Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

review

AP-1 as a regulator of cell life and


death
Eitan Shaulian*§ and Michael Karin†‡
*Department of Experimental Medicine and Cancer Research, School of Medicine, the Hebrew University, Jerusalem, 91120, Israel
†Laboratory of Gene Regulation and Signal Transduction, Department of Pharmacology, University of California San Diego, La Jolla, CA 92093, USA
e-mail: ‡karinoffice@ucsd.edu or §eshaulian@md.huji.ac.il

The transcription factor AP-1 (activator protein-1) is involved in cellular proliferation, transformation and death.
Using mice and cells lacking AP-1 components, the target-genes and molecular mechanisms mediating these
processes were recently identified. Interestingly, the growth-promoting activity of c-Jun is mediated by repression of
tumour suppressors, as well as upregulation of positive cell cycle regulators. Mostly, c-Jun is a positive regulator of
cell proliferation, whereas JunB has the converse effect. The intricate relationships between the different Jun pro-
teins, their activities and the mechanisms that mediate them will be discussed.

P-1 was one of the first mammalian transcription factors to

A be identified1, but its physiological functions are still being


unravelled. AP-1 activity is induced by a plethora of physio-
logical stimuli and environmental insults. In turn, AP-1 regulates a
c-Jun–c-Fos
JunB–c-Fos ?
p53
p21 p16

JunB–JunB ?
wide range of cellular processes, including cell proliferation, death, JunB–c-Jun ?
Cyclin D1 Proliferation
survival and differentiation. However, despite increasing knowl- Cdk
edge regarding the physiological functions of AP-1, the target-genes
G1 S
mediating these functions are not always obvious. This review is
focused on recent findings, which suggest how AP-1 proteins regu-
late cell proliferation and survival. Figure 1 A model describing the effects of AP-1 proteins on the major cell
The ability of a single transcription factor to control an eclectic cycle regulators. Putative JunB-containing dimers may account for the ability of
collection of biological processes stems primarily from its structur- AP-1 to both promote and inhibit cell cycle progression.
al and regulatory complexity. AP-1 is not a single protein, but a
menagerie of dimeric basic region-leucine zipper (bZIP) proteins
that belong to the Jun (c-Jun, JunB, JunD), Fos (c-Fos, FosB, Fra-1
and Fra2), Maf (c-Maf, MafB, MafA, MafG/F/K and Nrl) and ATF and viral infections, and a variety of physical and chemical stresses.
(ATF2, LRF1/ATF3, B-ATF, JDP1, JDP2) sub-families, which recog- These stimuli activate mitogen activated protein kinase (MAPK)
nize either 12-O-tetradecanoylphorbol-13-acetate (TPA) response cascades13 that enhance AP-1 activity through the phosphorylation
elements (5′-TGAG/CTCA-3′) or cAMP response elements (CRE, of distinct substrates. Serum and growth factors that potently induce
5′-TGACGTCA-3′)2. This review is focused mainly on the functions AP-1 do so by activating the extracellular-signal-regulated kinase
of Jun and Fos. c-Jun is the most potent transcriptional activator in (ERK) subgroup of MAPKs, whose members translocate to the
its group3, whose transcriptional activity is attenuated and some- nucleus to phosphorylate, and thereby potentiate, the transcription-
times antagonized by JunB4,5. The Fos proteins, which cannot al activity of ternary complex factors (TCFs) that bind to fos pro-
homodimerize, form stable heterodimers with Jun proteins and moters14. Furthermore, the ERKs directly phosphorylate Fra1 and 2
thereby enhance their DNA binding activity1. Whereas c-Fos and in response to serum stimulation, possibly enhancing their DNA
FosB contain transcriptional activation domains, Fra1, Fra2 and the binding in conjunction with c-Jun15. An additional contribution to
alternative splicing variants of FosB, FosB2 and DFosB, do not. AP-1 induction by serum growth factors comes from ERK5-
The regulation of AP-1 activity is complex, and we refer the read- induced phosphorylation of the transcription factor monocyte-
er to earlier and more comprehensive reviews6,7. Briefly, regulation specific enhancer binding factor 2c (MEF2C), whose activation
occurs through: first, changes in jun and fos gene transcription and increases c-Jun expression16.
mRNA turnover; second, effects on Jun and Fos protein turnover; The induction of AP-1 by pro-inflammatory cytokines and
third, post-translational modifications of Jun and Fos proteins that genotoxic stress is mostly mediated by the JNK and p38 MAPK cas-
modulate their transactivation potential; fourth, interactions with cades13. Once activated, the JNKs translocate to the nucleus17, where
other transcription factors that can either synergize or interfere they phosphorylate c-Jun and thereby enhance its transcriptional
with AP-1 activity2. In addition to being a transcriptional activator, activity6. The JNKs also phosphorylate and potentiate the activity
there is increasing evidence that some of the biological effects of of ATF2, which heterodimerizes with c-Jun to bind divergent AP-1
AP-1 are mediated by gene repression8–10. These effects may depend sites in the c-jun promoter18. Importantly, the induction of c-Jun
on the interactions of AP-1 proteins with transcriptional corepres- expression by certain genotoxic stresses, such as short-wavelength
sors11 and may also be affected by the nature of the AP-1 target ultraviolet (UV) radiation, is much more robust and persistent
site5,12. than the induction seen after mitogenic stimulation10,19.
Constitutive expression of activated oncogenes, such as Ha-ras, also
results in an elevation of AP-1 activity20, mostly through persistent
Signalling pathways that regulate AP-1 activity activation of ERK and JNK21,22. The contribution of p38 to AP-1
AP-1 activity is induced by growth factors, cytokines, neurotrans- induction can be mediated by the direct phosphorylation and acti-
mitters, polypeptide hormones, cell–matrix interactions, bacterial vation of ATF2, MEF2C and TCFs23.

NATURE CELL BIOLOGY VOL 4 MAY 2002 http://cellbio.nature.com E131

© 2002 Nature Publishing Group


review

Table 1 AP-1-regulated genes that significantly affect cell cycle progression and apoptosis.

