Next Article in Journal
Actin in Action: Imaging Approaches to Study Cytoskeleton Structure and Function
Next Article in Special Issue
Molecular Perspectives for mu/delta Opioid Receptor Heteromers as Distinct, Functional Receptors
Previous Article in Journal
Systems Biology as an Integrated Platform for Bioinformatics, Systems Synthetic Biology, and Systems Metabolic Engineering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacological Profiles of Oligomerized μ-Opioid Receptors

1
Center for Drug Abuse and Addiction, China Medical University Hospital, Taichung 40447, Taiwan
2
China Medical University, Taichung 40442, Taiwan
3
Graduate Institute of Clinical Medical Science, China Medical University, Taichung 40442, Taiwan
4
Center for Neuropsychiatric Research, National Health Research Institutes, Zhunan, Miaoli County 35053, Taiwan
*
Author to whom correspondence should be addressed.
Cells 2013, 2(4), 689-714; https://doi.org/10.3390/cells2040689
Submission received: 21 August 2013 / Revised: 30 September 2013 / Accepted: 9 October 2013 / Published: 11 October 2013
(This article belongs to the Special Issue Oligomerization & Trafficking of Opioid Receptors)

Abstract

:
Opioids are widely prescribed pain relievers with multiple side effects and potential complications. They produce analgesia via G-protein-protein coupled receptors: μ-, δ-, κ-opioid and opioid receptor-like 1 receptors. Bivalent ligands targeted to the oligomerized opioid receptors might be the key to developing analgesics without undesired side effects and obtaining effective treatment for opioid addicts. In this review we will update the biological effects of μ-opioids on homo- or hetero-oligomerized μ-opioid receptor and discuss potential mechanisms through which bivalent ligands exert beneficial effects, including adenylate cyclase regulation and receptor-mediated signaling pathways.

1. Introduction

Opioids are one of the most commonly prescribed pain relievers and have been used to treat pain for thousands of years. Considered as broad-spectrum analgesics acting at multiple sites in both the central and peripheral nervous systems, opioids also have multiple side effects and potential complications [1]. Concerns regarding tolerance to analgesic effects result in a reluctance to prescribe opioids for pain management. Adverse gastrointestinal and central nervous system events, including constipation [2,3], nausea, vomiting and sedation, are responsible for a large portion of patients discontinuing opioid treatment, often leading to inadequate pain relief and poor quality of life [4]. Hyperalgesia (increased pain sensitivity), decreased libido, and other hormonal effects, and depression may also occur [5]. Controversially, opioids are uniquely addictive, leading to misuse and diversion [6,7,8,9].
Opioids produce analgesia via G-protein-coupled opioid receptors [10]. Three conventional opioid receptors—μ (MOR), δ (DOR) and κ (KOR) [11,12,13,14]—and a non-opioid branch of opioid receptors—opioid receptor-like orphan receptor (ORL1) [15]—have been characterized based on their pharmacological, anatomical and molecular properties, with ORL1 displaying pharmacology distinct from those of conventional opioid receptors [16,17,18,19]. Activation of MOR, DOR, KOR or ORL1 produces common cellular actions by regulating the same secondary messengers, including inhibition of adenylate cyclase (AC) activity [20,21,22,23,24,25,26] and N-type [27] and L-type Ca2+ channels [28]. Activation of opioid receptors also increases phospholipase C activity, causes a transient increase in intracellular Ca2+ [29,30] and activates inwardly rectifying K+ channels [31,32] and mitogen-activated protein kinases (MAPK) [33,34].

2. Oligomerization of μ-Opioid Receptor

2.1. The Roles of MOR in the Physiological Effects of Opioids

Opioids, such as morphine and methadone, mediate their physiological effects by preferentially activating MOR in neurons. The MOR agonists display the best antinociceptive activity, but also the highest abuse liability [35]. Disruption of the MOR gene leads to a complete loss of the main biological actions of morphine, including analgesia, reward, withdrawal, respiratory depression, immunosuppression and constipation, demonstrating that both therapeutic and adverse effects of the prototypic opioid results from its interaction with MOR gene products [35,36,37,38,39,40,41,42]. For example, morphine conditioning (repeated low-dose injections, 3 mg/kg, subcutaneously (s.c.)) that prolonged the time spent in the morphine-associated compartment in wild-type mice failed to induce place preference in mutant mice [39]. An analgesic dose of morphine (6 mg/kg, s.c.) increased respiration time and decreased respiratory frequency in wild-type mice, whereas no change in respiratory parameters were detected in similarly treated MOR-deficient mice [36]. A single dose of morphine (15 mg/kg, s.c.) inhibited gastrointestinal motility in wild-type mice, but no such change was observed in mutant mice at doses up to 35 mg/kg [41].
Desensitization and internalization of MOR are potential regulatory mechanisms contributing to the development of tolerance to opioids [43,44]. Activation of MOR by an agonist may result in receptor phosphorylation mediated by G-protein receptor kinases (GRKs). Subsequently, β-arrestins bind to the phosphorylated MOR, making this complex unable to couple to G-proteins to activate downstream effectors, resulting in receptor desensitization [45]. This MOR complex is recruited to the clathrin-coated pit and then removed from the plasma membrane by endocytosis [46]. Besides rapidly reducing the number of receptors present at the cell surface, endocytosis also mediates receptor “resensitization”, which involves delivering MOR to an endosome-associated phosphatase and then returning the dephosphorylated MOR to the plasma membrane via a rapid recycling pathway [47,48,49,50,51]. The degree of desensitization of MOR signaling observed in locus coeruleus (LC) neurons correlates well with that of agonist-induced MOR endocytosis assessed in HEK293 cells, further suggesting that desensitization and internalization are closely linked [52].

2.2. Pharmacological Responses of Oligomerized MOR

Opioid receptors are members of the rhodopsin family of G-protein-coupled receptors (GPCRs), which associate with each other to form dimers and/or oligomers [53,54,55]. Human MOR form sodium dodecyl sulfate (SDS)-resistant homodimers, and increasing concentrations and the longer exposure of agonists reduce the levels of MOR dimers with a corresponding increase in that of MOR monomers. This antagonist-reversible effect is suggested to be related to the endocytosis of MOR, since monomerization proceeds internalization [56]. MOR has also been reported to hetero-oligomerize with various receptors, including DOR [57], KOR [58], ORL1 [59], the somatostatin subtype 2A (sst2A) receptor [60], the substance P receptor (NK1) [61], cannabinoid CB1 receptor (CB1R) [62] and metabotropic glutamate receptor 5 (mGluR5) [63]. Receptor hetero-oligomerization usually leads to alterations in MOR phosphorylation, internalization, desensitization, MAPK activation and coupling to voltage-dependent Ca2+ channels [64]. Our group also demonstrated that methadone and buprenorphine, MOR full and partial agonists, respectively, exert initially different (acute), yet eventually convergent (chronic), adaptive changes of AC activity in cells co-expressing, presumably heterodimeric, MOR and ORL1 receptors [65]. Hetero-oligomerization of opioid receptors generates novel ligand binding properties, in some cases resembling pharmacologically-defined opioid receptor subtypes, but not unique opioid receptor genes [57,66,67]. How oligomerized MOR affects the effects of MOR ligands is the focus of preclinical research and might indicate a way for developing analgesics without unwanted side effects.

2.2.1. MOR-DOR

Early studies have suggested that MOR and DOR cross-talk and affect each other’s properties [68]. Morphine binds to MOR and DOR and inhibits neurotransmitter release [69]. In transgenic animals lacking MOR, DOR ligand-mediated analgesia is changed [70]. In DOR knockout mice, supraspinal DOR-like analgesia is retained and morphine tolerance is lost [71]. DOR antagonists-treated mice exhibit reduced development of morphine tolerance and dependence [72]. Reduction in DOR by antisense oligonucleotides attenuates the development of morphine dependence [73]. Ligand binding assays show that MOR-selective ligands inhibit the binding of DOR-selective ligands, both competitively and noncompetitively [74,75]. Radioligand binding and electrophysiological studies suggested that MOR and DOR colocalize in the dorsal root ganglia [76,77,78]. Immunohistochemical studies revealed that MOR and DOR colocalize at the same axonal terminals of the superficial dorsal horn [79] and ultrastructurally in the plasmalemma of the dorsal horn neurons [80]. Additionally, several neuroblastomas co-express MOR and DOR [81,82,83,84]. These lines of evidence suggest the physical existences of MOR and DOR complexes.
Heterodimerization of MOR and DOR have been demonstrated and may provide foundations for more effective therapies. Co-expression of MOR and DOR in heterologous cells followed by selective immunoprecipitation results in the isolation of MOR-DOR heterodimers [57]. The bioluminescence resonance energy transfer (BRET) assay showed that MOR and DOR conjugated in living cells [85]. A combination of MOR and DOR selective agonists synergistically binds and potentiates signaling by activating the MOR-DOR heterodimer [57]. Signaling by clinically relevant MOR ligands, such as morphine, fentanyl and methadone, can be enhanced by DOR ligands [85]. Furthermore, morphine-mediated intrathecal analgesia is potentiated by a DOR antagonist [85]. Morphine and [D-Ala2-MePhe4-Glyol5]enkephalin (DAMGO), traditionally classified as MOR selective agonists, selectively activate MOR-DOR heteromeric opioid receptors with greater efficacy than homomeric opioid receptors [86]. This is consistent with studies implicating the involvement of both MOR and DOR in analgesia, tolerance and dependence [71,72,73,87,88,89].

2.2.2. MOR-KOR

MOR-KOR can form heterodimers with a similar affinity to that of MOR-DOR, as demonstrated by BRET, co-immunoprecipitation, receptor binding and G-protein coupling [58]. In males, spinal morphine antinociception requires the exclusive activation of spinal MOR; whereas in females, spinal morphine antinociception requires the concomitant activation of spinal MOR and KOR [90]. Expression of a MOR-KOR heterodimer is more prevalent in the spinal cord of proestrus vs. diestrus females and males. Spinal cord MOR-KOR heterodimers, likely to be the molecular transducer for the female-specific dynorphin/KOR component of spinal morphine antinociception, represent a unique pharmacological target for female-specific pain control [91].

2.2.3. MOR-ORL1

MOR and ORL1 are coexpressed in the dorsal horn of the spinal cord, the hippocampal formation, the caudate/putamen [92,93,94] and several subpopulations of central nervous system (CNS) neurons involved in nociception [95]. Behaviorally, mice lacking the ORL1 gene partially lose tolerance liability to morphine analgesia [96] and show marked attenuation of morphine-induced physical dependence, manifested as naloxone-precipitated withdrawal symptoms after repeated morphine treatments [97]. Co-administration of ORL1 antagonist, J-113397, during conditioning facilitates morphine-induced conditioned place preference (CPP), and ORL1 knockout rats are more sensitive to the rewarding effect of morphine than wild-type control rats, suggesting that the ORL1 system is involved in attenuating the rewarding effect of μ-opioids and offers a therapeutic target for the treatment of drug abuse and addiction [98]. Functional interactions between MOR ligands and nociceptin, the endogenous ligand of ORL1, have been observed in the human neuroblastoma cell line BE(2)-C, which contains both MOR and ORL1 [99]. Immunoprecipitation assay demonstrated that MOR can physically associate with ORL1, resulting in a complex with a unique binding selectivity profile [100]. Furthermore, heterodimerization of MOR and ORL1 impairs the potency of MOR agonist, DAMGO, [59] and attenuates ORL1-mediated inhibition of N-type channels [101].

2.2.4. MOR- sst2A

MOR and sst2A receptors coexist and cooperate functionally in nociception pathways [102,103], and there is extensive cross-talk between opioid and somatostatin-mediated analgesia responses [104,105]. A particularly high degree of MOR-sst2A receptor colocalization was observed in the locus coeruleus [60], a brain region involved in opioid withdrawal syndrome [106], and the attenuation of opioid withdrawal by the sst2A receptor agonist octreotide may be due to the physical contacts of MOR and sst2A in the locus coeruleus [107]. Co-immunoprecipitation studies provided direct evidence for the heterodimerization of MOR and sst2A. MOR-sst2A heterodimerization did not substantially alter the ligand binding properties of the receptors, but selectively cross-modulates receptor phosphorylation, internalization and desensitization [60].

