ArticlePDF Available

Feedback inhibition by RALT controls signal output by the ErbB network

Authors:

Abstract and Figures

The ErbB-2 interacting protein receptor-associated late transducer (RALT) was previously identified as a feedback inhibitor of ErbB-2 mitogenic signals. We now report that RALT binds to ligand-activated epidermal growth factor receptor (EGFR), ErbB-4 and ErbB-2.ErbB-3 dimers. When ectopically expressed in 32D cells reconstituted with the above ErbB receptor tyrosine kinases (RTKs) RALT behaved as a pan-ErbB inhibitor. Importantly, when tested in either cell proliferation assays or biochemical experiments measuring activation of ERK and AKT, RALT affected the signalling activity of distinct ErbB dimers with different relative potencies. RALT deltaEBR, a mutant unable to bind to ErbB RTKs, did not inhibit ErbB-dependent activation of ERK and AKT, consistent with RALT exerting its suppressive activity towards these pathways at a receptor-proximal level. Remarkably, RALT deltaEBR retained the ability to suppress largely the proliferative activity of ErbB-2.ErbB-3 dimers over a wide range of ligand concentrations, indicating that RALT can intercept ErbB-2.ErbB-3 mitogenic signals also at a receptor-distal level. A suppressive function of RALT deltaEBR towards the mitogenic activity of EGFR and ErbB-4 was detected at low levels of receptor occupancy, but was completely overcome by saturating concentrations of ligand. We propose that quantitative and qualitative aspects of RALT signalling concur in defining identity, strength and duration of signals generated by the ErbB network.
In vivo analysis of RALT interaction with EGFR and ErbB-4. (a) NIH 3T3 and NIH-EGFR cells were infected with recombinant retrovirus directing the expression of either RALT or RALT EBR. Control cells were infected with empty virus. After 2 h serum starvation, cells were stimulated for 5 min with either carrier solution or EGF (10 ng/ml) prior to lysis. Equal amounts of each lysate were analysed by immunoblot with anti-receptor and anti-P-Tyr antibodies. Lysates were subjected to immunoprecipitation with anti-RALT 19C5/4 monoclonal antibody followed by blot analysis with anti-RALT and anti-receptor antibodies. (b) ErbB-4 was transiently expressed in 293 cells along with either RALT or RALT EBR. At 48 h after transfection, cells were lysed after stimulation for 5 min with either carrier solution or 20 ng/ml NRG-1. Equal amounts of each lysate were analysed by immunoblot with anti-ErbB-4 and anti-P-Tyr antibodies. Anti-RALT 19C5/4 immunoprecipitates were resolved on SDS–PAGE and immunoblotted with anti-RALT and anti-ErbB-4 antibodies. (c) RALT was coexpressed in 293 cells with either wt EGFR or EGFR mutants depicted in the left portion of the panel, namely EGFR K721A (KD, i.e. kinase-defective) and EGFR C, that lacks the COOH-terminal region. At 48 h after transfection, cells were stimulated for 5 min with either carrier solution or 10 ng/ml EGF. Equal amounts of each lysate were analysed by immunoblot with anti-RALT and anti-P-Tyr antibodies. Anti-EGFR immunoprecipitates were resolved on SDS–PAGE and immunoblotted with anti-RALT and anti-EGFR antibodies
… 
Content may be subject to copyright.
Feedback inhibition by RALT controls signal output by the ErbB network
Sergio Anastasi
1,6
, Loredana Fiorentino
1,6,7
, Monia Fiorini
1
, Rocco Fraioli
1
, Gianluca Sala
1
,
L Castellani
2,3
, Stefano Alema
`
4
, Maurizio Alimandi
5
and Oreste Segatto
n
,1
1
Regina Elena Cancer Institute, Via Delle Messi d’Oro, 156, Rome 00158, Italy;
2
University of Cassino, Cassino, Italy;
3
INFM Sez.
B, University of Rome Tor Vergata, Rome, Italy;
4
Istituto di Biologia Cellulare, CNR Monterotondo 00016, Italy;
5
Department of
Experimental Medicine, University of Rome La Sapienza, Viale Regina Elena 324, Rome 00161, Italy
The ErbB-2 interacting protein receptor-associated late
transducer (RALT) was previously identified as a feed-
back inhibitor of ErbB-2 mitogenic signals. We now
report that RALT binds to ligand-activated epidermal
growth factor receptor (EGFR), ErbB-4 and ErbB-
2.ErbB-3 dimers. When ectopically expressed in 32D
cells reconstituted with the above ErbB receptor tyrosine
kinases (RTKs) RALT behaved as a pan-ErbB inhibitor.
Importantly, when tested in either cell proliferation assays
or biochemical experiments measuring activation of ERK
and AKT, RALT affected the signalling activity of
distinct ErbB dimers with different relative potencies.
RALT DEBR, a mutant unable to bind to ErbB RTKs, did
not inhibit ErbB-dependent activation of ERK and AKT,
consistent with RALT exerting its suppressive activity
towards these pathways at a receptor-proximal level.
Remarkably, RALT DEBR retained the ability to
suppress largely the proliferative activity of ErbB-
2.ErbB-3 dimers over a wide range of ligand concentra-
tions, indicating that RALT can intercept ErbB-2.ErbB-3
mitogenic signals also at a receptor-distal level. A
suppressive function of RALT DEBR towards the
mitogenic activity of EGFR and ErbB-4 was detected at
low levels of receptor occupancy, but was completely
overcome by saturating concentrations of ligand. We
propose that quantitative and qualitative aspects of RALT
signalling concur in defining identity, strength and
duration of signals generated by the ErbB network.
Oncogene (2003) 22, 4221–4234. doi:10.1038/sj.onc.1206516
Keywords: ErbB; feedback inhibition; Mig-6; negative
signaling; RALT
Introduction
The family of ErbB receptor tyrosine kinases (RTKs)
consists of four members: ErbB-1 (also referred to as
epidermal growth factor receptor, EGFR) through
ErbB-4. In the human genome, a dozen 10 genes encode
ErbB ligands possessing distinct receptor specificity.
Regardless of the latter, all ErbB ligands trigger cognate
receptors via a common mechanism entailing the
assembly of homo- and heterodimeric receptor combi-
nations (reviewed in Olayioye et al., 2000; Yarden and
Sliwkowski, 2001). The ensuing combinatorial reper-
toire of ligand-driven receptor dimers provides the
family of ErbB RTKs with a great deal of signalling
versatility and, consequently, the ability to govern a
wide array of cellular programs. This is best exemplified
by genetic studies, which have shown that defined
combinations of ErbB receptors play nonredundant
roles during development and in postnatal life. In
human pathology, this is mirrored by the selective
derangement of EGFR and ErbB-2 activity during the
pathogenesis of a number of tumors (reviewed in
Olayioye et al., 2000; Yarden and Sliwkowski, 2001).
ErbB-1 and ErbB-4 share canonical features of RTKs
and can generate downstream signals upon ligand-
induced homodimerization. In contrast, ErbB-2 and
ErbB-3 are peculiar in that ligand-driven lateral inter-
actions with other ErbB RTKs represent a sine qua non
for initiation of their signalling activity (reviewed in
Olayioye et al., 2000; Yarden and Sliwkowski, 2001).
ErbB-3 is able to bind ligands, such as neuregulins
(NRGs), directly (Carraway III et al., 1994, 1997), but is
devoid of intrinsic catalytic activity (Guy et al., 1994).
Therefore, ErbB-3 acquires signalling capacity only in
the context of heterodimers with catalytically competent
ErbB RTKs (Pinkas-Kramarski et al., 1996). Conver-
sely, ErbB-2 is unable to bind ligands directly, but
behaves as a coreceptor for several ligands when
complexed with other ErbB RTKs (Graus-Porta et al.,
1995, 1997; Riese et al., 1995; Karunagaran et al., 1996;
Riese and Stern, 1998). In fact, ErbB-2 appears to be the
favorite partner of all other ErbB receptors and emerges
as the hierarchically dominant element in the assembly
of ErbB heterodimers (Graus-Porta et al., 1997). Of
particular relevance is the ErbB-2/ErbB-3 combination
that reconstitutes a high-affinity receptor for NRGs
endowed with potent oncogenic activity (Alimandi et al.,
1995; Wallasch et al., 1995).
Individual ErbB receptors may couple selectively to
some signalling pathways: for example, only EGFR is
Received 1 October 2002; revised 19 February 2003; accepted 20
February 2003
*Correspondence: O Segatto; E-mail: segatto@ifo.it
6
These two authors contributed equally to this work
7
Current address: The Burnham Institute, 10901 N Torrey Pines Rd,
La Jolla, CA 92037, USA
Oncogene (2003) 22, 4221–4234
&
2003 Nature Publishing Group
All rights reserved 0950-9232/03 $25.00
www.nature.com/onc
able to activate STAT1 and STAT3 (Olayioye et al.,
1999). Combinatorial dimerization among ErbB RTKs
generates additional signalling potential via integration
of the signalling competence of individual protomers as
well as the generation of novel signalling profiles
(Olayioye et al., 1998). Finally, ErbB RTKs share some
signalling pathways, such as those leading to Ras–ERK
and PI-3K–AKT activation. Importantly, distinct ErbB
dimers may activate these pathways with diverse profiles
of intensity and duration (Olayioye et al., 2000 and
references therein), which may be themselves critical
determinants of signal specificity (Marshall, 1995;
Ghiglione et al., 1999). This raises the question of how
strength and duration of ErbB signals are proofread in
order to (a) allow the correct execution of cellular
programs governed by precise thresholds of receptor
signals and (b) restrain the oncogenic potential of
ErbB activity.
It is increasingly appreciated that negative signalling
is instrumental in defining quantitative aspects of ErbB
outputs. Coupling to internalization/degradation path-
ways is thought to set thresholds of activity of ErbB
RTKs at the inception of receptor triggering (reviewed
in Waterman and Yarden, 2001). Much less studied are
the mechanisms that gauge and eventually extinguish
ErbB signals as they accumulate over time because of
sustained receptor activity. Genetic studies in flies
indicate that expression of Drosophila EGFR (DER)
feedback inhibitors such as Kekkon1,Argos and Sprouty
is controlled transcriptionally by DER via the Ras-ERK
pathway (Perrimon and McMahon, 1999). Importantly,
negative feedback regulation of DER signalling is
critical for ensuring that DER activity be confined
within appropriate spatial and temporal boundaries
(Perrimon and McMahon, 1999; Freeman, 2000). Thus,
it appears that the levels of expression of Aos, Spry and
Kek1 are a means for the cell to gauge the accumulation
of receptor signals, while their activity is uniquely suited
to proofread, tune and extinguish DER activity.
Consequently, it has been proposed that negative
feedback loops provide DER signalling with an essential
element of stability, which prevents perturbation of
DER function and allows for accurate reproduction of
signals responsible for complex developmental patterns
(Freeman, 2000).
Receptor-associated late transducer (RALT) has been
recently identified as a feedback inhibitor of mitogenic
signals propagated by ErbB-2 (Fiorentino et al., 2000)
and EGFR (Hackel et al., 2001). RALT expression is
activated by EGF, TGF-aand NRG-1 in fibroblasty
and epithelial cells (Fiorentino et al., 2000), via a
mechanism entailing transcriptional activation of the
Ralt gene by the Ras–Raf–ERK pathway (Fiorini et al.,
2002). Constitutive overexpression of RALT markedly
inhibits the mitogenic and transforming activity of
EGFR (Hackel et al., 2001) and ErbB-2 (Fiorentino
et al., 2000). Conversely, neutralization of endogenous
RALT function by microinjection of anti-RALT anti-
bodies enhances the mitogenic activity of an EGFR/
ErbB-2 chimeric receptor (Fiorini et al., 2002). These
data indicate that physiological levels of RALT are
involved in the suppression of ErbB-2 signals and imply
that RALT function may in fact play a necessary
role in negative regulation of ErbB-2 activity. Owing to
its ability to complex with ErbB-2- and SH3-containing
proteins, RALT has been proposed to function as a
receptor-proximal inhibitor, which links ErbB-2 to
effectors containing SH3 domains (Fiorentino et al.,
2000). However, the significance of the RALT/ErbB-2
physical interaction in the context of RALT-
mediated inhibition of ErbB-2 signalling has not been
formally addressed, nor is it known whether and to what
extent RALT is able to antagonize mitogenic signals
propagated by the entire ErbB signalling network.
