Morphine-induced hyperalgesia involves mu opioid receptors and the metabolite morphine-3-glucuronide

Laurie-Anne Roeckel, Valérie Utard, David Reiss, Jinane Mouheiche, Hervé Maurin, Anne Robé, Emilie Audouard, John N Wood, Yannick Goumon, Frédéric Simonin, Claire Gaveriaux-Ruff, Laurie-Anne Roeckel, Valérie Utard, David Reiss, Jinane Mouheiche, Hervé Maurin, Anne Robé, Emilie Audouard, John N Wood, Yannick Goumon, Frédéric Simonin, Claire Gaveriaux-Ruff

Abstract

Opiates are potent analgesics but their clinical use is limited by side effects including analgesic tolerance and opioid-induced hyperalgesia (OIH). The Opiates produce analgesia and other adverse effects through activation of the mu opioid receptor (MOR) encoded by the Oprm1 gene. However, MOR and morphine metabolism involvement in OIH have been little explored. Hence, we examined MOR contribution to OIH by comparing morphine-induced hyperalgesia in wild type (WT) and MOR knockout (KO) mice. We found that repeated morphine administration led to analgesic tolerance and hyperalgesia in WT mice but not in MOR KO mice. The absence of OIH in MOR KO mice was found in both sexes, in two KO global mutant lines, and for mechanical, heat and cold pain modalities. In addition, the morphine metabolite morphine-3beta-D-glucuronide (M3G) elicited hyperalgesia in WT but not in MOR KO animals, as well as in both MOR flox and MOR-Nav1.8 sensory neuron conditional KO mice. M3G displayed significant binding to MOR and G-protein activation when using membranes from MOR-transfected cells or WT mice but not from MOR KO mice. Collectively our results show that MOR is involved in hyperalgesia induced by chronic morphine and its metabolite M3G.

Conflict of interest statement

The authors declare that they have no competing interests.

Figures

Figure 1
Figure 1
WT but not MOR KO mice show hyperalgesia under repeated morphine analgesic tolerance conditions. (A) The experimental design shows the schedule for nociceptive measures (arrows). Following baseline (BL), mice received 3 mg/kg morphine (ip) on day-1 (d1) to evaluate morphine-induced analgesia. Mice received thereafter 20 mg/kg morphine or saline control each day until day-7 (d7). On d8, nociceptive levels were measured before morphine administration to evaluate hyperalgesia, and following 3 mg/kg morphine to measure analgesic tolerance. Maintenance of hyperalgesia (OIH) was scored on the indicated days and latent sensitization on day 27. (B) Tail pressure (n = 18–19/group) and (C) tail immersion (48 °C, n = 19–25/group) results show analgesic tolerance (upper panels) in WT mice with repeated morphine. Pressure and heat hyperalgesia in WT but not KO mice are shown with the same mouse groups in bottom panels. **p < 0.01; ***p < 0.001 compared to baseline or d1. (D) Analgesic tolerance and hyperalgesia occur in MOR-flox but not MOR-CMV mice. **p < 0.01; ***p < 0.001 compared to baseline or to d1 (tail immersion 48 °C, n = 8–9/group). Two-way repeated ANOVA, Newman-Keuls test. (E) WT but not MOR KO mice show cold allodynia under morphine analgesic tolerance conditions. Following baseline (BL) cold response scoring on the 5 °C cold plate, mice received 20 mg/kg morphine or saline control once a day until day 7 as described in (A) and paw responses to cold were measured to evaluate cold allodynia. n = 13–16/group. ***p < 0.001 compared to BL. ANOVA repeated measures, Newman-Keuls test. Detailed statistical analyses are presented in Supplementary Table S2.
