Next Article in Journal
Eletrophilic Chemistry of Tranilast Is Involved in Its Anti-Colitic Activity via Nrf2-HO-1 Pathway Activation
Next Article in Special Issue
Opioid Dosage Levels, Concurrent Risk Factors and Self-Perceptions among Chronic Pain, Opioid-Managed Individuals at Elevated Risk for Opioid Overdose
Previous Article in Journal
Umbilical Cord-Derived Wharton’s Jelly for Regenerative Medicine Applications: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Opioid Analgesia and Opioid-Induced Adverse Effects: A Review

1
School of Pharmacy and Pharmacology, University of Tasmania, Hobart, TAS 7001, Australia
2
School of Medicine, Institute for Mental and Physical Health and Clinical Translation, Deakin University, Geelong, VIC 3216, Australia
3
Department of Biotechnology & Genetic Engineering, University of Development Alternative, Dhanmondi, Dhaka 1207, Bangladesh
4
School of Allied Health Sciences, World Union for Herbal Drug Discovery (WUHeDD) and Research Excellence Center for Innovation and Health Products (RECIHP), Walailak University, Nakhon Si Thammarat 80160, Thailand
5
Department of Clinical Tropical Medicine, Faculty of Tropical Medicine, Mahidol University, Bangkok 10400, Thailand
6
Department of Pharmaceutical Chemistry, Institute of Pharmacy and Center for Molecular Biosciences (CMBI), University of Innsbruck, Innrain 80–82, 6020 Innsbruck, Austria
7
Medical School, University of Cyprus, Nicosia 1678, Cyprus
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2021, 14(11), 1091; https://doi.org/10.3390/ph14111091
Submission received: 23 September 2021 / Revised: 26 October 2021 / Accepted: 26 October 2021 / Published: 27 October 2021
(This article belongs to the Special Issue Medicinal Chemistry and Pharmacological Activities of Opioid Drugs)

Abstract

:
Opioids are widely used as therapeutic agents against moderate to severe acute and chronic pain. Still, these classes of analgesic drugs have many potential limitations as they induce analgesic tolerance, addiction and numerous behavioural adverse effects that often result in patient non-compliance. As opium and opioids have been traditionally used as painkillers, the exact mechanisms of their adverse reactions over repeated use are multifactorial and not fully understood. Older adults suffer from cancer and non-cancer chronic pain more than younger adults, due to the physiological changes related to ageing and their reduced metabolic capabilities and thus show an increased number of adverse reactions to opioid drugs. All clinically used opioids are μ-opioid receptor agonists, and the major adverse effects are directly or potentially connected to this receptor. Multifunctional opioid ligands or peripherally restricted opioids may elicit fewer adverse effects, as shown in preclinical studies, but these results need reproducibility from further extensive clinical trials. The current review aims to overview various mechanisms involved in the adverse effects induced by opioids, to provide a better understanding of the underlying pathophysiology and, ultimately, to help develop an effective therapeutic strategy to better manage pain.

1. Introduction

1.1. Pain

According to the International Association for the Study of Pain (IASP), pain is defined as an unpleasant sensory and emotional experience associated with actual potential tissue damage or described in terms of such damage [1]. Pain can be categorised as nociceptive, neuropathic or nociplastic pain (a combination of both that cannot be entirely explained as nociceptive or neuropathic). Nociceptive pain is generated as a warning signal transmitted to the brain about the possible damage of a non-neural tissue [1,2]. In contrast, neuropathic pain usually results from damage to neural tissue by a disease, toxin or infection [1,3]. The third type of pain, the nociplastic pain, is a complex pain, not completely defined but probably caused by an alteration of neurons’ pain response and an increased sensitivity of the central nervous system (CNS). Nociceplastic pain generally presents over 3 months of duration with symptoms such as hyperalgesia and regional pain sensations and is commonly observed in patients with cancer and other long-term chronic disorders [4,5,6].
The opioid system is a physiological control system that modulates pain, emotions, immune defence and various other physiological responses. The opioid system involves the communication and coordination of a significant number of endogenous opioid peptides and several types of opioid receptors in the CNS and peripheral nervous system. This system also significantly modulates numerous sensory, emotional, motivational and cognitive functions, as well as addictive behaviours [7,8,9]. It is also involved in other physiological functions, including responses to stress, respiration, gastrointestinal transit, endocrine and immune functions [10]. These responses are orchestrated by opioid ligands that bind to specific opioid receptors to induce analgesia and behavioural effects in vivo. Therefore, to understand the pharmacological effects of specific opioids, it is first essential to clarify the specific roles of each opioid receptor type.

1.2. Opioid Receptors

The presence of opioid receptors was first proposed in 1954 [11]. However, the first evidence of the multiplicity of opioid receptors was only described in 1976 [12]. According to the International Union of Basic and Clinical Pharmacology (IUPHAR) and the British Pharmacological Society (BPS) joint IUPHAR/BPS Guide to Pharmacology, opioid receptors are classified into μ (Mu: MOP), δ (delta: DOP) and κ (kappa: KOP) receptors, as well as the non-classical nociception (NOP) receptor [13] (Table 1).
Opioid receptors belong to the family of seven-transmembrane helical G protein-coupled receptors (GPCRs) and share about 60% homology in the amino acid composition. These receptors display an extracellular N-terminus and an intracellular C-terminus and are coupled with heterotrimeric Gi/Go proteins [15,16,17]. Opioid ligands bind to opioid receptors by establishing ligand–receptor interactions in the binding pockets of the receptor, which are situated in the transmembrane helices. The binding pocket of opioid receptors can be divided into two distinct regions; the lower part (intracellular side) of the receptor is highly conserved for opioids (non-specific ‘message’ region), and the higher part of the pocket (extracellular side) contains divergent residues that confer selectivity (‘address’ region) to opioid receptor types; binding also depends on the type of the opioid ligand [18,19]. In 2012, the first molecular structures of all four opioid receptors were described in several reports [18,20,21,22].
Although all types of opioid receptor types modulate analgesia, the MOP receptor is thought to be dominant for its pain-relieving effects [23,24,25,26]. The major limitation of targeting the MOP receptor for analgesia is that it is also responsible for the induction of tolerance [27] and other undesirable adverse effects including addiction [28,29], dependence, respiratory depression [30] and constipation [31]. The MOP receptor is expressed in the brain, spinal cord and elsewhere in the body, and the adverse effects are relevant to its site of activation [28]. For example, in the gut, MOP receptor activation can cause constipation. However, the most important activation site is in the brain, as the MOP receptor drives hedonic reward, reinforcing, addictive, tolerance, dependence and withdrawal symptoms [32]. It is presumed that peripherally restricted MOP receptor agonists (that do not pass the blood–brain barrier) mediate local analgesia (effective against inflammatory or neuropathic pain) with reduced centrally mediated adverse effects [28]. MOP receptor-related adverse events are of great clinical concern and justify the characterisation of other opioid receptor types as suitable drug targets to induce analgesia. Unfortunately, the other three opioid receptor types (DOP, KOR and NOP receptors) do not have the same efficacy in mediating analgesia compared to the MOP receptor. DOP receptor agonists are generally less effective to treat acute thermal pain compared to inflammatory [33,34,35], neuropathic [36,37] and cancer-associated bone pain [38]. SNC80 and deltorphin II, two selective DOP receptor agonists, show significant anti-hyperalgesic effects, but these agonists are less potent or less efficacious in inducing thermal antinociceptive effects [34]. In addition, the use of DOP receptor agonists is limited, since DOP receptor-induced analgesia appears to require the presence of a pro-inflammatory state [39,40]. While DOP receptor agonists only produce moderate analgesia in non-human primates [41,42,43], despite being effective in rodent models of chronic pain [44], they are associated with convulsions in mice [45] and non-human primates [41,42,43]. Additionally, KOP receptor agonists are reported to reduce visceral [46,47], inflammatory [48,49] and neuropathic pain [50,51], but they also produce CNS-associated adverse events (i.e., dysphoria, psychotomimesis) [52,53,54]. While selective KOP and DOP receptor agonists lack some of the MOP receptor-mediated liabilities, such as constipation, respiratory depression and addiction, they display a side effect profile of their own [55]. Several NOP receptor agonists are reported to have antinociceptive effects in rodent [56,57] and primate models [58,59,60] and are associated with a reduced risk for abuse [61]. However, systemic administration of NOP agonists did not produce spinal analgesia in rodents [62,63], while showing efficacy after intrathecal administration in primates and rodent models of neuropathic pain [57,58,61,64]. Overall, MOP receptor agonists, despite their adverse effects, remain the most efficacious drugs in providing pain relief and are thus widely used in the clinic [26,65].
In the investigation on the role of specific opioid receptors and their ligands in pain modulation, antinociceptive tolerance and adverse behavioural effects, the generation of knockout animals has provided significant knowledge on the in vivo physiological role of the opioid system. For example, in MOP receptor knockout mice, MOP receptor agonist-induced antinociception and their associated side effects (e.g., hyperlocomotion, respiratory depression, inhibition of gastrointestinal tract transit, reward and withdrawal effects) were effectively abolished [23,66,67]. At the same time, morphine efficiently induced analgesia in DOP [68] and KOP [69] receptor knockout mice, albeit with reduced adverse effects (i.e., tolerance and withdrawal response). Similarly, KOP receptor agonists are also reported to induce analgesia in MOP [70] and DOP [68] receptor knockout mice, while predictably, in KOP receptor knockout animals, this effect was not observed [69]. However, DOP receptor agonists show only reduced levels of analgesia in DOP receptor knockout mice [68], although a mixed effect (decreased/maintained) on analgesia was observed in MOP receptor knockout mice [70,71].

1.3. Endogenous Opioid Ligands

Opioid ligands are from both endogenous and exogenous origins. Evidence of the existence of endogenous ligands for the opioid receptors was obtained in the 1970s, and the structures of [Leu]- and [Met]-enkephalin were reported in 1975 [72]. Endogenous opioid peptides are found in the CNS and peripheral nervous system and in the gastrointestinal tract [73]. These peptides are derived from the four different precursors pro-enkephalin, pro-dynorphin, pro-opiomelanocortin and prepro-nociceptin [15,74,75,76]. Pro-enkephalin contains two 267 amino acid polypeptides [77], and mainly produces the pentapeptides [Leu]- and [Met]-enkephalins [78,79] with selectivity for MOP and DOP receptors. Dynorphins are mainly big dynorphin, dynorphin A, dynorphin B and α-neo-endorphin and interact mainly with the KOP receptor [80,81]. Endorphins are derived from pro-opiomelanocortin [82] and are expressed as α-, β- and γ-endorphins [83]. While endorphins activate the MOP receptor, the prepro-nociceptin-derived neuropeptide nociceptin/orphanin FQ binds to the NOP receptor [7]. Endogenous opioids affect a multitude of physiological functions, such as pain modulation and analgesia, stress and emotional responses, tolerance and dependence, learning and memory, addiction, sexual activity and control of hormone levels, neurological disorders, eating and drinking behaviour, gastrointestinal, renal and hepatic functions, cardiovascular responses, respiration, thermoregulation and immunological responses [84,85].

1.4. Exogenous Opioid Ligands

Over more than 8000 years, the poppy plant (Papaver somniferum) and the opioids derived from it have been used for pain relief. In a Sumerian ideogram, the poppy plant was known as a “plant of joy” [85]. Crude opium admixtures were widely used in different British and German medicines in the 16th century, and effects like pain tolerance and physical dependence on opioids were noted at this time. In 1805, Friedrich Sertürner isolated morphium (morphine) and named it after the Greek god Morpheus (the “God of sleep and dreams”). Within two decades after the initial isolation of morphine, commercial production of morphine started, and morphine became available on the European market. Subsequently, after the invention of hypodermic syringes in the middle of the 19th century, morphine was injected systematically into painful areas [85].
Currently, different alkaloids extracted from the poppy plant (Papaver somniferum), including opium, morphine and codeine, are still used for pain relief, mood disorders and palliative care. In addition, several semi-synthetic and synthetic opioids, such as buprenorphine, dextropropoxyphene, hydromorphone, oxycodone, pethidine, fentanyl, methadone, tapentadol and tramadol, are widely used in patients that suffer from surgical or chronic pain [86].

1.5. Opioid Receptor-Mediated Signalling Pathways

Once an agonist binds to a GPCR, it activates a G-protein (a heterotrimeric protein composed of three different subunits (α, β and γ)). The α subunit (which can be Gs, Gi, Go, Gq) exchanges a bound guanosine diphosphate (GDP) for a guanosine triphosphate (GTP), which activates the subunit and dissociates it from the βγ subunits. The α subunit goes on to activate/inhibit different downstream signal transduction pathways. For example, it inhibits (Giα) or activates (Gsα) adenylyl cyclase, which leads to the activation or inhibition of the production of cyclic adenosine monophosphate (cAMP) from ATP [87,88]. This activation modulates voltage-gated calcium, sodium and potassium channels [89,90] and cellular levels of cAMP or the activity of protein kinase A (PKA) [91] (Figure 1). Activated PKA translocates to the nucleus, where it induces the phosphorylation of cAMP response element-binding protein (CREB) (Figure 1). Phosphorylated CREB facilitates the desired gene expression levels as the promoters contain cAMP response elements (CRE) and control different cellular functions [92] (Figure 1). Increased activity of CREB has been observed in cancer cells, chronic inflammatory or neuropathic pain conditions, and blocking CREB can inhibit cell proliferation, differentiation and survival, as well as peripheral neuropathy [93,94,95,96] (Figure 1). However, another study showed that CREB has a dual role in cell proliferation (stimulation or inhibition), and its function depends on the pathway of activation [97]. For example, cAMP-activated PKA-phosphorylated CREB stimulates mitosis (cell proliferation), but growth factor-activated CREB inhibits mitosis [97]. Opioid peptides activate membrane-bound receptors on sensory nerve fibres and produce acute analgesia by reducing the excitability of sensory neurons. Activation of the MOP receptor by both endogenous and exogenous opioids on post-synaptic neurons dissociates the Gα subunit from the Gβγ subunits of the G protein, which increases potassium (K+) conductance in neurons. The resulting efflux of K+ ions hyperpolarise the neuronal cells and reduces their excitability [98,99,100]. Opioids binding to opioid receptors also inhibits calcium (Ca2+) and sodium (Na+) ion influx, which reduces Ca2+- and Na+-induced depolarisation of neurons [65,95,101,102].
Different ligands can act through the same receptors but trigger distinct intracellular transduction pathways, and this phenomenon is termed biased agonism [103,104,105,106]. Biased agonism of the MOP receptor is nowadays investigated to avoid antinociceptive tolerance and other adverse effects, with a particular focus on the development of ligands with no β-arrestin-2 (β-arr2) recruitment [107,108,109]. Increased antinociceptive response and reduced adverse effects were observed for ligands with decreased β-arr2 recruitment or in β-arr2 knockout mice [109,110,111,112,113,114]. Biased agonists for the KOP receptor with reduced β-arr2 recruitment show antinociception and antipruritic effects, with fewer adverse effects than selective unbiased KOP receptor agonists [115,116]. Oliceridine, a MOP receptor biased ligand, displayed similar efficacy on MOP receptor binding, but showed a six-fold reduced efficacy on β-arr2 recruitment [117]. In the clinic, oliceridine was demonstrated to be an effective painkiller with a reduced side effect profile compared to morphine [117,118].

