Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Jan 12.
Published in final edited form as: Biochem Biophys Res Commun. 2006 Nov 16;352(2):397–403. doi: 10.1016/j.bbrc.2006.11.040

Heterodimerization of the α and β isoforms of the human Thromboxane receptor enhances isoprostane signalling.

Stephen J Wilson 1, Kevin McGinley 1, Albert J Huang 1, Emer M Smyth 1,
PMCID: PMC1766557  NIHMSID: NIHMS15449  PMID: 17134677

SUMMARY

Isoprostanes are free radical catalyzed products of arachidonic acid that are elevated in pro-oxidant disease states. Two isoprostanes, 8-isoprostaglandin F (iPFIII) and 8-isoprostaglandin E2 (iPE2III) act at the receptor for thromboxane A2 (the TP) to mediate pro-atherogenic effects in vivo. We confirmed dimerization of the human TP isoforms, TPα and TPβ, and determined the impact on isoprostane signaling. No overt changes in ligand binding at the TP were observed as a result of TPα/TPβ coexpression. The response to iPFIII or iPE2III was enhanced in HEK293 cells stably coexpressing TPα and TPβ, as measured by inositol phosphate generation or intracellular calcium mobilization, relative to cells expressing TPα or TPβ individually. In contrast, the response to traditional thromboxane analogs was unaltered. Augmented isoprostane signaling was similarly observed in HEK 293 cell transiently transfected with TPα and TPβ. These results indicate that TPα/TPβ dimerization enhances isoprostane-mediated signal transduction.

Keywords: Thromboxane, G-protein coupled receptor, Heterodimer, Isoprostanes

INTRODUCTION

Isoprostanes are free radical catalyzed products of arachidonic acid that are elevated in many oxidant-related disease states including atherosclerosis [1]. Isoprostane levels are elevated within developing atherosclerotic lesions in mice and normalization of isoprostane levels correlates with disease regression [2]. At least two isoprostanes, 8-isoprostaglandin F (iPFIII) and 8-isoprostaglandin E2 (iPE2III), act at the receptor for thromboxane A2 (the TP) in vivo to induce vasoconstriction and platelet aggregation [3]. Significantly, antagonism of TP, but not aspirin-mediated inhibition of TxA2 synthesis, reduced atherosclerosis in mice, suggesting that mediators other than TxA2, possibly isoprostanes, act at the TP to propagate the disease [4]. Additionally, increased lesion formation in response to exogenous iPFIII, was observed in two mouse models of atherosclerosis and was prevented by a TP antagonist [5]. Thus, isoprostanes represent both a marker and mediator of vascular disease.

The TP is a G-protein coupled receptor (GPCR) encoded by a single gene [6] that is alternatively spliced at the carboxyl terminus resulting in two isoforms TPα (343) and TPβ (407). The two isoforms are identical for the first 328 amino acids diverging only in their C-terminal regions[7]. Both TPα and TPβ are coupled primarily to phospholipase C, while the former may activate and the latter inhibit, adenylyl cyclase (AC) activity[7]. Additionally, dissimilar patterns of constitutive and agonist-dependent internalization of the membrane expressed TPα and TPβ have been observed, reportedly because of their differential interactions with regulatory proteins Rab11 and arrestin[8, 9]. However despite advances elucidating their individual functional and regulatory characteristics, the distinction between TPα and TPβ with respect to their ultimate physiological or pathophysiological roles remains unclear.

Appreciation of the complexity inherent in GPCR signaling increased markedly with the recent realization that receptors from closely related and distinct subfamilies can interact physically forming heterodimers [10]. GPCR heterodimerization results frequently in substantial modifications of receptor function[10]. Indeed, we demonstrated previously that heterodimerization of TPα with the human receptor for prostacyclin (the IP) dramatically altered TP signaling [11]. More recently, TPα and TPβ were shown to heterodimerize with consequent alterations in receptor regulation and signaling in response to a TxA2 analog [12, 13]. The vasopressor response to iPFIII was augmented in mice transgenic for TPβ expression in the vasculature [3] leading us to question the consequences of TPα/TPβ heterodimerization for isoprostane signaling.

