The EMBO Journal Vol.17 No.6 pp.1740–1749, 1998
Differential targeting of MAP kinases to the
ETS-domain transcription factor Elk-1
Shen-Hsi Yang, Alan J.Whitmarsh
1
,
Roger J.Davis
1
and Andrew D.Sharrocks
2
Department of Biochemistry and Genetics, The Medical School,
University of Newcastle upon Tyne, Newcastle upon Tyne NE2 4HH,
UK and
1
Program in Molecular Medicine, Department of Biochemistry
and Molecular Biology, Howard Hughes Medical Institute, University
of Massachusetts Medical School, Worcester, MA 01605, USA
2
Corresponding author
The activation of MAP kinase (MAPK) signal transduc-
tion pathways results in the phosphorylation of tran-
scription factors by the terminal kinases in these
cascades. Different pathways areactivated by mitogenic
and stress stimuli, which lead to the activation of
distinct groups of target proteins. The ETS-domain
transcription factor Elk-1 is a substrate for three
distinct classes of MAPKs. Elk-1 contains a targeting
domain, the D-domain, which is distinct from the
phosphoacceptor motifs and is required for efficient
phosphorylation and activation by the ERK MAPKs.
In this study, we demonstrate that members of the
JNK subfamily of MAPKs are also targeted to Elk-1
by this domain. Targeting via this domain is essential
for the efficient and rapid phosphorylation and activa-
tion of Elk-1 both in vitro and in vivo. The ERK and
JNK MAPKs use overlapping yet distinct determinants
in the D-domain for targeting to Elk-1. In contrast,
members of the p38 subfamily of MAPKs are not
targeted to Elk-1 via this domain. Our data therefore
demonstrate that different classes of MAPKs exhibit
differential requirements for targeting to Elk-1.
Keywords: Elk-1/ETS-domain proteins/MAP kinase/
TCFs/transcription factor
Introduction
The MAP kinase (MAPK) pathways play major roles
in converting mitogenic and stress stimuli into nuclear
responses (reviewed in Treisman, 1996; Whitmarsh and
Davis, 1996). In humans, at least three parallel pathways
exist which can be classified according to the sequence
conservation in the terminal MAPKs. The ERK pathway
primarily transmits mitogenic and differentiation stimuli,
whereas the JNK and p38 pathways primarily transduce
stress and cytokine stimuli to the nucleus. These pathways
are conserved in a diverse range of organisms including
Saccharomyces cerevisiae, Drosophila melanogaster and
Caenorhabditis elegans (reviewed in Treisman, 1996).
Several distinct MAPKs have been identified in each class
of pathway. For example, the ERK subclass contains
ERK1 and ERK2, the JNK subclass contains JNK1, JNK2
and JNK3, and the p38 subclass contains p38α, p38β,
1740
© Oxford University Press
p38γ and p38δ (Gupta et al., 1996; Jiang et al., 1996;
Lechner et al., 1996; Stein et al., 1997; Cuenda et al.,
1997; Goedert et al., 1997; Wang et al., 1997; Enslen
et al., 1998; reviewed in Whitmarsh and Davis, 1996;
Robinson and Cobb, 1997). In addition, multiple MAPK
isoforms are generated by alternative splicing which is
best illustrated by the JNK MAPKs (Gupta et al., 1996).
The large number of MAPKs potentially allows differential
phosphorylation of nuclear substrates and thereby specific
responses to upstream signals. However, the substrate
specificity determinants for these related kinases are just
beginning to be understood.
A combination of in vitro and in vivo approaches has
led to the identification of several nuclear targets for
MAPK pathways. For example, c-Myc (Gupta and Davis,
1994) and Spi-B (Mao et al., 1996) are ERK substrates,
c-Jun (Derijard et al., 1994; Kyriakis et al., 1994) and
NFAT4 (Chow et al., 1997) represent JNK targets, whereas
CHOP (Wang and Ron, 1996) and MEF2C (Han et al.,
1997) are targets for the p38 pathways. ATF-2 (Gupta
et al., 1995; Livingstone et al., 1995; Van Dam et al.,
1995) and ATFa (Gupta et al., 1995; Bocco et al., 1996)
are targets for both the JNK and p38 stress-activated
pathways. Members of the ternary complex factor (TCF)
sub-family of ETS-domain transcription factors are also
targets of MAPK pathways (reviewed in Treisman, 1996;
Whitmarsh and Davis, 1996). The TCF Elk-1 is a target
for all three pathways (Price et al., 1996; Janknecht and
Hunter, 1997a; Whitmarsh et al., 1997; reviewed in
Treisman, 1996). However, another family member,
SAP-1, appears to only be phosphorylated and activated
efficiently by the ERK and p38 pathways (Price et al.,
1996; Whitmarsh et al., 1995, 1997; Strahl et al., 1996),
although it can also act as a JNK substrate (Janknecht and
Hunter, 1997b). Phosphorylation of the TCFs takes place
at multiple residues in the conserved C-terminal transcrip-
tional activation domain (C-domain) which leads to
enhanced DNA binding and TCF-mediated transcriptional
activation (reviewed in Treisman, 1994, 1996; Price
et al., 1996).
It is becoming apparent that the substrate specificity of
MAPKs is determined by two components: the sequence
and local context of the phosphoacceptor motifs, and
binding of the kinase to docking sites on the substrate.
For example, in the case of phosphorylation of c-Jun by
the JNK MAPKs, the local context of the phosphoacceptor
motifs plays a major role in substrate specificity determina-
tion in combination with targeting via a kinase docking
domain on the transcription factor (Derijard et al., 1994;
Kallunki et al., 1994, 1996; Sluss et al., 1994; Dai et al.,
1995; Gupta et al., 1996). A short region of c-Jun, the
δ-domain, appears to be sufficient for this interaction
(Derijard et al., 1994; Kallunki et al., 1994, 1996; Sluss
et al., 1994; Dai et al., 1995; Gupta et al., 1996). Similar
MAPK targeting of Elk-1
interactions occur between JNK MAPKs and ATF-2 via
a short motif which is distinct from the phosphoacceptor
sites (Gupta et al., 1995, 1996; Livingstone et al., 1995).
It has recently been demonstrated that the ERK MAPKs
are also targeted to a nuclear substrate, Elk-1, by a docking
motif known as the D-domain which is conserved amongst
the TCFs (Yang et al., 1998). The D-domain is located
N-terminally from the transcriptional activation domain
and is required for efficient phosphorylation of Elk-1
within this adjacent domain and hence enhancement of its
transcriptional activation potential (Yang et al., 1998).
In this study, we have investigated the targeting of a
panel of mitogen- and stress-activated MAPKs to Elk-1.
Elk-1 has been implicated as a substrate for members of
all three classes of MAPKs and hence represents an
excellent candidate to investigate targeting of MAPKs.
The ERK and JNK MAPKs are both targeted to Elk-1
via the D-domain. Targeting is essential for efficient
phosphorylation of Elk-1 in vitro and in vivo. However,
different residues in the D-domain are important for ERK
and JNK binding. In contrast, the p38 MAPKs are not
targeted to Elk-1 via the D-domain. Our results therefore
demonstrate that Elk-1 contains a docking site that contains
specificity determinants which allow recognition by two
different classes of MAPKs but exclude binding of a third
class. Such specificity determinants within transcription
factor substrates are likely to play a pivotal role in
producing unique nuclear responses to the activation of
diverse MAPK signal transduction pathways.
Results
Differential requirement for the Elk-1 D-domain for
phosphorylation by MAPKs in vitro
Elk-1 has been shown to be an in vitro substrate for all
three classes of MAPKs and to be regulated by these
kinase cascades in vivo (Price et al., 1996; Janknecht and
Hunter, 1997a; Whitmarsh et al., 1997; reviewed in
Treisman, 1996; Whitmarsh and Davis, 1996). It has
recently been demonstrated that Elk-1 contains an ERK
targeting motif that maps to the D-domain which is located
N-terminally from the transcriptional activation domain
(Figure 1A). The integrity of this domain is required to
allow efficient Elk-1 phosphorylation in vitro (Yang et al.,
1998). In order to investigate whether the efficiency of
Elk-1 phosphorylation by members of the stress-activated
MAPKs is enhanced by the presence of the D-domain,
we investigated the ability of a representative member of
each subclass, JNK-1 and p38γ, to phosphorylate a series
of truncated Elk-1 derivatives (Figure 1).