2. Opioid-Induced Adverse Effects

2.1. Analgesic Tolerance

The development of analgesic tolerance to opioids after repetitive administration is one of the major limitations for their chronic use in the clinic. Morphine is one of the most effective and widely prescribed drugs against severe pain [26,119]. However, long-term morphine treatment is discouraged in the clinic due to the risk of adverse effects, including analgesic tolerance [120,121]. Tolerance manifests as decreased drug efficacy following repeated administration [122]. Therefore, to maintain efficacy, dose increments are required, which in turn contribute to generating cellular desensitisation, tolerance, physical dependence and behavioural withdrawal symptoms. In addition, increased morphine dosing is frequently required due to amplified disease progression rather than analgesic tolerance [123].
The clinical management of analgesic tolerance involves opioid rotation and the combination of opioids with adjuvants [124,125,126,127]. Adjuvants, such as gabapentin, pregabalin, dexamethasone, naproxen, ibuprofen, carbamazepine, aspirin, venlafaxine and acetaminophen, are combined with opioid analgesics in patients that require long-term analgesic treatment [128,129]. Similarly, in preclinical studies, a combination of opioids and non-opioid adjuvants or combinations of opioid agonist and antagonist are used to prevent antinociceptive tolerance [130,131,132,133]. The activation of the opioid receptors leads to receptor phosphorylation by GPCR kinases, which promotes the interaction with β-arr [134,135]. Both phosphorylation and interaction with β-arr are required for subsequent receptor internalisation [134,135]. This internalised receptor can be proteolytically degraded. However, receptors can also be recycled in endosomes to be returned to the cell membrane [135,136]. This process is called receptor trafficking. In addition, de novo receptor synthesis ensures that new opioid receptors are produced and transported to the cell membrane via the trans-Golgi network [135]. Prolonged treatment with opioids increases the number of inactive (phosphorylated) receptors on the membrane, as well as the number of de novo synthesised receptors [135,136]. A clinical study on the use of opioids in cancer-related chronic pain showed that chronic administration of opioids induces increased methylation of the MOP receptor gene (OPRM1) on peripheral leucocytes and causes analgesic tolerance, but the article reported that a preclinical study on mice showed that targeted re-expression of the MOP receptor (by gene therapy) in cancer cells can reverse analgesic tolerance [137].
Specifically, chronic exposure to morphine leads to the selective recruitment of β-arr2 but not of β-arr1 [138]. In contrast to the interaction with β-arr1, which leads to receptor recycling, β-arr2 does not results in opioid receptor recycling but increases the number of inactive receptors on the cell membrane. This process is associated with insufficient analgesia [138]. Although the molecular mechanisms that lead to opioid tolerance are not entirely clear, both desensitisation and trafficking are assumed to be the key factors that lead to insufficient analgesia [107,135,138,139]. Although it is essential to delineate the exact molecular mechanisms resulting in opioid tolerance [138,140,141,142,143], it is also important to understand how chronic morphine dosing itself can influence analgesic tolerance and associated behavioural dependence [144,145].
The antinociceptive effects of morphine and other opioids in preclinical studies are commonly measured as central (brain and spinal cord) or peripheral antinociception [146,147]. The commonly used tail-flick test potentially measures spinal-mediated nociception, while the hot-plate assay largely measures supraspinal-mediated nociception [147,148]. Generally, one such antinociception test is performed in preclinical studies with repeated morphine treatment. As a result, the progression of antinociceptive tolerance measured by a single pain assay may be different when using another assay [145].

2.2. Addiction and Physical Dependence

Apart from analgesic tolerance, long-term opioid treatment also causes behavioural adverse effects like physical dependence and addiction to these drugs. Physical dependence consists of craving for a drug either for pleasure or to avoid the occurrence of withdrawal symptoms following a reduction of the treatment dose or the intake of an opioid receptor antagonist [149,150]. Addiction indicates a loss of control of opioids use [150]. Physical dependence is associated with the upregulation of cAMP and noradrenergic signalling in the locus coeruleus (LC) neurons of the dorsal pontine tegmentum of the brainstem [151,152,153]. The molecular mechanism that initiates physical dependence and reward is associated with repeated opioid treatment [151,154,155]. Briefly, MOP receptor binding with opioids, like morphine, causes dopamine release by dopaminergic neurons in the VTA, VTA neurons transfer dopamine to the NAc, and this induces a pleasure feeling [154] (Figure 2). After chronic intake of opioids, a larger amount of opioids is required gradually to stimulate the VTA neurons and sustain the release of a similar amount of dopamine in the NAc. Thus, patients become dependent and tend to take more drugs to feel better [154]. The LC region of the brain that controls noradrenaline release is responsible for the dependence and reward processes [154].
Endogenous opioids bind to opioid receptors in neuronal cell bodies of the LC and stimulate adenylyl cyclase to convert ATP to cAMP, but acute opioid intake, e.g., of morphine or heroine, inhibits the conversion, and as a result, less cAMP is produced. As noradrenaline release is stimulated by cAMP, less noradrenaline is released in the LC [154]. Noradrenaline stimulates wakefulness, respiration, and several other processes. Repeated opioids intake causes desensitisation of the opioid receptors; thus, neuronal cells produce a similar amount of ATP and cAMP in the presence of a higher concentration of opioids in the LC [154]. Chronic morphine administration increases the levels of type I and VII of adenylyl cyclase, PKA subunits and several phosphoproteins (e.g., CREB) and results in the hyperactivation of the cAMP pathway [152] (Figure 2). Then, if the patient stops taking opioids, this causes a massive release of noradrenaline in the LC neurons, which causes anxiety, nervousness and muscle cramps [154]. Clinical guidelines for long-term opioid use propose a “start low and go slow” dosing regimen to prevent addiction, physical dependence, overdosing or abuse [120,156,157,158,159,160]. Therefore, clinical guidelines propose administering the smallest effective dose [161], rather than aiming for adequate long-term pain relief [162].

2.3. Constipation

Constipation is a very common unwanted side effect of opioids and is caused by the activation of the MOP receptor in the enteric nervous system [163,164]. Opioids bind to MOP receptors in enteric neurons and delay gastrointestinal (GI) transit time, which also stimulates non-propulsive GI motility and pylorus and ileocecal sphincters [134]. Morphine treatment increases the expression of aquaporin-3 (AQP3) water channels in the colon by increased secretion of serotonin (5-HT), which increases water absorption from the luminal part to the vascular part of the colon [165] (Figure 3). As a result, constipation develops by increased fluid absorption from the large intestine along with less electrolyte secretion by the intestinal lumen [164] (Figure 3). In contrast, chronic morphine treatment does not produce tolerance to reduced GI motility in the lower GI tract, while it induces analgesic tolerance and leaves GI motility unaffected in the upper GI tract [166]. As a result, patients over a long-term opioid treatment continuously suffer from constipation. Constipation affects about 40% of patients with chronic oral opioid treatment, and therefore, different laxatives and non-medication approaches (e.g., fibrous diet, hydration) are used to provide comfort to the patients [149,167,168,169]. In addition, opioids combined with a low dose of opioid antagonists, such as naloxone, methylnaltrexone or alvimopan, are effective in reducing constipation without affecting pain relief and induce fewer withdrawal symptoms [169,170,171,172].

2.4. Nausea and Vomiting

Nearly 20% of patients under long-term opioid treatment experience nausea and vomiting [173]. The actual mechanism of opioid-induced nausea and vomiting is not clear, but the activation of opioid receptors (MOP or DOP) present in the chemoreceptor trigger zone, vestibular apparatus (MOP) and GI tract (MOP, DOP or KOP) is probably involved in the induction of nausea and vomiting [134]. At present, it is thought that these adverse effects are a direct consequence of opioid-induced effects in the area postrema of the brainstem, an area rich in dopamine, opioid and serotonin receptors [174,175]. In the clinic, 5-HT3 and NK1 receptor antagonists are used to prevent opioid-induced emesis, which could indicate that several non-opioid receptors (e.g., dopamine (D2), 5-HT3 and histamine (H1)) might interact with opioid receptors in those brain areas that control nausea and vomiting [134,176,177,178]. Although patients treated with oral morphine experience chronic nausea and vomiting, opioid rotation or changing the route of administration (e.g., oral to subcutaneous) appear helpful to reduce these adverse effects [168,179,180].

2.5. Respiratory Depression

Respiratory depression occurs less frequently compared to other adverse effects, but typically, it can have fatal consequences [149,181]. Similar to the other side effects, opioid-induced respiratory depression is mediated by the MOP receptor [182,183,184]. For example, fentanyl does not induce respiratory depression in MOP receptor knockout mice, indicating that the MOP receptor is responsible for respiratory depression [185]. Neurons of the pre-Bötzinger complex, a sub-region of the ventrolateral medulla, are responsible for controlling autonomic neuronal functions, including normal respiration [184]. The neurons of the pre-Bötzinger complex express a variety of receptors including neurokinin-1, serotonin (5-HT) and MOP receptors [184]. Inhibition of neurons that generate respiratory rhythms in the pre-Bötzinger complex cause respiratory depression [186] (Figure 4). MOP receptor activation inhibits adenylyl cyclase and reduces the synthesis of intracellular cAMP, which is thought to depress the respiratory neurons, as reduced cAMP levels in the cytoplasm reduces neuronal excitability by an unknown mechanism [134] (Figure 4). On the other hand, serotonin receptors in this region stimulate respiration [186,187]. The 5-HT1(a) receptors are expressed widely on respiratory neurons and are stimulated by reduced cAMP levels that activate the glycine receptor type α3 (GlyRα3) [188]. The activated GlyRα3 receptor inhibits neurons contributing to respiratory depression (Figure 4). This effect is independent of the MOP receptor-induced signal transduction pathway [188]. Therefore, multiple non-opioid receptors together with the MOP receptor are involved in the control of respiration and opioid-induced respiratory depression. Although high-dose opioid users are at risk of respiratory depression [189], a selective peripherally selective opioid antagonist can effectively reduce the incidence of respiratory depression without significant withdrawal symptoms [190].

2.6. Other Adverse Effects

In addition to analgesic tolerance, physical dependence and addiction as major adverse effects of long-term opioid treatment, this discussion also needs to address other behavioural side effects observed in the clinic [191]. Morphine-induced biphasic behavioural effects are well known from preclinical studies and include initial motor suppression and subsequent hyper-excitation [192,193,194,195,196,197,198]. An open-field arena is widely used to assess motor behaviour and typically includes horizontal movement, rearing (vertical movement) and turning behaviour. Morphine-induced horizontal locomotion, turning and circling behaviours are related to the dopaminergic system [195,199,200,201]. Morphine treatment induces the dopamine receptor-1 (D1)-dependent βarr-2/phospho-ERK (βarr2/pERK) signalling complex, which stimulates morphine-induced horizontal locomotion. However, the effects were absent in D1 and D2 receptor knockout mice [202]. Acute morphine administration induces phosphorylation of dopamine- and cAMP-regulated phosphoprotein of 32 kDa (DARPP-32), which activates the D1 receptor on dopaminergic neurons of NAc, substantia nigra and dorsal striatum and stimulates locomotor activity [202,203,204,205]. The D1, D2 and D3 receptors are responsible for the control of locomotion, learning and memory-related functions [206]. Opioids-induced activation of μ1-opioid receptor decreases gamma-aminobutyric (GABA) release in co-localised GABAA receptors in the ventral tegmental area (VTA), but stimulates dopamine release in the NAc, as the dopaminergic neurons in the NAc emerge from the VTA [207] (Figure 2). The VTA, NAc (ventral striatum) and substantia nigra dopaminergic system is also involved in motivation-, reward- and addiction-related behaviour [208,209,210,211]. Noticeably, the effects of chronic morphine treatment on cAMP/PKA/DARPP-32 signalling are not fully understood at present. However, rearing behaviour can indicate increased exploration and reduced anxiety, which are related to GABA inhibitory neurotransmission [212,213,214]. Activation of the CREB transcription factor regulates anxiety-related behaviours, as CREB-deficient mice show an increased anxiogenic response [215]. However, the behavioural changes in response to chronic morphine treatment are independent of MOP receptor, cyclin-dependent kinase 5 (cdk5) or adenylyl cyclase activities in relevant areas of the brain [216]. Studies suggested that the morphine-induced behavioural effects are probably derived from its binding to the KOP receptor but not to the MOP receptor [217,218]. The likely multiple mechanisms that link chronic morphine treatment to its behavioural effects are not completely understood but may be controlled by a combination of dopaminergic, GABAergic, opioidergic and additional unknown neuronal signals [107,202,203,204,212,213,214]. The combination of multiple independent behavioural measurements is generally regarded as the most reliable approach to assess the total motor effects induced by opioids [193,194,219,220,221].
Drowsiness, lethargy, hyperalgesia and pruritus are also common adverse effects of opioids [222,223,224]. Suppression of motor behaviour causes drowsiness and lethargy; however, the exact mechanisms of this effect are not known. Morphine is responsible for generating an itching skin sensation by signalling through spinal heteromers of opioid and itch-mediating GPCRs [225,226]. Withdrawal after chronic morphine exposure also induces histamine-induced itching or scratching responses [227]. Furthermore, morphine can induce increased pain sensation (hyperalgesia) via the MOP receptor [228]. However, the molecular mechanism of morphine-induced hyperalgesia is not well understood but might be related to the upregulation of protein kinase C gamma (PKCγ) and the N-methyl-D-aspartate (NMDA) receptor subtype NR1 in the spinal cord [229,230]. It is also thought that different MOP receptor isoforms, functional interactions with other GPCRs or opioid metabolites, such as morphine 3-glucuronide that interacts with GABA or NMDA receptors, could be responsible for this adverse effect [224,228,231].