In the present study we examined if TPα/TPβ heterodimerization altered receptor responsiveness to isoprostanes. We confirmed heterodimerization of co-expressed TPα/TPβ and found that coexpression significantly augmented the response to the isoprostanes, iPFIII and iPE2III, but not to traditional TxA2 analogs. These data suggest that the TPα/TPβ heterodimer may constitute a modified TP receptor with increased sensitivity to isoprostanes.

EXPERIMENTAL PROCEDURES

Epitope tagging of hIP and hTP

The 9 amino acid hemagglutinin epitope (HA; YPYDVPDYA), or 10 amino acid Myc epitope (EQKLISEEDL), were inserted between the N-terminal initiator methionine and the second amino acid of the hTPα or hTPβ to generate HAhTPα or MychTPβ. Generation of HAhTPα was as described previously [11]. To generate MychTPβ, 5′-oligonucleotides which contained 3 miscellaneous bases, 6 bases encoding a HindIII site, the 3 bases immediately 5′ of the initiator methionine, 3 bases encoding a methionine, the epitope tag coding sequence and 21 bases encoding amino acids 2–8 were generated. 3′-oligonucleotides were complementary to the receptor coding sequence downstream of a unique restriction site (Nar I site for hTPβ). Using the hTPβ cDNAs as templates, polymerase chain reactions were carried out to generate the 5′-MychTPβ fragments. The resulting products were cloned into PCR 2.1 (Qiagen, CA) and, following verification of the sequence, were excised using HindIII/Nar I. Using the same enzymes the 3′ fragment in pcDNA3.1 (Hygro for MychTPβ) were generated and the two receptor pieces were ligated to each other. The integrity of the splice site was verified by sequencing.

Cell Culture and Transfection

HEK293 cells (American Type Tissue Culture Collection, Rockville, MD) were maintained, and stably transfected with Myc-tagged TPβ or HA-tagged TPα, as previously described[11]. For transient transfections 50–80% confluent HEK293 were transfected using Fugene 6 (Roche Biochemicals, IN) according to the manufacturers instructions.

Co-immunoprecipitation

Immunoprecipitation was carried out as previously described [11]. Briefly HEK 293 cells stably expressing HAhTPα or both HAhTPα and MychTPβ, were treated with 3 mM DSP (Pierce Biotechnology, IL) for 30min and lysed in buffer A (150 mM NaCl; 1 mM EDTA; 20 mM Tris HCL [pH 8.0]; 10% glycerol and a cocktail of protease inhibitors) for 2 h at 4ºC. Supernatants were precleared with protein G-Sepharose. Anti-Myc G-Sepharose was prepared by adding 9 μg anti-Myc ascites per lysate to 10% protein G-Sepharose followed by 1 h rotation. MychTPβ was immunoprecipitated from precleared lysates by adding 150 μl of anti-Myc-protein G-Sepharose to each lysate and rotating for 16 h. Protein G was precipitated at 14,000 rpm for 1 min, washed three times with Buffer A and resuspended in 10 μl of sample buffer (Nupage). Immunobloting for HA was carried out using biotinylated anti-HA antibody (1:500) followed by peroxidase labeled streptavidin.

Membrane preparation and radioligand binding

Membrane preparation and radioligand binding were performed as described previously [11]. Displacement experiments were initiated by the addition of [3H]-SQ 29548 (40 nM) allowed to continue for 30 min at 30°C, prior to termination by the addition of 3 ml ice cold-wash buffer (10 mM HEPES, pH 7.4, 0.01% bovine serum albumin), followed by immediate filtration through GF/C filters. Radioactivity associated with the filters was quantified by scintillation counting. Nonspecific binding was measured in the presence of a 500-fold excess of unlabeled SQ 29548.

Inositol Phosphate Production

Inositol phosphate production was assessed as described previously [11]. Briefly, cells were labeled overnight with 2 μCi/ml [3H]-myoinositol, pretreated with 20 mM LiCl for 30min and stimulated appropriately. Total inositol phosphates were extracted with formic acid for 30 min at room temperature, neutralized using 5 M ammonia and recovered by anion exchange using Dowex 1-X8 AG anion exchange resin.