Gene product Activity c-Jun JunB JunD


Cyclin D1(reviewed, ref. 24) Proliferation ↑ Up Down –
p53 (ref. 8) Proliferation ↓ Down – –
Apoptosis ↑
p21Cip1 (ref. 10) Proliferation ↓ Down – –
p16INK4 (ref. 9) Proliferation ↓ Down Up
Apoptosis ↑
p19ARF (ref. 43) Proliferation ↓ – – Down
Apoptosis ↑
GM-CSF (ref. 41) Proliferation ↑ Up Down
KGF (ref. 32) Proliferation ↑ Up Down –
HB-EGF (ref. 53) Proliferation ↑ Up – –
FL1 (ref. 57) Proliferation ↑ Up Up Up
FasL (ref. 24) Apoptosis ↑ Up – –
Fas (ref. 88) Apoptosis ↑ Down – –
Bc13 (ref. 98) Apoptosis ↓ Up – –

AP-1 in cell proliferation and transformation defects. These cells can be passed in culture only once or twice
Initial studies examining the function of AP-1 proteins in cell pro- before entering a premature senescence8,30, and even after immor-
liferation and oncogenic transformation, based on tissue culture talization, they continue to proliferate slower than wild type cells8.
models such as rodent and avian fibroblasts, suggested that c-Jun JNK-mediated phosphorylation of c-Jun partially stimulates cell
and the Fos proteins positively regulate cell proliferation, whereas proliferation, as c-junAla63/73 fibroblasts have a proliferation defect
JunB is thought to be a negative regulator. Recent studies, however, that is less severe than that of c-jun−/− fibroblasts29. c-Jun induction
support a more complicated scenario. Below, we describe some of is also required for cell cycle re-entry of UV-irradiated fibrob-
the experimental data implicating AP-1 proteins in growth control lasts10. Although wild type fibroblasts undergo a transient cell cycle
and transformation and discuss the possible mechanisms that arrest after exposure to UV, c-jun−/− cells undergo prolonged
mediate these effects. growth arrest and fail to resume proliferation. By contrast, cells
Inhibition of fos and jun expression in mouse fibroblasts and ery- that constitutively express c-Jun fail to arrest and continue to cycle
throleukaemia cells using antisense RNA demonstrated their after UV exposure10. Other cell types, such as foetal hepatoblasts
requirement for proliferation and cell cycle progression24. Similarly, derived from c-jun−/− foetuses, also display reduced proliferation31.
microinjection of antibodies against Fos and Jun prevents serum- More importantly, liver regeneration after partial hepatecotomy is
stimulated quiescent mouse fibroblasts from re-entering the cell impaired in the absence of c-Jun (Behrens et al., see note added in
cycle25. Interestingly, non-crossreactive antibodies against any Jun proof) Collectively, these results suggest that AP-1 activity is
protein, including JunB, inhibit cell cycle entry, whereas simultane- important for wound healing responses. Indeed, a recent study
ous inactivation of several Fos proteins is required to achieve a simi- demonstrated a reduced healing rate of skin injuries in mice whose
lar effect25, suggesting a unique and critical role for each Jun protein, epidermis was rendered deficient in c-Jun (R. Johnson, personal
but a redundant function for the Fos proteins. Studies using fibrob- communication). A c-Jun deficiency in mouse fibroblasts also
lasts derived from fos and jun knockout mice provide partial genetic results in a reduced ability to support keratinocyte proliferation in
support for these conclusions. Fibroblasts deficient in c-Fos or FosB culture32.
alone proliferate normally, and only cells lacking both proteins have In common with c-fos−/−fosB−/− double knockout fibroblasts, c-
a reduced proliferative capacity26,27. Interestingly, c-fos−/−fosB−/− dou- jun−/− MEFs also have reduced expression of cyclin D1 (ref. 33).
ble-knockout mice, but not single knockouts, are smaller than their However, this deficiency does not fully account for their prolifera-
wild type counterparts26. A similar correlation between reduced tive defect, as the defect is only partially suppressed by ectopic
fibroblast proliferation and smaller cell size was made for junD−/− expression of cyclin D1 in c-jun−/− MEFs33. By contrast, the defect is
mice28, or knockin mice that express a phosphorylation-defective completely suppressed through loss of the p53 tumour suppressor
mutant form of c-Jun, in which the major JNK phosphorylation sites protein8. Importantly, c-Jun functions as a direct repressor of p53
(Ser 63 and Ser 73) were replaced by alanine (c-junAla63/73 mice)29. gene transcription, and therefore p53 is elevated in c-jun−/− fibrob-
The availability of Fos-deficient mice allowed a rigorous exami- lasts through increased p53 gene transcription8. Furthermore, c-Jun
nation of the role of Fos proteins in cyclin D1 induction, thereby also decreases the transcriptional activity of p53 itself, down-mod-
providing a connection to the cell cycle machinery. AP-1 proteins ulating its ability to activate the p21Cip1 gene10. p21Cip1 is a CDK
were shown to bind directly to the cyclin D1 promoter and activate inhibitor (CDI) that is responsible for the antiproliferative activity
it in transient transfection experiments24. These results were con- of p53 (refs 34,35). Correspondingly, c-jun−/− fibroblasts have lower
firmed in vivo with the demonstration that the induction of cyclin levels of G1 CDK activity, not because of a cyclin D1 deficiency, but
D1 by serum is more efficient in wild type fibroblasts than in c-fos− because of elevated p21Cip1 expression8. Similarly, v-Jun activates the
/−
fosB−/− double mutants26. Thus, the induction of cyclin D1 pro- cyclin E–CDK complex in chicken embryo fibroblasts (CEF), with-
vides one route for AP-1 proteins to stimulate cell proliferation. out inducing expression of cyclin D or E36. Hence, c-Jun can stimu-
Among the different AP-1 deficiencies, mouse embryo fibroblasts late cell cycle progression through two mechanisms: induction of
(MEFs) from c-jun−/− embryos have the most severe proliferation cyclin D1 transcription and repression of p21Cip1 transcription.