Figure 2
Figure 2
MOR KO mice do not show OIH persistence. Persistence of hyperalgesia after cessation from seven-day 20 mg/kg morphine in WT but not MOR-KO mice. (B, C) On day-27 when WT mice had recovered from OIH (Fig. 1A), the opioid antagonist naltrexone induced hyperalgesia i.e. latent sensitization in WT (B) but no MOR KO (C) mice. Tail immersion (48 °C, n = 11–14/group for OIH persistence and n = 6–8/group for latent sensitization following 20 mg/kg morphine; n = 8–10/group for persistence of OIH following 60 mg/kg morphine). **p < 0.01; ***p < 0.001; repeated measures ANOVA followed by Newman-Keuls test. Detailed statistical analyses are presented in Supplementary Table S3.
Figure 3
Figure 3
Morphine induced hyperalgesia in female and male WT mice but not MOR KO mice. Analgesia, analgesic tolerance and OIH were measured following the protocol described in Fig. 1A (see sex-grouped analysis in Fig. 1) on female (left panels) and male (right panels) mice with the tail immersion (AD), pressure analgesimeter (EH) and cold plate (I,J) tests for heat, pressure and cold hypersensitivities, respectively. Data are expressed as mean ± SEM. n = 7–11 mice/group. *, **, ***p < 0.05, 0.01 and 0.001 compared to the corresponding group (ANOVA repeated measures, Newman-Keuls. BL, Baseline. Detailed statistical analyses are presented in Supplementary Table S4.
Figure 4
Figure 4
MOR KO mice show no morphine-induced hyperalgesia following partial sciatic nerve ligation (pSNL). (A) The experimental design shows the schedule for nociceptive measures (arrows). Following determination of baseline (BL) and neuropathic hypersensitivity (14 days post-pSNL, partial sciatic nerve ligation) on a 5 °C cold plate, WT and MOR KO mice were treated for 7 days with either morphine or saline control solution. On d8, allodynia was measured with the cold plate, von Frey filaments and heat plantar tests. (B) Cold pSNL allodynia and OIH (n = 10–12/group), two-way ANOVA, Newman-Keuls test. Morphine induced OIH in WT mice,*p < 0.05 compared to post-pSNL. (C) Mechanical pSNL allodynia (n = 10–12/group), two-way ANOVA, Newman-Keuls test. pSNL induced mechanical allodynia; WT saline group and WT morphine group, p < 0.001 pSNL vs BL; KO saline group and KO morphine group, p < 0.01 pSNL vs BL. Morphine induced OIH in WT mice, *p < 0.05 compared to post-pSNL. (D) Heat pSNL allodynia (n = 10–12/group), two-way ANOVA. Morphine induced no OIH. pSNL induced cold, mechanical and heat allodynia. Detailed statistical analyses are presented in Supplementary Table S5.
Figure 5
Figure 5
MOR KO mice show no morphine-3-glucuronide-induced hyperalgesia. M3G concentration in mouse plasma (A), brain (B) and spinal cord (C) 2 hr following 10 mg/kg morphine administration. M3G was quantified using LC-MS/MS. Data are expressed as mean ± SEM. n = 9–11 mice/group. ***p < 0.001 in KO mice as compared to WT mice. (D,E) Acute M3G induces heat and touch hyperalgesia in WT but not KO mice. (n = 9–10/group) one-way ANOVA, Newman-Keuls test *p < 0.05; **p < 0.01 compared to baseline (BL). (F,G) Acute M3G induces heat and touch hypersensitivity in MOR Flox and cKO mice (n = 6–9/group) one-way ANOVA, Newman-Keuls test #0.05 < p < 0.1; *p < 0.05; ***p < 0.001 compared to baseline (BL). Detailed statistical analyses are presented in Supplementary Table S6.