Calcium measurement

Intracellular calcium (Ca2+i) was assessed in stably transfected HEK cells, as previously described [14]. Cells were loaded for 1h at 37oC with 5uM Fura-2/AM (Molecular Probes) in phenol red free-RPMI 1640 culture media. Cells were washed, incubated for 5 min in PBS containing 1mM EDTA and 5mM EGTA and harvested. Cells were resuspended at 106 cells/ml RPMI1640 and fluorescence detected spectrophotometrically. Approximate values for Ca2+i were calculated using a Kd of 224 nM for Fura-2.

RESULTS

TPα/TPβ dimerization

We confirmed TPα/TPβ heterodimer formation which has been demonstrated previously [12, 13]; following immunoprecipitation of MycTPβ, immunobloting with anti-HA revealed the presence of HAhTPα in lysates derived from cells co-expressing HAhTPα and MychTPβ, but not lysates from cells expressing HAhTPα alone (Fig. 1A). The co-immunoprecipitated partner appeared primarily in the oligomeric form (Fig. 1A). This is not unexpected as nonreducing conditions were used for optimal visualization of oligomeric receptor. Under reducing conditions both monomeric and oligomeric forms of coimmunoprecipitated HAhTPα were observed (data not shown).

Fig. 1. HAhTPα and MychTPβ expression and dimer formation in stably expressing HEK cells.

Fig. 1

(A) TPα/TPβ-HEK or TPα-HEK cell lysates were subjected to co-immunoprecipitation using an anti-Myc antibody. Immunoprecipitates were resolved by 10% non-reducing SDS-PAGE and co-immunoprecipitated HAhTPα detected using an anti-HA antibody. Co-immunoprecipitation was only observed when HAhTPα and MychTPβ were coexpressed. (B) Lysates from cells stably expressing HAhTPα, MychTPβ or coexpressing both HAhTPα and MychTPβ, were resolved by SDS-PAGE, transferred to nitrocellulose, and immunoblotted with an anti-HA or anti-Myc antibody as described under “Experimental Procedures”. Molecular masses are in kDa. Western blots are representative of three independent experiments.

TPα/TPβ coexpression facilitates isoprostane-mediated inositol phosphate generation

To determine if TPα/TPβ dimerization specifically alters signaling in response to isoprostanes, we treated HEK 293 cells stably expressing comparable levels of TPα or TPβ individually or in combination (Fig. 1B), with increasing concentrations of the isoprostanes iPE2III or iPFIII, or with the TP specific agonists, IBOP and U46619. Treatment with the IBOP or U46619 induced a robust increase in inositol phosphate generation, which did not differ significantly between cells expressing TPα or TPβ individually, and those coexpressing both receptors (Fig. 2A and B). In contrast, inositol phosphate generation following treatment with the iPE2III or iPFIII, while less than that elicited by the TP-agonists, was significantly enhanced in cells co-expressing TPα and TPβ (Fig. 2C and D).

Fig. 2. Isoprostane-mediated inositol phosphate generation in cells stably coexpressing HAhTPα and MychTPβ.

Fig. 2

TPα-HEK (closed circles), TPβ-HEK (closed triangles) or TPα/TPβ-HEK (open diamonds) were treated with U46619 (A), IBOP (B), iPE2III (C) or iPFIII (D) for 10 min and inositol phosphates quantified as described under “Experimental Procedures”. Data are presented as Mean fold over basal ± S.E. from three to five experiments. iPE2III and iPF2αIII, p<0.05 TPα or TPβ vs TPα/TPβ as assessed by two way ANOVA.