E132 NATURE CELL BIOLOGY VOL 4 MAY 2002 http://cellbio.nature.com

© 2002 Nature Publishing Group


review

a even more effectively than wild type cells32. The effects of Jun on
FasL keratinocyte proliferation are probably mediated through modula-
tion of granulocyte-macrophage colony stimulating factor (GM-
Bim Apoptosis
JNK CSF) and keratinocyte growth factor (KGF) expression, with c-Jun
UV c-Jun
Bcl3
Σ Survival
as a positive regulator and JunB as a negative regulator of their
genes32. These results provide an example of a physiologically rele-
Fas vant antagonism between c-Jun and JunB.
However, new evidence supports the possibility that the differ-
ences in JunB and c-Jun function may be cell-type-specific or
depend on the relative expression levels of these proteins.
UV Replacement of c-jun by junB prevents most of the developmental
b
and cellular defects associated with a loss of c-Jun42. These results
Repaired Cell cycle suggest that in the absence of c-Jun, JunB may function as a posi-
c-Jun p53 damage re-entry tive growth regulator. However, in the presence of c-Jun, JunB has
Sustained opposite effects. It is possible that the anti-proliferative activity of
p21 Apoptosis
damage JunB requires the formation of growth inhibitory c-Jun–JunB het-
erodimers (Fig. 1).
The effects of JunD on cellular proliferation are even harder to
G1 S G2 M interpret. Certain findings suggest that JunD is a positive regulator
of cell proliferation: junD−/− MEFs have a proliferation defect which
results in premature senescence that is less severe than that
Figure 2. The effects of c-Jun in apoptosis. a, A model describing the direct
observed in c-jun−/− MEFs43. This defect correlates with increased
effects of c-Jun on apoptosis through transcriptional regulation of pro- and anti-
expression of p19ARF, a tumour suppressor gene that enhances p53
apoptotic gene products. c-Jun-mediated induction of pro-apoptotic genes (sur-
activity by sequestering Mdm2 (ref. 44). Immortalized junD−/− cells,
rounded by the oval) enhances apoptosis, whereas upregulation of anti-apoptotic
however, display a different phenotype, which includes increased
gene products and down-regulation of pro-apoptotic gene products (surrounded by
proliferation and higher levels of cyclin D143. As this phenotype is
the square) support cell survival. The sum of all signals (Σ) will determine whether
reminiscent of fibroblasts in which c-Jun is constitutively
the cells undergo apoptosis or survive. b, A model describing the indirect effects of
expressed8, c-Jun might be overexpressed in immortalized junD−/−
c-Jun on apoptosis. c-Jun activation by UV results in downregulation of p21Cip1
cells to compensate for the loss of JunD. The tumour suppressor
expression to allow growth-arrested cells to proliferate. However, the actual fate of
Menin interacts with JunD and inhibits its transcriptional activity,
cells exposed to UV will be determined by the extent of the UV-induced cell damage.
whereas tumour-derived mutants of Menin lack this activity45.
Thus, inhibition of JunD might be involved, in some cells, in the
suppression of neoplastic growth. However, other studies have
Expression of inducible dominant negative c-Jun in HT1080 demonstrated that JunD may partially suppress Ha-Ras-induced
fibrosarcoma cells demonstrated that the second mechanism is transformation and that its expression decreases in Ha-Ras-trans-
more responsive to c-Jun. Elevated p21Cip1 expression and reduced formed fibroblasts46. In conclusion, the role of JunD in the control
cell proliferation were obtained by dominant negative levels, which of cell proliferation is somewhat nebulous. It is safe to conclude,
were insufficient to induce an increase in cyclin D1 expression37. however, that JunD does not have as potent pro-proliferative func-
The antibody microinjection experiments described above may tion as c-Jun.
lead one to conclude that all Jun proteins positively regulate cell In light of their ability to regulate cell cycle progression, it is not
proliferation. However, such a conclusion is inconsistent with earli- surprising that Jun and Fos are also involved in transformation.
er results demonstrating that JunB antagonizes the transforming Both, c-fos and c-jun are cellular homologues of retroviral onco-
activity of c-Jun4,38. Recent experiments conducted in vivo support genes47,48. Furthermore, AP-1 activity is induced by various tumour
these earlier studies and suggest that JunB is a negative regulator of promoters and activated oncoproteins1. Certain AP-1 proteins can
proliferation, at least in certain cell types and in the presence of c- co-operate with activated oncogenes, such as Ha-Ras, in the trans-
Jun expression. Unlike the slower proliferation of MEFs deficient in formation of rodent fibroblasts4,49. Interestingly, all Jun proteins,
c-Jun or JunD, junB−/− MEFs proliferate normally in culture39. including JunB, can co-operate with Ha-Ras in rat embryo fibrob-
However, MEFs derived from junB transgenic mice that overexpress lasts (REF) transformation, although c-Jun is certainly the most
junB had limited proliferative capacity40. Furthermore, after immor- potent among them49. In addition, c-Jun is required for Ha-Ras-
talization they continued to proliferate slower than wild type fibrob- induced transformation, and c-jun−/− cells are refractory to the
lasts, caused by an extended G1 transition time9. Consistent with its transforming activity of oncogenic Ras50. This oncogenic coopera-
ability to antagonize c-Jun, JunB represses the cyclin D1 promoter40. tion may be partially dependent on the ability of c-Jun to be phos-
In addition, JunB induces transcription of the p16INK4a gene, coding phorylated by JNK51.
for another CDI that inhibits Rb phosphorylation by CDKs and Several genes involved in the transforming activity of Jun pro-
thereby prevents G1 to S phase transition9. p16INK4a is highly teins in CEF were recently identified. JAC (Jun-activated gene in
expressed in JunB transgenic fibroblasts and its deletion negates the CEF) and heparin-binding epidermal growth factor-like growth
inhibitory effect of JunB9. Further support for the anti-proliferative factor (HB-EGF) are c-Jun and v-Jun target-genes that are capable
function of JunB was also obtained by introducing a ubiquitously of transforming CEFs52,53. However, their transforming activity is
expressed JunB transgene into a junB−/− background. The JunB lower than that of c-Jun52,53 and they can not induce tumorigenici-
transgene suppressed the embryonic lethality associated with loss of ty when expressed individually, suggesting that several c-Jun target-
JunB, but serendipitous extinction of the JunB transgene in myeloid genes co-operate to exert its full transforming activity. The identity
cells, which rendered them JunB deficient41, prevented the induction or the transforming activity of the mammalian homologues of JAC
of p16Ink4a expression associated with terminal differentiation of and HB-EGF have yet to be determined, in order to establish their
granulocytes, resulting in the perpetual proliferation and eventual possible role in mammalian cell transformation by c-Jun. HB-EGF
transformation of granulocytic progenitors41. Some of the negative is known to exert mitogenic activity on keratinocytes, hepatocytes,
proliferative effects of JunB are not cell-autonomous. Unlike smooth muscle cells and fibroblasts.
c-Jun-deficient fibroblasts, which poorly support keratinocyte pro- Consistent with its effects on DNA binding, c-Fos enhances the
liferation, junB−/− fibroblasts support keratinocyte proliferation transforming ability of c-Jun4,5 and overexpression of either protein