2.2.5. MOR-NK1

MOR and NK1, the principle receptor for substance P, coexist and functionally collaborate in brain regions governing pain perception [108,109,110,111,112,113,114,115]. Moreover, MOR and NK1 are highly expressed in the nucleus accumbens, which mediates the motivational properties of drugs of abuse, including opioids. Interestingly, the rewarding effects of morphine are absent in NK1-deficient mice [113,114]. Using co-immunoprecipitation and BRET, MOR and NK1 were shown to form heterodimers [61]. MOR-NK1 heterodimerization does not significantly change the ligand binding and signaling properties of MOR, but dramatically altered its trafficking and resensitization profiles [61]. Altered β-arrestin trafficking in cells coexpressing MOR and NK1 could thus impact on opioid resensitization and the long-term cellular effects of opioids [61].

2.2.6. MOR-CB1R

Opioids and cannabinoids share several pharmacological effects, such as antinociception, hypothermia, hypotension and sedation [116,117]. The synergy of the pharmacological effects of opioids and cannabinoids is attributed to the cross-talk between MOR and CB1R [118,119]. Furthermore, a CB1R agonist, delta 9-tetrahydrocannabinol, can enhance the potency of morphine [120]. MOR and CB1R colocalize in dendritic spines in the caudate putamen and dorsal horn of the spinal cord [121,122,123]. Coexpression of MOR and CB1R leads to MOR-CB1R heterodimerization, as revealed by BRET [124] and fluorescence resonance energy transfer (FRET) [62] assays. Guanosine 5'-O-(3-thiotriphosphate) (GTPγS) binding and MAPK phosphorylation assays demonstrated that MOR signaling is attenuated by CB1R agonist, and this effect is reciprocal in both heterologous cells and endogenous tissues coexpressing MOR and CB1R [124]. An electrophysiological analysis has also been developed to determine the functional coupling of the heterodimerized MOR-CB1R [62].

2.2.7. MOR-mGluR5

MOR activity can be affected by presynaptic modulation through glutamate receptors. Glutamate elicits and modulates responses in the CNS via two groups of receptors, ionotropic (ligand-gated ion channels) and G-protein-coupled metabotropic receptors (mGluRs) [125,126,127]. mGluR5 plays major roles in modulatory CNS pathways and has pharmacological implications in pain [128,129]. Moreover, the inhibition of mGluR5 can modulate opioid analgesia and opioid tolerance. The specific mGluR5 antagonist, 2-methyl-6-(phenylethynyl)pyridine (MPEP), can block hyperalgesia and nociceptive behavior, and the co-administration of MPEP with morphine could suppress the loss of morphine-induced anti-nociception and inhibit the development of morphine-induced tolerance [130,131]. In human embryonic kidney (HEK) 293 cells coexpressing MOR and mGluR5, DAMGO-induced MOR phosphorylation, internalization and desensitization are attenuated by MPEP treatment [63]. Co-immunoprecipitation data further indicate MOR-mGluR5 heterodimerization [63]. Interestingly, the allosteric mGluR5 inhibitor MPEP is not simply acting as a blocker of glutamate signaling, but changes the conformation of MOR-mGluR5 heterodimer to affect MOR phosphorylation and desensitization [63].

3. Bivalent Ligands of Oligomerized μ-Opioid Receptor

Bivalent ligands are compounds that contain two recognition sites, or pharmacophores, joined through a connecting spacer. Pharmacophores are ensembles of steric and electronic features necessary for optimal supramolecular interactions with a specific biological target to exert its biological response. When talking about ligand-receptor interactions, pharmacophores are the molecular moieties of the ligands required for recognition by their corresponding receptors.
The endogenous opioid peptide family is characterized by a common tetrapeptide sequence (Tyr-Gly-Gly-Phe) and comprises over a dozen ligands [132,133,134], including β-endorphin [135,136], enkephalins [137] and dynorphins [138]. Two additional endogenous ligands, endomorphin-1 and -2, with a tetrapeptide sequence (Tyr-Pro-X-Phe-NH2 X = Trp or Phe) different from that of classical opioid peptides have also been reported [137]. Increasing the distance between enkephalin pharmacophores results in DOR-selective compounds, while reducing the distance makes it more MOR-selective [139,140]. Synthesis of bivalent ligands has been one of the most promising ways of developing new opioid analogues since the 1980s [139,140,141,142,143,144,145,146].
Ligand-induced clustering of opioid receptors was observed on the surface of neuroblastoma cells [147]. This observation prompted scientists to design double pharmacophore ligands—bivalent ligands—as probes for bridging hypothetical dimeric opioid receptors [148]. Convincing evidence for homo- and hetero-dimers among opioid receptors has been presented since the late 1990s [149]. Portoghese’s group envisioned that a bivalent ligand with a spacer of optimal length would exhibit greater potency than that derived from the sum of its two monovalent pharmacophores [144,150]. Based on molecular modeling of interlocking transmembrane (TM) helices in a homodimeric MOR-MOR, it was proposed a decade ago that interlocking dimers with a TM5,6-interface between the 7TM domains may be the dominant form of dimers [151].
The recognition sites of opioid agonists and antagonists on dimeric receptors are believed to be separate. The MOR protection experiment was performed using MOR-selective agonists (morphine) and antagonists (naloxone) to determine their effectiveness in blocking irreversible MOR antagonism by β-funaltrexamine (β-FNA) in the guinea pig ileum preparation. Relatively high concentrations (0.5–1 μM) of MOR agonists were needed to protect MOR against inactivation, while antagonists in the low nM range were effective in blocking the irreversible effect of β-FNA. It was suggested that MOR activation and antagonism are mediated through separate negative allosterically coupled recognition sites in a dimer [152]. According to their theory, an opioid could modulate its own effects on opioid receptors via regulating its own concentration. At low concentration, single occupancy of the dimer would occur; at higher concentration, occupation of the second site would dampen the overall binding and activation of the opioid receptor through negative cooperativity. The frequently-seen bell-shaped concentration-response curve is consistent with this proposal, and exogenous antagonists were envisioned to have greater affinity for this second site [148]. Site-directed mutagenesis, combined with the classical structure-activity relationship approach, led to the identification of amino acid residues on opioid receptors and chemical groups on ligands participating in molecular recognition [148,153].
Several lines of evidence support the homo-oligomerization of MORs. The human MORs form SDS-resistant homodimers, and increasing concentrations and longer exposure of both peptide (DAMGO) and alkaloid (ohmefentanyl, etorphine and morphine) agonists reduce the levels of dimers with a corresponding increase in those of monomers [56]. Complementation of function after coexpression of pairs of nonfunctional MORs that contain distinct inactivating G-protein mutations linked to the C-terminal tails suggests that MORs exist as dimers [154]. This receptor homodimerization is facilitated by a cholesterol-palmitoyl interaction in the MOR complex [155]. The synthesis of homo-bivalent ligands of mixed KOR/MOR agonists/antagonists has been achieved [156,157,158,159,160,161]. However, none of the homo-bivalent ligands have been applied to study the homo-oligomerization of MORs.
Heterodimeric opioid receptors raised the issue of pharmacological selectivity. The transition from MOR to DOR agonism upon chronic exposure of mice to MOR agonists (methadone and heroin) might reflect changes in the distribution of MOR-DOR heterodimers [162]. If there is an increase of the MOR-DOR density, MOR agonists that bind to MOR-selective sites of the dimmers could be antagonized by the interaction of a DOR antagonist at the DOR-selective site in the heterodimer. The presence of heterodimeric receptors has important influence on the interpretation of experimental data and in the screening strategy using homogeneous populations of cloned receptors. Additionally, dimeric receptors may activate transduction pathways different from those elicited by monomers [148].
Bivalent ligands are not necessary bi-functional ligands, and vice versa. Bivalent ligands are compounds that have two pharmacophores with a long linker. Earlier studies on GPCR bivalent ligands did not aim to target dimeric complexes, and many such ligands have relatively short linking groups between the pharmacophores. This suggests that these bivalent compounds may be interacting with neighboring binding sites on a single receptor rather than bridging sites across a receptor dimer. In contrast, bi-functional ligands serve as agonists and/or antagonists for different monomer receptors because of their multiple functional groups within the molecule. Considering their activities of recognizing heterodimeric receptors and the pharmacological potential of being a new-generation analgesic, we also provided an example in this review.

3.1. MOR-DOR Bivalent Ligands

The cross-talk between MOR and DOR and its possible significance in morphine tolerance and physical dependence have been suggested [71,73,88,163,164,165,166,167]. The chronic effects of morphine can be blocked by DOR antagonists without significantly compromising its antinociceptive action [72,167,168,169]. Following the aforementioned findings, mixed MOR agonist/DOR antagonist ligands have been designed to develop analgesics devoid of the side effects of traditional opioids [170,171,172]. Bivalent ligands of opioid alkaloids [143,144,145,150,173,174,175,176] and peptide agonists [139,140,177,178] derived from enkephalins have been shown to have increased opioid receptor selectively and potency compared to corresponding monovalent counterparts.
Considering the evidence for MOR-DOR heterodimers in cultured cells [57,67,85], similar interactions may also occur in vivo to mediate the synergy of MOR and DOR agonists. Bivalent ligands containing MOR agonist and DOR antagonist pharmacophores, μ-δ agonist-antagonist (MDAN) series, have been designed to address the MOR-DOR interaction through which MOR agonist-induced tolerance and dependence are attenuated [87]. The pharmacophores were derived from the MOR agonist, oxymorphone, and DOR antagonist, NTI. A transition in the behavioral pharmacology is believed to be a function of the spacer length of the bivalent ligand that reflects the bridging of the MOR-DOR heterodimer. According to this theory, if the MOR-DOR heterodimers mediate tolerance and dependence at the molecular level, changes in tolerance and dependence as a function of spacer length could be a manifestation of bridging. Indeed, the bivalent ligand with the shortest length, MDAN-16, developed a certain degree of dependence, whereas the remaining members of the series with longer spacers were essentially without dependence development. When the spacer length was longer than 22 Å (MDAN-19 to -21), neither tolerance nor dependence was observed. It appears that bridging MOR-DOR heterodimers by MDAN ligands negatively modulates putative signal transducers, thereby reducing tolerance and dependence [87]. In HEK293 cells, MDAN-21 prevents endocytosis, while MDAN-16 gives rise to robust internalization, of MOR-DOR heterodimers [179]. This dramatically divergent internalization of MOR-DOR heterodimer elicited by MDAN bivalent ligands of different spacer lengths is relevant to the role of receptor internalization in tolerance [179].

3.2. MOR-KOR Bivalent Ligands

The bivalent approach was applied to synthesize mixed MOR-KOR ligands as potential analgesics with decreased side effect [180]. Compared to MOR agonists, KOR agonists lack respiratory depressant, constipating and strong addictive (euphoria and physical dependence) properties. However, clinical trials with KOR agonists have been aborted, because of the occurrence of unacceptable sedative and dysphoric side effects. It is now recognized that KOR agonists with some MOR activity produce fewer adverse side effects than highly selective KOR agonists, so the mixed MOR-KOR ligands might act as clinically useful analgesics [181].
A series of bivalent ligands containing a KOR antagonist pharmacophore, 5’-guanidinonaltrindole (5’-GNTI), and a MOR antagonist pharmacophore, β-naltrexamine (β-NTX), linked through a spacer of varying length have been synthesized and characterized in vitro [182]. The design rationale is based on the desire to maintain a favorable hydrophilic and lipophilic balance coupled with flexibility. This includes a spacer that contains (1) glycine units that maintain a favorable hydrophilic-lipophilic balance, (2) a succinyl unit that contributes to the flexibility for favorable interaction with heterodimers and (3) an alkylamine moiety attached to 5’-GNTI that permits variation of the spacer length by one atom increments. The antagonist activities of the bivalent ligands were evaluated by measuring the inhibition of Ca2+ release in HEK293 cells stably expressing KOR and MOR singly or simultaneously. The selective agonists used to evaluate the antagonism selectivity of the target compounds were U69593 (KOR) [183] and DAMGO (MOR) [184]. One bivalent ligand, KMN-21 (with a spacer length of 21 atoms), significantly antagonized both U69593- and DAMGO-induced Ca2+ release in cells containing coexpressed MOR and KOR [182]. It is noteworthy that the KOR-DOR bivalent ligand antagonist, KDN-21, also contains a 21-atom spacer, suggesting common bridging modes for MOR-KOR and DOR-KOR heterodimeric receptors [185].
Another monovalent bi-functional ligand, N-naphthoyl-β-naltrexamine (NNTA), which selectively activates heterodimeric MOR-KOR in HEK293 cells and induces potent antinociception in mice, has been synthesized [186]. In the mouse tail-flick assay, intrathecal (i.t.) NNTA produced antinociception that was ~100-fold greater that intracerebroventricular (i.c.v.) administration. No tolerance was induced by i.t. administration, but marginal (three-fold) tolerance was observed by i.c.v. administration. Neither significant physical dependence nor place preference was produced in the ED50 dose range. These results suggest an approach to potent analgesics with fewer deleterious side effects [186].