To confirm these findings, HEK 293 cells were transiently transfected with a triple HA tagged (3xHA) hTPα (10 μg; UMR cDNA Resource Center), or cotransfected with 3xHAhTPα and MychTPβ (5μg of each construct), and inositol phosphate generation in response to IBOP or iPFIII examined. Similar to the results observed in stably transfected cells, the response to the isoprostane iPE2III (Fig. 3B), but not the TP-agonist IBOP (Fig. 3A), was significantly enhanced by TPα/TPβ coexpression (1.14 ± 0.06 vs 1.48 ± 0.8 fold over basal), demonstrating that this was not an artifact of stable receptor overexpression. Thus, it appears that coexpression of TPα and TPβ promotes specifically the TP-mediated response to isoprostanes.

Fig. 3. Isoprostane-mediated inositol phosphate generation in cells transiently transfected with HAhTPα and MychTPβ.

Fig. 3

HEK 293 cells transiently transfected with 10 μg of 3xHAhTPα (closed circles) or cotransfected with 5 μg each of 3xHAhTPα and MychTPβ (10μg total; open diamonds) were treated with U46619 (A) or iPE2III (B) for 1h min and inositol phosphates quantified as described under “Experimental Procedures”. Data are presented as Mean fold over basal ± S.E. from three to five experiments. iPE2III, p<0.0001 TPα vs TPα/TPβ as assessed by two way ANOVA.

TPα/TPβ coexpression facilitates isoprostane-mediated Ca2+i mobilization

We next sought to confirm enhanced isoprostane responsiveness of the TPα/TPβ coexpressing cells, using an alternate index of TP signaling, namely Ca2+i mobilization. Similar to the inositol phosphate experiments, the increase in Ca2+i observed in response to U46619 (Fig. 4A) did not differ significantly between TPα-HEK, TPβ-HEK and TPα/β-HEK cells. Once again, the response to iPFIII (Fig. 4B) or iPE2III (data not shown) was significantly enhanced in cells coexpressing TPα and TPβ.

Fig. 4. Isoprostane-mediated Ca2+i in cells stably coexpressing HAhTPα and MychTPβ.

Fig. 4

TPα-HEK (closed circles), TPβ-HEK (closed triangles) or TPα/TPβ-HEK (open diamonds) were treated with U46619 (A) or iPF2αIII (B) for 10 min and intracellular Ca2+i quantified as described under “Experimental Procedures”. Data are presented as Mean fold over basal ± S.E. from four to six experiments. iPF2αIII, p<0.0001 TPα or TPβ vs TPα/TPβ as assessed by two way ANOVA.

Receptor-ligand interactions

One of the characteristics frequently modified as a result of GPCR heterodimerization is the affinity of the individual receptors for their respective ligands [15,16]. We examined if altered ligand binding to the TPα/TPβ heterodimer might underlie the changes we observed in TP-signaling. Membranes from TPα-HEK, TPβ-HEK or TPα/β-HEK cells were labeled with the TP specific antagonist 3H-SQ 29548, which is used commonly to delineate the TP binding site [11, 17], and displacement by the traditional TP-ligands SQ 29548 and U46619, or the isoprostanes iPE2III and iPFIII, assessed.

No difference in the order of displacement was observed between TPα and TPβ: the rank order of potency for the displacing ligands was SQ 29548 > U46619 >> iPE2III >> iPFIII. Coexpression of TPα and TPβ did not alter affinity for SQ 29548 or U46619 as assessed by displacement (Fig. 5). Significantly, and in agreement with previous reports, iPE2III bound very weakly, while iPFIII did not bind at all to TPα or TPβ individually and this was unaltered by receptor coexpression. Thus, although isoprostane signaling was enhanced in TPα/β-HEK cells relative to cells expressing the receptors individually, this did not appear to reflect alterations in the ability of isoprostanes to ligate to the receptor.

Fig. 5. Effect of TPα/TPβ coexpression on TP ligand binding.

Fig. 5

Radioligand binding studies were carried out using membrane extracts from (A) TPα-HEK, (B) TPβ-HEK or (C) TPα/TPβ-HEK cells, labeled with [3H]-SQ 29548 as described under “Experimental Procedures”. Displacement by increasing concentrations of the TP ligands SQ 29548, U46619 and the isoprostanes iPE2III and iPF2αIII to displace [3H]-SQ 29548 from TP was examined. Data are mean ± S.E. percent of 3H-SQ 29548 displacement from 3 experiments.

