementary Table 1). Two-dimensional far-
western blotting of mitotic 14-3-3simmunoprecipitates showed
that several of these proteins, including eukaryotic initiation factor
4B (eIF4B), eIF-2aand elongation factor 1a(EF1a), bound directly
to 14-3-3s, whereas other 14-3-3s-associated proteins such as eIF4G
seemed to be interacting indirectly (Fig. 1e).
It is well established that translation in mammalian cells is markedly
altered during and immediately after mitosis
10
, with a pronounced
suppression of cap-dependent translation and a corresponding
enhancement of cap-independent translation
11
. The biological mech-
anism and significance of this translational switch, however, are
unclear. Given the abundance of proteins involved in translation in
the 14-3-3simmunoprecipitations, we investigated whether 14-3-3s
has a direct function in this process by using RNA interference
(RNAi). Synchronous 14-3-3sstable knockdown and control cells
were pulse-labelled with [
35
S]methionine, and newly translated pro-
teins were analysed by SDS–polyacrylamide gel electrophoresis
(PAGE) and autoradiography (Fig. 2a, d, and Supplementary Fig.
3a). As expected, in control short hairpin RNA (shRNA)-treated
HeLa and U2OS cells, radiolabel incorporation into nascent polypep-
tides was suppressed at 12 and 16 h after release from the double
thymidine block (Fig. 2a, e, lanes marked ‘con’)
11
. Remarkably, no
suppression of mitotic translation was observed in either of these
cell types when they were depleted of 14-3-3s(lanes marked s’).
Identical results were also obtained with two other distinct 14-3-3s
shRNAs (data not shown). In contrast to 14-3-3sdepletion, RNAi-
mediated knockdown of 14-3-3bhad no effect on the suppression
of translation during and immediately after mitosis (Fig. 2b, d).
Furthermore, the aberrant mitotic translation seen in the 14-3-3s
knockdown cells did not result from a failure of these cells to enter
mitosis (Fig. 2c). Although the global pattern of cell cycle progres-
sion did not change, we observed an increase in the population of 4n
DNA-containing cells and a significant increase in mitotic index
(Fig. 2c, and Supplementary Fig. 3b, c). Coexpression of an shRNA-
resistant construct of 14-3-3stogether with 14-3-3sshRNA restored
the suppression of mitotic translation (Fig. 2e, lanes marked s
r
’,
and Fig. 2f, top). These findings indicate that 14-3-3sis important
1
Center for Cancer Research, Department of Biology and
2
Biological Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA.
3
Department of
Biochemistry, McGill Cancer Centre, McGill University, Montreal, Quebec, H3G 1Y6, Canada.
Vol 446
|
15 March 2007
|
doi:10.1038/nature05584
329
Nature
©2007 Publishing Group
in the physiological downregulation of new protein synthesis during
and immediately after mitosis.
The inhibition of translation observed during mitosis has prev-
iously been shown to result from a global decrease in cap-dependent
translation, whereas cap-independent translation is subsequently
increased
12,13
. We used two different IRES-containing bi-cistronic
vectors, a viral-based IRES from HIV
14
, and a normal cellular IRES
from p27
Kip1
(ref. 15), in a direct investigation of the relative effects of
14-3-3son cap-dependent versus cap-independent translation in
synchronized U2OS and HeLa cells. As shown in Fig. 2g, 14-3-3s-
depleted cells failed to show the robust mitotic-specific increase in
the ratio of cap-independent to cap-dependent translation that
occurred in control cells. Thus, a loss of 14-3-3sresulted in both
aberrant persistence of cap-dependent translation and prevented the
normal enhancement of cap-independent to cap-dependent trans-
lation during mitosis.
The relative amounts of different eukaryotic initiation factors in
cells varies more than 100-fold
16
, and they have distinct roles in the
cap recognition process
17,18
. To investigate which eIF targets of 14-3-
3sare potentially responsible for the 14-3-3s-mediated suppression
of cap-dependent mitotic translation, we immunodepleted mitotic
extracts with the use of anti-14-3-3santibodies and probed the
depleted lysates for residual eIFs. Of the eIFs examined, the most
pronounced result was obtained with eIF4B, which was completely
absent from the 14-3-3s-depleted mitotic lysates (Fig. 1c). eIF4B
facilitates the ATP-dependent helicase activity of eIF4A to promote
ribosome recruitment required for cap-dependent translation
19
.We
further observed that transient overexpression of exogenous Flag-
tagged eIF4B was able to overcome the ability of endogenous 14-3-
3sto suppress mitotic translation in normal cells but had no effect on
the increased translation seen in 14-3-3sknockdown cells (Fig. 2e,
lanes marked ‘4B’, and Fig. 2f, bottom). Taken together, the data in
Figs 1 and 2 indicate that 14-3-3sis important in the physiological
downregulation of new protein synthesis during and immediately
after mitosis by suppressing cap-dependent translation through
binding to eIF4B.
Because loss of 14-3-3soccurs early in the process of tumori-
genesis, we also examined the 14-3-3sknockdown cells for a corres-
ponding cellular phenotype that might correlate with aberrant
regulation of mitotic translation. In the 14-3-3sknockdown cells
we observed significantly increased numbers of cells displaying
persistent cytokinetic bridges, binucleate cells (often containing mid-
body remnants), and cells that seemed to have ‘fused’ after mitosis
with widely separated nuclei, probably reflecting a failure late in
cytokinesis (Fig. 3a–c; P,0.05 for 14-3-3sshRNA versus control
cells, Student’s t-test, two-tailed; and Supplementary Fig. 3d–f).
Intriguingly, these phenotypes seemed to correlate with a failure of
Polo-like kinase-1 (Plk1), a critical mitotic kinase involved in the
completion of cytokinesis
20,21
, to localize to the midbody at the end
of mitosis (Fig. 4a). Both the localization of Plk1 to the midbody and
the mitotic phenotype of 14-3-3s-depleted cells could be signifi-
cantly restored by treatment of the cells with rapamycin, an inhibitor
of cap-dependent, but not cap-independent, translation
22
, immedi-
ately before mitotic entry (Figs 3b, c and 4a).