Figure 6
Figure 6
M3G binding and signaling to MOR (A,B) [3H]-diprenorphine binding on membranes from MOR expressing HEK293-Glo cells (A) and [3H]-DAMGO binding on brain membranes preparations from WT mice (B). Membranes were incubated with increasing doses of DAMGO, morphine, fentanyl or M3G in assay buffer containing a fixed dose of opioid radioligand. 100% represents maximal radioligand binding in the absence of competitor. Results are presented as means ± SEM of 2 or 3 experiments. (CE) MOR agonist-induced [35S]-GTPγS binding to brain membranes preparations from WT (C) or KO mice (D). Membranes were incubated with increasing doses of DAMGO, morphine or M3G agonists (10−9 to 10−4M) in assay buffer containing [35S]-GTPγS. Basal level (100%) represents [35S]-GTPγS binding in the absence of agonist. DAMGO, morphine and M3G significantly stimulated [35S]-GTPγS binding to membranes from WT but not KO mice. Results are presented as means ± SEM of 3–5 experiments on 3 independent membrane preparations per genotype. (E) The selective MOR antagonist CTOP inhibits DAMGO, morphine and M3G-induced [35S]-GTPγS binding to membranes from WT mice. Brain membranes from WT mice were incubated with increasing doses of the mu opioid antagonist CTOP combined with 100uM of DAMGO, morphine or M3G. Activation results are presented as means ± SEM of 6–10 experiments from 3 independent membrane preparations. (F) Effect of increasing concentrations of DAMGO, morphine, fentanyl and M3G on forskolin-stimulated cAMP production in HEK293 stably expressing GloSensor and MOR receptor. In the cases indicated, 1 µM of naloxone had been added to the cells 15 min prior to the agonist. Dose−response curves were normalised to maximal DAMGO activity. Evaluations were performed three times in duplicate for M3G, two times in duplicates for other agonists. Results are presented as means ± SEM. (G) eYFP-labelled Beta-arrestin-2 translocation to Rluc-MOR in HEK293 cells after 5 to 10 min of cells activation by DAMGO, morphine, fentanyl or M3G at 37 °C. Agonist specific BRET1 ratio were determined by subtracting BRET1 ratio of non activated cells, and normalised to maximal DAMGO-triggered effect. Presented results are means ± SEM of 2 to 4 experiments. (H–J) Dynamic mass redistribution (DMR) signals observed in HEK293-Glo-cells after activation by various concentrations of DAMGO, morphine or M3G. Baseline of buffer-treated cells has been subtracted. Evaluations were performed three times in duplicate or triplicates. Presented figures show a representative experiment.

References

    1. Wolkerstorfer A, Handler N, Buschmann H. New approaches to treating pain. Bioorg Med Chem Lett. 2016;26:1103–1119. doi: 10.1016/j.bmcl.2015.12.103.
    1. Gaveriaux-Ruff C. Opiate-induced analgesia: contributions from mu, delta and kappa opioid receptors mouse mutants. Curr Pharm Des. 2013;19:7373–7381. doi: 10.2174/138161281942140105163727.
    1. Trang T, et al. Pain and Poppies: The Good, the Bad, and the Ugly of Opioid Analgesics. J Neurosci. 2015;35:13879–13888. doi: 10.1523/JNEUROSCI.2711-15.2015.
    1. Roeckel LA, Le Coz GM, Gaveriaux-Ruff C, Simonin F. Opioid-induced hyperalgesia: Cellular and molecular mechanisms. Neuroscience. 2016;338:160–182. doi: 10.1016/j.neuroscience.2016.06.029.
    1. Rivat C, Ballantyne J. The dark side of opioids in pain management: basic science explains clinical observation. Pain reports. 2016;1 doi: 10.1097/PR9.0000000000000570.
    1. Arout CA, Edens E, Petrakis IL, Sofuoglu M. Targeting Opioid-Induced Hyperalgesia in Clinical Treatment: Neurobiological Considerations. CNS Drugs. 2015;29:465–486. doi: 10.1007/s40263-015-0255-x.
    1. Hutchinson MR, et al. Exploring the neuroimmunopharmacology of opioids: an integrative review of mechanisms of central immune signaling and their implications for opioid analgesia. Pharmacol Rev. 2011;63:772–810. doi: 10.1124/pr.110.004135.
    1. Lewis SS, et al. Evidence that intrathecal morphine-3-glucuronide may cause pain enhancement via toll-like receptor 4/MD-2 and interleukin-1beta. Neuroscience. 2010;165:569–583. doi: 10.1016/j.neuroscience.2009.10.011.