DISCUSSION

Studies report ligation of both existing membrane [11, 18] and nuclear prostaglandin receptors [19] by isoprostanes. However, the concept that specific isoprostane receptors may exist has been suggested [20]. This is based partially on the apparent inability of isoprostanes to ligate or signal efficiently through either TP isoform in vitro [20], despite evidence that their in vivo actions are via the TP[3]. However, to date, no molecular evidence for the existence of a distinct receptor for isoprotanes has been found. GPCR heterodimerization frequently alters many of the indices used to characterize receptors, including ligand binding, signaling and regulation [10]. Indeed it has been proposed that the κ2 opioid receptor, described solely on a pharmacological basis, may in fact be the δ-κ opioid receptor heterodimer [21]. Therefore, we considered whether the phenomenon of GPCR heterodimerization may provide resolution of the pharmacological discrepancies in isoprostane studies.

Dimerization of the alpha and beta isoforms of the TP receptor was recently shown to mediate alterations in both receptor regulation and signaling [12, 13], however, the impact of this interaction on the response to isoprostanes was not examined. Previously, we demonstrated that dimerization of IP and TPα mediated significant alterations in TP signaling and may result in the formation of a novel isoprostane binding site [11]. We confirmed TPα/TPβ heterodimer formation (Fig. 1A) and assessed if the TPα/TPβ heterodimerization might facilitate isoprostane signaling in a similar manner.

In agreement with previous reports, the isoprostanes iPFIII and iPE2III stimulated TP-mediated inositol phosphate generation to a lesser extent than the TP-agonists IBOP and U46619 in cells expressing either TPα or TPβ individually (Fig. 2). However, while coexpression of TPα and TPβ did not alter significantly the response to TP activation by IBOP or U46619, inositol phosphate generation (Fig. 2A and B) following treatment with iPE2III and iPFIII was significantly enhanced in cells stably expressing both receptors (Fig. 2C and D). We examined Ca2+i mobilization as an additional index of TP signaling and found a similar selective enhancement of isoprostane signaling (Fig. 4).

The dose response curves for IBOP- and U46619-induced inositol phosphate generation and Ca2+i mobilization were super-imposable, arguing against significant differences in the total receptor expression level across the three stably transfected lines. Moreover no overt differences in TP expression were evident in the binding studies (Fig. 5) or by Western blot analysis (Fig. 1). Therefore, we did not assess directly by saturation ligand binding the total TP expression levels achieved in the three stably transfected cell lines. Instead, we sought to confirm our findings, and establish that variations in total receptor expression levels were not responsible for the enhanced TPα/TPβ isoprostane sensitivity, in an independent set of transiently transfected cells. HEK 293 cells were transiently transfected with 10μg of TPα or cotransfected with 5μg each of TPα and TPβ (10μg total), thereby ensuring equivalent expression levels between the two conditions. The response to TP agonists was lower in the transient model compared to the stably transfected cells, a reflection of the lower overall expression levels that are achieved in transiently compared with stably transfected cells. Importantly, as for the stably transfected cell models, the isoprostane-induced inositol phosphate response was selectively augmented in the TPα/β co-expressing cells (Fig. 3). Taken together these results suggest strongly that interaction between the two TP isoforms specifically promotes isoprostane signaling.

Reports now abound of GPCR heterodimers, formed upon receptor coexpression, mediating alterations in the signaling characteristics of the individual receptors. Dimerization of the β1 and β2 adrenergic receptors enhances cAMP formation in response to isoproternol and has been implicated in regulating cardiac contractility[22]. Indeed, we have previously demonstrated that heterodimerization of TP with the human prostacyclin receptor facilitates cAMP generation in response to TP activation [11]. As such, modulation of isoprostane signaling upon TPα/TPβ dimerization is not unprecedented and likely has physiological or pathophysiological consequences.