N-terminal truncations of Elk-1 up to amino acid 310
(Elk-310) did not alter the efficiency of Elk-1 phosphoryl-
ation by ERK2. However, truncations up to and beyond
amino acid 330 (Elk-330 and Elk-349), which delete the
D-domain, resulted in a significant reduction in Elk-1
phosphorylation (Figure 1A; Figure 1B, lanes 1–4; Yang
et al., 1998). Similarly, the efficiency of phosphorylation
of the proteins Elk-330 and Elk-349 by JNK-1 was
dramatically reduced (Figure 1A; Figure 1B, lanes 5–8).
In contrast, the efficiency of phosphorylation of all the
Elk-1 deletion proteins by p38γ was virtually identical
(Figure 1A; Figure 1B, lanes 9–12). Taken together, these
results indicate that the integrity of the D-domain is
1741
Fig. 1. Differential requirement for the D-domain for efficient Elk-1
phosphorylation by MAPKs. (A) A diagram illustrating a series of
truncated Elk-1 proteins fused to GST. The black box represents the
D-domain of Elk-1 (amino acids 312–334), and the numbers of the
N- and C-terminal amino acids in the Elk-1 moiety are indicated. The
degree of phosphorylation of each protein in the immune complex
protein kinase assay (relative to Elk-310) is indicated. (B) The
phosphorylation of GST–Elk fusion proteins by MAPKs was examined
using either activated ERK2 or epitope-tagged JNK-1 and p38γ which
had been immunopurified from UV-activated COS-1 cells. The activity
of each protein kinase towards the substrate GST–Elk205 was
standardized with respect to activated ERK2 (1U, NEB). Kinase
assays were performed for 15 min at 30°C with equal molar quantities
(5 pmol) of GST–Elk-1 fusion proteins as substrates.
required for efficient phosphorylation of Elk-1 by the
ERK and JNK MAPKs but not by p38γ.
In order to determine whether the differential require-
ment for the D-domain is conserved within each class of
MAPK, we tested the ability of several members of
each subfamily to phosphorylate full-length Elk-1 in the
presence or absence of the D-domain. The activity of each
MAPK towards Elk-1 was initially standardized (Figure
2B, lanes 1–4; Figure 2C). The kinetics of phosphorylation
of wild-type Elk-1 were virtually indistinguishable (Figure
2B, lanes 1–4), although graphical analysis indicated that
phosphorylation by ERK-2, JNK-1 and, to a lesser extent,
JNK-2 was non-linear and rapidly reached maximal levels
(Figure 2C). In contrast, differential phosphorylation of
Elk-1D (which lacks the D-domain) by individual
MAPKs was observed (Figure 2B, lanes 5–8; Figure 2C).
The kinetics and overall efficiency of phosphorylation
of Elk-1D by p38α, p38β
2
and p38γ were virtually
indistinguishable from wild-type Elk-1 (compare Figure
2B, lanes 1–4 and 5–8; Figure 2C). In contrast, the kinetics
of Elk-1D phosphorylation by ERK2, JNK-1 and JNK-2
were delayed and the overall efficiency greatly reduced.
It appears, therefore, that the differential requirement for
the D-domain for efficient phosphorylation of Elk-1 is a
common property of all members of a particular subclass
of MAPK.
Identification of important residues in the
D-domain
The D-domain plays an important role in enhancing Elk-1
phosphorylation by the ERK and JNK MAPKs. In order
S.-H.Yang et al.
Fig. 2. Kinetic analysis of Elk-1 phosphorylation by MAPKs in the
presence and absence of the D-domain. (A) Diagrammatic illustration
of full-length Elk-1 and Elk-1 with a nine amino acid deletion in the
D-domain (Elk-1D). Bacterially expressed FLAG epitope-tagged full-
length Elk-1 and Elk-1D proteins were purified and equimolar
concentrations were used in all assays. (B) Full-length Elk-1 (lanes 1–
4) and Elk-1D (lanes 5–8) (3.75 pmol of each) were phosphorylated
by MAPKs (as indicated on the right panel of the figure) for the times
indicated above each lane. The activity of each kinase was
standardized to ERK-2 as described in Figure 1. (C) Graphical
representation of the data from (B). Open and closed squares denote
wild-type Elk-1 and Elk-1D respectively used as MAPK substrates.
Data are presented relative to phosphorylation of wild-type Elk-1 after
120 min (taken as 100).
to investigate whether identical residues within this domain
are involved in this function, pairs of amino acids were
mutated to alanine residues (Figure 3A). Such mutations
should preserve any structural motifs which are present but
remove side chains which are available for intermolecular
interactions. The mutant proteins were tested as substrates
for a panel of MAPKs (Figure 3B). Mutations in the N-
terminal half of the D-domain resulted in significant
reductions in the efficiency of phosphorylation by ERK2
(Figure 3B, mutants M1 and M2; Yang et al., 1998).
Similarly, reductions in the efficiency of phosphorylation
of the M1 and M2 mutants by JNK-1 and JNK-2 were
observed (Figure 3B, lanes 2 and 3). However, in the case
of JNK-1 and JNK-2, additional important residues were
identified by the M3 mutant which was also phosphorylated
to a reduced level (Figure 3B, lane 4). In contrast, none
of the mutations resulted in largedecreases in the efficiency
1742
Fig. 3. Identification of residues within the D-domain required for
efficient phosphorylation by MAPKs. (A) Amino acid sequence of the
wild-type (WT) and D-domain mutants R314A/K315A (M1), R317A/
L319A (M2), L323A/S324A (M3), L327A/L328A (M4) and L319A/
L321A (M5) are indicated. Numbers above the sequence represent the
N- and C-terminal residues in the D-domain. (B) Kinase assays of
wild-type and mutant GST–Elk205 fusion proteins by MAPKs were
carried out as described in Figure 1. Amounts of phosphorylation of
the mutant M1, M2, M3, M4 and M5 Elk-1 proteins relative to WT
Elk-1 (taken as 1) are 0.47 6 0.06, 0.40 6 0.04, 1.10 6 0.22, 1.12 6
0.28 and 0.23 6 0.11 (ERK2; Yang et al., 1998); 0.45 6 0.06, 0.39 6
0.11, 0.63 6 0.10, 1.19 6 0.20 and 0.21 6 0.06 (JNK-1); 0.39 6
0.04, 0.28 6 0.06, 0.46 6 0.09, 1.01 6 0.01 and 0.14 6 0.05
(JNK-2); 0.79 6 0.07, 0.72 6 0.04, 1.0 6 0.09, 1.17 6 0.24 and
0.75 6 0.35 (p38α); 0.71 6 0.15, 0.66 6 0.04, 0.97 6 0.01, 1.05 6
0.07 and 0.77 6 0.09 (p38β
2
); and 0.97 6 0.08, 0.9 6 0.18, 0.98 6
0.17, 1.08 6 0.18 and 1.22 6 0.31 (p38γ).
of Elk-1 phosphorylation by p38γ and, in comparison with
the effect on ERK and JNK MAPKs, only small differences
were seen for phosphorylation of the M1 and M2 mutants
by p38α and p38β
2
(Figure 3B). Finally, an additional
mutant was created (M5) which has mutations in both of
the leucine residues which make up the conserved ‘LXL
central core of several MAPK targeting motifs (Yang
et al., 1998). The efficiency of phosphorylation of the M5
mutant by the ERK and JNK MAPKs is severely reduced
whereas little effect is seen on phosphorylation by the
p38 MAPKs (Figure 3B, lane 7).