3.3. MOR-ORL1 Bivalent Ligands

Bifunctional MOR-ORL1 agonists are hypothesized to be useful as non-addictive analgesics or medication for drug addicts [187]. The concept of using mixed-action opioids for pain management and drug abuse treatment is clinically validated [188,189,190]. Buprenorphine, a MOR partial agonist and KOR antagonist, also has low efficacy at ORL1 [191,192]. Its ORL1 agonist activity is responsible for the attenuation of its antinociceptive activity at high doses [193] and the reduction of cocaine use in dually addicted cocaine-opioid addicts [194]. If ORL1 agonist-MOR agonist activities are present in the same molecule, the ORL1 activity may modulate the rewarding effects of the MOR activity, thereby producing opioid analgesics that have a reduced addiction liability. MOR-ORL1 agonists may also be used as drug abuse medications with diminished dependence and withdrawal tendencies.
SR16435 [1-(1-(bicyclo[3.3.1]nonan-9-yl)piperidin-4-yl)indolin-2-one], a nonselective ORL1 and MOR partial agonist, has potent antinociceptive activity in acute thermal pain, as well as CPP, an effect mediated by its MOR activity [195]. It is speculated that the partial agonist efficacy of SR16435 at ORL1 is not sufficient to attenuate the rewarding effect of its MOR activity. The in vivo pharmacological profile of three other ORL1 agonists, SR14150, SR16507 and SR16835, with different selectivity and efficacy for ORL1 and MOR, have been determined in a model of acute nociception (the tail-flick assay) and a model of reward (place conditioning paradigm) in mice [187]. SR14150 is a high-affinity ORL1 partial agonist that has 20-fold selectivity over MOR. It has antinociceptive activity by acting as a MOR partial agonist, but is not rewarding in the place-conditioning paradigm, due to its ORL1 agonist activity that attenuates MOR-mediated reward. SR16507 has high binding affinity for both ORL1 and MOR, acting as a full ORL1 agonist and partial MOR agonist. It has potent antinociceptive activity, but produces CPP. SR16835 is a modestly (seven-fold) selective ORL1 full agonist, with very low efficacy at MOR. It does not produce MOR-mediated antinociception and CPP. The overall antinociceptive and antirewarding profiles of these ligands depend on their selectivity between MOR and ORL1, as well as intrinsic activity at these receptors. Notably, the ORL1-selective ORL1-MOR partial agonist, SR14150, has antinociceptive activity without rewarding effects and may lead to clinically useful treatments for pain and drug addiction [187].
The structure-activity relationship (SAR) for discovering bifunctional MOR-ORL1 ligands, starting from ORL1-selective scaffolds, has been explored [196]. Based on the 2-indolinone class of ORL1 ligands [197], a 2-D pharmacophore model (comprised of three pharmacophoric features common to most ORL1 ligands) was developed to determine SAR leading to (1) selectivity versus the classical opioid receptors and (2) functional efficacy, i.e., agonist or antagonist activity [198]. Most ORL1 ligands contain the piperidine ring, so the piperidine 4-position heterocyclic ring and the piperidine N-substituent were systematically analyzed. The compounds were tested for binding affinity at human ORL1, MOR, DOR and KOR transfected into Chinese hamster ovary (CHO) cells. Binding to the opioid receptors utilized the selective agonists, [3H]nociceptin, [3H]DAMGO, [3H]Cl-DPDPE (d-penicillamine(2,5)-enkephalin) and [3H]U69593, for ORL1, MOR, DOR and KOR, respectively. Functional activity was determined by stimulation of [35S]GTPγS binding to cell membranes [192,199,200]. Data from these 17 compounds indicate that modulation of the ORL1 agonist versus MOR agonist potency using SAR approaches will be useful for developing bivalent ORL1-MOR ligands for medication development [196].

3.4. MOR-mGluR5 Bivalent Ligands

The selective mGluR5 antagonist, MPEP, acts allosterically by binding to the 7TM domain of the receptor [201]. Co-administration of MPEP and morphine could enhance morphine antinociception and suppress morphine-induced tolerance and dependence, suggesting a design strategy for developing potent analgesics targeting both MOR and mGluR5 [130,202,203]. The bivalent ligands being synthesized [204] contain pharmacophores derived from the MOR agonist, oxymorphone, and the mGLuR5 antagonist, m-methoxy-MPEP (M-MPEP) [201], linked through spacers of varying lengths (10–24 atoms). The series of compounds was evaluated for antinociception using the tail flick and von Frey assays in mice pretreated with lipopolysaccharide (LPS). MMG22 (22-atom spacer) is the most potent member of the series (intrathecal ED50 (effective dose in 50% of the population) ~9 fmol). Since other members with shorter or longer spacers have ≥500-fold higher ED50s, the exceptional potency of MMG22 may result from the optimal bridging of the MOR-mGluR5 heteromer. MMG22 possesses a >106 therapeutic ratio, suggesting that it might be an excellent candidate for the management of chronic, intractable inflammatory pain via spinal administration [204].

4. Conclusions

Bivalent ligands of the opioid receptors have been prepared to improve the pharmacological properties of opioid ligands, because ligand-induced clustering of opioid receptors occurs on the cell surface [142,147,148,205]. These efforts have been validated with the characterization of homo- and hetero-oligomers of the opioid receptors [55,57,67,85,206,207]. Hetero-oligomers of MOR with other receptors display altered ligand-binding profiles and novel signaling properties relative to receptor monomers [55,206]. Bivalent opioid ligands capable of interacting simultaneously with different recognition sites in the hetero-oligomerized MOR complexes exhibit increased efficacy in signal amplification [205,206], thereby with (1) enhanced agonist or antagonist activity, (2) improved metabolic stability, (3) improved membrane permeability, (4) reduced opioid-induced tolerance, (5) diminished physical dependence and (6) a raised potential for the treatment of drug addiction [205].
The oligomerization of GPCRs is dynamic and regulated. Microscopically, bivalent ligands bind to and thereby stabilize or de-stabilize pre-existing dimers. Specific ligands may promote ligand-induced dimerization or inhibit further dimerization, depending on the intrinsic properties of the ligands. The pharmacological difference between bivalent ligand binding and co-application of two monovalent ligands is that the former interact with two adjacent receptors simultaneously, but the latter bind to two distant receptors sequentially. The downstream signaling process of these two events could be different spatially and temporally. The advantages of using bivalent ligands as therapeutic tools are less undesired side effects, whereas the disadvantages are non-specific and unexpected effects where the targeted heterodimers do not colocalize. Whether both ligand binding sites are equal in a dimer or have a certain degree of cooperativity depends on the structures of heterodimers and the orientation of the linked pharmacophores on the bivalent ligands. We speculate that most of the binding sites are not equal in the heterodimers and are somewhat cooperative when interacting with the bivalent ligands.
There are caveats that one should be aware of when evaluating the data of the oligomerized opioid receptors. Most studies for oligomerization employ in vitro methods of overexpression receptors in cell lines. The overexpression system apparently will have artificial results by forcing receptors together. Co-immunoprecipitation methods contain detergent, which could disrupt protein-protein interactions, thus disrupting receptor oligomerization. Moreover, the in vivo data are not solid for receptor oligomerization, and most of the studies are based on a heterologous expression system. The observed pharmacological profiles of the bivalent ligands linking a relatively non-selective agonist and an antagonist in vivo might not always reflect the effects of the heterodimers. In areas where the distributions of the two receptors are non-overlapping, the effects of the ligands observed may be owing to the agonistic or antagonistic effects on one of the receptors alone. For example, Pickel and her co-workers reported that mu-receptors were located in 21% of the dendritic profiles and 3% of the axon terminals containing CB1 receptors in the rat nucleus accumbens shell [123], a rather low percentage of neurons that express both receptors in question. Although their findings provide ultrastructural evidence that cannabinoid-opioid interactions may be mediated by activation of CB1 and MOR within the same neurons in the nucleus accumbens, for ligands that are not selective for heterodimers, their pharmacological effects actually reflect the activities of heterodimers and the homodimers or monomers in combination in vivo.
The high-resolution crystal structures of MOR, DOR, KOR and ORL1 in ligand-bound conformations have been resolved [208,209,210,211]. The MOR structure shows tightly coupled pairs of receptor molecules, held together predominantly by complementary interactions involving TM5 and TM6. This pairing might regulate MOR signaling [211,212]. In contrast, the KOR structure shows a dimeric arrangement involving interactions of TM1, TM2 and helix 8 (H8), similar to the alternative, less compact packing in the MOR structure [209,212]. The proposed roles of the TM5-TM6 and TM1-TM2-H8 interfaces in functionally relevant receptor-receptor interactions need to be addressed to reveal the role of oligomerization in the signaling of opioid receptors [212]. The mission of identifying the functionally relevant oligomerization interfaces continues and will provide further insight into the SAR of the bivalent ligands and the oligomerized opioid receptors.