3. Discussion and Future Directions

Activation of one opioid receptor type might likely affect the behaviour of another opioid receptor type in the same complex, such as in the form of receptor homodimers or heterodimers. Different interactions exist between two or multiple receptors [70,232,233,234], which depend on the pharmacological profile of the ligands that interact with these receptors [235,236]. The current literature suggests that most opioid receptor ligands are not extremely selective and could therefore bind to one/more off-target receptors to produce beneficial/therapeutic effects and unwanted adverse effects. For example, morphine at high doses can elicit analgesia in MOP receptor knockout mice by activating the KOP receptor [237]. Similarly, in DOP receptor knockout mice, a DOP receptor agonist effectively produces analgesia, while a non-specific opioid antagonist (naltrexone) could reverse this effect [68]. Although these data could be interpreted as non-specific interactions of the different DOP receptor agonists and antagonists, they have been interpreted as evidence of the presence of a different DOP receptor subtype [68]. However, recent studies suggest that opioid receptor subtypes may not actually exist, but these results rather reflect the presence of homo- or heteromeric receptor dimers [14]. In light of a reduced DOP receptor activity in MOP receptor knockout mice [70], it was hypothesised that the specific interaction between MOP and DOP receptors in specific neural pathways could modulate pain perception. In line with this hypothesis, co-administration of morphine and a DOP receptor antagonist induces analgesia, while surprisingly reducing tolerance in rodent models [130,131,132,133,238,239], which suggests that MOP and DOP receptor interactions regulate antinociceptive tolerance. It has to be noted that respiratory depression was not prevented under these conditions, which suggests that additional receptor interactions are likely to be involved [132].
These studies justify the approach to target two opioid receptors simultaneously to explore the molecular mechanisms that contribute to tolerance and develop alternative drug candidates with reduced risk for the development of tolerance. The challenge is to develop multi-target-specific ligands that are effective as analgesics, with a favourable side effect profile. Several strategies for the simultaneous targeting of multiple receptors can be envisaged: (i) co-administration of two selective drugs, (ii) administration of one non-receptor-selective drug or (iii) use of a single drug that specifically targets different receptors (i.e., multiple-receptors-selective ligand) [240]. Particularly, the third strategy promises clinical advantages by reducing drug–drug interactions, as well as by allowing pharmacokinetics and pharmacodynamics that will be easier to control [240,241,242].
Various new opioid receptor ligands have been designed to target two or more opioid receptor types simultaneously, and many of these are effective in producing an analgesic response in vivo. Rational drug design and structure–activity relationship studies have evaluated the pharmacology of many of those ligands that act simultaneously on two or more different opioid receptors [242,243,244,245,246] or a combination of an opioid receptor with a non-opioid receptor [247,248,249]. For example, MDAN-21 is a mixed MOP receptor agonist/DOP receptor antagonist that is 50-fold more potent than morphine and produces less tolerance [32]. This effect is likely the consequence of reduced internalisation of MOP–DOP receptor heterodimers due to the bridging of the two receptors [250]. Another recent example is KGFF09, a bifunctional G-protein-biased MOP agonist–neuropeptide FF antagonist effective against acute nociceptive and inflammatory pain and with improved acute and chronic side effects [251]. The association of the two properties within a single molecule gathers the beneficial effects of G-protein-biased MOP receptor agonism on acute side effects (respiratory depression) and those of neuropeptide FF receptor antagonism on chronic side effects (opioid-induced hyperalgesia, analgesic tolerance and withdrawal syndrome). The most advanced among bifunctional opioid ligands is cebranopadol, a MOP/NOP agonist in advanced clinical development for the treatment of acute and chronic pain [252].
Studies showed that the chronic administration of morphine increased the expression of MOP–DOP receptor heteromers in the rostral ventral medulla in the brainstem, which is responsible for the processing of nociceptive responses [142,253]. Different opioid ligands have been focused on targeting the MOP–DOP receptor dimers with the objective that the co-expression of the MOP and DOP receptors might reduce analgesic tolerance [142,254,255,256]. Some of these ligands showed selectivity for the MOP–DOP receptor dimer, as well as for the individual MOP or DOP receptors [254], but other ligands did not show selectivity towards the individual receptors [256]. MOP–DOP receptor heteromer-biased ligands can activate both opioid-mediated and β-arr2-mediated signalling. An anti-analgesic effect of the MOP–DOP heteromers was also observed [257]. The adverse effects profile of heteromer-selective ligands are not clearly known [253], but research related to opioids with selectivity for heteromers and/or individual receptors is advancing.
Altogether, all clinical available and experimental opioids have some sort of adverse effects, but their clinical application should be based on the needs of individual patients and on the measurement of the risk/benefit ratio of a particular drug. Additionally, the complexity of chronic pain syndromes requires tailored pharmacological interventions and innovative drugs to effectively and safety control pain. Several experimental drugs (e.g., opioids, ion channel inhibitors) can be administered in combination (if they prove to be safe) with clinical opioids (e.g., intrathecal ziconotide, an N-type calcium ion channel inhibitor, ω-conotoxins with intrathecal morphine) and can provide better pain relief and less adverse events than opioids alone in clinical trials [258,259]. Further extensive work on the efficacy of new opioids or combination therapies is necessary to manage opioid-related adverse effects in clinical settings.

Author Contributions

Conceptualisation, A.K.P. and N.D.; writing—original draft preparation, A.K.P.; writing—review and editing, C.M.S., M.R., M.S., N.D., N.G., P.W. and V.N.; visualisation, A.K.P.; supervision, M.S., P.W., N.G. and N.D.; funding acquisition, N.D., N.G. and P.W. All authors have read and agreed to the published version of the manuscript.