We searched for proteins that were translated in mitosis in a cap-
independent manner and were postulated to be involved in the mitotic
process. One of the candidates, PITSLRE/Cdk11, is a member of the
Cdc2-like protein kinase family that undergoes cap-independent
translation from an internal ribosome entry site during mitosis to
produce a 58-kDa isoform that facilitates proper mitotic progression
and termination
23–25
. In both U2OS cells and HeLa cells we observed
that 14-3-3sknockdown cells, but not control cells, failed to synthe-
size p58 PITSLRE in mitosis (Fig. 4b, c, 16 h lanes). Depletion of
PITSLRE/Cdk11 from U2OS cells caused an increase in mitotic con-
tent as well as the number of binucleate cells, exactly as observed in
14-3-3s-depleted cells (Supplementary Figs 3 and 4). Addition of
rapamycin to 14-3-3sknockdown cells restored the mitotic trans-
lation of p58 PITSLRE(Fig. 4b, c). Furthermore, transienttransfection
of a complementary DNA encoding the 58-kDa isoform of PITSLRE
into the 14-3-3sknockdown cells was sufficient to relocalize Plk1 at
the midbody and partly rescue the mitosis-defective cell phenotype,
reducing the number of binucleate and fused cells by slightly more
a
b
c
d
e
A048 12162024h
Time after release from T/T block
14-3-3σ
Other
RNA/
transcription
Mitotic/
histones
Structural
Translation
+++–
14-3-3σ depletion
01216h
eIF4B
eIF4G
14-3-3σ
Blot
σ far-
western Immunoblot Merge
EF1α
eIF2α
eIF4B
eIF4G
IgG
pI 311
0 h
12 h
16 h
14-3-3σ
14-3-3σ
14-3-3σ
eIF4B
eIF2α
EF1α
Lysates
Figure 1
|
14-3-3sbinds to its targets during mitosis. a, SDS–PAGE
analysis of 14-3-3simmunoprecipitates from synchronized U2OS cells.
Sypro-Ruby stain. T/T, double thymidine; A, asynchronous cells. b, Mitotic
ligands of 14-3-3sidentified by mass spectrometry categorized into five
major groups. c, 14-3-3squantitatively immunodepletes eIF4B but not
eIF4G from mitotic (12 h), but not from interphase (0 h), U2OS cell extracts.
d, Two-dimensional far-western blotting of 14-3-3simmunoprecipitates,
with purified 14-3-3sas a probe for direct binding. pI, isoelectric point.
e, eIF4B, eIF2aand EF1aas direct mitotic ligands of 14-3-3s. Two-
dimensional far-western blots (green) re-probed with antibodies against the
indicated proteins (red). eIF4G in the 14-3-3simmunoprecipitates does not
directly bind to 14-3-3s.
LETTERS NATURE
|
Vol 446
|
15 March 2007
330
Nature
©2007 Publishing Group
than 50% (Fig. 4d, e, and Supplementary Fig. 4i). Thus, 14-3-3sis
required in normal mitosis to suppress cap-dependent translation
through binding eIF4B, allowing the cap-independent translation of
critical mitotic regulators including p58 PITSLRE kinase.
Two recent reports
26,27
have suggested that the formation of
binucleate cells as a result of cytokinesis failure is an early event in
tumour formation and underlies the subsequent development of
genomic instability. We have observed that 14-3-3s, an important
tumour suppressor protein whose expression is lost in a variety of
epithelial tumours, has a critical function in regulating protein syn-
thesis during and immediately after mitosis and is required for the
IRES-dependent translation of p58 PITSLRE, a protein kinase critical
to the proper completion of cytokinesis. These findings therefore
establish a connection between the aberrant regulation of mitotic
translation and improper cytokinesis resulting in a phenotype that
is associated with early stages of human oncogenesis.
METHODS
Immunoprecipitation studies and mass spectrometry analysis. U2OS, HeLa
and HCT116 cell lysates were prepared and immunoprecipitated overnight at
4uC with an anti-14-3-3smonoclonal antibody (CS112)
28,29
. Immunopre-
cipitates were digested with trypsin and analysed on a QSTAR XL Pro quad-
rupole time-of-flight mass spectrometer (Applied Biosystems).
Two-dimensional gel electrophoresis. 14-3-3simmunoprecipitates were
resolved by isoelectric focusing with IPGphor strips as described previously
29
.
Separation in the second dimension was performed with 12% Anderson gels and
used for far-western blot analyses with bacterially purified 14-3-3s(about
50 mgml
21
) as a probe, followed by immunoblotting with CS112.
Lentiviral shRNA constructs. 14-3-3s, 14-3-3band p58 PITSLRE shRNAs were
subcloned into pLentilox 3.7 for lentiviral production as described
30
. Heter-
geneous populations stably expressing shRNA were established in HeLa and
U2OS cells.
[
35
S]Methionine pulse-labelling. Cells were labelled with 100 mCi of
[
35
S]methionine for 20 min, lysed, and precipitated with acetone. Resuspended
α-tubulin DAPI
Multinucleate
Fused
Merged
Binucleate Fused
Control
14-3-3σ knockdown
+Rap. *
+Rap. *
+Rap. *
+Rap. *
Percentage
10 12 14 16 18 20 24 10 12
Time (h)
14 16 18 20 24
8
4
0
12
16
8
6
4
2
0
a
bc
Figure 3
|
Depletion of 14-3-3sresults in impaired cytokinesis. a, 14-3-3s
knockdown U2OS cells, stained as indicated, show increased numbers of
binucleate cells (short arrows) and cells in terminal cytokinesis
(arrowheads). Close-up views demonstrate persistent midbody structures
(long arrow), and a ‘fused’ post-mitotic cell lacking distinct intercellular
boundaries. Scale bars, 10 mm. DAPI, 4,6-diamidino-2-phenylindole.
b,c, Rapamycin treatment before mitotic entry suppresses binucleate (b) and
fused-cell (c) formation in 14-3-3sknockdown cells. Synchronized U2OS
shRNA cells were scored at the indicated times after release. Green bars,
control shRNA; black bars, 14-3-3sshRNA; 1Rap., 14-3-3sshRNA cells
treated with 20 ng ml
21
rapamycin at 9 h. Asterisk, P,0.02 for rapamycin-
treated versus untreated 14-3-3sknockdown cells. Results are means and
s.d. from three independent experiments.