NATURE CELL BIOLOGY VOL 4 MAY 2002 http://cellbio.nature.com E133

© 2002 Nature Publishing Group


review

transforms immortalized rodent fibroblasts in culture38,54. Fra1 is Such observations, however, do not demonstrate whether Jun or
induced in response to expression of oncogenic Ha-ras, and togeth- Fos induction is functionally involved in triggering apoptosis.
er with c-Jun, generates stable AP-1 heterodimers that contribute to Some direct evidence for pro-apoptotic functions was derived from
Ras-mediated transformation55. c-Fos-induced transformation is transient overexpression experiments, in which c-Jun or c-Fos were
mediated, at least in part, by induction of the enzyme DNA methyl found to induce apoptosis in various cell lines69,72,73. However, as the
transferase 1 (DNMT1), which may downregulate the expression of expression levels achieved in such experiments are unusually high,
negative growth regulators through methylation of their promoter and could therefore result in non-specific effects, such results may
region56. As a c-Jun antagonist, cotransfected JunB attenuates the have no bearing on the physiological function of the overexpressed
ability of c-Jun to participate in REF transformation38. protein. A more relevant way to study the function of AP-1 in apop-
Furthermore, overexpression of JunB in mouse fibroblasts reduces tosis is to examine the sensitivity of animals and cells lacking AP-1
transformation by v-ras or v-src9. expression or activity to different pro-apoptotic stimuli and geno-
The fact that the c-Jun/JunB antagonism is probably context- toxins. Antisense oligonucleotides directed against c-fos and c-jun
dependent42 is also reflected by the fact that both c-Jun and JunB mRNAs were found to increase the survival of growth-factor-
increase angiogenesis in a cell culture model of fibrosarcoma, deprived lymphoid cells70. Similarly, inhibition of c-Jun activity
through activation of proliferin (FL1) gene expression. FL1, an through the expression of a dominant negative c-Jun mutant or
angiogenic factor that is primarily produced by trophoblast giant injection of neutralizing antibodies, can protect neuronal cells from
cells, is required for the migration of uterine vascular endothelial apoptosis induced by withdrawal of nerve growth factor (NGF) or
cells toward trophoblasts. Interestingly, FL1 expression correlates chronic depolarization68,69,74–76. In addition, light-induced apoptosis
with the stage of fibrosarcoma development, suggesting a role for c- of retinal photoreceptors is impaired in c-Fos-deficient mice77. The
Jun and JunB in tumour progression57. anti-apoptotic activity of dominant negative c-Jun depends, at least
The importance of the dimerization partner of c-Jun for its partially, on its ability to induce the expression of Bim78, a pro-
transformation-related activities was also addressed. It seems that apoptotic Bcl-2 family member that is critical for neuronal apopto-
c-Jun–ATF2 heterodimers induce autocrine growth, whereas c- sis. It is not clear, however, whether Bim transcription is directly
Jun–c-Fos heterodimers induce anchorage-independent growth58. regulated by c-Jun. A more established AP-1 target is the gene
However, the differentially regulated target-genes responsible for encoding Fas-ligand (FasL), whose product can promote apopto-
these activities remain to be identified. sis24. c-Jun-dependent FasL induction was demonstrated in several
Transgenic mice expressing c-Fos develop bone lesions as early experimental systems. However, it should be noted that the regula-
as 4 weeks after birth, and after 9–10 months these eventually tion of FasL transcription is complex and may only be promoted by
progress into bone tumours in 15% of animals59,60. Deletion of c-fos c-Jun under certain conditions.
from mice expressing a v-ras transgene inhibits the development of JNK-mediated phosphorylation is important for c-Jun-induced
fully malignant skin tumours in response to treatment with phor- apoptosis in neuronal cells. Expression of c-Jun that has been
bol ester tumour promoters61. Similar inhibition of malignant mutated in the JNK phospho-acceptor sites, c-JunAla63/73, can block
transformation is observed in mice expressing a dominant negative apoptosis induced by NGF withdrawal74. Most importantly, c-
c-jun transgene62. Thus, c-Fos and c-Jun are required for two-stage junAla63/73 mice are resistant to kainic acid-induced cytotoxicity29.
skin carcinogenesis. However, transgenic overexpression of c-Jun in Furthermore, targeting of JNK3, a JNK isoform specifically
the absence of additional oncogenic stimulation does not increase expressed in neuronal cells, reduced both AP-1 activation and
tumour incidence63. kainate-induced apoptosis79. In addition, inhibition of JNK activity
Skin carcinogenesis may require JNK-mediated phosphoryla- by the products of two NF-κB target-genes, Gadd45β and X-chro-
tion of c-Jun, because the incidence of skin tumours induced by a mosome-linked inhibitor of apoptosis (XIAP), provides partial
transgene coding for an activated form of the Ras exchange factor protection from tumour necrosis factor (TNF)-α -induced apopto-
SOS, is reduced in a c-junAla63/73 background51. A recent study sug- sis80,81. However, it should be mentioned that there are many phys-
gested that Ras-induced c-Jun phosphorylation synergizes with the iological and pathological conditions that result in JNK activation
prolyl isomerase Pin1 to augment the induction of cyclin D1 by c- and c-Jun N-terminal phosphorylation in neuronal cells, but do
Jun in breast cancer64. In conclusion, a persistent alteration in AP-1 not cause apoptosis82–84.
activity can result in enhanced oncogenic transformation through The exact function of AP-1 in cellular responses to genotoxic
deregulated expression of positive and negative growth regulators, stress has not been entirely resolved. Several studies suggest that c-
such as cyclin D1, and a number of autocrine growth factors. Jun is involved in the induction of UV-induced apoptosis. c-Jun-
deficient fibroblasts and jnk1−/−jnk2−/− double knockout MEFs are less
sensitive to UV-induced apoptosis10,85. Similarly, c-jun−/− fibroblasts
AP-1 proteins as regulators of cell death and survival are resistant to the apoptotic effects of alkylating agents, which may
Apoptosis suppresses oncogenic transformation, and controls be mediated by the induction of FasL86. In addition, dominant nega-
organogenesis and immune responses. AP-1 transcription factors tive c-Jun reduces apoptosis in human monoblastic leukaemia cells
are involved in both the induction and prevention of apoptosis, exposed to a variety of genotoxins87. However, other studies suggest
and the exact outcome is highly tissue- and developmental- that c-Jun protects cells against UV-induced cell death33,88. This pro-
stage-specific. The first indication for such an apoptotic function tective activity may be mediated through co-operation with STAT3
for AP-1 came from observations linking the induction of c-Fos to suppress transcription of Fas88. However, the cells used in this
and c-Jun to conditions that result in cell death. c-Fos, for study are likely to be p53-deficient, as they are derived from tumours.
instance, is persistently induced in the brains of mice treated with Because p53 is a regulator of Fas expression89, and is required for c-
kainic acid, a potent activator of glutamate receptors that induces Jun-induced apoptosis after exposure to UV10, it is expected to have
apoptosis of hippocampal neurons65. c-Fos induction is also an important effect on the link between c-Jun, Fas and apoptosis.
observed during tissue remodelling, such as mammary gland invo- The involvement of p53 in UV-induced apoptosis of cells expressing
lution66 or castration-induced prostate gland regression67. The normal levels of c-Jun is nebulous. Although known as a potent pro-
induction of apoptosis in neuronal and lymphoid cell cultures apoptotic factor, p53 elicits protective effects on mouse fibroblasts
through withdrawal of growth factors is preceded by the induction exposed to UV, possibly by participating in nucleotide excision repair
of AP-1 proteins68–70. Another setting in which robust c-Jun induc- (NER) and arresting cell proliferation until DNA is sufficiently
tion precedes the onset of apoptosis is the exposure of cells to repaired90,91. The relative resistance of senescent keratinocytes to UV-
genotoxic stress, such as alkylating agents (for example, MMS) or induced apoptosis supports the notion that replication is required
short-wavelength UV radiation10,19,71. for the induction of apoptosis in UV-exposed cells92.

