SciELO - Scientific Electronic Library Online

 
vol.33 número2Estimulación magnética transcraneal para síntomas negativos en la esquizofrenia: una revisiónSistemas de memoria: reseña histórica, clasificación y conceptos actuales. Segunda parte: Sistemas de memoria de largo plazo: Memoria episódica, sistemas de memoria no declarativa y memoria de trabajo índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

Links relacionados

  • No hay artículos similaresSimilares en SciELO

Compartir


Salud mental

versión impresa ISSN 0185-3325

Salud Ment vol.33 no.2 México mar./abr. 2010

 

Actualización por temas

 

Endomorphin peptides: pharmacological and functional implications of these opioid peptides in the brain of mammals. Part one

 

Las endomorfinas: Implicaciones farmacológicas y funcionales de estos péptidos opioides en el cerebro de los mamíferos

 

Philippe Leff Gelman,1* Norma Estela González Herrera,2 Maura Epifanía Matus Ortega,1 Lenin Pavón Romero,3 Carlos Téllez Santillán,3 Alberto Salazar Juárez,1 Benito Antón Palma1

 

1 Laboratorio de Neurobiología Molecular y Neuroquímica de Adicciones. Subdirección de Investigaciones Clínicas, Instituto Nacional de Psiquiatría Ramón de la Fuente Muñiz.

2 Laboratorio de Oncología Molecular. Sección de Posgrado e Investigación. Escuela Superior de Medicina, Instituto Politécnico Nacional.

3 Laboratorio de Psicoinmunología. Dirección de Neurociencias. Instituto Nacional de Psiquiatría Ramón de la Fuente Muñiz.

 

*Corresponding:
Philippe Leff.
Laboratorio de Neurobiología Molecular y Neuroquímica de Adicciones.
Subdirección de Investigaciones Clínicas,
Instituto Nacional de Psiquiatría Ramón de la Fuente Muñiz.
Calzada México–Xochimilco 101,
San Lorenzo Huipulco,
Tlalpan, 14370,
México D F,
email: pleff@imp.edu.mx

 

Recibido: 3 de noviembre de 2009.
Aceptado: 29 de enero de 2010.

 

Abstract

The present paper describes several aspects of the biological activities, physiological and behavioral responses displayed by the most recent discovered opioid peptides: endomorphins. Endormorphins comprise two endogenous C–terminal amide tetrapeptides, named as endomorphin–1 (EM1; Tyr–Pro–Trp–Phe–NH2) and endomorphin–2 (EM2; Tyr–Pro–Phe–Phe–NH2), which were discovered a decade ago (1997) by Zadina's group. Initially, they reported the identification of two endogenous opioid peptides that displayed high binding affinities and selectivities for the µ–opioid receptor among other identified and cloned opioid receptors. These led authors to support the hypothesis that endomorphin peptides represent the endogenous ligand agonists for the µ–opioid receptor. Both peptides were identified and isolated from bovine and human brains. They consist of four amino acids that share a 75% structural homology among amino acids, and which display the structural α–amidated form of C–terminal –Phe– residue, as demonstrated for many other bioactive neuropeptides.

These peptides are structurally distinct from other endogenous opioid substances identified in the brain of mammals, although they share some similarities with other amide terapeptides such as Tyr–W–MIF–1, found also in the mammalian brain. Here, we review the structure–relationship activity of both endomorphin molecules comparing their binding properties to different opioid receptors. Both EM1/EM2 peptides appear to be vulnerable to enzymatic degradation when exposed to the activities of different proteolytic enzymes, as occurs with many other neuroactive peptides found in the SNC of mammals. Immunohistochemical studies showed the wide and asymmetric distribution of both EM1–2 peptides in the brain, leading to the extensive pharmacological, cellular, and physiological studies that demonstrated the wide and varied bioactivities displayed by these peptides at both central and peripheral tissues. These studies led several authors to suggest the potential endogenous role of these peptides in major physiological processes (e.g. analgesia or antinociception). Based on the generation of specific (rabbit) polyclonal antibodies and the use of combined radioimmunoassay (RIA) techniques and immunohistochemical procedures, it was shown the wide distribution of EM1–2–LI (endomorphin1–2–like immunoreactivities) throughout the brain of different species (e.g. rat, primate, human), particularly co–localized in specific areas where µ–opioid receptor has been shown to be expressed. IHC mapping of endomorphin material in the CNS showed a parallelism with the neuroanatomical distribution of other endogenous opioid peptides (e.g. Met/Leu–enk, Dynorphin A, β–endorphin) previously reported.

These studies showed for instance that, whereas EM1–LI was shown to be widely and densely distributed throughout the brain, particularly in forebrain structures (e.g. nucleus accumbens [NAc]; cortex [Cx]; amygdale [AMG]; thalamus [Th], the hypothalamus [Hyp], the striatum [CPu]), including the upper brainstem (BS); and dorsal root ganglia (DRG); EM2–LI is highly expressed in spinal cord and lower brainstem. Interesting enough is the demonstration of the expression of EM1–2–LI outside the CNS (e.g. spleen, thymus and blood), and detected in immune cells (e.g. macrophages/monocytes, lymphocytes, and polymophonuclear leucocytes) surrounding inflammatory foci. Pharmacological studies showed that these peptides displace with high potency several µ–opioid receptor ligands agonists in a concentration–dependent manner. Moreover, EM1–2 peptides have been shown to modulate the release of several conventional transmitters from neurons (e.g. DA, NA, 5–HT, ACh) besides on active neurohormones. Additionally, in vitro and in vivo studies showed that both EM–1/EM–2 peptides produce their pharmacological and biological effects by stimulating either µ1 or µ2–opioid receptors, which mediate the distinct pharmacological activities detected for each peptide. Cellular studies showed that both EM–1/EM–2 peptides induce a potent granule/vesicle endocytosis and trafficking of µ–opioid receptor in cells transfected with the µ–opioid receptor cDNA; following some endocytosis responses and µ–opioid receptor trafficking mechanisms shown in enteric neurons; cells previously reported to express naturally µ–opioid binding sites on cells.

Endomorphins have been shown to induce potent antinociceptive responses after ICV or IT administration into mice; to modulate nociceptive transmission and pain sensation into the brain after stimulating peripheral nociceptors on primary neuronal afferents; and to generate cross–tolerance between endomorphin peptides and between EM1 and opiate compounds, such as morphine.

Key words: Endomorphins, peptides, nociception, analgesia, muopioid receptor, physiology, immunoreactivity.

 

Resumen

Este artículo resume varios aspectos de las múltiples actividades biológicas, celulares, efectos farmacológicos, respuestas fisiológicas y conductuales de dos nuevas sustancias peptídicas de naturaleza opioide, descubiertas recientemente y denominadas endomorfinas. Las endomorfinas son dos péptidos opioides, clasificados como endomorfina–1 (EM1, Tyr–Pro–Trp–Phe–NH2) y endomorfina–2 (EM2, Tyr–Pro–Phe–Phe–NH2), cuyas secuencias peptídicas fueron identificadas y aisladas del cerebro de bovino y humano por el grupo de Zadina en 1997. Estudios de unión radioligando–receptor demostraron que estos péptidos se unen con alta afinidad de unión al receptor opioide µ en relación con su capacidad de unión a otros subtipos de receptores opioides (kappa [κ], delta [δ] ), previamente identificados en el SNC de mamíferos. Ambos péptidos están compuestos por cuatro aminoácidos y son estructuralmente distintos de las demás sustancias opioides endógenas conocidas.

Esta revisión detalla con precisión diversos aspectos de la farmacología y actividades celulares de estos opioides y sus implicaciones en la modulación de distintas circuitos o vías neurales y funcionamiento del SNC de los mamíferos, respectivamente. Los estudios relacionados con la función estructura–actividad de estos péptidos han mostrado que, al igual que la mayoría de los péptidos bioactivos endógenos de naturaleza opioide y no opioide, son vulnerables a la escisión peptídica por cortes enzimáticos mediante la exposición a distintas enzimas proteolíticas que pudiesen participar en la degradación endógena de las endomorfinas, y la obtención de diversos productos de degradación. Asimismo, este artículo menciona la amplia distribución neuroanatómica que poseen las endomorfinas en distintas regiones del cerebro, particularmente en aquellas que regulan el procesamiento y la transmisión de la información nociceptiva y que, por tanto, reflejan el papel potencial de estos péptidos en procesos fisiológicos de analgesia, entre muchos otros (memoria y otro aprendizaje). En este contexto, diferentes estudios basados en el empleo de ensayos inmunológicos (radioinmunoensayos [RIA] y técnicas de inmunohistoquímica [IHC]) que requieren el uso de anticuerpos específicos generados contra las secuencias consenso de las endomorfinas mostraron una amplia distribución de material inmunoreactivo a endomorfina (vg., EM1–LI, EM2–LI) en tejidos neurales de humano, bovino y roedores. Por ejemplo, la EM1–LI mostró una distribución relativamente abundante en una gran mayoría de las regiones del SNC de mamíferos estudiados, particularmente en la región rostral y superior del tallo cerebral, así como en el núcleo accumbens (NAc), la corteza prefrontal y frontal (PFCx), la amígdala (AMG), el tálamo (TH), el hipotálamo (HPT), el estriado (CPu) y fibras nerviosas de la raíz del ganglio dorsal (DRG). En contraste, la expresión de EMZ mostró ser muy abundante en la región de la médula espinal y en la región caudal del tallo cerebral.

La distribución de material inmunoreactivo a EM1–2 en el SNC de mamíferos mostró similitudes en cuanto a la distribución neuroanatómica reportada para otros péptidos opioides endógenos, previamente identificados (vg., encefalinas, dinorfinas, endorfinas). Así mismo, estudios paralelos lograron identificar la presencia de EM1–2–LI en órganos periféricos (vg., bazo, timo, células inflamatorias del tipo de macrófagos–monocitos, linfocitos y leucocitos PMN) y en plasma. Más aún, diversos estudios farmacológicos han mostrado que las actividades biológicas y respuestas fisiológicas de las EM1–2 están mediadas a través de la estimulación de los subtipos de receptores opioides µ1 y µ2. Estudios de inmunohistoquímica (IHC) demostraron la colocalización del receptor opioide µ y las EM1–2 en diversas regiones del SNC de mamiferos. Esto ha permitido proponer que las EM1–2 representan una nueva familia de péptidos opioides con funciones neuromoduladoras relevantes en el SNC, las cuales intervienen en la regulación de los procesos biológicos de percepción del dolor; respuestas de estrés; funciones límbicas de placer y recompensa inducidas por incentivos naturales y/o sustancias psicotrópicas; funciones de estado de alerta y vigilia, funciones cognitivas (de aprendizaje y memoria) y actividades de regulación neuroendócrina.

Además, diversos estudios celulares han mostrado que ambos péptidos opioides son capaces de inducir la internalización aguda o endocitosis del receptor opioide µ en células somáticas transfectadas con el ADN (ADNc) que codifica este mismo receptor opioide. Al igual que otros péptidos opioides (v.g., encefalinas), diversos estudios mostraron el catabolismo enzimático de estos péptidos amidados mediante la actividad de enzimas proteolíticas (v.g., carboxipeptidasa Y, aminopeptidasa M), lo que ha permitido sugerir que estos péptidos opioides son degradados por rutas de degradación enzimática similares que rigen para múltiples péptidos bioactivos moduladores en el SNC de los mamíferos. Al igual que otros péptidos endógenos, ambas endomorfinas mostraron la capacidad de modular la liberación neuronal de neurotransmisores (DA, NA, 5–HT, ACh) y hormonas peptídicas en áreas específicas del cerebro de los mamíferos. Asimismo, ambos péptidos mostraron una capacidad de generar efectos antinociceptivos potentes en forma dosis–dependiente posterior a su administración ICV o IT en animales experimentales, además de generar respuestas de tolerancia cruzada entre ambas endomorfinas y/o entre la EM1 y alcaloides opiáceos del tipo de la morfina.

Palabras clave: Endomorfina, péptidos, nocicepción, analgesia, receptor opioide mu, fisiología, inmuno reactividad.