Funding

M.S. thanks the Austrian Science Fund (FWF: TRP19-B18, I2463-B21 and I4697) for the financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IASP. Pain, IASP Terminology. International Association for the Study of Pain (IASP). Available online: https://www.iasp-pain.org/resources/terminology/#pain (accessed on 26 October 2021).
  2. Woolf, C.J.; Bennett, G.J.; Doherty, M.; Dubner, R.; Kidd, B.; Koltzenburg, M.; Lipton, R.; Loeser, J.D.; Payne, R.; Torebjork, E. Towards a mechanism-based classification of pain? Pain 1998, 77, 227–229. [Google Scholar] [CrossRef]
  3. Woolf, C.J.; Mannion, R.J. Neuropathic pain: Aetiology, symptoms, mechanisms, and management. Lancet 1999, 353, 1959–1964. [Google Scholar] [CrossRef]
  4. Fitzcharles, M.-A.; Cohen, S.P.; Clauw, D.J.; Littlejohn, G.; Usui, C.; Häuser, W. Nociplastic pain: Towards an understanding of prevalent pain conditions. Lancet 2021, 397, 2098–2110. [Google Scholar] [CrossRef]
  5. Nijs, J.; Lahousse, A.; Kapreli, E.; Bilika, P.; Saraçoğlu, I.; Malfliet, A.; Coppieters, I.; De Baets, L.; Leysen, L.; Roose, E.; et al. Nociplastic pain criteria or recognition of central sensitization? Pain phenotyping in the past, present and future. J. Clin. Med. 2021, 10, 3203. [Google Scholar] [CrossRef]
  6. IASP. Central Sensitization, IASP Terminology. Available online: https://www.iasp-pain.org/resources/terminology/?navItemNumber=576#Centralsensitization (accessed on 25 August 2021).
  7. Cai, Z.; Ratka, A. Opioid system and Alzheimer’s disease. Neuromol. Med. 2012, 14, 91–111. [Google Scholar] [CrossRef] [PubMed]
  8. Melzack, R. From the gate to the neuromatrix. Pain 1999, 82, S121–S126. [Google Scholar] [CrossRef]
  9. Holden, J.E.; Jeong, Y.; Forrest, J.M. The endogenous opioid system and clinical pain management. AACN Clin. Issues Adv. Pr. Acute Crit. Care 2005, 16, 291–301. [Google Scholar] [CrossRef] [PubMed]
  10. Le Merrer, J.; Becker, J.A.; Befort, K.; Kieffer, B.L. Reward processing by the opioid system in the brain. Physiol. Rev. 2009, 89, 1379–1412. [Google Scholar] [CrossRef] [PubMed]
  11. Beckett, A.H.; Casy, A.F. Synthetic analgesics: Stereochemical considerations. J. Pharm. Pharmacol. 2011, 6, 986–1001. [Google Scholar] [CrossRef]
  12. Martin, W.R.; Eades, C.G.; Thompson, J.A.; Huppler, R.E.; Gilbert, P.E. The effects of morphine-and nalorphine-like drugs in the nondependent and morphine-dependent chronic spinal dog. J. Pharmacol. Exp. Ther. 1976, 197, 517–532. [Google Scholar] [PubMed]
  13. Borsodi, A.; Bruchas, M.; Caló, G.; Chavkin, C.; Christie, M.J.; Civelli, O.; Connor, M.; Cox, B.M.; Devi, L.A.; Evans, C.; et al. Opioid Receptors, Introduction. IUPHAR/BPS Guide to PHARMACOLOGY. Available online: http://www.guidetoimmunopharmacology.org/GRAC/FamilyIntroductionForward?familyId=50 (accessed on 26 October 2021).
  14. Alexander, S.P.; Benson, H.E.; Faccenda, E.; Pawson, A.J.; Sharman, J.; Spedding, M.; Peters, J.A.; Harmar, A.J.; Collaborators, C. The concise guide to PHARMACOLOGY 2013/14: G Protein-Coupled receptors. Br. J. Pharmacol. 2013, 170, 1459–1581. [Google Scholar] [CrossRef] [PubMed]
  15. Waldhoer, M.; Bartlett, S.E.; Whistler, J.L. Opioid receptors. Annu. Rev. Biochem. 2004, 73, 953–990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chen, Y.; Mestek, A.; Liu, J.; Yu, L. Molecular cloning of a rat kappa opioid receptor reveals sequence similarities to the mu and delta opioid receptors. Biochem. J. 1993, 295, 625–628. [Google Scholar] [CrossRef] [PubMed]
  17. Mollereau, C.; Parmentier, M.; Mailleux, P.; Butour, J.-L.; Moisand, C.; Chalon, P.; Caput, D.; Vassart, G.; Meunier, J.-C. ORL1, a novel member of the opioid receptor family. FEBS Lett. 1994, 341, 33–38. [Google Scholar] [CrossRef]
  18. Granier, S.; Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Weis, W.; Kobilka, B.K. Structure of the δ-opioid receptor bound to naltrindole. Nat. Cell Biol. 2012, 485, 400–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Filizola, M.; Devi, L.A. How opioid drugs bind to receptors. Nat. Cell Biol. 2012, 485, 314–317. [Google Scholar] [CrossRef] [PubMed]
  20. Manglik, A.; Kruse, A.C.; Kobilka, T.S.; Thian, F.S.; Mathiesen, J.M.; Sunahara, R.K.; Pardo, L.; Weis, W.; Kobilka, B.K.; Granier, S. Crystal structure of the µ-opioid receptor bound to a morphinan antagonist. Nat. Cell Biol. 2012, 485, 321–326. [Google Scholar] [CrossRef] [Green Version]
  21. Thompson, A.A.; Liu, W.; Chun, E.; Katritch, V.; Wu, H.; Vardy, E.; Huang, X.-P.; Trapella, C.; Guerrini, R.; Calo, G.; et al. Structure of the nociceptin/orphanin FQ receptor in complex with a peptide mimetic. Nat. Cell Biol. 2012, 485, 395–399. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, H.; Wacker, D.; Mileni, M.; Katritch, V.; Han, G.W.; Vardy, E.; Liu, W.; Thompson, A.A.; Huang, X.-P.; Carroll, F.I.; et al. Structure of the human κ-opioid receptor in complex with JDTic. Nat. Cell Biol. 2012, 485, 327–332. [Google Scholar] [CrossRef]
  23. Matthes, H.W.D.; Maldonado, R.; Simonin, F.; Valverde, O.; Slowe, S.; Kitchen, I.; Befort, K.; Dierich, A.; Le Meur, M.; Dollé, P.; et al. Loss of morphine-induced analgesia, reward effect and withdrawal symptoms in mice lacking the µ-opioid-receptor gene. Nat. Cell Biol. 1996, 383, 819–823. [Google Scholar] [CrossRef] [PubMed]
  24. Kieffer, B.L. Opioids: First lessons from knockout mice. Trends Pharmacol. Sci. 1999, 20, 19–26. [Google Scholar] [CrossRef]
  25. Sora, I.; Takahashi, N.; Funada, M.; Ujike, H.; Revay, R.S.; Donovan, D.M.; Miner, L.L.; Uhl, G.R. Opiate receptor knockout mice define mu receptor roles in endogenous nociceptive responses and morphine-induced analgesia. Proc. Natl. Acad. Sci. USA 1997, 94, 1544–1549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Corbett, A.D.; Henderson, G.; McKnight, A.T.; Paterson, S.J. 75 years of opioid research: The exciting but vain quest for the Holy Grail. Br. J. Pharmacol. 2006, 147, S153–S162. [Google Scholar] [CrossRef] [Green Version]
  27. Williams, J.T.; Ingram, S.L.; Henderson, G.; Chavkin, C.; Von Zastrow, M.; Schulz, S.; Koch, T.; Evans, C.J.; Christie, M.J. Regulation of µ-opioid receptors: Desensitization, phosphorylation, internalization, and tolerance. Pharmacol. Rev. 2013, 65, 223–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Machelska, H.; Celik, M.Ö. Advances in achieving opioid analgesia without side effects. Front. Pharmacol. 2018, 9, 1388. [Google Scholar] [CrossRef]
  29. Darcq, E.; Kieffer, B.L. Opioid receptors: Drivers to addiction? Nat. Rev. Neurosci. 2018, 19, 499–514. [Google Scholar] [CrossRef]
  30. Romberg, R.; Sarton, E.; Teppema, L.; Matthes, H.W.; Kieffer, B.L.; Dahan, A. Comparison of morphine-6-glucuronide and morphine on respiratory depressant and antinociceptive responses in wild type and mu-opioid receptor deficient mice. Br. J. Anaesth. 2003, 91, 862–870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Mori, T.; Shibasaki, Y.; Matsumoto, K.; Shibasaki, M.; Hasegawa, M.; Wang, E.; Masukawa, D.; Yoshizawa, K.; Horie, S.; Suzuki, T. Mechanisms that underlie μ-opioid receptor agonist–induced constipation: Differential involvement of μ-opioid receptor sites and responsible regions. J. Pharmacol. Exp. Ther. 2013, 347, 91–99. [Google Scholar] [CrossRef]
  32. Daniels, D.J.; Lenard, N.R.; Etienne, C.L.; Law, P.-Y.; Roerig, S.C.; Portoghese, P.S. Opioid-induced tolerance and dependence in mice is modulated by the distance between pharmacophores in a bivalent ligand series. Proc. Natl. Acad. Sci. USA 2005, 102, 19208–19213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Pradhan, A.A.A.; Walwyn, W.; Nozaki, C.; Filliol, D.; Erbs, E.; Matifas, A.; Evans, C.; Kieffer, B.L. Ligand-directed trafficking of the-opioid receptor in vivo: Two paths toward analgesic tolerance. J. Neurosci. 2010, 30, 16459–16468. [Google Scholar] [CrossRef]
  34. Fraser, G.L.; Gaudreau, G.-A.; Clarke, P.B.S.; Ménard, D.P.; Perkins, M.N. Antihyperalgesic effects of δ opioid agonists in a rat model of chronic inflammation. Br. J. Pharmacol. 2000, 129, 1668–1672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Cahill, C.M.; Morinville, A.; Hoffert, C.; O’Donnell, D.; Beaudet, A. Up-regulation and trafficking of delta opioid receptor in a model of chronic inflammation: Implications for pain control. Pain 2003, 101, 199–208. [Google Scholar] [CrossRef]
  36. Gaveriaux-Ruff, C.; Nozaki, C.; Nadal, X.; Hever, X.C.; Weibel, R.; Matifas, A.; Reiss, D.; Filliol, D.; Nassar, M.A.; Wood, J.N.; et al. Genetic ablation of delta opioid receptors in nociceptive sensory neurons increases chronic pain and abolishes opioid analgesia. Pain 2011, 152, 1238–1248. [Google Scholar] [CrossRef] [PubMed]
  37. Kabli, N.; Cahill, C.M. Anti-Allodynic effects of peripheral delta opioid receptors in neuropathic pain. Pain 2007, 127, 84–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Brainin-Mattos, J.; Smith, N.D.; Malkmus, S.; Rew, Y.; Goodman, M.; Taulane, J.; Yaksh, T.L. Cancer-related bone pain is attenuated by a systemically available δ-opioid receptor agonist. Pain 2006, 122, 174–181. [Google Scholar] [CrossRef] [PubMed]
  39. Vanderah, T.W. Delta and kappa opioid receptors as suitable drug targets for pain. Clin. J. Pain 2010, 26, S10–S15. [Google Scholar] [CrossRef] [PubMed]
  40. Gendron, L.; Mittal, N.; Beaudry, H.; Walwyn, W. Recent advances on the δ opioid receptor: From trafficking to function. Br. J. Pharmacol. 2015, 172, 403–419. [Google Scholar] [CrossRef] [Green Version]
  41. Dykstra, L.A.; Schoenbaum, G.M.; Yarbrough, J.; McNutt, R.; Chang, K.J. A novel delta opioid agonist, BW373U86, in squirrel monkeys responding under a schedule of shock titration. J. Pharmacol. Exp. Ther. 1993, 267, 875–882. [Google Scholar] [PubMed]
  42. Negus, S.S.; Butelman, E.R.; Chang, K.J.; DeCosta, B.; Winger, G.; Woods, J.H. Behavioral effects of the systemically active delta opioid agonist BW373U86 in rhesus monkeys. J. Pharmacol. Exp. Ther. 1994, 270, 1025–1034. [Google Scholar]
  43. Negus, S.S.; Gatch, M.; Mello, N.K.; Zhang, X.; Rice, K. Behavioral effects of the delta-selective opioid agonist SNC80 and related compounds in rhesus monkeys. J. Pharmacol. Exp. Ther. 1998, 286, 362–375. [Google Scholar]
  44. Nozaki, C.; Le Bourdonnec, B.; Reiss, D.; Windh, R.T.; Little, P.J.; Dolle, R.E.; Kieffer, B.L.; Gaveriaux-Ruff, C. δ-Opioid mechanisms for ADL5747 and ADL5859 effects in mice: Analgesia, locomotion, and receptor internalization. J. Pharmacol. Exp. Ther. 2012, 342, 799–807. [Google Scholar] [CrossRef] [Green Version]
  45. Broom, D.C.; Nitsche, J.F.; Pintar, J.E.; Rice, K.C.; Woods, J.H.; Traynor, J.R. Comparison of receptor mechanisms and efficacy requirements for δ-agonist-induced convulsive activity and antinociception in mice. J. Pharmacol. Exp. Ther. 2002, 303, 723–729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Eisenach, J.C.; Carpenter, R.; Curry, R. Analgesia from a peripherally active κ-opioid receptor agonist in patients with chronic pancreatitis. Pain 2003, 101, 89–95. [Google Scholar] [CrossRef]
  47. Riviere, P. Peripheral kappa-opioid agonists for visceral pain. Br. J. Pharmacol. 2004, 141, 1331–1334. [Google Scholar] [CrossRef] [Green Version]
  48. Moon, S.W.; Park, E.H.; Suh, H.R.; Ko, D.H.; Kim, Y.I.; Han, H.C. The contribution of activated peripheral kappa opioid receptors (kORs) in the inflamed knee joint to anti-nociception. Brain Res. 2016, 1648, 11–18. [Google Scholar] [CrossRef]
  49. Bileviciute-Ljungar, I.; Spetea, M. Contralateral but not systemic administration of the κ-opioid agonist U-50,488H induces anti-nociception in acute hindpaw inflammation in rats. Br. J. Pharmacol. 2001, 132, 252–258. [Google Scholar] [CrossRef] [PubMed]
  50. Keita, H.; Kayser, V.; Guilbaud, G. Antinociceptive effect of a kappa-opioid receptor agonist that minimally crosses the blood-brain barrier (ICI 204448) in a rat model of mononeuropathy. Eur. J. Pharmacol. 1995, 277, 275–280. [Google Scholar] [CrossRef]
  51. Bileviciute-Ljungar, I.; Spetea, M. Contralateral, ipsilateral and bilateral treatments with the κ-opioid receptor agonist U-50,488H in mononeuropathic rats. Eur. J. Pharmacol. 2004, 494, 139–146. [Google Scholar] [CrossRef] [PubMed]
  52. Dykstra, L.A.; Gmerek, D.E.; Winger, G.; Woods, J.H. Kappa opioids in rhesus monkeys. I. Diuresis, sedation, analgesia and discriminative stimulus effects. J. Pharmacol. Exp. Ther. 1987, 242, 413–420. [Google Scholar]
  53. Butelman, E.R.; Harris, T.J.; Kreek, M.-J. Effects of E-2078, a stable dynorphin A(1-8) analog, on sedation and serum prolactin levels in rhesus monkeys. Psychopharmacology 1999, 147, 73–80. [Google Scholar] [CrossRef] [PubMed]
  54. Ko, M.C.H.; Johnson, M.D.; Butelman, E.R.; Willmont, K.J.; Mosberg, H.I.; Woods, J.H. Intracisternal nor-binaltorphimine distinguishes central and peripheral kappa-opioid antinociception in rhesus monkeys. J. Pharmacol. Exp. Ther. 1999, 291, 1113–1120. [Google Scholar] [PubMed]
  55. Peng, X.; Neumeyer, J.L. Kappa receptor bivalent ligands. Curr. Top. Med. Chem. 2007, 7, 363–373. [Google Scholar] [CrossRef] [PubMed]
  56. Calo’, G.; Rizzi, A.; Cifani, C.; Di Bonaventura, M.V.M.; Regoli, D.; Massi, M.; Salvadori, S.; Lambert, D.G.; Guerrini, R. UFP-112 a potent and long-lasting agonist selective for the nociceptin/orphanin FQ receptor. CNS Neurosci. Ther. 2010, 17, 178–198. [Google Scholar] [CrossRef] [PubMed]
  57. Rizzi, A.; Spagnolo, B.; Wainford, R.; Fischetti, C.; Guerrini, R.; Marzola, G.; Baldisserotto, A.; Salvadori, S.; Regoli, D.; Kapusta, D.R.; et al. In vitro and in vivo studies on UFP-112, a novel potent and long lasting agonist selective for the nociceptin/orphanin FQ receptor. Peptides 2007, 28, 1240–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Hu, E.; Calò, G.; Guerrini, R.; Ko, M.-C. Long-lasting antinociceptive spinal effects in primates of the novel nociceptin/orphanin FQ receptor agonist UFP-112. Pain 2010, 148, 107–113. [Google Scholar] [CrossRef] [Green Version]