The latter is an especially relevant issue, given the
high degree of signal integration among ErbB
RTKs.
Here, we report that RALT functions as a pan-ErbB
inhibitor. Remarkably, the potency of RALT suppres-
sive activity impacts differentially on individual ErbB
receptors. We propose that RALT activity adds a novel
layer of regulation to the specification of signal output
by the ErbB network.
Results
RALT is not a promiscuous RTK inhibitor
Despite its suppressive activity towards mitogenic and
transforming signals triggered by ErbB-2 and EGFR,
ectopically expressed RALT altered neither mitogenic
responses of murine fibroblasts to serum nor cell
transformation induced by v-ras and v-src (Fiorentino
et al., 2000; Hackel et al., 2001). The issue of RALT
specificity was further addressed in the experiments
shown in Figure 1. NIH-EGFR/ErbB-2 cells were
infected with either Pinco or Pinco–RALT retroviral
stocks. RALT expression was induced in Pinco
cells by EGF (Figure 1a) as well as by bFGF and
PDGF (data not shown, see also Fiorini et al., 2002)
with comparable kinetics. Pinco–RALT cells showed
constitutive expression of ectopic RALT at levels 3–4-
fold higher than the endogenous species (Figure 1a).
Whereas ErbB-2-dependent mitogenic signals were
clearly suppressed by RALT overexpression (42–52%
reduction over a 0.3–5 ng ml EGF concentration range),
proliferative responses to PDGF and bFGF were
comparable in Pinco and Pinco–RALT cells over a
wide range of ligand concentration (Figure 1b); also
proliferative responses to PMA were unaffected by
RALT overexpression (data not shown). These results
indicate that RALT is unlikely to behave as a pan-RTK
inhibitor.
RALT binds in vitro to all members of the ErbB family
Based on the above conclusions, we focused our
attention on the function of RALT within the ErbB
signalling network. Thus, we asked whether RALT
could enter in a complex with ErbB RTKs other than
ErbB-2.
Negative signalling of RALT to the ErbB network
S Anastasi el al
4222
Oncogene
Previous studies have shown that the ErbB-2-binding
activity of RALT is confined to the RALT COOH
terminus (residues 262–459); within this 197 amino-acid-
long stretch, the sequence comprised between positions
282 and 396 (GST-RALT cl.52) was sufficient to bind
ErbB-2 in GST pull down assays and in yeast two-
hybrid experiments (Fiorentino et al., 2000) (see also
Figures 2a and 3). GST-RALT cl.52 was able to
precipitate ErbB-2 and ErbB-3 from lysates of NIH-
E2.E3 transfectants; ErbB-3 could be precipitated only
from lysates of NRG-1-stimulated cells, whereas ErbB-2
bound to GST-RALT cl.52 irrespective of NRG-1
stimulation (Figure 2a). In the absence of ErbB-2
coexpression, ErbB-3 was not precipitated by GST-
RALT cl.52, even under conditions of optimal stimula-
tion with NRG-1. On the contrary, ErbB-2 bound
efficiently to GST-RALT also in the absence of
ectopically expressed ErbB-3. The ligand-independent
binding of ErbB-2 to GST-RALT cl.52 is most likely a
consequence of the constitutive activation of ErbB-2 in
NIH 3T3 transfectants, as reported previously (Lonardo
et al., 1990) and demonstrated herein by the anti-P-Tyr
immunoblot shown in Figure 2a. Binding to GST-
RALT cl.52 was also observed with EGFR (Figure 2b)
and ErbB-4 (Figure 2c) solubilized from NIH-E1 and
NIH-E4 transfectants, respectively. Binding of ErbB-4
to GST–RALT cl.52 was unaffected by EGF stimula-
tion but enhanced by NRG-1 stimulation. Conversely,
EGF but not NRG-1 stimulation enhanced binding of
EGFR to GST–RALT cl.52.
Figure 1 Biological analysis of RALT signalling in mitogenic
programs activated by PDGF and bFGF. NIH-EGFR/ErbB-2
cells were seeded in 24-well plates and infected with either Pinco or
Pinco–RALT retrovirus stocks. Infection efficiency, as monitored
by FACS assessment of GFP-positive cells was 495%. (a)
Following infection, cells were made quiescent by 24 h serum
deprivation and subsequently challenged with 5 ng/ml EGF for the
indicated time. Cell lysates were subjected to immunoblot analysis
with either the S1 antiserum that recognizes mouse, rat and human
RALT species or MoAb 19C5/4 specific for the ectopically
expressed rat RALT species. (b) Following infection, cells were
switched to either serum-free medium (SFM) or SFM containing
escalating concentrations of EGF (0.3, 1 and 5 ng/ml), bFGF (0.3,
1 and 5 ng/ml) or PDGF (0.2, 2 and 10 ng/ml). Assays were
performed in quadruplicate wells. After 44 h, cultures were labelled
for 4 h with 1 mCi/ml [
3
H]methyl-thymidine and cell proliferation
measured as fold increase of TCA-precipitated radioactivity in
growth factor-treated wells versus control cultures. The s.d. was
o15%. Data are representative of one out of three experiments
that gave similar results. White bars indicate Pinco cells, gray bars
indicate Pinco–RALT cells
Figure 2 In vitro analysis of RALT interaction with ErbB RTKs.
(a) Recombinant GST-RALT cl.52 fusion protein (referred to as
GST–RALT) spanning positions 282–396 of RALT was purified
onto glutathione–sepharose beads and used in pull-down assays
against ErbB receptors solubilized from the indicated NIH 3T3
transfectants. Lysates of parental NIH 3T3 were used as control.
Where indicated, cells were treated with 20 ng/ml recombinant
NRG-1bfor 5 min prior to lysis. Protein complexes were subjected
to immunoblot analysis with the indicated antibodies. The bottom
portion of the gel was stained with Coomassie blue to control for
equal input of GST–RALT. Reference lysates (immunoblotted
with anti-ErbB-3 and anti-P-Tyr antibodies) corresponded to 5%
of input lysate in GST pull-down experiments. Control experiments
(data not shown, see also Figure 3b) indicated that GST alone did
not bind to ErbB RTKs. (band c) NIH-EGFR (b) and NIH-ErbB-
4(c) transfectants were serum starved for 16 h and lysed either
without further treatment or after stimulation for 5 min with 20 ng/
ml of the indicated ligands. Lysates were subjected to pull-down
assays with GST–RALT and immunoblotted with anti-EGFR (b)
and anti-ErbB-4 antibodies (c). Reference lysates corresponded to
5% of input lysate in GST pull-down experiments. Input GST–
RALT was visualized by Coomassie stain
Negative signalling of RALT to the ErbB network
S Anastasi el al
4223
Oncogene
Identification of the minimal ErbB-2-binding region of
RALT
In order to define the minimal ErbB-2-binding region
(EBR) of RALT, we generated a panel of GST–RALT
fusion proteins (Figure 3a) and assayed their ability to
precipitate ErbB-2 from lysates of NIH-ErbB-2 trans-
fectants. Neither the NH2 (aa. 282–343) nor the COOH
(aa. 335–396) halves of GST–RALT cl.52 were capable
of binding to ErbB-2 (data not shown), whereas
sequences comprised between positions 282 and 312
were dispensable for directing the interaction of GST
RALT cl.52 with ErbB-2 (Figure 3b). Further deletion
analysis defined the sequence comprised between posi-
tions 323 and 372 of RALT as the minimal EBR
(Figure 3b). We took advantage of a conveniently
located StuI site in the Ralt cDNA sequence to generate
a RALT mutant lacking residues 315–361 (RALT
DEBR, Figure 3a), which could be expressed in various
cell lines as efficiently as full-length RALT (see below).
Analysis of RALT interaction with ErbB receptors in
living cells
We next asked whether RALT interacts with ErbB
RTKs in living cells in an EBR-dependent fashion. To
address these questions, we performed coimmunopreci-
pitation assays using lysates of either 293 or NIH 3T3
cells engineered to express ErbB RTKs along with either
RALT or RALT DEBR (Figures 4 and 5). Neither
RALT nor RALT DEBR had detrimental consequences
on steady-state levels of coexpressed ErbB receptors
(Figures 4 and 5). Importantly, the DEBR mutant was
able to interact with Grb-2 as efficiently as RALT
(Figure 4b). We infer that the DEBR mutation is
unlikely to cause gross conformational alterations of
the RALT protein, since the RALT/Grb-2 interaction
requires that Grb-2 SH3 domains bind to Pro-rich
sequences located in close proximity to the RALT EBR
module (Fiorentino et al., 2000).
RALT, but not RALT DEBR, was found in a
complex with ErbB-2 in 293 cells, as revealed by
probing anti-RALT immunoprecipitates with anti-
ErbB-2 antibodies. This immunoreactivity was contin-
gent on coexpression of RALT and ErbB-2 (Figure 4a).
ErbB-3 could be coimmunoprecipitated with anti-
RALT antibodies only if coexpressed with ErbB-2 and
solubilized from cells stimulated with NRG-1. Coim-
munoprecipitation of RALT and ErbB-3 required
integrity of the EBR (Figure 4a). Thus, it appears that
ErbB-3 coimmunoprecipitates with RALT only when
recruited to ErbB-2.ErbB-3 dimers generated by NRG-1
stimulation. Interestingly, in E2.E3/RALT transfectants
the amount of ErbB-2 immunoreactivity recovered in
anti-RALT immunoprecipitates from NRG-stimulated
cells was consistently lower than that obtained from
unstimulated cells (Figure 4a). We infer that in
unstimulated ErbB-2.ErbB-3 transfectants, RALT is
likely to be complexed to activated ErbB-2 homodimers
generated by ErbB-2 overexpression. NRG-1 stimula-
tion drives redistribution of activated ErbB-2 molecules
into complexes with ErbB-3, thus causing preferential
association of RALT with ErbB-2.ErbB-3 heterodimers
rather than with ErbB-2 homodimers.
EGFR (Figure 5a, c) and ErbB-4 (Figure 5b) were
also able to associate with RALT. These interactions
were largely dependent on ligand stimulation of the
receptors and absolutely contingent on the presence of
the EBR module in the RALT protein. A catalytically
inactive EGFR mutant carrying the K-A mutation at
position 721 was unable to recruit RALT in a physical
complex both in anti-EGFR immunoprecipitations
followed by anti-RALT immunoblot (Figure 5c), as
well as in anti-RALT immunoprecipitations followed by
anti-EGFR immunoblot (data not shown). Moreover,
we did not detect RALT/EGFR K721A complexes even
upon marked receptor overexpression, that is, under
conditions causing constitutive, ligand-indepen-
dent RALT/EGFR association (data not shown). In
Figure 3 Mapping of the minimal EBR in the RALT protein. (a,
top) Schematic outline of domain organization of RALT and
RALT DEBR mutant. NLS indicates a putative nuclear localiza-
tion signal. (Bottom) Schematic representation of RALT fragments
used as GST fusion products in pull-down assays reported in panel
b. The ability of each recombinant polypeptide to interact in vitro
with the ErbB-2 protein is also indicated. Note that RALT
fragments are drawn to a scale larger than the one used to draw
RALT and RALT DEBR in the top portion of the panel. (b) GST
fusion proteins (5 mg for each pull-down experiment) spanning the
indicated portions of the RALT protein were produced in E. coli,
purified onto glutathione–agarose beads and assayed for their
ability to capture ErbB-2 protein solubilized from NIH-ErbB-2
transfectants. Protein complexes were analysed by immunoblot
with anti-ErbB-2 antibodies. GST was used as control; reference
lysate corresponded to 5% of input lysate in pull-down assays
Negative signalling of RALT to the ErbB network
S Anastasi el al
4224
Oncogene
contrast, the deletion of the EGFR COOH tail, which
contains all the major autophosphorylation sites of the
receptor (DC mutant), did not affect RALT/EGFR
complex formation (Figure 5c).
Collectively, the experiments described in Figures 3–5
indicate that RALT interacts with different ErbB RTKs
via the EBR module. Importantly, when comparing the
amount of ErbB receptors precipitated by either GST–
RALT or anti-RALT antibodies to the amount of ErbB
receptors present in 5% of input lysate, we consistently
observed that RALT interacts with EGFR, ErbB-4 and
ErbB-2.ErbB-3 dimers with comparable efficiencies.
RALT is spatially controlled by its interaction with
ligand-activated ErbB RTKs
EGF-dependent activation of an EGFR/ErbB-2 chi-
meric receptor leads to the redistribution of RALT from
a cytosolic location to intracellular membranes (Fior-
entino et al., 2000). The inability of the DEBR mutant to
associate with ErbB RTKs allowed us to test whether
physical interaction with ErbB receptors is implicated in
spatial control of the RALT protein. MoAb 19C5/4
allows immunohistochemical detection of RALT and
RALT DEBR with similar efficiency. Since Ab 19C5/4
does not crossreact with murine RALT (Fiorentino
et al., 2000, see also Figure 1a), we could selectively
image ectopic RALT proteins in NIH-EGFR and NIH-
EGFR/ErbB-2 cells. Of note, the degree of RALT
overexpression in these cells was modest (3–4-fold
higher than that of the endogenous species, see
Figure 1a for reference). Immunofluorescence imaging
of RALT in quiescent cells indicated that both RALT
(Figure 6e, m) and RALT DEBR (Figure 6g, o) were
distributed in the cytosol. RALT relocated to a vesicular
compartment following activation of EGFR for 15 min
(Figure 6n) and EGFR/ErbB-2 for 15 min (not shown)
or 60 min (Figure 6f), showing extensive colocalization
with ErbB receptors (Figure 6j). Strikingly, ligand
activation of neither EGFR nor EGFR/ErbB-2 caused
relocation of RALT DEBR (Figure 6h, p). We conclude
that RALT relocation induced by activation of ErbB
RTKs requires that RALT forms a physical complex
with ligand-bound ErbB RTKs. We also infer that the
DEBR mutation does not abolish coimmunoprecipita-
tion between RALT and ErbB receptors by simply
causing a reduction of the affinity of RALT DEBR for
ErbB RTKs.
RALT inhibits ErbB signals leading to activation of ERKs
and AKT
ErbB family members share the ability to trigger the
Ras-ERK and the PI-3-AKT pathways (Yarden and
Sliwkowski, 2001). Thus, we surmised that evaluation of
ERK and AKT activities triggered by engagement of
ErbB RTKs could provide a convenient biochemical
readout for comparative analysis of RALT signalling to
distinct ErbB dimers.
For these studies we used 32D cells reconstituted with
defined ErbB RTKs. 32D.ErbB transfectants are re-
lieved from IL-3 dependency if cultured in media
containing ErbB ligands, thus providing a sensitive
assay of the signalling activity of defined ErbB dimers
(Alimandi et al., 1997). 32D.ErbB transfectants were
further engineered to express either RALT or RALT
DEBR. Of note, 32D cells do not express endogenous
RALT protein, as indicated by immunoblot analysis of
lysates prepared from 32D and 32D.E1 cells triggered
with either IL-3 (growth medium) or EGF (Figure 7c).
Figure 4 In vivo analysis of physical interaction between RALT
and the ErbB-2.ErbB-3 NRG-1 receptor. (a) ErbB-2 and ErbB-3
were transiently expressed in 293 cells either alone or in
combination. Where indicated, either RALT or RALT DEBR
were coexpressed with ErbB RTKs. After 2 h serum starvation,
cells were stimulated at 371C with either carrier solution or 20 ng/
ml NRG-1bbefore lysis. Equal amounts of each lysate were
analysed by immunoblot with anti-ErbB-2 and anti-ErbB-3
antibodies to control for receptor expression. Lysates were
subjected to immunoprecipitation with anti-RALT 19C5/4 mono-
clonal antibody followed by blot analysis with anti-RALT and
anti-ErbB-3 antibodies. Filters were subsequently reprobed with
anti-ErbB-2 antibodies. (b) The DEBR mutation does not affect the
ability of RALT to interact with Grb-2. HA-tagged GRB-2 was
expressed in 293 cells either alone or in combination with RALT or
RALT DEBR. Equal amounts of each lysate and anti-RALT
immunoprecipitates were immunoblotted with anti-HA tag 12CA/5
and anti-RALT 19C5/4 monoclonal antibodies
Negative signalling of RALT to the ErbB network
S Anastasi el al
4225
Oncogene
This was confirmed by Northern blot analysis (data not
shown), in agreement with the absence of Mig-6 mRNA
expression in cells of the hemopoietic lineage reported in
the Genecard database (http://bioinfo.weizmann.ac.il/
cards/). Thus, by expressing RALT proteins in
32D.ErbB transfectants we could (a) assess whether
reconstitution of RALT expression is per se capable of
restoring a pathway of negative signalling to ErbB
RTKs and (b) study the activity of the DEBR mutant
without the possible interference of endogenous RALT
protein.
Stable RALT and RALT DEBR derivatives of
32D.E1, 32D.E2.E3 and 32D.E4 cells were generated
by transfection with Pinco–RALT and Pinco–RALT
DEBR vectors, respectively. Control cells were drug-
selected after transfection with the backbone Pinco
Figure 5 In vivo analysis of RALT interaction with EGFR and ErbB-4. (a) NIH 3T3 and NIH-EGFR cells were infected with
recombinant retrovirus directing the expression of either RALT or RALT DEBR. Control cells were infected with empty virus. After
2 h serum starvation, cells were stimulated for 5 min with either carrier solution or EGF (10 ng/ml) prior to lysis. Equal amounts of
each lysate were analysed by immunoblot with anti-receptor and anti-P-Tyr antibodies. Lysates were subjected to immunoprecipitation
with anti-RALT 19C5/4 monoclonal antibody followed by blot analysis with anti-RALT and anti-receptor antibodies. (b) ErbB-4 was
transiently expressed in 293 cells along with either RALT or RALT DEBR. At 48 h after transfection, cells were lysed after stimulation
for 5 min with either carrier solution or 20ng/ml NRG-1. Equal amounts of each lysate were analysed by immunoblot with anti-ErbB-4
and anti-P-Tyr antibodies. Anti-RALT 19C5/4 immunoprecipitates were resolved on SDS–PAGE and immunoblotted with anti-
RALT and anti-ErbB-4 antibodies. (c) RALT was coexpressed in 293 cells with either wt EGFR or EGFR mutants depicted in the left
portion of the panel, namely EGFR K721A (KD, i.e. kinase-defective) and EGFR DC, that lacks the COOH-terminal region. At 48 h
after transfection, cells were stimulated for 5 min with either carrier solution or 10 ng/ml EGF. Equal amounts of each lysate were
analysed by immunoblot with anti-RALT and anti-P-Tyr antibodies. Anti-EGFR immunoprecipitates were resolved on SDS–PAGE
and immunoblotted with anti-RALT and anti-EGFR antibodies
Negative signalling of RALT to the ErbB network
S Anastasi el al
4226
Oncogene
vector. Neither RALT nor RALT DEBR affected the
steady-state levels of ErbB RTKs in 32D transfectants
(Figure 7a). For each of the 32D.ErbB transfectants
mentioned above, we selected cell populations expres-
sing comparable amounts of RALT and RALT DEBR.
Levels of RALT proteins were similar in 32D.E2.E3 and
32D.E4 cells. When compared to 32D.E4 and
32D.E2.E3 cells, 32D.E1 transfectants contained lower
levels of ectopic RALT and RALT DEBR (Figure 7b).
In 32D.E1 Pinco cells EGF stimulation (3 ng/ml) led
to sustained activation of ERKs throughout the 6 h time
interval of the experiment. AKT was also activated
robustly in the initial 60 min of EGF stimulation, after
which time anti-phospho-AKT reactivity declined
(Figure 8a). In EGF-treated 32D.E1 RALT DEBR cells
strength and duration of ERK and AKT activation were
quite similar to those observed in Pinco controls. In
contrast, expression of RALT in 32D.E1 cells led to
premature termination of ERK and AKT activation,
with no reactivity to anti-P-ERK and anti-P-AKT being
observed past the 1 h and 15 min time point, respectively
(Figure 8a). These differences were not accounted for by
changes in either EGFR expression or EGFR phos-
phorylation on tyrosine residues (Figure 8a).
Stimulation of Pinco and RALT DEBR derivatives of
32D.E2.E3 cells with 10 ng/ml NRG-1 generated rather
similar kinetics of activation of AKT and ERKs
(Figure 8b). While the profile of ERK activity in
32D.E1 Pinco and 32D.E2.E3 Pinco transfectants were
comparable, activated ErbB-2.ErbB-3 dimers generated
a more sustained AKT signal (compare Figure 8a, b).
This observation is consistent with ErbB-3 providing
powerful activation of PI-3K once transphosphorylated
by catalytically competent ErbB RTKs. Expression of
ectopic RALT in 32D.E2.E3 cells caused premature
extinction of both ERK and AKT activity (Figure 8b).
ERK activity in E2.E3 RALT cells returned to back-
ground levels past the 3 h time point, that is, later than
what we observed in E1 RALT transfectants (Figure 8a).
AKT activity was not detectable past the 3 h time point
Figure 6 Distinct spatial distribution of RALT and RALT DEBR in response to activation of ErbB receptors. Subconfluent cultures
of serum-deprived NIH-EGFR/ErbB-2 (a–l) and NIH-EGFR (m–p) cells ectopically expressing either RALT or RALT DEBR were
processed for immunocytochemistry before (a,e,i,c,g,k,m,o) or after stimulation with EGF (10 ng/ml) for 60 (b,f,j,d,h,l) or 15 min
(n,p). NIH-EGFR/ErbB-2 cells (a–l) were double stained for ErbB-2 (a–d, green) and either RALT (e,f, red) or RALT DEBR (g,h,
red). Upon EGF stimulation, RALT (f) relocates to vesicular perinuclear structures where it colocalizes with the EGFR/ErbB-2
chimera (b, merge in j), while RALT DEBR (h) does not show significant relocation and does not colocalizes with EGFR/ErbB-2 (d,
merge in l). Also in EGF-treated NIH-EGFR cells (mp), RALT (m,n) relocates from the cytosol to a vesicular compartment, whereas
spatial distribution RALT DEBR (o,p) is not detectably altered by EGFR activation. Bars: 10 mm.
Negative signalling of RALT to the ErbB network
S Anastasi el al
4227
Oncogene
in RALT cells, whereas it was still high after 9 h of
NRG-1 stimulation in both E2.E3 Pinco and E2.E3
RALT DEBR cells. Neither RALT nor RALT DEBR
altered significantly the profile of expression and
activation of ErbB-2 and ErbB-3 (Figure 8b).
At variance with what observed in E1 and E2.E3
transfectants, expression of RALT in 32D.E4 cells led to
a severe inhibition of NRG-1-dependent ERK activity
since the inception of receptor triggering and through-
out the 6 h time frame of the experiment (Figure 8c).
Pronounced attenuation of anti-P-AKT reactivity in
32D.E4 RALT cell lysates was also observed at the
earliest time point of NRG-1 stimulation, with P-AKT
levels being reduced by 80% past the 60 min time point
(Figure 8c). These drastic effects of RALT activity on
ErbB-4 signalling could not be accounted for by obvious
alterations in levels of ErbB-4 expression and tyrosine
phosphorylation (Figure 8c). The DEBR mutation
caused RALT to lose its ability to suppress ErbB-4-
dependent activation of ERKs and AKT (Figure 8c). In
fact, the activity of ERKs and AKT was reproducibly
enhanced by the expression of RALT DEBR, raising the
possibility that the DEBR mutant exerts a dominant-
negative function (see the Discussion section).
Collectively, the data presented in Figure 8 indicate
that genetic reconstitution of RALT expression in
32D.E1, 32D.E4 and 32D.E2.E3 cells is sufficient to
suppress ErbB-dependent activation of ERKs and AKT.
RALT suppressed AKT and ERK activities with kinetics
that were significantly different from one another in
distinct 32D.ErbB transfectants. However, such inhibi-
tory function of RALT was in all cases strictly dependent
on the integrity of the EBR module and, by inference,
required physical interaction and spatial segregation of
RALT with activated ErbB receptors.
Impact of RALT on mitogenic signals generated by ErbB
RTKs
We next studied the impact of RALT and RALT DEBR
signalling on mitogenic programs governed by ErbB
RTKs. Ectopically expressed RALT did not influence
Figure 7 Analysis of expression of RALT, RALT DEBR and ErbB RTKs in 32D transfectants. (a) 32D transfectants expressing
EGFR, ErbB-4 and the ErbB-2.ErbB-3 combination were supertransfected with Pinco, Pinco–RALT and Pinco–RALT DEBR
expression vectors. Stable Pinco, RALT and RALT DEBR derivatives were established by puromycin selection. Lysates prepared from
each of the transfectants were immunoblotted with antibodies specific for each of the ErbB family members. Filters were also probed
with anti-SHC antibodies to control for equal loading of lysates. (b) lysates from 32D transfectants described in (a) were analysed
simultaneously for RALT expression by immunoblot in order to compare relative levels of RALT expression. Filters were stripped and
reprobed with anti-SHC antibodies to control for sample loading. (c) 32D and 32D.E1 cells were deprived of IL-3 for 2 h and then
stimulated for the indicated time with either IL-3 (growth medium) or EGF (10 ng/ml). Equal amounts of lysates were immunoblotted
with the indicated antibodies. A lysate from 32D.E1 RALT cells was used as positive control for RALT immunodetection
Negative signalling of RALT to the ErbB network
S Anastasi el al
4228
Oncogene
the ability of 32D derivatives to proliferate in response
to IL-3 (data not shown). In contrast, EGF-dependent
proliferation of 32D.E1 RALT cells was severely
reduced in comparison to their Pinco counterpart
(Figure 9a). The suppressive effect of RALT on EGFR
mitogenic signals was almost complete at suboptimal
concentrations of EGF; saturating doses of EGF
alleviated RALT-mediated suppression, rescuing mito-
genic activity to 25–30% of that of Pinco cells
(Figure 9a). Similar results were obtained with TGF-a
(data not shown). Annexin V staining indicated that
EGF stimulation rescued 32D.E1 Pinco cells from
apoptotic cell death induced by IL-3 deprivation. In
contrast, EGF-treated 32D.E1 RALT cells underwent
apoptosis at rates similar to those of 32D.E1 Pinco and
32D.E1 RALT cells cultured in medium devoid of both
EGF and IL-3 (data not shown). Thus, in 32D.E1 cells
RALT inhibits EGFR signals necessary for both cell
survival and cell proliferation, but does not exert a
direct proapoptotic activity. The RALT DEBR mutant
showed a bimodal behavior. At concentrations of EGF
equal to or lower than 3 ng/ml, expression of RALT
DEBR partially suppressed the mitogenic responses of
32D.E1 cells (Figure 9a). On the other hand, at EGF
concentrations higher than 3 ng/ml, which in this assay
corresponded approximately to the ED
50
of EGF, the
DEBR mutant enhanced EGF-dependent proliferation
of 32D.E1 cells.
Expression of RALT in 32D.E2.E3 cells led to the
suppression of mitogenic responses elicited by NRG-1
(Figure 9b). The suppressive activity of RALT was
strongest at concentrations of NRG-1 equal to or lower
Figure 8 Interaction of RALT with ErbB receptors affects strength and duration of ERK and AKT activation. Pinco, RALT and
RALT DEBR derivatives of 32D.E1 (a), 32D.E2.E3 (b) and 32D.E4 (c) transfectants described in Figure 7 were deprived of IL-3 for
2 h. Following stimulation with 3 ng/ml EGF (a) or 10 ng/ml NRG-1b(b,c) for the indicated time, cells were harvested and lysed in
Laemmli buffer. Lysates were analysed by immunoblot with the indicated antibodies. (a–c) Show a representative experiment of at
least three independent assays for E1, E2.E3 and E4 transfectants
Negative signalling of RALT to the ErbB network
S Anastasi el al
4229
Oncogene
than its ED
50
, with higher NRG-1 doses somehow
overcoming the inhibitory activity of RALT. RALT
activity prevented the NRG-1-dependent rescue of
32D.E2.E3 cells from the apoptotic effect of IL-3
deprivation, but was not proapoptotic per se (data not
shown). Strikingly, RALT DEBR was as efficient as
RALT in suppressing NRG-1-driven proliferation of
32D.E2.E3 cells at all of the NRG-1 concentrations
tested, including doses in excess of those yielding
optimal proliferation of 32D.E2.E3 Pinco cells
(Figure 9b). Expression of RALT in 32D.E4 cells led
to complete suppression of NRG-1-dependent cell
proliferation (Figure 9c), even at high ligand concentra-
tions. Similar results were obtained with b-cellulin (data
not shown). As observed with E1 and E2.E3 transfec-
tants, RALT did not exert a direct proapoptotic activity
(data not shown). The DEBR mutation deprived RALT
of the ability to suppress proliferation of 32D.E4 cells
over a wide range of NRG-1 concentrations (Figure 9c).
In fact, at optimal ligand concentrations, RALT DEBR
significantly enhanced NRG-1-dependent proliferation
of 32D.E4 cells.
Discussion
Quantitative aspects of RTK outputs are important in
establishing the identity of receptor signals and restrain-
ing their oncogenic potential. They are defined by the
balanced interplay between mechanisms of signal gen-
eration and signal extinction. Experimental manipula-
tions leading to either ablation (Fiorini et al., 2002) or
enhancement of RALT activity (Fiorentino et al., 2000)
in murine fibroblasts are consistent with RALT being
physiologically involved in negative signalling to the
ErbB-2 receptor. Strikingly, RALT appears to be a
rather selective inhibitor, as its overexpression is unable
to suppress cell proliferation driven by PDGF, bFGF
(Figure 1) and serum (Fiorentino et al., 2000), or cell
transformation induced by v-src and v-ras (Fiorentino
et al., 2000). This contrasts with the relative promiscuity
of other receptor-proximal feedback inhibitors, such as
mSprouty and SOCS proteins (reviewed in Fiorini et al.,
2001). The aim of this study was to analyse how RALT
function impacts on the activity of the ErbB network.
Physical interaction of RALT with ErbB receptors
RALT is able to form a physical complex in living cells
with ligand-activated ErbB RTKs, including EGFR,
ErbB-4 and the high-affinity NRG-1 receptor generated
by ErbB-2.ErbB-3 dimerization. The ErbB-binding
activity of RALT can be reconstituted in vitro using a
minimal 47 amino-acid region that binds distinct ErbB
family members with comparable affinities.
Binding of RALT to the EGFR in intact cells requires
that the receptor be catalytically active, as demonstrated
by the failure of kinase-inactive EGFR to form a
complex with RALT. Furthermore, the most distal 214
residues of the EGFR are dispensable for the RALT/
EGFR interaction. Hence, autophosphorylation is not
required for EGFR to recruit RALT and RALT must
bind to the EGFR upstream of position 972.
The requirements of ErbB-1 for RALT binding
recapitulate those described for ErbB-2/RALT complex
formation (Fiorentino et al., 2000) and may in fact
Figure 9 Impact of RALT signalling on mitogenic activity of
defined ErbB dimers. Pinco, RALT and RALT DEBR derivatives
of 32D.E1 (a), 32D.E2.E3 (b) and 32D.E4 (c) transfectants
described in Figure 6 were seeded at 1 10
4
cells/well in IL-3-free
medium and either kept in IL-3-free medium or supplemented with
either EGF (a) or NRG-1b(b,c) to the indicated final
concentrations; control wells were added IL-3. After 44 h, cells
were pulse-labelled for 4 h with 1 mCi/ml [
3
H]methyl-thymidine.
EGF- and NRG-1b-dependent cell proliferation was calculated as
percentage over IL-3-dependent proliferation. A representative
experiment of four independent assays for E1, E2.E3 and E4
transfectants is shown. Each experimental point was determined in
quadruplicate wells, with standard deviation not exceeding 15%
Negative signalling of RALT to the ErbB network
S Anastasi el al
4230
Oncogene
underlie the entire spectrum of ErbB/RALT interac-
tions. Thus, the interaction of RALT with ErbB-4 is also
dependent on ligand activation of the receptor, while
ErbB-1 and ErbB-2 may bind RALT in a ligand-
independent fashion when constitutively activated.
Conversely, catalytically impaired ErbB-3 cannot bind
RALT directly and is found in a complex with RALT
only if recruited into a heterodimer with ErbB-2.
Recently, in structure–function studies based on the
yeast two-hybrid assay Hackel et al. (2001) found that
RALT is unable to bind to kinase-inactive EGFR.
These authors also suggested that RALT binds to the
sequence comprised between positions 985–995 of
activated EGFR. This conclusion was based on the
observation that EGFR baits spanning positions 646–
995 were able to trans-activate reporter genes in a
RALT-dependent fashion, an activity that was lost in
the 646–985 and 640–940 EGFR baits (Hackel et al.,
2001). In contrast, our coimmunoprecipitation experi-
ments show that EGFR truncation at position 972 still
allows efficient recruitment of RALT (Figure 5c). The
latter result is compatible with yeast two-hybrid experi-
ments that provisionally delimited the RALT-binding
surface of ErbB-2 to the NH2 half of its kinase domain,
that is, well apart from the ErbB-2 COOH tail
(Fiorentino et al., 2000). Clearly, only the precise
mapping of the RALT-binding surface in different ErbB
RTKs will allow to solve these discrepancies. Based on
our data, we propose that RALT binds to the catalytic
domain of kinase-active ErbB RTKs following a ligand-
induced conformational change of the receptor. This
model is consistent with the inability of RALT to be
recruited to catalytically impaired ErbB-3 in the absence
of ErbB-2. Upon triggering of ErbB RTKs, RALT
molecules undergo redistribution from a cytosolic
location to a vesicular compartment, which likely
corresponds to endocytic vesicles. This process shows
an absolute requirement for the integrity of the EBR
module of RALT. Spatial segregation with ligand-
engaged ErbB receptors appears to be directly linked
to some aspects of RALT signalling to the ErbB
network, that is, suppression of activation of ERKs
and AKT (see below for further discussion).
RALT is a negative regulator of mitogenic signals
generated by the ErbB signalling network
In murine fibroblasts, RALT suppresses mitogenic
signals generated by a recombinant EGFR.ErbB-2
chimeric receptor (Fiorentino et al., 2000). Under
physiological conditions, the signalling activity of
ErbB-2 is dictated by its lateral interactions with other
ErbB family members. We report here that RALT
suppresses the mitogenic activity of NRG-1-activated
ErbB-2.ErbB-3 dimers as well as that of EGFR and
ErbB-4 homodimers. RALT suppresses mitogenic sig-
nalling by ErbB-2.ErbB-4 dimers activated by either
EGF or NRG-1 (our unpublished observation) as well
as the transforming activity of overexpressed EGFR in
Rat1 fibroblasts (Hackel et al., 2001). Thus, RALT
affects signalling by all ErbB RTK combinations tested
so far and it does so in several cellular model systems
and different biological assays. These findings, along
with the demonstration that RALT expression is
induced by growth factors capable of activating
different ErbB RTKs (Fiorentino et al., 2000) are
consistent with RALT being a feedback inhibitor of
the entire ErbB signalling network. Comparable levels
of RALT suppressed the mitogenic activity of ErbB-4
much more efficiently than that of E2.E3 dimers. This
was mirrored by remarkable differences in the ability of
RALT to interfere with E4 and E2.E3 signals leading to
ERK and AKT activation (see below for further
discussion). The potency of RALT signalling to EGFR
could not be directly compared to RALT activity
towards E4 and E2.E3, given the lower level of RALT
expression in 32D.E1 cells. We notice, however, that
NRG-driven proliferation of 32D.E4 cells was ablated
by RALT activity even when cells were cultured with
high concentrations of NRG-1. In contrast, saturating
doses of ligand partially alleviated RALT suppression of
E1 and E2.E3 mitogenic signals. Thus, the sensitivity
of EGFR to RALT suppressive activity resembled that
of E2.E3 dimers rather than that of E4.
The differential sensitivity of ErbB RTKs to RALT
signalling may impact on signal specification by the
ErbB network whenever a promiscuous ligand activates
more than one type of ErbB dimers in the same target
cell. For instance, NRG-1-driven RALT expression may
be used to tune signals generated by the ErbB-2.ErbB-3
combination, while aborting the function of concomi-
tantly activated ErbB-4 homodimers.
Role of RALT in determining strength and duration of
AKT and ERK activity
During development, distinct thresholds of RTK-
induced ERK activity are instrumental in translating
morphogenetic gradients into correspondingly graded
transcriptional responses (reviewed in Hazzalin and
Mahadevan, 2002). In proliferating cells, growth factor
stimulation and ensuing RTK activity are required
throughout the G1 phase of the mitotic cell cycle
(reviewed in Jones and Kazlauskas, 2001). Besides
regulating transcriptional responses (Cook et al., 1999;
Kops et al., 2002; Murphy et al., 2002) the Ras-ERK
and PI-3K-AKT pathways couple RTK activity to the
regulation of the cell cycle machinery (reviewed in
Lawlor and Alessi, 2001; Pruitt and Der, 2001). Thus,
intensity and duration of Ras-ERK and PI-3K-AKT
signals concur in defining the biological outcome of
ErbB activity in normal and transformed cells (Zhou
et al., 2001a, b; Neve et al., 2002; Shin et al., 2002;
Viglietto et al., 2002).
Reconstitution of RALT expression in 32D cells leads
to suppression of EGFR, ErbB-4 and ErbB-2.ErbB-3
signals culminating in activation of ERKs and AKT.
However, the RALT-enforced suppression of ERK and
AKT activity displays temporal profiles that are
different in distinct 32D.ErbB transfectants. RALT
expression in 32D.E4 cells led to an abrupt and robust
inhibition of ERK and AKT activity since the inception
Negative signalling of RALT to the ErbB network
S Anastasi el al
4231
Oncogene
of receptor triggering. In contrast, suppression of ERK
and AKT activation by RALT in E1 and E2.E3 cells
was clearly observed only at later stages of receptor
stimulation. Additionally, the timing of suppression of
ERKs and AKT was further delayed in E2.E3 cells as
compared to EGFR transfectants. Delayed suppression
of ErbB-dependent signals by RALT is not confined to
the 32D cell background: RALT overexpression caused
delayed suppression of ERK activity triggered by an
EGFR/ErbB-2 chimeric receptor expressed in murine
fibroblasts, while leaving serum- and PMA-driven
responses unaffected (Fiorentino et al., 2000).
Our results indicate that RALT activity impacts in a
selective fashion on strength and duration of ERK and
AKT signals generated by E1, E4 and E2.E3 dimers.
While it is not yet clear how RALT discerns among
different ErbB RTKs, these findings suggest that RALT
may protect cells against perturbations induced by
additive or superadditive activation of these pathways.
As an example, the buffering function of RALT is likely
to prevent further boosting of ERK and AKT activity
generated by E2.E3 dimers, should ErbB-4 be concomi-
tantly activated.
Mechanisms of suppression of ErbB RTKs signalling by
RALT
Initial studies modelled RALT as an adaptor that links
cytosolic effectors to activated ErbB receptors, thereby
suppressing receptor-proximal events (Fiorentino et al.,
2000; Hackel et al., 2001). In order to validate this
model, we have compared the biochemical and biologi-
cal activities of RALT to those of RALT DEBR, a
mutant which retains the property to couple to SH3-
containing effectors but is unable to complex with and
be spatially regulated by ErbB RTKs. Our data indicate
that the DEBR mutation deprives RALT of its ability to
suppress ErbB-dependent activation of ERKs and
AKT. We conclude that RALT must be recruited to
ErbB RTKs in order to interfere with these signalling
pathways.
Since RALT inhibits multiple signalling pathways once
relocated to ErbB RTKs, we infer that it acts as a general
suppressor of receptor-proximal events. Indeed, Hackel
et al. (2001) have reported that overexpression of RALT
enhances ligand-dependent downregulation of the EGFR.
Several SH3-containing proteins play essential roles in
receptor endocytosis, including Grb-2, intersectin, synda-
pin and endophilin (Simpson et al., 1999). Thus, RALT
may suppress ErbB RTKs by coupling them to elements
of the endocytic machinery such as Grb-2 (Fiorentino
et al., 2000) and intersectin (our unpublished data). This
model would provide a rationale for the seemingly
paradoxical enhancement of EGFR and ErbB-4 mito-
genic activity observed upon expression of RALT DEBR
in 32D.E1 and 32D.E4 cells (Figure 9). Such a dominant-
negative effect may be explained by RALT DEBR acting
as a cytosolic trap for spatially regulated components of
the endocytic machinery.
RALT DEBR is able to suppress the mitogenic
activity of ErbB-2.ErbB-3 dimers over a wide range of
ligand concentrations and under conditions that gen-
erate normal profiles of ERK and AKT signals. We
propose that the DEBR mutant intercepts, at a receptor-
distal level, signals yet unidentified but necessary to the
mitogenic activity of E2.E3 dimers. It is likely that these
signals are generated in parallel to the ERK and AKT
pathways and eventually converge with the latter to fully
implement the E2.E3 mitogenic program. Intriguingly,
the EBR-independent component of RALT signalling
impacts on the mitogenic activity of E1 and E4 only at
low levels of receptor occupancy. We hypothesize that
high levels of EGFR and ErbB-4 activity relieve the
dependency for signals intercepted by RALT DEBR,
thus allowing the above-described dominant-negative
activity of RALT DEBR to be unmasked.
Are there structural determinants of RALT that could
account for EBR-independent signals? RALT has been
reported to bind to 14-3-3 proteins (Makkinje et al.,
2000), which are implicated in nuclear–cytoplasmic
traffic. We have found that both RALT and RALT
DEBR accumulate in the nuclei of cells treated with
leptomycin B (unpublished data). Intriguingly, a RALT
fragment spanning positions 1–150 behaves as a potent
transcriptional activator when fused to the DNA-
binding domain of GAL-4 (unpublished data). Hence,
dynamic regulation of its nuclear localization may
confer to RALT the ability to regulate gene activity
during G1 progression in an EBR-independent fashion.
Conclusions
RALT appears to be a versatile and selective feedback
inhibitor of signals propagated by ErbB RTKs. Its
expression is controlled at the transcriptional and post-
translational level: such an integrated control provides
for timely adjustments of RALT levels and, by
inference, of its signalling activity (Fiorini et al., 2002).
The RALT feedback loop is therefore suited to proof-
read and extinguish signals generated by ErbB RTKs.
The ability of RALT to integrate negative signalling to
combinatorial ErbB dimers and to affect in a differential
manner strength and duration of signals generated by
distinct ErbB RTKs is likely to be a mechanism through
which the identity of ErbB signals is preserved and
global output from the ErbB network is tightly
controlled.
Materials and methods
Materials
EGF was from Upstate Biotechnology, NRG-1b, bFGF and
PDGF were from R&D, glutathione–agarose, IPTG, puromy-
cin and transferrin were from Sigma. Tissue culture media and
sera were from Life Technologies.
Recombinant DNA methodologies
pcDNA3 vectors for wt EGFR, EGFR KD and EGFR DC
have been described (Levkowitz et al., 1998) and were obtained
Negative signalling of RALT to the ErbB network
S Anastasi el al
4232
Oncogene
from Y. Yarden. LTR-2-based expression vectors for ErbB-1–
ErbB-4 have been previously described (Di Fiore et al., 1990;
Alimandi et al., 1995, 1997). cDNA fragments encoding
discrete portions of RALT EBR were generated by polymerase
chain reaction (PCR) amplification using synthetic oligonu-
cleotide primers of 22 bases. Amplification products contained
50BamHI and 30EcoRI tags to allow cloning into the pGEX
4 T vector (Pharmacia) and expression in Escherichia coli. The
DEBR mutant was generated by fusing an EcoRI–StuI fragment
of RALT cDNA encompassing aa. 1–313 to a PCR-generated
StuI–XbaI fragment spanning positions 362–459 of RALT.
Sequence analysis confirmed the fidelity of PCR amplification
and that the fusion maintained the correct ORF skipping aa.
315–361. For expression in mammalian cells, RALT DEBR was
cloned in Pinco vector (Fiorentino et al., 2000).
Cell lines, cell culture conditions and gene transfer procedures
NIH-3T3 cells and their derivatives were grown in Dulbecco’s
minimal essential medium (D-MEM) containing 10% (vol/vol)
newborn calf serum (NCS). 293 human embryonal kidney cells
were grown in D-MEM containing 10% fetal calf serum
(FCS). Transfections of 293 cells were performed as described
(Fiorentino et al., 2000). Cells were processed for immuno-
precipitation experiments 36–48 h post-transfection. NIH 3T3,
NIH-EGFR and NIH-EGFR/ErbB-2 cells were infected with
Pinco, Pinco-RALT and Pinco–RALT DEBR retrovirus
stocks to obtain Pinco, RALT and RALT DEBR derivatives.
Flow-cytometry analysis of GFP-positive cells indicated that
infection efficiencies were routinely 495%. 32D cells and their
derivatives were cultured in RPMI medium supplemented with
10% FCS and 5% (vol/vol) conditioned medium from WEHI
cells as source of IL-3. Expression of ectopic RALT and
RALT DEBR in 32D.E1 (Di Fiore et al., 1990), 32D.E4 (Tang
et al., 1998) and 32D.E2.E3 cells (Alimandi et al., 1997) was
achieved by electroporation of Pinco–RALT and Pinco–
RALT DEBR vectors as described (Alimandi et al., 1997).
Transfected cells were selected in 1 mg/ml puromycin. Prolif-
eration assays with 32D cells were performed as described
(Alimandi et al., 1997). Briefly, cells were seeded at
110
4
cells/well in 96 flat bottom well plates in IL-3-free
medium containing either ErbB ligands or 5% WEHI medium.
Control cells were cultivated in IL-3-free medium. After 44 h,
cells were pulsed for 4 h with 1 mCi/ml [
3
H]methyl-thymidine
and TCA-precipitated radioactivity measured in a scintillation
counter. Assays were performed in quadruplicate wells. ErbB-
driven cell proliferation was calculated as percentage over IL-
3-driven proliferation. Mitogenic assays of NIH-EGFR/ErbB-
2 cells were performed as described (Fiorentino et al., 2000).
Briefly, 5 10
3
cells/well were seeded in 24-well plates and
subjected to three cycles of infection with either Pinco or
PincoRALT retrovirus stocks over a 24 h period. Cells were
then switched to serum-free medium containing defined
growth factors. Control wells were added to serum-free
medium only. Cell proliferation was assessed after 48 h by
pulsing cells for 4 h with 1 mCi/ml [
3
H]methyl-thymidine and
counting TCA-precipitated radioactivity in a scintillation
counter.
Immunochemical procedures
For immunoblot analysis, cells were lysed as described
(Fiorentino et al., 2000). The S1 antiserum and the 19C5/4
monoclonal antibody were used at 2 mg/ml. Anti-ERK, anti-p-
ERK, anti-AKT and anti-p-AKT antibodies (New England
Biolabs), anti-P-Tyr MoAb 4G10 (UBI), anti-SHC antiserum
(Transduction Laboratories) and affinity-purified anti-ErbB
receptors antibodies (Santa Cruz Biotechnology) were used at
1mg/ml. GST pull-down assays were performed as described
(Fiorentino et al., 2000). For coimmunoprecipitation assays,
lysates were incubated for 2 h at 41C with MoAb 19C5/4
bound to ImmunoPure resin (Pierce).
Immunofluorescence and confocal microscopy analysis
Cultures of NIH-EGFR and NIH-EGFR/ErbB-2 cells in-
fected with Pinco–RALT and Pinco–RALT DEBR recombi-
nant retroviruses were deprived of serum for 24 h prior to
stimulation with EGF (10 ng/ml) for different lengths of time.
Cells were fixed with 4% (vol/vol) paraformaldehyde in PBS
for 10 min at 201C and permeabilized with 0.25% (vol/vol)
Triton X-100 in PBS for 10 min. Comparable results were
obtained when cells were fixed and permeabilized with a 1 : 1
methanol/acetone solution for 10 min at 201C. Immunostain-
ing was performed with MoAb 19C5/4 (for RALT proteins),
MoAb 108 (for EGFR) and polyclonal antibody M6 (for the
EGFR/ErbB-2 chimera) as described (Fiorentino et al., 2000).
Secondary antibodies (FITC- and TRITC-conjugated donkey
anti-rabbit, anti-goat and anti-mouse antibodies) were from
Jackson Immunoresearch. Samples were routinely examined
with a Zeiss microscope equipped with 40 and 50 water-
immersion objectives. Confocal analysis was carried out with a
Leica NBT system, equipped with 40 1.00–0.5 and 100 1.3–
0.6 oil-immersion lenses.
Acknowledgements
S Anastasi dedicates this work to the loving memory of his
mother. We thank Y Yarden and J Sap for reagents and C Full
for tissue culture work. MF is the recipient of an AIRC
fellowship. This work was supported by grants from: EC (5th
program) and AIRC to OS, AIRC and CNR PF- Biotecno-
logie to SA, AIRC to MA.
References
Alimandi M, Romano A, Curia MC, Muraro R, Fedi P,
Aaronson SA, Di Fiore PP and Kraus MH. (1995).
Oncogene,10, 1813–1821.
Alimandi M, Wang LM, Bottaro D, Lee CC, Kuo A, Frankel
M, Fedi P, Tang C, Lippman M and Pierce JH. (1997).
EMBO J.,16, 5608–5617.
Carraway III KL, Sliwkowski MX, Akita R, Platko JV, Guy
PM, Nuijens A, Diamonti AJ, Vandlen RL, Cantley LC and
Cerione RA. (1994). J. Biol. Chem.,269, 14303–14306.
Carraway III KL, Weber JL, Unger MJ, Ledesma J, Yu N,
Gassmann M and Lai C. (1997). Nature,387,
512–516.
Cook SJ, Aziz N and McMahon M. (1999). Mol. Cell. Biol.,
19, 330–341.
Di Fiore PP, Segatto O, Taylor WG, Aaronson SA and Pierce
JH. (1990). Science,248, 79–83.
Fiorentino L, Pertica C, Fiorini M, Talora C, Crescenzi M,
Castellani L, Alema S, Benedetti P and Segatto O. (2000).
Mol. Cell. Biol.,20, 7735–7750.
Fiorini M, Alimandi M, Fiorentino L, Sala G and Segatto O.
(2001). FEBS Lett.,490, 132–141.
Fiorini M, Ballaro C, Sala G, Falcone G, Alema S and Segatto
O. (2002). Oncogene,21, 6530–6539.
Freeman M. (2000). Nature,408, 313–319.
Negative signalling of RALT to the ErbB network
S Anastasi el al
4233
Oncogene
Ghiglione C, Perrimon N and Perkins LA. (1999). Dev. Biol.,
205, 181–193.
Graus-Porta D, Beerli RR, Daly JM and Hynes NE. (1997).
EMBO J.,16, 1647–1655.
Graus-Porta D, Beerli RR and Hynes NE. (1995). Mol. Cell.
Biol.,15, 1182–1191.
Guy PM, Platko JV, Cantley LC, Cerione RA and
Carraway III KL. (1994). Proc. Natl. Acad. Sci. USA,91,
8132–8136.
Hackel PO, Gishizky M and Ullrich A. (2001). Biol. Chem.,
382, 1649–1662.
Hazzalin CA and Mahadevan LC. (2002). Nat. Rev. Mol. Cell.
Biol.,3, 30–40.
Jones SM and Kazlauskas A. (2001). FEBS Lett.,490, 110–
116.
Karunagaran D, Tzahar E, Beerli RR, Chen X, Graus-Porta
D, Ratzkin BJ, Seger R, Hynes NE and Yarden Y. (1996).
EMBO J.,15, 254–264.
Kops GJ, Medema RH, Glassford J, Essers MA, Dijkers PF,
Coffer PJ, Lam EW and Burgering BM. (2002). Mol. Cell.
Biol.,22, 2025–2036.
Lawlor MA and Alessi DR. (2001). J. Cell Sci.,114, 2903–
2910.
Lee KF, Simon H, Chen H, Bates B, Hung MC and Hauser C.
(1995). Nature,378, 394–398.
Levkowitz G, Waterman H, Zamir E, Kam Z, Oved S,
Langdon WY, Beguinot L, Geiger B and Yarden Y. (1998).
Genes Dev.,12, 3663–3674.
Lonardo F, Di Marco E, King CR, Pierce JH, Segatto O,
Aaronson SA and Di Fiore PP. (1990). New Biol.,2, 992–
1003.
Makkinje A, Quinn DA, Chen A, Cadilla CL, Force T,
Bonventre JV and Kyriakis JM. (2000). J. Biol. Chem.,275,
17838–17847.
Marshall CJ. (1995). Cell,80, 179–185.
Murphy LO, Smith S, Chen RH, Fingar DC and Blenis J.
(2002). Nat. Cell Biol.,4, 556–564.
Neve RM, Holbro T and Hynes NE. (2002). Oncogene,21,
4567–4576.
Olayioye MA, Beuvink I, Horsch K, Daly JM and Hynes NE.
(1999). J. Biol. Chem.,274, 17209–17218.
Olayioye MA, Graus-Porta D, Beerli RR, Rohrer J, Gay B
and Hynes NE. (1998). Mol. Cell. Biol.,18, 5042–5051.
Olayioye MA, Neve RM, Lane HA and Hynes NE. (2000).
EMBO J.,19, 3159–3167.
Perrimon N and McMahon AP. (1999). Cell,97, 13–16.
Pinkas-Kramarski R, Soussan L, Waterman H, Levkowitz G,
Alroy I, Klapper L, Lavi S, Seger R, Ratzkin BJ, Sela M and
Yarden Y. (1996). EMBO J.,15, 2452–2467.
Pruitt K and Der CJ. (2001). Cancer Lett.,171, 1–10.
Riese DJ and Stern DF. (1998). Bioessays,20, 41–48.
Riese DJ, van Raaij TM, Plowman GD, Andrews GC and
Stern DF. (1995). Mol. Cell. Biol.,15, 5770–5776.
Shin I, Yakes FM, Rojo F, Shin NY, Bakin AV, Baselga J and
Arteaga CL. (2002). Nat. Med.,8, 1145–1152.
Simpson F, Hussain NK, Qualmann B, Kelly RB, Kay BK,
McPherson PS and Schmid SL. (1999). Nat. Cell Biol.,1,
119–124.
Tang CK, Goldstein DJ, Payne J, Czubayko F, Alimandi M,
Wang LM, Pierce JH and Lippman ME. (1998). Cancer
Res.,58, 3415–3422.
Viglietto G, Motti ML, Bruni P, Melillo RM, D’Alessio A,
Califano D, Vinci F, Chiappetta G, Tsichlis P, Bellacosa A,
Fusco A and Santoro M. (2002). Nat. Med.,8, 1136–1144.
Wallasch C, Weiss FU, Niederfellner G, Jallal B, Issing W and
Ullrich A. (1995). EMBO J.,14, 4267–4275.
Waterman H and Yarden Y. (2001). FEBS Lett.,490,
142–152.
Yarden Y and Sliwkowski MX. (2001). Nat. Rev. Mol. Cell.
Biol.,2, 127–137.
Zhou BP, Liao Y, Xia W, Spohn B, Lee MH and Hung MC.
(2001a). Nat. Cell Biol.,3, 245–252.
Zhou BP, Liao Y, Xia W, Zou Y, Spohn B and Hung MC.
(2001b). Nat. Cell Biol.,3, 973–982.
Negative signalling of RALT to the ErbB network
S Anastasi el al
4234
Oncogene
... Still on the ErbB signaling, our results showed a higher relative expression of ERRFI1 at all times of treatment with L. laeta or L. intermedia SMases D, showing the highest expression peak at 2 h (Fig. 7b). ERRFI1 is known to be an inhibitor of the ErbB signaling pathway through its binding to the EGFR kinase domain, impairing receptor dimerization and thus its catalytic activity, as also as mediating EGFR lysosomal degradation (Anastasi et al. 2003;Zhang et al. 2007;Frosi et al. 2010;Cairns et al. 2018). ERRFI1 has been also reported to induce apoptosis and inhibit cell proliferation in some types of cancers (Lin et al. 2011;Cui et al. 2021). ...
... ERRFI1 has been also reported to induce apoptosis and inhibit cell proliferation in some types of cancers (Lin et al. 2011;Cui et al. 2021). Its overexpression is enough to avoid auto-phosphorylation and the subsequent signaling (Anastasi et al. 2003;Xu et al. 2005). Taken together, our previous results and the data showed here, the model that is drawn suggests that although the action of ADAM17 and ADAM19 contributes to the ectodomain shedding of the EGFR ligands, in the presence of SMases D, these same proteases act in the cleavage of EGFR, resulting in the inhibition of ErbB signaling. ...
Article
Full-text available
Sphingomyelinase D (SMase D), the main toxic component of Loxosceles venom, has a well-documented role on dermon-ecrotic lesion triggered by envenomation with these species; however, the intracellular mechanisms involved in this event are still poorly known. Through differential transcriptomics of human keratinocytes treated with L. laeta or L. intermedia SMases D, we identified 323 DEGs, common to both treatments, as well as upregulation of molecules involved in the IL-1 and ErbB signaling. Since these pathways are related to inflammation and wound healing, respectively, we investigated the relative expression of some molecules related to these pathways by RT-qPCR and observed different expression profiles over time. Although, after 24 h of treatment, both SMases D induced similar modulation of these pathways in keratinocytes, L. intermedia SMase D induced earlier modulation compared to L. laeta SMase D treatment. Positive expression correlations of the molecules involved in the IL-1 signaling were also observed after SMases D treatment, confirming their inflamma-tory action. In addition, we detected higher relative expression of the inhibitor of the ErbB signaling pathway, ERRFI1, and positive correlations between this molecule and pro-inflammatory mediators after SMases D treatment. Thus, herein, we describe the cell pathways related to the exacerbation of inflammation and to the failure of the wound healing, highlighting the contribution of the IL-1 signaling pathway and the ERRFI1 for the development of cutaneous loxoscelism.
... The protein can dock at the activation interface of EGFR kinase domain by using its evolutionarily conserved ErbB-binding region (EBR) module to disrupt the asymmetric domain dimerization (Hackel et al. 2001;Anastasi et al. 2007) and has been reported to cause phenotypes due to EGFR-driven excess cell proliferation upon genetic ablation (Ferby et al. 2006). Anastasi et al. found that RALT behaved as a pan-ErbB inhibitor and affected the signaling activity of distinct ErbB kinases with different relative potencies, which can bind to the homo-and hetero-dimers of EGFR:EGFR, ErbB2:ErbB3 and ErbB4:ErbB4 and suppress their downstream signaling activation at cellular level (Anastasi et al. 2003). In addition, overexpression of RALT in mouse fibroblasts was observed to inhibit ErbB2-induced cell proliferation/transformation and to sustain extracellular signalregulated activation by ErbB2 (Fiorentino et al. 2000). ...
... As might be expected, the protein can generally target ErbB kinases effectively, with ∆U bnd > − 40 kcal/ mol for EGFR, ErbB2 and ErbB4, although it has only a moderate potency to ErbB3 (∆U bnd = − 31.5 kcal/mol). The calculated results are consistent with the previous cellularlevel observation that the RALT can affect the signaling activity of different ErbB kinases with varying relative capabilities (Anastasi et al. 2003). In contrast, the binding capability of RALT EBR domain to most RTKs is moderate or modest, while only exhibiting strong or high binding to few kinases, indicating that the protein can selectively inhibit a number of specific kinases as its cognate targets. ...
Article
Full-text available
The ErbB family of receptor tyrosine kinases (RTKs) contains four members: EGFR, ErbB2, ErbB3 and ErbB4; they are involved in the tumorigenesis of diverse cancers and can be inhibited natively by receptor-associated late transducer (RALT), a negative feedback regulator of ErbB signaling in human hepatocytes and hepatocellular carcinoma. Although the biological effects of RALT on EGFR kinase have been widely documented previously, the binding behavior of RALT to other ErbB/RTK kinases still remains largely unexplored. Here, the intermolecular interactions of RALT ErbB-binding region (EBR) as well as its functional sections and peptide segments with ErbBs and other human RTKs were systematically investigated at molecular and structural levels, from which we were able to identify those potential kinase targets of RALT protein, and to profile the affinity, specificity and cross-reactivity of RALT EBR domain and its sub-regions against various RTKs. It is revealed that RALT can target all the four ErbB kinases with high affinity for EGFR/ErbB2/ErbB4 and moderate affinity for ErbB3, but generally exhibits modest affinity to other RTKs, albeit few kinases such as LTK, EPHB6, MET and MUSK were also top-ranked as the unexpected targets of RALT. Peptide segments covering the key binding regions of RALT EBR domain were identified with computational alanine scanning, which were then optimized to obtain a number of designed peptide mutants with improved selectivity between different top-ranked RTKs.
... ERRFI1 (ERBB receptor feedback inhibitor 1) is an early response gene encoding a non-kinase scaffold adaptor protein induced by various stimuli such as hormones and stresses [10]. ERRFI1 is considered a negative regulator of EGFR because it can directly bind to EGFR, inhibit the catalytic activity of EGFR, and mediate EGFR lysosomal degradation [11][12][13]. Considering the involvement of ERRFI1 in the signaling of the EGFR family, its potential role in cancer draws great attention. Studies indicate that ERRFI1 is downregulated in lung cancer, breast cancer, thyroid cancer, and many other cancer types [14][15][16]. ...
... ERRFI1 is induced by various stimuli such as hormones and stresses [10]. It is reported that ERRFI1 decreases tumor formation by inhibiting cell proliferation and increasing apoptosis [12,19,34]. ERRFI1 has been shown to promote apoptosis in lung cancer cells via the ERK pathway [19]. ...
Article
Full-text available
Tryptophan metabolism is an essential regulator of tumor immune evasion. However, the effect of tryptophan metabolism on cancer cells remains largely unknown. Here, we find that tumor cells have distinct responses to tryptophan deficiency in terms of cell growth, no matter hepatocellular carcinoma (HCC) cells, lung cancer cells, or breast cancer cells. Further study shows that ERRFI1 is upregulated in sensitive HCC cells, but not in resistant HCC cells, in response to tryptophan deficiency, and ERRFI1 expression level positively correlates with HCC patient overall survival. ERRFI1 knockdown recovers tryptophan deficiency-suppressed cell growth of sensitive HCC cells. In contrast, ERRFI1 overexpression sensitizes resistant HCC cells to tryptophan deficiency. Moreover, ERRFI1 induces apoptosis by binding PDCD2 in HCC cells, PDCD2 knockdown decreases the ERRFI1-induced apoptosis in HCC cells. Thus, we conclude that ERRFI1-induced apoptosis increases the sensitivity of HCC cells to tryptophan deficiency and ERRFI1 interacts with PDCD2 to induce apoptosis in HCC cells.
... Делеція гена Mig6 у миші гіперактивує передачу сигналів EGFR, що призводить до гіперплазії шкіри та спонтанного утворення пухлини в шкірі, легенях та інших тканинах. Нещодавнє дослідження показало, що втрата Mig6 сприяє ініціації та [23]. Геномні зміни в позаклітинних, карбокси-термінальних і каталітичних областях EGFR були виявлені в гліобластомі, а ампліфікації EGFR є відмінною рисою класичного підтипу цього захворювання [24]. ...
Article
Проблематика. Рецептори епідермального фактора росту та їх ліганди, зокрема епідермальний фактор росту та гепарин-зв'язуючий EGF-подібний фактор росту, відіграють ключову роль у регуляції клітинних процесів, таких як проліферація, диференціація, міграція та виживання. Їхні властивості роблять їх перспективними для використання у регенеративній медицині та біоінженерії. Мета роботи. Основною метою аналітичного огляду був аналіз молекулярно біологічних властивостей епідермального фактору росту та участь у регенеративних процесах організму людини. Методика реалізації. В роботі проаналізовано молекуляроно біологічні властивості та основне клінічне застосування епідермального фактора росту у регенеративній медицині та біоінженерії. Екзогенне введення епідермального фактора росту показало сприятливий антиапоптотичний і антиоксидантний ефект і, як було встановлено, зменшує пошкодження тканин, викликаних ішемією в різних органах, таких як серце, кишечник і нирки. Епідермальний фактор росту використовують для лікування алопеції та дерматиту після хіміотерапії, опікової хвороби, виразок діабетичної стопи, післяопераційних виразок, мукозиту ротової порожнини, виразок глотки та перфорації барабанної перетинки. Використання епідермального фактора росту показує певний успіх у реструктуризації шкірного покриву, прискоренні загоєння виразок. При використанні епідермального фактора росту, як лікувальної терапії слід враховувати особистий анамнез і сімейну генетичну основу. У біоінженерії епідермальний фактор росту використовується для створення біосумісних імплантатів, спрямованих на підтримку регенерації тканин. Використання рекомбінантних химер епідермального фактора росту дозволяє оптимізувати біоактивність імплантатів. Висновок. Загалом, розуміння властивостей та механізмів дії рецепторів та лігандів епідермального фактора росту сприятиме розвитку нових методів лікування та біоінженерних технологій з метою покращення здоров'я та якості життя пацієнтів. Ключові слова: епідермальний фактор росту, гепарин зв’язувальний фактор росту, епідерміс, біоінженерія
... In this study, we demonstrated that glucolipotoxicity and ER stress attenuate EGFR activation in pancreatic beta cells via the stress-responsive EGFR inhibitor, Mig6. Mig6 was initially characterized as an endogenous EGFR feedback inhibitor but has also been suggested to impair other RTKs [22][23][24], yet in our work here, we have been unable to ascribe the actions of Mig6 to HGF or IGF-1 signaling in 832/13 rat pancreatic beta cells. After mitogen stimulation, Mig6 is activated to abolish EGFR signaling transmission via a two-tiered mechanism: (1) Mig6 binds to the EGFR intracellular kinase domain and inhibits kinase dimerization and activation, and (2) Mig6 facilitates EGFR endo-lysosomal sorting and degradation [21]. ...
Article
Full-text available
A loss of functional beta cell mass is a final etiological event in the development of frank type 2 diabetes (T2D). To preserve or expand beta cells and therefore treat/prevent T2D, growth factors have been considered therapeutically but have largely failed to achieve robust clinical success. The molecular mechanisms preventing the activation of mitogenic signaling pathways from maintaining functional beta cell mass during the development of T2D remain unknown. We speculated that endogenous negative effectors of mitogenic signaling cascades impede beta cell survival/expansion. Thus, we tested the hypothesis that a stress-inducible epidermal growth factor receptor (EGFR) inhibitor, mitogen-inducible gene 6 (Mig6), regulates beta cell fate in a T2D milieu. To this end, we determined that: (1) glucolipotoxicity (GLT) induces Mig6, thereby blunting EGFR signaling cascades, and (2) Mig6 mediates molecular events regulating beta cell survival/death. We discovered that GLT impairs EGFR activation, and Mig6 is elevated in human islets from T2D donors as well as GLT-treated rodent islets and 832/13 INS-1 beta cells. Mig6 is essential for GLT-induced EGFR desensitization, as Mig6 suppression rescued the GLT-impaired EGFR and ERK1/2 activation. Further, Mig6 mediated EGFR but not insulin-like growth factor-1 receptor nor hepatocyte growth factor receptor activity in beta cells. Finally, we identified that elevated Mig6 augmented beta cell apoptosis, as Mig6 suppression reduced apoptosis during GLT. In conclusion, we established that T2D and GLT induce Mig6 in beta cells; the elevated Mig6 desensitizes EGFR signaling and induces beta cell death, suggesting Mig6 could be a novel therapeutic target for T2D.
... Previously we identified MIG-6 as a target of PGR in the uterus necessary to retain P4 responsiveness for endometrial homeostasis and successful embryo implantation 20,32 . Past studies also showed that MIG-6 binds to ERBB2 and inhibits its signaling in vitro [53][54][55] . In the present study, we delineate the action of PGR through MIG-6 to suppress ERBB2 expression in vivo in the uterus. ...
Article
Full-text available
Female subfertility is highly associated with endometriosis. Endometrial progesterone resistance is suggested as a crucial element in the development of endometrial diseases. We report that MIG-6 is downregulated in the endometrium of infertile women with endometriosis and in a non-human primate model of endometriosis. We find ERBB2 overexpression in the endometrium of uterine-specific Mig-6 knockout mice (Pgrcre/+Mig-6f/f; Mig-6d/d). To investigate the effect of ERBB2 targeting on endometrial progesterone resistance, fertility, and endometriosis, we introduce Erbb2 ablation in Mig-6d/d mice (Mig-6d/dErbb2d/d mice). The additional knockout of Erbb2 rescues all phenotypes seen in Mig-6d/d mice. Transcriptomic analysis shows that genes differentially expressed in Mig-6d/d mice revert to their normal expression in Mig-6d/dErbb2d/d mice. Together, our results demonstrate that ERBB2 overexpression in endometrium with MIG-6 deficiency causes endometrial progesterone resistance and a nonreceptive endometrium in endometriosis-related infertility, and ERBB2 targeting reverses these effects. Female subfertility is highly associated with endometriosis. Here the authors show that progesterone-induced MIG-6 is reduced in endometrium of infertile women and non-human primates with endometriosis, and in a mouse model find that Erbb2 is the key mediator of Mig-6 loss induced endometriosis-related infertility.
... MIG6 is a cytoplasmic protein that interacts with the tyrosine kinase domain of EGFR, ERBB2, and ERBB4, and inhibits receptor dimerization (Hackel et al., 2001;Anastasi et al., 2003;Zhang et al., 2007;Nagashima et al., 2009). The transient localization of SHC in the EGFR-overexpressing cell may indicate dimerization and/or a clustering state of ERBB receptors that distinctly regulates their affinities for adaptor proteins. ...
Article
Full-text available
p52SHC (SHC) and GRB2 are adaptor proteins involved in the RAS/MAPK (ERK) pathway mediating signals from cell-surface receptors to various cytoplasmic proteins. To further examine their roles in signal transduction, we studied the translocation of fluorescently-labeled SHC and GRB2 to the cell surface, caused by the activation of ERBB receptors by heregulin (HRG). We simultaneously evaluated activated ERK translocation to the nucleus. Unexpectedly, the translocation dynamics of SHC were sustained when those of GRB2 were transient. The sustained localization of SHC positively correlated with the sustained nuclear localization of ERK, which became more transient after SHC knockdown. SHC-mediated PI3K activation was required to maintain the sustainability of the ERK translocation regulating MEK but not RAF. In cells overexpressing ERBB1, SHC translocation became transient, and the HRG-induced cell fate shifted from a differentiation to a proliferation bias. Our results indicate that SHC and GRB2 functions are not redundant, but that SHC plays the critical role in the temporal regulation of ERK activation.
Article
Full-text available
Purpose: TGFβ-induced epithelial-to-myofibroblast transition (EMyT) of lens cells has been linked to the most common vision-disrupting complication of cataract surgery-namely, posterior capsule opacification (PCO; secondary cataract). Although inhibitors of the ErbB family of receptor tyrosine kinases have been shown to block some PCO-associated processes in model systems, our knowledge of ErbB signaling in the lens is very limited. Here, we investigate the expression of ErbBs and their ligands in primary cultures of chick lens epithelial cells (dissociated cell-derived monolayer cultures [DCDMLs]) and how TGFβ affects ErbB function. Methods: DCDMLs were analyzed by immunofluorescence microscopy and Western blotting under basal and profibrotic conditions. Results: Small-molecule ErbB kinase blockers, including the human therapeutic lapatinib, selectively inhibit TGFβ-induced EMyT of DCDMLs. Lens cells constitutively express ErbB1 (EGFR), ErbB2, and ErbB4 protein on the plasma membrane and release into the medium ErbB-activating ligand. Culturing DCDMLs with TGFβ increases soluble bioactive ErbB ligand and markedly alters ErbBs, reducing total and cell surface ErbB2 and ErbB4 while increasing ErbB1 expression and homodimer formation. Similar, TGFβ-dependent changes in relative ErbB expression are induced when lens cells are exposed to the profibrotic substrate fibronectin. A single, 1-hour treatment with lapatinib inhibits EMyT in DCDMLs assessed 6 days later. Short-term exposure to lower doses of lapatinib is also capable of eliciting a durable response when combined with suboptimal levels of a mechanistically distinct multikinase inhibitor. Conclusions: Our findings support ErbB1 as a therapeutic target for fibrotic PCO, which could be leveraged to pharmaceutically preserve the vision of millions of patients with cataracts.
Article
The main factors contributing to the unfavorable outcome in the clinical treatment of patients with nasopharyngeal carcinoma (NPC) patients are radiation resistance and recurrence. This study aimed to investigate the sensitivity and molecular foundation of cytokeratin 13 (CK13) in the radiotherapy of NPC. To achieve this, a human NPC cell line overexpressing CK13, HNE-3-CK13, was constructed. The effects of CK13 overexpression on cell viability and apoptosis under radiotherapy conditions were evaluated using the CCK-8 assay, immunofluorescence, and western blotting (WB). Next-generation sequencing was performed to identify the downstream genes and signaling pathways of CK13 that mediate radiotherapy response. The potential role of the candidate gene ERRFI1 in CK13-induced enhancement of radiosensitivity was investigated through rescue experiments using clone formation and WB. The effects of ERRFI1 on cell viability, cell apoptosis, cell cycle, and the related key genes were further evaluated using CCK-8, immunofluorescence, flow cytometry, quantitative polymerase chain reaction and WB. The results showed that CK13 overexpression in HNE-3 significantly inhibited cell survival under radiotherapy and promoted apoptosis marker γH2AX expression, leading to a significant increase of ERRFI1. Knockdown of ERRFI1 rescued the decreased cell viability and proliferation and the increased cell apoptosis that were caused by CK13 overexpression-mediated radiotherapy sensitization of NPC cells. In this process, EGFR, AKT, and GSK-3β were found involved. In the end, ERRFI1 was proven to inhibit expression levels of CDK1, CDK2, cyclin B1, and cyclin D1, resulting an increased G2/M cell ratio. Overexpression of CK13 enhances the radiosensitivity of NPC cells, which is characterized by decreased cell viability and proliferation and increased apoptosis. This regulation may affect the survival of HNE-3 cells by increasing the expression of ERRFI1 and activating the EGFR/Akt/GSK-3β signaling pathway, providing new potential therapeutic targets for the treatment of NPC.
Article
Ack1 is a proto‐oncogenic tyrosine kinase with homology to the tumor suppressor Mig6, an inhibitor of the epidermal growth factor receptor (EGFR). The residues critical for binding of Mig6 to EGFR are conserved within the Mig6 homology region (MHR) of Ack1. We tested whether intramolecular interactions between the Ack1 MHR and kinase domain (KD) are regulated by phosphorylation. We identified two Src phosphorylation sites within the MHR (Y859, Y860). Addition of Src‐phosphorylated MHR to the Ack1 KD enhanced enzymatic activity. Co‐expression of Src in cells led to increased Ack1 activity; mutation of Y859/Y860 blocked this increase. Collectively, the data suggest that phosphorylation of the Ack1 MHR regulates its kinase activity. Phosphorylation of Y859/Y860 occurs in cancers of the brain, breast, colon, and prostate, where genomic amplification or somatic mutations of Ack1 play a role in disease progression. Our findings suggest that MHR phosphorylation could contribute to Ack1 dysregulation in tumors. The proto‐oncogenic tyrosine kinase Ack1 is autoinhibited through its C‐terminal Mig6 homology region (MHR) under basal conditions. Phosphorylation by Src within the MHR (Y859, Y860) enhances Ack1 catalytic activity. Thus, MHR phosphorylation could contribute to the normal regulation of Ack1 and its dysregulation in human cancer.
Article
Full-text available
Ligand-induced down-regulation of two growth factor receptors, EGF receptor (ErbB-1) and ErbB-3, correlates with differential ability to recruit c-Cbl, whose invertebrate orthologs are negative regulators of ErbB. We report that ligand-induced degradation of internalized ErbB-1, but not ErbB-3, is mediated by transient mobilization of a minor fraction of c-Cbl into ErbB-1-containing endosomes. This recruitment depends on the receptor's tyrosine kinase activity and an intact carboxy-terminal region. The alternative fate is recycling of internalized ErbBs to the cell surface. Cbl-mediated receptor sorting involves covalent attachment of ubiquitin molecules, and subsequent lysosomal and proteasomal degradation. The oncogenic viral form of Cbl inhibits down-regulation by shunting endocytosed receptors to the recycling pathway. These results reveal an endosomal sorting machinery capable of controlling the fate, and, hence, signaling potency, of growth factor receptors.
Article
Full-text available
THE receptor erbB2/neu is a member of the epidermal growth factor receptor (EGFR or erbB) family that also includes erbB3 and erbB41. Amplification of the erbB2/neu gene is found in many cancer types and its overexpression is correlated with a poor prognosis for breast and ovarian cancer patients2. Investigation of the biology of erbB2 led to the identification of a family of ligands termed neuregulins which included the neu-differentiation factors3,4, the heregulins5, a ligand with acetylcholine-receptor-inducing activity6 and glial growth factor7. Several lines of evidence suggest that heterodimerization of erbB2 with other erbB receptors is required for neuregulin signalling1. Here we investigate the developmental role of erbB2 in mammalian development in mice carrying an erbB2 null allele. We find that mutant embryos die before Ell, probably as a result of dysfunctions associated with a lack of cardiac trabeculae. Development of cranial neural-crest-derived sensory ganglia was markedly affected. Dil retrograde tracing revealed that the development of motor nerves was also compromised. Our results demonstrate the importance of erbB2 in neural and cardiac development.
Article
Full-text available
Interleukin-3 (IL-3)-dependent murine 32D cells do not detectably express epidermal growth factor receptors (EGFRs) and do not proliferate in response to EGF, heregulin (HRG) or other known EGF-like ligands. Here, we report that EGF specifically binds to and can be crosslinked to 32D transfectants co-expressing ErbB2 and ErbB3 (32D.E2/E3), but not to transfectants expressing either ErbB2 or ErbB3 individually. [125I]EGF-crosslinked species detected in 32D.E2/E3 cells were displaced by HRG and betacellulin (BTC) but not by other EGF-like ligands that were analyzed. EGF, BTC and HRG also induced receptor tyrosine phosphorylation, activation of downstream signaling molecules and proliferation of 32D.E2/E3 cells. 32D transfectants were also generated which expressed an ErbB3–EGFR chimera alone (32D.E3-E1) or in combination with ErbB2 (32D.E2/E3-E1). While HRG stimulation of 32D.E3-E1 cells resulted in DNA synthesis and receptor phosphorylation, EGF and BTC were inactive. However, EGF and BTC were as effective as HRG in mediating signaling when ErbB2 was co-expressed with the chimera in the 32D.E2/E3-E1 transfectant. These results provide evidence that ErbB2/ErbB3 binding sites for EGF and BTC are formed by a previously undescribed mechanism that requires co-expression of two distinct receptors. Additional data utilizing MDA MB134 human breast carcinoma cells, which naturally express ErbB2 and ErbB3 in the absence of EGFRs, supported the results obtained employing 32D cells and suggest that EGF and BTC may contribute to the progression of carcinomas that co-express ErbB2 and ErbB3.
Article
Full-text available
In Metazoans a number of cellular functions are controlled by receptor tyrosine kinases (RTKs) during development and in postnatal life. The execution of these programs requires that signals of adequate strength are delivered for the appropriate time within precise spatial boundaries. Several RTK inhibitors have been identified in invertebrate and mammalian organisms. Because they are involved in tuning and termination of receptor signals, negative regulators of RTK activity fulfill a fundamental function in the control of receptor signaling. © 2001 Federation of European Biochemical Societies.
Article
Full-text available
ErbB-2 becomes rapidly phosphorylated and activated following treatment of many cell lines with epidermal growth factor (EGF) or Neu differentiation factor (NDF). However, these factors do not directly bind ErbB-2, and its activation is likely to be mediated via transmodulation by other members of the type I/EGF receptor (EGFR)-related family of receptor tyrosine kinases. The precise role of ErbB-2 in the transduction of the signals elicited by EGF and NDF is unclear. We have used a novel approach to study the role of ErbB-2 in signaling through this family of receptors. An ErbB-2-specific single-chain antibody, designed to prevent transit through the endoplasmic reticulum and cell surface localization of ErbB-2, has been expressed in T47D mammary carcinoma cells, which express all four known members of the EGFR family. We show that cell surface expression of ErbB-2 was selectively suppressed in these cells and that the activation of the mitogen-activated protein kinase pathway and p70/p85S6K, induction of c-fos expression, and stimulation of growth by NDF were dramatically impaired. Activation of mitogen-activated protein kinase and p70/p85S6K and induction of c-fos expression by EGF were also significantly reduced. We conclude that in T47D cells, ErbB-2 is a major NDF signal transducer and a potentiator of the EGF signal. Thus, our observations demonstrate that ErbB-2 plays a central role in the type I/EGFR-related family of receptors and that receptor transmodulation represents a crucial step in growth factor signaling.
Article
Full-text available
In the present study we demonstrate that erbB-3 and erbB-2 cooperate in neoplastic transformation. Under conditions in which neither gene alone induced transformation, they readily transformed NIH3T3 cells if co-expressed. Furthermore, at high expression levels of ErbB2 which cause transformation, ErbB3 enhanced focus formation by one order of magnitude. Synergy required an intact ErbB2 extracellular domain and tyrosine kinase activity. Cooperation between ErbB3 and ErbB2 involved heterodimerization and increased tyrosine phosphorylation of ErbB3. Signaling by the heterodimer resulted in increased PI 3-kinase recruitment as well as quantitative and qualitative differences in substrate phosphorylation. Evidence for signaling by an active ErbB3-ErbB2 heterodimer in four mammary tumor cell lines indicated relevance of this mechanism for human neoplasia. Our detection of the NDF/heregulin transcript in NIH3T3 cells implicates an autocrine loop involving this ligand in signaling by the ErbB3-ErbB2 heterodimer in the model system, whereas heregulin-independent mechanisms likely exist for cooperative signaling by ErbB3 and ErbB2 chronically activated in some human mammary carcinomas.
Article
The serine/threonine protein kinase PKB (also known as Akt) is thought to be a key mediator of signal transduction processes. The identification of PKB substrates and the role PKB phosphorylation plays in regulating these molecules have been a major focus of research in recent years. A recently developed motif-profile scoring algorithm that can be used to scan the genome for potential PKB substrates is therefore a useful tool, although additional considerations, such as the evolutionary conservation of the phosphorylation site, must also be taken into account. Recent evidence indicates that PKB plays a key role in cancer progression by stimulating cell proliferation and inhibiting apoptosis and is also probably a key mediator of insulin signalling. These findings indicate that PKB is likely to be a hot drug target for the treatment of cancer, diabetes and stroke. There are, however, a number of pitfalls of methodologies currently employed to study PKB function, and therefore caution should be used in interpretation of such experiments.
Article
Recent years have witnessed tremendous growth in the epidermal growth factor (EGF) family of peptide growth factors and the ErbB family of tyrosine kinases, the receptors for these factors. Accompanying this growth has been an increased appreciation for the roles these molecules play in tumorigenesis and in regulating cell proliferation and differentiation during development. Consequently, a significant question has been how diverse biological responses are specified by these hormones and receptors. Here we discuss several characteristics of hormone-receptor interactions and receptor coupling that contribute to specificity: 1) a single EGF family hormone can bind multiple receptors; 2) a single ErbB family receptor can bind multiple hormones; 3) there are three distinct functional groups of EGF family hormones; 4) EGF family hormones can activate receptors in trans, and this heterodimerization diversifies biological responses; 5) ErbB3 requires a receptor partner for signaling; and 6) ErbB family receptors differentially couple to signaling pathways and biological responses. BioEssays 20:41–48, 1998. © 1998 John Wiley & Sons, Inc.
Article
Overexpression of the erbB-2/neu gene is frequently detected in human cancers. When overexpressed in NIH 3T3 cells, the normal erbB-2 product, gp185erbB-2, displays potent transforming ability as well as constitutively elevated levels of tyrosine kinase activity in the absence of exogenously added ligand. To investigate the basis for its chronic activation we sought evidence of a ligand for gp185erbB-2 either in serum or produced by NIH 3T3 cells in an autocrine manner. We demonstrate that a putative ligand for gp185erbB-2 is not contained in serum. Chimeric molecules composed of the extracellular domain of gp185erbB-2 and the intracellular portion of the epidermal growth factor receptor (EGFR) did not show any transforming ability or constitutive autophosphorylation when they were expressed in NIH 3T3 cells. However, they were able to transduce a mitogenic signal when triggered by a monoclonal antibody directed against the extracellular domain of erbB-2. These results provide evidence against the idea that an erbB-2 ligand is produced by NIH 3T3 cells. Furthermore, we obtained direct evidence of the constitutive enzymative activity of gp185erbB-2 by demonstrating that the erbB-2 kinase remained active in a chimeric configuration with the extracellular domain of the EGFR, in the absence of any detectable ligand for the EGFR. Thus, under conditions of overexpression, the normal gp185erbB-2 is a constitutively active kinase able to transform NIH 3T3 cells in the absence of ligand.
Article
The epidermal growth factor (EGF) receptor (EGFR) can efficiently couple with mitogenic signaling pathways when it is transfected into interleukin-3 (IL-3)-dependent 32D hematopoietic cells. When expression vectors for erbB-2, which is structurally related to EGFR, or its truncated counterpart, delta NerbB-2, were introduced into 32D cells, neither was capable of inducing proliferation. This was despite overexpression and constitutive tyrosine kinase activity of their products at levels associated with potent transformation of fibroblast target cells. Thus, EGFR and erbB-2 couple with distinct mitogenic signaling pathways. The region responsible for the specificity of intracellular signal transduction was localized to a 270-amino acid stretch encompassing their respective tyrosine kinase domains. Thus, tissue- or cell-specific regulation of growth factor receptor signaling can occur at a point after the initial interaction of growth factor with receptor. Such specificity in signal transduction may account for the selection of certain oncogenes in some malignancies.