 

INTRODUCTION

Extensive studies on opiate receptor pharmacology led to the initial identification of the three major families of endogenous opioid ligands referred to as endorphins,1,2 enkephalins,3 and dynorphins,4,5 and most recently the nociceptin/OFQ peptide.6,7 Cloning and molecular characterization of the genes that encode the different propeptide precursors —namely pro–opiomelonocortin,2 proenkephalin A,8–10 prodynorphin,11 and pronociceptin/ OFQ peptide—12–14 showed that these large opioid propeptide precursors contain different copies of active peptide sequences flanked by pairs of basic amino acids (e.g. Arg–Arg or Arg–Lys), which upon tissue specific endo and exopeptidases15–17 mature opioid peptides are released from LDCV vesicles into the extracellular space by stimuli–coupling dependent mechanisms.18–21 Activation of the G–protein coupled membrane opioid receptors22–26 by endogenous opioid peptides results in the expression of a vast number of pharmacological, physiological, neuroendocrine, and behavioral responses in the CNS of mammals and humans.27

 

I. STRUCTURE–RELATIONSHIP ACTIVITY OF ENDOMORPHINS

Naturally occurring opioid peptides identified and cloned from mammalian tissues —such as casomorphin heptapeptide (Tyr–Pro–Phe–Pro–Gly–Pro–Ile), obtained from tryptic digests of β–casein;28 hemorphin–4 (Tyr–Pro–Trp–Thr), obtained from hemoglobin digests;29 the amide tetrapeptides, Tyr–MIF–11 (Tyr–Pro–Leu–Gly–NH2), and Tyr–W–MIF–1 (Tyr–Pro–Trp–Gly–NH2), isolated from the bovine and human brains, respectively— display preferential binding affinities and selectivity for the µ–opioid receptor when compared to both δ– and κ–opioid receptors, respectively.30,31 Nonetheless the previous observation, these peptides displayed low–to–medium binding affinities (Ki = 10–80 nM) for the µ–opioid receptor.32

A large number of synthetic analogs of Tyr–W–MIF–1 peptide (Tyr–Pro–Trp–Gly–NH2), which that contain all possible natural amino acid substitutions at position 4 with exception of the Gly residue, were used to screen out peptide compounds capable of displaying high and selective binding properties at the µ–opioid receptor (Kid ≤1.0 nM). Zadina's group reported the identification of a potent peptide sequence (Tyr–Pro–Trp–Phe–NH2), named as endomorphin 1 (EM1), which displayed the highest binding affinity (Ki=360 pM, 0.36 nM) and selectivity for the µ–opioid receptor (e.g. 4000–15,000–fold over δ– and κ–opioid receptors, respectively), when compared to other endogenous opioid ligands, including β–endorphin.33,34 This peptide sequence identified and isolated from cortical tissues of mammalian brains (e.g. bovine and human).33,35 This newly identified opioid peptide was extremely potent in the guinea pig ileum assay, a classic test for µ–opioid receptor agonist activity33 shown to display a potent and specific antinociceptive effect in vivo, as shown in the tail–flick test33,35 and thereby modulate both supraspinal and spinal antinociception) after ICV or IT injection of the peptide.36–40 A structurally close related peptide sequence, named as, endomorphin–2 (EM2) (Tyr–Pro–Phe–Phe–NH2) was subsequently isolated from mammalian tissues33 and shown to be almost as potent as endomorphin–1.33,35

Thus, endomorphins molecules represent the first peptides isolated from brain shown to bind the µ–opioid receptor with high affinity and selectivity and categorized as the natural opioid peptide ligands for the µ–opioid receptor in mammals, including humans.34 However, despite the whole range of activities reported for these peptides (see below), no existing evidences have been reported on the cloning of specific mRNA(s) encoding a precursor molecule(s) for these opioide peptides, as demonstrated for other major family of opioid peptides found in mammals (e.g. proenkephalin A, prodynorphin, POMC, proNOC/ OFQ peptide).8,12,41

Structurally, the peptide sequences encoded by each endomorphin molecules (EM1, EM2) have been shown to be completely different from the consensus motif encoded by several endogenous opioid peptides (e.g. endorphins, enkephalins, and dynorphins) as depicted in figure 1. As shown, classical opioid peptides share the Tyr–Gly–Gly–Phe motif at the N–terminus domain of each peptide, whereas EM1 and EM2 contain two pharmacophoric amino acid residues (Tyr [1] and Trp[3] in EM1; Phe[3] replacing Trp [3] in EM2) joined together via a Pro(2) amino acid residue, used as molecular spacer in both peptides. Both phenolic groups (Tyr, Phe) and/or aromatic rings (Trp, Phe) encoded by each amino acid residue represent structural and chemical requirements for opioid receptor recognition.32

Using nuclear magnetic resonance approaches to reveal amino acid conformation of endomorphin molecules and the structural domains required to recognize opioid receptors, these studies revealed that both Tyr(1) and Trp(3) side chains display opposite orientations with respect to Pro(2), which provides the necessary stereochemical properties of endomorphin–1 for binding the µ–opioid re–ceptor.42 Moreover, using pseudo–proline containing analogs of EM2, it was shown that the Tyr–Pro amide bond forms a cis–conformation structure along the peptide backbone, allowing its molecular conformation to bind and recognize the µ–opioid receptor.43

 

II. DISTRIBUTION OF ENDOMORPHINS IN CENTRAL AND PERIPHERAL TISSUES OF MAMMALS

Immunological assays (e.g. radioimmunoassays, immuno–cytochemical analyses) using different rabbit antisera raised against synthetic peptides of EM1 and EM2 showed that endomorphin1–2–like immunoreactivities (EM1–2–LIs) are widely distributed throughout the neuroaxis of the rat, primate, and human brains,35,44–50 displaying a preferential localization in areas where MOR–LI (µ–opioid receptor–immunoreactivity) is highly expressed (see representative immunoreactive areas in figure 2).51 Although EM1–2–LIs were localized in cells and fibers in similar regions of the rat CNS (e.g. stria terminalis [ST]; periaqueductal gray [PAG]; locus coeruleus [LC]; parabrachial nucleus [PBN]; hypothalamus [HPTH]; nucleus of the solitary tract [NTS]), major differences about the neural distribution of these peptides were detected.44,47,49 For instance, EM1–LI was shown to be widely and densely distributed in several forebrain structures (e.g. nucleus accumbens [NAC]; the cortex [Cx]; amygdale [AM]; thalamus [TH]; the hypothalamus [HTH]; striatum [ST]), together with the upper brainstem (e.g., nucleus of the solitary tract [NTS]) and the dorsal root ganglia,47,50 whereas EM2–LI was localized mostly in the spinal cord and lower brainstem,47,49 showing densely stained cell bodies and axon fibers within the nucleus of the solitary tract (NTS); substantia gelatinosa of the medulla and superficial layers of the spinal cord/dorsal horn, including the lateral hypothalamus.44,47,50 Although these studies showed that the anatomical distribution of EM1–2–like immunoreactive material along the rat CNS parallels the regional localization of several endogenous opioid peptides and their cognate opioid receptors;52,53 differences observed between the endomorphin material, enkephalin,54,55 and dynorphin56 peptide systems showed that EM2–LI in the mammalian brain47,50 is highly analogous to the distribution of the µ–opioid receptor peptide ligand, β–endorphin (BE).57

Moreover, it has been shown that endomorphins do not co–localize with brain areas expressing µ–opioid receptors (e.g. striatum) containing low–to negligible amounts of EM1–2–LI material.49 This asymmetric localization between endomorphins peptides and their corresponding binding sites in areas of the rat brain does not preclude the synaptic activation of membrane bound opioid receptors. However, µ–opioid receptors and EM1–2–LI fibers have been shown to co–localize in limbic structures of the rat brain (e.g. septal nuclei, the bed nucleus of the stria terminalis, NAC, amygdaloid complex, hypothalamic nuclei), whereas in other rat brain regions (e.g. amygdala, thalamus, hypothalamus, PAG), showing for instance the co–localization of low levels of EM1–2–LI in areas displaying high density of opioid–binding sites (figure 2).47,49,58

Very few reports have documented the extraneuronal distribution of endomorphin material in peripheral tissues,32 in contrast with the vast reports describing the distribution and expression of µ–opioid receptors in extraneuronal tissues.27 For instance, using RIA assays in combination with reversed phase high–performance liquid chromatography techniques (RP–HPLC),59 it was found the expression of detectable amounts of EM1–LI and EM2–LI in human spleen, as compared to high levels of EM1–2–like immunoreactive material detected in non–neuronal tissues of the rat, such as spleen and thymus, including plasma. These results contrast the low amounts of EM1–2–LI detected in anterior and posterior rat pituitaries, showing that secretion of opioid material from the pituitary gland does not contribute at all to the significant high levels of EM1–2 detected in plasma.59 Based on these results, authors proposed that EM1–2–like immunoreactive material secreted from the nerve fibers and/or presynaptic terminals localized in the spinal cord may explain the high content of endomorphin material observed in human and/or rat plasma samples.59 Moreover, endomorphins were shown to be expressed by immune cells (e.g., macrophages/monocytes, lymphocytes, and polymorphonuclear leukocytes),60,61 including inflamed subcutaneous tissue.62 Endomorphins have been found to alter a variety of immune parameters and functions,63–66 as well as to inhibit spleen cell antibody formation when added to in vitro culture systems at femtomolar concentrations (10–13 M to 10–15 M) and reverted after application of specific EM1/EM2 polyclonal rabbit antisera.67

 

III. BINDING ASSAYS AND CELLULAR ACTIONS OF ENDOMORPHINS

A large number of pharmacological and molecular studies have clearly demonstrated that the µ–opioid receptor mediates the spectrum of opiate–related activities of endomorphins in the CNS of mammals.32 The proposed hypothesis suggesting that endomorphin molecules are the endogenous ligands for the µ–opioid receptor, aroused from classic binding assays and functional studies, which showed that both endomorphin peptides displaced DAMGO (Tyr–DAla–Gly–MePhe–Gly–ol); naloxone, and other µ–opioid receptor–selective ligands, in a concentration–dependent manner68 and stimulate the binding of [35S]guanosine 5'–O–(3–thio)triphosphate ([35S]γGTP) in riched mu–opioid receptors–membrane preparations, isolated from thalamus69–71 periacuectal gray area (PAG)72 and pons/medulla.73 Moreover, autoradiographic studies using radiolabeled tracers of endomorphin molecules showed that these ligands matched the neuroanatomical distribution of DAMGO–binding sites (a conventional µ–opioid receptor selective ligand used as standard in binding studies) in the rodent's brain, postulating thus the role of endomorphins «as the preferential endogenous ligands of the µ–opioid receptor in the brain».32

However, based on the lower efficacy of endomorphins in stimulating µ–opioid receptors in several bioassays (e.g. ([35S]γGTP binding) compared to DAMGO and morphine, both peptides have been regarded as partial agonists of µ–opioid receptors.71,74 The extent to which these relative efficacies of endomorphins apply in the expression of functional bioactivities is still unclear and further analysis is needed to solve this issue.

Moreover, pharmacological studies showed that both EM1 and EM2 produce their biological effects by functionally stimulating different µ–opioid receptors subtypes, µ1 (MOR1)/µ2 (MOR2)–opioid receptors, which appear to be responsible for the distinct pharmacological activities mediated by opiate alkaloids, including both EM1–2 peptides.38,40,75 For instance, the µ1 receptor antagonist, naloxonazine, displayed a potent efficacy in blocking the antinociception response induced by EM2 compared to that exhibited by EM1. Converserly, the µ–opioid receptor antagonist, β–funaltrexamine, was capable of inhibiting the antinociceptive responses induced by both EM1–2 molecules.38,40,75

Pharmacological and molecular studies, showed that spinal administration of specific antisense oligodeoxy–nucleotides (used to knockdown the expression of the µ–opioid receptor genes) differentially attenuated the antinociception induced by either EM1 and/or EM2 peptides.76,77 These studies proved that endomorphin molecules activate and modulate µ2–receptor induced–responses (e.g. spinal analgesia, respiratory depression, and inhibition of gastrointestinal motility),78 whereas EM2 appears to modulate µ1–dependent responses, which include supraspinal analgesia, acetylcholine (ACh), and prolactin release from hypothalamus.78 Thus, these data provided significant insights on the differential regulation of µ–opioid receptor subtypes and their interaction with endomorphin peptides.

Quite interesting to note is that in vivo studies using radioligand binding techniques, combined with autoradio–graphic procedures, demonstrated that endomorphin peptides labeled some binding sites shown for DAMGO in specific reports of the rat brain,36,70,71 support the endogenous role of EMI–2 as p–opioid receptor peptide ligands (partial agonists of µ–opioid receptor) involved in the regulation of several µ1/µ2 opiate–receptor related bioactivities.32

Furthermore, a vast number of cellular and pharmacological studies showed that chronic exposure of µ–opioid receptor–agonists mediate long–lasting changes and cellular adaptations that underlie the development of opiate tolerance and physical dependence in animals and humans.7981 These observations led pharmacologists to limit a wide range of µ–opioid receptor–binding agents, clinically used to treat pain–related disorders.82,83 In vitro studies have shown that chronic administrations of EM1–2 peptides stimulated the development of tolerance in specific neuronal–tumor cell lines (e.g. SH–SY5Y neuroblastoma cells).84,85 For instance, chronic exposure of EM1/EM2 peptides on Chinese hamster ovary cells (functionally transfected with a µ–opioid receptor cDNA) induced a naloxone–dependent increase of forskolin–stimulated adenylyl cyclase activity above baseline responses, whereas in monkey kidney cells a specific EM1–2–dependent increase of type I and V–adenylyl cyclase isozymes was observed.86

In extension to the aforementioned studies, intracellular signaling pathways and trafficking molecules have been shown to mediate vesicle–endocytosis in brain cells expressing µ–opioid receptors, exposed to both opiates and opioid peptide ligand agonists.87,88 For instance, endomorphins have been shown to induce potent granule/vesicle endocytosis and trafficking of µ–opioid receptor in cells (transfected with the µ–opioid receptor cDNA), inducing similar sort of endocytosis and trafficking mechanisms for µ–opioid receptors in enteric neurons (shown to synthesize and express naturally opioid–receptor binding sites).85 This peptide–inducing regulated endocytosis and trafficking of µ–opioid receptor in cells has been shown to mediate desensitization, resensitization, and down–regulation of distinct molecular mechanisms that mediate the cell responsiveness to ligand stimulation, involved in the development of opioid tolerance and drug addiction.89

In vivo experiments showed that endomorphin peptides induced a naloxone–precipitated withdrawal response in rats, effects similar to those induced using same dose–unit of morphine.90 Using a preestablished scoring system used to detect specific withdrawal signs in rats (e.g. chewing, sniffing, grooming, wet–dog shakes, stretching, yawning, rearing, jumping, teeth grinding, ptosis, diarrhea, penile erection), Chen et al. demonstrated that endomorphin pre–treatment for 5 days (20 µg, ICV administration), followed by naloxone (4mg/kg, IP) administration, rats displayed physical signs of opioid dependence, exhibiting different potency for different withdrawal signs.90

The documented µ–opioid receptor ligand agonists (involved in the development of tolerance and physical dependence in animals and humans) have been suggested to operate through the interaction of conventional transmitter systems (e.g. dopamine[DA];91–93 norepinephrine [NE];94–97 cholinergic [ACh];98,99 benzodiazepine;100 GABA101 and N–methyl–D–aspartate [NMDA],102 imidazoline103 and glutamatergic neural pathways).104 However, no data on the interactions between endomorphin peptides and aforementioned transmission systems have been reported.32

Nitric oxide (NO) has been suggested to participate in the cellular mechanisms involved in peptide(s)–inducing physical dependence, based on observations that NO synthase inhibitors attenuated some signs of naloxone–precipitated withdrawal and cell–hyperactivity at the locus coeruleus (LC).105 In addition, the intrathecal (IT) infusion of morphine was shown to increase the N–methyl–D–aspartate (NMDA) binding activity, besides of inducing an up–regulation of NO synthase expression in neurons as well.106

 

IV. DEGRADATION OF ENDOMORPHINS: PEPTIDE CATABOLITES AND BIOACTIVITIES

Endomorphins have been shown to be vulnerable to enzymatic cleavage and several enzymes have been proposed as participants in endomorphin degradation.107 Studies performed to analyze the enzymatic degradation of endomorphins upon exposure to the activity of proteolytic enzymes (e.g. carboxipeptidase Y and A, proteinase A, and aminopeptidase M) showed a chromatographic profile of recovered endomorphin metabolites eluted from the HPLC column (figures 3A–B).108 These studies led to the proposal that enzymatic cleavage mediated by aminopeptidase M (EC 3.4.11.2) and/or aminopeptidase P (EC 3.4.11.9) degrade N–terminal Tyr–Pro peptide bonds, thus releasing the N–terminal amino acid (Tyr) and generating a tripeptide fragment from endomorphin molecules.109,110 Moreover, it has been shown that peptide sequences containing N–terminal hydrophobic residues followed by a Proline (Pro) residue; the preceding two amino acids (e.g., H2N–Tyr–Pro) in the peptide moiety may be released by aminopeptidase M, as an intact dipeptide (e.g. Tyr–Pro).111 In addition, enzymes —such as carboxypeptidase Y (a serine peptidase, EC 3.4.16.5) and/ or proteinase A (a non–pepsin–like acid endopeptidase [EC 3.4.23.6]), shown to display deamidase activities— have been postulated to catalyze the C–terminal amide group and catalyze the hydrolysis of the Xaa(3)–Phe(4) peptide bond (Xaa= natural amino acid), possibly by hydrolyzing EM1–2 peptide sequences into peptide acids, after releasing ammonia and cleaving off the C–terminal Phe(4) amino acid residue (figures 3A–B).108,112

Other enzymes, shown to display peptidase activities on distinct polypeptides, have been included as potential degrading enzymes of endomorphins. For instance, the membrane–bound serine proteinase, dipeptidyl–peptidase IV (DPPIV) (EC.3.4.14.5), has been shown to remove N–terminal dipeptide sequences from different polypeptides that contain the Pro residue at the penultimate position (H2N–Tyr–Pro…..Xn–COO).113 In vivo studies have shown that two groups of enzymes (e.g. proteinase A, aminopeptidase M/ P), working together in related catalytic pathways, may catalyze the degradation of endomorphins from formed dipeptides into single amino acids. In this context, the serine proteinase/DPP–IV system could trigger the catalytic activity of EM1/EM2 by cleaving the Pro(2)–Trp(3) and Pro(2)–Phe(3) peptide bonds, respectively, followed by the aminopeptidase M/P activities, recruited to catalyze secondary peptide cleavages. That is, hydrolyzing released dipeptides (Tyr–Pro/ Trp–Phe in EM1; Tyr–Pro/Phe–Phe in EM2) into single amino acids.114 Although the Tyr(1)–Pro(2) peptide bond might also be cleaved during the first step of enzyme processing pathway (dipeptide formation), some authors argued that EM1 appears to be more resistant to enzymatic degradation in vivo than EM2.115 Such proposals appear to be in good agreement with the observations showing that the spinal antinociceptive effects induced by EM1 are significantly longer than those induced by EM2.116 In a similar context, the analgesia induced by EM1 required a longer pretreatment time than that required for EM2, before tolerance is detected in mice.117 In addition, the antidepressant–like activity induced by EM2 results in a relatively shorter duration than that detected for EM1 in mice, after ICV administration of each peptide (figures 3A–B).32

One crucial aspect to elucidate the byproducts formed during the degradation of endomorphins is that the catabolites produced during peptidase activity could display related bioactivities similar to the parent compound. Thus, competition between endomorphin byproducts and full peptide sequences for the m–opioid receptor binding sites would lead to the expression of cellular and related opioid–bioactivities.32,118 In such context, radioligand binding studies demonstrated that the primary degradation products of EM1 (Tyr–Pro–Trp–Phe–OH and/or Pro–Trp–Phe–OH) displayed a very low m–opioid receptor binding affinity, showing no substantial activation/stimulation on Go/Gi proteins and non–antinociceptive activity after ICV administration of cleaved catabolites of EM1 in mice (figures 3A–B).118 Parallel studies used to evaluate the relationship between the binding activity of EM1/EM2 peptides and byproducts of peptide catabolism showed that peptidase inhibitors (e.g. diprotin A [Ile–Pro–Ile] a specific DPP–IV inhibitor) have no effect on the binding properties displayed by [3H]–EM2 on rat brain membrane preparations. However, when no inhibitor was added to the incubation mixture, 40% of the radioligand was degraded by endogenous membrane–bound proteinases.119 Moreover, ICV administration of a different DPP–IV inhibitor (e.g. Alapyrrolidonyl–2–nitrile) produced a significant increase in the antinociception activity induced by EM2, as shown for the altered analgesic related–parameters (e.g., magnitude, duration, and potency) detected in the rat tail–flick test.120 Similar EM2– induced analgesic effects were observed with diprotin A (a DPP–IV inhibitor), which produced a five–fold increase in potency with respect to EM2 itself in the paw withdrawal test114 or with actinonin (a selective serine–proteinase/peptidase inhibitor), which blocked the peptide catabolism of EM1, when incubated in a 6–day–old rat spinal cord homogenate, producing a potent antinociception effect compared to the peptide alone.121

 

V. ENDOMORPHINS REGULATING CONVENTIONAL NEUROTRANSMITTERS SYSTEMS AND RELEASE OF PEPTIDE HORMONES

Both exogenous and endogenous opioid receptor ligands display their biological activites by modifying the function and release of conventional transmitters in the CNS of mammals, including the evoked–release of various neurohormones from neuroendocrine tissues.32 Opioid ligands regulating neurotransmitter release have been extensively reviewed and reported elsewhere.32,122,123

In this context, µ–opioid receptor agonists have been shown to regulate the neuronal release of dopamine (DA);124–127 norepinerphrine (NA);125,128,129 serotonin (5–HT),122,123,125 and acetylcholine (ACh),130,131 in areas where µ–opioid receptors have been found to be co–localized with each neurotransmitter in specific areas of the mammalian brain.32

a) Dopamine (DA) transmission system

Immunohistochemical studies (IHC) revealed that µ–opioid receptors are densely expressed on dopamine neurons in the NAC and striatum in the CNS of rodents, including primates and humans,132–134 and they play an important regulatory role on the DA transmission system in these brain structures.32 In the striatum, opioid ligands have been shown to act on presynaptic opioid–receptors, producing indirect effects on DA turnover. This effect on DA activity varies according to the sort of opioid receptor bound by ligands.135–137 For instance, striatal administration of ligand agonists acting on postsynaptic µ–opioid receptors induced an increased DA efflux from neuronal cells138 without affecting the DA uptake or the levels of DA metabolites (DOPAC, HVA).139 Interesting enough is that the modulatory activity of DA release from striatal neurons by µ–opioid receptor agonists, requires the integrity of both cholinergic and GABAergic neurons,140 including the nigrostriatal neural pathway, where µ–opioid receptor agonists (e.g. morphiceptin) acting at a presynaptic level prevented the neuronal release of DA.139,141

In the accumbal region (NAC), microdialysis studies showed that µ–opioid receptor ligands agonists infused into the NAC induced an increased accumbal DA release or DA efflux in rats142 as shown for EM1–2 peptides, which produced a dose–dependent increase in DA efflux from NAC in freely moving rats.143 Whereas an accumbal infusion of EM1 showed a dose–dependent increase of DA efflux, which was abolished by intraccumbal perfusion of CTOP or by systemic administration of naloxone and/or naloxonazine,143 DA efflux induced by local infusion of EM2 was not abolished by selective µ–opioid receptor antagonists. This suggess that the EM2–induced effects may be mediated through different receptors other than the µ–opioid receptor.143 These studies support previous data showing that EM2 induces a dose–dependent stimulation of DA release and levels of DA metabolites (e.g. DOPAC, HVA) from the accumbal shell region126 as a result of the disinhibition of local GABA neurons.126

b) Noradrenergic (NA) transmission system

Noradrenergic system arises from a clustered group of NA–containing cell bodies, that arises from the locus coeruleous (LC) localized at the rostral area of the pons,144 and which projects axon fibers to forebrain areas and limbic structures. NA projecting fibers from LC represent the primary source of NA in cortical and hippocampal areas of the brain.145–147 The LC–NA neurons innervate almost all regions of the neuraxis, influencing several bioactivities and behaviors.148–150 LC–NA neurons also provide descending fibers to both medulla and spinal cord involved antinociceptive bioactivies.151–153 Neuroanatomical studies using autoradiographic and electrophysiological procedures using intracellular recording techniques revealed the expression of a high–density of µ–opioid receptor at the LC,154–156 where endogenous opioid peptides (e.g. morphiceptin, hemorphin–4, Tyr–MIF–1, Tyr–W–MIF–1), including endomorphins and non–natural synthetic peptides (e.g., Tyr–D–Arg–Phe–Sar; Tyr–D–Arg–Phe–Lys–NH2), shown to bind and activate µ–opioid receptors, were shown to modify the physiological activity of LC neurons.32,155

All peptides tested displayed opiate–alkaloid activities on LC–NA cells,155 producing a dose– and time– dependent electrophysiological responses (e.g. decreased neuronal excitability; inhibition of spontaneous cell–firing, opioid–induced hyperpolarization effect mediated through the opening of inward–rectification potassium channels, and other membrane bioactivies).155 Among the tested peptides, EM1 and EM2 represented the most effective peptides displaying equipotent electrophysiological responses (e.g. hyperpolarizations) on LC–NA neurons at very low concentrations, when compared to β–endorphin, DAMGO, or morphine.32,155,157 These results led authors to assume that the presence of the aromatic group on the Phe residue at position 4, in addition of the structural C–terminal amidation found in both EM1–2 peptides, represent crucial structural features by which these peptides exert their electrophysiological responses on LC–NA neurons after binding their cognate µ–opioid receptors on these cells.32,155

Ultrastructural studies showed that EM1–LI was mostly localized in axon terminals forming synaptic specializations between tyrosine hydroxylase (TH) and CRF–containing dendrites in peri–LC and Barrington's nucleus (localized at the brainstem dorsal pontine tegmentum) in the rat brain.157 Both areas were shown to express high–densities of µ–opioid receptors158,159 and were suggested to be targeted by EM1–2 peptides, producing a significant reduction of LC–NA neuronal excitability.157 Thus, EM1 could modulate LC neurons directly by tonically inhibiting neurons within the Barrington's nucleus.157 Futhermore, physiological studies showed the interaction between EM1 and NA at the level of the spinal cord, showing that the IT administration of EM1 in rats induced an evoked–release of NA from presynaptic axon terminals that arise from LC–NA descending neural pathways impinging into the dorsal horn of the rat's spinal cord.160 Thus, EM1 produced a potent analgesia activity (measured in different behavioral models of analgesia) by activating µ2–opioid receptors.32,160

c) Serotonin (5–HT) transmission system

The serotonin (5–HT) transmission system has been shown to mediate a vast number of physiological and behavioral responses in mammals (e.g. nociceptive, appetitive, emotional, stress and anxiety, motor, cognitive, and autonomic functions).161–165 Several bioactivities mediated by serotoninergic tone in the brain have been found to be influenced by the endogenous opioid system (EOS).32 The serotonin (5–HT) transmission system is functionally dependent on the serotoninergic tone of the CNS, which is regulated by the firing activity of cell bodies (somata) which are clustered in different nuclei along the midline of the brainstem; namely, the dorsal raphe nucleus (DRN) and the median raphe nucleus (MRN).162 Both nuclei have been shown to project axon fibers to almost all forebrain structures along the neuroaxis (e.g. mesolimbic areas, hippocampus, and cortex), where 5–HT axons projecting to terminal fields regulate the expression of different bioactivities and behavioral repertoires required for survival and adaptive functions in species.162 Thus, the 5–H/DRN located at the ventral aspect of the PAG has been shown to be crucial for the integration stress–responses,166 and pain and stressful stimuli activating peptidergic neurons at the PAG have been shown to modulate the firing activity of 5–HT neurons (at the somatodendritic level) via activation of µ–opioid receptors expressed at synaptic endings in short axons, innervating the 5–HT/DRN.167,168 Finally, 5–HT/DRN neurons project axon fibers to several limbic structures (e.g. amygdale, PFCx, hippocampus, NAC) involved in handling emotional and stress–related responses.167,169,170 However, the cellular mechanisms by which opiod agonists affect the activity 5–HT neurons have not been completely elucidated.

For instance, previous electrophysiological studies showed, on the one hand, that µ–opioid receptor agonists induced significant inhibitory effects on cell–firing of DRN/ 5–HT neurons.171 On the other hand, in vivo microdialysis showed that opioid agonists induced an increased activity on 5–HT neurons.167,172 Several mechanisms have been offered to explain such inconsistencies based on the structural organization of the DRN. This brainstem nucleus receives neural inputs from local GABAergic neurons and from descending glutamatergic axon fibers from the cortex, which control the cellular excitability of 5–HT neurons and the net equilibrium between both neural systems, will provide the result on the neuronal release or efflux of 5–HT into the external milieu.162 In this context, it has been shown that µ–receptor agonists move the net equilibrium of both neural systems into an increased release and extracellular levels of 5–HT at the DRN, possibly mediated through an opioid–enhanced activity of the glutamatergic (excitatory) input on 5–HT/DRN neurons and/or opioid–induced disinhibitory effect of GABAergic neurons.122

Microdialysis studies revealed that endomorphins peptides influenced the activity of DRN/5–HT neurons by increasing the net efflux of 5–HT (increased extracellular levels of the aminergic transmitter) within the DRN. This effect was blocked by selective µ–opioid receptor antagonists such as β–funaltrexamine.122 In a similar context, endomorphins have shown to modulate 5–HT activity in other brain areas. For instance, EM1–2 peptides infused into the VTA induced a significant decreased of 5–HT activity in the rat PFCx (cortex) and NAC (ventral striatum), thus suggesting that µ–opioid receptors could potentially mediate a neuronal depletion of 5–HT in these brain areas.125 However, no effect on the 5–HT metabolite, 5–hydroxyindoleacetic acid (5–HIAA), was found in the NAC.126 In forebrain areas, DAMGO and EM1 produced a strong, CTOP–reversible suppression of the 5–HT2A–induced excitatory post–synaptic currents on layer V–pyramidal cells, localized at the medial prefrontal cortex (PFCx).173 These studies showed that µ–opioid receptor agonists act presynaptically in cortical layer V pyramidal cells,174 and the interaction of 5–HT2A and µ–opioid receptor ligands appear to ocurr on projecting glutamatergic afferents sent from thalamic or amygdale nuclei (e.g. basolateral nucleus)175–177 that locally synapse layer V–piramidal cells.174

d) Acetylcholine (ACh) transmission system

With regard to the ACh–transmision system, several studies showed that EM1–2 peptides modulate cholinergic neurotransmission in peripheral tissues, such as the respiratory system and the gastrointestinal (GI) tract (see corresponding sections xiv and xv in the following chapter, Part II, in this journal ) acting on prejunctional µ–opioid receptors that inhibit the release of ACh that mediate the contractile muscle–activity and peristalsis in the GI tract.32 In a similar context, recent experiments showed that EM1 regulates ACh–transmission system in the rat inner hair cells.178 Using patch–clamp recording techniques, these authors showed that EM1 inhibited the nicotinic receptor–induced ACh–evoked currents in both frog saccular hair cells and rat inner hair cells, respectively.32,178

e) Release of neuropeptides hormones

Electrophysiological studies in vivo showed that EOS regulates the electrical activity of the neuron cells that synthesize and secrete oxytocin (OT) (from supraoptic nuclei [SON] localized within the hypothalamus) and arginine vasopressin (AVP) (from the anterior paraventricular nucleus of the hypothalamus [PVN]).179,180 Previous studies showed that µ–opioid receptor agonists produced a potent inhibition on OT–secreting cells (SON), whereas neuronal–activity of AVP–secreting cells (PVN) is decreased by the same receptor agonists.181 In a similar context, endormorphin peptides showed to display related activities on OT/SON and AVP/ PVN cells, respectively, showing that ICV administration of EM1 into rats induced a more potent inhibition effect of OT cells than on AVP cells.182 These effects detected on SON versus PVN–cells were attributed to the number of cell–surface opioid receptors expressed on each neurosecretory cell.182 Recent studies showed that the oxytocin (OT) induces important antinociceptive effects in rodents and humans,183–188 as shown after ICV,189 IP,183 or IT/IC184 administration of the peptide hormone, in addition of the injection of the peptide directly into the PAG and NRM (nucleus raphe magnus)190,191 (sections x–xiii in the following chapter, Part II, in this journal).

Recent studies shed important data concerning the interaction between µ–opioid receptors and OT–inducing antinociceptive responses in the rat CNS.192,193 These studies showed that µ–opioid receptors antagonists (e.g. β–FNA) inhibited OT–induced antinociceptive effects in response to the application of either thermal or mechanical stimuli in rats. So far, no experimental data have elucidated yet any role for the endogenous opioid system (e.g. β–endorphin), including both EM1–2 peptides, mediating the OT–induced antinociception responses in animals.32

Most of the experimental data reporting on the interaction between the AVP/PVN system and µ–opioid receptor agonists have been shown in stress–related behaviors32 (section xiii, in the following paper Part II, in this journal). However, no experimental works have clearly elucidated the mechanisms by which EOS and endormorphin peptide interact with the AVP/PVN peptide system. Nonetheless, recent documents showed that acute stress leads to an increase synaptic concentration of AVP in the rat amygdala or hypothalamus;194,195 where AVP released from PVN– neuronal cells was shown to activate its cognate membrane receptor (e.g. V1 receptor) and mediate crucial anxiogenic and depressive behavioral responses in rodents.195

With regard to somatostatin, µ–opioid receptors ligand agonists,196,197 including endomorphins,198 have been reported to regulate gastric–somatostatin and neuroendocrine activity at the GI tract (section xv, in the following chapter Part II, in this journal). For instance, EM1 was shown to exert an important inhibition on somatostatin–induced stimulation in the perfused rat stomach, which was partially blocked with CTOP.198 These results posit that EM1 acting on µ–opioid receptors mediate in large part the opioid–inhibitory effect on somatostatin activity.198 Mediated by the EM1 peptide, this inhibitory effect might be similar for other regulatory peptide–hormones in several systems.32 However, the interaction between peptide–hormone activity and EM1–2 peptides needs further investigation.

 

VI. ENDOMORPHINS MODULATING NOCICEPTIVE TRANSMISSION AND TOLERANCE

With regard to µ–opioid receptor–mediating pain modulation, in vivo studies showed that the intracerebro–ventricular (ICV) administration of endomorphins produced a potent antinociceptive effect in wild type mice,33,34,68 showing no significant effect in µ–opioid receptor mutant knockout mice.34,199–201 In a similar context, intrathecal (IT) administration produced significant antinociception effects in adult rodents exposed to the tail–flick, paw–withdrawal, tail pressure, and/or flexor–reflex tests, respectively (figures 4A–B).31,33,36,116,117,202,203

Based on the aforementioned description of the neuroanatomical distribution of endomorphins and µ–opioid receptors in the rodent, primate and human CNS (see section ii); several authors have categorized the functional role of EM1–2 peptides as neurotransmitters or neuromodulators, regulating different biological and physiological processes, as previously described.32 Such modulatory–related activities mediated by endomorphins include pain perception, stress–related responses, complex neural functions —such as reward, arousal, and vigilance—, as well as autonomic, cognitive, neuroendocrine, and limbic homeostasis.32

Endomorphin peptides have been implicated in the modulation of neural pathways of pain and/or nociceptive transmission in several areas of mammals brain,32 particularly, the spino(trigemino)–ponto–amygdaloid pathway, where nuclei and neuronal elements (e.g. caudal nucleus of spinal trigeminal tract, parabrachial nucleus, the NTS, PAG, nucleus ambiguous, LC, and midline thalamic nuclei) have been shown to express EM1–2–LI material (see reviews in204,205) (figures 2 and 4A–B). Based on the anatomical distribution of EM2–LI34,62 and its co–localization with µ–opioid receptors —densely expressed in spinal cord neurons and along primary afferent fibers penetrating the superficial layers of the dorsal horn—,46 EM2 was found to modulate nociceptive transmission pathway, modulating the release of excitatory transmitters (e.g. glutamate [Glu], substance P, γ–aminobutyric acid [GABA], glycine [Gly], including the calcitonin gene–related peptide [CGRP peptide]) and by stimulating presynaptic µ–opioid receptors expressed in axon terminals along the Lissahuer tract that penetrates the superficial dorsal horn and established synaptic contacts with intrinsic and relaying neurons of the spinal cord (figure 4A).206,207

Supporting evidences for the modulatory role of EM2 on primary nociceptive transmission pathways come from electrophysiological studies that showed the evoked–release of EM2 from dense–cored vesicles from spinal neurons after electrical stimulation from the dorsal root ganglion fibers;208,209 in addition to the EM–2–induced hyperpolarizing responses on cell–membranes of intrinsic spinal neurons, and the decreased excitability of postsynaptic µ–opioid receptors in intrinsic relay neurons as well.210

Furthermore, supraspinal (ICV) and spinal (IT) administration of EM1/EM2 showed that both amide tetrapeptides influence several neurotransmitter systems, in a similar fashion as demonstrated for morphine and DAMGO (preferential agonists of the µ–opioid receptor subtypes MOR1/MOR2) in the rat's CNS.33,34 In such a context, anatomical and pharmacological studies have extensively shown that neurotransmitters systems and chemical substrates that mediate mesencephalic analgesia induced by morphine (at the rostral ventromedial medulla) include the serotonin (5–HT),211,212 the inhibitory GABAergicsystem;213,214 the Glutamatergic and N–methyl–D–aspartate excitatory system,215,216 and the neurotensin peptide system as well (figures 4A–B).217

Pharmacological studies showed that both NA and 5–HT (acting on spinal cord, a2– and 5–HT receptors subtypes) mediate antinoceptive/analgesia responses induced by ICV administration of DAMGO,218,219 and conversely, the induced–depletion of both neurotransmitters (NA, 5–HT) after IT administration of 6–hydroxydopamine (6–OHDA) and/or 5,7–dihydroxytryptamine (5,7–DHT), respectively, or IT administration of yohimbine and methysergide (used to block spinal a2–adrenoreceptors and serotonin receptors) respectively; attenuated the antinociception response induced after ICV administration of morphine.220223 In a similar context, the antinoceptive responses induced after ICV administration of EM1/EM2 peptides were completely abolished after depleting (> 90%) the spinal NA system with IT administered 6–OHDA (3–day pretreatment period with neurotoxic compound).129 However, administration of 5,7–DHT (which depleted both 5–HT [390%] and NA [ 25%]) attenuated, but did not block the endomorphin–induced antinociception. These results clearly showed the relevance of the NA system over the 5–HT system in mediating the induced–antinociception by the supraspinal (ICV) administration of endomorphin molecules (figure 4B).32,129

Complementary results showed that the antinoception–induced effect detected after the spinal administration of EM1/EM2 peptides into mice was abolished with µ–recep–tor antagonists (e.g. naloxone, β–funaltrexamine), but not by the neurotoxic agents, 6–OHDA or 5,7–DHT, respectively,129,202 suggesting that the spinal induced–antinociception by EM1/EM2 peptides is caused by the direct stimulation of µ–opioid receptors, without affecting the NA or 5–HT transmission.32,129,202 Moreover, parallel studies showed that the supraspinal antinociception induced by ICV administration of EM2, but not EM1, were shown to be mediated by the activity of dynorphin A1–17 and Metenkephalin peptides, acting on their cognate κ– and δ2–opioid receptor subtypes, respectively.40 Thus, these data demonstrated functional differences in the spinal and supraspinal antinociceptive effects induced by EM1 and EM2 substances.40

Several studies provided strong evidence that the peripheral administration of endomorphins produce similar analgesic effect shown after IT or ICV administration of same opioid tetrapeptides. Albeit that the EM1/EM2 peptide–induced analgesia was demonstrated to be mediated through the activation of µ– and δ–opioid receptors expressed on primary neuronal afferents penetrating the dorsal horn (see above and figures 4A–B),224–226 the precise mechanisms and neural pathways mediating antinociception from peripherally–administered opioids are still unclear. In this context, several studies showed that EM1 produced a dose–dependent and naloxone–dependent reversible analgesia after i.p. administration in rats, exhibiting a delayed and less pronounced analgesic response when compared to the faster and increased antinociceptive effect induced from ICV or IT administered opioid peptides.227 The reduced analgesic effect observed after peripheral administration of opioid peptides could be explained by the rapid enzymatic degradation of EM1/ EM2 peptides in peripheral tissues (by unspecific membrane proteasas or esterases), besides the low permeation capability of these peptides to cross the brain–blood barrier (BBB), producing a small amount of active substances to elicit a centrally–mediated analgesia response.228,229 Whether peripheral administered–opioid peptides produced their antinociceptive/analgesia effect via this route or after permeating the BBB, is still a controversial issue, which requires further investigation (figures 4A–B).32

With regard to opioid–peptides inducing tolerance, in vivo studies showed that pre–treatment with different doses of EM1 or EM2 peptides in rodents (mice, rats)37,84,117 or the initial administration of single high doses of EM1 or EM2 peptides230–233 produced acute tolerance to the antinociceptive responses, induced by EM1/EM2 peptides after IT or ICV administration. Although tolerance development by EM1/ EM2 was much faster than that induced by morphine, tolerance induced by EM1 required a longer pretreatment time–period when compared to pretreatment with EM2, before the acute antinociceptive tolerance was observed. Such variations in the EM1/EM2 producing acute antinociceptive tolerance responses were proposed to be due to several pharmacological and cell–related mechanisms (e.g. different peptide half–lives, differences in µ–opioid receptor selectivity, divergent neuronal mechanisms), but not to the degree of how EM1–2 peptides stimulate membrane–bound µ–opioid receptors or induced µ–opioid receptor desensitization.232 Moreover, cross–tolerance studies (evaluating the induced–antinociceptive effects of EM1/EM2 peptides and morphine in both tail–flick and paw pressure tests in rats) showed that animals displayed a tolerance effect to EM2, exhibiting a partial antinociceptive cross–tolerance to EM1. However, rats displaying tolerance to EM1 showed no cross–tolerance to EM2.231

Similarly, these studies showed a cross–tolerance between morphine and EM1, but not for EM2, thus suggesting that morphine/EM1 mediate their analgesic/ tolerance effects via a common target opioid receptor (e.g. the µ2–opioid receptor subtype), whereas EM2 appears to mediate their opioid–related responses via a different membrane–bound opioid receptor (e.g. the µ1–opioid receptor subtype) or via a spliced variant of the µ2–opioid receptor and/or the same receptor displaying a different physical state of the functionally expressed µ–opioid receptor subtype.231

Moreover, cross–tolerance studies (evaluating the induced–antinociceptive effects of EM1–2 peptides and morphine in both tail–flick and paw pressure tests in rats) showed that animals made tolerant to EM2 exhibited a partial antinociceptive cross–tolerance to EM1, whereas rats tolerant to EM1 showed no cross–tolerance to EM2.231 Similarly, these studies showed a cross–tolerance between morphine and EM1, but not for EM2, thus suggesting that morphine/EM1 mediate their analgesic/tolerance effects via a common target opioid receptor (e.g. the µ2–opioid receptor subtype), whereas EM2 could mediate their opioid–related responses via a different membrane–bound opioid receptor (e.g., the µ1–opioid receptor subtype) or via a spliced variant of the µ2–opioid receptor and/or the same receptor displaying a different physical state of the functionally expressed µ–opioid receptor subtype.231 Albeit the several studies used to characterize the development of tolerance to the antinociceptive effects mediated by EM1–2 peptides in animal models of pain, tolerance responses to other endomorphin–inducing activities have not been elucidated and/or compared to morphine, its parent compound. For instance, tolerance does not develop for all of the alkaloid–mediating or inducing behavioral or functional responses (e.g. tolerance does not develop during morphine–inducing locomotor activity responses) (see corresponding sections vii–ix in the following chapter, part II, in this journal).234

 

ACKNOWLEDGEMENTS

This review is dedicated to the memory of profesor Ramon de la Fuente Muñiz, whose vision and knowledge in distinct areas on the neuroscience research led to the support and development of several projects in our lab. Moreover, we gratefully thank the following institutions for the funding of the present manuscript: National Institute of Psychiatry (Instituto Nacional de Psiquiatria) Mexico–Project INPRF–NC092318.1; INP–2040, ICYTDF–2007; CONACYT–SALUD–2003–C01–14; CONACyT/SEP–COI–47804; CONACyT–FOSSIS/SALUD–2007–COI–69373; Megaproyecto UNAM MP6–16; Fundación Gonzalo Rio Arronte A.C. CONADIC and BIRMEX are also acknowledged for supporting in the preparation of this manuscript.

 

REFERENCES

1. Birdsall NJ, Bradbury AF, Burgen AS et al. Interactions of peptides derived from the C–fragment of beta–lipotropin with brain opiate receptors [proceedings]. Br J Pharmacol 1976;58:460P–461P.        [ Links ]

2. Li CH, Chung D. Isolation and structure of an untriakontapeptide with opiate activity from camel pituitary glands. Proc Natl Acad Sci USA 1976;73:1145–1148.        [ Links ]

3. Hughes J, Smith TW, Kosterlitz HW et al. Identification of two related pentapeptides from the brain with potent opiate agonist activity. Nature 1975;258:577–580.        [ Links ]

4. Chavkin C, James IF, Goldstein A. Dynorphin is a specific endogenous ligand of the kappa opioid receptor. Science 1982;215:413–415.        [ Links ]

5. Goldstein A, Tachibana S, Lowney LI et al. Dynorphin–(1–13), an extraordinarily potent opioid peptide. Proc Natl Acad Sci USA 1979;76:6666–6670.        [ Links ]

6. Meunier JC, Mollereau C, Toll L et al. Isolation and structure of the endogenous agonist of opioid receptor–like ORL1 receptor. Nature 1995;377:532–535.        [ Links ]

7. Reinscheid RK, Nothacker HP, Bourson A et al. Orphanin FQ: a neuropeptide that activates an opioidlike G protein–coupled receptor. Science 1995;270:792–794.        [ Links ]

8. Comb M, Seeburg PH, Adelman J et al. Primary structure of the human Met– and Leu–enkephalin precursor and its mRNA. Nature 1982;295:663–666.        [ Links ]

9. Gubler U, Seeburg P, Hoffman BJ et al. Molecular cloning establishes proenkephalin as precursor of enkephalin–containing peptides. Nature 1982;295:206–208.        [ Links ]

10. Noda M, Furutani Y, Takahashi H et al. Cloning and sequence analysis of cDNA for bovine adrenal preproenkephalin. Nature 1982;295:202–206.        [ Links ]

11. Horikawa S, Takai T, Toyosato M et al. Isolation and structural organization of the human preproenkephalin B gene. Nature 1983;306:611–614.        [ Links ]

12. Nothacker HP, Reinscheid RK, Mansour A et al. Primary structure and tissue distribution of the orphanin FQ precursor. Proc Natl Acad Sci USA 1996;93:8677–8682.        [ Links ]

13. Mollereau C, Simons MJ, Soularue P et al. Structure, tissue distribution, and chromosomal localization of the prepronociceptin gene. Proc Natl Acad Sci USA 1996;93:8666–8670.        [ Links ]

14. Okuda–Ashitaka E, Tachibana S, Houtani T et al. Identification and characterization of an endogenous ligand for opioid receptor homologue ROR–C: its involvement in allodynic response to innocuous stimulus. Brain Res Mol Brain Res 1996;43:96–104.        [ Links ]

15. Fugere M, Day R. Inhibitors of the subtilase–like pro–protein convertases (SPCs). Curr Pharm Des 2002;8:549–562.        [ Links ]

16. Muller L, Lindberg I. The cell biology of the prohormone convertases PC1 and PC2. Prog Nucleic Acid Res Mol Biol 1999;63:69–108.        [ Links ]

17. Perone MJ, Castro MG. Prohormone and proneuropeptide synthesis and secretion. Histol Histopathol 1997;12:1179–1188.        [ Links ]

18. Branchaw JL, Hsu SF, Jackson MB. Membrane excitability and secretion from peptidergic nerve terminals. Cell Mol Neurobiol 1998;18:45–63.        [ Links ]

19. Burgoyne RD, Morgan A. Secretory granule exocytosis. Physiol Rev 2003;83:581–632.        [ Links ]

20. Glombik MM, Gerdes HH. Signal–mediated sorting of neuropeptides and prohormones: secretory granule biogenesis revisited. Biochimie 2000;82:315–326.        [ Links ]

21. Tse FW, Tse A. Regulation of exocytosis via release of Ca(2+) from intracellular stores. Bioessays 1999;21:861–865.        [ Links ]

22. Bunzow JR, Saez C, Mortrud M et al. Molecular cloning and tissue distribution of a putative member of the rat opioid receptor gene family that is not a mu, delta or kappa opioid receptor type. FEBS Lett 1994;347:284–288.        [ Links ]

23. Kosterlitz HW. Endogenous opioids and their receptors. Pol J Pharmacol Pharm 1987;39:571–576.        [ Links ]

24. Minami M, Satoh M. Molecular biology of the opioid receptors: structures, functions and distributions. Neurosci Res 1995;23:121–145.        [ Links ]

25. Mollereau C, Parmentier M, Mailleux P et al. ORL1, a novel member of the opioid receptor family. Cloning, functional expression and localization. FEBS Lett 1994;341:33–38.        [ Links ]

26. Reisine T, Law SF, Blake A et al. Molecular mechanisms of opiate receptor coupling to G proteins and effector systems. Ann NY Acad Sci 1996;780:168–175.        [ Links ]

27. Olson GA, Olson RD, Vaccarino AL et al. Endogenous opiates: 1997. Peptides 1998;19:1791–1843.        [ Links ]

28. Henschen A, Lottspeich F, Brantl V et al. Novel opioid peptides derived from casein (beta–casomorphins). II. Structure of active components from bovine casein peptone. Hoppe Seylers Z Physiol Chem 1979;360:1217–1224.        [ Links ]

29. Brantl V, Gramsch C, Lottspeich F et al. Novel opioid peptides derived from hemoglobin: hemorphins. Eur J Pharmacol 1986;125:309–310.        [ Links ]

30. Erchegyi J, Kastin AJ, Zadina JE. Isolation of a novel tetrapeptide with opiate and antiopiate activity from human brain cortex: Tyr–Pro–Trp–Gly–NH2 (Tyr–W–MIF–1). Peptides 1992;13:623–631.        [ Links ]

31. Horvath A, Kastin AJ. Isolation of tyrosine–melanocyte–stimulating hormone release–inhibiting factor 1 from bovine brain tissue. J Biol Chem 1989;264:2175–2179.        [ Links ]

32. Fichna J, Janecka A, Costentin J et al. The endomorphin system and its evolving neurophysiological role. Pharmacol Rev 2007;59:88–123.        [ Links ]

33. Zadina JE, Hackler L, Ge LJ et al. A potent and selective endogenous agonist for the mu–opiate receptor. Nature 1997;386:499–502.        [ Links ]

34. Zadina JE, Martin–Schild S, Gerall AA et al. Endomorphins: novel endogenous mu–opiate receptor agonists in regions of high mu–opiate receptor density. Ann NY Acad Sci 1999;897:136–144.        [ Links ]

35. Hackler L, Zadina JE, Ge LJ et al. Isolation of relatively large amounts of endomorphin–1 and endomorphin–2 from human brain cortex. Peptides 1997;18:1635–1639.        [ Links ]

36. Goldberg IE, Rossi GC, Letchworth SR et al. Pharmacological characterization of endomorphin–1 and endomorphin–2 in mouse brain. J Pharmacol Exp Ther 1998;286:1007–1013.        [ Links ]

37. Horvath G, Szikszay M, Tomboly C et al. Antinociceptive effects of intrathecal endomorphin–1 and –2 in rats. Life Sci 1999;65:2635–2641.        [ Links ]

38. Sakurada S, Zadina JE, Kastin AJ et al. Differential involvement of mu–opioid receptor subtypes in endomorphin–1– and –2–induced antinociception. Eur J Pharmacol 1999;372:25–30.        [ Links ]

39. Soignier RD, Vaccarino AL, Brennan AM et al. Analgesic effects of endomorphin–1 and endomorphin–2 in the formalin test in mice. Life Sci 2000;67:907–912.        [ Links ]

40. Tseng LF, Narita M, Suganuma C et al. Differential antinociceptive effects of endomorphin–1 and endomorphin–2 in the mouse. J Pharmacol Exp Ther 2000;292:576–583.        [ Links ]

41. Reinscheid RK, Nothacker H, Civelli O. The orphanin FQ/nociceptin gene: structure, tissue distribution of expression and functional implications obtained from knockout mice. Peptides 2000;21:901–906.        [ Links ]

42. Paterlini MG, Avitabile F, Ostrowski BG et al. Stereochemical requirements for receptor recognition of the mu–opioid peptide endomorphin–1. Biophys J 2000;78:590–599.        [ Links ]

43. Keller M, Boissard C, Patiny L et al. Pseudoproline–containing analogues of morphiceptin and endomorphin–2: evidence for a cis Tyr–Pro amide bond in the bioactive conformation. J Med Chem 2001;44:3896–3903.        [ Links ]

44. Anton B, Martinez I, Calva JC et al. Inmunohistochemistry for Endomorphins in the rat CNS. Neurosci Abstr 1998; 24[1], 851. Ref Type: Abstract        [ Links ]

45. Sanchez–Islas I, Ridaura C, Acevedo R et al. Comparative immunohistochemistry for endomorphins and mu opioid receptor in the Human Pons–Medulla and Spinal Cord. Neurosci Abstr 2000; 30[1], 634. Ref Type: Abstract        [ Links ]

46. Martin–Schild S, Zadina JE, Gerall AA et al. Localization of endomorphin–2–like immunoreactivity in the rat medulla and spinal cord. Peptides 1997;18:1641–1649.        [ Links ]

47. Martin–Schild S, Gerall AA, Kastin AJ et al. Differential distribution of endomorphin 1– and endomorphin 2–like immunoreactivities in the CNS of the rodent. J Comp Neurol 1999;405:450–471.        [ Links ]

48. Pierce TL, Grahek MD, Wessendorf MW. Immunoreactivity for endomorphin–2 occurs in primary afferents in rats and monkey. Neuroreport 1998;9:385–389.        [ Links ]

49. Pierce TL, Wessendorf MW. Immunocytochemical mapping of endomorphin–2–immunoreactivity in rat brain. J Chem Neuroanat 2000;18: 181–207.        [ Links ]

50. Schreff M, Schulz S, Wiborny D et al. Immunofluorescent identification of endomorphin–2–containing nerve fibers and terminals in the rat brain and spinal cord. Neuroreport 1998;9:1031–1034.        [ Links ]

51. Abbadie C, Rossi GC, Orciuolo A et al. Anatomical and functional correlation of the endomorphins with mu opioid receptor splice variants. Eur J Neurosci 2002;16:1075–1082.        [ Links ]