    1. Due MR, et al. Neuroexcitatory effects of morphine-3-glucuronide are dependent on Toll-like receptor 4 signaling. J Neuroinflammation. 2012;9 doi: 10.1186/1742-2094-9-200.
    1. Bai L, et al. Toll-like receptor 4-mediated nuclear factor-kappaB activation in spinal cord contributes to chronic morphine-induced analgesic tolerance and hyperalgesia in rats. Neurosci Bull. 2014;30:936–948. doi: 10.1007/s12264-014-1483-7.
    1. Johnson JL, et al. Codeine-induced hyperalgesia and allodynia: investigating the role of glial activation. Transl Psychiatry. 2014;4 doi: 10.1038/tp.2014.121.
    1. Ellis A, et al. Morphine amplifies mechanical allodynia via TLR4 in a rat model of spinal cord injury. Brain Behav Immun. 2016;58:348–356. doi: 10.1016/j.bbi.2016.08.004.
    1. Ferrini F, et al. Morphine hyperalgesia gated through microglia-mediated disruption of neuronal Cl(-) homeostasis. Nat Neurosci. 2013;16:183–192. doi: 10.1038/nn.3295.
    1. Mattioli TA, et al. Toll-like receptor 4 mutant and null mice retain morphine-induced tolerance, hyperalgesia, and physical dependence. PLoS One. 2014;9 doi: 10.1371/journal.pone.0097361.
    1. Skolnick P, Davis H, Arnelle D, Deaver D. Translational potential of naloxone and naltrexone as TLR4 antagonists. Trends Pharmacol Sci. 2014;35:431–432. doi: 10.1016/j.tips.2014.06.008.
    1. Mogil JS. Perspective: Equality need not be painful. Nature. 2016;535 doi: 10.1038/535S7a.
    1. Melchior M, Poisbeau P, Gaumond I, Marchand S. Insights into the mechanisms and the emergence of sex-differences in pain. Neuroscience. 2016;338:63–80. doi: 10.1016/j.neuroscience.2016.05.007.
    1. Doyle HH, Murphy AZ. Sex differences in innate immunity and its impact on opioid pharmacology. J Neurosci Res. 2017;95:487–499. doi: 10.1002/jnr.23852.
    1. Rosen S, Ham B, Mogil JS. Sex differences in neuroimmunity and pain. J Neurosci Res. 2017;95:500–508. doi: 10.1002/jnr.23831.
    1. Matthes HW, 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. doi: 10.1038/383819a0.
    1. Weibel R, et al. Mu opioid receptors on primary afferent nav1.8 neurons contribute to opiate-induced analgesia: insight from conditional knockout mice. PLoS One. 2013;8 doi: 10.1371/journal.pone.0074706.
    1. Fukagawa H, Koyama T, Kakuyama M, Fukuda K. Microglial activation involved in morphine tolerance is not mediated by toll-like receptor 4. J Anesth. 2013;27:93–97. doi: 10.1007/s00540-012-1469-4.
    1. Corder G, et al. Constitutive mu-opioid receptor activity leads to long-term endogenous analgesia and dependence. Science. 2013;341:1394–1399. doi: 10.1126/science.1239403.
    1. Khabbazi S, Xie N, Pu W, Goumon Y, Parat MO. The TLR4-Active Morphine Metabolite Morphine-3-Glucuronide Does Not Elicit Macrophage Classical Activation In Vitro. Front Pharmacol. 2016;7 doi: 10.3389/fphar.2016.00441.
    1. Frolich N, et al. Distinct pharmacological properties of morphine metabolites at G(i)-protein and beta-arrestin signaling pathways activated by the human mu-opioid receptor. Biochem Pharmacol. 2011;81:1248–1254. doi: 10.1016/j.bcp.2011.03.001.
    1. Corder G, et al. Loss of mu opioid receptor signaling in nociceptors, but not microglia, abrogates morphine tolerance without disrupting analgesia. Nat Med. 2017;23:164–173. doi: 10.1038/nm.4262.