  59. Ko, M.-C.; Woods, J.H.; Fantegrossi, W.E.; Galuska, C.M.; Wichmann, J.; Prinssen, E.P. Behavioral effects of a synthetic agonist selective for nociceptin/orphanin FQ peptide receptors in monkeys. Neuropsychopharmacology 2009, 34, 2088–2096. [Google Scholar] [CrossRef]
  60. Varty, G.B.; Lu, S.X.; Morgan, C.A.; Cohen-Williams, M.E.; Hodgson, R.A.; Smith-Torhan, A.; Zhang, H.; Fawzi, A.B.; Graziano, M.P.; Ho, G.D.; et al. The anxiolytic-like effects of the novel, orally active nociceptin opioid receptor agonist 8-[bis(2-methylphenyl)methyl]-3-phenyl-8-azabicyclo[3.2.1]octan-3-ol (SCH 221510). J. Pharmacol. Exp. Ther. 2008, 326, 672–682. [Google Scholar] [CrossRef] [Green Version]
  61. Lin, A.P.; Ko, M.-C. The therapeutic potential of nociceptin/orphanin FQ receptor agonists as analgesics without abuse liability. ACS Chem. Neurosci. 2012, 4, 214–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Dautzenberg, F.M.; Wichmann, J.; Higelin, J.; Py-Lang, G.; Kratzeisen, C.; Malherbe, P.; Kilpatrick, G.; Jenck, F. Pharmacological characterization of the novel nonpeptide orphanin FQ/nociceptin receptor agonist Ro 64-6198: Rapid and reversible desensitization of the ORL1 receptor in vitro and lack of tolerance in vivo. J. Pharmacol. Exp. Ther. 2001, 298, 812–819. [Google Scholar]
  63. Reiss, D.; Wichmann, J.; Tekeshima, H.; Kieffer, B.L.; Ouagazzal, A.-M. Effects of nociceptin/orphanin FQ receptor (NOP) agonist, Ro64-6198, on reactivity to acute pain in mice: Comparison to morphine. Eur. J. Pharmacol. 2008, 579, 141–148. [Google Scholar] [CrossRef]
  64. Obara, I.; Przewlocki, R.; Przewlocka, B. Spinal and local peripheral antiallodynic activity of Ro64-6198 in neuropathic pain in the rat. Pain 2005, 116, 17–25. [Google Scholar] [CrossRef] [PubMed]
  65. Pathan, H.; Williams, J. Basic opioid pharmacology: An update. Br. J. Pain 2012, 6, 11–16. [Google Scholar] [CrossRef] [Green Version]
  66. Schuller, A.G.; King, M.A.; Zhang, J.; Bolan, E.; Pan, Y.; Morgan, D.; Chang, A.; Czick, M.E.; Unterwald, E.M.; Pasternak, G.W.; et al. Retention of heroin and morphine–6β–glucuronide analgesia in a new line of mice lacking exon 1 of MOR–1. Nat. Neurosci. 1999, 2, 151–156. [Google Scholar] [CrossRef] [PubMed]
  67. Loh, H.H.; Liu, H.C.; Cavalli, A.; Yang, W.; Chen, Y.F.; Wei, L.N. mu Opioid receptor knockout in mice: Effects on ligand-induced analgesia and morphine lethality. Brain Res. Mol. Brain Res. 1998, 54, 321–326. [Google Scholar] [CrossRef]
  68. Zhu, Y.; King, M.A.; Schuller, A.G.; Nitsche, J.F.; Reidl, M.; Elde, R.P.; Unterwald, E.; Pasternak, G.W.; Pintar, J.E. Retention of supraspinal delta-like analgesia and loss of morphine tolerance in δ opioid receptor knockout mice. Neuron 1999, 24, 243–252. [Google Scholar] [CrossRef] [Green Version]
  69. Simonin, F.; Valverde, O.; Smadja, C.; Slowe, S.; Kitchen, I.; Dierich, A.; Le Meur, M.; Roques, B.P.; Maldonado, R.; Kieffer, B.L. Disruption of the kappa-opioid receptor gene in mice enhances sensitivity to chemical visceral pain, impairs pharmacological actions of the selective kappa-agonist U-50,488H and attenuates morphine withdrawal. EMBO J. 1998, 17, 886–897. [Google Scholar] [CrossRef] [Green Version]
  70. Matthes, H.W.; Smadja, C.; Valverde, O.; Vonesch, J.L.; Foutz, A.S.; Boudinot, E.; Denavit-Saubie, M.; Severini, C.; Negri, L.; Roques, B.P.; et al. Activity of the delta-opioid receptor is partially reduced, whereas activity of the kappa-receptor is maintained in mice lacking the mu-receptor. J. Neurosci. 1998, 18, 7285–7295. [Google Scholar] [CrossRef] [PubMed]
  71. Hosohata, Y.; Vanderah, T.W.; Burkey, T.H.; Ossipov, M.H.; Kovelowski, C.J.; Sora, I.; Uhl, G.R.; Zhang, X.; Rice, K.C.; Roeske, W.R.; et al. Delta-opioid receptor agonists produce antinociception and [35S]GTPgammaS binding in mu receptor knockout mice. Eur. J. Pharmacol. 2000, 388, 241–248. [Google Scholar] [CrossRef]
  72. Lefebvre, R. Opioid peptides and their receptors. In Comparative Veterinary Pharmacology, Toxicology and Theraphy; Van Miert, A.S.J.P.A.M., Bogaert, M.G., Debackere, M., Eds.; Springer: Berlin/Heidelberg, Germany, 1986; pp. 447–453. [Google Scholar] [CrossRef]
  73. Bergström, J.; Ahmed, M.; Li, J.; Ahmad, T.; Kreicbergs, A.; Spetea, M. Opioid peptides and receptors in joint tissues: Study in the rat. J. Orthop. Res. 2006, 24, 1193–1199. [Google Scholar] [CrossRef]
  74. Callahan, P.; Pasternak, G.W. Opiates, opioid peptides, and their receptors. J. Cardiothorac. Anesth. 1987, 1, 569–576. [Google Scholar] [CrossRef]
  75. Chaturvedi, K. Opioid peptides, opioid receptors and mechanism of down regulation. Indian J. Exp. Boil. 2003, 41, 5–13. [Google Scholar]
  76. Snyder, S.H.; Childers, S.R. Opiate receptors and opioid peptides. Annu. Rev. Neurosci. 1979, 2, 35–64. [Google Scholar] [CrossRef] [PubMed]
  77. Marino, R.; Struck, J.; Hartmann, O.; Maisel, A.S.; Rehfeldt, M.; Magrini, L.; Melander, O.; Bergmann, A.; Di Somma, S. Diagnostic and short-term prognostic utility of plasma pro-enkephalin (pro-ENK) for acute kidney injury in patients admitted with sepsis in the emergency department. J. Nephrol. 2015, 28, 717–724. [Google Scholar] [CrossRef] [PubMed]
  78. Roberts, E.; Shoureshi, P.; Kozak, K.; Szynskie, L.; Baron, A.; Lecaude, S.; Dores, R.M. Tracking the evolution of the proenkephalin gene in tetrapods. Gen. Comp. Endocrinol. 2007, 153, 189–197. [Google Scholar] [CrossRef] [PubMed]
  79. Gonzalez Nunez, V.; Gonzalez Sarmiento, R.; Rodriguez, R.E. Characterization of zebrafish proenkephalin reveals novel opioid sequences. Brain Res. Mol. Brain Res. 2003, 114, 31–39. [Google Scholar] [CrossRef]
  80. Merg, F.; Filliol, D.; Usynin, I.; Bazov, I.; Bark, N.; Hurd, Y.L.; Yakovleva, T.; Kieffer, B.L.; Bakalkin, G. Big dynorphin as a putative endogenous ligand for the kappa-opioid receptor. J. Neurochem. 2006, 97, 292–301. [Google Scholar] [CrossRef]
  81. Gein, S. Dynorphins in regulation of immune system functions. Biochemistry 2014, 79, 397–405. [Google Scholar] [CrossRef] [PubMed]
  82. Diamant, M.; Henricks, P.A.; Nijkamp, F.P.; de Wied, D. Beta-endorphin and related peptides suppress phorbol myristate acetate-induced respiratory burst in human polymorphonuclear leukocytes. Life Sci. 1989, 45, 1537–1545. [Google Scholar] [CrossRef]
  83. Lolait, S.J.; Clements, J.A.; Markwick, A.J.; Cheng, C.; McNally, M.; Smith, I.; Funder, J.W. Pro-opiomelanocortin messenger ribonucleic acid and posttranslational processing of beta endorphin in spleen macrophages. J. Clin. Investig. 1986, 77, 1776–1779. [Google Scholar] [CrossRef] [Green Version]
  84. Bodnar, R.J. Endogenous opiates and behavior: 2012. Peptides 2013, 50, 55–95. [Google Scholar] [CrossRef]
  85. Fine, P.G.; Portenoy, R.K. A Clinical Guide to Opioid Analgesia; The McGraw-Hill Companies: New York, NY, USA, 2004; pp. 1–7. [Google Scholar]
  86. AMH. Analgesics (Ch 3). In Australian Medicines Handbook (AMH). Available online: https://amhonline.amh.net.au/chapters/chap-03?menu=vertical (accessed on 22 October 2017).
  87. Pasternak, G.W. Opiate pharmacology and relief of pain. J. Clin. Oncol. 2014, 32, 1655–1661. [Google Scholar] [CrossRef] [Green Version]
  88. Pena, D.A.; Duarte, M.L.; Pramio, D.T.; Devi, L.A.; Schechtman, D. Exploring morphine-triggered PKC-targets and their interaction with signaling pathways leading to pain via TrkA. Proteomes 2018, 6, 39. [Google Scholar] [CrossRef] [Green Version]
  89. Law, P.Y. Opioid receptor signal transduction mechanism. In The Opiate Receptors, 2nd ed.; Pasternak, G.W., Ed.; Humana Press: New York, NY, USA, 2011; pp. 195–238. [Google Scholar]
  90. Witkowski, G.; Szulczyk, P. Opioid μ receptor activation inhibits sodium currents in prefrontal cortical neurons via a protein kinase A- and C-dependent mechanism. Brain Res. 2006, 1094, 92–106. [Google Scholar] [CrossRef] [PubMed]
  91. Seseã, A.E.; Vega, R.; Soto, E.; Seseña, E. Activation of μ-opioid receptors inhibits calcium-currents in the vestibular afferent neurons of the rat through a cAMP dependent mechanism. Front. Cell. Neurosci. 2014, 8, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Altarejos, J.Y.; Montminy, M. CREB and the CRTC co-activators: Sensors for hormonal and metabolic signals. Nat. Rev. Mol. Cell Biol. 2011, 12, 141–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Lei, H.; Zhang, Y.; Huang, L.; Xu, S.; Li, J.; Yang, L.; Wang, L.; Xing, C.; Wang, X.; Peng, Y. L-3-n-Butylphthalide regulates proliferation, migration, and differentiation of neural stem cell in vitro and promotes neurogenesis in APP/PS1 mouse model by regulating BDNF/TrkB/CREB/Akt pathway. Neurotox. Res. 2018, 34, 477–488. [Google Scholar] [CrossRef] [PubMed]
  94. Xia, Y.; Zhan, C.; Feng, M.; Leblanc, M.; Ke, E.; Yeddula, N.; Verma, I.M. Targeting CREB Pathway Suppresses Small Cell Lung Cancer. Mol. Cancer Res. 2018, 16, 825–832. [Google Scholar] [CrossRef] [Green Version]
  95. Smith, T.H.; Grider, J.R.; Dewey, W.L.; Akbarali, H.I. Morphine decreases enteric neuron excitability via inhibition of sodium channels. PLoS ONE 2012, 7, e45251. [Google Scholar] [CrossRef] [Green Version]
  96. Shirahama, M.; Ushio, S.; Egashira, N.; Yamamoto, S.; Sada, H.; Masuguchi, K.; Kawashiri, T.; Oishi, R. Inhibition of Ca2+/Calmodulin-Dependent protein kinase II reverses oxaliplatin-induced mechanical allodynia in rats. Mol. Pain 2012, 8, 26. [Google Scholar] [CrossRef] [Green Version]
  97. Hudson, C.; Kimura, T.E.; Duggirala, A.; Sala-Newby, G.B.; Newby, A.C.; Bond, M. Dual Role of CREB in the regulation of VSMC Proliferation: Mode of activation determines pro-or anti-mitogenic function. Sci. Rep. 2018, 8, 4904. [Google Scholar] [CrossRef] [PubMed]
  98. Chen, Y.L.; Law, P.-Y.; Loh, H.H. The other side of the opioid story: Modulation of cell growth and survival signaling. Curr. Med. Chem. 2008, 15, 772–778. [Google Scholar] [CrossRef] [PubMed]
  99. Knapp, C.M. Opiates. In Encyclopedia of the Human Brain; Academic Press: New York, NY, USA, 2002; pp. 729–739. [Google Scholar]
  100. Winters, B.; Gregoriou, G.C.; Kissiwaa, S.A.; Wells, O.A.; Medagoda, D.I.; Hermes, S.M.; Burford, N.T.; Alt, A.; Aicher, S.A.; Bagley, E.E. Endogenous opioids regulate moment-to-moment neuronal communication and excitability. Nat. Commun. 2017, 8, 14611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Ghelardini, C.; Di Cesare Mannelli, L.; Bianchi, E. The pharmacological basis of opioids. Clin. Cases Miner. Bone Metab. 2015, 12, 219–221. [Google Scholar] [CrossRef] [PubMed]
  102. Hashimoto, K.; Amano, T.; Kasakura, A.; Uhl, G.R.; Sora, I.; Sakai, N.; Kuzumaki, N.; Suzuki, T.; Narita, M. μ-Opioid receptor-independent fashion of the suppression of sodium currents by μ-opioid analgesics in thalamic neurons. Neurosci. Lett. 2009, 453, 62–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Thompson, G.L.; Lane, J.R.; Coudrat, T.; Sexton, P.; Christopoulos, A.; Canals, M. Biased agonism of endogenous opioid peptides at the μ-Opioid receptor. Mol. Pharmacol. 2015, 88, 335–346. [Google Scholar] [CrossRef] [Green Version]
  104. Wisler, J.W.; Xiao, K.; Thomsen, A.; Lefkowitz, R.J. Recent developments in biased agonism. Curr. Opin. Cell Biol. 2014, 27, 18–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Kenakin, T. Biased agonism. F1000 Biol. Rep. 2009, 1, 87. [Google Scholar] [CrossRef]
  106. Kenakin, T.; Watson, C.; Muniz-Medina, V.; Christopoulos, A.; Novick, S. A simple method for quantifying functional selectivity and agonist bias. ACS Chem. Neurosci. 2012, 3, 193–203. [Google Scholar] [CrossRef] [Green Version]
  107. Al-Hasani, R.; Bruchas, M.R. Molecular mechanisms of opioid receptor-dependent signaling and behavior. Anesthesiology 2011, 115, 1363–1381. [Google Scholar] [CrossRef] [Green Version]
  108. Molinari, P.; Vezzi, V.; Sbraccia, M.; Grò, C.; Riitano, D.; Ambrosio, C.; Casella, I.; Costa, T. Morphine-like opiates selectively antagonize receptor-arrestin interactions. J. Biol. Chem. 2010, 285, 12522–12535. [Google Scholar] [CrossRef] [Green Version]
  109. Bohn, L.M.; Lefkowitz, R.J.; Gainetdinov, R.R.; Peppel, K.; Caron, M.G.; Lin, F.T. Enhanced morphine analgesia in mice lacking beta-arrestin 2. Science 1999, 286, 2495–2498. [Google Scholar] [CrossRef]
  110. Soergel, D.G.; Subach, R.A.; Burnham, N.; Lark, M.W.; James, I.E.; Sadler, B.M.; Skobieranda, F.; Violin, J.D.; Webster, L.R. Biased agonism of the μ-opioid receptor by TRV130 increases analgesia and reduces on-target adverse effects versus morphine: A randomized, double-blind, placebo-controlled, crossover study in healthy volunteers. Pain 2014, 155, 1829–1835. [Google Scholar] [CrossRef] [PubMed]
  111. DeWire, S.; Yamashita, D.S.; Rominger, D.H.; Liu, G.; Cowan, C.L.; Graczyk, T.M.; Chen, X.-T.; Pitis, P.M.; Gotchev, D.; Yuan, C.; et al. A G Protein-Biased Ligand at the μ-Opioid receptor is potently analgesic with reduced gastrointestinal and respiratory dysfunction compared with morphine. J. Pharmacol. Exp. Ther. 2013, 344, 708–717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Raehal, K.M.; Walker, J.K.L.; Bohn, L. Morphine Side Effects in β-Arrestin 2 Knockout Mice. J. Pharmacol. Exp. Ther. 2005, 314, 1195–1201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Groer, C.E.; Tidgewell, K.; Moyer, R.A.; Harding, W.W.; Rothman, R.B.; Prisinzano, T.E.; Bohn, L.M. An opioid agonist that does not induce μ-Opioid receptor—arrestin interactions or receptor internalization. Mol. Pharmacol. 2007, 71, 549–557. [Google Scholar] [CrossRef] [Green Version]