52. Hokfelt T, Ljungdahl A, Terenius L et al. Immunohistochemical analysis of peptide pathways possibly related to pain and analgesia: enkephalin and substance P. Proc Natl Acad Sci USA 1977;74:3081–3085.        [ Links ]

53. Uhl GR, Goodman RR, Kuhar MJ et al. Immunohistochemical mapping of enkephalin containing cell bodies, fibers and nerve terminals in the brain stem of the rat. Brain Res 1979;166:75–94.        [ Links ]

54. Sar M, Stumpf WE, Miller RJ et al. Immunohistochemical localization of enkephalin in rat brain and spinal cord. J Comp Neurol 1978;182:17–37.        [ Links ]

55. Watson SJ, Akil H, Sullivan S et al. Immunocytochemical localization of methionine enkephalin: preliminary observations. Life Sci 1977;21:733–738.        [ Links ]

56. Gramsch C, Hollt V, Pasi A et al. Immunoreactive dynorphin in human brain and pituitary. Brain Res 1982;233:65–74.        [ Links ]

57. Finley JC, Lindstrom P, Petrusz P. Immunocytochemical localization of beta–endorphin–containing neurons in the rat brain. Neuroendocrinology 1981;33:28–42.        [ Links ]

58. Zadina JE. Isolation and distribution of endomorphins in the central nervous system. Jpn J Pharmacol 2002;89:203–208.        [ Links ]

59. Jessop DS, Major GN, Coventry TL et al. Novel opioid peptides endomorphin–1 and endomorphin–2 are present in mammalian immune tissues. J Neuroimmunol 2000;106:53–59.        [ Links ]

60. Jessop DS, Richards LJ, Harbuz MS. Opioid peptides endomorphin–1 and endomorphin–2 in the immune system in humans and in a rodent model of inflammation. Ann NY Acad Sci 2002;966:456–463.        [ Links ]

61. Seale JV, Jessop DS, Harbuz MS. Immunohistochemical staining of endomorphin 1 and 2 in the immune cells of the spleen. Peptides 2004;25:91–94.        [ Links ]

62. Mousa SA, Machelska H, Schafer M et al. Immunohistochemical localization of endomorphin–1 and endomorphin–2 in immune cells and spinal cord in a model of inflammatory pain. J Neuroimmunol 2002;126:5–15.        [ Links ]

63. Azuma Y, Wang PL, Shinohara M et al. Immunomodulation of the neutrophil respiratory burst by endomorphins 1 and 2. Immunol Lett 2000;75:55–59.        [ Links ]

64. Azuma Y, Ohura K. Endomorphins 1 and 2 inhibit IL–10 and IL–12 production and innate immune functions, and potentiate NF–kappaB DNA binding in THP–1 differentiated to macrophage–like cells. Scand J Immunol 2002;56:260–269.        [ Links ]

65. Azuma Y, Ohura K. Endomorphin–2 modulates productions of TNF–alpha, IL–1beta, IL–10, and IL–12, and alters functions related to innate immune of macrophages. Inflammation 2002;26:223–232.        [ Links ]

66. Azuma Y, Ohura K, Wang PL et al. Endomorphins delay constitutive apoptosis and alter the innate host defense functions of neutrophils. Immunol Lett 2002;81:31–40.        [ Links ]

67. Anton B, Leff P, Calva JC et al. Endomorphin 1 and endomorphin 2 suppress in vitro antibody formation at ultra–low concentrations: antipeptide antibodies but not opioid antagonists block the activity. Brain Behav Immun 2008;22:824–832.        [ Links ]

68. Horvath G. Endomorphin–1 and endomorphin–2: pharmacology of the selective endogenous mu–opioid receptor agonists. Pharmacol Ther 2000;88:437–463.        [ Links ]

69. Fichna J, do–Rego JC, Kosson P et al. [(35)S]GTPgammaS binding stimulated by endomorphin–2 and morphiceptin analogs. Biochem Biophys Res Commun 2006;345:162–168.        [ Links ]

70. Kakizawa K, Shimohira I, Sakurada S et al. Parallel stimulations of in vitro and in situ [35S]GTPgammaS binding by endomorphin 1 and DAMGO in mouse brains. Peptides 1998;19:755–758.        [ Links ]

71. Sim LJ, Liu Q, Childers SR et al. Endomorphin–stimulated [35S]GTPgammaS binding in rat brain: evidence for partial agonist activity at mu–opioid receptors. J Neurochem 1998;70:1567–1576.        [ Links ]

72. Narita M, Mizoguchi H, Narita M et al. G protein activation by endomorphins in the mouse periaqueductal gray matter. J Biomed Sci 2000;7: 221–225.        [ Links ]

73. Mizoguchi H, Narita M, Wu H et al. Differential involvement of mu(1)–opioid receptors in endomorphin– and beta–endorphin–induced G–protein activation in the mouse pons/medulla. Neuroscience 2000;100:835–839.        [ Links ]

74. Alt A, Mansour A, Akil H et al. Stimulation of guanosine–5'–O–(3–[35S]thio)triphosphate binding by endogenous opioids acting at a cloned mu receptor. J Pharmacol Exp Ther 1998;286:282–288.        [ Links ]

75. Sakurada S, Hayashi T, Yuhki M et al. Differential antagonism of endomorphin–1 and endomorphin–2 spinal antinociception by naloxonazine and 3–methoxynaltrexone. Brain Res 2000;881:1–8.        [ Links ]

76. Garzon J, Rodriguez–Munoz M, Lopez–Fando A et al. RGSZ1 and GAIP regulate mu– but not delta–opioid receptors in mouse CNS: role in tachyphylaxis and acute tolerance. Neuropsychopharmacology 2004;29: 1091–1104.        [ Links ]

77. Wu HE, Mizoguchi H, Terashvili M et al. Spinal pretreatment with antisense oligodeoxynucleotides against exon–1, –4, or –8 of mu–opioid receptor clone leads to differential loss of spinal endomorphin–1–and endomorphin–2–induced antinociception in the mouse. J Pharmacol Exp Ther 2002;303:867–873.        [ Links ]

78. Pasternak GW. Pharmacological mechanisms of opioid analgesics. Clin Neuropharmacol 1993;16:1–18.        [ Links ]

79. Defer N, Best–Belpomme M, Hanoune J. Tissue specificity and physiological relevance of various isoforms of adenylyl cyclase. Am J Physiol Renal Physiol 2000;279:F400–F416.        [ Links ]

80. Heyne A, May T, Goll P et al. Persisting consequences of drug intake: towards a memory of addiction. J Neural Transm 2000;107:613–638.        [ Links ]

81. Law PY, Wong YH, Loh HH. Molecular mechanisms and regulation of opioid receptor signaling. Annu Rev Pharmacol Toxicol 2000;40:389–430.        [ Links ]

82. Schiller PW, Weltrowska G, Berezowska I et al. The TIPP opioid peptide family: development of delta antagonists, delta agonists, and mixed mu agonist/delta antagonists. Biopolymers 1999;51:411–425.        [ Links ]

83. Shen J, Benedict GA, Gallagher A et al. Role of cAMP–dependent protein kinase (PKA) in opioid agonist–induced mu–opioid receptor down–regulation and tolerance in mice. Synapse 2000;38:322–327.        [ Links ]

84. Higashida H, Hoshi N, Knijnik R et al. Endomorphins inhibit high–threshold Ca2+ channel currents in rodent NG108–15 cells overexpressing mu–opioid receptors. J Physiol 1998;507(Pt 1):71–75.        [ Links ]

85. McConalogue K, Grady EF, Minnis J et al. Activation and internalization of the mu–opioid receptor by the newly discovered endogenous agonists, endomorphin–1 and endomorphin–2. Neuroscience 1999;90: 1051–1059.        [ Links ]

86. Nevo I, Vidor–Reiss T, Levy R et al. Acute and chronic activation of the mu–opioid receptor with the endogenous ligand endomorphin differentially regulates adenylyl cyclase isozymes. Neuropharmacology 2000;39: 364–371.        [ Links ]

87. Minnis JG, Patierno S, Kohlmeier SE et al. Ligand–induced mu opioid receptor endocytosis and recycling in enteric neurons. Neuroscience 2003;119:33–42.        [ Links ]

88. Sternini C, Brecha NC, Minnis J et al. Role of agonist–dependent receptor internalization in the regulation of mu opioid receptors. Neuroscience 2000;98:233–241.        [ Links ]

89. Bohm SK, Grady EF, Bunnett NW. Regulatory mechanisms that modulate signalling by G–protein–coupled receptors. Biochem J 1997;322(Pt1): 1–18.        [ Links ]

90. Chen JC, Tao PL, Li JY et al. Endomorphin–1 and –2 induce naloxone–precipitated withdrawal syndromes in rats. Peptides 2003;24:477–481.        [ Links ]

91. El–Kadi AO, Sharif SI. The role of dopamine in the expression of morphine withdrawal. Gen Pharmacol 1998;30:499–505.        [ Links ]

92. Samini M, Fakhrian R, Mohagheghi M et al. Comparison of the effect of levodopa and bromocriptine on naloxone–precipitated morphine withdrawal symptoms in mice. Hum Psychopharmacol 2000;15:95–101.        [ Links ]

93. Tokuyama S, Ho IK, Yamamoto T. A protein kinase inhibitor, H–7, blocks naloxone–precipitated changes in dopamine and its metabolites in the brains of opioid–dependent rats. Brain Res Bull 2000;52:363–369.        [ Links ]

94. Fuentealba JA, Forray MI, Gysling K. Chronic morphine treatment and withdrawal increase extracellular levels of norepinephrine in the rat bed nucleus of the stria terminalis. J Neurochem 2000;75:741–748.        [ Links ]

95. Fuertes G, Milanes MV, Rodriguez–Gago M et al. Changes in hypothalamic paraventricular nucleus catecholaminergic activity after acute and chronic morphine administration. Eur J Pharmacol 2000;388:49–56.        [ Links ]

96. Laorden ML, Fuertes G, Gonzalez–Cuello A et al. Changes in catecholaminergic pathways innervating paraventricular nucleus and pituitary–adrenal axis response during morphine dependence: implication of al–pha(1)– and alpha(2)–adrenoceptors. J Pharmacol Exp Ther 2000;293:578–584.        [ Links ]

97. Milanes MV, Laorden ML, Chapleur–Chateau M et al. Alterations in corticotropin–releasing factor and vasopressin content in rat brain during morphine withdrawal: correlation with hypothalamic noradrenergic activity and pituitary–adrenal response. J Pharmacol Exp Ther 1998; 285:700–706.        [ Links ]

98. Buccafusco JJ, Zhang LC, Shuster LC et al. Prevention of precipitated withdrawal symptoms by activating central cholinergic systems during a dependence–producing schedule of morphine in rats. Brain Res 2000; 852:76–83.        [ Links ]

99. Zhang LC, Buccafusco JJ. Prevention of morphine–induced muscarinic (M2) receptor adaptation suppresses the expression of withdrawal symptoms. Brain Res 1998;803:114–121.        [ Links ]

100. Tejwani GA, Sheu MJ, Sribanditmongkol P et al. Inhibition of morphine tolerance and dependence by diazepam and its relation to mu–opioid receptors in the rat brain and spinal cord. Brain Res 1998;797:305–312.        [ Links ]

101. Sayin U, Atasoy S, Uzbay IT et al. Gamma–vinyl–GABA potentiates the severity of naloxone–precipitated abstinence signs in morphine–dependent rats. Pharmacol Res 1998;38:45–51.        [ Links ]

102. Popik P, Mamczarz J, Fraczek M et al. Inhibition of reinforcing effects of morphine and naloxone–precipitated opioid withdrawal by novel glycine site and uncompetitive NMDA receptor antagonists. Neuropharmacology 1998;37:1033–1042.        [ Links ]

103. Su RB, Li J, Gao K et al. Influence of idazoxan on analgesia, tolerance, and physical dependence of morphine in mice and rats in vivo. Acta Pharmacol Sin 2000;21:1011–1015.        [ Links ]

104. Rockhold RW, Liu N, Coleman D et al. The nucleus paragigantocellularis and opioid withdrawal–like behavior. J Biomed Sci 2000;7:270–276.        [ Links ]

105. Pineda J, Torrecilla M, Martin–Ruiz R et al. Attenuation of withdrawal–induced hyperactivity of locus coeruleus neurones by inhibitors of nitric oxide synthase in morphine–dependent rats. Neuropharmacology 1998;37:759–767.        [ Links ]

106. Wong CS, Hsu MM, Chou YY et al. Morphine tolerance increases (3H)MK–801 binding affinity and constitutive neuronal nitric oxide synthase expression in rat spinal cord. Br J Anaesth 2000;85:587–591.        [ Links ]

107. Tomboly C, Peter A, Toth G. In vitro quantitative study of the degradation of endomorphins. Peptides 2002;23:1573–1580.        [ Links ]

108. Peter A, Toth G, Tomboly C et al. Liquid chromatographic study of the enzymatic degradation of endomorphins, with identification by electrospray ionization mass spectrometry. J Chromatogr A 1999;846:39–48.        [ Links ]

109. Dua AK, Pinsky C, LaBella FS. Peptidases that terminate the action of enkephalins. Consideration of physiological importance for amino–, carboxy–, endo–, and pseudoenkephalinase. Life Sci 1985;37:985–992.        [ Links ]

110. Harbeck HT, Mentlein R. Aminopeptidase P from rat brain. Purification and action on bioactive peptides. Eur J Biochem 1991;198:451–458.        [ Links ]

111. Bairoch A. The ENZYME data bank in 1995. Nucleic Acids Res 1996;24:221–222.        [ Links ]

112. Berne PF, Schmitter JM, Blanquet S. Peptide and protein carboxyl–terminal labeling through carboxypeptidase Y–catalyzed transpeptidation. J Biol Chem 1990;265:19551–19559.        [ Links ]