These results demonstrate that residues in the D-domain
play different roles in directing phosphorylation of Elk-1
by MAPKs. Residues in the N-terminal end of this domain
are important for both the ERK and JNK MAPKs, with
further C-terminal residues also being important for the
JNKs. In contrast, mutations within the D-domain have
minimal effects on phosphorylation by the p38 MAPKs,
consistent with the lack of effect of deleting this domain.
Differential binding of MAPKs to the Elk-1
D-domain
It previously has been demonstrated that the ability of
ERK2 to phosphorylate Elk-1 correlates with its ability
to bind to its substrate via the D-domain (Yang et al.,
1998). Moreover, JNKs have also been demonstrated to
MAPK targeting of Elk-1
interact physically with their substrate c-Jun via the
δ-domain (Derijard et al., 1994; Kallunki et al., 1994;
Sluss et al., 1994; Dai et al., 1995) and with Elk-1 (Gille
et al., 1996), although in the latter case the binding site
was not determined. In comparison with ERK2, binding
of JNKs to Elk-1 was barely detectable using the GST
pull-down assay (Gupta et al., 1996; data not shown),
suggesting that the kinetics of interaction differ between
these two different classes of kinases. In order to investi-
gate binding of the kinases to Elk-1 under the same
conditions, we therefore adopted a different approach
using a peptide competition assay. Peptides were synthe-
sized which correspond to Elk-1 amino acids 311–328
and encompass the D-domain (Figure 4A). Increasing
amounts of these peptides were included in kinase assays
to compete for binding of the MAPKs to Elk-1 via the
D-domain. Mutant D-domain peptides and a peptide from
JIP-1, a specific JNK inhibitor (Dickens et al., 1997),
were used as controls (Figure 4A).
The wild-type and M3 mutant D-domain peptides both
acted as inhibitors of ERK2 in a concentration-dependent
manner (Figure 4B, lanes 2–4 and 8–10). In contrast, the
M5 and JIP peptides did not act as competitors (Figure
4B, lanes 5–7 and 12–13). These results are consistent
with the observation that the M3 mutation affects neither
the efficiency of phosphorylation nor binding by ERK2,
whereas the M5 mutation inhibits both these functions
(Figure 3; Yang et al., 1998; data not shown). Similarly,
the wild-type peptide inhibits phosphorylation by the
JNK kinases whereas the M5 peptide is an ineffective
competitor. In comparison with ERK2, the wild-type
peptide inhibits phosphorylation by JNK-1 and JNK-2 at
a 10-fold lower concentration (Figure 4B, compare lanes
3 and 4). Significantly, inhibition of JNK-1 and JNK-2 by
the M3 peptide is reduced in comparison with the wild-
type peptide (Figure 4B, compare lanes 3 and 9), which
correlates with the reduction in phosphorylation observed
in the M3 mutant protein (Figure 3). The control JIP
peptide acts as an efficient JNK inhibitor in this assay
(Figure 4B, lanes 12 and 13). In contrast, none of the
peptides used act as efficient inhibitors of the p38 MAPKs
(Figure 4B).
In order to confirm that the peptides are blocking
substrate binding rather than impairing the catalytic
activity of the MAPKs, they were used as competitors in
kinase reactions containing ERK2 and JNK-2 and the
substrates myelin basic protein (MBP) and c-Jun respect-
ively. The binding of MAPKs to a docking domain on
MBP is not thought to occur and, consistent with this
concept, phosphorylation of MBP by ERK2 was not
inhibited by the presence of either the wild-type, M3
or M5 peptides (Figure 4C). The JIP peptide inhibited
phosphorylation of c-Jun by JNK-2 (Figure 4D, lanes 8
and 9). This inhibition previously has been attributed to
competition for binding of the kinase to the c-Jun δ-domain
(Dickens et al., 1997). Similarly, the wild-type D-domain
peptide inhibited phosphorylation of c-Jun by these kin-
ases, albeit to a lesser extent (Figure 4D, lanes 2 and 3).
Mutant D-domain peptides which are defective in JNK
binding are also defective in inhibiting phosphorylation
of c-Jun by JNKs (Figure 4D, lanes 4–7). Thus, the
D-domain peptide is acting as expected to disrupt inter-
1743
Fig. 4. The Elk-1 D-domain acts as a binding site for the ERK and
JNK MAPKs. (A) The sequences of peptide competitors and a
diagrammatic illustration of the substrate (GST–Elk310, 5 pmol) used
in the kinase reactions. Identical or highly conserved (Arg/Lys)
residues conserved in the Elk-1- and JIP-derived peptides are indicated
by dashes. (B) Phosphorylation of Elk-1 by MAPKs in the presence of
competitor peptides. The peptide competition assay was based on the
kinase assays described in Figure 1 except that the MAPKs were pre-
incubated in the absence (lanes 1 and 11) or presence of competitor
peptides (10- to 1000-fold excess over Elk-1 substrate), 50 pmol (lanes
2, 5, 8 and 12), 500 pmol (lanes 3, 6, 9 and 13) and 5 nmol (lanes 4,
7 and 10), respectively. (C) and (D) Competition assays were carried
out as in (B) by using 5 pmol of MBP and GST–c-Jun as substrates,
respectively. The assays were performed in the absence (C and D, lane
1) or presence of peptide competitors, 50 pmol (C, lanes 2, 5 and 8;
D, lanes 2, 4, 6 and 8), 500 pmol (C, lanes 3, 6 and 9; D, lanes 3, 6, 9
and 12) and 5 nmol (C, lanes 4, 7 and 10), respectively. Increases in
the concentration of added competitor peptides are indicated
schematically above each set of lanes.
actions between the JNKs and c-Jun but is an ineffective
competitor of phosphorylation of MBP by ERK2.
Collectively, these data demonstrate that the D-domain
acts as a binding site for both the ERK and JNK MAPKs,
but phosphorylation of Elk-1 by the p38 MAPKs does not
take place via this mechanism.
S.-H.Yang et al.
The JNK targeting domains of Elk-1 and c-Jun act
as discrete interchangeable modules
The Elk-1 D-domain is sufficient to bind to ERK2 in a
heterologous context when fused to another protein (Yang
et al., 1998). In order to investigate whether the D-domain
can act in a similar manner to allow JNK targeting,
reciprocal chimeric proteins were created in which the
MAPK targeting domains of c-Jun and Elk-1 were
exchanged (Figure 5A). Deletion of the Elk-1 D-domain
(Elk-330; Figure 1; Figure 5B, lanes 1 and 2) or the
δ-domain from c-Jun (cJun43–223; Figure 5B, lanes 7
and 8) resulted in a loss in the efficiency of phosphorylation
by both JNK-1 and JNK-2. However, addition of the
reciprocal targeting domains to these truncated proteins
converted them back to efficient JNK substrates (cJunδ
Elk-330 and ElkD–cJun; Figure 5B, lanes 4 and 5).
Interestingly, of the two chimeric proteins, only cJunδ
Elk-330 exhibited enhanced phosphorylation by ERK2
1744
despite the D-domain in the ElkD–cJun chimera represent-
ing an ERK-binding motif (Figure 5B, top panel, lanes 3
and 4). This is in agreement with the observation that the
local context of the phosphoacceptor motifs also plays
a major role in determining the specificity of c-Jun
phosphorylation (Gupta et al., 1996; Kallunki et al., 1996).
Finally, in order to rule out that the D-domain itself was
phosphorylated by the MAPKs, two control constructs
were made and tested. Neither ElkD (which contains
the D-domain alone) nor ElkD–cJun(S63A/S73A) (which
lacks the primary JNK phosphoacceptor motifs) represent
good ERK or JNK substrates (Figure 5B, lanes 3 and 6).