ab
cd
e
shRNA
150-
kDa kDa
75-
50-
25-
Con
σ
Con
σ
Con
σ
Con
σ
1 h 12 h 16 h 20 h
012 16h h012 16
14-3-3β
Control
shRNA
150-
75-
50-
25-
14-3-3σ
Control
4 h
8 h
10 h
12 h
14 h
16 h
24 h
2n4n2n 4n
Blot
shRNA Cσβ
14-3-3β
14-3-3σ
β-actin
0 h 12 h
Con
σ
σ
Con
σ
σ
Con
σ
shRNA
Plasmid vv vv 4B4B
σrσr
0121610
012
Time (h)
169
Ratio of IRES-dependent to
cap-dependent translation
0
4
1
2
3
Blot
Blot 14-3-3σ
shRNA
shRNA
Con
σCon
σ
++–
++
Vector
Flag–eIF4B
Flag
β-actin
++
+––
++–
+
pTre vector:
pTre sigma (σr)
Con
σ
Con
σ
σ
σ
0 h 12 h
12 h
fg
0
1.0
2.0
1.5
2.5
0.5
Figure 2
|
14-3-3sknockdown cells fail to suppress cap-dependent
translation during and immediately after mitosis. a,b, Control, 14-3-3s
(a) and 14-3-3b(b) shRNA HeLa cells were pulse-labelled with
[
35
S]methionine. New protein synthesis was assayed by SDS–PAGE and
autoradiography. Numbers at the left indicate molecular masses. c, Cell cycle
distribution of control and 14-3-3sshRNA-treated cells. d, Efficiency of
14-3-3 knockdown. C, control. e, Overexpression of eIF4B overcomes
14-3-3s-dependent suppression of translation in mitosis. V, vector; 4B,
Flag-eIF4B; s
r
, RNAi-resistant 14-3-3s. Synchronized cells pulse-labelled as
in bat indicated times. f, Expression of s
r
and Flag–eIF4B in e.g, Ratio of
cap-independent to cap-dependent translation in synchronized control
(light grey bars) and 14-3-3s(dark grey bars) knockdown cells assayed by
firefly/Renilla luciferase activity at the indicated times. Top, U2OS cells and
HIV IRES; bottom, HeLa cells and p27
Kip1
IRES. Results are means and
s.e.m. from duplicate experiments.
NATURE
|
Vol 446
|
15 March 2007 LETTERS
331
Nature
©2007 Publishing Group
proteins were separated on 12% SDS–PAGE gels and analysed by autoradio-
graphy. Where indicated, cells were treated with rapamycin at a final concentra-
tion of 20 ng ml
21
9 h after release from a double thymidine block.
Additional information and detailed protocols are provided in Supplementary
Methods.
Received 13 July 2006; accepted 9 January 2007; corrected 10 April 2007.
1. Vercoutter-Edouart, A. S. et al. Proteomic analysis reveals that 14-3-3s
is down-regulated in human breast cancer cells. Cancer Res. 61, 76
80
(2001).
2. Ferguson, A. T. et al. High frequency of hypermethylation at the 14-3-3slocus
leads to gene silencing in breast cancer. Proc. Natl Acad. Sci. USA 97, 6049
6054
(2000).
3. Umbricht, C. B. et al. Hypermethylation of 14-3-3s(stratifin) is an early event in
breast cancer. Oncogene 20, 3348
3353 (2001).
4. Urano, T. et al. Efp targets 14-3-3sfor proteolysis and promotes breast tumour
growth. Nature 417, 871
875 (2002).
5. Moreira, J. M., Gromov, P. & Celis, J. E. Expression of the tumor suppressor protein
14-3-3sis down-regulated in invasive transitional cell carcinomas of the urinary
bladder undergoing epithelial-to-mesenchymal transition. Mol. Cell. Proteomics 3,
410
419 (2004).
6. Iwata, N. et al. Frequent hypermethylation of CpG islands and loss of expression of
the 14-3-3sgene in human hepatocellular carcinoma. Oncogene 19, 5298
5302
(2000).
7. Suzuki, H. et al. Inactivation of the 14-3-3sgene is associated with 59
CpG island hypermethylation in human cancers. Cancer Res. 60, 4353
4357
(2000).
8. Villaret, D. B. et al. Identification of genes overexpressed in head and neck
squamous cell carcinoma using a combination of complementary DNA
subtraction and microarray analysis. Laryngoscope 110, 374
381 (2000).
9. Gasco, M. et al. Coincident inactivation of 14-3-3sand p16INK4a is an early event
in vulval squamous neoplasia. Oncogene 21, 1876
1881 (2002).
10. Prescott, A. R. Synthesis of RNA and protein during mitosis in mammalian tissue
culture cells. Exp. Cell Res. 26, 260
268 (1962).
11. Pyronnet, S., Dostie, J. & Sonenberg, N. Suppression of cap-dependent translation
in mitosis. Genes Dev. 15, 2083
2093 (2001).
12. Pyronnet, S., Pradayrol, L. & Sonenberg, N. A cell cycle-dependent internal
ribosome entry site. Mol. Cell 5, 607
616 (2000).
13. Qin, X. & Sarnow, P. Preferential translation of internal ribosome entry site-
containing mRNAs during the mitotic cycle in mammalian cells. J. Biol. Chem. 279,
13721
13728 (2004).
14. Brasey, A. et al. The leader of human immunodeficiency virus type 1 genomic RNA
harbors an internal ribosome entry segment that is active during the G2/M phase
of the cell cycle. J. Virol. 77, 3939
3949 (2003).
15. Kullmann, M., Gopfert, U., Siewe, B. & Hengst, L. B. ELAV/Hu proteins inhibit p27
translation via an IRES element in the p27 59UTR. Genes Dev. 16, 3087
3099
(2002).
16. Hershey, J. W. Expression of initiation factor genes in mammalian cells. Biochimie
76, 847
852 (1994).
17. Holcik, M. & Sonenberg, N. Translational control in stress and apoptosis. Nature
Rev. Mol. Cell Biol. 6, 318
327 (2005).
18. Merrick, W. C. Cap-dependent and cap-independent translation in eukaryotic
systems. Gene 332, 1
11 (2004).
19. Rogers, G. W. Jr, Richter, N. J. & Merrick, W. C. Biochemical and kinetic
characterization of the RNA helicase activity of eukaryotic initiation factor 4A. J.
Biol. Chem. 274, 12236
12244 (1999).