  114. Maguma, H.T.; Dewey, W.L.; Akbarali, H.I. Differences in the characteristics of tolerance to μ-opioid receptor agonists in the colon from wild type and β-arrestin2 knockout mice. Eur. J. Pharmacol. 2012, 685, 133–140. [Google Scholar] [CrossRef] [Green Version]
  115. Spetea, M.; Eans, S.O.; Ganno, M.L.; Lantero, A.; Mairegger, M.; Toll, L.; Schmidhammer, H.; McLaughlin, J.P. Selective κ receptor partial agonist HS666 produces potent antinociception without inducing aversion after i.c.v. administration in mice. Br. J. Pharmacol. 2017, 174, 2444–2456. [Google Scholar] [CrossRef]
  116. Morgenweck, J.; Frankowski, K.J.; Prisinzano, T.; Aubé, J.; Bohn, L.M. Investigation of the role of βarrestin2 in kappa opioid receptor modulation in a mouse model of pruritus. Neuropharmacology 2015, 99, 600–609. [Google Scholar] [CrossRef] [Green Version]
  117. Jin, Z.; Zhu, M.; Gupta, A.; Page, C.; Gan, T.J.; Bergese, S.D. Evaluating oliceridine as a treatment option for moderate to severe acute post-operative pain in adults. Expert Opin. Pharmacother. 2021, 1–9. [Google Scholar] [CrossRef]
  118. Hammer, G.B.; Khanna, A.K.; Michalsky, C.; Wase, L.; Demitrack, M.A.; Little, R.; Fossler, M.J.; Ayad, S. Oliceridine exhibits improved tolerability compared to morphine at equianalgesic conditions: Exploratory analysis from two phase 3 randomized placebo and active controlled trials. Pain Ther. 2021, 1–11. [Google Scholar] [CrossRef]
  119. Devereaux, A.L.; Mercer, S.L.; Cunningham, C.W. DARK classics in chemical neuroscience: Morphine. ACS Chem. Neurosci. 2018, 9, 2395–2407. [Google Scholar] [CrossRef]
  120. TG. Principles of Nonsteroidal Anti-Inflammatory Drug Use for Musculoskeletal Conditions in Adults. In eTG Complete Melbourne: Therapeutic Guidelines Limited. Available online: https://tgldcdp.tg.org.au/index (accessed on 29 December 2020).
  121. CDC. Prescription Opioids: Side Effects. Centers for Disease Control and Prevention. Available online: https://www.cdc.gov/drugoverdose/opioids/prescribed.html (accessed on 29 December 2020).
  122. WHO. Guidelines for the Psychosocially Assisted Pharmacological Treatment of Opioid Dependence. World Health Organization, Geneva. 2009. Available online: http://apps.who.int/iris/bitstream/10665/43948/1/9789241547543_eng.pdf (accessed on 26 October 2021).
  123. Collin, E.; Poulain, P.; Gauvain-Piquard, A.; Petit, G.; Pichard-Leandri, E. Is disease progression the major factor in morphine ‘tolerance’ in cancer pain treatment? Pain 1993, 55, 319–326. [Google Scholar] [CrossRef]
  124. Freye, E.; Anderson-Hillemacher, A.; Ritzdorf, I.; Levy, J.V. Opioid rotation from high-dose morphine to transdermal buprenorphine (Transtec®) in chronic pain patients. Pain Pract. 2007, 7, 123–129. [Google Scholar] [CrossRef] [PubMed]
  125. Rhondali, W.; Tremellat, F.; LeDoux, M.; Ciais, J.-F.; Bruera, E.; Filbet, M. Methadone rotation for cancer patients with refractory pain in a palliative care unit: An observational study. J. Palliat. Med. 2013, 16, 1382–1387. [Google Scholar] [CrossRef] [PubMed]
  126. Leppert, W.; Kowalski, G. Long-term administration of high doses of transdermal buprenorphine in cancer patients with severe neuropathic pain. OncoTargets Ther. 2015, 8, 3621–3627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Sutou, I.; Nakatani, T.; Hashimoto, T.; Saito, Y. Fentanyl Tolerance in the Treatment of Cancer Pain: A Case of Successful Opioid Switching from Fentanyl to Oxycodone at a Reduced Equivalent Dose. J. Pain Palliat. Care Pharmacother. 2015, 29, 161–165. [Google Scholar] [CrossRef]
  128. Shinde, S.; Gordon, P.; Sharma, P.; Gross, J.; Davis, M.P. Use of non-opioid analgesics as adjuvants to opioid analgesia for cancer pain management in an inpatient palliative unit: Does this improve pain control and reduce opioid requirements? Support. Care Cancer 2014, 23, 695–703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Raptis, E.; Vadalouca, A.; Stavropoulou, E.; Argyra, E.; Melemeni, A.; Siafaka, I. Pregabalin Vs. Opioids for the treatment of neuropathic cancer pain: A prospective, head-to-head, randomized, open-label study. Pain Pract. 2013, 14, 32–42. [Google Scholar] [CrossRef]
  130. Abdelhamid, E.E.; Sultana, M.; Portoghese, P.S.; Takemori, A.E. Selective blockage of delta opioid receptors prevents the development of morphine tolerance and dependence in mice. J. Pharmacol. Exp. Ther. 1991, 258, 299–303. [Google Scholar] [PubMed]
  131. Fundytus, M.E.; Schiller, P.W.; Shapiro, M.; Weltrowska, G.; Coderre, T.J. Attenuation of morphine tolerance and dependence with the highly selective delta-opioid receptor antagonist TIPP[psi]. Eur. J. Pharmacol. 1995, 286, 105–108. [Google Scholar] [CrossRef]
  132. Hepburn, M.J.; Little, P.J.; Gingras, J.; Kuhn, C.M. Differential effects of naltrindole on morphine-induced tolerance and physical dependence in rats. J. Pharmacol. Exp. Ther. 1997, 281, 1350–1356. [Google Scholar]
  133. Roy, S.; Guo, X.; Kelschenbach, J.; Liu, Y.; Loh, H.H. In vivo activation of a mutant-opioid receptor by naltrexone produces a potent analgesic effect but no tolerance: Role of-Receptor activation and-receptor blockade in morphine tolerance. J. Neurosci. 2005, 25, 3229–3233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Imam, M.Z.; Kuo, A.; Ghassabian, S.; Smith, M.T. Progress in understanding mechanisms of opioid-induced gastrointestinal adverse effects and respiratory depression. Neuropharmacology 2018, 131, 238–255. [Google Scholar] [CrossRef]
  135. Bie, B.; Pan, Z.Z. Trafficking of central opioid receptors and descending pain inhibition. Mol. Pain 2007, 3, 37. [Google Scholar] [CrossRef] [Green Version]
  136. Bs, A.C.B.; Whistler, J.L. How to design an opioid drug that causes reduced tolerance and dependence. Ann. Neurol. 2010, 67, 559–569. [Google Scholar] [CrossRef] [Green Version]
  137. Viet, C.T.; Dang, D.; Aouizerat, B.E.; Miaskowski, C.; Ye, Y.; Viet, D.T.; Ono, K.; Schmidt, B.L. OPRM1 methylation contributes to opioid tolerance in cancer patients. J. Pain 2017, 18, 1046–1059. [Google Scholar] [CrossRef] [PubMed]
  138. Groer, C.E.; Schmid, C.L.; Jaeger, A.M.; Bohn, L.M. Agonist-directed Interactions with Specific β-Arrestins Determine μ-Opioid Receptor Trafficking, Ubiquitination, and Dephosphorylation. J. Biol. Chem. 2011, 286, 31731–31741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Arttamangkul, S.; Heinz, D.A.; Bunzow, J.R.; Song, X.; Williams, J.T. Cellular tolerance at the µ-opioid receptor is phosphorylation dependent. eLife 2018, 7, e34989. [Google Scholar] [CrossRef] [PubMed]
  140. Martini, L.; Whistler, J.L. The role of mu opioid receptor desensitization and endocytosis in morphine tolerance and dependence. Curr. Opin. Neurobiol. 2007, 17, 556–564. [Google Scholar] [CrossRef]
  141. Bourova, L.; Vosahlikova, M.; Kagan, D.; Dlouha, K.; Novotny, J.; Svoboda, P. Long-term adaptation to high doses of morphine causes desensitization of mu-OR- and delta-OR-stimulated G-protein response in forebrain cortex but does not decrease the amount of G-protein alpha subunits. Med. Sci. Monit. 2010, 16, 260–270. [Google Scholar]
  142. Gupta, A.; Mulder, J.; Gomes, I.; Rozenfeld, R.; Bushlin, I.; Ong, E.; Lim, M.; Maillet, E.; Junek, M.; Cahill, C.M. Increased abundance of opioid receptor heteromers following chronic morphine administration. Sci. Signal. 2010, 3, ra54. [Google Scholar] [CrossRef] [Green Version]
  143. Leah, P.M.; Heath, E.M.L.; Balleine, B.W.; Christie, M. Chronic morphine reduces surface expression of δ-Opioid receptors in subregions of rostral striatum. Neurochem. Res. 2015, 41, 500–509. [Google Scholar] [CrossRef]
  144. Paul, A.; Gueven, N.; Dietis, N. Profiling the effects of repetitive morphine administration on motor behavior in rats. Molecules 2021, 26, 4355. [Google Scholar] [CrossRef] [PubMed]
  145. Paul, A.K.; Gueven, N.; Dietis, N. Morphine dosing strategy plays a key role in the generation and duration of the produced antinociceptive tolerance. Neuropharmacology 2017, 121, 158–166. [Google Scholar] [CrossRef] [PubMed]
  146. McQuay, H. Central analgesics. Acta Neurochir. Suppl. 1987, 38, 41–43. [Google Scholar]
  147. Le Bars, D.; Gozariu, M.; Cadden, S.W. Animal models of nociception. Pharmacol. Rev. 2001, 53, 597–652. [Google Scholar] [PubMed]
  148. Gårdmark, M.; Höglund, A.U.; Hammarlund-Udenaes, M. Aspects on Tail-Flick, Hot-Plate and electrical stimulation tests for morphine antinociception. Pharmacol. Toxicol. 1998, 83, 252–258. [Google Scholar] [CrossRef] [PubMed]
  149. Labianca, R.; Sarzi-Puttini, P.; Zuccaro, S.M.; Cherubino, P.; Vellucci, R.; Fornasari, D. Adverse effects associated with non-opioid and opioid treatment in patients with chronic pain. Clin. Drug Investig. 2012, 32, 53–63. [Google Scholar] [CrossRef]
  150. Juurlink, D.N.; Dhalla, I.A. Dependence and addiction during chronic opioid therapy. J. Med. Toxicol. 2012, 8, 393–399. [Google Scholar] [CrossRef] [Green Version]
  151. Ballantyne, J.C.; LaForge, S.K. Opioid dependence and addiction during opioid treatment of chronic pain. Pain 2007, 129, 235–255. [Google Scholar] [CrossRef] [PubMed]
  152. Nestler, E.J.; Aghajanian, G.K. Molecular and Cellular Basis of Addiction. Science 1997, 278, 58–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Counts, S.E.; Mufson, E.J. Chapter 12-Locus coeruleus. In The Human Nervous System, 3rd ed.; Mai, J.K., Paxinos, G., Eds.; Academic Press: San Diego, CA, USA, 2012; pp. 425–438. [Google Scholar]
  154. Kosten, T.R.; George, T.P. The neurobiology of opioid dependence: Implications for treatment. Sci. Pract. Perspect. 2002, 1, 13–20. [Google Scholar] [CrossRef] [Green Version]
  155. Listos, J.; Łupina, M.; Talarek, S.; Mazur, A.; Orzelska-Górka, J.; Kotlińska, J. The mechanisms involved in morphine addiction: An overview. Int. J. Mol. Sci. 2019, 20, 4302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Dowell, D.; Haegerich, T.M.; Chou, R. CDC Guideline for prescribing opioids for chronic pain—United States, 2016. JAMA 2016, 315, 1624–1645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. (PHE) Public Health England; Faculty of Medicine of the Royal College of Anaesthetists; Royal College of General Practitioners; The British Pain Society. Managing Persistent Pain in Secure Settings. 2013. Available online: http://www.nta.nhs.uk/uploads/persistentpain.pdf (accessed on 24 September 2016).
  158. NOUGG. Canadian Guideline for Safe and Effective Use of Opioids for Chronic Non-Cancer Pain. Canada. National Opioid Use Guideline Group (NOUGG). 2010. Available online: http://nationalpaincentre.mcmaster.ca/opioid/ (accessed on 24 September 2016).
  159. American Society of Anesthesiologists Task Force on Chronic Pain Management. American Society of Regional Anesthesia and Pain Medicine Practice Guidelines for Chronic Pain Management. Anesthesiology 2010, 112, 810–833. [Google Scholar] [CrossRef] [Green Version]
  160. Trescot, A.M.; Helm, S.; Hansen, H.; Benyamin, R.; Glaser, S.E.; Adlaka, R.; Patel, S.; Manchikanti, L. Opioids in the management of chronic non-cancer pain: An update of American Society of the Interventional Pain Physicians’ (ASIPP) Guidelines. Pain Physician 2008, 11, S5–S62. [Google Scholar]
  161. Kirpalani, D. How to maximize patient safety when prescribing opioids. PM&R 2015, 7, S225–S235. [Google Scholar] [CrossRef]
  162. Stein, C.; Reinecke, H.; Sorgatz, H. Opioid use in chronic noncancer pain: Guidelines revisited. Curr. Opin. Anaesthesiol. 2010, 23, 598–601. [Google Scholar] [CrossRef] [PubMed]
  163. Streicher, J.M.; Bilsky, E.J. Peripherally acting μ-Opioid receptor antagonists for the treatment of opioid-related side effects: Mechanism of action and clinical implications. J. Pharm. Pr. 2017, 31, 658–669. [Google Scholar] [CrossRef] [PubMed]
  164. Camilleri, M. Opioid-Induced constipation: Challenges and therapeutic opportunities. Am. J. Gastroenterol. 2011, 106, 835–842. [Google Scholar] [CrossRef] [PubMed]
  165. Kon, R.; Ikarashi, N.; Hayakawa, A.; Haga, Y.; Fueki, A.; Kusunoki, Y.; Tajima, M.; Ochiai, W.; Machida, Y.; Sugiyama, K. Morphine-Induced constipation develops with increased aquaporin-3 expression in the colon via increased serotonin secretion. Toxicol. Sci. 2015, 145, 337–347. [Google Scholar] [CrossRef] [Green Version]
  166. Akbarali, H.I.; Inkisar, A.; Dewey, W.L. Site and mechanism of morphine tolerance in the gastrointestinal tract. Neurogastroenterol. Motil. 2014, 26, 1361–1367. [Google Scholar] [CrossRef]
  167. Swegle, J.M.; Logemann, C. Management of common opioid-induced adverse effects. Am. Fam. Physician 2006, 74, 1347–1354. [Google Scholar]
  168. Cherny, N.; Ripamonti, C.; Pereira, J.; Davis, C.; Fallon, M.; McQuay, H.; Mercadante, S.; Pasternak, G.; Ventafridda, V. For the expert working group of the european association of palliative care network strategies to manage the adverse effects of oral morphine: An evidence-based report. J. Clin. Oncol. 2001, 19, 2542–2554. [Google Scholar] [CrossRef] [Green Version]
  169. Prichard, D.; Norton, C.; Bharucha, A.E. Management of opioid-induced constipation. Br. J. Nurs. 2016, 25, S4–S5, S8–S11. [Google Scholar] [CrossRef] [Green Version]
  170. Meissner, W.; Schmidt, U.; Hartmann, M.; Kath, R.; Reinhart, K. Oral naloxone reverses opioid-associated constipation. Pain 2000, 84, 105–109. [Google Scholar] [CrossRef]
  171. Shimoyama, N.; Shimoyama, M. Treatment of constipation in chronic pain patients. Masui. Jpn. J. Anesthesiol. 2013, 62, 822–828. [Google Scholar]
  172. Burness, C.B.; Keating, G.M. Oxycodone/naloxone prolonged-release: A review of its use in the management of chronic pain while counteracting opioid-induced constipation. Drugs 2014, 74, 353–375. [Google Scholar] [CrossRef]