In order to investigate binding of the kinases and the
relative strengths of the different binding motifs, peptide
competition assays were carried out with the chimeric
Elk-1–c-Jun proteins (Figure 5C). The wild-type D-domain
and JIP peptides acted as efficient competitors of ElkD–
cJun phosphorylation by JNKs (Figure 5C, lanes 2 and
5). Competition was also exhibited by the M3 peptide,
albeit to a lesser extent with JNK-2, whereas the M5
peptide was an ineffectual competitor (Figure 5C, lanes 3
and 4). These results are essentially the same as observed
with wild-type Elk-1 (Figure 4), thereby demonstrating that
the D-domain acts in a similar manner in a heterologous
context. Competition assays were also carried out with
cJunδ–Elk-330, and again the JIP peptide acted as an
efficient competitor of phosphorylation by JNKs (Figure
5C, lane 10). The wild-type and M3 D-domain peptides
also acted as competitors, but to a lesser extent than
observed with ElkD–cJun (Figure 5C, compare lanes 7
and 9 with lanes 2 and 4), indicating that the δ-domain
acts as a stronger JNK-binding motif. Furthermore, com-
petition for JNK-1 was more pronounced than for JNK-2
(Figure 5C, lanes 7 and 9; compare JNK-1 and JNK-2
panels). This is consistent with the observation that the
c-Jun δ-domain preferentially binds JNK-2 (Kallunki et al.,
1994; Sluss et al., 1994; Gupta et al., 1996) whereas the
Elk-1 D-domain preferentially binds JNK-1 (Figure 4;
data not shown). Residual ERK phosphorylation of cJunδ
Elk-330 was inhibited by the wild-type and M3 peptides
Fig. 5. The JNK-binding domain of Elk-1 (D-domain) and c-Jun
(δ-domain) can be functionally interchanged. (A) A diagram
illustrating fusions of GST to Elk-310 (amino acids 310–428), Elk-330
(amino acids 330–428), ElkD (amino acids 310–329), cJunδ–Elk-330
chimera (amino acids 1–58 of c-Jun, Elk-1 amino acids 330–428),
ElkD–cJun chimera (amino acids 321–348 of Elk-1, c-Jun amino acids
55–223), ElkD–cJun (S63A/S73A), cJun43–223 (c-Jun amino acids
43–223) and c-Jun (amino acids 1–223). The numbers of the N- and
C-terminal c-Jun (normal type) and Elk-1 (italics) amino acids are
indicated above and below each construct respectively. The locations
of the major JNK phosphorylation sites in c-Jun (Ser63 and Ser73) are
indicated as the most functionally important phosphoacceptor sites in
Elk-1 (Ser383 and Ser389). The sequences of the Elk-1 D-domain and
c-Jun δ-domain are shown and are indicated by solid and open boxes,
respectively. The Elk-1 transcriptional activation domain (C-) is shown
by a grey box. Identical or highly conserved (Arg/Lys) residues
conserved between the two domains are indicated by dashes. The
central ‘LXL motif is bracketed. (B) Kinase assays of GST fusion
proteins as substrates for the indicated MAPKs were carried out using
5 pmol of each protein for 30 min reactions as described in Figure 1.
(C) Phosphorylation of chimeric Elk-1–cJun chimeras by MAPKs in
the presence of competitor peptides. Peptide competition assays were
carried out as described in Figure 4. Five pmol of each chimeric
protein were used in the reactions with 500 pmol of the indicated
peptides. The exposure of the ERK2 panel was selected in order to
show equivalent ERK and JNK phosphorylation of cJunδ–Elk-330.
MAPK targeting of Elk-1
Fig. 6. Role of the D-domain in Elk-1 phosphorylation in vivo in IL-1-stimulated CHO and NIH-3T3 cells. CHO (A–D) or NIH-3T3 (E–G) cells
were transfected with 2 µg of CMV-driven expression vectors encoding either wild-type Elk-1 or Elk-1D. Total cell extracts were taken at the
indicated times after IL-1 stimulation and analysed by Western blot using the anti-Phospho Elk-1(383) antibody to detect phosphorylation of Ser383
(A and E) or anti-FLAG antibody to examine the total levels of Elk-1 and Elk-1D in each sample (B and F). Samples were also analysed by gel
retardation analysis in the presence of core
SRF
and the c-fos SRE (SRE*) (C, D, G and H). Anti-phospho Elk-1(383) antibody was also included.
Supershifted bands representing complexes containing phosphorylated Elk-1 derivatives are shown in (C) and (G). Ternary complexes with core
SRF
and a 134 bp fragment of c-fos promoter containing the SRE (SRE*) are shown in (D) and (H). The locations of unphosphorylated ternary complex
(3°I) and multiple phosphorylated forms of ternary complex (3°II) are indicated. (C and D) and (G and H) are from separate experiments using the
same extracts.
but not by the M5 and JIP peptides (Figure 5C, top
panel, lanes 6–10) as predicted from previous experiments
(Figure 4).
Collectively, these data demonstrate that the Elk-1
D-domain can act as a JNK-binding motif in a
heterologous context. However, differences in the relative
strength of kinase binding are apparent, with the c-Jun
δ-domain acting as a stronger JNK-2-binding site than
the Elk-1 D-domain and the Elk-1 D-domain preferably
binding JNK-1 rather than JNK-2.
Differential targeting of the JNK and p38 MAPKs
to Elk-1 in vivo
Evidence has been gathered to suggest that Elk-1 is a
substrate for both the JNK (Cavigelli et al., 1995; Gille
et al., 1995; Whitmarsh et al., 1995, 1997) and p38 (Price
et al., 1996; Whitmarsh et al., 1997) MAPK subclasses
in vivo. However, the response of Elk-1 to these different
stress-activated MAPK pathways differs depending on the
cell type and stimulus. For example, UV activates Elk-1
via the p38 pathway in HeLa and NIH-3T3 cells (Price
et al., 1996). In contrast, interleukin-1 (IL-1) activates
Elk-1 via the JNK pathway in CHO cells, but by a
combination of the JNK and p38 pathways in NIH-3T3
cells (Whitmarsh et al., 1997). We took advantage of this
cell type-specific response to IL-1 to investigate the role
of the D-domain in targeting stress-activated MAPKs to
Elk-1 in vivo.
CHO and NIH-3T3 cells were transfected with either
Elk-1 or Elk-1D, stimulated with IL-1, and cell extracts
subsequently were prepared over a 15 min time period.
1745
The phosphorylation status of the Elk-1 derivatives was
monitored indirectly by gel retardation analysis (Figure
6D and H) and directly by supershift analysis of the
Elk-1-containing ternary complexes using an anti-phospho
Elk-1(Ser383) antibody (Figure 6C and G) which was
also used in Western blotting experiments (Figure 6A and
E). In CHO cells, the phosphorylation of Ser383 in Elk-1
occurred rapidly within 5 min of stimulation (Figure 6A
and C, lane 3), and was accompanied by a shift in the
mobility of the ternary DNA-bound complex (Figure
6D, lane 3). In comparison, the speed and degree of
phosphorylation of Ser383 and induction of a lower
mobility ternary complex containing Elk-1D was
reduced. Maximal phosphorylation of Ser383 was not
reached for 15 min (Figure 6A and C), and the stimulation
of a lower mobility DNA-bound complex was reduced in
comparison with wild-type Elk-1 (Figure 6D, compare
lanes 1–5 and 6–10). Taken together with the observation
that JNKs represent the major IL-1-activated protein
kinases in CHO cells that activate Elk-1 (Whitmarsh et al.,
1997), these data provide strong evidence for a role of
the D-domain in targeting these cascades to Elk-1 in vivo.
In contrast, in NIH-3T3 cells, phosphorylation of Ser383
in response to IL-1 stimulation took place in Elk-1 and
Elk-1D with virtually identical kinetics (Figure 6E and
G). However, in comparison with CHO cells, the magni-
tude of Ser383 phosphorylation was reduced, which is
consistent with the observation that minimal effects on
the mobility of ternary complexes containing either Elk-1
or Elk-1D were observed (Figure 6H). Together with the
observation that p38 is a major effector of IL-1 signalling
S.-H.Yang et al.