Many studies investigating the characteristics of newly discovered GPCR heterodimers describe an altered receptor-ligand-effector profile [10], suggesting that TPα/TPβ heterodimer formation may generate an altered receptor with distinct ligand binding characteristics. However, we observed no distinct differences in the ability of iPFIII or iPE2III to displace the TP antagonist SQ 29548 in membranes generated from TPα–, TPβ– or TPα/TPβ–HEK cells. Thus, despite the fact that both isoprostanes signaled more efficiently via the TPα/TPβ heterodimer we observed no elevation in their apparent ability to bind to the heterodimer. We previously reported that 3H-SQ 29548 did not fully occupy the binding site for iPE2III in the TPα/IP heterodimer [11]. It is likely that a similar sub-optimal labeling of the TPα/TPβ isoprostane binding site occurred in the present study giving rise to the inconsistency between the signaling and radioligand studies. Thus, while it is apparent that TPα/TPβ heterodimerization does not alter the “traditional” TP binding site, it remains possible that an altered binding site for isoprostanes is created. An alternative explanation may lie in reports that downstream G protein coupling is modified when GPCR heterodimerize. Thus, if the TPα/TPβ heterodimer were more efficiently coupled to Gq in the co-transfected cells elevated inositol phosphate and calcium signals might be expected. However, the absence of a similarly enhanced response to the traditional TP agonists IBOP and U46619 argues against this interpretation and for the generation of an altered binding site that is not adequately labeled by 3H-SQ 29548. Binding studies were attempted with 3H-iPFIII to resolve this issue further. However, because of the extremely low specific activity of the commercially available radioligand useful data was not obtained.

The mechanism through which heterodimerization alters the ability of the signaling of the constituent receptors is currently an area of intense investigation. Several studies have demonstrated that G-proteins exist in a stable complex with their cognate receptor [23, 24], suggesting that receptor-effector coupling is mediated through dynamic conformational changes. In accord with this concept, a specific ligand-receptor interaction may mediate agonist activation of one effector pathway, yet act as antagonist for alternate pathway [25], a property likely to be modulated by dimerization. As such the TPα/TPβ heterodimer may stabilize a receptor-G-protein-effector configuration which specifically facilitates the isoprostane-induced conformational changes required for TP activation of PLC.

Evidence from several studies implicates isoprostanes, acting at the TP, in the development of atherosclerosis [4, 5, 26]. However, the question remains, given their poor efficacy as TP agonists, how is this effected? GPCR dimerization is increasingly being implicated in the regulation of physiological and pathophysiological processes, including atherosclerosis. [27]. Our study demonstrates that TPα/TPβ dimerization enhances isoprostane-mediated signal transduction, and would likely maximize the deleterious effects of isoprostanes in the setting of cardiovascular disease.