20. Neef, R. et al. Phosphorylation of mitotic kinesin-like protein 2 by polo-like kinase 1
is required for cytokinesis. J. Cell Biol. 162, 863
875 (2003).
21. Liu, X., Zhou, T., Kuriyama, R. & Erikson, R. L. Molecular interactions of Polo-like-
kinase 1 with the mitotic kinesin-like protein CHO1/MKLP-1. J. Cell Sci. 117,
3233
3246 (2004).
22. Beretta, L., Gingras, A. C., Svitkin, Y. V., Hall, M. N. & Sonenberg, N. Rapamycin
blocks the phosphorylation of 4E-BP1 and inhibits cap-dependent initiation of
translation. EMBO J. 15, 658
664 (1996).
23. Li, T., Inoue, A., Lahti, J. M. & Kidd, V. J. Failure to proliferate and mitotic arrest of
CDK11
p110/p58
-null mutant mice at the blastocyst stage of embryonic cell
development. Mol. Cell. Biol. 24, 3188
3197 (2004).
24. Petretti, C. et al. The PITSLRE/CDK11p58 protein kinase promotes
centrosome maturation and bipolar spindle formation. EMBO Rep. 7, 418
424
(2006).
25. Cornelis, S. et al. Identification and charac terization of a novel cell cycle-regulated
internal ribosome entry site. Mol. Cell 5, 597
605 (2000).
26. Shi, Q. & King, R. W. Chromosome nondisjunction yields tetraploid
rather than aneuploid cells in human cell lines. Nature 437, 1038
1042
(2005).
27. Fujiwara, T. et al. Cytokinesis failure generating tetraploids promotes
tumorigenesis in p53-null cells. Nature 437, 1043
1047 (2005).
28. Brunet, A. et al. 14-3-3 transits to the nucleus and participates in dynamic
nucleocytoplasmic transport. J. Cell Biol. 156, 817
828 (2002).
29. Wilker, E. W., Grant, R. A., Artim, S. C. &Yaffe, M. B. A structuralbasis for 14-3-3s
functional specificity. J. Biol. Chem. 280, 18891
18898 (2005).
30. Rubinson, D. A. et al. A lentivirus-based system to functionally silence genes in
primary mammalian cells, stem cells and transgenic mice by RNA interference.
Nature Genet. 33, 401
406 (2003).
Supplementary Information is linked to the online version of the paper at
www.nature.com/nature.
Acknowledgements We thank P. Stern, D. Lowery and W. Merrick for reagents and
technical assistance. This work was supported by postdoctoral fellowships from
the Anna Fuller Fund and the NIH to E.W.W., an EMBO long-term fellowship to
M.A.T.M.v.V., the David H. Koch Cancer Research Fund, NIH grants, and a
Burroughs-Wellcome Career Development Award to M.B.Y.
Author Information Reprints and permissions information is available at
www.nature.com/reprints. The authors declare no competing financial interests.
Correspondence and requests for materials should be addressed to M.B.Y.
(myaffe@mit.edu).
a
b
Anti-Plk1 Phalloidin DAPI Merged
Control
Mock-
treated+ Rapamycin
14-3-3σ knockdown
Rapamycin
0 h 10.5 h 16 h
++–– –– ––
CKDKD C KD
16 h 18 h
Blot:
anti-p58
Blot:
anti-p58
+
–– ––
Rapamycin
c
d
e
GFP
GFP
DAPI Merged
Anti-Plk1 DAPI Mer
g
ed
14-3-3σ knockdown
+p58
10
8
6
4
2
0
Percentage
Multinucleate Fused
CC CKD KDKDKDKD Control
σ shRNA
σ shRNA + p58
Figure 4
|
Depletion of 14-3-3sblocks IRES-dependent mitotic translation
of p58 PITSLRE kinase and results in failure of Plk1 to localize at the
midbody. a, Lack of localization of Plk1 to the midbody and cell fusion with
failure to complete cytokinesis in 14-3-3sknockdown U2OS cells is reversed
by rapamycin before mitotic entry. Scale bar, 5 mm. b,c, Synchronized
control (C) or 14-3-3sshRNA-treated (KD) U2OS (b) and HeLa
(c) cells, with and without rapamycintreatment, lysed at the indicated times
and immunoblotted for endogenous p58 PITSLRE expression
25
. IRES-
dependent translation of p58 PITSLRE in 14-3-3sknockdown cells was
restored by rapamycin. d,e, Forced expression of p58 PITSLRE reverts the
mitotic phenotype of 14-3-3sknockdown cells. GFP-expressing 14-3-3s
knockdown U2OS cells were transfected with p58 PITSLRE or vector
control, then imaged (d) and scored (e) 18 h after release from a double-
thymidine block. Results in eare means and s.d. from three independent
experiments.
LETTERS NATURE
|
Vol 446
|
15 March 2007
332
Nature
©2007 Publishing Group
... Earlier studies support the notion that mitotic translational repression is achieved via the inhibition of the initiation stage through downregulation of cap-dependent translation (Cornelis et al. 2000) (Pyronnet, Dostie, and Sonenberg 2001) (Qin and Sarnow 2004) (Wilker et al. 2007). On the other hand, more recent work provides evidence that blockade of translational elongation could also be contributing to mitotic translational repression (Sivan, Kedersha, and Elroy-Stein 2007)(Sivan and Elroy-Stein 2008). ...
Preprint
Full-text available
Modulation of protein synthesis according to the physiological cues is maintained through tight control of Eukaryotic Elongation Factor 2 (eEF2), whose unique translocase activity is essential for cell viability. Phosphorylation of eEF2 at its Thr56 residue inactivates this function in translation. In our previous study we reported a novel mode of post-translational modification that promotes higher efficiency in T56 phosphorylation. Cyclin A/CDK2-mediated phosphorylation of eEF2 at the S595 residue is required for more potent phosphorylation at the Thr56, suggesting CDK2 takes a role in robust suppression of protein synthesis. In the current study, we analyzed the cell cycle, proliferation, cell death, migration, colony formation, autophagy, and response to Cisplatin properties of the point-mutant variant of HCT116 cells that express the CDK2 mutant (S595A-eEF2) of eEF2. The knocked in S595A mutation resulted in decreased levels of T56 phosphorylation of eEF2, which appears to have similar biological consequences to other experimental manipulations such as silencing the activity of the kinase for the Thr56 residue, eEF2 Kinase (eEF2K). Our findings indicate that interfering with the inhibition of eEF2 results in elevated protein synthesis in HCT116 cells and is associated with the progression of malignancy in the colorectal cancer cell line, where eEF2K activity could provide a tumor suppressive role.
