<iframe src="https://www.googletagmanager.com/ns.html?id=GTM-KCV32QR" height="0" width="0" style="display:none;visibility:hidden">

Functional isolation of activated and unilaterally phosphorylated heterodimers of ERBB2 and ERBB3 as scaffolds in ligand-dependent signaling

Edited by John Kuriyan, University of California, Berkeley, CA, and approved May 26, 2012 (received for review January 24, 2012)
June 25, 2012
109 (33) 13237-13242
Commentary
Pari passu dimers of dimers
Mark X. Sliwkowski

Abstract

The EGFR (ERBB) family provides a model system for receptor signaling, oncogenesis, and the development of targeted therapeutics. Heterodimers of the ligand-binding–deficient ERBB2 (HER2) receptor and the kinase impaired ERBB3 (HER3) create a potent mitogenic signal, but the phosphorylation of ERBB2 in this context presents a challenge to established models of phosphorylation in trans. Higher order complexes of ERBB receptors have been observed biophysically and offer a theoretical route for ERBB2 phosphorylation, but it is not clear whether such complexes provide functionality beyond the constituent dimers. We now show that a previously selected inhibitory RNA aptamer that targets the extracellular domain (ECD) of ERBB3 acts by sterically disrupting these higher order interactions. Ligand binding, heterodimerization, phosphorylation of ERBB3, and AKT signaling are only minimally affected, whereas ERBB2 phosphorylation and MAPK signaling are selectively inhibited. The mapping of the binding site and creation of aptamer-resistant point mutants are consistent with a model of side-by-side oriented heterodimers to facilitate proxy phosphorylation, even at very low endogenous levels of receptors (below 10,000 receptors per cell). Additional modes of signaling with relevance to pathological ERBB expression states emerge at high receptor levels. Hence, higher order complexes of nonoverexpressed ERBB receptors are an integral and qualitatively distinct part of normal ERBB2/ERBB3 signaling. This mechanism of activation has implications for models of allosteric control, specificity of interactions, possible mechanisms of cross-talk, and approaches to therapeutic intervention that at present often generate experimental and clinical outcomes that do not reconcile with purely canonical, dimer-based models.
Biochemical and structural analysis of the EGFR or ERBB (ErbB) family of receptor tyrosine kinases has provided a wealth of molecular details that have contributed significantly to our understanding of cell surface signaling and its deregulation in a broad range of diseases. The longstanding mechanistic model of receptor tyrosine phosphorylation in trans within ligand-activated dimers has undergone significant expansion in recent years. Beyond the regulation at the level of dimers, higher order clustering phenomena have been reported both for inactive and active receptor states (18). However, a critical and so far inaccessible question has been whether higher order complexes create qualitatively distinct signals that cannot emanate from dimers. The functional asymmetry of the closely related ERBB2/ERBB3 heterodimer presents an opportunity for experimental dissection but also a long-standing challenge to existing signaling models. ERBB2 is an orphan receptor that tyrosine phosphorylates its heterodimerization partners. ERBB3 is itself catalytically impaired but binds ligand, and its kinase domain allosterically activates its partners (9). This functional asymmetry is underscored by the fact that the majority of MAPK signaling emanates from ERBB2, whereas ERBB3 dominates signaling through the PI3K/AKT pathway. Ligand specificity sets neuregulin (NRG)-activated ERBB2/ERBB3 functionally apart from the EGF-activated ERBB2/EGFR.
Paradoxically, the neuregulin-dependent activation of ERBB2/ERBB3 heterodimers results in very efficient phosphorylation of ERBB2, making ERBB2/ERBB3 the most mitogenic receptor pair in the ERBB family (10, 11). However, the phosphorylation mechanism is not understood. Recent studies have shown that ERBB3 does bind ATP (12, 13) and features a low but specific catalytic activity in vitro (12). However, the ATP-bound state surprisingly retains a conformation associated with an inactive state (13). The in vitro phosphoryl transfer is very inefficient compared with EGFR and resistant to existing kinase inhibitors of ligand-induced ERBB2/ERBB3 signaling in a cell culture setting (12). Hence, the primary function of ATP binding by ERBB3 remains an open question. Alternatively, phosphorylation of the C-terminal tail of ERBB2 could conceivably occur in an intramolecular fashion after allosteric activation has occurred in trans.
Phosphorylation between activated ERBB2 receptors within higher order complexes, previously termed proxy phosphorylation (14), provides an alternative conceptual framework. Proxy phosphorylation would presumably involve ligand-activated heterodimers and is therefore mechanistically distinct from the ligand-independent autophosphorylation of overexpressed ERBB2, which is generally thought to involve transient homodimerization. However, whereas various biophysical studies have shown that higher order clusters exist, this hypothesis has so far eluded direct experimental evaluation. More importantly, we so far have no direct evidence that higher order clusters, as detected biophysically for the ERBB receptor family, create unique functionality that goes beyond an assembly of functionally autonomous dimers. A mechanistic understanding of the functional role of higher order complexes was so far limited by our inability to dissect the proxy and canonical dimer component in a molecularly defined manner.
The starting point for our analysis was the inhibition of ligand-induced signaling by A30, a SELEX-derived RNA aptamer against ERBB3 extracellular domains (ECDs). It was characterized initially in vitro and in cell culture with respect to its ERBB3-binding specificity and ability to inhibit NRG-mediated growth stimulation (15). However, subsequent studies of its mode of inhibition did not reconcile with conventional dimerization models. The present analysis reveals instead that A30 blocks the interactions of activated but functionally nonautonomous heterodimers, thus preventing higher order complexes that are essential for the proxy phosphorylation of ERBB2, even at low endogenous receptor levels. Hence, higher order association and proxy phosphorylation do not only occur, but are in fact indispensable and qualitatively unique components of ERBB2/ERBB3 signaling.

Results

The ERBB3-Directed Aptamer Selectively Inhibits Ligand-Induced Phosphorylation of ERBB2 but Not the Formation of Active Heterodimers.

Most available cellular data on ERBB receptors are derived from overexpression systems. This finding reflects both technical reasons and the high importance of receptor amplification in cancer. Comparatively little is known about the rules that govern ERBB signaling at the very low receptor levels found in all but hematopoetic cells. Low endogenous ERBB levels play a critical role in the development of several organs and subsequent cardiac maintenance. To better understand both “normal” ERBB signaling and its distortion in a cancer setting, we studied ERBB2/ERBB3 signaling in MCF7 cells, a well established model system for non-ERBB2 overexpressing breast cancers with FACS-confirmed expression levels of both receptors at or below 10,000 receptors per cell. In contrast to autophosphorylation in ERBB2 overexpressing cells, ERBB2/ERBB3 signaling in MCF7 is very sensitive to nanomolar ligand concentrations. The inhibition of ligand binding or receptor heterodimerization is expected to interfere with the activation of both receptors. However, a TyrP immunoprecipitation of ligand-treated samples shows that the lower levels of remaining tyrosine phosphorylation after A30 treatment reside almost exclusively on ERBB3 (Fig. 1A). For cells with low receptor levels, mild detergent and ligand treatment can significantly influence the recovery of receptors from the membrane compartment. We confirmed the preferential inhibition of ERBB2 phosphorylation by direct SDS lysis and immunoblotting for ERBB2 pTyr1248 (Fig. 1B), pTyr1139 (e.g., Fig. 1C), and pTyr1289 for ERBB3. This approach confirms a selective inhibition of ERBB2 phosphorylation. Whereas ERBB2 phosphorylation is almost completely inhibited, ligand-induced coimmunoprecipitation of ERBB2 and ERBB3 is retained as is the phosphorylation of ERBB3 in those complexes (Fig. 1C). Compared to deliberately partial inhibition by pertuzumab (2C4), an antibody that targets the dimerization loop of ERBB2 and blocks heterodimerization (16, 17), inhibition by A30 is distinctly more asymmetric in nature (Fig. 1D). This asymmetric inhibition by A30 translates into downstream signaling (Fig. 2). The inhibition of phosphorylation at ERBB2/Y1139, one of several sites implicated in the initiation of MAPK activation, correlates with a pronounced inhibition of MAPK signaling. This is in contrast to a lack of inhibition for the ERBB3-dependent activation of AKT. Combined, these findings suggest that A30 exerts its primary inhibitory effect at a level other than ligand binding to ERBB3, heterodimerization, or the allosteric activation of ERBB2 in heterodimers. Instead, the asymmetric inhibition of ERBB2 phosphorylation creates an activated heterodimer with equally asymmetric downstream signaling properties.
Fig. 1.
A30 preferentially inhibits ERBB2 phosphorylation but not heterodimerization with ERBB3. Stimulation of MCF7 cells with NRG (10 nM, 15 min) in the presence or absence of A30. Note that whereas ligand stimulation rapidly alters receptor levels and receptor detection in parental cells, receptor levels in ligand-treated samples with or without A30 are equivalent. (A) Inhibition of total tyrosine phosphorylation (Upper) and relative inhibition of ERBB2 and ERRB3 phosphorylation (receptor IB after TyrP-IP under complex disrupting conditions, Lower). (B) A30 inhibition probed directly by site-specific tyrosine phosphorylation (ERBB2/pTyr1248 and ERBB3/pTyr1289). (C) Coimmunoprecipitation of ERBB2 and ERBB3 and the pERBB3/(pTyr1289) in recovered complexes are minimally impacted compared with ERBB2 phosphorylation (pTyr1139) inhibition. (D) In contrast to A30, the partial inhibition by the dimerization blocking anti-ERBB2 antibody pertuzumab (2C4) blocks the phosphorylation of ERBB2 (pTyr1248) and ERBB3 (pTyr1289) proportionally.
Fig. 2.
The preferential inhibition of ERBB2 correlates with a pronounced inhibition of MAPK activation, whereas both ERBB3 phosphorylation and AKT activation are largely resistant to A30. Note that pERBB3 initially increases at low aptamer concentrations to drop to a plateau of 20% inhibition. MCF7 parental cells were stimulated with 10 nM NRG for 10 min in the presence of the indicated concentrations of A30.

Where Does A30 Bind Relative to Known Interacting Surfaces of ERBB Receptors?

A 43-nucleotide minimal aptamer segment (mA30) is sufficient for specific ERBB3 binding and targeted photo cross-linking (18). However, whereas mA30 blocks the inhibitory properties of the 78-nucleotide-long A30, it is a much less potent inhibitor of ERBB2 phosphorylation (Fig. S1). These early observations implicated steric interference with complexes beyond the heterodimer. For an initial approximation of the portion of the ERBB3 ECD involved in A30 binding, we photo crosslinked thiouracil substituted A30 and insect cell expressed extracellular domains of ERBB3 (Fig. S2). This in vitro study with purified components narrowed the putative binding region for A30 to a segment near the junction of domain III and IV that is accessible in both the extended and tethered conformation. Because no structure is currently available for the ERBB3 receptor ECD in the extended conformation, we used the available structure of tethered ERBB3 for the initial selection of residues for mutagenesis. A striking feature of the domain III–IV junction is its neutral-to-positive surface charge (Fig. 3A). Whereas charge-based interactions alone cannot account for the high specificity and nanomolar affinity of A30, electrostatics exert strong “steering” forces during a SELEX procedure. Combined, these considerations further narrowed the list of candidate residues, making an analysis of cell-surface–expressed receptors by mutagenesis feasible.
Fig. 3.
A30 targets the domain III–IV junction region. (A) Location of residues in the tethered conformation of ERBB3 (25) that were selected for mutagenesis on the basis of in vitro photo cross-linking data (Fig. S2). Front and back representations of the molecular surfaces are colored by surface charge. (B) The contiguous surface presented by H446/H447 and R471/R472 contributes to A30 binding. Charge inversion at R471/R472, located in the domain III–IV connecting loop, eliminates A30 binding. Measurements represent fluorescence intensity ratios for A30 binding per surface localized receptors within a uniform range of overall receptor levels. Readouts are shown in Fig. 3C. Data (with indicated SDs) are the average for 200–300 individual live cells carrying stably expressed ERBB3 constructs. (C) Representative fluorescence data for clusters of attached live cells showing loss of A30 binding to ERBB3–R471/472E.
For live cell binding studies, we focused on four locations where mutations were not likely to significantly alter the underlying backbone conformation yet have a significant impact on the immediate surrounding surface. Binding was measured on large numbers of individual live cells using a triple color fluorescent tag assay that quantifies receptor levels (C-terminal GFP fusion, green), surface expression (N-terminal FLAG epitope tag, blue), and aptamer binding (mA30-biotin and streptavidin-Texas Red) on MCF7 cells that stably express the mutant receptors in a large excess over endogenous ERBB3. Fig. 3B shows the ratio of aptamer binding to surface receptors for 200–300 individual cells. Whereas the removal of the positive surface charge at lysine 453 and arginine 456 results in a modest (10%) increase in A30 binding, the removal of two negative charges at glutamic acid 460 and 461 moderately diminishes binding (7%). Both differences are statistically significant at P < 0.01. More extensive and statistically highly significant inhibition of binding (P < 0.001) was observed after mutating histidines 446/447 (20%) or arginines 471/472 (22%) to alanines. Those four residues form a contiguous surface patch that is spatially close to glutamic acids 460 and 461. H446/H447 represent the C-terminal “cap” of domain III and R471/R472 is located directly in the loop region between domains III and IV. The R471/472 site was selected for charge reversal, resulting in a nearly complete loss of A30 binding (90% inhibition). Inhibition is independent of receptor density (Fig. S3). The impact of mutagenesis on fully adherent cells is shown in Fig. 3C, demonstrating the loss of A30 binding to the R471/472E mutant at comparable expression and cell surface presentation levels. The stably expressed R471/472E mutant displays wild-type equivalent responsiveness to ligand, both in terms of ERBB2 tyrosine phosphorylation and downstream activation of MAPK (Fig. 4A) but is resistant to inhibition by A30. This response pattern suggests that the mutagenesis impacts A30 binding but is alone insufficient to disrupt proxy phosphorylation.
Fig. 4.
(A) Loss of A30 binding creates functional but A30-resistant ERBB3 receptors. Both ERBB2 and MAPK phosphorylation are insensitive to A30 for R471/472E, whereas ligand responsiveness is unaltered. (B) Transient overexpression of ERBB3–Dendra (versus Dendra control) creates an increase in ligand-dependent ERBB3 phosphorylation (Tyr1289) that is A30 sensitive, whereas the A30-insensitive pERBB3 remains constant (relative pERBB3 = pERBB3/tubulin with the average arbitrary ratio for stimulated but uninhibited sample = 100%). (C) Inhibition of constitutive and ligand-dependent ERBB receptor phosphorylation in ERBB2 overexpressing BT474 breast cancer cells after treatment with A30 (100 nM), pertuzumab (25 μg/mL, 2C4) and ligand (10 nM NRG) as indicated above lanes. Cells were pretreated with A30 and 2C4 for 30 min were appropriate, followed by 10-min ligand stimulation.

Overexpression Shifts the Dynamics of Proxy Phosphorylation.

To evaluate whether the mechanism of signaling and inhibition by A30 is sensitive to receptor levels, we transiently overexpressed wild-type ERBB3 in MCF7 cells (Fig. 4B). Ligand-independent activation is minimally enhanced (measurable only at longer exposures), whereas the ligand-induced phosphorylation of ERBB3 shows both a pronounced increase and enhanced A30 sensitivity. While the A30 inhibition of ERBB3 phosphorylation at endogenous receptor levels plateaus at 20% (Fig. 2), the ligand-dependent phosphorylation of overexpressed ERBB3 is suppressed by ∼60%. By contrast, the level of A30-insensitive phosphorylation is almost constant, regardless of ERBB3 levels. Combined, these inhibitor and ligand responses suggest that upon overexpression, a shift in the mechanism of ERBB3 phosphorylation occurs toward a more A30-sensitive interaction.
Whereas levels of ERBB3 that are orders of magnitude higher than ERBB2 are unnatural, the inverse scenario is a hallmark of many ERBB2 overexpressing cancers. We therefore evaluated the impact of A30 and pertuzumab on the ERBB2-overexpressing breast cancer cell line BT474 (Fig. 4C) and SKBr3 (Fig. S4), each harboring well over one million ERBB2 receptors in large excess over ERBB3. The phosphorylation of ERBB3 but not ERBB2 is stimulated by ligand, and a pertuzumab concentration that achieves the maximal achievable inhibition (Fig. S5) is more effective than A30 in suppressing the ligand-induced phosphorylation of ERBB3. Arguably more relevant to the cancer cell scenario, each inhibitor suppresses constitutive ERBB3 phosphorylation only partially, but both act synergistically, resulting in almost complete suppression of constitutive ERBB3 phosphorylation. Whereas largely ineffective in the inhibition of ERBB2 autophosphorylation, A30 is also surprisingly effective in further enhancing the ability of pertuzumab to suppress the constitutive and lapatinib-sensitive (Fig. S6) phosphorylation of ERBB2.

Discussion

Our current study of the mode of action of A30 provides insights into the mechanism of ERBB signaling beyond conventional models of functionally autonomous heterodimers. A30 does not interfere with ligand binding to ERBB3, receptor heterodimerization, or AKT signaling. The largely unimpeded tyrosine phosphorylation of ERBB3, even within coimmunoprecipitated heterodimers, indicates that the allosteric activation of the kinase activity of ERBB2 by ERBB3 is also not inhibited. However, under these conditions, ERBB2 tyrosine phosphorylation and much of the associated MAPK signaling is inhibited. Hence, aptamer binding creates the heterodimers that one would have theoretically expected on the basis of the impaired kinase activity of ERBB3 and a mechanism of trans-phosphorylation.
For the ERBB2/ERBB3 system, both EM (5) and homo-FRET correlated anisotropy measurements (7) indicate that rather than a formation of dimers from monomers, activation involves a rearrangement of clusters. Depending on the receptor species, clusters may be defined by proximity or, as in the case of ligand-free ERBB3, through actual oligomerization. For EGFR, independent studies have demonstrated the ligand-dependent emergence of tetramers as a dominant species in a population (4). However, the mechanistic contribution of higher order complexes has so far been unclear, and studies have been limited to high expression levels out of necessity. Assuming catalytically fully inactive ERBB3, Epstein and colleagues proposed proxy phosphorylation of ERBB2 in the past as a logical conclusion on the basis of the understanding of the individual receptor properties (14). However, a study of the intermediate building blocks had been out of reach. The current study uniquely demonstrates through inhibitor-driven experimental dissection, that self-standing ERBB2/ERBB3 heterodimers constitute activated signaling units, but with signaling properties that are imbalanced with regard to the classic, two-legged MAPK and AKT signaling that is a well-established hallmark of ERBB2/ERBB3 activation in a cellular setting. Hence, higher order complexes do not only exist for ERBB2/ERBB3, they are a basic mechanistic component needed to qualitatively create the ERBB2/ERBB3 signaling as we know it. Furthermore, these secondary interactions are crucial for normal signaling, even at low endogenous receptor levels that have so far been out of reach for the direct biophysical studies that first reported on the existence of higher order clusters.
What is the nature of the higher order complex? Our study was deliberately aimed at the nature of normal, ligand-dependent ERBB signaling at low endogenous receptor levels. Whereas the exact stoichiometry of the relevant higher order complexes is not readily accessible at low endogenous levels on technical grounds, we can address whether the mapping of the aptamer binding site unto the heterodimer is consistent with specific models derived from EGFR at high expression levels. For EGFR, FRET distance measurements suggest a poorly understood dependency of higher order interactions on the cellular context (19) and a likely side-by-side orientation of canonical dimers to form transient tetramers (6). To interrogate whether our inhibition data are consistent with such a side-by-side arrangement of heterodimers, we mapped the binding site of A30 on ERBB3. The residues that convey aptamer resistance are located distant from both the ligand-binding site and the canonical dimerization interface when mapped onto the crystallographically available tethered conformation. Within a homology model of the ERBB2/ERBB3 heterodimer, based on the crystal structure of ligand-bound dimeric EGFR ECDs (20, 21), the A30 binding site is located at the edge of a surface on ERBB3 with a distinctly neutral-to-positive surface charge. This patch stands out against the negative surface charge of ERBB2 and the remainder of ERBB3 (Fig. 5A). Within a side-by-side arrangement, such a charge distribution would favor transient interactions between dimers (Fig. 5B). A minimal version of A30 binds ERBB3 with high affinity but poorly inhibits ERBB2 phosphorylation. The tail region of full-length A30, which does not contribute to binding but almost doubles the size of the aptamer, confers a strong increase in inhibitory potency (Fig. S1). Hence, the mapping of the A30 binding site and functional inhibition data are consistent with a molecular model of side-by-side oriented heterodimers as the next, but not necessarily final level in the receptor assembly.
Fig. 5.
(A) Homology model of the ERBB2/ERBB3 heterodimer places the A30 interface on the side of the ERBB2/ERBB3 heterodimer, distant from the canonical dimerization interface and on the edge of a unique positively charged surface patch within ERBB3. (B) Cartoon of the heterodimer in the plasma membrane (PM) with indicated A30-binding site and charge complimentary interface on ERBB3 in dark blue. Cytoplasmic components are not drawn to scale. The structure-derived cartoon outline for the Top view highlights the interlocking canonical dimer interface, the charge complimentary interface on ERBB3 (blue), and the binding sites for A30 and NRG. (C) Ligand-enhanced coimmunoprecipitation of overexpressed V5 and Dendra tagged ERBB2 in the presence of low endogenous ERBB3 in MCF7 cells. The V5 epitope tag was immunoprecipitated. Ligand or A30 were added as indicated. (D) Proposed modes of receptor interactions and impact of A30 at low endogenous receptor levels compared with ERBB3 overexpression. Arrows indicate the directionality of tyrosine phosphorylation with arrow thickness indicating relative contributions.
Whereas ERBB2 alone does not bind ligand, the proposed model predicts a ligand-dependent increase in higher order complexes that contain two ERBB2 receptors. Although the proposed tetrameric state is expected to be relatively unstable, we tested the ability to coimmunoprecipitate two differentially tagged species of ERBB2 in a ligand-dependent but A30-sensitive manner. While heterodimers are readily recovered in CoIPs (e.g. Fig. 1C), the recovery of coimmunoprecipitated ERBB2 species is low. Nevertheless, we observed a significant ligand-dependent increase in coimmunoprecipitated Dendra and V5-tagged ERBB2 species (Fig. 5C). This CoIP was suppressed by the addition of A30. This observation matches data in a recent study by Ghosh et al. on the disruption of ERBB2 homodimers by trastuzumab (22). Whereas the authors used AP1510-mediated dimerization of ERBB2–FKBP fusion proteins, “negative” controls using NRG stimulation instead did not match AP1510 in forced dimerization but detected a 60% increase of ERBB2–ERBB2-containing complexes based on fluorescence proximity assays.
Our primary model system for biochemical dissection was MCF7 cells, which are often classified as ERBB2 negative by cancer overexpression standards. The mode of signaling we describe appears primarily geared toward maximizing activation at the low receptor levels present in most tissues. Upon overexpression, the mechanism of signaling appears to change qualitatively, and a second mode of activation emerges that uses both interfaces in a dimerization partner-independent manner (Fig. 4B). For overexpressed ERBB3, this mode of phosphorylation stands out through its sensitivity to A30 and strict dependency on ligand, suggesting that it makes use of interactions with a preassembled and preactivated heterodimeric scaffold. Fig. 5D compares the proposed flow of phosphorylation under conditions of balanced receptor levels versus overexpressed ERBB3.
A model in which ERBB2 can use two alterative interfaces for signaling also matches a comparative study of the ERBB2-directed, therapeutic antibodies pertuzumab and trastuzumab (Herceptin). It has long been known that both antibodies are not redundant but synergistic in targeting overexpressed ERBB2. Whereas trastuzumab is inefficient in interfering with ligand-induced heterodimerization (17), it is surprisingly more efficient than pertuzumab in blocking constitutive ERBB3 phoshphorylation (23). The large size of Herceptin and the spatially flexible nature of the segment of domain IV that it targets limited the mechanistic exploration of this observation. Our observed synergy of pertuzumab and A30 would involve A30 targeting the secondary interface in a similar manner than Herceptin, except on the side of ERBB3. In addition, A30 is a much smaller reagent that binds to a region of the ERBB3 receptor for which the placement in the receptor dimer is structurally definable by homology modeling. Interestingly, A30 amplified the ability of pertuzumab to block constitutive ERBB2 phosphorylation. This may suggest that at high levels of ERBB2, ERBB3 may not only be a target of constitutive phosphorylation and driver for enhanced cancer cell survival. Instead it may also serve ligand independently as a scaffold that facilitates efficient autophosphorylation of ERBB2 through two alternative approaches. Thus, whereas the primary objective of our study was the dissection of normal ERBB2/ERBB3 signaling, it has direct applicability to the distortion of ERBB signaling that results from overexpression and that cannot readily be explained within the confines of the canonical dimer model.
Regardless of whether tetramers are an endpoint or an intermediate for higher order complexes, it is important to realize that such association states are in competition with other states of clustering or oligomerization. Whereas ERBB2 does not form stable homooligomers, homo-FRET correlated anisotropy measurements indicate that it is organized in spatial clusters at high expression levels, and the average cluster size decreases upon addition of ligand, whereas the opposite appears true for EGFR (7). In the absence of ligand, ERBB3 is unique in its tendency to form actual oligomers that sequester the receptor away from its heterodimerization partners, and ERBB3 directed aptamers destabilize ERBB3 homooligomerization (18).
The lack of functional autonomy of the heterodimer does challenge our existing model for the regulation of ERBB receptors. What are the determinants of specificity in signal propagation once a ligand-bound dimer has established a preactivated interaction scaffold? Does allosteric kinase domain activation occur between dimers? Do allosteric acceptor and donor kinase domains trade places in “symmetric” EGFR dimers, or does a dedicated acceptor kinase domain need proxy phosphorylation in the reported tetrameric states to become phosphorylated itself? These questions require that the receptors in their entirety, including cytoplasmic kinase domain interactions, are integrated into more dynamic models of complex formation. Already existing data on clinical antibodies suggest that such expanded models may help in demystifying many of the observed and clinically highly relevant inconsistencies with traditional dimer-based models of signaling.

Materials and Methods

Reagents.

Antibodies were obtained from Santa Cruz Biotechnology (ERBB3/C17 and ERBB2/C18), Upstate Biotechnology (pTyr/4G10, pERBB2(Tyr1248), Cell Signaling Technologies [pERBB3(Tyr1289), pMAPK(Thr202/Tyr204), MAPK, pAKT(Ser473), and AKT], Epitomics [pERBB2(Tyr1139), Biogenex (ERBB2/CB11), Invitrogen (V5-HRP and V5), and Evrogen (Dendra)]. The purification of ERBB3–ECD constructs and the Trx–NRG fusion protein of the EGF-like domain of NRG1-β1 with thioredoxin have been described previously (3). Pertuzumab was provided by Mark Sliwkowski (Genentech Inc., South San Francisco, CA).

Cell Culture and Transfection.

MCF7 cells were maintained in RPMI-1640 [10% (vol/vol) FBS, 5% (vol/vol) CO2]. GFP fusion constructs of ERBB3 were expressed in the pFLAG-MYC-CMV-19 expression vector (Sigma), and cotransfected with puromycin resistance marker. Stably expressing cell lines were sorted for receptor–GFP fusions by FACS before expansion. For transient ERBB3 overexpression, MCF7 cells were transfected with either Dendra2 (Evrogen) or pFlag-ERBB3-Dendra2.

Aptamer Synthesis.

A30 was transcribed as previously described (15) using the RiboMAX large-scale RNA production system from Promega. For inhibition studies, the minimal aptamer (mA30) was generated using a PCR template with a shortened 5′ end. Synthetic minimal aptamer with a 3′ biotin was used for fluorescence microscopy studies.

Aptamer Inhibition of Ligand Stimulation.

Indicated cell lines at 60–70% confluency were treated with aptamer in RNaseOUT supplemented RPMI-1640 for 30 min at 37 °C, stimulated with Trx-NRG (10 nM) for 10 min at room temperature, and lysed in SDS sample buffer (57 mM Tris-HCl, 10% (vol/vol) glycerol, 3.3% (wt/vol) SDS, 0.17 mg/mL bromophenol blue, 1.7 mg/mL DTT) before Western blot analysis with the indicated antibodies. For coimmunoprecipitation, cells were lysed in mild lysis buffer [20 mM Tris, 137 mM NaCl, 1% (vol/vol) Triton X-100, 10% (vol/vol) glycerol, 5 mM EDTA, 1 mM sodium orthovanadate, 1 mM PMSF, 1× mixture]. Lysates were passed 10 times through a fine-gauge needle and incubated for 5 min at 37 °C. Cell debris was removed by centrifugation at 10,000 × g before immunoprecipitation of supernatants at room temperature for 2 h, five washes in mild lysis buffer and denaturation in 1× SDS sample buffer containing DTT. Inhibition studies in BT474 and SKBr3 cells were carried out as described above for MCF7 cells. For ERBB2 coimmunoprecipitation studies, MCF7 cells were cotransfected with ERBB2 constructs carrying either an N-terminal V5 epitope tag or C-terminal Dendra.

Molecular Modeling.

Primary sequences were aligned using clustalW. Structural models were created using Swiss Model builder (24) and surface charges were calculated in Swiss viewer using the default parameters. The EGF dimer was used as a structural framework to model the ERBB2/ERBB3 heterodimer. The C terminus of domain IV of ERBB3 (obtained from the tethered conformation) was projected into the N-terminal segment of domain IV of EGFR in the dimer structure.

Live Cell Aptamer Binding Studies.

MCF7 cells with stably overexpressed mutants of ERBB3–GFP were established after FACS sorting for GFP. Cells were dissociated with Versene (PBS/EDTA) and incubated for 10 min with 100 nM 3′ biotinylated minimal A30 and anti-FLAG(M2) antibody (Sigma) followed by 20 min of staining with streptavidin-Texas Red conjugate and Pacific-Blue–labeled secondary. Cells were allowed to settle and images of the uniformly rounded cells were acquired on a Zeiss Axiovert 200 M fluorescence microscope at low magnification in a rapid and automated manner using Openlab command script execution. Following density slicing, cells (<5%) with GFP signal outside two SDs of the mean (mainly due to well boundary effects) were excluded from the analysis. The ratio of surface localization (Pacific Blue) and aptamer binding (Texas Red) was analyzed as discussed. The statistical significance was analyzed using an uncoupled dual distribution t test, with P < 0.01 considered significant and P < 0.001 very significant.

Acknowledgments

We thank Dr. Vineet Gupta for detailed discussions and suggestions in the preparation of this manuscript. This work was supported by grants from the National Institutes of Health/National Cancer Institute (CA98881-05), Susan B. Komen Foundation (BCTR0504291), and the Braman Family Breast Cancer Institute.

Supporting Information

Supporting Information (PDF)
Supporting Information

References

1
TW Gadella, TM Jovin, Oligomerization of epidermal growth factor receptors on A431 cells studied by time-resolved fluorescence imaging microscopy. A stereochemical model for tyrosine kinase receptor activation. J Cell Biol 129, 1543–1558 (1995).
2
P Nagy, et al., Activation-dependent clustering of the erbB2 receptor tyrosine kinase detected by scanning near-field optical microscopy. J Cell Sci 112, 1733–1741 (1999).
3
R Landgraf, D Eisenberg, Heregulin reverses the oligomerization of HER3. Biochemistry 39, 8503–8511 (2000).
4
AH Clayton, et al., Ligand-induced dimer-tetramer transition during the activation of the cell surface epidermal growth factor receptor-A multidimensional microscopy analysis. J Biol Chem 280, 30392–30399 (2005).
5
S Yang, et al., Mapping ErbB receptors on breast cancer cell membranes during signal transduction. J Cell Sci 120, 2763–2773 (2007).
6
AH Clayton, SG Orchard, EC Nice, RG Posner, AW Burgess, Predominance of activated EGFR higher-order oligomers on the cell surface. Growth Factors 26, 316–324 (2008).
7
A Szabó, G Horváth, J Szöllosi, P Nagy, Quantitative characterization of the large-scale association of ErbB1 and ErbB2 by flow cytometric homo-FRET measurements. Biophys J 95, 2086–2096 (2008).
8
T Duke, I Graham, Equilibrium mechanisms of receptor clustering. Prog Biophys Mol Biol 100, 18–24 (2009).
9
X Zhang, J Gureasko, K Shen, PA Cole, J Kuriyan, An allosteric mechanism for activation of the kinase domain of epidermal growth factor receptor. Cell 125, 1137–1149 (2006).
10
PM Guy, JV Platko, LC Cantley, RA Cerione, KL Carraway, Insect cell-expressed p180erbB3 possesses an impaired tyrosine kinase activity. Proc Natl Acad Sci USA 91, 8132–8136 (1994).
11
LN Klapper, et al., The ErbB-2/HER2 oncoprotein of human carcinomas may function solely as a shared coreceptor for multiple stroma-derived growth factors. Proc Natl Acad Sci USA 96, 4995–5000 (1999).
12
F Shi, SE Telesco, Y Liu, R Radhakrishnan, MA Lemmon, ErbB3/HER3 intracellular domain is competent to bind ATP and catalyze autophosphorylation. Proc Natl Acad Sci USA 107, 7692–7697 (2010).
13
N Jura, Y Shan, X Cao, DE Shaw, J Kuriyan, Structural analysis of the catalytically inactive kinase domain of the human EGF receptor 3. Proc Natl Acad Sci USA 106, 21608–21613 (2009).
14
GC Huang, X Ouyang, RJ Epstein, Proxy activation of protein ErbB2 by heterologous ligands implies a heterotetrameric mode of receptor tyrosine kinase interaction. Biochem J 331, 113–119 (1998).
15
CH Chen, GA Chernis, VQ Hoang, R Landgraf, Inhibition of heregulin signaling by an aptamer that preferentially binds to the oligomeric form of human epidermal growth factor receptor-3. Proc Natl Acad Sci USA 100, 9226–9231 (2003).
16
MC Franklin, et al., Insights into ErbB signaling from the structure of the ErbB2-pertuzumab complex. Cancer Cell 5, 317–328 (2004).
17
DB Agus, et al., Targeting ligand-activated ErbB2 signaling inhibits breast and prostate tumor growth. Cancer Cell 2, 127–137 (2002).
18
E Park, R Baron, R Landgraf, Higher-order association states of cellular ERBB3 probed with photo-cross-linkable aptamers. Biochemistry 47, 11992–12005 (2008).
19
KB Whitson, JM Beechem, AH Beth, JV Staros, Preparation and characterization of Alexa Fluor 594-labeled epidermal growth factor for fluorescence resonance energy transfer studies: Application to the epidermal growth factor receptor. Anal Biochem 324, 227–236 (2004).
20
H Ogiso, et al., Crystal structure of the complex of human epidermal growth factor and receptor extracellular domains. Cell 110, 775–787 (2002).
21
TP Garrett, et al., Crystal structure of a truncated epidermal growth factor receptor extracellular domain bound to transforming growth factor alpha. Cell 110, 763–773 (2002).
22
R Ghosh, et al., Trastuzumab has preferential activity against breast cancers driven by HER2 homodimers. Cancer Res 71, 1871–1882 (2011).
23
TT Junttila, et al., Ligand-independent HER2/HER3/PI3K complex is disrupted by trastuzumab and is effectively inhibited by the PI3K inhibitor GDC-0941. Cancer Cell 15, 429–440 (2009).
24
T Schwede, J Kopp, N Guex, MC Peitsch, SWISS-MODEL: An automated protein homology-modeling server. Nucleic Acids Res 31, 3381–3385 (2003).
25
HS Cho, DJ Leahy, Structure of the extracellular region of HER3 reveals an interdomain tether. Science 297, 1330–1333 (2002).

Information & Authors

Information

Published in

Go to Proceedings of the National Academy of Sciences
Go to Proceedings of the National Academy of Sciences
Proceedings of the National Academy of Sciences
Vol. 109 | No. 33
August 14, 2012
PubMed: 22733765

Classifications

Submission history

Published online: June 25, 2012
Published in issue: August 14, 2012

Keywords

  1. oligomers
  2. proxy activation

Acknowledgments

We thank Dr. Vineet Gupta for detailed discussions and suggestions in the preparation of this manuscript. This work was supported by grants from the National Institutes of Health/National Cancer Institute (CA98881-05), Susan B. Komen Foundation (BCTR0504291), and the Braman Family Breast Cancer Institute.

Notes

This article is a PNAS Direct Submission.
See Commentary on page 13140.

Authors

Affiliations

Qian Zhang
Department of Biochemistry and Molecular Biology, Sylvester Comprehensive Cancer Center, Miller School of Medicine, University of Miami, Miami, FL 33136; and
Euisun Park
Department of Medicine, Hematology–Oncology, and Biological Chemistry, University of California, Los Angeles, CA 90095
Present address: SK Chemicals, Seongnam-si, Gyeonggi-do 463-400, Korea.
Kian Kani
Department of Medicine, Hematology–Oncology, and Biological Chemistry, University of California, Los Angeles, CA 90095
Present address: Center for Applied Molecular Medicine, University of Southern California, Los Angeles, CA 90033.
Ralf Landgraf3 [email protected]
Department of Biochemistry and Molecular Biology, Sylvester Comprehensive Cancer Center, Miller School of Medicine, University of Miami, Miami, FL 33136; and
Department of Medicine, Hematology–Oncology, and Biological Chemistry, University of California, Los Angeles, CA 90095

Notes

3
To whom correspondence should be addressed. E-mail: [email protected].
Author contributions: Q.Z., E.P., and R.L. designed research; Q.Z., E.P., and R.L. performed research; K.K. contributed new reagents/analytic tools; Q.Z., E.P., and R.L. analyzed data; and Q.Z. and R.L. wrote the paper.

Competing Interests

The authors declare no conflict of interest.

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Citation statements




Altmetrics

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Get Access

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to get full access to it.

    Single Article Purchase

    Functional isolation of activated and unilaterally phosphorylated heterodimers of ERBB2 and ERBB3 as scaffolds in ligand-dependent signaling
    Proceedings of the National Academy of Sciences
    • Vol. 109
    • No. 33
    • pp. 13133-13464

    Media

    Figures

    Tables

    Other

    Share

    Share

    Share article link

    Share on social media