Movatterモバイル変換


[0]ホーム

URL:


Jump to content
WikipediaThe Free Encyclopedia
Search

κ-opioid receptor

From Wikipedia, the free encyclopedia
Protein-coding gene in the species Homo sapiens, named for ketazocine

OPRK1
Available structures
PDBOrtholog search:PDBeRCSB
List of PDB id codes

4DJH

Identifiers
AliasesOPRK1, K-OR-1, KOR, KOR-1, OPRK, opioid receptor kappa 1, KOR1, KOP
External IDsOMIM:165196;MGI:97439;HomoloGene:20253;GeneCards:OPRK1;OMA:OPRK1 - orthologs
Gene location (Human)
Chromosome 8 (human)
Chr.Chromosome 8 (human)[1]
Chromosome 8 (human)
Genomic location for OPRK1
Genomic location for OPRK1
Band8q11.23Start53,225,724bp[1]
End53,251,637bp[1]
Gene location (Mouse)
Chromosome 1 (mouse)
Chr.Chromosome 1 (mouse)[2]
Chromosome 1 (mouse)
Genomic location for OPRK1
Genomic location for OPRK1
Band1 A1|1 1.89 cMStart5,658,689bp[2]
End5,676,354bp[2]
RNA expression pattern
Bgee
HumanMouse (ortholog)
Top expressed in
  • endothelial cell

  • pons

  • lateral nuclear group of thalamus

  • testicle

  • Brodmann area 23

  • entorhinal cortex

  • middle temporal gyrus

  • nucleus accumbens

  • spinal ganglia

  • prefrontal cortex
Top expressed in
  • Decidua basalis

  • arachnoid mater

  • gastrula

  • superior frontal gyrus

  • wall of uterus

  • lumbar subsegment of spinal cord

  • embryo

  • submandibular gland

  • embryo

  • endometrium
More reference expression data
BioGPS
More reference expression data
Gene ontology
Molecular function
Cellular component
Biological process
Sources:Amigo /QuickGO
Orthologs
SpeciesHumanMouse
Entrez

4986

18387

Ensembl

ENSG00000082556

ENSMUSG00000025905

UniProt

P41145

P33534

RefSeq (mRNA)

NM_001282904
NM_000912
NM_001318497

NM_001204371
NM_011011
NM_001318735

RefSeq (protein)

NP_000903
NP_001269833
NP_001305426

NP_001191300
NP_001305664
NP_035141

Location (UCSC)Chr 8: 53.23 – 53.25 MbChr 1: 5.66 – 5.68 Mb
PubMed search[3][4]
Wikidata
View/Edit HumanView/Edit Mouse

Theκ-opioid receptor orkappa opioid receptor, abbreviatedKOR orKOP for itsligandketazocine, is aG protein-coupled receptor that in humans is encoded by theOPRK1gene. The KOR is coupled to theG proteinGi/G0 and is among relatedreceptors that bindopioid-like compounds in thebrain and are responsible for mediating the effects of these compounds. These include alteringnociception,mood,reward system, andmotor control.[5][6]

KOR is one of the twoopioid receptors that binddynorphinopioid peptides as the primaryendogenousligands,[7][8] the other receptor being newly deorphanizedGPR139 receptor.[9] In additionoxytocin was found to be apositive allosteric modulator of KOR and a variety of naturalalkaloids,terpenes and synthetic ligands bind to the receptor.[10][11]

Dysregulation of this receptor system has been implicated in multiplepsychiatric disorders including: multipledepressive andanxiety disorders,[6][12][13]disorders of diminished motiviation,[14]schizophrenia,[15][16]borderline personality disorder,[17]bipolar disorder,[18]substance use disorder.[19][20]

Tissue distribution

[edit]

Central nervous system

[edit]

Brain

[edit]

KORs are widely distributed throughout thebrain.[6][21][22][23] Theclaustrum represents thebrain region with the highest density of KORexpression.[24][25][26] The claustrum-prelimbic cortex circuit operates viadynorphin and KOR signaling to modulatecognitive andaffective functions.[27]

AdditionalCNS regions expressing moderate to high KOR densities include theprefrontal cortex,periaqueductal gray, dorsalraphe nuclei (dorsal),ventral tegmental area,substantia nigra,dorsal striatum (putamen,caudate), ventral striatum (nucleus accumbens,olfactory tubercle),amygdala,bed nucleus of the stria terminalis,hippocampus (pyramidal and molecular layers, granular cell layer of thedentate gyrus),hypothalamus,thalamus (centromedian, paraventricular, and centrolateral nuclei),locus coeruleus,spinal trigeminal nucleus,parabrachial nucleus, andsolitary nucleus.[6][22][28][29]

Positron emission tomography (PET) imaging studies with the KOR-selectiveradioligand [11C]GR-103545 in non-humanprimates showed highbinding potential (BPND > 1.3) in thepituitary gland, followed byinsula, claustrum, andorbitofrontal cortex, with moderate binding (BPND 0.9–1.3) in nucleus accumbens, amygdala, and hippocampus.[30][31] [3H]bremazocine binding showed elevated densities along the ventral edge of the nucleus accumbens and ventralputamen regions.[32] Inprairie voles nucleus accumbens shell, claustrum, andventral pallidum exhibited the highest functional KOR activity in bothsexes.[29]

There is evidence that distribution and/or function of this receptor may differ betweensexes.[18][30][33][34]

Spinal cord

[edit]

Inspinal cord, KOR is expressed in thesubstantia gelatinosa and superficial laminae of thedorsal horn, where they modulate thermalnociception and chemicalviscelar pain.[22][35] They are concentrated in the upper laminae of the dorsal horn (laminae I–III) and within theposterolateral tract.[35] The highest density was localized within the inner segment of lamina II, forming a dense band immediately dorsal to lamina III.[35][36] 53% of KOR binding sites in the superficial dorsal horn (laminae I–II) are localizedpresynaptically on primaryafferent terminals, with the remainder distributedpostsynaptically.[37]

Peripheral nervous system

[edit]

Dorsal root ganglia

[edit]

KOR is present indorsal root ganglia (DRG) in moderate expression levels in human tissue.[38] KOR is expressed inpeptidergic primary afferents genes encodingcalcitonin gene-related peptide (CGRP) andsubstance P, as well as in populations of low-thresholdmechanoreceptors that innervatehair follicles.[39] In human DRGneurons, approximately 25%cells expressOPRK1 mRNA.[39]

Immune cells

[edit]

Inimmune cells, KOR is distributed in specificleukocyte populations. Approximately 50% of residentperitonealmacrophages express KOR, while expression decreases during lymphocyte maturation, with less than 25% of splenicT-helper orT-cytotoxiclymphocytes and only 16% ofsplenicB lymphocytes displaying receptor expression.[40][41]

Gastrointestinal tract

[edit]

In thegastrointestinal tract, KOR is expressed onmyenteric andsubmucosal plexus neurons, where they modulateintestinal motility andsecretion.[23][42] Both KOR and MOR mRNAs are expressed in all investigated gastrointestinal regions in one study, with thestomach and proximalcolon displaying the highest expression levels, and theduodenum exhibiting the lowest.[42] KOR in the proximal colon represented 40% of the amount found in the brain.[42]

Myenteric plexusperikarya of the rat stomach and colon areimmunoreactive for KOR, with numerousnerve fibers distributed in the longitudinal and circular muscle layers, whereassmooth muscle cells lack receptor immunoreactivity.[42][43] A higher number of neurons expressing KOR-like immunoreactivity are visualized in the myenteric plexus with a smaller number in thesubmucosal plexus, unlike the distribution pattern of MORs.[43]

Cardiovascular system

[edit]

KORs are expressed in humancardiac tissue, including cardiomyocytes, where they exert negativeinotropic andlusitropic effects throughpertussis toxin-sensitiveGi/o protein signaling.[44][45][46][47]

Renal system

[edit]

Healthy humankidney expresses KOR, yet detailed cellular localization within specificnephron segments aren't investigated.[48]

Subtypes

[edit]

Based onreceptor binding studies, three variants of the KOR: κ1, κ2, and κ3 have been characterized viaradioligand binding and regionalCNS mapping.[49][50] However, only one encodingcDNA has been cloned.,[51] hence these subtypes likely arise from interactions of the KORprotein with othermembrane-associated proteins rather thangene duplication.[52] Historically the understanding that KORs are encoded by a single gene reopened the question of how one receptor system could be involved in such a multiplicity of interactions and disparate profiles.[53]

Function

[edit]

General

[edit]

KORagonism seems to functionally oppose multiple effects mediated byμ-opioid receptors (MOR) andδ-opioid receptors (DOR), includinganalgesia,tolerance,euphoria, andmemory regulation.[54] Activation of KOR bydynorphins duringstress exposure has been shown to inducedysphoria,aversion, and negativeaffective states in both human and non-human subject.[55][56] This contrasts with activation of MOR, which is associated withmood elevation and producinghedonic effects.[57][58] Consequently, the KOR system has traditionally been conceptualized as mediatinganti-reward processes andnegative reinforcement, representing a functional counterpart to MOR in terms of behavioral and affective outcomes.[59][60] However, recent research highlights a more nuanced role for KOR signaling, implicating it in a spectrum of complex behaviors and neural processes that extend beyond a strictly dichotomous and unidemensional frameworks, including functions independent ofhedonic tone within reward processing.[53][61][62][63]

Centrally active KOR agonists have distinct, atypicaldissociativehallucinogenic effects, as exemplified bysalvinorin A (the active constituent inSalvia divinorum). The experiences include:dissociation, incapacitation,psychotomimesis, profound alterations ininteroception, somatic sensations,visual andauditory hallucinations,synesthesia (particularly visual-proprioceptive binding),sedation,anelgesia,anti-inflammation,neuroprotection,memory impairment,anti-addiction, aversion, dysphoria,anxiogeny, bothantidepressant anddepressogenic effect.[64][65][66]

Signaling bias

[edit]

Main section:§ Signaling dynamics

Many functional differences between KOR agonists can be explained bybiased signaling, whereby different agonists preferentially activate distinctsignaling pathways downstream of the receptor.[67][68][69] Evidence suggests thatG protein signaling primarily mediates the therapeutic analgesic andantipruritic effects of KOR agonists, whilstβ-arrestin2-dependent signaling throughp38 MAPK activation mediates adverse dysphoric, sedative, and aversive effects.[59][70]

Limitations

[edit]

Studying the exact functions mediated by KOR is limited by the non-selectivity and signaling biases of the compounds used in the research and naturally occuring in the human body.[69] Dynorphinpeptides,endogenous agonists of KOR, especiallybig dynorphin, are direct complex modulators of theNMDA receptor.[69] Certain dynorphin peptides also have affinity for the MOR and DOR and influence other pathways that are not directly coupled to KOR.[71][53] KOR activation in the context ofin vivo stress responses could be biased for β-arrestin2 and other pathways related to dysphoria due to the presence ofcorticotropin-releasing hormone (CRF).[12][72] Salvinorin A as well as other KOR agonists have been found to possess properties such asdopamine D2 receptor agonism with lower, but non-negligibleaffinity andpotency.[73][74] Salvinorin A is a balancedG protein and β-arrestin2 agonist.[75][76]

Pain

[edit]

Similarly toμ-opioid receptor (MOR), KOR activation produces antinociceptive effects. KOR agonists are potentlyanalgesic and have been employed clinically forpain management, but they produce characteristicadverse effects which while limit theirabuse potential, they also limit their therapeutic utility.[13][77]

The receptor mediates acute thermal and mechanicalpain processing. The anelgesic actions of KOR occur at bothspinal and supraspinal sites.[35] In thespinal cord,presynaptic activation suppressesnociceptive transmission through inhibition ofcalcium influx and reduction ofneurotransmitter release from primarysensory neurons.[78] In models ofchronic inflammation, KORmRNA levels increased in laminae I–II of the spinal dorsal hornipsilateral to inflamedpaws ofrats.[79]

Neuropathic pain followingperipheral nerve injury is accompanied by sustained elevation of dynorphin levels in the spinal dorsal horn, resulting in tonic KOR activation that contributes to pain inhibition.[80] Theprodynorphin-derived opioid system within the spinal cord exhibits bothpronociceptive and antinociceptive functions. Acute KOR activation produces pain reversal and chronic stimulation leads to receptortolerance and hyperalgesia withallodynia. Mechanisms such as activation ofNMDA receptors onspinal interneurons, and increasing glutamate and substance P release from primary afferent terminals might play a role.[80]

KOR also mediates theaffective-motivational dimensions of pain.[59] At the supraspinal level, KOR activation in theventral tegmental area,periaqueductal gray, and other pain-modulatorynuclei influences both pain perception and pain-related motivated behavior.[59] The engagement of KOR duringchronic pain states, particularlyneuropathic pain, has been implicated in the high comorbidity between chronic pain andmood disorders, as dynorphin-mediated KOR signaling inlimbic andreward-related brain regions drives negativeemotional states andanhedonia.[59]

Memory

[edit]

Receptor activation is linked to impairing multiplememory processes, includingworking memory,spatial learning, and fearmemory consolidation, by inhibitingsynaptic plasticity such aslong-term potentiation (LTP) in regions like theamygdala andhippocampus.[81][82][83] In models ofamnesia, endogenous receptor activation leads to reactivation of memory traces, prolonging retention latency in inhibitory avoidance tasks, withantagonism often protecting against stress-induced deficits.[83] Receptor activation by dynorphins reduces the intensity of the emotional aspect of memories.[84]

Neuroendocrine

[edit]

KOR agonists increase serumprolactin levels by tonic inhibition ofhypothalamicdopaminergic systems.[85] This response occurs following administration of bothcentrally penetrating andperipherally restricted KORagonists.[86][87]

Activation of KOR producediuretic effects through negative regulation ofvasopressin, also known asantidiuretichormone (ADH).[88][89] This waterdiuresis is characterized by increased urine volume and decreasedurineosmolality without prominent alterations inelectrolyte excretion. Both centrally and peripherally acting KOR agonists promote diuresis through mechanisms including decreasedantidiuretic hormone secretion from the hypothalamus andposterior pituitary, reducedrenal responsiveness to antidiuretic hormone, and modulation of renalsympathetic nerve activity.[90] KOR signaling in renal tissue may also modulate responses to metabolic stress and inducepathophysiological processes inkidney disease.[48]

Activation of the receptor increasesadrenocorticotropic hormone (ACTH) andcortisol levels in humans and non-human primates through activation of thehypothalamic-pituitary-adrenal axis (HPA).[91] Administration of the selective agonistU50,488 dose-dependently stimulates ACTH and cortisol release, an effect specific to KOR activation and not observed followingμ-opioid (MOR) orδ-opioid receptor (DOR) stimulation.[91]

KOR exhibits coexpression withoxytocin andvasopressin in theparaventricular nucleus (PVN) andsupraoptic nucleus (SON) of the hypothalamus.[92][93][94] Functional KORs are present onnerve terminals of both oxytocin and vasopressin neurons in the ratneurohypophysis, where agonists inhibitpotassium-evoked hormone release.[95] Both dynorphin-(1-8) and -(1-17) suppress stimulated oxytocin release from isolatedneurosecretory endings, with effects on the initial and secondary peaks of hormone secretion, while exerting no influence on vasopressin release under similar conditions.[96][97] These interactions extend to plasma hormone levels, where KOR agonists decrease circulating oxytocin concentrations, whileantagonists increase oxytocin release, suggesting that KOR signaling mediates negative regulation of oxytocinsecretion duringstress or physiological challenges.[97][98]

Mood and stress

[edit]

The involvement of KOR instress, as well as in consequences ofchronic stress such asdepression,anxiety,anhedonia,psychosis, and modulatingdrug-seeking behavior, has been made clear.[12][13][15][99][100] KOR plays an important and varied role in regulating variousaffective state andstress responses in humans through multiple complex processes.[12][19][100][101][102] There have been numerous studies implicating KOR in pathology of variouspsychiatric disorders.[6][13][15][99][103][104]

CRF-dynorphin-KOR cascade

[edit]

Diverse stressors initiate CRF release which subequently leads to dynorphin release and KOR activation inlimbic circuits. This integrated stress response is mediated primarily bycorticotropin-releasing factor (CRF), one of the mainneuropeptide integrators of the stress response.[12][98][100]

Physical stressors trigger CRF release from thehypothalamicparaventricular nucleus (PVN). Cold exposure provokes CRF secretion from the hypothalamus and produces increases in plasmaglucocorticoids.[105][106] Acute physical stressors such as forced swimming, inescapable footshock, and restraint stress similarly lead to CRF release.Acute stress induces rapid increases in plasmacorticosterone levels that are dependent on CRFsecretion.[98][105][107]Intravenous CRF administration induces rapid increases in KORphosphorylation instriatal,VTA,amygdaloid,hippocampal, andnucleus accumbens (NAcc) components of stress and anxiety circuits.[101] These CRF-induced increases are absent inprodynorphin (PDYN)knockout mice.[108]Social defeat stress model activates CRF anddynorphin.Learned helplessness also engages these systems.Fear conditioning and fear-related stress produce KOR-dependent behavioral responses through CRFreceptor. The uncontrollability and inescapability of the stressor substantially augments the response. Even brief, non-intensive stressors produce significant neurobiological and behavioral effects when unpredictable and uncontrollable, whereas identical stressors with controllability demonstrate attenuated activation of stress systems.[100][109]

CRF, produced in thePVN, activatesCRF₁ receptors andCRF₂ receptors distributed across limbic circuits. CRF₁ receptor activation, which mediates rapid and intense stress responses, triggers acute dynorphin release inlimbic stress-responsive regions including the NAcc,basolateral amygdala,dorsal raphe nucleus (DRN), hippocampus, andbed nucleus of the stria terminalis (BNST).[12][107][108] CRF₂ receptor activation, generally associated with slower, later-phase stress response components, also induces dynorphin-dependent aversive responses such asconditioned place aversion (CPA). Subsequently dynorphin activates KORs expressed onGABAergic anddopaminergicneurons, encoding theaversive anddysphoric qualities of stress exposure.[12][108][110]

Acutely stress-induced dynorphin release and KOR activation have evolutionarily adaptive functions. KOR-mediatedanalgesia facilitates physical escape responses to threat and concurrent KOR-induced dysphoria and aversion promoteavoidance and active coping. However, during the delayed temporal phase following acute stress exposure (hours to days), stress-induced KOR signaling initiatesintracellular signaling cascades includingp38 MAPK andextracellular signal-regulated kinases (ERK) which phosphorylatetranscription factors such ascAMP response element-binding protein (CREB) and alter dynorphin and KORgene expression itself, establishing a self-amplifying cycle.[111]

Chronic social defeat stress produces a counterintuitive long-lastingdownregulation ofprodynorphinmRNA levels in the NAcc (occurring by day 10 of chronic exposure), and this downregulation is reversed by chronic treatment with standardantidepressant medication (imipramine).[112] Despite this molecular downregulation, behavioral signs of stress-induced dysphoria, anhedonia, and anxiety persist and even intensify with repeated stress exposure, indicating that the coupling between dynorphin release and KOR phosphorylation, as well as the downstream consequences of KOR activation, may become sensitized through counter-adaptations in post-receptor signaling or in competing inhibitory circuits.[112] This process involves:§ Signaling after internalisation.