    1. Law PY, Reggio PH, Loh HH. Opioid receptors: toward separation of analgesic from undesirable effects. Trends Biochem Sci. 2013;38:275–282. doi: 10.1016/j.tibs.2013.03.003.
    1. Pasternak GW, Pan YX. Mu opioids and their receptors: evolution of a concept. Pharmacol Rev. 2013;65:1257–1317. doi: 10.1124/pr.112.007138.
    1. Convertino M, et al. mu-Opioid receptor 6-transmembrane isoform: A potential therapeutic target for new effective opioids. Prog Neuropsychopharmacol Biol Psychiatry. 2015;62:61–67. doi: 10.1016/j.pnpbp.2014.11.009.
    1. Gris P, et al. A novel alternatively spliced isoform of the mu-opioid receptor: functional antagonism. Mol Pain. 2010;6 doi: 10.1186/1744-8069-6-33.
    1. Crain SM, Shen KF. Ultra-low concentrations of naloxone selectively antagonize excitatory effects of morphine on sensory neurons, thereby increasing its antinociceptive potency and attenuating tolerance/dependence during chronic cotreatment. Proc Natl Acad Sci USA. 1995;92:10540–10544. doi: 10.1073/pnas.92.23.10540.
    1. Crain SM, Shen KF. Neuraminidase inhibitor, oseltamivir blocks GM1 ganglioside-regulated excitatory opioid receptor-mediated hyperalgesia, enhances opioid analgesia and attenuates tolerance in mice. Brain Res. 2004;995:260–266. doi: 10.1016/j.brainres.2003.09.068.
    1. Oladosu FA, et al. Mu Opioid Splice Variant MOR-1K Contributes to the Development of Opioid-Induced Hyperalgesia. PLoS One. 2015;10 doi: 10.1371/journal.pone.0135711.
    1. Machelska H, Celik MO. Recent advances in understanding neuropathic pain: glia, sex differences, and epigenetics. F1000Res. 2016;5 doi: 10.12688/f1000research.9621.1.
    1. Sorge RE, et al. Different immune cells mediate mechanical pain hypersensitivity in male and female mice. Nat Neurosci. 2015;18:1081–1083. doi: 10.1038/nn.4053.
    1. Taves S, et al. Spinal inhibition of p38 MAP kinase reduces inflammatory and neuropathic pain in male but not female mice: Sex-dependent microglial signaling in the spinal cord. Brain Behav Immun. 2016;55:70–81. doi: 10.1016/j.bbi.2015.10.006.
    1. Holtman JR, Jr., Wala EP. Characterization of morphine-induced hyperalgesia in male and female rats. Pain. 2005;114:62–70. doi: 10.1016/j.pain.2004.11.014.
    1. Arout CA, Caldwell M, Rossi G, Kest B. Spinal and supraspinal N-methyl-D-aspartate and melanocortin-1 receptors contribute to a qualitative sex difference in morphine-induced hyperalgesia. Physiol Behav. 2015;147:364–372. doi: 10.1016/j.physbeh.2015.05.006.
    1. Juni A, Klein G, Kowalczyk B, Ragnauth A, Kest B. Sex differences in hyperalgesia during morphine infusion: effect of gonadectomy and estrogen treatment. Neuropharmacology. 2008;54:1264–1270. doi: 10.1016/j.neuropharm.2008.04.004.
    1. Juni A, et al. Sex-specific mediation of opioid-induced hyperalgesia by the melanocortin-1 receptor. Anesthesiology. 2010;112:181–188. doi: 10.1097/ALN.0b013e3181c53849.
    1. Melik Parsadaniantz S, Rivat C, Rostene W, Reaux-Le Goazigo A. Opioid and chemokine receptor crosstalk: a promising target for pain therapy? Nat Rev Neurosci. 2015;16:69–78. doi: 10.1038/nrn3858.