Acknowledgments

This work was supported by the National Health Research Institutes (PD-102-PP-16 and NHRI-102A1-PDCO-1312141) and China Medical University Hospital (DMR-101-123 and DMR-102-029).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruan, X. Drug-related side effects of long-term intrathecal morphine therapy. Pain Phys. 2007, 10, 357–366. [Google Scholar]
  2. McCarberg, B.H. Overview and treatment of opioid-induced constipation. Postgrad. Med. 2013, 125, 7–17. [Google Scholar] [CrossRef]
  3. Ford, A.C.; Brenner, D.M.; Schoenfeld, P.S. Efficacy of Pharmacological Therapies for the Treatment of Opioid-Induced Constipation: Systematic Review and Meta-Analysis. Am. J. Gastroenterol. 2013, 108, 1566–1574. [Google Scholar] [CrossRef]
  4. Ahlbeck, K. Opioids: a two-faced Janus. Curr. Med. Res. Opin. 2011, 27, 439–448. [Google Scholar] [CrossRef]
  5. Jan, S.A. Introduction: landscape of opioid dependence. J. Manag. Care Pharm. 2010, 16, S4–8. [Google Scholar]
  6. McCarberg, B.H. Chronic pain: Reducing costs through early implementation of adherence testing and recognition of opioid misuse. Postgrad. Med. 2011, 123, 132–139. [Google Scholar] [CrossRef]
  7. McCarberg, B.H. A critical assessment of opioid treatment adherence using urine drug testing in chronic pain management. Postgrad. Med. 2011, 123, 124–131. [Google Scholar]
  8. McCarberg, B.H. Pain management in primary care: strategies to mitigate opioid misuse, abuse, and diversion. Postgrad. Med. 2011, 123, 119–130. [Google Scholar] [CrossRef]
  9. Juurlink, D.N.; Dhalla, I.A. Dependence and addiction during chronic opioid therapy. J. Med. Toxicol. 2012, 8, 393–399. [Google Scholar] [CrossRef]
  10. Cox, B.M.; Borsodi, A.; Caló, G.; Chavkin, C.; Christie, M.J.; Civelli, O.; Devi, L.A.; Evans, C.; Henderson, G.; Höllt, V.; Kieffer, B.; Kitchen, I.; Kreek, M.J.; Liu-Chen, L.Y.; Meunier, J.C.; Portoghese, P.S.; Shippenberg, T.S.; Simon, E.J.; Toll, L.; Traynor, J.R.; Ueda, H.; Wong, Y.H. Opioid receptors. IUPHAR database (IUPHAR-DB). Available online: http://www.iuphar-db.org/DATABASE/FamilyMenuForward?familyId=50 (accessed on 07 October2013).
  11. Dhawan, B.N.; Cesselin, F.; Raghubir, R.; Reisine, T.; Bradley, P.B.; Portoghese, P.S.; Hamon, M. International Union of Pharmacology. XII. Classification of opioid receptors. Pharmacol. Rev. 1996, 48, 567–592. [Google Scholar]
  12. Knapp, R.J.; Malatynska, E.; Collins, N.; Fang, L.; Wang, J.Y.; Hruby, V.J.; Roeske, W.R.; Yamamura, H.I. Molecular biology and pharmacology of cloned opioid receptors. FASEB J. 1995, 9, 516–525. [Google Scholar]
  13. Satoh, M.; Minami, M. Molecular pharmacology of the opioid receptors. Pharmacol. Ther. 1995, 68, 343–364. [Google Scholar] [CrossRef]
  14. Alfaras-Melainis, K.; Gomes, I.; Rozenfeld, R.; Zachariou, V.; Devi, L. Modulation of opioid receptor function by protein-protein interactions. Front. Biosci. 2009, 14, 3594–3607. [Google Scholar]
  15. Mollereau, C.; Parmentier, M.; Mailleux, P.; Butour, J.L.; Moisand, C.; Chalon, P.; Caput, D.; Vassart, G.; Meunier, J.C. ORL1, a novel member of the opioid receptor family. Cloning, functional expression and localization. FEBS Lett. 1994, 341, 33–38. [Google Scholar] [CrossRef]
  16. Chiou, L.C.; Liao, Y.Y.; Fan, P.C.; Kuo, P.H.; Wang, C.H.; Riemer, C.; Prinssen, E.P. Nociceptin/orphanin FQ peptide receptors: pharmacology and clinical implications. Curr. Drug Targets 2007, 8, 117–135. [Google Scholar] [CrossRef]
  17. Connor, M.; Christie, M.D. Opioid receptor signalling mechanisms. Clin. Exp. Pharmacol. Physiol. 1999, 26, 493–499. [Google Scholar] [CrossRef]
  18. Henderson, G.; McKnight, A.T. The orphan opioid receptor and its endogenous ligand--nociceptin/orphanin FQ. Trends Pharmacol. Sci. 1997, 18, 293–300. [Google Scholar] [CrossRef]
  19. Calo, G.; Guerrini, R.; Rizzi, A.; Salvadori, S.; Regoli, D. Pharmacology of nociceptin and its receptor: a novel therapeutic target. Br. J. Pharmacol. 2000, 129, 1261–1283. [Google Scholar] [CrossRef]
  20. Evans, C.J.; Keith, D.E., Jr.; Morrison, H.; Magendzo, K.; Edwards, R.H. Cloning of a delta opioid receptor by functional expression. Science 1992, 258, 1952–1955. [Google Scholar]
  21. Kieffer, B.L.; Befort, K.; Gaveriaux-Ruff, C.; Hirth, C.G. The delta-opioid receptor: isolation of a cDNA by expression cloning and pharmacological characterization. Proc. Natl. Acad. Sci. USA 1992, 89, 12048–12052. [Google Scholar] [CrossRef]
  22. Li, S.; Zhu, J.; Chen, C.; Chen, Y.W.; Deriel, J.K.; Ashby, B.; Liu-Chen, L.Y. Molecular cloning and expression of a rat kappa opioid receptor. Biochem. J. 1993, 295( Pt 3), 629–633. [Google Scholar]
  23. Chen, Y.; Mestek, A.; Liu, J.; Yu, L. Molecular cloning of a rat kappa opioid receptor reveals sequence similarities to the mu and delta opioid receptors. Biochem. J. 1993, 295( Pt 3), 625–628. [Google Scholar]
  24. Fukuda, K.; Kato, S.; Mori, K.; Nishi, M.; Takeshima, H. Primary structures and expression from cDNAs of rat opioid receptor delta- and mu-subtypes. FEBS Lett. 1993, 327, 311–314. [Google Scholar] [CrossRef]
  25. Meng, F.; Xie, G.X.; Thompson, R.C.; Mansour, A.; Goldstein, A.; Watson, S.J.; Akil, H. Cloning and pharmacological characterization of a rat kappa opioid receptor. Proc. Natl. Acad. Sci. USA 1993, 90, 9954–9958. [Google Scholar]
  26. Yasuda, K.; Raynor, K.; Kong, H.; Breder, C.D.; Takeda, J.; Reisine, T.; Bell, G.I. Cloning and functional comparison of kappa and delta opioid receptors from mouse brain. Proc. Natl. Acad. Sci. USA 1993, 90, 6736–6740. [Google Scholar]
  27. Tallent, M.; Dichter, M.A.; Bell, G.I.; Reisine, T. The cloned kappa opioid receptor couples to an N-type calcium current in undifferentiated PC-12 cells. Neuroscience 1994, 63, 1033–1040. [Google Scholar] [CrossRef]
  28. Piros, E.T.; Prather, P.L.; Law, P.Y.; Evans, C.J.; Hales, T.G. Voltage-dependent inhibition of Ca2+ channels in GH3 cells by cloned mu- and delta-opioid receptors. Mol. Pharmacol. 1996, 50, 947–956. [Google Scholar]
  29. Johnson, P.S.; Wang, J.B.; Wang, W.F.; Uhl, G.R. Expressed mu opiate receptor couples to adenylate cyclase and phosphatidyl inositol turnover. Neuroreport 1994, 5, 507–509. [Google Scholar] [CrossRef]
  30. Spencer, R.J.; Jin, W.; Thayer, S.A.; Chakrabarti, S.; Law, P.Y.; Loh, H.H. Mobilization of Ca2+ from intracellular stores in transfected neuro2a cells by activation of multiple opioid receptor subtypes. Biochem. Pharmacol. 1997, 54, 809–818. [Google Scholar] [CrossRef]
  31. Henry, D.J.; Grandy, D.K.; Lester, H.A.; Davidson, N.; Chavkin, C. Kappa-opioid receptors couple to inwardly rectifying potassium channels when coexpressed by Xenopus oocytes. Mol. Pharmacol. 1995, 47, 551–557. [Google Scholar]
  32. Ikeda, K.; Kobayashi, K.; Kobayashi, T.; Ichikawa, T.; Kumanishi, T.; Kishida, H.; Yano, R.; Manabe, T. Functional coupling of the nociceptin/orphanin FQ receptor with the G-protein-activated K+ (GIRK) channel. Brain Res. Mol. Brain Res. 1997, 45, 117–126. [Google Scholar] [CrossRef]
  33. Fukuda, K.; Kato, S.; Morikawa, H.; Shoda, T.; Mori, K. Functional coupling of the delta-, mu-, and kappa-opioid receptors to mitogen-activated protein kinase and arachidonate release in Chinese hamster ovary cells. J. Neurochem. 1996, 67, 1309–1316. [Google Scholar]
  34. Li, L.Y.; Chang, K.J. The stimulatory effect of opioids on mitogen-activated protein kinase in Chinese hamster ovary cells transfected to express mu-opioid receptors. Mol. Pharmacol. 1996, 50, 599–602. [Google Scholar]
  35. Kieffer, B.L. Opioids: first lessons from knockout mice. Trends Pharmacol. Sci. 1999, 20, 19–26. [Google Scholar] [CrossRef]
  36. Matthes, H.W.; Smadja, C.; Valverde, O.; Vonesch, J.L.; Foutz, A.S.; Boudinot, E.; Denavit-Saubie, M.; Severini, C.; Negri, L.; Roques, B.P.; Maldonado, R.; Kieffer, B.L. Activity of the delta-opioid receptor is partially reduced, whereas activity of the kappa-receptor is maintained in mice lacking the mu-receptor. J. Neurosci. 1998, 18, 7285–7295. [Google Scholar]
  37. Sora, I.; Takahashi, N.; Funada, M.; Ujike, H.; Revay, R.S.; Donovan, D.M.; Miner, L.L.; Uhl, G.R. Opiate receptor knockout mice define mu receptor roles in endogenous nociceptive responses and morphine-induced analgesia. Proc. Natl. Acad. Sci. USA 1997, 94, 1544–1549. [Google Scholar]
  38. Tian, M.; Broxmeyer, H.E.; Fan, Y.; Lai, Z.; Zhang, S.; Aronica, S.; Cooper, S.; Bigsby, R.M.; Steinmetz, R.; Engle, S.J.; et al. Altered hematopoiesis, behavior, and sexual function in mu opioid receptor-deficient mice. J. Exp. Med. 1997, 185, 1517–1522. [Google Scholar] [CrossRef]
  39. Matthes, H.W.; Maldonado, R.; Simonin, F.; Valverde, O.; Slowe, S.; Kitchen, I.; Befort, K.; Dierich, A.; Le Meur, M.; Dolle, P.; et al. Loss of morphine-induced analgesia, reward effect and withdrawal symptoms in mice lacking the mu-opioid-receptor gene. Nature 1996, 383, 819–823. [Google Scholar] [CrossRef]
  40. Loh, H.H.; Liu, H.C.; Cavalli, A.; Yang, W.; Chen, Y.F.; Wei, L.N. mu Opioid receptor knockout in mice: Effects on ligand-induced analgesia and morphine lethality. Brain Res Mol Brain Res 1998, 54, 321–326. [Google Scholar] [CrossRef]
  41. Roy, S.; Liu, H.C.; Loh, H.H. mu-Opioid receptor-knockout mice: the role of mu-opioid receptor in gastrointestinal transit. Brain Res. Mol. Brain Res. 1998, 56, 281–283. [Google Scholar] [CrossRef]
  42. Gaveriaux-Ruff, C.; Matthes, H.W.; Peluso, J.; Kieffer, B.L. Abolition of morphine-immunosuppression in mice lacking the mu-opioid receptor gene. Proc. Natl. Acad. Sci. USA 1998, 95, 6326–6330. [Google Scholar] [CrossRef]
  43. Whistler, J.L.; Chuang, H.H.; Chu, P.; Jan, L.Y.; von Zastrow, M. Functional dissociation of mu opioid receptor signaling and endocytosis: Implications for the biology of opiate tolerance and addiction. Neuron 1999, 23, 737–746. [Google Scholar] [CrossRef]
  44. Williams, J.T.; Christie, M.J.; Manzoni, O. Cellular and synaptic adaptations mediating opioid dependence. Physiol. Rev. 2001, 81, 299–343. [Google Scholar]
  45. Ferguson, S.S. Evolving concepts in G protein-coupled receptor endocytosis: the role in receptor desensitization and signaling. Pharmacol. Rev. 2001, 53, 1–24. [Google Scholar]