E134 NATURE CELL BIOLOGY VOL 4 MAY 2002 http://cellbio.nature.com

© 2002 Nature Publishing Group


review
2. Chimenov, Y. & Kerppola, T. K. Close encounters of many kinds: Fos–Jun interactions that mediate
A clearer example of anti-apoptotic activity of c-Jun is provided transcription regulatory specificity. Oncogene 6, 533–542 (2001).
by hepatocytes derived from c-jun−/− embryos, which undergo mas- 3. Ryseck, R. P. & Bravo, R. cJun, JunB, and JunD differ in their binding affinities to the AP-1 and CRE
sive apoptosis31,93. Surprisingly, a junB transgene can prevent liver consensus sequences: effect of Fos proteins. Oncogene 6, 533–542 (1991).
degeneration in c-jun−/− embryos42. The brains of jnk1−/−jnk2−/− dou- 4. Schutte, J. et al. JunB inhibits and cFos stimulates the transforming and transactivating activities of
cJun. Cell 59, 987–997 (1989).
ble knockout embryos have an abnormal morphology caused by 5. Chiu, R., Angel, P. & Karin, M. JunB differs in its biological properties from, and is a negative regu-
disregulated apoptosis, which is decreased in some areas of the lator of c-Jun. Cell 59, 979–986 (1989).
brain and increased in others94. Increased spontaneous apoptosis is 6. Karin, M. The regulation of AP-1 activity by mitogen-activated protein kinases. J. Biol. Chem. 270,
exhibited by the same c-jun−/− fibroblasts that are resistant to UV- 16483–16486 (1995).
7. Karin, M., Liu, Z.-G. & Zandi, E. AP-1 function and regulation. Curr. Opin. Cell Biol. 9, 240–246
induced apoptosis10. Furthermore, re-introduction of c-Jun reduces (1997).
this spontaneous apoptosis but simultaneously enhances UV- 8. Schreiber, M. et al. Control of cell cycle progression by c-Jun is p53 dependent. Genes Dev. 13,
induced apoptosis10. Some in vitro studies suggests that JNK is not 607–619 (1999).
required for TNF-induced apoptosis95 and even inhibits both TNF- 9. Passegue, E. & Wagner, E. F. JunB suppresses cell proliferation by transcriptional activation of
96 p16(INK4a) expression. EMBO J. 19, 2969–2979 (2000).
and Fas-induced97 apoptosis. Some of the protective effects of AP- 10. Shaulian, E. et al. The mammalian UV response: c-Jun induction is required for exit from p53-
1 might be mediated through induction of Bcl-3, whose product imposed growth arrest. Cell 103, 897–907 (2000).
functions as a survival factor for growth-factor-deprived T cells98. 11. Pessah, M. et al. c-Jun interacts with the corepressor TG-interacting factor (TGIF) to suppress
Two models can explain the diverse effects of AP-1 on cell death Smad2 transcriptional activity. Proc. Natl Acad. Sci. USA 98, 6198–6120 (2001).
12. Hsu, J. C., Cressman, D. E. & Taub, R. Promoter-specific trans-activation and inhibition mediated
and survival. The first model (Fig. 2a) suggests that the induction of
by JunB. Cancer Res. 53, 3789–3794 (1993).
AP-1 results in activation of various genes, such as FasL, Bim, or Bcl3, 13. Chang, L. & Karin, M. Mammalian MAP kinase signalling cascades. Nature 410, 37–40 (2001).
whose products are either positive or negative regulators of apoptosis. 14. Hill, C. S., Wynne, J. & Treisman, R. Serum-regulated transcription by serum response factor (Srf) -
It is the balance between the pro-apoptotic and anti-apoptotic target- a novel role for the DNA binding domain. EMBO J. 13, 5421–5432 (1994).
genes that determines whether the final outcome will be cell survival 15. Gruda, M. C., Kovary, K., Metz, R. & Bravo, R. Regulation of Fra-1 and Fra-2 phosphorylation dif-
fers during the cell cycle of fibroblasts and phosphorylation in vitro by MAP kinase affects DNA
or cell death. This balance may vary from one cell type to another, and binding activity. Oncogene 9, 2537–2547 (1994).
may be dependent on the type and duration of stimulus used to acti- 16. Kato, Y. et al. BMK1/ERK5 regulates serum-induced early gene expression through transcription
vate AP-1, as well as on the activation of other transcription factors. factor MEF2C. EMBO J. 16, 7054–7066 (1997).
According to the second model (Fig. 2b), AP-1 functions as a homeo- 17. Cavigelli, M., Dolfi, F., Claret, F. X. & Karin, M. Induction of c-fos expression through JNK-mediat-
ed TCF/Elk-1 phosphorylation. EMBO J. 14, 5957–5964 (1995).
static regulator that keeps cells in a certain proliferative steady state. 18. Gupta, S., Campbell, D., Dérijard, B. & Davis, R. J. Transcription factor ATF2: regulation by the
Changes in environmental conditions may enhance AP-1 activity. JNK signal transduction pathway. Science 267, 389–393 (1995).
Robust or persistent activation of AP-1 in cells containing damaged 19. Devary, Y., Gottlieb, R. A., Lau, L. & Karin, M. Rapid and preferential activation of the c-jun gene
DNA causes defective replication and may trigger apoptosis through during the mammalian UV response. Mol. Cell. Biol. 11, 2804–2811 (1991).
20. Schonthal, A., Herrlich, P., Rahmsdorf, H. J. & Ponta, H. Requirement for fos gene expression in the
the same mechanisms that induce cell death after constitutive expres-
transcriptional activation of collagenase by other oncogenes and phorbol esters. Cell 54, 325–334
sion of oncogenes. However, the activation of AP-1 in cells that are (1988).
able to proliferate promotes cell proliferation and survival. 21. Hibi, M., Lin, A., Smeal, T., Minden, A. & Karin, M. Identification of an oncoprotein-responsive
and UV-responsive protein kinase that binds and potentiates the c-Jun activation domain. Genes
Dev. 7, 2135–2148 (1993).
22. Dérijard, B. et al. JNK1: A protein kinase stimulated by UV light and Ha-Ras that binds and phos-
Future prospects phorylates the c-Jun activation domain. Cell 76, 1025–1037 (1994).
In the last few years, the progress in our understanding of the phys- 23. Han, J., Jiang, Y., Li, Z., Kravchenko, V. V. & Ulevitch, R. J. Activation of the transcription factor
iological and pathological activities of AP-1 transcription factors MEF2C by the MAP kinase p38 in inflammation. Nature 386, 296–299 (1997).
was highly dependent on the use of genetically altered mouse 24. Shaulian, E. & Karin, M. AP-1 in cell proliferation and survival. Oncogene 20, 2390–2400 (2001).
25. Kovary, K. & Bravo, R. The jun and fos protein families are both required for cell cycle progression
strains and cell cultures derived from them. The results obtained in fibroblasts. Mol. Cell. Biol. 11, 4466–4472 (1991).
from these experiments have shed light on a few key questions 26. Brown, J. R. et al. Fos family members induce cell cycle entry by activating cyclin D1. Mol. Cell Biol.
regarding AP-1 activity and provide some explanations of its role in 18, 5609–5619 (1998).
cell proliferation. However, other major questions regarding the 27. Brusselbach, S. et al. Cell proliferation and cell cycle progression are not impaired in fibroblasts and
ES cells lacking c-Fos. Oncogene 10, 79–86 (1995).
activity of individual AP-1 proteins and their specificity remain
28. Thepot, D. et al. Targeted disruption of the murine junD gene results in multiple defects in male
unanswered. For example, the ability of c-Jun and JunB to elicit reproductive function. Development 127, 143–153 (2000).
opposite transcriptional responses in the presence of apparently 29. Behrens, A., Sibilia, M. & Wagner, E. F. Amino-terminal phosphorylation of c-Jun regulates stress-
similar AP-1 recognition sites found in the control regions of dif- induced apoptosis and cellular proliferation. Nature Genet. 21, 326–329 (1999).
ferent genes remains enigmatic, as is the ability of JunD to posi- 30. Johnson, R. S., van Lingen, B., Papaioannou, V. E. & Spiegelman, B. W. A null mutation at the c-jun
locus causes embryonic lethality and retarded cell growth in culture. Genes Dev. 7, 1309–1317 (1993).
tively regulate the proliferation of MEFs while negatively regulating 31. Eferl, R. et al. Functions of c-jun in liver and heart development. J. Cell Biol. 145, 1049–1061
proliferation in immortalized fibroblasts. Furthermore, the molec- (1999).
ular mechanisms that underlie the ability of c-Jun and other AP-1 32. Szabowski, A. et al. c-Jun and JunB antagonistically control cytokine-regulated mesenchymal–epi-
proteins to repress gene transcription have yet to be identified. dermal interaction in skin. Cell 103, 745–755 (2000).
33. Wisdom, R., Johnson, R. S. & Moore, C. c-Jun regulates cell cycle progression and apoptosis by dis-
Most importantly, an understanding of the network of dynamic tinct mechanisms. EMBO J.18, 188–197 (1999).
interactions between AP-1 proteins and other transcription factors 34. Xiong, Y. et al. p21 is a universal inhibitor of cyclin kinases. Nature 366, 701–704 (1993).
is far from complete. It is expected, however, that the generation of 35. El-Deiry, W. S. et al. WAF1, a potential mediator of p53 tumor suppression. Cell 75, 817–825
additional mouse mutations, in which the coding regions of one (1993).
36. Clark, W. et al. v-Jun overrides the mitogen dependence of S-phase entry by deregulating
AP-1 protein are replaced with those of another, together with new
retinoblastoma protein phosphorylation and E2F-pocket protein interactions as a consequence of
technological advances that should facilitate the identification of a enhanced cyclin E–cdk2 catalytic activity. Mol. Cell Biol. 20, 2529–2542 (2000).
large number of AP-1 target-genes, should eventually provide us 37. Hennigan, R. F. & Stambrook, P. J. Dominant negative c-jun inhibits activation of the cyclin D1
with answers to many of these questions. Such an analysis should and cyclin E kinase complexes. Mol. Biol. Cell 12, 2352–2363 (2001).
also be facilitated by the availability of specific pharmacological 38. Schutte, J., Minna, J. D. & Birrer, M. Deregulated expression of human {Ic-jun} transforms primary
rat embryo cells in cooperation with and activated c-Ha-{Iras} gene and transforms Rat-la cells as a
inhibitors of the many signalling pathways that modulate the single gene. Proc. Natl Acad. Sci. USA 86, 2257–2261 (1989).
expression and activity of AP-1 proteins. 39. Schorpp-Kistner, M., Wang, Z. Q., Angel, P. & Wagner, E. F. JunB is essential for mammalian pla-
Note added in proof: the article by Behrens et al. has subsequently centation. EMBO J. 18, 934–948 (1999).
been published (Behrens, A. et al. Impaired postnatal hepatocyte pro- 40. Bakiri, L., Lallemand, D., Bossy-Wetzel, E. & Yaniv, M. Cell cycle-dependent variations in c-Jun and
JunB phosphorylation: a role in the control of cyclin D1 expression. EMBO J. 19, 2056–2068
liferation and liver regeneration in mice lacking c-Jun in the liver. (2000).
EMBO J. 21, 1782–1790 (2002). 41. Passegue, E., Jochum, W., Schorpp-Kistner, M., Mohle-Steinlein, U. & Wagner, E. F. Chronic
1. Angel, P. & Karin, M. The role of Jun, Fos and the AP-1 complex in cell-proliferation and transfor- myeloid leukemia with increased granulocyte progenitors in mice lacking junB expression in the
mation. Biochem. Biophys. Acta 1072, 129–157 (1991). myeloid lineage. Cell 104, 21–32 (2001).

NATURE CELL BIOLOGY VOL 4 MAY 2002 http://cellbio.nature.com E135

© 2002 Nature Publishing Group


review
42. Passegue, E., Jochum, W., Behrens, A., Ricci, R. & Wagner, E. JunB can substitute for c-Jun in mouse cell lines. J. Biol. Chem. 267, 18278–18283 (1992).
development and cell proliferation. Nature Genet. 30, 158–166 (2002). 71. Karin, M. Mitogen-activated protein kinase cascades as regulators of stress responses. Ann. NY
43. Weitzman, J. B., Fiette, L., Matsuo, K. & Yaniv, M. JunD protects cells from p53-dependent senes- Acad. Sci. 851, 139–146 (1998).
cence and apoptosis. Mol. Cell 6, 1109–1119 (2000). 72. Bossy-Wetzel, E., Bakiri, L. & Yaniv, M. Induction of apoptosis by the transcription factor c-Jun.
44. Sherr, C. J. Tumor surveillance via the ARF-p53 pathway. Genes Dev. 12, 2984–2991 (1998). EMBO J. 16, 1695–1709 (1997).
45. Agarwal, S. K. et al. Menin interacts with the AP1 transcription factor JunD and represses JunD- 73. Preston, G. A. et al. Induction of apoptosis by c-Fos protein. Mol. Cell Biol. 16, 211–218 (1996).
activated transcription. Cell 96, 143–152 (1999). 74. Le-Niculescu, H. et al. Withdrawal of survival factors results in activation of the JNK pathway in
46. Pfarr, C. M. et al. Mouse JunD negatively regulates fibroblast growth and antagonizes transforma- neuronal cells leading to Fas ligand induction and cell death. Mol. Cell. Biol. 19, 751–763 (1999).
tion by ras. Cell 76, 747–760 (1994). 75. Xia, Z., Dickens, M., Raingeaud, J., Davis, R. J. & Greenberg, M. E. Opposing effects of ERK and
47. Maki, Y., Bos, T. J., Davis, C., Starbuck, M. & Vogt, P. K. Avian sarcoma cirus 17 carries the Jun JNK-p38 MAP kinases on apoptosis. Science 270, 1326–1331 (1995).
oncogene. Proc. Natl Acad. Sci. USA 84, 2848–2852 (1987). 76. J., W., Neame, S. J., Paquet, L., Bernard, O. & Ham, J. Dominant-negative c-Jun promotes neuronal
48. van Straaten, F., Muller, R., Curran, T., Van Beveren, C. & Verma, I. M. Complete nucleotide survival by reducing BIM expression and inhibiting mitochondrial cytochrome c release. Neuron
sequence of a human c-onc gene: deduced amino acid sequence of the human c-fos protein. Proc. 29, 629–643 (2001).
Natl Acad. Sci. USA 80, 3183–3187 (1983). 77. Hafezi, F. et al. The absence of c-fos prevents light-induced apoptotic cell death of photoreceptors
49. Vandel, L. et al. Stepwise transformation of rat embryo fibroblasts: c-Jun, JunB, or JunD can coop- in retinal degeneration in vivo. Nature Med. 3, 346–349 (1997).
erate with Ras for focus formation, but a c-Jun-containing heterodimer is required for immortal- 78. Whitfield, J., Neame, S. J., Paquet, L., Bernard, O. & Ham, J. Dominant-negative c-Jun promotes
ization. Mol. Cell. Biol. 16, 1881–1888 (1996). neuronal survival by reducing BIM expression and inhibiting mitochondrial cytochrome c release.
50. Johnson, R., Spiegelman, B., Hanahan, D. & Wisdom, R. Cellular transformation and malignancy Neuron 29, 629–643 (2001).
induced by ras require c-jun. Mol. Cell. Biol. 16, 4504–4511 (1996). 79. Yang, D. et al. Absence of excitotoxicity-induced apoptosis in the hippocampus of mice lacking the
51. Behrens, A., Jochum, W., Sibilia, M. & Wagner, E. F. Oncogenic transformation by ras and fos is Jnk3 gene. Nature 389, 865–870 (1997).
mediated by c-Jun N-terminal phosphorylation. Oncogene 19, 2657–2663 (2000). 80. De Smaele, E. et al. Induction of gadd45β by NF-κB downregulates pro-apoptotic JNK signalling.
52. Hartl, M., Reiter, F., Bader, A. G., Castellazzi, M. & Bister, K. JAC, a direct target of oncogenic tran- Nature 414, 308–313 (2001).
scription factor Jun, is involved in cell transformation and tumorigenesis. Proc. Natl Acad. Sci. USA 81. Tang, G. et al. Inhibition of JNK activation through NF-κB target genes. Nature 414, 313–317
98, 13601–13606 (2001). (2001).
53. Fu, S., Bottoli, I., Goller, M. & Vogt, P. K. Heparin-binding epidermal growth factor-like growth factor, a 82. Xu, X., Raber, J., Yang, D., Su, B. & Mucke, L. Dynamic regulation of c-Jun N-terminal kinase activi-
v-Jun target gene, induces oncogenic transformation. Proc. Natl Acad. Sci. USA 96, 5716–5721 (1999). ty in mouse brain by environmental stimuli. Proc. Natl Acad. Sci. USA 94, 12655–12660 (1997).
54. Miller, A., Curran, T. & Verma, I. cFos can induce cellular transformation: novel mechanism of acti- 83. Kenney, A. M. & Kocsis, J. D. Peripheral axotomy induces long-term c-Jun amino-terminal kinase-1
vating a cellular oncogene. Cell 36, 51–60 (1984). activation and activator protein-1 binding activity by c-Jun and junD in adult rat dorsal root gan-
55. Mechta, F., Lallemand, D., Pfarr, C. M. & Yaniv, M. Transformation by ras modifies AP1 composi- glia in vivo. J. Neurosci. 18, 1318–1328 (1998).
tion and activity. Oncogene 14, 837–847 (1997). 84. Herdegen, T. et al. Lasting N-terminal phosphorylation of c-Jun and activation of c-Jun N-terminal
56. Bakin, A. V. & Curran, T. Role of DNA 5-methylcytosine transferase in cell transformation by fos. kinases after neuronal injury. J. Neurosci. 18, 5124–5135 (1998).
Science 283, 387–390 (1999). 85. Tournier, C. et al. Requirement of JNK for stress-induced activation of the cytochrome c-mediated
57. Toft, D. J., Rosenberg, S. B., Bergers, G., Volpert, O. & Linzer, D. I. Reactivation of proliferin gene death pathway. Science 288, 870–874 (2000).
expression is associated with increased angiogenesis in a cell culture model of fibrosarcoma tumor 86. Kolbus, A. et al. c-Jun-dependent CD95-L expression is a rate-limiting step in the induction of
progression. Proc. Natl Acad. Sci. USA 98, 13055–13059 (2001). apoptosis by alkylating agents. Mol. Cell. Biol. 20, 575–582 (2000).
58. van Dam, H. et al. Autocrine growth and anchorage independence: two complementing Jun-con- 87. Verheij, M. et al. Requirement for ceramide-initiated SAPK/JNK signalling in stress-induced apop-
trolled genetic programs of cellular transformation. Genes Dev. 12, 1227–1239 (1998). tosis. Nature 380, 75–79 (1996).
59. Ruther, U., Garber, C., Komitowski, D., Muller, R. & Wagner, E. F. Deregulated c-fos expression 88. Ivanov, V. N. et al. Cooperation between STAT3 and c-jun suppresses Fas transcription. Mol. Cell 7,
interferes with normal bone development in transgenic mice. Nature 325, 412–416 (1987). 517–528 (2001).
60. Ruther, U., Komitowski, D., Schubert, F. R. & Wagner, E. F. c-fos expression induces bone tumors in 89. Owen-Schaub, L. B. et al. Wild-type human p53 and a temperature-sensitive mutant induce
transgenic mice. Oncogene 4, 861–865 (1989). Fas/APO-1 expression. Mol. Cell. Biol. 15, 3032–3049 (1995).
61. Saez, E. et al. c-fos is required for malignant progression of skin tumors. Cell 82, 721–732 (1999). 90. Smith, M. L. & Fornace, A. J. Jr. p53-mediated protective responses to UV irradiation. Proc. Natl
62. Young, M. R. et al. Transgenic mice demonstrate AP-1 (activator protein-1) transactivation is Acad. Sci. USA 94, 12255–12257 (1997).
required for tumor promotion. Proc. Natl Acad. Sci. USA 96, 9827–9832 (1999). 91. Bissonnette, N. & Hunting, D. J. p21-induced cycle arrest in G(1) protects cells from apoptosis
63. Jochum, W., Passegue, E. & Wagner, E. F. AP-1 in mouse development and tumorigenesis. Oncogene induced by UV-irradiation or RNA polymerase II blockage. Oncogene 16, 3461–3469 (1998).
20, 2401–2412 (2001). 92. Gniadecki, R., Hansen, M. & Wulf, H. C. Resistance of senescent keratinocytes to UV-induced
64. Wulf, G. M. et al. EMBO J. 20, 3459–3472 (2001). apoptosis. Cell. Mol. Biol. 46, 121–127 (2000).
65. Smeyne, R. Continuous c-Fos expression procedes programmed cell death in vivo. Nature 363, 93. Hilberg, F., Aguzzi, A., Howells, N. & Wagner, E. F. c-jun is essential for normal mouse development
166–169 (1993). and hepatogenesis. Nature 365, 179–181 (1993).
66. Marti, A. et al. Protein kinase A and AP-1 (c-Fos/JunD) are induced during apoptosis of mouse 94. Kuan, C. Y. et al. The Jnk1 and Jnk2 protein kinases are required for regional specific apoptosis
mammary epithelial cells. Oncogene 9, 1213–1223 (1994). during early brain development. Neuron 22, 667–676 (1999).
67. Buttyan, R., Zakeri, Z., Lockshin, R. & Wolgemuth, D. Cascade induction of c-fos, c-myc, and heat 95. Liu, Z.-G., Hu, H., Goeddel, D. V. & Karin, M. Dissection of TNF receptor 1 effector functions: JNK
shock 70K transcripts during regression of the rat ventral prostate gland. Mol. Endocrinol. 2, activation is not linked to apoptosis, while NF-κB activation prevents cell death. Cell 87, 565–576
650–657 (1988). (1996).
68. Estus, S., et al. Altered gene expression in neurons during programmed cell death: identification of 96. Natoli, G. et al. Activation of SAPK/JNK by TNF receptor 1 through a noncytotoxic TRAF2-
c-jun as necessary for neuronal apoptosis. J. Cell Biol. 127, 1717–1727 (1994). dependent pathway. Science 275, 200–203 (1997).
69. Ham, J. et al. A c-Jun dominant negative mutant protects sympathetic neurons against pro- 97. Nishina, H. et al. Stress-signaling kinase Sek1 protects thymocytes from apoptosis mediated by
grammed cell death. Neuron 14, 927–939 (1995). CD95 and CD3. Nature 385, 350–353 (1997).
70. Colotta, F., Polentarutti, N., Sironi, M. & Mantovani, A. Expression and involvement of c-fos and c- 98. Rebollo, A. et al. Bcl-3 expression promotes cell survival following interleukin-4 deprivation and is
jun protooncogenes in programmed cell death induced by growth factor deprivation in lymphoid controlled by AP1 and AP1-like transcription factors. Mol. Cell. Biol. 20, 3407–3416 (2000).

E136 NATURE CELL BIOLOGY VOL 4 MAY 2002 http://cellbio.nature.com

© 2002 Nature Publishing Group

You might also like