Fig. 7. The role of the Elk-1 D-domain in IL-1-inducible
transcriptional activation in vivo.(A) A diagrammatic representation of
wild-type and mutant Elk-1 proteins fused to the DNA-binding domain
of GAL4. CHO (B and C) or NIH-3T3 cells (D and E) were
transfected with CMV promoter-driven constructs encoding GAL4
fusions to either wild-type or mutant Elk-1 derivatives and a GAL4-
driven luciferase reporter plasmid. Cells were either unstimulated or
stimulated with IL-1. Transfection efficiency was monitored by using
the β-galactosidase expression vector pCH110. The luciferase activities
relative to the unstimulated cells of each wild-type or mutant (means
6 standard deviations; n 5 3) are presented. Expression levels of the
GAL4 fusion proteins in CHO (C) and NIH-3T3 (E) cell lines were
examined by Western blotting using total cell extracts with an anti-
GAL4 antibody.
to Elk-1 in NIH-3T3 cells (Whitmarsh et al., 1997), these
results are consistent with the notion that the Elk-1 D-
domain is not required for targeting by p38 in vivo.
The response of GAL4 fusion proteins containing the
Elk-1 transcriptional activation domain to IL-1 stimulation
in these cell types was also investigated. Wild-type Elk-1
and proteins containing mutations in the D-domain were
constructed which either did not affect (M4) or reduced
(M5) the efficiency of phosphorylation by the JNKs
in vitro (Figure 7A). These proteins were all expressed at
equivalent levels in CHO and NIH-3T3 cells (Figure 7C
and E). In CHO cells, the response of the M5 mutant to
IL-1 stimulation was severely reduced (65% reduction)
whereas, in comparison, the M4 mutant was virtually
unaffected (15% reduction; Figure 7B). However, in NIH-
3T3 cells, the response of the M5 mutant was only slightly
reduced (20% reduction) and again the M4 mutant was
virtually unaffected (Figure 7D). These results therefore
add further weight to the notion that the D-domain is
required for targeting of JNK MAPKs but not the p38
MAPKs to Elk-1 in vivo.
Discussion
MAPK signalling cascades play a pivotal role in converting
extracellular signals into specific nuclear responses
1746
(reviewed in Treisman, 1996; Whitmarsh and Davis,
1996). In order to achieve such specific responses, MAPKs
must recognize their substrates with high specificity. A
general mechanism is emerging to generate such specificity
in which initial binding of the kinases to a docking domain
is followed by recognition of the local context of the
phosphorylation motifs (Gupta et al., 1996; Kallunki et al.,
1996; Yang et al., 1998). In the case of c-Jun, the δ-
domain directs binding by the JNK subclass of MAPKs
(Derijard et al., 1994; Kallunki et al., 1994; Sluss et al.,
1994; Dai et al., 1995). Similarly, the Elk-1 D-domain
directs binding of the ERK subclass of MAPKs (Yang
et al., 1998). However, in this study, we demonstrate that
the Elk-1 D-domain possesses dual specificity and is also
a target for the JNK MAPKs. This domain does not,
however, allow promiscuous MAPK binding but discrim-
inates against binding by the p38 MAPK subclass. This
is the first reported example of such a motif which allows
diverse MAPK pathways to be integrated via a single
transcription factor.
MAP kinase targeting domains
Short motifs which bind to the JNK MAPKs have been
identified in the JNK inhibitor protein JIP-1 (Dickens
et al., 1997) and the transcription factor c-Jun (Derijard
et al., 1994; Kallunki et al., 1994; Sluss et al., 1994; Dai
et al., 1995; Gupta et al., 1996). In this study, we have
identified a further motif in Elk-1 which targets JNKs to
transcription factors. Similarities between the binding
motifs from c-Jun and Elk-1 (Figure 5A) and JIP-1 and
Elk-1 (Figure 4A) are apparent, with the consensus motif
R
/
K
XXXXL
N
/
E
L representing a core binding site. The
JNK-binding site on ATF-2 also contains a motif which
loosely conforms to this consensus (KHEMTLKF) (Gupta
et al., 1995; Livingstone et al., 1995). However, examina-
tion of the binding motifs of several JNK substrates reveals
little similarity outside this central core motif although
more extensive similarities between pairs of proteins can
be observed (e.g. c-Jun/JIP-1 and Elk-1/NFAT4; Figure
8). Moreover, c-Jun preferentially binds to JNK-2
(Kallunki et al., 1994; Sluss et al., 1994; Gupta et al., 1996)
whereas Elk-1 preferentially binds to JNK-1 (Figures 4
and 5). Similarly, ATF-2 exhibits preferential binding of
certain JNK isoforms (Gupta et al., 1996). The non-
conserved residues surrounding this central core motif
must dictate these subtly different binding preferences.
Further specificity determinants must be built into the Elk-
1 D-domain as this is also bound efficiently by the ERK
MAPKs (Yang et al., 1998). Similarly, further specificity
determinants must be built into the ATF-2 docking site as
this serves as both a JNK- and p38-binding motif (N.Jones,
personal communication). Binding of the ERK and JNK
MAPKs to Elk-1 probably occurs with different kinetics
as, although binding can be detected by a competition
assay (Figure 4), only ERK binding can be detected by a
GST pull-down assay (Yang et al., 1998; data not shown).
Moreover, as wild-type D-domain peptides displace JNKs
more readily than ERK-2 from Elk-1, this suggests that
the JNKs may bind with high affinity but rapidly associate
and dissociate from the substrate in the presence of ATP
as observed with p38 binding to MEF2C (Han et al.,
1997). Stable complexes between Elk-1 and JNKs have
been demonstrated previously (Gille et al., 1996), but we
MAPK targeting of Elk-1
Fig. 8. Similarity amongst JNK-binding motifs. The sequences of the
JNK-binding domains found in Elk-1, c-Jun, JunB, JIP-1, NFAT4,
ATF-2 and ATFa are shown. The sequence of the D-domain of SAP-1
is also shown. The asterisk denotes that SAP-1 is predominantly an
ERK substrate and is poorly phosphorylated by JNKs. The MAPK-
binding motifs in Elk-1 and ATF-2 exhibit dual specificity with both
JNK and ERK MAPKs targeted to Elk-1 (this study) and both JNK
and p38 MAPKS targeted to ATF-2 via a single motif (N.Jones,
personal communication). Amino acid numbers of the N- and
C-terminal residues are given based on their location in either human
c-Jun, JunB, Elk-1, NFAT4, ATF-2, ATFa, SAP-1 or mouse JIP-1
proteins. Identical or highly conserved (Arg/Lys) amino acids
comprising the R/KXXXXLXL motif are highlighted. Brackets
indicate the conserved central MAPK-binding motif.
are unable to detect such complexes under conditions in
which we detect ERK2 binding. However, although we
can detect stable complexes of ERK2 and Elk-1 both
in vitro (Yang et al., 1998) and in vivo when both partners
are overexpressed (data not shown), the physiologically
relevant binding event is likely to involve a transient
interaction of the kinase and substrate after cellular stimu-
lation. An emerging theme amongst MAPK targeting
motifs is that variations on a simple core consensus can
dictate binding of specific subsets of transcription factors.
Specific rules are difficult to discern from the currently
characterized motifs, although all are short sequences (18
amino acids in the case of the D-domain) and the central
‘LXL’ motif clearly plays a pivotal role in kinase targeting
to transcription factors (e.g. M5 mutant, Figures 3, 4 and
7). However, MAPKs may also be targeted to substrates
by different motifs, as binding of the ERK and p38
MAPKs to Mnk1 and Mnk2 kinases does not appear to
be mediated by sequences related to the Elk-1 D-domain
(Waskiewiecz et al., 1997).
In contrast to the targeting of ERK and JNK MAPKs
to Elk-1 via the D-domain, this domain does not play a
role in the binding of p38 MAPKs (Figure 4). The
D-domain, therefore, does not represent a promiscuous
MAPK targeting motif but allows discrimination between
different classes of MAPKs.
Stress-activated MAPKs and Elk-1 activation
Elk-1 can be phosphorylated and activated by both the
JNK (Cavigelli et al., 1995; Gille et al., 1995; Whitmarsh
et al., 1995, 1997) and p38 (Price et al., 1996; Whitmarsh
et al., 1997) MAPK subclasses in vitro and in vivo.