113. Kato T, Nagatsu T, Kimura T et al. Studies on substrate specificity of X–prolyl dipeptidyl–aminopeptidase during new chromogenic substrates, X–Y–p–nitroanilides. Experientia 1978;34:319–320.        [ Links ]

114. Sakurada C, Sakurada S, Hayashi T et al. Degradation of endomorphin–2 at the supraspinal level in mice is initiated by dipeptidyl peptidase IV: an in vitro and in vivo study. Biochem Pharmacol 2003;66:653–661.        [ Links ]

115. Fujita T, Kumamoto E. Inhibition by endomorphin–1 and endomorphin–2 of excitatory transmission in adult rat substantia gelatinosa neurons. Neuroscience 2006;139:1095–1105.        [ Links ]

116. Grass S, Xu IS, Wiesenfeld–Hallin Z et al. Comparison of the effect of intrathecal endomorphin–1 and endomorphin–2 on spinal cord excitability in rats. Neurosci Lett 2002;324:197–200.        [ Links ]

117. Stone LS, Fairbanks CA, Laughlin TM et al. Spinal analgesic actions of the new endogenous opioid peptides endomorphin–1 and –2. Neuroreport 1997;8:3131–3135.        [ Links ]

118. Szatmari I, Biyashev D, Tomboly C et al. Influence of degradation on binding properties and biological activity of endomorphin 1. Biochem Biophys Res Commun 2001;284:771–776.        [ Links ]

119. Spetea M, Monory K, Tomboly C et al. In vitro binding and signaling profile of the novel mu opioid receptor agonist endomorphin 2 in rat brain membranes. Biochem Biophys Res Commun 1998;250:720–725.        [ Links ]

120. Shane R, Wilk S, Bodnar RJ. Modulation of endomorphin–2–induced analgesia by dipeptidyl peptidase IV. Brain Res 1999;815:278–286.        [ Links ]

121. Sugimoto–Watanabe A, Kubota K, Fujibayashi K et al. Antinociceptive effect and enzymatic degradation of endomorphin–1 in newborn rat spinal cord. Jpn J Pharmacol 1999;81:264–270.        [ Links ]

122. Tao R, Auerbach SB. Opioid receptor subtypes differentially modulate serotonin efflux in the rat central nervous system. J Pharmacol Exp Ther 2002;303:549–556.        [ Links ]

123. Sbrenna S, Marti M, Morari M et al. Modulation of 5–hydroxytryptamine efflux from rat cortical synaptosomes by opioids and nociceptin. Br J Pharmacol 2000;130:425–433.        [ Links ]

124. Bujdoso E, Jaszberenyi M, Gardi J et al. The involvement of dopamine and nitric oxide in the endocrine and behavioural action of endomorphin–1. Neuroscience 2003;120:261–268.        [ Links ]

125. Chen JC, Liang KW, Huang EY. Differential effects of endomorphin–1 and –2 on amphetamine sensitization: neurochemical and behavioral aspects. Synapse 2001;39:239–248.        [ Links ]

126. Huang EY, Chen CM, Tao PL. Supraspinal anti–allodynic and rewarding effects of endomorphins in rats. Peptides 2004;25:577–583.        [ Links ]

127. Ukai M, Lin HP. Endomorphins 1 and 2 induce amnesia via selective modulation of dopamine receptors in mice. Eur J Pharmacol 2002;446:97–101.        [ Links ]

128. Al–Khrasani M, Elor G, Yusuf AM et al. The effect of endomorphins on the release of 3H–norepinephrine from rat nucleus tractus solitarii slices. Regul Pept 2003;111:97–101.        [ Links ]

129. Hung KC, Wu HE, Mizoguchi H et al. Intrathecal treatment with 6–hydroxydopamine or 5,7–dihydroxytryptamine blocks the antinociception induced by endomorphin–1 and endomorphin–2 given intracere–broventricularly in the mouse. J Pharmacol Sci 2003;93:299–306.        [ Links ]

130. Domino EF. Opiate interactions with cholinergic neurons. Adv Biochem Psychopharmacol 1979;20:339–355.        [ Links ]

131. Jhamandas K, Sutak M. Action of enkephalin analogues and morphine on brain acetylcholine release: differential reversal by naloxone and an opiate pentapeptide. Br J Pharmacol 1980;71:201–210.        [ Links ]

132. Eghbali M, Santoro C, Paredes W et al. Visualization of multiple opioid–receptor types in rat striatum after specific mesencephalic lesions. Proc Natl Acad Sci USA 1987;84:6582–6586.        [ Links ]

133. Fallon JH, Leslie FM. Distribution of dynorphin and enkephalin peptides in the rat brain. J Comp Neurol 1986;249:293–336.        [ Links ]

134. Mansour A, Khachaturian H, Lewis ME et al. Anatomy of CNS opioid receptors. Trends Neurosci 1988;11:308–314.        [ Links ]

135. Clouet DH, Ratner M. Catecholamine biosynthesis in brains of rats treated with morphine. Science 1970;168:854–856.        [ Links ]

136. Gauchy C, Agid Y, Glowinski J et al. Acute effects of morphine on dopamine synthesis and release and tyrosine metabolism in the rat striatum. Eur J Pharmacol 1973;22:311–319.        [ Links ]

137. Smith CB, Sheldon MI, Bednarczyk JH et al. Morphine–induced increases in the incorporation of 14 C–Tyrosine into 14 C–Dopamine and 14 C–norepinephrine in the mouse brain: antagonism by naloxone and tolerance. J Pharmacol Exp Ther 1972;180:547–557.        [ Links ]

138. Dourmap N, Michael–Titus A, Costentin J. Differential effect of intras–triatal kainic acid on the modulation of dopamine release by mu– and delta–opioid peptides: a microdialysis study. J Neurochem 1992;58:709–713.        [ Links ]

139. Das D, Rogers J, Michael–Titus AT. Comparative study of the effects of mu, delta and kappa opioid agonists on 3H–dopamine uptake in rat striatum and nucleus accumbens. Neuropharmacology 1994;33:221–226.        [ Links ]

140. Dourmap N, Clero E, Costentin J. Involvement of cholinergic neurons in the release of dopamine elicited by stimulation of mu–opioid receptors in striatum. Brain Res 1997;749:295–300.        [ Links ]

141. Yonehara N, Clouet DH. Effects of delta and mu opiopeptides on the turnover and release of dopamine in rat striatum. J Pharmacol Exp Ther 1984;231:38–42.        [ Links ]

142. Spanagel R, Herz A, Shippenberg TS. Opposing tonically active endogenous opioid systems modulate the mesolimbic dopaminergic pathway. Proc Natl Acad Sci USA 1992;89:2046–2050.        [ Links ]

143. Okutsu H, Watanabe S, Takahashi I et al. Endomorphin–2 and endomorphin–1 promote the extracellular amount of accumbal dopamine via nonopioid and mu–opioid receptors, respectively. Neuropsychopharmacology 2006;31:375–383.        [ Links ]

144. Nestler EJ, Alreja M, Aghajanian GK. Molecular and cellular mechanisms of opiate action: studies in the rat locus coeruleus. Brain Res Bull 1994;35:521–528.        [ Links ]

145. Lewis DA, Campbell MJ, Foote SL et al. The distribution of tyrosine hydroxylase–immunoreactive fibers in primate neocortex is widespread but regionally specific. J Neurosci 1987;7:279–290.        [ Links ]

146. McCormick DA, Pape HC, Williamson A. Actions of norepinephrine in the cerebral cortex and thalamus: implications for function of the central noradrenergic system. Prog Brain Res 1991;88:293–305.        [ Links ]

147. Ston–Jones G, Foote SL, Segal M. Impulse conduction properties of noradrenergic locus coeruleus axons projecting to monkey cerebrocortex. Neuroscience 1985;15:765–777.        [ Links ]

148. Simpson KL, Altman DW, Wang L et al. Lateralization and functional organization of the locus coeruleus projection to the trigeminal somato–sensory pathway in rat. J Comp Neurol 1997;385:135–147.        [ Links ]

149. Waterhouse BD, Lin CS, Burne RA et al. The distribution of neocortical projection neurons in the locus coeruleus. J Comp Neurol 1983;217:418–431.        [ Links ]

150. Waterhouse BD, Border B, Wahl L et al. Topographic organization of rat locus coeruleus and dorsal raphe nuclei: distribution of cells projecting to visual system structures. J Comp Neurol 1993;336:345–361.        [ Links ]

151. Jones SL. Descending noradrenergic influences on pain. Prog Brain Res 1991;88:381–394.        [ Links ]

152. Lipp J. Possible mechanisms of morphine analgesia. Clin Neuropharmacol 1991;14:131–147.        [ Links ]

153. Proudfit HK. Pharmacologic evidence for the modulation of nociception by noradrenergic neurons. Prog Brain Res 1988;77:357–370.        [ Links ]

154. Mansour A, Lewis ME, Khachaturian H et al. Pharmacological and anatomical evidence of selective mu, delta, and kappa opioid receptor binding in rat brain. Brain Res 1986;399:69–79.        [ Links ]

155. Yang YR, Chiu TH, Chen CL. Structure–activity relationships of naturally occurring and synthetic opioid tetrapeptides acting on locus coeru–leus neurons. Eur J Pharmacol 1999;372:229–236.        [ Links ]

156. Mansour A, Fox CA, Akil H et al. Opioid–receptor mRNA expression in the rat CNS: anatomical and functional implications. Trends Neurosci 1995;18:22–29.        [ Links ]

157. Peoples JF, Wessendorf MW, Pierce T et al. Ultrastructure of endomorphin–1 immunoreactivity in the rat dorsal pontine tegmentum: evidence for preferential targeting of peptidergic neurons in Barrington's nucleus rather than catecholaminergic neurons in the perilocus coeruleus. J Comp Neurol 2002;448:268–279.        [ Links ]

158. Van Bockstaele EJ, Colago EE, Moriwaki A et al. Mu–opioid receptor is located on the plasma membrane of dendrites that receive asymmetric synapses from axon terminals containing leucine–enkephalin in the rat nucleus locus coeruleus. J Comp Neurol 1996;376:65–74.        [ Links ]

159. Van Bockstaele EJ, Colago EE, Cheng P et al. Ultrastructural evidence for prominent distribution of the mu–opioid receptor at extrasynaptic sites on noradrenergic dendrites in the rat nucleus locus coeruleus. J Neurosci 1996;16:5037–5048.        [ Links ]

160. Hao S, Takahata O, Iwasaki H. Intrathecal endomorphin–1 produces antinociceptive activities modulated by alpha 2–adrenoceptors in the rat tail flick, tail pressure and formalin tests. Life Sci 2000;66:L195–L204.        [ Links ]

161. Driessen B, Reimann W. Interaction of the central analgesic, tramadol, with the uptake and release of 5–hydroxytryptamine in the rat brain in vitro. Br J Pharmacol 1992;105:147–151.        [ Links ]

162. Frazer A, Henler J. Serotonin. In: Siegel G, Agranoff B, Albers R et al. (eds.). Basic Neurochemistry: Molecular, Cellular and Medical aspects. New York, NY: Lippincott– Raven Press; 1999.        [ Links ]

163. Kawahara H, Yoshida M, Yokoo H et al. Psychological stress increases serotonin release in the rat amygdala and prefrontal cortex assessed by in vivo microdialysis. Neurosci Lett 1993;162:81–84.        [ Links ]

164. Petty F, Kramer G, Wilson L. Prevention of learned helplessness: in vivo correlation with cortical serotonin. Pharmacol Biochem Behav 1992;43:361–367.        [ Links ]

165. Pini LA, Vitale G, Ottani A et al. Naloxone–reversible antinociception by paracetamol in the rat. J Pharmacol Exp Ther 1997;280:934–940.        [ Links ]

166. Roche M, Commons KG, Peoples A et al. Circuitry underlying regulation of the serotonergic system by swim stress. J Neurosci 2003;23:970–977.        [ Links ]

167. Tao R, Auerbach SB. Involvement of the dorsal raphe but not median raphe nucleus in morphine–induced increases in serotonin release in the rat forebrain. Neuroscience 1995;68:553–561.        [ Links ]

168. Wang QP, Nakai Y. The dorsal raphe: an important nucleus in pain modulation. Brain Res Bull 1994;34:575–585.        [ Links ]

169. Grahn RE, Maswood S, McQueen MB et al. Opioid–dependent effects of inescapable shock on escape behavior and conditioned fear responding are mediated by the dorsal raphe nucleus. Behav Brain Res 1999;99:153–167.        [ Links ]

170. Ma QP, Han JS. Neurochemical and morphological evidence of an antinociceptive neural pathway from nucleus raphe dorsalis to nucleus accumbens in the rabbit. Brain Res Bull 1992;28:931–936.        [ Links ]

171. Haigler HJ. Morphine: effects on serotonergic neurons and neurons in areas with a serotonergic input. Eur J Pharmacol 1978;51:361–376.        [ Links ]

172. Kalyuzhny AE, Arvidsson U, Wu W et al. mu–Opioid and delta–opioid receptors are expressed in brainstem antinociceptive circuits: studies using immunocytochemistry and retrograde tract–tracing. J Neurosci 1996;16:6490–6503.        [ Links ]

173. Marek GJ, Aghajanian GK. 5–Hydroxytryptamine–induced excitatory postsynaptic currents in neocortical layer V pyramidal cells: suppression by mu–opiate receptor activation. Neuroscience 1998;86:485–497.        [ Links ]