  173. Moore, R.A.; McQuay, H.J. Prevalence of opioid adverse events in chronic non-malignant pain: Systematic review of randomised trials of oral opioids. Arthritis Res. 2005, 7, R1046–R1051. [Google Scholar] [CrossRef] [Green Version]
  174. Watcha, M.F.; White, P.F. Postoperative nausea and vomiting. Its etiology, treatment, and prevention. Anesthesiology 1992, 77, 162–184. [Google Scholar] [CrossRef] [PubMed]
  175. Lehnen, N.; Heuser, F.; Saglam, M.; Schulz, C.M.; Wagner, K.J.; Taki, M.; Kochs, E.F.; Jahn, K.; Brandt, T.; Glasauer, S.; et al. Opioid-inducednausea involves a vestibular problem preventable by head-rest. PLoS ONE 2015, 10, e0135263. [Google Scholar] [CrossRef] [Green Version]
  176. Rojas, C.; Slusher, B.S. Mechanisms and latest clinical studies of new NK1 receptor antagonists for chemotherapy-induced nausea and vomiting: Rolapitant and NEPA (netupitant/palonosetron). Cancer Treat. Rev. 2015, 41, 904–913. [Google Scholar] [CrossRef]
  177. Rojas, C.; Raje, M.; Tsukamoto, T.; Slusher, B.S. Molecular mechanisms of 5-HT3 and NK1 receptor antagonists in prevention of emesis. Eur. J. Pharmacol. 2014, 722, 26–37. [Google Scholar] [CrossRef]
  178. Koju, R.B.; Gurung, B.S.; Dongol, Y. Prophylactic administration of ondansetron in prevention of intrathecal morphine-induced pruritus and post-operative nausea and vomiting in patients undergoing caesarean section. BMC Anesthesiol. 2015, 15, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Ashby, M.A.; Martin, P.; Jackson, K.A. Opioid substitution to reduce adverse effects in cancer pain management. Med. J. Aust. 1999, 170, 68–71. [Google Scholar] [CrossRef]
  180. De Stoutz, N.D.; Bruera, E.; Suarez-Almazor, M. Opioid rotation for toxicity reduction in terminal cancer patients. J. Pain Symptom Manag. 1995, 10, 378–384. [Google Scholar] [CrossRef]
  181. Sultan, P.; Gutierrez, M.C.; Carvalho, B. Neuraxial morphine and respiratory depression. Drugs 2011, 71, 1807–1819. [Google Scholar] [CrossRef]
  182. Kuo, A.; Wyse, B.D.; Meutermans, W.; Smith, M.T. In vivo profiling of seven common opioids for antinociception, constipation and respiratory depression: No two opioids have the same profile. Br. J. Pharmacol. 2014, 172, 532–548. [Google Scholar] [CrossRef] [Green Version]
  183. Boom, M.; Niesters, M.; Sarton, E.; Aarts, L.; Smith, T.W.; Dahan, A. Non-analgesic effects of opioids: Opioid-induced respiratory depression. Curr. Pharm. Des. 2012, 18, 5994–6004. [Google Scholar] [CrossRef]
  184. Kamei, J.; Ohsawa, M.; Hayashi, S.-S.; Nakanishi, Y. Effect of chronic pain on morphine-induced respiratory depression in mice. Neuroscience 2011, 174, 224–233. [Google Scholar] [CrossRef]
  185. Hill, R.; Santhakumar, R.; Dewey, W.; Kelly, E.; Henderson, G. Fentanyl depression of respiration: Comparison with heroin and morphine. Br. J. Pharmacol. 2020, 177, 254–265. [Google Scholar] [CrossRef] [PubMed]
  186. Manzke, T.; Guenther, U.; Ponimaskin, E.G.; Haller, M.; Dutschmann, M.; Schwarzacher, S.; Richter, D.W. 5-HT 4(a) Receptors avert opioid-induced breathing depression without loss of analgesia. Science 2003, 301, 226–229. [Google Scholar] [CrossRef] [PubMed]
  187. Pattinson, K.T.S. Opioids and the control of respiration. Br. J. Anaesth. 2008, 100, 747–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Manzke, T.; Niebert, M.; Koch, U.; Caley, A.; Vogelgesang, S.; Bischoff, A.-M.; Hülsmann, S.; Ponimaskin, E.; Müller, U.; Smart, T.; et al. Die von serotoninrezeptor 1A modulierte dephosphorylierung des glyzinrezeptors α3. Der. Schmerz 2011, 25, 272–281. [Google Scholar] [CrossRef]
  189. Dunn, E.B.; Wolfe, J.J. Finding an optimal dose: Considerations in accurate opioid dispensing. Vet. Hum. Toxicol. 2000, 42, 36–38. [Google Scholar]
  190. Lewanowitsch, T.; Irvine, R.J. Naloxone methiodide reverses opioid-induced respiratory depression and analgesia without withdrawal. Eur. J. Pharmacol. 2002, 445, 61–67. [Google Scholar] [CrossRef]
  191. Droney, J.M.; Gretton, S.K.; Sato, H.; Ross, J.R.; Branford, R.; Welsh, K.I.; Cookson, W.; Riley, J. Analgesia and central side-effects: Two separate dimensions of morphine response. Br. J. Clin. Pharmacol. 2013, 75, 1340–1350. [Google Scholar] [CrossRef] [Green Version]
  192. Babbini, M.; Davis, W.M. Time-dose relationships for locomotor activity effects of morphine after acute or repeated treatment. Br. J. Pharmacol. 1972, 46, 213–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Hollais, A.W.; Patti, C.L.; Zanin, K.A.; Fukushiro, D.F.; Berro, L.F.; Carvalho, R.C.; Kameda, S.R.; Frussa-Filho, R. Effects of acute and long-term typical or atypical neuroleptics on morphine-induced behavioural effects in mice. Clin. Exp. Pharmacol. Physiol. 2014, 41, 255–263. [Google Scholar] [CrossRef]
  194. Patti, C.L.; Frussa-Filho, R.; Silva, R.; Carvalho, R.C.; Kameda, S.R.; Takatsu-Coleman, A.L.; Cunha, J.L.; Abílio, V.C. Behavioral characterization of morphine effects on motor activity in mice. Pharmacol. Biochem. Behav. 2005, 81, 923–927. [Google Scholar] [CrossRef]
  195. Rodrı́guez-Arias, M.; Broseta, I.; Aguilar, M.; Miñarro, J. Lack of specific effects of selective D1 and D2 dopamine antagonists vs. risperidone on morphine-induced hyperactivity. Pharmacol. Biochem. Behav. 2000, 66, 189–197. [Google Scholar] [CrossRef]
  196. Domino, E.F.; Vasco, M.R.; Wilson, A.E. Mixed depressant and stimulant actions of morphine and their relationship to brain acetylcholine. Life Sci. 1976, 18, 361–376. [Google Scholar] [CrossRef] [Green Version]
  197. Brady, L.S.; Holtzman, S.G. Locomotor activity in morphine-dependent and post-dependent rats. Pharmacol. Biochem. Behav. 1981, 14, 361–370. [Google Scholar] [CrossRef]
  198. Tulunay, F.C.; Ayhan, I.H.; Sparber, S.B. The effects of morphine and delta-9-tetrahydrocannabinol on motor activity in rats. Psychopharmacology 1982, 78, 358–360. [Google Scholar] [CrossRef] [PubMed]
  199. Murphy, N.P.; Lam, H.A.; Maidment, N.T. A comparison of morphine-induced locomotor activity and mesolimbic dopamine release in C57BL6, 129Sv and DBA2 mice. J. Neurochem. 2008, 79, 626–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Christie, J.E.; Crow, T.J. Turning behaviour as an index of the action of amphetamines and ephedrines on central dopamine-containing neurones. Br. J. Pharmacol. 1971, 43, 658–667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Barber, D.; Blackburn, T.; Greenwood, D. An automatic apparatus for recording rotational behaviour in rats with brain lesions. Physiol. Behav. 1973, 11, 117–120. [Google Scholar] [CrossRef]
  202. Urs, N.M.; Daigle, T.L.; Caron, M.G. A Dopamine D1 receptor-dependent β-arrestin signaling complex potentially regulates morphine-induced psychomotor activation but not reward in mice. Neuropsychopharmacology 2010, 36, 551–558. [Google Scholar] [CrossRef] [Green Version]
  203. Ryczko, D.; Dubuc, R. Dopamine and the brainstem locomotor networks: From lamprey to human. Front. Neurosci. 2017, 11, 295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Chastain, L.G.; Qu, H.; Bourke, C.H.; Iuvone, P.M.; Dobner, P.R.; Nemeroff, C.B.; Kinkead, B. Striatal dopamine receptor plasticity in neurotensin deficient mice. Behav. Brain Res. 2014, 280, 160–171. [Google Scholar] [CrossRef] [Green Version]
  205. Borgkvist, A.; Usiello, A.; Greengard, P.; Fisone, G. Activation of the cAMP/PKA/DARPP-32 signaling pathway is required for morphine psychomotor stimulation but not for morphine reward. Neuropsychopharmacology 2007, 32, 1995–2003. [Google Scholar] [CrossRef] [Green Version]
  206. Mishra, A.; Singh, S.; Shukla, S. Physiological and functional basis of dopamine receptors and their role in neurogenesis: Possible implication for Parkinson’s disease. J. Exp. Neurosci. 2018, 12, 1179069518779829. [Google Scholar] [CrossRef] [Green Version]
  207. Horsfall, J.T.; Sprague, J.E. The pharmacology and toxicology of the ‘Holy Trinity’. Basic Clin. Pharmacol. Toxicol. 2016, 120, 115–119. [Google Scholar] [CrossRef]
  208. Anderson, E.; Hearing, M. Chapter 4-Neural circuit plasticity in addiction. In Neural Mechanisms of Addiction; Torregrossa, M., Ed.; Academic Press: Cambridge, MA, USA, 2019; pp. 35–60. [Google Scholar]
  209. Oliva, I.; Wanat, M.J. Ventral tegmental area afferents and drug-dependent behaviors. Front. Psychiatry 2016, 7, 30. [Google Scholar] [CrossRef] [Green Version]
  210. Adinoff, B. Neurobiologic processes in drug reward and addiction. Harv. Rev. Psychiatry 2004, 12, 305–320. [Google Scholar] [CrossRef] [PubMed]
  211. Gardner, E.L.; Ashby, C.R. Heterogeneity of the mesotelencephalic dopamine fibers: Physiology and pharmacology. Neurosci. Biobehav. Rev. 2000, 24, 115–118. [Google Scholar] [CrossRef]
  212. Lever, C.; Burton, S.; O’Keefe, J. Rearing on hind legs, environmental novelty, and the hippocampal formation. Rev. Neurosci. 2006, 17, 111–133. [Google Scholar] [CrossRef]
  213. Rodriguiz, R.M.; Wetsel, W.C. Frontiers in neuroscience assessments of cognitive deficits in mutant mice. In Animal Models of Cognitive Impairment; Levin, E.D., Buccafusco, J.J., Eds.; CRC Press/Taylor & Francis, Taylor & Francis Group, LLC.: Boca Raton, FL, USA, 2006. [Google Scholar]
  214. Alves, R.; de Carvalho, J.G.B.; Venditti, M.A.C. High-and Low-Rearing rats differ in the brain excitability controlled by the allosteric benzodiazepine site in the GABAA receptor. J. Behav. Brain Sci. 2012, 2, 315–325. [Google Scholar] [CrossRef] [Green Version]
  215. Valverde, O.; Mantamadiotis, T.; Torrecilla, M.; Ugedo, L.; Pineda, J.; Bleckmann, S.; Gass, P.; Kretz, O.; Mitchell, J.M.; Schütz, G.; et al. Modulation of anxiety-like behavior and morphine dependence in CREB-deficient mice. Neuropsychopharmacology 2004, 29, 1122–1133. [Google Scholar] [CrossRef]
  216. Contet, C.; Filliol, D.; Matifas, A.; Kieffer, B.L. Morphine-induced analgesic tolerance, locomotor sensitization and physical dependence do not require modification of mu opioid receptor, cdk5 and adenylate cyclase activity. Neuropsychopharmacology 2008, 54, 475–486. [Google Scholar] [CrossRef]
  217. Zan, G.-Y.; Wang, Q.; Wang, Y.-J.; Liu, Y.; Hang, A.; Shu, X.-H.; Liu, J.-G. Antagonism of κ opioid receptor in the nucleus accumbens prevents the depressive-like behaviors following prolonged morphine abstinence. Behav. Brain Res. 2015, 291, 334–341. [Google Scholar] [CrossRef]
  218. Aceves, M.; Bancroft, E.; Aceves, A.R.; Hook, M.A. Nor-binaltorphimine blocks the adverse effects of morphine after spinal cord injury. J. Neurotrauma 2017, 34, 1164–1174. [Google Scholar] [CrossRef] [Green Version]
  219. Walsh, R.N.; Cummins, R.A. The open-field test: A critical review. Psychol. Bull. 1976, 83, 482–504. [Google Scholar] [CrossRef]
  220. Paul, A.K.; Gueven, N.; Dietis, N. Age-dependent antinociception and behavioral inhibition by morphine. Pharmacol. Biochem. Behav. 2018, 168, 8–16. [Google Scholar] [CrossRef]
  221. Paul, A.K.; Gueven, N.; Dietis, N. Data on prolonged morphine-induced antinociception and behavioral inhibition in older rats. Data Brief. 2018, 19, 183–188. [Google Scholar] [CrossRef]
  222. Dominguez, J.E.; Habib, A. Prophylaxis and treatment of the side-effects of neuraxial morphine analgesia following cesarean delivery. Curr. Opin. Anaesthesiol. 2013, 26, 288–295. [Google Scholar] [CrossRef]
  223. Raffaeli, W.; Marconi, G.; Fanelli, G.; Taddei, S.; Borghi, G.B.; Casati, A. Opioid-related side-effects after intrathecal morphine. Eur. J. Anaesthesiol. 2006, 23, 605–610. [Google Scholar] [CrossRef]
  224. Benyamin, R.; Trescot, A.M.; Datta, S.; Buenaventura, R.; Adlaka, R.; Sehgal, N.; Glaser, S.E.; Vallejo, R. Opioid complications and side effects. Pain Physician 2008, 11, S105–S120. [Google Scholar] [CrossRef]
  225. Miyamoto, T.; Patapoutian, A. Why does morphine make you itch? Cell 2011, 147, 261–262. [Google Scholar] [CrossRef] [Green Version]
  226. Liu, X.; Liu, Z.-C.; Sun, Y.-G.; Ross, M.; Kim, S.; Tsai, F.-F.; Li, Q.-F.; Jeffry, J.; Kim, J.-Y.; Loh, H.H.; et al. Unidirectional cross-activation of GRPR by MOR1D uncouples itch and analgesia induced by opioids. Cell 2011, 147, 447–458. [Google Scholar] [CrossRef] [Green Version]
  227. Abe, K.; Kobayashi, K.; Yoshino, S.; Taguchi, K.; Nojima, H. Withdrawal of repeated morphine enhances histamine-induced scratching responses in mice. Drug Chem. Toxicol. 2015, 38, 167–173. [Google Scholar] [CrossRef]
  228. Roeckel, L.-A.; Le Coz, G.-M.; Gaveriaux-Ruff, C.; Simonin, F. Opioid-induced hyperalgesia: Cellular and molecular mechanisms. Neuroscience 2016, 338, 160–182. [Google Scholar] [CrossRef]
  229. Swartjes, M.; Mooren, R.A.G.; Waxman, A.R.; Arout, C.; Van De Wetering, K.; Hartigh, J.D.; Beijnen, J.H.; Kest, B.; Dahan, A. Morphine induces hyperalgesia without involvement of µ-Opioid receptor or morphine-3-glucuronide. Mol. Med. 2012, 18, 1320–1326. [Google Scholar] [CrossRef]
  230. Song, L.; Wu, C.; Zuo, Y. Melatonin prevents morphine-induced hyperalgesia and tolerance in rats: Role of protein kinase C and N-methyl-D-aspartate receptors. BMC Anesthesiol. 2015, 15, 12. [Google Scholar] [CrossRef] [Green Version]
  231. Elhabazi, K.; Ayachi, S.; Ilien, B.; Simonin, F. Assessment of morphine-induced hyperalgesia and analgesic tolerance in mice using thermal and mechanical nociceptive modalities. J. Vis. Exp. 2014, 89, e51264. [Google Scholar] [CrossRef] [Green Version]
  232. Van Rijn, R.; Whistler, J.L.; Waldhoer, M. Opioid-receptor-heteromer-specific trafficking and pharmacology. Curr. Opin. Pharmacol. 2010, 10, 73–79. [Google Scholar] [CrossRef] [Green Version]