Serotonergic pathway regulation

[edit]

KOR activation suppressesserotonergic tone through multiple mechanisms, including regulation of theserotonin transporter (SERT). Dynorphin, released from local GABAergic neurons within reward-related regions, binds to KORs expressed on serotonergic terminals projecting from the DRN to regions including such as NAcc,prefrontal cortex, and other limbic structures associated with mood regulation, reducing the availability of serotonin to activatepostsynaptic5-HT1A receptors in these targets.[113][114][115]Agonist-induced binding to these receptors triggers rapid,concentration-dependent downregulation of SERT function throughCaMKII andAkt.[116][117] This downregulation occurs through increased internalization of SERT from theplasma membrane viadynamin-dependentendocytosis, coupled with increased phosphorylation of thetransporter protein, reducing its functional availability and depressing serotonergicneurotransmission in hedonic circuits.[116] Under conditions of chronic stress, sustained KOR-mediated suppression could result in blunted responsiveness of postsynaptic neurons to residual serotonin, effectively rendering them hyporesponsive.[56][118]

In the NAcc, stress-dependent upregulation of postsynaptic5-HT1B receptors co-expressed ondirect pathway neurons expressing prodynorphin is an additional downstream mechanism. 5-HT1B receptor, also coupled toGi/o proteins, mediates serotonin-dependent inhibition ofdopaminergic neuronexcitability and modulate the balance between reward approach andbehavioral inhibition. Chronic stress-induced elevation of 5-HT1B expression in these accumbens neurons paradoxically increases sensitivity todopamine suppression and amplifies the anhedonic phenotype despite the simultaneous reduction in baseline serotonin availability.[119]

Dopaminergic pathway and reward suppression

[edit]

Themesolimbic dopaminergic circuit functions as a substrate for KOR-regulated mood homeostasis.[20][120]Dynorphin is synthesized and released bydopamine D1 receptor-expressingmedium spiny neurons within the NAcc, establishing a localnegative feedback loop that suppressesdopamine release.[121] KOR activation on dopamine terminals inhibits dopamine release through multiple mechanisms: increasedpotassium conductance viaG protein-coupledinward-rectifier potassium (GIRK) channels, suppression ofcalcium entry, activation ofprotein kinase C-β (PKCβ),c-Jun N-terminal kinase (JNK), and ERK,[122] as well as facilitation ofdopamine transporter (DAT) function through ERK1/2-dependent pathways that accelerate dopamine reuptake.[123] Additionally, KOR activation on local dynorphin-expressing neurons produces presynaptic inhibition of both glutamatergic and GABAergic afferents onto D1 receptor-expressing medium spiny neurons, with preferential suppression of amygdala inputs to D1-MSNs while facilitating integration of hippocampal/amygdalar inputs onto D2 receptor-expressing neurons through disinhibition.[124]

In the caudal NAcc shell, KOR-induced dopamine suppression triggersanxiogenic behaviors accompanied by reducedlocomotor activity.[125] Conversely, in the rostral shell, KOR activation produces attenuated dopaminergic suppression with diminished aversive behavioral consequences.[126] This topographic architecture extends to the NAcc core, where KOR-mediated dopamine inhibition similarly manifests with greater intensity in the caudal relative to rostral subregion.[127] The DRN to ventral tegmental area (VTA) circuit is an additional stress-responsive pathway whereby prodynorphin-expressing neurons release dynorphin at dopaminergic terminals, enabling KOR-dependent suppression of dopamine neuron excitability during acute stressors.[122][126][127][128]

Addiction

[edit]

The KOR system is involved in increaseddrug-seeking behavior.[12] KORagonists have been investigated for their therapeutic potential in the treatment of addiction.[129] and evidence points towardsdynorphinpeptides, theendogenous KOR agonists, to be the body's natural addiction control mechanism.[130]Childhood stress andabuse are well-known predictors ofdrug abuse which is reflected in alterations of the MOR and KOR systems.[131] In experimental "addiction" models the KOR has also been shown to influence stress-inducedrelapse to drug seeking behavior. For the drug-dependent individual, risk of relapse is a major obstacle to becoming drug-free. Recent reports demonstrated that KORs are required for stress-induced reinstatement ofcocaine seeking.[132][133]

Thenucleus accumbens (NAcc) and broaderstriatum are among the brain regions most strongly associated with addiction, although other structures that project to and from the NAcc also play critical roles in addictive processes. Though many other changes occur, addiction is often characterized by the reduction in the availability ofdopamine D2 receptors in the NAcc.[134] In addition to decreasing NAcc D2 binding,[135][136] cocaine is also known to produce a variety of changes to theprimate brain such as increases ofprodynorphinmRNA in caudate putamen and decreases of the same polypeptide in thehypothalamus. The administration of a KOR agonist produced an opposite effect, causing an increase in D2receptor availability in the NAcc.[137]

Additionally, while cocaineoverdose victims showed a large increase in KORs (doubled) in the NAcc,[138] KOR agonist administration is shown to be effective in decreasing cocaine seeking and self-administration.[139] Furthermore, while cocaine abuse is associated with loweredprolactin response,[140] KOR activation causes a release of prolactin,[87] ahormone known for its important role in learning,neuronal plasticity andmyelination.[141]

It has also been reported that the KOR system is critical for stress-induced drug-seeking. In animal models, stress has been demonstrated to potentiate cocaine reward behavior in a kappa opioid-dependent manner.[142][143] These effects are likely caused by stress-induced drug craving that requires activation of the KOR system. Although seemingly paradoxical, it is well known that drug taking results in a change fromhomeostasis toallostasis. It has been suggested thatwithdrawal-induceddysphoria or stress-induced dysphoria may act as a driving force by which the individual seeks alleviation via drug taking.[144] The rewarding properties of drug are altered, and it is clear KOR activation following stress modulates thevalence of drug to increase its rewarding properties and cause potentiation of reward behavior, or reinstatement to drug seeking. The stress-induced activation of KORs is likely due to multiple signaling mechanisms. The effects of KOR agonism ondopamine systems are well documented, and recent work also implicates thep38 MAPK cascade and pCREB in KOR-dependent behaviors.[145][111]

While the predominant drugs of abuse examined have been cocaine (44%),ethanol (35%), and opioids (24%).[146] As these are different classes ofdrugs of abuse working through different receptors (increasing dopamine directly and indirectly, respectively) albeit in the same systems produce functionally different responses. Conceptually then pharmacological activation of KOR can have marked effects in any of thepsychiatric disorders (clinical depression,bipolar disorder,anxiety disorder, etc.) as well as variousneurological disorders (i.e.Parkinson's disease andHuntington's disease).[20][102] Not only aregenetic differences in dynorphin receptor expression a marker foralcohol dependence, but a single dose of a KORantagonist markedly increasedalcohol consumption inrats.[147] There are numerous studies that reflect a reduction inself-administration of alcohol,[148] andheroin dependence has also been shown to be effectively treated with KOR agonism by reducing the immediate rewarding effects[149] and by causing the curative effect ofupregulation (increased production) of MORs[150] that have beendownregulated during opioid abuse.

The anti-rewarding properties of KOR agonists are mediated through both chronic and acute effects. The immediate effect of KOR agonism leads to reduction ofdopamine release in the NAcc during self-administration of cocaine[151] and over the chronic period upregulates receptors that have been downregulated during substance abuse such as the MOR and the D2 receptor. These receptors modulate the release of otherneurochemicals such asserotonin in the case of MOR agonists, andacetylcholine in the case of D2. These changes can account for the physical and psychological remission of the pathology of addiction. The longer effects of KOR agonism (30 minutes or greater) have been linked to KOR-dependent stress-induced potentiation and reinstatement of drug seeking. It is hypothesized that these behaviors are mediated by KOR-dependent modulation of dopamine, serotonin, ornorepinephrine and/or via activation of downstreamsignal transduction pathways.

Of significant note, while KOR activation blocks many of the behavioral and neurochemical responses elicited by drugs of abuse as stated above. These results are indicative of the KOR induced negative affective states counteracting the rewarding effects of drugs of abuse. Implicating the KOR/dynorphin system as an anti-reward system, supported by the role of KOR signaling and stress, mediating both stress-induced potentiation of drug reward and stress-induced reinstatement of seeking behavior.[20][102] This in turn addresses what was thought to be paradoxical above. That is, rather, KOR signaling is activated/upregulated by stress, drugs of abuse and agonist administration - resulting in negative affective state. As such drug addiction is maintained by avoidance of negative affective states in stress, craving, and drug withdrawal.[152] Consistent with KOR induced negative affective states and role in drug addiction, KOR antagonists are efficacious at blocking negative affect induced by drug withdrawal and at decreasing escalated drug intake in pre-clinical trial involving extended drug access.[20][146][102]

Traditional models of KOR function in drug addiction have postulated that KOR signaling is associated with dysphoria andaversion, thought to underlie the stress-induced exacerbation of addiction. However, recent research in animal models has proposed alternative models, suggesting that KOR-mediated responses may not act directly on negative valence systems but modulate related processes such as novelty processing.[153][154] Studies in humans came to similar conclusions that KORs may modulate various aspects of reward processing in a manner that is independent of the hedonic valence traditionally ascribed to them.[14][155] This broadens the potential understanding of KORs in addiction beyond a unidimensional framework, implicating their role in complex behaviors and treatment approaches that do not align strictly with stress or aversion. These emerging perspectives may inform the development of novel pharmacotherapies targeting KORs for the treatment ofsubstance use disorders, as they highlight the receptor's multifaceted role in addiction.

Consciousness and altered states

[edit]
Main articles:Claustrum andConsciousness

Claustral theories

[edit]

Theclaustrum is the region of thebrain in which the KOR is most densely expressed.[24][26][156] Historically, it has been proposed on the basis of the claustrum's structural and connectivity characteristics that this region orchestrates diverse brain functions and serves as a critical substrate forconsciousness.[24][26] Clinical observations supported this hypothesis:lesions of the claustrum in humans are associated with disruption of consciousness andcognition, andelectrical stimulation of theinsula-claustrum border has been found to produce immediate loss of consciousness in humans, with recovery upon cessation of stimulation.[26][157] Earlier theories proposed that inhibition of the claustrum (as well as, "additionally, the deep layers of the cortex, mainly in prefrontal areas") by activation of KORs in these areas is primarily responsible for the profound consciousness-altering atypicaldissociativehallucinogen effects ofsalvinorin A and other KORagonists.

According to Addy et al.:[156]

Theories suggest the claustrum may act to bind and integrate multisensory information, or else to encode sensory stimuli as salient or nonsalient (Mathur, 2014). One theory suggests the claustrum harmonizes and coordinates activity in various parts of the cortex, leading to the seamless integrated nature of subjective conscious experience (Crick and Koch, 2005; Stiefel et al., 2014). Disrupting claustral activity may lead to conscious experiences of disintegrated or unusually bound sensory information, perhaps including synesthesia. Such theories are in part corroborated by the fact that [salvia divinorum], which functions almost exclusively on the KOR system, can cause consciousness to be decoupled from external sensory input, leading to experiencing other environments and locations, perceiving other "beings" besides those actually in the room, and forgetting oneself and one's body in the experience.

From this perspective, disrupting claustral activity might lead to conscious experiences of disintegrated or unusually bound sensory information, includingsynesthesia.[156] However, even early formulations acknowledged that their assumptions are merely tentative and that "KORs are not exclusive to the claustrum; there is also a fairly high density of receptors located in theprefrontal cortex,hippocampus,nucleus accumbens andputamen", and that "disruptions to other brain regions could also explain the consciousness-altering effects [of salvinorin A]".[26]

Current neuroimaging evidence

[edit]

The task of elucidating the exact role of claustrum in mediating sensory information and conscioussness remains a topic of active debate.[158] And findings on whether distruptions of claustral activity lead to the loss of consciousness are conflicting.[26][159][160][161]

Recent imaging studies have confirmed the suspected complexity and multi-regional character of specifically KOR-mediated alterations, and argued that the neural substrates involvecortico-thalamic integration anddefault mode network (DMN) disruption rather than claustrum-centric mechanism.[65] Salvinorin A induces decreases in default mode network connectivity, specifically within the medial prefrontal cortex andposterior cingulate cortex and increased between-network connectivity with reduced dynamic connectivity stability.[65] While both salvinorin A andpsilocybin attenuate default mode network connectivity, their effects on thalamocortical networks differ; salvinorin A-induced thalamic modulation is independent of5-HT2A receptor activation.[162]

Thethalamus, especially thecentromedian,paraventricular, andcentrolateral nucleus, expresses high KOR density and mediates corticalarousal, viscero-limbic integration, and relay of sensory andinteroceptive information to cortical processing hierarchies. KOR activation within these thalamic nuclei reduces the relay ofexteroceptive andinteroceptive information to the cortex, producing the characteristic dissociation from external reality and loss of contact withself-representation andbody schema.[64][65] Salvinorin A induces prominentauditory phenomena and gating of audio-visual information at the perceptual threshold, coupled with unusual modifications of interoceptive awareness and body ownership that exhibit inverted-U-shapeddose-response relationship. Low to moderate doses enhance sensations and perceived body-safety, whereas high doses producedepersonalization, loss of body awareness,out-of-body experiences, and subjective feelings of existing in alternative spatial or dimensional realities, sometimes as objects or alternatively otherliving organisms.[65]

KOR activation also suppresses activity in sensory-integration regions, includingparietal andtemporal areas involved in body schema codification and multisensory binding, whilst simultaneously disrupting medial prefrontal cortex-mediated self-referential processing within the DMN.[64] The claustrum is embedded within cortico-claustro-cortical loops that depend on maintained thalamic-cortical communication; consequently, thalamic KOR activation may disrupt claustral function indirectly through compromisedafferent andefferent signaling rather than through direct local inhibition.[162] Collectively, it is likely that KOR-mediated experiences of dimensionality alterations, synesthesia, and modified temporal perception represent emergent properties of disrupted hierarchical sensory integration at thalamic and cortical levels coupled with claustrum activity rather than direct consequences of that single region or modulation of its pathways.[65][162]

Heart

[edit]

In thecardiovascular system, KOR activation produces negativeinotropic andlusitropic effects incardiac tissue throughpertussis toxin-sensitiveGi/o protein signaling.[44] KOR activation duringmyocardial infarctionreperfusion reduces infarct size throughERK1/2-dependent mechanisms, suggestingcardioprotective effects.[47][163] In cardiomyopathic hearts, KOR-mediated cardiac depression is augmented through increased inhibition ofcAMP accumulation and decreased amplitude of systolicCa2+ transients.[44]Ventricular arrhythmias resulting fromadministration of certainantagonists are attributed to the presence of KORs in theheart.[164]

Other

[edit]

A variety of other effects of KOR are known

Ligands

[edit]
22-Thiocyanatosalvinorin A (RB-64) is afunctionally-selective κ-opioid receptor agonist.

Agonists

[edit]

The synthetic alkaloidketazocine[172] andterpenoid natural productsalvinorin A[173] are potent and selective KORagonists. The KOR also mediates thedysphoria andhallucinations seen with opioids such aspentazocine.[174]

Benzomorphans
Morphinans
Arylacetamides
Peptides (endo-/exogenous)
Terpenoids
Others/unsorted

Nalfurafine (Remitch), which was introduced in 2009, is the first selective KOR agonist to enter clinical use.[180][181]

Antagonists

[edit]

Allosteric modulators

[edit]

Positive allosteric modulators

[edit]

Negative allosteric modulators

[edit]
  • c[D-Trp-Phe-β-Ala-β-Ala][188]

Genetics

[edit]

The humanOPRK1gene is located onchromosome 8 and comprises fourexons separated by threeintrons, spanning approximately 25kilobases.[189] The gene utilizes at least threetranscription initiation sites, generatingmRNAs with5′-UTRs of 215–299nucleotides, with the predominantisoform containing 238 nucleotides of 5′-UTRsequence.[189] The exon-intron organization is conserved between human,mouse, andratOPRK1 genes.[189]

Polymorphisms

[edit]

Single nucleotide polymorphisms (SNPs) genetic variations withinOPRK1, have been associated with susceptibility tosubstance use disorders andstress-related behaviors. The G36T SNP (rs1051660) is more frequent inheroin-dependent individuals compared to healthy controls.[190] Other study found an association ofOPRK1 variants withcocaine dependence andrelapse susceptibility.[191]

Epigenetics

[edit]

Early life stress

[edit]

Epigenetic modifications includingDNA methylation andhistone acetylation regulateOPRK1gene expression in response to environmental factors such asearly life stress andpsychological trauma. Decreased DNA methylation in intron 2 ofOPRK1, functioning as agene enhancer, has been observed in theanterior insula of individuals with histories ofchildhood abuse, which correlates with alteredreceptor expression and stress responsivity.[192]

Postmortem samples fromsuicide completers with a history of severechild abuse (CA) had higher rates of KOR downregulation relative to controls and suicide completers without CA history, an effect not accompanied by alterations in multiple other genes.[192] Hypomethylation ofOPRK1 intron 2 was associated with the CA group, as low levels ofDNA methylation facilitateglucocorticoid binding and subsequent regulation ofOPRK1transcription. Additionally, a specific insertion deletion (INDEL)polymorphism, rs35566036, in theOPRK1promoter region occurred more frequently in suicide completers withmajor depressive disorder relative to healthy controls.[192]

Borderline personality disorder

[edit]

Similar epigenetic alterations inOPRK1 methylation patterns have been linked toborderline personality disorder (BPD), where an imbalance between opioid receptor systems could cause symptoms such as chronicdysphoria,suicidality, and emotional instability.[17][193][194] In individuals with BPD, decreased DNA methylation (hypomethylation) in a differentially methylated region (DMR) located within thepromoter region, specifically at a cluster of five adjacentCpG sites (CG34–CG38) positioned immediately upstream of core CpG islands (CGI-1 and CGI-2), results in enhancedgene transcription and elevated KOR expression.[17]

The DMR hypomethylation in BPD is strategically positioned on the "falling slope" of the gene's methylation gap; a transition zone between the sparsely methylated CpG island promoter and densely methylated downstream regions. This location amplifies the functional consequences of hypomethylation by progressively steepening the methylation gradient, further facilitating transcription initiation at multiple transcription start sites (TSS) distributed throughout the CGI promoter region. Consequently, the decreased methylation rates in the DMR are associated with increasedOPRK1 mRNA transcription and heightened KORprotein expression in peripheralwhite blood cells and, by extension, in central brain regions involved inemotion regulation.[17]

Symptom severity in BPD correlates with DMR hypomethylation levels. As DMR methylation rates decrease (become more hypomethylated), BPD symptom severity measured by the Borderline Symptom List (BSL-23) increases. Additionally, heightened traitimpulsivity, measured by theBarratt Impulsivity Scale, and particularly its motor impulsivity subscale, showsinverse relationships with DMR methylation levels.[17]

The epigenetic imbalance may also impact social attachment and interpersonal functioning through effects onmu-opioid receptor (MOR).Childhood neglect produces chronic basal understimulation of MORs, which mediatereward and socialmotivation. Paradoxically, prolonged MOR understimulation may trigger compensatory MORupregulation in regions such as theamygdala andorbitofrontal cortex. This MORhypersensitization, with its heightened responsivity to negative affective stimuli, may in turn provoke strong counter-activating KOR responses, resulting in the increasedOPRK1 expression observed epigenetically. This KOR-MOR imbalance, where relative KOR overactivity combines with contextually inappropriate MOR hyperexcitability, likely affects BPD's dysregulation of interpersonal relationships and affective instability.[17]

Substances

[edit]
Alcohol
[edit]

Repeatedalcohol exposure alters bothDNA methylation andhydroxymethylation of theOPRK1 promoter in thenucleus accumbens, a key reward centre in alcohol-preferringrodent models.[195] Chronic intermittentethanol exposure reduces both5-methylcytosine and5-hydroxymethylcytosine percentages in theOPRK1 promoter, leading to changes in receptor expression correlated with addiction-relatedmotivational andreward behaviours.[195]

Stimulants
[edit]

Cocaine andmethamphetamine exposure induce epigenetic modifications of theOPRK1 andPDYN loci through both histone remodeling and DNA methylation pathways.[196] Acute cocaine and methamphetamine increasehistone H4acetylation andhistone acetyltransferase (HAT) activity in thestriatum, facilitating increasedPDYN andOPRK1 transcription that initiatesdynorphin-mediated counter-inhibition ofdopamine release.[196] This acute epigenetic activation could be interpreted as a compensatory mechanism attempting to restore dopaminergic homeostasis during drug-induced dopaminergic overstimulation.

In chronic context cocaine and methamphetamine exposure reverse this epigenetic profile through increasedDNA methyltransferase (DNMT) activity andhistone deacetylase (HDAC)-mediated repression of plasticity genes, including decreasedOPRK1 transcription.[196] Epigenetic silencing of adaptability genes consolidates compulsive drug-seeking behaviour whilst simultaneously dysregulating the KOR-mediated feedback system, facilitating withdrawal-related dysphoria and relapse vulnerability.[197]

Opioids
[edit]

In humansaddicted to opioids, epigenetic modifications of theOPRK1 gene, including altered DNA methylation profiles inperipheral blood cells, correlate with substance use severity andwithdrawal symptoms.[198]

Gastrointenstinal tract

[edit]

Age-related changes inOPRK1 gene expression were observed in mousegastrointestinal tract, with mRNA expression significantly decreased in the distalileum in 12-month-old mice compared to 6-month-old animals, though nostatistically significant differences were detected in the stomach and colon.[199] Protein expression of dynorphin in the colon was lower in older mice.[199]

Receptor oligomers

[edit]

Heteromerization with otherG protein-coupled receptors (GPCRs) produces complexes with differingligandselectivity and signaling properties.[39][52] They show alteredG protein coupling,receptor trafficking, andtissue distribution compared tohomodimers.[200] Targeting specific KOR-containingheteromers with bivalentligands may yieldanalgesics with fewerdysphoric effects, which could be relevant foraddiction research andtherapy.[201]

Heterodimer of KOR withδ-opioid receptor (DOR) is proposed to underlie the pharmacologically defined κ1 subtype and explain region-specific effects likeanalgesia ordysphoria.[202][203] Besides KOR-DOR the receptor heterodimerizes withμ-opioid (preferentially forms in females),[204]nociceptin (NOP),[205]orexin receptor 1 (OX1),[206]dopamine transporter (DAT),[53]neurotensin 1,[202]bradykinin B2,[207]beta-2 adrenergic receptors.[208] With others possible but not yet definitely established.[209]

Signaling dynamics

[edit]

Transduction

[edit]

Upon activation by dynorphin, KORs bind topertussis toxin-sensitive heterotrimericGαi proteins, initiating a pattern of signaling events within the cell, including inhibition ofadenylate cyclase activity, increase inK+ conductance, decrease incalcium conductance, emptying of intracellular calcium storage. KOR activation is coupled to theG proteinGi/G0, which subsequently increasesphosphodiesterase activity. Phosphodiesterases break downcAMP, producing an inhibitory effect in neurons.[210][211][212] KORs also couple toinward-rectifier potassium[213] and toN-type calcium ion channels.[214] Studies have also shown that agonist-induced stimulation of the KOR, like otherG-protein coupled receptors, can result in the activation ofmitogen-activated protein kinases (MAPK). These includeextracellular signal-regulated kinase (ERK),p38 mitogen-activated protein kinases, andc-Jun N-terminal kinases.[70][145][215][216][217][218]

Interactions

[edit]

KOR has been shown tointeract withsodium-hydrogen antiporter 3 regulator 1,[219][220]ubiquitin C,[221]5-HT1A receptor,[222] andRGS12.[223]

G protein and β-arrestin pathways

[edit]

KOR activation initiates bothG protein-mediated andβ-arrestin-dependent signaling pathways.[224] Following agonist binding, activated Gαi subunits inhibit adenylyl cyclase activity, whilstGβγ dimers activateG protein-coupled inwardly rectifying potassium channels (GIRKs) and inhibitcalcium channels.[225] G protein signaling also initiates early-phasephosphorylation of ERK through Gβγ-mediated activation ofphosphoinositide 3-kinase (PI3Ks).[225]

After G protein activation,G protein-coupled receptor kinases (GRKs) phosphorylate the receptor, which promotes recruitment ofβ-arrestins. Their recruitment mediates receptordesensitization,internalization, anddownregulation, whilst also initiating distinct signaling cascades independent of G protein activation.[61]β-arrestin2 is the dominant isoform mediating KOR desensitization,β-arrestin1 recruitment to KOR is possible but appears weaker and less functionally significant.[69] β-Arrestin2-mediated signaling includes late-phase ERK phosphorylation and activation ofp38 MAPK andc-Jun N-terminal kinase (JNK).[224][225]

Repeated stress produces dynorphin-dependent activation of both KOR and p38 MAPK withinGABAergic neurons localized to the nucleus accumbens, prefrontal cortex, and hippocampus.[111] This p38 activation is dependent uponG protein-coupled receptor kinase 3 (GRK3) andβ-arrestin2 recruitment and occurs through Ser369 phosphorylation of KOR itself. Inhibition of p38 MAPK selectively blocks stress-induced immobility and conditioned place aversion while preserving analgesia and non-selective learning processes, isolating p38 signaling as specifically responsible for dysphoric-like behavioral responses.[111]

Evidence suggests that G protein signaling mediates the therapeuticanalgesic andantipruritic effects of KOR agonists, whilst β-arrestin2-dependent signaling through p38 MAPK activation mediates adversedysphoric,sedative, andaversive effects.[59][111] Experiments in β-arrestin2knockout mice demonstrated that theantipruritic effects of KOR agonists are preserved in the absence of β-arrestin2, whilstconditioned place aversion requires both GRK3 and β-arrestin2.[56][226]

In cells coexpressingorexin receptor 1 (OX1) and KOR, OX1 activation attenuates KOR-mediated Gαi inhibition of cAMP but increases β-arrestin2 recruitment and p38 activation via a JNK-dependent pathway, shifting KOR signaling toward non-G protein pathways. This unidirectional crosstalk promotes preferential β-arrestin/p38 signaling over Gαi, without affecting KOR ligand binding or OX1 Gαq coupling.[227]

mTOR pathway

[edit]

U50,488, but notnalfurafine, activates themammalian target of rapamycin (mTOR) pathway in thestriatum andcortex following administration. Inhibition of the mTOR pathway withrapamycin abolished U50,488-induced aversion in theconditioned place preference test without affecting analgesic, antipruritic, sedative, ormotor incoordination effects.[228]

Other study have implicatedprotein kinase C (PKC) in regulating behavioral responses and signaling pathways.[229] PKC inhibition maintains the analgesic and antipruritic properties of KOR agonists whilst reducing adverse effects including conditioned place aversion,anxiogenesis, and motor incoordination. At 5 minutes following KOR activation, PKC regulates GRK5/6 andWnt signaling pathways, whilst at 30 minutes PKC influences mTOR pathways andcannabinoid receptor 1.[229]

Conformational states

[edit]

Simulations identified three distinct active-stateconformational states of KOR: the canonical active state, an alternative state, and an occluded state.[69] The alternative state, characterized by specifictransmembrane domain conformations, correlates withβ-arrestin2-biased signaling. The occluded state, in which the intracellular portion oftransmembrane helix 7 rotates clockwise towardtransmembrane helix 2, appears to favorG protein coupling whilst disfavoring β-arrestin recruitment.[69]

Specific residues within the receptorbinding pocket differentially influence G protein versus β-arrestin signaling. Disruption of the ionic interaction by certain agonists increases the distance between the extracellular ends of transmembrane helices 5 and 6, contributing to ligand-specifictransducer coupling preferences.[69]

Signaling after internalisation

[edit]

KOR undergoes agonist-mediated GRK-dependent phosphorylation followed by β-arrestin recruitment, initiatingclathrin-mediatedendocytosis.[230] KOR trafficking differs compared to otheropioid receptors. Whereas theμ-opioid receptor (MOR) contains aC-terminal LENL recycling motif that engagesretromer complexes for rapid plasma membrane recycling, and theδ-opioid receptor (DOR) undergoes predominantlylysosomal degradation followinginternalization, KOR requires aPDZ domain-binding sequence for post-endocytic sorting.[231] Following internalization, KOR rapidly accumulates in earlyendosomes, where it remains partially dissociated from β-arrestin, allowing continued G protein coupling and signaling in compartment-specific contexts.[231]

KOR-mediated signaling persists within late endosomes and lysosomes despite agonist-induced translocation from the plasma membrane, representing a form of sustained "post-internalization" signaling distinct from plasma membrane coupling.[232]Dynorphin A maintains prolonged adenylyl cyclase suppression when KOR is sequestered within late endosomal and lysosomal compartments, which suggests that dynorphin isoforms differentially stabilize intracellular receptor conformations suited to late-compartment signaling.[230] This property distinguishes KOR from MOR, which primarily signals from endosomal compartments when β-arrestin-bound, and from classical recycling receptors that rapidly regain surface expression. The intracellular KOR signaling axis involves continued Gi/o coupling on late endosomal membranes, sustained suppression of adenylyl cyclase and cAMP production, and prolonged recruitment ofERK pathway components through Gβγ-dependent mechanisms, thereby establishing abiochemical niche for chronic dynorphin signaling distinct from acute plasma membrane responses.[232] This signaling permits differential integration of intracellular second messenger systems and transcriptional responses compared to plasma membrane-restricted coupling.[233]

Clinical significance

[edit]

Pain

[edit]

KORagonists have been clinically employed asanalgesics, with examples includingbutorphanol,nalbuphine,levorphanol,levallorphan,pentazocine,phenazocine, andeptazocine.[234] Unlike MOR agonists, KOR agonists do not causerespiratory depression and have lowerabuse potential, butcentrally-mediated side effects such asdysphoria,hallucinations, anddissociation have limited their clinical utility.[234]

Nalorphine andnalmefene are dual MORantagonists and KOR agonists used clinically asantidotes for opioidoverdose, but the specific role of KOR activation to their efficacy remains uncertain as KOR agonists do not reverserespiratory depression induced by MOR activation and thus cannot serve as standalone antidotes for this purpose.[235]

Peripherally selective KOR agonists display analgesic efficacy mediated throughanti-inflammatory effects onimmune cells andnociceptors.[236]CR665 anddifelikefalin (CR845, FE-202845) have been investigated clinically; marking the first peripherally-restricted KOR agonist to reach regulatory approval, though none have yet been approved specifically for pain indication.[237][238][239] Recent evidence supports the therapeutic potential of mixed KOR/MOR agonists and KOR-biased ligands as adjuncts to conventional analgesics in inflammatory andcancer pain, with particular promise for chronicneuropathic pain syndromes.[234]

Major depressive disorder

[edit]

The mechanistic rationale for KORantagonism inmajor depressive disorder (MDD) derives from the observation thatchronic stress anddepression are associated with higher activity of the KOR system.[18][240] KOR activation suppressesdopamine release and prevents dopamine rebound after stress exposure, thereby leading toanhedonia and depressive phenotypes. KOR antagonists reverse this pathway by disinhibiting dopaminergic tone and restoring reward sensitivity.[13][72][132][241] KOR-mediated upregulation ofpro-inflammatory signaling inmicroglia likely drives the depression pathophysiology, and antagonism may provide benefits.[241]

Buprenorphine/samidorphan (ALKS-5461) displayed antidepressant efficacy inrandomized controlled trials as an adjunctive therapy and has shown durable effects with a favorable safety profile including low abuse potential and minimalwithdrawal symptoms.[242][243][244] Aphase 2 study demonstrated significant reduction of depressive symptoms and improvement inanhedonia whenaticaprant was added to existing antidepressant therapy.[245] A phase 3clinical trials (KOASTAL-1 and additional studies) ofnavacaprant failed to achievestatistically significant superiority overplacebo across the broader MDD population, with its development discontinued for this purpose in early 2025.[246]

Main sections: §§ Signaling after internalisation​ andConformational states

The persistent signaling that is present after internalisation could be the reason for ineffectivity of common KOR antagonists given that they work on receptor's outer membrane.[232] Antagonists also tend to preserves conformation states; in this case presumably "alternative state" due to the endogenous bias for theβ-arrestin signaling during stress responses.[69]

Anxiety disorders

[edit]

KOR antagonists have demonstratedanxiolytic efficacy inpreclinical stress models and early clinical evaluation. Early generation antagonists such asJDTic andnor-BNI produced anxiolytic-like effects inGAD,PTSD, andpanic disorder models. However their long duration of action and off-target toxicities limited clinical development.[164][242][247] Contemporary short-acting antagonists such as aticaprant are being evaluated for anxiety indications given their improved pharmacokinetic profiles and reduced toxicity burden.[245]

Schizophrenia

[edit]

Persistent KOR signaling has been implicated in thepathophysiology ofschizophrenia, in the generation of both positive and negative symptoms, and as an explanation for treatment-resistantpsychosis.[15] Mechanistically, chronic KOR activation produces long-term sensitization ofdopamine D2 receptors in thestriatum, which manifests as supersensitized D2 receptor states that amplify phasic dopamine signaling and hyperresponsivity to dopaminergic stimuli.[248][249] This mechanism could interact with the underlying excessivestriatal dopamine transmission in schizophrenia, potentiating positive symptoms includingdelusions andhallucinations, and explaining whydopamine D2 receptor antagonists (antipsychotics) remain effective. Apart from striatal mechanisms, KOR signaling modulates corticalglutamate andGABAhomeostasis through KOR activation onGABAergic terminals ofdynorphin-expressing neurons inprefrontal cortex which suppresses GABA release and disrupting the balance of cortical inhibition-excitation that might drivecognitive dysfunction and negative symptoms.[249] Thus, KOR antagonism may provide a complementary strategy to D2 antagonism by simultaneously reducing D2 receptor sensitization to normalize striatal dopamine responsivity and restoring cortical inhibition-excitation balance to ameliorate cognitive dysfunction.[15][249]

Borderline personality disorder

[edit]

Main section:§ Borderline personality disorder

Recent epigenetic findings suggest that KOR antagonists, which block the hyperactive KOR system, might be a viable pharmacological approach forborderline personality disorder (BPD) treatment, particularly for anhedonia, suicidality, and dissociative symptoms. Current early evidence supports the efficacy ofnaltrexone andnalmefene in reducing suicidal ideation, non-suicidalself-injury,binge eating, and dissociation in patients with BPD.[17][194]

Bipolar disorder

[edit]

In a small clinical study, pentazocine, a KOR agonist, rapidly reduced acutemanic symptoms inbipolar disorder patients.[18] The therapeutic mechanism is postulated to involve KOR agonist-mediated suppression of excessivedopaminergic signaling inreward pathways and striatal circuits that drive manic hyperactivity and impulsivity.[18] Complete desensitization of KOR renders the receptor unable to gate dopaminergic signaling, thereby lifting the inhibitory constraint and disinhibiting phasic dopamine andnorepinephrine release. Temporary KOR sensitization during acute mania may reverse this disinhibition.[232]

Addiction and withdrawal

[edit]