    1. Thomas J, Mustafa S, Johnson J, Nicotra L, Hutchinson M. The relationship between opioids and immune signalling in the spinal cord. Handb Exp Pharmacol. 2015;227:207–238. doi: 10.1007/978-3-662-46450-2_11.
    1. Rubovitch V, Pick CG, Sarne Y. Is withdrawal hyperalgesia in morphine-dependent mice a direct effect of a low concentration of the residual drug? Addict Biol. 2009;14:438–446. doi: 10.1111/j.1369-1600.2009.00164.x.
    1. Handal M, Grung M, Skurtveit S, Ripel A, Morland J. Pharmacokinetic differences of morphine and morphine-glucuronides are reflected in locomotor activity. Pharmacol Biochem Behav. 2002;73:883–892. doi: 10.1016/S0091-3057(02)00925-5.
    1. Drewe J, et al. Effect of P-glycoprotein modulation on the clinical pharmacokinetics and adverse effects of morphine. Br J Clin Pharmacol. 2000;50:237–246. doi: 10.1046/j.1365-2125.2000.00226.x.
    1. Mignat C, Jansen R, Ziegler A. Plasma and cerebrospinal fluid concentrations of morphine and morphine glucuronides in rabbits receiving single and repeated doses of morphine. J Pharm Pharmacol. 1995;47:171–175. doi: 10.1111/j.2042-7158.1995.tb05772.x.
    1. Juni A, Klein G, Kest B. Morphine hyperalgesia in mice is unrelated to opioid activity, analgesia, or tolerance: evidence for multiple diverse hyperalgesic systems. Brain Res. 2006;1070:35–44. doi: 10.1016/j.brainres.2005.11.054.
    1. Mignat C, Wille U, Ziegler A. Affinity profiles of morphine, codeine, dihydrocodeine and their glucuronides at opioid receptor subtypes. Life Sci. 1995;56:793–799. doi: 10.1016/0024-3205(95)00010-4.
    1. Thompson GL, Kelly E, Christopoulos A, Canals M. Novel GPCR paradigms at the mu-opioid receptor. Br J Pharmacol. 2015;172:287–296. doi: 10.1111/bph.12600.
    1. Yang Z, et al. Reverse of Acute and Chronic Morphine Tolerance by Lithocholic Acid via Down-Regulating UGT2B7. Front Pharmacol. 2016;7
    1. Su W, Pasternak GW. The role of multidrug resistance-associated protein in the blood-brain barrier and opioid analgesia. Synapse. 2013;67:609–619. doi: 10.1002/syn.21667.
    1. Yang ZZ, et al. siRNA capsulated brain-targeted nanoparticles specifically knock down OATP2B1 in mice: a mechanism for acute morphine tolerance suppression. Sci Rep. 2016;6 doi: 10.1038/srep33338.
    1. Khabbazi S, Goumon Y, Parat MO. Morphine Modulates Interleukin-4- or Breast Cancer Cell-induced Pro-metastatic Activation of Macrophages. Sci Rep. 2015;5 doi: 10.1038/srep11389.
    1. Wan J, Ma J, Anand V, Ramakrishnan S, Roy S. Morphine potentiates LPS-induced autophagy initiation but inhibits autophagosomal maturation through distinct TLR4-dependent and independent pathways. Acta Physiol (Oxf) 2015;214:189–199. doi: 10.1111/apha.12506.
    1. Xie N, et al. Activation of mu-opioid receptor and Toll-like receptor 4 by plasma from morphine-treated mice. Brain Behav Immun. 2017;61:244–258. doi: 10.1016/j.bbi.2016.12.002.
    1. Gessi S, et al. The activation of mu-opioid receptor potentiates LPS-induced NF-kB promoting an inflammatory phenotype in microglia. FEBS Lett. 2016;590:2813–2826. doi: 10.1002/1873-3468.12313.
    1. Lavin Y, et al. Tissue-resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell. 2014;159:1312–1326. doi: 10.1016/j.cell.2014.11.018.
    1. Denk F, Crow M, Didangelos A, Lopes DM, McMahon SB. Persistent Alterations in Microglial Enhancers in a Model of Chronic Pain. Cell Rep. 2016;15:1771–1781. doi: 10.1016/j.celrep.2016.04.063.
    1. Lagerstrom MC, et al. VGLUT2-dependent sensory neurons in the TRPV1 population regulate pain and itch. Neuron. 2010;68:529–542. doi: 10.1016/j.neuron.2010.09.016.
    1. Usoskin D, et al. Unbiased classification of sensory neuron types by large-scale single-cell RNA sequencing. Nat Neurosci. 2015;18:145–153. doi: 10.1038/nn.3881.
    1. Vanderah TW, et al. Tonic descending facilitation from the rostral ventromedial medulla mediates opioid-induced abnormal pain and antinociceptive tolerance. J Neurosci. 2001;21:279–286.
    1. Cahill CM, Walwyn W, Taylor AM, Pradhan AA, Evans CJ. Allostatic Mechanisms of Opioid Tolerance Beyond Desensitization and Downregulation. Trends Pharmacol Sci. 2016;37:963–976. doi: 10.1016/j.tips.2016.08.002.
    1. Gerhold KJ, Drdla-Schutting R, Honsek SD, Forsthuber L, Sandkuhler J. Pronociceptive and Antinociceptive Effects of Buprenorphine in the Spinal Cord Dorsal Horn Cover a Dose Range of Four Orders of Magnitude. J Neurosci. 2015;35:9580–9594. doi: 10.1523/JNEUROSCI.0731-14.2015.
    1. Bobeck EN, Ingram SL, Hermes SM, Aicher SA, Morgan MM. Ligand-biased activation of extracellular signal-regulated kinase 1/2 leads to differences in opioid induced antinociception and tolerance. Behav Brain Res. 2016;298:17–24. doi: 10.1016/j.bbr.2015.10.032.
    1. Manglik A, et al. Structure-based discovery of opioid analgesics with reduced side effects. Nature. 2016;537:185–190. doi: 10.1038/nature19112.
    1. Gendron L, Cahill CM, von Zastrow M, Schiller PW, Pineyro G. Molecular Pharmacology of delta-Opioid Receptors. Pharmacol Rev. 2016;68:631–700. doi: 10.1124/pr.114.008979.
    1. Samoshkin A, et al. Structural and functional interactions between six-transmembrane mu-opioid receptors and beta2-adrenoreceptors modulate opioid signaling. Sci Rep. 2015;5 doi: 10.1038/srep18198.
    1. McGrath JC, Drummond GB, McLachlan EM, Kilkenny C, Wainwright CL. Guidelines for reporting experiments involving animals: the ARRIVE guidelines. Br J Pharmacol. 2010;160:1573–1576. doi: 10.1111/j.1476-5381.2010.00873.x.
    1. Malmberg AB, Basbaum AI. Partial sciatic nerve injury in the mouse as a model of neuropathic pain: behavioral and neuroanatomical correlates. Pain. 1998;76:215–222. doi: 10.1016/S0304-3959(98)00045-1.
    1. Gaveriaux-Ruff C, et al. Genetic ablation of delta opioid receptors in nociceptive sensory neurons increases chronic pain and abolishes opioid analgesia. Pain. 2011;152:1238–1248. doi: 10.1016/j.pain.2010.12.031.
    1. Becker JA, et al. Ligands for kappa-opioid and ORL1 receptors identified from a conformationally constrained peptide combinatorial library. J Biol Chem. 1999;274:27513–27522. doi: 10.1074/jbc.274.39.27513.
    1. Michel G, et al. Plasma membrane translocation of REDD1 governed by GPCRs contributes to mTORC1 activation. J Cell Sci. 2014;127:773–787. doi: 10.1242/jcs.136432.
    1. Morse M, Tran E, Sun H, Levenson R, Fang Y. Ligand-directed functional selectivity at the mu opioid receptor revealed by label-free integrative pharmacology on-target. PLoS One. 2011;6 doi: 10.1371/journal.pone.0025643.

Source: PubMed

3
Suscribir