... 20,21 For example, 14-3-3σ (also known as stratifin or SFN) is originally characterized as a human mammary epithelial-specific marker that is downregulated in mammary carcinoma cells. 22 14-3-3σ has a unique role in regulating cell cycle progression [23][24][25] and epithelial differentiation because of their restricted expression, primarily in differentiated epithelial cells. 26,27 We previously identified that 14-3-3β interacts with FK506 binding protein (FKBP) 12 and synaptopodin (Synp) to maintain the structure of actin filaments in podocytes. ...
Article
Full-text available
14‐3‐3 proteins are a ubiquitously expressed family of adaptor proteins. Despite exhibiting high sequence homology, several 14‐3‐3 isoforms have isoform‐specific binding partners and roles. We reported that 14‐3‐3β interacts with FKBP12 and synaptopodin to maintain the structure of actin fibers in podocytes. However, the precise localization and differential role of 14‐3‐3 isoforms in kidneys are unclear. Herein, we showed that 14‐3‐3β in glomeruli was restricted in podocytes, and 14‐3‐3σ in glomeruli was expressed in podocytes and mesangial cells. Although 14‐3‐3β was dominantly co‐localized with FKBP12 in the foot processes, a part of 14‐3‐3β was co‐localized with Par3 at the slit diaphragm. 14‐3‐3β interacted with Par3, and FKBP12 bound to 14‐3‐3β competitively with Par3. Deletion of 14‐3‐3β enhanced the interaction of Par3 with Par6 in podocytes. Gene silencing for 14‐3‐3β altered the structure of actin fibers and process formation. 14‐3‐3β and synaptopodin expression was decreased in podocyte injury models. In contrast, 14‐3‐3σ in podocytes was expressed in the primary processes. 14‐3‐3σ interacted with vimentin but not with the actin‐associated proteins FKBP12 and synaptopodin. Gene silencing for 14‐3‐3σ altered the structure of vimentin fibers and process formation. 14‐3‐3σ and vimentin expression was increased in the early phase of podocyte injury models but was decreased in the late stage. Together, the localization of 14‐3‐3β at actin cytoskeleton plays a role in maintaining the foot processes and the Par complex in podocytes. In contrast, 14‐3‐3σ at vimentin cytoskeleton is essential for maintaining primary processes.
... Considering the steric hindrance effect of dCas13 on scanning 40S ribosomes, another mechanism of translational repression by the fused protein may further benefit translational repression ( Figure 5A). Here, we tested several translational suppressor proteins, including programmed cell death 4 (PDCD4) [94][95][96][97] , eIF4E-transporter protein (4E-T) 98-100 , eIF6 101,102 , 14-3-3σ 103 , poly(A)-binding protein-interacting protein 2 (PAIP2) [104][105][106][107][108][109] , Pelota (PELO) 110-114 , and 4EHP 115-120 , to dPspCas13b-NES targeting the start codon of the reporter and found that 4EHP exerted the most significant effects ( Figure S5B). The recruitment of 4EHP-fused dPspCas13b-NES not only to the start codon but also to the 5′ UTR outperformed that of the nonfused variant ( Figure 5B). ...
Preprint
Full-text available
Current gene silencing tools based on RNA interference (RNAi) or, more recently, clustered regularly interspaced short palindromic repeats (CRISPR)‒Cas13 systems, have critical drawbacks, such as off-target effects (RNAi) or collateral mRNA cleavage (CRISPR‒Cas13). Thus, a more specific method of gene knockdown is needed. Here, we developed "CRISPRδ", an approach for translational silencing, harnessing catalytically inactive Cas13 proteins (dCas13). Owing to its tight association with mRNA, dCas13 serves as a physical roadblock for scanning ribosomes during translation initiation and does not affect mRNA stability. Guide RNAs covering the start codon lead to the highest efficacy regardless of the translation initiation mechanism: cap-dependent or internal ribosome entry site (IRES)-dependent translation. Strikingly, genome-wide ribosome profiling revealed the extremely high gene knockdown specificity of CRISPRδ. Moreover, fusion of a translational repressor to dCas13 ensured further improvement of the knockdown efficacy. Our method provides a framework for translational repression-based gene silencing in eukaryotes.
... However, more studies are required to identify all TOR -responsive phosphorylation sites within RPS6.A second S6K1 downstream target is eIF4B. The Ser422 phosphorylation site of eIF4B is involved in mRNA translation control-phosphorylation of this site augments the helicase activity of eIF4A, and eIF4B recruitment to the 48S PICWILKER et al. ...
Thesis
Translation of mRNAs that harbor upstream open reading frames (uORFs) within their leader regions operates via a reinitiation mechanism. In plants, reinitiation is up regulated by the target of rapamycin (TOR) signaling via phosphorylation of the subunit h of initiation factor 3 (eIF3). The eif3h-1 mutant expressing the C-terminally truncated eIF3h while maintaining high translation initiation efficiency is not active in reinitiation. Cauliflower mosaic virus (CaMV) pregenomic polycistronic RNA is translated via an exceptional mechanism of reinitiation after long ORF translation under control of CaMV protein TAV, which ensures activation of TOR. To find the link between underlying mechanisms, we examined eIF3h function in cellular and viral context. Here we show that eIF3h, if phosphorylated, has a role in recruitment of eIF3 into actively translating ribosomes that is a prerequisite for formation of reinitiation-competent ribosomal complexes. C-terminal truncation of eIF3h abolished its integration into the eIF3 complex and eIF3 loading on polysomes as manifested by the eIF3 core subunit c. We also show that eIF3h as a putative target of TOR/S6K1 binds S6K1 in vitro. eIF3h phosphorylation is not required for eIF3 complex formation. We demonstrated that eIF3h is essential for TAV to activate reinitiation after long ORF translation. Protoplasts derived from eif3h-1 mutant failed to support TAV function in reinitiation, which is restored only upon overexpression of recombinant eIF3h indifferent to its phosphorylation status. eif3h-1 mutant defective in reinitiation was found resistant to CaMV infection suggesting that eIF3h is critical for virus amplification. In contrast, viruses that evolve translation initiation dependent on either cap or the internal ribosome entry site infect reinitiation deficient mutant. Thus, we conclude that TAV exploits the basic cell reinitiation machinery, particularly TOR and eIF3h, to overcome cellular barriers to reinitiation after long ORF translation.