Aticaprant was well-tolerated incocaine use disorder (CUD) patients.[250] Apositron emission tomography (PET) study in CUD patients utilizing a KOR selective agonist [11C]GR-103545radioligand showed CUD individuals with higher KOR availability were more prone to stress-induced relapse.[251] A subsequent PET scan following a three-day cocaine binge showed a decrease in KOR availability, interpreted as increased endogenous dynorphin competing with the radioligand at the KORbinding sites.[251] These findings are in support of the negative affect state and further implicate the KOR/dynorphin system clinically and therapeutically relevant in humans with CUD. Taken together, in drug addiction the KOR system is implicated as a homeostatic mechanism to counteract the acute effects of drugs of abuse. Chronic drug use and stress up-regulate the system in turn leading to a dysregulated state which induces negative affective states and stress reactivity.[102]

KOR agonists have also been investigated for their therapeutic potential in the treatment ofaddiction, particularlysubstance use disorders.Ibogaine, atypical KOR agonist withG-protein-biased signaling and complex pharmacodynamics involving multiple neurotransmitter systems.[252] Ibogaine's primary active metabolite,noribogaine, acts as a moderate KOR agonist selective for G protein and a potentserotonin reuptake inhibitor.[252] This mechanism, combined with activity at5-HT2A,5-HT2C,σ2, andNMDA receptors, likely leads its anti-addictive effects.[253] The precise extent to which KOR agonism underlies ibogaine's anti-addictive properties is unclear.[252]

In animal models, ibogaine administration has been shown to reduceself-administration ofopioids,stimulants, andalcohol, amelioratewithdrawal symptoms, and decrease drug-seeking behavior.[254][255] A 2022systematic review of 24 studies involving 705 participants found that both ibogaine and noribogaine show promise in treating substance use disorders and comorbid depressive symptoms.[252]

Pruritis

[edit]

KOR agonists suppressitching, and the selective KOR agonistnalfurafine is used clinically as anantipruritic.[166][167] Peripheral agonistdifelikefalin also have been approved in the US and Europe for moderate-to-severe pruritis.[256][257]

In a mouse model, agonism of inhibitory, GABAergic KOR-containing neurons in therostral ventromedial medulla activates a top-down mechanism of inhibiting pain and itch perception from the spinal cord simultaneously.[258]

Gut

[edit]

Eluxadoline is a peripherally restricted KOR agonist as well as MOR agonist and DOR antagonist that has been approved for the treatment ofdiarrhea-predominantirritable bowel syndrome.Asimadoline andfedotozine are selective and similarly peripherally restricted KOR agonists that were also investigated for the treatment of irritable bowel syndrome and reportedly demonstrated at least some efficacy for this indication but were ultimately never marketed.[23]

Heart

[edit]

Cardiomyopathichamsterhearts show augmented negative inotropic responses to KOR agonists inheart failure, mediated through increased inhibition ofcAMP accumulation and decreased amplitude of the systolicCa2+.[44] KOR activation in rat models ofmyocardial infarction during reperfusion reduces infarct size throughERK1/2-dependent butPI3K-AKT-independent pathways, suggestingcardioprotective effects.[163]

See also

[edit]

References

[edit]
  1. ^abcGRCh38: Ensembl release 89: ENSG00000082556Ensembl, May 2017
  2. ^abcGRCm38: Ensembl release 89: ENSMUSG00000025905Ensembl, May 2017
  3. ^"Human PubMed Reference:".National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. ^"Mouse PubMed Reference:".National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. ^Martin WR, Eades CG, Thompson JA, Huppler RE, Gilbert PE (June 1976). "The effects of morphine- and nalorphine- like drugs in the nondependent and morphine-dependent chronic spinal dog".The Journal of Pharmacology and Experimental Therapeutics.197 (3):517–532.doi:10.1016/s0022-3565(25)30536-7.ISSN 0022-3565.
  6. ^abcdeLobe MM, Verma S, Patil VM, Iyer MR (April 2025). "A review of kappa opioid receptor antagonists and their clinical trial landscape".European Journal of Medicinal Chemistry.287 117205.doi:10.1016/j.ejmech.2024.117205.PMID 39893986.
  7. ^James IF, Chavkin C, Goldstein A (1982). "Selectivity of dynorphin for kappa opioid receptors".Life Sciences.31 (12–13):1331–1334.doi:10.1016/0024-3205(82)90374-5.PMID 6128656.
  8. ^Merg F, Filliol D, Usynin I, Bazov I, Bark N, Hurd YL, et al. (April 2006)."Big dynorphin as a putative endogenous ligand for the κ-opioid receptor".Journal of Neurochemistry.97 (1):292–301.doi:10.1111/j.1471-4159.2006.03732.x.ISSN 0022-3042.
  9. ^Li X, Winters ND, Pandey S, Lankford C, Stoveken HM, Smith E, et al. (July 2025)."Homeostatic scaling of dynorphin signaling by a non-canonical opioid receptor".Nature Communications.16 (1) 6786.Bibcode:2025NatCo..16.6786L.doi:10.1038/s41467-025-62133-x.PMC 12287315.PMID 40701991.
  10. ^abMiyano K, Yoshida Y, Hirayama S, Takahashi H, Ono H, Meguro Y, et al. (October 2021)."Oxytocin Is a Positive Allosteric Modulator of κ-Opioid Receptors but Not δ-Opioid Receptors in the G Protein Signaling Pathway".Cells.10 (10): 2651.doi:10.3390/cells10102651.PMC 8534029.PMID 34685631.
  11. ^Muratspahić E, Tomašević N, Nasrollahi-Shirazi S, Gattringer J, Emser FS, Freissmuth M, et al. (August 2021)."Plant-Derived Cyclotides Modulate κ-Opioid Receptor Signaling".Journal of Natural Products.84 (8):2238–2248.Bibcode:2021JNAtP..84.2238M.doi:10.1021/acs.jnatprod.1c00301.PMC 8406418.PMID 34308635.
  12. ^abcdefghLand BB, Bruchas MR, Lemos JC, Xu M, Melief EJ, Chavkin C (January 2008)."The dysphoric component of stress is encoded by activation of the dynorphin kappa-opioid system".The Journal of Neuroscience.28 (2):407–414.doi:10.1523/JNEUROSCI.4458-07.2008.PMC 2612708.PMID 18184783.
  13. ^abcdeKhan MI, Sawyer BJ, Akins NS, Le HV (December 2022). "A systematic review on the kappa opioid receptor and its ligands: New directions for the treatment of pain, anxiety, depression, and drug abuse".European Journal of Medicinal Chemistry.243 114785.doi:10.1016/j.ejmech.2022.114785.PMID 36179400.
  14. ^abPizzagalli DA, Smoski M, Ang YS, Whitton AE, Sanacora G, Mathew SJ, et al. (September 2020)."Selective kappa-opioid antagonism ameliorates anhedonic behavior: evidence from the Fast-fail Trial in Mood and Anxiety Spectrum Disorders (FAST-MAS)".Neuropsychopharmacology.45 (10):1656–1663.doi:10.1038/s41386-020-0738-4.PMC 7419512.PMID 32544925.
  15. ^abcdeClark SD, Abi-Dargham A (October 2019). "The Role of Dynorphin and the Kappa Opioid Receptor in the Symptomatology of Schizophrenia: A Review of the Evidence".Biological Psychiatry.86 (7):502–511.doi:10.1016/j.biopsych.2019.05.012.PMID 31376930.
  16. ^Clark SD (2022)."The Role of Dynorphin and the Kappa Opioid Receptor in Schizophrenia and Major Depressive Disorder: A Translational Approach".Handbook of Experimental Pharmacology.271:525–546.doi:10.1007/164_2020_396.ISBN 978-3-030-89073-5.PMID 33459877.
  17. ^abcdefgGescher DM, Schanze D, Vavra P, Wolff P, Zimmer-Bensch G, Zenker M, et al. (December 2024)."Differential methylation of OPRK1 in borderline personality disorder is associated with childhood trauma".Molecular Psychiatry.29 (12):3734–3741.doi:10.1038/s41380-024-02628-z.PMC 11609100.PMID 38862675.
  18. ^abcdeRasakham K, Liu-Chen LY (January 2011)."Sex differences in kappa opioid pharmacology".Life Sciences.88 (1–2):2–16.doi:10.1016/j.lfs.2010.10.007.PMC 3870184.PMID 20951148.
  19. ^abAnderson RI, Becker HC (August 2017)."Role of the Dynorphin/Kappa Opioid Receptor System in the Motivational Effects of Ethanol".Alcoholism, Clinical and Experimental Research.41 (8):1402–1418.doi:10.1111/acer.13406.PMC 5522623.PMID 28425121.
  20. ^abcdeKarkhanis A, Holleran KM, Jones SR (2017). "Dynorphin/Kappa Opioid Receptor Signaling in Preclinical Models of Alcohol, Drug, and Food Addiction".International Review of Neurobiology.136:53–88.doi:10.1016/bs.irn.2017.08.001.ISBN 978-0-12-812473-4.PMID 29056156.
  21. ^Mansour A, Fox CA, Akil H, Watson SJ (January 1995). "Opioid-receptor mRNA expression in the rat CNS: anatomical and functional implications".Trends in Neurosciences.18 (1):22–29.doi:10.1016/0166-2236(95)93946-u.ISSN 0166-2236.
  22. ^abcWang YH, Sun JF, Tao YM, Chi ZQ, Liu JG (September 2010)."The role of kappa-opioid receptor activation in mediating antinociception and addiction".Acta Pharmacologica Sinica.31 (9):1065–1070.doi:10.1038/aps.2010.138.PMC 4002313.PMID 20729876.
  23. ^abcdSzczepaniak A, Machelak W, Fichna J, Zielińska M (October 2022). "The role of kappa opioid receptors in immune system - An overview".European Journal of Pharmacology.933 175214.doi:10.1016/j.ejphar.2022.175214.PMID 36007608.
  24. ^abcStiefel KM, Merrifield A, Holcombe AO (2014)."The claustrum's proposed role in consciousness is supported by the effect and target localization of Salvia divinorum".Frontiers in Integrative Neuroscience.8: 20.doi:10.3389/fnint.2014.00020.PMC 3935397.PMID 24624064.
  25. ^Mantas I, Flais I, Masarapu Y, Ionescu T, Frapard S, Jung F, et al. (September 2024)."Claustrum and dorsal endopiriform cortex complex cell-identity is determined by Nurr1 and regulates hallucinogenic-like states in mice".Nature Communications.15 (1) 8176.Bibcode:2024NatCo..15.8176M.doi:10.1038/s41467-024-52429-9.PMC 11408527.PMID 39289358.
  26. ^abcdefChau A, Salazar AM, Krueger F, Cristofori I, Grafman J (November 2015). "The effect of claustrum lesions on human consciousness and recovery of function".Consciousness and Cognition.36:256–264.doi:10.1016/j.concog.2015.06.017.PMID 26186439.S2CID 46139982.
  27. ^Wang YJ, Zan GY, Xu C, Li XP, Shu X, Yao SY, et al. (November 2023)."The claustrum-prelimbic cortex circuit through dynorphin/κ-opioid receptor signaling underlies depression-like behaviors associated with social stress etiology".Nature Communications.14 (1) 7903.Bibcode:2023NatCo..14.7903W.doi:10.1038/s41467-023-43636-x.PMC 10689794.PMID 38036497.
  28. ^Tempel A, Zukin RS (June 1987)."Neuroanatomical patterns of the mu, delta, and kappa opioid receptors of rat brain as determined by quantitative in vitro autoradiography".Proceedings of the National Academy of Sciences of the United States of America.84 (12):4308–4312.Bibcode:1987PNAS...84.4308T.doi:10.1073/pnas.84.12.4308.PMC 305074.PMID 3035579.
  29. ^abMartin TJ, Sexton T, Kim SA, Severino AL, Peters CM, Young LJ, et al. (December 2015)."Regional differences in mu and kappa opioid receptor G-protein activation in brain in male and female prairie voles".Neuroscience.311:422–429.doi:10.1016/j.neuroscience.2015.10.047.PMC 4663147.PMID 26523979.
  30. ^abJohnson BN, Kumar A, Su Y, Singh S, Sai KK, Nader SH, et al. (January 2023)."PET imaging of kappa opioid receptors and receptor expression quantified in neuron-derived extracellular vesicles in socially housed female and male cynomolgus macaques".Neuropsychopharmacology.48 (2):410–417.doi:10.1038/s41386-022-01444-9.PMC 9751296.PMID 36100655.
  31. ^Almeida AJ, Hobson BA, Savidge LE, Manca C, Paulus JP, Bales KL, et al. (August 2025)."Mapping Kappa Opioid Receptor Binding in Titi Monkeys with [11C]GR103545 PET".Molecular Imaging.24 15353508251341082.doi:10.1177/15353508251341082.ISSN 1535-3508.
  32. ^Vonkeman HE, Voorn P, Brady LS, Berendse HW, Richfield EK (October 1996). "Opioid receptor ligand binding in the human striatum: II. Heterogeneous distribution of kappa opioid receptor labeled with [3H]bremazocine".The Journal of Comparative Neurology.374 (2):223–229.doi:10.1002/(SICI)1096-9861(19961014)374:2<223::AID-CNE5>3.0.CO;2-4.PMID 8906495.
  33. ^Chartoff EH, Mavrikaki M (2015)."Sex Differences in Kappa Opioid Receptor Function and Their Potential Impact on Addiction".Frontiers in Neuroscience.9: 466.doi:10.3389/fnins.2015.00466.PMC 4679873.PMID 26733781.
  34. ^Siciliano CA, Calipari ES, Yorgason JT, Lovinger DM, Mateo Y, Jimenez VA, et al. (April 2016)."Increased presynaptic regulation of dopamine neurotransmission in the nucleus accumbens core following chronic ethanol self-administration in female macaques".Psychopharmacology.233 (8):1435–1443.doi:10.1007/s00213-016-4239-4.PMC 4814331.PMID 26892380.
  35. ^abcdFaull R, Villiger J (February 1987)."Opiate receptors in the human spinal cord: A detailed anatomical study comparing the autoradiographic localization of [3H]diprenorphine binding sites with the laminar pattern of substance P, myelin and nissl staining".Neuroscience.20 (2):395–407.doi:10.1016/0306-4522(87)90100-X.ISSN 0306-4522.
  36. ^Gouardères C, Cros J, Quirion R (July 1985). "Autoradiographic localization of mu, delta and kappa opioid receptor binding sites in rat and guinea pig spinal cord".Neuropeptides.6 (4):331–342.doi:10.1016/0143-4179(85)90006-X.
  37. ^Besse D, Lombard MC, Zajac JM, Roques BP, Besson JM (June 1990). "Pre- and postsynaptic distribution of mu, delta and kappa opioid receptors in the superficial layers of the cervical dorsal horn of the rat spinal cord".Brain Research.521 (1–2):15–22.doi:10.1016/0006-8993(90)91519-M.PMID 2169958.
  38. ^Peng J, Sarkar S, Chang SL (August 2012)."Opioid receptor expression in human brain and peripheral tissues using absolute quantitative real-time RT-PCR".Drug and Alcohol Dependence.124 (3):223–228.doi:10.1016/j.drugalcdep.2012.01.013.PMC 3366045.PMID 22356890.
  39. ^abcSnyder LM, Chiang MC, Loeza-Alcocer E, Omori Y, Hachisuka J, Sheahan TD, et al. (September 2018)."Kappa Opioid Receptor Distribution and Function in Primary Afferents".Neuron.99 (6): 1274–1288.e6.doi:10.1016/j.neuron.2018.08.044.PMC 6300132.PMID 30236284.
  40. ^Ignatowski TA, Bidlack JM (August 1999). "Differential kappa-opioid receptor expression on mouse lymphocytes at varying stages of maturation and on mouse macrophages after selective elicitation".The Journal of Pharmacology and Experimental Therapeutics.290 (2):863–870.doi:10.1016/S0022-3565(24)34975-4.PMID 10411603.
  41. ^Bidlack JM (September 2000)."Detection and function of opioid receptors on cells from the immune system".Clinical and Diagnostic Laboratory Immunology.7 (5):719–723.doi:10.1128/cdli.7.5.719-723.2000.PMC 95944.PMID 10973443.
  42. ^abcdFickel J, Bagnol D, Watson SJ, Akil H (June 1997). "Opioid receptor expression in the rat gastrointestinal tract: a quantitative study with comparison to the brain".Brain Research. Molecular Brain Research.46 (1–2):1–8.doi:10.1016/s0169-328x(96)00266-5.PMID 9191072.
  43. ^abBagnol D, Mansour A, Akil H, Watson SJ (November 1997). "Cellular localization and distribution of the cloned mu and kappa opioid receptors in rat gastrointestinal tract".Neuroscience.81 (2):579–591.doi:10.1016/s0306-4522(97)00227-3.PMID 9300443.
  44. ^abcdBolte C, Newman G, Schultz JE (October 2009)."Kappa and delta opioid receptor signaling is augmented in the failing heart".Journal of Molecular and Cellular Cardiology.47 (4):493–503.doi:10.1016/j.yjmcc.2009.06.016.PMC 2829247.PMID 19573531.
  45. ^Headrick JP, See Hoe LE, Du Toit EF, Peart JN (April 2015)."Opioid receptors and cardioprotection - 'opioidergic conditioning' of the heart".British Journal of Pharmacology.172 (8):2026–2050.doi:10.1111/bph.13042.PMC 4386979.PMID 25521834.
  46. ^Katari V, Dalal KK, Kondapalli N, Paruchuri S, Thodeti CK (August 2025)."Opioid receptors in cardiovascular function".British Journal of Pharmacology.182 (16):3710–3725.doi:10.1111/bph.70097.PMID 40456685.
  47. ^abZimlichman R, Gefel D, Eliahou H, Matas Z, Rosen B, Gass S, et al. (March 1996). "Expression of opioid receptors during heart ontogeny in normotensive and hypertensive rats".Circulation.93 (5):1020–1025.doi:10.1161/01.CIR.93.5.1020.
  48. ^abDidik S, Golosova D, Zietara A, Bohovyk R, Ahrari A, Levchenko V, et al. (August 2025)."Renal Implications of Kappa Opioid Receptor Signaling in Sprague-Dawley Rats".Function.6 (4) zqaf028.doi:10.1093/function/zqaf028.PMC 12316101.PMID 40586679.
  49. ^de Costa BR, Rothman RB, Bykov V, Jacobson AE, Rice KC (February 1989). "Selective and enantiospecific acylation of kappa opioid receptors by (1S,2S)-trans-2-isothiocyanato-N-methyl-N-[2-(1-pyrrolidinyl) cyclohexy l] benzeneacetamide. Demonstration of kappa receptor heterogeneity".Journal of Medicinal Chemistry.32 (2):281–283.doi:10.1021/jm00122a001.PMID 2536435.
  50. ^Rothman RB, France CP, Bykov V, De Costa BR, Jacobson AE, Woods JH, et al. (August 1989). "Pharmacological activities of optically pure enantiomers of the kappa opioid agonist, U50,488, and its cis diastereomer: evidence for three kappa receptor subtypes".European Journal of Pharmacology.167 (3):345–353.doi:10.1016/0014-2999(89)90443-3.hdl:2027.42/27799.PMID 2553442.
  51. ^Mansson E, Bare L, Yang D (August 1994)."Isolation of a Human κ Opioid Receptor cDNA from Placenta".Biochemical and Biophysical Research Communications.202 (3):1431–1437.Bibcode:1994BBRC..202.1431M.doi:10.1006/bbrc.1994.2091.ISSN 0006-291X.
  52. ^abJordan BA, Devi LA (June 1999)."G-protein-coupled receptor heterodimerization modulates receptor function".Nature.399 (6737):697–700.Bibcode:1999Natur.399..697J.doi:10.1038/21441.PMC 3125690.PMID 10385123.
  53. ^abcdFarahbakhsh ZZ, Siciliano CA (October 2025)."Kappa opioid receptor control of motivated behavior revisited".Neuropsychopharmacology.doi:10.1038/s41386-025-02226-9.PMID 41083597.
  54. ^Pan ZZ (March 1998). "mu-Opposing actions of the kappa-opioid receptor".Trends in Pharmacological Sciences.19 (3):94–98.doi:10.1016/S0165-6147(98)01169-9.PMID 9584625.
  55. ^Xuei X, Dick D, Flury-Wetherill L, Tian HJ, Agrawal A, Bierut L, et al. (November 2006)."Association of the kappa-opioid system with alcohol dependence".Molecular Psychiatry.11 (11):1016–1024.doi:10.1038/sj.mp.4001882.PMID 16924269.
  56. ^abcHuang CH, Chiang YW, Liang KC, Thompson RF, Liu IY (April 2010). "Extra-cellular signal-regulated kinase 1/2 (ERK1/2) activated in the hippocampal CA1 neurons is critical for retrieval of auditory trace fear memory".Brain Research.1326:143–151.doi:10.1016/j.brainres.2010.02.033.PMID 20188711.
  57. ^Chartoff EH, Connery HS (May 2014)."It's MORe exciting than mu: crosstalk between mu opioid receptors and glutamatergic transmission in the mesolimbic dopamine system".Frontiers in Pharmacology.5: 116.doi:10.3389/fphar.2014.00116.PMC 4034717.PMID 24904419.
  58. ^Swingler M, Donadoni M, Unterwald EM, Maggirwar SB, Sariyer IK (June 2025)."Molecular and cellular basis of mu-opioid receptor signaling: mechanisms underlying tolerance and dependence development".Frontiers in Neuroscience.19 1597922.doi:10.3389/fnins.2025.1597922.PMC 12234547.PMID 40630856.
  59. ^abcdefCahill CM, Taylor AM, Cook C, Ong E, Morón JA, Evans CJ (November 2014)."Does the kappa opioid receptor system contribute to pain aversion?".Frontiers in Pharmacology.5: 253.doi:10.3389/fphar.2014.00253.PMC 4233910.PMID 25452729.
  60. ^Lutz PE, Kieffer BL (March 2013)."Opioid receptors: distinct roles in mood disorders".Trends in Neurosciences.36 (3):195–206.doi:10.1016/j.tins.2012.11.002.PMC 3594542.PMID 23219016.
  61. ^abBrust TF (2022). "Biased Ligands at the Kappa Opioid Receptor: Fine-Tuning Receptor Pharmacology".Handbook of Experimental Pharmacology.271:115–135.doi:10.1007/164_2020_395.ISBN 978-3-030-89073-5.PMID 33140224.
  62. ^Estave PM, Spodnick MB, Karkhanis AN (2022). "KOR Control over Addiction Processing: An Exploration of the Mesolimbic Dopamine Pathway". In Liu-Chen LY, Inan S (eds.).The Kappa Opioid Receptor. Handbook of Experimental Pharmacology. Vol. 271. Springer International Publishing. pp. 351–377.doi:10.1007/164_2020_421.ISBN 978-3-030-89074-2.PMC 8192597.PMID 33301050.
  63. ^Wang YJ, Hang A, Lu YC, Long Y, Zan GY, Li XP, et al. (November 2016). "κ Opioid receptor activation in different brain regions differentially modulates anxiety-related behaviors in mice".Neuropharmacology.110 (Pt A):92–101.doi:10.1016/j.neuropharm.2016.04.022.PMID 27106167.
  64. ^abcMaqueda AE, Valle M, Addy PH, Antonijoan RM, Puntes M, Coimbra J, et al. (June 2015)."Salvinorin-A Induces Intense Dissociative Effects, Blocking External Sensory Perception and Modulating Interoception and Sense of Body Ownership in Humans".The International Journal of Neuropsychopharmacology.18 (12) pyv065.doi:10.1093/ijnp/pyv065.PMC 4675976.PMID 26047623.
  65. ^abcdefDoss MK, May DG, Johnson MW, Clifton JM, Hedrick SL, Prisinzano TE, et al. (October 2020)."The Acute Effects of the Atypical Dissociative Hallucinogen Salvinorin A on Functional Connectivity in the Human Brain".Scientific Reports.10 (1) 16392.doi:10.1038/s41598-020-73216-8.PMC 7532139.PMID 33009457.
  66. ^Zürrer WE, Wettstein L, Aicher HD, Scheidegger M, Ineichen BV (October 2025)."The translational potential of salvinorin A: systematic review and meta-analysis of preclinical studies".Translational Psychiatry.15 (1) 401.doi:10.1038/s41398-025-03638-3.PMC 12514287.PMID 41073402.
  67. ^Valentino RJ, Volkow ND (December 2018)."Untangling the complexity of opioid receptor function".Neuropsychopharmacology.43 (13):2514–2520.doi:10.1038/s41386-018-0225-3.PMC 6224460.PMID 30250308.