174. Vogt BA, Wiley RG, Jensen EL. Localization of Mu and delta opioid receptors to anterior cingulate afferents and projection neurons and input/output model of Mu regulation. Exp Neurol 1995;135:83–92.        [ Links ]

175. Bacon SJ, Headlam AJ, Gabbott PL et al. Amygdala input to medial prefrontal cortex (mPFC) in the rat: a light and electron microscope study. Brain Res 1996;720:211–219.        [ Links ]

176. Delfs JM, Kong H, Mestek A et al. Expression of mu opioid receptor mRNA in rat brain: an in situ hybridization study at the single cell level. J Comp Neurol 1994;345:46–68.        [ Links ]

177. Mansour A, Fox CA, Burke S et al. Mu, delta, and kappa opioid receptor mRNA expression in the rat CNS: an in situ hybridization study. J Comp Neurol 1994;350:412–438.        [ Links ]

178. Lioudyno MI, Verbitsky M, Glowatzki E et al. The alpha9/alpha10–containing nicotinic ACh receptor is directly modulated by opioid peptides, endomorphin–1, and dynorphin B, proposed efferent cotransmitters in the inner ear. Mol Cell Neurosci 2002;20:695–711.        [ Links ]

179. Brown CH, Leng G. In vivo modulation of post–spike excitability in vasopressin cells by kappa–opioid receptor activation. J Neuroendocrinol 2000;12:711–714.        [ Links ]

180. Hatton GI. Emerging concepts of structure–function dynamics in adult brain: the hypothalamo–neurohypophysial system. Prog Neurobiol 1990;34:437–504.        [ Links ]

181. Pumford KM, Russell JA, Leng G. Effects of the selective kappa–opioid agonist U50,488 upon the electrical activity of supraoptic neurones in morphine–tolerant and morphine–naive rats. Exp Brain Res 1993;94:237–246.        [ Links ]

182. Doi N, Brown CH, Cohen HD et al. Effects of the endogenous opioid peptide, endomorphin 1, on supraoptic nucleus oxytocin and vasopres–sin neurones in vivo and in vitro. Br J Pharmacol 2001;132:1136–1144.        [ Links ]

183. Agren G, Uvnas–Moberg K, Lundeberg T. Olfactory cues from an oxytocin–injected male rat can induce anti–nociception in its cagemates. Neuroreport 1997;8:3073–3076.        [ Links ]

184. Arletti R, Benelli A, Bertolini A. Influence of oxytocin on nociception and morphine antinociception. Neuropeptides 1993;24:125–129.        [ Links ]

185. Caron RW, Leng G, Ludwig M et al. Naloxone–induced supersensitivity of oxytocin neurones to opioid antagonists. Neuropharmacology 1998;37:887–897.        [ Links ]

186. Kang YS, Park JH. Brain uptake and the analgesic effect of oxytocin—its usefulness as an analgesic agent. Arch Pharm Res 2000;23:391–395.        [ Links ]

187. Louvel D, Delvaux M, Felez A et al. Oxytocin increases thresholds of colonic visceral perception in patients with irritable bowel syndrome. Gut 1996;39:741–747.        [ Links ]

188. Petersson M, Alster P, Lundeberg T et al. Oxytocin increases nociceptive thresholds in a long–term perspective in female and male rats. Neurosci Lett 1996;212:87–90.        [ Links ]

189. Liu ZY, Zhuang DB, Lunderberg T et al. Involvement of 5–hydroxytryptamine(1A) receptors in the descending anti–nociceptive pathway from periaqueductal gray to the spinal dorsal horn in intact rats, rats with nerve injury and rats with inflammation. Neuroscience 2002;112:399–407.        [ Links ]

190. Robinson DA, Wei F, Wang GD et al. Oxytocin mediates stress–induced analgesia in adult mice. J Physiol 2002;540:593–606.        [ Links ]

191. Zhang Y, Lundeberg T, Yu L. Involvement of neuropeptide Y and Y1 receptor in antinociception in nucleus raphe magnus of rats. Regul Pept 2000;95:109–113.        [ Links ]

192. Wang JW, Lundeberg T, Yu LC. Antinociceptive role of oxytocin in the nucleus raphe magnus of rats, an involvement of mu–opioid receptor. Regul Pept 2003;115:153–159.        [ Links ]

193. Zubrzycka M, Fichna J, Janecka A. Inhibition of trigemino–hypoglossal reflex in rats by oxytocin is mediated by mu and kappa opioid receptors. Brain Res 2005;1035:67–72.        [ Links ]

194. Ebner K, Wotjak CT, Landgraf R et al. Forced swimming triggers vaso–pressin release within the amygdala to modulate stress–coping strategies in rats. Eur J Neurosci 2002;15:384–388.        [ Links ]

195. Wigger A, Sanchez MM, Mathys KC et al. Alterations in central neuropeptide expression, release, and receptor binding in rats bred for high anxiety: critical role of vasopressin. Neuropsychopharmacology 2004;29:1–14.        [ Links ]

196. Fickel J, Bagnol D, Watson SJ et al. Opioid receptor expression in the rat gastrointestinal tract: a quantitative study with comparison to the brain. Brain Res Mol Brain Res 1997;46:1–8.        [ Links ]

197. McIntosh CH, Jia X, Kowk YN. Characterization of the opioid receptor type mediating inhibition of rat gastric somatostatin secretion. Am J Physiol 1990;259:G922–G927.        [ Links ]

198. Lippl F, Schusdziarra V, Allescher HD. Effect of endomorphin on somatostatin secretion in the isolated perfused rat stomach. Neuropeptides 2001;35:303–309.        [ Links ]

199. Narita M, Mizoguchi H, Narita M et al. Absence of G–protein activation by mu–opioid receptor agonists in the spinal cord of mu–opioid receptor knockout mice. Br J Pharmacol 1999;126:451–456.        [ Links ]

200. Sakurada S, Hayashi T, Yuhki M. Differential antinociceptive effects induced by intrathecally–administered endomorphin–1 and endomorphin–2 in mice. Jpn J Pharmacol 2002;89:221–223.        [ Links ]

201. Tseng LF. The antinociceptive properties of endomorphin–1 and endomorphin–2 in the mouse. Jpn J Pharmacol 2002;89:216–220.        [ Links ]

202. Ohsawa M, Mizoguchi H, Narita M et al. Differential antinociception induced by spinally administered endomorphin–1 and endomorphin–2 in the mouse. J Pharmacol Exp Ther 2001;298:592–597.        [ Links ]

203. Sakurada S, Hayashi T, Yuhki M et al. Differential antinociceptive effects induced by intrathecally administered endomorphin–1 and endomorphin–2 in the mouse. Eur J Pharmacol 2001;427:203–210.        [ Links ]

204. Przewlocki R, Labuz D, Mika J et al. Pain inhibition by endomorphins. Ann N Y Acad Sci 1999;897:154–164.        [ Links ]

205. Przewlocki R, Przewlocka B. Opioids in chronic pain. Eur J Pharmacol 2001;429:79–91.        [ Links ]

206. Wu SY, Ohtubo Y, Brailoiu GC et al. Effects of endomorphin on substantia gelatinosa neurons in rat spinal cord slices. Br J Pharmacol 2003;140:1088–1096.        [ Links ]

207. Yajiri Y, Huang LY. Actions of endomorphins on synaptic transmission of Adelta–fibers in spinal cord dorsal horn neurons. J Biomed Sci 2000;7: 226–231.        [ Links ]

208. Wang QP, Zadina JE, Guan JL et al. Morphological studies of the endomorphinergic neurons in the central nervous system. Jpn J Pharmacol 2002;89:209–215.        [ Links ]

209. Williams CA, Wu SY, Dun SL et al. Release of endomorphin–2 like substances from the rat spinal cord. Neurosci Lett 1999;273:25–28.        [ Links ]

210. Wu SY, Dun SL, Wright MT et al. Endomorphin–like immunoreactivity in the rat dorsal horn and inhibition of substantia gelatinosa neurons in vitro. Neuroscience 1999;89:317–321.        [ Links ]

211. Kiefel JM, Cooper ML, Bodnar RJ. Serotonin receptor subtype antagonists in the medial ventral medulla inhibit mesencephalic opiate analgesia. Brain Res 1992;597:331–338.        [ Links ]

212. Kiefel JM, Cooper ML, Bodnar RJ. Inhibition of mesencephalic morphine analgesia by methysergide in the medial ventral medulla of rats. Physiol Behav 1992;51:201–205.        [ Links ]

213. McGowan MK, Hammond DL. Intrathecal GABAB antagonists attenuate the antinociception produced by microinjection of L–glutamate into the ventromedial medulla of the rat. Brain Res 1993;607:39–46.        [ Links ]

214. McGowan MK, Hammond DL. Antinociception produced by microinjection of L–glutamate into the ventromedial medulla of the rat: mediation by spinal GABAA receptors. Brain Res 1993;620:86–96.        [ Links ]

215. Spinella M, Cooper ML, Bodnar RJ. Excitatory amino acid antagonists in the rostral ventromedial medulla inhibit mesencephalic morphine analgesia in rats. Pain 1996;64:545–552.        [ Links ]

216. van PH, Frenk H. The role of glutamate in opiate descending inhibition of nociceptive spinal reflexes. Brain Res 1990;524:101–105.        [ Links ]

217. Urban MO, Smith DJ. Role of neurotensin in the nucleus raphe magnus in opioid–induced antinociception from the periaqueductal gray. J Pharmacol Exp Ther 1993;265:580–586.        [ Links ]

218. Tseng LF, Tang R. Different mechanisms mediate beta–endorphin– and morphine–induced inhibition of the tail–flick response in rats. J Pharmacol Exp Ther 1990;252:546–551.        [ Links ]

219. Tseng LF, Collins KA. Different mechanisms mediating tail–flick inhibition induced by beta–endorphin, DAMGO and morphine from ROb and GiA in anesthetized rats. J Pharmacol Exp Ther 1991;257:530–538.        [ Links ]

220. Rodriguez FD, Rodriguez RE. Intrathecal administration of 5,6–DHT or 5,7–DHT reduces morphine and substance P–antinociceptive activity in the rat. Neuropeptides 1989;13:139–146.        [ Links ]

221. Sawynok J, Reid A, Nance D. Spinal antinociception by adenosine analogs and morphine after intrathecal administration of the neurotoxins capsaicin, 6–hydroxydopamine and 5,7–dihydroxytryptamine. J Pharmacol Exp Ther 1991;258:370–380.        [ Links ]

222. Suh HH, Fujimoto JM, Tseng LL. Differential mechanisms mediating beta–endorphin– and morphine–induced analgesia in mice. Eur J Pharmacol 1989;168:61–70.        [ Links ]

223. Zhong FX, Ji XQ, Tsou K. Intrathecal DSP4 selectively depletes spinal noradrenaline and attenuates morphine analgesia. Eur J Pharmacol 1985;116:327–330.        [ Links ]

224. Craft RM, Henley SR, Haaseth RC et al. Opioid antinociception in a rat model of visceral pain: systemic versus local drug administration. J Pharmacol Exp Ther 1995;275:1535–1542.        [ Links ]

225. Kolesnikov Y, Pasternak GW. Topical opioids in mice: analgesia and reversal of tolerance by a topical N–methyl–D–aspartate antagonist. J Pharmacol Exp Ther 1999;290:247–252.        [ Links ]

226. Nozaki–Taguchi N, Yaksh TL. Characterization of the antihyperalgesic action of a novel peripheral mu–opioid receptor agonist—loperamide. Anesthesiology 1999;90:225–234.        [ Links ]

227. Li ZH, Shan LD, Jiang XH et al. Analgesic effect of endomorphin–1. Acta Pharmacol Sin 2001;22:976–980.        [ Links ]

228. Hau VS, Huber JD, Campos CR et al. Effect of guanidino modification and proline substitution on the in vitro stability and blood–brain barrier permeability of endomorphin II. J Pharm Sci 2002;91:2140–2149.        [ Links ]

229. Spampinato S, Qasem AR, Calienni M et al. Antinociception by a peripherally administered novel endomorphin–1 analogue containing betaproline. Eur J Pharmacol 2003;469:89–95.        [ Links ]

230. Hung KC, Wu HE, Mizoguchi H et al. Acute antinociceptive tolerance and unidirectional cross–tolerance to endomorphin–1 and endomorphin–2 given intraventricularly in the rat. Eur J Pharmacol 2002;448:169–174.        [ Links ]

231. Labuz D, Przewlocki R, Przewlocka B. Cross–tolerance between the different mu–opioid receptor agonists endomorphin–1, endomorphin–2 and morphine at the spinal level in the rat. Neurosci Lett 2002;334:127–130.        [ Links ]

232. Wu HE, Hung KC, Mizoguchi H et al. Acute antinociceptive tolerance and asymmetric cross–tolerance between endomorphin–1 and endomorphin–2 given intracerebroventricularly in the mouse. J Pharmacol Exp Ther 2001;299:1120–1125.        [ Links ]

233. Wu HE, Darpolor M, Nagase H et al. Acute antinociceptive tolerance and partial cross–tolerance to endomorphin–1 and endomorphin–2 given intrathecally in the mouse. Neurosci Lett 2003;348:139–142.        [ Links ]

234. Spanagel R, Sillaber I, Zieglgansberger W et al. Acamprosate suppresses the expression of morphine–induced sensitization in rats but does not affect heroin self–administration or relapse induced by heroin or stress. Psychopharmacology (Berl) 1998;139:391–401.        [ Links ]

Creative Commons License Todo el contenido de esta revista, excepto dónde está identificado, está bajo una Licencia Creative Commons