However, the response of Elk-1 to these different stress-
activated MAPK pathways differs depending on the cell
type and stimulus. IL-1 activates Elk-1 via the JNK
pathway in CHO cells but by a combination of the JNK
and p38 pathways in NIH-3T3 cells (Whitmarsh et al.,
1747
1997), whereas UV activates Elk-1 via a combination of
the p38 and ERK pathways in HeLa and NIH-3T3 cells
(Price et al., 1996). In the present study, several lines of
evidence indicate that phosphorylation of Elk-1 requires
targeting of the kinase by binding to the D-domain for
the JNK but not the p38 MAPKs. First, deletions of the
D-domain reduce the efficiency of Elk-1 phosphorylation
by members of the JNK but not the p38 MAPK subclasses
(Figures 1 and 2). Secondly, point mutations in the
D-domain cause minimal reductions in the efficiency of
Elk-1 phosphorylation by the p38 subclass in comparison
with the JNK MAPKs (Figure 3). Thirdly, peptide competi-
tion assays demonstrate a role for the D-domain in binding
the JNK but not the p38 MAPKs (Figure 4). Interestingly,
although not as marked as the effect on the ERK and JNK
MAPKs, subtle changes in the proficiency of Elk-1 as a
substrate for p38α and, to a lesser extent, for p38β2 are
uncovered in the M1 and M2 mutants (Figure 3). As
deletion of the D-domain does not affect phosphorylation
of Elk-1 by the p38 MAPKs (Figures 1 and 2), this result
may reflect that these mutations cause subtle conforma-
tional changes in Elk-1 which may restrict kinase access
to the phosphoacceptor motifs. However, as the D-domain
peptide acts as a weak competitor of Elk-1 phosphorylation
by p38β
2
(Figure 4), this might reflect that the Elk-1-
binding site on p38 MAPKs retains some similarity to the
other MAPKs and hence binds weakly to Elk-1 via
the D-domain. Finally, in vivo, the efficiency of Elk-1
phosphorylation and activation by p38 MAPK is not
dependent upon the integrity of the D-domain (Figures 6
and 7). The latter result is also consistent with the
observation that stimulation of Elk-1 phosphorylation and
activity in vivo by the p38 MAPK pathways is not as
pronounced as activation by either the JNK (compare
Figure 6D and H and Figure 7B and D; Whitmarsh et al.,
1997) or ERK pathways (Yang et al., 1998). Furthermore,
in response to UV stimulation in HeLa and NIH-3T3
cells, the p38 pathway appears insufficient for full Elk-1
activation and requires cooperation with the ERK pathway
(Price et al., 1996). Finally, in 293 cells, Elk-1 does not
appear to be activated by p38 MAPKs (Janknecht and
Hunter, 1997a). The reduced activation by and lack of
targeting of p38 MAPKs to Elk-1 in vitro and in vivo in
comparison with the strong activation elicited by the ERK
and JNK MAPKs suggests that the residual activation by
p38 may be a consequence of overexpressing pathway
components and overriding normal specificity determin-
ants. The p38 MAPKs may, however, phosphorylate Elk-1
in a more constitutive manner which does not require rapid
and efficient kinase targetingto the substrate. Alternatively,
sustained p38 activation may be required to activate Elk-
1. Further studies are required to differentiate between
these possibilities. Finally, p38 may bind to a docking site
located elsewhere on Elk-1. However, none of the deleted
proteins we have tested exhibit markedly reduced phos-
phorylation by p38 isoforms (Figure 1; data not shown).
Moreover, no binding of p38α, β
2
and γ to Elk-1 could
be detected in GST pull-down assays carried out in the
absence of ATP (data not shown). Together these results
strongly suggest that p38 does not require or bind to a
docking motif on Elk-1.
In conclusion, our data contribute to the understanding
of how MAPKs recognize and phosphorylate their nuclear
S.-H.Yang et al.
targets. Elk-1 contains a MAPK-binding motif which
allows efficient targeting and subsequent transcription
factor activation by both the ERK and JNK subclasses of
MAPKs. Similarly, both JNK and p38 subclasses of
MAPKs are targeted via a single binding motif to ATF-2
(N.Jones, personal communication). Future studies are
likely to uncover further motifs which direct binding of
the p38 subclass of MAPKs to transcription factors (e.g.
MEF2C; Han et al., 1997). Investigation of these targeting
domains will provide significant insights into how the
specificity of signalling via MAPK cascades is achieved.
Materials and methods
Plasmid constructs
The following plasmids were used for expressing GST fusion proteins in
Escherichia coli. pAS407 (encoding GST–Elk205; Elk-1 amino acids
205–428), pAS545 (encoding GST–Elk310; Elk-1 amino acids 310–428),
pAS406 (encoding GST–Elk330; Elk-1 amino acids 330–428), pAS405
(encodingGST–Elk349;Elk-1 aminoacids 349–428)and pAS547(encod-
ing GST–ElkD–cJun; Elk-1 amino acids 310–348 fused to c-Jun amino
acids55–223)havebeendescribedpreviously(Yanget al.,1998).Plasmids
encoding GST–cJun1–223 and GST–cJun43–223 have been described
previously (Hibi et al., 1993). pAS767, encoding GST fused to Elk-1
aminoacids307–329(GST–ElkD), wasconstructedbyinsertingaBamHI–
EcoRI-cleaved PCR-derived fragment into the same sites of pGEX-3X.
pAS768, encoding c-Jun amino acids 1–58 fused to Elk-1 amino acids
330–428 (cJunδ–Elk-330), was constructed by ligating a BamHI–BglII-
cleavedPCRfragment(encodingc-Junaminoacids1–58)intotheBglIIsite
of pAS406. pAS548, pAS549, pAS550, pAS564 and pAS769 (encoding
GST–Elk205 mutants) are derivatives of pAS407 with the site-directed
mutations R314A/K315A (GST–Elk205M1), R317A/L319A (GST–
Elk205M2), L323A/S324A (GST–Elk205M3), L327A/L328A (GST–
Elk205M4) and L319A/L321A (GST–Elk205M5), respectively. pAS565
(GST–Elk307M2), pAS566 (GST–Elk307M3), pAS567 (GST–
Elk310
S383A/S389A
) and pAS568 (GST–Elk310
M2/S383A/S389A
) were
described previously (Yang et al., 1998). Mutations were introduced by a
two-step PCR protocolusinga mutagenic primer and two flanking primers
as described previously (Shore et al., 1996).
pAS278 and pAS380 [encoding full-length His/FLAG-tagged Elk-1
and the same protein with an internal deletion of amino acids 312–321
(Elk-1D)] were used for expression of Elk-1 derivatives in E.coli (Yang
et al., 1998).
The following plasmids were constructed for use in mammalian cell
transfections. pG5E1b contains five GAL4 DNA-binding sites cloned
upstream of a minimal promoter element and the firefly luciferase gene
(Seth et al., 1992). pSG424 encodes the GAL4 DNA-binding domain
(Sadowski and Ptashne, 1989). The vectors pCDNA3-F-JNK1 (Derijard
et al., 1994), pCDNA3-F-JNK2 (Sluss et al., 1994), pCMV5-F-p38α
(Raingeaud et al., 1995), pCDNA3-F-p38β
2
(Enslen et al., 1998)
and pCDNA3-F-p38γ (J.Raingeaud and R.J.Davis, unpublished data)
encoding flag-tagged MAPKs have been described previously. pAS572
(pCMV-GAL4-Elk205) and pAS577 (pCMV-GAL4-Elk205M4) were
described previously (Yang et al., 1998). pAS770 (GAL4-Elk205M5)
was constructed by ligating BamHI–XbaI fragments from pAS769 into
the same sites of pSG424. pAS771 (pCMV-GAL4-Elk205M5) was
constructed by ligating the HindIII–XbaI fragments from pAS770 into
the same sites of pCMV5. The cytomegalovirus (CMV) promoter-driven
expression vectors, pAS383 and pAS387, encoding full-length Elk-1
and Elk-1D, respectively, with C-terminal FLAG tags were described
previously (Yang et al., 1998). All plasmid constructs encoding Elk-1-
derived proteins made by PCR were verified by automated dideoxy
sequencing.
Protein expression and purification
GST fusion proteins were expressed in the E.coli JM101 strain and
purified asdescribedpreviously (Shore et al.,1995). Full-length hexahisti-
dine-tagged polypeptides were expressed in E.coli BL21(DE3)pLysS
with the pET vector system and quantified as described previously (Yang
et al., 1998).
Tissue culture, cell transfection and reporter gene assays
COS-1 cells were maintained in Dulbecco’s modified Eagle’s medium
(DMEM) supplemented with 10% fetal bovine serum (FBS) (Gibco-
1748
BRL). CHO cells were maintained in F12 medium supplemented with
5% FBS. NIH-3T3 cells were maintained in DMEM supplemented with
10% FBS. Transfection experiments were carried out using Superfect
transfection reagent (Qiagen) as described previously (Yang et al., 1998).
For reporter gene assays, GAL4-driven promoters were co-transfected
with various vectors encoding GAL4–Elk-1 fusion proteins. The activities
of the GAL4 DNA-binding domain (amino acids 1–147) and GAL4–
Elk-1 deletion fusion proteins (50 ng of plasmid DNA) were measured
in co-transfection assays in all cell lines using 1 µg of reporter plasmid
pG5E1bLuc. Transfection efficiencies were normalized by measuring
the activity from a co-transfected plasmid (1 µg) which expresses
β-galactosidase (pCH110, Pharmacia KB Biotechnology Inc.). Cell
extracts were prepared, and luciferase and β-galactosidase assays were
carried out as described previously (Yang et al., 1998).
Protein kinase assays
In order to prepare recombinant JNK and p38 MAPKs, COS-1 cells
were transfected with constructs encoding FLAG epitope-tagged MAPKs.
Kinases were activated by stimulation with UV light for 30 min. Purified
kinases were then eluted from beads by competing with 0.1 mg/ml of
FLAG peptide. Recombinant active ERK2 was obtained from New
England Biolabs (NEB) and MBP was obtained from Sigma. The kinase
assays were carried out in 20 µl reaction volumes as described previously
(Yang et al., 1998). The phosphorylation of substrate proteins was
examined following SDS–PAGE by autoradiography, and quantified by
phosphorimaging (Fuji BAS1500; TINA 2.08e software). Data were
quantified by phosphorimaging and the data presented graphically after
curve fitting with the appropriate equation using BIOSOFT Fig.P or
Microsoft Excel software. Peptide competition experiments were carried
out essentially as described above, except that pre-incubation with
50–5000 pmol of the peptide competitors with MAPKs was carried out
before the kinase reactions. Final peptide concentrations were
2.5–25 µM (10- to 100-fold excess over Elk-1 substrate).
Western blot analysis
FLAG-tagged Elk-1 and Elk-1D in extracts from CHO or NIH-3T3
cells were detected by immunoblot analysis using a mouse monoclonal
anti-M2 FLAG antibody (Kodak) or anti-phosphoplus™ Elk-1 (S383)
antibody (NEB). GAL4 fusion proteins were detected using the anti-
GAL4 antibody directed against the amino-terminal DNA-binding
domain (Santa Cruz). Immune complexes were detected by using
horseradish peroxidase-conjugated secondary antibody followed by ECL
(Amersham).
Gel retardation assays
Gel retardation assays were performed with a
32
P-labelled 134 bp c-fos
promoter fragment containing the SRE (SRE*) as described previously
(Whitmarsh et al., 1995). Total cell extracts from transfected cells
containing ~0.02 pmol of Elk-1 or Elk-1D were used in DNA-binding
reactions. Binding reactions on SRE-containing sites also contained
purified bacterially expressed core
SRF
(Shore and Sharrocks, 1994).
Antibody supershift experiments were described previously (Yang et al.,
1998). Protein–DNA complexes were analysed on non-denaturing 5%
polyacrylamide gels cast in 0.53 Tris–borate–EDTA and visualized by
autoradiography and phosphorimaging.
Figure generation and data quantification
All figures were generated electronically from scanned images of
autoradiographic images by using Picture Publisher (Micrografix) and
Powerpoint version 7.0 (Microsoft) software. Final images are represent-
ative of the original autoradiographic images. Phosphorimager data were
quantified using Tina software (version 2.08e).
Acknowledgements
We thank Margaret Bell and Catherine Pyle for excellent technical and
secretarial assistance, Bob Liddell for DNA sequencing and oligonucleo-
tide synthesis and Arthur Moir for peptide synthesis. We are grateful to
Steve Yeaman, Janet Quinn and members of our laboratories for
comments on the manuscript and for stimulating discussions, and to Nic
Jones for communicating data prior to publication. This work was
supported by the North of England Cancer Research Campaign, the
Wellcome Trust and the National Cancer Institute (USA). R.J.D. is an
investigator of the Howard Hughes Medical Institute. A.D.S. is supported
by the Lister Institute of Preventative Medicine.
MAPK targeting of Elk-1
References
Bocco,J.L., Bahr,A., Goetz,J., Hauss,C., Kallunki,T., Kedinger,C. and
Chatton,B. (1996) In vivo association of ATFa with JNK/SAP kinases
activities. Oncogene, 12, 1971–1980.
Cavigelli,M., Dolfi,F., Claret,F.X. and Karin,M. (1995) Induction of c-fos
expression through JNK-mediatedTCF/Elk-1 phosphorylation. EMBO.
J., 14, 5957–5964.
Chow,C.-W., Rincon,M., Cavanagh,J., Dickens,M. and Davis,R.J. (1997)
Nuclear accumulation of NFAT4 opposed by the JNK signal
transduction pathway. Science, 278, 1638–1641.
Cuenda,A., Cohen,P., Buee-Scherrer,V. and Goedert,M. (1997) Activation
of stress-activated protein kinase-3 (SAPK3) by cytokines and cellular
stresses is mediated via SAPKK3 (MKK6); comparison of the
specificities of SAPK3 and SAPK2 (RK/p38). EMBO J., 16, 295–305.
Dai,T., Rubie,E., Franklin,C.C., Kraft,A., Gillespie,D.A.F., Avruch,J.,
Kyriakis,J.M. and Woodgett,J.R. (1995) Stress-activated protein
kinases bind directly to the δ domain of c-Jun in resting cells:
implications for repression of c-jun function. Oncogene, 10, 849–855.
De
´
rijard,B., Hibi,M., Wu,I-H., Barrett,T., Su,B., Deng,T., Karin,M. and
Davis,R.J. (1994) JNK1: a protein kinase stimulated by UV light and
Ha-Ras that binds and phosphorylates the c-jun activation domain.
Cell, 76, 1025–1037.
Dickens,M., Rogers,J., Cavanagh,J., Raitano,A., Xia,Z., Halpern,J.R.,
Greenberg,M.E., Sawyers,C. and Davis,R.J. (1997) A cytoplasmic
inhibitor of the JNK signal transduction pathway. Science, 277,
693–696.
Enslen,H., Raingeaud,J. and Davis,R.J. (1998) Selective activation of
p38 MAP kinase isoforms by the MAP kinase kinases MKK3 and
MKK6. J. Biol. Chem., 273, 1741–1748.
Gille,H., Strahl,T. and Shaw,P.E. (1995) Activation of ternary complex
factor Elk-1 by stress-activated protein kinases. Curr. Biol., 5, 1191–
1200.
Goedert,M., Cuenda,A., Craxton,M., Jakes,R. and Cohen,P. (1997)
Activation of the novel stress-activated protein kinase SAPK4 by
cytokines and cellular stresses is mediated by SKK3 (MKK6);
comparison of its substrate specificity with that of other SAP kinases.
EMBO J., 16, 3563–3571.
Gupta,S. and Davis,R.J. (1994) MAP kinase binds to the NH
2
-terminal
activation domain of c-Myc. FEBS Lett., 353, 281–285.
Gupta,S., Campbell,D., De
´
rijard,B. and Davis,R.J. (1995) Transcription
factor ATF2 regulation by the JNK signal transduction pathway.
Science, 267, 389–393.
Gupta,S., Barrett,T., Whitmarsh,A.J., Cavanagh,J., Sluss,H.K.,
De
´
rijard,B. and Davis,R.J. (1996) Selective interaction of JNK protein
kinase isoforms with transcription factors. EMBO J., 15, 2760–2770.
Han,J., Jiang,Y., Li.,Z., Kravchenko,V.V. and Ulevitch,R.J. (1997)
Activation of the transcription factor MEF2C by the MAP kinase p38
in inflammation. Nature, 386, 296–299.
Hibi,M., Lin,A., Smeal,T., Minden,A. and Karin,M. (1993) Identification
of an oncoprotein- and UV-responsive protein kinase that binds and
potentiates the c-Jun activation domain. Genes Dev., 7, 2135–2148.
Janknecht,R. and Hunter,T. (1997a) Convergence of MAP kinase
pathways on the ternary complex factor SAP-1a. EMBO J., 16,
1620–1627.
Janknecht,R. and Hunter,T. (1997b) Activation of the SAP-1a
transcription factor by the c-Jun N-terminal kinase (JNK) mitogen-
activated protein kinase. J. Biol. Chem., 272, 4219–4224.
Jiang,Y., Chen,C., Li,Z., Guo,W., Gegner,J.A., Lin,S. and Han,J. (1996)
Characterization of the structure and function of a new mitogen-
activated protein kinase (p38) J. Biol. Chem., 271, 17920–17926.
Kallunki,T., Su,B. Tsigelny,I., Sluss,H.K., De
´
rijard,B., Moore,G.,
Davis,R.J. and Karin,M. (1994) JNK2 contains a specificity-
determining region responsible for efficient c-jun binding and
phosphorylation. Genes Dev., 8, 2996–3007.
Kallunki,T., Deng,T., Hibi,M. and Karin,M. (1996) c-Jun can recruit
JNK to phosphorylate dimerization partners via specific docking
interactions. Cell, 87, 929–939.
Kyriakis,J.M., BanerjeeP., Nikolakaki,E., Dai,T., Rubie,E.A.,
Ahmad,M.F., Avruch,J. and Woodgett,J.R. (1994) The stress-activated
protein kinase subfamily of c-Jun kinases. Nature, 369, 156–160.
Lechner,C., Zahalka,M.A., Giot,J.-F., Møller,N.P.H. and Ullrich,A.
(1996) ERK6, a mitogen-activated protein kinase involved in C2C12
myoblast differentiation. Proc. Natl Acad. Sci. USA, 93, 4355–4359.
Livingstone,C., Patel,G., and Jones,N. (1995) ATF-2 contains a
phosphorylation-dependent transcriptional activation domain. EMBO
J., 14, 1785–1797.
1749
Mao,C., Ray-Gallet,D., Tavitian,A. and Moreau-Gachelin,F. (1996)
Differential phosphorylations of Spi-B and Spi-1 transcription factors.
Oncogene, 12, 863–873.
Price,M.A., Cruzalegui,F.H. and Treisman,R. (1996) The p38 and ERK
MAP kinase pathways co-operate to activate ternary complex factors
and c-fos transcription in response to UV light. EMBO J., 15,
6552–6563.
Raingeaud,J., Gupta,S., Rogers,J.S., Dickens,M., Han,J., Ulevitch,R.J.
and Davis,R.J. (1995) Pro-inflammatory cytokines and environmental
stress cause p38 mitogen-activated protein kinase activation by dual
phosphorylation on tyrosine and threonine. J. Biol. Chem., 270,
7420–7426.
Robinson,M.J. and Cobb,M.H. (1997) Mitogen-activated protein kinase
pathways. Curr. Opin. Cell Biol., 9, 180–186.
Sadowski,I. and Ptashne,M. (1989) A vector for expressing GAL4(1–
147) fusions in mammalian cells. Nucleic Acids Res., 17, 7539.
Seth,A., Gonzalez,F.A., Gupta,S., Raden,D.L. and Davis,R.J. (1992)
Signal transduction within the nucleus by mitogen-activated protein
kinase. J. Biol. Chem., 267, 24796–24804.
Shore,P. and Sharrocks,A.D. (1994) The transcription factors Elk-1 and
serum response factor interact by direct protein–protein contacts
mediated by a short region of Elk-1. Mol. Cell. Biol., 14, 3283–3291.
Shore,P., Bisset,L., Lakey,J., Waltho,J.P., Virden,R. and Sharrocks,A.D.
(1995) Characterization of the Elk-1 ETS DNA-binding domain.
J. Biol. Chem., 270, 5805–5811.
Shore,P., Whitmarsh,A.J., Bhaskaran,R., Waltho,J., Davis,R.J. and
Sharrocks,A.D. (1996) Determinants of DNA binding specificity of
ETS-domain transcription factors. Mol. Cell. Biol., 16, 3338–3349.
Sluss,H.K., Barrett,T., Derijard,B. and Davis,R.J. (1994) Signal
transduction by tumor necrosis factor mediated by JNK protein
kinases. Mol. Cell. Biol., 14, 8376–8384.
Stein,B., Yang,M.X., Young,D.B., Janknecht,R., Hunter,T., Murray,B.W.
and Barbosa,M.S. (1997) p38-2, a novel mitogen-activated protein
kinase with distinct properties. J. Biol. Chem., 272, 19509–19517.
Strahl,T., Gille,H. and Shaw,P.E. (1996) Selective response of ternary
complex factor Sap-1a to different mitogen-activated protein kinase
sub-groups. Proc. Natl Acad. Sci. USA, 93, 11563–11568.
Treisman,R. (1994) Ternary complex factors: growth regulated
transcriptional activators. Curr. Opin. Genet. Dev., 4, 96–101.
Treisman,R. (1996) Regulation of transcription by MAP kinase cascades.
Curr. Opin. Cell Biol., 8, 205–215.
Van Dam,H., Wilhelm,D., Herr,I., Steffen,A., Herrlich,P. and Angel,P.
(1995) ATF-2 is preferentially activated by stress-activated protein
kinases to mediate c-jun induction in response to genotoxic agents.
EMBO J., 14, 1798–1811.
Wang,X. and Ron,D. (1996) Stress-induced phosphorylation and
activation of the transcription factor CHOP (GADD153) by p38 MAP
kinase. Science, 272, 1347–1349.
Wang,X.S. et al. (1997) Molecular cloning and characterization of a
novel p38 mitogen-activated protein kinase. J. Biol. Chem., 272,
23668–23674.
Waskiewicz,A.J., Flynn,A., Proud,C.G. and Cooper,J.A. (1997) Mitogen-
activated protein kinases activate the serine-threonine kinases Mnk1
and Mnk2. EMBO J., 16, 1909–1920.
Whitmarsh,A.J. and Davis,R.J. (1996) Transcription factor AP-1
regulation by mitogen-activated protein kinase signal transduction
pathways. J. Mol. Med., 74, 589–607.
Whitmarsh,A.J., Shore,P., Sharrocks,A.D. and Davis,R.J. (1995)
Integration of MAP kinase signal transduction pathways at the serum
response element. Science, 269, 403–407.
Whitmarsh,A.J., Yang,S.-H., Su,M.S.-S., Sharrocks,A.D. and Davis,R.J.
(1997) Role of p38 and JNK MAP kinases in the activation of ternary
complex factors. Mol. Cell. Biol., 17, 2360–2371.
Yang,S.-H., Whitmarsh,A.J., Davis,R.J. and Sharrocks,A.D. (1998) The
Elk-1 ETS-domain transcription factor containsaMAP kinase targeting
motif. Mol. Cell. Biol., 18, 710–720.
Received November 25, 1997; revised January 19, 1998;
accepted January 20, 1998