Article
Full-text available
Current gene silencing tools based on RNA interference (RNAi) or, more recently, clustered regularly interspaced short palindromic repeats (CRISPR)‒Cas13 systems have critical drawbacks, such as off-target effects (RNAi) or collateral mRNA cleavage (CRISPR‒Cas13). Thus, a more specific method of gene knockdown is needed. Here, we develop CRISPRδ, an approach for translational silencing, harnessing catalytically inactive Cas13 proteins (dCas13). Owing to its tight association with mRNA, dCas13 serves as a physical roadblock for scanning ribosomes during translation initiation and does not affect mRNA stability. Guide RNAs covering the start codon lead to the highest efficacy regardless of the translation initiation mechanism: cap-dependent, internal ribosome entry site (IRES)-dependent, or repeat-associated non-AUG (RAN) translation. Strikingly, genome-wide ribosome profiling reveals the ultrahigh gene silencing specificity of CRISPRδ. Moreover, the fusion of a translational repressor to dCas13 further improves the performance. Our method provides a framework for translational repression-based gene silencing in eukaryotes.
Article
A family of structurally related cyclin‐dependent protein kinases (CDKs) drives many aspects of eukaryotic cell function. Much of the literature in this area has considered individual members of this family to act primarily either as regulators of the cell cycle, the context in which CDKs were first discovered, or as regulators of transcription. Until recently, CDK7 was the only clear example of a CDK that functions in both processes. However, new data points to several “cell‐cycle” CDKs having important roles in transcription and some “transcriptional” CDKs having cell cycle‐related targets. For example, novel functions in transcription have been demonstrated for the archetypal cell cycle regulator CDK1. The increasing evidence of the overlap between these two CDK types suggests that they might play a critical role in coordinating the two processes. Here we review the canonical functions of cell‐cycle and transcriptional CDKs, and provide an update on how these kinases collaborate to perform important cellular functions. We also provide a brief overview of how dysregulation of CDKs contributes to carcinogenesis, and possible treatment avenues. This article is categorized under: RNA Interactions with Proteins and Other Molecules > RNA‐Protein Complexes RNA Processing > 3′ End Processing RNA Processing > Splicing Regulation/Alternative Splicing
Article
Full-text available
Long ignored as a vestigial remnant of cytokinesis, the mammalian midbody (MB) is released post-abscission inside large extracellular vesicles called MB remnants (MBRs). Recent evidence suggests that MBRs can modulate cell proliferation and cell fate decisions. Here, we demonstrate that the MB matrix is the site of ribonucleoprotein assembly and is enriched in mRNAs that encode proteins involved in cell fate, oncogenesis, and pluripotency, which we are calling the MB granule. Both MBs and post-abscission MBRs are sites of spatiotemporally regulated translation, which is initiated when nascent daughter cells re-enter G1 and continues after extracellular release. MKLP1 and ARC are necessary for the localization and translation of RNA in the MB dark zone, whereas ESCRT-III is necessary to maintain translation levels in the MB. Our work reveals a unique translation event that occurs during abscission and within a large extracellular vesicle.
Article
Full-text available
14-3-3 proteins play critical roles in controlling multiple aspects of the cellular response to stress and DNA damage including regulation of metabolism, cell cycle progression, cell migration, and apoptotic cell death by binding to protein substrates of basophilic protein kinases following their phosphorylation on specific serine/threonine residues. Although over 200 mammalian proteins that bind to 14-3-3 have been identified, largely through proteomic studies, in many cases the relevant protein kinase responsible for conferring 14-3-3-binding to these proteins is not known. To facilitate the identification of kinase-specific 14-3-3 clients, we developed a biochemical approach using high-density protein filter arrays and identified the translational regulatory molecule PABPC1 as a substrate for Chk1 and MAPKAP Kinase-2 (MK2) in vitro , and for MK2 in vivo , whose phosphorylation results in 14-3-3-binding. We identify Ser-470 on PABPC1 within the linker region connecting the RRM domains to the PABC domain as the critical 14-3-3-binding site, and demonstrate that loss of PABPC1 binding to 14-3-3 results in increased cell proliferation and decreased cell death in response to UV-induced DNA damage.
Article
Objective: Hepatocellular carcinoma (HCC) is the second leading cancer cause of death worldwide. SFN plays a vital role in some malignancies. The purpose of this study was to investigate the role of SFN in the development of HCC. Methods: The bioinformatics database was used to detect the expression of SFN and its prognosis in HCC patients. And the protein-protein interaction network was established. IHC and Elisa were used to analyze the expression level and clinical characteristics of SFN in HCC patients. Subsequently, siRNA knockdown of SFN expression in HCC cell lines was used to explore whether SFN could promote the development of HCC. Results: SFN was highly expressed in the tissues and serum of hepatocellular carcinoma, and its expression level was correlated with the tumor which was single or not in patients. Bioanalysis and histochemistry results showed that CDC25B was co-expressed with SFN in HCC, which may be the upstream and downstream signaling molecule of SFN. Knockdown of SFN can inhibit cell proliferation, migration, invasion and promote apoptosis. Conclusions: Our results suggest that SFN may play an important role in HCC progression and may interact with CDC25B to promote malignant progression of HCC, providing a molecular target for future HCC therapy.
Preprint
Full-text available
The midbody (MB) is a transient structure at the spindle midzone that is required for cytokinesis, the terminal stage of cell division. Long ignored as a vestigial remnant of cytokinesis, we now know MBs are released post-abscission as extracellular vesicles called MB remnants (MBRs) and can modulate cell proliferation, fate decisions, tissue polarity, neuronal architecture, and tumorigenic behavior. Here, we demonstrate that the MB matrix—the structurally amorphous MB core of unknown composition—is the site of ribonucleoprotein assembly and is enriched in mRNAs that encode proteins involved in cell fate, oncogenesis, and pluripotency, that we are calling the MB granule. Using a quantitative transcriptomic approach, we identified a population of mRNAs enriched in mitotic MBs and confirmed their presence in signaling MBR vesicles released by abscission. The MB granule is unique in that it is translationally active, contains both small and large ribosomal subunits, and has both membrane-less and membrane-bound states. Both MBs and post-abscission MBRs are sites of spatiotemporally regulated translation, which is initiated when nascent daughter cells re-enter G1 and continues after extracellular release. We demonstrate that the MB is the assembly site of an RNP granule. MKLP1 and ARC are necessary for the localization and translation of RNA in the MB dark zone, whereas ESCRT-III was necessary to maintain translation levels in the MB. Our data suggest a model in which the MB functions as a novel RNA-based organelle with a uniquely complex life cycle. We present a model in which the assembly and transfer of RNP complexes are central to post-mitotic MBR function and suggest the MBR serves as a novel mode of RNA-based intercellular communication with a defined biogenesis that is coupled to abscission, and inherently links cell division status with signaling capacity. To our knowledge, this is the first example of an autonomous extracellular vesicle with active translation activity.
Article
Full-text available
The immunosuppressant drug rapamycin blocks progression of the cell cycle at the G1 phase in mammalian cells and yeast. Here we show that rapamycin inhibits cap-dependent, but not cap-independent, translation in NIH 3T3 cells. Cap-dependent translation is also specifically reduced in extracts from rapamycin-treated cells, as determined by in vitro translation experiments. This inhibition is causally related to the dephosphorylation and consequent activation of 4E-BP1, a protein recently identified as a repressor of the cap-binding protein, eIF-4E, function. These effects of rapamycin are specific as FK506, a structural analogue of rapamycin, had no effect on either cap-dependent translation or 4E-BP1 phosphorylation. The rapamycin-FK506 binding protein complex is the effector of the inhibition of 4E-BP1 phosphorylation as excess of FK506 over rapamycin reversed the rapamycin-mediated inhibition of 4E-BP1 phosphorylation. Thus, inactivation of eIF-4E is, at least in part, responsible for inhibition of cap-dependent translation in rapamycin-treated cells. Furthermore, these results suggest that 4E-BP1 phosphorylation is mediated by the FRAP/TOR signalling pathway.
Article
Full-text available
We have identified 14-3-3 σ (σ) as a gene whose expression is lost in breast carcinomas, primarily by methylation-mediated silencing. In this report, we investigated the timing of loss of σ gene expression during breast tumorigenesis in vivo. We analysed the methylation status of σ in breast cancer precursor lesions using microdissection for selective tissue sampling. We found hypermethylation of σ in 24 of 25 carcinomas (96%), 15 of 18 (83%) of ductal carcinoma in situ, and three of eight (38%) of atypical hyperplasias. None of the five hyperplasias without atypia showed σ-hypermethylation. Unexpectedly, patients with breast cancer showed σ hypermethylation in adjacent histologically normal breast epithelium, while this was never observed in individuals without evidence of breast cancer. Also, samples of periductal stromal breast tissue were consistently hypermethylated, underscoring the importance of selective tissue sampling for accurate assessment of 14-3-3-σ methylation in breast epithelium. These results suggest that hypermethylation of 14-3-3-σ occurs at an early stage in the progression to invasive breast cancer, and may occur in apparently normal epithelium adjacent to breast cancer. These results provide evidence that loss of expression of σ is an early event in neoplastic transformation.
Article
Full-text available
Cap-dependent translation is mediated by eIF4F, a protein complex composed of three subunits as follows: eIF4E, which recognizes the mRNA 5′ cap structure; eIF4A, an RNA-helicase; and eIF4G, a scaffolding protein that binds eIF4E, eIF4A, and the eIF4E-kinase Mnk1 simultaneously. eIF4E is hypophosphorylated and cap-dependent translation is reduced at mitosis. Here, we show that 4E-BP1, a suppressor of eIF4E function, is also hypophosphorylated in mitosis, resulting in disruption of the eIF4F complex. Consequently, eIF4E is sequestered from the eIF4G/Mnk1 complex. These results explain the specific inhibition of cap-dependent translation in mitosis and also explain how eIF4E is rendered hypophosphorylated during mitosis. Furthermore, eIF4E interaction with eIF4GII is strongly decreased coincident with hyperphosphorylation of eIF4GII. Thus, inhibition of cap-dependent translation in mitosis results from a combination of phosphorylation modifications leading to eIF4F complex disruption.
Article
Full-text available
The immunosuppressant drug rapamycin blocks progression of the cell cycle at the G1 phase in mammalian cells and yeast. Here we show that rapamycin inhibits cap-dependent, but not cap-independent, translation in NIH 3T3 cells. Cap-dependent translation is also specifically reduced in extracts from rapamycin-treated cells, as determined by in vitro translation experiments. This inhibition is causally related to the dephosphorylation and consequent activation of 4E-BP1, a protein recently identified as a repressor of the cap-binding protein, eIF-4E, function. These effects of rapamycin are specific as FK506, a structural analogue of rapamycin, had no effect on either cap-dependent translation or 4E-BP1 phosphorylation. The rapamycin-FK506 binding protein complex is the effector of the inhibition of 4E-BP1 phosphorylation as excess of FK506 over rapamycin reversed the rapamycin-mediated inhibition of 4E-BP1 phosphorylation. Thus, inactivation of eIF-4E is, at least in part, responsible for inhibition of cap-dependent translation in rapamycin-treated cells. Furthermore, these results suggest that 4E-BP1 phosphorylation is mediated by the FRAP/TOR signalling pathway.
Article
Full-text available
Eukaryotic initiation factor (eIF) 4A is the prototypic member of the DEAD box family of proteins and has been proposed to act as an RNA helicase to unwind secondary structure in the 5'-untranslated region of eukaryotic mRNAs. Previous studies have shown that the RNA helicase activity of eIF4A is dependent on the presence of a second initiation factor, eIF4B. In this report, eIF4A has been demonstrated to function independently of eIF4B as an ATP-dependent RNA helicase. The biochemical and kinetic properties of this activity were examined. By using a family of RNA duplexes with an unstructured single-stranded region followed by a duplex region of increasing length and stability, it was observed that the initial rate of duplex unwinding decreased with increasing stability of the duplex. Furthermore, the maximum amount of duplex unwound also decreased with increasing stability. Results suggest that eIF4A acts in a non-processive manner. eIF4B and eIF4H were shown to stimulate the helicase activity of eIF4A, allowing eIF4A to unwind longer, more stable duplexes with both an increase in initial rate and maximum amount of duplex unwound. A simple kinetic model is proposed to explain the mechanism by which eIF4A unwinds RNA duplex structures in an ATP-dependent manner.
Article
Full-text available
A cDNA encoding human eukaryotic initiation factor (eIF) 4H was subcloned into a bacterial expression plasmid for purification of recombinant protein. Recombinant human eIF4H (heIF4H) was purified to greater than 95% homogeneity and shown to have similar physical characteristics to eIF4H purified from rabbit reticulocyte lysate as described previously. Functional studies have revealed that recombinant heIF4H functions identically to rabbit eIF4H in stimulating protein synthesis, and the ATP hydrolysis and helicase activities of eIF4A. More detailed enzymatic studies revealed that eIF4H increases the affinity of eIF4A for RNA by 2-fold, but has no effect on the binding of ATP by eIF4A. eIF4H stimulates the helicase activity of eIF4A at least 4-fold, and it is postulated that this stimulation occurs through increasing the processivity of eIF4A. Northern blot analysis shows that eIF4H is expressed ubiquitously in human tissues, and displays different levels of expression in given tissues relative to eIF4B. Secondary structure analysis of heIF4H by circular dichroism suggest that eIF4H has a mostly beta-sheet structure, which appears similar to other RNA recognition motif-containing proteins. Finally, it is suggested that eIF4H functions in translation initiation through protein-protein interactions that possibly stabilize conformational changes that occur in eIF4A during RNA binding, ATP hydrolysis, and RNA duplex unwinding.
Article
Full-text available
Expression of 14-3-3 final sigma (final sigma) is induced in response to DNA damage, and causes cells to arrest in G(2). By SAGE (serial analysis of gene expression) analysis, we identified final sigma as a gene whose expression is 7-fold lower in breast carcinoma cells than in normal breast epithelium. We verified this finding by Northern blot analysis. Remarkably, final sigma mRNA was undetectable in 45 of 48 primary breast carcinomas. Genetic alterations at final sigma such as loss of heterozygosity were rare (1/20 informative cases), and no mutations were detected (0/34). On the other hand, hypermethylation of CpG islands in the final sigma gene was detected in 91% (75/82) of breast tumors and was associated with lack of gene expression. Hypermethylation of final sigma is functionally important, because treatment of final sigma-non-expressing breast cancer cell lines with the drug 5-aza-2'-deoxycytidine resulted in demethylation of the gene and synthesis of final sigma mRNA. Breast cancer cells lacking final sigma expression showed increased number of chromosomal breaks and gaps when exposed to gamma-irradiation. Therefore, it is possible that loss of final sigma expression contributes to malignant transformation by impairing the G(2) cell cycle checkpoint function, thus allowing an accumulation of genetic defects. Hypermethylation and loss of final sigma expression are the most consistent molecular alterations in breast cancer identified so far.
Article
RNA interference (RNAi) has recently emerged as a specific and efficient method to silence gene expression in mammalian cells either by transfection of short interfering RNAs (siRNAs; ref. 1) or, more recently, by transcription of short hairpin RNAs (shRNAs) from expression vectors and retroviruses. But the resistance of important cell types to transduction by these approaches, both in vitro and in vivo, has limited the use of RNAi. Here we describe a lentiviral system for delivery of shRNAs into cycling and non-cycling mammalian cells, stem cells, zygotes and their differentiated progeny. We show that lentivirus-delivered shRNAs are capable of specific, highly stable and functional silencing of gene expression in a variety of cell types and also in transgenic mice. Our lentiviral vectors should permit rapid and efficient analysis of gene function in primary human and animal cells and tissues and generation of animals that show reduced expression of specific genes. They may also provide new approaches for gene therapy.
Article
This review focuses on how cells establish the levels of initiation factors, within the broader context of determining levels of the translational machinery. Most initiation factor polypeptides are moderately abundant proteins with concentrations approaching those of ribosomes. eIF4A and eIF5A are more abundant than ribosomes, whereas eIF4F alpha and eIF2B are considerably less abundant than the other factors. The cloning of cDNAs generates hybridization probes for monitoring the levels and activities of factor mRNAs, and the cloning of their genes is just beginning to provide insight into promoter structures and regulation. Initiation factor gene expression appears to be coordinately regulated in many cases, and preferential synthesis is seen in mitogen-activated T-cells. The gene for eIF2 alpha has been best characterized, and mechanisms that provide for the coordinated synthesis of eIF2 subunits are emerging. Recombinant DNA methods also allow investigators to manipulate the levels of expression of specific factor genes by overexpression or antisense repression. Such approaches provide a means to investigate in vivo the mechanisms of action of the initiation factors and their roles in regulating translation rates.
Article
To discover unique genes specific for squamous cell carcinoma of the head and neck for eventual development as tumor markers and vaccine candidates. Molecular biological analysis of fresh-frozen head and neck squamous cell cancer (HNSCC). A subtractive library was made from two HNSCC and six normal tissues using a polymerase chain reaction (PCR)-based approach. Genes from this library were PCR amplified and placed on a microarray glass slide. RNA was prepared or obtained from 16 fresh-frozen HNSCC and 22 normal tissue sources. Fluorescent probes were made from the polyA+ RNA derived from the tumor and normal tissues. The probes were hybridized to the glass slides and excited by a tuneable laser. One hundred seven of the genes showing the highest differential fluorescence value between tumor and normal tissue were identified by sequence analysis. Thirteen independent genes were found to be overexpressed in tumor tissues. Of these, nine were previously known: keratins K6 and K16, laminin-5, plakophilin-1, matrix metalloproteinase-2 (MMP), vascular endothelial growth factor, connexin 26, 14-3-3 sigma, and CaN19. The level of polyA+ RNA of these genes in the tumors was significantly different from the levels in normal tissue (P < .05). Four previously unidentified genes were also discovered to have increased expression in tumor tissue. Comparing the total tumor group (n = 16) to the normal group (n = 22), only one of these genes showed significant overexpression. We report the identification of nine known genes that are significantly overexpressed in HNSCC as compared to normal tissue using subtractive and microarray technology. In addition, we present four previously unidentified genes that are overexpressed in a subset of tumors. These genes will be developed as tumor markers and vaccine candidates.