Footnotes

This work was supported by NHLBI/National Institute of Health Grant HL066233

References

  • 1.Griendling KK, FitzGerald GA. Oxidative stress and cardiovascular injury: Part II: animal and human studies. Circulation. 2003;108:2034–2040. doi: 10.1161/01.CIR.0000093661.90582.c4. [DOI] [PubMed] [Google Scholar]
  • 2.Pratico D, Tangirala RK, Rader DJ, Rokach J, FitzGerald GA. Vitamin E suppresses isoprostane generation in vivo and reduces atherosclerosis in ApoE-deficient mice. Nat Med. 1998;4:1189–1192. doi: 10.1038/2685. [DOI] [PubMed] [Google Scholar]
  • 3.Audoly LP, Rocca B, Fabre JE, Koller BH, Thomas D, Loeb AL, Coffman TM, FitzGerald GA. Cardiovascular responses to the isoprostanes iPF(2alpha)-III and iPE(2)- III are mediated via the thromboxane A(2) receptor in vivo. Circulation. 2000;101:2833–2840. doi: 10.1161/01.cir.101.24.2833. [DOI] [PubMed] [Google Scholar]
  • 4.Cayatte AJ, Du Y, Oliver-Krasinski J, Lavielle G, Verbeuren TJ, Cohen RA. The thromboxane receptor antagonist S18886 but not aspirin inhibits atherogenesis in apo E-deficient mice: evidence that eicosanoids other than thromboxane contribute to atherosclerosis. Arterioscler Thromb Vasc Biol. 2000;20:1724–1728. doi: 10.1161/01.atv.20.7.1724. [DOI] [PubMed] [Google Scholar]
  • 5.Tang M, Cyrus T, Yao Y, Vocun L, Pratico D. Involvement of Thromboxane Receptor in the Proatherogenic Effect of Isoprostane F2{alpha}-III: Evidence From Apolipoprotein E- and LDL Receptor-Deficient Mice. Circulation. 2005;112:2867–2874. doi: 10.1161/CIRCULATIONAHA105.562223. [DOI] [PubMed] [Google Scholar]
  • 6.Nusing RM, Hirata M, Kakizuka A, Eki T, Ozawa K, Narumiya S. Characterization and chromosomal mapping of the human thromboxane A2 receptor gene. J Biol Chem. 1993;268:25253–25259. [PubMed] [Google Scholar]
  • 7.Narumiya S, FitzGerald GA. Genetic and pharmacological analysis of prostanoid receptor function. J Clin Invest. 2001;108:25–30. doi: 10.1172/JCI13455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Hamelin E, Theriault C, Laroche G, Parent J-L. The Intracellular Trafficking of the G Protein-coupled Receptor TP{beta} Depends on a Direct Interaction with Rab11. J Biol Chem. 2005;280:36195–36205. doi: 10.1074/jbc.M503438200. [DOI] [PubMed] [Google Scholar]
  • 9.Parent JL, Labrecque P, Driss Rochdi M, Benovic JL. Role of the differentially spliced carboxyl terminus in thromboxane A2 receptor trafficking: identification of a distinct motif for tonic internalization. J Biol Chem. 2001;276:7079–7085. doi: 10.1074/jbc.M009375200. [DOI] [PubMed] [Google Scholar]
  • 10.Devi LA. Heterodimerization of G-protein-coupled receptors: pharmacology, signaling and trafficking. Trends Pharmacol Sci. 2001;22:532–537. doi: 10.1016/s0165-6147(00)01799-5. [DOI] [PubMed] [Google Scholar]
  • 11.Wilson SJ, Roche AM, Kostetskaia E, Smyth EM. Dimerization of the human receptors for prostacyclin and thromboxane facilitates thromboxane receptor-mediated cAMP generation. J Biol Chem. 2004;279:53036–53047. doi: 10.1074/jbc.M405002200. [DOI] [PubMed] [Google Scholar]
  • 12.Laroche G, Lepine MC, Theriault C, Giguere P, Giguere V, Gallant MA, de Brum-Fernandes A, Parent JL. Oligomerization of the alpha and beta isoforms of the thromboxane A(2) receptor: Relevance to receptor signaling and endocytosis. Cell Signal. 2005 doi: 10.1016/j.cellsig.2005.02.008. [DOI] [PubMed] [Google Scholar]
  • 13.Sasaki M, Miyosawa K, Ohkubo S, Nakahata N. Physiological significance of thromboxane A(2) receptor dimerization. J Pharmacol Sci. 2006;100:263–270. doi: 10.1254/jphs.fp0050839. [DOI] [PubMed] [Google Scholar]
  • 14.Walsh MT, Foley JF, Kinsella BT. The alpha, but not the beta, isoform of the human thromboxane A2 receptor is a target for prostacyclin-mediated desensitization. J Biol Chem. 2000;275:20412–20423. doi: 10.1074/jbc.M907881199. [DOI] [PubMed] [Google Scholar]
  • 15.Jordan BA, Devi LA. G-protein-coupled receptor heterodimerization modulates receptor function. Nature. 1999;399:697–700. doi: 10.1038/21441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Pfeiffer M, Koch T, Schroder H, Laugsch M, Hollt V, Schulz S. Heterodimerization of somatostatin and opioid receptors cross-modulates phosphorylation, internalization, and desensitization. J Biol Chem. 2002;277:19762–19772. doi: 10.1074/jbc.M110373200. [DOI] [PubMed] [Google Scholar]
  • 17.Turek JW, Halmos T, Sullivan NL, Antonakis K, Le Breton GC. Mapping of a ligand-binding site for the human thromboxane A2 receptor protein. J Biol Chem. 2002;277:16791–16797. doi: 10.1074/jbc.M105872200. [DOI] [PubMed] [Google Scholar]
  • 18.Kunapuli P, Lawson JA, Rokach JA, Meinkoth JL, FitzGerald GA. Prostaglandin F2alpha (PGF2alpha) and the isoprostane, 8, 12-iso- isoprostane F2alpha-III, induce cardiomyocyte hypertrophy. Differential activation of downstream signaling pathways. J Biol Chem. 1998;273:22442–22452. doi: 10.1074/jbc.273.35.22442. [DOI] [PubMed] [Google Scholar]
  • 19.McNamara P, Lawson JA, Rokach J, FitzGerald GA. PPAR activation by isoprostanes: a signaling mechanism for oxidant injury. Circulation. 1999;100:2532. [Google Scholar]
  • 20.Fukunaga M, Yura T, Grygorczyk R, Badr KF. Evidence for the distinct nature of F2-isoprostane receptors from those of thromboxane A2. Am J Physiol. 1997;272:F477–483. doi: 10.1152/ajprenal.1997.272.4.F477. [DOI] [PubMed] [Google Scholar]
  • 21.Bhushan RG, Sharma SK, Xie Z, Daniels DJ, Portoghese PS. A bivalent ligand (KDN-21) reveals spinal delta and kappa opioid receptors are organized as heterodimers that give rise to delta(1) and kappa(2) phenotypes. Selective targeting of delta-kappa heterodimers. J Med Chem. 2004;47:2969–2972. doi: 10.1021/jm0342358. [DOI] [PubMed] [Google Scholar]
  • 22.Zhu WZ, Chakir K, Zhang S, Yang D, Lavoie C, Bouvier M, Hebert TE, Lakatta EG, Cheng H, Xiao RP. Heterodimerization of beta1- and beta2-adrenergic receptor subtypes optimizes beta-adrenergic modulation of cardiac contractility. Circ Res. 2005;97:244–251. doi: 10.1161/01.RES.0000176764.38934.86. [DOI] [PubMed] [Google Scholar]
  • 23.Nobles M, Benians A, Tinker A. Heterotrimeric G proteins precouple with G protein-coupled receptors in living cells. Proc Natl Acad Sci U S A. 2005;102:18706–18711. doi: 10.1073/pnas.0504778102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Gales C, Van Durm JJJ, Schaak S, Pontier S, Percherancier Y, Audet M, Paris H, Bouvier M. Probing the activation-promoted structural rearrangements in preassembled receptor-G protein complexes. Nat Struct Mol Biol. 2006;13:778–786. doi: 10.1038/nsmb1134. [DOI] [PubMed] [Google Scholar]
  • 25.Galandrin S, Bouvier M. Distinct Signaling Profiles of {beta}1 and {beta}2 Adrenergic Receptor Ligands towards Adenylyl Cyclase and Mitogen-Activated Protein Kinase Reveals the Pluridimensionality of Efficacy. Mol Pharmacol. 2006 doi: 10.1124/mol.106.026716. [DOI] [PubMed] [Google Scholar]
  • 26.Egan KM, Wang M, Lucitt MB, Zukas AM, Pure E, Lawson JA, FitzGerald GA. Cyclooxygenases, thromboxane, and atherosclerosis: plaque destabilization by cyclooxygenase-2 inhibition combined with thromboxane receptor antagonism. Circulation. 2005;111:334–342. doi: 10.1161/01.CIR.0000153386.95356.78. [DOI] [PubMed] [Google Scholar]
  • 27.AbdAlla S, Lother H, Langer A, el Faramawy Y, Quitterer U. Factor XIIIA transglutaminase crosslinks AT1 receptor dimers of monocytes at the onset of atherosclerosis. Cell. 2004;119:343–354. doi: 10.1016/j.cell.2004.10.006. [DOI] [PubMed] [Google Scholar]

RESOURCES