  46. Law, P.Y.; Erickson, L.J.; El-Kouhen, R.; Dicker, L.; Solberg, J.; Wang, W.; Miller, E.; Burd, A.L.; Loh, H.H. Receptor density and recycling affect the rate of agonist-induced desensitization of mu-opioid receptor. Mol. Pharmacol. 2000, 58, 388–398. [Google Scholar]
  47. Koch, T.; Schulz, S.; Schroder, H.; Wolf, R.; Raulf, E.; Hollt, V. Carboxyl-terminal splicing of the rat mu opioid receptor modulates agonist-mediated internalization and receptor resensitization. J. Biol. Chem. 1998, 273, 13652–13657. [Google Scholar]
  48. Lefkowitz, R.J. G protein-coupled receptors. III. New roles for receptor kinases and beta-arrestins in receptor signaling and desensitization. J. Biol. Chem. 1998, 273, 18677–18680. [Google Scholar] [CrossRef]
  49. Sternini, C.; Spann, M.; Anton, B.; Keith, D.E., Jr.; Bunnett, N.W.; von Zastrow, M.; Evans, C.; Brecha, N.C. Agonist-selective endocytosis of mu opioid receptor by neurons in vivo. Proc. Natl. Acad. Sci. USA 1996, 93, 9241–9246. [Google Scholar] [CrossRef]
  50. Finn, A.K.; Whistler, J.L. Endocytosis of the mu opioid receptor reduces tolerance and a cellular hallmark of opiate withdrawal. Neuron 2001, 32, 829–839. [Google Scholar] [CrossRef]
  51. Connor, M.; Osborne, P.B.; Christie, M.J. Mu-opioid receptor desensitization: is morphine different? Br. J. Pharmacol. 2004, 143, 685–696. [Google Scholar] [CrossRef]
  52. Alvarez, V.A.; Arttamangkul, S.; Dang, V.; Salem, A.; Whistler, J.L.; Von Zastrow, M.; Grandy, D.K.; Williams, J.T. mu-Opioid receptors: Ligand-dependent activation of potassium conductance, desensitization, and internalization. J. Neurosci. 2002, 22, 5769–5776. [Google Scholar]
  53. Angers, S.; Salahpour, A.; Bouvier, M. Dimerization: an emerging concept for G protein-coupled receptor ontogeny and function. Annu. Rev. Pharmacol. Toxicol. 2002, 42, 409–435. [Google Scholar] [CrossRef]
  54. Devi, L.A. Heterodimerization of G-protein-coupled receptors: Pharmacology, Signaling and trafficking. Trends Pharmacol. Sci. 2001, 22, 532–537. [Google Scholar] [CrossRef]
  55. George, S.R.; O'Dowd, B.F.; Lee, S.P. G-protein-coupled receptor oligomerization and its potential for drug discovery. Nat. Rev. Drug Discov. 2002, 1, 808–820. [Google Scholar] [CrossRef]
  56. Li-Wei, C.; Can, G.; De-He, Z.; Qiang, W.; Xue-Jun, X.; Jie, C.; Zhi-Qiang, C. Homodimerization of human mu-opioid receptor overexpressed in Sf9 insect cells. Protein Pept. Lett. 2002, 9, 145–152. [Google Scholar] [CrossRef]
  57. Gomes, I.; Jordan, B.A.; Gupta, A.; Trapaidze, N.; Nagy, V.; Devi, L.A. Heterodimerization of mu and delta opioid receptors: A role in opiate synergy. J. Neurosci. 2000, 20, RC110. [Google Scholar]
  58. Wang, D.; Sun, X.; Bohn, L.M.; Sadee, W. Opioid receptor homo- and heterodimerization in living cells by quantitative bioluminescence resonance energy transfer. Mol. Pharmacol. 2005, 67, 2173–2184. [Google Scholar] [CrossRef]
  59. Wang, H.L.; Hsu, C.Y.; Huang, P.C.; Kuo, Y.L.; Li, A.H.; Yeh, T.H.; Tso, A.S.; Chen, Y.L. Heterodimerization of opioid receptor-like 1 and mu-opioid receptors impairs the potency of micro receptor agonist. J. Neurochem. 2005, 92, 1285–1294. [Google Scholar] [CrossRef]
  60. 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. [Google Scholar]
  61. Pfeiffer, M.; Kirscht, S.; Stumm, R.; Koch, T.; Wu, D.; Laugsch, M.; Schroder, H.; Hollt, V.; Schulz, S. Heterodimerization of substance P and mu-opioid receptors regulates receptor trafficking and resensitization. J. Biol. Chem. 2003, 278, 51630–51637. [Google Scholar] [CrossRef]
  62. Hojo, M.; Sudo, Y.; Ando, Y.; Minami, K.; Takada, M.; Matsubara, T.; Kanaide, M.; Taniyama, K.; Sumikawa, K.; Uezono, Y. mu-Opioid receptor forms a functional heterodimer with cannabinoid CB1 receptor: Electrophysiological and FRET assay analysis. J. Pharmacol. Sci. 2008, 108, 308–319. [Google Scholar] [CrossRef]
  63. Schroder, H.; Wu, D.F.; Seifert, A.; Rankovic, M.; Schulz, S.; Hollt, V.; Koch, T. Allosteric modulation of metabotropic glutamate receptor 5 affects phosphorylation, internalization, and desensitization of the micro-opioid receptor. Neuropharmacology 2009, 56, 768–778. [Google Scholar] [CrossRef]
  64. Walwyn, W.; John, S.; Maga, M.; Evans, C.J.; Hales, T.G. Delta receptors are required for full inhibitory coupling of mu-receptors to voltage-dependent Ca(2+) channels in dorsal root ganglion neurons. Mol. Pharmacol. 2009, 76, 134–143. [Google Scholar] [CrossRef]
  65. Lee, C.W.; Yan, J.Y.; Chiang, Y.C.; Hung, T.W.; Wang, H.L.; Chiou, L.C.; Ho, I.K. Differential pharmacological actions of methadone and buprenorphine in human embryonic kidney 293 cells coexpressing human mu-opioid and opioid receptor-like 1 receptors. Neurochem. Res. 2011, 36, 2008–2021. [Google Scholar] [CrossRef]
  66. Jordan, B.A.; Devi, L.A. G-protein-coupled receptor heterodimerization modulates receptor function. Nature 1999, 399, 697–700. [Google Scholar] [CrossRef]
  67. George, S.R.; Fan, T.; Xie, Z.; Tse, R.; Tam, V.; Varghese, G.; O'Dowd, B.F. Oligomerization of mu- and delta-opioid receptors. Generation of novel functional properties. J. Biol. Chem. 2000, 275, 26128–26135. [Google Scholar]
  68. Traynor, J.R.; Elliott, J. delta-Opioid receptor subtypes and cross-talk with mu-receptors. Trends Pharmacol. Sci. 1993, 14, 84–86. [Google Scholar] [CrossRef]
  69. Macdonald, R.L.; Nelson, P.G. Specific-opiate-induced depression of transmitter release from dorsal root ganglion cells in culture. Science 1978, 199, 1449–1451. [Google Scholar]
  70. Sora, I.; Funada, M.; Uhl, G.R. The mu-opioid receptor is necessary for [D-Pen2,D-Pen5]enkephalin-induced analgesia. Eur. J. Pharmacol. 1997, 324, R1–R2. [Google Scholar] [CrossRef]
  71. Zhu, Y.; King, M.A.; Schuller, A.G.; Nitsche, J.F.; Reidl, M.; Elde, R.P.; Unterwald, E.; Pasternak, G.W.; Pintar, J.E. Retention of supraspinal delta-like analgesia and loss of morphine tolerance in delta opioid receptor knockout mice. Neuron 1999, 24, 243–252. [Google Scholar] [CrossRef]
  72. Abdelhamid, E.E.; Sultana, M.; Portoghese, P.S.; Takemori, A.E. Selective blockage of delta opioid receptors prevents the development of morphine tolerance and dependence in mice. J. Pharmacol. Exp. Ther. 1991, 258, 299–303. [Google Scholar]
  73. Sanchez-Blazquez, P.; Garcia-Espana, A.; Garzon, J. Antisense oligodeoxynucleotides to opioid mu and delta receptors reduced morphine dependence in mice: Role of delta-2 opioid receptors. J. Pharmacol. Exp. Ther. 1997, 280, 1423–1431. [Google Scholar]
  74. Rothman, R.B.; Westfall, T.C. Allosteric modulation by leucine-enkephalin of [3H]naloxone binding in rat brain. Eur. J. Pharmacol. 1981, 72, 365–368. [Google Scholar] [CrossRef]
  75. Rothman, R.B.; Bowen, W.D.; Schumacher, U.K.; Pert, C.B. Effect of beta-FNA on opiate receptor binding: Preliminary evidence for two types of mu receptors. Eur. J. Pharmacol. 1983, 95, 147–148. [Google Scholar] [CrossRef]
  76. Fields, H.L.; Emson, P.C.; Leigh, B.K.; Gilbert, R.F.; Iversen, L.L. Multiple opiate receptor sites on primary afferent fibres. Nature 1980, 284, 351–353. [Google Scholar] [CrossRef]
  77. Egan, T.M.; North, R.A. Both mu and delta opiate receptors exist on the same neuron. Science 1981, 214, 923–924. [Google Scholar]
  78. Zieglgansberger, W.; French, E.D.; Mercuri, N.; Pelayo, F.; Williams, J.T. Multiple opiate receptors on neurons of the mammalian central nervous system. In vivo and in vitro studies. Life Sci. 1982, 31, 2343–2346. [Google Scholar] [CrossRef]
  79. Arvidsson, U.; Riedl, M.; Chakrabarti, S.; Lee, J.H.; Nakano, A.H.; Dado, R.J.; Loh, H.H.; Law, P.Y.; Wessendorf, M.W.; Elde, R. Distribution and targeting of a mu-opioid receptor (MOR1) in brain and spinal cord. J. Neurosci. 1995, 15, 3328–3341. [Google Scholar]
  80. Cheng, P.Y.; Liu-Chen, L.Y.; Pickel, V.M. Dual ultrastructural immunocytochemical labeling of mu and delta opioid receptors in the superficial layers of the rat cervical spinal cord. Brain Res. 1997, 778, 367–380. [Google Scholar] [CrossRef]
  81. Yu, V.C.; Richards, M.L.; Sadee, W. A human neuroblastoma cell line expresses mu and delta opioid receptor sites. J. Biol. Chem. 1986, 261, 1065–1070. [Google Scholar]
  82. Kazmi, S.M.; Mishra, R.K. Comparative pharmacological properties and functional coupling of mu and delta opioid receptor sites in human neuroblastoma SH-SY5Y cells. Mol. Pharmacol. 1987, 32, 109–118. [Google Scholar]
  83. Baumhaker, Y.; Wollman, Y.; Goldstein, M.N.; Sarne, Y. Evidence for mu-, delta-, and kappa-opioid receptors in a human neuroblastoma cell line. Life Sci. 1993, 52, PL205–PL210. [Google Scholar] [CrossRef]
  84. Palazzi, E.; Ceppi, E.; Guglielmetti, F.; Catozzi, L.; Amoroso, D.; Groppetti, A. Biochemical evidence of functional interaction between mu- and delta-opioid receptors in SK-N-BE neuroblastoma cell line. J. Neurochem. 1996, 67, 138–144. [Google Scholar]
  85. Gomes, I.; Gupta, A.; Filipovska, J.; Szeto, H.H.; Pintar, J.E.; Devi, L.A. A role for heterodimerization of mu and delta opiate receptors in enhancing morphine analgesia. Proc. Natl. Acad. Sci. USA 2004, 101, 5135–5139. [Google Scholar]
  86. Yekkirala, A.S.; Kalyuzhny, A.E.; Portoghese, P.S. Standard opioid agonists activate heteromeric opioid receptors: Evidence for morphine and [d-Ala(2)-MePhe(4)-Glyol(5)]enkephalin as selective mu-delta agonists. ACS Chem. Neurosci. 2010, 1, 146–154. [Google Scholar] [CrossRef]
  87. Daniels, D.J.; Lenard, N.R.; Etienne, C.L.; Law, P.Y.; Roerig, S.C.; Portoghese, P.S. Opioid-induced tolerance and dependence in mice is modulated by the distance between pharmacophores in a bivalent ligand series. Proc. Natl. Acad. Sci. USA 2005, 102, 19208–19213. [Google Scholar]
  88. Kest, B.; Lee, C.E.; McLemore, G.L.; Inturrisi, C.E. An antisense oligodeoxynucleotide to the delta opioid receptor (DOR-1) inhibits morphine tolerance and acute dependence in mice. Brain Res. Bull. 1996, 39, 185–188. [Google Scholar] [CrossRef]
  89. Lenard, N.R.; Daniels, D.J.; Portoghese, P.S.; Roerig, S.C. Absence of conditioned place preference or reinstatement with bivalent ligands containing mu-opioid receptor agonist and delta-opioid receptor antagonist pharmacophores. Eur. J. Pharmacol. 2007, 566, 75–82. [Google Scholar] [CrossRef]
  90. Liu, N.J.; von Gizycki, H.; Gintzler, A.R. Sexually dimorphic recruitment of spinal opioid analgesic pathways by the spinal application of morphine. J. Pharmacol. Exp. Ther. 2007, 322, 654–660. [Google Scholar] [CrossRef]
  91. Chakrabarti, S.; Liu, N.J.; Gintzler, A.R. Formation of mu-/kappa-opioid receptor heterodimer is sex-dependent and mediates female-specific opioid analgesia. Proc. Natl. Acad. Sci. USA 2010, 107, 20115–20119. [Google Scholar] [CrossRef]
  92. Anton, B.; Fein, J.; To, T.; Li, X.; Silberstein, L.; Evans, C.J. Immunohistochemical localization of ORL-1 in the central nervous system of the rat. J. Comp. Neurol. 1996, 368, 229–251. [Google Scholar] [CrossRef]
  93. Moriwaki, A.; Wang, J.B.; Svingos, A.; van Bockstaele, E.; Cheng, P.; Pickel, V.; Uhl, G.R. mu Opiate receptor immunoreactivity in rat central nervous system. Neurochem. Res. 1996, 21, 1315–1331. [Google Scholar] [CrossRef]
  94. Daunais, J.B.; Letchworth, S.R.; Sim-Selley, L.J.; Smith, H.R.; Childers, S.R.; Porrino, L.J. Functional and anatomical localization of mu opioid receptors in the striatum, amygdala, and extended amygdala of the nonhuman primate. J. Comp. Neurol. 2001, 433, 471–485. [Google Scholar] [CrossRef]
  95. Houtani, T.; Nishi, M.; Takeshima, H.; Sato, K.; Sakuma, S.; Kakimoto, S.; Ueyama, T.; Noda, T.; Sugimoto, T. Distribution of nociceptin/orphanin FQ precursor protein and receptor in brain and spinal cord: a study using in situ hybridization and X-gal histochemistry in receptor-deficient mice. J. Comp. Neurol. 2000, 424, 489–508. [Google Scholar] [CrossRef]
  96. Ueda, H.; Yamaguchi, T.; Tokuyama, S.; Inoue, M.; Nishi, M.; Takeshima, H. Partial loss of tolerance liability to morphine analgesia in mice lacking the nociceptin receptor gene. Neurosci. Lett. 1997, 237, 136–138. [Google Scholar] [CrossRef]
  97. Ueda, H.; Inoue, M.; Takeshima, H.; Iwasawa, Y. Enhanced spinal nociceptin receptor expression develops morphine tolerance and dependence. J. Neurosci. 2000, 20, 7640–7647. [Google Scholar]
  98. Rutten, K.; De Vry, J.; Bruckmann, W.; Tzschentke, T.M. Pharmacological blockade or genetic knockout of the NOP receptor potentiates the rewarding effect of morphine in rats. Drug Alcohol Depend. 2011, 114, 253–256. [Google Scholar]
  99. Mandyam, C.D.; Altememi, G.F.; Standifer, K.M. beta-Funaltrexamine inactivates ORL1 receptors in BE(2)-C human neuroblastoma cells. Eur. J. Pharmacol. 2000, 402, R1–R37. [Google Scholar] [CrossRef]
  100. Pan, Y.X.; Bolan, E.; Pasternak, G.W. Dimerization of morphine and orphanin FQ/nociceptin receptors: Generation of a novel opioid receptor subtype. Biochem. Biophys. Res. Commun. 2002, 297, 659–663. [Google Scholar] [CrossRef]
  101. Evans, R.M.; You, H.; Hameed, S.; Altier, C.; Mezghrani, A.; Bourinet, E.; Zamponi, G.W. Heterodimerization of ORL1 and opioid receptors and its consequences for N-type calcium channel regulation. J. Biol. Chem. 2010, 285, 1032–1040. [Google Scholar]
  102. Schulz, S.; Schreff, M.; Koch, T.; Zimprich, A.; Gramsch, C.; Elde, R.; Hollt, V. Immunolocalization of two mu-opioid receptor isoforms (MOR1 and MOR1B) in the rat central nervous system. Neuroscience 1998, 82, 613–622. [Google Scholar]
  103. Schulz, S.; Schreff, M.; Schmidt, H.; Handel, M.; Przewlocki, R.; Hollt, V. Immunocytochemical localization of somatostatin receptor sst2A in the rat spinal cord and dorsal root ganglia. Eur. J. Neurosci. 1998, 10, 3700–3708. [Google Scholar] [CrossRef]
  104. Betoin, F.; Ardid, D.; Herbet, A.; Aumaitre, O.; Kemeny, J.L.; Duchene-Marullaz, P.; Lavarenne, J.; Eschalier, A. Evidence for a central long-lasting antinociceptive effect of vapreotide, an analog of somatostatin, involving an opioidergic mechanism. J. Pharmacol. Exp. Ther. 1994, 269, 7–14. [Google Scholar]
  105. Bereiter, D.A. Morphine and somatostatin analogue reduce c-fos expression in trigeminal subnucleus caudalis produced by corneal stimulation in the rat. Neuroscience 1997, 77, 863–874. [Google Scholar] [CrossRef]
  106. Rasmussen, K.; Beitner-Johnson, D.B.; Krystal, J.H.; Aghajanian, G.K.; Nestler, E.J. Opiate withdrawal and the rat locus coeruleus: Behavioral, Electrophysiological, and biochemical correlates. J. Neurosci. 1990, 10, 2308–2317. [Google Scholar]
  107. Bell, J.R.; Young, M.R.; Masterman, S.C.; Morris, A.; Mattick, R.P.; Bammer, G. A pilot study of naltrexone-accelerated detoxification in opioid dependence. Med. J. Aust. 1999, 171, 26–30. [Google Scholar]
  108. Aicher, S.A.; Punnoose, A.; Goldberg, A. mu-Opioid receptors often colocalize with the substance P receptor (NK1) in the trigeminal dorsal horn. J. Neurosci. 2000, 20, 4345–4354. [Google Scholar]
  109. Aicher, S.A.; Sharma, S.; Cheng, P.Y.; Liu-Chen, L.Y.; Pickel, V.M. Dual ultrastructural localization of mu-opiate receptors and substance p in the dorsal horn. Synapse 2000, 36, 12–20. [Google Scholar] [CrossRef]
  110. Foran, S.E.; Carr, D.B.; Lipkowski, A.W.; Maszczynska, I.; Marchand, J.E.; Misicka, A.; Beinborn, M.; Kopin, A.S.; Kream, R.M. A substance P-opioid chimeric peptide as a unique nontolerance-forming analgesic. Proc. Natl. Acad. Sci. USA 2000, 97, 7621–7626. [Google Scholar] [CrossRef]
  111. Herrero, J.F.; Laird, J.M.; Lopez-Garcia, J.A. Wind-up of spinal cord neurones and pain sensation: Much ado about something? Prog. Neurobiol. 2000, 61, 169–203. [Google Scholar] [CrossRef]
  112. Jasmin, L.; Tien, D.; Weinshenker, D.; Palmiter, R.D.; Green, P.G.; Janni, G.; Ohara, P.T. The NK1 receptor mediates both the hyperalgesia and the resistance to morphine in mice lacking noradrenaline. Proc. Natl. Acad. Sci. USA 2002, 99, 1029–1034. [Google Scholar]
  113. Murtra, P.; Sheasby, A.M.; Hunt, S.P.; De Felipe, C. Rewarding effects of opiates are absent in mice lacking the receptor for substance P. Nature 2000, 405, 180–183. [Google Scholar] [CrossRef]
  114. Ripley, T.L.; Gadd, C.A.; De Felipe, C.; Hunt, S.P.; Stephens, D.N. Lack of self-administration and behavioural sensitisation to morphine, but not cocaine, in mice lacking NK1 receptors. Neuropharmacology 2002, 43, 1258–1268. [Google Scholar] [CrossRef]
  115. Spike, R.C.; Puskar, Z.; Sakamoto, H.; Stewart, W.; Watt, C.; Todd, A.J. MOR-1-immunoreactive neurons in the dorsal horn of the rat spinal cord: Evidence for nonsynaptic innervation by substance P-containing primary afferents and for selective activation by noxious thermal stimuli. Eur. J. Neurosci. 2002, 15, 1306–1316. [Google Scholar]
  116. Braida, D.; Pozzi, M.; Cavallini, R.; Sala, M. Conditioned place preference induced by the cannabinoid agonist CP 55,940: Interaction with the opioid system. Neuroscience 2001, 104, 923–926. [Google Scholar]
  117. Manzanares, J.; Corchero, J.; Fuentes, J.A. Opioid and cannabinoid receptor-mediated regulation of the increase in adrenocorticotropin hormone and corticosterone plasma concentrations induced by central administration of delta(9)-tetrahydrocannabinol in rats. Brain Res. 1999, 839, 173–179. [Google Scholar] [CrossRef]
  118. Ledent, C.; Valverde, O.; Cossu, G.; Petitet, F.; Aubert, J.F.; Beslot, F.; Bohme, G.A.; Imperato, A.; Pedrazzini, T.; Roques, B.P.; et al. Unresponsiveness to cannabinoids and reduced addictive effects of opiates in CB1 receptor knockout mice. Science 1999, 283, 401–404. [Google Scholar] [CrossRef]
  119. Ghozland, S.; Matthes, H.W.; Simonin, F.; Filliol, D.; Kieffer, B.L.; Maldonado, R. Motivational effects of cannabinoids are mediated by mu-opioid and kappa-opioid receptors. J. Neurosci. 2002, 22, 1146–1154. [Google Scholar]
  120. Cichewicz, D.L. Synergistic interactions between cannabinoid and opioid analgesics. Life Sci. 2004, 74, 1317–1324. [Google Scholar] [CrossRef]
  121. Rodriguez, J.J.; Mackie, K.; Pickel, V.M. Ultrastructural localization of the CB1 cannabinoid receptor in mu-opioid receptor patches of the rat Caudate putamen nucleus. J. Neurosci. 2001, 21, 823–833. [Google Scholar]
  122. Salio, C.; Fischer, J.; Franzoni, M.F.; Mackie, K.; Kaneko, T.; Conrath, M. CB1-cannabinoid and mu-opioid receptor co-localization on postsynaptic target in the rat dorsal horn. Neuroreport 2001, 12, 3689–3692. [Google Scholar] [CrossRef]
  123. Pickel, V.M.; Chan, J.; Kash, T.L.; Rodriguez, J.J.; MacKie, K. Compartment-specific localization of cannabinoid 1 (CB1) and mu-opioid receptors in rat nucleus accumbens. Neuroscience 2004, 127, 101–112. [Google Scholar] [CrossRef]
  124. Rios, C.; Gomes, I.; Devi, L.A. mu opioid and CB1 cannabinoid receptor interactions: Reciprocal inhibition of receptor signaling and neuritogenesis. Br. J. Pharmacol. 2006, 148, 387–395. [Google Scholar] [CrossRef]
  125. Seeburg, P.H. The TINS/TiPS Lecture. The molecular biology of mammalian glutamate receptor channels. Trends Neurosci. 1993, 16, 359–365. [Google Scholar] [CrossRef]
  126. Conn, P.J.; Pin, J.P. Pharmacology and functions of metabotropic glutamate receptors. Annu. Rev. Pharmacol. Toxicol. 1997, 37, 205–237. [Google Scholar] [CrossRef]
  127. Bordi, F.; Ugolini, A. Involvement of mGluR(5) on acute nociceptive transmission. Brain Res. 2000, 871, 223–233. [Google Scholar] [CrossRef]
  128. Spooren, W.P.; Gasparini, F.; Salt, T.E.; Kuhn, R. Novel allosteric antagonists shed light on mglu(5) receptors and CNS disorders. Trends Pharmacol. Sci. 2001, 22, 331–337. [Google Scholar] [CrossRef]
  129. Green, M.D.; Yang, X.; Cramer, M.; King, C.D. In vitro metabolic studies on the selective metabotropic glutamate receptor sub-type 5 (mGluR5) antagonist 3-[(2-methyl-1,3-thiazol-4-yl) ethynyl]-pyridine (MTEP). Neurosci. Lett. 2006, 391, 91–95. [Google Scholar] [CrossRef]
  130. Gabra, B.H.; Smith, F.L.; Navarro, H.A.; Carroll, F.I.; Dewey, W.L. mGluR5 antagonists that block calcium mobilization in vitro also reverse (S)-3,5-DHPG-induced hyperalgesia and morphine antinociceptive tolerance in vivo. Brain Res. 2008, 1187, 58–66. [Google Scholar] [CrossRef]
  131. Lee, H.J.; Choi, H.S.; Ju, J.S.; Bae, Y.C.; Kim, S.K.; Yoon, Y.W.; Ahn, D.K. Peripheral mGluR5 antagonist attenuated craniofacial muscle pain and inflammation but not mGluR1 antagonist in lightly anesthetized rats. Brain Res. Bull. 2006, 70, 378–385. [Google Scholar] [CrossRef]
  132. Hughes, J.; Smith, T.W.; Kosterlitz, H.W.; Fothergill, L.A.; Morgan, B.A.; Morris, H.R. Identification of two related pentapeptides from the brain with potent opiate agonist activity. Nature 1975, 258, 577–580. [Google Scholar] [CrossRef]
  133. Birdsall, N.J.; Hulme, E.C. C fragment of lipotropin has a high affinity for brain opiate receptors. Nature 1976, 260, 793–795. [Google Scholar] [CrossRef]
  134. Cox, B.M.; Goldstein, A.; Hi, C.H. Opioid activity of a peptide, beta-lipotropin-(61–91), derived from beta-lipotropin. Proc. Natl. Acad. Sci. USA 1976, 73, 1821–1823. [Google Scholar] [CrossRef]
  135. Li, C.H.; Chung, D.; Doneen, B.A. Isolation, characterization and opiate activity of beta-endorphin from human pituitary glands. Biochem. Biophys. Res. Commun. 1976, 72, 1542–1547. [Google Scholar] [CrossRef]
  136. Li, C.H.; Lemaire, S.; Yamashiro, D.; Doneen, B.A. The synthesis and opiate activity of beta-endorphin. Biochem. Biophys. Res. Commun. 1976, 71, 19–25. [Google Scholar] [CrossRef]
  137. Zadina, J.E.; Hackler, L.; Ge, L.J.; Kastin, A.J. A potent and selective endogenous agonist for the mu-opiate receptor. Nature 1997, 386, 499–502. [Google Scholar] [CrossRef]
  138. Goldstein, A.; Tachibana, S.; Lowney, L.I.; Hunkapiller, M.; Hood, L. Dynorphin-(1–13), an extraordinarily potent opioid peptide. Proc. Natl. Acad. Sci. USA 1979, 76, 6666–6670. [Google Scholar] [CrossRef]
  139. Costa, T.; Wuster, M.; Herz, A.; Shimohigashi, Y.; Chen, H.C.; Rodbard, D. Receptor binding and biological activity of bivalent enkephalins. Biochem. Pharmacol. 1985, 34, 25–30. [Google Scholar] [CrossRef]
  140. Shimohigashi, Y.; Costa, T.; Chen, H.C.; Rodbard, D. Dimeric tetrapeptide enkephalins display extraordinary selectivity for the delta opiate receptor. Nature 1982, 297, 333–335. [Google Scholar] [CrossRef]
  141. Lipkowski, A.W.; Konecka, A.M.; Sroczynska, I.; Przewlocki, R.; Stala, L.; Tam, S.W. Bivalent opioid peptide analogues with reduced distances between pharmacophores. Life Sci. 1987, 40, 2283–2288. [Google Scholar] [CrossRef]
  142. Lipkowski, A.W.; Konecka, A.M.; Sroczynska, I. Double-enkephalins--synthesis, activity on guinea-pig ileum, and analgesic effect. Peptides 1982, 3, 697–700. [Google Scholar] [CrossRef]
  143. Hazum, E.; Chang, K.J.; Leighton, H.J.; Lever, O.W., Jr.; Cuatrecasas, P. Increased biological activity of dimers of oxymorphone and enkephalin: Possible role of receptor crosslinking. Biochem. Biophys. Res. Commun. 1982, 104, 347–353. [Google Scholar] [CrossRef]
  144. Portoghese, P.S.; Ronsisvalle, G.; Larson, D.L.; Yim, C.B.; Sayre, L.M.; Takemori, A.E. Opioid agonist and antagonist bivalent ligands as receptor probes. Life Sci. 1982, 31, 1283–1286. [Google Scholar] [CrossRef]
  145. Portoghese, P.S.; Larson, D.L.; Yim, C.B.; Sayre, L.M.; Ronsisvalle, G.; Lipkowski, A.W.; Takemori, A.E.; Rice, K.C.; Tam, S.W. Stereostructure-activity relationship of opioid agonist and antagonist bivalent ligands. Evidence for bridging between vicinal opioid receptors. J. Med. Chem. 1985, 28, 1140–1141. [Google Scholar] [CrossRef]
  146. Schiller, P.W.; Nguyen, T.M.; Lemieux, C.; Maziak, L.A. A novel side-chain-linked antiparallel cyclic dimer of enkephalin. FEBS Lett. 1985, 191, 231–234. [Google Scholar] [CrossRef]
  147. Hazum, E.; Chang, K.J.; Cuatrecasas, P. Opiate (Enkephalin) receptors of neuroblastoma cells: Occurrence in clusters on the cell surface. Science 1979, 206, 1077–1079. [Google Scholar]
  148. Portoghese, P.S. From models to molecules: Opioid receptor dimers, Bivalent ligands, and selective opioid receptor probes. J. Med. Chem. 2001, 44, 2259–2269. [Google Scholar] [CrossRef]
  149. Jordan, B.A.; Cvejic, S.; Devi, L.A. Opioids and their complicated receptor complexes. Neuropsychopharmacology 2000, 23, S5–S18. [Google Scholar] [CrossRef]
  150. Erez, M.; Takemori, A.E.; Portoghese, P.S. Narcotic antagonistic potency of bivalent ligands which contain beta-naltrexamine. Evidence for bridging between proximal recognition sites. J. Med. Chem. 1982, 25, 847–849. [Google Scholar] [CrossRef]
  151. Gouldson, P.R.; Snell, C.R.; Bywater, R.P.; Higgs, C.; Reynolds, C.A. Domain swapping in G-protein coupled receptor dimers. Protein Eng. 1998, 11, 1181–1193. [Google Scholar] [CrossRef]
  152. Portoghese, P.S.; Takemori, A.E. Different receptor sites mediate opioid agonism and antagonism. J. Med. Chem. 1983, 26, 1341–1343. [Google Scholar] [CrossRef]
  153. Hjorth, S.A.; Thirstrup, K.; Grandy, D.K.; Schwartz, T.W. Analysis of selective binding epitopes for the kappa-opioid receptor antagonist nor-binaltorphimine. Mol. Pharmacol. 1995, 47, 1089–1094. [Google Scholar]
  154. Pascal, G.; Milligan, G. Functional complementation and the analysis of opioid receptor homodimerization. Mol. Pharmacol. 2005, 68, 905–915. [Google Scholar]
  155. Zheng, H.; Pearsall, E.A.; Hurst, D.P.; Zhang, Y.; Chu, J.; Zhou, Y.; Reggio, P.H.; Loh, H.H.; Law, P.Y. Palmitoylation and membrane cholesterol stabilize mu-opioid receptor homodimerization and G protein coupling. BMC Cell Biol. 2012, 13, 6. [Google Scholar] [CrossRef]
  156. Zhang, B.; Zhang, T.; Sromek, A.W.; Scrimale, T.; Bidlack, J.M.; Neumeyer, J.L. Synthesis and binding affinity of novel mono- and bivalent morphinan ligands for kappa, mu, and delta opioid receptors. Bioorganic Med. Chem. 2011, 19, 2808–2816. [Google Scholar] [CrossRef]
  157. Neumeyer, J.L.; Zhang, A.; Xiong, W.; Gu, X.H.; Hilbert, J.E.; Knapp, B.I.; Negus, S.S.; Mello, N.K.; Bidlack, J.M. Design and synthesis of novel dimeric morphinan ligands for kappa and micro opioid receptors. J. Med. Chem. 2003, 46, 5162–5170. [Google Scholar] [CrossRef]
  158. Peng, X.; Knapp, B.I.; Bidlack, J.M.; Neumeyer, J.L. Pharmacological properties of bivalent ligands containing butorphan linked to nalbuphine, naltrexone, and naloxone at mu, delta, and kappa opioid receptors. J. Med. Chem. 2007, 50, 2254–2258. [Google Scholar] [CrossRef]
  159. Peng, X.; Knapp, B.I.; Bidlack, J.M.; Neumeyer, J.L. Synthesis and preliminary in vitro investigation of bivalent ligands containing homo- and heterodimeric pharmacophores at mu, delta, and kappa opioid receptors. J. Med. Chem. 2006, 49, 256–262. [Google Scholar] [CrossRef]
  160. Decker, M.; Fulton, B.S.; Zhang, B.; Knapp, B.I.; Bidlack, J.M.; Neumeyer, J.L. Univalent and bivalent ligands of butorphan: Characteristics of the linking chain determine the affinity and potency of such opioid ligands. J. Med. Chem. 2009, 52, 7389–7396. [Google Scholar] [CrossRef]
  161. Mathews, J.L.; Peng, X.; Xiong, W.; Zhang, A.; Negus, S.S.; Neumeyer, J.L.; Bidlack, J.M. Characterization of a novel bivalent morphinan possessing kappa agonist and micro agonist/antagonist properties. J. Pharmacol. Exp. Ther. 2005, 315, 821–827. [Google Scholar] [CrossRef]
  162. Rady, J.J.; Holmes, B.B.; Portoghese, P.S.; Fujimoto, J.M. Morphine tolerance in mice changes response of heroin from mu to delta opioid receptors. Proc. Soc. Exp. Biol. Med. Soc. Exp. Biol. Med. (New York, N.Y.) 2000, 224, 93–101. [Google Scholar] [CrossRef]
  163. Vaught, J.L.; Takemori, A.E. Differential effects of leucine and methionine enkephalin on morphine-induced analgesia, acute tolerance and dependence. J. Pharmacol. Exp. Ther. 1979, 208, 86–90. [Google Scholar]
  164. Vaught, J.L.; Rothman, R.B.; Westfall, T.C. Mu and delta receptors: their role in analgesia in the differential effects of opioid peptides on analgesia. Life Sci. 1982, 30, 1443–1455. [Google Scholar] [CrossRef]
  165. Heyman, J.S.; Jiang, Q.; Rothman, R.B.; Mosberg, H.I.; Porreca, F. Modulation of mu-mediated antinociception by delta agonists: Characterization with antagonists. Eur. J. Pharmacol. 1989, 169, 43–52. [Google Scholar] [CrossRef]
  166. Heyman, J.S.; Vaught, J.L.; Mosberg, H.I.; Haaseth, R.C.; Porreca, F. Modulation of mu-mediated antinociception by delta agonists in the mouse: Selective potentiation of morphine and normorphine by [D-Pen2,D-Pen5]enkephalin. Eur. J. Pharmacol. 1989, 165, 1–10. [Google Scholar] [CrossRef]
  167. Miyamoto, Y.; Portoghese, P.S.; Takemori, A.E. Involvement of delta 2 opioid receptors in the development of morphine dependence in mice. J. Pharmacol. Exp. Ther. 1993, 264, 1141–1145. [Google Scholar]
  168. Miyamoto, Y.; Bowen, W.D.; Portoghese, P.S.; Takemori, A.E. Lack of involvement of delta-1 opioid receptors in the development of physical dependence on morphine in mice. J. Pharmacol. Exp. Ther. 1994, 270, 37–39. [Google Scholar]
  169. Fundytus, M.E.; Schiller, P.W.; Shapiro, M.; Weltrowska, G.; Coderre, T.J. Attenuation of morphine tolerance and dependence with the highly selective delta-opioid receptor antagonist TIPP[psi]. Eur. J. Pharmacol. 1995, 286, 105–108. [Google Scholar] [CrossRef]
  170. Schiller, P.W.; Fundytus, M.E.; Merovitz, L.; Weltrowska, G.; Nguyen, T.M.; Lemieux, C.; Chung, N.N.; Coderre, T.J. The opioid mu agonist/delta antagonist DIPP-NH(2)[Psi] produces a potent analgesic effect, no physical dependence, and less tolerance than morphine in rats. J. Med. Chem. 1999, 42, 3520–3526. [Google Scholar] [CrossRef]
  171. Wells, J.L.; Bartlett, J.L.; Ananthan, S.; Bilsky, E.J. In vivo pharmacological characterization of SoRI 9409, a nonpeptidic opioid mu-agonist/delta-antagonist that produces limited antinociceptive tolerance and attenuates morphine physical dependence. J. Pharmacol. Exp. Ther. 2001, 297, 597–605. [Google Scholar]
  172. Ananthan, S.; Kezar, H.S., 3rd; Carter, R.L.; Saini, S.K.; Rice, K.C.; Wells, J.L.; Davis, P.; Xu, H.; Dersch, C.M.; Bilsky, E.J.; et al. Synthesis, Opioid receptor binding, and biological activities of naltrexone-derived pyrido- and pyrimidomorphinans. J. Med. Chem. 1999, 42, 3527–3538. [Google Scholar] [CrossRef]
  173. Portoghese, P.S.; Ronsisvalle, G.; Larson, D.L.; Takemori, A.E. Synthesis and opioid antagonist potencies of naltrexamine bivalent ligands with conformationally restricted spacers. J. Med. Chem. 1986, 29, 1650–1653. [Google Scholar] [CrossRef]
  174. Portoghese, P.S.; Larson, D.L.; Sayre, L.M.; Yim, C.B.; Ronsisvalle, G.; Tam, S.W.; Takemori, A.E. Opioid agonist and antagonist bivalent ligands. The relationship between spacer length and selectivity at multiple opioid receptors. J. Med. Chem. 1986, 29, 1855–1861. [Google Scholar] [CrossRef]
  175. Portoghese, P.S.; Larson, D.L.; Ronsisvalle, G.; Schiller, P.W.; Nguyen, T.M.; Lemieux, C.; Takemori, A.E. Hybrid bivalent ligands with opiate and enkephalin pharmacophores. J. Med. Chem. 1987, 30, 1991–1994. [Google Scholar] [CrossRef]
  176. Takemori, A.E.; Yim, C.B.; Larson, D.L.; Portoghese, P.S. Long-acting agonist and antagonist activities of naltrexamine bivalent ligands in mice. Eur. J. Pharmacol. 1990, 186, 285–288. [Google Scholar] [CrossRef]
  177. Costa, T.; Shimohigashi, Y.; Krumins, S.A.; Munson, P.J.; Rodbard, D. Dimeric pentapeptide enkephalin: a novel probe of delta opiate receptors. Life Sci. 1982, 31, 1625–1632. [Google Scholar] [CrossRef]
  178. Sasaki-Yagi, Y.; Kimura, S.; Imanishi, Y. Binding to opioid receptors of enkephalin derivatives taking alpha-helical conformation and its dimer. Int. J. Pept. Protein Res. 1991, 38, 378–384. [Google Scholar] [CrossRef]
  179. Yekkirala, A.S.; Kalyuzhny, A.E.; Portoghese, P.S. An Immunocytochemical-Derived Correlate for Evaluating the Bridging of Heteromeric Mu-Delta Opioid Protomers by Bivalent Ligands. ACS Chem. Biol. 2013, 8, 1412–1416. [Google Scholar] [CrossRef]
  180. Peng, X.; Neumeyer, J.L. Kappa receptor bivalent ligands. Curr. Top. Med. Chem. 2007, 7, 363–373. [Google Scholar] [CrossRef]
  181. Archer, S.; Glick, S.D.; Bidlack, J.M. Cyclazocine revisited. Neurochem. Res. 1996, 21, 1369–1373. [Google Scholar] [CrossRef]
  182. Zhang, S.; Yekkirala, A.; Tang, Y.; Portoghese, P.S. A bivalent ligand (KMN-21) antagonist for mu/kappa heterodimeric opioid receptors. Bioorganic Med. Chem. Lett. 2009, 19, 6978–6980. [Google Scholar] [CrossRef]
  183. Lahti, R.A.; Mickelson, M.M.; McCall, J.M.; Von Voigtlander, P.F. [3H]U-69593 a highly selective ligand for the opioid kappa receptor. Eur. J. Pharmacol. 1985, 109, 281–284. [Google Scholar] [CrossRef]
  184. Handa, B.K.; Land, A.C.; Lord, J.A.; Morgan, B.A.; Rance, M.J.; Smith, C.F. Analogues of beta-LPH61–64 possessing selective agonist activity at mu-opiate receptors. Eur. J. Pharmacol. 1981, 70, 531–540. [Google Scholar] [CrossRef]
  185. Daniels, D.J.; Kulkarni, A.; Xie, Z.; Bhushan, R.G.; Portoghese, P.S. A bivalent ligand (KDAN-18) containing delta-antagonist and kappa-agonist pharmacophores bridges delta2 and kappa1 opioid receptor phenotypes. J. Med. Chem. 2005, 48, 1713–1716. [Google Scholar] [CrossRef]
  186. Yekkirala, A.S.; Lunzer, M.M.; McCurdy, C.R.; Powers, M.D.; Kalyuzhny, A.E.; Roerig, S.C.; Portoghese, P.S. N-naphthoyl-beta-naltrexamine (NNTA), a highly selective and potent activator of mu/kappa-opioid heteromers. Proc. Natl. Acad. Sci. USA 2011, 108, 5098–5103. [Google Scholar] [CrossRef]
  187. Toll, L.; Khroyan, T.V.; Polgar, W.E.; Jiang, F.; Olsen, C.; Zaveri, N.T. Comparison of the antinociceptive and antirewarding profiles of novel bifunctional nociceptin receptor/mu-opioid receptor ligands: Implications for therapeutic applications. J. Pharmacol. Exp. Ther. 2009, 331, 954–964. [Google Scholar] [CrossRef]
  188. Hoskin, P.J.; Hanks, G.W. Opioid agonist-antagonist drugs in acute and chronic pain states. Drugs 1991, 41, 326–344. [Google Scholar] [CrossRef]
  189. Mello, N.K.; Kamien, J.B.; Lukas, S.E.; Drieze, J.; Mendelson, J.H. The effects of nalbuphine and butorphanol treatment on cocaine and food self-administration by rhesus monkeys. Neuropsychopharmacology 1993, 8, 45–55. [Google Scholar] [CrossRef]
  190. Bowen, C.A.; Negus, S.S.; Zong, R.; Neumeyer, J.L.; Bidlack, J.M.; Mello, N.K. Effects of mixed-action kappa/mu opioids on cocaine self-administration and cocaine discrimination by rhesus monkeys. Neuropsychopharmacology 2003, 28, 1125–1139. [Google Scholar]
  191. Huang, P.; Kehner, G.B.; Cowan, A.; Liu-Chen, L.Y. Comparison of pharmacological activities of buprenorphine and norbuprenorphine: norbuprenorphine is a potent opioid agonist. J. Pharmacol. Exp. Ther. 2001, 297, 688–695. [Google Scholar]
  192. Spagnolo, B.; Calo, G.; Polgar, W.E.; Jiang, F.; Olsen, C.M.; Berzetei-Gurske, I.; Khroyan, T.V.; Husbands, S.M.; Lewis, J.W.; Toll, L.; et al. Activities of mixed NOP and mu-opioid receptor ligands. Br. J. Pharmacol. 2008, 153, 609–619. [Google Scholar] [CrossRef]
  193. Lutfy, K.; Eitan, S.; Bryant, C.D.; Yang, Y.C.; Saliminejad, N.; Walwyn, W.; Kieffer, B.L.; Takeshima, H.; Carroll, F.I.; Maidment, N.T.; Evans, C.J. Buprenorphine-induced antinociception is mediated by mu-opioid receptors and compromised by concomitant activation of opioid receptor-like receptors. J. Neurosci. 2003, 23, 10331–10337. [Google Scholar]
  194. Montoya, I.D.; Gorelick, D.A.; Preston, K.L.; Schroeder, J.R.; Umbricht, A.; Cheskin, L.J.; Lange, W.R.; Contoreggi, C.; Johnson, R.E.; Fudala, P.J. Randomized trial of buprenorphine for treatment of concurrent opiate and cocaine dependence. Clin. Pharmacol. Ther. 2004, 75, 34–48. [Google Scholar] [CrossRef]
  195. Khroyan, T.V.; Zaveri, N.T.; Polgar, W.E.; Orduna, J.; Olsen, C.; Jiang, F.; Toll, L. SR 16435 [1-(1-(bicyclo[3.3.1]nonan-9-yl)piperidin-4-yl)indolin-2-one], a novel mixed nociceptin/orphanin FQ/mu-opioid receptor partial agonist: analgesic and rewarding properties in mice. J. Pharmacol. Exp. Ther. 2007, 320, 934–943. [Google Scholar]
  196. Zaveri, N.T.; Jiang, F.; Olsen, C.; Polgar, W.E.; Toll, L. Designing bifunctional NOP receptor-mu opioid receptor ligands from NOP receptor-selective scaffolds. Part I. Bioorganic Med. Chem. Lett. 2013, 23, 3308–3313. [Google Scholar] [CrossRef]
  197. Zaveri, N.T.; Jiang, F.; Olsen, C.M.; Deschamps, J.R.; Parrish, D.; Polgar, W.; Toll, L. A novel series of piperidin-4-yl-1,3-dihydroindol-2-ones as agonist and antagonist ligands at the nociceptin receptor. J. Med. Chem. 2004, 47, 2973–2976. [Google Scholar] [CrossRef]
  198. Zaveri, N.; Jiang, F.; Olsen, C.; Polgar, W.; Toll, L. Small-molecule agonists and antagonists of the opioid receptor-like receptor (ORL1, NOP): ligand-based analysis of structural factors influencing intrinsic activity at NOP. AAPS J. 2005, 7, E345–352. [Google Scholar] [CrossRef]
  199. Toll, L.; Berzetei-Gurske, I.P.; Polgar, W.E.; Brandt, S.R.; Adapa, I.D.; Rodriguez, L.; Schwartz, R.W.; Haggart, D.; O'Brien, A.; White, A.; et al. Standard binding and functional assays related to medications development division testing for potential cocaine and opiate narcotic treatment medications. NIDA Res. Monogr. 1998, 178, 440–466. [Google Scholar]
  200. Dooley, C.T.; Spaeth, C.G.; Berzetei-Gurske, I.P.; Craymer, K.; Adapa, I.D.; Brandt, S.R.; Houghten, R.A.; Toll, L. Binding and in vitro activities of peptides with high affinity for the nociceptin/orphanin FQ receptor, ORL1. J. Pharmacol. Exp. Ther. 1997, 283, 735–741. [Google Scholar]
  201. Gasparini, F.; Andres, H.; Flor, P.J.; Heinrich, M.; Inderbitzin, W.; Lingenhohl, K.; Muller, H.; Munk, V.C.; Omilusik, K.; Stierlin, C.; et al. [(3)H]-M-MPEP, a potent, subtype-selective radioligand for the metabotropic glutamate receptor subtype 5. Bioorganic Med. Chem. Lett. 2002, 12, 407–409. [Google Scholar] [CrossRef]
  202. Lee, J.S.; Ro, J.Y. Peripheral metabotropic glutamate receptor 5 mediates mechanical hypersensitivity in craniofacial muscle via protein kinase C dependent mechanisms. Neuroscience 2007, 146, 375–383. [Google Scholar] [CrossRef]
  203. Aoki, T.; Narita, M.; Shibasaki, M.; Suzuki, T. Metabotropic glutamate receptor 5 localized in the limbic forebrain is critical for the development of morphine-induced rewarding effect in mice. Eur. J. Neurosci. 2004, 20, 1633–1638. [Google Scholar] [CrossRef]
  204. Akgun, E.; Javed, M.I.; Lunzer, M.M.; Smeester, B.A.; Beitz, A.J.; Portoghese, P.S. Ligands that interact with putative MOR-mGluR5 heteromer in mice with inflammatory pain produce potent antinociception. Proc. Natl. Acad. Sci. USA 2013, 110, 11595–11599. [Google Scholar]
  205. Ballet, S.; Pietsch, M.; Abell, A.D. Multiple ligands in opioid research. Protein Pept. Lett. 2008, 15, 668–682. [Google Scholar] [CrossRef]
  206. Levac, B.A.; O'Dowd, B.F.; George, S.R. Oligomerization of opioid receptors: Generation of novel signaling units. Curr. Opin. Pharmacol. 2002, 2, 76–81. [Google Scholar] [CrossRef]
  207. Rios, C.D.; Jordan, B.A.; Gomes, I.; Devi, L.A. G-protein-coupled receptor dimerization: Modulation of receptor function. Pharmacol. Ther. 2001, 92, 71–87. [Google Scholar] [CrossRef]
  208. Thompson, A.A.; Liu, W.; Chun, E.; Katritch, V.; Wu, H.; Vardy, E.; Huang, X.P.; Trapella, C.; Guerrini, R.; Calo, G.; Roth, B.L.; Cherezov, V.; Stevens, R.C. Structure of the nociceptin/orphanin FQ receptor in complex with a peptide mimetic. Nature 2012, 485, 395–399. [Google Scholar] [CrossRef]
  209. Wu, H.; Wacker, D.; Mileni, M.; Katritch, V.; Han, G.W.; Vardy, E.; Liu, W.; Thompson, A.A.; Huang, X.P.; Carroll, F.I.; et al. Structure of the human kappa-opioid receptor in complex with JDTic. Nature 2012, 485, 327–332. [Google Scholar] [CrossRef]
  210. Granier, S.; Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Weis, W.I.; Kobilka, B.K. Structure of the delta-opioid receptor bound to naltrindole. Nature 2012, 485, 400–404. [Google Scholar] [CrossRef]
  211. Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Mathiesen, J.M.; Sunahara, R.K.; Pardo, L.; Weis, W.I.; Kobilka, B.K.; Granier, S. Crystal structure of the micro-opioid receptor bound to a morphinan antagonist. Nature 2012, 485, 321–326. [Google Scholar] [CrossRef]
  212. Filizola, M.; Devi, L.A. Structural biology: How opioid drugs bind to receptors. Nature 2012, 485, 314–317. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Lee, C.W.-S.; Ho, I.-K. Pharmacological Profiles of Oligomerized μ-Opioid Receptors. Cells 2013, 2, 689-714. https://doi.org/10.3390/cells2040689

AMA Style

Lee CW-S, Ho I-K. Pharmacological Profiles of Oligomerized μ-Opioid Receptors. Cells. 2013; 2(4):689-714. https://doi.org/10.3390/cells2040689

Chicago/Turabian Style

Lee, Cynthia Wei-Sheng, and Ing-Kang Ho. 2013. "Pharmacological Profiles of Oligomerized μ-Opioid Receptors" Cells 2, no. 4: 689-714. https://doi.org/10.3390/cells2040689

Article Metrics

Back to TopTop