  68. ^Bohn LM, Aubé J (July 2017)."Seeking (and Finding) Biased Ligands of the Kappa Opioid Receptor".ACS Medicinal Chemistry Letters.8 (7):694–700.doi:10.1021/acsmedchemlett.7b00224.PMC 5512133.PMID 28740600.
  69. ^abcdefghEl Daibani A, Paggi JM, Kim K, Laloudakis YD, Popov P, Bernhard SM, et al. (March 2023)."Molecular mechanism of biased signaling at the kappa opioid receptor".Nature Communications.14 (1) 1338.Bibcode:2023NatCo..14.1338E.doi:10.1038/s41467-023-37041-7.PMC 10008561.PMID 36906681.
  70. ^abBruchas MR, Macey TA, Lowe JD, Chavkin C (June 2006)."Kappa opioid receptor activation of p38 MAPK is GRK3- and arrestin-dependent in neurons and astrocytes".The Journal of Biological Chemistry.281 (26):18081–18089.doi:10.1074/jbc.M513640200.PMC 2096730.PMID 16648139.
  71. ^Hauser KF, Aldrich JV, Anderson KJ, Bakalkin G, Christie MJ, Hall ED, et al. (January 2005)."Pathobiology of dynorphins in trauma and disease".Frontiers in Bioscience.10 (1–3):216–235.doi:10.2741/1522.PMC 4304872.PMID 15574363.
  72. ^abHauger RL, Risbrough V, Oakley RH, Olivares-Reyes JA, Dautzenberg FM (October 2009)."Role of CRF receptor signaling in stress vulnerability, anxiety, and depression".Annals of the New York Academy of Sciences.1179 (1):120–143.Bibcode:2009NYASA1179..120H.doi:10.1111/j.1749-6632.2009.05011.x.PMC 3115623.PMID 19906236.
  73. ^Seeman P, Guan HC, Hirbec H (August 2009). "Dopamine D2High receptors stimulated by phencyclidines, lysergic acid diethylamide, salvinorin A, and modafinil".Synapse.63 (8):698–704.doi:10.1002/syn.20647.PMID 19391150.
  74. ^Zjawiony JK, Machado AS, Menegatti R, Ghedini PC, Costa EA, Pedrino GR, et al. (March 2019)."Cutting-Edge Search for Safer Opioid Pain Relief: Retrospective Review of Salvinorin A and Its Analogs".Frontiers in Psychiatry.10 157.doi:10.3389/fpsyt.2019.00157.PMC 6445891.PMID 30971961.
  75. ^White KL, Scopton AP, Rives ML, Bikbulatov RV, Polepally PR, Brown PJ, et al. (January 2014)."Identification of novel functionally selective κ-opioid receptor scaffolds".Molecular Pharmacology.85 (1):83–90.doi:10.1124/mol.113.089649.PMC 3868907.PMID 24113749.
  76. ^Paton KF, Biggerstaff A, Kaska S, Crowley RS, La Flamme AC, Prisinzano TE, et al. (2020)."Evaluation of Biased and Balanced Salvinorin A Analogs in Preclinical Models of Pain".Frontiers in Neuroscience.14 765.doi:10.3389/fnins.2020.00765.PMC 7385413.PMID 32792903.
  77. ^abBrust TF, Morgenweck J, Kim SA, Rose JH, Locke JL, Schmid CL, et al. (November 2016)."Biased agonists of the kappa opioid receptor suppress pain and itch without causing sedation or dysphoria".Science Signaling.9 (456): ra117.doi:10.1126/scisignal.aai8441.PMC 5231411.PMID 27899527.
  78. ^Attali B, Saya D, Nah SY, Vogel Z (January 1989)."Kappa opiate agonists inhibit Ca2+ influx in rat spinal cord-dorsal root ganglion cocultures. Involvement of a GTP-binding protein".The Journal of Biological Chemistry.264 (1):347–353.doi:10.1016/S0021-9258(17)31264-4.PMID 2535841.
  79. ^Maekawa K, Minami M, Masuda T, Satoh M (February 1996)."Expression of μ- and κ-, but not δ-, opioid receptor mRNAs is enhanced in the spinal dorsal horn of the arthritic rats".Pain.64 (2):365–371.doi:10.1016/0304-3959(95)00132-8.ISSN 0304-3959.
  80. ^abXu M, Petraschka M, McLaughlin JP, Westenbroek RE, Caron MG, Lefkowitz RJ, et al. (May 2004)."Neuropathic pain activates the endogenous kappa opioid system in mouse spinal cord and induces opioid receptor tolerance".The Journal of Neuroscience.24 (19):4576–4584.doi:10.1523/JNEUROSCI.5552-03.2004.PMC 2376823.PMID 15140929.
  81. ^Ilyutchenok RY, Dubrovina NI (December 1995). "Memory retrieval enhancement by kappa opioid agonist and mu, delta antagonists".Pharmacology, Biochemistry, and Behavior.52 (4):683–687.doi:10.1016/0091-3057(95)00099-I.PMID 8587905.
  82. ^Huge V, Rammes G, Beyer A, Zieglgänsberger W, Azad SC (February 2009). "Activation of kappa opioid receptors decreases synaptic transmission and inhibits long-term potentiation in the basolateral amygdala of the mouse".European Journal of Pain.13 (2):124–129.doi:10.1016/j.ejpain.2008.03.010.PMID 18439862.
  83. ^abWallace TL, Martin WJ, Arnsten AF (November 2022)."Kappa opioid receptor antagonism protects working memory performance from mild stress exposure in Rhesus macaques".Neurobiology of Stress.21 100493.doi:10.1016/j.ynstr.2022.100493.PMC 9755019.PMID 36532373.
  84. ^Bilkei-Gorzo A, Mauer D, Michel K, Zimmer A (February 2014). "Dynorphins regulate the strength of social memory".Neuropharmacology.77:406–413.doi:10.1016/j.neuropharm.2013.10.023.PMID 24184385.
  85. ^Ur E, Wright DM, Bouloux PM, Grossman A (March 1997)."The effects of spiradoline (U-62066E), a kappa-opioid receptor agonist, on neuroendocrine function in man".British Journal of Pharmacology.120 (5):781–784.doi:10.1038/sj.bjp.0700971.PMC 1564535.PMID 9138682.
  86. ^Butelman ER, Ball JW, Kreek MJ (October 2002). "Comparison of the discriminative and neuroendocrine effects of centrally penetrating kappa-opioid agonists in rhesus monkeys".Psychopharmacology.164 (1):115–120.doi:10.1007/s00213-002-1195-y.PMID 12373425.
  87. ^abButelman ER, Kreek MJ (July 2001). "kappa-Opioid receptor agonist-induced prolactin release in primates is blocked by dopamine D(2)-like receptor agonists".European Journal of Pharmacology.423 (2–3):243–249.doi:10.1016/S0014-2999(01)01121-9.PMID 11448491.
  88. ^Yamada K, Imai M, Yoshida S (January 1989). "Mechanism of diuretic action of U-62,066E, a kappa opioid receptor agonist".European Journal of Pharmacology.160 (2):229–237.doi:10.1016/0014-2999(89)90495-0.PMID 2547626.
  89. ^Inan S (2022). "Kappa Opioid Agonist-Induced Diuresis: Characteristics, Mechanisms, and Beyond".Handbook of Experimental Pharmacology.271:401–417.doi:10.1007/164_2020_399.ISBN 978-3-030-89073-5.PMID 33483878.
  90. ^Brooks DP, Valente M, Petrone G, Depalma PD, Sbacchi M, Clarke GD (March 1997). "Comparison of the water diuretic activity of kappa receptor agonists and a vasopressin receptor antagonist in dogs".The Journal of Pharmacology and Experimental Therapeutics.280 (3):1176–1183.doi:10.1016/S0022-3565(24)36542-5.PMID 9067301.
  91. ^abPascoe JE, Williams KL, Mukhopadhyay P, Rice KC, Woods JH, Ko MC (May 2008)."Effects of mu, kappa, and delta opioid receptor agonists on the function of hypothalamic-pituitary-adrenal axis in monkeys".Psychoneuroendocrinology.33 (4):478–486.doi:10.1016/j.psyneuen.2008.01.006.PMC 2443864.PMID 18325678.
  92. ^Shuster SJ, Riedl M, Li X, Vulchanova L, Elde R (February 2000). "The kappa opioid receptor and dynorphin co-localize in vasopressin magnocellular neurosecretory neurons in guinea-pig hypothalamus".Neuroscience.96 (2):373–383.doi:10.1016/S0306-4522(99)00472-8.PMID 10683577.
  93. ^Watson SJ, Akil H, Fischli W, Goldstein A, Zimmerman E, Nilaver G, et al. (April 1982). "Dynorphin and vasopressin: common localization in magnocellular neurons".Science.216 (4541):85–87.Bibcode:1982Sci...216...85W.doi:10.1126/science.6121376.PMID 6121376.
  94. ^Meister B, Villar MJ, Ceccatelli S, Hökfelt T (January 1990). "Localization of chemical messengers in magnocellular neurons of the hypothalamic supraoptic and paraventricular nuclei: an immunohistochemical study using experimental manipulations".Neuroscience.37 (3):603–633.doi:10.1016/0306-4522(90)90094-K.PMID 1701038.
  95. ^Zhao BG, Chapman C, Bicknell RJ (October 1988). "Functional κ-opioid receptors on oxytocin and vasopressin nerve terminals isolated from the rat neurohypophysis".Brain Research.462 (1):62–66.doi:10.1016/0006-8993(88)90585-9.ISSN 0006-8993.
  96. ^Falke N (May 1988). "Dynorphin (1-8) inhibits stimulated release of oxytocin but not vasopressin from isolated neurosecretory endings of the rat neurohypophysis".Neuropeptides.11 (4):163–167.doi:10.1016/0143-4179(88)90070-4.PMID 2901684.
  97. ^abBrown CH, Ludwig M, Tasker JG, Stern JE (June 2020)."Somato-dendritic vasopressin and oxytocin secretion in endocrine and autonomic regulation".Journal of Neuroendocrinology.32 (6) e12856.doi:10.1111/jne.12856.PMC 9134751.PMID 32406599.
  98. ^abcBales KL, Rogers FD (August 2022)."Interactions between theκ opioid system, corticotropin-releasing hormone and oxytocin in partner loss".Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences.377 (1858) 20210061.doi:10.1098/rstb.2021.0061.PMC 9272146.PMID 35858099.
  99. ^abJi MJ, Yang J, Gao ZQ, Zhang L, Liu C (February 2021)."The Role of the Kappa Opioid System in Comorbid Pain and Psychiatric Disorders: Function and Implications".Frontiers in Neuroscience.15 642493.doi:10.3389/fnins.2021.642493.PMC 7943636.PMID 33716658.
  100. ^abcdLimoges A, Yarur HE, Tejeda HA (October 2022)."Dynorphin/kappa opioid receptor system regulation on amygdaloid circuitry: Implications for neuropsychiatric disorders".Frontiers in Systems Neuroscience.16 963691.doi:10.3389/fnsys.2022.963691.PMC 9579273.PMID 36276608.
  101. ^abVan't Veer A, Carlezon WA (October 2013)."Role of kappa-opioid receptors in stress and anxiety-related behavior".Psychopharmacology.229 (3):435–452.doi:10.1007/s00213-013-3195-5.PMC 3770816.PMID 23836029.
  102. ^abcdeTejeda HA, Bonci A (June 2019). "Dynorphin/kappa-opioid receptor control of dopamine dynamics: Implications for negative affective states and psychiatric disorders".Brain Research.1713:91–101.doi:10.1016/j.brainres.2018.09.023.PMID 30244022.S2CID 52339964.
  103. ^Jacobson ML, Browne CA, Lucki I (January 2020)."Kappa Opioid Receptor Antagonists as Potential Therapeutics for Stress-Related Disorders".Annual Review of Pharmacology and Toxicology.60 (1):615–636.doi:10.1146/annurev-pharmtox-010919-023317.ISSN 0362-1642.
  104. ^Crowley NA, Bloodgood DW, Hardaway JA, Kendra AM, McCall JG, Al-Hasani R, et al. (March 2016)."Dynorphin Controls the Gain of an Amygdalar Anxiety Circuit".Cell Reports.14 (12):2774–2783.doi:10.1016/j.celrep.2016.02.069.PMC 4814306.PMID 26997280.
  105. ^abHatalski CG, Guirguis C, Baram TZ (September 1998)."Corticotropin releasing factor mRNA expression in the hypothalamic paraventricular nucleus and the central nucleus of the amygdala is modulated by repeated acute stress in the immature rat".Journal of Neuroendocrinology.10 (9):663–669.doi:10.1046/j.1365-2826.1998.00246.x.PMC 3382972.PMID 9744483.
  106. ^Xu X, Zheng S, Ren J, Li Z, Li J, Xu Z, et al. (January 2024)."Hypothalamic CRF neurons facilitate brain reward function".Current Biology.34 (2): 389–402.e5.Bibcode:2024CBio...34..389X.doi:10.1016/j.cub.2023.12.046.PMC 10842365.PMID 38215742.
  107. ^abKnoll AT, Carlezon WA (February 2010)."Dynorphin, stress, and depression".Brain Research.1314:56–73.doi:10.1016/j.brainres.2009.09.074.PMC 2819644.PMID 19782055.
  108. ^abcBruchas MR, Land BB, Lemos JC, Chavkin C (December 2009)."CRF1-R activation of the dynorphin/kappa opioid system in the mouse basolateral amygdala mediates anxiety-like behavior".PloS One.4 (12) e8528.Bibcode:2009PLoSO...4.8528B.doi:10.1371/journal.pone.0008528.PMC 2795205.PMID 20052275.
  109. ^Tafet GE, Ortiz Alonso T (2025)."Learned helplessness and learned controllability: from neurobiology to cognitive, emotional and behavioral neurosciences".Frontiers in Psychiatry.16 1600165.doi:10.3389/fpsyt.2025.1600165.PMC 12417506.PMID 40933805.
  110. ^Chen Y, Wang CY, Zan GY, Yao SY, Deng YZ, Shu XL, et al. (March 2023)."Upregulation of dynorphin/kappa opioid receptor system in the dorsal hippocampus contributes to morphine withdrawal-induced place aversion".Acta Pharmacologica Sinica.44 (3):538–545.doi:10.1038/s41401-022-00987-3.PMC 9958091.PMID 36127507.
  111. ^abcdeBruchas MR, Land BB, Aita M, Xu M, Barot SK, Li S, et al. (October 2007)."Stress-induced p38 mitogen-activated protein kinase activation mediates kappa-opioid-dependent dysphoria".The Journal of Neuroscience.27 (43):11614–11623.doi:10.1523/JNEUROSCI.3769-07.2007.PMC 2481272.PMID 17959804.
  112. ^abDonahue RJ, Landino SM, Golden SA, Carroll FI, Russo SJ, Carlezon WA (October 2015)."Effects of acute and chronic social defeat stress are differentially mediated by the dynorphin/kappa-opioid receptor system".Behavioural Pharmacology.26 (7 Spec No):654–663.doi:10.1097/FBP.0000000000000155.PMC 4586263.PMID 26110224.
  113. ^Lemos JC, Roth CA, Messinger DI, Gill HK, Phillips PE, Chavkin C (September 2012)."Repeated stress dysregulates κ-opioid receptor signaling in the dorsal raphe through a p38α MAPK-dependent mechanism".The Journal of Neuroscience.32 (36):12325–12336.doi:10.1523/JNEUROSCI.2053-12.2012.PMC 3477582.PMID 22956823.
  114. ^Tao R, Auerbach SB (July 2005). "mu-Opioids disinhibit and kappa-opioids inhibit serotonin efflux in the dorsal raphe nucleus".Brain Research.1049 (1):70–79.doi:10.1016/j.brainres.2005.04.076.PMID 15935332.
  115. ^Pomrenze MB, Cardozo Pinto DF, Neumann PA, Llorach P, Tucciarone JM, Morishita W, et al. (December 2022)."Modulation of 5-HT release by dynorphin mediates social deficits during opioid withdrawal".Neuron.110 (24): 4125–4143.e6.doi:10.1016/j.neuron.2022.09.024.PMC 9789200.PMID 36202097.
  116. ^abSundaramurthy S, Annamalai B, Samuvel DJ, Shippenberg TS, Jayanthi LD, Ramamoorthy S (February 2017)."Modulation of serotonin transporter function by kappa-opioid receptor ligands".Neuropharmacology.113 (Pt A):281–292.doi:10.1016/j.neuropharm.2016.10.011.PMC 5148672.PMID 27743931.
  117. ^West AM, Holleran KM, Jones SR (January 2023)."Kappa Opioid Receptors Reduce Serotonin Uptake and Escitalopram Efficacy in the Mouse Substantia Nigra Pars Reticulata".International Journal of Molecular Sciences.24 (3): 2080.doi:10.3390/ijms24032080.PMC 9916942.PMID 36768403.
  118. ^Lanfumey L, Pardon MC, Laaris N, Joubert C, Hanoun N, Hamon M, et al. (November 1999). "5-HT1A autoreceptor desensitization by chronic ultramild stress in mice".Neuroreport.10 (16):3369–3374.doi:10.1097/00001756-199911080-00021.PMID 10599847.
  119. ^Fontaine HM, Silva PR, Neiswanger C, Tran R, Abraham AD, Land BB, et al. (March 2022)."Stress decreases serotonin tone in the nucleus accumbens in male mice to promote aversion and potentiate cocaine preference via decreased stimulation of 5-HT1B receptors".Neuropsychopharmacology.47 (4):891–901.doi:10.1038/s41386-021-01178-0.PMC 8882182.PMID 34564712.
  120. ^Margolis EB, Karkhanis AN (October 2019)."Dopaminergic cellular and circuit contributions to kappa opioid receptor mediated aversion".Neurochemistry International.129 104504.doi:10.1016/j.neuint.2019.104504.PMC 6702044.PMID 31301327.
  121. ^Al-Hasani R, McCall JG, Shin G, Gomez AM, Schmitz GP, Bernardi JM, et al. (September 2015)."Distinct Subpopulations of Nucleus Accumbens Dynorphin Neurons Drive Aversion and Reward".Neuron.87 (5):1063–1077.doi:10.1016/j.neuron.2015.08.019.PMC 4625385.PMID 26335648.
  122. ^abPolter AM, Barcomb K, Chen RW, Dingess PM, Graziane NM, Brown TE, et al. (April 2017)."Constitutive activation of kappa opioid receptors at ventral tegmental area inhibitory synapses following acute stress".Elife.6 e23785.doi:10.7554/eLife.23785.PMC 5389861.PMID 28402252.
  123. ^Kivell B, Uzelac Z, Sundaramurthy S, Rajamanickam J, Ewald A, Chefer V, et al. (November 2014)."Salvinorin A regulates dopamine transporter function via a kappa opioid receptor and ERK1/2-dependent mechanism".Neuropharmacology.86:228–240.doi:10.1016/j.neuropharm.2014.07.016.PMC 4188751.PMID 25107591.
  124. ^Tejeda HA, Counotte DS, Oh E, Ramamoorthy S, Schultz-Kuszak KN, Bäckman CM, et al. (August 2013)."Prefrontal cortical kappa-opioid receptor modulation of local neurotransmission and conditioned place aversion".Neuropsychopharmacology.38 (9):1770–1779.doi:10.1038/npp.2013.76.PMC 3717537.PMID 23542927.
  125. ^Castro DC, Berridge KC (March 2014)."Opioid hedonic hotspot in nucleus accumbens shell: mu, delta, and kappa maps for enhancement of sweetness "liking" and "wanting"".The Journal of Neuroscience.34 (12):4239–4250.doi:10.1523/JNEUROSCI.4458-13.2014.PMC 3960467.PMID 24647944.
  126. ^abPirino BE, Spodnick MB, Gargiulo AT, Curtis GR, Barson JR, Karkhanis AN (December 2020)."Kappa-opioid receptor-dependent changes in dopamine and anxiety-like or approach-avoidance behavior occur differentially across the nucleus accumbens shell rostro-caudal axis".Neuropharmacology.181 108341.doi:10.1016/j.neuropharm.2020.108341.PMC 8424943.PMID 33011200.
  127. ^abKarkhanis AN, West AM, Jones SR (November 2023)."Kappa opioid receptor agonist U50,488 inhibits dopamine more in caudal than rostral nucleus accumbens core".Basic & Clinical Pharmacology & Toxicology.133 (5):526–534.doi:10.1111/bcpt.13929.PMC 10972693.PMID 37539456.
  128. ^Margolis EB, Wallace TL, Van Orden LJ, Martin WJ (2020)."Differential effects of novel kappa opioid receptor antagonists on dopamine neurons using acute brain slice electrophysiology".PloS One.15 (12) e0232864.Bibcode:2020PLoSO..1532864M.doi:10.1371/journal.pone.0232864.PMC 7771853.PMID 33373369.
  129. ^Hasebe K, Kawai K, Suzuki T, Kawamura K, Tanaka T, Narita M, et al. (October 2004). "Possible pharmacotherapy of the opioid kappa receptor agonist for drug dependence".Annals of the New York Academy of Sciences.1025 (1):404–413.Bibcode:2004NYASA1025..404H.doi:10.1196/annals.1316.050.PMID 15542743.S2CID 85031737.
  130. ^Frankel PS, Alburges ME, Bush L, Hanson GR, Kish SJ (July 2008)."Striatal and ventral pallidum dynorphin concentrations are markedly increased in human chronic cocaine users".Neuropharmacology.55 (1):41–46.doi:10.1016/j.neuropharm.2008.04.019.PMC 2577569.PMID 18538358.
  131. ^Michaels CC, Holtzman SG (April 2008). "Early postnatal stress alters place conditioning to both mu- and kappa-opioid agonists".The Journal of Pharmacology and Experimental Therapeutics.325 (1):313–318.doi:10.1124/jpet.107.129908.PMID 18203949.S2CID 30383220.
  132. ^abBeardsley PM, Howard JL, Shelton KL, Carroll FI (November 2005). "Differential effects of the novel kappa opioid receptor antagonist, JDTic, on reinstatement of cocaine-seeking induced by footshock stressors vs cocaine primes and its antidepressant-like effects in rats".Psychopharmacology.183 (1):118–126.doi:10.1007/s00213-005-0167-4.PMID 16184376.S2CID 31140425.
  133. ^Redila VA, Chavkin C (September 2008)."Stress-induced reinstatement of cocaine seeking is mediated by the kappa opioid system".Psychopharmacology.200 (1):59–70.doi:10.1007/s00213-008-1122-y.PMC 2680147.PMID 18575850.
  134. ^Blum K, Braverman ER, Holder JM, Lubar JF, Monastra VJ, Miller D, et al. (November 2000). "Reward deficiency syndrome: a biogenetic model for the diagnosis and treatment of impulsive, addictive, and compulsive behaviors".Journal of Psychoactive Drugs.32 (Suppl):i–iv,1–112.doi:10.1080/02791072.2000.10736099.PMID 11280926.S2CID 22497665.
  135. ^Stefański R, Ziółkowska B, Kuśmider M, Mierzejewski P, Wyszogrodzka E, Kołomańska P, et al. (July 2007). "Active versus passive cocaine administration: differences in the neuroadaptive changes in the brain dopaminergic system".Brain Research.1157:1–10.doi:10.1016/j.brainres.2007.04.074.PMID 17544385.S2CID 42090922.
  136. ^Moore RJ, Vinsant SL, Nader MA, Porrino LJ, Friedman DP (September 1998)."Effect of cocaine self-administration on dopamine D2 receptors in rhesus monkeys".Synapse.30 (1):88–96.doi:10.1002/(SICI)1098-2396(199809)30:1<88::AID-SYN11>3.0.CO;2-L.PMID 9704885.S2CID 22569502.
  137. ^D'Addario C, Di Benedetto M, Izenwasser S, Candeletti S, Romualdi P (January 2007). "Role of serotonin in the regulation of the dynorphinergic system by a kappa-opioid agonist and cocaine treatment in rat CNS".Neuroscience.144 (1):157–164.doi:10.1016/j.neuroscience.2006.09.008.PMID 17055175.S2CID 34243587.
  138. ^Mash DC, Staley JK (June 1999). "D3 dopamine and kappa opioid receptor alterations in human brain of cocaine-overdose victims".Annals of the New York Academy of Sciences.877 (1):507–522.Bibcode:1999NYASA.877..507M.doi:10.1111/j.1749-6632.1999.tb09286.x.PMID 10415668.S2CID 25867468.
  139. ^Schenk S, Partridge B, Shippenberg TS (June 1999). "U69593, a kappa-opioid agonist, decreases cocaine self-administration and decreases cocaine-produced drug-seeking".Psychopharmacology.144 (4):339–346.doi:10.1007/s002130051016.PMID 10435406.S2CID 19726351.
  140. ^Patkar AA, Mannelli P, Hill KP, Peindl K, Pae CU, Lee TH (August 2006)."Relationship of prolactin response to meta-chlorophenylpiperazine with severity of drug use in cocaine dependence".Human Psychopharmacology.21 (6):367–375.doi:10.1002/hup.780.PMID 16915581.S2CID 21895907.
  141. ^Gregg C, Shikar V, Larsen P, Mak G, Chojnacki A, Yong VW, et al. (February 2007)."White matter plasticity and enhanced remyelination in the maternal CNS".The Journal of Neuroscience.27 (8):1812–1823.doi:10.1523/JNEUROSCI.4441-06.2007.PMC 6673564.PMID 17314279.
  142. ^McLaughlin JP, Marton-Popovici M, Chavkin C (July 2003)."Kappa opioid receptor antagonism and prodynorphin gene disruption block stress-induced behavioral responses".The Journal of Neuroscience.23 (13):5674–5683.doi:10.1523/JNEUROSCI.23-13-05674.2003.PMC 2104777.PMID 12843270.
  143. ^McLaughlin JP, Li S, Valdez J, Chavkin TA, Chavkin C (June 2006)."Social defeat stress-induced behavioral responses are mediated by the endogenous kappa opioid system".Neuropsychopharmacology.31 (6):1241–1248.doi:10.1038/sj.npp.1300872.PMC 2096774.PMID 16123746.
  144. ^Koob GF (July 2008)."A role for brain stress systems in addiction".Neuron.59 (1):11–34.doi:10.1016/j.neuron.2008.06.012.PMC 2748830.PMID 18614026.
  145. ^abBruchas MR, Xu M, Chavkin C (September 2008)."Repeated swim stress induces kappa opioid-mediated activation of extracellular signal-regulated kinase 1/2".Neuroreport.19 (14):1417–1422.doi:10.1097/WNR.0b013e32830dd655.PMC 2641011.PMID 18766023.
  146. ^abBanks ML (2020). "The Rise and Fall of Kappa-Opioid Receptors in Drug Abuse Research".Substance Use Disorders. Handbook of Experimental Pharmacology. Vol. 258. pp. 147–165.doi:10.1007/164_2019_268.ISBN 978-3-030-33678-3.PMC 7756963.PMID 31463605.
  147. ^Mitchell JM, Liang MT, Fields HL (November 2005). "A single injection of the kappa opioid antagonist norbinaltorphimine increases ethanol consumption in rats".Psychopharmacology.182 (3):384–392.doi:10.1007/s00213-005-0067-7.PMID 16001119.S2CID 38011973.
  148. ^Walker BM, Koob GF (February 2008)."Pharmacological evidence for a motivational role of kappa-opioid systems in ethanol dependence".Neuropsychopharmacology.33 (3):643–652.doi:10.1038/sj.npp.1301438.PMC 2739278.PMID 17473837.
  149. ^Xi ZX, Fuller SA, Stein EA (January 1998). "Dopamine release in the nucleus accumbens during heroin self-administration is modulated by kappa opioid receptors: an in vivo fast-cyclic voltammetry study".The Journal of Pharmacology and Experimental Therapeutics.284 (1):151–161.doi:10.1016/S0022-3565(24)37186-1.PMID 9435173.
  150. ^Narita M, Khotib J, Suzuki M, Ozaki S, Yajima Y, Suzuki T (June 2003). "Heterologous mu-opioid receptor adaptation by repeated stimulation of kappa-opioid receptor: up-regulation of G-protein activation and antinociception".Journal of Neurochemistry.85 (5):1171–1179.doi:10.1046/j.1471-4159.2003.01754.x.PMID 12753076.S2CID 26034314.
  151. ^Maisonneuve IM, Archer S, Glick SD (November 1994). "U50,488, a kappa opioid receptor agonist, attenuates cocaine-induced increases in extracellular dopamine in the nucleus accumbens of rats".Neuroscience Letters.181 (1–2):57–60.doi:10.1016/0304-3940(94)90559-2.PMID 7898771.S2CID 25258989.
  152. ^Wee S, Koob GF (June 2010)."The role of the dynorphin-kappa opioid system in the reinforcing effects of drugs of abuse".Psychopharmacology.210 (2):121–135.doi:10.1007/s00213-010-1825-8.PMC 2879894.PMID 20352414.
  153. ^Farahbakhsh ZZ, Song K, Branthwaite HE, Erickson KR, Mukerjee S, Nolan SO, et al. (May 2023)."Systemic kappa opioid receptor antagonism accelerates reinforcement learning via augmentation of novelty processing in male mice".Neuropsychopharmacology.48 (6):857–868.doi:10.1038/s41386-023-01547-x.PMC 10156709.PMID 36804487.
  154. ^Braden K, Castro DC (May 2023)."Kappa opioid receptors as modulators of novelty processing".Neuropsychopharmacology.48 (6):848–849.doi:10.1038/s41386-023-01561-z.PMC 10156680.PMID 36922627.
  155. ^Krystal AD, Pizzagalli DA, Smoski M, Mathew SJ, Nurnberger J, Lisanby SH, et al. (May 2020)."A randomized proof-of-mechanism trial applying the 'fast-fail' approach to evaluating κ-opioid antagonism as a treatment for anhedonia".Nature Medicine.26 (5):760–768.doi:10.1038/s41591-020-0806-7.PMC 9949770.PMID 32231295.
  156. ^abcAddy PH, Garcia-Romeu A, Metzger M, Wade J (April 2015). "The subjective experience of acute, experimentally-induced Salvia divinorum inebriation".Journal of Psychopharmacology.29 (4):426–435.doi:10.1177/0269881115570081.PMID 25691501.
  157. ^Koubeissi MZ, Bartolomei F, Beltagy A, Picard F (August 2014). "Electrical stimulation of a small brain area reversibly disrupts consciousness".Epilepsy & Behavior.37:32–35.doi:10.1016/j.yebeh.2014.05.027.PMID 24967698.
  158. ^Calarco N (April 2024)."Mapping the claustrum to elucidate consciousness".Nature Reviews Psychology.3 (6): 374.doi:10.1038/s44159-024-00313-0.ISSN 2731-0574.
  159. ^Narikiyo K, Mizuguchi R, Ajima A, Shiozaki M, Hamanaka H, Johansen JP, et al. (June 2020). "The claustrum coordinates cortical slow-wave activity".Nature Neuroscience.23 (6):741–753.doi:10.1038/s41593-020-0625-7.PMID 32393895.S2CID 256840965.
  160. ^Bickel S, Parvizi J (August 2019)."Electrical stimulation of the human claustrum".Epilepsy & Behavior.97:296–303.doi:10.1016/j.yebeh.2019.03.051.PMID 31196825.S2CID 182952015.
  161. ^Cheng Y, Zheng Y, Ren L (January 2025)."The Disruption of the Claustrum-Anterior Cingulate Cortex Pathway Impaired the Human Consciousness".Brain Stimulation.18 (1): 609.doi:10.1016/j.brs.2024.12.1141.ISSN 1935-861X.
  162. ^abcBagdasarian FA, Hansen HD, Chen J, Yoo CH, Placzek MS, Hooker JM, et al. (July 2024). "Acute Effects of Hallucinogens on Functional Connectivity: Psilocybin and Salvinorin-A".ACS Chemical Neuroscience.15 (14):2654–2661.doi:10.1021/acschemneuro.4c00245.PMID 38916752.
  163. ^abKim JH, Jang YH, Chun KJ, Kim J, Park YH, Kim JS, et al. (May 2011)."Kappa-opioid receptor activation during reperfusion limits myocardial infarction via ERK1/2 activation in isolated rat hearts".Korean Journal of Anesthesiology.60 (5):351–356.doi:10.4097/kjae.2011.60.5.351.PMC 3110294.PMID 21716908.
  164. ^abChavkin C, Martinez D (August 2015)."Kappa Antagonist JDTic in Phase 1 Clinical Trial".Neuropsychopharmacology.40 (9):2057–2058.doi:10.1038/npp.2015.74.PMC 4613619.PMID 26174493.
  165. ^Ito H, Navratilova E, Vagnerova B, Watanabe M, Kopruszinski C, Moreira de Souza LH, et al. (March 2023)."Chronic pain recruits hypothalamic dynorphin/kappa opioid receptor signalling to promote wakefulness and vigilance".Brain.146 (3):1186–1199.doi:10.1093/brain/awac153.PMC 10169443.PMID 35485490.
  166. ^abKamimura K, Yokoo T, Kamimura H, Sakamaki A, Abe S, Tsuchiya A, et al. (2017)."Long-term efficacy and safety of nalfurafine hydrochloride on pruritus in chronic liver disease patients: Patient-reported outcome based analyses".PloS One.12 (6) e0178991.Bibcode:2017PLoSO..1278991K.doi:10.1371/journal.pone.0178991.PMC 5467861.PMID 28604788.
  167. ^abInui S (2015-05-11)."Nalfurafine hydrochloride to treat pruritus: a review".Clinical, Cosmetic and Investigational Dermatology.8:249–255.doi:10.2147/CCID.S55942.PMC 4433050.PMID 26005355.
  168. ^Baskin DS, Hosobuchi Y, Grevel JC (January 1986). "Treatment of experimental stroke with opiate antagonists. Effects on neurological function, infarct size, and survival".Journal of Neurosurgery.64 (1):99–103.doi:10.3171/jns.1986.64.1.0099.PMID 3941353.
  169. ^Zeynalov E, Nemoto M, Hurn PD, Koehler RC, Bhardwaj A (March 2006)."Neuroprotective effect of selective kappa opioid receptor agonist is gender specific and linked to reduced neuronal nitric oxide".Journal of Cerebral Blood Flow and Metabolism.26 (3):414–420.doi:10.1038/sj.jcbfm.9600196.PMID 16049424.
  170. ^Feiner JR, Weiskopf RB (January 2017)."Evaluating Pulmonary Function: An Assessment of PaO2/FIO2".Critical Care Medicine.45 (1):e40 –e48.doi:10.1097/CCM.0000000000002017.PMC 5203937.PMID 27618274.
  171. ^Tortella FC, Robles L, Holaday JW (April 1986). "U50,488, a highly selective kappa opioid: anticonvulsant profile in rats".The Journal of Pharmacology and Experimental Therapeutics.237 (1):49–53.doi:10.1016/S0022-3565(25)24991-6.PMID 3007743.
  172. ^Pasternak GW (June 1980)."Multiple opiate receptors: [3H]ethylketocyclazocine receptor binding and ketocyclazocine analgesia".Proceedings of the National Academy of Sciences of the United States of America.77 (6):3691–3694.Bibcode:1980PNAS...77.3691P.doi:10.1073/pnas.77.6.3691.PMC 349684.PMID 6251477.
  173. ^Roth BL, Baner K, Westkaemper R, Siebert D, Rice KC, Steinberg S, et al. (September 2002)."Salvinorin A: a potent naturally occurring nonnitrogenous kappa opioid selective agonist".Proceedings of the National Academy of Sciences of the United States of America.99 (18):11934–11939.Bibcode:2002PNAS...9911934R.doi:10.1073/pnas.182234399.PMC 129372.PMID 12192085.
  174. ^Holtzman SG (February 1985). "Drug discrimination studies".Drug and Alcohol Dependence.14 (3–4):263–282.doi:10.1016/0376-8716(85)90061-4.PMID 2859972.
  175. ^Nielsen CK, Ross FB, Lotfipour S, Saini KS, Edwards SR, Smith MT (December 2007). "Oxycodone and morphine have distinctly different pharmacological profiles: radioligand binding and behavioural studies in two rat models of neuropathic pain".Pain.132 (3):289–300.doi:10.1016/j.pain.2007.03.022.PMID 17467904.S2CID 19872213.
  176. ^abWhite KL, Robinson JE, Zhu H, DiBerto JF, Polepally PR, Zjawiony JK, et al. (January 2015)."The G protein-biased κ-opioid receptor agonist RB-64 is analgesic with a unique spectrum of activities in vivo".The Journal of Pharmacology and Experimental Therapeutics.352 (1):98–109.doi:10.1124/jpet.114.216820.PMC 4279099.PMID 25320048.
  177. ^Wang Y, Chen Y, Xu W, Lee DY, Ma Z, Rawls SM, et al. (March 2008)."2-Methoxymethyl-salvinorin B is a potent kappa opioid receptor agonist with longer lasting action in vivo than salvinorin A".The Journal of Pharmacology and Experimental Therapeutics.324 (3):1073–1083.doi:10.1124/jpet.107.132142.PMC 2519046.PMID 18089845.
  178. ^Munro TA, Duncan KK, Xu W, Wang Y, Liu-Chen LY, Carlezon WA, et al. (February 2008)."Standard protecting groups create potent and selective kappa opioids: salvinorin B alkoxymethyl ethers".Bioorganic & Medicinal Chemistry.16 (3):1279–1286.doi:10.1016/j.bmc.2007.10.067.PMC 2568987.PMID 17981041.
  179. ^Baker LE, Panos JJ, Killinger BA, Peet MM, Bell LM, Haliw LA, et al. (April 2009)."Comparison of the discriminative stimulus effects of salvinorin A and its derivatives to U69,593 and U50,488 in rats".Psychopharmacology.203 (2):203–211.doi:10.1007/s00213-008-1458-3.PMID 19153716.
  180. ^Patrick GL (10 January 2013).An Introduction to Medicinal Chemistry. OUP Oxford. pp. 657–.ISBN 978-0-19-969739-7.
  181. ^Nagase H (21 January 2011).Chemistry of Opioids. Springer. pp. 34, 48,57–60.ISBN 978-3-642-18107-8.
  182. ^Katavic PL, Lamb K, Navarro H, Prisinzano TE (August 2007)."Flavonoids as opioid receptor ligands: identification and preliminary structure-activity relationships".Journal of Natural Products.70 (8):1278–1282.doi:10.1021/np070194x.PMC 2265593.PMID 17685652.
  183. ^abCasal-Dominguez JJ, Furkert D, Ostovar M, Teintang L, Clark MJ, Traynor JR, et al. (March 2014)."Characterization of BU09059: a novel potent selective κ-receptor antagonist".ACS Chemical Neuroscience.5 (3):177–184.doi:10.1021/cn4001507.PMC 3963132.PMID 24410326.
  184. ^"CVL-354".adisinsight.springer.com. Retrieved2 February 2023.
  185. ^Hartung AM, Beutler JA, Navarro HA, Wiemer DF, Neighbors JD (February 2014)."Stilbenes as κ-selective, non-nitrogenous opioid receptor antagonists".Journal of Natural Products.77 (2):311–319.Bibcode:2014JNAtP..77..311H.doi:10.1021/np4009046.PMC 3993902.PMID 24456556.
  186. ^Wold EA, Chen J, Cunningham KA, Zhou J (January 2019)."Allosteric Modulation of Class A GPCRs: Targets, Agents, and Emerging Concepts".Journal of Medicinal Chemistry.62 (1):88–127.doi:10.1021/acs.jmedchem.8b00875.PMC 6556150.PMID 30106578.BMS-986187 (179), with a chemically novel core compared to previous BMS series, was discovered as an effective PAM at the DOR and at the κ-opioid receptor (KOR) rather than the MOR with an approximately 20-to 30-fold higher affinity in the allosteric ternary complex model.261
  187. ^Livingston KE, Stanczyk MA, Burford NT, Alt A, Canals M, Traynor JR (February 2018)."Pharmacologic Evidence for a Putative Conserved Allosteric Site on Opioid Receptors".Molecular Pharmacology.93 (2):157–167.doi:10.1124/mol.117.109561.PMC 5767684.PMID 29233847.
  188. ^Zhao J, Baiula M, Cuna E, Francescato M, Matalińska J, Lipiński PF, et al. (October 2024)."Identification of c[D-Trp-Phe-β-Ala-β-Ala], the First κ-Opioid Receptor-Specific Negative Allosteric Modulator".ACS Pharmacology & Translational Science.7 (10):3192–3204.doi:10.1021/acsptsci.4c00372.PMC 11475277.PMID 39416958.
  189. ^abcYuferov V, Fussell D, LaForge KS, Nielsen DA, Gordon D, Ho A, et al. (December 2004)."Redefinition of the human kappa opioid receptor gene (OPRK1) structure and association of haplotypes with opiate addiction".Pharmacogenetics.14 (12):793–804.doi:10.1097/00008571-200412000-00002.PMC 6141019.PMID 15608558.
  190. ^Demircan G, Ozkan-Kotilolgu S (2023)."The frequency of OPRK1 G36T and OPRM1 A118G opioid receptor gene polymorphisms in heroin-dependent individuals and non-dependent healthy subjects in Turkey".Turkish Journal of Clinical Psychiatry.26 (3):163–169.doi:10.5505/kpd.2023.73693.
  191. ^Xu K, Seo D, Hodgkinson C, Hu Y, Goldman D, Sinha R (August 2013)."A variant on the kappa opioid receptor gene (OPRK1) is associated with stress response and related drug craving, limbic brain activation and cocaine relapse risk".Translational Psychiatry.3 (8): e292.doi:10.1038/tp.2013.62.PMC 3756290.PMID 23962922.
  192. ^abcLutz PE, Gross JA, Dhir SK, Maussion G, Yang J, Bramoulle A, et al. (November 2018). "Epigenetic Regulation of the Kappa Opioid Receptor by Child Abuse".Biological Psychiatry.84 (10):751–761.doi:10.1016/j.biopsych.2017.07.012.PMID 28886759.
  193. ^Stanley B, Siever LJ (January 2010)."The interpersonal dimension of borderline personality disorder: toward a neuropeptide model".The American Journal of Psychiatry.167 (1):24–39.doi:10.1176/appi.ajp.2009.09050744.PMC 4176877.PMID 19952075.
  194. ^abDavis M, Wilson O, DeBonee S, Nabulsi N, Matuskey D, Esterlis I (May 2022)."P204. Role of Kappa Opioid Receptor Availability in Borderline Personality Disorder and Suicidal Behavior: Preliminary Results From an in Vivo [11C]EKAP PET Study".Biological Psychiatry.91 (9):S169 –S170.doi:10.1016/j.biopsych.2022.02.438.ISSN 0006-3223.
  195. ^abNiinep K, Anier K, Eteläinen T, Piepponen P, Kalda A (2021)."Repeated Ethanol Exposure Alters DNA Methylation Status and Dynorphin/Kappa-Opioid Receptor Expression in Nucleus Accumbens of Alcohol-Preferring AA Rats".Frontiers in Genetics.12 750142.doi:10.3389/fgene.2021.750142.PMC 8652212.PMID 34899839.
  196. ^abcMurphy MD, Heller EA (December 2022)."Convergent actions of stress and stimulants via epigenetic regulation of neural circuitry".Trends in Neurosciences.45 (12):955–967.doi:10.1016/j.tins.2022.10.001.ISSN 0166-2236.PMC 9671852.PMID 36280459.
  197. ^Yuferov V, Nielsen DA, Levran O, Randesi M, Hamon S, Ho A, et al. (April 2011)."Tissue-specific DNA methylation of the human prodynorphin gene in post-mortem brain tissues and PBMCs".Pharmacogenetics and Genomics.21 (4):185–196.doi:10.1097/FPC.0b013e32833eecbc.PMC 3017726.PMID 20808262.
  198. ^Mansouri N, Omarmeli V, Sharafshah A, Assefi M, Lewandrowski KU (2023)."Investigation of Methylation Levels in OPRK1 Gene Promoter among Smokers and Opium-Addicts underwent Methadone Maintenance Treatment"(PDF).Annals of Clinical and Medical Case Reports.10 (19):01–08.doi:10.47829/ACMCR.2023.101902 (inactive 9 November 2025).ISSN 2639-8109.{{cite journal}}: CS1 maint: DOI inactive as of November 2025 (link)
  199. ^abSzymaszkiewicz A, Talar M, Włodarczyk J, Świerczyński M, Bartoszek A, Krajewska J, et al. (March 2022)."The Involvement of the Endogenous Opioid System in the Gastrointestinal Aging in Mice and Humans".International Journal of Molecular Sciences.23 (7): 3565.doi:10.3390/ijms23073565.PMC 8998735.PMID 35408926.
  200. ^van Rijn RM, Whistler JL, Waldhoer M (February 2010)."Opioid-receptor-heteromer-specific trafficking and pharmacology".Current Opinion in Pharmacology.10 (1):73–79.doi:10.1016/j.coph.2009.09.007.PMC 2900797.PMID 19846340.
  201. ^Le Naour M, Lunzer MM, Powers MD, Kalyuzhny AE, Benneyworth MA, Thomas MJ, et al. (August 2014)."Putative kappa opioid heteromers as targets for developing analgesics free of adverse effects".Journal of Medicinal Chemistry.57 (15):6383–6392.doi:10.1021/jm500159d.ISSN 1520-4804.PMC 4136663.PMID 24978316.
  202. ^abLiu H, Tian Y, Ji B, Lu H, Xin Q, Jiang Y, et al. (November 2016). "Heterodimerization of the kappa opioid receptor and neurotensin receptor 1 contributes to a novel β-arrestin-2-biased pathway".Biochimica et Biophysica Acta.1863 (11):2719–2738.doi:10.1016/j.bbamcr.2016.07.009.PMID 27523794.
  203. ^Johnston JM, Aburi M, Provasi D, Bortolato A, Urizar E, Lambert NA, et al. (March 2011)."Making Structural Sense of Dimerization Interfaces of Delta Opioid Receptor Homodimers".Biochemistry.50 (10):1682–1690.doi:10.1021/bi101474v.ISSN 0006-2960.PMC 3050604.PMID 21261298.
  204. ^Chakrabarti S, Liu NJ, Gintzler AR (November 2010)."Formation of mu-/kappa-opioid receptor heterodimer is sex-dependent and mediates female-specific opioid analgesia".Proceedings of the National Academy of Sciences of the United States of America.107 (46):20115–20119.Bibcode:2010PNAS..10720115C.doi:10.1073/pnas.1009923107.PMC 2993367.PMID 21041644.
  205. ^Donica CL, Awwad HO, Thakker DR, Standifer KM (May 2013)."Cellular mechanisms of nociceptin/orphanin FQ (N/OFQ) peptide (NOP) receptor regulation and heterologous regulation by N/OFQ".Molecular Pharmacology.83 (5):907–918.doi:10.1124/mol.112.084632.PMC 3629824.PMID 23395957.
  206. ^Chen J, Zhang R, Chen X, Wang C, Cai X, Liu H, et al. (July 2015)."Heterodimerization of human orexin receptor 1 and kappa opioid receptor promotes protein kinase A/cAMP-response element binding protein signaling via a Gαs-mediated mechanism".Cellular Signalling.27 (7):1426–1438.doi:10.1016/j.cellsig.2015.03.027.ISSN 0898-6568.
  207. ^Ji B, Liu H, Zhang R, Jiang Y, Wang C, Li S, et al. (February 2017). "Novel signaling of dynorphin at κ-opioid receptor/bradykinin B2 receptor heterodimers".Cellular Signalling.31:66–78.doi:10.1016/j.cellsig.2017.01.005.PMID 28069442.
  208. ^Jordan BA, Trapaidze N, Gomes I, Nivarthi R, Devi LA (January 2001)."Oligomerization of opioid receptors with beta 2-adrenergic receptors: a role in trafficking and mitogen-activated protein kinase activation".Proceedings of the National Academy of Sciences of the United States of America.98 (1):343–348.doi:10.1073/pnas.98.1.343.PMC 14592.PMID 11134510.
  209. ^Xiao J, Zeng S, Wang X, Babazada H, Li Z, Liu R, et al. (2016)."Neurokinin 1 and opioid receptors: relationships and interactions in nervous system".Translational Perioperative and Pain Medicine.1 (3):11–21.PMC 5388438.PMID 28409174.
  210. ^Lawrence DM, Bidlack JM (September 1993). "The kappa opioid receptor expressed on the mouse R1.1 thymoma cell line is coupled to adenylyl cyclase through a pertussis toxin-sensitive guanine nucleotide-binding regulatory protein".The Journal of Pharmacology and Experimental Therapeutics.266 (3):1678–1683.doi:10.1016/S0022-3565(25)39354-7.PMID 8103800.
  211. ^Konkoy CS, Childers SR (January 1993). "Relationship between kappa 1 opioid receptor binding and inhibition of adenylyl cyclase in guinea pig brain membranes".Biochemical Pharmacology.45 (1):207–216.doi:10.1016/0006-2952(93)90394-C.PMID 8381004.
  212. ^Schoffelmeer AN, Rice KC, Jacobson AE, Van Gelderen JG, Hogenboom F, Heijna MH, et al. (September 1988). "Mu-, delta- and kappa-opioid receptor-mediated inhibition of neurotransmitter release and adenylate cyclase activity in rat brain slices: studies with fentanyl isothiocyanate".European Journal of Pharmacology.154 (2):169–178.doi:10.1016/0014-2999(88)90094-5.PMID 2906610.
  213. ^Henry DJ, Grandy DK, Lester HA, Davidson N, Chavkin C (March 1995). "Kappa-opioid receptors couple to inwardly rectifying potassium channels when coexpressed by Xenopus oocytes".Molecular Pharmacology.47 (3):551–557.doi:10.1016/S0026-895X(25)08575-X.PMID 7700253.
  214. ^Tallent M, Dichter MA, Bell GI, Reisine T (December 1994)."The cloned kappa opioid receptor couples to an N-type calcium current in undifferentiated PC-12 cells".Neuroscience.63 (4):1033–1040.doi:10.1016/0306-4522(94)90570-3.PMID 7700508.S2CID 22003522.
  215. ^Bohn LM, Belcheva MM, Coscia CJ (February 2000)."Mitogenic signaling via endogenous kappa-opioid receptors in C6 glioma cells: evidence for the involvement of protein kinase C and the mitogen-activated protein kinase signaling cascade".Journal of Neurochemistry.74 (2):564–573.doi:10.1046/j.1471-4159.2000.740564.x.PMC 2504523.PMID 10646507.
  216. ^Belcheva MM, Clark AL, Haas PD, Serna JS, Hahn JW, Kiss A, et al. (July 2005)."Mu and kappa opioid receptors activate ERK/MAPK via different protein kinase C isoforms and secondary messengers in astrocytes".The Journal of Biological Chemistry.280 (30):27662–27669.doi:10.1074/jbc.M502593200.PMC 1400585.PMID 15944153.
  217. ^Kam AY, Chan AS, Wong YH (July 2004). "Kappa-opioid receptor signals through Src and focal adhesion kinase to stimulate c-Jun N-terminal kinases in transfected COS-7 cells and human monocytic THP-1 cells".The Journal of Pharmacology and Experimental Therapeutics.310 (1):301–310.doi:10.1124/jpet.104.065078.PMID 14996948.S2CID 39445016.
  218. ^Bruchas MR, Yang T, Schreiber S, Defino M, Kwan SC, Li S, et al. (October 2007)."Long-acting kappa opioid antagonists disrupt receptor signaling and produce noncompetitive effects by activating c-Jun N-terminal kinase".The Journal of Biological Chemistry.282 (41):29803–29811.Bibcode:2007JBiCh.28229803B.doi:10.1074/jbc.M705540200.PMC 2096775.PMID 17702750.
  219. ^Huang P, Steplock D, Weinman EJ, Hall RA, Ding Z, Li J, et al. (June 2004)."kappa Opioid receptor interacts with Na(+)/H(+)-exchanger regulatory factor-1/Ezrin-radixin-moesin-binding phosphoprotein-50 (NHERF-1/EBP50) to stimulate Na(+)/H(+) exchange independent of G(i)/G(o) proteins".The Journal of Biological Chemistry.279 (24):25002–25009.doi:10.1074/jbc.M313366200.PMID 15070904.
  220. ^Li JG, Chen C, Liu-Chen LY (July 2002)."Ezrin-radixin-moesin-binding phosphoprotein-50/Na+/H+ exchanger regulatory factor (EBP50/NHERF) blocks U50,488H-induced down-regulation of the human kappa opioid receptor by enhancing its recycling rate".The Journal of Biological Chemistry.277 (30):27545–27552.doi:10.1074/jbc.M200058200.PMID 12004055.
  221. ^Li JG, Haines DS, Liu-Chen LY (April 2008)."Agonist-promoted Lys63-linked polyubiquitination of the human kappa-opioid receptor is involved in receptor down-regulation".Molecular Pharmacology.73 (4):1319–1330.doi:10.1124/mol.107.042846.PMC 3489932.PMID 18212250.
  222. ^Maraschin JC, Almeida CB, Rangel MP, Roncon CM, Sestile CC, Zangrossi H, et al. (June 2017). "Participation of dorsal periaqueductal gray 5-HT1A receptors in the panicolytic-like effect of the κ-opioid receptor antagonist Nor-BNI".Behavioural Brain Research.327:75–82.doi:10.1016/j.bbr.2017.03.033.PMID 28347824.S2CID 22465963.
  223. ^Gross JD, Kaski SW, Schmidt KT, Cogan ES, Boyt KM, Wix K, et al. (September 2019)."Role of RGS12 in the differential regulation of kappa opioid receptor-dependent signaling and behavior".Neuropsychopharmacology.44 (10):1728–1741.doi:10.1038/s41386-019-0423-7.PMC 6785087.PMID 31141817.
  224. ^abBruchas MR, Chavkin C (June 2010)."Kinase cascades and ligand-directed signaling at the kappa opioid receptor".Psychopharmacology.210 (2):137–147.doi:10.1007/s00213-010-1806-y.PMC 3671863.PMID 20401607.
  225. ^abcLiu-Chen LY, Huang P (November 2022)."Signaling underlying kappa opioid receptor-mediated behaviors in rodents".Frontiers in Neuroscience.16 964724.doi:10.3389/fnins.2022.964724.PMC 9670127.PMID 36408401.
  226. ^Morgenweck J, Frankowski KJ, Prisinzano TE, Aubé J, Bohn LM (December 2015)."Investigation of the role of βarrestin2 in kappa opioid receptor modulation in a mouse model of pruritus".Neuropharmacology.99:600–609.doi:10.1016/j.neuropharm.2015.08.027.PMC 4739521.PMID 26318102.
  227. ^Robinson JD, McDonald PH (July 2015)."The orexin 1 receptor modulates kappa opioid receptor function via a JNK-dependent mechanism".Cellular Signalling.27 (7):1449–1456.doi:10.1016/j.cellsig.2015.03.026.PMC 5549559.PMID 25857454.
  228. ^Liu JJ, Chiu YT, DiMattio KM, Chen C, Huang P, Gentile TA, et al. (April 2019)."Phosphoproteomic approach for agonist-specific signaling in mouse brains: mTOR pathway is involved in κ opioid aversion".Neuropsychopharmacology.44 (5):939–949.doi:10.1038/s41386-018-0155-0.PMC 6462019.PMID 30082888.
  229. ^abLiu JJ, Chiu YT, Chen C, Huang P, Mann M, Liu-Chen LY (December 2020)."Pharmacological and phosphoproteomic approaches to roles of protein kinase C in kappa opioid receptor-mediated effects in mice".Neuropharmacology.181 108324.doi:10.1016/j.neuropharm.2020.108324.PMC 8148297.PMID 32976891.
  230. ^abKunselman JM, Gupta A, Gomes I, Devi LA, Puthenveedu MA (April 2021)."Compartment-specific opioid receptor signaling is selectively modulated by different dynorphin peptides".Elife.10 e60270.doi:10.7554/eLife.60270.PMC 8112862.PMID 33908346.
  231. ^abBowman SL, Puthenveedu MA (2015)."Postendocytic Sorting of Adrenergic and Opioid Receptors: New Mechanisms and Functions".Progress in Molecular Biology and Translational Science.132:189–206.doi:10.1016/bs.pmbts.2015.03.005.PMC 6317848.PMID 26055059.
  232. ^abcdRappoport A (2024),A Dynorphin Theory of Depression and Bipolar Disorder,arXiv:2408.06763, retrieved2025-11-06
  233. ^Crilly SE, Puthenveedu MA (June 2021)."Compartmentalized GPCR Signaling from Intracellular Membranes".The Journal of Membrane Biology.254 (3):259–271.doi:10.1007/s00232-020-00158-7.PMC 8141539.PMID 33231722.
  234. ^abcGarcía-Domínguez M (October 2025)."κ-Opioid Receptor Agonists as Robust Pain-Modulating Agents: Mechanisms and Therapeutic Potential in Pain Modulation".Journal of Clinical Medicine.14 (20): 7263.doi:10.3390/jcm14207263.PMC 12564769.PMID 41156133.
  235. ^Kaye AD (2015). Vadivelu N, Urman RD (eds.).Substance Abuse: Inpatient and Outpatient Management for Every Clinician. SpringerLink Bücher. New York, NY s.l: Springer New York Imprint: Springer.ISBN 978-1-4939-1951-2.{{cite book}}: CS1 maint: publisher location (link)
  236. ^Kaski SW, White AN, Gross JD, Siderovski DP (February 2021)."Potential for Kappa-Opioid Receptor Agonists to Engineer Nonaddictive Analgesics: A Narrative Review".Anesthesia and Analgesia.132 (2):406–419.doi:10.1213/ANE.0000000000005309.PMC 7992303.PMID 33332902.
  237. ^Santino F, Gentilucci L (January 2023)."Design of κ-Opioid Receptor Agonists for the Development of Potential Treatments of Pain with Reduced Side Effects".Molecules.28 (1): 346.doi:10.3390/molecules28010346.PMC 9822356.PMID 36615540.
  238. ^US Food and Drug Administration. Drugs@FDA: FDA-Approved Drugs. Accessed October 18, 2025.
  239. ^European Medicines Agency. Kapruvia. Published February 22, 2022. Accessed October 18, 2025.
  240. ^Zan GY, Wang Q, Wang YJ, Liu Y, Hang A, Shu XH, et al. (September 2015). "Antagonism of κ opioid receptor in the nucleus accumbens prevents the depressive-like behaviors following prolonged morphine abstinence".Behavioural Brain Research.291:334–341.doi:10.1016/j.bbr.2015.05.053.PMID 26049060.S2CID 32817749.
  241. ^abWong S, Le GH, Vasudeva S, Teopiz KM, Phan L, Meshkat S, et al. (October 2024). "Preclinical and clinical efficacy of kappa opioid receptor antagonists for depression: A systematic review".Journal of Affective Disorders.362:816–827.doi:10.1016/j.jad.2024.07.030.PMID 39019223.
  242. ^abUrbano M, Guerrero M, Rosen H, Roberts E (May 2014). "Antagonists of the kappa opioid receptor".Bioorganic & Medicinal Chemistry Letters.24 (9):2021–2032.doi:10.1016/j.bmcl.2014.03.040.PMID 24690494.
  243. ^Todtenkopf MS, Dean RL, Eyerman DJ, Cunningham JI, Deaver DR (October 2014)."P.2.a.010 Buprenorphine in combination with samidorphan (ALKS 33) results in antidepressive-like effects in two distinct rat models".European Neuropsychopharmacology. Abstracts of the 27 ECNP Congress.24:S366 –S367.doi:10.1016/S0924-977X(14)70585-1.ISSN 0924-977X.
  244. ^Thase ME, Stanford AD, Memisoglu A, Martin W, Claxton A, Bodkin JA, et al. (December 2019)."Results from a long-term open-label extension study of adjunctive buprenorphine/samidorphan combination in patients with major depressive disorder".Neuropsychopharmacology.44 (13):2268–2276.doi:10.1038/s41386-019-0451-3.PMC 6897901.PMID 31254971.
  245. ^abSchmidt ME, Kezic I, Popova V, Melkote R, Van Der Ark P, Pemberton DJ, et al. (August 2024)."Efficacy and safety of aticaprant, a kappa receptor antagonist, adjunctive to oral SSRI/SNRI antidepressant in major depressive disorder: results of a phase 2 randomized, double-blind, placebo-controlled study".Neuropsychopharmacology.49 (9):1437–1447.doi:10.1038/s41386-024-01862-x.PMC 11251157.PMID 38649428.
  246. ^Neumora Therapeutics. Navacaprant phase 3 KOASTAL-1 trial results. January 2025.
  247. ^Knoll AT, Meloni EG, Thomas JB, Carroll FI, Carlezon WA (December 2007)."Anxiolytic-Like Effects of κ-Opioid Receptor Antagonists in Models of Unlearned and Learned Fear in Rats".The Journal of Pharmacology and Experimental Therapeutics.323 (3):838–845.doi:10.1124/jpet.107.127415.ISSN 0022-3565.
  248. ^Mayer FP, Stewart A, Varman DR, Moritz AE, Foster JD, Owens AW, et al. (October 2025)."Kappa opioid receptor antagonism restores phosphorylation, trafficking and behavior induced by a disease-associated dopamine transporter variant".Molecular Psychiatry.30 (10):4651–4664.doi:10.1038/s41380-025-03055-4.PMC 12436197.PMID 40442453.
  249. ^abcAbi-Dargham A (2021)."Relevance of the Kappa Dynorphin System to Schizophrenia and Its Therapeutics".Journal of Psychiatry and Brain Science.6.doi:10.20900/jpbs.20210015.PMC 8442669.PMID 34532594.
  250. ^Reed B, Butelman ER, Fry RS, Kimani R, Kreek MJ (March 2018)."Repeated Administration of Opra Kappa (LY2456302), a Novel, Short-Acting, Selective KOP-r Antagonist, in Persons with and without Cocaine Dependence".Neuropsychopharmacology.43 (4): 928.doi:10.1038/npp.2017.245.PMC 5809793.PMID 29422497.
  251. ^abMartinez D, Slifstein M, Matuskey D, Nabulsi N, Zheng MQ, Lin SF, et al. (September 2019)."Kappa-opioid receptors, dynorphin, and cocaine addiction: a positron emission tomography study".Neuropsychopharmacology.44 (10):1720–1727.doi:10.1038/s41386-019-0398-4.PMC 6785004.PMID 31026862.
  252. ^abcdMash DC, Duque L, Page B, Allen-Ferdinand K (2018)."Ibogaine Detoxification Transitions Opioid and Cocaine Abusers Between Dependence and Abstinence: Clinical Observations and Treatment Outcomes".Frontiers in Pharmacology.9 529.doi:10.3389/fphar.2018.00529.PMC 5996271.PMID 29922156.
  253. ^Glick SD, Maisonneuve IS (May 1998). "Mechanisms of antiaddictive actions of ibogaine".Annals of the New York Academy of Sciences.844 (1):214–226.Bibcode:1998NYASA.844..214G.doi:10.1111/j.1749-6632.1998.tb08237.x.PMID 9668680.
  254. ^Glick S, Rossman K, Steindorf S, Maisonneuve I, Carlson J (April 1991)."Effects and aftereffects of ibogaine on morphine self-administration in rats".European Journal of Pharmacology.195 (3):341–345.doi:10.1016/0014-2999(91)90474-5.ISSN 0014-2999.
  255. ^Cappendijk SL, Dzoljic MR (September 1993). "Inhibitory effects of ibogaine on cocaine self-administration in rats".European Journal of Pharmacology.241 (2):261–265.doi:10.1016/0014-2999(93)90212-z.ISSN 0014-2999.
  256. ^"Drugs@FDA: FDA-Approved Drugs".www.accessdata.fda.gov. Retrieved2025-10-18.
  257. ^"Kapruvia".European Medicines Agency (EMA). 2022-02-22. Retrieved2025-10-18.
  258. ^Nguyen E, Smith KM, Cramer N, Holland RA, Bleimeister IH, Flores-Felix K, et al. (July 2022)."Medullary kappa-opioid receptor neurons inhibit pain and itch through a descending circuit".Brain.145 (7):2586–2601.doi:10.1093/brain/awac189.PMC 9612802.PMID 35598161.

External links

[edit]
Neurotransmitter
Adrenergic
Purinergic
Serotonin
Other
Metabolites and
signaling molecules
Eicosanoid
Other
Peptide
Neuropeptide
Other
Miscellaneous
Taste, bitter
Orphan
Other
Adhesion
Orphan
Other
Taste, sweet
Other
Frizzled
Smoothened
G protein-coupled receptor
Hormone receptors
Hypothalamic
Pituitary
Other
Opioid receptors
Other neuropeptide receptors
Type I cytokine receptor
Enzyme-linked receptor
Other
μ-opioid
(MOR)
Agonists
(abridged;
full list)
Antagonists
δ-opioid
(DOR)
Agonists
Antagonists
κ-opioid
(KOR)
Agonists
Antagonists
Nociceptin
(NOP)
Agonists
Antagonists
Others
Retrieved from "https://en.wikipedia.org/w/index.php?title=Κ-opioid_receptor&oldid=1323450804"
Categories:
Hidden categories:

[8]ページ先頭

©2009-2025 Movatter.jp