  233. Walwyn, W.; John, S.; Maga, M.; Evans, C.J.; Hales, T.G. δ Receptors are required for full inhibitory coupling of μ receptors to voltage-dependent Ca2+ channels in dorsal root ganglion neurons. Mol. Pharmacol. 2009, 76, 134–143. [Google Scholar] [CrossRef] [Green Version]
  234. Alfaras-Melainis, K.; Gomes, I.; Rozenfeld, R.; Zachariou, V.; Devi, L. Modulation of opioid receptor function by protein-protein interactions. Front. Biosci. 2009, 14, 3594–3607. [Google Scholar] [CrossRef] [Green Version]
  235. Ballet, S.; Pietsch, M.; Abell, A.D. Multiple ligands in opioid research. Protein Pept. Lett. 2008, 15, 668–682. [Google Scholar] [CrossRef]
  236. Lezoualc, H.F.; Jockers, R. Bivalent ligands as specific pharmacological tools for G protein-coupled receptor dimers. Curr. Drug Discov. Technol. 2008, 5, 312–318. [Google Scholar] [CrossRef]
  237. Yamada, H.; Shimoyama, N.; Sora, I.; Uhl, G.R.; Fukuda, Y.; Moriya, H.; Shimoyama, M. Morphine can produce analgesia via spinal kappa opioid receptors in the absence of mu opioid receptors. Brain Res. 2006, 1083, 61–69. [Google Scholar] [CrossRef]
  238. Abul-Husn, N.; Sutak, M.; Milne, B.; Jhamandas, K. Augmentation of spinal morphine analgesia and inhibition of tolerance by low doses of μ- and δ -opioid receptor antagonists. Br. J. Pharmacol. 2007, 151, 877–887. [Google Scholar] [CrossRef] [Green Version]
  239. McNaull, B.; Trang, T.; Sutak, M.; Jhamandas, K. Inhibition of tolerance to spinal morphine antinociception by low doses of opioid receptor antagonists. Eur. J. Pharmacol. 2007, 560, 132–141. [Google Scholar] [CrossRef]
  240. Dietis, N.; Guerrini, R.; Calo, G.; Salvadori, S.; Rowbotham, D.J.; Lambert, D.G. Simultaneous targeting of multiple opioid receptors: A strategy to improve side-effect profile. Br. J. Anaesth. 2009, 103, 38–49. [Google Scholar] [CrossRef] [Green Version]
  241. Morphy, R.; Rankovic, Z. Designed multiple ligands. An emerging drug discovery paradigm. J. Med. Chem. 2005, 48, 6523–6543. [Google Scholar] [CrossRef] [PubMed]
  242. Schiller, P.W. Bi- or multifunctional opioid peptide drugs. Life Sci. 2010, 86, 598–603. [Google Scholar] [CrossRef] [Green Version]
  243. Toll, L.; Khroyan, T.V.; Polgar, W.E.; Jiang, F.; Olsen, C.; Zaveri, N.T. Comparison of the antinociceptive and antirewarding profiles of novel bifunctional nociceptin receptor/μ-opioid receptor ligands: Implications for therapeutic applications. J. Pharmacol. Exp. Ther. 2009, 331, 954–964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Turnaturi, R.; Aricò, G.; Ronsisvalle, G.; Parenti, C.; Pasquinucci, L. Multitarget opioid ligands in pain relief: New players in an old game. Eur. J. Med. Chem. 2016, 108, 211–228. [Google Scholar] [CrossRef] [PubMed]
  245. Anand, J.P.; Montgomery, D. Multifunctional opioid ligands. Gastrointest. Regul. Pept. 2018, 247, 21–51. [Google Scholar] [CrossRef]
  246. Cunningham, C.W.; Elballa, W.M.; Vold, S.U. Bifunctional opioid receptor ligands as novel analgesics. Neuropharmacology 2019, 151, 195–207. [Google Scholar] [CrossRef]
  247. Agnes, R.S.; Ying, J.; Kövér, K.E.; Lee, Y.S.; Davis, P.; Ma, S.-W.; Badghisi, H.; Porreca, F.; Lai, J.; Hruby, V.J. Structure–activity relationships of bifunctional cyclic disulfide peptides based on overlapping pharmacophores at opioid and cholecystokinin receptors. Peptides 2008, 29, 1413–1423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Kulkarni, V.V.; Lee, Y.S.; Salibay, C.; Davis, P.; Ma, S.-W.; Porreca, F.; Hruby, V.J. Novel analogues of bifunctional ligands for opioid and melanocortin 4 receptor. In Peptides for Youth; Springer: New York, NY, USA, 2009; pp. 195–196. [Google Scholar] [CrossRef]
  249. Yamamoto, T.; Nair, P.; Jacobsen, N.E.; Kulkarni, V.; Davis, P.; Ma, S.-W.; Navratilova, E.; Yamamura, H.I.; Vanderah, T.W.; Porreca, F.; et al. Biological and conformational evaluation of bifunctional compounds for opioid receptor agonists and neurokinin 1 receptor antagonists possessing two penicillamines. J. Med. Chem. 2010, 53, 5491–5501. [Google Scholar] [CrossRef] [Green Version]
  250. Yekkirala, A.S.; Kalyuzhny, A.E.; Portoghese, P.S. An immunocytochemical-derived correlate for evaluating the bridging of heteromeric mu-delta opioid protomers by bivalent ligands. ACS Chem. Biol. 2013, 8, 1412–1416. [Google Scholar] [CrossRef]
  251. La Rochelle, A.D.; Guillemyn, K.; Dumitrascuta, M.; Martin, C.; Utard, V.; Quillet, R.; Schneider, S.; Daubeuf, F.; Willemse, T.; Mampuys, P.; et al. A bifunctional-biased mu-opioid agonist–neuropeptide FF receptor antagonist as analgesic with improved acute and chronic side effects. Pain 2018, 159, 1705–1718. [Google Scholar] [CrossRef] [PubMed]
  252. Tzschentke, T.M.; Linz, K.; Koch, T.; Christoph, T. Cebranopadol: A novel first-in-class potent analgesic acting via NOP and opioid receptors. Antisense Res. Appl. 2019, 254, 367–398. [Google Scholar] [CrossRef]
  253. Fujita, W.; Gomes, I.; Devi, L.A. Heteromers of μ-δ opioid receptors: New pharmacology and novel therapeutic possibilities. Br. J. Pharmacol. 2014, 172, 375–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Gomes, I.; Jordan, B.A.; Gupta, A.; Trapaidze, N.; Nagy, V.; Devi, L.A. Heterodimerization of mu and delta opioid receptors: A role in opiate synergy. J. Neurosci. 2000, 20, RC110. [Google Scholar] [CrossRef] [Green Version]
  255. Fujita, W.; Gomes, I.; Devi, L.A. Revolution in GPCR signalling: Opioid receptor heteromers as novel therapeutic targets: IUPHAR Review 10. Br. J. Pharmacol. 2014, 171, 4155–4176. [Google Scholar] [CrossRef] [Green Version]
  256. Gomes, I.; Fujita, W.; Gupta, A.; Saldanha, S.A.; Negri, A.; Pinello, C.E.; Eberhart, C.; Roberts, E.; Filizola, M.; Hodder, P.; et al. Identification of a-opioid receptor heteromer-biased agonist with antinociceptive activity. Proc. Natl. Acad. Sci. USA 2013, 110, 12072–12077. [Google Scholar] [CrossRef] [Green Version]
  257. Milan-Lobo, L.; Enquist, J.; Van Rijn, R.M.; Whistler, J.L. Anti-analgesic effect of the mu/delta opioid receptor heteromer revealed by ligand-biased antagonism. PLoS ONE 2013, 8, e58362. [Google Scholar] [CrossRef]
  258. Wallace, M.S.; Kosek, P.S.; Staats, P.; Fisher, R.; Schultz, D.M.; Leong, M. Phase II, Open-label, multicenter study of combined intrathecal morphine and ziconotide: Addition of ziconotide in patients receiving intrathecal morphine for severe chronic pain. Pain Med. 2008, 9, 271–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Reig, E.; Abejón, D.; Krames, E.S. Chapter 35-Intrathecal non-opioid analgesics for the control of pain. In Neuromodulation; Krames, E.S., Peckham, P.H., Rezai, A.R., Eds.; Academic Press: San Diego, CA, USA, 2009; pp. 467–481. [Google Scholar]
Figure 1. Effects of opioid agonists on pain and signal transduction. Abbreviations: ATP, adenosine triphosphate; ADP, adenosine diphosphate; GABA, gamma-aminobutyric acid; AMP, adenosine monophosphate; cAMP, cyclic adenosine monophosphate; PKA, protein kinase A; CRE, cAMP-responsive element; CREB, cAMP-responsive element binding protein; GTP, Guanosine triphosphate; GDP, Guanosine diphosphate; PDEs, Cyclic nucleotide phosphodiesterases; Ca2+/CaMK-II, Ca2+/calmodulin-dependent protein kinase II. Symbols: solid arrow, strong activation; dashed arrow, moderate activation; solid T-shaped line, strong inhibition; dashed T-shaped line, moderate inhibition, line with ×, blocking; upward arrow, increased effect; downward arrow, decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Figure 1. Effects of opioid agonists on pain and signal transduction. Abbreviations: ATP, adenosine triphosphate; ADP, adenosine diphosphate; GABA, gamma-aminobutyric acid; AMP, adenosine monophosphate; cAMP, cyclic adenosine monophosphate; PKA, protein kinase A; CRE, cAMP-responsive element; CREB, cAMP-responsive element binding protein; GTP, Guanosine triphosphate; GDP, Guanosine diphosphate; PDEs, Cyclic nucleotide phosphodiesterases; Ca2+/CaMK-II, Ca2+/calmodulin-dependent protein kinase II. Symbols: solid arrow, strong activation; dashed arrow, moderate activation; solid T-shaped line, strong inhibition; dashed T-shaped line, moderate inhibition, line with ×, blocking; upward arrow, increased effect; downward arrow, decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Pharmaceuticals 14 01091 g001
Figure 2. Effects of opioids on motor behaviour. Abbreviations: GABA, gamma-aminobutyric; cAMP, cyclic adenosine monophosphate; PKA, protein kinase A; CRE, cAMP-responsive element; CREB, cAMP-responsive element binding protein; D1, dopamine receptor-1; DARPP-32, dopamine- and cAMP-regulated phosphoprotein of 32 kDa; NMDA, N-methyl-D-aspartate receptor; PDEs, Cyclic nucleotide phosphodiesterases; NA, noradrenaline, VTA, ventral tegmental area; NAc, nucleus accumbens. Symbols: solid arrow, strong activation; dashed arrow, moderate activation; solid T-shaped line, strong inhibition; dashed T-shaped line, moderate inhibition, upward arrow, increased effect; downward arrow, decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Figure 2. Effects of opioids on motor behaviour. Abbreviations: GABA, gamma-aminobutyric; cAMP, cyclic adenosine monophosphate; PKA, protein kinase A; CRE, cAMP-responsive element; CREB, cAMP-responsive element binding protein; D1, dopamine receptor-1; DARPP-32, dopamine- and cAMP-regulated phosphoprotein of 32 kDa; NMDA, N-methyl-D-aspartate receptor; PDEs, Cyclic nucleotide phosphodiesterases; NA, noradrenaline, VTA, ventral tegmental area; NAc, nucleus accumbens. Symbols: solid arrow, strong activation; dashed arrow, moderate activation; solid T-shaped line, strong inhibition; dashed T-shaped line, moderate inhibition, upward arrow, increased effect; downward arrow, decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Pharmaceuticals 14 01091 g002
Figure 3. Effects of opioids in the gastrointestinal tract. Abbreviations: 5-HT, serotonin; GI, gastrointestinal. Symbols: solid arrow, strongly connected; dashed arrow, moderately connected; upward arrow (symbol), increased effect; downward arrow (symbol), decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Figure 3. Effects of opioids in the gastrointestinal tract. Abbreviations: 5-HT, serotonin; GI, gastrointestinal. Symbols: solid arrow, strongly connected; dashed arrow, moderately connected; upward arrow (symbol), increased effect; downward arrow (symbol), decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Pharmaceuticals 14 01091 g003
Figure 4. Effects of opioids on the respiratory function. Abbreviations: 5-HT, serotonin; GlyRα3, glycine receptor type-α3. Symbols: solid arrow, strongly connected; dashed arrow, moderately connected; upward arrow (symbol), increased effect; downward arrow (symbol), decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Figure 4. Effects of opioids on the respiratory function. Abbreviations: 5-HT, serotonin; GlyRα3, glycine receptor type-α3. Symbols: solid arrow, strongly connected; dashed arrow, moderately connected; upward arrow (symbol), increased effect; downward arrow (symbol), decreased effect. The figure was made with www.biorender.com (accessed on 19 October 2021).
Pharmaceuticals 14 01091 g004
Table 1. Major types of opioid receptors.
Table 1. Major types of opioid receptors.
Receptor
Nomenclature 1
GeneMost Common
Location in the CNS 2
Most Common
Roles and Functions
Selective
Agonist 3
Selective
Antagonist 4
μ, mu, MOPOPRM1Thalamus, amygdala, dorsal horn, cerebral cortex, striatum, hippocampus, locus coeruleusAnalgesia, intestinal transit, feeding, mood, hormone secretion, thermoregulation, cardiovascular functionDAMGO,
sufentanil,
PL017
CTAP,
CTOP,
β-FNA
δ, delta, DOPOPRD1Olfactory bulb, thalamus, cortex, caudate putamen, nucleus accumbens (NAc), amygdala, dorsal hornAnalgesia, mood,
gastrointestinal motility, behaviour, cardiovascular regulation
DPDPE,
[D-Ala2]deltorphin I,
[D-Ala2]deltorphin II
SNC80
Naltrindole,
TIPPᴪ,
Naltriben
κ, kappa, KOPOPRK1Olfactory bulb, NAc, cerebral cortex, claustrum, amygdala, caudate nucleus, hypothalamus, subthalamic nucleus, thalamus, corpus callosum.Analgesia in inflammation, diuresis, feeding, neuroprotection,
neuroendocrine functions
Enadoline,
U50488,
U69593,
salvinorin A
norBNI,
GNTI
N/OFQ, NOPOPRL1Hippocampus, hypothalamus, amygdala, substantia nigra, dorsal horn, lateral septumSpinal analgesia, anxiety, mood, memory, feeding, locomotor activityUFP-102,
Ro64-6198,
N/OFQ-(1-13)-NH2,
UFP-112
UFP-101,
SB 612111,
J-113397,
JTC-801
1,2,3,4 Further detailed information can be obtained from IUPHAR/BPS guide to pharmacology (http://www.guidetopharmacology.org; accessed on: 26 October 2021) and Alexander and colleagues [14]. Abbreviations: β-funaltrexamine (β-FNA), [D-Ala2, N-MePhe4, Gly-ol]-enkephalin (DAMGO), norbinaltorphimine (norBNI), [D-Pen2, D-Pen5] enkephalin (DPDPE), nociceptin/orphanin-FQ (N/OFQ), H-Tyr–Tic–Phe–Phe-OH (TIPP).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Paul, A.K.; Smith, C.M.; Rahmatullah, M.; Nissapatorn, V.; Wilairatana, P.; Spetea, M.; Gueven, N.; Dietis, N. Opioid Analgesia and Opioid-Induced Adverse Effects: A Review. Pharmaceuticals 2021, 14, 1091. https://doi.org/10.3390/ph14111091

AMA Style

Paul AK, Smith CM, Rahmatullah M, Nissapatorn V, Wilairatana P, Spetea M, Gueven N, Dietis N. Opioid Analgesia and Opioid-Induced Adverse Effects: A Review. Pharmaceuticals. 2021; 14(11):1091. https://doi.org/10.3390/ph14111091

Chicago/Turabian Style

Paul, Alok K., Craig M. Smith, Mohammed Rahmatullah, Veeranoot Nissapatorn, Polrat Wilairatana, Mariana Spetea, Nuri Gueven, and Nikolas Dietis. 2021. "Opioid Analgesia and Opioid-Induced Adverse Effects: A Review" Pharmaceuticals 14, no. 11: 1091. https://doi.org/10.3390/ph14111091

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop